content
stringlengths
1
15.9M
\section{#1}} \def\Subsection#1{{\vskip-0.8cm}\hbox{\,}\subsection{#1}} \journalid{337}{15 January 1989} \articleid{11}{14} \lefthead {DE OLIVEIRA-COSTA {{\frenchspacing\it et al. \frenchspacing}}} \righthead{DIFFUSE GALACTIC EMISSION} \begin{document} \title{Cross-correlation of Tenerife data with Galactic templates --- evidence for spinning dust?} \author{ Ang\'elica de Oliveira-Costa$^{1,2}$, Max Tegmark$^{1,3}$, Carlos M. Gutierrez$^{4}$, Aled W. Jones$^{5}$, R. D. Davies$^{6}$, A. N. Lasenby$^{5}$, R. Rebolo$^{4}$ \& R. A. Watson$^{6,4}$} \begin{abstract} The recent discovery of dust-correlated diffuse microwave emission has prompted two rival explanations: free-free emission and spinning dust grains. We present new detections of this component at 10 and 15 GHz by the switched-beam Tenerife experiment. The data show a turnover in the spectrum and thereby supports the spinning dust hypothesis. We also present a significant detection of synchrotron radiation at 10 GHz, useful for normalizing foreground contamination of CMB experiments at high-galactic latitudes. \end{abstract} \keywords{cosmic microwave background -- diffuse radiation -- radiation mechanisms: thermal and non-thermal -- methods: data analysis} \section{INTRODUCTION} Understanding the diffuse microwave emission from the Galaxy is crucial for doing cosmology with Cosmic Microwave Background (CMB) anisotropies. Although three components of Galactic emission have been firmly identified (synchrotron~ and free-free radiation, and thermal emission from dust particles), it is important to better quantify their frequency dependence and spatial distribution (see, {\frenchspacing\it e.g.}, Kogut {\frenchspacing\it et al. \frenchspacing} 1996a; Tegmark {\frenchspacing\it et al. \frenchspacing} 1999 and references therein). Cross-correlations of CMB data with far-IR maps have shown the existence of a microwave emission component whose spatial distribution is traced by these maps (Kogut {\frenchspacing\it et al. \frenchspacing} 1996b; Lim {\frenchspacing\it et al. \frenchspacing} 1996; de Oliveira-Costa {\frenchspacing\it et al. \frenchspacing} 1997; Leitch {\frenchspacing\it et al. \frenchspacing} 1997; de Oliveira-Costa {\frenchspacing\it et al. \frenchspacing} 1998, hereafter dOC98). Although this emission component has a spectral index suggestive of free-free emission (Kogut 1999), the correlations between H$\alpha$ and dust maps have been found to be marginal (McCullough 1997; Kogut 1997). The source of this correlated emission is therefore an open question. Recent work suggests that it originates from spinning dust grain emission (Draine and Lazarian 1998), which should have a spectral index of $-3.3 < \beta_{spin} < -4$ between 19 and 53~GHz. Since spinning dust grains have a predicted turndown in their emission at frequencies below 20~GHz, a cross-correlation analysis with lower frequency measurements may help to discriminate between free-free and spinning dust emission models. The purpose of this letter is to evaluate the Galactic contribution in the Tenerife 10 and 15~GHz data sets by cross-correlating them with the DIRBE dust maps and with the Haslam and Reich and Reich synchrotron~ maps. \bigskip {\small{ \noindent $^{1}$Institute for Advanced Study, Olden Lane, Princeton, NJ 08540; <EMAIL> \noindent $^{2}$Princeton University, Dept. of Physics, Princeton, NJ 08544 \noindent $^{3}$Hubble Fellow \noindent $^{4}$Instituto de Astrofisica de Canarias, 38200 La Laguna, Tenerife, Spain \noindent $^{5}$Mullard Radio Astronomy Observatory, Cavendish Laboratory, Madingley Road, Cambridge, CB3 0HE, UK \noindent $^{6}$University of Manchester, Nuffield Radio Astronomy Laboratories, Jodrell Bank, Macclesfield, Cheshire, SK11 9DL, UK }} \noindent Our analysis is based on the Tenerife measurements made up until the end of 1997 (Guti\'errez {\frenchspacing\it et al. \frenchspacing} 1999, hereafter G99). The Tenerife switched-beam experiment is a sky survey between $0^{\circ} \le {\rm RA} \le 360^{\circ}$ and $30^{\circ} \le \delta \le 45^{\circ}$ (Figure~1 shows a sample strip at $\delta=32^\circ$) carried out at an angular resolution of $~ 5.1^{\circ}$ FWHM with an instrument that uses a double-differencing technique. Data was taken at frequencies 10 and 15~GHz (we omit the 33~GHz data here due to lack of low latitude sky coverage), which are treated separately in order to gain additional frequency information on Galactic emission. \vspace{-2cm} \centerline{{\vbox{\epsfxsize=9cm\epsfbox{fig1new.ps}}}} \vspace{-2cm} \noindent{\small Fig.~1 --- Scan of the Tenerife region $0^{\circ} < {\rm RA} < 360^{\circ}$ for $\delta = 32^{\circ}$. The thin curve shows Tenerife 15~GHz data, while the thicker curve represents DIRBE 100$\mu$m~ convolved with the Tenerife triple-beam, rescaled for best fit. The spikes at RA$\sim 80^{\circ}$ and $300^{\circ}$ correspond to the Galactic plane crossings. \label{figTen1} } \section{METHOD} The Tenerife data consists of $N=2880$ pixels at each frequency with double-differenced sky temperatures $y_i$ and noise $n_i$. We assume that this data is a linear superposition of CMB fluctuations $x_{CMB}^i$ and $M$ Galactic components whose angular distributions are traced in part by external foreground templates. Writing these contributions as $N$-dimensional vectors, we obtain \beq{signals} {\bf y} = {\bf X}{\bf a} + {\bf x}_{CMB} + {\bf n}, \end{equation} \goodbreak \noindent where ${\bf X}$ is an $N\times M$ matrix whose rows contain the various foreground templates convolved with the Tenerife triple-beam ({\frenchspacing\it i.e.}, ${\bf X}_{ij}$ would be the $i^{th}$ observation if the sky had looked like the $j^{th}$ foreground template), and ${\bf a}$ is a vector of size $M$ that gives the levels at which these foreground templates are present in the Tenerife data. We treat ${\bf n}$ and ${\bf x}_{CMB}$ as uncorrelated random vectors with zero mean, and the matrix ${\bf X}$ as constant; thus the data covariance matrix is given by \beq{varCMB} {\bf C} \equiv \expec{{\bf y} {\bf y}^T} - \expec{{\bf y}} \expec{{\bf y}^T} = \expec{{\bf x}_{CMB} {\bf x}_{CMB}^T} + \expec{{\bf n} {\bf n}^T}. \end{equation} The Tenerife noise covariance matrix $\expec{{\bf n} {\bf n}^T}$ is, to good approximation, diagonal (see Guti\'errez de la Cruz {\frenchspacing\it et al. \frenchspacing} 1995), while the CMB covariance matrix has the form \beq{CMBcovarEq} \expec{{\bf x}_{CMB} {\bf x}_{CMB}^T}_{ij} = \sum_{k=-1}^1 \sum_{l=-1}^1 w_k w_l ~ c(\hat{\rbf}_{ik} \cdot \hat{\rbf}_{jl}), \end{equation} where the vector ${\bf w}\equiv(-0.5,1,-0.5)$ gives the weights of the three lobes in the Tenerife triple beam, $\hat{\rbf}_{i0}$ is the direction that the central lobe pointed to during the $i^{th}$ observation, and $\hat{\rbf}_{i,\pm 1}$ is at the same declination but offset by $\pm 8.1^\circ$ on the sky. We take the CMB correlation function to be \beq{corrFunc} c(\hat{\rbf}_{ik} \cdot \hat{\rbf}_{jl}) = c(\cos\theta) \equiv \sum_{\ell=2}^\infty B_\ell^2 C_\ell P_\ell(\cos\theta) \end{equation} for a flat power spectrum $C_\ell\propto 1/\ell(\ell+1)$ normalized to a Q$_{flat}=(5C_2/4\pi)^{1/2}=20${$\mu$K} (G99). Finally, we model the Tenerife beam as a Fisher function (Fisher {{\frenchspacing\it et al. \frenchspacing}} 1987) \beq{rot2} B(\hat{\rbf}\cdot\hat{\rbf}') = \frac{\exp(\hat{\rbf} \cdot\hat{\rbf}' / \sigma^2)} {4 \pi \sigma^2\sinh (\sigma^{-2})} \end{equation} with FWHM$=\sqrt{8\ln 2}\sigma=5.1^\circ$, which gives a window function $B_\ell\approx e^{-\sigma^2\ell(\ell+1)/2}$. The Fisher function is locally Gaussian and integrates to unity over the unit sphere. Since our goal is to measure ${\bf a}$, both ${\bf n}$ and ${\bf x}_{CMB}$ act as unwanted noise in \eq{signals}. Minimizing $ \chi^2 \equiv ({\bf y} - {\bf X}{\bf a})^T {\bf C}^{-1} ({\bf y} - {\bf X}{\bf a}) $ yields the minimum-variance estimate of ${\bf a}$, \beq{alpha} \albfHat = \left[{\bf X}^T ~ {\bf C}^{-1} ~ {\bf X}\right]^{-1} {\bf X}^T ~ {\bf C}^{-1} ~ {\bf y}, \end{equation} with covariance matrix \beq{varalpha} {{\bf\Sigma}}\equiv\expec{\albfHat^2} - \expec{\albfHat}^2 = \left[{\bf X}^T ~ {\bf C}^{-1} ~ {\bf X}\right]^{-1}. \end{equation} The error bars on individual correlations are therefore $\Delta \widehat{a}_i={\bf\Sigma}_{ii}^{1/2}$. This includes the effect of chance alignments between the CMB and the various template maps, since the CMB anisotropy term is incorporated in $\expec{{\bf x}_{CMB} {\bf x}_{CMB}^T}$. \section{DATA ANALYSIS \& RESULTS} Due to the double-differencing technique, the Tenerife data are insensitive to the monopole ($\ell=0$) and dipole ($\ell=1$). Accordingly, when we convolve the template maps with the Tenerife triple beam function, we are removing the mean of the templates as well as large angular scale structure. As a consequence, our results depend predominantly on the small scale intensity variations in the templates ($\ell\sim 10-30$) and are insensitive to the zero levels and gradients in the CMB and the template maps. \bigskip {\footnotesize\center{Table~1. -- Correlations for 10 and 15~GHz data.} \vspace{-0.1cm} \begin{center} \begin{tabular}{llrrrr} \hline \multicolumn{1}{c}{$b$ \& $\nu$} & \multicolumn{1}{c}{Template$^{(a)}$} & \multicolumn{1}{r}{$\albfHat\pm\Delta\albfHat^{(b)}$}& \multicolumn{1}{r}{${\albfHat\over\Delta\albfHat} $}& \multicolumn{1}{c}{$\sigma_{Gal} $}& \multicolumn{1}{c}{$\Delta T$ [$\mu$K]$^{(c)}$}\\ \hline \hline $|b| > 20^{\circ}$ &100$\mu$m~ & 49.8$\pm$11.2 &{\bf 4.5 }& 0.8 & 38.1$\pm$ 8.5 \\ 10 GHz &Has & 29.5$\pm$ 5.9 &{\bf 5.0 }& 1.0 & 30.8$\pm$ 6.1 \\ &R\&R & 0.8$\pm$ 0.3 &{\bf 2.8 }&23.2 & 17.8$\pm$ 6.4 \\ \hline &100$\mu$m~ & 71.8$\pm$ 4.5 &{\bf 5.9 }& 0.7 & 52.8$\pm$ 3.3 \\ 15 GHz &Has & 0.8$\pm$ 3.4 & 0.2 & 1.1 & 0.8$\pm$ 3.7 \\ &R\&R & 0.1$\pm$ 0.2 & 0.3 &24.6 & 1.3$\pm$ 3.7 \\ \hline \hline $|b| > 30^{\circ}$ &100$\mu$m~ & $-$8.3$\pm$31.9 &$-$0.3 & 0.2 & $-$1.9$\pm$ 7.2 \\ 10 GHz &Has & 35.8$\pm$ 8.8 &{\bf 4.1 }& 0.9 & 31.1$\pm$ 7.6 \\ &R\&R & 0.5$\pm$ 0.4 & 1.2 &19.4 & 9.3$\pm$ 7.8 \\ \hline &100$\mu$m~ & 94.9$\pm$15.3 &{\bf 6.2 }& 0.3 & 24.9$\pm$ 4.0 \\ 15 GHz &Has & $-$8.3$\pm$ 5.0 &$-$1.7 & 0.9 & $-$7.3$\pm$ 4.4 \\ &R\&R & $-$0.3$\pm$ 0.2 &$-$1.4 &19.8 & $-$6.0$\pm$ 4.4 \\ \hline \hline $|b| > 40^{\circ}$ &100$\mu$m~ & 84.6$\pm$54.8 & 1.5 & 0.2 & 14.6$\pm$ 9.6 \\ 10 GHz &Has & 24.1$\pm$11.4 &{\bf 2.1 }& 0.8 & 19.4$\pm$ 9.1 \\ &R\&R & 0.4$\pm$ 0.5 & 0.6 &19.1 & 6.7$\pm$10.5 \\ \hline &100$\mu$m~ & 72.4$\pm$33.3 &{\bf 2.2 }& 0.2 & 12.3$\pm$ 5.7 \\ 15 GHz &Has &$-$10.8$\pm$ 6.3 &$-$1.7 & 0.8 & $-$8.7$\pm$ 5.1 \\ &R\&R & $-$0.4$\pm$ 0.3 &$-$1.1 &19.5 & $-$7.0$\pm$ 6.2 \\ \hline \end{tabular} \end{center} } \vspace{-0.1cm} \noindent{\small $^{(a)}$ The DIRBE and Haslam correlations listed in this table correspond to joint 100$\mu{\rm m}-$Has fits, whereas the R\&R numbers correspond to a joint 100$\mu{\rm m}-$R\&R fit. \\ $^{(b)}$$\albfHat$ has units $\mu$K~ (MJy/sr)$^{-1}$ for the 100$\mu$m~ template, $\mu$K/K for the Has template and $\mu$K/mK for the R\&R template. We use antenna temperature throughout. \\ $^{(c)}$ $\Delta T \equiv (\albfHat\pm\Delta\albfHat) \sigma_{Gal}$.} \bigskip \subsection{Correlations \& Variances}\label{corrANDvar} We cross-correlate the Tenerife data with two different synchrotron~ templates: the 408~MHz survey (Haslam {\frenchspacing\it et al. \frenchspacing} 1981) and the 1420~MHz survey (Reich and Reich 1988), hereafter Has and R\&R, respectively. To study dust and/or free-free emission, we cross-correlate the Tenerife data with three Diffuse Infrared Background Experiment (DIRBE) sky maps at wavelengths 100, 140 and 240~$\mu{\rm m}$ (Boggess {\frenchspacing\it et al. \frenchspacing} 1992). The extent of point source contamination in the Tenerife data is discussed and estimated in G99, and therefore will not be addressed in this $Letter$. In practice, we just remove the estimated point sources contribution before calculating the correlations. For definiteness, we use the DIRBE 100$\mu{\rm m}$ channel when placing all limits below since it is the least noisy of the three DIRBE channels, and the Haslam map since it is the synchrotron~ template at lowest frequency. Table~1 shows the coefficients $\albfHat$ and the corresponding fluctuations in antenna temperature in the Tenerife data ($\Delta T = \albfHat \sigma_{Gal}$, where $\sigma_{Gal}$ is the standard deviation of the template map). The analysis is done for three different cuts: $20^{\circ}, 30^{\circ}$ and the {\it Tenerife cut} (which consists of data with $160^{\circ} < {\rm RA} < 250^{\circ}$, corresponding to Galactic latitudes $|b| \mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$} 40^{\circ}$). Note that the fits are done jointly for $M=2$ templates. The DIRBE and Haslam correlations listed in Table~1 correspond to joint 100$\mu{\rm m}-$Has fits, whereas the R\&R numbers correspond to a joint 100$\mu{\rm m}-$R\&R fit. Statistically significant ($>2\sigma$) correlations are listed in boldface. The two synchrotron~ templates are found to be correlated with the 10~GHz data at low Galactic latitudes and this correlation persists with the Has template even at higher latitudes, while the 15~GHz data is not correlated with these templates even for a $20^{\circ}$ cut. The 100$\mu{\rm m}$ and 10~GHz data are correlated only at lower Galactic latitudes ($20^{\circ}$ cut), while the correlation with the 15~GHz data persists to higher latitudes. The same analysis was carried out for the DIRBE 140$\mu{\rm m}$ and 240$\mu{\rm m}$ maps, giving very similar results. \subsection{Latitude Dependence} To investigate the dependence of the correlation on Galactic latitude, we sliced the maps into six regions of equal area, each corresponding to a range of latitude $|b|$. Figure~2 shows the results for the 100$\mu{\rm m}$ map. Note that $\albfHat$ from the 100$\mu{\rm m}$$-$15~GHz correlation is similar in all latitude bands, suggesting that the particular ISM properties that are responsible for the correlation do not vary strongly with latitude. \vspace{-2.0cm} \centerline{{\vbox{\epsfxsize=9cm\epsfbox{slicesnew3.ps}}}} \vspace{-1.5cm} \noindent{\small Fig.~2 --- Dependence of $\albfHat$ on Galactic latitude for the 100$\mu{\rm m}$$-$correlated emission at 15 GHz. \label{figTen2} } \bigskip \subsection{Spectral Index} Writing the frequency dependence as $\delta T\propto\nu^{\beta}$, we obtain interesting limits on the spectral index $\beta$. For synchrotron radiation, both the Has$-$10~GHz and the R\&R$-$10~GHz correlations are consistent with $\beta\sim -3$ in all latitude slices. The most favored values between 408~MHz and 10~GHz values ($-3.2 \mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$} \beta \mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$} -3.4$) are slightly steeper than the canonical sub-GHz slope of $-2.7\mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$} \beta \mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$} -2.9$ (Davies {{\frenchspacing\it et al. \frenchspacing}} 1998; Platania {{\frenchspacing\it et al. \frenchspacing}} 1998). Such a steepening of the spectrum at higher frequencies is consistent with a steepening of the spectrum of cosmic ray electrons at higher energies (Rybicki and Lightman 1979). For the emission component correlated with the 100$\mu$m map, the spectral index between 10 and 15~GHz is $0.80^{+.77}_{-.66}$ for $|b| > 20^{\circ}$, and the spectrum is also clearly rising at $|b| \mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$} 30^{\circ}$ (see Table~1). Calibration uncertainties are below 5\% in the Tenerife data at both 10 and 15 GHz, and therefore do not affect the conclusions that we will draw from this spectral behavior. \subsection{Effect of Template Correlations}\label{multifrequencyfit} \Eq{varalpha} shows that we can interpret ${\bf\Sigma}^{-1}$ as measuring correlations between the various templates. Table~2 shows the dimensionless correlation coefficients $r \equiv {\bf\Sigma}^{-1}_{ij} ({\bf\Sigma}^{-1}_{ii}{\bf\Sigma}^{-1}_{jj})^{-0.5}$, using the 10 GHz pixel weighting, and enables a number of conclusions to be drawn: \begin{enumerate} \item The three DIRBE maps are, as expected, highly correlated, {\frenchspacing\it i.e.}, they trace the same warm dust emission; \item The two synchrotron~ maps are less well correlated, with only $r^2\sim 50\%$ of the variance in the Has map being traced by the R\&R map at $|b| > 20^{\circ}$ cut, and even less at higher latitudes\footnote{ Systematic effects (such as striping) are present in the synchrotron~ radio maps and may be partially responsible for the uncorrelated part of their signal (Davies {{\frenchspacing\it et al. \frenchspacing}} 1998). In addition, detector noise in the template maps will of course also reduce their correlation with each other and with the Tenerife data, but is in our case negligible compared to the Tenerife noise.}; and \item All three DIRBE maps are seen to be almost uncorrelated with both radio maps, with the common variance $r^2$ being only a few percent. This means that ${\bf\Sigma}$ is almost diagonal, and that simply performing a separate correlation analysis for each component (with $N=1$, as was done in {\frenchspacing\it e.g.}, de Oliveira-Costa {{\frenchspacing\it et al. \frenchspacing}} 1997; dOC98) will give almost identical results to a joint ($N>1$) analysis. We tested this, and indeed obtained results in good agreement with those presented in subsection \ref{corrANDvar}). \end{enumerate} \vspace{-0.1cm} {\footnotesize\center{Table~2. -- Correlation between foreground templates$^{(a)}$.} \vspace{-0.1cm} \begin{center} \begin{tabular}{lcccccc} \hline & &100$\mu$m~ &140$\mu$m~ &240$\mu$m~ &Has &R\&R \\ &&&&&&\\ 100$\mu$m~ & &1 &0.97 &0.96 &0.12 &0.14 \\ 140$\mu$m~ & &0.91 &1 &0.98 &0.08 &0.11 \\ 240$\mu$m~ & &0.91 &0.88 &1 &0.07 &0.11 \\ Has & &0.11 &0.12 &0.11 &1 &0.71 \\ R\&R & &0.15 &0.11 &0.17 &0.53 &1 \\ \hline \end{tabular} \end{center} } \vspace{-0.1cm} \noindent{\small $^{(a)}$ The upper right triangle of the table uses a $|b| > 20^{\circ}$ cut and the lower left triangle a $|b| > 30^{\circ}$ cut.} \medskip \subsection{Fake Skies}\label{fakeskies} We tested our software by analyzing constrained realizations of the CMB and the Tenerife instrument noise. Cholesky decomposing the covariance matrix as ${\bf C} = {\bf L} {\bf L}^T$, we generated fake skies using ${\bf y} = {\bf L} {\bf z} + {\bf X}{\bf a}$, where ${\bf z}$ is a vector of independent Gaussian random variables with $\expec{{\bf z}}={\bf 0}$ and $\expec{{\bf z}\zbf^t}={\bf I}$ (the identity matrix), which gives $\expec{{\bf y} {\bf y}^T}-\expec{{\bf y}}\expec{{\bf y}}^t = {\bf L} \langle {\bf z} {\bf z}^T \rangle {\bf L}^T = {\bf L} {\bf I} {\bf L}^T = {\bf C}$. Analyzing 1000 such realizations, we recovered unbiased estimates $\albfHat$ with a variance in agreement with \eq{varalpha}. \subsection{Sky Rotations} Systematic errors or unmodeled foregrounds may add non-Gaussian fluctuations to the the data set ${\bf y}$, making the true error bars larger than suggested by \eq{varalpha}. To address this issue, we repeated the fit done in \eq{alpha} after replacing the template map with a set of ``control'' patches selected from different portions in the sky. We repeated the analysis with 24 $\times$ 2 $\times$ 2 = 96 transformed maps, rotated around the Galactic axis by multiples of 15$^{\circ}$ and/or flipped vertically and/or horizontally. The correct Haslam template has the highest of all 96 correlations with the 10~GHz data. Even for a $|b| \mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$} 40^{\circ}$ cut, 95 of the 96 patches (or 99\% of them) are less correlated with the the 10~GHz data than the correct 100$\mu{\rm m}$ patch. The same test was carried out for other Galactic cuts and the 15~GHz data, giving similar results. These results show that our correlations are not due to systematic errors or chance alignments between the CMB and the various template maps. We also applied the rotation test to the foreground template correlations in Table~2. Although as mentioned above, the dust-synchrotron correlations are tiny ($r\ll 1$) and negligible for our purposes, they are still statistically significant; for a 40$^{\circ}$ Galactic cut, the correct synchrotron template (Has or R\&R) has the highest of all 96 correlations with the DIRBE map. \section{CONCLUSIONS} We have detected high-latitude galactic foreground signals in the Tenerife data at high ($\mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$} 5\sigma$) statistical significance, summarized in Figure~3. The synchrotron signal is consistent with previous upper limits (Kogut {{\frenchspacing\it et al. \frenchspacing}} 1996b) and detections (dOC98), and provides a useful normalization for modeling microwave foregrounds. We also verify that the 15 GHz {\it Tenerife-cut} data is useful for CMB cosmology; although the DIRBE-correlated component is statistically significant, it is much smaller in amplitude than the primary cosmological signal. Our results indicate that the DIRBE-correlated signal may turn out to be the dominant foreground seen by the the MAP satellite. What is its physical origin? A ``smoking gun'' indication of spinning dust grains would be a turnover in the spectrum, since all Draine and Lazarian models show such a feature whereas the antenna temperature of free-free emission continues to rise with $\beta\sim -2.1$ towards lower frequencies. Figure~3 combines our present results with previous measurements. The lower panel shows that the synchrotron spectrum is well fit by a power law $\nu^\beta$ with $\beta\sim -3$ (cyan line). In the upper panel, the dashed curve shows the spectrum for a best fit linear combination of vibrational (green) and rotational (yellow) dust emission. The vibrational spectrum assumes a 20K dust temperature and an emissivity of 2. The rotational spectrum is the warm interstellar medium model of Draine and Lazarian (1998) from {\it www.astro.princeton.edu/$\sim$draine/}. A joint three-parameter fit, including free-free emission with spectrum $\nu^{-2.15}$ as a third component (magenta), prefers negligible amounts of free-free above 10 GHz. On purely physical grounds, free-free emission must of course be present at some level. However, if the 10 GHz Tenerife correlation were entirely due to free-free emission, this component would explain only about 10\% of the variance observed at 15 GHz and a percent of the variance in the Saskatoon data. The interpretation of such fits that directly compare data points from different experiments is complicated by conversions between different angular scales (dOC98). However, a direct comparison between the 10 and 15 GHz Tenerife points from this work is straightforward, since the angular resolution and window functions are the same, and gives a spectral index of $\beta=0.80^{+.77}_{-.66}$. This is clearly inconsistent with free-free emission, and provides evidence that the main culprit is indeed spinning dust. \bigskip The authors wish to thank Bruce Draine, Lyman Page, David Spergel and David Wilkinson for helpful comments. Support for this work was provided by NASA though grants NAG5-6034 and Hubble Fellowship HF-01084.01-96A (from STScI, operated by AURA, Inc. under NASA contract NAS5-26555), and NSF grant PHY-9600015. AWJ acknowledges King's College, Cambridge, for support in the form of a Research Fellowship. \medskip \centerline{{\vbox{\epsfxsize=9cm\epsfbox{fig3new.ps}}}} \noindent{\small Fig.~3 --- Frequency dependence of DIRBE-correlated emission (top) and Haslam-correlated emission (bottom). The DIRBE-correlated emission is seen to be approximately fit by a combination of spinning and vibrating dust (dashed curve), whereas free-free emission alone (falling straight line) cannot explain the drop from 15 to 10 GHz. The slope of the Haslam-correlated emission is seen to be weekly constrained, fitting {e.g.} a single $\beta=-3.26$ power law down to 408 MHz (where $\alpha=10^6\mu$K$/$K$=1$ by definition). Tenerife data is from Table 1 for a $20^\circ$ cut. Upper limits are 2-$\sigma$. Note that the error bars on the 15 GHz point (top) are too small to be visible. \label{figTen3} } \bigskip
\section{Introduction} An outstanding problem in extra-galactic astronomy is understanding the parameters that determine the structure and evolution of galaxies. A first step towards understanding the physical properties of galaxies is classification based on morphology. Spiral galaxies were first classified by Hubble according to the size of the bulge, the tightness with which the spiral arms are wound, and the resolution of individual \hii regions (\markcite{Hubble1936}Hubble 1936; \markcite{devauc1959} de Vaucouleurs 1959). Several observational studies have elucidated differences among galaxies along the Hubble sequence (for a review see \markcite{Roberts1994} Roberts \& Haynes 1994; \markcite{Kennicutt1998} Kennicutt 1998). Observations plus modeling of broad band colors of galaxies reveal that the bulge dominated Sa galaxies are red compared to the disk dominated Sc galaxies, suggesting an older population \markcite{Larson1974}(Larson \& Tinsley 1974). Similarly, \markcite{Roberts1969}Roberts(1969) measured the atomic gas content of 75 spiral galaxies and found that the ratio, M(HI)/L(B), decreases systematically from the late-type spirals to the early-type Sa galaxies, suggesting early types to be deficient in hydrogen gas, compared to the late types. The star formation history of galaxies can be traced using the integrated \ha equivalent widths, where the \ha emission line flux is normalized by the past star formation rate through the red continuum. \markcite{KK1983}Kennicutt \& Kent (1983) measured the \ha equivalent widths for $\sim 200$ spiral galaxies and demonstrated that the \ha equivalent widths decrease systematically from late-type spirals to early-type spiral galaxies, suggesting early-types to be deficient in massive young stars compared to the late types. There are other observational results, however, which suggest that early-type spirals are not so quiescent. A recent Hubble Space Telescope(HST) study of the bulges of 75 spiral galaxies (\markcite{Carollo1998}Carollo {\it et al.} 1998) reveals a wide variety of activity, including star formation, hidden underneath the bulges of early-type spirals. \markcite{YK1989}Young \& Knezek (1989) have showed that the dominant phase of the interstellar medium in Sa-Sab types is molecular, not atomic and that the molecular fraction is much higher in the early-types compared to the later types. The result, however, has been recently challenged by \markcite{Casoli1998}Casoli {\it et al.} (1998), who find that molecular gas comprises only about one third to one fourth of the total gas content of spirals of types Sa through Sc. A recent analysis of the Infrared Astronomical Satellite (IRAS) database by \markcite{DH1997}Devereux \& Hameed (1997) suggests that the global massive star formation rates, as determined by 60 micron luminosity functions, are comparable in early and late-type spirals. Similarly, far infrared to blue luminosity ratios of a large sample of nearby spiral galaxies do not show any morphological dependence (\markcite{Tomita1996}Tomita {\it et al.} 1996; \markcite{DH1997}Devereux \& Hameed 1997). Evidently, the IRAS data has revealed a previously unsuspected population of early-type spirals with high massive star formation rates. The IRAS results, do not support previous claims, based on \ha equivalent widths, that massive star formation rates increase along the Hubble sequence from Sa to Sc(\markcite{Kennicutt1983}Kennicutt 1983; \markcite{Kennicutt1994}Kennicutt {\it et al.} 1994). Part of the problem is that the sample of early-type spirals selected by \markcite{Kennicutt1983}Kennicutt (1983) and \markcite{Kennicutt1994}Kennicutt (1994) is small in number and is biased towards galaxies with low values of L(FIR)/L(B) (\markcite{DH1997}Devereux \& Hameed 1997; \markcite{Usui1998}Usui {\it et al.} 1998). We are, therefore, conducting an \ha imaging survey of {\it all known} nearby early-type spiral galaxies in order to better understand the differences between the IRAS results and those of existing \ha studies. High resolution \ha images of nearby galaxies provide important information about the morphology and luminosity of the ionized hydrogen gas. Surprisingly few \ha images of early-type spirals exist in the published literature. The general notion that early-type spirals do not have significant massive star formation is, at least partially, responsible for the dearth of \ha observations (\markcite{Young1996}Young {\it et al.} 1996). The continuum morphology of early-type spirals is dominated by an 'inert' stellar bulge which can hide star forming complexes that lie underneath them. CCD imaging allows the hidden \hii regions to be revealed by subtracting the overwhelming continuum light. Despite the dearth of \ha images, our appreciation of the heterogeneous nature of early-type spirals has evolved considerably in the past quarter century. \markcite{vandenbergh1976}Van den Bergh(1976) and \markcite{Kormendy1977}Kormendy(1977) found no \hii regions in NGC 4594 and NGC 2841, leading them to speculate that the IMF in early-type spirals may be biased against the formation of massive stars. Later studies, however, showed that \hii regions are indeed present in these particular galaxies (\markcite{Schweizer1978}Schweizer 1978; \markcite{HK1983}Hodge \& Kennicutt 1983; \markcite{Kennicutt1988}Kennicutt 1988). A detailed study of \hii regions in the disks of seven Sa galaxies by \markcite{Caldwell1991}Caldwell {\it et al.} (1991) found that \hii regions are quite abundant in the disks but are significantly smaller than those in late-type spirals. Specifically, \markcite{Caldwell1991}Caldwell {\it et al.} found that there are no \hii regions in the disks of early-type spirals with luminosities $ >10^{39} erg s^{-1}$. The purpose of the present paper is to report that \hii regions are not only abundant in early-type spirals but some contain giant \hii regions that are comparable in size and luminosity to giant \hii regions seen in late-type spirals. Our findings support recent \ha observations by \markcite{Young1996}Young {\it et al.} (1996) and \markcite{Usui1998}Usui {\it et al.} (1998), who have also identified numerous early-type spirals with star formation rates comparable to the most prolifically star forming late-type spirals. In order to quantify the diverse star forming capabilities of early-type spirals, we are conducting a systematic program of \ha imaging. The results are presented here for twenty seven galaxies imaged to date. The sample is described in section 2 and the observations are described in section 3. The results are presented in section 4, followed by discussion in section 5. \section{The Sample} An \ha imaging survey is being conducted to investigate the star forming capabilities of nearby early-type (Sa-Sab) spiral galaxies. The target galaxies have been selected from the Nearby Galaxy Catalog (NBG)(\markcite{Tully1988}Tully 1988) which is the largest complete compilation of bright galaxies with velocity less than $3000 km/s$, corresponding to a distance of 40 Mpc ($H_{0}=75 kms^{-1}/Mpc$). The target galaxies are listed in Table 1 with some useful observables. Distances in the NBG catalog are based on a Virgo-centric in-fall model (\markcite{Tully1988}Tully 1988). The NBG catalog also lists morphological types for each galaxy The complete sample includes all (57) bright, m(B) $\le$ 12.1 magnitude, non-interacting early-type (Sa-Sab) spirals known within 40 Mpc. For the purposes of the present work, ``interacting'' galaxies are defined as those which have cataloged companions within 6\arcmin of each other. The goal of the survey is to image the complete sample of nearby early-type spirals. Imaging the largest complete sample will circumvent incompleteness corrections and minimize statistical errors in quantifying the incidence of nuclear starbursts, nuclear emission line spirals, nuclear point sources, and other morphological peculiarities in nearby early-type spirals. So far we have imaged twenty-one galaxies, the results of which are presented in this paper. As part of a complimentary study we are also obtaining \ha images of twenty-one far-infrared luminous early-type spirals, identified by \markcite{DH1997} Devereux and Hameed (1997). Fifteen of these galaxies have m(B) $\le$ 12.1 and, hence, are already included in the complete sample described above. The remaining six galaxies have been imaged and they are included in this paper also. \section{Observations} Observations of northern hemisphere galaxies were obtained with the Astrophysical Research Consortium (ARC) 3.5m telescope at Apache Point Observatory (APO) in New Mexico. Southern hemisphere galaxies were observed with the 1.5m telescope at Cerro Tololo Inter-American Observatory (CTIO) in Chile. Six northern hemisphere early-type spiral galaxies were observed using the Double Imaging Spectrograph (DIS) at APO between August 1996 and January 1997. The Texas Instruments CCD chip has a pixel scale of $0.61\arcsec pixel^{-1}$ and a $4.2\arcmin$ field of view. Two red-shifted narrow band H$\alpha$ + [NII] (6570\AA \& 6610\AA, $\Delta\lambda$ = 72\AA) filters and a line free red continuum (6450\AA,$ \Delta\lambda$ = 120\AA) filter were used to obtain the line and continuum images, respectively. Three exposures were obtained through each of the line and continuum filters. Details of the observations are summarized in Table 2. Twenty-one southern hemisphere galaxies were observed using the Cassegrian Focus CCD Imager (CFCCD) on the CTIO 1.5m telescope. CFCCD uses a $2048 \times 2048$ Tektronics chip and has a pixel scale of $0.43\arcsec pixel^{-1}$ at f/7.5, yielding a field of view of $14.7' \times 14.7'$. All of the galaxies were imaged using the narrow band H$\alpha$ + [NII] filter at 6606\AA ($\Delta\lambda$ = 75\AA) with the exception of NGC 5728 for which we used the H$\alpha$ + [NII] filter at 6649\AA ($\Delta\lambda$ = 76\AA). The narrow band line free red continuum filter for all the galaxies was centered at 6477\AA ($\Delta\lambda$ = 75\AA). Three exposures of 900 seconds were obtained through each of the line and the continuum filters for all the southern hemisphere galaxies (see Table 2). \subsection{Data Reduction} The Image Reduction Analysis Facility (IRAF) software package was used to process the images. The images were bias subtracted, and then flat fielded using twilight flats taken on the same night as the galaxy data. Sky subtraction was achieved by measuring the sky level around the galaxy and then subtracting it from the images. Images were then registered and median combined to improve the S/N ratio and eliminate cosmic rays. The continuum image was then scaled to the line plus continuum image by measuring the integrated fluxes of several ($\ge 10$) stars common to both images. The final H$\alpha$ image was obtained by subtracting the scaled continuum image from the H$\alpha$+continuum image to remove foreground stars and the galaxy continuum. The H$\alpha$ images were flux calibrated using observations of the standard stars G1912B2B and BD $+28\arcdeg 4211$ (\markcite{Massey1988}Massey {\it et al.} 1988) for the northern hemisphere galaxies and LTT 3218, LTT 1020, and LTT 7987 (\markcite{Hamuy1994}Hamuy {\it et al.} 1994) for the southern hemisphere galaxies (see Table 2). The observing conditions were photometric, with $<10 \%$ variations in the standard star fluxes, for all of the imaging observations. There are several factors that contribute to the uncertainties in the determination of \ha flux measurements. Large systematic errors may be introduced by the continuum subtraction procedure. Small random errors are introduced by read noise in the detector and photon noise but these are negligible compared to the uncertainty introduced by the continuum subtraction procedure. There are two factors that contribute to the uncertainty in determining the continuum level. First, one assumes that the foreground stars used for estimating continuum level have the same intrinsic color as the galaxy. Second, one must assume that the galaxy has the same color everywhere. Both assumptions are unlikely to be correct. Unfortunately the uncertainty in the \ha flux depends sensitively and non-linearly on the continuum level subtracted. It has been determined empirically that $2\%-4\%$ errors in the continuum level correspond to $10\%-50\%$ errors in the \ha flux, depending on the relative contribution of the continuum light. The \ha fluxes are presented in Table 3 along with the calculated \ha luminosities. The \ha fluxes include contributions from the satellite [NII] lines at $\lambda\lambda$ 6548,6584 and have not been corrected for Galactic or internal extinction.\markcite{KK1983} Kennicutt and Kent (1983) suggest that the average extinction in their sample is typically $\sim$1 mag. The \ha extinction, however, is expected to be more for galaxies that have high inclinations or those that harbor nuclear starbursts. We have not applied extinction corrections due to the uncertainties involved in determining its true value. Consequently the \ha fluxes and luminosities presented in Table 3 are lower limits to the intrinsic \ha fluxes and luminosities. \subsection{Comparison with Previous Measurements} Seven of the galaxies presented in this paper already have published \ha fluxes. Figure 1 compares our measured values in the same apertures as those in the literature. The agreement between the measurements is, in general, good. NGC 1433 and NGC 7552, in particular, provide a good test as both ours and the comparison observations(\markcite{Crocker1996}Crocker {\it et al.} 1996 and \markcite{Lehnert1995}Lehnert \& Heckman 1995 respectively) were made with the 1.5m telescope at CTIO and the measurements are in good ($<11\%$) agreement with each other. The biggest discrepancy is with NGC 1022, where our \ha flux is $68\%$ higher than the value published by \markcite{KK1983}Kennicutt \& Kent (1983). However, our measured flux for NGC 1022 is within $8\%$ of the value obtained by \markcite{Usui1998}Usui {\it et al.} (1998). \section{Results} \subsection{Classification of Early-type Spirals} A study of the \hii region luminosity functions in the disks of seven Sa galaxies by \markcite{Caldwell1991}Caldwell {\it et al.}(1991) showed that while \hii regions are abundant, there are none with \ha luminosities $> 10^{39} erg s^{-1}$. Our results confirm that the \ha luminosity of \hii regions in most early-type spirals is less than $10^{39} erg s^{-1}$. However, we also find that a significant fraction ($15-20\%$) of early-type spirals do have at least one \hii region in the disk with $L_{H\alpha} \ge 10^{39} erg s^{-1}$. The latter result is not necessarily in contrast to Caldwell's study for two reasons. First, Caldwell {\it et al.} had a small sample containing only seven galaxies and could have easily missed early-type spirals with giant \hii regions. Second, their sample contained only Sa galaxies, whereas early-type spirals discussed in this paper include both Sa and Sab types. Early-type spirals have been divided in two categories based on the \ha luminosity of the largest \hii region present in the disk. The \ha luminosity of the individual \hii regions in the disks of all category 1 galaxies is less than $10^{39} erg s^{-1}$, whereas category 2 galaxies contain at least one \hii region in the disk with $L_{H\alpha} \ge 10^{39} erg s^{-1}$. Figure 2 shows the range of global \ha luminosities for the two categories of Sa-Sab galaxies. Perhaps not surprisingly, category 2 galaxies dominate at high, $L_{H\alpha}>1.7 \times 10^{41}\lum$, luminosities. A two-tailed Kolmogorov-Smirnov(KS) test indicates that the two categories of galaxies are not derived from the same population at a confidence level greater than $99\%$. Similarly, Figure 3 shows the range of \ha equivalent widths, where the continuum-free line flux is normalized by the red continuum, for the two categories of early-type spirals. The K-S test again reveals that the two categories of galaxies are not derived from the same population at a confidence level greater than $99\%$. In order to verify that the difference in the distribution of global \ha luminosities between the two categories is not related to the nuclear properties, the histograms of nuclear(1 kpc) \ha luminosities are compared in Figure 4. The K-S test reveals that the difference between the distributions in Figure 4 is insufficient to reject the null hypothesis that the two categories are drawn from the same population. Thus the difference between the two categories of early-type spirals is primarily a global phenomenon unrelated to their nuclei. In addition to the statistical tests described above, we varied the \ha luminosity of the largest \hii region in each galaxy by 20$\%$ to further verify the dichotomy in the properties of category 1 and 2 early-type spirals. Indeed, a 20 percent variation, which represents a typical uncertainty in \ha measurements of individual \hii regions, has no effect on the distribution of galaxies between the two categories. We do want to stress, however, that the purpose of this paper is to introduce early-type spirals with giant \hii regions and to show that Sa-Sab galaxies are heterogeneous in nature. The sub-classification of early-type spirals into two categories has been done solely to investigate the conditions that may lead to the formation of luminous \hii regions. The continuum and continuum-subtracted \ha images are shown in Figures 5, 6, \& 7 and the observed properties of Sa-Sab galaxies are described in more detail below. \begin{center} {\it Category 1 Early-type Spirals} \end{center} \hii regions in category 1 galaxies are either small or totally absent from the spiral arms (see Figure 5). By definition, all \hii regions in the disk of category 1 spirals have $L _{H\alpha}< 10^{39}erg s^{-1}$. Consistent with earlier studies (\markcite{Kennicutt1988}Kennicutt 1988; \markcite{Caldwell1991}Caldwell {\it et al.} 1991; \markcite{Bresolin1997}Bresolin \& Kennicutt 1997), we find that \hii regions in most early-type spirals contain only a few massive stars, whereas \hii regions in late-type spirals can contain hundreds or even thousands of stars. The seventeen galaxies included in category 1 resemble what most astronomers identify as classical early-type spirals. Morphologically, most category 1 galaxies appear undisturbed in the continuum image, but the \ha images reveal very diverse nuclear properties. \noindent {\it Extended nuclear emission line region (ENER):} There are seven category 1 galaxies (NGC 1350, NGC 1371, NGC 1398, NGC 1433, NGC 1515, NGC 1617, NGC 3169) in which an extended nuclear emission line region(ENER) has been detected. The detailed morphology of the ENER gas is difficult to discern at the distances of these galaxies, but they appear to be similar to the nuclear emission line spirals discovered in the centers of M81 (\markcite{Jacoby1989}Jacoby {\it et al.} 1989; \markcite{Devereux1995}Devereux {\it et al.} 1995) and M31 (\markcite{Jacoby1985}Jacoby {\it et al.} 1985; \markcite{Devereux1996}Devereux {\it et al.} 1996). It is clear that the filamentary emission line gas is quite different from the clumpy \hii regions in the spiral arms. The ENER is most likely shock-ionized or photo-ionized by UV radiation from bulge post asymptotic giant branch stars (\markcite{Devereux1995}Devereux {\it et al.} 1995; \markcite{Heckman1996}Heckman 1996). All category 1 galaxies hosting extended nuclear emission line regions have the lowest nuclear \ha luminosities. \noindent {\it Nuclear Point Sources:} Of the remaining category 1 galaxies, five (NGC 2273, NGC 5188, NGC 5728, NGC 7172, NGC 7213) have an unresolved \ha source at the nucleus. Apart from 5188, all four galaxies have been spectroscopically identified as Seyferts(NGC 2273: \markcite{Ho1997a}Ho {\it et al.} 1997a; NGC 5728: \markcite{Phillips1983}Phillips {\it et al.} 1983; NGC 7172: \markcite{Sharples1984}Sharples {\it et al.} 1984; NGC 7213: \markcite{Filippenko1984}Filippenko \& Halpern 1984). NGC 2273, NGC 5728 and NGC 7172 have faint \hii regions in their disk, whereas NGC 7213 has a number of \hii regions in the circumnuclear region, but the disk is almost devoid of any \hii regions. NGC 7213 also has an extended nuclear emission line region region surrounding the point source. NGC 5188 has prominent \hii regions and its nucleus has been spectroscopically classified as an \hii region by \markcite{VV1986}Veron-Cetty \& Veron (1986). \noindent {\it Nuclear starbursts:} The nuclei in four of the remaining category 1 galaxies (NGC 3471, NGC 1482, NGC 3885, NGC 1022) are resolved and contribute more than $50\%$ of the total \ha luminosity. NGC 1022(\markcite{Ashby1995}Ashby {\it et al.} 1995), NGC 3471(\markcite{Balzano1983}Balzano 1983) and NGC 3885 (\markcite{Lehnert1995}Lehnert \& Heckman 1995) have been spectroscopically identified as nuclear starbursts. No spectral confirmation is available in the literature for NGC 1482 but it has been included as a nuclear starburst galaxy based on similarities in morphology and \ha luminosity to NGC 1022, NGC 3471, and NGC 3885. It is worthwhile to mention that all four galaxies also have very high far infrared luminosities, $\ge 10^{10}L_{\sun}$ (\markcite{DH1997}Devereux \& Hameed 1997), comparable to the prototypical starburst galaxy M82 (\markcite{Rieke1980}Rieke {\it et al.} 1980). Dust lanes are prominent in all four galaxies and there are virtually no \hii regions in their disks. NGC 3717 has a high inclination and could not be classified into any of the above sub-groups. Spectroscopically, the nucleus of NGC 3717 has been classified as an \hii region by \markcite{VV1986}Veron-Cetty \& Veron (1986). \begin{center} {\it Category 2 Early-type Spirals} \end{center} Eight of the galaxies presented in this paper have at least one disk \hii region with $L_{H\alpha} \ge 10^{39} erg s^{-1}$ defining a new category of early-type spirals (see Figure 6). Dust lanes or tidal tails are present in all category 2 galaxies. Despite the fact that extinction is an important factor in these dusty galaxies, early-type spirals with the highest \ha luminosity belong exclusively to category 2(Figure 2). Additionally, all category 2 galaxies have far-infrared luminosities in excess of $10^{10} L_{\sun}$ (\markcite{DH1997}Devereux \& Hameed 1997). The discovery of a significant number of early-type spirals with \ha luminosities comparable to those produced by the most prolifically star forming late-type spirals is surprising, as early-type spirals are widely believed to have massive star formation rates that are considerably lower. Based on the available images and far-infrared luminosities of the entire sample, it is estimated that category 2 galaxies could represent $15\%-20\%$ of all early-type spirals in the nearby($D \le 40 Mpc$) universe, a significant group that can no longer be ignored as oddities. The continuum images of two category 2 galaxies (NGC 986, NGC 7552) reveal the possible existence of previously uncataloged dwarf galaxies that appear to be interacting with the larger galaxies. Redshift information is not yet available for either of the dwarf galaxies so the physical association cannot be confirmed. Nevertheless, if the dwarf galaxy seen at the end of the northern spiral arm in NGC 7552 is indeed associated with NGC 7552, it has an absolute magnitude of -12.95 in R which is comparable to some of the dwarf galaxies seen around M31 and the Milky Way. There is no detectable \ha emission from the dwarf companion of NGC 7552. In the case of NGC 986, the dwarf galaxy is located at the end of the north-eastern spiral arm and appears to be in the process of being tidally disrupted. The distorted shape of the dwarf along with the contribution of stars in the spiral arms of NGC 986 makes it difficult to measure its magnitude with any certainty. Nevertheless, the presence of interacting dwarfs exclusively in category 2 galaxies may be related to the existence of giant \hii regions in these galaxies. Two category 2 galaxies (NGC 7582, NGC 6810) have been optically classified as Seyferts. The presence of Seyferts in both categories of early-type spirals suggest that the incidence of Seyfert activity is unrelated to the phenomenon that is responsible for the global differences between category 1 and category 2 galaxies. \begin{center} {\it Unclassified} \end{center} Surprisingly, NGC 660 and NGC 2146 apparently have no disk \hii regions with $L_{H\alpha} \ge 10^{39}\lum$ but both have high far infrared luminosity (\markcite{DH1997}Devereux \& Hameed 1997) and a morphology that appears to be even more disturbed than any of the category 2 galaxies (see Figure 7). The apparent absence of luminous \hii regions may be a consequence of the 13 and 9 magnitudes of extinction in the V band for NGC 660 and NGC 2146 respectively (\markcite{Young1988}Young {\it et al.} 1988). Nevertheless, we have not classified these two galaxies as they cannot be unambiguously placed into either category. \subsection{\hii Region Luminosity Functions} In order to understand some of the properties of \hii regions in the two categories of early-type spirals, we derived \ha luminosity functions for the \hii regions in a category 1 galaxy, NGC 1398, and a category 2 galaxy, NGC 7552. The two galaxies were chosen because they are at a comparable distance (16.1 Mpc \& 19.5 Mpc for NGC 1398 \& NGC 7552 respectively), they have relatively low inclinations, and they provide a good contrast between the two categories of early-type spirals. \noindent{\it Measurement of \hii region \ha fluxes} Individual \hii regions in the two galaxies were identified computationally using the DAOfind routine incorporated in an IDL program,'{\rm H{\small II}{\it phot}}', developed by \markcite{Thilker1999}Thilker {\it et al.} 1999. The program identifies local maxima by using Gaussian kernels comparable, and slightly larger than the effective resolution, on the original image and also on several smoothed versions. Earlier studies of luminosity functions have mostly assumed a symmetrical shape for the \hii regions.(e.g. \markcite{Kennicutt1989}Kennicutt {\it et al.} 1989; \markcite{Caldwell1991}Caldwell {\it et al.} 1991). In nature, however, \hii regions are observed to be asymmetrical. The attribute of {\rm H{\small II}{\it phot}} is that it provides the freedom for \hii regions to acquire any shape. The \ha flux was measured by summing the pixel values within each region. The local background level for each region was determined by the mode of the 'region-hugging' annulus having a specified physical width. The details of the '{\rm H{\small II}{\it phot}}' program will be published elsewhere (\markcite{Thilker1999}Thilker {\it et al.} 1999). \noindent{\it Luminosity Functions} The observed \ha fluxes were converted into \ha luminosities using the distances for NGC 1398 and NGC 7552 listed in Table 1. The striking difference between the \ha images of NGC 1398 and NGC 7552 (see Figures 5 and 6 respectively) is also reflected in the widely dissimilar \hii region luminosity functions as illustrated in Figure 8. In both cases, the {\it bright} \hii regions have a luminosity function that is well represented by a power-law: $$ N(L) \propto L^{\alpha}dL $$ with exponent values of -2.4($\pm 0.2$) and -2.0($\pm 0.1$) for NGC 1398 and NGC 7552 respectively. Our slope for NGC 1398 is in good agreement with the slope obtained by \markcite{Caldwell1991}Caldwell {\it et al.} for the same galaxy, whereas our value is steeper for NGC 7552 than the -1.7 slope determined by \markcite{Feinstein1997}Feinstein (1997). The two categories of early-type spirals, as defined in this paper, are based on the \ha luminosity of the largest \hii region in the disk. The \hii region luminosity functions show that \hii regions, in general, are more luminous in the category 2 galaxy, NGC 7552, as compared to the category 1 galaxy, NGC 1398. We also, cautiously, report the difference in the shape of the two luminosity functions. NGC 1398 has a steep power law distribution. The \hii region luminosity function for NGC 7552, on the other hand, is a Gaussian with a turnover luminosity located at least a factor of six higher than in NGC 1398. The two galaxies are at about the same distance and were observed on the same night with the 1.5m telescope at CTIO for the same integration time. Similarly the data was reduced in an identical manner and \hii region luminosity functions were determined with '{\rm H{\small II}{\it phot}}' using the same parameters for the two galaxies. So there is little doubt that the the difference in the shapes of the two luminosity functions is real and confirms the visual impression given by \ha images of NGC 1398 and NGC 7552 in Figures 5 and 6 respectively. Thus the difference in global \ha luminosities between the two categories of early-type spirals is not attributable to just one anomalous \hii region, but rather suggests intrinsic differences in the ensemble of \hii regions found in the two categories of early-type spirals. A subsequent paper will present a more detailed analysis of the \hii region luminosity functions for nearby early-type spirals. \section{Discussion} Early-type spirals have, in the past, been associated with low massive star formation rates. The results presented in this paper, however, indicate that early-type spirals are a heterogeneous group of galaxies with a significant number of them showing star forming regions that are comparable in luminosity to those found in late-type spirals. Early-type spirals have been divided into two categories based on the \ha luminosity of their largest \hii region in the disk. \hii regions in all category 1 galaxies have $L_{H\alpha} < 10^{39}\lum$, whereas category 2 early-type spirals have at least one \hii region with $L_{H\alpha} \ge 10^{39}\lum$. The purpose of the following discussion is to elaborate on the diversity of early-type spirals as revealed by the \ha and the continuum images. \subsection{\it{Nuclear Diversity in Category 1 Galaxies }} Most of the category 1 early-type spirals appear to be morphologically undisturbed. Despite the similarities in the continuum image, they have very diverse nuclear \ha properties ranging from nuclear starbursts to low luminosity extended nuclear emission line regions (ENERs) to unresolved point sources. A recent, comprehensive, spectroscopic survey of the nuclei of nearby galaxies by \markcite{Ho1997}Ho {\it et al.} (1997) revealed a similarly wide variety of spectroscopic phenomena including \hii nuclei in $22\%$, LINERs in $36\%$, and Seyfert nuclei in $18\%$ of early-type spirals. In addition, they have found that LINERs and Seyferts reside most frequently in early-type spiral galaxies. Unfortunately, we are presently unable to correlate all of the images presented in this paper with Ho's spectroscopic survey as most of our images are for the southern hemisphere whereas Ho's survey is for the northern hemisphere. However, we do have spectroscopic classifications for the nuclei of 17 galaxies that can be used to determine the correspondence between spectroscopy and morphology (Table 4). All category 1 galaxies with unresolved \ha nuclear point sources have been optically classified as Seyferts except NGC 5188, which is classified as ``H'' by \markcite{VV1986}Veron Cetty \& Veron (1986). Spectroscopic information is available for only 4 ENER category 1 early-type spirals but all 4 have a LINER or a Seyfert-like spectrum, classified as ``N'' by \markcite{VV1986}Veron-Cetty \& Veron (1986). The spectra of Seyfert-like, ``N'', galaxies have faint \ha and [NII] lines, $H\alpha < 1.2 \times [NII] \lambda 6583$, and no other detectable emission lines. The poor S/N of the data, make it impossible to distinguish between a Seyfert 2 and a LINER (\markcite{VV1986}Veron-Cetty \& Veron 1986), however, we suspect that all ENER category 1 galaxies with ``N'' classification are LINERs. Conversely, we also suspect that not all LINERs are AGNs, as advocated recently by \markcite{Ho1997}Ho {\it et al.} (1997). The re-discovery of a possible correspondence between LINERs and extended emission line gas in the nuclei of spiral galaxies is not surprising. \markcite{Keel1983a}Keel (1983a), in an \ha imaging survey of the nuclei of spiral galaxies hosting LINER spectra, found extended emission in most cases. In addition, he found that LINER type spectra are ubiquitous in the nuclei of spiral galaxies not containing an \hii region of much higher emission luminosity, which would make the low ionization region undetectable if present (\markcite{Keel1983b}Keel 1983b). Similarly, we have been able to detect ENERs only in galaxies with few or no \hii regions in the nuclear region. M31, the nearest spiral galaxy, hosts an extended nuclear emission line region(\markcite{RF1971}Rubin \& Ford 1971; \markcite{Jacoby1985}Jacoby {\it et al.} 1985) and provides an excellent opportunity to study the circumnuclear region in detail. The ENER in M31 spans 2 kpc (\markcite{Devereux1994}Devereux {\it et al.} 1994) and has a complex filamentary structure that is similar to the ENERs observed in more distant category 1 early-type spirals. The spectroscopic analysis of the ENER in M31 shows that it is indeed a LINER (\markcite{Heckman1996}Heckman 1996). The absence of a UV nuclear point source rules out an AGN as a source for photoionization of the emission line gas in M31. Similarly, HST observations have shown that the nuclear spiral in M31 is not ionized by massive stars (\markcite{King1992}King {\it et al.} 1992; \markcite{Devereux1994}Devereux {\it et al.} 1994). Post-asymptotic giant branch(PAGB) stars provide a possible source of ionization for the gas in the nuclear region (\markcite{Devereux1995}Devereux {\it et al.} 1995; \markcite{Heckman1996}Heckman 1996) and \markcite{Binette1994}Binette {\it et al.} (1994) have shown that photoionization by PAGB stars can lead to a LINER spectrum. Thus a LINER spectrum does not necessarily betray an AGN as advocated recently by Ho {\it et al.} (1996). M81 is the nearest category 1 Sa-Sab spiral and it hosts a low luminosity LINER/Seyfert 1 nucleus (\markcite{FS1988}Filippenko \& Sargent 1988; \markcite{Ho1996}Ho {\it et al.} 1996). The presence of a broad line region(\markcite{PT1981}Peimbert \& Torres-Peimbert 1981), a compact X-ray (\markcite{Fabbiano1988}Fabbiano 1988; \markcite{Petre1993}Petre {\it et al.} 1993) and radio source (\markcite{Beitenholz1996}Beitenholz {\it et al.}), and a UV nuclear source (\markcite{Devereux1997}Devereux {\it et al.} 1997) strongly argue for the presence of a low luminosity AGN (\markcite{Ho1996}Ho {\it et al.} 1996; \markcite{Maoz1995}Maoz {\it et al.} 1995). High resolution \ha imaging has also revealed an extended nuclear emission line spiral which is strikingly similar to the one observed in M31 (\markcite{Jacoby1989}Jacoby {\it et al.} 1989; \markcite{Devereux1995}Devereux {\it et al.} 1995). The spiral is not ionized by the spectroscopically identified AGN, as the Seyfert contributes $< 5\%$ of the total \ha luminosity of the spiral (\markcite{Devereux1995}Devereux {\it et al.} 1995; \markcite{Devereux1997}Devereux {\it et al.} 1997). Furthermore, the extended emission line gas, like M31, has a LINER spectrum(Heckman, private communication). Recent HST observations have shown that the nuclear spiral in M81 is not ionized by massive stars (\markcite{Devereux1995}Devereux {\it et al.} 1995, \markcite{Devereux1997}1997). The source of ionization for the nuclear spiral is unknown, but it is most likely shock-ionized or photo-ionized by UV radiation from bulge post asymptotic giant branch stars (\markcite{Devereux1995}Devereux {\it et al.} 1995). M81 is likely to be a composite object; a truly compact AGN surrounded by an extended emission line region like the one in M31. The two nearest LINERS, M31 and M81, illustrate the need for both spectroscopic and imaging observations in order to distinguish between true AGNs and ENERs with spectra that look like AGN. The spectroscopic survey conducted by \markcite{Ho1997a}Ho {\it et al.} (1997a) included most of our northern hemisphere early-type spirals, which we are in the process of imaging. The new images will allow us to determine the extent to which the spectroscopically identified AGNs are responsible for ionizing the extended emission line gas seen under the bulges of early-type spirals. \subsection{\it{Morphological Peculiarities in Category 2 Galaxies}} In recent years it has been suggested that interactions may play an important role in the formation and subsequent evolution of early-type spirals (e.g. \markcite{Schweizer1990}Schwiezer 1990, \markcite{Pfenniger1991}Pfenniger 1991, \markcite{Zaritsky1995}Zaritsky 1995). Furthermore, there is growing observational evidence to suggest that early-type spirals are perhaps the most dynamic of all the nearby galaxy systems. A recent HST survey of the bulges of nearby spiral galaxies has revealed that a significant fraction ($\sim 40\%$) of early-type spirals show little or no morphological evidence for a smooth, $R^{1/4}$ law (\markcite{Carollo1998}Carollo {\it et al.} 1998). Indeed, \markcite{SS1988}Schwiezer \& Seitzer (1988) discovered ripples around some early-type spirals indicating that a major accretion event occurred in the past 1-2 Gyr. Similarly, the discovery of counter-rotating gas and star disks in the early-type spirals NGC 4826 (\markcite{Braun1992}Braun {\it et al.} 1992) and NGC 7217 (\markcite{MK1994}Merrifield \& Kuijken 1994) provide additional evidence for past interactions. Perhaps galaxies classified as category 2 are the early-type spirals interacting in the current epoch. Note that we had excluded all galaxies that had cataloged companions within 6\arcmin and yet we still see signs of morphological disturbance in most of the category 2 galaxies. The discovery of previously unknown inter-loping dwarf galaxies in two of the category 2 early-type spirals provides direct observational evidence for on-going interactions. Other morphological peculiarities, such as tidal tails and dust lanes, are present in all category 2 galaxies providing further indirect evidence for interactions. Numerous \ha and far-infrared studies indicate that interactions elevate massive star formation rates(e.g. \markcite{Bushouse1987}Bushouse 1987, \markcite{Kennicutt1987}Kennicutt {\it et al.} 1987, \markcite{LK1995}Liu \& Kennicutt 1995, \markcite{Young1996}Young {\it et al.} 1996). All of the category 2 galaxies have far-infrared luminosities in excess of $ 10^{10}L_{\sun}$, and populate the high end of the local \ha luminosity function, both measures indicative of high rates of massive star formation that is clearly seen in \ha images in the form of giant \hii regions. It is interesting to note that the distinction between the two categories of early-type spirals is based only on the luminosity of the largest \hii region in the disk, and yet this simple segregation leads to other clear differences in the range of global \ha luminosities, galaxy morphology, and the range of individual \hii region luminosities. The presence of morphological peculiarities in a significant fraction of early-type spirals may be linked to the idea of forming big bulges through minor mergers. According to the numerical simulations of \markcite{MH1994}Mihos \& Hernquist (1994), the merger of a dwarf galaxy with a disk galaxy can lead to the formation of a stellar bulge. A series of such mergers or accretion events could form the large bulges found in early-type spirals (\markcite{Pfenniger1993}Pfenniger 1993; \markcite{MH1994}Mihos \& Hernquist 1994; \markcite{RT1997}Rich \& Terndrup 1997). Classical models of galaxy formation (\markcite{Eggen1962}Eggen {\it et al} 1962) predict bulges of galaxies to be old and metal rich. While the Galactic Bulge provides support for this theory, it may not be representative of all bulges. Recent observations of bulges have revealed them to be diverse and heterogeneous in nature (for a review see \markcite{Wyse1997}Wyse {\it et al.} 1997). Work on color gradients by \markcite{BP1994}Balcells \& Palatier (1994) show that there is little or no color change between the disk and bulge for some early-type spirals. Bulge populations seem to resemble their parent disks, suggesting that bulges may not be significantly older. Similarly, blue bulges observed in several S0 galaxies (\markcite{Schweizer1990}Schweizer 1990) (NGC 5102, NGC 3156, IC 2035 etc.) lend further support to the idea of bulge formation through minor mergers. Further work on stellar populations is needed to understand and quantify the diverse properties exhibited by the bulges in nearby early-type spiral galaxies. \subsection{{\it Influence of Bars on Early-type Spirals}} Several studies have been conducted to determine the influence of bars on star formation rates (e.g. \markcite{Hawarden1986}Hawarden {\it et al.} 1986; \markcite{Devereux1987}Devereux 1987; \markcite{Puxley1988}Puxley {\it et al} 1988; \markcite{Arsenault1989}Arsenault 1989; \markcite{RD1994}Ryder \& Dopita 1994; \markcite{Huang1996}Huang {\it et al.} 1996; \markcite{Tomita1996}Tomita {\it et al.} 1996; \markcite{Ho1997b}Ho {\it et al} 1997b ). Whereas, \ha and far-infrared studies of large galaxy samples (\markcite{RD1994}Ryder \& Dopita 1994; \markcite{Tomita1996}Tomita {\it et al.} 1996 respectively) find no correlation of global star formation with the presence of bars, numerical models suggest that stellar bars can strongly perturb gas flows in disk, and trigger nuclear star formation. Indeed, nuclear star formation rates are measured to be preferentially higher for galaxies that have bars, especially in early-type spirals. \markcite{Devereux1987}Devereux (1987) measured strong nuclear 10$\mu m$ emission in 40$\%$ of barred early-type spirals, indicating high star formation rates. A similar excess was not observed in late-type barred spirals. To provide some measure of comparison the {\it nuclear} star formation rates in some barred early-type spirals are comparable to the globally integrated star formation rates of late-type spirals. Enhanced nuclear star formation in early-type spirals may be related to the location of the inner Lindblad resonance (ILR) (\markcite{Devereux1987}Devereux 1987) which is dependent on the central mass distribution. The ILR in bulge-dominated Sa-Sab galaxies is expected to be located well inside the bar and close to the nucleus (\markcite{EE1985}Elmegreen \& Elmegreen 1985), allowing a stellar bar to transfer gas from the disk region into the nuclear region more efficiently, prompting a nuclear starburst. However, the presence of a bar in an early-type spiral does not necessarily mean enhanced nuclear star formation. NGC 1398 and NGC 1433 (Figure 5) have prominent bars and yet they have little or no massive star formation in the nuclear region. Similarly, bars are not essential for the presence of a nuclear starburst as is illustrated by the starburst in the nucleus of non-barred galaxy, NGC 3885. \ha observations of a large sample of early-type spirals provide an excellent opportunity to study the effects of bars on these bulge-dominated Sa-Sab galaxies. Unfortunately, with our current data, we don't yet have enough statistics to analyze the influence of bars with any certainty. In our sample of 27 galaxies presented here, 17 have bars (including intermediate cases), 5 don't have bars, and no information is available for the remaining 5 galaxies (see Table 4). We will discuss early-type spirals in the context of bars in a later paper when we have obtained \ha images for the complete sample of nearby early-type spirals. \subsection{{\it Star Formation in Early-type Spirals}} The star forming properties of galaxies can also be quantified using \ha equivalent widths, where the \ha flux is normalized by the red light representing the older stellar population. In essence, the \ha equivalent width measures the ratio of current star formation to past star formation. Figure 9 compares \ha equivalent widths for the galaxies presented in this paper with the early-type spirals observed by \markcite{KK1983}Kennicutt \& Kent (1983) and \markcite{Usui1998}Usui {\it et al.} (1998). Contrary to popular perception, \ha equivalent widths of early-type spirals presented in this paper reveal a significant fraction of galaxies with high ratios of present to past star formation. The comparison in Fig. 9 suggests that \markcite{KK1983}Kennicutt \& Kent's sample contains early-type spirals with preferentially low massive star formation rates. This has also been noted by \markcite{Usui1998}Usui {\it et al.} who have obtained \ha fluxes of galaxies with high L(fir)/L(blue) luminosity ratios and their results are also shown in Figure 9. The images presented in this paper reveal early-type spirals to encompass a wide range of \ha equivalent widths spanning that measured by both Kennicutt \& Kent and Usui {\it et al}. In an attempt to reconcile the differences in \ha equivalent widths between this study and \markcite{KK1983}Kennicutt \& Kent (1983), we have examined the morphological classifications of the galaxies. We have noted that \markcite{KK1983}Kennicutt \& Kent used classifications from the Revised Shapley-Ames Catalog (RSA) (\markcite{ST1981}Sandage \& Tammann 1981). On the other hand classifications in the NBG catalog (\markcite{Tully1988}Tully 1988), which are used for this paper, are derived mostly from the Second Reference Catalog (RC2) (\markcite{devauc1976} de Vaucouleurs {\it et al.} 1976). Table 5 lists the morphology of the galaxies included in the present study as they appear in the Tully catalog (\markcite{Tully1988}Tully 1988), the Second Reference Catalog (RC2) (\markcite{devauc1976}de Vaucouleurs 1976), and the Revised Shapley-Ames Catalog (\markcite{ST1981}Sandage \& Tammann 1981). Part of the difference in the \ha equivalent widths measured for early-type spirals by us and \markcite{KK1983}Kennicutt \& Kent can be traced to a difference in galaxy classification. Almost all of the category 2 Sa-Sab galaxies have been classified as Sb's in the RSA catalog and would have been regarded as such in the study of \markcite{KK1983}Kennicutt \& Kent. The systematic difference between RC2 and RSA classifications is surprising and may be traced to subtle variations in the application of the criteria used for classifying spiral galaxies. The classification criteria for the RC2 are explained in {\it Handbuch der Physik} (\markcite{devauc1959}de Vaucouleurs 1959), whereas RSA classifications are elaborated in {\it the Hubble Atlas of Galaxies} (\markcite{Sandage1961}Sandage 1961). In the RC2 catalog, sub-types of spirals, Sa-Sc, are defined by the ``{\it relative importance of the nucleus (decreasing from a to c) and the degree of unwinding and resolution of the arms (increasing from a to c)}'' (\markcite{devauc1959}de Vaucouleurs 1959). Spiral galaxies in RSA catalog are classified according to the same criteria: ``{\it the openness of the spiral arms, the degree of resolution of the arms into stars, and the relative size of the unresolved nuclear region}'' (\markcite{Sandage1961}Sandage 1961). However, the third criterion (bulge size) was not considered important for actual classifications in RSA; ``{\it For many years it was thought that the third classification criterion of the relative size of the unresolved nuclear region usually agreed with the criterion of the arms. Inspection of large numbers of photographs shows that, although there is a general correlation of the criteria, there are Sa galaxies that have small nuclear regions. The assignment of galaxies to the Sa, Sb, or Sc type is based here primarily on the characteristics of the arms}'' (Sandage 1961). The preceding sentence may provide an explanation for the classification difference found for category 2 galaxies in the RC2 and RSA catalogs. The differences can be quite extreme, for example, NGC 7552 is considered a prototypical SBab galaxy by \markcite{devauc1959}de Vaucouleurs (1959) but an SBbc by Sandage in the RSA catalog(\markcite{ST1981}Sandage \& Tammann 1981). Over the years, many studies have been conducted analyzing star forming properties of spiral galaxies along the Hubble sequence. \markcite{KK1983}Kennicutt \& Kent (1983), \markcite{dejong1984}de Jong (1984), \markcite{Sandage1986}Sandage (1986), \markcite{RL1986}Rieke \& Lebofsky (1986), \markcite{PR1989}Pompea \& Rieke (1989), \& \markcite{Caldwell1991}Caldwell {\it et al.} (1991) used morphological classifications from RSA catalog and found a dependence of star formation on Hubble types Sa-Scd. However, this is not totally surprising since RSA classifications are primarily based on the arm characteristics, which are tightly correlated with star formation. On the other hand, \markcite{DY1990}Devereux and Young (1990), \markcite{IF1992}Isobe \& Fiegelson (1992), \markcite{Tomita1996}Tomita {\it et al.} (1996), \markcite{Young1996}Young {\it et al.} (1996), Devereux and Hameed (1997), and Usui {\it etal} (1998) have used RC2 classifications and find that star formation rates in spiral(Sa-Scd) galaxies are independent of Hubble type. Similar results were obtained by \markcite{Gavazzi1986}Gavazzi (1986) and \markcite{Bothun1989}Bothun {\it et al.}(1989), who used classifications given in the UGC catalog. Thus it appears that perceptions concerning star formation rates among spiral galaxies along the Hubble sequence depends on the choice of catalog. Furthermore, it may be that other Hubble-type correlations also depend on the choice of catalog. We do not want to claim that the classifications of one catalog are better than the other. However, we do want to stress the importance of the NBG catalog in the study of nearby galaxies. As described earlier, the NBG catalog is the most complete compilation of nearby galaxies. It contains all known galaxies within 40 Mpc that are brighter than 12th magnitude. Any systematic study aimed at understanding the properties of nearby galaxies will rely on the NBG catalog. Results presented in this paper have exposed the subjective nature of the prevailing classification schemes. Perhaps it is time to start classifying galaxies quantitatively using measures of their physical properties. \section{Summary \& Conclusions} Preliminary results of an on-going \ha imaging survey of nearby, bright, early-type (Sa-Sab) spirals have revealed them to be a heterogeneous class of galaxies. Furthermore, our images have identified a significant number of early-type spirals with massive star formation rates comparable to the most prolifically star forming Sc galaxies. An analysis of the \ha images suggests that early-type spirals can be divided into two categories based on the \ha luminosity of the largest \hii region in the disk. The first category includes galaxies for which the individual \hii regions have \ha luminosity, $L_{H\alpha} < 10^{39} \lum$. Previously it was thought that all \hii regions in early-type spirals have $L_{H\alpha} < 10^{39}\lum$, but our new observations have revealed a second category which includes $15-20\%$ of early-type spirals. Most of the category 1 galaxies appear morphologically undisturbed. Despite the similarities in the continuum images, category 1 galaxies exhibit diverse nuclear properties. The early-type spirals with the lowest \ha luminosities host extended nuclear emission line regions whereas the galaxies that have the highest \ha luminosities contain nuclear starbursts. We also derived \hii region luminosity functions for one galaxy from each category. Our results suggest that, overall, \hii regions in category 1 early-type spirals are less luminous than \hii regions in category 2 early type spirals and that the difference between the two categories is not due to the existence of a single ``anomalous'' \hii region with $L_{H\alpha} \ge 10^{39}\lum$. Dust lanes and tidal tails are present in the continuum images of all category 2 galaxies suggesting a recent interaction, which may also explain the presence of giant \hii regions in these galaxies. The heterogeneous nature of early-type spirals, as revealed by the continuum and \ha images, illustrates the problems associated with classifying galaxies morphologically. The results presented in this paper constitute part of an on-going \ha survey to image a complete sample of nearby, bright, early-type spirals. It is anticipated that the survey will be the first to quantify the diverse nature of early-type spirals by providing a statistically significant set of observations. \acknowledgments The authors would like to thank Rene' Walterbos for providing the \ha filter set for the APO observations. Thanks also to the TACs at NOAO and APO for the generous allocation of observing time and their staffs for expert assistance at the telescopes. N.D. gratefully acknowledges the National Geographic Society for travel support to CTIO. S.H. would like to thank NOAO for travel support to CTIO, Jon Holtzman, Charles Hoopes, and Bruce Greenawalt for useful comments on the paper, and David Thilker for providing the \hii region LF program, '{\rm H{\small II}{\it phot}}'. We would also like to thank the referee, Marcella Carollo, for useful comments that improved the presentation of the paper. {\rm H{\small II}{\it phot}} was created by David Thilker under support from NASA's Graduate Student Researcher Program (NGT-51640). The IDL source code and explanatory documentation will soon be available by request. Contact <EMAIL> for details.
\section{Introduction} The random phase approximation (RPA) is widely used throughout solid state physics. It properly describes the screening when the kinetic energy is much larger than the Coulomb energy. In the strongly correlated limit, however, it is qualitatively wrong. Little is known for the intermediate regime, where kinetic and Coulomb energy are comparable, and perturbative methods fail since there is no small parameter. In such a situation quantum Monte Carlo is the method of choice. Our goal is to investigate the screening in a strongly correlated system close to the Mott transition. To be specific we focus on the doped Fullerenes, like K$_3$C$_{60}$. These materials are strongly correlated and close to a Mott transition \cite{c60mott}, but they are also superconductors with quite large transition temperatures ($T_c\approx30\,\mathrm{K}$ for Rb$_3$C$_{60}$) \cite{rmp}. The superconductivity is driven by the coupling to intramolecular phonons. These phonons mediate an effective electron--electron attraction which is, however, counteracted by the electron--electron repulsion, which is large in such strongly correlated systems. Unlike conventional superconductors, in the doped Fullerenes this repulsion is not reduced much by retardation effects. Therefore efficient screening is important for reducing the electron--electron repulsion sufficiently to allow for an electron--phonon driven superconductivity \cite{mu,screening}. Using quantum Monte Carlo we find that even for quite strong correlations the RPA gives a surprisingly accurate description of the static screening on the metallic side of a Mott transition until the system is close to the transition \cite{screening}. Besides the immediate consequences this result has for our understanding of the superconductivity in the doped Fullerides, it should also have quite general implications for the physics of systems close to a Mott transition. We start by introducing the model Hamiltonian used to describe the doped Fullerenes. We briefly discuss the L\"owdin downfolding technique that gives a general prescription for constructing low-energy Hamiltonians. Next we introduce the screening problem and discuss the results of the Monte Carlo calculations. Finally we review the quantum Monte Carlo methods used. We discuss the choice of the trial wave function and introduce a very efficient method for optimizing the Gutzwiller-type parameters. \section{Model Hamiltonian} Fullerites\index{C$_{60}$} are crystals made of C$_{60}$ molecules on an fcc lattice. They are characterized by very weak inter-molecular interactions. Therefore the discrete molecular levels merely broaden into narrow, well separated bands \cite{ldabands}. The valence band originates from the lowest unoccupied molecular orbital, which is a 3-fold degenerate $t_{1u}$ orbital\index{$t_{1u}$ orbital}. In doped Fullerenes, there are alkali atoms sitting in the space between the C$_{60}$ molecules. They do not affect the band structure around the Fermi energy very much. Only the filling of the $t_{1u}$ band changes, since each alkali atom donates its valence electron. Hence for K$_3$C$_{60}$ the (3-fold degenerate) $t_{1u}$ band is half-filled. We are mainly interested in the properties of these valence-band electrons. To simplify the description we therefore want to get rid of the other bands. They can be projected out by the L\"owdin downfolding\index{L\"owdin downfolding} technique \cite{lowdin}. The basic idea is to partition the Hilbert space into a subspace that contains the degrees of freedom that we are interested in (in our case the `$t_{1u}$-subspace') and the rest of the Hilbert space: ${\cal H}={\cal H}_0 \oplus {\cal H}_1$. We can then write the Hamiltonian of the system as \begin{equation} H=\left(\begin{array}{cc}H_{00}&H_{01}\\ H_{10}&H_{11}\end{array}\right)\;, \end{equation} where $H_{ii}$ is the projection of the Hamiltonian onto subspace ${\cal H}_i$, while the $H_{ij}$ $(i\ne j)$ contain the hybridization matrix elements between the two subspaces. Writing Green's function $G=(E-H)^{-1}$ in the same way, we can calculate the projection of $G$ onto ${\cal H}_0$ \cite{invpart}: \begin{equation} G_{00}=\Big(E-\underbrace{[H_{00}+ H_{01}\,(E-H_{11})^{-1}H_{10}]}_{H_{\rm eff}(E)}\Big)^{-1} \; . \end{equation} We see that the physics of the full system is described by an effective Hamiltonian $H_{\rm eff}(E)$ that operates on the subspace ${\cal H}_0$ only. We have, however, to pay a price for this drastic simplification: the effective Hamiltonian is energy dependent. In practice one approximates it with an energy-independent Hamiltonian $H_{\rm eff}(E_0)$. This works well if we are only interested in energies close to $E_0$. In solid C$_{60}$ we have the fortunate situation that the bands retain the character of the molecular orbitals, since the hybridization matrix elements are small compared to the energy separations of the orbitals. In fact we can neglect the other bands altogether and get the hopping matrix elements $t_{in,\,jn'}$ between the $t_{1u}$ orbitals $n$ and $n'$ on molecules $i$ and $j$ directly from a tight-binding parameterization \cite{TBparam,A4C60}. To demonstrate how well this works, Fig.\ \ref{c60bands} shows the comparison of the ab initio $t_{1u}$ band structure with the band structure obtained from the tight-binding \index{tight-binding} Hamiltonian with only $t_{1u}$ orbitals. \begin{figure}[t] \includegraphics[width=.75\textwidth]{c60band.eps} \caption[]{Band structure ($t_{1u}$ band) of solid C$_{60}$ (fcc) (a) as calculated ab initio using the local density approximation \cite{ldabands} and (b) using a tight-binding Hamiltonian with only $t_{1u}$ orbitals \cite{TBparam}.} \label{c60bands} \end{figure} A realistic description of the electrons in the $t_{1u}$ band also has to include the correlation effects which come from the Coulomb repulsion of electrons in $t_{1u}$ orbitals on the same molecule. The resulting Hamiltonian which describes the interplay of the hopping of electrons and their Coulomb repulsion has the form \begin{equation}\label{Hamil} H=\sum_{\langle ij\rangle} \sum_{nn'\sigma} t_{in,jn'}\; c^\dagger_{in\sigma} c^{\phantom{\dagger}}_{jn'\sigma} +\;U\sum_i\hspace{-0.5ex} \sum_{(n\sigma)<(n'\sigma')}\hspace{-1ex} n_{i n\sigma} n_{i n'\sigma'} . \end{equation} The on-site Coulomb interaction\index{on-site Coulomb interaction} $U$ can be calculated within density functional theory \cite{calcU}. It is given by the increase in the energy of the $t_{1u}$ level per electron that is added to one molecule of the system. It is important to avoid double counting in the calculation of $U$. While the relaxation of the occupied orbitals and the polarization of neighboring molecules have to be included in the calculation, excitations within the $t_{1u}$ band must be excluded, since they are contained explicitly in the Hamiltonian (\ref{Hamil}). The results are consistent with experimental estimates \cite{expU,lof}: $U\approx 1.2-1.4\,\mathrm{eV}$. For comparison, the width of the $t_{1u}$ band is in the range $W\approx 0.5-0.85\;eV$. To properly describe K$_3$C$_{60}$ the effect of the orientational disorder \cite{TBparam,Mazin} of the C$_{60}$ molecules in the crystal are built into the hopping matrix elements $t_{in,jn'}$. Multiplet effects are not included, since they tend to be counteracted by the Jahn-Teller effect, which is also neglected. In K$_3$C$_{60}$ the system has three electrons per molecule. In the limit of weak correlations ($U=0$), this corresponds to a metal with a half-filled conduction band. In the atomic limit ($U\to\infty$) the Coulomb energy dominates, forcing every molecule to be occupied by exactly three electrons, and suppressing any hopping. This is a Mott insulator. We therefore expect a metal-insulator transition for some finite value of the Coulomb interaction $U$. For the model Hamiltonian (\ref{Hamil}) with parameters describing K$_3$C$_{60}$ it occurs for $U\approx 1.5-1.75\,\mathrm{eV}$ \cite{c60mott}. Given the estimates for the true value of $U$, K$_3$C$_{60}$ is therefore close to the Mott transition, in a correlated metallic state. \section{Screening of a Point Charge} We now investigate how efficient the screening is in a strongly correlated system like K$_3$C$_{60}$. To be specific we analyze how a test charge $q$ sitting on one molecule is screened by the conduction electrons in the $t_{1u}$ band. To describe the influence of the test charge situated on the molecule at site $c$ we include an additional term \begin{equation}\label{Hq} H_q = q\,U\;\sum_{n\sigma} n_{cn\sigma} \end{equation} in the Hamiltonian (\ref{Hamil}). Determining the electron density at site $c$ for the system without test charge and for the system with a finite $q$ we find the screening \begin{equation} {\Delta n\over q}= {n_c(0)-n_c(q) \over q}\;. \end{equation} Let's first discuss the screening in the RPA. In the random phase approximation\index{random phase approximation} it only costs kinetic energy to screen the test charge. In the limit where a typical Coulomb integral $U$ is large compared with the band width $W$, the kinetic energy cost of screening is relatively small compared with the potential energy gain. Therefore, within the random phase approximation, the screening is very efficient for large $U$. This means that as the test charge $q$ is introduced, almost the same amount of electronic charge moves away from the site: $\Delta n \approx q$ for large $U$ (see Fig.\ \ref{extrapolate}). The random phase approximation neglects, however, that when an electron leaves a site it has to find another site with a missing electron or there is an increase in Coulomb energy of the order of $U$. Thus the RPA is accurate for small values of $U/W$, while it is {\em qualitatively wrong} for large $U/W$. It is not clear what happens when Coulomb energy $U$ and band width $W$ are comparable. \begin{figure}[bt] \rotatebox{270}{\includegraphics[width=.58\textwidth]{extrapolate.eps}} \caption[]{ Screening charge $\Delta n$ on the site of the test charge ($q=0.25\,e$) as a function of $U/W$, extrapolated to infinite cluster size. The full curve shows the screening charge in the RPA, obtained from Hartree calculations for the Hamiltonian (\ref{scrhamil}). The crosses with errorbars give the results of the QMC calculations. The RPA screening remains rather accurate up to $U/W\sim 2$, but fails badly for larger values of $U/W$. The screening is very efficient for $U/W\sim 0.5-2.0$. } \label{extrapolate} \end{figure} To address this question we have performed quantum Monte Carlo calculations for the combined Hamiltonian \begin{equation}\label{scrhamil} H=\!\sum_{\langle ij\rangle} \sum_{nn'\sigma} t_{in,jn'} c^\dagger_{in\sigma} c^{\phantom{\dagger}}_{jn'\sigma} +U\sum_i\hspace{-1.0ex} \sum_{(n\sigma)<(n'\sigma')}\hspace{-2ex} n_{i n\sigma} n_{i n'\sigma'} +q U \sum_{n\sigma} n_{cn\sigma} \end{equation} Since we are interested in the linear response, we should calculate the effect of an infinitesimally small test charge $q$. Because of the statistical error in a Monte Carlo calculation it is, however, difficult to determine the response to a small perturbation. To get a good signal-to-noise ratio, we would therefore like to use as large a test charge as possible. To estimate how large we can make $q$ and still be in the linear response regime, we have performed Lanczos calculations for a small system of 4 molecules, where exact calculations are possible. Checking the response for different test charges we find that for $q\leq 0.25\,e$ the response is practically linear. The quantum Monte Carlo calculations were then performed for large clusters of $N_\mathrm{mol}=$32, 48, 64, 72, and 108 molecules, corresponding to systems with 96, 144, 192, 216, and 324 electrons, and $q=0.25$. To extrapolate the results $\Delta n(N_\mathrm{mol})$ to infinite systems size we used a finite-size scaling of the form $\Delta n=\Delta n(N_{\mathrm{mol}})+\alpha/N_\mathrm{mol}$. The results are shown in Fig.\ \ref{extrapolate}. For rather small values of $U/W$ ($\sim 0.5-1.0$), the RPA somewhat underestimates the screening. Such a behavior is also found in the electron gas \cite{Hedin}. For intermediate values of $U/W$ ($\sim 1.0-2.0$) the RPA gives surprisingly accurate results. For larger $U/W$ the screening rapidly breaks down, and the RPA becomes qualitatively wrong, as discussed above. The efficient screening almost up to the Mott transition which occurs for $U/W\sim 2.5$ \cite{c60mott} has profound implications for the superconductivity in the alkali doped Fullerenes. We will only give qualitative arguments here, a more detailed discussion can be found in \cite{screening}. In BCS theory the superconducting transition temperature is given by \begin{equation}\label{Tc} T_c\propto e^{-1/(\lambda-\mu^\ast)}\;, \end{equation} where $\lambda=N(0)\,V$ describes the electron--phonon coupling, which mediates the effective electron--electron attraction, $\mu^\ast=N(0)\,U^\ast$ is the Coulomb pseudopotential that describes the repulsive Coulomb interaction between electrons, and $N(0)$ is the density of states at the Fermi energy. Since the electrons couple to intramolecular phonons, the coupling constant $V$ is to a good approximation a molecular property. Therefore increasing the density of states $N(0)$ will increase the electron--phonon coupling $\lambda$. This can be achieved by increasing the lattice constant of the solid Fullerene. Taking the C$_{60}$ molecules further apart will decrease the hopping matrix elements $t_{in,jn'}$ thus narrowing the $t_{1u}$ band and correspondingly increasing the density of states. If also $U^\ast$ was a molecular quantity, i.e.\ independent of the lattice constant, and if $V>U^\ast$ which, of course, is the prerequisite for superconductivity, then according to (\ref{Tc}) increasing the lattice constant would raise $T_c$. Such a variation of the transition temperature with the lattice constant is indeed found experimentally \cite{rmp}. Here the variation of the lattice constant $a$ is achieved by applying hydrostatic pressure (to reduce $a$) or by using \lq larger' alkali metals like Rb and Cs to increase the lattice constant compared to K$_3$C$_{60}$ (chemical pressure). This suggests that by inserting even more bulky ions or molecules, like NH$_3$, into the doped Fullerenes, could be increase $T_c$ even further. As it turns out, however, $T_c$ rapidly {\em decreases} with $a$ when the lattice constant becomes too large, and eventually superconductivity disappears. This drop in $T_c$ can be understood as a natural consequence of the breakdown of the screening close to the Mott transition. In the doped Fullerenes $U^\ast$ is essentially given by $U\,(1-\Delta n/q)$, where $U$ is the unscreened Coulomb matrix element (cf.\ the interaction term in the Hamiltonian) and $\Delta n/q$ describes the screening by the $t_{1u}$ electrons. $U$ is practically independent of the lattice constant, while the screening efficiency changes with the band width as shown in Fig.\ \ref{extrapolate}. For $U/W$ in the region $1.0-2.0$ the screening is almost RPA-like and does not vary very much, which means that in this region $U^\ast$ is almost independent of the lattice constant $a$. In this region $T_c$ thus increases with $a$. For even smaller band width, or correspondingly larger lattice constants, the screening rapidly becomes inefficient. Now $U^\ast$ increases with the lattice constant, leading to a decrease of $\lambda-\mu^\ast$. Hence for $a$ too large, $T_c$ rapidly decreases with $a$ and eventually vanishes. Since the screening only breaks down close to the metal-insulator transition, we have the interesting situation that {\em $T_c$ peaks close to a Mott transition!} The efficient screening found in the Monte Carlo calculations also explains why the coupling to the alkali-phonons is weak. Each C$_{60}$ molecule is surrounded by 14 alkali ions with very weak force constants. When an electron arrives on a molecule one would therefore expect that the surrounding alkali ions respond strongly. This was, however, not confirmed by experiment \cite{alkaliisotope}. That result can be naturally understood as an effect of the efficient screening: When an electron arrives on a molecule other electrons leave, effectively leaving the molecule almost neutral. The alkali ions then only see a small change in the net charge and therefore respond weakly, leading to a small electron--phonon coupling. But being molecular crystals, doped Fullerenes also have intramolecular phonons. Some of those intramolecular phonons shift the $t_{1u}$ levels in such a way that the center of gravity of the energy levels is not changed. These are the modes that are not screened by the transfer of charged. They are therefore the modes that drive superconductivity in the Fullerides. \section{Quantum Monte Carlo} We now turn to the question of how the results shown in Fig.\ \ref{extrapolate} were obtained. To keep the notation simple we will discuss the different methods for a simple Hubbard model (only one orbital per site, next neighbor hopping matrix elements $t_{ij}=-t$): \begin{equation}\label{Hubbard} H=-t\;\sum c^\dagger_i c_j + U\sum n_{i\uparrow} n_{i\downarrow} \; . \end{equation} The generalization to Hamiltonians like (\ref{Hamil}) is straightforward. The first step in the quantum Monte Carlo approach is to identify a trial function $\Psi_T$. For the Hubbard model that function should balance the opposing tendencies of the hopping term and the interaction: Without interaction (i.e.\ for $U=0$) the ground state of the Hamiltonian (\ref{Hubbard}) is the Slater determinant $\Phi$ that maximizes the kinetic energy. Without hopping ($t=0$) the interaction is minimized. Since only doubly occupied sites, i.e.\ sites with $n_{i\uparrow}=1$ and $n_{i\downarrow}=1$, contribute to the Coulomb energy, the electrons are distributed as uniformly as possible over the lattice to minimize the number of double occupancies. A good compromise between these two extremes is to start from the non-interacting wavefunction $\Phi$ but reduce the weight of configurations $R$ with large double occupancies $D(R)$. This leads (up to normalization) to the Gutzwiller wavefunction\index{Gutzwiller wavefunction} \cite{GWF}: \begin{equation}\label{GWF} \Psi_T(R) = g^{D(R)}\;\Phi(R) , \end{equation} with $g\in(0,1]$ the Gutzwiller parameter. In configuration space the Coulomb term in the Hamiltonian is given by $U\,D(R)$. Thus the Gutzwiller factor reflects the interaction term in the Hubbard Hamiltonian. In this spirit we can also construct trial functions for Hamiltonians with additional terms like the screening Hamiltonian (\ref{scrhamil}). We add a second Gutzwiller factor that reflects the interaction with the test charge $q\,U\,n_c(R)$: \begin{equation}\label{scrwf} \Psi_T(R) = g^{D(R)}\,h^{n_c(R)}\;\Phi(R) \;. \end{equation} Changing the additional Gutzwiller factor $h$ we can vary the occupation of the site with the test charge. \subsection{Variational Monte Carlo} Since the Gutzwiller factor, like the interaction term, is diagonal in configuration space, we have to perform a sum over all configurations $R$ in order to calculate the energy expectation value for a Gutzwiller wavefunction: \begin{equation}\label{Evar} E_T = {\langle\Psi_T|H|\Psi_T\rangle \over \langle\Psi_T|\Psi_T\rangle} = {\sum_R E_{\rm loc}(R)\;\Psi_T^2(R) \over \sum_R \Psi_T^2(R)} \;, \end{equation} where we have introduced the local energy for a configuration $R$ \begin{equation}\label{Eloc} E_{\rm loc}(R) = \sum_{R'} {\langle\Psi_T|R'\rangle\,\langle R'|H|R\rangle \over \langle\Psi_T|R\rangle} = \sum_{R'}\!'\;t\;{\Psi_T(R')\over\Psi_T(R)} + U\,D(R) \;. \end{equation} Since the number of configurations $R$ grows exponentially with system-size, the summations in (\ref{Evar}) can only be done explicitly for very small systems. For larger problems we use variational Monte Carlo\index{variational Monte Carlo} \cite{VMC}. The idea is to perform a random walk in the space of configurations, with transition probabilities $p(R\to R')$ chosen such that the configurations $R_\mathrm{VMC}$ in the random walk have the probability distribution function $\Psi_T^2(R)$. Then \begin{equation}\label{Evmc} E_\mathrm{VMC} = {\sum_{R_\mathrm{VMC}} E_{\rm loc}(R_\mathrm{VMC})\over\sum_{R_\mathrm{VMC}} 1} \approx {\sum_R E_{\rm loc}(R)\;\Psi_T^2(R) \over \sum_R \Psi_T^2(R)} = E_T . \end{equation} The transition probabilities can be determined from detailed balance \begin{equation}\label{detailedbalance} \Psi_T^2(R)\,p(R\to R') = \Psi_T^2(R')\,p(R'\to R) \end{equation} which is fulfilled by the choice $p(R\to R')={1/N}\;\min[1,\Psi_T^2(R')/\Psi_T^2(R)]$, with $N$ being the maximum number of possible transitions. It is sufficient to consider only transitions between configurations that are connected by the Hamiltonian, i.e.\ transitions in which one electron hops to a neighboring site. The standard prescription is then to propose a transition $R\to R'$ with probability $1/N$ and accept it with probability $\min[1,\Psi_T^2(R')/\Psi_T^2(R)]$. This works well for $U$ not too large. For strongly correlated systems, however, the random walk will stay for long times in configurations with a small number of double occupancies $D(R)$, since most of the proposed moves will increase $D$ and hence be rejected with probability $\approx 1-g^{D(R')-D(R)}$. There is, however, a way to integrate-out the time the walk stays in a given configuration. To see how, we first observe that for the local energy (\ref{Eloc}) the ratio of the wavefunctions for all transitions induced by the Hamiltonian have to be calculated. This in turn means that we also know all transition probabilities $p(R\to R')$. We can therefore eliminate any rejection (i.e.\ accept with probability one) by proposing moves with probabilities \begin{equation} \tilde{p}(R\to R') = {p(R\to R')\over\sum_{R'} p(R\to R')} = {p(R\to R')\over 1-p_{\rm stay}(R)} \;. \end{equation} Checking detailed balance (\ref{detailedbalance}) we find that now we are sampling configurations $\tilde{R}_\mathrm{VMC}$ from the probability distribution function $\Psi_T^2(R)\,(1-p_{\rm stay}(R))$. To compensate for this we assign a weight $w(R)=1/(1-p_{\rm stay}(R))$ to each configuration $R$. The energy expectation value is then given by \begin{equation} E_T \approx {\sum_{\tilde{R}_\mathrm{VMC}} w(\tilde{R}_\mathrm{VMC})\, E_{\rm loc}(\tilde{R}_\mathrm{VMC}) \over \sum_{\tilde{R}_\mathrm{VMC}} w(\tilde{R}_\mathrm{VMC})} \;. \end{equation} The above method is quite efficient since it ensures that in every Monte Carlo step a new configuration is created. Instead of staying in a configuration where $\Psi_T$ is large, this configuration is weighted with the expectation value of the number of steps the simple Metropolis algorithm would stay there. This is particularly convenient for simulations of systems with strong correlations: Instead of having to do longer and longer runs as $U$ is increased, the above method produces, for a fixed number of Monte Carlo steps, results with comparable error estimates. \subsubsection*{Correlated sampling} So far we have only specified the form of the trial function $\Psi_T$. The goal of a variational calculation is now to identify the parameters that result in the best trial function. A criterion for a good trial function is e.g.\ a low variational energy. To find the wavefunction that minimizes the variational energy we could perform independent VMC calculations for a set of different trial functions. It is, however, difficult to compare the energies from these calculations since each VMC result comes with its own statistical errors. This problem can be avoided with correlated sampling\index{correlated sampling} \cite{corrsmpl}. The idea is to use the same random walk for calculating the expectation value of all the different trial functions. This reduces the {\em relative} errors and hence makes it easier to find the minimum. Let us assume we have generated a random walk $R_\mathrm{VMC}$ using $\Psi_T$ as the trial function. Using the same random walk, we can then estimate the energy expectation value (\ref{Evmc}) for a different trial function $\tilde{\Psi}_T$, by introducing the reweighting factors $\tilde{\Psi}_T^2(R)/\Psi_T^2(R)$: \begin{equation}\label{corrsmpl} \tilde{E}_T \approx {\sum_{R_\mathrm{VMC}} \tilde{E}_{\rm loc}(R)\,\tilde{\Psi}_T^2(R)/\Psi_T^2(R) \over \sum_{R_\mathrm{VMC}} \tilde{\Psi}_T^2(R)/\Psi_T^2(R) }\;. \end{equation} with \begin{eqnarray} \tilde{E}_{\rm loc}(R) &=&\sum_{R'} t {\tilde{\Psi}_T(R')\over\tilde{\Psi}(R)} + U\;D(R) \\ &=&\sum_{R'} t {\Psi_T(R')\over\Psi_T(R)}\; {\tilde{\Psi}_T(R')/\Psi_T(R') \over \tilde{\Psi}_T(R)/\Psi_T(R)} + U\;D(R) \nonumber \end{eqnarray} We notice that (also in $\tilde{E}_{\rm loc}$) the new trial function $\tilde{\Psi}_T$ appears only in ratios with the old $\Psi_T$. For trial functions (\ref{GWF}) that differ only in the Gutzwiller factor this means that the Slater determinants cancel, leaving only powers $(\tilde{g}/g)^{D(R)}$. Since $D(R)$ is {\em integer} we can then rearrange the sums in (\ref{corrsmpl}) into polynomials in $\tilde{g}/g$. To find the optimal Gutzwiller parameter we then pick a reasonable $g$, perform a VMC run for $\Psi_T(g)$ during which we also estimate the coefficients for these polynomials. We can then calculate $E(\tilde{g})$ by simply evaluating the ratio of the polynomials. Since there are typically only of the order of some ten non-vanishing coefficients this method is very efficient. The idea of rewriting the sum over configurations into a polynomial can be easily generalized to trial functions with more correlation factors of the type $r^{c(R)}$, as long as the correlation function $c(R)$ is integer-valued on the space of configurations. As a specific example of how the method works in practice, Fig.\ \ref{screen} shows the result of a correlated sampling run for the trial function (\ref{scrwf}). \begin{figure}[t] \rotatebox{270}{\includegraphics[width=.5\textwidth]{Evmc.eps}} \rotatebox{270}{\includegraphics[width=.5\textwidth]{Edmc.eps}} \caption[]{\label{screen} Correlated sampling for the parameters $g$ and $h$ in the generalized Gutzwiller wavefunction $|\Psi_T\rangle= g^D h^{n_c}\,|\Phi\rangle$, cf.\ eqn.\ (\ref{scrwf}), in variational (upper plot) and fixed-node diffusion Monte Carlo (lower plot). The plots show the energy as a function of the Gutzwiller parameters $g$ and $h$, both as surfaces and contours. The calculations were done for an fcc cluster of 64 molecules with $96+96$ electrons (half-filled $t_{1u}$-band), an on-site Hubbard interaction $U=1.25\,\mathrm{eV}$, and a test charge of $q=1/4\,\mathrm{e}$. } \end{figure} \subsection{Fixed-node diffusion Monte Carlo} Diffusion Monte Carlo\index{diffusion Monte Carlo} \cite{GFMC} allows us, in principle, to sample the true ground state of a Hamiltonian. The basic idea is to use a projection operator that has the lowest eigenstate as a fixed point. For a lattice problem where the spectrum is bounded $E_n\in[E_0,E_{\rm max}]$, the projection is given by \begin{equation}\label{proj} |\Psi^{(n+1)}\rangle = [1-\tau(H-E_0)]\;|\Psi^{(n)}\rangle \;;\quad |\Psi^{(0)}\rangle=|\Psi_T\rangle . \end{equation} If $\tau<2/(E_{\rm max}-E_0)$ and $|\Psi_T\rangle$ has a non-vanishing overlap with the ground state, the above iteration converges to $|\Psi_0\rangle$. There is no time-step error involved. Because of the prohibitively large dimension of the many-body Hilbert space, the matrix-vector product in (\ref{proj}) cannot be done exactly. Instead, we rewrite the equation in configuration space \begin{equation}\label{iter} \sum |R'\rangle\langle R'|\Psi^{(n+1)}\rangle = \sum_{R,R'} |R'\rangle \underbrace{\langle R'|1-\tau(H-E_0)|R\rangle}_{=:F(R',R)} \langle R|\Psi^{(n)}\rangle \end{equation} and perform the propagation in a stochastic sense: $\Psi^{(n)}$ is represented by an ensemble of configurations $R$ with weights $w(R)$. The transition matrix element $F(R',R)$ is rewritten as a transition probability $p(R\to R')$ times a normalization factor $m(R',R)$. The iteration (\ref{iter}) is then stochastically performed as follows: For each $R$ we pick a new configuration $R'$ with probability $p(R\to R')$ and multiply its weight by $m(R',R)$. Then the new ensemble of configurations $R'$ with their respective weights represents $\Psi^{(n+1)}$. Importance sampling\index{Importance sampling} decisively improves the efficiency of this process by replacing $F(R',R)$ with $G(R',R)=\langle\Psi_T|R'\rangle\,F(R',R)/\langle R|\Psi_T\rangle$, so that transitions from configurations where the trial function is small to configurations with large trial function are enhanced: \begin{equation} \sum |R'\rangle\langle\Psi_T| R'\rangle\langle R'|\Psi^{(n+1)}\rangle = \sum_{R,R'} |R'\rangle\,G(R',R)\, \langle\Psi_T|R\rangle\,\langle R|\Psi^{(n)}\rangle . \end{equation} Now the ensemble of configurations represents the product $\Psi_T\,\Psi^{(n)}$. After a large number $n$ of iterations the ground state energy is then given by the mixed estimator\index{mixed estimator} \begin{equation}\label{mixedest} E_0 = {\langle\Psi_T|H|\Psi^{(n)}\rangle \over \langle\Psi_T|\Psi^{(n)}\rangle} \approx {\sum_R E_{\rm loc}(R)\;w(R) \over \sum_R w(R)} . \end{equation} As long as the evolution operator has only non-negative matrix elements $G(R',R)$, all weights $w(R)$ will be positive. If, however, $G$ has negative matrix elements there will be both configurations with positive and negative weight. Their contributions to the estimator (\ref{mixedest}) tend to cancel so that eventually the statistical error dominates, rendering the simulation useless. This is the infamous sign problem\index{sign problem}. A straightforward way to get rid of the sign problem is to remove the offending matrix elements from the Hamiltonian, thus defining a new Hamiltonian $H_{\rm eff}$ by \begin{equation} \langle R'|H_{\rm eff}| R\rangle = \left\{ \begin{array}{cc} 0 & \mbox{ if $G(R',R)<0$} \\ \langle R'|H| R\rangle & \mbox{ else} \end{array}\right. \end{equation} For each off-diagonal element $\langle R'|H| R\rangle$ that has been removed, a term is added to the diagonal: \begin{equation} \langle R|H_{\rm eff}|R\rangle = \langle R|H|R\rangle + \sum_{R'} \Psi_T(R')\langle R'|H|R\rangle/\Psi_T(R) . \end{equation} This is the fixed-node approximation\index{fixed-node approximation} for lattice Hamiltonians \cite{FNDMC}. $H_{\rm eff}$ is by construction free of the sign problem and variational, i.e.\ $E_0^{\rm eff}\ge E_0$. The equality holds if $\Psi_T(R')/\Psi_T(R)=\Psi_0(R')/\Psi_0(R)$ for all $R$, $R'$ for which $G(R',R)$ is negative. Fixed-node diffusion Monte Carlo for a lattice Hamiltonian thus means that we choose a trial function from which we construct an effective Hamiltonian and determine its ground state by diffusion Monte Carlo. Because of the variational property, we want to pick the $\Psi_T$ such that $E_0^{\rm eff}$ is minimized, i.e. we want to optimize the trial function, or, equivalently, the effective Hamiltonian. Also here we can use the concept of correlated sampling. For optimizing the Gutzwiller parameter $g$ we can even exploit the idea of rewriting the correlated sampling sums into polynomials in $\tilde{g}/g$, that we already have introduced in VMC. There is, however, a problem arising from the fact that the weight of a given configuration $R^{(n)}$ in iteration $n$ is given by the product $w(R^{(n)})=\prod_{i=1}^n m(R^{(i)},R^{(i-1)})$. Each individual normalization factor $m(R',R)$ can be written as a finite polynomial, but the order of the polynomial for $w(R^{(n)})$ increases steadily with the number of iterations. It is therefore not practical to try to calculate the ever increasing number of coefficients for the correlated sampling function $E^{(n)}(\tilde{g})$. But since we still can easily calculate the coefficients for the $m(R',R)$, we may use them to evaluate $E^{(n)}(\tilde{g})$ in each iteration on a set of predefined values $\tilde{g}_i$ of the Gutzwiller parameter. Figure \ref{screen} shows an example, again for the more general trial function (\ref{scrwf}). \subsubsection{Extrapolated estimator} So far we have only considered estimators for the total energy. For determining the screening, however, we need to know the charge density $n_c$ on site $c$. It is no problem to calculate the expectation value $n_c(\mathrm{VMC})=\langle\Psi_T|\hat{n}_c|\Psi_T\rangle$ in variational Monte Carlo. Diffusion Monte Carlo gives, however, only the mixed estimator $n_c(\mathrm{DMC})=\langle\Psi_T|\hat{n}_c|\Psi_0\rangle$. Since the density operator does not commute with the Hamiltonian, the mixed estimator is different from expectation $n_c(\mathrm{QMC})=\langle\Psi_0|\hat{n}_c|\Psi_0\rangle$. For a good choice of the trial function the true expectation value can be determined using the extrapolated estimator\index{extrapolated estimator} $n_c(\mathrm{QMC})\approx 2\;n_c(\mathrm{DMC})-n_c(\mathrm{VMC})$. To test the accuracy of this approach to the screening problem, which besides the extrapolated estimator also involves the fixed-node approximation, we have compared the results of the quantum Monte Carlo calculations with the exact results from exact diagonalization for a small system of four molecules (12 electrons). As illustrated by Fig.\ \ref{lcheck}, the QMC calculations for determining the screening charge are very accurate. \begin{figure}[t] \rotatebox{270}{\includegraphics[width=.5\textwidth]{lcheck.eps}} \caption[]{ Screening charge $\Delta n$ on the site of the test charge ($q=0.25\,\mathrm{e}$) as a function of $U/W$, where $U$ is the Coulomb interaction and $W$ is the band width. Exact diagonalization and QMC calculations have been performed for four molecules (12 electrons). The figure shows that the QMC calculations are quite accurate over the whole range of $U/W$.} \label{lcheck} \end{figure} \subsubsection*{Acknowledgment.} This work has been supported by the Alexander-von-Humboldt-Stiftung under the Feodor-Lynen-Program and the Max-Planck-Forschungspreis, and by grant No.\ DEFG02-96ER45439 from the Department of Energy.
\section{Introduction} The mechanisms responsible for gamma-ray bursts (GRBs) are not yet known with any certainty, but as evidence for a cosmological origin mounts (e.g., Metzger et al. 1997; Frail et al. 1997; Bloom, et al. 1998; Bloom et al. 1999a), models based upon black hole accretion disks (BHAD models) have gained prominence. Such models have several advantages, especially their large inherent energy and potentially short time scales. In the neutrino-powered GRB paradigm, BHAD models can overcome the strong angle dependence of neutrino annihilation and extract some of the gravitational energy of the disk, possibly without ejecting too much baryonic matter (Paczynski 1991; Woosley 1993; Ruffert et al. 1997; Popham, Woosley, \& Fryer 1998; MacFadyen \& Woosley 1999). BHAD models also provide favorable conditions for magnetic fields to extract energy, either from the gravitational potential energy of the accretion torus itself, or from the rotational energy of spinning black holes (Blandford \& Znajek 1977; MacDonald et al. 1986; Paczynski 1991, 1998; Narayan et al. 1992; Hartmann \& Woosley 1995; Meszaros \& Rees 1997a; Katz 1997; Livio, Ogilvie, \& Pringle 1999). The merger of double neutron star (DNS) systems is probably the best studied representative of this class, but a variety of progenitor systems lead to similar conditions: mergers of other compact systems such as binaries consisting of a neutron star and a black hole (BH/NS) or of a white dwarf and black hole (WD/BH), the inspiral of a compact object into its companion's helium core during common envelope evolution (helium-mergers), and the accretion of material into a black hole in a ``failed supernova'' (collapsar). Estimates of the rates for these various occurrences are plagued by a variety of uncertainties in the supernova explosion mechanism, stellar evolution, and binary evolution. Although the current literature abounds with population synthesis calculations, especially of DNS systems, and the closely-related BH/NS binaries (Clark, van den Heuvel, \& Sutantyo 1979; Lipunov, Postnov, \& Prokhorov 1987; Hills, Bender \& Webbink 1991; Tutukov \& Yungelson 1993; Lipunov et al. 1995; Portegies-Zwart \& Spreeuw 1996; Fryer, Burrows, \& Benz 1998; Portegies-Zwart \& Yungelson 1998, Bethe \& Brown 1998; Belczy\'nski \& Bulik 1998; Bloom, Sigurdsson, \& Pols 1999b), little attention has been directed towards either exploring the uncertainties in the models or in examining the other kinds of models. In this paper, we shall study the rate of formation of all varieties of BHAD progenitors, including our best estimates of the (admittedly uncertain) physics. We will show that the discrepancy between rates calculated from population studies and extrapolations from the data of DNS systems are reconcilable if one includes both the uncertainties in the population synthesis calculations and the error in the measured rates. BHAD models other than DNS will also be considered within the same parameterized framework. Comparisons with observed compact systems can be used to constrain some of the parameters, e.g., the pulsar velocity distribution and the observed properties of X-ray binaries and DNS systems. However, even with these constraints, we shall find that most rates for GRB progenitor formation are uncertain by over two orders of magnitude. From these studies, we shall also determine the velocity distributions of the GRB progenitor systems and their corresponding merger time scales. By evolving these systems in representative gravitational potentials for both high and low mass galaxies, we can determine the distances each GRB progenitor travels before producing a GRB. A fraction of the compact binary systems receive large enough systemic velocities to be ejected from their host galaxies, but many remain bound even in such low mass galaxies as the host of GRB970508 (Bloom et al. 1998). In $\S$2, we describe formation scenarios for each of the black hole accretion disk GRB progenitor systems. Section 3 describes our population synthesis code and the parameters of population synthesis, noting the effects of the uncertainty in each. Section 4 gives our predictions, including the effects of varying systemic velocities and merger time scales to obtain both the spatial and temporal distribution of GRBs from each progenitor. We conclude with a summary of the merits of the different models. \section{Black Hole Accretion Disk progenitors} Modeling the progenitors of BHAD derived GRBs involves all the uncertainties inherent in any study of compact object formation. Moreover, since the systems must be in short-period binaries to merge, the problem is compounded by uncertainties in the parameters of binaries (mass ratios, separations) and their evolution. Before delving farther into these uncertainties, it is first helpful to have in mind some general scenarios (Table 1, Fig. 1). \subsection{Double Neutron Star Binaries}\label{sec:dns} When two neutron stars merge, they quickly form an object too large to be supported by nuclear and degeneracy pressure (for most choices of equation of state). A black hole forms on a dynamic time scale, but a significant amount of mass, typically 0.03 -- 0.3 M$_\odot$, has too much angular momentum to enter the hole promptly. An accretion disk forms. What happens next depends upon the viscosity and entropy of this disk, the evolution of its magnetic field, and the interaction of the disk with the rotating black hole. All three may be quite complicated. If the viscosity is low and the temperature inadequate to drive rapid energy loss by neutrinos, the disk may persist for a long time, perhaps over 10 s in the models envisioned by Meszaros \& Rees (1997a). Any relativistic outflow that develops must then be a consequence of MHD processes in the disk or of the magnetic extraction of rotational energy from the hole. These same MHD processes can also function in a viscous disk (indeed some appreciable magnetic field must be present to provide the viscosity), but the time scale is much shorter and the temperature higher. Most models of DNS mergers that have been carried out on the computer are of this latter variety (Davies et al. 1994; Rasio \& Shapiro 1994; Ruffert, Janka, Sch\"afer 1995; Janka \& Ruffert 1996; Wilson, Mathews, \& Marronetti 1996; Mathews \& Wilson 1997; Ruffert et al. 1997, Ruffert \& Janka 1998; Rossweg at al. 1999). In the ``hot model'' The entire accretion process then lasts for 10-100\,ms and most of the binding energy of the disk is radiated as neutrinos. Indeed the neutrino luminosity can be so great that a significant fraction, on the order of 1\%, is converted, via neutrino annihilation, to thermal energy along the rotational axis. This is enough to power energetic jets that produce a GRB either by internal or external shocks. The hot DNS model might therefore be more appropriate for the short, hard subset of GRBs (Fishman \& Meegan 1995) which are very difficult to make other ways (except in the closely allied BH/NS model). Observationally, these short hard bursts are not yet constrained, either in energy or host, by the observations of Beppo-Sax, which hitherto has employed a 5 s trigger time. The progenitor systems here are double neutron stars in binaries so close that merger occurs in a Hubble time as a consequence of gravitational radiation. DNS systems have a long observable lifetime because the primary's neutron star accretes matter and becomes a recycled radio pulsar. Recycled pulsars are thought to be produced in binaries where accretion from a companion spins up the neutron star whose magnetic field has partially decayed (see reviews by Verbunt 1993; Phinney \& Kulkarni 1994). Due to the low magnetic field strengths, spin down of these pulsars occurs slowly, and consequently recycled radio pulsars have observable lifetimes $\gtrsim 10-100$ times longer than normal pulsars. It is not surprising then that essentially all pulsars observed in double neutron star systems are recycled pulsars. The standard scenario for forming close orbit DNS systems (Scenario I, Fig. 2) begins with two massive stars. The more massive star (primary) evolves off the main sequence, overfills its Roche-lobe, and transfers mass to its companion (secondary). The primary then evolves to the end of its life, forming a neutron star in a supernova explosion. The binary system remains bound, unless large kicks\footnote{The pulsar velocity distribution requires that neutron stars are somehow accelerated to mean velocities of roughly $\sim 450 {\rm \, km \, s^{-1}}$, most likely at or near the time of birth. We discuss this in more detail in \S \ref{sec:sn}.} are imparted to the neutron star during formation, and a binary consisting of a neutron star and massive companion is formed. This system passes through an X-ray binary phase which then evolves through a common envelope as the secondary star expands. During this common envelope phase in the standard model, the neutron star spirals into the massive secondary and the orbital energy released ejects the secondary's hydrogen mantle, forming a neutron star/helium star binary. Mass accretion onto the neutron star primary is alleged to add angular momentum to the neutron star, ``recycling'' it as a pulsar. After the explosion of the helium star as a supernova, a close DNS binary remains which, in time, merges via gravitational wave emission. However, recent calculations of neutron star accretion (Chevalier 1993, 1996; Houck \& Chevalier 1992; Brown 1995; Fryer, Benz, \& Herant 1996) reveal that, during the common envelope phase, the neutron star accretes at the Bondi-Hoyle rate, not the Eddington rate. During its spiral in, the neutron star can accrete over a solar mass (Bethe \& Brown 1998) and collapse to form a black hole. Thus, the ``standard'' scenario for DNS systems may in fact form BH/NS binaries. How, then, are close DNS systems formed? Brown (1995) suggested an alternate scenario (Scenario II, Fig. 3) in which the initial binary system is comprised of two massive stars of nearly equal mass. The secondary (whose mass is assumed to be within 5\% of the primary) evolves off the main sequence before the explosion of the primary as a supernova. The two stars then enter a common envelope phase with two helium cores orbiting within one combined hydrogen envelope. After the hydrogen envelope is ejected, first one, then the other, helium stars explode. If the kicks imparted to the neutron stars during the supernova explosion are sufficiently low, a double neutron star system can form. To generate the observed systems, after the first supernova, the newly formed neutron star must accrete material, possibly from its companion's helium star wind, to become the observed ``recycled'' radio pulsar. If accretion is insufficient, a ``silent'' double neutron star binary is formed in which neither neutron star is recycled. Such systems, though only briefly detectable as radio pulsras, might still evolve to produce GRBs. A third scenario (Scenario III, Fig. 4) avoids any common envelope phase, but requires that the neutron star formed in the explosion of the secondary receive a kick that places it into an orbit that will allow the two neutron stars to merge within a Hubble time. However, in general, to avoid a common envelope phase, the pre-supernova orbital separation must be extremely wide (greater than 1 AU). The range of kick magnitudes and directions which will produce such orbits is so small that, as we shall show in this paper, this formation scenario is probably not important. From the number of known Galactic DNS binaries (PSR 1913+16, PSR 1534+12), a formation rate can be estimated. However, correcting for time evolution of pulsar luminosities, beaming fractions, and distance estimations proves tricky for a sample of only two merging double neutron star systems. From the number of observed DNS systems ($N_{\rm obs}$), and including an estimate of all the selection effects: beaming fraction, fraction of galaxy currently observed, etc., the formation rate is calculated as \begin{equation}\label{eq:rate} {\rm Rate}_{\rm DNS}=f_{\rm b} \frac{N_{\rm obs}} {T_{\rm PSR}}, \end{equation} where $T_{\rm PSR}$ is the observable lifetime of the pulsar and all other uncertainties have been represented by a single factor, $f_{\rm b}$, whose value may be uncertain by 1-2 orders of magnitude. With best guesses for $f_{\rm b}$, the current rate inferred for our Galaxy from the observations fall in the range: $\sim 10^{-6}-10^{-5} \, {\rm yr^{-1}}$ (Phinney 1991; Narayan, Piran, \& Shemi 1991; Curran \& Lorimer 1995; van den Heuvel \& Lorimer 1996). Bailes (1996) estimated an upper limit for the formation rate of these systems using the fact that all of the observed pulsars in DNS systems are recycled and none of the companion neutron stars are observed as normal pulsars. Although normal pulsars have much shorter lifetimes, they dominate the observed pulsar population because they do not require binary accretion to ``recycle'' them. The formation rate of single pulsars is roughly $\sim0.008$\,yr$^{-1}$ and, at the time of the calculation by Bailes (1996), roughly 650 normal radio pulsars were observed. Using eq. (\ref{eq:rate}), we can then estimate the value of $f_{\rm b}/T_{\rm PSR}$. If we assume that the collapse and pulsar formation of the secondary is not too dissimilar from that of single stars (since the core of a massive star is not affected greatly by its binary companion, this assumption may not be too bad), then we can use the value of $f_{\rm b}/T_{\rm PSR}$ derived from the pulsar population to place an upper limit on the DNS formation rate. Since the number of normal secondary pulsars is 0 ($<1$), the current DNS formation rate in our Galaxy is then (Bailes 1996) \begin{equation} {\rm Rate}_{\rm DNS}< \left( \frac{f_{\rm b}}{T_{\rm PSR}} \right)_{\rm PSR}=10^{-5} \, {\rm yr^{-1}}. \end{equation} Since this calculation, the number of observed pulsars, the number of DNS systems has increased, and the pulsar lifetime has changed, and this upper limit has risen to $10^{-4} \, {\rm yr^{-1}}$ (Arzoumanian, Cordes, \& Wasserman 1999). \subsection{Black Hole + Neutron Star Binaries}\label{sec:bhns} As BH/NS binaries merge, the neutron star is tidally disrupted. Some of the material accretes directly, producing little emission, but the remainder forms an accretion disk of a few tenths of solar mass which can power a GRB (Lee \& Kluzniak 1995; Eberl 1998; Eberl et al. 1999). Like the DNS mergers, if the disk is hot and viscous, the relativistic outflows created by BH/NS mergers are probably of short duration, but somewhat more energetic than their DNS counterparts because of the larger mass reservoir. Merging BH/NS binaries are formed by pathways that are similar to DNS binaries. In general, the collapse of a massive star can form black holes in two distinct ways (Woosley \& Weaver 1995; Fryer 1999). The star can explode, but with such a weak shock that much of the stellar mantle of heavy elements falls back onto the neutron star, causing it to further collapse into a black hole. Alternatively, especially for the more massive stars, the ram pressure of the infalling stellar material may be too large to allow the shock to propagate at all and the star may collapse directly into a black hole. The two limits for black hole formation from fallback and direct collapse are, respectively, $M_{\rm BH} \approx 25$M$_{\scriptscriptstyle \odot}$\,, $M_{\rm Coll} \approx 40$M$_{\scriptscriptstyle \odot}$\,(Fryer 1999). The ``standard'' formation scenario for BH/NS systems (Scenario IV, Fig. 5) begins with two massive stars, the primary having a mass greater than $M_{\rm BH}$. As with the standard DNS formation scenario, the primary evolves off the main sequence, overfills its Roche-lobe, and transfers mass to its companion. The primary continues to evolve to the end of its life, forming a black hole and, possibly, a supernova explosion. With the exception of extremely large kicks, the system remains bound, creating a binary consisting of a black hole and a massive star. This system passes through an X-ray binary phase (e.g., Cyg X-1, LMC X-1) and then evolves through a common envelope as the secondary star expands. During this common envelope phase (in the standard model) the black hole spirals into the massive secondary and ejects the secondary's hydrogen envelope. After the secondary's supernova explosion, a BH/NS binary forms which then merges via gravitational wave emission. BH/NS binaries can also form from neutron star binaries in which the neutron star of the primary gains too much matter during common envelope phase and collapses into a black hole (\S \ref{sec:dns}, Scenario V, Fig. 6). This formation scenario produces binaries consisting of a low-mass black hole ($\sim 3 M_\odot$) and a neutron star which could generate the most energetic BH/NS GRBs (Eberl 1998, Eberl et al. 1999). Finally, just as with the DNS systems, a BH/NS binary can form under a third mechanism avoiding any common envelope evolution for appropriate kick magnitudes and directions (Scenario VI, Fig. 7). Like the DNS kick formation scenario, the kick scenario is probably not important for BH/NS formation. Although BH/NS binaries have not yet been observed, their formation rate is probably comparable or even larger than that of DNS systems. As with the double neutron stars, we can determine an observational upper limit to the BH/NS rate in the Galaxy (assuming the observational biases are similar to those for single pulsars) of $\sim 10^{-4}$\,yr$^{-1}$. \subsection{Black Hole -- White Dwarf Binaries} As WD/BH binaries merge, the white dwarf is tidally disrupted and most of its matter is converted into an accretion disk around the black hole. The accretion of this disk material onto the black hole can drive a gamma-ray burst (Fryer et al. 1999). Although the accretion disk is more massive than that formed in DNS mergers, its radius is much larger so the accretion time scale is longer and neutrino energy transport less efficient. However, the black hole is rapidly rotating and up to 10$^{53}$ erg of either rotational energy or disk binding energy is potentially available on a viscous time scale for the disk. MHD powered jets might provide long-duration GRBs. Not surprisingly, there is also more than one way to produce a binary containing a white dwarf and a black hole. Our first scenario begins with a binary consisting of a primary with mass greater than $M_{\rm BH}$ and a low-mass ($< 10$M$_{\scriptscriptstyle \odot}$) companion (Scenario VII, Fig. 8). After the primary collapses to form a black hole, one has a binary consisting of a black hole and a main sequence star. If the mass of the main sequence companion is less than $\sim 2-3 M_\odot$, magnetic braking will drive the binary together, forming a low-mass X-ray binary (LMXB). More massive companions will not become X-ray binaries until they evolve off the main sequence (e.g. LMC X-3, 2023+338 Nova Cyg, J0422+32 Nova Per). For masses of $5-10 {\rm M_\odot}$, a common envelope phase is likely to occur which decreases the orbital separation. These stars eventually evolve into high mass white dwarfs, a fraction of which will merge with the black hole primary in a Hubble time. Alternatively, these systems may form from neutron star binaries with low-mass ($< 10 M_\odot$) companions (Scenario VIII, Fig. 9). If the neutron star enters a common envelope phase when the low-mass companion expands into a giant, it can accrete too much material and collapse to a black hole. BH/WD binaries may also form avoiding the first common envelope phase (where the low-mass companion spirals into the massive primary) entirely with a ``well-placed'' kick (Scenario IX, Fig. 10). Just as with Scenario VII, the system evolves briefly through an X-ray binary phase and possible common envelope phase which tightens the orbit. Just as with BH/NS binaries, there are no observed closely interacting BH/WD binaries and our estimates must be very uncertain. The lack of a detectable system so far implies only \begin{equation} {\rm Rate}_{\rm WD/BH} \lesssim \frac{V_{\rm Galaxy}} {V_{\rm WD \, Obs}} T_{\rm WD}^{-1} \approx \frac{(10 {\rm \, kpc})^2 (1 {\rm \, kpc})}{(10 \, {\rm pc})^3} \times (10^{10} \, {\rm yr})^{-1} \approx 0.01 \, {\rm yr^{-1}}. \end{equation} In this paper, we will show in fact that the predicted rate is $\sim 5$ orders of magnitude less than this upper limit. \subsection{Collapsars} The only BHAD GRB progenitor which does not necessarily require a binary system (though, as we shall see, being in a binary still helps) is the collapsar (Woosley 1993). Collapsars are formed when the collapse of the iron core of a rotating massive supernova progenitor proceeds directly to a black hole. As the stellar mantle falls into the newly formed black hole, angular momentum slows the collapse along the equator, ultimately forming an accretion disk which powers the GRB (Woosley 1993, MacFadyen \& Woosley 1999). Collapsars can produce long duration bursts ($\gtaprx10$\,s) with a large range of burst energies by either neutrino or MHD processes. In order that the jet not dissipate its energy in the hydrogen envelope of a red supergiant star, it is necessary that the collapsar occur in a hydrogen stripped, Wolf-Rayet star. The collapsar model requires that no supernova shock is launched, and hence collapsars form only from stars whose masses exceed $M_{\rm Coll}$. One obvious consequence is that GRBs from this model would closely trace star formation. There will also be significant amounts of circum-burster material from the pre-SN stellar wind and the gas in the surrounding star forming medium. The afterglow (and even some bursts) would be produced by shock interactions with this wind, not with the ordinary interstellar medium. Scenario X (Fig. 11) is the simplest possibility for producing a collapsar. The iron core of a single very massive star collapses promptly to a black hole. Current simulations of collapsars (MacFadyen \& Woosley 1999) suggest that a successful collapsar model requires specific angular momenta of the massive star cores to be in the range: $j \approx 3 \times 10^{16} - 2 \times 10^{17}$ and that the massive star has lost most of its hydrogen envelope, but retains a sufficiently large helium core to collapse and power a GRB. If the angular momentum is below this range, an accretion disk does not form and the material falls more or less directly into the black hole. If the angular momentum is too large, the disk does not efficiently dissipate energy through neutrino emission and a weak outflow instead of an accretion disk results. The requisite high mass of the helium star may be more easily attained in a star of low metallicity (MacFadyen \& Woosley 1999). The biggest uncertainty in estimating the rate of collapsar models, however, is the angular momentum. There are diverse views regarding how much angular momentum is retained in the stellar cores of massive stars when they die. Heger (1998) modeled the evolution of rotating massive stars and found that the cores of most stars naturally have angular momenta close to the range required by MacFadyen \& Woosley (1999). However, Spruit \& Phinney (1998) have argued that magnetic fields will couple the cores of stars to their envelopes, causing the star to rigidly rotate. As the star became a red supergiant (or even a blue one), the core rotation would essentially halt. No GRB could be produced if this occurred. However, the results of Spruit \& Phinney rest upon uncertain assumptions involving the evolution of magnetic field instabilities, especially those responsible for producing a radial component in the region where the core and envelope are interacting. Here we will assume that there is no such magnetic coupling and all massive stars are rotating within the range required by MacFadyen \& Woosley (1999). It will turn out however that, because of the stellar mass loss prescription we adopt, massive stars other than those interacting binaries, will produce a minority of collapsar progenitors. If the stellar envelope is removed very early by a companion star, perhaps the rotation of the core is not slowed as much. However, one may want to multiply our final numbers for collapsar rates by an uncertain factor less than one to account for inadequate angular momentum. Binary companions can also remove hydrogen envelopes during a common envelope phase. After common envelope evolution, which uncovers the helium core, the primary star (if massive enough) collapses into a black hole. If the angular momentum of the primary is appropriate, a collapsar is formed (Scenario XI, Fig. 12). Note that this system may remain in a binary and later evolve into a BH/NS or WD/BH binary. Even if both Scenarios X and XI fail (e.g. because of magnetic braking of the core's rotation; Spruit \& Phinney 1998), it is still possible to form a collapsar in a binary in which the two stars are of comparable mass, and evolve off the main sequence together. If the two helium cores merge during their common envelope evolution, they will form a massive helium core with a large amount of angular momentum even if their individual cores were rotating slowly (Scenario XII, Fig. 13). Scenario XII by itself then might be thought of as giving a lower bound to the collapsar rate. \subsection{Helium Star Mergers} The helium star merger model (Fryer \& Woosley 1998) is another outcome possible for the common envelope phase so central to making DNS and BH/NS binaries. If the inspiralling compact object does not have sufficient orbital energy to eject its companion's hydrogen envelope, it moves on into the helium core (e.g., in Scenarios I, IV, V; see Figs. 2,5,6). As the helium core and compact object merge, the Bondi-Hoyle accretion rate may become extremely high ($1 M_\odot$\,s$^{-1}$) and the compact object quickly accretes enough to become a black hole if it has not already. The angular momentum of the helium core/black hole binary is injected into the helium core, forming a massive disk ($\sim 4 M_\odot$) around a spinning black hole (Scenario XIV). This BHAD GRB model has not been simulated in detail, but it is likely to produce bursts of long duration similar to that of the collapsar model, though probably longer. Unlike the collapsar model, there is no concern about having enough angular momentum. Indeed the problem may be too much of it. Alternatively, the merging process can be initiated by a kick that places the neutron star inside its companion, ultimately leading to the merger of the neutron star with its companion's helium core, again forming an accretion disk around a black hole. However, the odds of such a well-placed kick are again slim, and this scenario never occurred in our simulations and hence we do not list it separately. \section{The Population Synthesis Calculation} For our numerical calculations, we used a modified version of the Monte Carlo population synthesis code of Fryer, Burrows, \& Benz (1998). This code has previously been tested against, and, within the errors, agrees with the low-mass X-ray binary formation calculations of Kalogera \& Webbink (1998) and the high-mass X-ray binary formation simulations of Dalton \& Sarazin (1995). Calculations of compact object formation begin with a set of initial binary conditions, each of which brings some element of uncertainty to the result: the initial mass function, the fraction of binary stars, the binary mass ratio distribution, the distribution of orbital separations, and initial eccentricities. Monte-Carlo calculations then evolve the binary stars and algorithms for both the single-star evolution (e.g., He core mass, radii during giant phase, winds) and the binary-star evolution (e.g., mass transfer, common envelope evolution) must also be assumed. Finally, one must make some approximations to the supernova explosion mechanism and its results - especially neutron star (and black hole) kicks. We allow all these parameters to range over a reasonable range of uncertainties, applying constraints where we can, to derive a range in resultant GRB rate. The effects of the free parameters are demonstrated in Tables 2-7. An estimated range of rates, including all the uncertainties, is given Table 8. With such a large number of free parameters, we cannot survey the entire parameter space, so we instead chose a ``standard'' model from which to examine the effects of each individually. Our standard model uses the following assumptions (to be described in greater detail in subsequent subsections): $\alpha_{\rm IMF}=2.7$, circular initial orbits, a logarithmic distribution of initial separations [$P(A) \propto 1/A$], a flat probability distribution for the mass ratio, a binary fraction of 40\%, a supernova rate of $0.02\,$yr$^{-1}$, isotropic supernova kicks given to both neutron stars and black holes, a lower mass limit for black hole formation of $25 M_\odot$, stellar radii for mass-transfer and common-envelope evolution set to 1/4 that given by Kalogera \& Webbink (1998), helium star masses as given by Kalogera \& Webbink (1998) but with wind mass-loss rates set to 10$\%$ of those used by Woosley, Langer, \& Weaver (1995), mass transfer parameters ($\alpha_{\rm MT}$, $\beta_{\rm MT}$) of 1.,0.8 respectively, and common envelope efficiency, ($\alpha_{\rm CE}$), equal to 0.5. For other parameters not discussed in this section, we used the standard values given in Fryer, Burrows, \& Benz (1998). Unless otherwise stated, each Monte-Carlo simulation models 20 million massive binaries and the plots and tables include only those binaries which will merge within a Hubble time. With 20 million binaries, a supernova rate of $0.02\,$yr$^{-1}$ and a binary fraction of 40\%, a binary merger rate of 1 Myr$^{-1}$ corresponds to 5000 systems and statistical errors of roughly 2\%. How calculations normalized to the supernova rate in our Galaxy are translated into universal rates based upon a history of star formation is discussed in $\S$4.2. \subsection{Initial Conditions} The initial conditions in our population synthesis calculations include distributions of the zero-age main-sequence masses, of orbital separations, and of orbital eccentricities for the binaries. The primary mass is determined using an initial mass function (IMF): \begin{equation} f(M_{p,0}) = f(1) M_{p,0}^{-\alpha_{\rm IMF}} {\rm stars \, yr^{-1} \, M^{-1}_\odot}, \end{equation} where $M_{p,0}$ is the initial mass of the primary, and $\alpha_{\rm IMF}$ is typically set to 2.7 (Scalo 1986). Rather than fix $f(1)$, we fix the supernova rate (\S \ref{sec:sn}). Since we are only interested in compact binaries, we require that the primaries have a mass of at least $10 M_\odot$. We allow massive stars up to 150 $M_\odot$ to be formed, but our choice of the initial mass function severely restricts the role these massive stars play in our population synthesis. Repeating the calculation with $\alpha_{\rm IMF}=2.35$ yields similar (within a factor of 3) merger rates (Table 2). Secondary masses are much more difficult to determine. The standard technique (see Hogeveen for a review 1990) prescribes a mass ratio ($q=\frac{M_{s}}{M_{p}}$) distribution $P(q)$ by \begin{equation} P(q) \propto q^{-\alpha_{\rm MR}}. \end{equation} Because it is extremely difficult to detect low-mass companions of massive stars, the value of $\alpha_{\rm MR}$ ranges from -1 or 0 (Garmany, Conti, \& Massey 1980) to 2.7 for $q>0.35$ and 0. for $q<0.35$ (Hogeveen 1990). The discrepancy in these mass ratios arises not from different data, but a different interpretation of the data. Of the 67 massive stars measured by Garmany, Conti, \& Massey (1980), 30\% contained close binaries with mass ratios exceeding 0.4, and, according to their analysis, 40\% of massive stars were in binaries. Hogeveen (1990), on the other hand, argues that Garmany, Conti, \& Massey (1980) underestimated the number of binaries with low-mass companions and instead estimates a much larger fraction of binary massive stars ($\sim $70\%) peaked toward extreme mass ratios ($\alpha_{\rm MR}=2.7$). For the merger rates in Table 2, we always assume a binary fraction of 40\%. However it is likely that the binary fraction, and hence the merger rates, for the $\alpha_{\rm MR}=2.7$ calculations are a factor of 2 higher. Even with this increase of 2 of the binary fraction, these extreme mass ratios predict over an order of magnitude fewer DNS binaries (which, dominated by Scenario II, require nearly equal mass binaries) and nearly a factor of 3 more WD/BH binaries (Table 2). In this case, WD/BH mergers are more common than DNS mergers! For initial orbital separations, we use the distribution introduced by Kraicheva et al. (1979): \begin{equation} P(A_0) \propto 1/A_0 \end{equation} with constraints on the innermost initial separation (equal to 3 times the sum of the stellar radii) and on the outermost initial separation (equal to $10^4 R_\odot$). Increasing this inner separation by a factor of 2, or increasing the outer separation by 2 orders of magnitude, raises (or lowers for the outer limit) the merger rates by less than 20-30\%. Likewise, assuming the initial eccentricity is evenly distributed between 0 and 0.9 (versus assuming initially circular orbits) also has less than a 40\% effect (Table 2). With the exception of the mass ratio distribution, uncertainties in the initial conditions have little effect on the merger rates of neutron star binaries. For our standard set of parameters, we use, as do most of the previous calculations of DNS binaries, a flat mass ratio distribution ($\alpha_{\rm MR}=0$) with a binary mass fraction of 40\%. However, the standard formation scenario for low-mass X-ray binaries requires that the mass ratio distribution is peaked toward extreme mass ratios (Kalogera \& Webbink 1998). Unless low-mass X-ray binaries are formed via some alternative mechanism, $\alpha_{\rm MR}$ must be higher ($\sim 2.7$), and we are overestimating the DNS merger rate by over an order of magnitude and underestimating the WD/BH merger rate by a factor of 3. For the rest of our standard set of parameters, we use the Scalo (1986) initial mass function, initially circular orbits, and the distribution of separations and inner and outer limits discussed in the previous paragraph. \subsection{Uncertainties in the Supernova Model}\label{sec:sn} To form the binaries containing compact objects that might eventually produce BHAD GRBs, we must not only follow the evolution of binary stars, but also take into account the many uncertainties associated with the supernova process. One key quantity, to which all our formation scenarios are proportional, is the supernova rate and its history in the universe. We assume a current Galactic supernova rate of 0.02\,yr$^{-1}$, consistent with the results of Cappellaro et al. (1997), but even that is uncertain by at least a factor of 2. As one moves to the distant past, the rate is presumably larger for our Galaxy, but then galaxies too were evolving in size and luminosity. We discuss in \S 4 the scaling of our supernova rate to distant redshifts. Another source of great uncertainty in estimating the formation rate of DNS and BH/NS systems is an inadequate understanding of the kicks imparted to neutron stars (or black holes) at birth. Growing evidence suggests that neutron stars receive a momentum boost, or ``kick'', ($v_{\rm kick}^{\rm mean} \sim 300-500$\,km\,s$^{-1}$) during their formation, probably due to some asymmetry in the explosion itself. Evidence includes the pulsar velocity distribution (Dewey \& Cordes 1987; Lyne \& Lorimer 1994; Fryer, Burrows, Benz 1998; Cordes \& Chernoff 1998), associations of neutron stars with supernova remnants (Caraveo 1993; Frail, Goss \& Whiteoak 1994 although see Gaensler \& Johnston 1995), observations of bow shocks produced by neutron stars as they plow through the interstellar medium (Cordes, Romani, \& Lundgren 1993), and the formation scenarios for DNS systems and many other close binary systems(DNS systems: Flannery \& van den Heuvel 1975; Burrows \& Woosley 1986; Yamaoka, Shigeyama, \& Nomoto 1993; Fryer \& Kalogera 1997; spin/orbit angles in binaries: Kaspi {\it et al.} 1996; Wasserman,Cordes, \& Chernoff 1996; eccentric binaries: van den Heuvel \& Rappaport 1986). Although small kicks ($\sim 200$\,km\,s$^{-1}$) are required to match the observed DNS systems, larger kicks tend to disrupt the binaries. Recent analyses of the pulsar velocity distribution suggests a bimodal kick distribution with roughly half of the neutron stars receiving kicks smaller than $200\,$km\,s$^{-1}$ and the other half receiving kicks greater than $500-600$\,km\,s$^{-1}$ (Fryer, Burrows, \& Benz 1998). The binary systems whose neutron stars receive very large kicks become unbound. The remaining systems may form DNS and BH/NS binaries. The choice of kick magnitude has a strong effect on the formation rate of these two compact binary progenitors (Fig. 14). During the supernova explosion, the binary only remains bound only if $M_{\rm ejecta} < 0.5 M_{\rm sys}r/a_0$ where $M_{\rm sys}$, $r$, and $a_0$ are the total mass, orbital separation, and semi-major axis of the binary (Hills 1983). With the circular orbits important for close-binary formation (common envelope evolution circularizes the orbits), this equation becomes $M_{\rm ejecta} < 0.5 M_{\rm sys}$. When a kick is imparted onto the neutron star during the supernova explosion, the criterion for bound systems is: \begin{equation}\label{eq:kick} M_{\rm ejecta} < 0.5 M_{\rm sys} \left[ 1-2 (v_{\rm kick}/v_{\rm orb}) cos \theta cos \phi - (v_{\rm kick}/v_{\rm orb})^2 \right], \end{equation} where $v_{\rm kick}$ and $v_{\rm orb}$ are the magnitudes of the kick and orbital velocities respectively. $\theta$ and $\phi$ are the polar angle directions of the kick velocity with respect the the orbital velocity. Although for some systems, a kick might actually prevent the disruption of the binary (this is important for low-mass X-ray binary formation: Kalogera \& Webbink 1998; Kalogera 1998, Fryer, Burrows, \& Benz 1998), in general, the kick tends to unbind binaries. For an isotropic kick distribution, the probability that a system remains bound after the supernova explosion is given by (Brandt \& Podsiadlowski 1995): \begin{equation} P(v_{\rm kick},v_{\rm orb},M_{\rm ejecta},M_{\rm sys})= 0.5+0.5\frac{v_{\rm orb}}{v_{\rm kick}}\left[ \frac{M_{\rm sys}-M_{\rm ejecta}}{M_{\rm sys}}-0.5 \left( \frac{v_{\rm kick}}{v_{\rm orb}} \right)^2 -0.5 \right]. \end{equation} We performed a series of Monte-Carlo simulations assuming delta function kicks (Fig. 14). Note that the merger rates of DNS binaries initially increase as the mean kick velocities increase from $0 \, {\rm km \, s^{-1}}$. The actual binary {\it formation} rates are much higher for lower kicks (Fig. 15), but higher kicks are required to form the close binaries which will merge within a Hubble time. The presupernova orbital separation must be sufficiently wide to avoid a common envelope phase which would destroy the binary and a kick is required to tighten the orbits and make close merging binaries. In DNS binary formation, the primary's neutron star must avoid common envelope evolution with its secondary's helium star (Fryer \& Kalogera 1997) or it will accrete too much material and become a BH/NS GRB. Similarly, low-mass stars must have sufficiently wide orbital separations so that they do not merge with their massive companions before the massive star collapses to form a black hole. Clearly, the radius of the helium star is important for these calculations. We will return to this effect in \S \ref{sec:sevo}. The actual kick distribution is probably not a delta function, but spans wide range of velocities. In Table 3, we give the merger rates for both a Maxwellian velocity distribution (with a mean of 450$\, {\rm km \, s^{-1}}$) and a double peaked distribution similar to that described by Fryer, Burrows, \& Benz (1998) with two peaks near 100$\, {\rm km \, s^{-1}}$ and 600$\, {\rm km \, s^{-1}}$ with $\sigma=50,150 {\rm km \, s^{-1}}$ respectively (``FBB kick distribution''). Note that although the FBB and Maxwellian kick distributions have roughly the same average magnitude (and both roughly explain the pulsar velocity distribution), the FBB kick distribution produces many more binary systems. Because the merger rates depend so sensitively upon the kick distribution, we will employ several kick velocities in our studies. At the present time, no known mechanism provides the large kicks implied by the pulsar observations. It is therefore doubly difficult to predict the distribution for black holes since the velocities of single black holes have not been measured. Kick mechanisms driven by asymmetries in the supernova explosion seem likely to impart smaller kick velocities to black holes, if only because they are more massive. If anisotropic neutrino emission is involved, for example, the black hole may also form before much neutrino momentum has been emitted and therefore experience less recoil. As one might expect, setting the black hole kick magnitudes to 10\% that of neutron stars implies much larger black hole binary merger rates (Table 3). Another uncertainty in supernova explosions is the limiting main sequence mass star that will form a black hole, either by fallback ($M_{\rm BH}$) or prompt collapse ($M_{\rm Coll}$). Current supernova models (Fryer 1999) suggest that $M_{\rm BH}$ is not much above $25 {\rm M_\odot}$ which agrees roughly with observations (Maeder 1992; Kobulnicky \& Skillman 1997; Portegies Zwart, Verbunt, \& Ergma 1992; Ergma \& van den Heuvel 1998). In Fig. 16 (and Table 3), we show results from simulations with $M_{\rm BH}$ ranging from $25 M_\odot$ to $85 M_\odot$ and $M_{\rm Coll}=M_{\rm BH}+15$M$_{\scriptscriptstyle \odot}$\, using the FBB kick distribution. A higher mass limit leads to the formation of fewer black holes, and hence, fewer BH/WD binaries and collapsars. The merger rate of BH/NS binaries does not depend sensitively on the black hole mass limit. This is because the dominant formation scenario (Sc. V) requires the primary to collapse into a neutron star which then accretes matter in a common envelope phase and collapses to a black hole (see \S \ref{sec:bev}). If hyper-critical accretion does not occur, formation scenario V does not work and, as shown in Fig. 16, the formation rate of BH/NS binaries depends more sensitively upon the critical black hole mass limit. For our standard set of simulations, we set $M_{\rm BH}=25 $M$_{\scriptscriptstyle \odot}$ and $M_{\rm Coll}=40 $M$_{\scriptscriptstyle \odot}$. In our simulations, we also assume that massive stars collapse to form black holes with masses set to 1/3 the primary star's mass at the time of collapse. For stars with masses between $M_{\rm BH}$ and $M_{\rm Coll}$, a supernova ejects some of the mass. Even beyond $M_{\rm Coll}$, not all of the star will collapse to form a black hole. As the gamma-ray burst forms, it also may drive off part of the helium core. The resultant mass distribution of black holes in binaries with massive companions is double peaked (Fig. 17). The large peak at 2.4\,M$_{\scriptscriptstyle \odot}$ ($> 50$\% of the black holes) is caused by those black holes formed by hyper-critical accretion onto neutron stars in common envelope evolution. These low-mass black holes will produce the most energetic BH/NS merger GRBs. Their actual masses depend sensitively upon common envelope evolution, but will probably not exceed 4\,M$_{\scriptscriptstyle \odot}$. The second, spread-out distribution, is even more difficult to quantify, as it depends not only upon binary evolution, but also the rates of stellar winds and the mass of the ejecta in supernova explosions. In fact, assuming that the mass of the black hole is only 1/3 that of the primary star's mass is probably a lower limit to the black hole mass (Fryer 1999). Setting the black hole mass equal to the collapsing star's mass slightly increases the rates of black hole binaries for low kicks. However, for high kicks, the high momentum imparted to these more massive systems tends to unbind the system. \subsection{Stellar Evolution}\label{sec:sevo} Uncertainties in stellar models are often neglected but, as we shall show in this section, they can have a large effect on the rates for the BHAD models. The most important uncertainty in the stellar models is the stellar radius, especially as the star moves off the main sequence. The radius of a star is important because it determines if, and when, a binary system passes through a common envelope or mass transfer phase. Accurate stellar radii are difficult to determine observationally and current theoretical models concentrate on the cores of massive stars. Many of the theoretically predicted radii depend upon artificial outer boundary conditions. Indeed, especially for massive giant stars and for Wolf Rayet stars, the radius is not a very precisely defined quantity. Observations tell us only about the radius of the photosphere. This size can be much larger than the radius which is important for population synthesis calculations, that is, the radius where the stellar density is sufficient to drive common envelope evolution (see \S \ref{sec:bev}). For both giant and helium stars, our standard model assumes that the common envelope radius is 25\% the radius given in the stellar model fits by Kalogera \& Webbink (1998). If we raise this to the values given by Kalogera \& Webbink (1998), both the rates of DNS and BH/NS binaries decrease over an order of magnitude in some cases (Table 4). Increasing stellar radii causes more binaries to go into common envelope phases, but, more importantly, increasing helium-star radii cause more binaries to merge (producing fewer compact binaries and more helium merger GRBs). In addition, many of the formation scenarios (e.g. I, II, IV, V) produce compact stars (neutron star, black hole) around helium-star companions. As the helium-star radius increases, the orbital separations of compact binaries become wider and consequently fewer {\it merging} compact binaries are produced. Thus, for a given kick distribution, increasing the helium-star radius causes both the number of binaries and the fraction of those binaries which will merge in a Hubble time to decrease (see Fryer \& Kalogera 1997). Winds affect both helium star masses and radii. For instance, if we use a mass loss rate like that of Woosley, Langer, \& Weaver (1993,1995), helium masses are relatively small (e.g., a massive star forming a $20 M_\odot$ helium core will end its life with a $3.5 M_\odot$ helium core), and their radii are greater. For larger helium star radii, fewer close binaries form, and the merger rates of both DNS systems and BH/NS systems decrease (Table 4). In addition, high winds raise the critical mass for black hole formation (less massive stars simply do not have massive enough cores at time of collapse to prevent a supernova explosion) which reduces the rates of all black hole systems (the collapsar rate decreases by an order of magnitude). Note that this mass loss is an extreme case and some helium stars must not have such drastic winds if the formation scenarios for black hole X-ray binaries such as V404 Cyg, Cyg X-1, and XN Mus 92, are correct (White \& Paradijs 1996). In our simulations, we parameterize the mass loss rate as a fraction of the rate used by Woosley, Langer, \& Weaver (1993,1995): $\alpha_{\rm Wind}=$(Mass Loss)/(Mass Loss$_ {\rm WLW}$). Current mass loss estimates suggest that the WLW models overestimate the maximum mass-loss by a factor of two to three. The effect of mass-loss is minimal unless $\alpha_{\rm Wind}$ exceeds 0.6 (Fig. 18). If the mass loss never exceeds this value, only the collapsar model is affected at all by mass loss. As the mass loss rises from $\alpha_{\rm Wind}=0.$ to $\alpha_{\rm Wind}=0.6$, the collapsar rate is cut by a factor of 2-3. We have simplified the uncertainties by assuming the error in the red supergiant wind is the same as that of Wolf-Rayet winds. We decrease the mass loss in the red supergiant wind by the same factor that we decrease the mass loss for helium stars by lumping the mass loss uncertainty into one parameter. With this assumption, we are probably underestimating the number of collapsars formed via winds (Sc. X), perhaps grossly so. Using just one parameter, those stars with winds strong enough to uncover the helium core also lose too much of their helium core to collapse directly into a black hole. Observations suggest that single stars above ($\sim 40$\,M$_{\scriptscriptstyle \odot}$\,) must lose their hydrogen envelope (Chiosi \& Maeder 1986), which would require a high value for the wind parameter: $\alpha_{\rm Wind}>0.8$. However, to match the black hole binaries, the wind parameter for helium stars must be low: $\alpha_{\rm Wind}<0.5$ (Kalogera 1999). Assuming that black hole binaries are formed in systems which do not go into common envelope evolution until after central helium exhaustion, Wellstein \& Langer (1999) are able to explain the black hole binaries but maintain a high mass loss rate. If their models are correct, very few massive stars, single or binary, would have enough mass prior to collapse to form black holes directly, and the formation rate of collapsars in the present-day universe would drop dramatically. However, since mass loss is metal dependent, the collapsar rate might be larger at higher redshifts than it is today (MacFadyen \& Woosley 1999). \subsection{Binary Evolution}\label{sec:bev} For binaries to merge and form GRBs, the progenitor systems must have short periods. consequently, most will undergo some sort of active binary evolution. We have assumed that if the mass of the expanding star ($M_{\rm exp}$) is less than $\sim$2.5 times the mass of its companion ($M_{\rm comp}$), the companion is unable to incorporate more than a small fraction of the overflowing mass (Webbink 1979). A fraction ($\beta_{\rm MT}$) of the expanding star's envelope is accreted by the companion, and the remainder is lost from the system, carrying away a specific angular momentum ($\alpha_{\rm MT}$). The orbital separation of the system ($A$) in terms of its initial separation ($A_0$) after mass transfer is given by (Podsiadlowski, Joss, \& Hsu 1992): \begin{equation} \frac{A}{A^0}=\frac{M_{\rm exp}+M_{\rm comp}}{M_{\rm exp}^0+M_{\rm comp}^0} \left ( \frac{M_{\rm exp}}{M_{\rm exp}^0} \right)^{C_1} \left ( \frac{M_{\rm comp}}{M_{\rm comp}^0} \right)^{C_2} \end{equation} where the values of the constants are: \begin{equation} C_1 \equiv 2 \alpha_{\rm MT} (1-\beta_{\rm MT}) - 2 \end{equation} \begin{equation} C_2 \equiv \frac{-2 \alpha_{\rm MT}}{\beta_{\rm MT}} (1-\beta_{\rm MT}) - 2 \end{equation} and \begin{equation} M_{\rm comp}=\beta_{\rm MT} (M_{\rm exp}^0-M_{\rm exp}) + M_{\rm comp}^0, \end{equation} and where the superscript 0 denotes pre-mass transfer phase values. Although the values of $\alpha_{\rm MT}$ and $\beta_{\rm MT}$ are relatively uncertain, varying $\alpha_{\rm MT}$ from $0.5-2.0$ does not change the GRB rates by more than 30\%(Table 5). However, changing the amount of material lost by the system during mass transfer has a similar effect to that of changing the mass loss by winds. Increasing $\beta_{\rm MT}$ from 0.5 to 1.0 decreases the BH/NS merger, collapsar, and helium merger rates by up to an order of magnitude (Table 5). If $M_{\rm exp} \gtrsim 2.5 M_{\rm comp}$ (if the companion is a compact object, this factor of 2.5 reduces to 1.0), the companion can not adjust to the accreting mass and instead enters into a common envelope phase. Once a common envelope is created, the orbital separation decreases rapidly, probably in $1-1000$\,yr (see Sandquist et al. 1998 and references therein). This inspiral continues until the binary is able to shed the expanding star's hydrogen envelope leaving a bare helium core. One can estimate the final orbital separation ($A_f$) in terms of the initial separation ($A_i$) of the binary after common envelope evolution (Webbink 1984): \begin{equation} \frac{A_{f}}{A_{i}}=\frac{\alpha_{\rm CE}\, r_{L}\, q}{2} \left(\frac{M_{He}} {M_{s}-M_{He} + \frac{1}{2}\alpha_{\rm CE} r_{L} M_{NS}} \right), \end{equation} where $M_{s}$, $M_{He}$, and $M_{NS}$ are, respectively, the masses of the secondary, the secondary's helium core, and of the neutron star and $r_{L}=R_L/A_i$ is the dimensionless Roche lobe radius of the secondary. $\alpha_{\rm CE}$ is a parameter designed to represent how efficiently the orbital energy ejects the hydrogen envelope and includes assumptions about the binding energy of the expanding star and the amount of orbital energy injected, per gram, into the hydrogen envelope. The value of $\alpha_{\rm CE}$ is not well constrained, nor is it constant. It depends both upon the mass and evolutionary stage of the expanding star when a common envelope phase occurs. Varying the common envelope efficiency over a range (0.25-1.0) varies the GRB rates by less than a factor of 3 (Table 6), and one might naively assume that common envelope evolution does not affect the GRB rates. However, common envelope evolution may be drastically different from that given by the simple $\alpha_{\rm CE}$ prescription. Nearly all past population synthesis calculations ignore the possible merger of an inspiralling compact object with the core of its companion in a common envelope phase. In most cases, previous population synthesis codes simply did not check to see when the merging compact object fell within the helium core radius. However, such an assumption may not be entirely incorrect. Terman, Taam, \& Hernquist (1995) found that, if the star is sufficiently evolved to produce a steep density gradient between the helium core and the hydrogen envelope, the inspiral of the compact object might halt just above the helium core, preventing the merger of the helium core and compact object. Although their findings were greatly influenced by inadequate resolution ($\sim 3000$ particles) and the subsequent effects of the numerical artificial viscosity inherent in their smooth particle hydrodynamical technique (Rasio \& Livio 1996), nature may still produce similar results. At this time, common envelope evolution is not well enough understood to derive from it any reliable constraints. If we ignore the non-zero helium radius, the rate of DNS and BH/NS mergers increases nearly up to $\sim 10^{-4} \, {\rm yr^{-1}}$(Table 7). However, since this particular orbital evolution is currently only justified by an incorrect hydrodynamical simulation, this high merger rate should be considered with suspicion. Another critical aspect of common envelope evolution is the amount of accretion experienced by a compact object spiraling into the hydrogen envelope of a companion star. In past calculations, it has been assumed that in this state a neutron star will accrete at the Eddington rate. However, studies of the conditions arising near the base of a neutron star suggest that neutron stars in common envelopes will actually accrete at the Bondi-Hoyle rate since the accreting material loses its energy via neutrino emission (Chevalier 1993, 1996; Houck \& Chevalier 1992; Brown 1995; Fryer, Benz, \& Herant 1996). Assuming all neutron stars inspiral to the surface of the expanding star's helium core, Bethe \& Brown (1998) estimated that all neutron stars in common envelope would accrete roughly $1 M_\odot$ of material and collapse to form black holes. This collapse can only be avoided only if: angular momentum in the accretion hydrogen is sufficient to lower the accretion rate (Chevalier 1996), neutrino driven explosions eject the accreting hydrogen allowing only small spurts of accretion (Fryer, Benz \& Herant 1996), or the star simply doesn't inspiral to the surface of the helium core where the densities of the hydrogen envelope lead to high temperatures at the surface of the neutron star. Although angular momentum can halt accretion along the axis, it is unlikely to prevent the rapid accretion onto the neutron star simply because accretion is unimpeded along the poles. Especially at the high densities near the helium core surface (where Bethe \& Brown 1998 predict most of the accretion to occur), it is unlikely that angular momentum will limit the accretion rate sufficiently to prevent neutrino cooling (Fryer \& Kalogera 1998). Similarly, although neutrino driven explosions may prevent accretion in the outer layers of a hydrogen giant, near the helium core surface, this explosion mechanism is ineffective (Fryer, Benz, Herant 1996). However, if the neutron star does not inspiral all the way down to the helium core surface, it may not accrete enough to become a black hole and will remain a neutron star. Note that by ignoring hypercritical accretion and assuming that compact objects will not spiral into the helium core produces a DNS merger rate comparable to the rates of past work, most of which made these two, physically unlikely assumptions(Table 7). In addition, these binaries have much shorter orbits on average, and hence shorter merger times. In our standard model, we assumed that all the common envelope neutron stars form black holes and, hence, scenario I (see \S \ref{sec:dns}) is excluded from these simulations. \subsection{Summary and Constraints} Table 8 summarizes the range of results from all the population synthesis calculations. The basic uncertainties in the rates can be grouped into three categories: kicks, progenitor evolution, and orbital separations. Except for the collapsar model, all the BHAD GRB models depend sensitively (up to three orders of magnitude in the rate) upon the magnitude and distribution of the kick velocities. Uncertainty here dominates the errors in these BHAD GRB rates. On the observational front, additional data on pulsar velocities would be useful. From the theorist we still seek an understanding of why the kicks are so large. Uncertainties in the progenitor evolution include the mass ratio distribution in binaries, which determines the mass of the secondary and the compact object into which the secondary will evolve, and the mass limits for black hole and collapsar formation. The mass ratio distribution is difficult to determine from first principles, but it better observations could certainly constrain this uncertainty. The mass limit for collapsars on the other hand is something theorists can work on. Together, these uncertainties affect all the GRB rates, and can vary the WD/BH merger and collapsar rates by $\sim 2$ orders of magnitude. Finally, we have lumped all the uncertainties of binary evolution into the seemingly innocuous category: orbital separation distribution. This category is dominated by uncertainties in the stellar radii and the common envelope evolution. Most of the high DNS rates listed in Lipunov, Postnov, \& Prokhorov (1987) are derived by assuming that neutron stars do not accrete hypercritically and do not merge with their companion's helium core (most previous work used population synthesis codes which do not even determine whether the neutron star spirals into its companion's helium core during a common envelope phase). If we take these assumptions as incorrect and instead rely upon the $\alpha_{\rm CE}$ prescription for common envelope evolution, the uncertainties in this category result in roughly an order of magnitude decrease in the DNS GRB rates. Comparison of population synthesis models with observations of compact binary systems can place some constraints upon the allowed parameter space. For instance, in Fig. 19 we plot the simulated distribution of eccentricities and separations of double neutron star systems superimposed on the 4 observed galactic double neutron star systems\footnote{van Kerkwijk \& Kulkarni 1999 have discovered an observational candidate of PSR 2303+46 ($A=30$R$_{\scriptscriptstyle \odot}$\,,$e=0.66$), which would mean that this system consists of a neutron star and a white dwarf and is not a DNS system at all.}. Although with so few systems and no knowledge of the observational biases, we do not dare do any statistics on this sample, we can probably rule out very large stellar radii and low kick velocities ($v_{\rm kick}=50$) because such a set of parameters would have a very small probability of producing the observed systems (see Fryer \& Kalogera 1998 for more details). The remaining three simulations all reproduce the observed systems, but predict a range of merger rates: $10^{-7}$\,yr$^{-1}<\,{\rm Merger Rate}_{\rm DNS}\, <5 \times 10^{-5}$\,yr$^{-1}$. Similarly, the existence of short-period black hole binaries requires that not all helium stars have the strong winds proposed by Woosley, Langer \& Weaver (1995). Additional constraints exist and may well provide more stringent limits on the binary formation rate. Until these categories improve and better limits are constructed, we must be satisfied with the rates in Table 8. However, the rates in table 8 provide useful constraints nonetheless. To get the extremely high merger rates of DNS and BH/NS binaries, one must assume that compact objects do not merge with their companion's helium core during a common envelope phase. Otherwise, the rate of mergers is at least an order of magnitude lower which, although this decrease may prove troublesome for LIGO and VIRGO, it is certainly sufficient to explain GRBs (see \S 4). Collapsars and He-mergers also have rates large enough to explain the GRB statistics. It is unlikely that BH/WD mergers explain a sizable fraction of GRBs, but they may be required to explain some peculiar GRBs. \section{Implications For Observations} Table 9 summarizes the event rate, distance traveled before producing a burst, and mean redshift for the models we have studied. In addition we give estimates of the burst energy and duration (from other works, e.g., Popham et al. 1999; Meszaros \& Rees 1997a). There is effectively no delay for collapsars and He-mergers which produce bursts occurring, respectively, within $\sim$10\,Myr and $\sim$ 50\,Myr of their formation. However most DNS and BH/NS systems experience a considerable delay between their formation and merger and, during that tie, may travel a considerable distance. How far depends upon the mass of the host galaxy. Galaxies similar to the Milky Way have escape velocities of order 400-700 km/s, depending on the location inside the galaxy. The angular separation between GRBs and the centers of their host galaxies is thus an important diagnostic (e.g., Portegies-Zwart \& Yungelson 1998; Bloom, Sigurdsson, \& Pols 1999b; Bulik, Belczy\'nski, \& Zbijewski 1999). There is of course the possibility of a strong selection effect. It might be that GRBs which occur far from their host galaxy are less likely to produce observable afterglows because of the low density of the ISM in galactic halos (Meszaros \& Rees 1997b). This would then mean that we have to multiply the event rates by a corresponding factor much less than one. In this section, we restrict our analysis to 5 simulations that we feel represent the range of most-likely population synthesis parameters. These include the standard set of parameters, a simulation with winds set to 50\% that of Woosley, Langer \& Weaver (1995), and a simulation with the stellar radii raised to those of Kalogera \& Webbink (1998) all using the FBB kick distribution. Two simulations were run using a Maxwellian kick distribution, one with the standard set of parameters, and one with the stellar radii equal to Kalogera \& Webbink (1998). However the latter of these produced so few binaries that we do not include it in the figures (although the results are in the tables). \subsection{The angular separation between OTs and host candidates} There now exists a small, but growing sample of optical GRB afterglows (OTs) associated with host galaxies. In almost every case the angular offset of the OT from the photometric center of the host is of order arcseconds. A linear distance of 10 kpc corresponds to an angular separation of 0.7 h$_{100}$ F(z) arcsec, where F $\sim 3$ for redshifts in the range 1-3 (e.g., Peebles 1993) and h$_{100}$ is the Hubble constant in units of 100 km\,s$^{-1}$\,Mpc$^{-1}$. Host redshifts are known for GRB970508 (z = 0.838; Metzger {\it et al.} 1997; Bloom {\it et al.} 1998), GRB971214 (z = 3.42; Kulkarni {\it et al.} 1998), GRB980703 (z = 0.966; Djorgovski {\it et al.} 1998), and GRB980613 (z = 1.096; Djorgovski {\it et al.} 1999). These observations clearly argue against any model that invokes massive black holes located at the centers of active or normal galaxies. They also have important implications for models discussed in this paper. We have modeled the spatial distribution of DNS and BH/NS GRBs by calculating orbits in three representative galactic mass models of Miyamoto-Nagai type (see Hartmann, Epstein, \& Woosley 1990 for details of the dynamic model): a massive galaxy resembling the Milky Way (with V$_{\rm rot}$(R$_\odot$) = 220 km/s), a system with a mass four times smaller, and a system with 1/100th the mass of the Milky Way (essentially a negligible mass). The average initial positions of the GRB progenitors in each class were taken to be an axisymmetric ring with galactocentric radius R = 5 kpc with a Gaussian spread of 1 kpc around that radius. The orbits were following the formation of the primary's compact object. For the DNS and BH/NS binaries, the binaries, this process required two steps, the evolution before and after the collapse of the secondary. Systemic velocities were added as isotropic vectors, while the initial velocity was assumed to be that for circular motion with V$_{\rm rot}$. The orbits were followed for a period given from the Monte Carlo population synthesis model. Several thousand orbits were used for each BHAD model. At the moment of final merger the distance to the galactic center was recorded. Fig. 20 shows the resulting distance distribution for 4 different sets of population synthesis parameters. Integrating these curves, it is clear that the models predict a substantial fraction of DNS and BH/NS GRBs at large distances from their hosts (see Fig. 21, Table 10). In Fig. 21, we plot the distribution of DNS and BH/NS binaries separately, but clearly the two GRB progenitors are very similar reflecting their similar formation history. In addition, because many of the binaries are composed of low-mass black holes (Fig. 17), the velocity boost received to the system as the secondary forms a neutron star is very similar to that received in the case of DNS systems. Our results agree well with those of Bulik, Belczy\'nski, \& Zbijewski (1999) which have a wider spread than that of Bloom et al. (1999b). The population synthesis studies of Bloom et al. (1999b) probably did not include a non-zero helium star radius, allowing them to form DNS binaries with shorter periods, and hence shorter merger times (see \S 3.4). Fig. 21 also shows the distribution of separations compared to the the observations of OTs corresponding to long hard bursts. It appears that DNS and BH/NS mergers are ruled out as the likely explanation for most of these events. The observations instead support the idea that long duration bursts are either collapsars, helium-mergers, or WD/BH mergers, all of which tend to occur well within the galaxy (see Fig. 21). Indeed, collapsars should all occur at their star formation site. The spread in Fig. 21 reflects only the assumption regarding galactocentric radius for star formation (and is thus also the birth line for the progenitors of DNS and BH/NS systems). Ideally this radius should be adjusted to the size of the observed GRB host galaxy, so the agreement with the ``promptly dying'' models is really better than the figure indicates. \subsection{GRBs and the Star Formation History of the Universe} Since only very massive stars are believed to produce black holes, all of our models are in one way or another dependent upon the star formation history of the universe. Helium-mergers and collapsars trace it most directly since the delay time between star birth and the GRB is, by cosmic standards, instantaneous. The relatively infrequent WD/BH mergers may be delayed by a longer time ($\sim 100$\,Myr) but still should trace star formation directly. DNS and BH/NS systems, on the other hand have a delay and though they reflect the star formation rate, they do so with a spread of delays, depending especially on the gravitational radiation time scale. Fig. 22 shows the distribution of DNS+BH/NS merger times for our four different sets of population synthesis parameters. Note that all four have some short-period mergers, but the typical case gives a GRB that lags behind the star formation rate. Fig. 23 shows the GRB redshift history for an Einstein-deSitter universe using two different star formation histories under the functional form given by Watanabe et al. (1998): flattened - $A=3., \, z_p=1.5, \, B=0.$ and peaked - $A=1., \, z_p=1., \, B=0.5$. The flattened and peaked formation histories represent extremes in the star formation. Recent studies of the cosmic star formation rate indicates that the SFR(z) density function does not turn over past z $\sim$ 1.5 (e.g., Pascarelle, Lanzetta, $\&$ Fernandez-Soto 1998), but instead remains roughly constant to redshifts of $\sim$ 5. Beyond a redshift of 5, very little is known about the star formation history. At some redshift, the star formation rate must drop. Differences in the population synthesis models cause considerable scatter in the median redshift of GRBs (Table 11). For a flattened star formation rate, the median redshift varies from 0.85-1.2. Note that the peaked star formation rate leads to much lower median burst redshifts (0.54-0.81). Collapsars and helium-mergers will produce bursts which trace the star formation more directly and hence their median redshifts are higher: 1.8, 0.96 respectively for the flattened, peaked star formation histories (Table 10). The median redshift of collapsar GRBs is likely to be even higher, as its rate increases with decreasing mass loss. The low metallicity stars born at high redshift are likely to have weaker winds and, hence, produce more collapsars. The current redshift estimates of z=1-3 from the observations (e.g., Hogg and Fruchter 1999) lie within our two star formation rates for all of the GRB models. Unfortunately, the uncertainty in the star formation rate beyond a redshift of 1.5 makes it impossible to rule out any of the models. However, even with the uncertainties in the dominant BHAD model and the population synthesis models, the two star formation rates can be easily distinguished. If an increased sample of GRB redshifts pushes toward a median GRB redshift of 2, then no BHAD GRB model can explain the burst distribution with the peaked star formation history. In this way, GRBs can be used to place constraints on the star formation history. Another constraint on the star formation history is the fact that no lensed GRBs have been observed (Holz, Miller, \& Quashnock 1999). Using their analysis of constant redshift GRB distributions, we can constrain the GRB histories calculated here. For a flat SFR history combined with our standard model and the FBB kick distribution, only 30\% of the DNS+BH/NS GRBs have redshifts greater than 4, and less than 10\% have redshift greater than 8. The flat SFR history again must be restricted if the he-mergers and collapsars are the dominant long-duration bursts. The peaked SFR history predicts few lensed GRBs, and hence a lensed event would argue against this star formation history. Table 11 gives the total number of bursts per day for the three star formation rates. To then determine how many bursts we will observe requires a convolution with the detection efficiencies and the beaming factor, which lower this rate by a factor of 30-1000. Within the uncertainties in the conversion from actual bursts to observed bursts, the star formation history, and the rates, all of the BHAD GRB progenitors can explain the observed bursts. \section{BHAD GRB Models} The merger rates, durations, energies, median redshifts and locations of the BHAD GRB models are summarized in Table 9. Our estimated rates, median redshifts and locations, coupled with the previously calculated energies and durations (Popham, Woosley, \& Fryer 1998; Ruffert \& Janka (1998); Janka, Ruffert, \& Eberl 1998) compare well with the current observational limits. From the data in Table 9, we can construct a general picture of GRB progenitors in which short-duration bursts are dominated by DNS and BH/NS mergers, and long-duration bursts are dominated by collapsars and helium mergers. If this picture is correct, we can make several observational predictions which will be tested in the next 10 years. We predict that all long-duration bursts should reside in their host galaxy, whereas some short-duration bursts should occur beyond their host galaxy. In addition, the median redshift of short-duration bursts should be less than that of their long-duration counterparts. \subsection{Short-Duration Bursts} DNS and BH/NS binary mergers are capable of producing short duration bursts. Their total rate very likely lies in the range of 0.1-30\,Myr$^{-1}$ in the Milky way, which corresponds to an overall rate in the universe of 1-1500\,day$^{-1}$. To determine an observed rate, one must then include the beaming factor and detector efficiencies. Note that this rate is much less than has been predicted in previous calculations (Lipunov, Postnov, \& Prokhorov 1987). Although this lower rate does not prevent DNS and BH/NS binary mergers from playing a major role in explaining GRBs, it has discouraging implications for the predicted rate of gravity wave events detectable with LIGO and VIRGO (for a review, see Finn 1999). DNS and BH/NS binaries are the only BHAD GRBs which can occur outside of their host galaxy and the only models capable of giving bursts much shorter than one second. Roughly 3-8\% of these bursts occur more than a few arc minutes from a host galaxy of Milky Way mass. This fraction can increase beyond 20\% for lower massed host galaxies. In addition, the median redshift of DNS and BH/NS merger GRBs is less, in some cases significantly less, than that of the median star formation redshift. The current bursts whose locations have been ``pinpointed'' by BeppoSAX are all long-duration bursts and hence, no constraints can be placed on the locations and redshift of short-duration bursts. However, since none of the observed long-duration bursts have been found outside of their host galaxy, the evidence is beginning to suggest that DNS and BH/NS binaries are limited to short-duration bursts. But much more data is required to draw any firm conclusions from the observations. DNS and BH/NS binaries also experience a significant delay (up to a Hubble time) between their formation and their merger to form a GRB. This delay causes their median redshift to be lower by 20-50\% than that of stars, and hence, the long-duration gamma-ray bursts. The major uncertainty in these calculations comes from uncertainties in common envelope evolution: accretion during common envelope evolution, the amount of inspiral that occurs and even stellar radii, which determine if a binary goes into a common envelope. Whereas hypercritical accretion requires a new mechanism for DNS formation (Brown 1995), it creates a new formation mechanism for BH/NS systems (Bethe \& Brown 1998). The inspiral in common envelope helps to produce short-period binaries, but if the inspiraling compact object merges with its companion, it forms a helium-merger GRB, not a DNS or BH/NS binary. \subsection{Long-Duration Bursts} Collapsars, helium-mergers, and WD/BH binaries all definitely make long-duration bursts. Because we do not know how many stars will have the necessary angular momentum and mass to produce collapsars, the collapsar rate is particularly uncertain. A relatively conservative estimate comes from employing only Scenario XII (Fig. 13) - the late time merger of two helium stars. Using standard parameters (\S 3) this mechanism alone gives a daily rate of 400\,day$^{-1}$ (Table 3). A very liberal estimate, valid if magnetic fields do not slow the rotation of the helium core, additionally includes Scenario XI and a comparable number of single star collapsars (Scenario X with low helium core mass loss rates and high red giant mass loss rates). This gives a value about two orders of magnitude larger. Likewise, since the merger of black holes with helium stars has not been studied in detail, its rate is uncertain (10-4000\,day$^{-1}$). The WD/BH binary rate is significantly smaller (0.1-50\,day$^{-1}$). Collapsars and helium-mergers occur in their host galaxy and trace the star formation rate. Hence, any long duration burst occurring outside of its host galaxy would present a problem for the simple picture assuming DNS and BH/NS mergers produce only short-duration bursts. Collapsars and He-mergers also occur nearly simultaneously with the star formation rate and will have a higher median redshift than short-duration bursts (Table 9). These bursts may also provide ideal diagnostics of the high redshift star formation history. A much larger sample of well localized GRBs is essential to test these concepts, and to lend support or to disprove the BHAD scenario for cosmological GRBs. Just like DNS and BH/NS systems, calculations of helium mergers and WD/BH binaries depend most sensitively upon the common envelope evolution. The uncerainties in collapsars are due primarily to the mass loss and spin rate of their progenitors. The mass loss depends upon stellar winds and the spin rate depends upon the coupling of the core to the stellar envelope. \acknowledgements This research has been supported by NASA (NAG5-2843, MIT SC A292701, and NAG5-8128), the NSF (AST-97-31569), and the US DOE ASCI Program (W-7405-ENG-48). It is a pleasure to thank Alex Heger for his invaluable advice. We are grateful for conversations with Alex Heger, Andrew MacFadyen, Norbert Langer, Bob Popham, and Thomas Janka that helped to elucidate some of the uncertain aspects of gamma-ray bursts, massive stellar evolution and hyperaccreting black holes.
\section{Full Lorentz Invariance: Proper and Improper} For many decades there was a strong theoretical and/or aesthetic prejudice for fundamental physical laws that were symmetric under both spatial and temporal reflection (parity and time-reversal invariance). When the $V - A$ character of weak interactions was established in the late 1950's, exact parity invariance was apparently empirically falsified.\cite{wu,ly} Soon after, the discovery of $CP$ violation apparently falsified exact time-reversal invariance.\cite{cf} It seemed as if only the Proper Lorentz Transformations were true symmetries of Nature. However, Lee and Yang recognised from the outset that parity can be retained as an exact symmetry, despite the $V - A$ character of weak interactions, provided that the ordinary particle spectrum is doubled.\cite{ly} From a modern theoretical and phenomenological perspective this requires every ordinary lepton, quark, gauge boson and Higgs boson to be paired with a mirror analogue.\cite{flv1} Since the ordinary and mirror particle sectors are but weakly coupled to each other, the resulting scenario is phenomenologically viable. The violation of parity invariance is therefore, remarkably, still an open question. A simple argument to be presented below shows that if the above type of exact parity symmetry exists in Nature, then so necessarily does a form of time reversal invariance. So, it is possible for the full Lorentz Group to be a completely unbroken symmetry of Nature!\cite{flv1} Of particular interest is the fact that the observed atmospheric and solar neutrino anomalies may be the first experimental manifestation of the mirror matter sector (mirror neutrinos to be specific).\cite{flv2,f,fv} In order to further test this proposal, neutral current based measurements probing both the atmospheric and solar anomalies are vital. Such measurements will determine whether the relevant ordinary neutrinos are transforming into other ordinary neutrinos, or into something more exotic such as mirror or sterile neutrinos. The gauge theoretic construction of a theory with exact parity symmetry is very easy to understand.\cite{flv1,fv} Consider a theory defined by a parity {\it violating} Lagrangian ${\cal L}$ which has gauge group $G$. This theory may, for instance, be the minimal Standard Model, or, more pertinently, the Standard Model augmented by nonzero neutrino masses and mixings. For every field $\psi$ in ${\cal L}$ introduce a mirror or parity partner $\psi'$. For spin-$1/2$ fields this requires of course that the $\psi'$ have opposite chirality to the $\psi$. The fields $\psi'$ are singlets under $G$ but transform under a gauge group $G'$ which is isomorphic to $G$, while the $\psi$'s are correspondingly required to be singlets under $G'$. The fields $\psi$ and $\psi'$ are placed into identical multiplets under their respective gauge groups $G$ and $G'$, and the discrete parity symmetry (schematically $\psi \leftrightarrow \psi'$) is enforced. The resulting Lagrangian is \begin{equation} {\cal L}_{\rm total}(\psi,\psi') = {\cal L}(\psi) + {\cal L}'(\psi') + {\cal L}_{\rm int}(\psi,\psi'), \end{equation} where ${\cal L}'$ is exactly the same function of the $\psi'$ fields as ${\cal L}$ is of the $\psi$ fields.\footnote{The dependence of the Lagrangian on first derivatives of the fields is of course understood.} The extremely important interaction term ${\cal L}_{\rm int}$ describes any gauge and parity invariant renormalisable coupling terms between the ordinary and mirror sectors. The above procedure was first carried out for the minimal Standard Model in the first paper quoted under Ref.\cite{flv1}, where it was shown that parity was a symmetry of the vacuum as well as the Lagrangian for a large region in Higgs potential parameter space. We will focus on this parameter space region from now on.\footnote{By extending the Higgs sector, it is possible to spontaneously break the parity symmetry together with the electroweak and mirror-electroweak gauge symmetries.\cite{bm}} We call the resulting theory the ``Exact Parity Model (EPM)''. The ordinary and mirror sectors are coupled by gravitation and ${\cal L}_{\rm int}$. The gravitational coupling is very interesting from the point of view of cosmology and the dark matter problem,\cite{bk} but will not be further discussed here. The nongravitational effects in ${\cal L}_{\rm int}$ in general feature photon -- mirror photon, $Z$ -- mirror $Z$ and Higgs -- mirror Higgs mixing. These particles are singled out because they are neutral under the electromagnetic and colour forces and their mirror analogues, so there are no exact conservation laws to prevent mixing. (Electron -- mirror electron mixing is forbidden by both ordinary and mirror electric charge conservation, for instance.) Unfortunately, cosmological constraints from Big Bang Nucleosynthesis make it unlikely that these effects will be seen in the laboratory.\cite{cg} We now come to the crux of the matter: {\it since neutrinos and mirror neutrinos are electrically neutral and colourless, they will in general mix if they also have nonzero masses.} Furthermore, we will see in the next section that the exact parity symmetry forces the mixing angle between an ordinary neutrino and its mirror partner to be the maximal value of $\pi/4$. I close this section with two brief comments. (i) Let the exact parity symmetry be denoted by $P'$. Note that it is different from the usual (broken) parity symmetry $P$.\footnote{For instance, the left-handed electron is transformed into the right-handed mirror-electron by $P'$, whereas it is transformed into the right-handed electron by $P$.} However, standard $CPT$ is still of course an exact symmetry of the theory. We can therefore define a non-standard time-reversal invariance $T'$ through $CPT = P'T'$ that must necessarily be exact if $P'$ is exact. The full Lorentz Group, including all Improper Transformations, is thus a symmetry of the theory.\cite{flv1} (ii) It is amusing to compare the `exact parity' idea with spacetime supersymmetry. Both extend the Proper Lorentz or Poincar\'e Group, and both require degree-of-freedom doubling. A crucial difference, though, is that phenomenology forces spacetime supersymmetry to be broken. The resulting proliferation of soft-supersymmetry breaking parameters has no analogue in the EPM. \section{Phenomenology of Mirror Neutrinos} Under the exact parity symmetry $P'$, an ordinary neutrino field $\nu_{\alpha}$ (where $\alpha = e, \mu, \tau$) transforms into its mirror partner field $\nu'_{\alpha}$ as per \begin{equation} \nu_{\alpha L} \to \gamma_0 \nu'_{\alpha R}. \end{equation} From basic quantum mechanics, we know that the exact $P'$ symmetry forces the parity eigenstates to also be mass eigenstates. In the absence of interfamily mixing, this means that the two mass eigenstates $\nu_{\alpha \pm}$ per family take the form \begin{equation} | \nu_{\alpha \pm} \rangle = \frac{1}{\sqrt{2}} \left( |\nu_{\alpha}\rangle \pm |\nu'_{\alpha}\rangle \right). \end{equation} The positive and negative parity states, $\nu_{\alpha +}$ and $\nu_{\alpha -}$ respectively, in general have arbitrary masses. The oscillation parameter $\Delta m^2_{\alpha+ \alpha-} \equiv |m^2_{\nu_{\alpha +}} - m^2_{\nu_{\alpha -}}|$ is therefore free. The mixing angle, however, is forced by $P'$ symmetry to have the maximal value of $\pi/4$.\cite{flv2,f,fv} Interestingly, SuperKamiokande and other experiments have observed a disappearence of atmospheric muon-neutrinos in a manner which favours maximal mixing with another flavour $\nu_x$.\cite{sktalk} Current results preclude $x = e$, but allow both $x = \tau$ and $x = s$\cite{all} (where $s$ stands for ``sterile''). It is natural in the EPM to identify $\nu_x$ with the effectively sterile mirror muon-neutrino $\nu'_{\mu}$.\cite{f,fv} The maximal mixing angle for the $\nu_{\mu} - \nu'_{\mu}$ subsystem is a simple and characteristic prediction of the EPM that is strongly supported by experiment.\footnote{For attempts to explain the large mixing angle in the case of $\nu_x$ identified as $\nu_{\tau}$ see, for instance, Ref.\cite{a}.} The $\Delta m^2_{\mu+ \mu-}$ oscillation parameter is adjusted to agree with the measurements. This requires it to be in the $10^{-3} - 10^{-2}$ eV$^2$ range.\cite{sktalk,fvy1} The experimental discrimination between $\nu_x = \nu_s$ and $\nu_x = \nu_{\tau}$ is a vital further test of this proposal. Hope for progress in this area in the immediate future lies with SuperKamiokande atmospheric neutrino data and the K2K long baseline experiment.\cite{s} The basic requirement is a neutral current measurement, since $\nu_\tau$ is sensitive to this interaction while $\nu_s$ and $\nu'_\mu$ are not. SuperKamiokande has quoted a measured value for the atmospheric neutrino induced $\pi^0/e$ ratio (see Ref.\cite{sktalk} for the present status) that cannot discriminate between the two possibilities because of a large theoretical uncertainty in the $\pi^0$ production cross-section. The measurement of this cross-section by the K2K long baseline experiment is thus of great importance. This may allow a discrimination based on a zenith-angle averaged atmospheric neutrino induced $\pi^0/e$ ratio by SuperKamiokande within about a year from the time of writing.\cite{l} It should be noted, however, that the $\nu_\mu \to \nu'_\mu\, (\nu_s)$ case predicts that the actual $\pi^0/e$ ratio will be about $0.8$ times the no-oscillation or $\nu_{\mu} \to \nu_\tau$ expectation.\cite{fpriv} If the SuperKamiokande central value were to be around $0.9$, then the remaining systematic error would still be too large to discriminate between the possibilities. A cleaner discrimination, which however requires significantly greater statistics, lies in the future through the $\pi^0$ up-down asymmetry.\cite{dg} The K2K experiment could in principle discriminate between the two possibilities on its own by comparing the neutral- to charged-current rates at the near and far detectors. However, inadequate statistics at the far detector (SuperKamiokande) may preclude a useful result. However, provided $\Delta m^2_{\mu+\mu-}$ is sufficiently large, it should at the very least confirm $\nu_\mu$ disappearence. Looking slightly further into the future, the long baseline experiment MINOS and the proposed CERN--Gran-Sasso long baseline experiments should provide important information.\cite{lbl} The solar neutrino anomaly provides further motivation for the maximal mixing feature of the EPM.\cite{flv2,fv} Consider the maximally mixed $\nu_e - \nu'_e$ subsytem in the zero interfamily mixing limit. For the $10^{-3} \stackrel{<}{\sim} \Delta m^2_{e+e-}/{\rm eV}^2 \stackrel{<}{\sim} 10^{-10}$ range, the maximal $\nu_e \to \nu'_e$ oscillations are consistent with disappearence experiments and lead to an energy-independent day-time solar neutrino flux reduction by $50\%$. This is consistent with four out of the five solar event rate meansurements relative to the latest standard solar model calculations.\cite{solar} (The Chlorine experiment sees a greater than $50\%$ deficit.) Since this talk was given, Guth et al.\cite{g}\ have emphasised that the night-time oscillation-affected solar neutrino rate differs from the day-time rate, even if the vacuum mixing angle is maximal. This leads to some energy-dependence in the night-time flux suppression, and provides an interesting further test. Preliminary calculations show that the day-night asymmetry for the $\nu_e \to \nu'_e$ case (or the $\nu_e \to \nu_s$ case with maximal mixing) is potentially observable for the range $6 \times 10^{-8} \stackrel{<}{\sim} \Delta m^2_{e+e-}/{\rm eV}^2 \stackrel{<}{\sim} 2 \times 10^{-5}$, with the range $2 \times 10^{-7} \stackrel{<}{\sim} \Delta m^2_{e+e-}/{\rm eV}^2 \stackrel{<}{\sim} 8 \times 10^{-6}$ already disfavoured by the data.\cite{cfv} The future KAMLAND experiment will probe the $10^{-3} \stackrel{<}{\sim} \Delta m^2_{e+e-}/{\rm eV}^2 \stackrel{<}{\sim} {\rm few}\ \times 10^{-5}$ regime by looking for $\overline \nu_e$ disappearence.\cite{kamland}\footnote{The $\nu_e \to \nu'_e$ mode also has potentially observable consequence for atmospheric $\nu_e$'s for this parameter range.\cite{bfv}} Another extremely important future test is the neutral to charged current event rate ratio that will be measured by SNO.\cite{sno} The mirror electron-neutrino $\nu'_e$ is effectively a sterile flavour, so SNO should measure the ``standard'' value for this ratio. If $\Delta m^2_{e+e-}/{\rm eV}^2$ is in the $10^{-10} - 10^{-11}$ range, then ``just-so'' oscillations result.\cite{justso} One amusing possibility\cite{cfv} is the following: as $\Delta m^2_{e+e-}$ is reduced from the range considered in the previous paragraph into the just-so regime, the energy at which the averaged oscillations give way to coherent just-so behaviour decreases. For some value, this transition will happen within the energy range probed by SuperKamiokande. This could possibly be the origin of the mysterious high-energy spectral feature reported by SuperKamiokande!\cite{solar} (This type of idea was first examined in the context of $\nu_e$ oscillations into an active flavour in Ref.\cite{bfl}.) So, putting the above in a nutshell, we have KAMLAND probing the high range for $\Delta m^2_{e+e-}$, the day-night asymmetry being used in the intermediate range, and just-so signatures such as seasonal variation and Boron neutrino energy spectrum distortion probing the low $\Delta m^2_{e+e-}$ regime. The range between about $6 \times 10^{-8}$ eV$^2$ and the beginning of the just-so region appears to have no characteristic signature other than the $50\%$ energy independent flux suppression. Furthermore, the crucial neutral current measurement at SNO will test the general idea that solar neutrinos are disappearing into sterile states of some sort for the whole $\Delta m^2_{e+e-}$ range of interest. The above analysis saw interfamily mixing set to zero. Certainly, small interfamily neutrino mixing is well motivated by the small mixing observed for the quark sector. However, it is unlikely that this mixing exactly vanishes. I will now comment on three possible consequences of interfamily mixing. First, the LSND anomaly\cite{lsnd} can be trivially accomodated within the EPM by switching on $\nu_e - \nu_{\mu}$ mixing with the appropriate parameter choices.\cite{fv} The LSND parameter regime does not significantly modify the solar and atmospheric neutrino scenario discussed above. Second, the solar neutrino flux depletion can be due to an amalgam of vacuum $\nu_e \to \nu'_e$ oscillations and MSW interfamily transitions.\cite{vw} This leads to characteristic energy-dependent flux depletions depending on the precise oscillation parameter range chosen. Further, the neutral to charged current induced event rate ratio to be measured by SNO can take on values intermediate between the extreme cases of $\nu_e \to \nu_{\rm active}$ only and $\nu_e \to \nu_s$ only. Third, it turns out that small interfamily mixing is well motivated from cosmology, a topic I very briefly review in the next section. \section{Cosmology} The tale of how neutrino oscillations affect early universe cosmology is long and complicated. I will pass over it lightly here, just for the sake of completeness, without much in the way of explanations. Please consult, for example, Refs.\cite{cosmo1,cosmo2,cosmo3,epm} for further details. Cosmology and ordinary-mirror (and ordinary-sterile) neutrino oscillations present challenges to each other. On the one hand, it had long been thought that sterile neutrinos ought to mix but weakly with ordinary neutrinos lest the reasonably successful Big Bang Nucleosynthesis (BBN) predictions be spoiled. In particular, it was thought that a $\nu_{\mu} \to \nu_s$ solution to the atmospheric neutrino problem would have necessarily implied the thermal equilibration of the sterile flavour prior to the BBN epoch, and thus would have increased the expansion rate of the universe. Recall that the expansion rate of the universe during BBN is driven by the relativistic degrees of freedom in the plasma, with ``neutrino flavour number $N_{\nu}$'' being a convenient measure. In the minimal Standard Model $N_{\nu} = 3$, while one thermally equilibrated sterile flavour in addition to the ordinary neutrinos produces $N_\nu = 4$. There is some confusion in interpreting primordial element abundance data at present, but it is arguable that $N_\nu < 4$ is preferred.\cite{Nnu} So, it had been thought that a large region of active-sterile oscillation parameter space was at least disfavoured by BBN. This problem was seen to be much more acute for the EPM than for models with a single extra sterile state, because of the {\it three} mirror neutrino flavours as well as the mirror photons, electrons and positrons. Prior wisdom would have concluded that the EPM ruined BBN and was therefore unlikely to be true. Thus cosmology challenged sterile and mirror neutrino models. On the other hand, the discovery of relic neutrino asymmetry amplification, driven by the ordinary-mirror or ordinary-sterile neutrino transitions themselves, showed that the previous pessimism was misplaced: a very natural mechanism for reconciling BBN with sterile or mirror neutrinos, born out of the apparently problematic neutrino scenario itself, actually existed all along but had been missed.\cite{cosmo2} The basic point is that large relic neutrino asymmetries (neutrino chemical potentials) will, in a certain large region of oscillation parameter space, be inevitably created via a positive feedback process from the tiny $CP$ asymmetry (baryon asymmetry for instance) of the high-temperature background plasma. The large matter (Wolfenstein) effective potentials so induced then damp further ordinary-mirror or ordinary-sterile transitions and lead to quite acceptable BBN predictions (for the appropriate region of parameter space). The full story of BBN in the presence of ordinary-mirror/sterile transitions is complicated because many different oscillation modes are in general involved. We have discussed above how the excitation of mirror or sterile neutrinos prior to BBN increases the expansion rate as quantified through $N_\nu$. But there is another important effect: a fairly large electron neutrino asymmetry will be created before and during BBN given appropriate oscillation parameters. This asymmetry will directly affect BBN reaction rates and will alter the primordial Helium abundance so as to mimic {\it either} a negative {\it or} a positive contribution to the effective neutrino number during BBN. A detailed numerical calculation is often necessary to determine the final BBN outcome. Such calculations have been performed for a couple of models featuring a single sterile flavour.\cite{cosmo3} They have demonstrated that strong ordinary-sterile neutrino mixing can be reconciled with BBN for realistic sterile neutrino models via the interesting physics just discussed. The first full analysis of neutrino asymmetry evolution and BBN in the Exact Parity Model was completed after this Symposium.\cite{epm} It demonstrated that the EPM scenario outlined above is consistent with primordial element abundance measurements for a large region of oscillation parameter space. It turns out that this parameter space region requires some small interfamily mixing. The challenge for observational cosmology, then, is to pin down cosmological parameters precisely enough to test early universe neutrino physics in some detail. Continuing primordial element abundance measurements will help, but much dramatic new information is likely from the cosmic microwave background anistropy measurements promised by the future MAP and PLANCK satellite missions.\cite{kk} \section{Conclusion} The Exact Parity Model predicts that if ordinary neutrinos mix with their mirror partners, then they mix maximally. This has been proposed as a very natural and simple explanation of the very large mixing angle deduced from atmospheric $\nu_{\mu}$ disappearence measurements. In addition, maximal oscillations of the $\nu_e$ into its mirror partner are well motivated by most of the solar neutrino data. With small interfamily mixing switched on, the LSND anomaly can be explained by ordinary $\nu_e \to \nu_\mu$ oscillations. The EPM offers a theoretically elegant solution to all of the neutrino puzzles within a model that had as its original motivation the retention of the full Lorentz Group as an exact symmetry of Nature. The model also has some very interesting consequences for early universe cosmology, particularly the process of Big Bang Nucleosynthesis. In addition to the important results that continue to be produced by SuperKamiokande, we await with interest upcoming experiments -- such as K2K, SNO, KAMLAND and others -- that will provide further crucial tests of the Exact Parity Model. \section{Acknowledgements} I would very much like to thank Professor Milla Baldo Ceolin for organising this stimulating symposium in such an inspirational setting. I would also like to thank the participants of the symposium for interesting discussions and for their informative presentations. I warmly acknowledge my very wise long time collaborator Robert Foot, and my present students Nicole Bell, Roland Crocker and Yvonne Wong for many fruitful scientific discussions.
\section{Introduction} The role of gas flows in spiral arms upon the galactic magnetic field evolution is a lively debated issue. Theories of the field amplification by small scale turbulent motions (e.g. axisymmetric dynamo, Wielebinski \&~Krause 1993), which well reproduce the polarization properties of galaxies with weak density waves (Urbanik et~al. 1997), do not need any spiral arm flows. In other theories (e.g. Chiba 1993), the density wave perturbations are the main agent amplifying the galactic magnetic field. The importance of the latter process compared to the dynamo action may be a function of the density wave strength. Though successful attempts to model the magnetic field evolution driven by the dynamo and spiral arms or bars have been made (Mestel \&~Subramanian 1991, Subramanian \&~Mestel 1993, Moss 1997, Moss et~al. 1998), this issue has still very poor observational grounds. Existing observations of large, nearby spirals (Beck et~al. 1996) do not allow to state whether in case of strong density waves the magnetic field evolution becomes dominated by processes in spiral arms. Strong density wave signatures are present in M51 and in the inner disk of M81. In M51 the polarization B-vectors follow local structural details of dust lanes (Neininger \&~Horellou 1996), as expected for a magnetic field dominated by density-wave compression. However, the spiral arms are too tightly wound to separate this field component from a possible dynamo-generated one, which is usually distributed more uniformly in the disk. No clear magnetic field component related to density waves has been identified in M81, however its inner disk shows a subtle network of local compression regions filling the whole interarm space (Visser 1980). NGC~6946, NGC~4254 and the outer parts of M81 do not show strong density wave signatures. Their magnetic fields form either broad ''magnetic arms`` in the middle of the interarm space (NGC~6946, Beck \&~Hoernes 1996) or smoothly fill the interarm space (M81, Krause et~al. 1989). In NGC~4254 a coherent spiral pattern of polarization B-vectors exists even in regions of completely chaotic optical structures (Soida et~al. 1996). The nearby well-studied spirals also have enhanced star formation in spiral arms, destroying the regular fields. We note that the detection of smoothly distributed dynamo-type fields needs a very good sensitivity to extended polarized emission, which is not always ensured by high-resolution studies. In this paper we present observations with good sensitivity to smooth, extended structures to check whether a galaxy with very strong signs of density waves may have a global magnetic field dominated by the component caused by density wave action. We obtained 10.55~GHz total power and polarization maps of the spiral NGC~3627 interacting within the Leo Triplet (Haynes et~al. 1979). The galaxy has a bar and two spiral arms with a broad interarm space discernible even with a modest resolution (see Fig.~1). The western arm contains a long dust lane tracing large-scale gas (and possibly frozen-in field) compression, accompanied by little star formation (cf. H$\alpha$ map by Smith et~al. 1994). The middle part of the arm is unusually straight, bending sharply in the outer disk. The eastern arm has a heavy dust lane in its southern part, breaking into a subtle network of filaments in its northern half. The southern dust lane segment is accompanied by a chain of bright star-forming regions. Reuter et~al. (1996) found perturbations of the galaxy's CO velocity field possibly due to streaming motions related to spiral arms. NGC~3627 has been also observed in the far infrared by Sievers et~al. (1994). A total power map at 1.49~GHz using the VLA D-array was made by Condon (1987). \section{Observations and data reduction} \begin{figure*} \resizebox{12cm}{!}{\includegraphics{Fig01.ps}} \parbox[b]{55mm}{ \caption{ The total power contour map of NGC~3627 at 10.55~GHz with B-vectors of polarized intensity superimposed onto the H$\alpha$ image from Smith et~al. (1994). The resolution is 1\farcm 13. The contour levels are 1.5, 4, 6.5~mJy/b.a., then 9, 14, 19$\ldots$ etc. mJy/b.a.. The first contour corresponds to about 2.5$\sigma$ r.m.s. noise. The dashed contour shows the zero level } \label{fig1} } \end{figure*} The total power and polarization observations at 10.55 GHz were performed in May 1993, as well as in April and May 1994 using the four-horn system in the secondary focus of the Effelsberg 100-m MPIfR telescope (Schmidt et~al. 1993). With 300~MHz bandwidth and $\sim 40$~K system noise temperature, the r.m.s. noise for 1~sec integration and combination of all horns is $\sim 2$~mJy/beam area in total power and $\sim 1$~mJy/beam area in polarized intensity. Each horn was equipped with two total power receivers and an IF polarimeter resulting in 4 data channels containing the Stokes parameters I, Q and U. The telescope pointing was corrected by making cross-scans of Virgo~A at time intervals of about 2~hours. As flux calibrator the highly polarized source 3C286 was observed. A total power flux density of 4450~mJy at 10.55~GHz has been adopted using the formulae by Baars et~al. (1977). The same calibration factors were used for total power and polarized intensity, yielding a mean degree of polarization of 12.2\%, in reasonable agreement with other published values (Tabara \&~Inoue 1980). In total 29 coverages of NGC~3627 in the azimuth-elevation frame were obtained. The data reduction process was performed using the NOD2 data reduction package (Haslam 1974). By combining the information from appropriate horns, using the ''software beam-switching`` technique (Morsi \&~Reich 1986) followed by a restoration of total intensities (Emerson et~al. 1979), we obtained for each coverage the I, Q and U maps of the galaxy. All coverages were then combined using the spatial frequency weighting method (Emerson \&~Gr\"ave 1988), yielding the final maps of total power, polarized intensity, polarization degree and polarization position angles. A digital filtering process, which removes spatial frequencies corresponding to noisy structures smaller than the telescope beam, was applied to the final maps. A special CLEAN procedure to remove instrumental polarization was applied to the polarization data. The original beam of our observations was 1\farcm 13. With the distance modulus of $-30\fm37$ given by Ryan \&~Visvanathan (1989), corresponding to a distance of 11.9~Mpc, our beamwidth is equivalent to 3.9~kpc in the sky plane. In the galaxy's disk plane this corresponds to 3.9 and 8 kpc along major and minor axes, respectively. \section{Results} \subsection{Total power emission} The total power map at the original resolution with B-vectors of polarized intensity is shown in Fig.~1. The map has an r.m.s. noise of 0.6~mJy/b.a. Bright total power peaks are found at the bar ends, where both the CO(1-0) and CO(2-1) maps by Reuter et~al. (1996) as well as the H$\alpha$ map by Smith et~al. (1994) show large accumulations of molecular gas and young star formation products. There is no indication of a bright central source. The outer disk shows a remarkable asymmetry. The total power emission is considerably more extended and decreases more smoothly towards the south than to the north. In this respect it resembles the optical, CO and H$\alpha$ morphology: the western arm running southwards extends to a considerably larger distance from the centre than does the eastern one. A slight extension towards the east at RA$_{1950}$ of about $11^{h}17^{m}45^{s}$ and Dec$_{1950}$ of about $13\degr 17\arcmin$ is also seen in the map by Urbanik et~al. (1985) and must be real. It has no optical counterpart but corresponds roughly to the region where Haynes et~al. (1979) found a counter-rotating HI plume, probably caused by tidal interactions within the Leo Triplet. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{Fig02.ps}} \caption{ The integrated radio spectrum of NGC~3627 } \label{fig2} \end{figure} The integration of the total power map in elliptical rings using an inclination of $67\fdg 5$ and a position angle of $173\degr$ (both taken from the Lyon-Meudon Extragalactic Database) yields an integrated flux density at 10.55~GHz of 103$\pm$10~mJy within the radius of 20~kpc, very close to the total flux obtained by Niklas et~al. (1995). This value has been combined with available data at lower frequencies collected in Table 1. All values have been converted to the flux density scale by Baars et~al. (1977). A weighted power law fit to the data yields a mean spectral index of 0.64$\pm$0.04 ($S_{\nu}\sim\nu^{-\alpha}$). As the deviations from a single power law are comparable to the errors in the observed flux densities (Fig.~2), a spectral index of 0.64 has been adopted for the whole frequency range between 80~MHz and 10.7~GHz. \begin{figure*} \resizebox{12cm}{!}{\includegraphics{Fig03.ps}} \parbox[b]{55mm}{ \caption{ The contour map of the polarized intensity of NGC~3627 at 10.55~GHz with B-vectors of the polarization degree superimposed onto an optical image (Arp 1966). The resolution is 1\farcm 13. The contour levels are 0.4~mJy/b.a., then from 0.6~mJy/b.a. with an increment of 0.4~mJy/b.a.. The first thin contour corresponds to about 2.2$\sigma$, and the first thick one to about 3.3$\sigma$ r.m.s. noise. The dashed contour shows the zero level } \label{fig3} } \end{figure*} \begin{table}[th] Table 1. The integrated radio spectrum of NGC~3627 \begin{center} \begin{tabular}{rrrl} \hline Frequ-&Flux &error & References \\ ency &density& & \\\ [GHz] & [mJy] & [mJy] & \\ \hline 0.080 & 2448 &$\pm$ 410 & Slee (1972) \\ 0.160 & 1680 &$\pm$ 315 & Huchtmeier (1975) \\ 0.430 & 946 &$\pm$ 106 & Lang \& Terzian (1969) \\ 0.430 & 938 &$\pm$ 20 & Israel \& van der Hulst (1983) \\ 0.611 & 915 &$\pm$ 105 & Lang \& Terzian (1969) \\ 0.835 & 580 &$\pm$ 20 & Israel \& van der Hulst (1983) \\ 1.415 & 551 &$\pm$ 100 & de la Beaujardiere et~al. (1968) \\ 1.415 & 512 &$\pm$ 51 & Hummel (1980) \\ 1.420 & 567 &$\pm$ 72 & Whiteoak (1970) \\ 2.695 & 359 &$\pm$ 28 & de Jong (1967) \\ 2.695 & 206 &$\pm$ 41 & Kazes et~al. (1970) \\ 5.000 & 187 &$\pm$ 19 & Whiteoak (1970) \\ 5.000 & 175 &$\pm$ 20 & Sramek (1975) \\ 10.550 &103 &$\pm$ 10 & This paper \\ \hline \end{tabular} \end{center} \end{table} \subsection{Polarized intensity} Our pola\-rized inten\-sity map has an r.m.s. noise of 0.18 mJy/b.a.. It shows two asymmetric lobes with B-vectors locally parallel to the principal arms (Fig.~3). The strongest peak of the polarized brightness, with the polarization degree reaching locally 25\%, is located west of the galaxy's centre, at the position of the unusually straight dust lane segment (see also Fig.~5). No bright star-forming regions are present there. The second, weaker peak does not coincide with a prominent dust lane but is located in the interarm region between the northern segment of the eastern arm and the bar where only small, barely visible dust filaments are present. No polarization was detected in the vicinity of a particularly heavy dust lane segment in the southern part of the eastern arm at RA$_{1950}$ of about $11^{h}17^{m}42^{s}$ and Dec$_{1950}$ of $13\degr 15\arcmin 30\arcsec$, accompanied by a chain of star-forming regions. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{Fig04.ps}} \caption{ The contour map of the polarized intensity of NGC 3627 with B-vectors proportional to the same quantity, convolved to a beam of 1\farcm 3. The first contour is 0.3~mJy/b.a., then contours start from 0.5~mJy/b.a. and increase by 0.4 mJy/b.a. The first thick contour corresponds to about 3$\sigma$ r.m.s. noise. The dashed contour denotes the zero level. The thick line terminated by dots marks the position of the bar as defined by the CO(1-0) maxima at its ends } \label{fig4} \end{figure} With a polarization degree less than 3\% the bar ends are generally weakly polarized. However, clear polarization patches in the vicinity of RA$_{1950}$ of $11^{h}17^{m}41\fs5$ Dec$_{1950}$ of $13\degr 14\arcmin 30\arcsec$ and RA$_{1950}$ of $11^{h} 17^{m} 35\fs3$ Dec$_{1950}$ of $13\degr 17\arcmin 32\arcsec$, surrounding the bar ends and being marginally significant in Fig.~3, exceed the $3\sigma$ noise level after convolving the data to a beamwidth of 1\farcm 3 (Fig.~4). The degree of polarization in these regions is about 5--6\%. The orientations of the polarization B-vectors cor\-rected to face-on position are shown in Fig.~5. West of the centre they run parallel to a straight segment of the dust lane, following its bend in the southern disk. In the polarized peak in the NE disk the B-vectors in the interarm region follow the direction of the dust lane which itself is only weakly polarized. In the above mentioned weak polarization patches near the bar ends, best visible in Fig.~4, the B-vectors tend to turn smoothly around the terminal points of the bar. No ordered optical or H$\alpha$ structures are present there. Close to the northern bar end the vector orientations smoothly join these in the western arm (see also Fig.~5). Near the southern bar end the B-vectors deviate strongly to the east with a large pitch angle. Across the unpolarized region in the southern part of the eastern arm their orientations jump by about $90\degr$. The question of possible geometrical depolarization at this position is discussed in detail in Sect.~4. \begin{figure*} \resizebox{12cm}{!}{\includegraphics{Fig05.ps}} \parbox[b]{55mm}{ \caption{ Overlay of B-vector directions computed from the polarization map with the original resolution of 1\farcm 13 onto an optical image of NGC~3627, both rectified to face-on position. The galaxy's major axis runs vertically, directions to north and west are shown by arrows } \label{fig5} } \end{figure*} The integration of the polarized intensity map shown in Fig.~3 in the same rings as described in Sect~3.1 yields an integrated polarized flux density of $6.0\pm 1.8$~mJy. This implies a mean polarization degree of $5.8\pm 1.8$\%. An application of the formula of Segalovitz et~al. (1976) yields a mean ratio of regular to total field strengths $B_u/B_t = 0.22\pm 0.04$. \section{Discussion} \subsection{Total magnetic field strength and distribution of total power brightness} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{Fig06.ps}} \caption{ Cross-section along the bar of NGC~3627 in the total power brightness (circles and error bars) compared to that in CO(1-0) (Reuter at~al. 1996, short-dashed line) and H$\alpha$ (Smith et~al. 1994, long dashed line), convolved to a common beam of 1\farcm 13 and each normalized to its maximum value. The CO(1-0) profile with the central peak subtracted from the original map is shown, too (solid line) } \label{fig6} \end{figure} The integrated radio spectrum of NGC~3627 does not show obvious deviations from a single power-law with a slope $\alpha=0.64$ (see Fig.~2). Using the integrated flux density, $\alpha=0.64$ and assuming the minimum-energy or energy equipartition condition we derive a mean total magnetic field strength of $13\pm 4\ \mu$G. The regular field component (assuming that it is entirely parallel to the disk) derived from the polarized intensity equals $3.5\pm 1.3\ \mu$G. This value refers to a magnetic field which is regular over scales larger than our beam of 4~kpc. In computing the above values we assumed a lower limit of the cosmic-ray spectrum of 300~MeV (cf. Beck 1991), a ratio of proton-to-electron density ratio of 100 (Pacholczyk 1970), and a nonthermal disk scaleheight of 1~kpc. The error in the total magnetic field strength includes an uncertainty by a factor of 2 of the proton-to-electron ratio, the disk thickness and the lower energy cutoff, as well as an unknown thermal fraction between 0\% and 40\%. The mean total magnetic field of NGC~3627 is stronger than average for spiral galaxies (Beck et~al. 1996), in spite of the low neutral gas content (Young et~al. 1983, Zhang et~al. 1993, see also Urbanik 1987). The bright total power sources at the ends of the bar lie at the positions of huge star-forming molecular complexes, also coincident with HI peaks (Zhang et~al. 1993). Their spectral index between 1.49~GHz and 10.55~GHz derived using Condon's (1987) map is 0.73--0.75, thus nonthermal emission is dominating. Fig.~6 shows the cross-sections along the bar of the total power intensity at 10.55~GHz as well as of the H$\alpha$ and CO(1-0) line emission (Smith et~al. 1994, Reuter et~al. 1996), both convolved to our resolution. All profiles were normalized to their maximum values. The CO profile has a peak at the galaxy's centre due to the emission from the central molecular complex which in the original maps (Reuter et~al. 1996) has the same peak brightness and extent as the CO complexes at the bar ends. However, the H$\alpha$ emission lacks the central peak being very weak in the nuclear region. An intense star formation in the central molecular complex with its manifestation in the H$\alpha$ line obscured by the dust (abundant in nuclear regions of spiral galaxies) is unlikely as the 10.55~GHz profile (Fig.~6) has a central depression, too. The total power minimum is even deeper than in CO(1-0) after subtracting completely the nuclear region from the original CO map of Reuter et~al. (1996). Thus the central molecular complex apparently forms stars at a much lower rate than aggregates of the cold gas at the bar ends. We also note that the HI map of Zhang et~al. (1993) shows a central depression, too. The H$\alpha$ emission from the southern bar end is considerably weaker than from the northern one. No such asymmetry exists in radio continuum. The CO(1-0) brightness (thus also the content of an opaque cold gas) is however somewhat higher at the southern than at the northern end of the bar. The mentioned asymmetry of the H$\alpha$ emission may thus be caused by a higher absorption in the southern bar end. The determination of the radial scale length $r_0$ of the nonthermal disk by fitting a beam-smoothed exponential model encounters severe problems because of the high inclination and emission excess at the bar ends. Nevertheless reasonable values are in the range 1\farcm 2--1\farcm 8 corresponding to 4.2--6.2~kpc at the distance of 11.9~Mpc. \subsection{The magnetic field structure } \begin{figure} \resizebox{\hsize}{!}{\includegraphics{Fig07.ps}} \caption{ A comparison of changes of the polarized brightness with the azimuth distance from the corresponding bar ends for both polarized lobes. The azimuth runs anticlockwise from either the northern bar end (solid line, eastern lobe) or from the southern end (dotted line, western lobe). The polarized intensity was integrated along a ring 24\arcsec wide with a face-on radius of 2\arcmin, having the same inclination and position angle as NGC~3627, divided into sectors with an azimuthal width of $10\degr$. Horizontal bars show the range of azimuthal angles corresponding to the beam size at various azimuthal distances from the bar end (labelled in degrees) } \label{fig7} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{Fig08.ps}} \caption{ The contours of polarized intensity for NGC~3627 at 10.55~GHz with the original resolution of 1\farcm 13 overlaid onto a)~the greyscale map of CO(1-0) of Reuter et~al. (1996), and b)~the H$\alpha$ image of Smith et~al. (1994). The polarized intensity contours are the same as in Fig.~3 } \label{fig8} \end{figure} The polarized brightness is strongly peaked at the middle of a straight portion of the dust lane in the western arm. To check whether the emission is resolved we tried to subtract the beam-smoothed point source at the position of the observed brightness maximum. We found that the polarized peak can be decomposed into an unresolved source with a polarized flux density of about 1.8~mJy and an extension along the southern part of the dust lane with a maximum polarized brightness of about 0.8~mJy/b.a.. The eastern lobe is however rather poorly resolved. \begin{figure*} \resizebox{12cm}{!}{\includegraphics{Fig09.ps}} \parbox[b]{55mm}{ \caption{ The model distributions of polarized intensity in NGC~3627 assuming: a)~a regular magnetic field parallel to the arms and concentrated in dust lanes, as well as one parallel to the bar, b) as above, without the contribution from the bar and no regular field in the strongly star-forming segment of the eastern arm, c)~a disk-parallel axisymmetric field with a constant pitch angle of $30\degr$, d)~the above axisymmetric field with no polarization at azimuthal angles corresponding to the star-forming arm segment. For spiral arm models (a~and b) the thick dotted lines mark the parts of spiral structure assumed to house regular fields, for axisymmetric models (c~and d) the spiral pattern is also shown for easier comparison with the observations. The models were made for $r_0=1\farcm 2$, $p_{1}=0$ and $p_{2}=0.5$. } \label{fig9} } \end{figure*} The eastern and western polarized lobes differ not only in their peak brightness but also in their positions and azimuthal extent relative to optical arms. Moving in the galactic disk along the azimuth anticlockwise from the corresponding bar ends, we observe initially a very similar increase of polarized brightness (Fig.~7). However, while the polarized intensity in the western arm continues to rise reaching a maximum at an azimuthal distance of about $100\degr$ from the southern bar end, the polarized brightness in the eastern arm drops at an azimuthal distance of $75\degr$ from the northern bar end, showing even a local minimum at $100\degr$. As the inner parts of both arms have a similar shape the unpolarized region in the eastern arm does not result from effects related to the spiral arm geometry. The statistical significance of the differences between the profiles was estimated by averaging them in non-overlapping azimuthal angle intervals corresponding to the beam size at the appropriate azimuthal distance from the bar end. This yielded for each profile 7 statistically independent points. Assuming that they represent independent random variables having the r.m.s. dispersion equal to the polarization map noise we found that the probability that the differences between profiles result purely from random fluctuations is smaller than $2\times 10^{-6}$. This result was checked to be independent of the starting point of averaging intervals. The dust lanes in the western arm and its segment in the inner part of the eastern one coincide with ridges of CO(1-0) emission (Reuter et~al. 1996, see also Fig.~8a) tracing very dense, narrow and elongated molecular gas complexes forming in density-wave compression regions. They are also visible in the HI map of Zhang et~al. (1993). While the western polarized lobe peaks on the CO ridge and extends along it, the eastern one falls on a hole in the CO emission. On the other hand, the CO (and HI) ridge in the southern part of the eastern arm, being even stronger than the western one, coincides with a completely unpolarized region. The narrow, elongated CO features are thus not always associated with highly polarized regions, as one would expect from pure compression of magnetic field by density waves. We note however, that the western CO ridge is accompanied by only isolated, small HII regions (Smith et~al. 1994, Fig.~8b) while the unpolarized CO ridge in the eastern arm hosts a chain of large complexes of H$\alpha$-emitting gas. Faraday effects at 10.55~GHz are negligible, thus the depolarization is primarily of geometrical nature. Tangling of the magnetic field by star-forming process inside the H$\alpha$-bright knot is insufficient: most of the star formation occurs outside of the dust lane, thus it cannot destroy a possible density wave-related field and occupies a too small volume to randomize a smoothly-distributed dynamo field in a whole disk quadrant. However, at the inclination of $67\fdg 5$ the polarization degree may be significantly lowered by vertical magnetic field fluctuations developing above the strongly star-forming chain in the eastern arm. They may be caused by vertical chimneys or superbubbles powered by multiple supernova explosions (see e.g. Mineshige et al. 1993, Tomisaka 1998). Some role of Parker instabilities (Parker 1966) cannot be excluded, too. The magnetic field structures stretching perpendicularly to the disk with a significant vertical field component, projected to the sky plane and seen by a large beam together with the disk-parallel field (either concentrated in the dust lane or filling the whole disk), may provide an efficient depolarizing agent. \subsection{Magnetic field models} To judge whether our polarization map is dominated by the density-wave magnetic field component or by the axisymmetric, dynamo-type we need the beam-smoothed models of polarized emission from magnetic fields of an assumed structure. Four kinds of models were computed using techniques described by Urbanik et~al. (1997): \begin{itemize} \item[a)] A model assuming a regular magnetic field concentrated in prominent dust lanes and running along them. In addition the magnetic field running along the bar could be switched on and off. \item[b)] The above model without polarized emission from the strongly star-forming segment of the eastern arm. \item[c)] A model assuming an axisymmetric, spiral, plane-pa\-ra\-llel field with a constant intrinsic pitch angle of $-30\degr$ (mean value for NGC~3627). \item[d)] The above model without polarized emission in the eastern region corresponding to the discussed star-for\-ming arm segment, as expected for strong vertical field fluctuations seen in projection together with a disk-para\-llel field. \end{itemize} In all models the adopted radial distribution of the total field strength and cosmic ray electron density was set to yield an exponential total power disk with a radial scale length $r_0$ (Sect.~4.1). The intrinsic degree of polarization was rising linearly from $p_1$ in the centre to $p_2$ in the disk outskirts. $r_0$, $p_1$~and $p_2$ were adjusted to yield the best qualitative agreement of the models with observations. The best results presented in Fig.~9a--d are as follows: \begin{itemize} \item[-] The presence of two polarized lobes with B-vectors running parallel to optical spiral arms is reproduced by both axisymmetric and spiral arm models. They both give the position of the western lobe on the middle of the spiral arm, in agreement with observations. \item[-] Both axisymmetric and spiral arm models need an extra depolarizing agent in the star-forming segment of the eastern arm, otherwise both models give the maximum of polarized intensity where observations show a complete lack of polarization (Fig.~9a and c). \item[-] Only the axisymmetric models (with and without an extra depolarization, Fig.~9c and d) correctly place the NE lobe in the interarm space. The spiral arm models (Fig.~9a and b) invariably give the position of the eastern polarized lobe on the position of the prominent dust lane which disagrees with observations. \item[-] Only the axisymmetric models reproduce the observed regions of a weak polarized signal encircling minima at the bar ends with B-vectors turning smoothly from one arm to the other. \item[-] Even with a suppressed polarization in the SE disk region the axisymmetric model yields similar peak amplitudes of both lobes and thus does not reproduce their observed asymmetry. The spiral arm model does this considerably better. \item[-] Another difficulty of the axisymmetric model is a too large extent of the modelled lobes into the outer disk compared to their rather peaked shape in NGC~3627 (especially of the western one). Varying $r_0$ and/or $p_2$ can make the western lobes more peaked but moves it to the interarm space, worsening the agreement with observations. \end{itemize} The last two difficulties of the axisymmetric model can be somewhat relaxed by adding an unresolved polarized source in the middle of the western arm, where its unusually straight part (see Fig.~5) and a steep HI gradient on this disk side are suggestive for an external gas compression (Haynes et~al. 1989). However, the difficulties of the spiral arm model can only be improved by adding a widespread, significant axisymmetric magnetic field. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{Fig10.ps}} \caption{ The azimuthal variations of the B-vector pitch angles $\psi$ corrected to face-on position, averaged in azimuthal sectors $5\degr$ wide along a ring with a width of 24\arcsec and a face-on radius of 2\arcmin. The data for NGC~3627 are shown by circles and bars, for the spiral arm model by a solid line and for the axisymmetric model with an unpolarized region in the eastern arm segment as a dashed line. A dotted line shows an attempt to remove the jumps of $\psi$ in the spiral model by dropping sectors of lowest signal from the analysis. For NGC~3627 only the data above the $1\sigma$ noise level were used to determine the pitch angle. } \label{fig10} \end{figure} Attempts to reproduce the variations of face-on corrected magnetic pitch angles $\psi$ with the azimuthal angle in the disk are shown in Fig.~10. The observed changes of $\psi$, and especially a jump near the azimuthal angle of $90\degr$, rule out a purely axisymmetric magnetic field. However, addition of the discussed unpolarized region to our axisymmetric model yields the jump at the correct position, though its exact shape in our simple model is still far from reality. The spiral arm models shown in Fig.~9a~and b also have some dip at about 90$^{o}$, however they yield abrupt jumps of $\psi$ at $150\degr$~and $330\degr$. These features do not depend on model parameters, nor on the inclusion or exclusion of low-brightness regions in model maps. They naturally result from the spiral arm shape and cannot be removed without changing the basic model assumptions. In the azimuthal angle range of $270\degr$~to $330\degr$ the spiral arm model deviates also from the data much more than that assuming the axisymmetric field. Despite very strong density waves NGC~3627 still shows clear signatures of axisymmetric, dynamo-type magnetic fields. At present it is hard to say whether it dominates the disk field, showing only locally effects of external compression, or whether it coexists with the density-wave component as an important constituent of the global magnetic field. A detailed discrimination between these possibilities needs observations with a considerably higher resolution complemented by extensive computations of a whole grid of detailed quantitative models of NGC~3627, which is beyond the scope of this paper. \section{Summary and conclusions} The strongly interacting Leo Triplet galaxy NGC~3627 has been observed at 10.55~GHz with the 100-m MPIfR radio telescope. Total power and polarization maps with a resolution of 1\farcm 13, very sensitive to extended, diffuse polarized emission were obtained. Their analysis in the context of optical, CO and H$\alpha$ data yielded the following results: \begin{itemize} \item[-] The total power map shows two bright maxima at the bar ends, coincident with strong CO and H$\alpha$ peaks. There is no evidence of a significant radio emission from the central region, thus a large central molecular complex (Reuter at~al. 1996) has star formation rate much lower than the molecular gas accumulations at the bar ends and its weak H$\alpha$ emission is not entirely due to a strong dust obscuration. However, differences in absorption could explain the asymmetry of H$\alpha$ emission between the bar ends. \item[-] The polarized emission forms two asymmetric lobes: a strong one peaking on the dust lane bent inwards in the western arm and extending along this arm while a weaker one is located in the interarm space in the NE disk. We also detected diffuse, extended, polarized emission encircling the bar ends away from spiral arms. The polarization B-vectors run parallel to the principal arms, twisting around the bar ends in weakly polarized regions. \item[-] The southern part of the eastern arm is completely depolarized. This region shows signs of strong density wave compression however, it contains a lot of ionized gas indicating strong star formation. At the galaxy's inclination, the development of vertical magnetic instabilities seen in projection together with the disk-parallel magnetic field could be a suitable depolarizing agent. \item[-] Attempts to qualitatively explain the distribution of polarized intensity and the B-vector geometry in NGC 3627 in terms of simple magnetic field models suggest the presence of a significant (if not dominant) axisymmetric, dynamo-type field. However, to best explain our polarization maps all models need an extra geometrical depolarization e.g. by vertical fields above the discussed star-forming segment of the eastern arm. To reproduce the polarization asymmetry the dynamo-generated field also requires an extra polarized component (probably due to external compression?) at the position of the straight western arm segment. \end{itemize} The present work demonstrates that even in galaxies with strong density waves observations sensitive to extended diffuse polarized emission cannot be fully explained by the density wave-related magnetic field component but show clear signatures of large-scale axisymmetric dynamo-type fields. On the other hand, in the same object the density-wave component may show up much better or become dominant in high-resolution interferometric observations underestimating the extended polarized emission. We believe that combined interferometric and single-dish data on such objects supported by extensive modelling might help to establish the mutual relationships and relative roles of turbulence and density-wave flows in galactic magnetic field evolution. \begin{acknowledgements} The Authors wish to express their thanks to Dr~Beverly Smith from IPAC for providing us with her H$\alpha$~map in a numerical format. We are grateful to numerous colleagues from the Max-Planck-Institut f\"ur Radioastronomie (MPIfR) in Bonn, in particular to Drs E.M.~Berkhuijsen and P.~Reich for their valuable discussions during this work. M.S. and M.U. are indebted to the Directors of the MPIfR for the invitations to stay at this Institute, where substantial parts of this work were done, and to Dr H-P.~Reuter for his assistance in using his CO maps. They are also grateful to colleagues from the Astronomical Observatory of the Jagiellonian University in Krak\'ow and in particular to Drs. K.~Otmianowska-Mazur and M.~Ostrowski for their comments. We thank to the anonymous referee for the valuable remarks. This work was supported by a grant from the Polish Research Committee (KBN), grants no. 578/P03/95/09. and 962/P03/97/12. Large parts of computations were made using the HP715 workstation at the Astronomical Observatory in Krak\'ow, partly sponsored by the ESO C\&EE grant A-01-116 and on the Convex-SPP machine at the Academic Computer Centre ''Cyfronet`` in Krak\'ow (grant no. KBN/C3840/UJ/011/1996 and KBN/SPP/UJ/011/1996). \end{acknowledgements}
\section{Introduction} Starbursts, especially in dwarf galaxies, play a key role in understanding galaxy formation and evolution. A strong burst of star formation is expected to occur during the initial virialization and baryonic dissipational collapse of a proto-galaxy. In hierarchical cosmological scenarios, the first systems to virialize (at $z \approx 3$) have masses similar to those of dwarf galaxies, which then merge to form larger galaxies (e.g.\ \cite{sz78}). Late epoch starbursts have been invoked to explain the excess of faint blue galaxies in the field (\cite{br92}; \cite{bf96}), and the increasing contribution of blue galaxies with $z$ in clusters (\cite{cs87}). Starburst driven galactic winds (\cite{ham90}) have a particularly important role in dwarf galaxy evolution. Mass loss is easier in the shallow potential of a dwarf galaxy, so winds may play an important role in regulating the gas content of dwarfs (\cite{ds86}). While the initial geometry of the ISM is also important (\cite{dYH94}), and extended dark matter halos complicate the mass-loss predictions, it is unlikely that a dwarf galaxy can retain the hot ($\sim 10^7$ K) metal enriched gas that inflates the Kpc scale \Hline{\mbox{$\alpha$}}\ bubbles commonly seen in starburst dwarfs (e.g.\ \cite{hdlfgw95}). Hence, starburst driven winds, especially from dwarfs, play an important role in enriching the intergalactic (and intra-cluster) medium, while keeping the metal content of the remaining ISM lower than closed-box model predictions. Dwarf galaxies may contribute strongly to QSO absorption line spectra, especially because of their winds, as well as their often-extended gaseous disks (e.g.\ DDO154: \cite{cb89}; NGC~2915: \cite{mcbf96}). While these issues illustrate the conceptual importance of starbursts and dwarf galaxies, our understanding of these objects are far from complete. Two questions in particular, still are not resolved. Firstly, what is the role of starbursts in dwarf galaxy evolution? In some models a starburst is a singular event, which through rapid star formation and consequent galactic wind, transforms a gas rich dwarf irregular into a low surface brightness, gas poor system (e.g.\ \cite{br92}). Hence starbursting dwarfs may be the link between gas rich and gas poor configurations. However observations of the stellar populations in even the lowest mass dwarfs show that they typically have had multi-episodic or an otherwise complex star formation history (e.g.\ \cite{dacosta97}, and references therein). Given the usual disk geometry of the ISM, even in dwarfs, and the geometric arguments of De Young \&\ Heckman (1994\markcite{dYH94}; cf. \cite{mf99}), it is not clear that winds alone can sweep the ISM out of a dwarf galaxy. This has also been discussed from an observational point of view by Welch {\it et al} (1998). Secondly, how do starbursts themselves evolve? The theoretical concept is that starbursts have a very short duration. For example the fading dwarf model of Babul and Ferguson (1996) requires a burst duration $dt \lesssim 10$ Myr. However observations of well known starbursts show them to be complex, consisting of multiple star clusters embedded in a dominant distribution of diffusely distributed high mass stars (\cite{mhlkrg95}). Can such structures be formed on such short timescales? In order to address these questions, we have commenced an observational study of nearby star-bursting dwarf galaxies. Not only may these galaxies be the active link between gas-rich and gas-poor dwarf galaxy configurations but they are the most numerous type of starbursting systems, hence allowing investigation of relatively nearby systems (e.g. within the Virgo Cluster distance: D $\leq$ 16 Mpc). Starbursting dwarf galaxies go by several names. These include blue compact dwarf (BCD), \mbox {H\thinspace{\footnotesize II}}\ or emission-line galaxies, and amorphous galaxies. These names indicate slightly different selection criteria. For example Thuan \&\ Martin (1981\markcite{tm81}) define BCDs by low luminosity, small sizes, blue colors and an emission line spectrum, and thus is both a color and spectroscopic classification. The \mbox {H\thinspace{\footnotesize II}}\ galaxy classification is purely spectroscopic, based usually on prism surveys for emission line sources (e.g.\ \cite{ml78}). The ``amorphous'' classification introduced by Sandage \&\ Brucato (1979\markcite{sb79}) is morphological. It is based on deep photographic imaging and requires an extended high surface brightness region in a host containing no spiral arms. Hence amorphous galaxies resemble E or S0 galaxies, but typically have blue colors. Our sample of twelve starburst dwarfs ($M_B \gtrsim - 18$) was selected primarily from blue galaxies classified as amorphous, or likely amorphous galaxies (by \cite{sb79}; \cite{gh87}; \cite{hvwg94}) because these galaxies include some of the nearest starburst dwarfs known. Two well known dwarf \mbox {H\thinspace{\footnotesize II}}\ galaxies, Haro~14, and II~Zw~40, were also included to round out our observing time. The basic properties of the sample galaxies are listed in Table \ref{tabStats}. Their proximity makes them an ideal class for studying starburst galaxies in detail. \begin{deluxetable}{lccccrcrrrcrrr} \footnotesize \tablewidth{0pt} \tablecaption{Basic Properties of Galaxies\label{tabStats}} \tablehead{ \colhead{Galaxy} & \colhead{$v_{\rm helio}$} & \colhead{$v_0$} & \colhead{$D$} & \colhead{scale} & \colhead{$r_{e,B}$} & \colhead{$r_{25}$} & \colhead{$b/a$} & \colhead{P.A.} & \colhead{$M_{B_0}$} & \colhead{$\mu_{e,B}$} & \colhead{$A_B$} & \colhead{[O/H]$_\odot$} \\ \colhead{} & \colhead{\scriptsize(km s$^{-1}$)} & \colhead{\scriptsize(km s$^{-1}$)} & \colhead{\scriptsize(Mpc)} & \colhead{\scriptsize(pc $\prime\prime^{-1}$)} & \colhead{\scriptsize(pc)} & \colhead{\scriptsize(kpc)} & \colhead{} & \colhead{\scriptsize(deg)} & \colhead{} & \colhead{\scriptsize(mag$/\sq^{\arcsec}$)} & \colhead{} & \colhead{} \\ \colhead{(1)} & \colhead{(2)} & \colhead{(3)} & \colhead{(4)} & \colhead{(5)} & \colhead{(6)} & \colhead{(7)} & \colhead{(8)} & \colhead{(9)} & \colhead{(10)} & \colhead{(11)} & \colhead{(12)} & \colhead{(13)} } \startdata Haro 14 &\phn944 & 1040 & 12.5 & \phn60 & \phn580 & 2.0 & 0.88 & 50 & --17.02 & 19.65 & 0.24 &\nodata & \nl NGC 625 &\phn386 &\phn333 &\phn4.1 &\phn20 &\phn580 & 1.8 & 0.39 & 92 & --16.31 & 21.20 & 0.18 &\nodata & \nl NGC 1510 &\phn968 &\phn830 & 11.0 &\phn53 &\phn370 & 1.9 & 0.85 & 90 & --16.75 & 18.93 & 0.16 & --0.7 & \nl NGC 1705 &\phn629 &\phn433 &\phn6.1 & \phn30 &\phn260 & 1.4 & 0.75 & 50 & --16.20 & 18.92 & 0.18 & --0.35 & \nl NGC 1800 &\phn803 &\phn664 &\phn9.2 & \phn45 &\phn520 & 1.9 & 0.63 & 113 & --16.72 & 19.73 & 0.13 & --0.46 & \nl NGC 2101 & 1192 &\phn973 & 13.6 & 66 & 1600 & 2.6 & 0.64 & 85 & --17.40 & 22.24 & 0.45 &\nodata & \nl II Zw 40 & \phn789 &\phn751 & 11.1 & \phn54 & \phn470 & 2.4 & \nodata & \nodata & --16.77 & 19.42 & 1.86 & --0.8 & \nl NGC 2915 & \phn460 &\phn183 &\phn2.9 &\phn14 & \phn190 & 0.7 & 0.54 & 129 & --14.73 & 20.02 & 0.79 & $\le-0.30$ & \nl NGC 3125 & 1110 &\phn827 & 13.8 &\phn67& \phn500 & 2.5 & 0.90 & 114 & --18.03 & 18.24 & 0.71 & --0.46 & \nl NGC 3955 & 1491 & 1227 & 21.1 & 102 & 1770 & 5.9 & 0.33 & 165 & --19.53 & 21.17 & 0.62 &\nodata & \nl NGC 4670 & 1069 & 1031 & 14.6 &\phn70 & \phn 540 & 2.8 & 0.83 & 90 & --17.78 & 18.81 & 0.12 & --0.50 & \nl NGC 5253 & \phn404 &\phn156 &\phn3.3 & \phn16 & \phn350 & 1.8 & 0.50 & 45 & --17.05 & 19.58 & 0.43 & --0.43 & \nl \enddata \tablecomments{ (2) Galaxy heliocentric velocity taken from the NASA/IPAC Extragalactic Database (NED). (3) Velocity relative to the centroid of the Local Group, computed from the heliocentric velocity in col. (2) following the precepts of the RSA. (4) Adopted distance to the galaxy. For NGC 5253, we have used a distance of 3.3 Mpc for the Centaurus group. For other galaxies, we used a Virgocentric infall model (cf. Schecter 1980) with the parameters $\gamma = 2, \omega_\odot = 222$ km s$^{-1}$, $v_{\rm{Virgo}}=976$ km s$^{-1}$ (Bingelli et al. 1986), and D$_{\rm{Virgo}} = 15.9$ Mpc (i.e., $H_0=75$ km s$^{-1}$ Mpc$^{-1}$). (5) scale in pc per arcsecs. (6) Effective (half-light) radius of galaxy derived from our B-band images. (7) Isophotal radius at a surface brightness level of 25 B mag arcsec$^{-2}$ (8) Ratio of semiminor to semimajor axes at $r_{25}$ dervied from our optical images. (9) Position angle of the optical major axis at large radii derived from our optical images. (10) Absolute total B magnitude. For most cases, this was determined from the growth curve of the galaxy. NGC 5253 took a large fraction of the frame, and the growth curve did not level off, so the flux outside of $r_{25}$ was extrapolated using the exponential fit and added to the flux inside $r_{25}$ to obtain the total magnitude. (11) Face on surface brightness within $r_e$ , corrected for Galactic extinction. (12) Galactic extinction for B (in magnitudes) based on the HI column densities in Stark et al. 1992, using the extinction curve provided in Mathis (1990). (13) The log of the metallicity of the ionized gas relative to solar (Meurer et. al (1994), Storchi-Bergmann et al. 1995, Telles 1996). } \end{deluxetable} The data we collected on these galaxies include broad band UBVI imaging, narrow band \Hline{\mbox{$\alpha$}}\ imaging, and long-slit \Hline{\mbox{$\alpha$}}\ spectroscopy. An initial paper, Marlowe {\it et~al.\/}\ (1995\markcite{mhws95}), presents \Hline{\mbox{$\alpha$}}\ imaging and spectroscopy of the seven galaxies in our sample with evidence of outflows. It shows that the bubbles have short ($\sim 10$ Myr) expansion timescales, and expansion velocities close to, but probably smaller than, the escape velocity of the galaxies. The energetics of the outflows were estimated and shown to be in accord with models of the young populations in the centers of these galaxies. In the first paper of this series (Marlowe {\it et~al.\/}\ 1997; hereafter \cite{pap1}) we presented the entire dataset on all the galaxies in the sample. The broad band images were used to construct surface brightness profiles, which show that these galaxies typically consist of a young blue core population, which we identify as the starburst, within an older redder envelope with an exponential radial surface brightness profile, which we identify as the underlying ``host'' galaxy. We also compared the global properties (e.g.\ magnitudes, colors, \Hline{\mbox{$\alpha$}}\ luminosities, \mbox {H\thinspace{\footnotesize I}}\ fluxes, etc\ldots) of our sample with samples of non-bursting gas rich dwarf irregulars (dI), gas poor dwarf ellipticals (dE), as well as other samples of dwarf starbursts (i.e.\ BCD and \mbox {H\thinspace{\footnotesize II}}\ galaxy samples). While there were clear differences (in the obvious sense) in the global properties of our sample compared to dEs and dIs, the differences with other dwarf starburst samples were small and could be traced to the original selection criteria. This strengthens the claim that the differences in the dwarf starburst classifications corresponds to little more than different naming schemes or slight differences in selection techniques. Here we consider the structure and stellar population of our sample in more detail. In \S\ref{secData} the data are summarized. In \S\ref{secStruct} we consider the structure of our sample galaxies and show that the structure of most of the galaxies in our sample is qualitatively and quantitatively the same as found in other dwarf starburst classifications, confirming that these classifications isolate the same physical phenomenon. In \S\ref{secHost} the structure and stellar content of the underlying envelope, or host galaxy, is used to test possible evolutionary connections between starburst dwarfs, dIs and dEs. We also fit models to the core and envelope colors to constrain the star formation history. Our conclusions are summarized in \S\ref{secConc}. \section{Data and Analysis}\label{secData} The complete dataset for this project consists of UBVI and \Hline{\mbox{$\alpha$}}\ Fabry-Perot images and \Hline{\mbox{$\alpha$}}\ \'Echelle spectra. The reduction and analysis of these data are discussed in detail in \cite{pap1}, and the reader is referred there for full details. Here we are primarily concerned with the broad band imaging data. The UBVI images were fully reduced, and calibrated as described in \cite{pap1}. The surface brightness profiles of most of our sample (8/12) appear to have two components, an outer region that is roughly exponential, which we call the envelope, and an inner region in excess of this inwardly extrapolated exponential, which we call the core. In all cases we fit an exponential to the outer portion of the galaxies between $1.5 r_{e,B}$ (where $r_{e,B}$ is the effective or half light radius in the $B$ band of the measured profile), and the outer limits of the photometry (where the surface brightness is roughly twice the sky level uncertainty). The exponential fit has two parameters, the scale length $\alpha^{-1}$, and the extrapolated face-on central surface brightness $\mu_{0,c}$ (Freeman, 1970). For a pure exponential profile $r_e = 1.68\alpha^{-1}$ and $\mu_e = \mu_{0,c} + 1.12$ mag. We subtract the exponential model from the surface brightness profile, which yields the core surface brightness profile. Figure \ref{figSampDecomp} shows the decomposition of NGC 1705's surface brightness into the core and envelope components as an example. Comparisons of the envelope and the core can be found in Table \ref{tabDCProp}. As Haro 14 and NGC 3955 have almost pure exponential profiles, the results for the core are highly uncertain. Therefore we present only their envelope properties and the inner colors (within $0.5r_{e,B}$) of the entire galaxy. Neither envelope-core decomposition, nor exponential fit, were done for NGC~2101 due to the poor quality of the outer profiles, resulting from scattered light from a bright foreground star. Here we present only the ``core'' properties as defined by the inner colors within $0.5r_{e,B}$. \begin{figure*} \centerline{\hbox{\psfig{figure=bag_f01.eps,height=8.5cm}}} \caption{Sample decomposition of NGC 1705's surface brightness profile into envelope and core components.\label{figSampDecomp} } \end{figure*} \begin{deluxetable}{lccrccccccrccc} \footnotesize \tablewidth{0pc} \tablecaption{Envelope vs. Core: Basic properties\label{tabDCProp}} \tablehead{ &\multicolumn{6}{c}{Envelope} & { } & \multicolumn{4}{c}{Core} \\ \cline{2-7} \cline{9-12} \\ \colhead{Galaxy}& \colhead{M$_B$} &\colhead{$\mu_{0,c}$} & \colhead{$\alpha^{-1}$}& \colhead{(U$-$B)}& \colhead{(B$-$V)} & \colhead{(V$-$I)}& & \colhead{M$_B$} & \colhead{(U$-$B)}& \colhead{(B$-$V)} & \colhead{(V$-$I)} \\ \colhead{}& \colhead{} &\colhead{\scriptsize(mag arcsec$^{-2}$)} & \colhead{\scriptsize(pc)} & \colhead{} &\colhead{} & \colhead{} && \colhead{} & \colhead{} &\colhead{} & \colhead{}\\ \colhead{(1)}& \colhead{(2)} &\colhead{(3)} & \colhead{(4)} & \colhead{(5)} &\colhead{(6)} & \colhead{(7)} && \colhead{(8)} &\colhead{(9)} & \colhead{(10)} & \colhead{(11)} } \startdata Haro 14 & --17.04 & 19.7 & 430 & 0.03 & 0.51 & 0.38 & &\nodata & (--0.29) & (0.31) & (0.54)\nl NGC 625 & --16.28 & 21.3 & 650 & --0.01 & 0.55 & 0.84 & & --14.71 & --0.49 & 0.22 & 0.49 \nl NGC 1510 & --16.17 & 20.9 & 530 & --0.04 & 0.51 & 0.78 & & --15.97 & --0.37 & 0.24 & 0.39 \nl NGC 1705 & --15.57 & 21.0 & 390 & --0.23 & 0.48 & 0.69 & & --15.53 & --0.84 & --0.03 & 0.50 \nl NGC 1800 & --16.28 & 21.3 & 610 & --0.05 & 0.56 & 0.74 & & --15.77 & --0.32 & 0.15 & 0.43 \nl NGC 2101 &\nodata &\nodata &\nodata &\nodata &\nodata &\nodata &&\nodata & (--0.49) & (0.35) & (0.16) \nl NGC 2915 & --14.43 & 20.7 & 200 & 0.04 & 0.50 & 0.92 & & --13.36 & --0.59 & 0.23 & 0.53 \nl NGC 3125 & --17.54 & 19.5 & 480 & --0.34 & 0.35 & 0.63 & & --17.42 & --0.45 & 0.31 & 0.71 \nl NGC 3955 & --19.38 & 20.9 & 2170 & 0.07 & 0.71 & 0.78 & &\nodata & (--0.17) & (0.45) & (0.93) \nl NGC 4670 & --17.38 & 20.3 & 690 & --0.27 & 0.45 & 0.76 & & --17.39 & --0.60 & 0.29 &\nodata \nl NGC 5253 & --16.57 & 20.7 & 550 & --0.07 & 0.46 & 0.76 & & --16.16 & --0.65 & 0.22 & 0.44 \nl \enddata \tablecomments{ Col(2)-- Absolute total B magnitude of the envelope, corrected for Galactic extinction. Col(3)--Face on envelope B band central surface brightness of the underlying exponential galaxy, corrected for Galactic extinction, derived from the exponential fit. Col(4) Scale length in the B band of the envelope derived from the exponential fit. Cols (5),(6),(7)--Total exponential envelope colors, corrected for Galactic extinction. Col(8)--Absolute total B magnitude of the ``core'' (the residual galaxy after the exponential envelope described above is subtracted), corrected for Galactic extinction. Cols (9),(10),(11)--Total core colors, corrected for Galactic extinction and emission-line contamination. Colors shown in parentheses are not based on a core-envelope decompostion, but are merely the inner colors ($r <$ 0.5$r_{e,B}$)}. \end{deluxetable} We caution that the core-envelope separation is prone to large systematic errors. Consider first the effect of the apertures used to extract the profiles. In Paper I (appendix) we showed that fits to surface brightness profiles derived from circular annuli when applied to a galaxy with elliptical isophotes will underestimate the central surface brightness and overestimate the scale length. Such an error will also effects the core/envelope ratio. In the worst case, Haro 14 appears to have a core (core/envelope $\sim 0.4$) in the circular annuli profile, but little or no core when elliptical annuli photometry is used (Paper I, fig.~7). We do considerably better by using elliptical annuli photometry, because the isophotes are better approximated by ellipses. Using variable ellipse parameters fitted to the isophotes yields profiles of higher fidelity (with respect to the isophotes) than those from averaged parameters. However it results in a conceptual problem: if the shape varies with radius what is the shape of the envelope where the core dominates? For our sample the difference in total flux of the envelope is less than 15\%\ when results from variable and constant shaped ellipse parameters are compared, and there is no bias in the sign of the difference. Hence we adopt the constant shaped elliptical annuli photometry for the exponential fits and the core-envelope separation. There is one exception to this: II~Zw~40 which is a merging dwarf system (\cite{bk88}; \cite{vzss98}) and has an ``$\times$'' morphology (\cite{pap1}). For it we adopt the concentric circular annuli profiles centered on the brightness peak. The radial fit range (set by the depth of the photometry) also effects the results. As a worst case consider NGC~2915, for which we derive $F_{\rm B,core}/F_{\rm B,env} = 0.37$ ($B$ band core/envelope flux ratio) from our photometry. Using the deeper photometry of Meurer et al.\ (1994\nocite{mmc94}) we derive $F_{\rm B,core})/F_{\rm B,env} = 0.89$. In this case the onset of the core is gradual, and the fit we present in \cite{pap1} includes part of the core seen in the Meurer et al.\ profile. Better agreement is found for NGC~1705, for which we derive $F_{\rm B,core}/F_{\rm B,env} = 0.96$, whereas Meurer et al.\ (1992) have $F_{\rm B,core}/F_{\rm B,env} = 0.91$. In summary, mismatches in isophote shapes could cause a ``random'' error of $\lesssim 0.15$ in $F_{\rm B,core}/F_{\rm B,env}$, while the limited depth of our photometry may result in the ratio being underestimated by $\lesssim 0.5$ in some cases. The latter effect will be smaller in the $I$ band where the core is weaker. As we are using the galaxy colors to make quantitative statements about the galaxy's age, we need to consider the effects of emission line contamination. While in some galaxies the corrections are negligible, in others, colors can be affected by as much as a few tenths of a magnitude. UV-optical spectra for six of our twelve galaxies, kindly provided by Daniela Calzetti (see \cite{calzetal}), were used to measure the equivalent widths of the emission lines that fall within each of our filters for these galaxies. The aperture size of the spectra is $10^{\prime\prime} \times 20^{\prime\prime}$, which encompasses most of the cores in the galaxies. Thus the correction would only be valid in the inner regions of the galaxy, so we only apply them to the core. Excluding the exceptional case of NGC 1705, the average value for the ratio of the half-light radii of the emission-line gas and optical continuum in these galaxies is about 50\% (see Paper I for details). Since the gas is usually more compact than the starlight, we therefore make no corrections to the envelope colors. Comparing the equivalent widths with the FWHM of the filter, and adjusting for throughput differences of the different emission lines, we estimate a correction for each filter: \begin{equation} \label{eqEMLCor} \Delta m = 2.5 \log\left( 1 + \left(\sum_i{\frac{EW_i}{{\rm FWHM}_{fil}} \frac{T_i}{T_{fil}}} \right)\right), \end{equation} where $\Delta m$ is the correction to filter $m$ in magnitudes, $EW_i$ is the equivalent width of emission line $i$ found in filter $m$, FWHM$_{fil}$ is the full width at half maximum of filter $m$, ${T_i}$ is the throughput of the filter at the wavelength of emission line $i$, and $T_{fil}$ is the average throughput of the filter within its FWHM bandpass. If we assume the corrections scale with the \Hline{\mbox{$\alpha$}}\ equivalent width (derived from our \Hline{\mbox{$\alpha$}}\ images), and that NGC 5253 and NGC 3125 have line ratios that are typical of the sample, we can estimate the corrections for those galaxies for which we do not have spectra. Of the remaining galaxies, NGC~2101, Haro~14, and II~Zw~40 have \Hline{\mbox{$\alpha$}}\ equivalent widths that suggest emission line contamination would be a significant problem. For Haro~14 and NGC~2101 we estimate the corrections by assuming they have spectra similar to NGC 3125. II~Zw~40, however, has an exceptionally high \Hline{\mbox{$\alpha$}}\ equivalent width ($\sim 400$\AA\ in the central region), and may have very different line ratios. Therefore we do not have much confidence in the accuracy of such corrections, and feel it best to make no attempt at correcting II Zw 40 without multiwavelength spectra. The final corrected core colors for Haro 14, NGC 2101, NGC 3125 and NGC 5253 are given in Table \ref{tabDCProp}. The Leitherer \&\ Heckman (1995) stellar population synthesis models used below include the contribution from the nebular continuum (see Leitherer \&\ Heckman 1995\markcite{lh95} text for discussion); therefore we do not correct for nebular continuum contamination. \section{The Structure of Blue Amorphous Galaxies}\label{secStruct} \subsection{Morphology and Structure}\label{ssecMorph} Our sample, primarily selected to consist of nearby blue dwarf amorphous galaxies, includes galaxies with a variety of morphologies. The dominant form is galaxies with smooth round outer isophotes and centrally concentrated star formation. These are iE or nE galaxies in the nomenclature Loose \&\ Thuan (1985\markcite{lt85}) adopt for describing BCDs. Indeed they find that for BCDs these are the most common types. Our sample also includes II~Zw~40 a BCD/\mbox {H\thinspace{\footnotesize II}}\ galaxy with irregular structure at all radii (type iI following \cite{lt85}). Also in our sample are two highly inclined disk galaxies (NGC~625, NGC~3955) and one galaxy (NGC~2101) with a morphology like a normal dwarf irregular, albeit of somewhat high surface brightness. As we will discuss below, (U--B) colors bluer than about --0.4 are indicative of a strong starburst (burst age younger than $\sim$ 10$^8$ years for continuous star-formation and younger than $\sim$ 10$^7$ years for an instantaneous burst). We find that this signature of a starburst is in most cases correlated with the galaxy surface brightness profile: galaxies whose core color (U--B) color indicates a starburst have the afore-mentioned core-envelope structure. Those that do are NGC~1705, NGC~2915, NGC~3125, NGC~4670, and NGC~5253. Moreover, in all these cases the core is bluer than the envelope, as can be seen in table~\ref{tabDCProp}, and the color profiles shown in \cite{pap1} (Fig.~9). Four of these galaxies have been {\em firmly\/} classified as amorphous by Sandage \&\ Brucato (1979\markcite{sb79}) or Gallagher \&\ Hunter (1987\markcite{gh87}). The exception is NGC~4670 which was classified amorphous in a later paper (\cite{hvwg94}). These galaxies all have high effective surface brigthnesses as well: Table 1 shows that $\mu_{e,B}$ is brighter than 20 mag arcsec$^{-2}$ in all cases. Of the remaining five galaxies with redder cores in (U--B), one - Haro 14 - has an almost purely exponential profile; one - NGC~2101 - has a central plateau (flattening) of its surface brightness profile; one - NGC~625 - has both a core-halo structure and a central plateau, and two - NGC~1510 and NGC~1800 have simple core-envelope structures. Of these, only NGC~1510 and NGC~1800 are amorphous galaxies (the others are all border-line amorphous galaxies judging by doubts expressed by Sandage \&\ Brucato [1979] or Gallagher \&\ Hunter [1987]). NGC~625 and NGC~2101 have considerably fainter effective surface brigthnesses than the other galaxies ($\mu_{e,B}$ = 21.2 and 22.2 mag arcsec$^{-2}$ respectively - see Paper I). Note that we exclude from the above discussion NGC~3955, which is much more luminous than the other galaxies ($M_B = -19.5$), shows spiral arms (contrary to the original definition of the Amorphous classification (\cite{sb79}), and whose central region is riddled by dust lanes whose associated obscuration affect its colors and radial surface brightness profile (inducing an apparent plateau). We conclude that photometric indicators of a starburst (the presence of a photometrically-distinct core whose colors are bluer than the envelope and are sufficiently blue to require a short-duration starburst) are usually in satisfactory agreement with one another. These characteristics are present with a high incidence rate among bona fide amorphous galaxies. \subsection{Connection with Outflows} In Marlowe {\it et al} (1995) we discussed H$\alpha$ \'echelle spectroscopy that provided kinematic evidence for outflowing gas in seven of the galaxies in our present sample (NGC~1705, NGC~1800, NGC~2915, NGC~3125, NGC~3955, NGC~4670, and NGC~5253). In contrast, the galaxies Haro~14, NGC~625, NGC~1510, NGC~2101, and II~Zw~40 did not reveal outflows. Is there some connection between the presence or absence of an outflow and the structure or stellar content of the galaxy (based upon our multi-color surface-photometry)? We exclude from consideration the luminous spiral NGC~3955 (see above) and II~Zw~40 (for which we do not have reliable colors). Of the remaining ten galaxies, the five with high effective surface brightnesses ($\mu_{e,B} \leq$ 20 mag arcsec$^{-2}$) and whose cores have ``starburst colors''((U--B) $< -0.4$) are all driving outflows of ionized gas (NGC~1705, NGC~2915, NGC~3125, NGC~4670, and NGC~5253). In marked contrast, only {\it one} of the five galaxies with lower effective surface brightnesses and/or redder cores is driving a weak outflow (NGC~1800). This suggests that the condition needed for driving a large-scale outflow of ionized gas in a gas-rich dwarf galaxy is a transient (10 to 100 Myr) episode of star-formation in the galaxy core in which the rate of star-formation per unit area is also high. Both the short timescale and high intensity of the star-formation are in accord with the usual definitions of starbursts and with theoretical understanding of starburst-driven outflows (cf. MacLow \& Ferrara 1998). \subsection{The Dwarf Starburst Galaxy Connection}\label{ssecEvolve} Here we compare our sample to other samples of dwarf starburst galaxies, in order to determine to what degree our sample is representative of the dwarf starburst phenomenon. The comparison will primarily be with the samples of Papaderos {\it et~al.\/}\ (1996a,b; hereafter P96) and Telles (hereafter T97: \cite{tmt97}; \cite{tt97}). The P96 sample was primarily selected from the Thuan \&\ Martin (1981) list of BCDs, while the T97 sample was selected from {\em The Spectrophotometric Catalog of \mbox {H\thinspace{\footnotesize II}}\ Galaxies\/} (Terlevich, 1991). Hence we are comparing our morphologically selected sample to samples of color and emission line selected dwarf starbursts. One concern is that the morphological selection may result in the less extreme, or older, starbursts since emission lines are not as important in the selection. If so, we may expect to see a difference in the burst strength, as parameterized by the core/envelope flux ratio, then seen in our sample. We may also expect our sample to have redder core colors. Figure \ref{figCoreFluxRatio} compares the ratio of the core to envelope fluxes of our sample in B versus $M_B$ against the P96 sample, and in I versus $M_I$ against the T97 sample. For the T97 sample the I core flux is estimated by subtracting the total luminosity derived from the exponential fit to the envelope from the galaxy total luminosity in those galaxies that showed a core component. While the I band is less sensitive to the massive stellar population, it is the only band in common with our sample for which the profile fits are available in the T97 sample. Note the surface brightness profiles of the T97 sample were extracted using circular annuli and this may artificially induce an apparent central excess (`core') in the surface-brightness profile (see \cite{pap1} Appendix for details). \begin{figure*} \centerline{\hbox{\psfig{figure=bag_f02a.eps,height=8.5cm} \psfig{figure=bag_f02b.eps,height=8.5cm}}} \caption{Core to envelope flux ratio plotted against absolute magnitude for starburst dwarfs. a) $B$ band values: blue amorphous galaxies (BAGs, diamonds) and BCDs (stars). b) $I$ band values: BAGs (diamonds) and HII galaxies (triangles). The uncertainty in these ratios is likely to be $\lesssim 0.15$ due to mismatches in isophote shape, while the limited depth of the photometry may result in some ratios being underestimated by $\lesssim 0.5$, in the $B$ band, with a lesser effect in the $I$ band. \label{figCoreFluxRatio}} \end{figure*} Figure~\ref{figCoreFluxRatio} shows two things. Firstly, while some galaxies in the P96 and T97 samples have a stronger core/envelope flux ratio, there is nevertheless a strong degree of overlap between the samples. Secondly, the cores do not totally dominate the galaxy. The cores in our sample galaxies with core-envelope structures amount to 0.2 to 1.1 mag brightening in the B band light, relative to the envelope alone. In the P96 sample the maximum ratio, 2.15, corresponds to a 1.25 mag enhancement relative to the envelope alone. Similarly, Sudarsky \&\ Salzer (1995\markcite{ss95}) report core enhancements of 0.2 - 0.6 mag in their sample of BCDs. Hence starbursts in dwarf galaxies represent only a modest $\lesssim$ 1 magnitude enhancement to their pre-burst progenitors. Even allowing for some downwards biasing in our core/envelope ratios due to the limited depth of our photometry (see \S\ref{secData}) the cores represent a $\lesssim$ 1.5 magnitude enhancement. \begin{figure*} \centerline{\hbox{\psfig{figure=bag_f03a.eps,height=8.5cm} \psfig{figure=bag_f03b.eps,height=8.5cm}}} \caption{Comparison of cores and envelopes. a) Envelope scale length plotted against $M_B$ of the core. b) $M_B$ of Envelope plotted against $M_B$ of the core. \label{figCoreEnv}} \end{figure*} In Figure \ref{figCoreEnv}a we compare the absolute magnitude of the core with the size (scale length) of the envelope for our sample and the P96 sample. Here the two samples form a sequence, with brighter cores forming in larger galaxies. Similarly, Figure \ref{figCoreEnv}b shows that brighter cores are associated with brighter underlying galaxies (envelopes). To some degree, both these correlations reflect a selection effect: bright or large underlying galaxies with weak or absent starburst cores certainly exist (even in our sample - e.g. NGC~625, Haro~14, NGC~2101), but are harder to distinguish. Thus, the absence of points in the upper right (lower right) of Figure \ref{figCoreEnv}a (Figure \ref{figCoreEnv}b) is a selection effect. On the other hand, the absence of very bright cores in small or faint galaxies is not a selection effect. Thus, these figures suggest that the maximum strength of the starburst is correlated with the size and luminosity of the host galaxy. Similar trends are present in samples of more powerful starbursts in brighter and more massive host galaxies (cf. Lehnert \& Heckman 1996; Heckman et al 1998). Comparing core colors, we see that the cores in our sample have a mean $\langle {\rm V-I}\rangle = 0.39$ (Table \ref{tabDCProp}) similar to that found in the sample of T97 $\langle {\rm V-I}\rangle = 0.45$. The envelopes of our sample have similar colors ($\langle {\rm V-I}\rangle = 0.73$) to the envelopes studied by T97: $\langle {\rm V-I}\rangle = 0.76$. Figure \ref{figDiskProps} compares the structural parameters of the exponential envelope ($\alpha^{-1}$, $\mu_{0,c}$) of various starburst dwarf samples as well as other samples of dwarf systems. In Figure \ref{figDiskProps2}, we show these parameters as a function of the absolute magnitude of the exponential envelope. For the T97 sample, exponential fits were only done in the I band. Thus for these objects we assume the B scale length is the same as the I scale length, and convert I magnitude values to B by assuming an average envelope color of (B--I)$=1.4$. Examination of our sample implies that these assumptions are reasonable ones For our own sample, we find that the B and I scale lengths agree well (the B scale lengths are systematically smaller, but only by an average of 9\%). The (B--I) colors of the envelopes range from 0.9 to 1.5 with a median of 1.3. On the scale of Figure \ref{figDiskProps} and Fig.~\ref{figDiskProps2}, these possible offsets between B and I values are unimportant. A more significant problem with the T97 sample is that the structural parameters for this sample were derived from circular aperture profiles. This biases the results yielding fainter central surface brightnesses and longer scale lengths relative to fits of surface brightness profiles extracted through matched elliptical annuli. The magnitude of the effect is likely to be less than 40\%\ in scale length, but up to 1.2 mag in $\mu_{0,c}$ (\cite{pap1} Appendix). \begin{figure*} \centerline{\hbox{\psfig{figure=bag_f04.eps,height=8.5cm}}} \caption{Structural properties of the exponential portion of surface brightness profiles of (mostly) dwarf galaxies, are shown in the upper left panel. The upper right panel shows the symbol correspondence, with the top three entries corresponding to the {\em envelopes} of starburst dwarfs: our sample, \protect\cite{pltf96}, and T97; also shown are dIs -- Patterson \&\ Thuan 1996 (PT)\protect\markcite{pt96}. dEs -- \protect\cite{cb87}. The quantities are measured in the B band except for those from T97 which are measured in the I band (see the text for our adopted transformation). The dotted line marks the parameters of exponential profiles having $M_B = -18$ mag. Galaxies above this line are too bright to be dwarfs. The lower two panels project the points into one dimensional distributions. \label{figDiskProps}} \end{figure*} \begin{figure*} \centerline{\hbox{\psfig{figure=bag_f05.eps,height=8.5cm}}} \caption{Structural properties of the exponential portion of surface brightness profiles of dwarf galaxies as a function of absolute blue magnitude. Samples and symbols as in Fig. \protect\ref{figDiskProps}. The quantities are measured in the B band except for those from T97 which are measured in the I band (see the text for our adopted transformation). The arrows indicate the fading of the envelope if star formation were cut-off for 10 Gyr (see text). The starbursts could fade into dEs, but not into the dIs, which are too blue for such fading. \label{figDiskProps2}} \end{figure*} Figure \ref{figDiskProps} and Fig.~\ref{figDiskProps2} show that the different starburst dwarf samples cover an overlapping area on this diagram, that is distinct from dI and dE galaxies (more on this in \S\ref{secHost}). Our sample tends to have central surface brightnesses somewhat brighter (by about 0.5 mag arcsec$^{-2}$ in the mean) compared to the other starburst dwarf samples. This may in large part be due to differences in measurement techniques, and may also reflect the greater mean distances of the T97 galaxies compared to ours (differences in the mean distances of the various subsamples may result in a systematic change in the measured parameters since the more distant galaxies are less well resolved). As mentioned the T97 sample may have systematically fainter $\mu_{0,c}$ due to their use of circularly averaged surface brightness profiles. In addition our images have somewhat limited depth and field of view. Hence the exponential fits are more weighted to small radii. If the edge of the core contaminates the fit there will be a bias towards high $\mu_{0,B}$ (\S\ref{secData}). \begin{figure*} \centerline{\hbox{\psfig{figure=bag_f06.eps,height=8.5cm}}} \caption{Comparison of H$\alpha$ equivalent widths. The solid histogram shows the distribution for the total galaxy H$\alpha$ equivalent width, while the dotted histogram shows the H$\alpha$ equivalent width assuming that all H$\alpha$ flux is associated with the core (where a core is present). Also shown is the average (large square) and standard deviation (error bars) of the H$\alpha$ equivalent width for HII galaxies (\protect\cite{terlea91}). \label{figEWComp}} \end{figure*} Figure \ref{figEWComp} compares the \Hline{\mbox{$\alpha$}}\ equivalent widths of our sample to that of \mbox {H\thinspace{\footnotesize II}}\ galaxies to further investigate core age differences. The histogram shows our sample while the point with error bars shows the average and standard deviation of \Hline{\mbox{$\alpha$}}\ equivalent widths given in {\em The Spectrophotometric Catalog of \mbox {H\thinspace{\footnotesize II}}\ Galaxies\/} (\cite{terlea91}). We note that the Terlevich et al. equivalent widths were mostly taken through apertures corresponding to $\sim 0.5$ kpc, which would more likely correspond to the core than to the entire galaxy, so we show our core upper limit \Hline{\mbox{$\alpha$}}\ equivalent widths (see Table \ref{tabEWComp}) in a dashed line for comparison. As can be seen, our sample and the \mbox {H\thinspace{\footnotesize II}}\ galaxy sample are broadly similar, although the latter typically have somewhat higher equivalent widths. Note that the subsample of \mbox {H\thinspace{\footnotesize II}}\ galaxies studied by T97 are selected to have the highest \Hline{\mbox{$\beta$}}\ equivalent widths, and so should have somewhat younger luminosity-weighted mean ages than the more comprehensive sample of Terlevich et al. (1991). \begin{deluxetable}{lrrrrr} \tablewidth{0pt} \tablecaption{H$\alpha$ Equivalent Widths: modeled vs. measured \label{tabEWComp}} \tablehead{ \colhead{Galaxy} & \colhead{$EW_{\rm LH,B}$} & \colhead{$EW_{\rm LH,C}$} & \colhead{$EW_{\rm core}$} & \colhead{$EW_{\rm spec}$} & \colhead{$EW_{\rm gal}$} \\ \colhead{(1)} & \colhead{(2)} & \colhead{(3)} & \colhead{(4)} & \colhead{(5)} & \colhead{(6)} } \startdata Haro 14 & 9 & 159 & \nodata & \nodata & 109 \nl NGC 625 & 40 & 177 & 243 & \nodata & 45 \nl NGC 1510 & 9 & 159 & 127 & 70 & 54 \nl NGC 1705 & 136 & 604 & 249 & 42 & 101 \nl NGC 1800 & 1 & 159 & 128 & 20 & 42 \nl NGC 2101 & 184 & 255 & \nodata & \nodata & 45 \nl NGC 2915 & 40 & 177 & 136 & \nodata & 35 \nl NGC 3125 & 40 & 177 & 301 & 218 & 168 \nl NGC 3955 & 9 & 159 & \nodata & \nodata & 14 \nl NGC 4670 & 27 & 177 & 149 & \nodata & 96 \nl NGC 5253 & 64 & 177 & 234 & 475 & 92 \nl \enddata \tablecomments{(2) EW (in \AA) predicted by the LH95 instantaneous burst models for each galaxy, based upon the ages in col (6) of Table \protect{\ref{tabAges}}. (3) EW (in \AA) predicted by the LH95 continuous star formation models, based upon the ages in col (9) of Table \protect{\ref{tabAges}}. (4) EW of the core, assuming that the entire H$\alpha$ flux is emitted by the burst in the core. The continuum level is determined by interpolating between the V and I fluxes of the core only, and thus provides an upper limit on the true core equivalent width. (5) EW as measured for the inner region of the galaxy, from the multiwavelength spectra provided by D. Calzetti, taken in a $10^{\prime\prime} \times 20^{\prime\prime}$ aperture. This would roughly correspond with the core in most cases (NGC 5253 has a significantly larger core area), and provides a comparison with the upper limit given in col (4). (6) EW of the galaxy as a whole. This is determined by dividing the total H$\alpha$ flux by the R flux of the entire galaxy. Again, the continuum level was found by interpolating between the V and I fluxes of the galaxy. } \end{deluxetable} Thus we conclude that our sample of predominantly morphologically selected dwarf starburst galaxies, i.e., blue amorphous galaxies, represent the same basic physical phenomena as other dwarf starburst samples, i.e.\ BCDs and \mbox {H\thinspace{\footnotesize II}}\ galaxies. There is a substantial overlap in such key paremeters as core colors, envelope colors, \Hline{\mbox{$\alpha$}}\ equivalent widths, core to envelope flux ratios, and envelope structural parameters. We note only one difference with other dwarf starburst samples: BCDs and \mbox {H\thinspace{\footnotesize II}}\ galaxies tend to have somewhat lower metallicities (oxygen abundance) than amorphous galaxies ($\sim 1/6Z_\odot$ versus $\sim 1/3Z_\odot$; \cite{pap1}). This is probably a second order effect and we do not consider it any further here. \subsection{Star Formation Timescales of Cores and Envelopes}\label{secCETimes} We estimate ages for the cores and envelopes by comparing the UBVI colors of the cores and envelopes with the models of Leitherer \&\ Heckman (1995, hereafter \cite{lh95}) and Bruzual \& Charlot (1996, hereafter \cite{bcmod93}, see Bruzual \& Charlot 1993 for details). The latter models extend to larger timescales ($t > 4 \times 10^8$) years, which turns out to be important for modeling the envelopes. Note that the relatively large difference between the \cite{lh95} and \cite{bcmod93} evolutionary tracks at ages $\leq 10$ Myr is due to the fact that \cite{lh95} take into account the contribution of the nebular continuum, whereas \cite{bcmod93} does not. The LH95 models we use employ a Salpeter IMF ranging from 0.1$M_\odot$ to an upper cutoff mass of 100 $M_\odot$. We interpolate between the models for solar and quarter solar metallicities to obtain a model of one third solar metallicity, the average metallicity of our sample (see Table \ref{tabStats}). These models are used to determine the core properties. The \cite{bcmod93} models used here employ a Salpeter IMF ranging from $0.1M_\odot$ to $100M_\odot$ and have 40\%\ solar metallicity. They are employed to determine the envelope properties. \begin{figure*} \centerline{\hbox{\psfig{figure=bag_f07a.eps,height=8.5cm} \psfig{figure=bag_f07b.eps,height=8.5cm}}} \caption{Comparison of our sample with constant star formation models in the (U--B) {\em vs.} (B--V) plane (Figure 7a) and the (U--B) {\em vs.} (V--I) plane (Figure 7b). The cores are plotted as open stars, the envelopes as asterisks. For galaxies with weak or absent cores (NGC~2101, NGC~3955, \&\ Haro~14) the inner colors are plotted as pentagons. The data are compared with the constant star formation rate models of \protect\cite{lh95} (solid line) and \protect\cite{bcmod93} (dashed line). The separtion of the large large symbols along the model lines represent one dex in time, and is labeled with log(t [years]). The smaller symbols represent 0.2 dex in time. The galaxies are labeled as follows: (a) Haro 14, (b) NGC 625, (c) NGC 1510, (d) NGC 1705, (e) NGC 1800, (f) NGC 2101, (g) NGC 2915, (h) NGC 3125, (i) NGC 3955, (j) NGC 4670, (k) NGC 5253. The core colors are corrected for emission line contamination as discussed in the text. II Zw 40 and the envelope of NGC 2101 were omitted: II Zw 40 is strongly affected by emission lines and the outer regions of NGC 2101 is badly contaminated by a nearby red star. Also shown is the reddening vector for E(B-V)=0.1. For some galaxies, \protect\cite{calzetal} give E(B-V) values internal to the galaxy as derived from the Balmer decrement: NGC 1510 (0.08), NGC 1705 (0.00), NGC 1800 (0.07), NGC 3125 (0.13), and NGC 5253 (0.00). \label{figSFHa} } \end{figure*} \begin{figure*} \centerline{\hbox{\psfig{figure=bag_f08a.eps,height=8.5cm} \psfig{figure=bag_f08b.eps,height=8.5cm}}} \caption{Comparison with instantaneous burst models in the (U--B) {\em vs.} (B--V) plane (Figure 8a) and the (U--B) {\em vs.} (V--I) plane (Figure 8b). Symbols as in Fig. \protect\ref{figSFHa}. The solid line is the \protect\cite{lh95} model and the dashed line the \protect\cite{bcmod93} model. The disagreement of the two models at times $<$ 10$^7$ years is mostly due to the inclusion of the nebular continuum in the \protect\cite{lh95} models, but not in the \protect\cite{bcmod93} models. \label{figSFHb}} \end{figure*} The comparison of data and models is shown in Fig.\ \ref{figSFHa} and Fig.\ \ref{figSFHb} for constant star formation and instantaneous burst models respectively. We plot the core (star) colors and and envelope (asterisk) colors for the sample, along with the constant star formation models (Fig.\ \ref{figSFHa}) and the instantaneous burst models (Fig.\ \ref{figSFHb}). Where no core subtraction could be done, the inner ($< 0.5 r_{e,B}$) colors were used instead (pentagons). Table \ref{tabAges} shows the ages derived from the models for an instantaneous burst as well as for continuous star formation based on the $U-B$, $B-V$, and $V-I$ colors of the cores and envelopes. To determine the ages, we find the ``distance'' \begin{equation} d_t=\sqrt{\Sigma\left(\rm color_{gal}-color_{mod,t}\right)^2} \end{equation} of the observed colors to the models for each time step $t$, then determined the age with the minimum ``distance'' for each galaxy. The ages and $d_t$ are given in table \ref{tabAges}. Because NGC 4670 was assigned an arbitrary I zero point (see \cite{pap1}), we derive its age based solely on the $U-B$, $B-V$ colors. We use the age and the absolute blue magnitude of the population to estimate some properties of the population, namely the total mass and the star formation rate (SFR), the latter for constant star formation models only. \begin{table*} \caption{~}\label{tabAges} \centerline{\hbox{\psfig{figure=bag_t04.eps}}} \end{table*} The models imply that while the envelopes are $\sim 10^{10}$ years old (assuming continuous star formation), the cores are $\sim 10^7$ if they are instantaneous bursts, or $\sim 10^8$ years old if they have been undergoing continuous star formation. Note that NGC 3955 is an exception to this: both its central and envelope colors are consistent with constant star formation over a Hubble time. This is consistent with its spiral morphology and structure - it doesn't have a core, unlike the majority of our sample. Also, we note that NGC 3955 has pronounced dust lanes and a large ratio of far-IR to \Hline{\mbox{$\alpha$}}\ luminosity, so its colors may be strongly affected by dust. We see that the cores are much too blue to be consistent with constant star formation over a Hubble time. While in most cases the instantaneous burst models provide a marginally better fit to the broad-band data (compare $d_t$ values in Table~\ref{tabAges}), they have a hard time reproducing the \Hline{\mbox{$\alpha$}}\ fluxes. This is because there should be little ionizing flux at the implied burst ages of $\sim$ 10 Myr. This is shown in Table~\ref{tabEWComp} where we give the \Hline{\mbox{$\alpha$}}\ equivalent widths predicted for each core by the best fitting \cite{lh95} models to the broad band data and compare them to the \Hline{\mbox{$\alpha$}}\ equivalent width based on \Hline{\mbox{$\alpha$}}\ and broad band images of the galaxy (see \cite{pap1} for details). In column (4), we derive the measured \Hline{\mbox{$\alpha$}}\ equivalent width assuming that all the \Hline{\mbox{$\alpha$}}\ emission is associated with the core derived from (H$\alpha_{total}/R_{core}$, where the R band flux $R_{core}$ is interpolated from the V and I fluxes derived for the core). Hence, here we are assuming that the nebulae are ionization bounded and that the core is the only, or dominant, ionizing population. If this is not the case, and the envelope also contributes significantly to the ionizing flux then this \Hline{\mbox{$\alpha$}}\ equivalent width can be considered an upper limit to the true equivalent width of the core. In column (6) we also give the equivalent width for the entire galaxy (derived from H$\alpha_{total}/R_{galaxy}$). Column (5) gives the equivalent widths of \Hline{\mbox{$\alpha$}}\ from the central region spectra of D.\ Calzetti (see \S\ref{secData}). If we assume that the core is producing all or most of the ionizing flux, then it is clear that there is better agreement with the constant star formation models in almost all cases. Notice, however, that the spectral equivalent width is significantly lower than the equivalent width we get by assuming that all the \Hline{\mbox{$\alpha$}}\ flux is produced by the core for three of the galaxies: NGC 1510, NGC 1800, and NGC 1705. While this may be due in part to the fact that the aperture of the spectra is not an exact match to the core, it could also indicate that a significant portion of the \Hline{\mbox{$\alpha$}}\ flux is being produced outside the core in these cases (e.g.\ NGC~1705 and NGC~1800). However this does not preclude our assumption that all the ionizing photons are produced in the core (i.e.\ some \mbox {H\thinspace{\footnotesize II}}\ emission is far from the ionizing source). Most of our sample have \Hline{\mbox{$\alpha$}}\ equivalent widths (measured with respect to the core) that are more consistent with constant star formation over $\sim 10^8$ year timescales, rather than the best fit instantaneous burst models. The exception is NGC~1705, which has an equivalent width that is too low to be consistent with the constant star formation model, but is comparable to the instantaneous burst model. This is in large part due to the presence of NGC1705-1. It is a ``super'' star cluster with an age of about 13 Myr (\cite{mfdc92}), and hence provides a strong UV-optical flux but little ionizing continuum. However, Meurer {\it et~al.\/}\ (1992) show that star formation is clearly continuing in the core which contains other smaller ionizing clusters (cf. Meurer {\it et~al.\/}\ 1995), and that the \Hline{\mbox{$\alpha$}}\ flux is consistent with what is expected from the stellar populations excluding NGC1705-1. When NGC1705-1's light is removed from the core photometry, its UBV colors are consistent with constant star formation over a $\sim$ Gyr timescale (\cite{mfdc92}). Another indication of a strong short duration burst in a galaxy is the presence of unusually strong Wolf-Rayet features in the spectrum. The presence of Wolf-Rayet stars in an integrated spectrum constrains a model of an instantaneous or nearly-instantaneous burst to an age between $3-7$ Myr (cf. \cite{lh95}). Four of the galaxies in our sample (NGC 1510, II Zw 40, NGC 3125, and NGC 5253) have broad emission lines at 4686\AA\ due to HeII (\cite{Conti91} and sources therein), usually the strongest line at optical wavelengths for Wolf-Rayet stars. However, a recent burst that produced the Wolf-Rayet stars in these galaxies may not account for their entire cores. A case in point is NGC~5253, probably the best studied ``Wolf-Rayet galaxy'' in our sample. Schaerer {\it et~al.\/}\ (1997\markcite{SCKM97}) find that the Wolf-Rayet features are localized to two knots, i.e.\ clusters, and derive ages of 2.8, and 4.3 Myr for them. Calzetti {\it et~al.\/}, (1997\markcite{C97}) studied the stellar populations in NGC~5253 at high resolution using broad and narrow band images taken with the Hubble Space Telescope. They confirm the ages of the Wolf-Rayet clusters, but show that other prominent clusters have ages up to $\sim 50$ Myr, and moreover, that the colors of the diffusely distributed stars between the clusters is consistent with continuous star formation for $\sim 500$ Myr timescales. This is fairly consistent with the 200 Myr duration we derive for the core (combined clusters and diffusely distributed stars) in the continuous star formation model. The overall picture we find is that the star formation histories of the ``starburst'' cores is complex, most likely consisting of continuous star formation, interspersed with strong enhancements as clusters form. While the timescale for cluster formation is probably $\leq$ few Myr, as implied by (localized) Wolf-Rayet features in some cases, the timescale for the whole of the core to form is less clear. At a minimum it should be the luminosity weighted mean age of $\sim 10$ Myr derived from instantaneous burst models. As discussed above, however, timescales of $\sim 100$ Myr are more likely for the total core. \section{Evolutionary connections} \label{secHost} Starbursting dwarfs are similar to dE and dI galaxies in having exponential surface brightness profiles, at least at large radii. If there are evolutionary connections between starburst dwarfs and the other types, then it is the envelope that tells us about the host, hence the progenitor and successor to the current starburst system. In Figure \ref{figDiskProps} and Fig.~\ref{figDiskProps2}, we compare the envelope structural properties of various samples of dwarfs. The dI sample is taken from Patterson \&\ Thuan (1996, hereafter \cite{pt96}). For the dEs, we chose the \cite{cb87} dwarf elliptical sample over the more extensive \cite{bc93} sample so as to compare results with exponential fits to elliptical annular surface brightness profiles where possible. Figure \ref{figDiskProps} shows a clear structural difference in the envelopes of starburst dwarfs compared to other dwarf galaxies. This is clear in both the $\alpha^{-1}$, $\mu_{0,c}$ plane and in the histograms of the two quantities. In terms of average quantities $\langle\log(\alpha^{-1} [{\rm kpc}]) \rangle$, $\langle \mu_{0,c} \rangle$, starburst dwarfs have -0.1, 21.3 (combined all three samples); dIs have 0.2, 23.6; and dEs have -0.3, 22.9. The median ($\log(\alpha^{-1} [{\rm kpc}]) $, $\mu_{0,c} $) quantities are -0.2, 21.2 for starburst dwarfs; 0.2, 23.7 for dIs; and -0.3, 22.8 for dEs. Considering the differences in their exponential envelopes, what evolutionary connections are allowed between dwarf galaxy classes? We now consider this question. \subsection{Progenitors of starburst dwarfs} The progenitors to starburst dwarfs must be gas rich systems. This rules out dEs, which are gas poor, as the progenitors. Dwarf irregulars have commonly been thought to be the most likely candidate to the preburst state (e.g.\ \cite{dp88}). However, the typical BCD has an envelope with scale length about a factor of two smaller than typical dIs, and central surface brightness about a factor of ten higher than typical dIs. Much of the difference in scale lengths may be attributed to selection effects. Both our sample and the P96 sample have a (loose) absolute magnitude limit ($M_B \gtrsim -18$), as does the PT sample of dI galaxies. This limiting magnitude (for a pure exponential profile) is shown as a dotted line in Fig.~\ref{figDiskProps}. Galaxies significantly above this line will not be selected except perhaps in the T97 sample; i.e.\ there is a bias against large scale length, high surface brightness galaxies -- they are not dwarfs. However, such galaxies do exist, and occasionally do make it into ``dwarf'' samples. However, there is still the factor of $\sim$ ten difference in central surface brightness to explain. Figure~\ref{figDiskProps} suggests a dividing line at $\mu_{0,c} \approx 22\, {\rm mag\, arcsec^{-2}}$; envelopes of starburst dwarfs relatively rarely have (extrapolated) central intensities less than this, while dIs rarely have higher central intensities. This too could in principle be due in part to a sample selection effect: do the galaxy catalogs that are used to generate samples of dIs (e.g. the UGC) sytematically exclude high surface brightness irregular galaxies from the list of objects classified as dIs? Is there are population of non-bursting, gas-rich, dwarfs with high central surface brightnesses and short scale lengths? The qualitative answer is ``yes'', since our sample includes objects like Haro~14, NGC~625, and NGC~2101 (none of which have strong blue starbursting cores). While it is beyond the scope of the present paper to quantify the importance of this effect, it is important to keep this caveat in mind. Selection effects aside, could a measurement bias cause the segregation of types in Fig.~\ref{figDiskProps}? The concern is that a low surface brightness envelope would be difficult to detect, with the available photometry, if a strong burst were present. This is unlikely to be a significant effect since neither our sample nor those of P96 nor T97 contain `bare cores' (i.e. an envelope is always detected around the bright blue starburst core). Nevertheless, we have performed some simple simulations to test this possible measurement effect. We assume a double exponential structure for burst plus envelope. We take $\alpha^{-1} = 120$ pc, $\mu_0 = 18.5\, {\rm mag\, arcsec^{-2}}$ for the burst, and vary the envelope structural parameters. We find that our photometry, which typically reaches $\mu(B) \approx 24.5\, {\rm mag\, arcsec^{-2}}$ should have no problem detecting the exponential profile of an envelope with $\mu_{0,c} = 22\, {\rm mag\, arcsec^{-2}}$ and $\alpha^{-1}$ in the range of 0.5 to 2 kpc. However at $\mu_{0,c} = 23\, {\rm mag\, arcsec^{-2}}$, it would be difficult to detect short scale length, $\alpha^{-1} \lesssim 0.5$ kpc, (face-on) envelopes with our photometry. Envelopes with $\mu_{0,c} \gtrsim 24\, {\rm mag\, arcsec^{-2}}$ would be difficult to detect, unless at a high inclination. The photometry of T97 frequently goes about a magnitude deeper (albeit in the V band). Consequently they should be able to detect the exponential envelopes with $\mu_{0,c}(V) = 24\, {\rm mag\, arcsec^{-2}}$ and $\alpha^{-1} \gtrsim 1$ kpc. The photometry of P96 is the deepest, typically reaching $\mu(B) \approx 26.5\, {\rm mag\, arcsec^{-2}}$. They should be able to detect exponential envelopes $\mu_{0,c}(B) = 24\, {\rm mag\, arcsec^{-2}}$ and $\alpha^{-1} \gtrsim 0.7$ kpc. Our conclusion is that enough of the envelope parameter space is available to our study or others that moderately low surface brightness envelopes should have been detected if they exist. Deeper observations and more detailed simulations would be useful for limiting and characterizing the envelope measurement biases. We are left with the result that there is a real difference between the envelope structure of starburst dwarfs and dwarf irregulars. Could this be due to a dynamical response of the envelope to the starburst core? If a dI where to have a central burst, gas will have to accrete into the center. The response of the envelope to the increased potential will be a contraction, and consequently an increased surface brightness. Conversely, a wind associated with the starburst will expel mass from the center, and the stellar distribution will expand and fade in surface brightness in response. For a fractional change of mass $f$, the expansion/contraction factor $e \approx (1-f)/(1-2f)$ if the mass loss/gain occurs on a timescale much less than the dynamical time, and $e \approx 1/(1-f)$ if the mass loss/gain occurs on a timescale much greater than the dynamical time (\cite{ya87}). Considering the fairly long core star formation timescales we derive, it is likely that the mass loss or gain is slow. The problem is that for evolution between dIs and starburst dwarfs we need $e \approx 3$, or $f \approx 2/3$. This huge change in mass does not seem feasible. First, consider the effects of winds. In Marlowe {\it et~al.\/}\ (1995), we discuss in detail the dynamics of those galaxies showing evidence of large scale outflows. In that paper, we roughly estimate the energy output of the starbursting region based on galaxy-wide parameters. With the improved population models for the cores, we can now refine our estimates for the mechanical energy output of the starbursts in those galaxies with a windy \Hline{\mbox{$\alpha$}}\ morphology. The new estimates are shown in Table \ref{tabMechLum}, which also includes the estimates for those galaxies that do not have any evidence of large scale outflows. The main conclusions of Marlowe {\it et~al.\/}\ (1995) are not changed significantly by these new energy estimates: the mechanical luminosities of the present bursts are insufficient to remove large amounts of gas from the galaxies in our sample, but {\it are} sufficient to drive the large-scale expansion seen in the H$\alpha$ bubbles and to allow much/most of the newly-created metals to be ejected from the galaxy: adopting the termin0logy of \cite{dYH94} and \cite{mf99}, the bursts can accomplish ``blow out'', but not ``blow away''. Second, it should be remembered that dwarf galaxies are dark matter dominated. Both dI and starburst dwarfs are known to have dark matter halos dominating even into the optical face of the galaxy (e.g.\ DDO 154: \cite{cb89}; NGC~2915: \cite{mcbf96}). Hence even if all the ISM were accreted or removed, there is likely to be little effect on the potential well depth, and hence little expansion or contraction. \begin{deluxetable}{lrrccccc} \small \tablecaption{Star formation and Energy Rates of the Starbursts \label{tabMechLum} } \tablehead{ \colhead{Galaxy} & \colhead{$L_{\rm Bol,B}$} & \colhead{$L_{\rm Bol,C}$} & \colhead{$M_{*,{\rm B}}$} & \colhead{SFR$_C$} & \colhead{SFR$_{LyC}$} & \colhead{$\log\dot{E}_{\rm B}$} & \colhead{$\log\dot{E}_{\rm C}$} \\ \colhead{} & \colhead{($10^8 L_\odot$)} & \colhead{($10^8 L_\odot$)} & \colhead{($10^6 M_\odot$)} & \colhead{($M_\odot$ yr$^{-1}$)} &\colhead{($M_\odot$ yr$^{-1}$)} & \colhead{(ergs s$^{-1}$)} & \colhead{(ergs s$^{-1}$)} \\ \colhead{(1)} & \colhead{(2)} & \colhead{(3)} & \colhead{(4)} & \colhead{(5)} & \colhead{(6)} & \colhead{(7)} & \colhead{(8)} } \startdata NGC 625 &3.1 & 3.2 & 0.61 & 0.038 & 0.040 & 40.1 & 40.0 \nl NGC 1510 &9.3 & 9.3 & 2.24 & 0.106 & 0.068 &40.7 & 40.5 \nl NGC 1705 &6.7 & 14.0 & 0.87 & 0.278 & 0.04\phn & 40.4 & 40.4 \nl NGC 1800 &7.7 & 7.7 & 4.05 & 0.088 & 0.058& 40.9 & 40.4 \nl NGC 2915 &0.9 & 0.9 & 0.18 & 0.011 & 0.007 & 39.5 & 39.5 \nl NGC 3125 &37.5 & 39.3 & 7.47 & 0.469 & 0.67\phn & 41.2 & 41.1 \nl NGC 4670 &38.4 & 38.3 & 7.56 & 0.457 & 0.34\phn & 41.2 & 41.1 \nl NGC 5253 &13.2 & 12.2 & 2.08 & 0.146 & 0.17\phn & 40.6 & 40.6 \nl \enddata \tablecomments{Values in this table (except Col. (6)) are derived from the LH95 models using the core blue luminosity (Table \protect\ref{tabDCProp} and the ages as derived in Table \protect\ref{tabAges}). (2) Bolometric luminosity of the galaxy core, assuming the instantaneous burst models of LH95 as described in \S\protect{\ref{secCETimes}}. (3) Bolometric luminosity of the galaxy core, assuming the constant star formation model of LH95, as described in \S\ref{secCETimes}. (4) Mass of gas converted into stars in the core, assuming LH95 instanteous burst models. (5) Star formation rate in the core, assuming LH95 constant star formation rate star formation models and IMF described in \S3.3. (6) Star formation rate calculated from the same metallicity and IMF as in (5), but using the Lyman continuum as calculated from the H$\alpha$ luminosities in Paper I (Table 7) assuming case B conditions at $T=10^4$ K (Osterbrock 1989). (7) Rate of energy deposited by the core, assuming LH95 instantaeous burst models. (8) Rate of energy deposited by the core, assuming LH95 constant star formation rate models. } \end{deluxetable} Another way to enhance the $\mu_{0,c}$ of the envelope is to have galaxy wide star formation correlated with the core formation. A hint of this is suggested in our sample as can be seen in Figure \ref{figCorr} which plots the envelope U-B colors versus the absolute blue magnitude of the cores: brighter cores are correlated with bluer envelopes. This suggests that whatever event triggered the intense star formation in the core may also have triggered increased star formation throughout the galaxy on a less intense scale. However, there are three inadequacies with this idea. First, the color profiles show the envelopes to be redder, and hence have a longer star formation timescale (in most cases roughly a Hubble time) than the core. Second, we still have the problem with scale: we need a factor of about ten enhancement in surface brightness. This seems excessively large considering the rather moderate contribution of the core. Third, there may be more contamination of the envelope by the core (starburst) light in the larger systems. \begin{figure*} \centerline{\hbox{\psfig{figure=bag_f09.eps,height=8.5cm}}} \caption{Envelope U--B color {\em vs.\/} core absolute blue magnitude. Blue amorphous galaxies (BAG): this sample (diamonds), dEs: \protect\cite{cb87} (circles). \label{figCorr}} \end{figure*} We are left with the likelihood that there is little evolution between dI and starburst dwarf morphologies. Furthermore, high surface brightness starbursts, as we know them, do not occur in low surface brightness ($\mu_{0,c} > 22\, {\rm mag\, arcsec^{-2}}$) host galaxies. Rather they require a relatively high surface brightness host ($\mu_{0,c} < 22\, {\rm mag\, arcsec^{-2}}$). What then are the progenitors of starbursting dwarfs? They must have moderately high central surface brightness and relatively short scale length. The most likely contenders, we believe, would have pure exponential or plateau profiles. Some of these objects are found amongst samples of dIs, and comprise the dIs with the highest surface-brightness (see Figure 5). Other progenitors may already have one of the various `starburst' dwarf designations. However, the `starburst' classification would be incorrect in these cases. It would be based mainly on high surface brightness (relative to typical dIs). Closer examination would show these galaxies to have modest emission line equivalent widths, small or absent color gradients between the central regions and the outer extremes, and colors throughout that are consistent with a current star-formation rate not too different from the past average over a Hubble time. This idea is similar to the scenario of Meurer {\it et~al.\/}\ (1994\markcite{mmc94}): at the peak of its star formation burst the system would look like a strong core/envelope system such as NGC~1705, while before or after the burst it would like a weak core system such as Haro 14, NGC~625, or NGC~2101. That is, we would expect the class of amorphous galaxies to contain systems in both the `burst' (strong blue core) and `inter-burst' (weak/absent core) states. Such galaxies might cycle between these two states intermittently until the gas is used up or expelled (see below). Since the starburst cores seem to involve only a few percent at most of the total galaxy stellar mass, there is no strong constraint on the duty cycle (e.g. the galaxy could go through many such weak bursts and spend much of its time in this phase). Salzer {\it et~al.\/}\ (1995) \markcite{salzetal95} presents a large survey of bursting (including weakly bursting galaxies) and ``normal'' dwarfs. The fractional number density of bursting dwarfs is about 1/3. This suggests that starburst dwarfs at $M_B \approx -16$ are ``on'' about 1/3 of the time. With a typical burst duration of $\sim 2 \times 10^8$ yr, the typical quiescent period would last $\sim 6 \times 10^8$ years. Thus, over 10 Gyr, a galaxy could conceivably burst about 10 times, with a total star forming time of a few $10^9$ years. This is on the order of the gas consumption times seen in most blue amorphous galaxies (cf. \cite{mhws95}). About 1/3 of the dIs studied by \cite{pt96} also have core/envelope structures, but the central intensity of even the {\em core\/} is much lower - typically $\mu_{0,c}(B) \gtrsim 21\, {\rm mag\, arcsec^{-2}}$. This suggests that dIs also have episodes of star formation. However the intensity of star formation is generally lower, and apparently does not usually produce the characteristics (high surface brightness in continuum and lines) which we normally associate with starbursts. There may be a good dynamical basis for a correlation between burst intensity and host structure. There is a correlation between dark matter halo central density $\rho_0$ and galaxy surface brightness - low surface brightness dIs have low $\rho_0$ (\cite{dbm97}) while high surface brightness starburst dwarfs have high $\rho_0$ (\cite{mcbf96}; \cite{mssk98}). In both cases the DM dominates, even into the centers of the galaxies. If efficient star formation, i.e. a burst of star formation, occurs when a disk becomes unstable to self gravitation and forms stars on the shortest possible timescale - the dynamical timescale, then we expect the surface brightness of the starburst $\propto \rho_0$ (\cite{mhlll97}). A (weak) correlation in this sense is also observed for starbursts in normal systems (\cite{mhlll97}). \subsection{Evolutionary endpoints of starburst dwarfs} In the previous section, we have explored the idea that the progenitors of the dwarf starburst systems are most likely the ``coreless'' amorphous galaxies like Haro~14, NGC~625, and NGC~2101 in the present sample and dIs of the highest surface brightness. We have also suggested that such galaxies may cycle repeatedly between ``on'' and ``off'' states (a recurring series of mild starbursts). Thus, on intermediate time-scales the evolutionary descendants of starburst dwarfs are likely to be objects like the three listed above. In this section we explore the plausibility of a more drastic and long-term evolutionary path which might occur if a starburst dwarf were able to consume or lose its gas supply. We first consider what the evolutionary path would be {\it if} one of our bursting dwarfs were to lose its gas and cease forming stars. We then consider whether this scenario is physically plausible. Dwarf ellipticals are the most likely candidates for the ultimate end product of a starbursting dwarf that loses all its gas. This possible evolutionary link has long been proposed (e.g.\ \cite{bmcm86}; \cite{dh91}, \cite{mfdc92}). The envelopes currently seen in most blue amorphous galaxies are too bright and too blue to be dE envelopes (however cf.\ NGC~2915). Starburst dwarf envelopes are consistent with constant star formation for several Gyr. If this star formation were turned off, however, after another several Gyr, the blue amorphous galaxy envelopes would fade several magnitudes and become redder. For example, a standard Bruzual \& Charlot (1996) model (Salpeter IMF from 0.1 to 125 M$_{\odot}$ with 40\% solar metallicity) for star-formation with a duration of 3 to 10 Gyr will fade by by about 2.4 magnitudes in the next 10 Gyr if star-formation is shut off. This would bring starburst dwarf envelopes well into the $\mu_{0,c}$ regime of dEs (Fig \ref{figDiskProps} and Fig.~\ref{figDiskProps2}). A fraction of the dEs in the Caldwell \&\ Bothun (1987\markcite{cb87}) sample have cores, and are designated as ``nucleated'' dEs. We plot the scale length versus $M_{B,core}$ and $M_{B,core}$ versus $M_{B,env}$ for these galaxies along with starburst dwarfs in Figure \ref{figdEcores}. There is more scatter for dEs, which is what one might expect if they are faded compact galaxies: the scale length would not change drastically, though the core magnitude would. The apparent lower limit to the scale length for dEs probably corresponds to a selection effect and seeing limitation since all these galaxies are in the Virgo and Fornax clusters. \begin{figure*} \centerline{\hbox{\psfig{figure=bag_f10a.eps,height=8.5cm} \psfig{figure=bag_f10b.eps,height=8.5cm}}} \caption{Comparison of dwarf ellipticals with cores to starbursting dwarfs. (a) Scale length vs. core absolute blue magnitude. Starbursting dwarfs: this sample (diamonds), P96 sample (stars), and dEs: \protect\cite{cb87} (large circles) and \protect\cite{bc93} (small circles). (b) M$_{B,core}$ vs. M$_{B,env}$ (symbols as in (a)). The vector shows the effect of cutting off star-formation in both the core and the envelope and allowing both to fade for 10 Gyr. The core fades more than the envelope because it has had a much shorter duration of star-formation (see text for details.) \label{figdEcores}} \end{figure*} If the cores of the amorphous galaxies and BCDs fade enough, the faded galaxies will move into the area occupied by these nucleated dEs. The \cite{bcmod93} models predict that after 10 Gyr, a 0.1 Gyr duration burst will fade about 6 magnitudes. Thus, if we were to shut off the star-formation for 10 Gyr in both the core and envelope of one of these starburst dwarfs, the stronger differential fading of the core (6 magnitudes) vs. the envelope (2.4 magnitudes) would move the galaxy along the core vs. envelope sequence in Figure \ref{figdEcores}b from the domain of the starburst dwarfs into the domain of the nucleated dEs. However, the fading of the core will not be so dramatic if cores result from repeated weak bursts rather than a single strong burst (the limiting case of many weak bursts will approximate a constant star-formation rate, which would then fade by only 2.4 magnitudes in 10 Gyr - see above). The major problem with starburst dwarfs evolving into dEs is gas removal. As previously noted, the mechanical luminosity of the cores is insufficient to remove all the ISM; at least not over the core star formation timescale. There is also the problem that dwarf ellipticals are found predominantly in clusters (\cite{vs91}), while starburst dwarfs are more numerous in the field (\cite{sal89}). One way around these problems is to have multiple bursts. The ISM consumed in star formation and ejected in winds would then be multiplied by the number of bursts. In clusters, evolution would occur at a more rapid pace as there are more frequent perturbing ``triggers''. In addition, ram pressure stripping of the ISM will play a more important role in clusters. This might explain the environmental differences between dEs and starburst dwarfs. The multiple burst scenario can also explain the correlation seen in dEs between exponential envelope color and the core strength shown in Fig.~\ref{figCorr}. This correlation is in the sense that brighter cores have redder envelopes, opposite in sense to what is seen for starburst dwarfs. \markcite{cb87}Caldwell \&\ Bothun (1987) attribute this to metallicity effects: dEs that have undergone multiple bursts before fading have accumulated more metals in the galaxy than those with fewer bursts. Multiple bursts would also build up the core strength as low mass stars from previous bursts would continue to contribute light to the core. It may not be possible to have such a correlation from cores formed in a single burst, since this would involve a much stronger burst, one more likely to expel a metal enriched ISM in a wind. We conclude that while it is possible in a photometric or purely phenomenological sense for the evolutionary endpoint of starburst dwarfs to be dEs, there is no compelling reason to expect that {\it typical starburst dwarfs in the present cosmological epoch} (such as those we have studied in this paper) will follow this pathway any time soon. In fact, we have argued that these starbursting dwarfs are experiencing rather mild recurring events, so that the depletion/loss of gas (and the resulting cessation of star-formation) will occur slowly. Turning this around, since the progenitors of present-day dEs were star-forming gas-rich dwarfs that succeeded in consuming/ejecting/losing their ISM, these progenitors must have undergone bursts of star-formation that were considerably more intense than the bursts seen in typical BCDs today. \section{Conclusions}\label{secConc} Our analysis of blue amorphous galaxies shows that their surface brightness and color profiles typically consist of a young burst occurring in an older, enveloping population (\cite{pap1}). The underlying galaxy surface brightness profile (which dominates beyond about 1.5 times the half-light radius) is consistent with an exponential profile. The young burst can be characterized by a central excess (the core) of this exponential profile. Some objects loosely classified as starburst dwarfs (our sample and others) instead have a purely exponential or central plateau structure, instead of the more common core - envelope structure. In these ``coreless'' objects there is little if any color gradient - the colors throughout are consistent with a roughly constant star-formation rate over the last few Gyr. We show that the underlying galaxies in our sample have colors consistent with roughly constant star formation over a Hubble time. The cores are younger and have broadband colors consistent with ages of $\sim 10$ Myr if formed in an instantaneous burst or $\sim 100$ Myr if they result from a constant rate star formation. Most cores have \Hline{\mbox{$\alpha$}}\ equivalent widths more consistent with the constant star formation rate models. The core typically provides about half the light of the galaxy, and comprises a few percent of the stellar mass, as derived from our population models (Table \ref{tabAges}) We thus conclude that the burst is a relatively modest event in the life of the underlying galaxy. These bursts are unlike the short duration dominant bursts hypothesized to occur in dwarf galaxies (\cite{bf96}) in order to explain the faint blue galaxy problem. The properties of our sample of predominantly morphologically selected dwarf starbursts has a high degree of overlap with those of other dwarf starburst samples, i.e.\ those going by the monikers BCD and \mbox {H\thinspace{\footnotesize II}}\ galaxies. They have similar core colors, envelope colors, \Hline{\mbox{$\alpha$}}\ equivalent widths, core to envelope flux ratios, and envelope structural parameters. Hence we conclude that all these starburst dwarf samples represent the same basic physical phenomenon: recent and intense high mass star formation occurring in a redder and presumably older host. Our study and the similar studies of Telles {\it et~al.\/}\ (1997a,b) and Papaderos {\it et~al.\/}\ (1996a,b) show that the structure of the envelope is different from the exponential profiles of most dI galaxies: starburst dwarfs mostly have face-on central surface brightness $\mu_{0,c}(B) \lesssim 22\, {\rm mag\, arcsec^{-2}}$, while dIs have $\mu_{0,c}(B) \gtrsim 22\, {\rm mag\, arcsec^{-2}}$. On average the difference in $\mu_{0,c}(B)$ between dI and starburst dwarfs amounts to $\sim 2.5\, {\rm mag\, arcsec^{-2}}$. While measurement bias is a concern (it is difficult to detect a compact low surface brightness envelope in the presence of a dominant starburst), these studies have explored enough of parameter space that low surface brightness envelopes to starburst dwarfs should have been detected if they existed. Instead only high surface brightness envelopes have been found. The large change in the structure of a dI required to transform it into a starburst dwarf envelope makes it unlikely that that such a transformation can occur as a dynamical response to the formation of the starburst core. Hence it is unlikely that typical dIs are the progenitors of starburst dwarfs. Instead it appears that episodes of star formation occur in both low surface brightness and high surface brightness hosts with the intensity of the star formation episode correlated with the host surface brightness. Only in high surface brightness hosts do these episodes reach a high enough intensity to be classified as a ``starburst''. We believe that the most likely progenitors (and immediate evolutionary descendants) of the starburst dwarfs are the ``coreless'' galaxies discussed above (which may be the same population as the members of the dI class with the highest surface brightness). These have envelopes like those of the starburst dwarfs but do not exhibit a very blue starbursting core. Even though they populate lists of ``starburst'' galaxies, we believe this is a misnomer: they enter catalogs because they are moderately blue and are actively forming stars, and have much higher surface-brightnesses than typical dIs. Since the bursts we have studied are modest events involving only a few percent of the stellar mass, high-surface-brightness gas-rich dwarfs could cycle between burst (strong core) and quiescent (weak/absent core) multiple times with a fairly large duty cycle (cf. \cite{salzetal95}) until they use up or expel the remaining gas. Ultimately the gas supply will be exhausted (and the star formation will then cease). The envelope properties would then fade on a timescale of Gyr into the region of the $M_B$--log$(\alpha^{-1})$ plane populated by dEs. One possible way to shut off star formation would be if the burst in the core blew away the ISM. This does not appear to be energetically feasible in a single burst. In addition the typical disk \mbox {H\thinspace{\footnotesize I}}\ geometry of dwarf galaxies makes it more difficult to blow out large a large fraction of the ISM (cf. \cite{dYH94}). An alternate method of turning starburst galaxies into gas poor dEs is through multiple bursts of star formation that both consume and expel the ISM until it is totally gone. If galaxies undergo bursts for about 1/3 of their lifetime (as suggested by the data in \cite{salzetal95}), most of the galaxies in this sample could turn their current HI gas into stars in about $10^{10}$ years. Thus, this evolutionary path is a slow one (typical starburst dwarfs in the present cosmological epoch will not start down the final path to dE status any time soon). Conversely, since present-day dEs most certainly did evolve from gas-rich dwarfs that experienced one or more intense bursts of star-formation in the distant past, these bursts must have been more powerful than the events occurring in typical present-day dwarfs. We thank the anonymous referee and the scientific editor, Greg Bothun, for their thoughtful and useful comments. This paper benefited from data and literature searches with NED, the NASA/IPAC Extragalactic Database, a facility operated by the Jet Propulsion Laboratory, Caltech, under contract with the National Aeronautics and Space Administration. This work was supported with NASA grant NAGW-3138.
\section{Introduction} Recently a new instant form of dynamics, the rest-frame Wigner-covariant instant form, was defined in Ref.\cite{lus} as an important tool for the program of obtaining a unified description of the four interactions with a canonical reduction to the independent degrees of freedom (Dirac's observables for a generalized Coulomb gauge; see Ref.\cite{dubna} for a review). This tool emerged from the study of the reformulation of ordinary theories in Minkowski spacetime on arbitrary spacelike hypersurfaces as a prerequisite to their description in curved spacetimes. Since this formulation is the classical background of the Tomonaga-Schwinger description of quantum field theory, this new instant form will also open the possibility to obtain a Wigner-covariant equal time quantization of Minkowski theories, to define a set of Tomonaga-Schwinger asymptotic states and to open new possibilities for the description of relativistic bound states, whose formulation in the framework of the standard Fock approach is still problematic (see Ref.\cite{weinberg} and the spurious solutions of the Bethe-Salpeter equation \cite{naka}). In the rest-frame instant form of every isolated system there is a 3+1 splitting of Minkowski spacetime with a foliation whose leaves, labelled by a scalar time parameter $\tau$, are the spacelike Wigner hyperplanes $\Sigma_{W\tau}$ orthogonal to the four-momentum of the isolated system (assumed timelike). In an arbitrary Lorentz frame the points of the Wigner hyperplane $\Sigma_{W\tau}$ have coordinates $z^{\mu}(\tau ,\vec \sigma )=x^{\mu}_s(\tau )+\epsilon^{\mu}_r (u(p_s)) \sigma^r$: the point $x^{\mu}_s(\tau )$ is an arbitrary origin for the 3-coordinates $\vec \sigma$ on $\Sigma_{W\tau}$ and $p^{\mu}_s$, its conjugate momentum, is orthogonal to $\Sigma_{W\tau}$ [$\epsilon^{\mu}_r(u(p_s))$ are three suitable spacelike four-vectors tangent to $\Sigma_{W\tau}$]. After the canonical reduction of the isolated system to the Wigner hyperplanes we remain with four universal first class constraints\cite{lus}:\hfill\break i) one, $p^2_s\approx M^2_{sys}$, identifying the invariant mass $M_{sys}$ of the isolated system as the Hamiltonian for the evolution in the rest-frame time;\hfill\break ii) three others defining the rest frame by the requirement that the (Wigner spin 1) total 3-momentum of the isolated system is zero, ${\vec P}_{sys} \approx 0$ [so that $p^{\mu}_s\approx M_{sys} u^{\mu}(p_s)$, with the unit four-vector $U^{\mu}(p_s)$ describing the orientation of the Wigner hyperplanes with respect to the arbitrary Lorentz frame]. Among the physical canonical degrees of freedom there are the coordinates of a point of the Wigner hyperplane, different from its covariant origin $x^{\mu}_s(\tau )$ of the internal 3-coordinates, with canonical noncovariant coordinates ${\tilde x}^{\mu}_s(\tau )$. This point, the classical analogue of the Newton-Wigner position operator, is decoupled from the system, describes its ``external" center-of-mass variable and may be interpreted as a decoupled point particle observer with his clock measuring the rest-frame time. Instead, on the Wigner hyperplane there are only physical canonical relative degrees of freedom, whose identification requires the addition of three gauge fixing constraints ${\vec X}_{sys}\approx 0$, so that the constraints ${\vec X}_{sys} \approx 0$ and ${\vec P}_{sys}\approx 0$ are second class and may be eliminated. These constraints ${\vec X}_{sys}\approx 0$ may be interpreted as the elimination of an ``internal" center-of-mass-like variable of the system, which is the gauge variable conjugated to the first class constraints ${\vec P}_{sys}\approx 0$, so not to have a double counting of the center-of-mass degrees of freedom. The constraints ${\vec X}_{sys}\approx 0$ eliminate this ``internal" center-of-mass-like variable from the physical degrees of freedom forcing it to coincide with the arbitrary origin $x^{\mu}_s(\tau )=z^{\mu}(\tau ,\vec \sigma ={\vec X}_{sys}=0)$ of the Wigner hyperplane. An open problem is the identification of these constraints ${\vec X}_{sys} \approx 0$ for the various isolated systems. In a future paper\cite{iten} it will be studied for a system of scalar particles. Instead, in this paper we will study this probem for the real Klein-Gordon field, as a particular aspect of the more general problem of finding canonical collective and relative variables for its configurations. In Refs.\cite{lon,lon1} Longhi and Materassi found a set of such variables in the standard Lorentz covariant approach. In this paper we shall reformulate their solution on the Wigner hyperplanes starting from the formulation of the Klein-Gordon field on arbitrary spacelike hypersurfaces in Minkowski spacetime\cite{albad}. Then, after having expressed the energy-momentum tensor of the Klein-Gordon field on the Wigner hyperplanes, where it assumes a form similar to the energy-momentum tensor of a relativistic perfect fluid, we shall study Dixon's multipoles\cite{dixon} of the field. This will open the way to the identification of the ``internal" center of mass ${\vec X}_{\phi}={\vec q}_{\phi}$ of the Klein-Gordon field and to the realization that the Longhi-Materassi collective variable is not the center of mass but a ``center of phase" of the field configuration. By considering a Klein-Gordon field configuration as a relativistic extended object, we can make a detailed study of the kinematical properties of the description of such an object on Wigner hyperplanes. Then we shall apply the same approach to the system of N scalar particles interacting with a real Klein-Gordon field and to the system of a complex charged Klein-Gordon field plus the electromagnetic field. In the latter case its two components with definite sign of energy and charge (identified by means of the Feshbach-Villars formalism in absence of interaction) generate two centers of phase, so that, besides the global center of phase of the complex field, there is a second collective variable describing the global action-reaction between the two components. A similar collective variable exists also for the system of N particles interacting with a real Klein-Gordon field. In Section II the real Klein-Gordon field is reformulated on spacelike hypersurfaces in Minkowski spacetime, following Ref.\cite{albad}, and then restricted to Wigner hyperplanes. In Section III its modulus-phase variables on the Wigner hyperplane are defined. In Section IV there is the definition of the collective and relative canonical variables and the inverse canonical transformation to express the original real Klein-Gordon field in terms of them. It is shown that each constant energy surface of the Klein-Gordon field is the disjoint union of symplectic manifolds. Also a set of canonical multipoles is defined for a certain class of field configurations and there are some comments on the self-interactions of the field. In Section V the energy-momentum tensor of a configuration of the real Klein-Gordon field is restricted to the Wigner hyperplanes and there is a study of Dixon's multipoles on it. This suggests a definition for the relativistic center of mass of the field configuration. In Section VI, by using the group-theoretical methods of Ref.\cite{pauri}, there is a complete discussion of the concepts of ``external" and ``internal" (with respect to the Wigner hyperplane) centers of mass of the field configuration. The Longhi-Materassi collective variable is interpreted as a ``center of phase" of the field configuration. In Section VII the isolated system of N scalar particles interacting with a real Klein-Gordon field is studied to visualize its collective variables and their interpretation. In Section VIII the previous analysis is extended to the charged Klein-Gordon field by defining the collective and relative canonical variables for its positive and negative frequency parts with the help of the Feshbach-Villars formalism\cite{fv,cs,gross}. The two collective variables can be combined to give the overall ``center of phase" of the field configuration plus a collective relative variable. Then there is a study of the coupling to the electromagnetic field and of the Dixon multipoles. In the Conclusions there are some comments about the utility of the multipolar expansions of the Klein-Gordon field in general relativity and about the problems existing for the quantization of the collective and relative variables. In Appendix A there are some notations for spacelike hypersurfaces. In Appendix B there is a review of Longhi-Materassi papers. In Appendix C there is a review of the Feshbach-Villars formalism and the definition of the Fourier coefficients of the Klein-Gordon field corresponding to the solutions of the square-root Klein-Gordon equation with both positive (or negative) energy and electric charge in the free case. In Appendix D there is the expression of the invariant mass for the charged Klein-Gordon field plus the electromagnetic field on the Wigner hyperplanes in terms of the Fourier coefficients defined in Appendix C. \vfill\eject \section{The real klein-Gordon field on spacelike hypersurfaces.} In Ref.\cite{albad} there is the description of scalar electrodynamics on spacelike hypersurfaces. In this Section we shall review this description restricting ourselves to a real Klein-Gordon field with a self-interaction $V(\phi )$ following the scheme of Ref.\cite{lus} [see Appendix A for the notations; $\dot \phi (\tau ,\vec \sigma )=\partial \phi (\tau ,\vec \sigma )/\partial \tau$]. The action of a scalar Klein-Gordon field reads \begin{eqnarray} S&=& \int d\tau d^3\sigma N(\tau ,\vec \sigma )\sqrt{\gamma (\tau ,\vec \sigma ) }\nonumber \\ &&{1\over 2} \Big[ g^{\tau\tau} {\dot \phi}^2+2 g^{\tau \check r} \dot \phi \partial_{\check r}\phi + g^{\check r\check s}\partial_{\check r}\phi \partial_{\check s}\phi -m^2\phi^2 -2V(\phi ) \Big] (\tau ,\vec \sigma )=\nonumber \\ &=&\int d\tau d^3\sigma \sqrt{\gamma (\tau ,\vec \sigma )} {1\over 2}\Big[ {1\over N}[\partial_{\tau} -N^{\check r}\partial_{\check r}]\phi [\partial_{\tau} -N^{\check s}\partial_{\check s}]\phi + \nonumber \\ &+&N [\gamma^{\check r\check s}\partial_{\check r}\phi \partial_{\check s}\phi -m^2\phi^2-2V(\phi ) ] \, \Big] (\tau ,\vec \sigma ). \label{II1} \end{eqnarray} \noindent where the configuration variables are $z^{\mu}(\tau ,\vec \sigma )$, $\phi (\tau ,\vec \sigma )={\tilde \phi}(z(\tau ,\vec \sigma ))$. The main difference between the standard theory and the one on spacelike hypersurfaces is that now the configuration variables are the fields $\phi (\tau ,\vec \sigma )=\tilde \phi (z(\tau ,\vec \sigma ))=(\tilde \phi \circ z) (\tau ,\vec \sigma )$, with $\tilde \phi (x)$ solution of the Klein-Gordon equation. These new fields $\phi =\tilde \phi \circ z$ contain the nonlocal information about the 3+1 splitting of Minkowski spacetime $M^4$ with a foliation of spacelike hypersurfaces $\Sigma_{\tau}$, obtained through an embedding $R\times \Sigma \rightarrow M^4$, $(\tau ,\vec \sigma ) \mapsto z^{\mu}(\tau ,\vec \sigma )$ [$\Sigma$ is an abstract 3-surface diffeomorphic to $R^3$]. The fields $\phi (\tau ,\vec \sigma )$ have a built-in definition of equal time associated with the Lorentz-scalar time parameter $\tau$ which labels the leaves of the foliation. Since $z^{\mu}_{\tau}=Nl^{\mu}+N^{\check r}z^{\mu}_{\check r}$, with the lapse and shift functions $N$, $N ^{\check r}$ functionals of $z^{\mu}(\tau ,\vec \sigma )$ through the metric $g_{AB}(\tau ,\vec \sigma )$, one has ${{\partial}\over {\partial z^{\mu}_{\tau}}}=l_{\mu}{{\partial}\over {\partial N}}+z_{\check s\mu}\gamma^{\check s\check r}{{\partial}\over {\partial N^{\check r}}}$. The canonical momenta are \begin{eqnarray} \pi(\tau ,\vec \sigma )&=&{{\partial L}\over {\partial \partial_{\tau} \phi (\tau ,\vec \sigma )}}={{\sqrt{\gamma}(\tau ,\vec \sigma )}\over {N(\tau ,\vec \sigma )}} \Big[ \dot \phi -N^{\check r} \partial_{\check r}\phi \Big] (\tau ,\vec \sigma ),\nonumber \\ &&\Rightarrow \, \dot \phi (\tau ,\vec \sigma )=\Big[ {N\over {\sqrt{\gamma}}} \pi +N^{\check r}\partial_{\check r}\phi \Big] (\tau ,\vec \sigma ),\nonumber \\ \rho_{\mu}(\tau ,\vec \sigma )&=&-{{\partial L}\over {\partial \partial_{\tau} z^{\mu}(\tau ,\vec \sigma )}}=\nonumber \\ &=&l_{\mu}(\tau ,\vec \sigma ) \Big[ {{\sqrt{\gamma}}\over 2} [{1\over {N^2}}(\dot \phi -N^{\check r}\partial_{\check r}\phi )^2- \gamma^{\check r\check s}\partial_{\check r}\phi \partial_{\check s}\phi +m^2\phi^2+2V(\phi )] \Big] (\tau ,\vec \sigma ) +\nonumber \\ &+&z_{\check s\mu}(\tau ,\vec \sigma )\gamma^{\check s\check r}(\tau ,\vec \sigma ) \Big[ {{ \sqrt{\gamma}}\over N}\partial_{\check r}\phi (\dot \phi - N^{\check u}\partial_{\check u}\phi ) \Big] (\tau ,\vec \sigma ). \label{II2} \end{eqnarray} We have the following primary constraints \begin{eqnarray} {\cal H}_{\mu}(\tau ,\vec \sigma )&=& \rho_{\mu}(\tau ,\vec \sigma )- \nonumber \\ &&-l_{\mu}(\tau ,\vec \sigma ) \Big[ {{\pi^2}\over {2\sqrt{\gamma}}}- {{\sqrt{\gamma}}\over 2} [\gamma^{\check r\check s}\partial_{\check r}\phi \partial_{\check s}\phi -m^2\phi^2-2V(\phi )] \Big] (\tau ,\vec \sigma )- \nonumber \\ &-&z_{\check s\mu}(\tau ,\vec \sigma )\gamma^{\check r\check s}(\tau ,\vec \sigma ) [ \pi \partial_{\check r}\phi ](\tau ,\vec \sigma ) \approx 0, \label{II3} \end{eqnarray} \noindent and the following Dirac Hamiltonian [$\lambda^{\mu}(\tau ,\vec \sigma )$ are Dirac multiplier] \begin{equation} H_D=\int d^3\sigma \lambda^{\mu} (\tau ,\vec \sigma ){\cal H}_{\mu}(\tau ,\vec \sigma ). \label{II4} \end{equation} By using the Poisson brackets \begin{eqnarray} \{ z^{\mu}(\tau ,\vec \sigma ),\rho_{\nu}(\tau ,{\vec \sigma}^{'}) \} &=&\eta ^{\mu}_{\nu} \delta^3(\vec \sigma -{\vec \sigma}^{'}),\nonumber \\ \{ \phi (\tau ,\vec \sigma ),\pi(\tau ,{\vec \sigma}^{'}) \} &=& \delta^3(\vec \sigma -{\vec \sigma}^{'}), \label{II5} \end{eqnarray} \noindent one finds that the time constancy of the primary constraints does not imply the existence of new secondary constraints. The constraints are first class with the algebra $\{ {\cal H}_{\mu}(\tau ,\vec \sigma ),{\cal H}_{\nu} (\tau ,\vec \sigma ) \} =0$. The conserved Poincare' generators are \begin{eqnarray} p_s^\mu \left( \tau \right) &=&\int \rho ^\mu \left( \tau ,\vec \sigma \right) d^3\sigma ,\nonumber \\ J_s^{\mu \nu }\left( \tau \right) &=&\int \left[ z^\mu \left( \tau ,\vec \sigma \right) \rho ^\nu \left( \tau ,\vec \sigma \right) -z^\nu \left( \tau ,\vec \sigma \right) \rho ^\mu \left( \tau ,\vec \sigma \right) \right] d^3\sigma . \label{II6} \end{eqnarray} Following Ref.\cite{lus}, we can restrict ourselves to spacelike hyperplanes $z^{\mu}(\tau ,\vec \sigma )=x_s^{\mu}(\tau )+b^{\mu}_{\check r}(\tau ) \sigma^{\check r}$, where the normal $l^{\mu}=\epsilon^{\mu} _{\alpha\beta\gamma}b_1^{\alpha}(\tau )b_2^{\beta}(\tau )b_3^{\gamma}(\tau )$ is $\tau$-independent [with $b^{\mu}_{\tau}=l^{\mu}$; $b^{\mu}_{\check A}(\tau )=\Big( b^{\mu}_{\tau}, b^{\mu}_{\check r}(\tau ) \Big)$ is an orthonormal tetrad]. Using the results of that paper we find that $\{ x^{\mu}_s,p^{\nu}_s \} {}^{*}=-\eta^{\mu\nu}$, $J^{\mu\nu}_s=x_s^{\mu}p_s^{\nu}-x_s^{\nu}p_s^{\mu}+S^{\mu\nu}_s$ and that the constraints are reduced to the following ten ones at the level of Dirac brackets [now $\gamma^{\check r\check s}=-\delta ^{\check r\check s}$] \begin{eqnarray} {\tilde {\cal H}}^{\mu}(\tau )&=& \int d^3\sigma {\cal H}^{\mu}(\tau ,\vec \sigma )=\nonumber \\ &=&p^{\mu}_s-b^{\mu}_{\check A}(\tau ){\tilde P}^{\check A}_{\phi}=p^{\mu}_s-l ^{\mu}{\tilde P}^{\tau}_{\phi}-b^{\mu}_{\check r}(\tau ) P^{\check r}_{\phi}= \nonumber \\ &=&p^{\mu}_s-l^{\mu} {1\over 2} \int d^3\sigma \Big[ \pi^2+(\vec \partial \phi )^2 +m^2\phi^2 +2 V(\phi )\Big] (\tau ,\vec \sigma ) -\nonumber \\ &-&b^{\mu}_{\check r}(\tau ) \int d^3\sigma [\pi \partial_{\check r}\phi ] (\tau ,\vec \sigma ) \approx 0,\nonumber \\ {\tilde {\cal H}}^{\mu\nu}(\tau )&=&b^{\mu}_{\check r}(\tau ) \int d^3\sigma \sigma^{\check r} {\cal H}^{\nu}(\tau ,\vec \sigma )-b^{\nu}_{\check r}(\tau ) \int d^3\sigma \sigma^{\check r} {\cal H}^{\mu}(\tau ,\vec \sigma )=\nonumber \\ &=&S^{\mu\nu}_s-\Big( b^{\mu}_{\check r}(\tau )l^{\nu}-b^{\nu}_{\check r}(\tau ) l^{\mu}\Big)\nonumber \\ && {1\over 2}\int d^3\sigma \sigma^{\check r} \Big[ \pi^2+(\vec \partial \phi )^2+m^2\phi^2+2 V(\phi )\Big] (\tau ,\vec \sigma ) -\nonumber \\ &-&\Big( b^{\mu}_{\check r}(\tau )b^{\nu}_{\check s}(\tau )-b^{\nu} _{\check r}(\tau )b^{\mu}_{\check s}(\tau )\Big) \int d^3\sigma \sigma^{\check r} [\pi \partial^{\check s}\phi ] (\tau ,\vec \sigma ) =\nonumber \\ &=&S^{\mu\nu}_s-b^{\mu}_{\check r}(\tau )b^{\nu}_{\check s}(\tau )S_{\phi} ^{\check r\check s} - \Big( b^{\mu}_{\check r}(\tau )l^{\nu}-b^{\nu}_{\check r}(\tau )l^{\mu}\Big) S^{\check r\tau}_{\phi} \approx 0, \label{II7} \end{eqnarray} \noindent where ${\tilde P}^{\check A}_{\phi}=({\tilde P}^{\tau}_{\phi}; P ^{\check r}_{\phi})$ is the (Lorentz-scalar) 4-momentum of the field configuration and $S^{\check r\check s}_{\phi}=J^{\check r\check s}_{\phi} {|}_{ {\vec P}_{\phi}=0}$, $S^{\tau \check r}_{\phi}=J^{o\check r}_{\phi} {|}_{ {\vec P}_{\phi}=0}$ its spin tensor [see Appendix B]. The configuration variables are reduced from $z^{\mu}(\tau ,\vec \sigma )$, $\phi (\tau ,\vec \sigma )$, to $x_s^{\mu}(\tau )$, to the six independent degrees of freedom hidden in the orthonormal tetrad $b^{\mu}_{\check A}(\tau ) $, to $\phi (\tau ,\vec \sigma )$, and to the associated momenta: $p^{\mu}_s$ conjugate to $x^{\mu}_s$; six degrees of freedom hidden in $S^{\mu\nu}_s$ as the momenta conjugate to the variables hidden in the orthonormal tetrad $b^{\mu}_{\check A}$ [see Ref.\cite{lus} for the associated Hanson-Regge Dirac brackets for these degrees of freedom]; $\pi (\tau ,\vec \sigma )$ conjugate to $\phi (\tau ,\vec \sigma )$. If we select all the configurations of the system with timelike total momentum [$p^2_s > 0$], we can restrict ourselves to the special Wigner hyperplanes $\Sigma_{W\tau}$ orthogonal to $p^{\mu}_s$. The procedure for doing this canonical reduction implies\cite{lus} to boost at rest the variables $b^{\mu}_{\check A}$, $S_s^{\mu\nu}$, with the standard Wigner boost $L^{\mu}{}_{\nu}(p_s,{\buildrel \circ \over p}_s)$ for timelike Poincar\'e orbits [${\buildrel \circ \over p}_s^{\mu} =(\epsilon_s=\eta_s\sqrt{p^2_s}; \vec 0)$; $\eta_s=sign\, p^o_s$] and then to add the gauge-fixings $b^{\mu}_{\check A}-L^{\mu}{}_{\nu =\check A}(p_s, {\buildrel \circ \over p}_s)=b^{\mu}_{\check A}-\epsilon^{\mu}_{\check A}(u(p _s))\approx 0$ [$u^{\mu}(p_s)=p^{\mu}_s/\epsilon_s$]. These gauge-fixings [only six of them are independent], together with ${\tilde {\cal H}}^{\mu\nu} (\tau )\approx 0$, form six pairs of second class constraints. The Dirac brackets with respect to these second class constraints [implying ${\tilde {\cal H}}^{\mu\nu}(\tau )\equiv 0$] admit the following canonical basis: $\phi (\tau ,\vec \sigma )$, $\pi (\tau ,\vec \sigma )$, ${\tilde x}^{\mu}_s(\tau )$ [it is not a 4-vector, but has only the covariance of the little group of timelike Poincar\'e orbits like the Newton-Wigner position operator], $p^{\mu}_s$. As shown in Ref.\cite{lus}, the indices $\check A=(\tau ,\check r)$ are replaced by $A=(\tau ,r)$; all the quantities $B^{\tau}$ are Lorentz scalars, while the quantities $\vec B=\{ B^r \}$ are spin-1 Wigner 3-vectors; instead $p^{\mu}_s$ is a Minkowski 4-vector giving the orientation of the Wigner hyperplane with respect to a given Lorentz frame. We get $J^{\mu\nu}_s={\tilde x}^{\mu}_sp^{\nu}_s-{\tilde x}^{\nu}_sp^{\mu}_s+ {\tilde S}_s^{\mu\nu}$ with the spin tensor ${\tilde S}^{\mu\nu}_s$ given in Eqs.(59) of Ref.\cite{lus}. Therefore, the final effect of these gauge-fixings is a canonical reduction to a phase space spanned only by the variables ${\tilde x} ^{\mu}_s(\tau )$, $p^{\mu}_s$, $\phi (\tau ,\vec \sigma )$, $\pi(\tau ,\vec \sigma )$, with standard Dirac brackets. Instead of ${\tilde x}^{\mu}_s$, $p^{\mu}_s$, describing a decoupled observer, one can use the canonical variables\cite{lus}: $\epsilon_s$, $T_s= p_s\cdot {\tilde x}_s/\epsilon_s=p_s\cdot x_s/\epsilon_s$, ${\vec k}_s={\vec p}_s/\epsilon_s$, ${\vec z}_s=\epsilon_s[{\vec {\tilde x}}_s-{{{\vec p}_s}\over {p^o_s}}{\tilde x}^o_s]$ [${\vec z}_s$ is the noncovariant 3-coordinate corresponding to the Newton-Wigner position operator]. The only surviving four constraints are [now ${\vec P}_{\phi}=\{ P^r_{\phi} \}$ is a spin-1 Wigner 3-vector] \begin{eqnarray} {\cal H}(\tau )&=&\epsilon_s -{\tilde P}^{\tau}_{\phi}=\nonumber \\ &=&\epsilon_s- {1\over 2} \int d^3\sigma \Big[ \pi^2+(\vec \partial \phi )^2+m^2\phi^2+2 V(\phi )\Big] (\tau ,\vec \sigma ) \approx 0,\nonumber \\ {\vec {\cal H}}_p(\tau )&=& {\vec P}_{\phi}= \int d^3\sigma [\pi \vec \partial \phi ](\tau ,\vec \sigma ) \approx 0. \label{II8} \end{eqnarray} \noindent where ${\tilde P}^A_{\phi}=({\tilde P}^{\tau}_{\phi}; {\vec P} _{\phi})$, ${\tilde P}^{\tau}_{\phi}=P^{\tau}_{\phi}+\int d^3\sigma V(\phi ) (\tau ,\vec \sigma )$, is the 4-momentum of the field configuration. The Dirac Hamiltonian is now $H_D=\lambda (\tau ){\cal H} +\vec \lambda (\tau )\cdot {\vec {\cal H}}_p$. By defining ${\bar S}^{AB}_s=\epsilon^A_{\mu}(u(p_s))\epsilon^B_{\nu}(u(p_s))S ^{\mu\nu}_s$, we can showRef.\cite{lus} that on the Wigner hyperplanes we have ${\bar S}^{AB}_s\equiv J^{AB}_{\phi}{|}_{P^D_{\phi}=0}= S^{AB}_{\phi}$ [$J^{AB}_{\phi}$ is the angular momentum of the field configuration] and ${\tilde S}^{ij}=\delta^{ir}\delta^{js}{\bar S}^{rs}_s\equiv \delta^{ir}\delta^{js}{\bar S}^{rs}_{\phi}$, ${\tilde S}^{oi}_s=-\delta^{ir} {\bar S}^{rs}_s\delta^{sj}p^j_s/(p^o_s+\epsilon_s)\equiv -\delta^{ir} {\bar S}^{rs}_{\phi}\delta^{sj}p^j_s/(p^o_s+\epsilon_s)$, so that $J^{\mu\nu}_s$ becomes independent from the boosts $S_{\phi}^{\tau r}$. Therefore, the generators of the realization of the Poincar\'e group in the rest-frame Wigner-covariant instant form of dynamics (``external" Poincar\'e algebra) are $p_s^{\mu}$ [or $\epsilon_s$, ${\vec k}_s$] and the Lorentz generators \begin{eqnarray} J^{ij}_s&=&{\tilde x}_s^ip_s^j-{\tilde x}_s^jp_s^i+\delta^{ir}\delta^{js} {\bar S}_s^{rs},\nonumber \\ J^{oi}_s&=&{\tilde x}_s^op_s^i-{\tilde x}^i_sp^o_s-{{\delta^{ir}{\bar S}_s^{rs} p_s^s}\over {p^o_s+\epsilon_s}},\nonumber \\ &&{}\nonumber \\ {\bar S}_s^{rs}&\equiv& S^{rs}_{\phi}=J^{rs}_{\phi}{|}_{{\vec P}_{\phi}=0}= \int d^3\sigma \{ \sigma^r[\pi \partial^s \phi ] (\tau ,\vec \sigma )-(r \leftrightarrow s) \} {|}_{{\vec P}_{\phi}=0}. \label{II9} \end{eqnarray} With the gauge fixing $\chi =T_s-\tau \approx 0$, we can eliminate the variables $\epsilon_s$, $T_s$, and find that the $\tau$-evolution (in the Lorentz scalar rest-frame time $T_s\equiv \tau$) is governed by the Hamiltonian \begin{eqnarray} H_R&=&M_{\phi}-\vec \lambda (\tau )\cdot {\vec {\cal H}}_p(\tau ),\nonumber \\ &&{}\nonumber \\ M_{\phi}&=&{\tilde P}^{\tau}_{\phi}= {1\over 2}\int d^3\sigma \Big[ \pi^2+(\vec \partial \phi )^2+m^2\phi^2 +2 V(\phi )\Big] (\tau ,\vec \sigma ), \label{II10} \end{eqnarray} \noindent where $M_{\phi}$ is the invariant mass of the field configuration [it replaces the non relativistic Hamiltonian $H_{rel}$ appearing in the separation of the center-of-mass motion $H={{ {\vec P}^2}\over {2M}}+H_{rel}$]. In the gauge $\vec \lambda (\tau )=0$, the Hamilton equations are [$\triangle =- {\vec \partial}^2$] \begin{eqnarray} \partial_{\tau} \phi (\tau ,\vec \sigma )\, &{\buildrel \circ \over =}\,& \pi (\tau ,\vec \sigma ),\nonumber \\ \partial_{\tau}\pi (\tau ,\vec \sigma )\, &{\buildrel \circ \over =}\,& [-\triangle -m^2] \phi (\tau ,\vec \sigma )-{{\partial V(\phi )} \over {\partial \phi}}(\tau ,\vec \sigma ),\nonumber \\ &&\Rightarrow [\partial_{\tau}^2+\triangle +m^2] \phi (\tau ,\vec \sigma )\, {\buildrel \circ \over =}\, -{{\partial V(\phi )}\over {\partial \phi}}(\tau ,\vec \sigma ). \label{II11} \end{eqnarray} We got a description in which the noncovariant canonical ``external" center-of-mass 3-variables ${\vec z}_s$, ${\vec k}_s$, move freely and are decoupled from the Klein-Gordon field variables $\phi (\tau ,\vec \sigma )$, $\pi (\tau ,\vec \sigma )$, living on the Wigner hyperplane and restricted by ${\vec P}_{\phi} \approx 0$. To reduce the field variables only to relative degrees of freedom one has to find an ``internal" collective center-of-mass-like variable ${\vec X} _{\phi}[\phi ,\pi]$ conjugate to ${\vec P}_{\phi}$, such that the gauge fixings ${\vec X}_{\phi}\approx 0$ force the position of the field ``internal" collective variable $\vec \sigma ={\vec X}_{\phi}$ to coincide with the origin $x^{\mu}_s(\tau )=z^{\mu}(\tau , \vec \sigma =0)$ of the Wigner hyperplane. In this way we get a decoupled point particle observer (the ``external" center of mass ${\tilde x}^{\mu}_s$ of the isolated system) whose scalar time $T_s\equiv \tau$ labels the evolution of the relative field variables on the Wigner hyperplanes, defined by the field configuration itself, foliating Minkowski spacetime. \vfill\eject \section{Fourier transform and modulus-phase variables on the Wigner hyperplane.} In the rest-frame Wigner-covariant instant form on Wigner hyperplanes [in which one considers only field configurations with a timelike 4-momentum], we can define the Fourier coefficients of the fields $\phi (\tau ,\vec \sigma )$, $\pi (\tau ,\vec \sigma )$ \begin{equation} \left\{ \begin{array}{l} a\left( \tau ,\vec q\right) =\int d^3\sigma \, \left[ \omega \left( q\right) \phi \left( \tau ,\vec \sigma \right) +i\pi \left( \tau ,\vec \sigma \right) \right] e^{i\left( \omega \left( q\right) \tau -\vec q\cdot \vec \sigma \right) }, \\ a^{*}\left( \tau ,\vec q\right) =\int d^3\sigma \, \left[ \omega \left( q\right) \phi \left( \tau ,\vec \sigma \right) -i\pi \left( \tau ,\vec \sigma \right) \right] e^{-i\left( \omega \left( q\right) \tau -\vec q\cdot \vec \sigma \right) }, \\ \\ \phi \left( \tau ,\vec \sigma \right) =\int d\tilde q \, \left[ a\left( \tau ,\vec q\right) e^{-i\left( \omega \left( q\right) \tau -\vec q\cdot \vec \sigma \right) }+a^{*}\left( \tau ,\vec q\right) e^{+i\left( \omega \left( q\right) \tau -\vec q\cdot \vec \sigma \right) }\right] , \\ \pi \left( \tau ,\vec \sigma \right) =-i\int d\tilde q \, \omega \left( q\right) \left[ a\left( \tau ,\vec q\right) e^{-i\left( \omega \left( q\right) \tau -\vec q\cdot \vec \sigma \right) }-a^{*}\left( \tau ,\vec q\right) e^{+i\left( \omega \left( \vec q\right) \tau -\vec q\cdot \vec \sigma \right) }\right] , \end{array} \right. \label{III1} \end{equation} \noindent with \begin{equation} \begin{array}{ccc} \omega \left( q\right) =\sqrt{m^2+ {\vec q}^2}, & d\tilde q=% \frac{d^3q}{\Omega \left( q\right) }, & \Omega \left( q\right) =(2\pi )^3\,2\omega \left( q\right) ,\quad q=|\vec q|=\sqrt{{\vec q}^2}. \end{array} \label{III2} \end{equation} \noindent Here $q_A\sigma^A=\omega ( q)\tau -\vec q\cdot \vec \sigma$ is defined by using $q^A=(q^{\tau}=\omega ( q);\, q^r)$ with $q^{\tau}$ Lorentz scalar and $\vec q$ a spin-1 Wigner 3-vector like $\vec \sigma$ [on arbitrary hyperplanes all the quantities $\tau$ ,$\vec \sigma$ ,$q^{\tau}$, $\vec q$, would be Lorentz scalars]. The Fourier coefficients are $\tau$-dependent only when there is a non null self interaction $V(\phi )$; they and their gradients are assumed to belong to the space $L_2(d\tilde q)$. From the Poisson brackets $\left\{ \phi \left( \tau ,\vec \sigma \right) ,\pi \left( \tau ,\vec \sigma^{\prime }\right) \right\} =\delta ^3\left( \vec \sigma -\vec \sigma^{\prime }\right) $, we get [leaving the factor $\Omega( q)$ to agree with the standard notations] \begin{equation} \left\{ a\left( \tau ,\vec q\right) ,a^{*}\left( \tau ,\vec k \right) \right\} =-i\Omega \left( q\right) \delta ^3\left( \vec q-\vec k\right) , \label{III3} \end{equation} The 4-momentum and angular momentum of the field configuration are [${\cal V}\left[ a,a^{*}\right] $ being the contribution of the potential $V\left( \phi \right) $ when present] \begin{eqnarray} {\tilde P}_{\phi}^\tau &=&P^{\tau}_{\phi}+{\cal V}\left[ a,a^{*}\right] = {1\over 2} \int d^3\sigma [ \pi^2+(\vec \partial \phi )^2+m^2\phi^2+2V(\phi )] (\tau ,\vec \sigma )=\nonumber \\ &=&\int d\tilde q \, \omega \left( q\right) a^{*}\left( \tau ,\vec q\right) a\left( \tau ,\vec q\right) +{\cal V}\left[ a,a^{*}\right] ,\nonumber \\ {\vec P}_{\phi}&=&\int d^3\sigma [\pi \vec \partial \phi ](\tau ,\vec \sigma ) =\int d\tilde q \, \vec q\, a^{*}\left( \tau ,\vec q\right) a\left( \tau ,\vec q\right) , \label{III4} \end{eqnarray} \begin{eqnarray} J_{\phi}^{rs}&=& \int d^3\sigma [\pi (\sigma^r\partial^s-\sigma^s\partial^r)\phi ](\tau ,\vec \sigma )=\nonumber \\ &=&-i\int d\tilde q \, a^{*}\left( \tau ,\vec q\right) \left( q^r\frac \partial {\partial q^s}-q^s\frac \partial {\partial q^r}\right) a\left( \tau ,\vec q\right) , \nonumber \\ J^{\tau r}_{\phi}&=&-\tau P^r_{\phi}+{1\over 2} \int d^3\sigma \, \sigma^r\,\, [\pi^2+(\vec \partial \phi )^2+m^2\phi^2](\tau ,\vec \sigma )=\nonumber \\ &=&- \tau P_{\phi}^r+i\int d\tilde q \, \omega \left( q\right) a^{*}\left( \tau ,\vec q\right) \frac \partial {\partial q^r}a\left( \tau ,\vec q\right) . \label{III5} \end{eqnarray} We want to define four variables $X_{\phi}^A[\phi ,\pi ]=(X_{\phi}^\tau ;{\vec X}_{\phi})$ canonically conjugated to $P_{\phi}^A[\phi ,\pi ]=(P_{\phi}^\tau ;{\vec P}_{\phi})$. First of all we make a canonical transformation to modulus-phase canonical variables \begin{equation} \left\{ \begin{array}{l} a\left( \tau ,\vec q\right) =\sqrt{I\left( \tau ,\vec q\right) }\, e^{\left[ i\varphi \left( \tau ,\vec q\right) \right]} , \\ a^{*}\left( \tau ,\vec q\right) =\sqrt{I\left( \tau ,\vec q\right) }\, e^{\left[ -i\varphi \left( \tau ,\vec q\right) \right]} , \\ \\ I\left( \tau ,\vec q\right) =a^{*}\left( \tau ,\vec q\right) a\left( \tau ,\vec q\right) , \\ \varphi \left( \tau ,\vec q\right) =\frac 1{2i}\ln \left[ \frac{a\left( \tau , \vec q\right) }{a^{*}\left( \tau ,\vec q\right) }\right] , \end{array} \right. \label{III6} \end{equation} \begin{equation} \left\{ I\left( \tau ,\vec q\right) ,\varphi \left( \tau ,\vec q^{\prime }\right) \right\} =\Omega \left( q\right) \delta ^3\left( \vec q-\vec q^{\prime }\right) . \label{III7} \end{equation} In terms of the original canonical variables $\phi$, $\pi$, we have \begin{eqnarray} I( \tau ,\vec q) &=&\int d^3\sigma \int d^3\sigma ^{\prime }\, e^{i\vec q\cdot ( \vec \sigma -\vec \sigma ^{\prime }) }[ \omega ( q) \phi ( \tau ,\vec \sigma ) -i\pi ( \tau ,\vec \sigma ) ] [ \omega ( q) \phi ( \tau ,\vec \sigma ^{\prime }) +i\pi ( \tau ,\vec \sigma ^{\prime }) ] , \nonumber \\ \varphi ( \tau ,\vec q) &=&\frac 1{2i}\ln\, \Big[ \frac{\int d^3\sigma \, [ \omega ( q) \phi ( \tau ,\vec \sigma ) +i\pi ( \tau ,\vec \sigma ) ] e^{i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }}{\int d^3\sigma ^{\prime }\, [ \omega ( q) \phi ( \tau ,\vec \sigma ^{\prime }) -i\pi ( \tau ,\vec \sigma ^{\prime }) ] e^{-i( \omega ( q) \tau -\vec q\cdot \vec \sigma ^{\prime }) }}\Big] =\nonumber \\ &=&\omega (q) \tau + \frac 1{2i}\ln\, \Big[ \frac{\int d^3\sigma \, [ \omega ( q) \phi ( \tau ,\vec \sigma ) +i\pi ( \tau ,\vec \sigma ) ] e^{-i\vec q\cdot \vec \sigma }}{\int d^3\sigma ^{\prime }\, [ \omega ( q) \phi ( \tau ,\vec \sigma ^{\prime }) -i\pi ( \tau ,\vec \sigma ^{\prime }) ] e^{ i\vec q\cdot \vec \sigma ^{\prime } }}\Big] . \label{III8} \end{eqnarray} \begin{equation} \left\{ \begin{array}{l} \phi \left( \tau ,\vec \sigma \right) =\int d\tilde q\, \sqrt{I\left( \tau ,\vec q\right) }\left[ e^{i\varphi \left( \tau ,\vec q\right) -i\left( \omega \left( q\right) \tau -\vec q\cdot \vec \sigma \right) }+e^{-i\varphi \left( \tau ,\vec q\right) +i\left( \omega \left( q\right) \tau -\vec q\cdot \vec \sigma \right) }\right] , \\ \pi \left( \tau ,\vec \sigma \right) =-i\int d\tilde q\, \omega \left( q\right) \sqrt{I\left( \tau ,\vec q\right) }\left[ e^{i\varphi \left( \tau ,\vec q\right) -i\left( \omega \left( q\right) \tau -\vec q\cdot \vec \sigma \right) }-e^{-i\varphi \left( \tau ,\vec q\right) +i\left( \omega \left( q\right) \tau -\vec q\cdot \vec \sigma \right) }\right] . \end{array} \right. \label{III9} \end{equation} The Poincar\'e charges of the field configuration take the form \begin{equation} \left\{ \begin{array}{l} {\tilde P}_{\phi}^\tau =P^{\tau}_{\phi}+{\cal V}\left[ I,\varphi \right] = \int d\tilde q\, \omega \left( q\right) I\left( \tau ,\vec q\right) +{\cal V}\left[ I,\varphi \right] , \\ \\ {\vec P}_{\phi}=\int d\tilde q\, \vec q\, I\left( \tau ,\vec q\right) , \end{array} \right. \label{III10} \end{equation} \begin{equation} \left\{ \begin{array}{l} J^{rs}_{\phi}=\int d\tilde q\, I\left( \tau ,\vec q\right) \left( q^r\frac \partial {\partial q^s}-q^s\frac \partial {\partial q^r}\right) \varphi \left( \tau ,\vec q\right) , \\ \\ J_{\phi}^{\tau r}=-\tau P_{\phi}^r-\int d\tilde q\, \omega \left( q\right) I\left( \tau ,\vec q\right) \frac \partial {\partial q^r}\varphi \left( \tau ,\vec q\right). \end{array} \right. \label{III11} \end{equation} The classical analogue of the occupation number is [$\triangle =-{\vec \partial}^2$] \begin{eqnarray} N_{\phi}&=&\int d\tilde q\, a^{*}(\tau ,\vec q) a(\tau ,\vec q) = \int d\tilde q\, I(\tau ,\vec q)=\nonumber \\ &=&{1\over 2} \int d^3\sigma [\pi {1\over {\sqrt{m^2+\triangle}}}+\phi \sqrt{m^2+\triangle} \phi ](\tau ,\vec \sigma ). \label{III12} \end{eqnarray} \section{Collective and relative canonical variables.} A sequence of canonical transformations from the field variables $\phi (x)$, $\pi (x)={{\partial \phi (x)}\over {\partial x^o}}$ to the Fourier coefficients $a(\vec k)$, $a^{*}(\vec k)$ [$k^{\mu}=(k^o=\omega (k=|\vec k|)= \sqrt{m^2+{\vec k}^2};\, \vec k)$], then to the modulus-phase (or action-angle) variables $I(\vec k)$, $\varphi (\vec k)$, and finally to a canonical basis $X^{\mu} _{\phi}[F]$, $P^{\mu}_{\phi}$, ${\cal H}(\vec k | F]$, ${\cal K}(\vec k | F]$ [depending on an arbitrary normalized weight function $F(P_{\phi},k)$] was found in Ref.\cite{lon} when the real Klein-Gordon field $\phi (x)$ satisfies certain conditions. The main results of that paper are reviewed in Appendix B. Here, we shall reformulate this approach on the Wigner hyperplane. Due to its peculiar Wigner covariance properties\cite{lus}, the weight function $F(P_{\phi},k)$ [with its generic non polynomial (for instance $F \approx e^{-P_{\phi}\cdot k}$) dependence upon $P^{\mu}_{\phi}$ due to manifest Lorentz covariance] may now be replaced by two separate scalar functions ${\tilde F}^{\tau}(P^{\tau}_{\phi},q)$, $\tilde F({\vec P}_{\phi},\vec q)$. Moreover, if we accept that these two functions are singular at $\vec q=0$, we can take them linear in $P^{\tau}_{\phi}$ and ${\vec P}_{\phi}$ respectively: ${\tilde F}^{\tau}(P^{\tau}_{\phi},q)=P^{\tau}_{\phi}\, F^{\tau}(q)$, $\tilde F({\vec P}_{\phi}, \vec q)=-{\vec P}_{\phi}\cdot \vec q F(q)$ with a suitable normalization for $F^{\tau}(q)$ and $F(q)$. From Appendix B, we see that we must require the following behaviours also on the Wigner hyperplane \begin{eqnarray} i)\, q\, \rightarrow \, \infty &&,\quad\quad \sigma > 0,\nonumber \\ &&|a(\tau ,\vec q)| \rightarrow q^{-{3\over 2}-\sigma},\quad |I(\tau ,\vec q)| \rightarrow q^{-3-\sigma},\quad |\varphi (\tau ,\vec q)| \rightarrow q,\nonumber \\ ii)\, q\, \rightarrow \, 0 &&,\quad\quad \epsilon > 0,\quad \eta > -\epsilon , \nonumber \\ &&|a(\tau ,\vec q)| \rightarrow q^{-{3\over 2}+\epsilon},\quad |I(\tau ,\vec q)| \rightarrow q^{-3+\epsilon},\quad |\varphi (\tau ,\vec q)| \rightarrow q^{\eta}. \label{IV1} \end{eqnarray} \noindent in order to have the Poincar\'e generators and the occupation number of Section III finite. Let us remark that the collective and relative canonical variables can be defined in closed form only in absence of self-interactions of the field, so that in this Section we shall put $V(\phi )={\cal V}[I,\varphi ]=0$. \subsection{Definition of the collective variables.} Let us define the four functionals of the phases \begin{eqnarray} &&X_{\phi}^\tau =\int d\tilde q \omega (q) F^\tau \left( q\right) \varphi \left( \tau ,\vec q\right) , \nonumber \\ &&{\vec X}_{\phi}=\int d\tilde q \vec q\, F\left( q\right) \varphi \left( \tau ,\vec q\right) ,\nonumber \\ &&\Rightarrow \lbrace X^r_{\phi},X^s_{\phi} \rbrace =0,\quad\quad \lbrace X_{\phi}^\tau ,X_{\phi}^r \rbrace =0. \label{IV2} \end{eqnarray} \noindent depending on two Lorentz scalar functions $F^{\tau}(q)$, $F(q)$, whose form will be restricted by the following requirements implying that $X^A _{\phi}$ and $P^A_{\phi}$ are canonical variables: \begin{equation} \begin{array}{cccc} \left\{ P_{\phi}^\tau ,X_{\phi}^\tau \right\} =1, & \left\{ P^r_{\phi},X_{\phi} ^s \right\} =-\delta ^{rs}, & \left\{ P_{\phi}^r,X_{\phi}^\tau \right\} =0, & \left\{ P_{\phi}^\tau ,X_{\phi}^r\right\} =0, \end{array} \label{IV3} \end{equation} Since $\left\{ P_{\phi}^\tau ,X_{\phi}^\tau \right\} =\int d\tilde q \omega^2 \left( q\right) F^\tau \left( q\right)$ and $\left\{ P_{\phi}^r,X_{\phi}^s\right\} =\int d\tilde q \, q^rq^s\, F\left( q \right) $, we must require the following normalizations for $F^{\tau }(q)$, $F(q)$ \begin{equation} \int d\tilde q \omega^2(q) F^\tau \left( q\right) =1 \label{IV4} \end{equation} \begin{equation} \int d\tilde q\, q^rq^s F\left( q\right) =-\delta ^{rs}. \label{IV5} \end{equation} Moreover, $\left\{ P_{\phi}^r,X_{\phi}^\tau \right\} =\int d\tilde q \omega (q) q^r F^\tau \left( q\right)$ and $\left\{ P_{\phi}^\tau ,X_{\phi} ^r\right\} =\int d\tilde q \omega (q)\, q^r F\left( q \right)$ , imply the following conditions \begin{equation} \int d\tilde q \omega (q)\, q^r F^\tau \left( q\right) =0, \label{IV6} \end{equation} \begin{equation} \int d\tilde q \omega (q)\, q^r F\left( q\right) =0. \label{IV7} \end{equation} \noindent which are automatically satisfied because $F^{\tau}(q)$, $F(q)$, $q=|\vec q|$, are even under $q^r\, \rightarrow -q^r$. A solution of Eqs.(\ref{IV4}) and (\ref{IV5}) is \begin{eqnarray} F^{\tau}(q) &=& {{16 \pi^2}\over {m\, q^2\, \sqrt{m^2+q^2} }} e^{-{{4\pi}\over {m^2}}\, q^2},\nonumber \\ &&{}\nonumber \\ F(q) &=& - {{48 \pi^2}\over {m\, q^4}} \sqrt{m^2+q^2}\, e^{-{{4\pi}\over {m^2}} q^2}. \label{IV8} \end{eqnarray} The singularity in $\vec q =0$ requires $\varphi (\tau ,\vec q)\, {\rightarrow}_{q\rightarrow 0}\, q^{\eta}$, $\eta > 0$, [and not $\eta > - \epsilon$ as in Eq.(\ref{IV1})] for the existence of $X^{\tau}_{\phi}$, ${\vec X}_{\phi}$. The analogue of the function $F(P_{\phi},k)$ of Ref.\cite{lon} is now \begin{eqnarray} {\cal F}\left( P_{\phi},\vec q\right) &=&F^\tau \left( q\right) \omega \left( q\right) P^\tau_{\phi} -F\left( q\right) \vec q\cdot {\vec P}_{\phi}, \nonumber \\ &&{}\nonumber \\ &&\int d\tilde q\, {\cal F}\left( P_{\phi},\vec q\right) \omega \left( q \right) =P_{\phi}^\tau , \nonumber \\ && \int d\tilde q\, {\cal F}\left( P_{\phi},\vec q\right) q^r=P_{\phi}^r. \label{IV9} \end{eqnarray} Let us remark that for field configurations $\phi (\tau ,\vec \sigma )$ such that the Fourier transform $\hat \phi (\tau ,\vec q)$ has compact support in a sphere centered at $\vec q=0$ of volume V, we get $X^{\tau}_{\phi}=-{1\over V} \int {{d^3q}\over {\omega (q)}} \varphi (\tau ,\vec q)$, ${\vec X}_{\phi}= {1\over V}\int d^3q {{3\vec q}\over {{\vec q}^2}} \varphi (\tau ,\vec q)$. \subsection{Auxiliary relative variables.} As in Ref.\cite{lon}, let us define an auxiliary relative action variable and an auxiliary relative phase variable \begin{equation} \hat I\left( \tau ,\vec q\right) =I\left( \tau ,\vec q\right) -F^\tau \left( q\right) P_{\phi}^\tau \omega \left( q\right) +F\left( q\right) \vec q\cdot {\vec P}_{\phi}, \label{IV10} \end{equation} \begin{equation} \hat \varphi \left( \tau ,\vec q\right) =\varphi \left( \tau ,\vec q\right) -\omega \left( q\right) X_{\phi}^\tau +\vec q\cdot {\vec X}_{\phi}. \label{IV11} \end{equation} The previous canonicity conditions on $F^{\tau}(q)$, $F(q)$, imply \begin{eqnarray} &&\int d\tilde q\, \omega \left( q\right) \hat I\left( \tau ,\vec q\right) =0, \nonumber \\ &&\int d\tilde q\, q^i\, \hat I\left( \tau ,\vec q\right) =0. \label{IV12} \end{eqnarray} \begin{eqnarray} &&\int d\tilde q F^\tau \left( \mathrm{q}\right) \omega \left( q\right) \hat \varphi \left( \tau ,\vec q\right) d\tilde q=0,\nonumber \\ &&\int d\tilde q F\left( q\right) q^i\, \hat \varphi \left( \tau ,\vec q\right)=0. \label{IV13} \end{eqnarray} These auxiliary variables have the following nonzero Poisson bracket \begin{equation} \left\{ \hat I\left( \tau ,\vec k\right) ,\hat \varphi \left( \tau ,\vec q \right) \right\} =\Delta \left( \vec k,\vec q\right) , \label{IV14} \end{equation} \noindent with \begin{equation} \Delta \left( \vec k,\vec q\right) =\Omega \left( k\right) \delta ^3\left( \vec k-\vec q\right) -F^\tau \left( k\right) \omega \left( k\right) \omega \left( q\right) +F\left( k\right) \vec k\cdot \vec q. \label{IV15} \end{equation} The distribution $\Delta \left( \vec k,\vec q\right) $ has the semigroup property \begin{equation} \int d\tilde q\Delta \left( \vec k,\vec q\right) \Delta \left( \vec q,\vec k^{\prime }\right) =\Delta \left( \vec k,\vec k^{\prime }\right) , \label{IV16} \end{equation} \noindent and satisfies the constraints: \begin{equation} \begin{array}{cc} \int d\tilde q\, \omega \left( q\right) \Delta \left( \vec q,\vec k\right) =0, & \int d\tilde q\, q^r\, \Delta \left( \vec q,\vec k\right)=0, \end{array} \label{IV17} \end{equation} \begin{equation} \begin{array}{cc} \int d\tilde q\, F^\tau \left( q\right) \omega \left( q\right) \Delta \left( \vec k,\vec q\right) =0, & \int d\tilde q\, q^r\, F\left( q\right) \Delta \left( \vec k,\vec q\right) =0. \end{array} \label{IV18} \end{equation} At this stage the canonical variables $I(\tau ,\vec q)$, $\varphi (\tau ,\vec q)$ for the Klein-Gordon field are replaced by the noncanonical set $X_{\phi}^\tau$ , $P_{\phi}^\tau$ , ${\vec X}_{\phi}$, ${\vec P}_{\phi}$, $\hat I(\tau ,\vec q)$, $\hat \varphi (\tau ,\vec q)$ with the Poisson brackets \begin{equation} \left\{ \begin{array}{l} \begin{array}{cccc} \left\{ P_{\phi}^\tau ,X_{\phi}^\tau \right\} =1, & \left\{ P_{\phi}^r, X_{\phi}^s\right\} =-\delta ^{rs}, & \left\{ P_{\phi}^r,X_{\phi}^\tau \right\} =0, & \left\{ P_{\phi}^\tau ,X_{\phi}^r\right\} =0, \end{array} \\ \\ \begin{array}{ccc} \left\{ X_{\phi}^r,X_{\phi}^s\right\} =0, & \left\{ X_{\phi}^\tau ,X_{\phi}^r \right\} =0, & \left\{ P^A_{\phi}, P^B_{\phi} \right\} =0,\quad A,B=(\tau , r), \end{array} \\ \\ \begin{array}{cccc} \left\{ P_{\phi}^\tau ,\hat I\left( \tau ,\vec q\right) \right\} =0, & \left\{ P_{\phi}^r,\hat I\left( \tau ,\vec q\right) \right\} =0, & \left\{ \hat I\left( \tau ,\vec q\right) ,X_{\phi}^\tau \right\} =0, & \left\{ \hat I\left( \tau ,\vec q\right) ,X_{\phi}^r\right\} =0, \end{array} \\ \\ \begin{array}{cccc} \left\{ X_{\phi}^r,\hat \varphi \left( \tau ,\vec q\right) \right\} =0, & \left\{ X_{\phi}^\tau ,\hat \varphi \left( \tau ,\vec q\right) \right\} =0, & \left\{ P^r_{\phi},\hat \varphi \left( \tau ,\vec q\right) \right\} =0, & \left\{ P_{\phi}^\tau ,\hat \varphi \left( \tau ,\vec q\right) \right\} =0, \end{array} \\ \\ \left\{ \hat I\left( \tau ,\vec k\right) ,\hat \varphi \left( \tau ,\vec q \right) \right\} =\Omega \left( k\right) \delta ^3\left( \vec k-\vec q\right) -F^\tau \left( k\right) \omega \left( k\right) \omega \left( q\right) +F\left( k\right) \vec k\cdot \vec q . \end{array} \right. \label{IV19} \end{equation} The generators of Lorentz group are already decomposed into two parts, the collective and the relative ones, each satisfying the Lorentz algebra and having vanishing mutual Poisson brackets \begin{equation} \left\{ \begin{array}{l} J^{rs}_{\phi}=L_{\phi}^{rs}+{\hat S}_{\phi}^{rs}, \\ \\ L_{\phi}^{rs}=X_{\phi}^rP_{\phi}^s-X_{\phi}^sP_{\phi}^r, \\ \\ {\hat S}_{\phi}^{rs}=\int d\tilde q\, \hat I\left( \tau ,\vec q\right) \left( q^r \frac \partial {\partial q^s}-q^s\frac \partial {\partial q^r}\right) \hat \varphi \left( \tau ,\vec q\right) , \end{array} \right. \label{IV20} \end{equation} \begin{equation} \left\{ \begin{array}{l} J^{\tau r}_{\phi}=L^{\tau r}_{\phi}+{\hat S}^{\tau r}_{\phi}, \\ \\ L_{\phi}^{\tau r}=[X_{\phi}^\tau -\tau ] P_{\phi}^r-X_{\phi}^r P_{\phi}^\tau , \\ \\ {\hat S}_{\phi}^{\tau r}=-\int d\tilde q\, \omega \left( q\right) \hat I\left( \tau ,\vec q\right) \frac \partial {\partial q^r}\hat \varphi \left( \tau ,\vec q\right) . \end{array} \right. \label{IV21} \end{equation} \subsection{Canonical relative variables.} We must now find the canonical relative variables hidden inside the auxiliary ones, which are not free but satisfy Eqs.(\ref{IV12}) and (\ref{IV13}). As in Ref.\cite{lon}, let us introduce the following differential operator [$\triangle_{LB}$ is the Laplace-Beltrami operator of the mass shell submanifold $H^1_3$ (see Appendix B and Ref.\cite{lon,lon1})] \begin{eqnarray} {\cal D}_{\vec q}&=&3-m^2 \triangle_{LB}=\nonumber \\ &=&3-m^2\left[ \sum\limits_{i=1}^3\left( \frac \partial {\partial q^i}\right) ^2+\frac 2{m^2}\sum\limits_{i=1}^3q^i\, \frac \partial {\partial q^i}+\frac 1{m^2}\left( \sum\limits_{i=1}^3q^i\, \frac \partial {\partial q^i}\right) ^2\right] , \label{IV22} \end{eqnarray} \noindent which is a scalar on the Wigner hyperplane since it is invariant under Wigner's rotations. Since $\omega \left( \mathrm{q}\right) $ and $\vec q$ are null modes of this operator\cite{lon}, we can put \begin{eqnarray} \hat I\left( \tau ,\vec q\right) &=&{\cal D}_{\vec q}{\bf H}( \tau ,\vec q) , \nonumber \\ {\bf H}\left( \tau ,\vec q\right) &=&\int d\tilde k\, {\cal G}\left( \vec q,\vec k\right) \hat I\left( \tau ,\vec k\right) , \label{IV23} \end{eqnarray} \noindent with ${\cal G}\left( \vec q,\vec k\right) $ being the Green function of ${\cal D}_{\vec q}$ [see Refs.\cite{lon,lon1} for its expression] \begin{equation} {\cal D}_{\vec q}{\cal G}\left( \vec q,\vec k\right) =\Omega \left( k \right) \delta ^3\left( \vec k-\vec q\right) . \label{IV24} \end{equation} Like in Ref.\cite{lon}, for each zero mode $f_o(\vec q)$ of ${\cal D}_{\vec q}$ [${\cal D}_{\vec q}\, f_o(\vec q)=0$] for which one has $| \int d\tilde q\, f_o(\vec q) \hat I(\tau ,\vec q) | < \infty$, we have by integration by parts \begin{eqnarray} \int d\tilde q && f_o(\vec q) \hat I(\tau ,\vec q) = \int d\tilde q\, f_o(\vec q) {\cal D}_{\vec q} {\bf H}(\tau ,\vec q)=\nonumber \\ &=&-{1\over {2(2\pi )^3}} \int d^3q\, {{\partial}\over {\partial q^r}} \Big( {{m^2\delta^{rs}+q^rq^s}\over {\omega (q)}} \Big[ f_o(\vec q) {{\partial}\over {\partial q^s}} {\bf H}(\tau ,\vec q)-{\bf H}(\tau ,\vec q) {{\partial}\over {\partial q^s}} f_o(\vec q)\Big] \Big) . \label{IV25} \end{eqnarray} The boundary conditions (ensuring finite Poincar\'e generators) \begin{eqnarray} {\bf H}(\tau ,\vec q)\, &{\rightarrow}_{q \rightarrow 0}\,& q^{-1+\epsilon}, \quad\quad \epsilon > 0,\nonumber \\ {\bf H}(\tau ,\vec q)\, &{\rightarrow}_{q \rightarrow \infty}\,& q^{-3-\sigma}, \quad\quad \sigma > 0, \label{IV26} \end{eqnarray} \noindent imply $\int d\tilde q\, f_o(\vec q) \hat I(\tau ,\vec q)=0$, or \begin{eqnarray} \int d\tilde q\, f_o(\vec q) I(\tau ,\vec q) &=& P^{\tau}_{\phi}\, \int d\tilde q\, \omega (q) f_o(\vec q) F^{\tau}(q)-\nonumber \\ &-&{\vec P}_{\phi}\, \cdot \, \int d\tilde q\, \vec q\, f_o(\vec q) F(q). \label{IV27} \end{eqnarray} We shall restrict ourselves to field configurations for which $I(\tau ,\vec q)\, {\rightarrow}_{q\rightarrow 0}\, q^{-3+\eta}$ with $\eta \in (0,1]$ to avoid the extra condition connected with the zero modes $f_o(\vec q)=v^{(o)}_{2,-3,lm}(\vec q)$ [see the $Q_{lm}$ of Appendix B]. Instead, for the zero modes $f_o(\vec q)= v^{(o)}_{1,-3,lm}(\vec q)= q^l\,\, {}_2F_1 ({{l-1}\over 2}, {{l+3}\over 2}; l+{3\over 2}; -q^2)\, Y_{lm}(\alpha ,\beta )$, [$\vec q=q (cos\, \alpha cos\, \beta, cos\, \alpha sin\, \beta , sin\, \alpha )$] of Ref.\cite{lon} we get from Eq.(\ref{IV27}): i) for l=0,1, the identities $P^{\tau}_{\phi}=P^{\tau}_{\phi}$, ${\vec P}_{\phi}={\vec P}_{\phi}$ [$v^{(o)}_{1,-3,00}$ and $v^{(o)}_{1,-3,1m}$ give $f_o(\vec q)=\omega (q), \vec q$]; ii) for $l \geq 2$ the conditions \begin{eqnarray} P_{lm}&=& \int d\tilde q\, v^{(o)}_{1,-3,lm}(\vec q) I(\tau ,\vec q) = \nonumber \\ &=& const. \int d\tilde q\, q^l\, {}_2F_1 ({{l-1}\over 2}, {{l+3}\over 2}; l+{3\over 2}; -q^2)\, Y_{lm}(\theta ,\varphi ) \nonumber \\ &&\int d^3\sigma \int d^3\sigma^{'} e^{i\vec q\cdot (\vec \sigma -{\vec \sigma} ^{'})} [\omega (q)\phi (\tau ,\vec \sigma )-i\pi (\tau ,\vec \sigma )][\omega (q)\phi (\tau ,{\vec \sigma}^{'})+i\pi(\tau ,{\vec \sigma}^{'})]=0. \label{IV28} \end{eqnarray} \noindent Since $\vec q\cdot (\vec \sigma -{\vec \sigma}^{'})=q cos\, \theta |\vec \sigma -{\vec \sigma}^{'}|$, one can check that $P_{l\, m\not= 0}$ is automatically zero. Moreover, in $P_{l0}$ the term $\omega (q) [\phi (\tau ,\vec \sigma )\pi(\tau ,{\vec \sigma}^{'})-\phi (\tau ,{\vec \sigma}^{'}) \pi (\tau ,\vec \sigma )]$ does not contribute for symmetry reasons. Therefore, the final restriction is \begin{eqnarray} P_{l0}&=&const. \int d\tilde q\, q^l\, {}_2F_1 ({{l-1}\over 2}, {{l+3}\over 2}; l+{3\over 2}; -q^2)\, Y_{l0}(\theta ,\varphi ) \nonumber \\ &&\int d^3\sigma \int d^3\sigma^{'} e^{i\vec q\cdot (\vec \sigma -{\vec \sigma} ^{'})} \Big[ (m^2+q^2)\phi (\tau ,\vec \sigma )\phi (\tau ,{\vec \sigma}^{'})+ \pi (\tau ,\vec \sigma )\pi (\tau ,{\vec \sigma}^{'})\Big] =0. \label{IV29} \end{eqnarray} These conditions on $\phi (\tau ,\vec \sigma )$ and $\pi (\tau ,\vec \sigma ) =\partial_{\tau} \phi (\tau ,\vec \sigma )$ identify the class of configurations of the Klein-Gordon field for which one can define the previous canonical transformation and for which there is no ambiguity in defining a unique realization of the Poincar\'e group (the BMS algebra degenerates in the Poincar\'e algebra in this case). We can satisfy the constraints on $\hat \varphi (\tau ,\vec q)$ with the definition [${\cal D}_{\vec q} \omega (q)={\cal D}_{\vec q} \vec q =0$] \begin{eqnarray} \hat \varphi (\tau ,\vec q) &=&\int d\tilde k\int d\tilde k^{\prime } {\bf K}(\tau ,\vec k) {\cal G}(\vec k,\vec k^{\prime}) \Delta (\vec k^{\prime}, \vec q) ,\nonumber \\ {\bf K}(\tau ,\vec q) &=&{\cal D}_{\vec q}\hat \varphi (\tau ,\vec q) = {\cal D}_{\vec q}\varphi (\tau ,\vec q),\nonumber \\ &&{}\nonumber \\ &&{\rightarrow}_{q\rightarrow \infty}\, q^{1-\epsilon},\quad \epsilon > 0, \quad\quad {\rightarrow}_{q\rightarrow 0}\, q^{\eta -2},\quad \eta > 0, \label{IV30} \end{eqnarray} \noindent which also implies \begin{equation} \left\{ \begin{array}{l} \begin{array}{cc} \left\{ {\bf H}\left( \tau ,\vec q\right) ,X_{\phi}^\tau \right\} =0, & \left\{ {\bf H}\left( \tau ,\vec q\right) ,P_{\phi}^\tau \right\} =0, \\ \left\{ {\bf H}\left( \tau ,\vec q\right) ,X_{\phi}^r\right\} =0, & \left\{ {\bf H} \left( \tau ,\vec q\right) ,P_{\phi}^r\right\} =0, \end{array} \\ \\ \begin{array}{cc} \left\{ {\bf K}\left( \tau ,\vec q\right) ,X_{\phi}^\tau \right\} =0, & \left\{ {\bf K}\left( \tau ,\vec q\right) ,P_{\phi}^\tau \right\} =0, \\ \left\{ {\bf K}\left( \tau ,\vec q\right) ,X_{\phi}^r\right\} =0, & \left\{ {\bf K} \left( \tau ,\vec q\right) ,P_{\phi}^r\right\} =0, \end{array} \\ \\ \left\{ {\bf H}\left( \tau ,\vec q\right) ,\mathbf{K}\left( \tau ,\vec q^{\prime }\right) \right\} =\Omega \left( q\right) \delta ^3\left( \vec q-\vec q^{\prime }\right) . \end{array} \right. \label{IV31} \end{equation} The final decomposition of Lorentz generators is \begin{equation} \left\{ \begin{array}{l} J_{\phi}^{rs}=L_{\phi}^{rs}+S_{\phi}^{rs}, \\ \\ L_{\phi}^{rs}=X_{\phi}^rP_{\phi}^s-X_{\phi}^sP_{\phi}^r, \\ \\ S_{\phi}^{rs}=\int d\tilde k{\bf H}\left( \tau ,\vec k\right) \left( k^r \frac \partial {\partial k^s}-k^s\frac \partial {\partial k^r}\right) {\bf K}\left( \tau ,\vec k\right) , \end{array} \right) \label{IV32} \end{equation} \begin{equation} \left\{ \begin{array}{l} J_{\phi}^{\tau r}=L_{\phi}^{\tau r}+S_{\phi}^{\tau r}, \\ \\ L_{\phi}^{\tau r}=(X_{\phi}^\tau -\tau )P_{\phi}^r-X_{\phi}^rP_{\phi}^\tau , \\ \\ S_{\phi}^{\tau r}=-\int d\tilde q\omega \left( q\right) {\bf H} \left( \tau ,\vec q\right) \frac \partial {\partial q^r}{\bf K} \left( \tau ,\vec q\right) . \end{array} \right) \label{IV33} \end{equation} \subsection{Field variables in terms of collective-relative variables.} We have found the canonical transformation \begin{eqnarray} I\left( \tau ,\vec q\right) &=& F^{\tau}(q)\omega (q) P^{\tau}_{\phi}-F(q)\vec q\cdot {\vec P}_{\phi}+ {\cal D}_{\vec q} {\bf H}(\tau ,\vec q),\nonumber \\ \varphi (\tau ,\vec q)&=&\int d\tilde k\int d\tilde k^{\prime } {\bf K} ( \tau ,\vec k) {\cal G}\left( \vec k,\vec k^{\prime }\right) \Delta \left( \vec k^{\prime },\vec q\right) +\omega \left( q\right) X_{\phi}^\tau -\vec q\cdot {\vec X}_{\phi},\nonumber \\ &&{}\nonumber \\ N_{\phi}&=&P^{\tau}_{\phi} \int d\tilde q\, \omega (q) F^{\tau}(q)-{\vec P} _{\phi}\cdot \,\, \int d\tilde q\, \vec q F(q) +\int d\tilde q\, {\cal D}_{\vec q} {\bf H}(\tau ,\vec q)=\nonumber \\ &=& \tilde c {{ P^{\tau}_{\phi}}\over m}+\int d\tilde q\, {\cal D}_{\vec q} {\bf H}(\tau ,\vec q),\nonumber \\ &&\tilde c = m \int d\tilde q \omega (q) F^{\tau}(q)=2\int_0^{\infty}{{dq}\over {\sqrt{m^2+q^2}}}e^{-{{4\pi}\over {m^2}}q^2}=2 e^{4\pi} \int^{\infty}_m {{dx}\over {\sqrt{x^2-m^2}}}e^{-{{4\pi}\over {m^2}}x^2},\nonumber \\ &&{} \label{IV34} \end{eqnarray} \noindent with the two functions $F^{\tau}(q)$, $F(q)$ given in Eqs.(\ref{IV8}). Its inverse is \begin{eqnarray} P^{\tau}_{\phi}&=&\int d\tilde q \omega (q) I(\tau ,\vec q)={1\over 2}\int d^3\sigma \Big[ \pi^2 +(\vec \partial \phi )^2+m^2\phi^2\Big] (\tau ,\vec \sigma ),\nonumber \\ {\vec P}_{\phi}&=&\int d\tilde q \vec q\, I(\tau ,\vec q)=\int d^3\sigma \Big[ \pi \vec \partial \phi \Big] (\tau ,\vec \sigma ),\nonumber \\ X^{\tau}_{\phi}&=&\int d\tilde q \omega (q) F^{\tau}(q) \varphi (\tau ,\vec q)= \tau +\nonumber \\ &+&{1\over {2\pi i m}} \int d^3q {{e^{-{{4\pi}\over {m^2}} q^2} }\over {q^2\, \omega (q)}} ln\, \Big[ {{\omega (q) \int d^3\sigma e^{i\vec q\cdot \vec \sigma} \phi (\tau ,\vec \sigma ) +i\int d^3\sigma e^{i\vec q\cdot \vec \sigma} \pi (\tau ,\vec \sigma )}\over {\omega (q) \int d^3\sigma e^{-i\vec q\cdot \vec \sigma} \phi (\tau ,\vec \sigma ) -i \int d^3\sigma e^{-i\vec q\cdot \vec \sigma} \pi (\tau ,\vec \sigma )}}\Big] =\nonumber \\ &{\buildrel {def} \over =}\,& \tau + {\tilde X}^{\tau}_{\phi},\quad\quad \Rightarrow \quad L^{\tau r}_{\phi}={\tilde X}^{\tau}_{\phi}P^r_{\phi}- X^r_{\phi}P^{\tau}_{\phi},\nonumber \\ &&{}\nonumber \\ {\vec X}_{\phi}&=&\int d\tilde q \vec q\, F(q) \varphi (\tau ,\vec q)= \nonumber \\ &=&{{2i}\over {\pi m}} \int d^3q\, {{q^i}\over {q^4}}\, e^{-{{4\pi}\over {m^2}} q^2} ln\, \Big[ {{\sqrt{m^2+q^2} \int d^3\sigma e^{i\vec q\cdot \vec \sigma} \phi (\tau ,\vec \sigma ) +i \int d^3\sigma e^{i\vec q\cdot \vec \sigma} \pi (\tau ,\vec \sigma )}\over {\sqrt{m^2+q^2} \int d^3\sigma e^{-i\vec q\cdot \vec \sigma} \phi (\tau ,\vec \sigma ) -i \int d^3\sigma e^{-i\vec q\cdot \vec \sigma} \pi (\tau ,\vec \sigma )}} \Big] ,\nonumber \\ &&{}\nonumber \\ {\bf H}(\tau ,\vec q)&=&\int d\tilde k {\cal G}(\vec q,\vec k) [I(\tau ,\vec k) -F^{\tau}(k)\omega (k) \int d{\tilde q}_1 \omega (q_1) I(\tau ,{\vec q}_1)+ \nonumber \\ &+&F(k) \vec k \cdot \, \int d{\tilde q}_1 {\vec q}_1\, I(\tau ,{\vec q}_1) ]= \nonumber \\ &=&\int d^3\sigma_1d^3\sigma_2 \Big[ \pi (\tau ,{\vec \sigma}_1) \pi (\tau ,{\vec \sigma}_2) \int d\tilde k {\cal G}(\vec q,\vec k) \int d{\tilde k}_1 \triangle (\vec k,{\vec k}_1)e^{i{\vec k}_1\cdot ({\vec \sigma}_1-{\vec \sigma}_2)}+\nonumber \\ &+&\phi (\tau ,{\vec \sigma}_1) \phi (\tau ,{\vec \sigma}_2) \int d\tilde k {\cal G}(\vec q,\vec k) \int d{\tilde k}_1 \omega^2(k_1)\triangle (\vec k,{\vec k}_1) e^{i{\vec k}_1\cdot ({\vec \sigma}_1-{\vec \sigma}_2)}-\nonumber \\ &-&i\Big( \pi (\tau ,{\vec \sigma}_1) \phi (\tau ,{\vec \sigma}_2)+\pi (\tau ,{\vec \sigma}_2) \phi (\tau ,{\vec \sigma}_1)\Big) \nonumber \\ && \int d\tilde k {\cal G}(\vec q,\vec k) \int d{\tilde k}_1 \omega (k_1) \triangle (\vec k.{\vec k}_1) e^{i{\vec k}_1\cdot ({\vec \sigma}_1-{\vec \sigma}_2)} \Big] ,\nonumber \\ &&{}\nonumber \\ {\bf K}(\tau ,\vec q)&=&{\cal D}_{\vec q} \hat \varphi (\tau ,\vec q)={\cal D} _{\vec q}\varphi (\tau ,\vec q)=\nonumber \\ &=&{1\over {2i}} {\cal D}_{\vec q} \ln \left[ \frac{\int d^3\sigma \, \left[ \omega \left( q\right) \phi \left( \tau ,\vec \sigma \right) +i\pi \left( \tau ,\vec \sigma \right) \right] e^{-i\vec q\cdot \vec \sigma }}{\int d^3\sigma ^{\prime }\, \left[ \omega \left( q\right) \phi \left( \tau ,\vec \sigma ^{\prime }\right) -i\pi \left( \tau ,\vec \sigma ^{\prime }\right) \right] e^{i\vec q\cdot \vec \sigma ^{\prime } }}\right] . \label{IV35} \end{eqnarray} \noindent We get the following expression of the other canonical variables $a(\tau ,\vec q)$, $\phi (\tau ,\vec \sigma )$, $\pi (\tau ,\vec \sigma )$, in terms of the final ones \begin{eqnarray} a(\tau ,\vec q)&=&\sqrt{F^{\tau}(q)\omega (q)P^{\tau} _{\phi}-F(q)\vec q\cdot {\vec P}_{\phi}+{\cal D}_{\vec q} {\bf H}(\tau , \vec q) }\nonumber \\ &&e^{i[\omega (q) X_{\phi}^\tau -\vec q\cdot {\vec X}_{\phi}] +i\int d\tilde k\int d\tilde k^{\prime }\, {\bf K}( \tau ,\vec k) {\cal G}\left( \vec k,\vec k^{\prime }\right) \Delta \left( \vec k^{\prime } ,\vec q\right) },\nonumber \\ &&{}\nonumber \\ N_{\phi}&=& \tilde c {{P^{\tau}_{\phi}}\over m}- +\int d\tilde k\, {\cal D}_{\vec k} {\bf H}(\tau ,\vec k), \label{IV36} \end{eqnarray} \begin{eqnarray} \phi (\tau ,\vec \sigma ) &=& \int d\tilde q\sqrt{F^{\tau}(q)\omega (q)P^{\tau} _{\phi}-F(q)\vec q\cdot {\vec P}_{\phi}+{\cal D}_{\vec q} {\bf H}(\tau , \vec q) }\nonumber \\ &&\left[ e^{-i[\omega (q) \left( \tau -X_{\phi}^\tau \right) - \vec q\cdot \left(\vec \sigma -{\vec X}_{\phi}\right)] +i\int d\tilde k\int d\tilde k^{\prime }\, {\bf K}( \tau ,\vec k) {\cal G}\left( \vec k,\vec k^{\prime }\right) \Delta \left( \vec k^{\prime } ,\vec q\right) }\right. + \nonumber \\ &&+\left. e^{i[\omega \left( q\right) \left( \tau -X_{\phi}^\tau \right) -\vec q\cdot \left( \vec \sigma -{\vec X}_{\phi}\right) ] -i\int d\tilde k\int d\tilde k^{\prime }\, {\bf K}\left( \tau ,\vec k\right) {\cal G}\left( \vec k,\vec k^{\prime }\right) \Delta \left( \vec k^{\prime },\vec q\right) }\right] =\nonumber \\ &=&2 \int d\tilde q {\bf A}_{\vec q}(\tau ;P^A_{\phi},{\bf H}]\, cos\, \Big[ \vec q\cdot \vec \sigma +{\bf B}_{\vec q}(\tau ;X^A_{\phi},{\bf K}] \Big] ,\nonumber \\ &&{}\nonumber \\ \pi (\tau ,\vec \sigma )&=&-i \int d\tilde q \omega (q)\sqrt{F^{\tau}(q) \omega (q)P^{\tau}_{\phi}-F(q)\vec q\cdot {\vec P}_{\phi}+{\cal D}_{\vec q} {\bf H}(\tau , \vec q) }\nonumber \\ &&\left[ e^{-i[\omega (q) \left( \tau -X_{\phi}^\tau \right)- \vec q\cdot \left(\vec \sigma -{\vec X}_{\phi}\right)] +i\int d\tilde k\int d\tilde k^{\prime }\, {\bf K}( \tau ,\vec k) {\cal G}\left( \vec k,\vec k^{\prime }\right) \Delta \left( \vec k^{\prime } ,\vec q\right) }\right. - \nonumber \\ &&-\left. e^{+i[\omega \left( q\right) \left( \tau -X_{\phi}^\tau \right) -\vec q\cdot \left( \vec \sigma -{\vec X}_{\phi}\right) ]-i\int d\tilde k\int d\tilde k^{\prime }\, {\bf K}\left( \tau ,\vec k\right) {\cal G}\left( \vec k,\vec k^{\prime }\right) \Delta \left( \vec k^{\prime },\vec q\right) }\right] =\nonumber \\ &=&-2\int d\tilde q \omega (q) {\bf A}_{\vec q}(\tau ;P^A_{\phi},{\bf H}]\, sin\, \Big[ \vec q\cdot \vec \sigma +{\bf B}_{\vec q}(\tau ;X^A_{\phi},{\bf K}] \Big] ,\nonumber \\ &&{}\nonumber \\ {\bf A}_{\vec q}(\tau ;P^A_{\phi},{\bf H}]&=&\sqrt{F^{\tau}(q)\omega (q) P^{\tau}_{\phi}-F(q)\vec q\cdot {\vec P}_{\phi}+{\cal D}_{\vec q}{\bf H}(\tau ,\vec q)}=\sqrt{I(\tau ,\vec q)},\nonumber \\ {\bf B}_{\vec q}(\tau ;X^A_{\phi},{\bf K}]&=&-\vec q\cdot {\vec X}_{\phi}- \omega (q)(\tau -X^{\tau}_{\phi})+\int d\tilde kd{\tilde k}^{'} {\bf K}(\tau ,\vec k){\cal G}(\vec k,{\vec k}^{'})\triangle ({\vec k}^{'},\vec q)= \nonumber \\ &=&\varphi (\tau ,\vec q)-\omega (q) \tau . \label{IV37} \end{eqnarray} The Klein-Gordon field configuration is described by:\hfill\break i) its energy $P^{\tau} _{\phi}$ and the conjugate field time-variable $X^{\tau}_{\phi}$, which is equal to $\tau$ plus some kind of internal time ${\tilde X}^{\tau}_{\phi}$; \hfill\break ii) the conjugate reduced canonical variables of a free point ${\vec X}_{\phi}$, ${\vec P}_{\phi}$;\hfill\break iii) an infinite set of canonically conjugate relative variables ${\bf H}(\tau ,\vec q)$, ${\bf K}(\tau ,\vec q)$. \hfill\break While the sets i) and ii) describe a ``monopole" field configuration see Section V), which depends only on 8 degrees of freedom like a scalar particle at rest [${\vec P}_{\phi}\approx 0$] and with mass $\epsilon_s\approx \sqrt{(P^{\tau}_{\phi})^2-{\vec P}_{\phi}^2}\approx P^{\tau}_{\phi}$, corresponding to the decoupled collective variables of the field configuration, the set iii) describes an infinite set of ``canonical relative variables" with respect to the relativistic collective variables of the sets i) and ii). The conditions ${\bf H}(\tau ,\vec q)={\bf K}(\tau ,\vec q)=0$ select the class of field configurations, solutions of the Klein-Gordon equation, which are of the ``monopole" type on the Wigner hyperplanes \begin{eqnarray} \phi_{mon}(\tau ,\vec \sigma )&=&2 \int d\tilde q \sqrt{F^{\tau}(q)\omega (q) P^{\tau}_{\phi}-F(q)\vec q\cdot {\vec P}_{\phi}} cos\, \Big[ \vec q\cdot (\vec \sigma -{\vec X}_{\phi})-\omega (q)(\tau -X^{\tau}_{\phi})\Big] \approx \nonumber \\ &\approx& 2 \sqrt{P^{\tau}_{\phi}} \int d\tilde q \sqrt{F^{\tau}(q)\omega (q)} cos\, \Big[ \vec q\cdot (\vec \sigma -{\vec X}_{\phi})-\omega (q)(\tau -X^{\tau}_{\phi})\Big] ,\nonumber \\ \pi_{mon}(\tau ,\vec \sigma )&=& -2 \int d\tilde q \sqrt{F^{\tau}(q)\omega (q)P^{\tau}_{\phi}-F(q)\vec q\cdot {\vec P}_{\phi}} sin\, \Big[ \vec q\cdot (\vec \sigma -{\vec X}_{\phi})-\omega (q)(\tau -X^{\tau}_{\phi})\Big] \approx \nonumber \\ &\approx& -2 \sqrt{P^{\tau}_{\phi}} \int d\tilde q \sqrt{F^{\tau}(q)\omega (q)} sin\, \Big[ \vec q\cdot (\vec \sigma -{\vec X}_{\phi})-\omega (q)(\tau -X^{\tau}_{\phi})\Big] . \label{IV38} \end{eqnarray} If we add the gauge-fixings ${\vec X}_{\phi}\approx 0$ to ${\vec P}_{\phi} \approx 0$ [this implies $\vec \lambda (\tau )=0$ in Eq.(\ref{II10})] and go to Dirac brackets, the rest-frame instant-form Klein-Gordon canonical variables in the gauge $\tau \equiv T_s=p_s\cdot x_s/ \epsilon_s$ (see the end of Section II) are [in the following formulas one has $T_s-X^{\tau}_{\phi}=-{\tilde X}^{\tau}_{\phi}$] \begin{eqnarray} a(T_s ,\vec q)&=&\sqrt{F^{\tau}(q)\omega (q)P^{\tau}_{\phi} +{\cal D}_{\vec q} {\bf H}(T_s , \vec q) }\nonumber \\ &&e^{i[\omega (q) {\tilde X}_{\phi}^\tau + \vec q\cdot \vec \sigma ]+ i\int d\tilde k\int d\tilde k^{\prime }\, {\bf K}(T_s ,\vec k) {\cal G}\left( \vec k,\vec k^{\prime }\right) \Delta \left( \vec k^{\prime } ,\vec q\right) },\nonumber \\ &&{}\nonumber \\ N_{\phi}&=&\tilde c {{P^{\tau}_{\phi}}\over m} +\int d\tilde q {\cal D}_{\vec q} {\bf H}(T_s,\vec q),\nonumber \\ &&{}\nonumber \\ \phi (T_s ,\vec \sigma ) &=& \int d\tilde q\sqrt{F^{\tau}(q)\omega (q)P^{\tau} _{\phi}+{\cal D}_{\vec q} {\bf H}(T_s , \vec q) } \nonumber \\ &&\left[ e^{i[\omega (q) {\tilde X}_{\phi}^\tau + \vec q\cdot \vec \sigma ]+ i\int d\tilde k\int d\tilde k^{\prime }\, {\bf K}( T_s ,\vec k) {\cal G}\left( \vec k,\vec k^{\prime }\right) \Delta \left( \vec k^{\prime } ,\vec q\right) }\right. + \nonumber \\ &&+\left. e^{-i[\omega \left( q\right) {\tilde X}_{\phi}^\tau +\vec q\cdot \vec \sigma ] -i\int d\tilde k\int d\tilde k^{\prime }\, {\bf K}\left( T_s ,\vec k\right) {\cal G}\left( \vec k,\vec k^{\prime }\right) \Delta \left( \vec k^{\prime },\vec q\right) }\right] =\nonumber \\ &=&2 \int d\tilde q\, {\bf A}_{\vec q}(T_s;P^{\tau} _{\phi},{\bf H}] \, cos\, \Big[ \vec q\cdot \vec \sigma +{\bf B}_{\vec q}(T_s; {\tilde X}^{\tau}_{\phi},{\bf K}]\, \Big] ,\nonumber \\ &&{}\nonumber \\ \pi (T_s ,\vec \sigma )&=&-i \int d\tilde q \omega (q)\sqrt{F^{\tau}(q) \omega (q)P^{\tau}_{\phi}+{\cal D}_{\vec q} {\bf H}(T_s , \vec q) }\nonumber \\ &&\left[ e^{i[\omega (q) {\tilde X}_{\phi}^\tau + \vec q\cdot\vec \sigma ] +i\int d\tilde k\int d\tilde k^{\prime }\, {\bf K}( T_s ,\vec k) {\cal G}\left( \vec k,\vec k^{\prime }\right) \Delta \left( \vec k^{\prime } ,\vec q\right) }\right. - \nonumber \\ &&-\left. e^{-i[\omega \left( q\right) {\tilde X}_{\phi}^\tau +\vec q\cdot \vec \sigma ] -i\int d\tilde k\int d\tilde k^{\prime }\, {\bf K}\left( T_s ,\vec k\right) {\cal G}\left( \vec k,\vec k^{\prime }\right) \Delta \left( \vec k^{\prime },\vec q\right) }\right] =\nonumber \\ &=&-2 \int d\tilde q\, \omega (q) {\bf A}_{\vec q}(T_s;P^{\tau} _{\phi},{\bf H}] \, sin\, \Big[ \vec q\cdot \vec \sigma +{\bf B}_{\vec q}(T_s; {\tilde X}^{\tau}_{\phi},{\bf K}]\, \Big] ,\nonumber \\ &&{}\nonumber \\ {\bf A}_{\vec q}(T_s;P^{\tau}_{\phi},{\bf H}]&=& \sqrt{ F^{\tau}(q) \omega (q) P^{\tau}_{\phi}+ {\cal D}_{\vec q} {\bf H}(T_s,\vec q)}= \sqrt{I(T_s,\vec q)},\nonumber \\ {\bf B}_{\vec q}(T_s;X^{\tau}_{\phi},{\bf K}]&=& \int d\tilde k d\tilde k^{'}\, {\bf K}(T_s,\vec k){\cal G}(\vec k,{\vec k}^{'}) \triangle ({\vec k}^{'},\vec q) +\omega (q) {\tilde X}^{\tau}_{\phi}=\nonumber \\ &=&\varphi (T_s,\vec q)-\omega (q) T_s,\nonumber \\ &&\Downarrow \nonumber \\ {{\partial \phi (T_s,\vec \sigma )}\over {\partial P^{\tau}_{\phi}}}&=& \int d\tilde q F^{\tau}(q)\omega (q) {{cos\, \Big[ \vec q\cdot \vec \sigma + {\bf B}_{\vec q}(T_s;{\tilde X}^{\tau}_{\phi},{\bf K}] \Big]}\over {{\bf A} _{\vec q}(T_s;P^{\tau}_{\phi},{\bf K}] }},\nonumber \\ {{\partial \phi (T_s,\vec \sigma )}\over {\partial X^{\tau}_{\phi}}}&=& 2\int d\tilde q \omega(q) {\bf A}_{\vec q}(T_s;P^{\tau}_{\phi},{\bf H}] sin\, \Big[ \vec q\cdot \vec \sigma + {\bf B}_{\vec q}(T_s;{\tilde X}^{\tau}_{\phi},{\bf K}] \Big],\nonumber \\ {{\delta \phi (T_s,\vec \sigma )}\over {\delta {\bf K}(T_s,\vec q)}}&=& 2 \int d\tilde k\, {\bf A}_{\vec k}(T_s;P^{\tau}_{\phi},{\bf H}]\nonumber \\ && sin\, \Big[ \vec k\cdot \vec \sigma +{\bf B}_{\vec k}(T_s;{\tilde X} ^{\tau}_{\phi},{\bf K}] \Big] \int d{\tilde k}^{'} {\cal G}(\vec q,{\vec k} ^{'})\triangle ({\vec k}^{'},\vec k),\nonumber \\ {{\delta \phi (T_s,\vec \sigma )}\over {\delta {\bf H}(T_s,\vec q)}}&=& \int d\tilde k \Big( {{cos\, \Big[ \vec k\cdot \vec \sigma + {\bf B}_{\vec k}(T_s;{\tilde X}^{\tau} _{\phi},{\bf K}] \Big]}\over {{\bf A}_{\vec k}(T_s;P^{\tau}_{\phi},{\bf H}]}} \Big) {\cal D}_{\vec k}\delta^3(\vec q-\vec k)=\nonumber \\ &=&-{\cal D}_{\vec q} \Big( {{cos\, \Big[ \vec q\cdot \vec \sigma + {\bf B}_{\vec q}(T_s;{\tilde X}^{\tau} _{\phi},{\bf K}] \Big]}\over {{\bf A}_{\vec q}(T_s;P^{\tau}_{\phi},{\bf H}]}} \Big) . \label{IV39} \end{eqnarray} From Eq.(\ref{II10}), the Hamiltonian is now $M_{\phi}=P^{\tau}_{\phi}$: it generates the following evolution in $T_s$ \begin{eqnarray} &&{{\partial}\over {\partial T_s}}\, X^{\tau}_{\phi}\, {\buildrel \circ \over =}\, \{ X^{\tau}_{\phi},P^{\tau}_{\phi} \} =-1,\quad\quad \Rightarrow \quad X^{\tau}_{\phi}\, {\buildrel \circ \over =}\, -T_s, \nonumber \\ &&{{\partial}\over {\partial T_s}}\, P^{\tau}_{\phi}\, {\buildrel \circ \over =}\, 0,\nonumber \\ &&{{\partial}\over {\partial T_s}}\, {\bf H}(T_s,\vec q)\, {\buildrel \circ \over =}\, 0,\nonumber \\ &&{{\partial}\over {\partial T_s}}\, {\bf K}(T_s,\vec q)\, {\buildrel \circ \over =}\, 0,\nonumber \\ &&\Rightarrow \,\, {{\partial}\over {\partial T_s}}{\bf A}_{\vec q}(T_s;P ^{\tau}_{\phi},{\bf H}]={{\partial}\over {\partial T_s}}{\bf B}_{\vec q}(T_s;{\tilde X}^{\tau}_{\phi},{\bf K}]\, {\buildrel \circ \over =}\, 0, \nonumber \\ &&{}\nonumber \\ &&{{\partial}\over {\partial T_s}}\, \phi (T_s,\vec \sigma )\, {\buildrel \circ \over =}\, -{{\partial}\over {\partial X^{\tau}_{\phi}}}\, \phi (T_s, \vec \sigma )=\pi (T_s,\vec \sigma ),\nonumber \\ &&{{\partial}\over {\partial T_s}}\, \pi (T_s,\vec \sigma )\, {\buildrel \circ \over =}\, -{{\partial}\over {\partial X^{\tau}_{\phi}}}\, \pi (T_s,\vec \sigma )=-[\triangle +m^2]\phi (T_s,\vec \sigma ),\nonumber \\ \Rightarrow&& ({{\partial^2}\over {\partial T_s^2}}-{{\partial^2}\over {\partial {\vec \sigma}^2}}+m^2) \phi (T_s,\vec \sigma )\, {\buildrel \circ \over =}\, 0. \label{IV40} \end{eqnarray} Therefore, in the free case ${\bf H}(T_s,\vec q)$, ${\bf K}(T_s,\vec q)$ are constants of the motion [complete integrability and Liouville theorem for the Klein-Gordon field]. Since the canonical variable $P^{\tau}_{\phi}$ is the Hamiltonian for the evolution in $T_s\equiv \tau$, we need the ``internal" variable $X^{\tau}_{\phi}=\tau +{\tilde X}^{\tau}_{\phi}$ [i.e. the ``internal time variable" ${\tilde X}^{\tau}_{\phi}$] to write Hamilton's equations ${{\partial}\over {\partial T_s}} F\, {\buildrel \circ \over =}\, \{ F,P^{\tau} _{\phi} \} =-{{\partial F}\over {\partial X^{\tau}_{\phi}}}= -{{\partial F}\over {\partial {\tilde X}^{\tau}_{\phi}}}$; in the free case we have ${{\partial}\over {\partial T_s}}\, {\buildrel \circ \over =}\, -{{\partial}\over {\partial X^{\tau}_{\phi}}}$ on $\phi (T_s,\vec \sigma )[X ^{\tau}_{\phi},P^{\tau}_{\phi},{\bf H},{\bf K}]$ and $\pi (T_s,\vec \sigma )[X ^{\tau}_{\phi},P^{\tau}_{\phi},{\bf H},{\bf K}]$, so that the evolution in the time $X^{\tau}_{\phi}=T_s+{\tilde X}^{\tau}_{\phi}$ [which takes place inside the Wigner hyperplane and which can be interpreted as an evolution in the internal time ${\tilde X}^{\tau}_{\phi}$] is equal and opposite to the evolution in the rest-frame time $T_s$ from a Wigner hyperplane to the next one in the free case. By adding the two second class constraints $X^{\tau}_{\phi}-T_s ={\tilde X} ^{\tau}_{\phi} \approx 0$, $P^{\tau}_{\phi}-const. \approx 0$, and by going to Dirac brackets, we get the rest-frame Hamilton-Jacobi formulation corresponding to the given constant value of the total energy [the field $\phi (T_s,\vec \sigma )$, which is $T_s$-independent since it depends only on the internal time ${\tilde X}^{\tau}_{\phi}$, now becomes also ${\tilde X}^{\tau}_{\phi}$- independent]. In this way we find a symplectic subspace (spanned by the canonical variables ${\bf H}$, ${\bf K}$) of each constant energy [$P^{\tau} _{\phi}=const.$] surface of the Klein-Gordon field. Each constant energy surface is not a symplectic manifold, but in this way it turns out to be the disjoint union (over ${\tilde X}^{\tau}_{\phi}$) of the symplectic manifolds determined by ${\tilde X}^{\tau}_{\phi}=const.$, $P^{\tau}_{\phi}=const.$ \subsection{A family of canonical multipoles.} From Eqs.(\ref{IV39}) we get [assuming that we can interchange the sums with the integrals] \begin{eqnarray} \phi (T_s,\vec \sigma )&=&2 \int d\tilde q {\bf A}_{\vec q}(T_s;P^{\tau}_{\phi}, {\bf H}] \Big( cos\, \vec q\cdot \vec \sigma \,\, cos\, {\bf B}_{\vec q}(T_s; {\tilde X}^{\tau}_{\phi},{\bf K}]-\nonumber \\ &-&sin\, \vec q\cdot \vec \sigma \,\, sin\, {\bf B} _{\vec q}(T_s;{\tilde X}^{\tau}_{\phi},{\bf K}] \Big) =\nonumber \\ &=& \sum_{k=0}^{\infty} (-)^k \Big[ {{\sigma^{r_1}...\sigma^{r_{2k+1}}}\over {(2k+1)!}} {\cal T}_{\phi ,O}^{r_1...r_{2k+1}}- {{\sigma^{r_1}...\sigma^{r_{2k}}}\over {(2k)!}} {\cal T}_{\phi ,E} ^{r_1..r_{2k}} \Big] ,\nonumber \\ &&{}\nonumber \\ \pi (T_s,\vec \sigma )&=&-2 \int d\tilde q\omega (q) {\bf A}_{\vec q}(T_s;P^{\tau}_{\phi}, {\bf H}] \Big( sin\, \vec q\cdot \vec \sigma \,\, cos\, {\bf B}_{\vec q}(T_s; {\tilde X}^{\tau}_{\phi},{\bf K}]+\nonumber \\ &+&cos\, \vec q\cdot \vec \sigma \,\, sin\, {\bf B} _{\vec q}(T_s;{\tilde X}^{\tau}_{\phi},{\bf K}] \Big) =\nonumber \\ &=& -\sum_{k=0}^{\infty} (-)^k \Big[ {{\sigma^{r_1}...\sigma^{r_{2k+1}}}\over {(2k+1)!}} {\cal T}_{\pi ,O}^{r_1...r_{2k+1}}- {{\sigma^{r_1}...\sigma^{r_{2k}}}\over {(2k)!}} {\cal T}_{\pi ,E} ^{r_1..r_{2k}} \Big] ,\nonumber \\ &&{}\nonumber \\ {\cal T}_{\phi ,O}^{r_1...r_{2k+1}}&=& 2\int d\tilde q\, q^{r_1}...q^{r_{2k+1}}\, {\bf A}_{\vec q}(T_s;P^{\tau}_{\phi},{\bf H}] cos\, {\bf B}_{\vec q}(T_s;{\tilde X}^{\tau}_{\phi},{\bf K}],\nonumber \\ {\cal T}_{\phi ,E}^{r_1..r_{2k}}&=& 2\int d\tilde q\, q^{r_1}...q^{r_{2k}}\, {\bf A}_{\vec q}(T_s;P^{\tau}_{\phi},{\bf H}] sin\, {\bf B}_{\vec q}(T_s;{\tilde X}^{\tau}_{\phi},{\bf K}],\nonumber \\ {\cal T}_{\pi ,O}^{r_1...r_{2k+1}}&=& 2\int d\tilde q\omega (q)\, q^{r_1}...q^{r_{2k+1}}\, {\bf A}_{\vec q}(T_s;P^{\tau}_{\phi},{\bf H}] sin\, {\bf B}_{\vec q}(T_s;{\tilde X}^{\tau}_{\phi},{\bf K}],\nonumber \\ {\cal T}_{\pi ,E}^{r_1..r_{2k}}&=& 2\int d\tilde q\omega (q)\, q^{r_1}...q^{r_{2k}}\, {\bf A}_{\vec q}(T_s;P^{\tau}_{\phi},{\bf H}] cos\, {\bf B}_{\vec q}(T_s;{\tilde X}^{\tau}_{\phi},{\bf K}] . \label{IV41} \end{eqnarray} In this way we have defined a set of canonical multipoles of the fields $\phi (T_s,\vec \sigma )$, $\pi (T_s,\vec \sigma )$, with respect to $\vec \sigma =0$ , i.e. with respect to the origin $x^{\mu}_s(\tau )$ of the Wigner hyperplanes (a noncanonical covariant centroid) and they are expressed in terms of the final canonical basis for the Klein-Gordon field. When the fields have a compact support W in momentum space, it can be shown that the only nonvanishing Poisson brackets of these multipoles are \begin{eqnarray} \{ {\cal T}_{\pi ,O}^{r_1...r_{2k+1}},{\cal T}_{\phi ,O}^{s_1...s_{2h+1}}\} &=& {\cal I}^{r_1...r_{2k+1},s_1...s_{2h+1}},\nonumber \\ \{ {\cal T}_{\pi ,E}^{r_1...r_{2k}},{\cal T}_{\phi ,E}^{s_1...s_{2h}}\} &=&- {\cal I}^{r_1...r_{2k},s_1...s_{2h}},\nonumber \\ &&{}\nonumber \\ {\cal I}^{r_1...r_m,s_1...s_n}&=&2 \int_Wd\tilde q \omega^2(q) q^{r_1}\cdots q^{r_m}q^{s_1}\cdots q^{s_n}. \label{IV42} \end{eqnarray} That is they form a closed algebra with a generalized Kronecker symbol, which could be quantized instead of the Fourier coefficients. \subsection{Effects of self interactions $V(\phi )$.} When there is an interaction term $V(\phi )$, the Hamiltonian becomes $M_{\phi}= {\tilde P}^{\tau}_{\phi}=P^{\tau}_{\phi}+\int d^3\sigma V(\phi )(T_s,\vec \sigma )$ and one has [see the difference between the rest-frame time evolution and the internal time one in presence of interactions] \begin{eqnarray} {{\partial}\over {\partial T_s}}\, X^{\tau}_{\phi}\, &{\buildrel \circ \over =}\,& -1-{{\partial}\over {\partial P^{\tau}_{\phi}}} \int d^3\sigma V(\phi (T_s,\vec \sigma ))=\nonumber \\ &=&-1-\int d\tilde k {{F^{\tau}(k)\omega (k)}\over { {\bf A}_{\vec k}(T_s;P ^{\tau}_{\phi},{\bf H}]}} \int d^3\sigma V^{'}(\phi (T_s,\vec \sigma )) cos\, [\vec k\cdot \vec \sigma +{\bf B}_{\vec k}(T_s;X^{\tau}_{\phi},{\bf K}] \, ],\nonumber \\ {{\partial}\over {\partial T_s}}\, {\tilde P}^{\tau}_{\phi}\, &{\buildrel \circ \over =}\,& 0,\nonumber \\ {{\partial}\over {\partial T_s}}\, P^{\tau}_{\phi}\, &{\buildrel \circ \over =}\,& {{\partial}\over {\partial X^{\tau}_{\phi}}}\, \int d^3\sigma V(\phi (T_s ,\vec \sigma ))=\nonumber \\ &=&-2 \int d\tilde k \omega (k) {\bf A}_{\vec k}(T_s;P^{\tau}_{\phi},{\bf H}] \int d^3\sigma V^{'}(\phi (T_s,\vec \sigma )) sin\, [\vec k\cdot \vec \sigma + {\bf B}(T_s;X^{\tau}_{\phi},{\bf K}]\, ],\nonumber \\ {{\partial}\over {\partial T_s}}\, {\bf H}(T_s,\vec q)\, &{\buildrel \circ \over =}\,& {{\delta}\over {\delta {\bf K}(T_s,\vec q)}} \int d^3\sigma V(\phi (T_s ,\vec \sigma ))=\nonumber \\ &=&-2 \int d^3\sigma V^{'}(\phi (T_s,\vec \sigma )) \int d\tilde k\, {\bf A} _{\vec k}(T_s;P^{\tau}_{\phi},{\bf H}]\nonumber \\ &&sin\, [\vec k\cdot \vec \sigma +{\bf B}_{\vec k}(T_s;X^{\tau}_{\phi}, {\bf K}]\, ] \int d\tilde k^{'}\, {\cal G}(\vec q,{\vec k}^{'})\triangle ({\vec k}^{'},\vec k) ,\nonumber \\ {{\partial}\over {\partial T_s}}\, {\bf K}(T_s,\vec q)\, &{\buildrel \circ \over =}\,& -{{\delta}\over {\delta {\bf H}(T_s,\vec q)}} \int d^3\sigma V(\phi (T_s,\vec \sigma ))=\nonumber \\ &=&-\int d^3\sigma V^{'}(\phi (T_s,\vec \sigma ))\, {\cal D}_{\vec q} \Big( {{cos\, \Big[ \vec q\cdot \vec \sigma +{\bf B}_{\vec q}(T_s;X^{\tau} _{\phi},{\bf K}] \Big]}\over {{\bf A}_{\vec q}(T_s;P^{\tau}_{\phi},{\bf H}]}} \Big), \label{IV43} \end{eqnarray} \noindent with $V^{'}(\phi )=\partial V(\phi )/\partial \phi$. In particular we get \begin{eqnarray} {{\partial}\over {\partial T_s}}{\bf A}_{\vec q}(T_s;P^{\tau}_{\phi},{\bf H}] &=&{1\over {2{\bf A}_{\vec q}(T_s;P^{\tau}_{\phi},{\bf H}]}} \Big[ F^{\tau}(q) \omega (q) {{\partial P^{\tau}_{\phi}}\over {\partial T_s}}+{\cal D}_{\vec q} {{\partial {\bf H}(T_s,\vec q)}\over {\partial T_s}} \Big] \, {\buildrel \circ \over =}\nonumber \\ &{\buildrel \circ \over =}\,& -{1\over {{\bf A}_{\vec q}(T_s;P^{\tau}_{\phi}, {\bf H}]}} \int d^3\sigma V^{'}(\phi (T_s,\vec \sigma )) \int d\tilde k \Big[ F^{\tau}(q)\omega (q)\omega (k)+\triangle (\vec q, \vec k)\Big] \nonumber \\ && {\bf A}_{\vec k}(T_s;P^{\tau}_{\phi},{\bf H}] sin\, \Big[ \vec k\cdot \vec \sigma +{\bf B}_{\vec k}(T_s;X^{\tau}_{\phi},{\bf K}]\Big] = \nonumber \\ &=&-{1\over {{\bf A}_{\vec q}(T_s;P^{\tau}_{\phi},{\bf H}]}} \int d\tilde k {\bf A}_{\vec k}(T_s;P^{\tau}_{\phi},{\bf H}] \int d^3\sigma V^{'}(\phi (T_s,\vec \sigma ))\nonumber \\ && sin\, \Big[ \vec k\cdot \vec \sigma +{\bf B}_{\vec k}(T_s;X^{\tau}_{\phi},{\bf K}]\Big] \Big[ \Omega (k)\delta^3(\vec q-\vec k) +F(q) \vec q\cdot \vec k\Big] ,\nonumber \\ {{\partial}\over {\partial T_s}}{\bf B}_{\vec q}(T_s;X^{\tau}_{\phi},{\bf K}] &=&-\omega (q)\Big( 1-{{\partial X^{\tau}_{\phi}}\over {\partial T_s}}\Big) + \int d\tilde k d{\tilde k}^{'} {{\partial {\bf K}(T_s,\vec k)}\over {\partial T_s}} {\cal G}(\vec k,{\vec k}^{'}) \triangle ({\vec k}^{'},\vec q)\, {\buildrel \circ \over =}\nonumber \\ &{\buildrel \circ \over =}\,& -2\omega (q)-\int d^3\sigma V^{'}(\phi (T_s,\vec \sigma ))\nonumber \\ &&\int d\tilde k \Big[ \omega (q)F^{\tau}(k)\omega (k)+\triangle (\vec k,\vec q) \Big] {{ cos\, \Big[ \vec k\cdot \vec \sigma +{\bf B}_{\vec k}(T_s;X^{\tau} _{\phi},{\bf K}]\Big] }\over {{\bf A}_{\vec k}(T_s;P^{\tau}_{\phi},{\bf H}]}} =\nonumber \\ &=&-2\omega (q)-\int d\tilde k {{\Omega (k)\delta^3(\vec q-\vec k)+F(k)\vec q\cdot \vec k}\over {{\bf A}(T_s;P^{\tau}_{\phi},{\bf H}]}}\nonumber \\ && \int d^3\sigma V^{'}(\phi (T_s,\vec \sigma )) cos\, \Big[ \vec k\cdot \vec \sigma +{\bf B} _{\vec k}(T_s;X^{\tau}_{\phi},{\bf K}]\Big] . \label{IV44} \end{eqnarray} For $V(\phi )={1\over {n+1}} \xi \phi^{n+1}$ we have:\hfill\break \hfill\break $V^{'}(\phi (T_s,\vec \sigma ))= \xi \phi^n(T_s,\vec \sigma )= 2^n \xi \int d{\tilde q}_1..d{\tilde q}_n {\bf A}_{{\vec q}_1}(T_s;P^{\tau}_{\phi},{\bf H}]....{\bf A}_{{\vec q}_n}(T_s; P^{\tau}_{\phi},{\bf H}] cos\, \Big( {\vec q}_1\cdot \vec \sigma +{\bf B}_{{\vec q}_1}(T_s;{\tilde X}^{\tau}_{\phi},{\bf K}]\Big) ... cos\, \Big( {\vec q}_n \cdot \vec \sigma +{\bf B}_{{\vec q}_n}(T_s;{\tilde X}^{\tau}_{\phi},{\bf K}] \Big)$. For the sine-Gordon case we have $V(\phi )={{\mu^4}\over {\lambda}}[cos\, ({{\sqrt{\lambda}}\over {\mu}}-1]$ and $V^{'}(\phi (T_s,\vec \sigma ))=$ \hfill\break \hfill\break ${{\mu^3}\over {\sqrt{\lambda}}} sin\, \Big({{\sqrt{\lambda}}\over {\mu}} \phi (T_s,\vec \sigma )\Big) ={{\mu^3}\over {\sqrt{\lambda}}} sin\, \Big( 2{{\sqrt{\lambda}}\over {\mu}} \int d\tilde q {\bf A}_{\vec q}(T_s;P ^{\tau}_{\phi},{\bf H}] cos\, \Big[ \vec q\cdot \vec \sigma +{\bf B}_{\vec q}(T_s;{\tilde X}^{\tau}_{\phi},{\bf K}]\Big] \Big)$. With a completely integrable interaction, it should exist a new canonical basis ${\hat {\tilde X}}^{\tau}_{\phi}$, ${\hat P}^{\tau}_{\phi}$, ${\hat {\bf H}}(T_s,\vec q)$, ${\hat {\bf K}}(T_s,\vec q)$ with ${\hat {\tilde X}} ^{\tau}_{\phi}$ the real ``internal" time like in the free case. For instance, in the new canonical basis of Eqs.(\ref{IV35}) one can define a family of completely integrable interactions associated with the Hamiltonian $M_{\phi}=P^{\tau} _{\phi}+\int d\tilde q [terms\, quadratic\, in\, {\bf H}(T_s,\vec q)\, and\, {\bf K}(T_s,\vec q)]$. However, their associated Lagrangian density as functionals of $\phi (\tau ,\vec \sigma )$ would be completely non local first because the canonical transformation $\phi , \pi \mapsto X^A_{\phi}, P^A_{\phi}, {\bf H}, {\bf K}$ is non local, and second because one would have to solve an integral equation to get the momenta $\pi (\tau ,\vec \sigma )$ in terms of the velocities $\dot \phi (\tau ,\vec \sigma )$ [${{\partial}\over {\partial T_s}} \phi (T_s,\vec \sigma )\, {\buildrel \circ \over =}\, \{ \phi (T_s,\vec \sigma ),M_{\phi} \}$]. \vfill\eject \section{The energy-momentum tensor on Wigner hyperplanes, Dixon's multipoles and the center of mass of a field configuration.} Let us now look at other properties of the Klein-Gordon field on the Wigner hyperplanes. In particular we are interested in identifying which kind of collective variables describe the center of mass of a field configuration and which is their relation, if any, with the previous collective variables. In so doing, we shall consider a field configuration as a relativistic extended body and we shall study its Dixon multipoles\cite{dixon}. The Euler-Lagrange equations from the action (\ref{II1}) are \begin{eqnarray} &&\Big( {{\partial {\cal L}}\over {\partial z^{\mu}}}-\partial_A {{\partial {\cal L}}\over {\partial z^{\mu}_A}}\Big) (\tau ,\vec \sigma )=\eta_{\mu\nu}\partial_A[\sqrt{g} T^{AB}[\phi ] z_B^{\nu}](\tau ,\vec \sigma )\, {\buildrel \circ \over =}\, 0,\nonumber \\ &&\Big( {{\partial {\cal L}}\over {\partial \phi}}-\partial_A{{\partial {\cal L}}\over {\partial \partial_A\phi}}\Big) (\tau ,\vec \sigma )\, {\buildrel \circ \over =}\, 0, \label{V00} \end{eqnarray} \noindent where we introduced the energy-momentum tensor \begin{eqnarray} T^{AB}(\tau ,\vec \sigma )[\phi ]&=&-\Big[ {2\over {\sqrt{g}}}{{\delta S}\over {\delta g_{AB}}}\Big] (\tau ,\vec \sigma )=\nonumber \\ &=&[\partial^A\phi \partial^B\phi -{1\over 2}g^{AB}(\pi^2-(\vec \partial \phi )^2-m^2\phi^2)](\tau ,\vec \sigma ). \label{V0} \end{eqnarray} When $\partial_A[\sqrt{g} z^{\mu}_B]=0$, as it happens on the Wigner hyperplanes in the gauge $T_s-\tau \approx 0$, $\vec \lambda (\tau )=0$, we get the conservation of the energy-momentum tensor $T^{AB}$, i.e. $\partial_AT^{AB}\, {\buildrel \circ \over =}\, 0$. Otherwise, there is compensation coming from the dynamics of the surface. The conserved, manifestly Lorentz covariant, energy-momentum tensor of the Klein-Gordon field $T^{\mu\nu}(x)[\tilde \phi ]=-{1\over 2} \eta^{\mu\nu}[\partial_{\alpha}\tilde \phi (x)\partial^{\alpha}\tilde \phi (x) -m^2{\tilde \phi}^2(x)]+\partial^{\mu}\tilde \phi (x)\partial^{\nu}\tilde \phi (x)$ becomes $T^{AB}(\tau ,\vec \sigma )[\phi ]$ in coordinates adapted to the hypersurface $\Sigma_{\tau}$ [$\sigma^A=(\tau ,\vec \sigma )$, $q^A=(\omega (q),\vec q)$], where \begin{eqnarray} T^{AB}(\tau ,\vec \sigma )[\phi ]&=&z^A_{\mu}(\tau ,\vec \sigma )z^B_{\nu}(\tau ,\vec \sigma )T^{\mu\nu}(z(\tau ,\vec \sigma ))[\tilde \phi ]=\nonumber \\ &=&z^A_{\mu}(\tau ,\vec \sigma )z^B_{\nu}(\tau ,\vec \sigma )T^{\mu\nu}(\tau ,\vec \sigma )[\phi =\tilde \phi \circ z]= \nonumber \\ &=&-{1\over 2}g^{AB}(\tau ,\vec \sigma )\Big[ \pi^2 - (\vec \partial \phi )^2-m^2\phi^2 \Big] (\tau ,\vec \sigma )+\partial^A\phi (\tau ,\vec \sigma ) \partial^B\phi (\tau ,\vec \sigma ). \label{V1} \end{eqnarray} By using the results in Ref.\cite{lus}, we have the following expression for the energy-momentum tensor 1) On arbitrary spacelike hypersurfaces we get \begin{eqnarray} T^{\mu\nu}(\tau ,\vec \sigma )[\phi ]&=&-{1\over 2}\eta^{\mu\nu}\Big[ \partial _A\phi (\tau ,\vec \sigma )\partial^A\phi (\tau ,\vec \sigma )-m^2\phi^2(\tau ,\vec \sigma )\Big]+\nonumber \\ &+&z^{\mu}_A(\tau ,\vec \sigma )z^{\nu}_B(\tau ,\vec \sigma ) \partial^A\phi (\tau ,\vec \sigma )\partial^B\phi (\tau ,\vec \sigma )= \nonumber \\ &=&\Big[ -{1\over 2}\eta^{\mu\nu} g^{\tau\tau}+(z^{\mu}_{\tau}g^{\tau\tau}+ z^{\mu}_ug^{u\tau})(z^{\nu}_{\tau}g^{\tau\tau}+z^{\nu}_vg^{v\tau})\Big] (\tau ,\vec \sigma ) \pi^2(\tau ,\vec \sigma )+\nonumber \\ &+&\Big[ -\eta^{\mu\nu}g^{\tau r}+(z^{\mu}_{\tau}g^{\tau\tau}+z^{\mu}_ug ^{u\tau})(z^{\nu}_{\tau}g^{\tau r}+z^{\nu}_vg^{vr})+\nonumber \\ &+&(z^{\mu}_{\tau}g ^{\tau r}+z^{\mu}_ug^{ur})(z^{\nu}_{\tau}g^{\tau\tau}+z^{\nu}_vg^{v\tau})\Big] (\tau ,\vec \sigma ) [\pi \, \partial_r\phi ](\tau ,\vec \sigma )+\nonumber \\ &+&\Big[-{1\over 2}\eta^{\mu\nu}g^{rs}+(z^{\mu}_{\tau}g^{\tau r}+z^{\mu}_ug ^{ur})(z^{\nu}_{\tau}g^{\tau s}+z^{\nu}_vg^{vs})\Big] (\tau ,\vec \sigma ) [\partial_r\phi \partial_s\phi ](\tau ,\vec \sigma )+\nonumber \\ &+&{1\over 2}\eta^{\mu\nu} m^2 \phi^2(\tau ,\vec \sigma ). \label{V2} \end{eqnarray} 2) On arbitrary spacelike hyperplanes, where \begin{eqnarray} z^{\mu}(\tau ,\vec \sigma )&=&x^{\mu}_s(\tau )+b^{\mu}_u(\tau )\sigma^u,\nonumber \\ z^{\mu}_r(\tau ,\vec \sigma )&=&b^{\mu}_r(\tau ),\quad\quad z^{\mu}_{\tau}(\tau ,\vec \sigma )={\dot x}_s^{\mu}(\tau )+{\dot b}^{\mu}_u (\tau )\sigma^u=l^{\mu}/\sqrt{g^{\tau\tau}}-g_{\tau r}z^{\mu}_r,\nonumber \\ &&{}\nonumber \\ g_{\tau \tau}&=& {\dot x}_s\cdot {\dot b}_r\sigma^r]^2,\quad\quad g_{\tau r}=b_{r\mu}[{\dot x}_s^{\mu}+ {\dot b}^{\mu}_s\sigma^s], \nonumber \\ g_{rs}&=&-\delta_{rs},\quad\quad \gamma^{rs}=-\delta^{rs},\quad\quad \gamma =1,\nonumber \\ g&=&g_{\tau\tau}+\sum_rg^2_{\tau r}, \nonumber \\ g^{\tau\tau}&=&1/[l_{\mu}({\dot x}_s^{\mu}+{\dot b}^{\mu}_u\sigma^u)]^2, \quad\quad g^{\tau r} =g^{\tau\tau}g_{\tau r}=b_{r\mu}({\dot x}_s^{\mu}+{\dot b}^{\mu}_u\sigma^u)/ [l_{\mu}({\dot x}_s^{\mu}+{\dot b}^{\mu}_u\sigma^u)]^2,\nonumber \\ g^{rs}&=&-\delta^{rs} +g^{\tau\tau}g_{\tau r}g_{\tau s}=-\delta^{rs}+ b_{r\mu}({\dot x}_s^{\mu}+{\dot b}^{\mu}_u\sigma^u) b_{r\nu}({\dot x}_s^{\nu}+{\dot b}^{\nu}_v\sigma^v) /[l_{\mu}({\dot x}_s^{\mu}+{\dot b}^{\mu}_u\sigma^u)]^2, \label{V3} \end{eqnarray} \noindent there is no relevant simplification for $T^{\mu\nu}(x^{\beta}_s(\tau )+b^{\beta}_u(\tau )\sigma ^u)[\phi]$. 3) On Wigner hyperplanes, where \begin{eqnarray} z^{\mu}(\tau ,\vec \sigma )&=&x^{\mu}_s(\tau )+\epsilon^{\mu}_u(u(p_s))\sigma^u,\nonumber \\ &&{}\nonumber \\ z^{\mu}_r&=&\epsilon^{\mu}_r(u(p_s),\quad\quad l^{\mu}=u^{\mu}(p_s),\quad\quad z^{\mu}_{\tau}={\dot x}^{\mu}_s(\tau ),\nonumber \\ g&=&[{\dot x}_s(\tau )\cdot u(p_s)],\quad\quad g_{\tau\tau}={\dot x}^2_s, \quad\quad g_{\tau r}={\dot x}_{s\mu} \epsilon^{\mu}_r(u(p_s)),\quad\quad g_{rs}=-\delta_{rs},\nonumber \\ g^{\tau\tau}&=&1/[{\dot x}_{s\mu}u^{\mu}(p_s)]^2,\quad\quad g^{\tau r}={\dot x}_{s\mu}\epsilon^{\mu} _r(u(p_s))/[{\dot x}_{s\mu}u^{\mu}(p_s)]^2,\nonumber \\ g^{rs}&=&-\delta^{rs}+{\dot x} _{s\mu}\epsilon^{\mu}_r(u(p_s)) {\dot x}_{s\nu}\epsilon^{\nu}_s(u(p_s)) /[{\dot x}_{s\mu}u^{\mu}(p_s)]^2, \label{V4} \end{eqnarray} \noindent we get \begin{eqnarray} T^{\mu\nu}[x^{\mu}_s(\tau )&+&\epsilon^{\mu}_u(u(p_s))\sigma^u][\phi ]= {1\over {[{\dot x}_s(\tau )\cdot u(p_s)]^2}}\Big[ -{1\over 2}\eta^{\mu\nu}+ \nonumber \\ &+&[{\dot x}^{\mu}_s(\tau )+{\dot x}_{s\rho}(\tau )\epsilon^{\rho}_u(u(p_s)) \epsilon^{\mu}_u(u(p_s))][{\dot x}^{\nu}_s(\tau )+{\dot x}_{s\sigma}(\tau ) \epsilon^{\sigma}_u(u(p_s))\epsilon^{\nu}_u(u(p_s))]\Big] \pi^2(\tau ,\vec \sigma )+\nonumber \\ &+&{1\over {[{\dot x}_s(\tau )\cdot u(p_s)]^2}} \Big[ {1\over 2}\eta^{\mu\nu}\Big( [{\dot x}_s(\tau )\cdot u(p_s)]^2\delta _{rs}-{\dot x}_{s\alpha}(\tau )\epsilon^{\alpha}_r(u(p_s)){\dot x} _{s\beta}(\tau )\epsilon^{\beta}_s(u(p_s))\big)+\nonumber \\ &+&[{\dot x}_s(\tau )\cdot u(p_s)]^2\Big( \epsilon^{\mu}_r(u(p_s))- {{{\dot x}_{s\beta}(\tau )\epsilon^{\beta}_r(u(p_s))}\over {[{\dot x}_s(\tau ) \cdot u(p_s)]^2}}[{\dot x}^{\mu}_s(\tau )+{\dot x}_{s\alpha}(\tau ) \epsilon^{\alpha}_u(u(p_s))\epsilon^{\mu}_u(u(p_s))]\Big)\nonumber \\ &&\Big( \epsilon^{\nu}_r(u(p_s))- {{{\dot x}_{s\beta}(\tau )\epsilon^{\beta}_r(u(p_s))}\over {[{\dot x}_s(\tau ) \cdot u(p_s)]^2}}[{\dot x}^{\nu}_s(\tau )+{\dot x}_{s\alpha}(\tau ) \epsilon^{\alpha}_v(u(p_s))\epsilon^{\nu}_v(u(p_s))]\Big) \Big] \nonumber \\ && [\partial_r\phi \partial_s\phi ](\tau ,\vec \sigma )+ {1\over 2}\eta^{\mu\nu} m^2 \phi^2(\tau ,\vec \sigma )+\nonumber \\ &+&{1\over {[{\dot x}_s(\tau )\cdot u(p_s)]^2}} \Big[ -\eta^{\mu\nu}{\dot x} _{s\rho}(\tau )\epsilon^{\rho}_r(u(p_s))-\nonumber \\ &-&[{\dot x}^{\mu}_s(\tau )+{\dot x}_{s\rho}(\tau )\epsilon^{\rho}_u(u(p_s)) \epsilon^{\mu}_u(u(p_s))]\nonumber \\ &&\Big( \epsilon^{\nu}_r(u(p_s))- {{{\dot x}_{s\beta}(\tau )\epsilon^{\beta}_r(u(p_s))}\over {[{\dot x}_s(\tau ) \cdot u(p_s)]^2}}[{\dot x}^{\nu}_s(\tau )+{\dot x}_{s\alpha}(\tau ) \epsilon^{\alpha}_v(u(p_s))\epsilon^{\nu}_v(u(p_s))]\Big)-\nonumber \\ &-&\Big( \epsilon^{\mu}_r(u(p_s))- {{{\dot x}_{s\beta}(\tau )\epsilon^{\beta}_r(u(p_s))}\over {[{\dot x}_s(\tau ) \cdot u(p_s)]^2}}[{\dot x}^{\mu}_s(\tau )+{\dot x}_{s\alpha}(\tau ) \epsilon^{\alpha}_u(u(p_s))\epsilon^{\mu}_u(u(p_s))]\Big) \nonumber \\ &&[{\dot x}^{\nu}_s(\tau )+{\dot x}_{s\rho}(\tau )\epsilon^{\rho}_v(u(p_s)) \epsilon^{\nu}_v(u(p_s))] [\pi \partial_r\phi ](\tau ,\vec \sigma ) . \label{V5} \end{eqnarray} Since we have \begin{eqnarray} {\dot x}_s^{\mu}(\tau )&=&-\lambda^{\mu}(\tau )=[u^{\mu}(p_s)u^{\nu}(p_s)-\epsilon^{\mu}_r(u(p_s))\epsilon^{\nu}_r(u(p_s))] {\dot x}_{s \nu}(\tau )=\nonumber \\ &=&- u^{\mu}(p_s) \lambda (\tau )+ \epsilon^{\mu}_r(u(p_s)) \lambda_r(\tau ),\nonumber \\ {\dot x}^2_s(\tau )&=& \lambda^2(\tau )-{\vec \lambda}(\tau ) > 0,\nonumber \\ U^{\mu}_s(\tau )&=& {{ {\dot x}^{\mu}_s(\tau )}\over { \sqrt{{\dot x}^2 _s(\tau )} }}={{-\lambda (\tau )u^{\mu}(p_s)+\lambda_r(\tau )\epsilon^{\mu}_r (u(p_s))}\over {\sqrt{\lambda^2(\tau )-{\vec \lambda}^2(\tau )} }}, \label{V6} \end{eqnarray} \noindent the timelike worldline described by the origin of the Wigner hyperplane is arbitrary (i.e. gauge dependent): $x^{\mu}_s(\tau )$ may be any covariant noncanonical centroid [the real ``external" center of mass is the canonical noncovariant ${\tilde x}^{\mu}_s(T_s)= x^{\mu} _s(T_s)-{1\over {\epsilon_s(p^o_s+\epsilon_s)}}\Big[ p_{s\nu}S^{\nu\mu}_s+\epsilon_s(S^{o\mu}_s+S^{o\nu}_s {{p_{s\nu}p^{\mu}_s}\over {\epsilon_s^2}}) \Big]$: it describes a decoupled point particle observer ; see Section VI]. In the gauge $T_s-\tau \approx 0$, ${\vec X}_{\phi}\approx 0$, implying $\lambda (\tau )=-1$, $\vec \lambda (\tau )=0$ [$g_{\tau\tau}=1$, $g_{\tau r}=0$], we get ${\dot x}^{\mu}_s(T_s)=u^{\mu}(p_s)$. Therefore, in this gauge, we have the centroid $x^{\mu}_s(T_s)=x_s^{({\vec X}_{\phi})\mu }(T_s)= x^{\mu}_s(0)+T_s u^{\mu}(p_s)$, which carries the Klein-Gordon ``internal" collective variable ${\vec \sigma}_{\phi}={\vec X}_{\phi}\approx 0$ (see Section VI). In this gauge we get the following form of the energy-momentum tensor [$\eta^{\mu\nu}=u^{\mu}(p_s)u^{\nu}(p_s)-\sum_{r=1}^3\epsilon^{\mu}_r(u(p_s)) \epsilon^{\nu}_r(u(p_s))$] \begin{eqnarray} T^{\mu\nu}[x^{\mu}_s(T_s)+\epsilon^{\mu}_u(u(p_s))\sigma^u][\phi ]&=& T^{\mu\nu}[x^{\mu}_s(T_s)+\epsilon^{\mu}_u(u(p_s))\sigma^u][X^A_{\phi}, P^A_{\phi},{\bf H},{\bf K}]=\nonumber \\ &=&{1\over 2}u^{\mu}(p_s)u^{\nu}(p_s)[\pi^2+(\vec \partial \phi )^2+ m^2\phi^2](T_s,\vec \sigma )+\nonumber \\ &+&\epsilon^{\mu}_r(u(p_s))\epsilon^{\nu}_s(u(p_s))[-{1\over 2}\delta_{rs} [\pi^2-(\vec \partial \phi )^2-m^2\phi^2]+\nonumber \\ &+&\partial_r\phi\partial_s\phi ](T_s, \vec \sigma )-\nonumber \\ &-&[u^{\mu}(p_s)\epsilon^{\nu}_r(u(p_s))+u^{\nu}(p_s)\epsilon^{\mu}_r(u(p_s))] [\pi \partial_r\phi ](T_s,\vec \sigma )=\nonumber \\ &=& \Big[ \rho [\phi ,\pi ] u^{\mu}(p_s)u^{\nu}(p_s)+{\cal P}[\phi ,\pi ] [\eta ^{\mu\nu}-u^{\mu}(p_s)u^{\nu}(p_s)]+\nonumber \\ &+&u^{\mu}(p_s)q^{\nu}[\phi ,\pi ]+ u^{\nu}(p_s)q^{\mu}[\phi ,\pi ]+\nonumber \\ &+&T^{rs}_{an\, stress}[\phi ,\pi ] \epsilon^{\mu}_r(u(p_s))\epsilon^{\nu} _s(u(p_s))\Big] (T_s,\vec \sigma ),\nonumber \\ &&{}\nonumber \\ \rho [\phi ,\pi ]&=&{1\over 2}[\pi^2+(\vec \partial \phi )^2+m^2\phi^2],\nonumber \\ {\cal P}[\phi ,\pi ]&=&{1\over 2} [\pi^2-{5\over 3}(\vec \partial \phi )^2- m^2\phi^2],\nonumber \\ q^{\mu}[\phi ,\pi ]&=&-\pi \partial_r\phi \epsilon^{\mu}_r(u(p_s)),\nonumber \\ T^{rs}_{an\, stress}[\phi ,\pi ]&=&-[\partial^r\phi \partial^s\phi -{1\over 3} \delta^{rs}(\vec \partial \phi )^2],\nonumber \\ &&{}{}{} \delta_{uv}T^{uv}_{an\, stress}[\phi ,\pi ]=0,\nonumber \\ &&{}\nonumber \\ T^{rs}_{stress}(T_s,\vec \sigma )[\phi ]&=&\epsilon^r_{\mu}(u(p_s))\epsilon^s _{\nu}(u(p_s))T^{\mu\nu}[x^{\mu}_s(T_s)+\epsilon^{\mu} _u(u(p_s))\sigma^u][\phi ]=\nonumber \\ &=&[\partial^r\phi \partial^s\phi ](T_s,\vec \sigma )- {1\over 2}\delta^{rs} [\pi^2-(\vec \partial \phi )^2-m^2\phi^2](T_s,\vec \sigma ),\nonumber \\ &&{}\nonumber \\ T^{\mu}{}_{\mu}[x^{\mu}_s(T_s)+\epsilon^{\mu}_u(u(p_s))\sigma^u][\phi ]&=& 2 [\pi^2-(\vec \partial \phi )^2-m^2 \phi^2](T_s, \vec \sigma ),\nonumber \\ &&{}\nonumber \\ T^{\mu\nu}[x^{\mu}_s(T_s)+\epsilon^{\mu}_u(u(p_s))\sigma^u][\phi ]\,\, u_{\nu}(p_s)&=&{1\over 2}[\pi^2+(\vec \partial \phi )^2+m^2\phi^2](T_s,\vec \sigma ) u^{\mu}(p_s)+\nonumber \\ &+&[\pi \partial^r\phi ](T_s,\vec \sigma )\epsilon^{\mu} _r(u(p_s)),\nonumber \\ &&{}\nonumber \\ P^{\mu}_T[\phi]&=&\int d^3\sigma T^{\mu\nu}[x^{\mu}_s(T_s)+\epsilon^{\mu} _u(u(p_s))\sigma^u][\phi ]u_{\nu}(p_s)=\nonumber \\ &=&P^{\tau}_{\phi} u^{\mu}(p_s)+P^r_{\phi}\epsilon^{\mu}_r(u(p_s))\approx P^{\tau}_{\phi} u^{\mu}(p_s) \approx p^{\mu}_s,\nonumber \\ M_{\phi}&=&P^{\mu}_T[\phi ] u_{\mu}(p_s) =P^{\tau}_{\phi}.\nonumber \\ \label{V7} \end{eqnarray} \noindent While the stress tensor of the Klein-Gordon field on the Wigner hyperplanes is $T^{rs}_{stress}(T_s,\vec \sigma )[\phi ]$, from the last line of the expression of the energy-momentum tensor we see that it acquires a form reminiscent of the energy-momentum tensor of an ideal relativistic fluid as seen from a local observer at rest (Eckart decomposition; see Ref.\cite{israel} ): i) the constant normal $u^{\mu}(p_s)$ to the Wigner hyperplanes replaces the hydrodynamic velocity field of the fluid; ii) $\rho [\phi ,\pi ](T_s,\vec \sigma ]$ is the energy density; iii) ${\cal P}[\phi ,\pi ](T_s,\vec \sigma )$ is the analogue of the pressure (sum of the thermodynamical pressure and of the non-equilibrium bulk stress or viscous pressure); iv) $q^{\mu}[\phi ,\pi ](T_s, \vec \sigma )$ is the analogue of the heat flow; v) $T^{rs}_{an\, stress}[\phi ,\pi ](T_s,\vec \sigma )$ is the shear (or anisotropic) stress tensor. We can now study the manifestly Lorentz covariant Dixon multipoles \cite{dixon} for the free real Klein-Gordon field on the Wigner hyperplanes in the gauge $\lambda (\tau )=-1$, $\vec \lambda (\tau )=0$ [so that ${\dot x}_s^{\mu}(T_s )=u^{\mu}(p_s)$, ${\ddot x}^{\mu}_s(T_s)=0$, $x^{\mu}_s(T_s)=u^{\mu}(p_s) T_s+x^{\mu}_s(0)$] with respect to the origin $x^{\mu}_s(T_s)$ [$\delta x^{\mu}_s(\vec \sigma )=\epsilon^{\mu}_u(u(p_s)) \sigma^u$; $(\mu_1..\mu_n)$ means symmetrization, while $[\mu_1..\mu_n]$ means antisymmetrization] \begin{eqnarray} t_T^{\mu_1...\mu_n\mu\nu}(T_s)&=&t_T^{(\mu_1...\mu_n)(\mu\nu)}(T_s)= \nonumber \\ &=&\int d^3\sigma \delta x^{\mu_1}_s(\vec \sigma )...\delta x^{\mu_n}_s(\vec \sigma ) T^{\mu\nu}[x^{\mu}_s(T_s)+\epsilon^{\mu}_u(u(p_s))\sigma^u][\phi ] =\nonumber \\ &=&\epsilon^{\mu_1}_{r_1}(u(p_s))...\epsilon^{\mu_n}_{r_n}(u(p_s))\nonumber \\ &&\Big[ u^{\mu}(p_s) u^{\nu}(p_s) {1\over 2}\int d^3\sigma \sigma^{r_1}... \sigma^{r_n}[\pi^2+(\vec \partial \phi )^2+m^2\phi^2](T_s,\vec \sigma )+ \nonumber \\ &+&\epsilon^{\mu}_r(u(p_s))\epsilon^{\nu}_s(u(p_s))\int d^3\sigma \sigma^{r_1} ...\sigma^{r_n}[-{1\over 2}\delta_{rs} [\pi^2-(\vec \partial \phi )^2-m^2\phi^2]+\nonumber \\ &+&\partial_r\phi\partial_s\phi ](T_s, \vec \sigma )+\nonumber \\ &+&[u^{\mu}(p_s)\epsilon^{\nu}_r(u(p_s))+u^{\nu}(p_s)\epsilon^{\mu}_r(u(p_s))] \int d^3\sigma \sigma^{r_1}...\sigma^{r_n}[\pi \partial^r\phi ](T_s,\vec \sigma ) \Big] =\nonumber \\ &=&\epsilon^{\mu_1}_{r_1}(u(p_s))...\epsilon^{\mu_n}_{r_n}(u(p_s)) \epsilon^{\mu}_A(u(p_s))\epsilon^{\nu}_B(u(p_s)) I_T^{r_1..r_nAB}(T_s)=\nonumber \\ &=&\epsilon^{\mu_1}_{r_1}(u(p_s))...\epsilon^{\mu_n}_{r_n}(u(p_s)) \Big[ u^{\mu}(p_s) u^{\nu}(p_s) I_T^{r_1...r_n\tau\tau}(T_s)+\nonumber \\ &+&\epsilon^{\mu}_r(u(p_s))\epsilon^{\nu}_s(u(p_s)) I_T^{r_1...r_nrs}(T_s)+\nonumber \\ &+&[u^{\mu}(p_s)\epsilon^{\nu}_r(u(p_s))+u^{\nu}(p_s)\epsilon^{\mu}_r(u(p_s))] I_T^{r_1...r_nr\tau}(T_s) \Big] ,\nonumber \\ &&{}\nonumber \\ u_{\mu_1}(p_s)&& t_T^{\mu_1...\mu_n\mu\nu}(T_s)=0,\nonumber \\ &&{}\nonumber \\ For && n=0\, (monopole)\quad I^{\tau\tau}_T(T_s)=P^{\tau}_{\phi},\quad\quad I_T^{r\tau}(T_s)=P^r_{\phi},\nonumber \\ &&{}\nonumber \\ t_T^{\mu_1...\mu_n\mu}{}_{\mu}(T_s)&=&\int d^3\sigma \delta x^{\mu_1}_s(\vec \sigma )...\delta x^{\mu_n}_s(\vec \sigma ) T^{\mu}{}_{\mu} [x^{\mu}_s(T_s)+\epsilon^{\mu}_u(u(p_s))\sigma^u][\phi ] =\nonumber \\ &=&\epsilon^{\mu_1}_{r_1}(u(p_s))...\epsilon^{\mu_n}_{r_n}(u(p_s))\nonumber \\ &&\int d^3\sigma \sigma^{r_1}...\sigma^{r_n} \Big[ 2[\pi^2-(\vec \partial \phi )^2]-m^2\phi^2\Big] (T_s,\vec \sigma )=\nonumber \\ &{\buildrel {def} \over =}&\epsilon^{\mu_1}_{r_1}(u(p_s))...\epsilon^{\mu_n} _{r_n}(u(p_s))I_T^{r_1...r_nA}{}_A(T_s)\nonumber \\ &&{}\nonumber \\ I_T^{r_1r_2A}{}_A(T_s)&=&{\check I}_T^{r_1r_2A}{}_A(T_s)-{1\over 3}\delta^{r_1r_2}\delta _{uv}I_T^{uvA}{}_A(T_s)=i^{r_1r_2}_T(T_s)-{1\over 2}\delta^{r_1r_2}\delta_{uv}i_T ^{uv}(T_s),\nonumber \\ {\check I}_T^{r_1r_2A}{}_A(T_s)&=&i_T^{r_1r_2}(T_s)-{1\over 3}\delta^{r_1r_2}\delta _{uv}i_T^{uv}(T_s),\quad\quad \delta_{uv}{\check I}_T^{uvA}{}_A(T_s)=0,\nonumber \\ i^{r_1r_2}_T(T_s)&=&I_T^{r_1r_2A}{}_A(T_s)-\delta^{r_1r_2}\delta_{uv}I_T^{uvA}{}_A(T_s)= \nonumber \\ &=&\int d^3\sigma (\sigma^{r_1}\sigma^{r_2}-\delta ^{r_1r_2}{\vec \sigma}^2) \Big[ 2[\pi^2-(\vec \partial \phi )^2]-m^2\phi^2\Big] (T_s,\vec \sigma ),\nonumber \\ &&{}\nonumber \\ {\tilde t}_T^{\mu_1...\mu_n}(T_s)&=& t_T^{\mu_1...\mu_n\mu\nu}(T_s) u_{\mu}(p_s)u_{\nu}(p_s)=\nonumber \\ &=&\epsilon^{\mu_1}_{r_1}(u(p_s))...\epsilon^{\mu_n}_{r_n}(u(p_s)) I_T^{r_1..r_n\tau\tau}(T_s),\nonumber \\ {\tilde t}_T^{\mu_1}(T_s)&=& \epsilon^{\mu_1}_{r_1}(u(p_s)) {1\over 2} \int d^3\sigma \sigma^{r_1} [\pi^2+(\vec \partial \phi )^2+m^2\phi^2](T_s ,\vec \sigma )= \nonumber \\ &=&\epsilon^{\mu_1}_{r_1}(u(p_s)) I_T^{r_1\tau\tau}(T_s)=-P^{\tau}_{\phi} \epsilon^{\mu_1}_{r_1}(u(p_s)) r^{r_1}_{\phi},\nonumber \\ &&{}\nonumber \\ {\tilde t}_T^{\mu_1\mu_2}(T_s)&=& \epsilon^{\mu_1}_{r_1}(u(p_s))\epsilon^{\mu_2}_{r_2}(u(p _s))I_T^{r_1r_2\tau\tau}(T_s),\nonumber \\ &&{}\nonumber \\ I^{r_1r_2\tau\tau}_T(T_s)&=&{\hat I}^{r_1r_2\tau\tau}_T(T_s)-{1\over 3}\delta^{r_1r_2} \delta_{uv}I^{uv\tau\tau}_T(T_s)={\tilde i}^{r_1r_2}_T(T_s)-{1\over 2}\delta ^{r_1r_2}\delta_{uv}{\tilde i}^{uv}_T(T_s),\nonumber \\ {\hat I}^{r_1r_2\tau\tau}_T(T_s)&=&{\tilde i}^{r_1r_2}_T(T_s)-{1\over 3}\delta^{r_1r_2} \delta_{uv}{\tilde i}^{uv}_T(T_s),\quad\quad \delta_{uv}{\hat I}^{uv\tau\tau}_T(T_s)=0, \nonumber \\ {\tilde i}_T^{r_1r_2}(T_s)&=&I_T^{r_1r_2\tau\tau}(T_s)-\delta^{r_1r_2}\delta _{uv}I_T^{uv\tau\tau}(T_s). \label{V8} \end{eqnarray} The Wigner covariant multipoles $I_T^{r_1..r_n\tau\tau}(T_s)$, $I_T^{r_1..r_nrs}(T_s)$, $I_T^{r_1..r_nr\tau}(T_s)$ are the mass, stress and momentum multipoles respectively. The quantities ${\check I}_T^{r_1r_2A}{}_A(T_s)$ and $i^{r_1r_2}_T (T_s)$ are the traceless quadrupole moment and the inertia tensor defined by Thorne in Ref.\cite{thorne}. The quantities $I_T^{r_1r_2\tau\tau}(T_s)$ and ${\tilde i}_T^{r_1r_2} (T_s)$ are Dixon's definitions of quadrupole moment and of tensor of inertia respectively. Moreover, Dixon's definition of ``center of mass" of an extended object is ${\tilde t}^{\mu_1}_T(T_s)=0$ or $I_T^{r\tau\tau}(T_s)=-P^{\tau}_{\phi} r^r_{\phi}=0$: therefore the quantity ${\vec r}_{\phi} $ defined in the previous equation is a noncanonical [$\{ r^r_{\phi},r^s_{\phi} \} =S^{rs}_{\phi}$] candidate for the ``internal" center of mass of the field configuration: its vanishing is a gauge fixing for ${\vec P}_{\phi}\approx 0$ and implies $x^{\mu}_s(T_s)=x^{({\vec r}_{\phi})\mu}_s(T_s)=x^{\mu}_s(0)+u^{\mu}(p_s) T_s$ (see next Section). When $I^{r\tau\tau}_T(T_s)=0$, the equations $0={{dI^{r\tau\tau}_T(T_s)}\over {dT_s}}=-P^{\tau}_{\phi} {{dr^r_{\phi}}\over {dT_s}}= {d\over {dT_s}} \int d^3\sigma \, \sigma^r[\pi^2+(\vec \partial \phi )^2+m^2\phi^2](T_s,\vec \sigma )\, {\buildrel \circ \over =}\, -P^r_{\phi}$ implies the correct momentum-velocity relation ${{{\vec P}_{\phi}}\over {P^{\tau}_{\phi}}}\, {\buildrel \circ \over =}\, {{d{\vec r}_{\phi}}\over {dT_s}} \approx 0$> These multipoles, whose Poisson brackets are non-trivial, are related to the previous ones of Eqs.(\ref{IV41}), with the Poisson brackets given in Eqs.(\ref{IV42}),in the following way (remember that these other multipoles exist only if the Klein-Gordon fields have compact support V in momentum space) \begin{eqnarray} I_T^{r_1...r_n\tau\tau}(T_s)&=&{1\over 2}\int d^3\sigma \sigma^{r_1}... \sigma^{r_n}[\pi^2+(\vec \partial \phi )^2+m^2\phi^2](T_s,\vec \sigma )= \nonumber \\ &=&2 \int d{\tilde q}_1d{\tilde q}_2 {\bf A}_{{\vec q}_1}(T_s;P^{\tau}_{\phi}, {\bf H}] {\bf A}_{{\vec q}_2}(T_s;P^{\tau}_{\phi},{\bf H}] \int d^3\sigma \sigma^{r_1}...\sigma^{r_n} \nonumber \\ &&\Big( [\omega (q_1)\omega (q_2) +{\vec q}_1\cdot {\vec q}_2] sin\, ({\vec q}_1\cdot \vec \sigma +{\bf B}_{{\vec q}_1}(T_s;{\tilde X}^{\tau}_{\phi}, {\bf K}]) sin\, ({\vec q}_2\cdot \vec \sigma +{\bf B}_{{\vec q}_2}(T_s;{\tilde X}^{\tau}_{\phi}, {\bf K}]) +\nonumber \\ &+&m^2 cos\, ({\vec q}_1\cdot \vec \sigma +{\bf B}_{{\vec q}_1}(T_s;{\tilde X}^{\tau}_{\phi}, {\bf K}]) cos\, ({\vec q}_2\cdot \vec \sigma +{\bf B}_{{\vec q}_2}(T_s;{\tilde X}^{\tau}_{\phi}, {\bf K}]) \Big) =\nonumber \\ &=&{1\over 2} \sum_{h,k=0}^{\infty} (-)^{h+k} \nonumber \\ &&\Big( {1\over {(2h+1)!(2k+1)!}} \Big[ V_1^{r_1...r_nu_1...u_{2h+1}v_1...v_{2k+1}} \nonumber \\ &&[{\cal T}_{\pi ,O}^{u_1...u _{2h+1}}{\cal T}_{\pi ,O}^{v_1...v_{2k+1}}+m^2{\cal T}_{\phi ,O}^{u_1...u _{2h+1}}{\cal T}_{\phi ,O}^{v_1...v_{2k+1}}]+\nonumber \\ &+&V_2^{r_1...r_nu_1...u_{2h+1}v_1...v_{2k+1}}{\cal T}_{\phi ,O}^{u_1...u _{2h+1}}{\cal T}_{\phi ,O}^{v_1...v_{2k+1}} \Big] -\nonumber \\ &-&{1\over {(2h+1)!(2k)!}} \Big[ V_1^{r_1...r_nu_1...u_{2h+1}v_1...v_{2k}} [{\cal T}_{\pi ,O}^{u_1...u _{2h+1}}{\cal T}_{\pi ,E}^{v_1...v_{2k}}+m^2{\cal T}_{\phi ,O}^{u_1...u _{2h+1}}{\cal T}_{\phi ,E}^{v_1...v_{2k}}]+\nonumber \\ &+&V_2^{r_1...r_nu_1...u_{2h+1}v_1...v_{2k}}{\cal T}_{\phi ,O}^{u_1...u_{2h+1}} {\cal T}_{\phi ,E}^{v_1...v_{2k}} \Big] -\nonumber \\ &-&{1\over {(2h)!(2k+1)!}} \Big[ V_1^{r_1...r_nu_1...u_{2h}v_1...v_{2k+1}} [{\cal T}_{\pi ,E}^{u_1...u _{2h}}{\cal T}_{\pi ,O}^{v_1...v_{2k+1}}+m^2{\cal T}_{\phi ,E}^{u_1...u _{2h}}{\cal T}_{\phi ,O}^{v_1...v_{2k+1}}]+\nonumber \\ &+&V_2^{r_1...r_nu_1...u_{2h}v_1...v_{2k+1}}{\cal T}_{\phi ,E}^{u_1...u_{2h}} {\cal T}_{\phi ,O}^{v_1...v_{2k+1}} \Big] +\nonumber \\ &+&{1\over {(2h)!(2k)!}} \Big[ V_1^{r_1...r_nu_1...u_{2h}v_1...v_{2k}} [{\cal T}_{\pi ,E}^{u_1...u _{2h}}{\cal T}_{\pi ,E}^{v_1...v_{2k}}+m^2{\cal T}_{\phi ,E}^{u_1...u _{2h}}{\cal T}_{\phi ,E}^{v_1...v_{2k}}]+\nonumber \\ &+&V_2^{r_1...r_nu_1...u_{2h}v_1...v_{2k}}{\cal T}_{\phi ,E}^{u_1...u_{2h}} {\cal T}_{\phi ,E}^{v_1...v_{2k}} \Big] \Big) ,\nonumber \\ I_T^{r_1...r_nrs}(T_s)&=&\int d^3\sigma \sigma^{r_1}...\sigma^{r_n}[-{1\over 2} \delta_{rs}[\pi^2-(\vec \partial \phi )^2-m^2\phi^2]+\partial_r\phi\partial_s \phi ](T_s,\vec \sigma )=\nonumber \\ &=&4\int d{\tilde q}_1d{\tilde q}_2 {\bf A}_{{\vec q}_1}(T_s;P^{\tau}_{\phi}, {\bf H}] {\bf A}_{{\vec q}_2}(T_s;P^{\tau}_{\phi},{\bf H}] \int d^3\sigma \sigma^{r_1}...\sigma^{r_n}\nonumber \\ &&\Big( -{1\over 2}\delta^{rs}\Big[ [\omega (q_1)\omega (q_2) -{\vec q}_1\cdot {\vec q}_2] \nonumber \\ &&sin\, ({\vec q}_1\cdot \vec \sigma +{\bf B}_{{\vec q}_1}(T_s;{\tilde X}^{\tau}_{\phi}, {\bf K}]) sin\, ({\vec q}_2\cdot \vec \sigma +{\bf B}_{{\vec q}_2}(T_s;{\tilde X}^{\tau}_{\phi}, {\bf K}])-\nonumber \\ &-&m^2 cos\, ({\vec q}_1\cdot \vec \sigma +{\bf B}_{{\vec q}_1}(T_s;{\tilde X}^{\tau}_{\phi}, {\bf K}]) cos\, ({\vec q}_2\cdot \vec \sigma +{\bf B}_{{\vec q}_2}(T_s;{\tilde X}^{\tau}_{\phi}, {\bf K}]) \Big] +\nonumber \\ &+& q^r_1 q^s_2 sin\, ({\vec q}_1\cdot \vec \sigma +{\bf B}_{{\vec q}_1}(T_s;{\tilde X}^{\tau}_{\phi}, {\bf K}]) sin\, ({\vec q}_2\cdot \vec \sigma +{\bf B}_{{\vec q}_2}(T_s;{\tilde X}^{\tau}_{\phi}, {\bf K}]) \Big) =\nonumber \\ &=&\sum_{h,k=0}^{\infty} (-)^{h+k}\nonumber \\ &&\Big( {1\over {(2h+1)!(2k+1)!}} \Big[-{1\over 2}\delta^{rs} V_1^{r_1...r_nu_1...u_{2h+1}v_1...v_{2k+1}} \nonumber \\ &&[{\cal T}_{\pi ,O}^{u_1...u _{2h+1}}{\cal T}_{\pi ,O}^{v_1...v_{2k+1}}-m^2{\cal T}_{\phi ,O}^{u_1...u _{2h+1}}{\cal T}_{\phi ,O}^{v_1...v_{2k+1}}]+\nonumber \\ &+&({1\over 2}\delta^{rs} V_2^{r_1...r_nu_1...u_{2h+1}v_1...v_{2k+1}}+V_3^{r_1...r_nu_1...u_{2h+1}v_1... v_{2k+1}rs}) {\cal T}_{\phi ,O}^{u_1...u_{2h+1}} {\cal T}_{\phi ,O}^{v_1...v_{2k+1}} \Big] -\nonumber \\ &-&{1\over {(2h+1)!(2k)!}} \Big[ -{1\over 2}\delta^{rs}V_1^{r_1...r_nu_1...u_{2h+1}v_1...v_{2k}} \nonumber \\ &&[{\cal T}_{\pi ,O}^{u_1...u _{2h+1}}{\cal T}_{\pi ,E}^{v_1...v_{2k}}-m^2{\cal T}_{\phi ,O}^{u_1...u _{2h+1}}{\cal T}_{\phi ,E}^{v_1...v_{2k}}]+\nonumber \\ &+&({1\over 2}\delta^{rs}V_2^{r_1...r_nu_1...u_{2h+1}v_1...v_{2k}}+V_3^{r_1... r_nu_1...u_{2h+1}v_1...v_{2k}rs}){\cal T}_{\phi ,O}^{u_1...u_{2h+1}} {\cal T}_{\phi ,E}^{v_1...v_{2k}} \Big] -\nonumber \\ &-&{1\over {(2h)!(2k+1)!}} \Big[ -{1\over 2}\delta^{rs}V_1^{r_1...r_nu_1...u_{2h}v_1...v_{2k+1}} \nonumber \\ &&[{\cal T}_{\pi ,E}^{u_1...u _{2h}}{\cal T}_{\pi ,O}^{v_1...v_{2k+1}}-m^2{\cal T}_{\phi ,E}^{u_1...u _{2h}}{\cal T}_{\phi ,O}^{v_1...v_{2k+1}}]+\nonumber \\ &+&({1\over 2}\delta^{rs}V_2^{r_1...r_nu_1...u_{2h}v_1...v_{2k+1}}+V_3^{r_1... r_nu_1...u_{2h}v_1...v_{2k+1}rs}){\cal T}_{\phi ,E}^{u_1...u_{2h}} {\cal T}_{\phi ,O}^{v_1...v_{2k+1}} \Big] +\nonumber \\ &+&{1\over {(2h)!(2k)!}} \Big[ -{1\over 2}\delta^{rs} V_1^{r_1...r_nu_1...u_{2h}v_1...v_{2k}} [{\cal T}_{\pi ,E}^{u_1...u _{2h}}{\cal T}_{\pi ,E}^{v_1...v_{2k}}-m^2{\cal T}_{\phi ,E}^{u_1...u _{2h}}{\cal T}_{\phi ,E}^{v_1...v_{2k}}]+\nonumber \\ &+&({1\over 2}\delta^{rs}V_2^{r_1...r_nu_1...u_{2h}v_1...v_{2k}}+V_3^{r_1... r_nu_1...u_{2h}v_1...v_{2k}rs}){\cal T}_{\phi ,E}^{u_1...u_{2h}} {\cal T}_{\phi ,E}^{v_1...v_{2k}} \Big] \Big) ,\nonumber \\ I_T^{r_1...r_nr\tau}(T_s)&=&\int d^3\sigma \sigma^{r_1}...\sigma^{r_n}[\pi \partial^r \phi ](T_s,\vec \sigma )=\nonumber \\ &=&4 \int d{\tilde q}_1d{\tilde q}_2 {\bf A}_{{\vec q}_1}(T_s;P^{\tau}_{\phi}, {\bf H}] {\bf A}_{{\vec q}_2}(T_s;P^{\tau}_{\phi},{\bf H}] \omega (q_1) q^r_2 \int d^3\sigma \sigma^{r_1}...\sigma^{r_n}\nonumber \\ &&sin\, ({\vec q}_1\cdot \vec \sigma +{\bf B}_{{\vec q}_1}(T_s;{\tilde X}^{\tau}_{\phi}, {\bf K}]) sin\, ({\vec q}_2\cdot \vec \sigma +{\bf B}_{{\vec q}_2}(T_s;{\tilde X}^{\tau}_{\phi}, {\bf K}]) =\nonumber \\ &=&\sum_{h,k=0}^{\infty} (-)^{h+k}\nonumber \\ &&\Big( {1\over {(2h+1)!(2k+1)!}} V_4^{r_1...r_nu_1...u_{2h+1}v_1...v_{2k+1}} {\cal T}_{\pi ,O}^{u_1...u _{2h+1}}{\cal T}_{\phi ,O}^{v_1...v_{2k+1}}-\nonumber \\ &-&{1\over {(2h+1)!(2k)!}} V_4^{r_1...r_nu_1...u_{2h+1}v_1...v_{2k}} {\cal T}_{\pi ,O}^{u_1...u _{2h+1}}{\cal T}_{\phi ,E}^{v_1...v_{2k}}-\nonumber \\ &-&{1\over {(2h)!(2k+1)!}} V_4^{r_1...r_nu_1...u_{2h}v_1...v_{2k+1}} {\cal T}_{\pi ,E}^{u_1...u _{2h}}{\cal T}_{\phi ,O}^{v_1...v_{2k+1}}+\nonumber \\ &+&{1\over {(2h)!(2k)!}} V_4^{r_1...r_nu_1...u_{2h}v_1...v_{2k}} {\cal T}_{\pi ,E}^{u_1...u _{2h}}{\cal T}_{\phi ,E}^{v_1...v_{2k}}\Big) ,\nonumber \\ &&{}\nonumber \\ &&where\nonumber \\ &&{}\nonumber \\ &&V_1^{r_1...r_nu_1...u_hv_1...v_k}=\int_V d^3\sigma \sigma^{r_1}...\sigma^{r_n} \sigma^{u_1}...\sigma^{u_h}\sigma^{v_1}...\sigma^{v_k},\nonumber \\ &&V_2^{r_1...r_nu_1...u_hv_1...v_k}=\int_V d^3\sigma \sigma^{r_1}...\sigma^{r_n} \vec \partial (\sigma^{u_1}...\sigma^{u_h}) \cdot \vec \partial (\sigma^{v_1} ...\sigma^{v_k}),\nonumber \\ &&V_3^{r_1...r_nu_1...u_hv_1...v_krs}=\int_V d^3\sigma \sigma^{r_1}...\sigma ^{r_n} \partial^r (\sigma^{u_1}...\sigma^{u_h})\, \partial^s (\sigma^{v_1}... \sigma^{v_k}),\nonumber \\ &&V_4^{r_1...r_nu_1...u_hv_1...v_kr}=\int_V d^3\sigma \sigma^{r_1}...\sigma ^{r_n}\sigma^{u_1}...\sigma^{u_h}\, \partial^r (\sigma^{v_1}...\sigma^{v_k}). \label{V9} \end{eqnarray} Then there are the related Dixon multipoles \begin{eqnarray} p^{\mu_1...\mu_n\mu}_T(T_s)&=&t_T^{\mu_1...\mu_n\mu\nu}(T_s) u_{\nu}(p_s)= p_T^{(\mu_1...\mu_n)\mu}(T_s)=\nonumber \\ &=&\epsilon^{\mu_1}_{r_1}(u(p_s))...\epsilon^{\mu_n}_{r_n}(u(p_s)) \epsilon^{\mu}_A(u(p_s)) I_T^{r_1..r_nA\tau}(T_s)=\nonumber \\ &=&\epsilon^{\mu_1}_{r_1}(u(p_s))...\epsilon^{\mu_n}_{r_n}(u(p_s)) \int d^3\sigma \sigma^{r_1}...\sigma^{r_n} \Big[ {1\over 2}[\pi^2+(\vec \partial \phi )^2+m^2\phi^2](T_s ,\vec \sigma ) u^{\mu}(p_s)+\nonumber \\ &+&[\pi \partial^r\phi ](T_s ,\vec \sigma )\epsilon^{\mu} _r(u(p_s)) \Big] =\nonumber \\ &=&2 \epsilon^{\mu_1}_{r_1}(u(p_s))...\epsilon^{\mu_n}_{r_n}(u(p_s)) \nonumber \\ &&\int d{\tilde q}_1 d{\tilde q}_2 {\bf A}_{{\vec q}_1}(T_s;P^{\tau}_{\phi}, {\bf H}] {\bf A}_{{\vec q}_2}(T_s;P^{\tau}_{\phi},{\bf H}] \int d^3\sigma \sigma^{r_1}...\sigma^{r_n}\nonumber \\ &&\Big[ u^{\mu}(p_s) \Big( (\omega (q_1) \omega (q_2) +{\vec q}_1\cdot {\vec q}_2) \nonumber \\ &&sin\, ({\vec q}_1\cdot \vec \sigma +{\bf B}_{{\vec q}_1}(T_s;{\tilde X} ^{\tau}_{\phi},{\bf K}]) sin\, ({\vec q}_2\cdot \vec \sigma +{\bf B}_{{\vec q}_2}(T_s;{\tilde X} ^{\tau}_{\phi},{\bf K}])+\nonumber \\ &+& m^2 cos\, ({\vec q}_1\cdot \vec \sigma +{\bf B}_{{\vec q}_1}(T_s;{\tilde X} ^{\tau}_{\phi},{\bf K}]) cos\, ({\vec q}_1\cdot \vec \sigma +{\bf B}_{{\vec q}_1}(T_s;{\tilde X} ^{\tau}_{\phi},{\bf K}]) \Big) -\nonumber \\ &-&\epsilon^{\mu}_r(u(p_s))\,\, q^r_2\,\, sin({\vec q}_1\cdot \vec \sigma +{\bf B}_{{\vec q}_1}(T_s;{\tilde X} ^{\tau}_{\phi},{\bf K}]) sin({\vec q}_2\cdot \vec \sigma +{\bf B}_{{\vec q}_2}(T_s;{\tilde X} ^{\tau}_{\phi},{\bf K}]) \Big] ,\nonumber \\ &&{}\nonumber \\ u_{\mu_1}(p_s)&& p_T^{\mu_1...\mu_n\mu}(T_s)=0,\nonumber \\ &&{}\nonumber \\ n=0&&\Rightarrow p^{\mu}_T(T_s)=P^{\mu}_T[\phi ]=\epsilon^{\mu}_A(u(p_s)) P^A_{\phi}\approx p^{\mu}_s,\nonumber \\ &&{}\nonumber \\ p_T^{\mu_1..\mu_n\mu}(T_s)u_{\mu}(p_s)&=&{\tilde t}_T^{\mu_1...\mu_n}(T_s) =\epsilon^{\mu_1}_{r_1}(u(p_s))...\epsilon^{\mu_n}_{r_n}(u(p_s)) I_T^{r_1..r_n\tau\tau}(T_s). \label{V10} \end{eqnarray} The spin dipole is defined as \begin{eqnarray} S^{\mu\nu}_T(T_s)[\phi ]&=&2 p_T^{[\mu\nu ]}(T_s)=2 \epsilon^{[\mu}_r(u(p_s))\epsilon^{\nu ]}_A(u(p_s)) I_T^{rA\tau}(T_s)=\nonumber \\ &=&[\epsilon^{\mu}_r(u(p_s)) u^{\nu}(p_s)-\epsilon^{\nu}_r(u(p_s)) u^{\mu} (p_s)] {1\over 2} \int d^3\sigma \sigma^r [\pi^2+(\vec \partial \phi )^2 +m^2\phi^2](T_s ,\vec \sigma )+\nonumber \\ &+&[\epsilon^{\mu}_r(u(p_s)) \epsilon^{\nu}_s(u(p_s))- \epsilon^{\nu}_r(u(p_s)) \epsilon^{\mu}_s(u(p_s))] \int d^3\sigma \sigma^r [\pi \partial^s\phi ](T_s,\vec \sigma )=\nonumber \\ &=&S^{\mu\nu}_s=\epsilon^{\mu}_r(u(p_s))\epsilon^{\nu}_s(u(p_s))S^{rs}_{\phi}+ [\epsilon^{\mu}_r(u(p_s)) u^{\nu}(p_s)-\epsilon^{\nu}_r(u(p_s)) u^{\mu} (p_s)]S^{\tau r}_{\phi}=\nonumber \\ &=&\epsilon^{\mu}_r(u(p_s))\epsilon^{\nu}_s(u(p_s)) \int d\tilde q {\bf H}(T_s,\vec q) \Big(q^r{{\partial}\over {\partial q^s}}-q^s{{\partial} \over {\partial q^r}}\Big) {\bf K}(T_s,\vec q)-\nonumber \\ &-&[\epsilon^{\mu}_r(u(p_s)) u^{\nu}(p_s)-\epsilon^{\nu}_r(u(p_s)) u^{\mu} (p_s)]\int d\tilde q\, \omega (q) {\bf H}(T_s,\vec q) {{\partial}\over {\partial q^r}} {\bf K}(T_s,\vec q),\nonumber \\ &&{}\nonumber \\ &&u_{\mu}(p_s) S^{\mu\nu}_T(T_s)[\phi ]=-\epsilon^{\nu}_r(u(p_s)) S^{\tau r}_{\phi}=-{\tilde t}^{\nu}_T(T_s)=P^{\tau}_{\phi}\epsilon^{\nu}_r(u(p_s)) r^r_{\phi}, \label{V11} \end{eqnarray} \noindent with $u_{\mu}(p_s)S^{\mu\nu}_T(T_s)[\phi ]=0$ when ${\tilde t}_T^{\mu_1}(T_s)=0$ and this condition can be taken as a definition of center of mass equivalent to Dixon's one. When this condition holds, the barycentrice spin dipole is $S^{\mu\nu}_T(T_s)[\phi ]=2 \epsilon^{[\mu}_r(u(p_s)) \epsilon^{\nu ]}_s(u(p_s)) I_T^{rs\tau}(T_s)$, so that $I^{[rs]\tau}_T(T_s)=\epsilon^r_{\mu}(u(p_s)) \epsilon^s_{\nu}(u(p_s)) S^{\mu\nu}_T(T_s)[\phi ]$. As shown in Ref.\cite{dixon}, if the Klein-Gordon field has a compact support W on the Wigner hyperplanes $\Sigma_{W\tau}$ and if $f(x)$ is a $C^{\infty}$ complex-valued scalar function on Minkowski spacetime with compact support [so that its Fourier transform $\tilde f(k)=\int d^4x f(x) e^{ik\cdot x}$ is a slowly increasing entire analytic function on Minkowski spacetime ($|(x^o+iy^o) ^{q_o}...(x^3+iy^3)^{q_3}f(x^{\mu}+iy^{\mu})| < C_{q_o...q_3} e^{a_o|y^o|+...+ a_3|y^3|}$, $a_{\mu} > 0$, $q_{\mu}$ positive integers for every $\mu$ and $C_{q_o...q_3} > 0$), whose inverse is $f(x)=\int {{d^4k}\over {(2\pi )^4}} \tilde f(k)e^{-ik\cdot x}$], we have \begin{eqnarray} < T^{\mu\nu},f >&=& \int d^4x T^{\mu\nu}(x) f(x)=\nonumber \\ &=&\int dT_s\int d^3\sigma f(x_s+\delta x_s)T^{\mu\nu}[x_s(T_s)+\delta x_s(\vec \sigma )][\phi ]=\nonumber \\ &=&\int dT_s \int d^3\sigma \int {{d^4k}\over {(2\pi )^4}} \tilde f(k) e^{-ik\cdot [x_s(T_s)+\delta x_s(\vec \sigma )]}T^{\mu\nu} [x_s(T_s)+\delta x_s(\vec \sigma )][\phi ]=\nonumber \\ &=&\int dT_s\int {{d^4k}\over {(2\pi )^4}} \tilde f(k) e^{-ik\cdot x_s(T_s)} \int d^3\sigma T^{\mu\nu}[x_s(T_s)+\delta x_s(\vec \sigma )][\phi ]\nonumber \\ &&\sum_{n=0}^{\infty} {{(-i)^n}\over {n!}} [k_{\mu}\epsilon^{\mu}_u(u(p_s)) \sigma^u]^n=\nonumber \\ &=&\int dT_s\int {{d^4k}\over {(2\pi )^4}} \tilde f(k) e^{-ik\cdot x_s(T_s)} \sum_{n=0}^{\infty} {{(-i)^n}\over {n!}} k_{\mu_1}...k_{\mu_n} t_T^{\mu_1... \mu_n\mu\nu}(T_s), \label{V12} \end{eqnarray} \noindent and, but only for $f(x)$ analytic in W \cite{dixon}, we get \begin{eqnarray} < T^{\mu\nu},f >&=& \int dT_s \sum_{n=0}^{\infty}{1\over {n!}} t_T^{\mu_1... \mu_n\mu\nu}(T_s) {{\partial^nf(x)}\over {\partial x^{\mu_1}...\partial x^{\mu _n}}}{|}_{x=x_s(T_s)},\nonumber \\ &&\Downarrow \nonumber \\ T^{\mu\nu}(x)[\phi ]&=& \sum_{n=0}^{\infty}{{(-1)^n}\over {n!}} {{\partial^n}\over {\partial x^{\mu_1}...\partial x^{\mu_n}}} \int dT_s \delta^4(x-x_s(T_s)) t_T^{\mu_1...\mu_n\mu\nu}(T_s). \label{V13} \end{eqnarray} For a non analytic $f(x)$ we have \begin{eqnarray} < T^{\mu\nu},f > &=& \int dT_s \sum^N_{n=0} {1\over {n!}} t_T^{\mu_1...\mu_n \mu\nu}(T_s) {{\partial^nf(x)}\over {\partial x^{\mu_1}...\partial x^{\mu _n}}}{|}_{x=x_s(T_s)}+\nonumber \\ &+&\int dT_s \int {{d^4k}\over {(2\pi )^4}} \tilde f(k) e^{-ik\cdot x_s(T_s)} \sum_{n=N+1}^{\infty} {{(-i)^n}\over {n!}} k_{\mu_1}...k_{\mu_n} t_T^{\mu_1... \mu_n\mu\nu}(T_s), \label{V14} \end{eqnarray} \noindent and, as shown in Ref.\cite{dixon}, from the knowledge of the moments $t_T^{\mu_1...\mu_n\mu}(T_s)$ for all $n > N$ we can get $T^{\mu\nu}(x)$ and, thus, all the moments with $n\leq N$. In Appendix C other types of Dixon's multipoles are analyzed. From this study it turns out that the multipolar expansion(\ref{V13}) may be rearranged with the help of the Hamilton equations implying $\partial_{\mu}T^{\mu\nu}\, {\buildrel \circ \over =}\, 0$, so that for analytic Klein-Gordon configurations from Eq.(\ref{c5}) we get \begin{eqnarray} T^{\mu\nu}(x)[\phi]\, &{\buildrel \circ \over =}\,& u^{(\mu}(p_s) \epsilon^{\nu )}_A(u(p_s)) \int dT_s\, \delta^4(x-x_s(T_s))\, P^A_{\phi}+\nonumber \\ &+&{1\over 2} {{\partial}\over {\partial x^{\rho}}} \int dT_s\, \delta^4(x-x_s(T_s))\, S_T^{\rho (\mu}(T_s)[\phi] u^{\nu )}(p_s)+\nonumber \\ &+&\sum_{n=2}^{\infty}{{(-1)^n}\over {n!}} {{\partial^n}\over {\partial x^{\mu_1}...\partial x^{\mu_n}}} \int dT_s\, \delta^4(x-x_s(T_s)) {\cal I}_T^{\mu_1..\mu_n\mu\nu}(T_s), \label{V15} \end{eqnarray} \noindent where for $n\geq 2$ $\quad {\cal I}_T^{\mu_1..\mu_n\mu\nu}(T_s) ={{4(n-1)}\over {n+1}} J_T^{(\mu_1..\mu_{n-1} | \mu | \mu_n)\nu}(T_s)$, with the quantities $J_T^{\mu_1..\mu_n\mu\nu\rho\sigma}(T_s)$ being the Dixon $2^{2+n}$-pole inertial moment tensors given in Eqs.(\ref{c7}) [the quadrupole and related inertia tensor are proportional to $I_T^{r_1r_2\tau\tau}(T_s)$]. The equations $\partial_{\mu}T^{\mu\nu}\, {\buildrel \circ \over =}\, 0$ imply the Papapetrou-Dixon-Souriau equations for the `pole-dipole' system $P_T^{\mu}(T_s)$ and $S^{\mu\nu}_T(T_s)[\phi ]$ [see Eqs.(\ref{c1}) and (\ref{c4})] \begin{eqnarray} {{d P^{\mu}_T(T_s)}\over {dT_s}}\, &{\buildrel \circ \over =}\,& 0,\nonumber \\ {{d S^{\mu\nu}_T(T_s)[\phi ]}\over {dT_s}}\, &{\buildrel \circ \over =}\,& 2 P^{[\mu}_T(T_s) u^{\nu ]}(p_s)=2 P^r_{\phi} \epsilon^{[\mu}_r(u(p_s)) u^{\nu ]}(p_s) \approx 0. \label{V16} \end{eqnarray} \vfill\eject \section{External and internal canonical center of mass, Moller's center of energy and Fokker-Pryce center of inertia} Let us now consider the problem of the definition of the relativistic center of mass of a Klein-Gordon field configuration, after having seen in the previous Section, Eq(\ref{V10}), Dixon's definition of this concept in the multipolar approach. Let us remark that in the approach leading to the rest-frame instant form of dynamics on Wigner's hyperplanes there is a splitting of this concept in an ``external" and an ``internal" one. One can either look at the isolated system from an arbitrary Lorentz frame or put himself inside the Wigner hyperplane. From outside one finds after the canonical reduction to Wigner hyperplane that there is an origin $x^{\mu}_s(\tau )$ for these hyperplanes (a covariant noncanonical centroid) and a noncovariant canonical coordinate ${\tilde x} ^{\mu}_s(\tau )$ describing an ``external" decoupled point particle observer with a clock measuring the rest-frame time $T_s$. Associated with them there is the ``external" realization (\ref{II9}) of the Poincar\'e group. Instead, all the degrees of freedom of the isolated system (here the Klein-Gordon field configuration) are described by canonical variables on the Wigner hyperplane restricted by the rest-frame condition ${\vec P}_{\phi} \approx 0$, implying that an ``internal" collective variable ${\vec X}_{\phi}$ is a gauge variable and that only relative variables are physical degrees of freedom (a form of weak Mach principle). Inside the Wigner hyperplane at $\tau =0$ there is another realization of the Poincar\'e group given by Eqs. (\ref{III4}), (\ref{III5}) with generators $P^{\tau}_{\phi}$, $P^r_{\phi}$, $J^{rs}_{\phi}$, $K^r_{\phi}=J^{\tau r}_{\phi}{|}_{\tau =0}=S^{\tau r}_{\phi}$, the ``internal" Poincar\' e algebra. By using the methods of Ref.\cite{pauri} (where there is a complete discussion of many definitions of relativistic center-of-mass-like variables) we can build the following three 'internal" (that is inside the Wigner hyperplane) Wigner 3-vectors corresponding to the 3-vectors 'canonical center of mass' ${\vec q}_{\phi}$, 'Moller center of energy' ${\vec r}_{\phi}$ and 'Fokker-Pryce center of inertia' ${\vec y}_{\phi}$ (the analogous concepts for the relativistic N-body problem are under study \cite{iten}). The noncanonical ``internal" M\o ller center of energy and the associated spin 3-vector are \begin{eqnarray} {\vec r}_{\phi}&=& - {{\vec K}_{\phi}\over {P^{\tau}_{\phi}}} = -{1\over {2P^{\tau}_{\phi}}} \int d^3\sigma \,\, \vec \sigma \, [\pi^2+(\vec \partial \phi )^2+m^2\phi^2](\tau ,\vec \sigma )=\nonumber \\ &=&-{2\over {P^{\tau}_{\phi}}} \int d{\tilde q}_1d{\tilde q}_2 {\bf A}_{{\vec q}_1}(\tau ;P^A_{\phi},{\bf H}] {\bf A}_{{\vec q}_2}(\tau ;P^A_{\phi},{\bf H}] \int d^3\sigma \, \vec \sigma \nonumber \\ &&\Big([\omega (q_1)\omega (q_2)+{\vec q}_1\cdot {\vec q}_2] sin\, ({\vec q}_1 \cdot \vec \sigma +{\bf B}_{{\vec q}_1}(\tau ;X^A_{\phi},{\bf K}]) sin\, ({\vec q}_2 \cdot \vec \sigma +{\bf B}_{{\vec q}_2}(\tau ;X^A_{\phi},{\bf K}]) +\nonumber \\ &+&m^2 cos\, ({\vec q}_1 \cdot \vec \sigma +{\bf B}_{{\vec q}_1}(\tau ;X^A _{\phi},{\bf K}]) cos\, ({\vec q}_2 \cdot \vec \sigma +{\bf B}_{{\vec q}_2}(\tau ;X^A_{\phi},{\bf K}]) \Big), \nonumber \\ &&{}\nonumber \\ {\vec \Omega}_{\phi} &=& {\vec J}_{\phi} -{\vec r}_{\phi}\times {\vec P} _{\phi},\nonumber \\ &&\{ r^r_{\phi},P^s_{\phi} \} =\delta^{rs},\quad\quad \{ r^r_{\phi},P^{\tau} _{\phi} \} ={{P^r_{\phi}}\over {P^{\tau}_{\phi}}},\nonumber \\ &&\{ r^r_{\phi},r^s_{\phi} \} =-{1\over {(P^{\tau}_{\phi})^2}} \epsilon^{rsu} \Omega^u_{\phi},\nonumber \\ &&\{ \Omega^r_{\phi},\Omega^s_{\phi} \} =\epsilon^{rsu}(\Omega^u_{\phi}-{1\over {(P^{\tau}_{\phi})^2}}({\vec \Omega}_{\phi} \cdot {\vec P}_{\phi})\,\, P_{\phi} ^u),\quad\quad \{ \Omega^r_{\phi},P^{\tau}_{\phi} \} =0. \label{VI1} \end{eqnarray} \noindent We see that ${\vec r}_{\phi}$ coincides with Dixon's definition of Eq.(\ref{V10}). The canonical ``internal" center of mass [$\{ q^r_{\phi},q^s_{\phi} \} =0$, $\{ q^r_{\phi},P^s_{\phi} \} = \delta^{rs}$, $\{ J^r_{\phi},q^s_{\phi} \} =\epsilon^{rsu}q^u_{\phi}$] is \begin{eqnarray} {\vec q}_{\phi}&=& {\vec r}_{\phi}- {{{\vec J}_{\phi}\times {\vec \Omega} _{\phi}}\over {\sqrt{(P^{\tau}_{\phi})^2-{\vec P}^2_{\phi}}(P^{\tau}_{\phi}+ \sqrt{(P^{\tau}_{\phi})^2-{\vec P}^2_{\phi}})}}= \nonumber \\ &=&-{{{\vec K}_{\phi}}\over {\sqrt{(P^{\tau}_{\phi})^2-{\vec P}^2_{\phi}}}}+ {{{\vec J}_{\phi}\times {\vec P}_{\phi}}\over {\sqrt{(P^{\tau}_{\phi})^2-{\vec P}^2_{\phi}}(P^{\tau}_{\phi}+\sqrt{(P^{\tau}_{\phi})^2 -{\vec P}^2_{\phi}})}}+\nonumber \\ &+&{{({\vec K}_{\phi}\cdot {\vec P}_{\phi})\,\, {\vec P}_{\phi}}\over {P^{\tau} _{\phi}\sqrt{(P^{\tau}_{\phi})^2-{\vec P}^2_{\phi}}\Big( P^{\tau}_{\phi}+ \sqrt{(P^{\tau}_{\phi})^2-{\vec P}^2_{\phi}}\Big) }},\nonumber \\ &&\approx {\vec r}_{\phi}\quad for\quad {\vec P}_{\phi}\approx 0;\quad\quad \{ {\vec q}_{\phi},P^{\tau}_{\phi} \} ={{{\vec P}_{\phi}}\over {P^{\tau} _{\phi}}}\approx 0,\nonumber \\ &&{}\nonumber \\ {\vec S}_{q \phi} &=&{\vec J}_{\phi}-{\vec q}_{\phi}\times {\vec P}_{\phi}= \nonumber \\ &=& {{P^{\tau}_{\phi}{\vec J}_{\phi}}\over {\sqrt{(P^{\tau}_{\phi})^2-{\vec P} ^2_{\phi}}}}+{{{\vec K}_{\phi}\times {\vec P}_{\phi}}\over {\sqrt{(P^{\tau} _{\phi})^2-{\vec P}^2_{\phi}}}}-{{({\vec J}_{\phi}\cdot {\vec P}_{\phi})\,\, {\vec P}_{\phi}}\over {\sqrt{(P^{\tau}_{\phi})^2-{\vec P}^2_{\phi}}\Big( P^{\tau}_{\phi}+\sqrt{(P^{\tau}_{\phi})^2-{\vec P}^2_{\phi}}\Big) }}\approx \nonumber \\ &\approx& {\vec S}_{\phi}, \quad for\quad {\vec P}_{\phi}\approx 0,\nonumber \\ &&\{ {\vec S}_{q \phi},{\vec P}_{\phi} \} =\{ {\vec S}_{q \phi},{\vec q}_{\phi} \} =0,\quad\quad \{ S^r_{q \phi},S^s_{q \phi} \} =\epsilon^{rsu}S^u_{q \phi}. \label{VI2} \end{eqnarray} The ``internal" noncanonical Fokker-Pryce center of inertia' ${\vec y}_{\phi}$ is \begin{eqnarray} {\vec y}_{\phi}&=& {\vec q}_{\phi}+{{{\vec S}_{q \phi}\times {\vec P}_{\phi}} \over {\sqrt{(P^{\tau}_{\phi})^2-{\vec P}_{\phi}^2} (P^{\tau}_{\phi}+ \sqrt{(P^{\tau}_{\phi})^2-{\vec P}_{\phi}^2})}} ={\vec r}_{\phi}+{{{\vec S}_{q \phi}\times {\vec P}_{\phi}}\over {P^{\tau} _{\phi}\sqrt{(P^{\tau}_{\phi})^2-{\vec P}_{\phi}^2}}},\nonumber \\ &&{}\nonumber \\ {\vec q}_{\phi}&=&{\vec r}_{\phi}+{{{\vec S}_{q \phi}\times {\vec P}_{\phi}} \over {P^{\tau}_{\phi}(P^{\tau}_{\phi}+\sqrt{(P^{\tau}_{\phi})^2-{\vec P} _{\phi}^2})}} = {{P^{\tau}_{\phi}{\vec r}_{\phi}+\sqrt{(P^{\tau}_{\phi})^2- {\vec P}_{\phi}^2} {\vec y}_{\phi}}\over {P^{\tau}_{\phi}+\sqrt{(P^{\tau} _{\phi})^2-{\vec P}_{\phi}^2}}},\nonumber \\ &&\{ y^r_{\phi},y^s_{\phi} \} ={1\over {P^{\tau}_{\phi}\sqrt{(P^{\tau}_{\phi}) ^2-{\vec P}_{\phi}^2} }}\epsilon^{rsu}\Big[ S^u_{q \phi}+{{ ({\vec S}_{q \phi} \cdot {\vec P}_{\phi})\, P^u_{\phi}}\over {\sqrt{(P^{\tau}_{\phi})^2-{\vec P} _{\phi}^2}(P^{\tau}_{\phi}+\sqrt{(P^{\tau}_{\phi})^2-{\vec P}_{\phi}^2})}} \Big] ,\nonumber \\ &&{}\nonumber \\ {\vec P}_{\phi}\approx 0 &\Rightarrow& {\vec q}_{\phi}\approx {\vec r}_{\phi} \approx {\vec y}_{\phi}. \label{VI3} \end{eqnarray} The Wigner 3-vector ${\vec q}_{\phi}$ is therefore the canonical 3-center of mass of the Klein-Gordon field [since ${\vec q}_{\phi}\approx {\vec r}_{\phi}$, it also describe that point $z^{\mu}(\tau ,{\vec q}_{\phi})=x^{\mu}_s(\tau )+ q^r_{\phi} \epsilon^{\mu}_r(u(p_s))$ where the energy of the field configuration is concentrated]. Instead, the previous variable ${\vec X}_{\phi}$ should be better named the 'center of phase' of the field configuration. There should exist a canonical transformation from the canonical basis ${\tilde X}^{\tau}_{\phi}$, $P^{\tau}_{\phi}$, ${\vec X}_{\phi}$, ${\vec P} _{\phi}$, ${\bf H}(\tau ,\vec k)$, ${\bf K}(\tau ,\vec k)$, to a new basis $q^{\tau}_{\phi}$, $P^{\tau \, '}_{\phi} =\sqrt{ (P^{\tau}_{\phi})^2-{\vec P}^2 _{\phi}}$ [since $\{ {\vec q}_{\phi},P^{\tau}_{\phi} \} = {\vec P}_{\phi}/P^{\tau}_{\phi}$ only weakly zero], ${\vec q}_{\phi}$, ${\vec P}_{\phi}$, ${\bf H}^{'}(\tau ,\vec k)$, ${\bf K}^{'}(\tau ,\vec k)$ containing relative variables with respect to the true center of mass of the field configuration. However it does not seem easy to identify this final canonical basis; in particular it is not clear how to find the `time' variable $q^{\tau}_{\phi}$. Maybe the methods used in Ref.\cite{iten} can be extended from particles to fields. The gauge fixing ${\vec q}_{\phi}\approx 0$ [it also implies $\vec \lambda (\tau )=0$ like ${\vec X}_{\phi}\approx 0$] forces all three internal center-of-mass variables to coincide with the origin $x^{\mu}_s$ of the Wigner hyperplane. We shall denote $x_s^{({\vec q}_{\phi})\mu}(\tau )=x^{\mu}_s(0)+\tau u^{\mu}(p_s)$ the origin in this gauge, because, as we shall see, in this case it enjoys of properties not present in the ${\vec X}_{\phi}\approx 0$ gauge [where we have that $x^{({\vec X}_{\phi})\mu}_s(\tau )=x ^{\mu}_s(0)+\tau u^{\mu}_s(p_s)$ is formally equal to $x^{({\vec q}_{\phi})\mu} _s(\tau )$, but, since ${\tilde t}_T^{\mu}(T_s)\not= 0$, $u_{\mu}(p_s) S_T^{\mu\nu}(T_s)[\phi ]\not= 0$, it has to be interpreted as a different centroid]. In the gauge ${\vec q}_{\phi}\approx 0$ the origin becomes the Dixon's center of mass which coincide with the internal Moller center of energy as shown in Eq.(\ref{V10}) [and then Eq.(\ref{V14}) implies that the origin is also the Tulczyjew\cite{tul} and the Pirani \cite{pirani} centroid (see Ref.\cite{bini} for a review of these concepts in relation with the Papapetrou-Dixon-Souriau pole-dipole approximation of an extended object), because they are defined by $p_{s\mu}S^{\mu\nu}_T(T_s)[\phi ]\approx 0$ and by $u_{\mu}(p_s) S_T^{\mu\nu}(T_s)[\phi ]\approx 0$ respectively and we have $p^{\mu}_s =\epsilon_s u^{\mu}(p_s)$]. Therefore, the worldline $x^{({\vec q}_{\phi})\mu} _s$ is the unique center-of-mass worldline of special relativity in the sense of Refs.\cite{bei}. The rest-frame instant form ``external" realization of the Poincar\'e algebra of Eq. (\ref{II9}) has the generators $p^{\mu}_s$, $J^{ij}_s={\tilde x}^i_sp^j_s-{\tilde x}^j_sp^i_s+\delta^{ir} \delta^{js} S^{rs}_{\phi}$, $K^i_s=J^{oi}_s={\tilde x}^o_sp^i_s- {\tilde x}^i_sp^o_s-{{\delta^{ir} S^{rs}_{\phi}\, p^s_s}\over {p^o_s+ \epsilon_s}}={\tilde x}^o_sp^i_s-{\tilde x}^i_sp^o_s+\delta^{ir}{{({\vec S} _{\phi}\times {\vec p}_s)^r}\over {p^o_s+\epsilon_s}}$ [for ${\tilde x}^o_s=0$ this is the Newton-Wigner decomposition of $J^{\mu\nu}_s$]. As already said the canonical variables ${\tilde x}^{\mu}_s$, $p^{\mu}_s$, may be replaced by the canonical pairs $\epsilon_s=\pm \sqrt{p^2_s}$, $T_s=p_s\cdot {\tilde x}_s/ \epsilon_s$ [to be gauge fixed with $T_s-\tau \approx 0$]; ${\vec k}_s={\vec p}_s/\epsilon_s=\vec u(p_s)$, ${\vec z}_s= \epsilon_s ({\vec {\tilde x}}_s- {{{\vec p}_s}\over {p^o_s}} {\tilde x}^o_s)\equiv \epsilon_s {\vec q}_s$. One can build three ``external" 3-variables, the canonical ${\vec Q}_s$, the Moller ${\vec R}_s$ and the Fokker-Pryce ${\vec Y} _s$ by using this rest-frame ``external" realization of the Poincar\'e algebra \begin{eqnarray} {\vec R}_s&=& -{1\over {p^o_s}}{\vec K}_s=({\vec {\tilde x}}_s-{{{\vec p}_s}\over {p^o_s}} {\tilde x}^o_s)-{{{\vec S}_{\phi}\times {\vec p}_s} \over {p^o_s(p^o_s+\epsilon_s)}},\nonumber \\ {\vec Q}_s&=&{\vec {\tilde x}}_s-{{{\vec p}_s}\over {p^o_s}}{\tilde x}^o_s= {{{\vec z}_s}\over {\epsilon_s}}= {\vec R}_s+{{ {\vec S}_{\phi}\times {\vec p}_s}\over {p^o_s(p^o_s+ \epsilon_s)}}={{p^o_s {\vec R}_s+\epsilon_s {\vec Y}_s}\over {p^o_s+\epsilon_s}} ,\nonumber \\ {\vec Y}_s&=&{\vec Q}_s+{{ {\vec S}_{\phi}\times {\vec p}_s}\over {\epsilon_s (p^o_s+\epsilon_s)}}={\vec R}_s+{{ {\vec S}_{\phi}\times {\vec p}_s}\over {p^o_s\epsilon_s}},\nonumber \\ &&\{ R^r_s,R^s_s \} =-{1\over {(p^o_s)^2}}\epsilon^{rsu}\Omega^u_s, \quad\quad {\vec \Omega}_s={\vec J}_s-{\vec R}_s\times {\vec p}_s,\nonumber \\ &&\{ Y^r_s,Y^s_s \} ={1\over {\epsilon_sp^o_s}}\epsilon^{rsu}\Big[ S^u_{\phi} +{{ ({\vec S}_{\phi}\cdot {\vec p}_s)\, p^u_s}\over {\epsilon_s(p^o_s+ \epsilon_s)}}\Big] ,\nonumber \\ &&{}\nonumber \\ &&{\vec p}_s\cdot {\vec Q}_s={\vec p}_s\cdot {\vec R}_s={\vec p}_s\cdot {\vec Y}_s={\vec k}_s\cdot {\vec z}_s,\nonumber \\ {\vec p}_s=0 &\Rightarrow& {\vec Q}_s={\vec Y}_s={\vec R}_s, \label{VI4} \end{eqnarray} \noindent with the same velocity and coinciding in the Lorentz rest frame where ${\buildrel \circ \over p}^{\mu}_s=\epsilon_s (1;\vec 0)$ In Ref.\cite{pauri} in a one-time framework without constraints and at a fixed time, it is shown that the 3-vector ${\vec Y}_s$ [but not ${\vec Q}_s$ and ${\vec R}_s$] satisfies the condition $\{ K^r_s,Y^s_s \} = Y^r_s\, \{ Y^s_s,p^o_s \}$ for being the space component of a 4-vector $Y^{\mu}_s$. In the enlarged canonical treatment including time variables, it is not clear which are the time components to be added to ${\vec Q}_s$, ${\vec R} _s$, ${\vec Y}_s$, to rebuild 4-dimesnional quantities ${\tilde x}^{\mu}_s$, $R^{\mu}_s$, $Y^{\mu}_s$, in an arbitrary Lorentz frame $\Gamma$, in which the origin of the Wigner hyperplane is the 4-vector $x^{\mu}_s = (x^o_s; {\vec x} _s)$. We have \begin{eqnarray} {\tilde x}^{\mu}_s(\tau )&=& ({\tilde x}^o_s(\tau ); {\vec {\tilde x}}_s(\tau ) )= x^{\mu}_s-{1\over {\epsilon_s(p^o_s+\epsilon_s)}}\Big[ p_{s\nu}S_s^{\nu\mu}+\epsilon_s(S^{o\mu}_s-S^{o\nu}_s{{p_{s\nu}p_s^{\mu}}\over {\epsilon^2_s}}) \Big],\quad\quad p^{\mu}_s,\nonumber \\ {\tilde x}^o_s&=&\sqrt{1+{\vec k}_s^2} (T_s+{{{\vec k}_s\cdot {\vec z}_s}\over {\epsilon_s}})=\sqrt{1+{\vec k}^2_s}(T_s+{\vec k}_s\cdot {\vec q}_s)\not= x^0_s,\quad\quad p^o_s=\epsilon_s\sqrt{1+{\vec k}_s^2},\nonumber \\ {\vec {\tilde x}}_s&=&{{ {\vec z}_s}\over {\epsilon_s}}+(T_s+{{{\vec k}_s\cdot {\vec z}_s}\over {\epsilon_s}}) {\vec k}_s={\vec q}_s+(T_s+{\vec k}_s\cdot {\vec q}_s){\vec k}_s,\quad\quad {\vec p}_s=\epsilon_s {\vec k}_s. \label{VI5} \end{eqnarray} \noindent for the non-covariant (frame-dependent) canonical center of mass and its conjugate momentum. Each Wigner hyperplane intersects the worldline of the arbitrary origin 4-vector $x^{\mu}_s(\tau )=z^{\mu}(\tau ,\vec 0)$ in $\vec \sigma =0$, the pseudo worldline of ${\tilde x}^{\mu}_s(\tau )=z^{\mu}(\tau ,{\tilde {\vec \sigma}})$ in some ${\tilde {\vec \sigma}}$ and the worldline of the Fokker-Pryce 4-vector $Y^{\mu}_s(\tau )=z^{\mu}(\tau ,{\vec \sigma}_Y)$ in some ${\vec \sigma}_Y$ [on this worldline one can put the ``internal center of mass" with the gauge fixing ${\vec q}_{\phi}\approx 0$ (${\vec q}_{\phi}\approx {\vec r}_{\phi}\approx {\vec y}_{\phi}$ due to ${\vec P}_{\phi}\approx 0$)]; one also has $R^{\mu}_s=z^{\mu}(\tau ,{\vec \sigma}_R)$. Since we have $T_s=u(p_s)\cdot x_s=u(p_s)\cdot {\tilde x}_s\equiv \tau$ on the Wigner hyperplane labelled by $\tau$, we require that also $Y^{\mu}_s$, $R^{\mu}_s$ have time components such that they too satisfy $u(p_s)\cdot Y_s=u(p_s)\cdot R_s=T_s\equiv \tau$. Therefore, it is reasonable to assume that ${\tilde x}^{\mu}_s$, $Y^{\mu}_s$ and $R^{\mu}_s$ satisfy the following equations consistently with Eqs.(\ref{VI1}), (\ref{VI2}) when $T_s\equiv \tau$ and ${\vec q}_{\phi}\approx 0$ \begin{eqnarray} {\tilde x}^{\mu}_s&=&( {\tilde x}^o_s; {\vec {\tilde x}}_s)=({\tilde x}^o_s; {{{\vec z}_s}\over {\epsilon_s}}+(T_s+{{{\vec k}_s\cdot {\vec z}_s}\over {\epsilon_s}}){\vec k}_s ) =x^{({\vec q}_{\phi})\mu}_s+\epsilon^{\mu}_u(u(p_s)) {\tilde \sigma}^u, \nonumber \\ Y^{\mu}_s&=&({\tilde x}^o_s;\, {1\over {\epsilon_s}}[{\vec z}_s+{{{\vec S}_{\phi}\times {\vec p}_s}\over {\epsilon_s[1+u^o(p_s)]}}]+(T_s+ {{{\vec k}_s\cdot {\vec z}_s}\over {\epsilon_s}}){\vec k}_s\, )=\nonumber \\ &=&{\tilde x}^{\mu}_s+\eta^{\mu}_r{{({\vec S}_{\phi}\times {\vec p}_s)^r}\over {\epsilon_s[1+u^o(p_s)]}}=\nonumber \\ &=&x^{({\vec q}_{\phi})\mu}_s+\epsilon^{\mu}_u(u(p_s)) \sigma^u_Y,\nonumber \\ R^{\mu}_s&=&( {\tilde x}^o_s;\, {1\over {\epsilon_s}}[{\vec z}_s- {{{\vec S}_{\phi}\times {\vec p}_s}\over {\epsilon_s u^o(p_s) [1+u^o(p_s)]}}]+(T_s+ {{{\vec k}_s\cdot {\vec z}_s}\over {\epsilon_s}}){\vec k}_s\, )=\nonumber \\ &=&{\tilde x}^{\mu}_s-\eta^{\mu}_r{{({\vec S}_{\phi}\times {\vec p}_s)^r} \over {\epsilon_su^o(p_s)[1+u^o(p_s)]}}=\nonumber \\ &=&x^{({\vec q}_{\phi})\mu}_s+\epsilon^{\mu}_u(u(p_s)) \sigma^u_R,\nonumber \\ &&{}\nonumber \\ T_s&=&u(p_s)\cdot x_s^{({\vec q}_{\phi})}=u(p_s)\cdot {\tilde x}_s=u(p_s)\cdot Y_s=u(p_s)\cdot R_s,\nonumber \\ &&{}\nonumber \\ {\tilde \sigma}^r&=&\epsilon_{r\mu}(u(p_s)) [x^{({\vec q}_{\phi})\mu}_s-{\tilde x}^{\mu}_s]= {{ \epsilon_{r\mu}(u(p_s)) [u_{\nu}(p_s)S^{\nu\mu}_s+S^{o\mu}_s]}\over {[1+u^o(p_s)]}}=\nonumber \\ &=&-S_{\phi}^{\tau r}+{{S_{\phi}^{rs}p^s_s}\over {\epsilon_s[1+u^o(p_s)]}} =\epsilon_s r^r_{\phi}+{{S_{\phi}^{rs}u^s(p_s)}\over {1+u^o(p_s)}} \approx \nonumber \\ &\approx& \epsilon_s q^r_{\phi}+{{S_{\phi}^{rs}u^s(p_s)}\over {1+ u^o(p_s)}}\approx {{S_{\phi}^{rs}u^s(p_s)}\over {1+u^o(p_s)}} ,\nonumber \\ \sigma^r_Y&=&\epsilon_{r\mu}(u(p_s))[x^{({\vec q}_{\phi})\mu}_s-Y^{\mu}_s]= {\tilde \sigma}^r-\epsilon_{ru}(u(p_s)){{({\vec S}_{\phi}\times {\vec p}_s)^u}\over {\epsilon_s[1+u^o(p_s)]}}=\nonumber \\ &=&{\tilde \sigma}^r+{{S^{rs}_{\phi}u^s(p_s)}\over {1+u^o(p_s)}}= \epsilon_s r^r_{\phi} \approx \epsilon_s q^r_{\phi} \approx 0,\nonumber \\ \sigma^r_R&=&\epsilon_{r\mu}(u(p_s)) [x^{({\vec q}_{\phi})\mu}_s-R^{\mu}_s]={\tilde \sigma}^r+ \epsilon_{ru}(u(p_s)) {{({\vec S}_{\phi}\times {\vec p}_s)^u} \over {\epsilon_su^o(p_s)[1+u^o(p_s)]}}=\nonumber \\ &=&{\tilde \sigma}^r-{{S_{\phi}^{rs}u^s(p_s)}\over {u^o(p_s)[1+ u^o(p_s)]}}=\epsilon_sr^r_{\phi}+{{[1-u^o(p_s)]S^{rs}_{\phi}u^s(p_s)}\over {u^o(p_s)[1+u^o(p_s)]}}\approx \nonumber \\ &\approx& {{[1-u^o(p_s)]S^{rs}_{\phi}u^s(p_s)}\over {u^o(p_s)[1+u^o(p_s)]}},\nonumber \\ &&{}\nonumber \\ &\Rightarrow& x^{({\vec q}_{\phi})\mu}_s(\tau ) = Y^{\mu}_s,\quad for\quad {\vec q}_{\phi}\approx 0, \label{VI6} \end{eqnarray} \noindent namely in the gauge ${\vec q}_{\phi}\approx 0$ the external Fokker-Pryce non canonical center of inertia coincides with the origin $x^{({\vec q}_{\phi})\mu}_s(\tau )$ carrying the ``internal" center of mass (coinciding with the ``internal" M\"oller center of energy and with the ``internal" Fokker-Pryce center of inertia) and also being the Pirani centroid and the Tulczyjew centroid. Therefore, if we would find the canonical basis $q^{\tau}_{\phi}$, $P^{\tau \, {'}}_{\phi}= \sqrt{(P^{\tau}_{\phi})^2-{\vec P}^2_{\phi}}$, ${\vec q}_{\phi}$, ${\vec P} _{\phi}$, ${\bf H}^{'}(\tau ,\vec q)$, ${\bf K}^{'}(\tau ,\vec q)$, then, in the gauge ${\vec q}_{\phi}\approx 0$ and $T_s \approx \tau$, the Klein-Gordon field configurations would have the four-momentum density peaked on the worldline $x^{({\vec q}_{\phi})\mu}_s(T_s)$; the canonical variables ${\bf H}^{'}(\tau ,\vec q)$, ${\bf K}^{'}(\tau ,\vec q)$ would characterize the relative motions with respect to the ``monopole" configuration describing the center of mass of the field configuration. The ``monopole" solutions of the Klein-Gordon equation would be identified by the conditions ${\bf H} ^{'}(\tau ,\vec q)={\bf K}^{'}(\tau ,\vec q)=0$ [formally they are given by Eqs.(\ref{IV38}) with $X^{\tau}_{\phi}$, ${\vec X}_{\phi}$ and $P^{\tau} _{\phi}$ replaced by $q^{\tau}_{\phi}$, ${\vec q}_{\phi}$ and $\sqrt{(P^{\tau} _{\phi})^2-{\vec P}^2_{\phi}}$]: these field configurations have the same independent degrees of freedom of a free scalar particle at rest with mass $P^{\tau \, {'}}_{\phi}\approx P^{\tau}_{\phi}$ [its conjugate ``time" $q^{\tau}_{\phi}$ would satisfy $\partial /\partial T_s\, {\buildrel \circ \over =}\, \partial /\partial q^{\tau}_{\phi}$ in the free case, see Subsection D of Section IV]. Remember that the canonical center of mass lies in between the Moller center of energy and the Fokker-Pryce center of inertia and that the noncovariance region around the Fokker-Pryce 4-vector extends to a worldtube with radius (the Moller radius) $|{\vec S}_{\phi}| / P^{\tau}_{\phi}$. \vfill\eject \section{Coupling to scalar particles.} In this Section we shall consider a Yukawa-type coupling of the Klein-Gordon field to the masses of scalar relativistic particles\cite{lus} to find which are the Dirac observables spanning the reduced phase space. The action is \begin{eqnarray} S&=& \int d\tau L(\tau )= \int d\tau d^3\sigma \Big( N(\tau ,\vec \sigma ) \sqrt{\gamma (\tau ,\vec \sigma )}\nonumber \\ &&{1\over 2} \Big[ g^{\tau\tau} {\dot \phi}^2+2 g^{\tau \check r} \dot \phi \partial_{\check r}\phi + g^{\check r\check s}\partial_{\check r}\phi \partial_{\check s}\phi -m^2\phi^2 \Big] (\tau ,\vec \sigma )-\nonumber \\ &-&\sum_{i=1}^N\delta^3(\vec \sigma -{\vec \eta}_i(\tau )) \Big[\eta_im_i+G\phi (\tau ,\vec \sigma )]\sqrt{[g_{\tau\tau}+2g_{\tau \check r}{\dot \eta}_i ^{\check r}(\tau )+g_{\check r\check s}{\dot \eta}_i^{\check r}(\tau ){\dot \eta}_i^{\check s}(\tau )\Big] (\tau ,\vec \sigma )} \Big)=\nonumber \\ &=&\int d\tau d^3\sigma \Big( \sqrt{\gamma (\tau ,\vec \sigma )} {1\over 2} \Big[ {1\over N}[\partial_{\tau} -N^{\check r}\partial_{\check r}]\phi \, [\partial_{\tau} -N^{\check s}\partial_{\check s}]\phi + \nonumber \\ &+&N [\gamma^{\check r\check s}\partial_{\check r}\phi \partial_{\check s}\phi -m^2\phi^2 ]\, \Big] (\tau ,\vec \sigma ) -\sum_{i=1}^N\delta^3(\vec \sigma -{\vec \eta}_i(\tau )) [\eta_im_i+G\phi (\tau ,\vec \sigma )]\nonumber \\ &&\sqrt{ [N^2+g_{\check r\check s}(N^{\check r}+{\dot \eta}_i^{\check r}(\tau )) (N^{\check s}+{\dot \eta}_i^{\check s}(\tau ))](\tau ,\vec \sigma )} \Big) , \nonumber \\ &&{} \label{VII1} \end{eqnarray} \noindent and the canonical momenta are \begin{eqnarray} \kappa_{i\check r}(\tau )&=&-{{\partial L(\tau )}\over {\partial {\dot \eta} _i^{\check r}}}= \Big[ \eta_im_i+G\phi (\tau ,{\vec \eta}_i(\tau ))\Big] \nonumber \\ &&{{ g_{\check r\check s} (N^{\check s}+{\dot \eta}_i^{\check s}(\tau ))}\over {\sqrt{N^2+g_{\check r\check s}(N^{\check r}+{\dot \eta}_i^{\check r}(\tau )) (N^{\check s}+{\dot \eta}_i^{\check s}(\tau ))} }}(\tau ,{\vec \eta}_i(\tau )), \nonumber \\ \pi(\tau ,\vec \sigma )&=&{{\partial L}\over {\partial \partial_{\tau} \phi (\tau ,\vec \sigma )}}=\nonumber \\ &=&{{\sqrt{\gamma}(\tau ,\vec \sigma )}\over {N(\tau ,\vec \sigma )}}\Big[ \dot \phi -N^{\check r}\partial_{\check r}\phi \Big] (\tau ,\vec \sigma ),\nonumber \\ \rho_{\mu}(\tau ,\vec \sigma )&=&-{{\partial L}\over {\partial \partial_{\tau} z^{\mu}(\tau ,\vec \sigma )}}=\nonumber \\ &=&l_{\mu}(\tau ,\vec \sigma ) \Big[\, \Big( {{\sqrt{\gamma}}\over 2} [{1\over {N^2}}(\dot \phi -N^{\check r}\partial_{\check r}\phi )^2- \gamma^{\check r\check s}\partial_{\check r}\phi \partial_{\check s}\phi +m^2\phi^2]\, \Big) (\tau ,\vec \sigma ) +\nonumber \\ &+&\sum_{i=1}^N\delta^3(\vec \sigma -{\vec \eta}_i(\tau )) {{[\eta_im_i+G\phi (\tau ,\vec \sigma )]\, N(\tau ,\vec \sigma )}\over { \sqrt{[N^2+g_{\check r\check s}(N^{\check r}+{\dot \eta}_i^{\check r}(\tau )) (N^{\check s}+{\dot \eta}_i^{\check s}(\tau ))](\tau ,\vec \sigma )} }}\, \Big] +\nonumber \\ &+&z_{\check s\mu}(\tau ,\vec \sigma )\gamma^{\check s\check r}(\tau ,\vec \sigma ) \Big[ \Big( {{ \sqrt{\gamma}}\over N}\partial_{\check r}\phi (\dot \phi -N^{\check u}\partial_{\check u}\phi )\, \Big) (\tau ,\vec \sigma )+\nonumber \\ &+&\sum_{i=1}^N\delta^3(\vec \sigma -{\vec \eta}_i(\tau )) {{[\eta_im_i+G\phi ] g_{\check r\check s}(N^{\check u}+{\dot \eta}_i^{\check u}(\tau ))}\over {\sqrt{N^2+g_{\check r\check s}(N^{\check r}+{\dot \eta}_i^{\check r}(\tau )) (N^{\check s}+{\dot \eta}_i^{\check s}(\tau ))} }} (\tau ,\vec \sigma )\, \Big] . \label{VII2} \end{eqnarray} We get the first class constraints \begin{eqnarray} {\cal H}_{\mu}(\tau ,\vec \sigma )&=&\rho_{\mu}(\tau ,\vec \sigma )- \nonumber \\ &&-l_{\mu}(\tau ,\vec \sigma ) \Big[ \Big( {{\pi^2}\over {2\sqrt{\gamma}}}- {{\sqrt{\gamma}}\over 2} [\gamma^{\check r\check s}\partial_{\check r}\phi \partial_{\check s}\phi -m^2\phi^2] \Big) (\tau ,\vec \sigma )+ \nonumber \\ &+&\sum_{i=1}^N\delta^3(\vec \sigma -{\vec \eta}_i(\tau ))\eta_i\sqrt{[\eta _im_i+G\phi (\tau ,\vec \sigma )]^2-\gamma^{\check r\check s}(\tau ,\vec \sigma )\kappa_{i\check r}(\tau )\kappa_{i\check s}(\tau )} \Big] -\nonumber \\ &-&z_{\check s\mu}(\tau ,\vec \sigma )\gamma^{\check r\check s}(\tau ,\vec \sigma ) \Big[ (\pi \partial_{\check r}\phi )(\tau ,\vec \sigma ) +\sum_{i=1}^N \delta^3(\vec \sigma -{\vec \eta}_i(\tau ))\kappa_{i\check r}(\tau )\Big] \approx 0. \label{VII3} \end{eqnarray} Following the procedure of Section II, one can arrive at the reduction on the Wigner hyperplanes, where the remaining four first class constraints and Dirac Hamiltonian are \begin{eqnarray} {\cal H}(\tau )&=&\epsilon_s - \Big[ \sum_{i=1}^N\eta_i\sqrt{[\eta_im_i+G\phi (\tau ,{\vec \eta}_i(\tau ))]^2+{\vec \kappa}_i^2(\tau )}+\nonumber \\ &+& {1\over 2} \int d^3\sigma [\pi^2+(\vec \partial \phi )^2+m^2\phi^2](\tau ,\vec \sigma ) \Big] = \epsilon_s-M \approx 0,\nonumber \\ {\vec {\cal H}}_p(\tau )&=& \sum_{i=1}^N {\vec \kappa}_i(\tau )+ \int d^3\sigma [\pi \vec \partial \phi ](\tau ,\vec \sigma ) \approx 0, \nonumber \\ &&{}\nonumber \\ H_D&=& \lambda (\tau ) {\cal H} + \vec \lambda (\tau )\cdot {\vec {\cal H}}_p. \label{VII4} \end{eqnarray} In the gauge $T_s-\tau \approx 0$ the Hamiltonian is $H_R=M -\vec \lambda (\tau )\cdot {\vec {\cal H}}_p$. Let us make a canonical transformation from the canonical basis ${\vec \eta}_i (\tau )$, ${\vec \kappa}_i(\tau )$, $\phi (\tau ,\vec \sigma )$, $\pi (\tau ,\vec \sigma )$, to the new basis ${\vec \eta}_{+}(\tau )$, ${\vec \kappa} _{+}(\tau )$, ${\vec \rho}_a(\tau )$, ${\vec \pi}_a(\tau )$, $X^{\tau}_{\phi}$, $P^{\tau}_{\phi}$, ${\vec X}_{\phi}$, ${\vec P}_{\phi}$, ${\bf H}(\tau ,\vec q)$, ${\bf K}(\tau ,\vec q)$ [$\phi$, $\pi$, are assumed to satisfy Eq.(\ref{IV29})] with the particle variables defined by \begin{eqnarray} {\vec \eta}_i&=&{\vec \eta}_{+}+{1\over {\sqrt{N}}} \sum_{a=1}^{N-1}{\hat \gamma}_{ai}{\vec \rho}_a,\nonumber \\ {\vec \kappa}_i&=&{1\over N} {\vec \kappa}_{+}+\sqrt{N} \sum_{a=1}^{N-1}{\hat \gamma}_{ai} {\vec \pi}_a,\nonumber \\ &&{}\nonumber \\ {\vec \eta}_{+}&=&{1\over N} \sum_{i=1}^N {\vec \eta}_i,\quad\quad {\vec \rho}_a=\sqrt{N} \sum_{i=1}^N {\hat \gamma} {\vec \eta}_i,\nonumber \\ {\vec \kappa}_{+}&=&\sum_{i=1}^N {\vec \kappa}_i,\quad\quad {\vec \pi}_a={1\over {\sqrt{N}}} \sum_{i=1}^N {\hat \gamma}_{ai} {\vec \kappa}_i,\nonumber \\ &&{}\nonumber \\ &&\sum_{i=1}^N{\hat \gamma}_{ai}=0,\quad\quad \sum_{a=1}^{N-1}{\hat \gamma}_{ai} {\hat \gamma}_{aj}=\delta_{ij}-{1\over N},\quad\quad \sum_{i=1}^N{\hat \gamma}_{ai}{\hat \gamma}_{bi}=\delta_{ab}. \label{VII5} \end{eqnarray} The variable ${\vec \eta}_{+}(\tau )$ is playing the role of a naive ``internal" center of mass for the particles on the Wigner hyperplane. From the discussion of the previous Section, it is clear that the real ``internal" center of mass of the N particles is a ${\vec q}_{+}$ defined like ${\vec q}_{\phi}$ of Eq.(\ref{VI2}) with ${\vec r}_{+}=\sum_{i=1}^N \sqrt{m^2_i+{\vec \kappa}^2_i}\, {\vec \eta}_i / \sum_{k=1}^N\sqrt{m^2_k+{\vec \kappa}^2_k}$. But, since it is not known (like it happens for the Klein-Gordon field) the canonical transformation ${\vec \eta}_i, {\vec \kappa}_i\, \mapsto {\vec q}_{+}, {\vec \kappa}_{+}, {\vec \rho}_{qa}, {\vec \pi}_{qa}$ identifying the real relative variables ${\vec \rho}_{qa}$, ${\vec \pi}_{qa}$ [see however Ref. \cite{iten}], we shall go on with the previous naive canonical transformation in the following discussion to see which kind of collective and relative variables emerge for the particles plus the real Klein-Gordon field. We also remark that we use the field ``center of phase" ${\vec X}_{\phi}$ and not the real internal center of mass ${\vec q}_{\phi}$, because the knowledge of the canonical basis containing ${\vec X}_{\phi}$ allows us to illustrate interpretational aspects which will hold also with the canonical basis containing ${\vec q}_{\phi}$ when it will be found. The constraints become \begin{eqnarray} {\cal H}(\tau )&=& \epsilon_s-\Big[ P^{\tau}_{\phi}+\sum_{i=1}^N \eta_i \Big( [\eta_im_i+G\phi (\tau,{\vec \eta}_i(\tau ))]^2+\nonumber \\ &+&[{1\over N}{\vec \kappa}_{+}(\tau )+\sqrt{N}\sum_{a=1}^{N-1}{\hat \gamma} _{ai} {\vec \pi}_a(\tau )]^2 \Big)^{1/2} \approx 0,\nonumber \\ {\vec {\cal H}}_p(\tau )&=&{\vec \kappa}_{+}+{\vec P}_{\phi}\approx 0, \label{VII6} \end{eqnarray} We can now replace the canonical variables ${\vec \eta}_{+}$, ${\vec \kappa} _{+}$, ${\vec X}_{\phi}$, ${\vec P}_{\phi}$ with the following canonical ones \begin{eqnarray} \vec Y&=&{1\over 2} ({\vec \eta}_{+}+{\vec X}_{\phi}),\nonumber \\ {\vec {\cal H}}_p&=&{\vec \kappa}_{+}+{\vec P}_{\phi} \approx 0,\nonumber \\ \vec \zeta &=&{\vec \eta}_{+}-{\vec X}_{\phi},\nonumber \\ {\vec \pi}_{\zeta}&=&{1\over 2} ({\vec \kappa}_{+}-{\vec P}_{\phi}),\nonumber \\ &&{}\nonumber \\ {\vec \eta}_{+}&=&{1\over 2}\vec \zeta +\vec Y,\nonumber \\ {\vec X}_{\phi}&=&-{1\over 2}\vec \zeta +\vec Y,\nonumber \\ {\vec \kappa}_{+}&=&{1\over 2} {\vec {\cal H}}_p+{\vec \pi}_{\zeta}\approx {\vec \pi}_{\zeta},\nonumber \\ {\vec P}_{\phi}&=&{1\over 2}{\vec {\cal H}}_p-{\vec \pi}_{\zeta}\approx -{\vec \pi}_{\zeta}. \label{VII7} \end{eqnarray} We see that $\vec Y$ is playing the role of the naive ``internal" center of mass of the full ``particles+field" system: it is the gauge variable conjugate to ${\vec {\cal H}}_p\approx 0$ The global relative variables $\vec \zeta (\tau )$, ${\vec \pi}_{\zeta}(\tau )$, describe the relative motion of the particle and the field naive centers of mass and rule the action-reaction between the two subsystems. With the natural gauge-fixings: $\vec Y\approx 0$, $T_s-\tau \approx 0$ [so that ${\vec \eta}_{+}\approx {1\over 2}\vec \zeta \approx -{\vec X}_{\phi}$] and the associated Dirac brackets, we get the reduced phase space ${\vec z}_s$, ${\vec k}_s$, ${\vec \zeta}$, ${\vec \pi}_{\zeta}$, ${\vec \rho}_a(\tau )$, ${\vec \pi}_a(\tau )$, $X^{\tau}_{\phi}$, $P^{\tau}_{\phi}$, ${\bf H}(T_s,\vec q)$, ${\bf K}(T_s,\vec q)$ with the evolution in $T_s\equiv \tau$ ruled by the Hamiltonian \begin{eqnarray} H&=&P^{\tau}_{\phi}+\sum_{i=1}^N \eta_i \Big( [\eta_im_i+G\phi (T_s,{1\over 2} \vec \zeta (T_s)+{1\over {\sqrt{N}}}\sum_{a=1}^{N-1}{\hat \gamma}_{ai}{\vec \rho}_a(T_s))]^2+\nonumber \\ &+&[{1\over N}{\vec \pi}_{\zeta}(T_s)+\sqrt{N}\sum_{a=1}^{N-1}{\hat \gamma} _{ai} {\vec \pi}_a(T_s)]^2 \Big)^{1/2}=\nonumber \\ &{\buildrel {def} \over =}\,& P^{\tau}_{\phi} +\sum_{i=1}^N \eta_i M_i, \nonumber \\ &&{}\nonumber \\ M^2_i&=&(\eta_im_i+G\phi )^2+[{{ {\vec \pi}_{\zeta}}\over N}+\sqrt{N}\sum_{a=1} ^{N-1}{\hat \gamma}_{ai} {\vec \pi}_a ]^2,\nonumber \\ &&{}\nonumber \\ \phi &=&\phi \Big( T_s,{1\over 2} \vec \zeta (T_s)+{1\over {\sqrt{N}}}\sum_{a=1}^{N-1}{\hat \gamma}_{ai}{\vec \rho}_a(T_s)\Big)=\nonumber \\ &=& 2 \int d\tilde q \sqrt{F^{\tau}(q)\omega (q)P^{\tau}_{\phi}+ F(q)\vec q\cdot {\vec \pi}_{\zeta}(T_s)+{\cal D}_{\vec q}{\bf H}(T_s,\vec q)}\nonumber \\ &&cos\, \Big[ \vec q\cdot \Big( \vec \zeta (T_s)+{1\over {\sqrt{N}}}\sum_{a=1}^{N-1}{\hat \gamma}_{ai}{\vec \rho}_a(T_s)\Big) -\nonumber \\ &-&\omega (q)(T_s-X^{\tau}_{\phi})+ \int d\tilde kd{\tilde k}^{'} {\bf K}(T_s, \vec k){\cal G}(\vec k,{\vec k}^{'})\triangle ({\vec k}^{'},\vec q) \Big]. \label{VII8} \end{eqnarray} The equations of motion for $X^{\tau}_{\phi}$, $P^{\tau}_{\phi}$, $\vec \zeta (\tau )$, ${\vec \pi}_{\zeta}(\tau )$, ${\vec \rho}_a(\tau )$, ${\vec \pi}_a(\tau )$, ${\bf H}(\tau ,\vec q)$, ${\bf K}(\tau ,\vec q)$ are [remember that $T_s-X^{\tau}_{\phi}=-{\tilde X}^{\tau}_{\phi}$] \begin{eqnarray} {{d \vec \zeta (T_s)}\over {dT_s}}&\, {\buildrel \circ \over =}\,& \{ \vec \zeta (T_s),H \} ={{\partial H}\over {\partial {\vec \pi}_{\zeta}(T_s)}}= \nonumber \\ &=&\sum_{i=1}^N {1\over {\eta_iM_i}} \Big[ {{ {\vec \pi}_{\zeta}}\over {N^2}}+ {1\over {\sqrt{N}}} \sum_{a=1}^{N-1}{\hat \gamma}_{ai}{\vec \pi}_a +\nonumber \\ &+&G(\eta_im_i+G\phi ) \int d\tilde q {{F(q)\vec q}\over {\sqrt{F^{\tau}(q) \omega (q) P^{\tau}_{\phi}+F(q)\vec q\cdot {\vec \pi}_{\zeta}(T_s)+{\cal D} _{\vec q}{\bf H}(T_s,\vec q)}}} \nonumber \\ &&cos\, \Big( \vec q\cdot (\vec \zeta (T_s)+ {1\over {\sqrt{N}}}\sum_{a=1}^{N-1}{\hat \gamma}_{ai}{\vec \rho}_{ai}(T_s))- \nonumber \\ &-&\omega (q)(T_s-X^{\tau}_{\phi})+\int d\tilde kd{\tilde k}^{'} {\bf K}(T_s, \vec k) {\cal G}(\vec k,{\vec k}^{'}) \triangle ({\vec k}^{'},\vec q)\Big) \Big] ,\nonumber \\ {{d {\vec \pi}_{\zeta}(T_s)}\over {dT_s}}&\, {\buildrel \circ \over =}\,& \{ {\vec \pi}_{\zeta}(T_s),H \} =-{{\partial H}\over {\partial \vec \zeta (T_s) }}=\nonumber \\ &=&2G \sum_{i=1}^N {{\eta_im_i+G\phi }\over {\eta_iM_i}} \int d\tilde q \vec q \sqrt{F^{\tau}(q)\omega (q) P^{\tau}_{\phi}+F(q)\vec q\cdot {\vec \pi} _{\zeta}(T_s)+{\cal D}_{\vec q}{\bf H}(T_s,\vec q)}\nonumber \\ &&sin\, \Big( \vec q\cdot (\vec \zeta (T_s)+ {1\over {\sqrt{N}}}\sum_{a=1}^{N-1}{\hat \gamma}_{ai}{\vec \rho}_{ai}(T_s))- \nonumber \\ &-&\omega (q)(T_s-X^{\tau}_{\phi})+\int d\tilde kd{\tilde k}^{'} {\bf K}(T_s, \vec k) {\cal G}(\vec k,{\vec k}^{'}) \triangle ({\vec k}^{'},\vec q)\Big) ,\nonumber \\ &&{}\nonumber \\ {{d {\vec \rho}_a(T_s)}\over {dT_s}}&\, {\buildrel \circ \over =}\,& \{ {\vec \rho}_a(T_s),H \} = {{\partial H}\over {\partial {\vec \pi}_a(T_s)}}= \nonumber \\ &=&\sum_{i=1}^N {{ {\hat \gamma}_{ai}}\over {\eta_iM_i}}({{ {\vec \pi} _{\zeta}}\over {\sqrt{N}}}+N\sum_{b=1}^{N-1} {\hat \gamma}_{bi} {\vec \pi}_b) ,\nonumber \\ {{d {\vec \pi}_a(T_s)}\over {dT_s}}&\, {\buildrel \circ \over =}\,& \{ {\vec \pi}_a(T_s),H \} =-{{\partial H}\over {\partial {\vec \rho}_a(T_s)}}= \nonumber \\ &=&{{2G}\over {\sqrt{N}}}\sum_{i=1}^N {\hat \gamma}_{ai} {{\eta_im_i+G\phi}\over {\eta_iM_i}} \int d\tilde q \vec q\, \sqrt{F^{\tau}(q) \omega (q) P^{\tau}_{\phi}+F(q)\vec q\cdot {\vec \pi}_{\zeta}(T_s)+{\cal D} _{\vec q}{\bf H}(T_s,\vec q)}\nonumber \\ &&sin\, \Big( \vec q\cdot (\vec \zeta (T_s)+ {1\over {\sqrt{N}}}\sum_{a=1}^{N-1}{\hat \gamma}_{ai}{\vec \rho}_{ai}(T_s))- \nonumber \\ &-&\omega (q)(T_s-X^{\tau}_{\phi})+\int d\tilde kd{\tilde k}^{'} {\bf K}(T_s, \vec k) {\cal G}(\vec k,{\vec k}^{'}) \triangle ({\vec k}^{'},\vec q)\Big) ,\nonumber \\ &&{}\nonumber \\ {{d X^{\tau}_{\phi}}\over {dT_s}}&\, {\buildrel \circ \over =}\,& \{ X^{\tau}_{\phi},H \} =- {{\partial H}\over {\partial P^{\tau}_{\phi}}}= \nonumber \\ &=&-1+G\sum_{i=1}^N {{\eta_im_i+G\phi}\over {\eta_iM_i}} \int d\tilde q {{F^{\tau}(q)\omega (q)}\over {\sqrt{F^{\tau}(q) \omega (q) P^{\tau}_{\phi}+F(q)\vec q\cdot {\vec \pi}_{\zeta}(T_s)+{\cal D} _{\vec q}{\bf H}(T_s,\vec q)} }}\nonumber \\ &&cos\, \Big( \vec q\cdot (\vec \zeta (T_s)+ {1\over {\sqrt{N}}}\sum_{a=1}^{N-1}{\hat \gamma}_{ai}{\vec \rho}_{ai}(T_s))- \nonumber \\ &-&\omega (q)(T_s-X^{\tau}_{\phi})+\int d\tilde kd{\tilde k}^{'} {\bf K}(T_s, \vec k) {\cal G}(\vec k,{\vec k}^{'}) \triangle ({\vec k}^{'},\vec q)\Big) ,\nonumber \\ {{d P^{\tau}_{\phi}}\over {dT_s}}&\, {\buildrel \circ \over =}\,& \{ P^{\tau}_{\phi},H \} ={{\partial H}\over {\partial X^{\tau}_{\phi}}}= \nonumber \\ &=&-2G\sum_{i=1}^N {{\eta_im_i+G\phi}\over {\eta_iM_i}} \int d\tilde q \omega (q) \sqrt{F^{\tau}(q) \omega (q) P^{\tau}_{\phi}+F(q)\vec q\cdot {\vec \pi}_{\zeta}(T_s)+{\cal D} _{\vec q}{\bf H}(T_s,\vec q)}\nonumber \\ &&sin\, \Big( \vec q\cdot (\vec \zeta (T_s)+ {1\over {\sqrt{N}}}\sum_{a=1}^{N-1}{\hat \gamma}_{ai}{\vec \rho}_{ai}(T_s))- \nonumber \\ &-&\omega (q)(T_s-X^{\tau}_{\phi})+\int d\tilde kd{\tilde k}^{'} {\bf K}(T_s, \vec k) {\cal G}(\vec k,{\vec k}^{'}) \triangle ({\vec k}^{'},\vec q)\Big) ,\nonumber \\ &&{}\nonumber \\ {{\partial {\bf H}(T_s,\vec q)}\over {\partial T_s}}&\, {\buildrel \circ \over =}\,& \{ {\bf H}(T_s,\vec q),H \} = {{\delta H}\over {\delta {\bf K}(T_s,\vec q)}}= \nonumber \\ &=&-2G \sum_{i=1}^N {{\eta_im_i+G\phi}\over {\eta_iM_i}} \int d\tilde kd{\tilde k}^{'} {\cal G}(\vec q,{\vec k}^{'}) \triangle ({\vec k}^{'},\vec k) \nonumber \\ &&\sqrt{F^{\tau}(q) \omega (q) P^{\tau}_{\phi}+F(q)\vec q\cdot {\vec \pi}_{\zeta}(T_s)+{\cal D} _{\vec q}{\bf H}(T_s,\vec q)}\nonumber \\ &&sin\, \Big( \vec k\cdot (\vec \zeta (T_s)+ {1\over {\sqrt{N}}}\sum_{a=1}^{N-1}{\hat \gamma}_{ai}{\vec \rho}_{ai}(T_s))- \nonumber \\ &-&\omega (k)(T_s-X^{\tau}_{\phi})+\int d{\tilde k}_1d{\tilde k}_2 {\bf K}(T_s, {\vec k}_1) {\cal G}({\vec k}_1,{\vec k}_2) \triangle ({\vec k}_2,\vec k)\Big) ,\nonumber \\ {{\partial {\bf K}(T_s,\vec q)}\over {\partial T_s}}&\, {\buildrel \circ \over =}\,& \{ {\bf K}(T_s,\vec q),H \} =- {{\delta H}\over {\delta {\bf H}(T_s,\vec q)}} =\nonumber \\ &=&-2G \sum_{i=1}^N {{\eta_im_i+G\phi}\over {\eta_iM_i}}\nonumber \\ &&{\cal D}_{\vec q} \Big[ cos\, \Big( \vec q\cdot (\vec \zeta (T_s)+ {1\over {\sqrt{N}}}\sum_{a=1}^{N-1}{\hat \gamma}_{ai}{\vec \rho}_{ai}(T_s))- \nonumber \\ &-&\omega (q)(T_s-X^{\tau}_{\phi})+\int d\tilde kd{\tilde k}^{'} {\bf K}(T_s, \vec k) {\cal G}(\vec k,{\vec k}^{'}) \triangle ({\vec k}^{'},\vec q)\Big) \nonumber \\ && \Big( \sqrt{F^{\tau}(q) \omega (q) P^{\tau}_{\phi}+F(q)\vec q\cdot {\vec \pi}_{\zeta}(T_s)+{\cal D} _{\vec q}{\bf H}(T_s,\vec q)} \Big) {}^{-1} \Big] . \label{VII9} \end{eqnarray} When the decomposition into center-of-mass and relative variables will be available for the transverse electromagnetic field in the Coulomb gauge, one will be able to make a similar treatment of the isolated system consisting of N charged particles (with Grassmann-valued electric charges) plus the electromagnetic field \cite{lus,albad}. \vfill\eject \section{Charged Klein-Gordon Field.} \subsection{On spacelike hypersurfaces.} Following Ref.\cite{albad}, let us now consider a charged Klein-Gordon field: $\phi ={1\over {\sqrt{2}}} (\phi_1+i\phi_2)$, $\phi^{*}={1\over {\sqrt{2}}}(\phi_1-i\phi_2)$ [$\phi_1= {1\over {\sqrt{2}}}(\phi +\phi^{*})$, $\phi_2={{-i}\over {\sqrt{2}}}(\phi - \phi^{*})$], whose action is \begin{eqnarray} S&=& \int d\tau d^3\sigma N(\tau ,\vec \sigma )\sqrt{\gamma (\tau ,\vec \sigma ) } \Big( g^{\tau\tau} {\dot \phi}^{*}\, \dot \phi +\nonumber \\ &&+g^{\tau \check r} \Big[ {\dot \phi}^{*}\, \partial _{\check r}\phi +\partial_{\check r}\phi^{*}\, \dot \phi \Big] + g^{\check r\check s}\partial_{\check r}\phi^{*}\, \partial _{\check s}\phi - m^2 \phi^{*}\phi \Big) (\tau ,\vec \sigma )=\nonumber \\ &&=\int d\tau d^3\sigma \sqrt{\gamma (\tau ,\vec \sigma )} \Big( {1\over N} [{\dot \phi}^{*} -N^{\check r}\partial_{\check r}\phi^{*}]\, [\dot \phi -N^{\check s}\partial_{\check s}\phi ]+\nonumber \\ &+&N \Big[ \gamma^{\check r\check s}\partial_{\check r}\phi^{*} \partial_{\check s}\phi -m^2\phi^{*}\phi \Big] \Big) (\tau ,\vec \sigma ). \label{VIII1} \end{eqnarray} The canonical momenta are \begin{eqnarray} \pi_{\phi}(\tau ,\vec \sigma )&=&{{\partial L}\over {\partial \partial_{\tau} \phi (\tau ,\vec \sigma )}}={1\over {\sqrt{2}}}(\pi_1-i\pi_2)(\tau ,\vec \sigma )=\nonumber \\ &=&{{\sqrt{\gamma (\tau ,\vec \sigma )}}\over {N(\tau ,\vec \sigma )}}\Big[ {\dot \phi}^{*}-N^{\check r}\partial _{\check r} \phi^{*}\Big] (\tau ,\vec \sigma ), \nonumber \\ \pi_{\phi^{*}}(\tau ,\vec \sigma )&=&{{\partial L}\over {\partial \partial _{\tau}\phi^{*}(\tau ,\vec \sigma )}}={1\over {\sqrt{2}}}(\pi_1+i\pi_2)(\tau ,\vec \sigma )=\nonumber \\ &=&{{\sqrt{\gamma (\tau ,\vec \sigma )}}\over {N(\tau ,\vec \sigma )}} \Big[ \dot \phi -N^{\check r}\partial _{\check r}\phi \Big] (\tau ,\vec \sigma ), \nonumber \\ \rho_{\mu}(\tau ,\vec \sigma )&=&-{{\partial L}\over {\partial \partial_{\tau} z^{\mu}(\tau ,\vec \sigma )}}=\nonumber \\ &=&l_{\mu}(\tau ,\vec \sigma ) \Big( {{\pi_{\phi} \pi_{\phi^{*}}}\over {\sqrt{\gamma}}}-\sqrt{\gamma} \Big[ \gamma ^{\check r\check s}\partial_{\check r}\phi^{*} \partial_{\check s}\phi - m^2\phi^{*}\phi \Big] \Big) (\tau ,\vec \sigma )+\nonumber \\ &&+z_{\check s\mu}(\tau ,\vec \sigma )\gamma^{\check r\check s}(\tau ,\vec \sigma ) \Big( \pi_{\phi^{*}}\partial_{\check r}\phi^{*}+ \pi_{\phi} \partial_{\check r}\phi \Big) (\tau ,\vec \sigma ). \label{VIII2} \end{eqnarray} Therefore, we have the following primary constraints \begin{eqnarray} {\cal H}_{\mu}(\tau ,\vec \sigma )&=& \rho_{\mu}(\tau ,\vec \sigma )- \nonumber \\ &&-l_{\mu}(\tau ,\vec \sigma ) \Big( {{\pi_{\phi} \pi_{\phi^{*}}}\over {\sqrt{\gamma}}}-\sqrt{\gamma} [\gamma^{\check r\check s} \partial_{\check r}\phi^{*} \partial_{\check s}\phi -m^2\phi^{*}\phi] \Big) (\tau ,\vec \sigma )+\nonumber \\ &&+z_{\check s\mu}(\tau ,\vec \sigma )\gamma^{\check r\check s}(\tau ,\vec \sigma ) \Big( \pi_{\phi^{*}}\partial_{\check r}\phi^{*}+ \pi_{\phi} \partial_{\check r}\phi \Big) (\tau ,\vec \sigma ) \approx 0, \label{VIII3} \end{eqnarray} \noindent and the following Dirac Hamiltonian \begin{equation} H_D=\int d^3\sigma \lambda^{\mu} (\tau ,\vec \sigma ){\cal H}_{\mu}(\tau ,\vec \sigma ). \label{VIII4} \end{equation} By using the Poisson brackets \begin{eqnarray} \{ z^{\mu}(\tau ,\vec \sigma ),\rho_{\nu}(\tau ,{\vec \sigma}^{'}) \} &=&\eta ^{\mu}_{\nu} \delta^3(\vec \sigma -{\vec \sigma}^{'}),\nonumber \\ \{ \phi (\tau ,\vec \sigma ),\pi_{\phi}(\tau ,{\vec \sigma}^{'}) \} &=& \{ \phi^{*}(\tau ,\vec \sigma ),\pi_{\phi^{*}}(\tau ,{\vec \sigma}^{'}) \} = \delta^3(\vec \sigma -{\vec \sigma}^{'}), \label{VIII5} \end{eqnarray} \noindent we find that the time constancy of the primary constraints implies the existence of no secondary constraint. The constraints turn out to be first class. After the restriction to spacelike hyperplanes $z^{\mu}(\tau ,\vec \sigma )=x_s^{\mu}(\tau )+b^{\mu}_{\check r}(\tau ) \sigma^{\check r}$, the constraints are reduced to the following ones \begin{eqnarray} {\tilde {\cal H}}^{\mu}(\tau )&=& \int d^3\sigma {\cal H}^{\mu}(\tau ,\vec \sigma )=\nonumber \\ &=&p^{\mu}_s-l^{\mu} \Big( \int d^3\sigma \Big[ \pi_{\phi^{*}} \pi_{\phi}+ \vec \partial \phi^{*} \cdot \vec \partial \phi +m^2\phi^{*}\phi \Big] (\tau ,\vec \sigma ) \Big) -\nonumber \\ &-&b^{\mu}_{\check r}(\tau ) \{ \int d^3\sigma [\pi_{\phi^{*}}\partial _{\check r}\phi^{*}+\pi_{\phi} \partial_{\check r}\phi ](\tau ,\vec \sigma ) \} \approx 0,\nonumber \\ {\tilde {\cal H}}^{\mu\nu}(\tau )&=&b^{\mu}_{\check r}(\tau ) \int d^3\sigma \sigma^{\check r} {\cal H}^{\nu}(\tau ,\vec \sigma )-b^{\nu}_{\check r}(\tau ) \int d^3\sigma \sigma^{\check r} {\cal H}^{\mu}(\tau ,\vec \sigma )=\nonumber \\ &=&S^{\mu\nu}_s-\Big( b^{\mu}_{\check r}(\tau )l^{\nu}-b^{\nu}_{\check r}(\tau ) l^{\mu}\Big) \nonumber \\ &&\int d^3\sigma \sigma^{\check r}\Big[ \pi_{\phi^{*}}\pi_{\phi}+\vec \partial \phi^{*}\cdot \vec \partial \phi +m^2\phi^{*}\phi \Big] (\tau , \vec \sigma ) +\nonumber \\ &+&\Big( b^{\mu}_{\check r}(\tau )b^{\nu}_{\check s}(\tau )-b^{\nu} _{\check r}(\tau )b^{\mu}_{\check s}(\tau )\Big) \nonumber \\ && \int d^3\sigma \sigma^{\check r} \Big[ \pi_{\phi^{*}}\partial_{\check s} \phi^{*}+\pi_{\phi} \partial_{\check s}\phi \Big] (\tau ,\vec \sigma ) \approx 0. \label{VIII6} \end{eqnarray} The only surviving constraints on the Wigner hyperplanes [with the reduced canonical variables ${\tilde x}^{\mu}_s(\tau )$, $p^{\mu}_s$, $\phi (\tau ,\vec \sigma )$, $\pi_{\phi}(\tau ,\vec \sigma )={\dot \phi}^{*} (\tau ,\vec \sigma )$, $\phi^{*}(\tau ,\vec \sigma )$, $\pi_{\phi^{*}}(\tau ,\vec \sigma )=\dot \phi (\tau ,\vec \sigma )$, satisfying standard Dirac brackets] are \begin{eqnarray} {\cal H}(\tau )&=& \epsilon_s- \int d^3\sigma \Big[ \pi_{\phi^{*}} \pi_{\phi}+\vec \partial \phi^{*} \cdot \vec \partial \phi +m^2\phi^{*}\phi \Big] (\tau ,\vec \sigma ) \approx 0,\nonumber \\ {\vec {\cal H}}_p(\tau )&=& \int d^3\sigma \Big[ \pi_{\phi^{*}} \vec \partial \phi^{*} +\pi_{\phi} \vec \partial \phi \Big] (\tau ,\vec \sigma ) \approx 0. \label{VIII7} \end{eqnarray} The Lorentz generators have the form of Eq.(\ref{II9}) with the spin tensor \begin{equation} {\bar S}_s^{rs}\equiv S^{rs}_{\phi}=J^{rs}_{\phi}{|}_{{\vec P}_{\phi}=0}= \int d^3\sigma \Big( \sigma^r\Big[ \pi_{\phi^{*}}\partial^s\phi^{*}+\pi _{\phi}\partial^s\phi \Big] (\tau ,\vec \sigma )-(r \leftrightarrow s) \Big). \label{VIII8} \end{equation} With the gauge fixing $\chi =T_s-\tau \approx 0$ the final Hamiltonian becomes \begin{eqnarray} H_R&=&M_{\phi}-\vec \lambda (T_s)\cdot {\vec {\cal H}}_p(T_s),\nonumber \\ M_{\phi}&=& \int d^3\sigma \Big[ \pi_{\phi^{*}} \pi_{\phi}+\vec \partial \phi^{*} \cdot \vec \partial \phi +m^2\phi^{*}\phi \Big] (T_s ,\vec \sigma ) \label{VIII9} \end{eqnarray} \subsection{Collective and relative variables.} A priori we have the following two possibilities for the description of the complex Klein-Gordon field $\phi (\tau ,\vec \sigma )$ in terms of a Fourier transform \begin{eqnarray} \phi ( \tau ,\vec \sigma ) &=&{1\over {\sqrt{2}}}[ \phi_1(\tau ,\vec \sigma ) +i\phi_2(\tau ,\vec \sigma )]=\nonumber \\ &=&{1\over {\sqrt{2}}} \int d\tilde q \Big(\, [ a_1( \tau ,\vec q) e^{-i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }+a_1^{*}(\tau ,\vec q) e^{+i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) } ] + \nonumber \\ &+&i [ a_2( \tau ,\vec q) e^{-i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }+a_2^{*}( \tau ,\vec q) e^{+i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }] \, \Big) = \nonumber \\ &=&\int d\tilde q [ a ( \tau ,\vec q) e^{-i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }+b^{*} ( \tau ,\vec q) e^{+i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }] ,\nonumber \\ \phi^{*}(\tau ,\vec \sigma )&=&{1\over {\sqrt{2}}}[\phi_1(\tau ,\vec \sigma )- i\phi_2(\tau ,\vec \sigma )]=\nonumber \\ &=&{1\over {\sqrt{2}}} \int d\tilde q \Big( \, [ a_1( \tau ,\vec q) e^{-i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }+a_1^{*}( \tau ,\vec q) e^{i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }] - \nonumber \\ &-&i [ a_2( \tau ,\vec q) e^{-i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }+a_2^{*}( \tau ,\vec q) e^{+i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }] \, \Big) =\nonumber \\ &=&\int d\tilde q [ b ( \tau ,\vec q) e^{-i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) } + a^{*} ( \tau ,\vec q) e^{+i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) } ] ,\nonumber \\ &&{}\nonumber \\ \pi ( \tau ,\vec \sigma )& =&{1\over {\sqrt{2}}}[\pi_1(\tau ,\vec \sigma )-i \pi_2(\tau ,\vec \sigma )]=\nonumber \\ &=&{{-i}\over {\sqrt{2}}} \int d\tilde q \Big( \, [ a_1( \tau ,\vec q) e^{-i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }-a_1^{*}( \tau ,\vec q) e^{+i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }] - \nonumber \\ &-&i [ a_2( \tau ,\vec q) e^{-i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }-a_2^{*}( \tau ,\vec q) e^{+i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }] \, \Big) =\nonumber \\ &=&-i\int d\tilde q\omega ( q) [ b( \tau ,\vec q) e^{-i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }-a^{*}( \tau ,\vec q) e^{+i( \omega ( \vec q) \tau -\vec q\cdot \vec \sigma ) }] ,\nonumber \\ \pi^{*}(\tau ,\vec \sigma )&=&{1\over {\sqrt{2}}}[\pi_1(\tau ,\vec \sigma )+i \pi_2(\tau ,\vec \sigma )]=\nonumber \\ &=&{{-i}\over {\sqrt{2}}} \int d\tilde q \Big( \, [ a_1( \tau ,\vec q) e^{-i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }-a_1^{*}( \tau ,\vec q) e^{+i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }] + \nonumber \\ &+&i [ a_2( \tau ,\vec q) e^{-i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }-a_2^{*}( \tau ,\vec q) e^{+i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }] \, \Big) =\nonumber \\ &=&-i\int d\tilde q\omega ( q) [ a( \tau ,\vec q) e^{-i( \omega ( \vec q) \tau -\vec q\cdot \vec \sigma ) } - b^{*}( \tau ,\vec q) e^{+i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) } ] ,\nonumber \\ &&{}\nonumber \\ a_i(\tau ,\vec q)&=&\sqrt{I_i(\tau ,\vec q)} e^{i\varphi_i(\tau ,\vec q)}, \quad\quad i=1,2,\nonumber \\ a( \tau ,\vec q) &=&{1\over {\sqrt{2}}} [a_1(\tau ,\vec q)+ia_2(\tau ,\vec q) ]={1\over {\sqrt{2}}} [\sqrt{I_1}e^{i\varphi_1}+i\sqrt{I_2} e^{i\varphi_2}](\tau ,\vec q)=\nonumber \\ &=&\int d^3\sigma [ \omega ( q) \phi ( \tau ,\vec \sigma ) +i\pi^{*} ( \tau ,\vec \sigma ) ] e^{i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }=\sqrt{I_a(\tau ,\vec q)}e^{i\varphi_a(\tau ,\vec q)}, \nonumber \\ b(\tau ,\vec q)&=&{1\over {\sqrt{2}}} [a_1(\tau ,\vec q)-ia_2(\tau ,\vec q)]= {1\over {\sqrt{2}}}[\sqrt{I_1}e^{i\varphi_1}-i\sqrt{I_2}e^{i\varphi_2}](\tau ,\vec q)=\nonumber \\ &=&\int d^3\sigma [\omega (q) \phi^{*}(\tau ,\vec \sigma )+i \pi (\tau ,\vec \sigma )] e^{i(\omega (q)\tau -\vec q\cdot \vec \sigma )}=\sqrt{I_b(\tau ,\vec q)}e^{i\varphi_b(\tau ,\vec q)},\nonumber \\ a^{*}( \tau ,\vec q) &=&{1\over {\sqrt{2}}}[a_1^{*}(\tau ,\vec q)-ia _2^{*}(\tau ,\vec q)]=\nonumber \\ &=&\int d^3\sigma [ \omega ( q) \phi^{*} ( \tau ,\vec \sigma ) -i\pi ( \tau ,\vec \sigma ) ] e^{-i( \omega ( q) \tau -\vec q\cdot \vec \sigma ) }, \nonumber \\ b^{*}(\tau ,\vec q)&=&{1\over {\sqrt{2}}}[a_1^{*}(\tau ,\vec q)+ ia_2^{*}(\tau ,\vec q)]=\nonumber \\ &=&\int d^3\sigma [\omega (q)\phi (\tau ,\vec \sigma )-i\pi^{*}(\tau ,\vec \sigma )] e^{-i(\omega (q)\tau -\vec q\cdot \vec \sigma )},\nonumber \\ &&{}\nonumber \\ I_i(\tau ,\vec q)&=&a^{*}_i(\tau ,\vec q) a_i(\tau ,\vec q),\nonumber \\ I_a(\tau ,\vec q)&=&a^{*}(\tau ,\vec q)a(\tau ,\vec q)={1\over 2}[a^{*}_1a_1+ a^{*}_2a_2+i(a^{*}_1a_2-a^{*}_2a_1)](\tau ,\vec q),\nonumber \\ I_b(\tau ,\vec q)&=&b^{*}(\tau ,\vec q)b(\tau ,\vec q)={1\over 2}[a^{*}_1a_1+ a^{*}_2a_2-i(a^{*}_1a_2-a^{*}_2a_1)],\nonumber \\ (I_a+I_b)&&(\tau ,\vec q)=[I_1+I_2](\tau ,\vec q),\nonumber \\ (I_a-I_b)&&(\tau ,\vec q)=i[a^{*}_1a_2-a^{*}_2a_1](\tau ,\vec q),\nonumber \\ &&{}\nonumber \\ N_{i \phi}&=&\int d\tilde q a^{*}_i(\tau ,\vec q) a_i(\tau ,\vec q)= \int d\tilde q I_i(\tau ,\vec \sigma ),\nonumber \\ N_{a \phi}&=&\int d\tilde q a^{*}(\tau ,\vec q)a(\tau ,\vec q) ={1\over 2}[N_{1 \phi}+N_{2 \phi}]+{i\over 2}\int d\tilde q [a^{*}_1a_2- a^{*}_2a_1](\tau ,\vec q)=\nonumber \\ &=& \int d\tilde q I_a(\tau ,\vec q)=\nonumber \\ &=&{1\over 2} \int d^3\sigma [\pi_{\phi^{*}} {1\over {\sqrt{m^2+\triangle}}} \pi +\phi^{*} \sqrt{m^2+\triangle} \phi +i(\pi_{\phi^{*}}\phi^{*}-\pi \phi ) ](\tau ,\vec \sigma ),\nonumber \\ N_{b \phi}&=&\int d\tilde q b^{*}(\tau ,\vec q) b(\tau ,\vec q) ={1\over 2}[N_{1 \phi}+N_{2 \phi}]-{i\over 2}\int d\tilde q [a^{*}_1a_2- a^{*}_2a_1](\tau ,\vec q)=\nonumber \\ &+&\int d\tilde q I_b(\tau ,\vec q)=\nonumber \\ &=&{1\over 2} \int d^3\sigma [\pi_{\phi^{*}} {1\over {\sqrt{m^2+\triangle}}} \pi +\phi^{*} \sqrt{m^2+\triangle} \phi -i(\pi_{\phi^{*}}\phi^{*}-\pi \phi ) ](\tau ,\vec \sigma ),\nonumber \\ &&{}\nonumber \\ N_{\phi}&=&N_{1 \phi}+N_{2 \phi} = N_{a \phi} + N_{b \phi}=\nonumber \\ &=&\int d^3\sigma [\pi_{\phi^{*}} {1\over {\sqrt{m^2+\triangle}}} \pi +\phi^{*} \sqrt{m^2+\triangle} \phi ](\tau ,\vec \sigma ) \label{VIII10} \end{eqnarray} However, the description of the conserved electric charge of the Klein-Gordon field privileges the use of the Fourier coefficients $a(\tau ,\vec q)$, $b(\tau ,\vec q)$ rather than of the $a_i(\tau ,\vec q)$'s, because we have \begin{equation} q_{\phi}=i\int d^3\sigma [\pi_{\phi^{*}}\phi^{*}-\pi \phi ](\tau ,\vec \sigma )= N_{b \phi} - N_{a \phi},\quad {{dq_{\phi}}\over {d\tau}}\, {\buildrel \circ \over =}\, 0; \label{VIII11} \end{equation} \noindent on the Wigner hyperplanes the conservation $\partial_{\mu}J^{\mu} _{\phi}\, {\buildrel \circ \over =}\, 0$ of the electromagnetic current $J^{\mu}_{\phi}(x)=-i [\partial^{\mu} {\tilde\phi}^{*} \tilde \phi -{\tilde \phi}^{*} \partial^{\mu} \tilde \phi ](x)$ is replaced by the existence of the Hamiltonian constant of motion $q_{\phi}$. The Fourier coefficients $a_i(\tau ,\vec \sigma )$, i=1,2, correspond to the description of the field as two real Klein-Gordon fields $\phi_i$. Instead, as shown in Appendix D, the Fourier coefficients ``a" and ``b" correspond to two Klein-Gordon fields $\phi_a$, $\phi_b$, with positive ($\phi_a$) or negative ($\phi_b$) energy and electric charge, which are nonlocal combinations of $\phi$ and $\pi_{\phi^{*}}$. The total 4-momentum of the Klein-Gordon field may be described either in terms of the ``a" and ``b" or of the ``$a_i$" \begin{eqnarray} P^{\tau}_{\phi}&=&\int d\tilde q \omega (q) [I_1+I_2](\tau ,\vec q)=P^{\tau} _{1 \phi}(\tau )+P^{\tau}_{2 \phi}(\tau )=\nonumber \\ &=&\int d\tilde q \omega (q) [I_a+I_b](\tau ,\vec q)=P^{\tau}_{a \phi}(\tau )+ P^{\tau}_{b \phi}(\tau )=\nonumber \\ &=&\int d^3\sigma \Big[ \pi_{\phi^{*}} \pi_{\phi}+\vec \partial \phi^{*} \cdot \vec \partial \phi +m^2\phi^{*}\phi \Big] (\tau ,\vec \sigma ),\nonumber \\ {\vec P}_{\phi}&=&\int d\tilde q \vec q [I_1+I_2](\tau ,\vec q)={\vec P} _{1 \phi}(\tau )+{\vec P}_{2 \phi}(\tau )=\nonumber \\ &=&\int d\tilde q \vec q [I_a+I_b](\tau ,\vec q)={\vec P}_{a \phi}(\tau )+ {\vec P}_{b \phi}(\tau )=\nonumber \\ &=&\int d^3\sigma \Big[ \pi_{\phi^{*}} \vec \partial \phi^{*} +\pi_{\phi} \vec \partial \phi \Big] (\tau ,\vec \sigma ), \label{VIII12} \end{eqnarray} \noindent but it is possible to define a global ``relative 4-momentum" of the field only in terms of the ``a" and ``b" \begin{eqnarray} Q^{\tau}_{\phi}(\tau )&=&{1\over 2}\int d\tilde q \omega (q) [I_a-I_b](\tau ,\vec q)=P^{\tau}_{a \phi}(\tau )-P^{\tau}_{b \phi}(\tau )=\nonumber \\ &=&{i\over 2} \int d^3\sigma [\pi_{\phi^{*}} \sqrt{m^2+\triangle} \phi^{*}- \pi \sqrt{m^2+\triangle} \phi ](\tau ,\vec \sigma )\not= \nonumber \\ &\not=& {1\over 2}\int d\tilde q \omega (q) [I_1-I_2](\tau ,\vec q)= P^{\tau}_{1 \phi}(\tau )-P^{\tau}_{2 \phi}(\tau ),\nonumber \\ {\vec Q}_{\phi}(\tau )&=&{1\over 2}\int d\tilde q \vec q [I_a-I_b](\tau ,\vec q)={\vec P}_{a \phi}(\tau )-{\vec P}_{b \phi}(\tau )=\nonumber \\ &=&-{i\over 2} \int d^3\sigma [\pi_{\phi^{*}} {{\vec \partial}\over {\sqrt{m^2 +\triangle}}} \pi -\phi^{*} \sqrt{m^2+\triangle} \vec \partial \phi ](\tau ,\vec \sigma )\not= \nonumber \\ &\not=&{1\over 2}\int d\tilde q \vec q [I_1-I_2](\tau ,\vec q)={\vec P} _{1 \phi}(\tau )-{\vec P}_{2 \phi}(\tau ),\nonumber \\ &&{}\nonumber \\ P^{\tau}_{a \phi}&=&{1\over 2}(P^{\tau}_{\phi}+Q^{\tau}_{\phi}),\quad\quad {\vec P}_{a \phi}={1\over 2}({\vec P}_{\phi}+{\vec Q}_{\phi}),\nonumber \\ P^{\tau}_{b \phi}&=&{1\over 2}(P^{\tau}_{\phi}-Q^{\tau}_{\phi}),\quad\quad {\vec P}_{b \phi}={1\over 2}({\vec P}_{\phi}-{\vec Q}_{\phi}), \label{VIII13} \end{eqnarray} The two partial centers of phase have coordinates \begin{eqnarray} X^{\tau}_{a\, \phi}&=& \int d\tilde q \omega (q) F^{\tau}(q) \varphi_a(\tau ,\vec q)={1\over {2i}} \int d\tilde q \omega (q) F^{\tau}(q) ln\, {{a(\tau ,\vec q)}\over {a^{*}(\tau ,\vec q)}}=\tau +\nonumber \\ &+&{1\over {2\pi i m}} \int d^3q {{e^{-{{4\pi}\over {m^2}} q^2} }\over {q^2\, \omega (q)}} ln\, \Big[ {{\omega (q) \int d^3\sigma e^{i\vec q\cdot \vec \sigma} \phi (\tau ,\vec \sigma ) +i\int d^3\sigma e^{i\vec q\cdot \vec \sigma} \pi_{\phi^{*}}(\tau ,\vec \sigma )}\over {\omega (q) \int d^3\sigma e^{-i\vec q\cdot \vec \sigma} \phi^{*}(\tau ,\vec \sigma ) -i \int d^3\sigma e^{-i\vec q\cdot \vec \sigma} \pi (\tau ,\vec \sigma )}}\Big]=\nonumber \\ &{\buildrel {def} \over =}& \tau +{\tilde X}^{\tau}_{a\, \phi},\nonumber \\ {\vec X}_{a\, \phi}&=& \int d\tilde q \, \vec q\, F(q) \varphi_a(\tau ,\vec q) ={1\over {2i}} \int d\tilde q\, \vec q\, F(q) ln\, {{a(\tau ,\vec q)}\over {a^{*}(\tau ,\vec q)}}=\nonumber \\ &=&{{2i}\over {\pi m}} \int d^3q\, {{\vec q}\over {q^4}}\, e^{-{{4\pi}\over {m^2}} q^2} ln\, \Big[ {{\sqrt{m^2+q^2} \int d^3\sigma e^{i\vec q\cdot \vec \sigma} \phi (\tau ,\vec \sigma ) +i \int d^3\sigma e^{i\vec q\cdot \vec \sigma} \pi_{\phi^{*}}(\tau ,\vec \sigma )}\over {\sqrt{m^2+q^2} \int d^3\sigma e^{-i\vec q\cdot \vec \sigma} \phi^{*}(\tau ,\vec \sigma ) -i \int d^3\sigma e^{-i\vec q\cdot \vec \sigma} \pi (\tau ,\vec \sigma )}} \Big] ,\nonumber \\ &&{}\nonumber \\ X^{\tau}_{b\, \phi}&=& \int d\tilde q \omega (q) F^{\tau}(q) \varphi_b(\tau ,\vec q)={1\over {2i}} \int d\tilde q \omega (q) F^{\tau}(q) ln\, {{b(\tau ,\vec q)}\over {b^{*}(\tau ,\vec q)}}=\tau +\nonumber \\ &+&{1\over {2\pi i m}} \int d^3q {{e^{-{{4\pi}\over {m^2}} q^2} }\over {q^2\, \omega (q)}} ln\, \Big[ {{\omega (q) \int d^3\sigma e^{i\vec q\cdot \vec \sigma} \phi^{*}(\tau ,\vec \sigma ) +i\int d^3\sigma e^{i\vec q\cdot \vec \sigma}\pi (\tau ,\vec \sigma )}\over {\omega (q) \int d^3\sigma e^{-i\vec q\cdot \vec \sigma} \phi (\tau ,\vec \sigma ) -i \int d^3\sigma e^{-i\vec q\cdot \vec \sigma} \pi_{\phi^{*}}(\tau ,\vec \sigma )}}\Big]=\nonumber \\ &{\buildrel {def} \over =}& \tau +{\tilde X}^{\tau}_{b\, \phi},\nonumber \\ {\vec X}_{b\, \phi}&=& \int d\tilde q \, \vec q\, F(q) \varphi_b(\tau ,\vec q) ={1\over {2i}} \int d\tilde q\, \vec q\, F(q) ln\, {{b(\tau ,\vec q)}\over {b^{*}(\tau ,\vec q)}}=\nonumber \\ &=&{{2i}\over {\pi m}} \int d^3q\, {{\vec q}\over {q^4}}\, e^{-{{4\pi}\over {m^2}} q^2} ln\, \Big[ {{\sqrt{m^2+q^2} \int d^3\sigma e^{i\vec q\cdot \vec \sigma} \phi^{*}(\tau ,\vec \sigma ) +i \int d^3\sigma e^{i\vec q\cdot \vec \sigma} \pi (\tau ,\vec \sigma )}\over {\sqrt{m^2+q^2} \int d^3\sigma e^{-i\vec q\cdot \vec \sigma} \phi (\tau ,\vec \sigma ) -i \int d^3\sigma e^{-i\vec q\cdot \vec \sigma} \pi_{\phi^{*}}(\tau ,\vec \sigma )}} \Big] . \label{VIII14} \end{eqnarray} Therefore, we shall define the canonical transformation of Section IV separately to $a,a^{*}\mapsto$ $P^A_{a \phi}$, $X^A_{a \phi} (X^{\tau}_{a\phi}= \tau +{\tilde X}^{\tau}_{a\phi})$, ${\bf H}_a$, ${\bf K}_a$ and to $b,b^{*}\mapsto$ $P^A_{b \phi}$, $X^A_{b \phi} (X^{\tau}_{b\phi}=\tau +{\tilde X}_{b\phi})$, ${\bf H}_b$, ${\bf K}_b$ and not to $a_i,a^{*}_i$, i=1,2, $\mapsto P^A_{i \phi}, X^A_{i \phi}, {\bf H}_i, {\bf K}_i$ [assuming that the fields $\phi_a$, $\pi_a$, $\phi_b$, $\pi_b$ satisfy Eq.(\ref{IV29})]. Then, we shall do the further canonical transformation from the two sets of collective variables $X^A_{a\phi}$, $P^A_{a\phi}$, $X^A_{b\phi}$, $P^A_{b\phi}$, to global center of phase variables $X^A_{\phi}$, $P^A_{\phi}$, and global relative variables $R^A_{\phi}$, $Q^A _{\phi}$ [$R^A_{\phi}$ describes the action-reaction between the centers of phase of positive and negative energy field configutations] \begin{eqnarray} X^A_{\phi}&=&{1\over 2}(X^A_{a\, \phi}+X^A_{b\, \phi})\, [= {1\over 2}(X^A_{1\phi}+X^A_{2\phi})],\quad\quad X^{\tau}_{\phi}=\tau +{\tilde X}^{\tau}_{\phi},\nonumber \\ P^A_{\phi}&=&P^A_{a\, \phi}+P^A_{b\, \phi}[=P^A_{1\phi}+P^A_{2\phi}], \nonumber \\ R^A_{\phi}&=&X^A_{a\, \phi}-X^A_{b\, \phi},\nonumber \\ Q^A_{\phi}&=&{1\over 2}(P^A_{a\, \phi}-P^A_{b\, \phi}). \label{VIII15} \end{eqnarray} The Poincar\'e generators are $P^{\tau}_{\phi}$, ${\vec P}_{\phi}$ and \begin{eqnarray} J^{rs}_{\phi}&=&\int d^3\sigma [\pi_{\phi^{*}}(\sigma^r\partial^s-\sigma^s \partial^r)\phi^{*}+\pi_{\phi}(\sigma^r\partial^s-\sigma^s\partial^r) \phi](\tau ,\vec \sigma )=\nonumber \\ &=&-i \sum_{i=1}^2 \int d\tilde q a^{*}_i(\tau ,\vec q)(q^r{{\partial}\over {\partial q^s}}-q^s{{\partial}\over {\partial q^r}}) a_i(\tau ,\vec q)= \nonumber \\ &=&-i \int d\tilde q \Big[ a^{*}(\tau ,\vec q)(q^r{{\partial}\over {\partial q^s}}-q^s{{\partial}\over {\partial q^r}})a(\tau ,\vec q)+b^{*}(\tau ,\vec q) (q^r{{\partial}\over {\partial q^s}}-q^s{{\partial}\over {\partial q^r}}) b(\tau ,\vec q)\Big],\nonumber \\ J^{\tau r}_{\phi}&=&-\tau P^r_{\phi}+{1\over 2}\int d^3\sigma \, \sigma^r [\pi _{\phi^{*}}\pi_{\phi}+\vec \partial \phi^{*}\cdot \vec \partial \phi +m^2 \phi^{*}\phi ](\tau ,\vec \sigma )=\nonumber \\ &=&-\tau P^r_{\phi}+i\sum_{i=1}^2 \int d\tilde q \omega (q) a^{*}_i(\tau ,\vec q) {{\partial}\over {\partial q^r}} a_i(\tau ,\vec q)=\nonumber \\ &=&-\tau P^r_{\phi}+i \int d\tilde q \omega (q) \Big[ a^{*}(\tau ,\vec q) {{\partial}\over {\partial q^r}}a(\tau ,\vec q)+b^{*}(\tau ,\vec q) {{\partial} \over {\partial q^r}} b(\tau ,\vec q). \label{VIII16} \end{eqnarray} To find the real center of mass ${\vec q}_{\phi}$ we must use Eq.(\ref{VI1}) of Section VI, namely $q^r_{\phi} \approx r^r_{\phi}=-J^{\tau r}_{\phi}/P^{\tau} _{\phi}{|}_{\tau =0}$. The original variables have the following expression \begin{eqnarray} a(\tau ,\vec k)&=&\sqrt{F^{\tau}(k)\omega (k)P^{\tau} _{a \phi}-F(k)\vec k\cdot {\vec P}_{a \phi}(\tau )+{\cal D}_k {\bf H}_a(\tau , \vec k) }\nonumber \\ &&e^{i\int d\tilde q\int d\tilde q^{\prime }\, {\bf K}_a( \tau ,\vec q) {\cal G}\left( \vec q,\vec q^{\prime }\right) \Delta \left( \vec q^{\prime } ,\vec k\right) +i\omega (k) X_{a \phi}^\tau (\tau ) - i\vec k\cdot {\vec X}_{a \phi}(\tau ) },\nonumber \\ &&{}\nonumber \\ b(\tau ,\vec q)&=&\sqrt{F^{\tau}(k)\omega (k)P^{\tau}_{b \phi}(\tau )- F(k)\vec k\cdot {\vec P}_{b \phi}(\tau )+{\cal D}_k {\bf H}_b(\tau , \vec k) } \nonumber \\ &&e^{i\int d\tilde q\int d\tilde q^{\prime }\, {\bf K}_b( \tau ,\vec q) {\cal G}\left( \vec q,\vec q^{\prime }\right) \Delta \left( \vec q^{\prime } ,\vec k\right) +i\omega (k) X_{b \phi}^\tau (\tau ) - i\vec k\cdot {\vec X}_{b \phi}(\tau ) } ,\nonumber \\ &&{}\nonumber \\ \phi (\tau ,\sigma )&=& \int d\tilde k \Big(\, \sqrt{F^{\tau}(k)\omega (k) P^{\tau}_{a \phi}(\tau )-F(k)\vec k\cdot {\vec P}_{a \phi}(\tau )+{\cal D} _k {\bf H}_a(\tau , \vec k) }\nonumber \\ &&e^{i\int d\tilde q\int d\tilde q^{\prime }\, {\bf K}_a( \tau ,\vec q) {\cal G}\left( \vec q,\vec q^{\prime }\right) \Delta \left( \vec q^{\prime } ,\vec k\right) +i\omega (k) {\tilde X}_{a \phi}^\tau (\tau ) + i\vec k\cdot \left(\vec \sigma -{\vec X}_{a \phi}(\tau )\right) }+\nonumber \\ &&+\sqrt{F^{\tau}(k)\omega (k)P^{\tau}_{b \phi}(\tau )- F(k)\vec k\cdot {\vec P}_{b \phi}(\tau )+{\cal D}_k {\bf H}_b(\tau , \vec k) } \nonumber \\ &&e^{-i\int d\tilde q\int d\tilde q^{\prime }\, {\bf K}_b( \tau ,\vec q) {\cal G}\left( \vec q,\vec q^{\prime }\right) \Delta \left( \vec q^{\prime } ,\vec k\right) -i\omega (k) {\tilde X}_{b \phi}^\tau (\tau ) - i\vec k\cdot \left(\vec \sigma -{\vec X}_{b \phi}(\tau )\right) } \Big)=\nonumber \\ &&=\int d\tilde k \Big(\, \sqrt{F^{\tau}(k)\omega (k)[{1\over 2}P^{\tau} _{\phi}+Q^{\tau}_{\phi}(\tau )]-F(k)\vec k\cdot [{1\over 2}{\vec P}_{\phi}+ {\vec Q}_{\phi}(\tau )]+{\cal D}_k {\bf H}_a(\tau , \vec k) }\nonumber \\ &&e^{+i\int d\tilde q\int d\tilde q^{\prime }\, {\bf K}_a( \tau ,\vec q) {\cal G}\left( \vec q,\vec q^{\prime }\right) \Delta \left( \vec q^{\prime } ,\vec k\right) +i\omega (k) [{\tilde X}_{\phi}^\tau +{1\over 2}{\tilde R}^{\tau} _{\phi}(\tau )] +i\vec k\cdot \left(\vec \sigma -[{\vec X}_{\phi}+{1\over 2} {\vec R}_{\phi}(\tau )]\right) }+\nonumber \\ &&+\sqrt{F^{\tau}(k)\omega (k)[{1\over 2}P^{\tau}_{\phi}-Q^{\tau}_{\phi}(\tau )] -F(k)\vec k\cdot [{1\over 2}{\vec P}_{\phi}-{\vec Q}_{\phi}(\tau )]+ {\cal D}_k {\bf H}_b(\tau , \vec k) }\nonumber \\ &&e^{-i\int d\tilde q\int d\tilde q^{\prime }\, {\bf K}_b( \tau ,\vec q) {\cal G}\left( \vec q,\vec q^{\prime }\right) \Delta \left( \vec q^{\prime } ,\vec k\right) -i\omega (k) [{\tilde X}_{\phi}^\tau -{1\over 2}{\tilde R}^{\tau} _{\phi}(\tau )] -i\vec k\cdot \left(\vec \sigma -[{\vec X}_{\phi}- {1\over 2}{\vec R}_{\phi}(\tau )]\right) } \Big)=\nonumber \\ &=&\int d\tilde k\Big[ {\bf A}_{a\vec k}(\tau ;{1\over 2}P^A_{\phi}+Q^A_{\phi}, {\bf H}_a] e^{i\Big( \vec k\cdot \vec \sigma +{\bf B}_{a\vec k}(\tau ;X^A _{\phi}+{1\over 2}R^A_{\phi},{\bf K}_a]\Big)} +\nonumber \\ &+&{\bf A}_{b\vec k}(\tau ;{1\over 2}P^A_{\phi}-Q^A_{\phi},{\bf H}_b] e^{-i\Big( \vec k\cdot \vec \sigma +{\bf B}_{b\vec k}(\tau ;X^A_{\phi}-{1\over 2}R^A_{\phi},{\bf K}_b]\Big)}\Big] ,\nonumber \\ &&{}\nonumber \\ \pi (\tau ,\vec \sigma )&=&-i \int d\tilde k \omega (k) \nonumber \\ &&\Big( \sqrt{F^{\tau}(k)\omega (k)[{1\over 2}P^{\tau}_{\phi}-Q^{\tau}_{\phi}(\tau )] -F(k)\vec k\cdot [{1\over 2}{\vec P}_{\phi}-{\vec Q}_{\phi}(\tau )]+ {\cal D}_k {\bf H}_b(\tau , \vec k) }\nonumber \\ &&e^{+i\int d\tilde q\int d\tilde q^{\prime }\, {\bf K}_b( \tau ,\vec q) {\cal G}\left( \vec q,\vec q^{\prime }\right) \Delta \left( \vec q^{\prime } ,\vec k\right) +i\omega (k) [{\tilde X}_{\phi}^\tau -{1\over 2}{\tilde R}^{\tau} _{\phi}(\tau )] +i\vec k\cdot \left(\vec \sigma -[{\vec X}_{\phi}- {1\over 2}{\vec R}_{\phi}(\tau )]\right) }-\nonumber \\ &&-\sqrt{F^{\tau}(k)\omega (k)[{1\over 2}P^{\tau} _{\phi}+Q^{\tau}_{\phi}(\tau )]-F(k)\vec k\cdot [{1\over 2}{\vec P}_{\phi}+ {\vec Q}_{\phi}(\tau )]+{\cal D}_k {\bf H}_a(\tau , \vec k) }\nonumber \\ &&e^{-i\int d\tilde q\int d\tilde q^{\prime }\, {\bf K}_a( \tau ,\vec q) {\cal G}\left( \vec q,\vec q^{\prime }\right) \Delta \left( \vec q^{\prime } ,\vec k\right) -i\omega (k) [{\tilde X}_{\phi}^\tau +{1\over 2}{\tilde R}^{\tau} _{\phi}(\tau )] -i\vec k\cdot \left(\vec \sigma -[{\vec X}_{\phi}+{1\over 2} {\vec R}_{\phi}(\tau )]\right) } \Big)=\nonumber \\ &=&i \int d\tilde k \omega (k)\nonumber \\ &&\Big[ {\bf A}_{a\vec k}(\tau ;{1\over 2}P^A_{\phi}+Q^A_{\phi}, {\bf H}_a] e^{-i\Big( \vec k\cdot \vec \sigma +{\bf B}_{a\vec k}(\tau ;X^A _{\phi}+{1\over 2}R^A_{\phi},{\bf K}_a]\Big)}-\nonumber \\ &-&{\bf A}_{b\vec k}(\tau ;{1\over 2}P^A_{\phi}-Q^A_{\phi},{\bf H}_b] e^{i\Big( \vec k\cdot \vec \sigma +{\bf B}_{b\vec k}(\tau ;X^A_{\phi}-{1\over 2}R^A_{\phi},{\bf K}_b]\Big)} \Big],\nonumber \\ &&{}\nonumber \\ &&{}\nonumber \\ &&{\bf A}_{a\vec k}(\tau ;{1\over 2}P^A_{\phi}+Q^A_{\phi},{\bf H}_a] = \nonumber \\ &&\sqrt{F^{\tau}(k)\omega (k)[{1\over 2}P^{\tau} _{\phi}+Q^{\tau}_{\phi}(\tau )]-F(k)\vec k\cdot [{1\over 2}{\vec P}_{\phi}+ {\vec Q}_{\phi}(\tau )]+{\cal D}_{\vec k} {\bf H}_a(\tau ,\vec k) },\nonumber \\ &&{\bf A}_{b\vec k}(\tau ;{1\over 2}P^A_{\phi}-Q^A_{\phi},{\bf H}_b] = \nonumber \\ &&\sqrt{F^{\tau}(k)\omega (k)[{1\over 2}P^{\tau} _{\phi}-Q^{\tau}_{\phi}(\tau )]-F(k)\vec k\cdot [{1\over 2}{\vec P}_{\phi}- {\vec Q}_{\phi}(\tau )]+{\cal D}_{\vec k} {\bf H}_b(\tau ,\vec k) },\nonumber \\ &&{\bf B}_{a\vec k}(\tau ;X^A_{\phi}+{1\over 2}R^A_{\phi},{\bf K}_a] =i\omega (k)[{\tilde X}^{\tau}_{\phi}+{1\over 2}R^{\tau}_{\phi}(\tau )]-i\vec k\cdot [{\vec X}_{\phi}+{1\over 2}{\vec R}_{\phi}(\tau )]+\nonumber \\ &+&\int d\tilde qd{\tilde q}^{'}\, {\bf K}_a(\tau ,\vec q) {\cal G}(\vec q ,{\vec q}^{'}) \triangle ({\vec q}^{'},\vec k),\nonumber \\ &&{\bf B}_{b\vec k}(\tau ;X^A_{\phi}-{1\over 2}R^A_{\phi},{\bf K}_a] =i\omega (k)[{\tilde X}^{\tau}_{\phi}-{1\over 2}R^{\tau}_{\phi}(\tau )]-i\vec k\cdot [{\vec X}_{\phi}-{1\over 2}{\vec R}_{\phi}(\tau )]+\nonumber \\ &+&\int d\tilde qd{\tilde q}^{'}\, {\bf K}_b(\tau ,\vec q) {\cal G}(\vec q ,{\vec q}^{'}) \triangle ({\vec q}^{'},\vec k), \label{VIII17} \end{eqnarray} In the gauge $\chi =T_s-\tau \approx 0$ we get \begin{eqnarray} \phi (T_s,\vec \sigma )&=& \int d\tilde k \Big(\, \sqrt{F^{\tau}(k)\omega (k)[{1\over 2}P^{\tau} _{\phi}+Q^{\tau}_{\phi}(T_s)]-F(k)\vec k\cdot {\vec Q}_{\phi}(T_s)+{\cal D}_k {\bf H}_a(T_s , \vec k) }\nonumber \\ &&e^{i\int d\tilde q\int d\tilde q^{\prime }\, {\bf K}_a( T_s ,\vec q) {\cal G}\left( \vec q,\vec q^{\prime }\right) \Delta \left( \vec q^{\prime } ,\vec k\right) +i\omega (k) [{\tilde X}_{\phi}^\tau +{1\over 2}{\tilde R}^{\tau} _{\phi}(T_s)] +i\vec k\cdot \left(\vec \sigma -{1\over 2} {\vec R}_{\phi}(T_s)\right) }+\nonumber \\ &&+\sqrt{F^{\tau}(k)\omega (k)[{1\over 2}P^{\tau}_{\phi}-Q^{\tau}_{\phi}(T_s)] +F(k)\vec k\cdot {\vec Q}_{\phi}(T_s)+ {\cal D}_k {\bf H}_b(T_s , \vec k) }\nonumber \\ &&e^{-i\int d\tilde q\int d\tilde q^{\prime }\, {\bf K}_b( T_s ,\vec q) {\cal G}\left( \vec q,\vec q^{\prime }\right) \Delta \left( \vec q^{\prime } ,\vec k\right) -i\omega (k) [{\tilde X}_{\phi}^\tau -{1\over 2}{\tilde R}^{\tau} _{\phi}(T_s)] -i\vec k\cdot \left(\vec \sigma + {1\over 2}{\vec R}_{\phi}(T_s)\right) } \Big),\nonumber \\ &&{}\nonumber \\ \pi (T_s ,\vec \sigma )&=&-i \int d\tilde k \omega (k)\nonumber \\ && \Big( \sqrt{F^{\tau}(k)\omega (k)[{1\over 2}P^{\tau}_{\phi}-Q^{\tau}_{\phi}(T_s)] +F(k)\vec k\cdot {\vec Q}_{\phi}(T_s)+ {\cal D}_k {\bf H}_b(T_s , \vec k) }\nonumber \\ &&e^{+i\int d\tilde q\int d\tilde q^{\prime }\, {\bf K}_b( T_s ,\vec q) {\cal G}\left( \vec q,\vec q^{\prime }\right) \Delta \left( \vec q^{\prime } ,\vec k\right) +i\omega (k) [{\tilde X}_{\phi}^\tau -{1\over 2}{\tilde R}^{\tau} _{\phi}(T_s)] +i\vec k\cdot \left(\vec \sigma + {1\over 2}{\vec R}_{\phi}(T_s)\right) }-\nonumber \\ &&-\sqrt{F^{\tau}(k)\omega (k)[{1\over 2}P^{\tau} _{\phi}+Q^{\tau}_{\phi}(T_s)]-F(k)\vec k\cdot {\vec Q}_{\phi}(T_s)+{\cal D}_k {\bf H}_a(T_s , \vec k) }\nonumber \\ &&e^{-i\int d\tilde q\int d\tilde q^{\prime }\, {\bf K}_a( T_s ,\vec q) {\cal G}\left( \vec q,\vec q^{\prime }\right) \Delta \left( \vec q^{\prime } ,\vec k\right) -i\omega (k) [{\tilde X}_{\phi}^\tau +{1\over 2}{\tilde R}^{\tau} _{\phi}(T_s)] -i\vec k\cdot \left(\vec \sigma -{1\over 2} {\vec R}_{\phi}(T_s)\right) } ),\nonumber \\ &&{}\nonumber \\ N_{\phi} &=& P^{\tau}_{\phi} \int d\tilde k \omega (k) F^{\tau}(k) +\int d\tilde k (-F(k) \vec k\cdot {\vec P}_{\phi} +{\cal D}_k[{\bf H}_a(T_s,\vec k) +{\bf H}_b(T_s,\vec k)])\approx \nonumber \\ &&\approx P^{\tau}_{\phi} \int d\tilde k \omega (k) F^{\tau}(k) +\int d\tilde k {\cal D}_k[{\bf H}_a(T_s,\vec k)+{\bf H}_b(T_s,\vec k)],\nonumber \\ q_{\phi}&=&2 Q^{\tau}_{\phi}(T_s) \int d\tilde k \omega (k) F^{\tau}(k) + \int d\tilde k (-2 F(k) \vec k\cdot {\vec Q}_{\phi}(T_s) +{\cal D}_k[{\bf H}_a(T_s,\vec k)-{\bf H}_b(T_s,\vec k)]).\nonumber \\ && \label{VIII18} \end{eqnarray} \subsection{The coupling to the electromagnetic field.} Let us consider the action describing a charged Klein Gordon field interacting with the electromagnetic field on spacelike hypersurfaces following the scheme of Ref.\cite{albad} \begin{eqnarray} S&=& \int d\tau d^3\sigma N(\tau ,\vec \sigma )\sqrt{\gamma (\tau ,\vec \sigma ) }\nonumber \\ &&\{ g^{\tau\tau} (\partial_{\tau}+ieA_{\tau}) \phi^{*}\, (\partial_{\tau}-ieA _{\tau}) \phi +\nonumber \\ &&+g^{\tau \check r} [(\partial_{\tau}+ieA_{\tau}) \phi^{*}\, (\partial _{\check r}-ieA_{\check r}) \phi +(\partial_{\check r}+ieA_{\check r}) \phi^{*}\, (\partial_{\tau}-ieA_{\tau}) \phi ] +\nonumber \\ &&+g^{\check r\check s}(\partial_{\check r}+ieA_{\check r}) \phi^{*}\, (\partial _{\check s}-ieA_{\check s}) \phi - m^2 \phi^{*}\phi -{1\over 4}g^{\check A\check C}g^{\check B\check D}F_{\check A\check B}F_{\check C\check D}\, \} (\tau ,\vec \sigma )=\nonumber \\ &&=\int d\tau d^3\sigma \sqrt{\gamma (\tau ,\vec \sigma )} \{ {1\over N} [\partial_{\tau}+ieA_{\tau} -N^{\check r}(\partial_{\check r}+ieA_{\check r})] \phi^{*}\nonumber \\ &&[\partial_{\tau}-ieA_{\tau} -N^{\check s}(\partial_{\check s}-ieA_{\check s})] \phi + N [\gamma^{\check r\check s}(\partial_{\check r}+ieA_{\check r})\phi^{*} (\partial_{\check s}-ieA_{\check s})\phi -m^2\phi^{*}\phi ]-\nonumber \\ &&-{1\over {2N}}(F_{\tau\check r}-N^{\check u}F_{\check u\check r})\gamma ^{\check r\check s}(F_{\tau \check s}-N^{\check v}F_{\check v\check s})-{N\over 4} \gamma^{\check r\check s}\gamma^{\check u\check v}F_{\check r\check u} F_{\check s\check v}\, \}(\tau ,\vec \sigma ). \label{VIII19} \end{eqnarray} \noindent where the configuration variables are $z^{\mu}(\tau ,\vec \sigma )$, $\phi (\tau ,\vec \sigma )={\tilde \phi}(z(\tau ,\vec \sigma ))$ and $A_{\check A}(\tau ,\vec \sigma )=z^{\mu}_{\check A}(\tau ,\vec \sigma ){\tilde A}_{\mu}(z (\tau ,\vec \sigma ))$ [$\tilde \phi (z)$ and ${\tilde A}_{\mu}(z)$ are the standard Klein-Gordon field and electromagnetic potential, which do not know the embedding of the spacelike hypersurface $\Sigma$ in Minkowski spacetime like $\phi$ and $A_{\check A}$]. The canonical momenta are \begin{eqnarray} \pi^{\tau}(\tau ,\vec \sigma )&=&{{\partial L}\over {\partial \partial_{\tau} A_{\tau}(\tau ,\vec \sigma )}}=0,\nonumber \\ \pi^{\check r}(\tau ,\vec \sigma )&=&{{\partial L}\over {\partial \partial _{\tau}A_{\check r}(\tau ,\vec \sigma )}}= -{{\sqrt{\gamma (\tau ,\vec \sigma )}}\over {N(\tau ,\vec \sigma )}}\gamma^{\check r\check s}(\tau ,\vec \sigma )(F_{\tau \check s}-N^{\check u}F_{\check u\check s})(\tau ,\vec \sigma ),\nonumber \\ \pi_{\phi}(\tau ,\vec \sigma )&=&{{\partial L}\over {\partial \partial_{\tau} \phi (\tau ,\vec \sigma )}}= {{\sqrt{\gamma (\tau ,\vec \sigma )}}\over {N(\tau ,\vec \sigma )}}[\partial_{\tau}+ieA_{\tau}-N^{\check r}(\partial _{\check r}+ieA_{\check r})](\tau ,\vec \sigma ) \phi^{*}(\tau ,\vec \sigma ), \nonumber \\ \pi_{\phi^{*}}(\tau ,\vec \sigma )&=&{{\partial L}\over {\partial \partial _{\tau}\phi^{*}(\tau ,\vec \sigma )}}= {{\sqrt{\gamma (\tau ,\vec \sigma )}}\over {N(\tau ,\vec \sigma )}} [\partial_{\tau}-ieA_{\tau}-N^{\check r}(\partial _{\check r}-ieA_{\check r})](\tau ,\vec \sigma ) \phi (\tau ,\vec \sigma ), \nonumber \\ \rho_{\mu}(\tau ,\vec \sigma )&=&-{{\partial L}\over {\partial \partial_{\tau} z^{\mu}(\tau ,\vec \sigma )}}=\nonumber \\ &=&l_{\mu}(\tau ,\vec \sigma ) \{ {{\pi_{\phi} \pi_{\phi^{*}}}\over {\sqrt{\gamma}}}-\sqrt{\gamma} [\gamma^{\check r\check s} (\partial_{\check r}+ieA_{\check r})\phi^{*} (\partial_{\check s}-ieA_{\check s})\phi -\nonumber \\ &&-m^2\phi^{*}\phi] +{1\over {2\sqrt{\gamma}}}\pi^{\check r}g_{\check r\check s}\pi^{\check s}-{{\sqrt{\gamma}}\over 4}\gamma^{\check r\check s}\gamma ^{\check u\check v}F_{\check r\check u}F_{\check s\check v} \}(\tau ,\vec \sigma )+\nonumber \\ &&+z_{\check s\mu}(\tau ,\vec \sigma )\gamma^{\check r\check s}(\tau ,\vec \sigma ) \{ \pi_{\phi^{*}}(\partial_{\check r}+ieA_{\check r})\phi^{*}+ \pi_{\phi} (\partial_{\check r}-ieA_{\check r})\phi-F_{\check r\check u} \pi^{\check u} \} (\tau ,\vec \sigma ). \label{VIII20} \end{eqnarray} Therefore, we have the following primary constraints \begin{eqnarray} \pi^{\tau}(\tau ,\vec \sigma )&\approx& 0,\nonumber \\ {\cal H}_{\mu}(\tau ,\vec \sigma )&=& \rho_{\mu}(\tau ,\vec \sigma )- \nonumber \\ &&-l_{\mu}(\tau ,\vec \sigma ) \{ {{\pi_{\phi} \pi_{\phi^{*}}}\over {\sqrt{\gamma}}}-\sqrt{\gamma} [\gamma^{\check r\check s} (\partial_{\check r}+ieA_{\check r})\phi^{*} (\partial_{\check s}-ieA_{\check s})\phi -\nonumber \\ &&-m^2\phi^{*}\phi] +{1\over {2\sqrt{\gamma}}}\pi^{\check r}g_{\check r\check s}\pi^{\check s}-{{\sqrt{\gamma}}\over 4}\gamma^{\check r\check s}\gamma ^{\check u\check v}F_{\check r\check u}F_{\check s\check v} \}(\tau ,\vec \sigma )+\nonumber \\ &&+z_{\check s\mu}(\tau ,\vec \sigma )\gamma^{\check r\check s}(\tau ,\vec \sigma ) \{ \pi_{\phi^{*}}(\partial_{\check r}+ieA_{\check r})\phi^{*}+ \pi_{\phi} (\partial_{\check r}-ieA_{\check r})\phi-F_{\check r\check u} \pi^{\check u} \} (\tau ,\vec \sigma ) \approx 0,\nonumber \\ && \label{VIII21} \end{eqnarray} \noindent and the following Dirac Hamiltonian [$\lambda (\tau ,\vec \sigma )$ and $\lambda^{\mu}(\tau ,\vec \sigma )$ are Dirac multiplier] \begin{equation} H_D=\int d^3\sigma [-A_{\tau}(\tau ,\vec \sigma ) \Gamma (\tau ,\vec \sigma )+ \lambda (\tau ,\vec \sigma )\pi^{\tau}(\tau ,\vec \sigma )+\lambda^{\mu} (\tau ,\vec \sigma ){\cal H}_{\mu}(\tau ,\vec \sigma )]. \label{VIII22} \end{equation} By using the Poisson brackets \begin{eqnarray} \{ z^{\mu}(\tau ,\vec \sigma ),\rho_{\nu}(\tau ,{\vec \sigma}^{'}) \} &=&\eta ^{\mu}_{\nu} \delta^3(\vec \sigma -{\vec \sigma}^{'}),\nonumber \\ \{ A_{\check A}(\tau ,\vec \sigma ),\pi^{\check B}(\tau ,{\vec \sigma}^{'}) \} &=&\eta^{\check B}_{\check A} \delta^3(\vec \sigma -{\vec \sigma}^{'}), \nonumber \\ \{ \phi (\tau ,\vec \sigma ),\pi_{\phi}(\tau ,{\vec \sigma}^{'}) \} &=& \{ \phi^{*}(\tau ,\vec \sigma ),\pi_{\phi^{*}}(\tau ,{\vec \sigma}^{'}) \} = \delta^3(\vec \sigma -{\vec \sigma}^{'}), \label{VIII23} \end{eqnarray} \noindent we find that the time constancy of the primary constraints implies the existence of only one secondary constraint \begin{equation} \Gamma (\tau ,\vec \sigma ) =\partial_{\check r}\pi^{\check r}(\tau ,\vec \sigma )+ie (\pi_{\phi^{*}} \phi^{*}-\pi_{\phi} \phi )(\tau ,\vec \sigma ) \approx 0. \label{VIII24} \end{equation} We can verify that these constraints are first class with the algebra given in Eqs.(125) of Ref.\cite{lus}. The Poincar\'e generators are like in Eq.(\ref{II6}). On spacelike hyperplanes $z^{\mu}(\tau ,\vec \sigma )=x_s^{\mu}(\tau )+ b^{\mu}_{\check r}(\tau )\sigma^{\check r}$, the constraints are reduced to the following ones \begin{eqnarray} \pi^{\tau}(\tau ,\vec \sigma )&\approx& 0,\nonumber \\ \Gamma (\tau ,\vec \sigma )&=&-\vec \partial \vec \pi (\tau ,\vec \sigma )+ ie [\pi_{\phi^{*}}\phi^{*}-\pi_{\phi}\phi ](\tau ,\vec \sigma )\approx 0, \nonumber \\ {\tilde {\cal H}}^{\mu}(\tau )&=& \int d^3\sigma {\cal H}^{\mu}(\tau ,\vec \sigma )=\nonumber \\ &=&p^{\mu}_s-l^{\mu} \{ {1\over 2} \int d^3\sigma [{\vec \pi}^2+{\vec B}^2] (\tau ,\vec \sigma )+\nonumber \\ &+&\int d^3\sigma [\pi_{\phi^{*}} \pi_{\phi}+(\vec \partial +ie\vec A) \phi^{*} \cdot (\vec \partial -ie\vec A)\phi +m^2\phi^{*}\phi ](\tau ,\vec \sigma ) \}-\nonumber \\ &-&b^{\mu}_{\check r}(\tau ) \{ \int d^3\sigma (\vec \pi \times \vec B)_{\check r}(\tau ,\vec \sigma )+\int d^3\sigma [\pi_{\phi^{*}}(\partial_{\check r}+ ieA_{\check r})\phi^{*}+\nonumber \\ &+&\pi_{\phi} (\partial_{\check r}-ieA_{\check r}) \phi ](\tau ,\vec \sigma ) \} \approx 0,\nonumber \\ {\tilde {\cal H}}^{\mu\nu}(\tau )&=&b^{\mu}_{\check r}(\tau ) \int d^3\sigma \sigma^{\check r} {\cal H}^{\nu}(\tau ,\vec \sigma )-b^{\nu}_{\check r}(\tau ) \int d^3\sigma \sigma^{\check r} {\cal H}^{\mu}(\tau ,\vec \sigma )=\nonumber \\ &=&S^{\mu\nu}_s-(b^{\mu}_{\check r}(\tau )l^{\nu}-b^{\nu}_{\check r}(\tau ) l^{\mu}) [{1\over 2}\int d^3\sigma \sigma^{\check r} ({\vec \pi}^2+{\vec B}^2) (\tau ,\vec \sigma )+\nonumber \\ &+&\int d^3\sigma \sigma^{\check r}[\pi_{\phi^{*}}\pi_{\phi}+(\vec \partial + ie\vec A)\phi^{*}\cdot (\vec \partial -ie\vec A)\phi +m^2\phi^{*}\phi ](\tau , \vec \sigma ) \}+\nonumber \\ &+&(b^{\mu}_{\check r}(\tau )b^{\nu}_{\check s}(\tau )-b^{\nu}_{\check r}(\tau )b^{\mu}_{\check s}(\tau )) \{ \int d^3\sigma \sigma^{\check r} (\vec \pi \times \vec B)_{\check s}(\tau ,\vec \sigma )+\nonumber \\ &+&\int d^3\sigma \sigma^{\check r} [\pi_{\phi^{*}}(\partial_{\check s}+ieA _{\check s})\phi^{*}+\pi_{\phi} (\partial_{\check s}-ieA_{\check s})\phi ] (\tau ,\vec \sigma ) \} \approx 0. \label{VIII25} \end{eqnarray} With the final restriction to Wigner hyperplanes, where the canonical variables are ${\tilde x} ^{\mu}_s(\tau )$, $p^{\mu}_s$, $A_{\tau}(\tau ,\vec \sigma )$, $\pi^{\tau} (\tau ,\vec \sigma )$, $\vec A(\tau ,\vec \sigma )$, $\vec \pi (\tau ,\vec \sigma )$, $\phi (\tau ,\vec \sigma )$, $\pi_{\phi}(\tau ,\vec \sigma )= [{\dot \phi}^{*}-ieA_{\tau}](\tau ,\vec \sigma )$, $\phi^{*}(\tau ,\vec \sigma )$, $\pi_{\phi^{*}}(\tau ,\vec \sigma )= [\dot \phi -ieA_{\tau}](\tau ,\vec \sigma )$, the only surviving constraints are \begin{eqnarray} \pi^{\tau}(\tau ,\vec \sigma )&\approx& 0,\nonumber \\ \Gamma (\tau ,\vec \sigma )&=&-\vec \partial \vec \pi (\tau ,\vec \sigma )+ ie [\pi_{\phi^{*}}\phi^{*}-\pi_{\phi}\phi ](\tau ,\vec \sigma )\approx 0, \nonumber \\ {\cal H}(\tau )&=& \epsilon_s-\{ {1\over 2} \int d^3\sigma ({\vec \pi}^2+{\vec B}^2)(\tau ,\vec \sigma )+\nonumber \\ &+&\int d^3\sigma [\pi_{\phi^{*}}\phi^{*} \pi_{\phi}+(\vec \partial +ie\vec A) \phi^{*} \cdot (\vec \partial -ie\vec A)\phi +m^2\phi^{*}\phi ](\tau ,\vec \sigma ) \} \approx 0,\nonumber \\ {\vec {\cal H}}_p(\tau )&=& \int d^3\sigma (\vec \pi \times \vec B)(\tau ,\vec \sigma )+\nonumber \\ &+&\int d^3\sigma [\pi_{\phi^{*}} (\vec \partial +ie\vec A)\phi^{*} +\pi_{\phi} (\vec \partial -ie\vec A)\phi ](\tau ,\vec \sigma ) \approx 0. \label{VIII26} \end{eqnarray} In the Lorentz generators of Eq.(\ref{II9}) the spin tensor has the form \begin{eqnarray} {\bar S}_s^{rs}&\equiv& S^{rs}_{\phi} =\int d^3\sigma \{ \sigma^r (\vec \pi \times \vec B)^s(\tau ,\vec \sigma )-\sigma^s (\vec \pi \times \vec B)^r(\tau ,\vec \sigma ) \}+ \nonumber \\ &+&\int d^3\sigma \{ \sigma^r[\pi_{\phi^{*}}(\partial^s+ieA^s)\phi^{*}+\pi _{\phi}(\partial^s-ieA^s)\phi](\tau ,\vec \sigma )-(r \leftrightarrow s) \}. \label{VIII27} \end{eqnarray} To make the reduction to Dirac's observables with respect to the electromagnetic gauge transformations, let us recall \cite{lusa,val} that the electromagnetic gauge degrees of freedom are described by the two pairs of conjugate variables $A_{\tau}(\tau ,\vec \sigma )$, $\pi_{\tau}(\tau ,\vec \sigma )[\approx 0]$, $\eta_{em}(\tau ,\vec \sigma )=-{1\over {\triangle}} {{\partial}\over {\partial \vec \sigma}}\cdot \vec A(\tau ,\vec \sigma )$, $\Gamma (\tau ,\vec \sigma ) [\approx 0]$, so that we have the decompositions \begin{eqnarray} A^r(\tau ,\vec \sigma )&=&{{\partial}\over {\partial \sigma^r}} \eta_{em}(\tau ,\vec \sigma )+A^r_{\perp}(\tau ,\vec \sigma ),\nonumber \\ \pi^r(\tau ,\vec \sigma )&=&\pi^r_{\perp}(\tau ,\vec \sigma )+\nonumber \\ &+&{1\over {\triangle}} {{\partial}\over {\partial \sigma^r}} [-\Gamma (\tau ,\vec \sigma )+ie(\pi_{\phi^{*}}\phi^{*}-\pi_{\phi}\phi )(\tau ,\vec \sigma )] \approx 0,\nonumber \\ &&{}\nonumber \\ \lbrace A^r_{\perp}(\tau ,\vec \sigma )&,&\pi^s_{\perp}(\tau ,{\vec \sigma}^{'}) \rbrace =-P^{rs}_{\perp}(\vec \sigma ) \delta^3(\vec \sigma -{\vec \sigma}^{'}), \label{VIII28} \end{eqnarray} \noindent where $P^{rs}_{\perp}(\vec \sigma )=\delta^{rs}+{{\partial^r \partial^s}\over {\triangle}}$, $\triangle =-{\vec \partial}^2$. Then, we have \begin{eqnarray} \int d^3\sigma &&{\vec \pi}^2(\tau ,\vec \sigma )= \int d^3\sigma {\vec \pi}^2 _{\perp}(\tau ,\vec \sigma )-\nonumber \\ &-&{{e^2}\over {4\pi}} \int d^3\sigma_1d^3\sigma_2 {{i(\pi_{\phi^{*}}\phi^{*}- \pi_{\phi}\phi )(\tau ,{\vec \sigma}_1)\, i(\pi_{\phi^{*}}\phi^{*}- \pi_{\phi}\phi )(\tau ,{\vec \sigma}_2)}\over {|{\vec \sigma}_1-{\vec \sigma} _2|}}. \label{VIII29} \end{eqnarray} Since we have \begin{eqnarray} \lbrace \phi (\tau ,\vec \sigma ),\Gamma (\tau ,{\vec \sigma}^{'})\rbrace &=& -ie \phi(\tau ,\vec \sigma ) \delta^3(\vec \sigma -{\vec \sigma}^{'}), \nonumber \\ \lbrace \phi^{*}(\tau ,\vec \sigma ),\Gamma (\tau ,{\vec \sigma}^{'})\rbrace &=& ie \phi^{*}(\tau ,\vec \sigma ) \delta^3(\vec \sigma -{\vec \sigma}^{'}), \nonumber \\ \lbrace \pi_{\phi}(\tau ,\vec \sigma ),\Gamma (\tau ,{\vec \sigma}^{'})\rbrace &=&ie \pi_{\phi}(\tau ,\vec \sigma ) \delta^3(\vec \sigma -{\vec \sigma}^{'}), \nonumber \\ \lbrace \pi_{\phi^{*}}(\tau ,\vec \sigma ),\Gamma (\tau ,{\vec \sigma}^{'}) \rbrace&=&-ie \pi_{\phi^{*}}(\tau ,\vec \sigma ) \delta^3(\vec \sigma -{\vec \sigma}^{'}), \label{VIII30} \end{eqnarray} \noindent the Dirac observables for the Klein-Gordon field are \begin{eqnarray} \hat \phi (\tau ,\vec \sigma )&=& e^{-ie\eta_{em}(\tau ,\vec \sigma )} \phi (\tau ,\vec \sigma ),\nonumber \\ {\hat \phi}^{*}(\tau ,\vec \sigma )&=& e^{ie\eta_{em}(\tau ,\vec \sigma )} \phi^{*}(\tau ,\vec \sigma ),\nonumber \\ {\hat \pi}_{\phi}(\tau ,\vec \sigma )&=&e^{ie\eta_{em}(\tau ,\vec \sigma )} \pi_{\phi}(\tau ,\vec \sigma ),\nonumber \\ {\hat \pi}_{\phi^{*}}(\tau ,\vec \sigma )&=&e^{-ie\eta_{em}(\tau ,\vec \sigma )} \pi_{\phi^{*}}(\tau ,\vec \sigma ),\nonumber \\ &&{}\nonumber \\ \lbrace \hat \phi (\tau ,\vec \sigma )&,&\Gamma (\tau ,{\vec \sigma}^{'})\rbrace =\lbrace {\hat \pi}_{\phi}(\tau ,\vec \sigma ),\Gamma (\tau ,{\vec \sigma}^{'}) \rbrace =0,\nonumber \\ \lbrace {\hat \phi}^{*}(\tau ,\vec \sigma )&,&\Gamma (\tau ,{\vec \sigma}^{'}) \rbrace =\lbrace {\hat \pi}_{\phi^{*}}(\tau ,\vec \sigma ),\Gamma (\tau ,{\vec \sigma}^{'})\rbrace =0,\nonumber \\ &&{}\nonumber \\ \lbrace \hat \phi (\tau ,\vec \sigma ),{\hat \pi}_{\phi}(\tau ,{\vec \sigma} ^{'}) \rbrace &=& \lbrace {\hat \phi}^{*}(\tau ,\vec \sigma ),{\hat \pi} _{\phi^{*}}(\tau ,{\vec \sigma}^{'}) \rbrace = \delta^3(\vec \sigma ,{\vec \sigma}^{'}). \label{VIII31} \end{eqnarray} Therefore, in the generalized Wigner-covariant Coulomb gauge $A_{\tau}(\tau ,\vec \sigma )=\eta_{em}(\tau ,\vec \sigma )=0$, we have ${\hat \phi}^{*} (\tau ,\vec \sigma )=[\hat \phi (\tau ,\vec \sigma )]^{*}$, ${\hat \pi}_{\phi ^{*}}(\tau ,\vec \sigma )=[{\hat \pi}_{\phi}(\tau ,\vec \sigma )]^{*}$ like in the free case. The constraints take the following form \begin{eqnarray} {\cal H}(\tau )&=&\epsilon_s- \{ \, {1\over 2}\int d^3\sigma ({\vec \pi}^2 _{\perp}+{\vec B}^2)(\tau ,\vec \sigma )+\nonumber \\ &+&\int d^3\sigma [{\hat \pi}_{\phi^{*}}{\hat \pi}_{\phi}+(\vec \partial +ie {\vec A}_{\perp}){\hat \phi}^{*}\cdot (\vec \partial -ie{\vec A}_{\perp})\hat \phi +m^2{\hat \phi}^{*}\hat \phi ](\tau ,\vec \sigma )-\nonumber \\ &-&{{e^2}\over {8\pi}}\int d^3\sigma_1d^3\sigma_2 {{i({\hat \pi}_{\phi^{*}}{\hat \phi}^{*}-{\hat \pi}_{\phi}\hat \phi )(\tau ,{\vec \sigma}_1)\, i({\hat \pi} _{\phi^{*}}{\hat \phi}^{*}-{\hat \pi}_{\phi}\hat \phi )(\tau ,{\vec \sigma_2)}} \over {|{\vec \sigma}_1-{\vec \sigma}_2|}} \} ,\nonumber \\ &&{}\nonumber \\ {\vec {\cal H}}_p(\tau )&=&\int d^3\sigma ({\vec \pi}_{\perp}\times \vec B) (\tau ,\vec \sigma )+\int d^3\sigma ({\hat \pi}_{\phi^{*}}\vec \partial {\hat \phi}^{*}+{\hat \pi}_{\phi}\vec \partial \hat \phi )(\tau ,\vec \sigma ) \approx 0, \label{VIII32} \end{eqnarray} \noindent where the Coulomb self-interaction appears in the invariant mass and where the 3 constraints defining the rest frame do not depend on the interaction since we are in an instant form of the dynamics. The final form of the rest-frame spin tensor is \begin{eqnarray} {\bar S}^{rs}_s&=& \int d^3\sigma \{ \sigma^r [({\vec \pi}_{\perp}\times \vec B)^s +{\hat \pi}_{\phi^{*}}\partial^s{\hat \phi}^{*}+{\hat \pi}_{\phi}\partial^s \hat \phi ] - (r\leftrightarrow s) \} (\tau ,\vec \sigma )=\nonumber \\ &=& S^{rs}_{\phi}(\tau )+\int d^3\sigma \{ \sigma^r [({\vec \pi}_{\perp}\times \vec B)^s- (r\leftrightarrow s) \} (\tau ,\vec \sigma ). \label{VIII33} \end{eqnarray} If we go to the gauge $\chi =T_s-\tau \approx 0$, we can eliminate the variables $\epsilon_s$, $T_s$, and the $\tau$-evolution (in the Lorentz scalar rest-frame time) is governed by the Hamiltonian \begin{eqnarray} H_R&=&M-\vec \lambda (\tau )\cdot {\vec {\cal H}}_p(\tau ),\nonumber \\ &&{}\nonumber \\ M&=&{1\over 2}\int d^3\sigma ({\vec \pi}^2 _{\perp}+{\vec B}^2)(\tau ,\vec \sigma )+\nonumber \\ &+&\int d^3\sigma [{\hat \pi}_{\phi^{*}}{\hat \pi}_{\phi}+(\vec \partial +ie {\vec A}_{\perp}){\hat \phi}^{*}\cdot (\vec \partial -ie{\vec A}_{\perp})\hat \phi +m^2{\hat \phi}^{*}\hat \phi ](\tau ,\vec \sigma )-\nonumber \\ &-&{{e^2}\over {8\pi}}\int d^3\sigma_1d^3\sigma_2 {{i({\hat \pi}_{\phi^{*}}{\hat \phi}^{*}-{\hat \pi}_{\phi}\hat \phi )(\tau ,{\vec \sigma}_1)\, i({\hat \pi} _{\phi^{*}}{\hat \phi}^{*}-{\hat \pi}_{\phi}\hat \phi )(\tau ,{\vec \sigma)_2)}} \over {|{\vec \sigma}_1-{\vec \sigma}_2|}}. \label{VIII34} \end{eqnarray} In the gauge $\vec \lambda (\tau )=0$, the Hamilton equations are \begin{eqnarray} \partial_{\tau} \hat \phi (\tau ,\vec \sigma )\, &{\buildrel \circ \over =}\,& {\hat \pi}_{\phi^{*}}(\tau ,\vec \sigma )+\nonumber \\ &+&{{ie^2}\over {4\pi}}\hat \phi (\tau ,\vec \sigma ) \int d^3\bar \sigma {{i({\hat \pi}_{\phi^{*}}{\hat \phi}^{*}-{\hat \pi}_{\phi}\hat \phi )(\tau , {\vec {\bar \sigma}})}\over {|\vec \sigma -{\vec {\bar \sigma}}|}}, \nonumber \\ \partial_{\tau}{\hat \pi}_{\phi^{*}}(\tau ,\vec \sigma )\, &{\buildrel \circ \over =}\,& [(\vec \partial -ie{\vec A}_{\perp}(\tau ,\vec \sigma ))^2-m^2] \hat \phi (\tau ,\vec \sigma )+\nonumber \\ &+&{{ie^2}\over {4\pi}}{\hat \pi}_{\phi^{*}}(\tau ,\vec \sigma )\int d^3\bar \sigma {{i({\hat \pi}_{\phi^{*}}{\hat \phi}^{*}-{\hat \pi}_{\phi}\hat \phi ) (\tau ,{\vec {\bar \sigma}})}\over {|\vec \sigma -{\vec {\bar \sigma}}|}}, \nonumber \\ &&{}\nonumber \\ \partial_{\tau}A^r_{\perp}(\tau ,\vec \sigma )\, &{\buildrel \circ \over =}\,& -\pi^r_{\perp}(\tau ,\vec \sigma ),\nonumber \\ \partial_{\tau}\pi^r_{\perp}(\tau ,\vec \sigma )\, &{\buildrel \circ \over =}\,& \triangle A^r_{\perp}(\tau ,\vec \sigma )+\nonumber \\ &+&ie P^{rs}_{\perp}(\vec \sigma ) [{\hat \phi}^{*}(\partial^s-ieA^s_{\perp}) \hat \phi -\hat \phi (\partial^s+ieA^s_{\perp}){\hat \phi}^{*}](\tau ,\vec \sigma ). \label{VIII35} \end{eqnarray} \noindent The equations for ${\hat \phi}^{*}$ and $\pi_{\phi}$ are the complex conjugate of those for $\hat \phi$ and for ${\hat \pi}_{\phi^{*}}$. By using the results of Ref.\cite{val}, we have the following inversion formula \begin{eqnarray} {\hat \pi}_{\phi^{*}}\, &{\buildrel \circ \over =}\,& \partial_{\tau} \hat \phi +ie^2 \hat \phi {1\over {\triangle}} i({\hat \pi}_{\phi^{*}}{\hat \phi}^{*} -{\hat \pi}_{\phi}\hat \phi )=\nonumber \\ &=&\partial_{\tau} \hat \phi +ie^2 \hat \phi {1\over {\triangle +2e^2{\hat \phi}^{*}\hat \phi}} i({\hat \phi}^{*} \partial_{\tau}\hat \phi -\hat \phi \partial_{\tau} {\hat \phi}^{*}), \label{VIII36} \end{eqnarray} \noindent since we have $i({\hat \phi}^{*}\partial_{\tau}\hat \phi -\hat \phi \partial_{\tau}{\hat \phi}^{*})=(1+2e^2{\hat \phi}^{*}\hat \phi {1\over {\triangle}}) i({\hat \pi}_{\phi^{*}}{\hat \phi}^{*}-{\hat \pi}_{\phi}\hat \phi )$ and where use has been done of the operator identity ${1\over A}{1\over {1+B{1\over A}}}={1\over A}[1-B{1\over A}+B{1\over A}B{1\over A}-...]={1\over {A+B}}$ (valid for B a small perturbation of A) for $A=\triangle$ and $B=2e^2 {\hat \phi}^{*}\hat \phi$. Using this formula, we get the following second order equations of motion \begin{eqnarray} &&\{ [\, \partial_{\tau}+ie^2{1\over {\triangle +2e^2{\hat \phi}^{*}\hat \phi}} \, i({\hat \phi}^{*} \partial_{\tau}\hat \phi -\hat \phi \partial_{\tau}{\hat \phi}^{*})]^2-(\vec \partial -ie{\vec A}_{\perp})^2+m^2\, \} \hat \phi {\buildrel \circ \over =}\, 0,\nonumber \\ &&{}\nonumber \\ &&[\partial_{\tau}^2+\triangle ] A^r_{\perp}\, {\buildrel \circ \over =}\, ie P^{rs}_{\perp}(\vec \sigma ) [{\hat \phi}^{*}(\partial^s -ieA^s_{\perp})\hat \phi -\hat \phi (\partial^s+ieA^s_{\perp}){\hat \phi}^{*}]. \label{VIII37} \end{eqnarray} We see that the non-local velocity-dependent self-energy is formally playing the role of a scalar potential. Due to Eqs.(\ref{VIII31}), in the gauge $A_{\tau}(\tau ,\vec \sigma )=\eta_{em} (\tau ,\vec \sigma )=0$ the Dirac observables $\hat \phi$, ${\hat \pi}_{\phi}$, ${\hat \phi}^{*}$, ${\hat \pi}_{\phi^{*}}$, admit a Fourier decomposition like in the free case, see Eqs.(\ref{VIII10}) and (\ref{d5}), with Fourier coefficients $\hat a$, $\hat b$ [the fields ${\hat \phi}_a$, ${\hat \phi}_b$ and the Fourier coefficients $\hat a$, $\hat b$, tend to the free ones $\phi_a$, $\phi_b$, $a$, $b$, when we turn off the electromagnetic interaction, switching off the electric charge] \begin{eqnarray} \hat a(\tau ,\vec q)&=&\sqrt{2m\omega (q)} \int d^3\sigma {\hat \phi}_a(\tau ,\vec \sigma )e^{i(\omega (q)\tau -\vec q\cdot \vec \sigma )}=\nonumber \\ &=&\int d^3\sigma [\omega (q) \hat \phi (\tau ,\vec \sigma )+i{\hat \pi} _{\phi^{*}}(\tau ,\vec \sigma )] e^{i(\omega (q)\tau -\vec q\cdot \vec \sigma )}=\nonumber \\ &=& \sqrt{{\hat I}_a(\tau ,\vec q)}e^{i{\hat \varphi}_a(\tau ,\vec q)}, \nonumber \\ {\hat a}^{*}(\tau ,\vec q)&=&\sqrt{2m\omega (q)} \int d^3\sigma {\hat \phi} ^{*}_a(\tau ,\vec \sigma )e^{-i(\omega (q)\tau -\vec q\cdot \vec \sigma )}=\nonumber \\ &=&\int d^3\sigma [\omega (q) {\hat \phi}^{*}(\tau ,\vec \sigma )+i{\hat \pi} _{\phi}(\tau ,\vec \sigma )] e^{-i(\omega (q)\tau -\vec q\cdot \vec \sigma )}=\nonumber \\ &=& \sqrt{{\hat I}_a(\tau ,\vec q)}e^{-i{\hat \varphi}_a(\tau ,\vec q)}, \nonumber \\ \hat b(\tau ,\vec q)&=&\sqrt{2m\omega (q)} \int d^3\sigma {\hat \phi}_b(\tau ,\vec \sigma )e^{i(\omega (q)\tau -\vec q\cdot \vec \sigma )}=\nonumber \\ &=&\int d^3\sigma [\omega (q) \hat \phi(\tau ,\vec \sigma )-i{\hat \pi} _{\phi^{*}}(\tau ,\vec \sigma )] e^{i(\omega (q)\tau -\vec q\cdot \vec \sigma )}=\nonumber \\ &=&\sqrt{{\hat I}_b(\tau ,\vec q)}e^{i{\hat \varphi}_b(\tau ,\vec q)}, \nonumber \\ {\hat b}^{*}(\tau ,\vec q)&=&\sqrt{2m\omega (q)} \int d^3\sigma {\hat \phi} ^{*}_b(\tau ,\vec \sigma )e^{-i(\omega (q)\tau -\vec q\cdot \vec \sigma )}=\nonumber \\ &=&\int d^3\sigma [\omega (q) {\hat \phi}^{*}(\tau ,\vec \sigma )+i{\hat \pi} _{\phi}(\tau ,\vec \sigma )] e^{-i(\omega (q)\tau -\vec q\cdot \vec \sigma )}=\nonumber \\ &=&\sqrt{{\hat I}_b(\tau ,\vec q)}e^{-i{\hat \varphi}_b(\tau ,\vec q)}, \nonumber \\ &&{}\nonumber \\ {\hat \phi}_a(\tau ,\vec \sigma )&=&{1\over {\sqrt{2m\sqrt{m^2+\triangle}}}} \Big[ \sqrt{m^2+\triangle} \hat \phi (\tau ,\vec \sigma )+i{\hat \pi}_{\phi^{*}}(\tau ,\vec \sigma )\Big]=\nonumber \\ &=&\sqrt{ {2\over m} } \int d\tilde q \sqrt{\omega (q)} \hat a (\tau ,\vec q) e^{-i(\omega (q)\tau -\vec q\cdot \vec \sigma)},\nonumber \\ {\hat \phi}_b(\tau ,\vec \sigma )&=&{1\over {\sqrt{2m\sqrt{m^2+\triangle}}}} \Big[ \sqrt{m^2+\triangle} \hat \phi (\tau ,\vec \sigma )+i{\hat \pi}_{\phi^{*}}(\tau ,\vec \sigma )\Big]=\nonumber \\ &=&\sqrt{ {2\over m} } \int d\tilde q \sqrt{\omega (q)} \hat b (\tau ,\vec q) e^{-i(\omega (q)\tau -\vec q\cdot \vec \sigma)},\nonumber \\ &&{}\nonumber \\ \hat \phi (\tau ,\vec \sigma )&=&[{\hat \phi}^{*}(\tau ,\vec \sigma )]^{*}= \sqrt{ {m\over 2} }{1\over {(m^2+\triangle )^{1/4}}} [{\hat \phi}_a+{\hat \phi} ^{*}_b](\tau ,\vec \sigma ),\nonumber \\ {\hat \pi}_{\phi^{*}}(\tau ,\vec \sigma )&=&[{\hat \pi}_{\phi}(\tau ,\vec \sigma )]^{*}=-i\sqrt{ {m\over 2} }(m^2+\triangle )^{1/4} [{\hat \phi}_a-{\hat \phi}^{*}_b](\tau ,\vec \sigma ). \label{VIII38} \end{eqnarray} Let us now assume that the fields satisfy Eq.(\ref{IV29}). Then, we can define the canonical transformations $\hat a, {\hat a}^{*} \mapsto$ ${\hat P}^A_{a\phi}$, ${\hat X} ^A_{a\phi} ({\hat X}^{\tau}_{a\phi}=\tau +{\hat {\tilde X}}^{\tau}_{a\phi})$, ${\hat {\bf H}}_a$, ${\hat {\bf K}}_a$ and $\hat b, {\hat b}^{*} \mapsto$ ${\hat P}^A_{b\phi}$, ${\hat X} ^A_{b\phi} ({\hat X}^{\tau}_{b\phi}=\tau +{\hat {\tilde X}}^{\tau}_{b\phi})$, ${\hat {\bf H}}_b$, ${\hat {\bf K}}_b$, and then ${\hat X}^A_{a\phi}$, ${\hat P}^A_{a\phi}$, ${\hat X}^A_{b\phi}$, ${\hat P}^A_{b\phi} \mapsto$ ${\hat X}^A_{\phi}$, ${\hat P}^A_{\phi}$, ${\hat R}^A_{\phi}$, ${\hat Q}^A _{\phi}$. We get \begin{eqnarray} {\hat P}^{\tau}_{\phi}&=&m \int d^3\sigma [{\hat \phi}^{*}_a \sqrt{m^2+ \triangle} {\hat \phi}_a+{\hat \phi}^{*}_b \sqrt{m^2+\triangle} {\hat \phi}_b] (\tau ,\vec \sigma ),\nonumber \\ {\hat {\vec P}}_{\phi}&=& i m \int d^3\sigma [{\hat \phi}^{*}_a \vec \partial {\hat \phi}_a+{\hat \phi}^{*}_b \vec \partial {\hat \phi}_b](\tau ,\vec \sigma ), \nonumber \\ {\hat Q}^{\tau}_{\phi}&=&{m\over 2} \int d^3\sigma [{\hat \phi}^{*}_a \sqrt{m^2+ \triangle} {\hat \phi}_a-{\hat \phi}^{*}_b \sqrt{m^2+\triangle} {\hat \phi}_b] (\tau ,\vec \sigma ),\nonumber \\ {\hat {\vec Q}}_{\phi}&=& i{m\over 2} \int d^3\sigma [{\hat \phi}^{*}_a \vec \partial {\hat \phi}_a-{\hat \phi}^{*}_b \vec \partial {\hat \phi}_b](\tau ,\vec \sigma ),\nonumber \\ {\hat X}^{\tau}_{\phi}&=&\tau +{\hat {\tilde X}}^{\tau}_{\phi}=\tau + \nonumber \\ &+&{1\over {4\pi i m}} \int d^3q {{e^{-{{4\pi}\over {m^2}} q^2} }\over {q^2\, \omega (q)}} \nonumber \\ &&\Big( ln\, {{\int d^3\sigma e^{i\vec q\cdot \vec \sigma} (m^2+\triangle )^{-1/4}[(\omega (q)+\sqrt{m^2+\triangle}){\hat \phi}_a+(\omega (q)-\sqrt{m^2+ \triangle}){\hat \phi}^{*}_b](\tau ,\vec \sigma )}\over {\int d^3\sigma e^{-i\vec q\cdot \vec \sigma} (m^2+\triangle)^{-1/4}[(\omega (q)+\sqrt{m^2+ \triangle}){\hat \phi}^{*}_a+(\omega (q)-\sqrt{m^2+\triangle}){\hat \phi}_b] (\tau ,\vec \sigma )}} +\nonumber \\ &&+ln\, {{\int d^3\sigma e^{i\vec q\cdot \vec \sigma} (m^2+\triangle )^{-1/4}[(\omega (q)+\sqrt{m^2-\triangle}){\hat \phi}^{*}_a+(\omega (q)+ \sqrt{m^2+\triangle}){\hat \phi}_b](\tau ,\vec \sigma )}\over {\int d^3\sigma e^{-i\vec q\cdot \vec \sigma} (m^2+\triangle)^{-1/4}[(\omega (q)-\sqrt{m^2+ \triangle}){\hat \phi}_a+(\omega (q)+\sqrt{m^2+\triangle}){\hat \phi}^{*}_b] (\tau ,\vec \sigma )}} \Big) ,\nonumber \\ {\hat {\vec X}}_{\phi}&=&{{2i}\over {\pi m}} \int d^3q {{\vec q}\over {q^4}} e^{-{{4\pi}\over {m^2}} q^2}\nonumber \\ &&\Big( ln\, {{\int d^3\sigma e^{i\vec q\cdot \vec \sigma} (m^2+\triangle )^{-1/4}[(\omega (q)+\sqrt{m^2+\triangle}){\hat \phi}_a+(\omega (q)-\sqrt{m^2+ \triangle}){\hat \phi}^{*}_b](\tau ,\vec \sigma )}\over {\int d^3\sigma e^{-i\vec q\cdot \vec \sigma} (m^2+\triangle)^{-1/4}[(\omega (q)+\sqrt{m^2+ \triangle}){\hat \phi}^{*}_a+(\omega (q)-\sqrt{m^2+\triangle}){\hat \phi}_b] (\tau ,\vec \sigma )}} -\nonumber \\ &&-ln\, {{\int d^3\sigma e^{i\vec q\cdot \vec \sigma} (m^2+\triangle )^{-1/4}[(\omega (q)+\sqrt{m^2-\triangle}){\hat \phi}^{*}_a+(\omega (q)+ \sqrt{m^2+\triangle}){\hat \phi}_b](\tau ,\vec \sigma )}\over {\int d^3\sigma e^{-i\vec q\cdot \vec \sigma} (m^2+\triangle)^{-1/4}[(\omega (q)-\sqrt{m^2+ \triangle}){\hat \phi}_a+(\omega (q)+\sqrt{m^2+\triangle}){\hat \phi}^{*}_b] (\tau ,\vec \sigma )}} \Big) ,\nonumber \\ {\hat R}^{\tau}_{\phi}&=& {1\over {2\pi i m}} \int d^3q {{e^{-{{4\pi}\over {m^2}} q^2} }\over {q^2\, \omega (q)}} \nonumber \\ &&\Big( ln\, {{\int d^3\sigma e^{i\vec q\cdot \vec \sigma} (m^2+\triangle )^{-1/4}[(\omega (q)+\sqrt{m^2+\triangle}){\hat \phi}_a+(\omega (q)-\sqrt{m^2+ \triangle}){\hat \phi}^{*}_b](\tau ,\vec \sigma )}\over {\int d^3\sigma e^{-i\vec q\cdot \vec \sigma} (m^2+\triangle)^{-1/4}[(\omega (q)+\sqrt{m^2+ \triangle}){\hat \phi}^{*}_a+(\omega (q)-\sqrt{m^2+\triangle}){\hat \phi}_b] (\tau ,\vec \sigma )}} -\nonumber \\ &&-ln\, {{\int d^3\sigma e^{i\vec q\cdot \vec \sigma} (m^2+\triangle )^{-1/4}[(\omega (q)+\sqrt{m^2-\triangle}){\hat \phi}^{*}_a+(\omega (q)+ \sqrt{m^2+\triangle}){\hat \phi}_b](\tau ,\vec \sigma )}\over {\int d^3\sigma e^{-i\vec q\cdot \vec \sigma} (m^2+\triangle)^{-1/4}[(\omega (q)-\sqrt{m^2+ \triangle}){\hat \phi}_a+(\omega (q)+\sqrt{m^2+\triangle}){\hat \phi}^{*}_b] (\tau ,\vec \sigma )}} \Big) ,\nonumber \\ {\hat {\vec R}}_{\phi}&=&{i\over {\pi m}} \int d^3q {{\vec q}\over {q^4}} e^{-{{4\pi}\over {m^2}} q^2}\nonumber \\ &&\Big( ln\, {{\int d^3\sigma e^{i\vec q\cdot \vec \sigma} (m^2+\triangle )^{-1/4}[(\omega (q)+\sqrt{m^2+\triangle}){\hat \phi}_a+(\omega (q)-\sqrt{m^2+ \triangle}){\hat \phi}^{*}_b](\tau ,\vec \sigma )}\over {\int d^3\sigma e^{-i\vec q\cdot \vec \sigma} (m^2+\triangle)^{-1/4}[(\omega (q)+\sqrt{m^2+ \triangle}){\hat \phi}^{*}_a+(\omega (q)-\sqrt{m^2+\triangle}){\hat \phi}_b] (\tau ,\vec \sigma )}} +\nonumber \\ &&+ln\, {{\int d^3\sigma e^{i\vec q\cdot \vec \sigma} (m^2+\triangle )^{-1/4}[(\omega (q)+\sqrt{m^2-\triangle}){\hat \phi}^{*}_a+(\omega (q)+ \sqrt{m^2+\triangle}){\hat \phi}_b](\tau ,\vec \sigma )}\over {\int d^3\sigma e^{-i\vec q\cdot \vec \sigma} (m^2+\triangle)^{-1/4}[(\omega (q)-\sqrt{m^2+ \triangle}){\hat \phi}_a+(\omega (q)+\sqrt{m^2+\triangle}){\hat \phi}^{*}_b] (\tau ,\vec \sigma )}} \Big) ,\nonumber \\ &&{}\nonumber \\ {\hat \phi}_a(\tau ,\vec \sigma )&=&\int d\tilde k \sqrt{ {{2\omega (k)}\over m}} {\bf A}_{a\, \vec k}(\tau ;{1\over 2}{\hat P}^A_{\phi}+{\hat Q}^A_{\phi}, {\hat {\bf B}}_a]\, e^{i(\vec k\cdot \vec \sigma +{\bf B}_{a\, \vec k}(\tau ; {\hat X}^A_{\phi}+{1\over 2}{\hat R}^A_{\phi},{\hat {\bf K}}_a])},\nonumber \\ {\hat \phi}_b(\tau ,\vec \sigma )&=&\int d\tilde k \sqrt{ {{2\omega (k)}\over m}} {\bf A}_{b\, \vec k}(\tau ;{1\over 2}{\hat P}^A_{\phi}+{\hat Q}^A_{\phi}, {\hat {\bf B}}_b]\, e^{i(\vec k\cdot \vec \sigma +{\bf B}_{b\, \vec k}(\tau ; {\hat X}^A_{\phi}+{1\over 2}{\hat R}^A_{\phi},{\hat {\bf K}}_b])}. \label{VIII39} \end{eqnarray} In terms of these variables Eqs.(\ref{VIII34}) and (\ref{VIII33}) become \begin{eqnarray} H_R&=&M-\vec \lambda (\tau )\cdot {\vec {\cal H}} _p(\tau ),\nonumber \\ &&{}\nonumber \\ M&=& {1\over 2} \int d^3\sigma ({\vec \pi}^2_{\perp}+{\vec B}^2)(\tau ,\vec \sigma )+{\hat P}^{\tau}_{\phi}(\tau )+\nonumber \\ &+&m \int d^3\sigma \Big[ ie {\vec A}_{\perp}(\tau ,\vec \sigma )\cdot {1\over {\sqrt{m^2+\triangle}}}[{\hat \phi}^{*}_a+{\hat \phi}_b](\tau ,\vec \sigma ) {{\vec \partial}\over {\sqrt{m^2+\triangle}}}[{\hat \phi}_a+{\hat \phi} _b^{*}](\tau ,\vec \sigma )+\nonumber \\ &+&{{e^2}\over 2}{\vec A}_{\perp}^2(\tau ,\vec \sigma ) {1\over {\sqrt{m^2+ \triangle}}} [{\hat \phi}^{*}_a+{\hat \phi}_b](\tau ,\vec \sigma ) {1\over {\sqrt{m^2+\triangle}}}[{\hat \phi}_a+{\hat \phi}_b^{*}](\tau ,\vec \sigma ) \Big]-\nonumber \\ &-&{{e^2m^2}\over {32\pi}} \int {{d^3\sigma_1d^3\sigma_2}\over {|{\vec \sigma} _1-{\vec \sigma}_2|}} \Big[ \sqrt{m^2+\triangle_1}[{\hat \phi}_a-{\hat \phi}_b ^{*}] {1\over {\sqrt{m^2+\triangle_1}}}[{\hat \phi}^{*}_a+{\hat \phi}_b]- \nonumber \\ &-&\sqrt{m^2+\triangle_1}[{\hat \phi}^{*}_a-{\hat \phi}_b] {1\over {\sqrt{m^2+\triangle_1}}}[{\hat \phi}_a+{\hat \phi}_b^{*}] \Big] (\tau ,{\vec \sigma}_1) \nonumber \\ &&\Big[ \sqrt{m^2+\triangle_2}[{\hat \phi}_a-{\hat \phi}_b ^{*}] {1\over {\sqrt{m^2+\triangle_2}}}[{\hat \phi}^{*}_a+{\hat \phi}_b]- \nonumber \\ &-&\sqrt{m^2+\triangle_2}[{\hat \phi}^{*}_a-{\hat \phi}_b] {1\over {\sqrt{m^2+\triangle_2}}}[{\hat \phi}_a+{\hat \phi}_b^{*}] \Big] (\tau ,{\vec \sigma}_2),\nonumber \\ {\vec {\cal H}}_p(\tau )&=&{\hat {\vec P}}_{\phi}(\tau )+\int d^3\sigma ({\vec \pi}_{\perp}\times \vec B)(\tau ,\vec \sigma )\approx 0,\nonumber \\ {\bar S}^{rs}_s&\equiv& {\hat S}^{rs}_{\phi}(\tau )+\int d^3\sigma \Big[ \sigma^r ({\vec \pi}_{\perp}\times \vec B)^s-\sigma^s({\vec \pi}_{\perp}\times \vec B)^r\Big] . \label{VIII40} \end{eqnarray} In Appendix Ethere is the expression of the interaction terms in M in terms of $\hat a$, $\hat b$. The fields ${\hat \phi}_a$, ${\hat \phi}_b$, are solutions of the coupled Hamilton equations generated by $H_R$ and these equations cannot be decoupled when the associated Feshbach-Villars Hamiltonian cannot be diagonalized: in these cases ${\hat \phi}_a$ [${\hat \phi}_b$] does not correspond to a positive [negative] energy solution. Only when we have the collective and relative canonical variables of the transverse electromagnetic fields ${\vec A}_{\perp}$, ${\vec \pi}_{\perp}$, we will be able to add the gauge fixings for the constraints ${\vec {\cal H}} _p\approx 0$, defining the rest frame, and to get the final form of the physical Hamiltonian for classical scalar electrodynamics in the rest-frame Wigner-covariant instant form of the dynamics. \subsection{The energy-momentum tensor.} The conserved energy-momentum tensor of the isolated system formed by the Klein-Gordon field plus the electromagnetic field is [$D^{(A)}_{\mu}\tilde \phi (x)=[\partial_{\mu}-ieA_{\mu}(x)]\tilde \phi (x)$; $\tilde \phi (z(\tau ,\vec \sigma ))=\phi (\tau ,\vec \sigma )$; ${\tilde A} _{\mu}(z(\tau ,\vec \sigma ))=z^A_{\mu}(\tau ,\vec \sigma )A_A(\tau ,\vec \sigma )$; $T^{\mu\nu}_{em}(x)={\tilde F}^{\mu\alpha}(x){\tilde F}_{\alpha}{} ^{\nu}(x)+{1\over 4}\eta^{\mu\nu}{\tilde F}^{\alpha\beta}(x){\tilde F} _{\alpha\beta}(x)$] \begin{eqnarray} \Theta^{\mu\nu}(x)&=&{\tilde F}^{\mu\alpha}(x){\tilde F}_{\alpha}{}^{\nu}(x)+ {1\over 4}\eta^{\mu\nu}{\tilde F}^{\alpha\beta}(x){\tilde F}_{\alpha\beta}(x) +\nonumber \\ &+&{(D^{(A)\mu}\tilde \phi (x))}^{*}\, D^{(A)\nu}\tilde \phi (x)+{(D^{(A)\nu} \tilde \phi (x))}^{*}D^{(A)\mu}\tilde \phi (x)-\nonumber \\ &-&\eta^{\mu\nu}[{(D^{(A)\alpha}\tilde \phi (x))}^{*}D^{(A)}{}_{\alpha} \tilde \phi (x)]=\nonumber \\ &=&\Big[ z^{\mu}_A z^{\nu}_B \Big( [g^{AD}g^{CE}g^{BF}+{1\over 4}g^{AB}g^{CD} g^{EF}] F_{DE} F_{CF} +g^{AB} m^2 \phi^{*} \phi +\nonumber \\ &+&[g^{AC}g^{BD}+g^{BC}g^{AD}-g^{AB}g^{CD}] (\partial_C+ieA_C)\phi^{*}\, (\partial_D-ieA_D) \phi \Big) \Big] (\tau ,\vec \sigma )=\nonumber \\ &=&T^{\mu\nu}_{em}(x) + T^{\mu\nu}_{\phi ,A}(x),\nonumber \\ &&{}\nonumber \\ \partial_{\nu}\Theta^{\nu\mu}(x)&{\buildrel \circ \over =}& 0,\nonumber \\ \Rightarrow && \partial_{\nu} T^{\mu\nu}_{\phi ,A}(x)\, {\buildrel \circ \over =}\, -{\tilde F}^{\mu\nu}(x) J_{\phi \nu}(x), \label{VIII41} \end{eqnarray} \noindent where the conserved electromagnetic current of the Klein-Gordon field $J^{\mu}_{\phi}$ is [$\partial_{\mu}J^{\mu}_{\phi}\, {\buildrel \circ \over =}\, 0$ is replaced by ${{dq_{\phi}}\over {dT_s}}\, {\buildrel \circ \over =}\, 0$ at the Hamiltonian level on the Wigner hyperplanes with $q_{\phi} =e$ given in Eq.(\ref{VIII11})] \begin{eqnarray} J^{\mu}_{\phi}(\tau ,\vec \sigma ) &=&-i[(\partial^{\mu}+ie{\tilde A}^{\mu}){\tilde \phi}^{*} \tilde \phi -{\tilde \phi}^{*} (\partial^{\mu}-ie{\tilde A}^{\mu})\tilde \phi ] (z)=\nonumber \\ &=&-i z^{\mu}_A(\tau ,\vec \sigma ) [(\partial^A+ieA^A)\phi^{*} \phi -\phi^{*} (\partial^A-ieA^A)\phi ](\tau ,\vec \sigma )=\nonumber \\ &=&iz^{\mu}_{\tau}(\tau ,\vec \sigma ) [\pi_{\phi^{*}}\phi^{*}-\pi_{\phi} \phi ](\tau ,\vec \sigma )+\nonumber \\ &+&iz^{\mu}_r(\tau ,\vec \sigma ) [\phi^{*} (\partial^r-ieA^r)\phi -(\partial^r +ieA^r)\phi^{*} \phi ](\tau ,\vec \sigma ). \label{VIII42} \end{eqnarray} On Wigner hyperplanes with $T_s=\tau$ in the Coulomb gauge we have \begin{eqnarray} {\hat J}^{\mu}_{\phi}(T_s ,\vec \sigma )&=& u^{\mu}(p_s) i[{\hat \pi}_{\phi ^{*}}{\hat \phi}^{*}-{\hat \pi}_{\phi}\hat \phi ](T_s,\vec \sigma )+\nonumber \\ &+&\epsilon^{\mu}_r(u(p_s)) i[{\hat \phi}^{*}(\partial^r-ieA^r_{\perp})\hat \phi -(\partial^r+ieA^r_{\perp}){\hat \phi}^{*} \hat \phi ](T_s,\vec \sigma )]. \label{VIII43} \end{eqnarray} On the Wigner hyperplanes we get the following expression for the energy-momentum tensor \begin{eqnarray} \Theta^{\mu\nu}&&[x^{\beta}_s(T_s)+\epsilon^{\beta}_u(u(p_s)) \sigma^u]= \nonumber \\ &=&u^{\mu}(p_s) u^{\nu}(p_s) \Big[ {1\over 2}({\vec \pi}^2+{\vec B}^2)+\pi _{\phi^{*}} \pi_{\phi} +(\vec \partial +ie\vec A)\phi^{*}\cdot (\vec \partial -ie\vec A)\phi +m^2\phi^{*}\phi \Big](T_s,\vec \sigma )+\nonumber \\ &+&\epsilon^{\mu}_r(u(p_s))\epsilon^{\nu}_s(u(p_s))\Big[ {1\over 2}\delta^{rs} ({\vec \pi}^2+{\vec B}^2)-(\pi^r\pi^s+B^rB^s)+\delta^{rs}(\pi_{\phi^{*}} \pi_{\phi}-m^2\phi^{*}\phi )+\nonumber \\ &+&(\delta^{ru}\delta^{sv}+\delta^{rv}\delta^{su}-\delta^{rs}\delta^{uv}) (\partial^u+ieA^u)\phi^{*}\, (\partial^v-ieA^v)\phi \Big] (T_s,\vec \sigma )+ \nonumber \\ &+&[u^{\mu}(p_s)\epsilon^{\nu}_r(u(p_s))+u^{\nu}(p_s)\epsilon^{\mu}_r(u(p_s))] \nonumber \\ &&\Big[ ({\vec \pi}\times \vec B)^r+\pi_{\phi^{*}}(\partial^r+ieA^r)\phi^{*}+ \pi_{\phi} (\partial^r-ieA^r)\phi \Big] (T_s,\vec \sigma ), \label{VIII44} \end{eqnarray} \noindent whose expression in the $A_{\tau}(T_s,\vec \sigma )=\eta_{em}(T_s, \vec \sigma )=0$ Coulomb gauge, where, from Eq.(\ref{VIII28}), we have $\pi^r =\pi^r_{\perp}+{e\over {\triangle}} {\hat {\cal Q}}$ with ${\hat {\cal Q}}=i [{\hat \pi}_{\phi^{*}}{\hat \phi}^{*}-{\hat \pi}_{\phi}\hat \phi ]$ [so that Eq.(\ref{VIII11}) becomes $q_{\phi}=\int d^3\sigma {\hat {\cal Q}}(\tau ,\vec \sigma )$], is \begin{eqnarray} {\hat \Theta}^{\mu\nu}&&[x^{\beta}_s(T_s)+\epsilon^{\beta}_u(u(p_s)) \sigma^u]= \nonumber \\ &=&u^{\mu}(p_s) u^{\nu}(p_s) \Big[ {1\over 2}({\vec \pi}^2_{\perp} +{\vec B}^2)+e {\vec \pi}_{\perp}\cdot {{\vec \partial}\over {\triangle}} {\hat {\cal Q}}+{{e^2}\over 2} ({{\vec \partial}\over {\triangle}}{\hat {\cal Q}})^2+ \nonumber \\ &+&{\hat \pi}_{\phi^{*}} {\hat \pi}_{\phi} +(\vec \partial +ie{\vec A}_{\perp}){\hat \phi}^{*}\cdot (\vec \partial -ie{\vec A}_{\perp})\hat \phi +m^2{\hat \phi}^{*}\hat \phi \Big](T_s,\vec \sigma )+\nonumber \\ &+&\epsilon^{\mu}_r(u(p_s))\epsilon^{\nu}_s(u(p_s))\Big[ {1\over 2}\delta^{rs} ({\vec \pi}^2_{\perp}+{\vec B}^2)-(\pi^r_{\perp}\pi^s_{\perp}+B^rB^s)+ \nonumber \\ &+&e\Big( \delta^{rs}{\vec \pi}_{\perp}\cdot {{\vec \partial}\over {\triangle}} {\hat {\cal Q}}-[\pi^r_{\perp}{{\partial^s}\over {\triangle}}+\pi^s_{\perp} {{\partial^r}\over {\triangle}}]{\hat {\cal Q}}\Big) +\nonumber \\ &+&e^2 \Big( {1\over 2}\delta^{rs} ({{\vec \partial}\over {\triangle}}{\hat {\cal Q}})^2-{{\partial^r}\over {\triangle}}{\hat {\cal Q}}{{\partial^s}\over {\triangle}}{\hat {\cal Q}}\Big)+\nonumber \\ &+&\delta^{rs}({\hat \pi}_{\phi^{*}}{\hat \pi}_{\phi}-m^2{\hat \phi}^{*}\hat \phi )+\nonumber \\ &+&(\delta^{ru}\delta^{sv}+\delta^{rv}\delta^{su}-\delta^{rs}\delta^{uv}) (\partial^u+ieA_{\perp}^u){\hat \phi}^{*}\, (\partial^v-ieA^v_{\perp})\hat \phi \Big] (T_s,\vec \sigma )+\nonumber \\ &+&[u^{\mu}(p_s)\epsilon^{\nu}_r(u(p_s))+u^{\nu}(p_s)\epsilon^{\mu}_r(u(p_s))] \nonumber \\ &&\Big[ ({\vec \pi}_{\perp}\times \vec B)^r+{\hat \pi}_{\phi^{*}}(\partial^r +ieA_{\perp}^r){\hat \phi}^{*}+ {\hat \pi}_{\phi} (\partial^r-ieA_{\perp}^r)\hat \phi \Big] (T_s,\vec \sigma ), \nonumber \\ &&{}\nonumber \\ {\hat \Theta}^{\mu}{}_{\mu}&&[x^{\beta}_s(T_s)+\epsilon^{\beta}_u(u(p_s)) \sigma^u]=\nonumber \\ &=&-2\Big[ {\hat \pi}_{\phi^{*}} {\hat \pi}_{\phi} -(\vec \partial +ie{\vec A}_{\perp}){\hat \phi}^{*}\cdot (\vec \partial -ie{\vec A}_{\perp})\hat \phi -2m^2{\hat \phi}^{*}\hat \phi \Big](T_s,\vec \sigma ),\nonumber \\ {\hat P}^{\mu}_{\Theta}&=&\int d^3\sigma {\hat \Theta}^{\mu\nu} u_{\nu}(p_s)= \nonumber \\ &=& M u^{\mu}(p_s) +{\cal H}_p^r(T_s) \epsilon^{\mu}_r(u(p_s))\approx M u^{\mu}(p_s),\nonumber \\ {\hat \Theta}^{rs}_S&=&\epsilon^r_{\mu}(u(p_s))\epsilon^s_{\nu}(u(p_s)){\hat \Theta}^{\mu\nu}=\nonumber \\ &=&\Big[ {1\over 2}\delta^{rs} ({\vec \pi}^2_{\perp}+{\vec B}^2)-(\pi^r_{\perp}\pi^s_{\perp}+B^rB^s)+ \nonumber \\ &+&e\Big( \delta^{rs}{\vec \pi}_{\perp}\cdot {{\vec \partial}\over {\triangle}} {\hat {\cal Q}}-[\pi^r_{\perp}{{\partial^s}\over {\triangle}}+\pi^s_{\perp} {{\partial^r}\over {\triangle}}]{\hat {\cal Q}}\Big) +\nonumber \\ &+&e^2 \Big( {1\over 2}\delta^{rs} ({{\vec \partial}\over {\triangle}}{\hat {\cal Q}})^2-{{\partial^r}\over {\triangle}}{\hat {\cal Q}}{{\partial^s}\over {\triangle}}{\hat {\cal Q}}\Big)+\nonumber \\ &+&\delta^{rs}({\hat \pi}_{\phi^{*}}{\hat \pi}_{\phi}-m^2{\hat \phi}^{*}\hat \phi )+\nonumber \\ &+&(\delta^{ru}\delta^{sv}+\delta^{rv}\delta^{su}-\delta^{rs}\delta^{uv}) (\partial^u+ieA_{\perp}^u){\hat \phi}^{*}\, (\partial^v-ieA^v_{\perp})\hat \phi \Big] (T_s,\vec \sigma ). \label{VIII45} \end{eqnarray} The Dixon multipoles with respect to the origin $x^{\mu}_s(\tau )$ are \begin{eqnarray} t^{\mu_1...\mu_n\mu\nu}_{\Theta}(T_s)&=&\int d^3\sigma \delta x^{\mu_1}_s(\vec \sigma )...\delta x^{\mu_n}_s(\vec \sigma ) {\hat \Theta}^{\mu\nu}[x_s ^{\beta}(T_s)+\epsilon^{\beta}_u(u(p_s))\sigma^u] =\nonumber \\ &=&\epsilon^{\mu_1}_{r_1}(u(p_s))...\epsilon^{\mu_n}_{r_n}(u(p_s))\epsilon^{\mu}_A(u(p_s)) \epsilon^{\nu}_B(u(p_s)) I_{\Theta}^{r_1..r_nAB}(T_s), \label{VIII46} \end{eqnarray} \noindent and all the derived multipoles for the interacting case can be obtained from these ones following the scheme of Section V. With the same techniques we can study the multipoles $t_{\phi ,A}^{\mu_1... \mu_n\mu\nu}(T_s)$ and all the related ones ${\hat P}^{\mu}_{\phi ,A}$, ${\hat S}^{\mu\nu}_{\phi ,A}$, ${\hat J}_{\phi ,A}^{\mu_1...\mu_n\mu\nu}$ of ${\hat T}^{\mu\nu}_{\phi ,A}$. The Dixon multipoles of the electromagnetic field in the rest-frame instant form will be studied in a future paper. Let us remark that in absence of the electromagnetic field, namely in the free theory of a complex Klein-Gordon field, we have \begin{eqnarray} \Theta^{\mu\nu}[x_s^{\beta}(T_s)+\epsilon^{\beta}_u(u(p_s))\sigma^u]&=& T^{\mu\nu}_{\phi}[x_s^{\beta}(T_s)+\epsilon^{\beta}_u(u(p_s))\sigma^u]= \nonumber \\ &=&u^{\mu}(p_s)u^{\nu}(p_s) [\pi_{\phi^{*}}\pi_{\phi}+\vec \partial \phi^{*} \cdot \vec \partial \phi +m^2\phi^{*}\phi ](T_s,\vec \sigma )+\nonumber \\ &+&[u^{\mu}(p_s)\epsilon^{\nu}_r(u(p_s))+u^{\nu}(p_s)\epsilon^{\mu}_r(u(p_s))] [\pi_{\phi^{*}}\partial^r\phi^{*}+\pi_{\phi}\partial^r\phi ](T_s,\vec \sigma )+ \nonumber \\ &+&\epsilon^{\mu}(u(p_s))\epsilon^{\nu}_s(u(p_s))[\delta^{rs}(\pi_{\phi^{*}}\pi _{\phi}-m^2\phi^{*}\phi )+\nonumber \\ &+&(\delta^{ru}\delta^{sv}+\delta^{rv}\delta^{su}- \delta^{rs}\delta^{uv})\partial^u\phi^{*}\partial^v\phi ](T_s,\vec \sigma ),\nonumber \\ &&{}\nonumber \\ t^{\mu_1...\mu_n\mu\nu}_{T*}(T_s)&=&\int d^3\sigma \delta x^{\mu_1}_s(\vec \sigma )...\delta x^{\mu_n}_s(\vec \sigma ) T_{\phi}^{\mu\nu}[x_s ^{\beta}(T_s)+\epsilon^{\beta}_u(u(p_s))\sigma^u] =\nonumber \\ &=&\epsilon^{\mu_1}_{r_1}(u(p_s))...\epsilon^{\mu_n}_{r_n}(u(p_s))\epsilon^{\mu}_A(u(p_s)) \epsilon^{\nu}_B(u(p_s)) I_{T*}^{r_1..r_nAB}(T_s), \label{VIII47} \end{eqnarray} \noindent and, again with the methods of Section V, we can define the Dixon multipoles [$t_{\phi}^{\mu_1...\mu_n\mu\nu}(T_s)$ and the related ones] of the complex Klein-Gordon field in absence of interaction. In particular, from Eq.(\ref{V8}) we get ${\tilde t}^{\mu_1}_{\phi}(T_s)=\epsilon^{\mu_1}_{r_1}(u(p_s)) I_{T*}^{r_1\tau\tau}(T_s)$ with $I^{r\tau\tau}_{T*}(T_s)\equiv -P^{\tau}_{\phi} r^r_{\phi} = \int d^3\sigma \, \sigma^r [\pi_{\phi^{*}}\pi_{\phi} +\vec \partial \phi^{*}\cdot \vec \partial \phi +m^2\phi^{*}\phi ](T_s,\vec \sigma )$ and its vanishing identifies Dixon's center of mass (M\"oller noncanonical center of energy) ${\vec r}_{\phi}$; then the canonical 3-center of mass ${\vec q}_{\phi}$ can be obtained with the methods of Setion VI. Then, as in the case of a real Klein-Gordon field, one can look for the canonical transformation from the center-of-phase canonical basis $X^A_{\phi}$, $P^A_{\phi}$, $R^A_{\phi}$, $Q^A_{\phi}$, ${\bf H}_a(\tau ,\vec k)$, ${\bf K}_a(\tau ,\vec k)$, to a center-of-mass canonical basis $q^{\tau}_{\phi}$, $P^{\tau {'}}_{\phi}=\sqrt{(P^{\tau}_{\phi})^2-{\vec P}^2_{\phi}}$, ${\vec q}_{\phi}$, ${\vec P}_{\phi}$, $R^{A {'}}_{\phi}$, $Q^{A {'}}_{\phi}$, ${\bf H}_a^{'}(\tau ,\vec k)$, ${\bf K}_a^{'}(\tau ,\vec k)$. Now ${\vec R}^{'}_{\phi}$ shouls describe the relative position of the ``centers of mass" for the $``a"$ and $``b"$ modes of the field configuration. In the free case Eqs.(\ref{d6}) show that the theory can be reformulated in a non-local way as the sum of two theories, one for each mode: see Ref.\cite{lam} for the Lagrangian associated to Eqs.(\ref{d6}) in the framework of pseudo-differential operators. This implies that the energy momentum tensor can be written as the sum of two pieces, one for each mode, and that the multipole expansion and the associated definition of center of mass can be applied to each piece. Already in the present framework, by using Eqs.(\ref{d5}) and by making integrations by parts, we get $I^{\tau\tau}_{T*}(T_s)=\int d^3\sigma [\pi_{\phi^{*}}\pi_{\phi}+\vec \partial \phi^{*}\cdot \vec \partial \phi +m^2\phi^{*}\phi ](T_s,\vec \sigma )= m\int d^3\sigma [\phi^{*}_a \sqrt{m^2+\triangle} \phi_a+\phi^{*}_b \sqrt{m^2+\triangle} \phi_b](T_s,\vec \sigma )=I^{\tau\tau}_{a*}(T_s) +I^{\tau\tau}_{b*}(T_s)$ and $I^{r\tau\tau}_{T*}(T_s)=\int d^3\sigma [\pi_{\phi^{*}}\partial^r\phi +\pi_{\phi}\partial^r\phi ](T_s,\vec \sigma )=im \int d^3\sigma [\phi^{*}_a\partial^r\phi_a+ \phi^{*}_b\partial^r\phi_b](T_s,\vec \sigma )=I^{r\tau\tau}_{a*}(T_s) +I^{r\tau\tau}_{b*}(T_s)\equiv -P^{\tau}_{a\phi} r^r_{a\phi}- P^{\tau}_{b\phi} r^r_{b\phi}$. Then we could evaluate ${\vec q}_{a\phi}$ and ${\vec q}_{b\phi}$ and we expect to have ${\vec q}_{a\phi}={\vec X}^{'}_{a\phi}={\vec q}_{\phi}+{1\over 2}{\vec R}^{'}_{\phi}$ and ${\vec q}_{b\phi}={\vec X}^{'}_{b\phi}={\vec q}_{\phi}-{1\over 2}{\vec R}^{'}_{\phi}$. The Dixon multipoles\cite{dixon} for the electromagnetic current ${\hat J}^{\mu}_{\phi}$ are \begin{eqnarray} && n \geq 0,\nonumber \\ {\hat j}_{\phi}^{\mu_1...\mu_n\mu}(T_s)&=&{\hat j}_{\phi}^{(\mu_1...\mu_n)\mu} (T_s)=\int d^3\sigma \delta x^{\mu_1}_s(\vec \sigma )...\delta x^{\mu_n}_s(\vec \sigma ) {\hat J}_{\phi}^{\mu}(T_s,\vec \sigma )=\nonumber \\ &=&\epsilon^{\mu_1}_{r_1}(u(p_s))...\epsilon^{\mu_n}_{r_n}(u(p_s))\nonumber \\ &&\Big[ u^{\mu}(p_s) \int d^3\sigma \sigma^{r_1}...\sigma^{r_n} i[{\hat \pi} _{\phi^{*}}{\hat \phi}^{*}-{\hat \pi}_{\phi}\hat \phi ](T_s,\vec \sigma )+ \nonumber \\ &+&\epsilon^{\mu}_r(u(p_s)) \int d^3\sigma \sigma^{r_1}...\sigma^{r_n} i[{\hat \phi}^{*}(\partial^r-ieA^r_{\perp})\hat \phi -(\partial^r+ieA^r_{\perp}){\hat \phi}^{*} \hat \phi ](T_s,\vec \sigma )\Big] ,\nonumber \\ &&u_{\mu_1}(p_s){\hat j}_{\phi}^{\mu_1...\mu_n\mu}(T_s)=0,\nonumber \\ &&{}\nonumber \\ && n=0,\nonumber \\ {\hat j}^{\mu}_{\phi}(T_s)&=& q_{\phi} u^{\mu}(p_s) +\epsilon^{\mu}_r(u(p_s)) \nonumber \\ &&\int d^3\sigma i[{\hat \phi}^{*}(\partial^r-ieA^r_{\perp})\hat \phi -(\partial^r+ieA^r_{\perp}){\hat \phi}^{*} \hat \phi ](T_s,\vec \sigma )=\nonumber \\ &=&q_{\phi}u^{\mu}(p_s)+\epsilon^{\mu}_r(u(p_s)) im \int d^3\sigma \Big( ({\hat \phi}^{*}_a+{\hat \phi}_b){{\partial^r}\over {\triangle}}({\hat \phi}_a+{\hat \phi}^{*}_b)-\nonumber \\ &-&ieA^r_{\perp} [(m^2+\triangle )^{-1/4}({\hat \phi}^{*}_a+{\hat \phi}_b)] [(m^2+\triangle )^{-1/4}({\hat \phi}_a+{\hat \phi}^{*}_b)]\Big) (T_s,\vec \sigma ), \nonumber \\ {\hat q}_{\phi}^{\mu_1...\mu_n}(T_s)&=&{\hat q}_{\phi}^{(\mu_1...\mu_n)}(T_s)= {\hat j}_{\phi}^{\mu_1...\mu_n\mu}(T_s) u_{\mu}(p_s)= \quad\quad (n > 0)\nonumber \\ &=&\epsilon^{\mu_1}_{r_1}(u(p_s))...\epsilon^{\mu_n}_{r_n}(u(p_s)) \int d^3\sigma \sigma^{r_1}...\sigma^{r_n} i[{\hat \pi} _{\phi^{*}}{\hat \phi}^{*}-{\hat \pi}_{\phi}\hat \phi ](T_s,\vec \sigma ), \nonumber \\ &&u_{\mu_1}(p_s) {\hat q}_{\phi}^{\mu_1...\mu_n}(T_s)=0. \label{VIII48} \end{eqnarray} In Ref.\cite{dixon} it is shown how to arrive to the following reconstruction of the electromagnetic current in terms of the multipoles (if $f(k)$ is analytic) \begin{eqnarray} <{\hat J}^{\mu}_{\phi},f >&=& \int d^4x {\hat J}^{\mu}_{\phi}(x) f(x)=\int dT_s \int {{d^4k}\over {(2\pi )^4}}\tilde f(k) e^{-ik\cdot x_s(T_s)} \nonumber \\ &&\int d^3\sigma {\hat J}^{\mu}_{\phi}[x_s(T_s)+\delta x_s(\vec \sigma )] \sum_{n=0}^{\infty} {{(-i)^n}\over {n!}} k_{\mu_1}...k_{\mu_n}{\hat j}_{\phi} ^{\mu_1...\mu_n\mu}(T_s)=\nonumber \\ &=&\int dT_s \sum_{n=0}^{\infty} {1\over {n!}} {\hat j}^{\mu_1...\mu_n\mu} _{\phi}(T_s){{\partial^nf(x)}\over {\partial x_s^{\mu_1}...\partial x^{\mu_n}_s}}{|}_{x=x_s(T_s)},\nonumber \\ &&{}\nonumber \\ {\hat J}^{\mu}_{\phi}(x)&=& \sum_{n=0}^{\infty} {{(-)^n}\over {n!}} {{\partial^n}\over {\partial x_s^{\mu_1}...\partial x^{\mu_n}_s}} \int dT_s \delta^4(x-x_s(T_s)) {\hat j}_{\phi}^{\mu_1...\mu_n\mu}(T_s). \label{VIII49} \end{eqnarray} It is also shown that if the multipoles ${\hat j}_{\phi}^{\mu_1...\mu_n\mu}(T _s)$ are known for all $n > N$ for some fixed N, then ${\hat J}_{\phi}^{\mu}$ and the multipoles with $n\leq N$ are completely determined. See Appendix C for other types of multipoles. \vfill\eject \section{Conclusions.} In this paper we have made a detailed study of the kinematical description of scalar Klein-Gordon fields on the Wigner hyperplanes of the rest-frame Wigner-covariant instant form of dynamics. We have considered a Klein-Gordon field configuration as a relativistic extended object and we utilized both phase space techniques from relativistic particle mechanics and multipolar expansions from relativistic fluidodynamics to study aspects of the Klein-Gordon field which are usually ignored notwithstanding the relevance of scalar fields in physics: Higgs particles, Bose-Einstein condensation, Brans-Dicke scalar-tensor general relativity and multipolar expansions for the theory of gravitational waves, boson stars. Simultaneously, we have used the Klein-Gordon field as an example to explore the description of isolated systems on the Wigner hyperplanes showing how the elusive concept of ``relativistic center of mass" has to be divided in an ``external" part (with respect to an arbitrary Lorentz frame) and in an ``internal" part (inside the Wigner hyperplane). In both cases a canonical noncovariant Newton-Wigner-like center of mass, a covariant noncanonical Fokker-Pryce center of inertia and a noncanonical noncovariant M\"oller center of energy may be defined only in terms of the generators of suitable realizations of the Poinca\'e algebra. The three ``internal" centers weakly coincide due to the three first class constraints defining the rest frame of the isolated system and are, therefore, gauge variables inside the Wigner hyperplanes. Namely there is the gauge freedom in the choice of the ``external" timelike worldline to which they have to be attached. Now, in the description of the Wigner hyperplane with respect to an arbitrary Lorentz frame in Minkowski spacetime this gauge freedom is reflected in the arbitrariness of the choice of a timelike worldline $x^{\mu}_s(\tau )$, with noncanonical 4-coordinates, to be used as origin of the ``internal" 3-coordinates. The natural gauge fixing for this gauge freedom is to put the three weakly coinciding ``internal" centers in this origin: in this way the origin $x^{\mu}_s(\tau )$ is forced to coincide with the ``external" covariant noncanonical Fokker-Pryce center of inertia and simultaneously to satisfy the conditions for being both the Pirani and the Tulczyjew centroid. Around this worldline there is a noncovariance worldtube (whose finite extension is measured by the M\"oller radius\cite{moller,dubna}) containing all the pseudoworldlines of the noncovariant ``external" canonical Newton-Wigner-like center of mass and noncanonical M\"oller center of energy. Naturally, one could fix the gauge freedom in a different way by identifying a different ``internal" collective 3-vector with the origin $x^{\mu}_s$: in such a case $x^{\mu}_s$ becomes one of the many possible covariant noncanonical centroids existing in literature and does not coincide with the ``external" Fokker-Pryce center of inertia. For instance this happens for the Klein-Gordon field, because we do not yet know its canonical basis containing the ``internal" center of mass 3-vector. In the Longhi-Materassi canonical decomposition of the Klein-Gordon field in collective and relative variables this 3-vector is replaced by another 3-vector, which can be named the ``internal" center of phase of the field configuration by its construction. The other main result of this paper is the identification of the canonical basis containing the ``internal" center of phase of the Klein-Gordon field and of its multipoles in the framework of the rest-frame Wigner-covariant instant form of dynamics. Therefore, many unrelated kinematical concepts find a well defined setting on the Wigner hyperplanes of the rest-frame Wigner-covariant instant form of dynamics, which seems to be the natural tool (like the separation of the center-of-mass motion in the nonrelativistic case) to be used in relativistic statistical mechanics and in lattice gauge theories due to its intrinsic Euclidean signature (without any Wick rotation) and to its associated description of the evolution by means of the Lorentz scalar rest-frame time measured by the clock of the decoupled (but noncovariant) point particle observer defined by the ``external" canonical center of mass. Even if much work is still needed to clarify all the aspects of the multipolar expansions of the Klein-Gordon field and of its properties as a relativistic extended object, especially for the charged fields of scalar electrodynamics, now we have a well defined framework in which to make further investigations and a first completely worked out canonical decomposition of the field in collective and relative variables. A final comment on quantization. The canonical basis of the Klein-Gordon field containing the center of phase does not seem a good candidate for the quantization of the field in the rest-frame instant form of dynamics both due to the complicated expression of the original fields in terms of the new variables and to the fact that there is no sound canonical quantization of phases and angles (different is the case with their exponentials) as can be seen from the various articles contained in Ref.\cite{phase}. Let us remark that the problem of the non measurability of absolute phases (see the previous reference) is connected with the open problem of the measurability of the ``external" decoupled noncovariant canonical center of mass of an isolated system (the quantization of these degrees of freedom would generate an equivalent of the ``wave function of the universe") as it is clear from the Longhi-Materassi canonical basis, in which the collective position variable is built starting from the phases of the Fourier coefficients of the Klein-Gordon field. \vfill\eject
\section{Introduction} Hickson Compact Groups (hereafter HCGs; Hickson 1982) are small systems of several galaxies (four or more) in an apparent close proximity in the sky. Their dynamical state has been the subject of a controversial debate. Comparison of the observational data with model calculations led Mamon (1986, 1987) to conclude that less than a half of HCGs could be considered bounded dense systems. However accumulated statistical evidences (Hickson \& Rood 1988; Hickson 1992) together with observational evidences, also coming from X-ray observations (Saracco \& Ciliegi 1995; Ponman et al. 1996; Pildis et al. 1995), clearly favor the view that the majority of HCGs are physical systems and not chance projections or transient systems. One of the expected consequences of the gravitational interactions between galaxies is the enhancement of the star formation rate (SFR) in the interacting systems (Joseph \& Wright 1985; Bushouse 1987; Laurikainen \& Moles 1989). The photometric and spectroscopic studies carried out so far (Rubin et al. 1991 \& Mendes de Oliveira et al. 1997; Moles et al. 1994 \& Mendes de Oliveira et al. 1994; Vilchez \& Iglesias Paramo 1998; Plana et al. 1998, Iglesias Paramo \& Vilchez 1999), aiming at establishing the fraction of interacting galaxies in HCGs, have often given contradictory results. Actually a possibility exists that only a fraction of the HCGs are bound systems and that bound HCGs evolve in different ways. Different evolutions could be due to both the different dynamical properties of HCGs and to their different 'birthplace', i.e. the environment in which they are embedded. Powerful signs of star formation activity are the ionization lines emitted by the heated gas surrounding the regions of star formation. Unlike the color indexes in the $U,B,V$ filters, that give indications about the past star formation ($> 10^8$ years), the H$_\alpha$ emission line at 6563 {\AA} can be used as a quantitative and spatial tracer of the rate of massive ($\geq$ 10 M$_\odot$) and therefore recent ($\leq 10^7$ years) star formation (Kennicutt 1983; Ryder \& Dopita 1994). Therefore, knowing the H$_\alpha$ emission of the HCG galaxies, it is possible in principle to carry out important knowledges about the present merger and interaction events in these systems. Up to now the only H$_\alpha$ images regarding HCG galaxies have been collected by Rubin et al. (1991) and more recently by Vilchez \& Iglesias Paramo (1998). They carried out H$_\alpha$ emission-line images respectively for 14 and 16 HCGs. While Vilchez \& Iglesias Paramo estimate the H$_\alpha$ flux for each of the 63 galaxies of their sample (Iglesias Paramo \& Vilchez 1999), Rubin et al. do not use flux calibrated and they take into account a sample constituted by disk galaxies only. We have recently obtained $H_{\alpha}$ fluxes and luminosities for a sample of 95 galaxies from calibrated observations of 31 HCGs (Severgnini et al. 1999). Here we present the preliminary results of the analysis performed on a subsample of 66 galaxies. \section{Observations} Observations have been carried out at the 2.1 meter telescope (design Ritchey-Chretien) at the National Observatory of Mexico in S. Pedro Martir during three different observing runs (November 1995, April 1996 and February 1997). The telescope was equipped with a Tektronix CCD of 1024x1024 pixels, each 24$\mu$m x 24$\mu$m. The telescope scale (13 arcsec/mm) and the pixel dimensions provide a pixel size of 0.3 arcsec/pix with a resulting field of view of 5.12$^\prime\times$5.12$^\prime$. During these three runs we observed 31 HCGs in H$_\alpha$ filters. The remaining 61 HCGs were not in ours sample because the adequate H$_\alpha$ interferometric filters were not available. This is the only criterion used to select the observed groups.\\ In order to calibrate our data, we have observed some spectrophotometric stars, equally spaced in time during each night, taken from the list of Massey $\&$ Strobel (1988). The standards were observed in all the H$_\alpha$ narrow-band filters used to observe HCGs. During the observations the seeing was in the range of 2 to 2.6 arcsec and the photometry was within 0.05 mag in all but one (worse) night. The mean limiting flux of the observations, at one sigma from the background and within the mean seeing disk (2.3 arcsec), is 9.22$\cdot10^{-17}$ erg cm$^{-2}$ s$^{-1}$. We estimated the $H_{\alpha}$ flux and luminosity for 66 galaxies, 12 out of which are upper limits. We adopt H$_0$=100 km/(s Mpc) and q$_0$=0.5.\\ In Figure 1 the distribution of $H_{\alpha}$ luminosity ($L_{H_\alpha}$) of the 54 detected galaxies is shown. $L_{H_\alpha}$ for each galaxy has been derived from using a $H_\alpha$ isophotal flux computed within the region defined by a detection threshold of one sigma above the background. A detailed description of the data reduction, the photometric calibration, the flux estimate and the luminosity derivation is given in Severgnini et al. (1999). \begin{figure} \centerline{\epsfig{file=luminosita.ps, width=16pc}} \caption{Distribution of $H_{\alpha}$ isophotal luminosity, corrected for Galactic Extinction (see Severgnini et al. 1999), of the 54 detected galaxies.} \end{figure} \begin{figure} \centerline{\epsfig{file=dv_ha.ps, width=16pc}} \caption{$H_{\alpha}$ luminosity of the 66 galaxies of our sample vs velocity dispersion of the groups. There is a clear correlation (the probability of the two not being correlated is P=0.001) which suggests an increasing of $H_{\alpha}$ luminosity of galaxies with decreasing of velocity dispersion of groups.} \end{figure} \begin{figure} \centerline{\epsfig{file=dv_hai_l2.ps,width=16pc}} \caption{$H_{\alpha}$ luminosity of galaxies vs velocity dispersion of the isolated HCGs ({\it upper panel}) and of the HCGs embedded in loose groups ({\it lower panel}). The velocity dispersion of HCG$_L$ is correlated to H$_\alpha$ luminosity of their galaxies (the probability of the two not being correlated is P=0.03), while the same correlation is absent for HCG$_I$ (P=0.60). In both the Figures 2 and 3, galaxies are binned in dynamical properties. The dots represent the mean luminosity and the error bars are the standard deviations of the values. The down arrows refer to upper limits inside of the bin. In Figure 2 the width of the bins is also shown.} \end{figure} \section{Results} We have investigated about the dependency of the H$_\alpha$ luminosity of galaxies on the dynamical properties of HCGs and on the environment in which groups are embedded. The statistical analysis we have performed has made use of the survival analysis (Isobe, Feigelson \& Nelson 1986) since some of the observed galaxies have not been detected in our observations. In an undisrupted galaxy it is expected that actual SFR ($<$10$^{6-7}$ years) is correlated to the quantity of young stellar population (O,B stars). To test if this correlation is present also for HCG galaxies we searched for a correlation between the $H_{\alpha}$ luminosity of galaxies and their $B$ absolute magnitude (M$_B$ ({\em "Atlas of Compact Groups of Galaxies"}, Hickson 1993)), good tracer of young stars. We found a significant correlation between these two quantities in the sense that: \begin{description} \item $\bullet$ galaxy having higher (lower) $H_{\alpha}$ luminosity have higher (lower) $B$ luminosity. \end{description} We have then searched for correlations between the dynamical parameters of HCGs (velocity dispersion, mass density, surface brightness and crossing time, as from Hickson et al. 1992) and the $H_{\alpha}$ luminosity of the member galaxies. Our analysis shows the presence of correlations between each of these parameters and the $H_{\alpha}$ luminosity of the galaxies. In particular we found that the $H_{\alpha}$ emission is correlated to crossing time and it is anti-correlated to velocity dispersion, mass density and surface brightness in the sense that: \begin{description} \item $\bullet$ HCGs having higher (lower) densities have higher (lower) velocity dispersions and the galaxies inside them have lower (higher) $H_{\alpha}$ luminosities. \end{description} In Figure 2 the relation between $H_{\alpha}$ luminosity of galaxies and the velocity dispersion of the groups is shown (the probability of the two not being correlated is 0.001). Such correlations would seem to point toward a dependence of the $H_{\alpha}$ luminosity of galaxies, and hence of their SFR, on the dynamical properties of HCGs. To test if different environments, where HCGs are actually found, affect the H$_\alpha$ of galaxies inside them we have divided our sample in two subsamples, following the Rood and Struble's (1994) classification: the first composed by the isolate compact groups (HCG$_I$) and the second one consisting of those compact groups embedded in loose groups (HCG$_L$). Through a statistical comparison we found that there aren't significant differences between $H_{\alpha}$ luminosity distributions of galaxies inside HCG$_I$ and HCG$_L$. The same result is found if we compare the dynamical parameters of isolate HCGs to those of HCGs in loose groups. Nevertheless, from our data we found that the velocity dispersion (Figure 3) and crossing time of HCG$_L$ are correlated to the $H_{\alpha}$ luminosity of their galaxies and that these correlations are absent for HCG$_I$. We thus can assert that inside HCG$_L$ there are correlations between two dynamical parameters and $H_{\alpha}$ emission that are not present inside HCG$_I$. These results suggest that, although the surrounding environment of HCGs does not directly influence the SFR of galaxies and the dynamical properties of groups, it modifies the relation between dynamical properties of groups and SFR of their galaxies.\\ Moreover we have tested that the correlation found between $H_{\alpha}$ and $B$ luminosity is present only for galaxies inside HCG$_L$ (Figure 4). This result suggests that the different HCG environments affect also the ratio between the population of young stars and actual star formation of galaxies.\\ Thus we assert that the surrounding environment of HCGs affects the evolution of the HCG galaxies. \begin{figure} \centerline{\epsfig{file=Mb_hail2.ps, width=16pc}} \caption{ {\it Upper panel:} The H$_\alpha$ luminosity of galaxies in HCG$_I$ vs their own absolute magnitude in the $B$ filter ($M_B$). {\it Lower panel:} The H$_\alpha$ luminosity of galaxies in HCG$_L$ vs their own absolute magnitude in the $B$ filter ($M_B$). $M_B$ of galaxies into HCG$_L$ is correlated to their own H$_\alpha$ luminosity (the probability of the two not being correlated is P=0.007) while such correlation is absent for HCG$_I$ (P=0.17). In both the figures, galaxies are binned in $M_B$ values.} \end{figure} Finally, we also investigated if the morphology of galaxies is influenced by different environments and if a particular HCG surrounding environment can favor the formation and evolution of a particular morphology. We divide all the galaxies of Hickson's sample in Ellipticals $+$ Lenticulars (E/S0) and Spirals $+$ Irregulars (S/I) finding the same fraction of E/SO and S/I inside HCG$_I$ and HCG$_L$ respectively. Thus we conclude that the environment in which HCGs are embedded does not influence the morphology of their member galaxies. \section{Discussion and Conclusions} We have studied H$_\alpha$ calibrated fluxes for a sample of 66 galaxies in 31 HCGs. 12 galaxies of our sample have not been detected in our images and this implies that they have a H$_\alpha$ emission lower than $f_{lim}=$9.22$\cdot 10^{-17}$ erg s$^{-1}$ cm$^{-2}$, being $f_{lim}$ the 1$\sigma$ limiting flux integrated within one seeing disk reached in our observations. From the analysis presented above, we could conclude that the $H_{\alpha}$ luminosity of galaxies is affected by the dynamical properties of compact groups. The correlations found would point towards a merger scenario in which $H_{\alpha}$ brightest galaxies are inside groups with higher probability of interaction i.e. lower values of velocity dispersion and mass density. It is worth noting however that $H_{\alpha}$ luminosity could be matched by mass of galaxies, that is the SFR could be higher (lower) because of the higher (lower) quantity of the available gas and not because of a real higher {\em efficiency} or {\em fraction}. In principle such degeneracy could be avoided normalizing the $H_{\alpha}$ luminosity by a mass tracer of galaxies which is very well represented by the luminosity in the near IR band (H or K band), as shown by Gavazzi et al. (1996). Nevertheless some authors (e.g. Iglesias Paramo et al. 1999) make use of the absolute B magnitude under the implicit assumption that $M\propto L_B$. On the other hand, since a correlation between $H_{\alpha}$ and B luminosities is expected under robust assumptions (and we found it) the use of $L_B$ to remove the degeneracy could lead to misleading results. Thus, even if we find correlations between $H_{\alpha}$ luminosity and dynamical parameters of the groups, further investigations are required in order to establish if the dynamics of HCGs affects the evolution of their member galaxies. Moreover the relations we find seem to depend on the environment in which HCGs are embedded, as confirmed by the correlations found only in the case of HCG$_{L}$ sample. Since the $H_{\alpha}$ luminosity is a good star formation tracer, this result suggests that also the environments surrounding HCGs could influence the evolution of galaxies into HCGs. The SFR for objects can be directly inferred by H$_\alpha$ luminosity using the result of Kennicut (1983), which relates the SFR to $H_{\alpha}$ luminosity through the relation \begin{equation} SFR(total)={{L(H_\alpha)}\over{1.12\cdot10^{41} {\rm erg~~s^{-1}}}} M_\odot yr^{-1} \end{equation} where a Salpeter initial mass function with an upper mass cutoff of 100 $M_\odot$ have been assumed. From the luminosities we infer a star formation rate for our detected galaxies in the range 0.01-2.88 $M_\odot$ yr$^{-1}$ in absence of internal extinction.
\section{#1}} \renewcommand{\theequation}{\thesection.\arabic{equation}} \setcounter{page}{1} \begin{document} \begin{titlepage} \begin{center} April, 1999 \hfill PM/99-19\\ \vskip .3 in {\large \bf Fractional Supersymmetry~ and}\\ {\large \bf $F^{\mathrm{th}}-$roots of Representations}\\ \vskip .3truecm { {\bf M. Rausch de Traubenberg}\footnote{<EMAIL>, <EMAIL>} \vskip 0.2 cm {\it Laboratoire de Physique Th\'eorique, Universit\'e Louis Pasteur}\\ {\it 3-5 rue de l'universit\'e, 67084 Strasbourg Cedex, France}\\ \vskip 0.2 cm and \\ {\it Laboratoire de Physique Math\'ematique et Th\'eorique, Universit\'e de Montpellier 2}\\ {\it place Eug\`ene Bataillon, case 70, 34095 Montpellier Cedex 5, France}\\ \vskip .8 cm {\bf M. J. Slupinski\footnote{<EMAIL>}}\\ {\it Institut de Recherches en Math\'ematique Avanc\'ee}\\ { \it Universit\'e Louis-Pasteur, and CNRS}\\ {\it 7 rue R. Descartes, 67084 Strasbourg Cedex, France}\\ } \end{center} \vskip .5 in \begin{abstract} A generalization of super-Lie algebras is presented. It is then shown that all known examples of fractional supersymmetry can be understood in this formulation. However, the incorporation of three dimensional fractional supersymmetry in this framework needs some care. The proposed solutions lead naturally to a formulation of a fractional supersymmetry starting from any representation ${\cal D}$ of any Lie algebra $g$. This involves taking the $F^{{\mathrm th}}-$roots of ${\cal D}$ in an appropriate sense. A fractional supersymmetry in any space-time dimension is then possible. This formalism finally leads to an infinite dimensional extension of $g$, reducing to the centerless Virasoro algebra when $g=sl(2,\hbox{\it I\hskip -2.pt R })$. \vskip1cm \noindent PACS: 02.20.Qs; 02.20.Tw;03.65.Fd;11.30.Ly \end{abstract} \end{titlepage} \renewcommand{\thepage}{\arabic{page}} \mysection{Introduction} Describing the laws of physics in terms of underlying symmetries has always been a powerful tool. In this respect, it is interesting to study the kind of symmetries which are allowed in space-time. Within the framework of Quantum Field Theory (unitarity of the $S$ matrix {\it etc}) it is generally admitted that we cannot go beyond supersymmetry (SUSY). However, the no-go theorem stating that supersymmetry is {\it the only non-trivial extension beyond the Poincar\'e algebra} is valid only if one considers Lie or Super-Lie algebras. Indeed, if one considers Lie algebras, the Coleman and Mandula theorem \cite{cm} allows only trivial extensions of the Poincar\'e symmetry, {\it i.e.} extra symmetries must commute with the Poincar\'e generators. In contrast, if we consider superalgebras, the theorem of Haag, Lopuszanski and Sohnius \cite{hls} shows that we can construct a unique (up to the number of supercharges) superalgebra extending the Poincar\'e Lie algebra non-trivially. It may seem that these two theorems encompass all possible symmetries of space-time. But, if one examines the hypotheses of the above theorems, one sees that it is possible to imagine symmetries which go beyond supersymmetry. Several possibilities have been considered in the literature \cite{ker, luis, fsusy, fsusy1d, fr,am, prs, fsusy2d, fvir, fsusy3d, fsusyh}, the intuitive idea being that the generators of the Poincar\'e algebra are obtained as an appropriate product of more fundamental additional symmetries. These new generators are in a representation of the Lorentz group which can be neither bosonic nor fermionic (bosonic charges close under commutators and generate a Lie algebra, whilst fermionic charges close under anticommutators and induce super-Lie algebras). In this paper we propose an algebraic structure, called an $F-$Lie algebra, which makes this idea precise in the context of fractional supersymmetry (FSUSY) of order $F$. Of course, when $F=1$ this is a Lie algebra, and when $F=2$ this is a Super-Lie algebra. We show that all examples of FSUSY considered in the literature can be described within this framework.\\ FSUSY ($F>2$) has been investigated in dimensions one, two and three. In $1D$ the algebraic structure is relatively simple \cite{fsusy1d,fr,am} (one just adds a new supercharge $Q$ such that $Q^F=\partial_t$). In two dimensions, one can add either two or an infinite number of additional generators \cite{prs,fsusy2d,fvir}. In three dimensions the situation is more complicated. We showed \cite{fsusy3d} that it is possible to inject equivariantly the vector representation of $so(1,2)$ in a quotient of the $F-$th symmetric product of an appropriate representation $D_{1/F}$ of $so(1,2)$. In other words, we were able to express the generators of space-time translations as symmetric $F-$ order polynomials in more fundamental generators but with the new supercharges satisfying extra constraints. We also constructed explicitly in \cite{fsusy3d} unitary representations of the corresponding algebraic structure which can be understood as relativistic anyons \cite{lm,b,jn,p}. However, it was not possible to consider translations as $F-$order symmetric products of the new supercharges without imposing extra constraints. In contrast, this problem exists neither in dimensions one and two \cite{fsusy1d,fr,am,prs, fsusy2d}, nor in any dimension when $F=2$ (SUSY). To understand the results of our paper \cite{fsusy3d} in terms of $F-$Lie algebras, we propose two solutions: (i) extending the vectorial representation or, (ii) extending the Poincar\'e algebra $B= t \oplus so(1,2)$ to $\hat B= \hat t \oplus { \mathrm Vir}$, where $so(1,2) \subset {\mathrm Vir}$ is the Virasoro (without central charge) algebra and $\hat t$ a representation of Vir which extends the vectorial representation of $so(1,2)$. Correspondingly we also have to extend the $D_{1/F}$ representation of $so(1,2)$ to $\hat D_{1/F}$. The problem encountered in $3D$ FSUSY and especially the solution we propose to solve it, enables us to define a general method of associating an FSUSY to any representation ${\cal D}$ of any Lie algebra $g$. This algebraic structure is in general associated to a non unitary infinite dimensional representation of $g$. Furthermore, as for $so(1,2)$, one can define an infinite dimensional Lie algebra $V(g)$ having $g$ as a sub-algebra and leading to an $F-$lie algebra. The content of this paper is as follows. In section two, we give a precise mathematical definition of the algebraic structure which underlies FSUSY. Several simple examples are then given. In section 3, we show how one can incorporate $3D$ FSUSY into this general mathematical description by extending the vectorial representation to an appropriate reducible (but indecomposable) representation. We then construct FSUSY starting from any semi-simple Lie algebra $g$ (playing the role of $so(1,2)$) and any representation ${\cal D}$ (playing the role of the vector representation). This construction involves taking the $F^{{\mathrm th}}-$root of ${\cal D}$ in some sense. In particular this means that one can construct FSUSY in all space-time dimensions. In section 4, we study an $F-$Lie algebra associated to an infinite dimensional algebra $V(g)$ having $g$ as a sub-algebra. For $g=so(1,2)$, $V(g)$ reduces to the centerless Virasoro algebra. \mysection{Algebraic Structure of Fractional Supersymmetry} In this section, we give the abstract mathematical structure which underlies this paper and which generalizes the theory of Lie super-algebras and their (unitary) representations. Let $F$ be a positive integer and $q=\exp{({2i \pi \over F})}$. We consider a complex vector space $S$ together with a linear map $\varepsilon$ from $S$ into itself satisfying $\varepsilon^F=1$. We set $A_k= S_{q^k}$ and $B=S_1$ (where $S_\lambda$ is the eigenspace corresponding to the eigenvalue $\lambda$ of $\varepsilon$) so that $S=B\oplus_{k=1}^{F-1} A_k$. The map $\varepsilon$ is called the grading. If $S$ is endowed with the following structures we will say that $S$ is a fractional super Lie algebra ($F$-Lie algebra for short): \begin{enumerate} \item $B$ is a Lie algebra and $A_k$ is a representation of $B$. We write these structures as a bracket $[b,X]$ with the understanding that $[b,X]=-[X,b]$ if $X \in A_k,~ b\in B$. It is clear that $[\varepsilon(X),\varepsilon(Y)]= \varepsilon\left([X,Y]\right)$. \item There are multilinear, $B-$equivariant ({\it i.e.} which respect the action of $B$) maps $\left\{~~, \cdots,~~ \right\}: {\cal S}^F\left(A_k\right) \rightarrow B$ from ${\cal S}^F\left(A_k\right)$ into $B$. In other words, we assume that some of the elements of the Lie algebra $B$ can be expressed as $F-$th order symmetric products of ``more fundamental generators''. Here ${ \cal S}^F(D)$ denotes the $F-$fold symmetric product of $D$. It is then easy to see that: \beqa \label{eq:epsi} \left\{\varepsilon(a_1), \cdots, \varepsilon(a_F)\right\}= \varepsilon\left(\left\{a_1, \cdots, a_F\right\}\right), \forall a_i \in A_k. \eeqa \item For $b_i \in B$ and $a_j \in A_k$ the following ``Jacobi identities'' hold: \beqa \label{eq:J} &&\left[\left[b_1,b_2\right],b_3\right] + \left[\left[b_2,b_3\right],b_1\right] + \left[\left[b_3,b_1\right],b_2\right] =0 \nonumber \\ &&\left[\left[b_1,b_2\right],a_3\right] + \left[\left[b_2,a_3\right],b_1\right] + \left[\left[a_3,b_1\right],b_2\right] =0 \nonumber \\ &&\left[b,\left\{a_1,\dots,a_F\right\}\right] = \left\{\left[b,a_1 \right],\dots,a_F\right\} + \dots + \left\{a_1,\dots,\left[b,a_F\right] \right\} \\ &&\sum\limits_{i=1}^{F+1} \left[ a_i,\left\{a_1,\dots, a_{i-1}, a_{i+1},\dots,a_{F+1}\right\} \right] =0. \nonumber \eeqa \noindent The first identity is the usual Jacobi identity for Lie algebras, the second says that the $A_k$ are representation spaces of $B$ and the third is just the Leibniz rule (or the equivariance of $\left\{~~, \cdots,~~ \right\}$). The fourth identity is the analogue of the graded Leibniz rule of Super-Lie algebras for $F-$Lie algebras \hskip -1 cm If we want to be able to talk about unitarity, we also require the following additional struc- \hskip -1 cm ture and in this case, $S$ is called an $F-$Lie algebra with adjoint. \item A conjugate linear map $\dag$ from $S$ into itself such that: \beqa \label{eq:conj} \begin{array}{ll} \mathrm{a)}&(s^\dag)^\dag=s, \forall s \in S \cr \mathrm{b)}& \left[a,b\right]^\dag= \left[b^\dag,a^\dag\right] \cr \mathrm{c)}&\varepsilon(s^\dag)=\varepsilon(s)^\dag \cr \mathrm{d)}& \left\{a_1,\cdots,a_F\right\}^\dag= \left\{\left(a_1\right)^\dag,\cdots,\left(a_F\right)^\dag\right\},~~~ \forall a \in A_k. \end{array} \eeqa From a) and c) we see that for $X \in B$ we have $X^\dag \in B$, and that for $X \in A_k$, we have $ X^\dag \in A_{F-k}.$ \end{enumerate} \noindent A unitary representation of an $F-$Lie algebra with adjoint $S$ is a linear map $\rho : ~ S \to \mathrm{End}(H)$, (where $H$ is a Hilbert space and ${\mathrm{End}}(H)$ the space of linear operators acting on $H$) and a unitary endomorphism $\hat \varepsilon$ such that $ \hat \varepsilon^F=1$ which satisfy \beqa \label{eq:rep} \begin{array}{ll} \mathrm{a)}& \rho\left(\left[x,y\right]\right)= \rho(x) \rho(y)- \rho(y)\rho(x) \cr \mathrm{b)}& \rho \left\{a_1.\cdots,a_F\right\}= {1 \over F !} \sum \limits_{\sigma \in S_F} \rho\left(a_{\sigma(1)}\right) \cdots \rho\left(a_{\sigma(F)}\right) \cr \mathrm{c)}& \rho(s)^\dag = \rho(s^\dag) \cr \mathrm{d)}& \hat \varepsilon \rho\left(s\right) \hat \varepsilon^{-1} = \rho\left(\varepsilon\left(s\right)\right) \end{array} \eeqa \noindent ($S_F$ being the group of permutations of $F$ elements). Note that with the normalisation of b), when $F=2$, one has $\rho(\{a_1,a_2\})=1/2(a_1 a_2 + a_2 a_1)$ instead of the usual $\rho(\{a_1,a_2\})=(a_1 a_2 + a_2 a_1)$. As a consequence of these properties, since the eigenvalues of $\hat \varepsilon$ are $\mathrm{F}^{\mathrm{th}}-$ roots of unity, we have the following decomposition of the Hilbert space $$H= \bigoplus \limits_{k=0}^{F-1} H_k,$$ \noindent where $H_k=\left\{\left|h\right> \in H ~:~ \hat \varepsilon\left|h\right>=q^k \left|h\right> \right\}$. The operator $N \in \mathrm{End}(H)$ (the set of linear operators acting on $H$) defined by $N\left|h\right>=k \left| h \right>$ if $\left|h\right> \in H_k$ is the ``number operator'' (obviously $q^N=\hat \varepsilon$). Since $\hat \varepsilon \rho(b)= \rho(b) \hat \varepsilon, \forall b \in B$ each $H_k$ provides a representation of the Lie algebra $B$. Furthermore, for $a \in A_\ell$, $\hat \varepsilon \rho(a)=q^\ell \rho(a) \hat \varepsilon$ and so we have $\rho(a) .H_k\ \subseteq H_{k+\ell ({\mathrm{mod~} F)}}$ \\ \noindent \underline{Remark 1}: \noindent For all $k=1,\cdots, F-1$ it is clear that the subspace $B \oplus A_k$ of $S$ satisfies (\ref{eq:epsi}-\ref{eq:J}) and the subspace $B \oplus A_k \oplus A_{-k}$ satisfies (\ref{eq:epsi}-\ref{eq:conj}) (when $S$ has an adjoint). \noindent \underline{Remark 2}: \noindent It is important to notice that bracket $\{ \cdots \}$ is a priori not defined for elements in different gradings. \noindent \underline{Remark 3}: \noindent If we set \beqa \label{eq:real} B_{{\tiny \hbox{\it I\hskip -2.pt R }}} =\Big \{b \in B: b^\dag=-b \Big \} \nonumber \\ A_{{\tiny \hbox{\it I\hskip -2.pt R }}} =\Big \{a \in A: a^\dag=a \Big \}, \eeqa \noindent then $S_{{\tiny \hbox{\it I\hskip -2.pt R }}} = B_{{\tiny \hbox{\it I\hskip -2.pt R }}} \oplus A_{{\tiny \hbox{\it I\hskip -2.pt R }}}$ is stable by $\varepsilon$ and satisfies (\ref{eq:epsi}-\ref{eq:J}). Here we use the normalizations conventionally used in mathematical literature (no $i$ factor in the structure constants of the algebra). For physicists, notice that if $b^\dag = -b$ then $(i b)$ is hermitian. \vskip .5truecm \subsection*{Example 1:} Obviously, a $1-$Lie algebra is just a Lie algebra. A $2-$Lie algebra is just a Lie super-algebra: $S=B\oplus A_{1}$, with even part $B$ and odd part $A_1$. In supersymmetry, because of the spin statistics theorem $A_1$ is a fermionic representation of $B$. Note that the Jacobi identities (\ref{eq:J}) above reduce to the standard Jacobi identities of a Super-Lie algebra. If we consider unitarity in SUSY, property (\ref{eq:conj}), (\ref{eq:rep}) above have also to be considered but for the super-Poincar\'e algebra, the nature of $\dag$ depends very much on the dimension of the space-time and the signature of the metric (Majorana, Weyl, Majorana-Weyl, $SU(2)-$Majorana and $SU(2)-$Majorana-Weyl conditions). \vskip .5truecm \subsection*{Example 2:} Let $V$ be finite dimensional complex vector space and let $\varepsilon:~V \rightarrow V$ be a linear operator satisfying $\varepsilon^F=1$. Then $$V=\oplus_{k=0}^{F-1} V_k,$$ \noindent where $V_k= \left\{ \left|v\right> \in V : \varepsilon \left|v\right>=q^k \left|v\right> \right\}$. We define $$A_k=\left\{ f \in \mbox{End}(V) : \varepsilon\circ f\circ \varepsilon^{-1}= q^k f\right\}$$ \noindent and $S=B \oplus_{k=1}^{F-1} A_k$ (with $B=A_0$). Since $A_k A_\ell \subset A_{k + \ell (\mbox{\tiny{mod~}} F)}$ one has \beqa \label{eq:Vect1} &\left[A_0, A_k \right] \subset A_k \\ \label{eq:Vect2} &{\underbrace{A_k A_k \cdots A_k}} \subset A_0. \\ & \hskip -1cm {\mathrm F-times} \nonumber \eeqa \noindent The bracket $[~,~]$ of $S$ is defined by (\ref{eq:Vect1}) and $\{ \cdots \}: S^F(A_k) \rightarrow B$ by \\ $\left\{a_1 \cdots a_F\right\} = 1/F! \sum \limits_{\sigma \in S_F} a_{\sigma(1)} \cdots a_{\sigma(F)}$. The first three Jacobi (\ref{eq:J}) identities are clearly satisfied, and calculation shows that the last Jacobi identity also holds. Thus $S$ is an $F-$Lie algebra. If $V$ is endowed with a hermitian metric and $\varepsilon$ is a unitary operator then adjunction defines an adjoint on the $F-$Lie algebra $S$. \vskip .5truecm \subsection*{Example 3:} In $1D$ \cite{fsusy1d,fr,am} the simplest $F-$Lie algebra is two dimensional, and is generated by the operators $\partial_t, Q$ with the relation $Q^F=\partial_t$. We take $B= \Big < \partial_t \Big >$, the translation in time, and $A_1 = \Big < Q \Big >$ . We obviously have $\varepsilon (\partial_t)=\partial_t$ and $\varepsilon(Q) = q Q$. An explicit representation in terms of generalized grassmann variables \cite{rr,r,re,frr,hq} can be constructed \cite{fsusy1d,fr,am}. It is possible to extend this $F-$Lie algebra to a $F-$Lie algebra with adjoint by the addition of one more generator, $Q^\dag$, such that $\varepsilon(Q^\dag)=q^{-1} Q^\dag$, $(Q^\dag)^F=(\partial_t)^\dag=-\partial_t$. Let us recall once again that there are no algebraic relations between $Q$ and $Q^\dag$. \vskip .5truecm \subsection*{Example 4:} In $2D$ there are several possible algebras. The simplest one is obtained by considering the $3$ generators $\partial_z, \partial_{\bar z}$ and $Q_z$. We set $B= \Big < \partial_z, \partial_{\bar z} \Big >$ and $A_1= \Big <Q_z \Big >$, the relations are $ \big ( Q_z \big)^F= \partial_z$ and $[\partial_z, Q_z]= [\partial_{\bar z}, Q_z]=0$. \\ This $F-$Lie algebra can be extended to a $4-$dimensional $F-$Lie algebra with adjoint by adding one more generator $Q_{\bar z} \in A_{-1}$ such that $\left(\partial_z\right)^\dag=\partial_{\bar z}$, $\big ( Q_{\bar z} \big)^F=\partial_{\bar z}$ and $\big ( Q_z \big)^\dag= Q_{\bar z}$ \cite{prs,fsusy2d}. \\ There is also a more complicated algebraic extension, involving an infinite number of generators which corresponds to an extension of the Virasoro algebra without central charge. In addition to the Virasoro generators $L_n, \overline{L}_n~ n \in \hbox{$Z$$ we add the generators $G_r, r \in \hbox{$Z$+1/F$, which correspond to the modes of a field of conformal weight $1+1/F$ and satisfy the following relations \cite{fvir} \beqa \label{eq:fv2} \left[L_n,L_m\right]& =& (n-m)L_{m+n} \nonumber \\ \left[\overline{L}_n,\overline{L}_m\right]& =& (n-m)\overline{L}_{m+n} \nonumber \\ \left[{L_n},\overline{L}_m\right]&=& 0 \\ \left[L_n,G_r\right] &=& ({n \over F}-r)G_{n+r} \nonumber \\ \left[\overline{L}_n,G_r\right] &=&0 \nonumber \\ \left\{G_{r_1},\cdots,G_{r_F}\right\}&=& L_{r_1 + \cdots + r_F}, \nonumber \eeqa \noindent Here we take $B= {\mathrm {Vir}} \oplus \overline{{\mathrm {Vir}}}$ and $A_1=\Big < G_r, r \in \hbox{$Z$ + 1/F \Big >$. In this extension, we have $L_1 \equiv \partial_z, \overline{L}_1 \equiv \partial_{\bar z}$ and $G_{{1 \over F}} \equiv Q_z$. We can also include an adjoint by adding $A_{-1}=\overline{A}_1$. As in $1D$, it is possible to construct an explicit realisation of the above algebras using generalised grassmann variables \cite{fvir,prs, fsusy2d}. In all the given examples, appropriate representations have been obtained in terms of adapted superfields \cite{fsusy1d,fr,am,prs,fsusy2d}. To our knowledge, unitarity remains an unsolved problem. \vskip .5truecm In three dimensions, the situation is much more complicated and we will study this in the next section. \subsection*{Example 5} Let $g$ be a complex Lie algebra and let ${\mathbf r}, {\mathbf r^\prime}$ be representations of $g$ such that there is a $g-$equivariant map $\mu: S^F(\mathbf{r}) \rightarrow \mathbf{r^\prime}$. We set $$S=B\oplus A_1 = (g \oplus {\mathbf r^\prime}) \oplus {\mathbf r}.$$ \noindent $B=g \oplus {\mathbf r^\prime}$ is a Lie algebra as the semi-direct product of $g$ and $\mathbf{r^\prime}$ (the latter with the trivial bracket). We can extend the action of $g$ on ${\mathbf r}$ to an action of $B$ on ${\mathbf r}$ by letting ${\mathbf r^\prime}$ act trivially on ${\mathbf r}$. This defines the bracket $[~,~]$ on $S$. For the map $\left\{\cdots \right\}$ we take $\mu$. The first three Jacobi identities (\ref{eq:J}) are clearly satisfied, and the fourth is also satisfied as each term in the expression of the L.H.S vanishes. For example, if $${\cal S}^F\left({\mathbf r}\right)= \oplus_k {\mathbf r_k},$$ \noindent is a decomposition into irreducible summands, then for a given $k$ $$S_k=(g \oplus {\mathbf r_k}) \oplus {\mathbf r},$$ is an $F-$Lie algebra. \\ As an illustration, if $g=so(1,2)$ and ${\mathbf r}={\mathbf 2} \oplus {\mathbf 1} $ (the spin representation plus the trivial representation), then ${\cal S}^3({\mathbf 2} \oplus {\mathbf 1}) ={\cal S}^3({\mathbf 2}) \oplus {\cal S}^2({\mathbf 2}) \oplus {\mathbf 2} \oplus {\mathbf 1} = {\mathbf 4} \oplus {\mathbf 3} \oplus {\mathbf 2} \oplus {\mathbf 1} $ and it is possible to obtain the spinorial or the vectorial representations of $so(1,2)$ from a symmetric product of order 3. This can be compared with the result of R. Kerner \cite{ker} where a cubic root of the Dirac equation is obtained. More generally, for any $F$: ${\cal S}^F({\mathbf 2} \oplus {\mathbf 1})={\cal S}^F({\mathbf 2}) \oplus {\cal S}^{F-1} ({\mathbf 2}) \oplus \cdots \oplus {\mathbf 1} = (\mathbf{F+1}) \oplus \mathbf{F} \oplus \cdots \oplus \mathbf{1}$. \mysection{Fractional Supersymmetry and finite dimensional \ Lie algebras} \subsection{Fractional Supersymmetry in three dimensions} In \cite{fsusy3d} we considered FSUSY in three dimensions. In order to understand our results in term of $F-$Lie algebras let us introduce a realisation of $so(1,2)$ which is convenient for explicit calculations. Let ${\cal F}$ be the vector space of functions on $\hbox{\it I\hskip -2.pt R }^{2,*}_+=\big\{(x,y) \in \hbox{\it I\hskip -2.pt R }^2: x,y > 0 \big \}$. Consider the linear operators acting on ${\cal F}$ given by \beqa \label{eq:so(1,2)} J_{-} &=& x \partial_y \nonumber \\ J_0 &=& {1 \over 2} (y \partial_y - x \partial_x) \\ J_{+} &=& y \partial_x. \nonumber \eeqa \noindent These operators satisfy the commutation relations \beqa \label{eq:corel} \big[J_{-}, J_+\big] &=& -2 J_0 \nonumber \\ \big [J_{0}, J_+\big] &=& J_+ \\ \big[J_{0}, J_{-}\big] &=& -J_{-}, \nonumber \eeqa \noindent and thus generate the Lie algebra $so(1,2)$. It is easy to check that the following subspaces of ${\cal F}$ are representations of $so(1,2)$: \beqa \label{eq:repp} {\cal D}_{-n}&=&\Big<~ x^{2n},x^{2n-1}y,\cdots, x y^{2n-1}, y^{2n} \Big >, \hskip .5cm \Big(n \in \hbox{\it I\hskip -2.pt N}/2 \Big) \nonumber \\ {\cal D}^+_{-\lambda}&=&\Big<~ x^{2 \lambda} \left( {y \over x }\right)^m, m \in \hbox{\it I\hskip -2.pt N} \Big >,\hskip .5cm \Big(\lambda \in \hbox{\it I\hskip -2.pt R } \setminus \hbox{\it I\hskip -2.pt N}/2\Big) \\ {\cal D}^-_{-\lambda}&=&\Big<~ y^{2 \lambda} \left( {x \over y }\right)^m, m \in \hbox{\it I\hskip -2.pt N} \Big >, \hskip .5cm \Big( \lambda \in \hbox{\it I\hskip -2.pt R } \setminus \hbox{\it I\hskip -2.pt N}/2 \Big). \nonumber \eeqa \noindent Of course other representations can also be obtained (for instance unbounded from below and above) but they are not useful for our purpose. The representation ${\cal D}_{-n}$ is the $(2n+1)-$dimensional irreducible representation and the representations ${\cal D}^\pm_{-\lambda}$ are infinite dimensional representations, bounded from below and above respectively. It is important to emphasize that the representations given in (\ref{eq:repp}) do not have the normalizations conventionally used in the literature and the basis is not orthonormal, but those normalizations are convenient for further developments. For a general classification of the representations of three-parameter Lie algebras, see {\it e.g.} \cite{wy} where analogous monomials (of the form $x^\alpha y^\beta \left({x \over y}\right)^m$, with $\alpha, \beta \in \hbox{\it l\hskip -5.5pt C\/}, m \in \hbox{$Z$)$ are considered. \\ In the paper \cite{fsusy3d} we introduced four representations, ${\cal D}^\pm_{-1/F,\pm}$. These are are related to the above representations by the following isomorphisms \beqa D^+_{-1/F,+} \cong D^+_{-1/F,-} \cong D^+_{-1/F} \nonumber \\ D^-_{-1/F,+} \cong D^-_{-1/F,-} \cong D^-_{-1/F}. \nonumber \eeqa \noindent In this article, for practical reasons we work only with the representations (\ref{eq:repp}). The multiplication map $m_n : {\cal F} \times \cdots \times {\cal F} \rightarrow {\cal F}$ given by \beq \label{eq:multi} \ m_n(f_1,\cdots, f_n)= f_1 \cdots f_n \eeq \noindent is multilinear and totally symmetric and hence induces a map $\mu_F$ from ${\cal S}^F({\cal F})$ into ${\cal F}$. Restricting to ${\cal S}^F\left(D^\pm_{-1/F}\right)$ one sees that \beqa \label{eq:FSUSY} S^F\big({\cal D}^{+}_{-1/F}\big)_{{\mathrm red}} \buildrel{\hbox{def}} \over = \mu_F\Big(S^F\big( D^{+}_{-1/F}\big)\Big) &=&\Big<x^2 \left({y \over x} \right)^m, ~ m\in \hbox{\it I\hskip -2.pt N} ~~\Big> \supset {\cal D}_{-1} \\ S^F\big({\cal D}^{-}_{-1/F}\big)_{{\mathrm red}} \buildrel{\hbox{def}} \over = \mu_F\Big(S^F\big( D^{-}_{-1/F}\big)\Big) &=&\Big<y^2 \left({x \over y} \right)^m, ~ m\in \hbox{\it I\hskip -2.pt N} ~~\Big> \supset {\cal D}_{-1} \nonumber \eeqa \noindent The simple observation of (\ref{eq:FSUSY}) together with example 5 in section 2 naturally lead to the $F-$Lie algebra \beqa \label{eq:F-so(1,2)} \Big(so(1,2) \oplus S^F\big({\cal D}^{\pm}_{-1/F}\big)_{{\mathrm red}} \Big) \oplus {\cal D}^\pm_{-1/F}. \eeqa \noindent In \cite{fsusy3d}, by considering an adapted conjugations $\dag: {\cal D}^+_{-1/F}= \left({\cal D}^-_{-1/F}\right)^\dag$ and from the Wigner induced representation we proved that the representations of 3d-FSUSY are unitary and induce a symmetry between relativistic anyons. Looking at the representations defined in (\ref{eq:FSUSY}) {\it i.e.} $S^F\big({\cal D}^{\pm}_{-1/F}\big)_{{\mathrm red}}$, one sees that, even though ${\cal D}_{-1}$ is a subspace stable under $so(1,2)$ there is no complement stable under $so(1,2)$ \cite{fsusy3d}. Indeed, these representations cannot be built from a primitive vector. This is due to the fact that $J_{+}^3 \Big(x^2\Big) =0$ and consequently we cannot reach $x^{-1}y^3$ from $x^2$ but conversely $J_{-}^3 $ $ \Big(x^{-1}y^3)= 6 x^2$ (such reducible but indecomposable representations also appear in \cite{wy}). This is the reson why there is no $F-$Lie algebra structure on $so(1,2) \oplus {\cal D}_{-1}$. \subsection{Extension to any Lie algebra} We consider now $g$ a complex semi-simple Lie algebra of rank $r$ and ${\cal D}$ an arbitrary representation. The purpose of this section, is to construct an $F-$Lie algebra $S=B \oplus A_1$ such that the Lie algebra $B$ contains the semi-direct product $g \oplus {\cal D}$. If $g=so(1,2)$ and ${\cal D}$ is the vector representation, this construction leads to the $F-$Lie algebra (\ref{eq:F-so(1,2)}) above. Let $h$ be a Cartan sub-algebra of $g$, let $\Phi \subset h^\star$ (the dual of $h$) be the corresponding set of roots and let $f_\alpha$ be the one dimensional root space associated to $\alpha \in \Phi$. We choose a basis $\{H_i, i=1, \cdots, r\}$ of $h$ and elements $E^\alpha \in f_\alpha$ such that the commutation relations become \beqa \label{eq:lie} \big[H_i,H_j \big] &=& 0 \nonumber \\ \big[H_i,E^\alpha \big] &=& \alpha^i E^\alpha \\ \big[E^\alpha, E^\beta\big] &=& \left \{ \begin{array}{ll} \epsilon\{\alpha,\beta\} E^{\alpha+\beta}& {\mathrm {~~if~~}} \alpha + \beta \in \Phi \cr {2\alpha.H \over \alpha.\alpha}& {\mathrm {~~if~~} } \alpha+\beta=0 \cr 0& {\mathrm {~~otherwise }} \end{array} \right. \nonumber \eeqa \noindent Recall that the real Lie algebra spanned by the $H_i$ and the $E^\alpha$ is the split real form of $g$, and that the real Lie algebra spanned by $iH_j$, $E^\alpha-E^{-\alpha}$ and $i(E^\alpha+E^{-\alpha})$ is the compact real form of $g$. We now introduce $\{\alpha_{(1)},\cdots, \alpha_{(r)}\}$ (the positive roots) a basis of simple roots. The weight lattice $\Lambda_W(g) \subset h^\star$ is the set of vectors $\mu$ such that $ {2 \alpha.\mu \over \alpha.\alpha} \in \hbox{$Z$$ and, as is well known, there is a basis of the weight lattice consisting of the fundamental weights $\{\mu_{(1)}, \cdots, \mu_{(r)} \}$ defined by $ {2 \mu_{(i)}.\alpha_{(j)} \over \alpha_{(j)}.\alpha_{(j)}}= \delta_{ij}$. A weight $\mu = \sum \limits_{i=1}^r n_i \mu_{(i)}$ is called dominant if all the $n_i \ge 0$ and it is well known that the set of dominant weights is in one to one correspondence with the set of (equivalence classes of) irreducible finite dimensional representations of $g$. Recall briefly how one can associate a representation of $g$ to $\mu \in h^\star$. In ${\cal{U}}(g)$, the universal enveloping algebra of $g$, let $I_\mu$ be the left-ideal generated by the elements $\Big\{E^\alpha (\alpha > 0), h_i-\mu(h_i). {\mathbf I} (h_i \in h)\Big\}$, where $ {\mathbf I}$ is the identity of ${\cal{U}}(g)$ and $h_i= 2 {\alpha_{(i)}.H_i \over \alpha_{(i)}^2}$. \noindent The Lie algebra $g$ acts on ${\cal{U}}(g)$ by left multiplication, $I_\mu$ is stable under this action and therefore the quotient ${\cal V}_\mu={\cal{U}}(g)/I_\mu$ is a representation space of $g$: ${\cal V}_\mu$ is a highest weight representation and is called the Verma module associated to $\mu$ \cite{kr}. If $\mu$ is dominant, then ${\cal V}_\mu$ has a unique maximal proper sub-representation $M_\mu$ and the quotient ${\cal D}_\mu= {\cal V}_\mu/M_\mu$ is an irreducible finite dimensional representation of $g$. To come back to our original problem, consider a finite dimensional irreducible representation ${\cal D}_\mu$ of $g$. If ${\cal V}_{\mu/F}$ is the Verma module associated to $\mu/F$, there is a $g-$equivariant map \beq \label{eq:sfmu} i: {\cal V}_\mu \rightarrow {\cal S}^F\left({\cal V}_{\mu/F} \right), \eeq \noindent because ${\mathbf I} \otimes \cdots \otimes {\mathbf I} \in {\cal S}^F\left({\cal V}_{\mu/F}\right)$ is a highest weight vector of weight $\mu$. Taking the quotient by $M_\mu$ one obtains a $g-$equivariant inclusion \beq \label{eq:dmu} {\cal D}_\mu \hookrightarrow {\cal S}^F\left({\cal V}_{\mu/F} \right)/i(M_\mu), \eeq \noindent since $D_\mu= {\cal V}_\mu/M_\mu$. Denoting $ {\cal S}^F\left({\cal V}_{\mu/F} \right)/i(M_\mu)$ by ${\cal S}^F\left({\cal D}_{\mu/F}\right)_{\mathrm{red}}$ then, as in example 5, \beq \label{eq:F-g} S=(g \oplus {\cal S}^F\left({\cal D}_{\mu/F}\right)_{\mathrm{red}}) \oplus {\cal V}_{\mu/F} \eeq \noindent is naturally an $F-$Lie algebra. We can reformulate this construction in less abstract terms. If $\left|\mu\right>$ is the primitive vector associated to a dominant weight $\mu$ ({\it i.e.} ($E^\alpha \left|\mu\right> =0, \alpha >0$ and $h_i\left|\mu\right>= n_i\left|\mu\right>= \mu(h_i)\left|\mu\right>$)), the representation ${\cal D}_\mu$ is generated by the action of the $E^\alpha, \alpha < 0$ on $\left|\mu\right>$. Because the representation is finite dimensional, corresponding to the highest weight state $\left|\mu\right>$ we have a lowest state $\left|\mu^\prime\right>$, $\mu^\prime= \sum \limits_{i=1}^r n^\prime_i \mu_{(i)}$ ($E^\alpha\left|\mu^\prime\right>=0, \alpha <0$). Of course the two representations built with $\mu$ or $\mu^\prime$ are the same. But, if the weight is not dominant the situation is more involved. For our purposes, to the representation $\cal{D}_\mu$ we associate two infinite dimensional representations: one associated to the weight ${\mu \over F}$ and one to ${\mu^\prime \over F}$, noted ${\cal{D}}_{{\mu \over F}}^\pm$ respectively (the first is bounded from below and the second from above). These two inequivalent representations are characterized by primitive vectors: \beqa \label{eq:muF} {\cal{D}}_{{\mu \over F}}^+: \left|{\mu \over F} \right>&& h_i\left|{\mu \over F} \right> = {n_i \over F} \left|{\mu \over F} \right>, E^\alpha \left|{\mu \over F} \right> =0, \alpha >0 \\ {\cal{D}}_{{\mu \over F}}^-: \left| {\mu^\prime \over F} \right>&& h_i\left|{\mu^\prime \over F} \right> = {n^\prime_i \over F} \left|{\mu^\prime \over F}\right>, E^\alpha \left|{\mu^\prime \over F} \right> =0, \alpha < 0. \nonumber \eeqa \noindent ${\cal D}^+_{\mu/F}$ is the Verma module ${\cal V}_{\mu/F}$ abstractly defined above: In these terms the projections $\mu_F$ from ${\cal S}^F\Big( {\cal D}^\pm_{\mu \over F}\Big) \longrightarrow {\cal{S}}^F\left({\cal{D}}_{\mu \over F}^\pm\right)_{{\mathrm red}}$ is given by $\mu_F \left({\cal S}^F\Big(\left|h_1\right>,\cdots,\left|h_F\right> \Big) \right)= \left|h_1 + \cdots + \cdots h_F \right>$. \noindent We observe that that ${\cal{D}}_\mu \subset {\cal{S}}^F\left({\cal{D}}_{\mu \over F}^\pm\right)_{{\mathrm red}}$ because $\mu_F \left({\cal S}^F\Big(\left|{\mu \over F}\right>,\cdots, \left|{\mu \over F}\right> \Big) \right)= \left|\mu\right>$ (or $\mu_F \left( {\cal S}^F\right. $ $\left. \Big( \left| {\mu^\prime \over F} \right>, \cdots, \left|{\mu^\prime \over F}\right> \Big) \right)= \left|\mu^\prime\right>$) is the primitive vector of ${\cal D}_\mu$. \section{Fractional supersymmetry and infinite dimensional algebras} In the previous section we constructed a canonical $F-$Lie algebra \beq \label{eq:flie} S=(g \oplus {\cal S}^F\left({\cal D}_{\mu/F}\right)_{\mathrm{red}}) \oplus {\cal D}_{\mu/F} \eeq \noindent associated to a finite dimensional Lie algebra $g$ and an irreducible finite dimensional representation ${\cal D}_\mu$. In this section we will show that one can extend the representation $g$ in $D_\mu$ to a representation of an infinite dimensional Lie algebra $V(g)$ in $\hat D_\mu$, and construct an $F-$Lie algebra containing (\ref{eq:flie}) as a sub-algebra. \subsection{$so(1,2)$ and the Virasoro algebra} It is well know that the Virasoro algebra admits $so(1,2)$ as a sub-algebra. The action (\ref{eq:so(1,2)}) of $so(1,2)$ on ${\cal F}$ extends to an action of the Virasoro algebra (without central extension) on ${\cal F}$ by setting \beqa \label{eq:vir} L_n={n+1 \over 2} \left( {y \over x} \right)^n x \partial_x +{n-1 \over 2} \left( {y \over x} \right)^n y \partial_y, n\in \hbox{$Z$. \eeqa \noindent One can verify the commutation relations \beqa \label{eq:comrel2} [L_n,L_m]=(n-m) L_{n+m} \eeqa \noindent and that $J_{-}=-L_{-1},J_0=-L_0,J_+=L_1$. In analogy with the representations (\ref{eq:repp}) of $so(1,2)$ we define representations of Vir as follows: \beqa \label{eq:repv} \hat{\cal D}_{-n}&=&\Big<~ f^{(-n)}_{m-n}=x^{2n} \left( {y \over x} \right)^m, m \in \hbox{$Z$ \Big >, \hskip .5cm \left(n \in \hbox{\it I\hskip -2.pt N}/2 \right) \nonumber \\ \hat {\cal D}^+_{-\lambda}&=&\Big<~ f^{(+,-\lambda)}_{m-\lambda}=x^{2 \lambda} \left( {y \over x }\right)^m, m \in \hbox{$Z$ \Big >, \hskip .5cm \left(\lambda \in \hbox{\it I\hskip -2.pt R } \setminus \hbox{\it I\hskip -2.pt N}/2 \right)\\ \hat {\cal D}^-_{-\lambda}&=&\Big<~ f^{(-,-\lambda)}_{\lambda -m}=y^{2 \lambda} \left( {x \over y} \right)^m, m \in \hbox{$Z$ \Big >, \hskip .5 cm \left(\lambda \in \hbox{\it I\hskip -2.pt R } \setminus \hbox{\it I\hskip -2.pt N}/2 \right). \nonumber \eeqa \noindent Then, one can check explicitly that the action of Vir on (\ref{eq:repv}) is given by \beqa \label{eq:viract} L_k \left(f^{(-n)}_{p}\right)&=&(kn-p) f^{(-n)}_{k+p} \nonumber \\ L_k \left(f^{(+,-\lambda)}_{p}\right)&=&(k\lambda-p) f^{(+,-\lambda)}_{k+p} \\ L_k \left(f^{(-,-\lambda)}_{p}\right)&=&(k\lambda-p) f^{(+,-\lambda)}_{k+p}, \nonumber \eeqa \noindent where the indices in (\ref{eq:repv}) are chosen in such a way that they correspond to the eigenvalues of $-L_0$ {\it i.e.} the helicity. In the language of conformal field theory, $f^{(-n)}_p, f^{(+,-\lambda)}_p$ and $f^{(-,-\lambda)}_p$ correspond to the modes of conformal fields of conformal weight $n+1$ and $\lambda+1$ respectively. Finally, we observe that the the representations (\ref{eq:repp}) are included in the corresponding representations (\ref{eq:repv}) and in each case that the action of Vir extends the action of $so(1,2)$. Let us remark that these representations are all unbounded from below and above ({\it i.e.} they cannot be obtained from primitive vectors or a highest/lowest weight state).\\ The fundamental property of these representations is that there is a Vir-equivariant map from $S^F\Big(\hat {\cal D}^+_{-1/F}\Big)$ and $S^F\Big(\hat {\cal D}^-_{-1/F}\Big)$ to $\hat {\cal D}_{-1}$. This is just the multiplication map $\mu_F: {\cal F} \rightarrow {\cal F}$ (see (\ref{eq:multi})) which is obviously Vir-equivariant. In fact, a direct calculation shows that \beqa \label{eq:FSUSY1} S^F\big(\hat{\cal D}^{\pm}_{-1/F}\big)_{{\mathrm red}} \buildrel{\hbox{def}} \over = \mu_F\Big(S^F\big(\hat{\cal D}^{\pm}_{-1/F}\big)\Big) \cong \hat {\cal D}_{-1}, \eeqa \noindent and hence that $S^F\big(\hat{\cal D}^{\pm}_{-1/F}\big)_{{\mathrm red}}$ and $\hat {\cal D}_{-1}$ are isomorphic. By the method explained in example 5, \beq \label{eq:falg} S=\left(\mathrm{Vir} \oplus \hat {\cal D}_{-1}\right) \oplus \hat D^{+}_{-1/F} \oplus \hat D^{-}_{-1/F}, \eeq \noindent is an $F-$Lie algebra. \noindent Denoting $\hat {\cal D}_{-1} = \Big<~P_{m-1}=f^{(-1)}_{m-1}, m\in \hbox{$Z$ ~~\Big>$, $\hat {\cal D}^\pm_{-1/F} = \Big<~Q^\pm_{\pm(m-1/F)}=f^{(\pm,-1/F)}_{\pm(m-1/F)}, m\in \hbox{$Z$ ~~\Big>$, the brackets in $S$ are given explicitly by the following formulae: \beqa \label{eq:F-vir} \Big[ L_n, L_m \Big] &=& (n-m) L_{n+m} \nonumber \\ \Big[ L_n, P_m \Big] &=&(n-m) P_{n+m} \nonumber \\ \Big[ L_n, Q^\pm_r \Big] &=&({n \over F}-r) Q^\pm_{n+r}\nonumber \\ \Big[ P_n, P_m \Big] &=& 0 \\ \Big[ P_n, Q^\pm_r \Big] &=& 0 \nonumber \\ \left\{Q^\pm_{r_1}, \cdots,Q^\pm_{r_F} \right\} \nonumber &=&P_{r_1+ \cdots r_F}. \eeqa \noindent \underline{Remark} Any weight $\mu$ of the Lie algebra $so(1,2)$ can be considered as a weight $\hat \mu$ of Vir if we set $\hat \mu(L_0)= -\mu(J_0)$ \cite{kr}. As in section 3.1, we can construct the associated Verma module and define the $F-$Lie algebra \beq \label{eq:F-Vir} S=({\mathrm{Vir}} \oplus {\cal S}^F\left({\cal V}_{\mu/F}\right)_{\mathrm{red}}) \oplus {\cal V}_{\mu/F} \eeq \noindent with obvious notation (see (\ref{eq:F-g}). Of course this construction is different from the previous one (\ref{eq:falg}) since ${\cal V}_{\mu/F}$ is a highest weight representation of Vir. \subsection{The construction of $V(g)$} In this section, we will construct an infinite dimensional Lie algebra $V(g)$ which contains the Lie algebra $g$ in the same way as Vir contains $sl(2)$. Let $g$ be a semi-simple complex Lie algebra, let $h$ be a Cartan sub-algebra, let $\Phi_+$ be the positive roots and let $(\alpha_{(1)},\cdots,\alpha_{(r)})$ be the positive simple roots. \\ We consider the vector space $V$ generated by \beqa \label{eq:gene} V=\Big<L_0^{\alpha_i},~~ L^\alpha_{n}:~~ i=1,\cdots,r, ~~\alpha \in \Phi_+,~~ n \in \hbox{$Z$^* \Big> \eeqa \noindent satisfying the commutation relations: \begin{enumerate} \item \begin{enumerate} \item for positive simple roots $\alpha_{(1)},\cdots,\alpha_{(r)}$, \beqa \label{eq:vir-g1} \left[L_0^{\alpha_i}, L_0^{\alpha_j} \right]=0. \eeqa \item for $n >0$, \beqa \label{eq:vir-g2} \left[{L_0^{\alpha_i} \over- n}, {L^\alpha_{\pm n} \over \pm n} \right] &=& \pm{ \alpha_{(i)} . \alpha \over \alpha_{(i)}^2}{L^\alpha_{\pm n} \over \pm n},~ i=1,\cdots,r, \alpha \in \Phi_+. \nonumber \\ \left[{L^\alpha_n\over n},{L^\beta_n \over n}\right] &=& \left \{ \begin{array}{ll} \epsilon\{\alpha,\beta\} {L^{\alpha+\beta}_n\over n}& {\mathrm {~~if~~}} \alpha + \beta \in \Phi_+ \cr 0& {\mathrm {~~otherwise }}. \end{array} \right. \nonumber \\ \left[{L^\alpha_n \over n}{,L^\beta_{-n} \over - n} \right] &=& \left \{ \begin{array}{ll} \epsilon(\alpha,-\beta){ L_n^{\alpha-\beta} \over n}& {\mathrm {~~if~~}} \alpha - \beta \in \Phi_+ \cr \epsilon(\alpha,-\beta) {L_{-n}^{-\alpha+\beta} \over -n}& {\mathrm {~~if~~}} -\alpha + \beta \in \Phi_+ \cr 2 {L^\alpha_0 \over -n} & {\mathrm {~~if~~} } \alpha=\beta. \cr \end{array} \right. \\ \left[{L^\alpha_{-n} \over -n},{L^\beta_{-n}\over -n}\right] &=& \left \{ \begin{array}{ll} \epsilon\{-\alpha,-\beta\} {L^{\alpha+\beta}_{-n}\over -n}& {\mathrm {~~if~~}} \alpha + \beta \in \Phi_+ \cr 0& {\mathrm {~~otherwise. }} \end{array} \right. \nonumber \eeqa \noindent Thus $g_n=\Big<L_0^{\alpha_i}, L_{\pm n}^{\alpha}: \alpha \in \Phi_+, i=1,\cdots,r\Big>$ is a Lie algebra isomorphic to $g$, with Cartan sub-algebra $\Big<L_0^{\alpha_i}, i=1,\cdots,r \Big>$, roots $\Phi$, simple positive roots $\alpha_{(1)},\cdots, \alpha_{(r)}$ and root spaces $\Big<L^\alpha_{\pm n} \Big>$. The isomorphism with (\ref{eq:lie}) is: \beqa \label{eq:gn} {L_0^{\alpha_i} \over- n} &\Leftrightarrow& {\alpha_{(i)} . H \over \alpha_{(i)}^2} \\ \nonumber {{L^\alpha_n} \over n} &\Leftrightarrow& E^\alpha \\ {{L^\alpha_{-n}} \over -n} &\Leftrightarrow& E^{-\alpha} \nonumber. \eeqa It is important to emphasize that the Cartan sub-algebra of each $g_n$ is independent of $n$. Consequently the relation $[L^\alpha_1,L^\alpha_{-1}]=2 L^\alpha_0$ defines with no ambiguity $L^\alpha_0$ when $\alpha$ is not a simple root (it is just a linear combination of the $L^{\alpha_i}_0$ where $\alpha_{(i)}, i=1,\cdots,r$ are the simple roots). \end{enumerate} \item With the notation $L^\alpha_0= {1 \over 2n} [L^\alpha_n,L^\alpha_{-n}]$ (this is independent of $n>0$ by (\ref{eq:vir-g2})) for $\alpha \in \Phi_+$: \beqa \label{eq:vir-g3} \big[L_n^\alpha,L_m^\alpha\big]=(n-m)L_{n+m}^\alpha. \eeqa \end{enumerate} \noindent The relations (\ref{eq:vir-g1}), (\ref{eq:vir-g2}) and (\ref{eq:vir-g3}) do not specify all commutators $[L^{\alpha}_n,L^\beta_m]$ ({\it e.g.} $[L^\alpha_1, L^\beta_2]$ if $\alpha \ne \beta$). In order to obtain a Lie algebra from (\ref{eq:vir-g1}), (\ref{eq:vir-g2}) and (\ref{eq:vir-g3}) we define \beqa \label{eq:nivvg} V(g) = T\Big(V\Big)/{\cal I}, \eeqa \noindent where $T\Big(V\Big)$ is the tensorial algebra on $V$ and ${\cal I}$ is the two-sided ideal generated by the relations (\ref{eq:vir-g1}), (\ref{eq:vir-g2}) and (\ref{eq:vir-g3}). $V(g)$ is an associative algebra but we will consider it as a Lie algebra for the induced Lie bracket ({\it i.e.} commutator). The universal property of $V(g)$ is the following: any linear map $f:V \rightarrow {\mathrm{End}}(H)$ such that the $f(L_n^\alpha)$ satisfy the relations (\ref{eq:vir-g1}), (\ref{eq:vir-g2}) and (\ref{eq:vir-g3}) extends to a unique Lie algebra homomorphism $\tilde f:V(g) \rightarrow{\mathrm{End}}(H)$. The relations (\ref{eq:vir-g1}), (\ref{eq:vir-g2}) and (\ref{eq:vir-g3}) can be arranged in the following diagram (for $g=su(3)$) \begin{figure}[!h] \epsfysize =10.cm $$\epsffile{vsu3.eps}$$ \caption{ Digram for the primitive generators for the $V(su(3))$ algebra. $\alpha, \beta$ are the primitive roots and $\gamma$ is the third positive root.} \end{figure} Of course similar diagrams can be constructed for higher rank Lie algebra, and the results can be easily extend to all Lie algebras $g$. This diagram has a concentric and radial structure. The concentric symmetry is just the manifestation that to any positive $n$ can be associated an algebra isomorphic to $g$ (see (\ref{eq:vir-g2})). The radial symmetries extend the $sl(2)_\alpha$ (generated by $(L^\alpha_0, L^\alpha_{\pm 1})$) algebra to ${\mathrm Vir}_\alpha$ a Virasoro algebra (see (\ref{eq:vir-g3})). Finally, composing concentric and radial symmetries through Jacobi identities generates extra symmetries {\it i.e.} the secondary, ternary ... generators. The generators $L_n^{\alpha_i}$ with $n>0$ and $\alpha_i$ a simple root are called the fundamental generators (or primary) of $V(g)$ in the sense that all the others can be obtained from them by taking commutators. The primary generator $L^\alpha_n$ is associated to the root $n \alpha$. Calculating secondary, ternary, ... generators through Jacobi identities induces generators associated to a root $\sum n_i \alpha_i$. For instance if $\alpha + \beta \in \Phi$, the root associated to $[L^\alpha_n,L^\beta_m]$ is $n \alpha + m \beta$. The Lie algebra $V(g)$ is clearly not a Kac-Moody algebra, but according to Kac classification of Lie algebras \cite{k}, $V(g)$ is an infinite dimensional Lie algebra with a non-polynomial degeneracy of the roots. Let us stress again that, in such diagrams, the Cartan sub-algebra is at the origine of the diagram, and the bracket between two generators is defined {\it only} between generators on the same circle, or the same radius. The decomposition \beqa \label{eq:vect} V=\Big<L_0^{\alpha_i}: i=1,\cdots,r\Big> \oplus \Big<L_n^\alpha: \alpha \in \Phi_+, n > 0 \Big> \oplus \Big<L_{-n}^\alpha: \alpha \in \Phi_+, n > 0 \Big> \eeqa \noindent should be thought of as a decomposition of $V$ into a ``Cartan`` sub-algebra and positive and negative ``root'' spaces. From this point of view, one expects to be able to construct representations of $V(g)$ from ``weights'', {\it i.e.} from linear forms on the Cartan sub-algebra $\Big<L_0^{\alpha_i}: i=1,\cdots,r\Big>$. Indeed, if $\mu \in \Big<L_0^{\alpha_i}: i=1,\cdots,r\Big>^*$ (the dual space of $\Big<L_0^{\alpha_i}: i=1,\cdots,r\Big>$) is a weight then ${\cal U}\Big( V(g)\Big)/I_\mu$ is a representation of $V(g)$ where $I$ is the two-sided ideal in ${\cal U} \Big(V(g)\Big)$ generated by the $L^{\alpha_i}_0 - \mu(L^{\alpha_i}_0) {\mathbf 1}~ (i=1,\cdots,r)$ and the $L^\alpha_n ~(n > 0, \alpha \in \Phi_+)$ (see section 3.2). As in section 3.2 \beq \label{eq:F-V(g)} S=(V(g) \oplus {\cal S}^F\left({\cal V}_{\mu/F}\right)_{\mathrm{red}}) \oplus {\cal V}_{\mu/F} \eeq \noindent is an $F-$Lie algebra (with the obvious notation). \subsection{Examples for $so(6)$} The Verma module construction in the previous sub-section gives many examples of $F-$Lie algebras. However, the representations of $V(g)$ obtained are all highest weight representations. In this section, in the spirit of section 4.1, we will construct explicitly non-highest weight representations of $V(so(6))$ and the corresponding $F-$Lie algebras. First of all, we introduce $H_1, H_2, H_3$ the generators of the Cartan sub-algebra. We denote $\pm e^i$ the eigenvalues of $H_i$, and $\Phi=\left\{\alpha_1=e^1-e^2, \alpha_2=e^2-e^3,\right.$ $\left.\alpha_3 = e^2+e^3, \beta_1=e^1+e^2,\right.$ $\left.\beta_2=e^1-e^3, \beta_3=e^1+e^3 \right \}$ the positive roots, where $\alpha_1, \alpha_2$ and $\alpha_3$ are the simple roots. Then we introduce the six $sl(2): \left\{E^{\pm \phi}, {\phi.H, \over \phi^2}, ~ \phi \in \Phi\right \}$. The three $sl(2)$ associated to the primitive roots are then generated by \beqa \label{eq:sl2} \begin{array}{ll} sl(2)_1:&\Big\{ E^{\pm \alpha_1}, {1 \over 2} \left(H_1-H_2\right) \Big \} \cr sl(2)_1:&\Big\{ E^{\pm \alpha_2}, {1 \over 2} \left(H_2-H_3\right) \Big \} \cr sl(2)_3:&\Big\{ E^{\pm \alpha_3}, {1 \over 2} \left(H_2+H_3\right) \Big \}. \end{array} \eeqa Secondly, we need to understand, using the results of section 4.1, how one can extend the representations (\ref{eq:repp}) of $so(1,2)$ to representations (\ref{eq:repv}) of Vir. For that purpose, we introduce commuting variables belonging to the $so(1,2)$ representation: \begin{itemize} \item if the representation is a spin-$s$ representation we define $\big\{x_{-s},\cdots,x_s\big\}={\cal D}_{-s}$; \item for infinite dimensional representation bounded from below/above we consider ${\cal D}_{\lambda}^\pm=\big\{x_p, p \in \pm(\hbox{\it I\hskip -2.pt N} - \lambda)\big\}$. \end{itemize} Next, using the results of section 4.1, we can extend any representations ${\cal D}$ of $so(1,2)$ to a representation $\hat{\cal D}$ of the Virasoro algebra. Explicitly: \begin{itemize} \item For $\hat {\cal D}_{-s}$, one can easily observe that $x_{-ns}$, for any positive $n$ when $s$ integer and for odd positive $n$ when $s$ half-integer, is a primitive vector for the $sl(2)-$ sub-algebra spanned by $\left\{L_{\pm n}/n, L_0/n\right\}$ so ${\cal D}_{-s}^n=\left\{x_{-ns},x_{-(n-1)s},\cdots, x_{ns} \right\}$ is isomorphic to a ${\cal D}_{-s}$ representation; \item in the case of infinite dimensional representations we have similar results if $\lambda$ is a rational number. In the case of interest $\lambda= 1/F$, we have primitive vector only when $n=\pm(pF+1)$. \end{itemize} To conclude with the construction of the representation of Vir, we just have to introduce $p_\ell$ the conjugate momentum of $x_k, [p_{-\ell},x_k] = \delta_{k \ell}$) \begin{itemize} \item For $\hat {\cal D}_{-s}=\big\{x_n, n \in \hbox{$Z$ + s \big\}$ the generators of the Virasoro algebra are $L_n= \sum \limits_{k \in \hbox{$Z$} \big(ns-m) x_{n+m} p_{-m}$; \item and for $\hat {\cal D}^\pm_\lambda=\big\{x_p, p \in \hbox{$Z$ \mp \lambda \big\}$: $L_n= \sum \limits_{k \in \hbox{$Z$ \mp \lambda} \big(n \lambda - k) x_{n+k} p_{-k}$. \end{itemize} In conformal field theory these expressions for the generators of the Virasoro algebra are known \cite{pol}. The $L_n$ are the modes of the stress-energy tensor (for the conformal or superconformal ghosts) associated to conjugate conformal fields of conformal weight $1+ \lambda$ (the $x$'s) and $-\lambda$ (the $p$'s) But it is important to emphasize that in contrast with the case of conformal field theory, here we do not have any normal ordering prescription and consequently no central charge. \subsubsection*{The spin representation of so(6)} The spinorial representation of chirality $+$ is obtained from the dominant weight $\mu_+= \mu_{(3)}={1 \over 2} (e^1+e^2+e^3)$ and the highest weight state is $\left| {1 \over 2},{1 \over 2},{1 \over 2}\right>$. Then by the action of $E^{-\alpha_i}$ we get the whole representation $${\cal D}_{\mu_+} = \left \{ \left| {1 \over 2},{1 \over 2},{1 \over 2} \right>, \left| {1 \over 2},-{1 \over 2},-{1 \over 2} \right>, \left| -{1 \over 2},{1 \over 2},-{1 \over 2} \right>, \left| -{1 \over 2},-{1 \over 2},{1 \over 2} \right> \right \}. $$ \noindent (In $\left|a_1,a_2,a_3 \right>, a_i$ is the eigenvalue of $H_i$). Now, it is easy to see that with respect to the three $sl(2)$ (noted $sl(2)_i$) we have three spinorial representations (for each $sl(2)_i$, we have $\Big({\cal D}_{-1/2}\Big)_i \subset {\cal D}_{\mu_+})$ \beqa \label{eq:spin} \begin{array}{ll} sl(2)_1~:&\Big(D_{-1/2}\Big)_1= \left\{ x^{(1)}_{1/2} \equiv\left| {1 \over 2},-{1 \over 2},-{1 \over 2} \right>, x^{(1)}_{-1/2} \equiv \left| -{1 \over 2},{1 \over 2},-{1 \over 2} \right> \right\}\cr sl(2)_2~:&\Big(D_{-1/2}\Big)_2=\left\{ x^{(2)}_{1/2}\equiv\left| -{1 \over 2},{1 \over 2},-{1 \over 2} \right>, x^{(2)}_{-1/2}\equiv \left| -{1 \over 2},-{1 \over 2},{1 \over 2} \right> \right\}\cr sl(2)_3~:&\Big(D_{-1/2}\Big)_3=\left\{ x^{(3)}_{1/2}\equiv \left| {1 \over 2},{1 \over 2},{1 \over 2} \right>, x^{(3)}_{-1/2}\equiv\left| {1 \over 2},-{1 \over 2},-{1 \over 2} \right> \right\} \end{array} \eeqa \noindent where in $x^{(i)}_m$, $i$ indicates to which $sl(2)$ the states belong and $m$ is the eigenvalue of ${1 \over 2} \alpha_i.H$. Now, having introduced these variables, we can straightforwardly write the generators associated to the three $sl(2)$. For $sl(2)_i$, we have (at this point we already take the normalizations appropriate for the $V(so(6))$ generalization) \beqa \label{eq:spin2} E^{-\alpha_i}=J^{(i)}_-&=&-x^{(i)}_{-1/2} p^{(i)}_{-1/2} \nonumber \\ {1 \over 2} \alpha_i.H=J_0^{(i)}&=&{1 \over 2} \left( x^{(i)}_{-1/2} p^{(i)}_{1/2}- x^{(i)}_{1/2} p^{(i)}_{-1/2} \right) \\ E^{\alpha_i}=J^{(i)}_+&=&x^{(i)}_{1/2} p^{(i)}_{1/2} \nonumber \eeqa \noindent with $p$ the conjugate momentum of $x$ ($[p_{-\ell},x_k]=\delta_{\ell k}$). Of course all the $x$ variables are not independent and the following identifications have to be made \beqa \label{eq:ident1} \begin{array}{ll} x^{(1)}_{-1/2}=x^{(2)}_{1/2}&p^{(1)}_{1/2}=p^{(2)}_{-1/2} \cr x^{(1)}_{1/2}=x^{(3)}_{-1/2}&p^{(1)}_{-1/2}=p^{(3)}_{1/2}, \end{array} \eeqa \noindent leading to dependent $sl(2)$. Now, the next point is to introduce new variables which allow us to define the $\hat {\cal D}_{\mu_+}$ representation of $V(so(6))$. In other words, we need to extend the $J^{(i)}_\pm, J^{(i)}_0$ operators to $L^{(i)}_n, n \in \hbox{$Z$$. This can be done, by considering the three $sl(2)_{2p+1}$ span by $L_{\pm n}^{(i)}/n, L_0^{(i)}/n$ with $n=2p+1, p \in \hbox{\it I\hskip -2.pt N}$. We extend the $\Big({\cal D}_{-1/2}\Big)_i$ rep. of $sl(2)_i$ into a rep. of ${\mathrm Vir}_i$. For that we introduce a highest weight state $\left|\mu_p\right>= \left|{2p+1 \over 2},{2p+1 \over 2}, {2p+1 \over 2} \right>$ (corresponding to a primitive vector of the $sl(2)_{2p+1}$), and construct explicitly the induced spinorial representation for $so(6)$: \beqa \label{eq:spinind} {\cal D}^p_{\mu_+} = &&\left \{ \left| {2p+1\over 2},{2p+1 \over 2},{2p+1 \over 2} \right>, \left| {2p+1 \over 2},-{2p+1 \over 2},-{2p+1 \over 2} \right>, \right. \\ &&\left| -{2p+1 \over 2},{2p+1 \over 2},-{2p+1 \over 2} \right>, \left.\left| -{2p+1 \over 2},-{2p+1 \over 2},{2p+1 \over 2} \right> \right\}. \nonumber \eeqa \noindent As previously, we interpret the various states as spinorial multiplets of the three $sl(2)$, and we define $6$ variables $x^{(i)}_{\pm(2p+1)}$. Now arguing that the $x^{(i)}_r, r \in \hbox{$Z$ + 1/2$ span a $\left(\hat {\cal D}_{-1/2}\right)_{i}$ representation of the Virasoro algebra, we set \beqa \label{eq:viri} L_n^{(i)} = \sum \limits_{r \in \hbox{$Z$ +1/2} ({n \over 2} - r) x^{(i)}_{n+r} p^{(i)}_{-r}. \eeqa \noindent To conclude the construction of the $\hat {\cal D}_{\mu_+}=\oplus_p {\cal D}_{\mu,+}^p$ representation, we make the same identifications as in (\ref{eq:ident1}) \beqa \label{eq:ident2} \begin{array}{ll} x^{(1)}_{-(2p+1)/2}=x^{(2)}_{(2p+1)/2}& p^{(1)}_{(2p+1)/2}=p^{(2)}_{-(2p+1)/2} \cr x^{(1)}_{(2p+1)/2}=x^{(3)}_{-(2p+1)/2}& p^{(1)}_{-(2p+1)/2}=p^{(3)}_{(2p+1)/2} \end{array} \eeqa This specified the $\hat {\cal D}_{\mu_+}$ representation. The remaining commutators of $V(so(6))$ can then be calculated explicitly. \subsubsection*{The Vector representation} For the vector representation the dominant weight is $\mu_v=\mu_{(1)}=e^1$. To construct $\hat {\cal D}_{\mu_v}$ from ${\cal} D_{\mu_v}$, we proceed along the same lines as for the spin representation, so we will be less explicit in our construction. Constructing ${\cal D}_{\mu_v}$, we have for each $sl(2)_i$ the following decomposition ${\cal D}_{\mu_v} \supset \Big({\cal D}_{-1/2}\Big) \oplus \Big({\cal D}^\prime_{-1/2}\Big)$. We define highest weights associated to the three $sl(2)$ generated by $(L_{\pm}^{(i)}/n, L_0^{(i))}/n)$ with $n=2p+1$. This induce a vector representation ${\cal D}^p_{\mu_v}$: \beqa \label{eq:revvect} {\cal D}^p_{\mu_v} =\left\{\left|\pm(2p+1),0,0\right>, \left|0,\pm(2p+1),0 \right>, \left|0,0\pm(2p+1)\right> \right \}. \eeqa \noindent We then identify six spinorial representations of the primitive $sl(2)$: \beqa \label{eq:spinv} \begin{array}{ll} sl(2)_1~:&\Big({\cal D}^p_{-1/2}\Big)_1= \left\{x^{(1)}_{(2p+1)/2}\equiv \left|2p+1,0,0\right>, x^{(1)}_{-(2p+1)/2}\equiv \left|0,2p+1,0\right> \right\}\cr &\Big({\cal D}^{\prime p}_{-1/2}\Big)_1= \left\{ x^{\prime~(1)}_{(2p+1)/2}\equiv \left|0,-(2p+1),0\right>, x^{\prime~(1)}_{-(2p+1)/2}\equiv \left|-(2p+1),0,0\right> \right\} \cr sl(2)_2~:&\Big({\cal D}^p_{-1/2}\Big)_2= \left\{x^{(2)}_{(2p+1)/2}\equiv \left|0,2p+1,0\right>, x^{(2)}_{-(2p+1)/2} \equiv \left|0,0,2p+1\right> \right\} \cr &\Big({\cal D}^{\prime p}_{-1/2}\Big)_2= \left\{x^{\prime~(2)}_{(2p+1)/2}\equiv \left|0,0,-(2p+1)\right>, x^{\prime~(2)}_{-(2p+1)/2}\equiv \left|0,-(2p+1),0\right>\right\} \cr sl(2)_3~:&\Big({\cal D}^p_{-1/2}\Big)_3= \left\{x^{(3)}_{(2p+1)/2}\equiv \left|0,2p+1,0\right>, x^{(3)}_{-(2p+1)/2}\equiv \left|0,0,-(2p+1)\right>\right\} \cr &\Big({\cal D}^{\prime p}_{-1/2}\Big)_3= \left\{x^{\prime~(3)}_{(2p+1)/2}\equiv \left|0,0,(2p+1)\right>, x^{\prime~(3)}_{-(2p+1)/2}\equiv \left|0,-(2p+1),0\right> \right\} \end{array} \eeqa \noindent The appropriate identifications can be read off from (\ref{eq:spinv}) \beqa \label{eq:identv} \begin{array}{ll} x^{(1)}_{-(2p+1)/2}=x^{(2)}_{(2p+1)/2}=x^{(3)}_{(2p+1)/2}& p^{(1)}_{(2p+1)/2}=p^{(2)}_{-(2p+1)/2}=p^{(3)}_{-(2p+1)/2} \cr x^{\prime~(1)}_{(2p+1)/2}=x^{\prime~(2)}_{-(2p+1)/2}=x^{\prime~(3)}_{-(2p+1)/2} & p^{\prime~(1)}_{-(2p+1)/2}=p^{\prime~(2)}_{(2p+1)/2}=p^{\prime~(3)}_{(2p+1)/2} \cr x^{(2)}_{-(2p+1)/2}= x^{\prime (3)}_{(2p+1)/2}& p^{(2)}_{(2p+1)/2}= p^{\prime (3)}_{-(2p+1)/2} \cr x^{\prime (2)}_{(2p+1)/2}=x^{(3)}_{-(2p+1)/2}& p^{\prime (2)}_{-(2p+1)/2}=p^{(3)}_{(2p+1)/2} \end{array} \eeqa \noindent So, finally we obtain the explicit expression for the primitive (associate to the simple roots of $so(6)$) generators of $V(so(6))$ for the $\hat {\cal D}_{\mu_v} = \oplus_p {\cal D}_{\mu_v}^p$ representation. \beqa \label{eq:virvec} L_n^{(i)}= \sum \limits_{r \in \hbox{$Z$ +1/2} (n- {r \over 2}) \left(x^{(i)}_{n+r} p^{(i)}_{-r} + x^{\prime~(i)}_{n+r} p^{\prime~(i)}_{-r} \right). \eeqa \subsubsection*{The ${\cal D}^+_{{\mu_+ \over F}}$ representation of $so(6)$} In this example, we show how one can obtain an explicit realization of the $\hat {\cal D}^+_{{\mu_+ \over F}}$ representation without giving any differential realization. Starting with the highest weight state $\left| {\mu_+ \over F} \right>= \left|{1 \over 2F},{1 \over 2F},{1 \over 2F}\right>$ we construct the ${\cal D}^+_{{\mu_+ \over F}}$ representation. ($\hat {\cal D}^-_{{\mu_+ \over F}}$ would have been obtained from $\left|-{1 \over 2F},-{1 \over 2F},{1 \over 2F}\right>$). The interesting point with such a representation is that the states of the third $sl(2)$ belong to ${\cal D}_{1/F}^+$ and of the first and second $sl(2)$ to finite dimensional representations. Now, to obtain the whole representation $\hat {\cal D}_{{\mu_+ \over F}}$ it is enough to find the primitive vector associated to $(L^{(i)}_{\pm (2Fp+1)}/(2Fp+1),L^{(i)}_0/(2Fp+1))$ (there are no other states annihilated by some $L_n$, see (\ref{eq:repv}) and (\ref{eq:viract})). This vector, $\left|{1+2Fp \over 2F},{1+2Fp \over 2F},{1+2Fp \over 2F}\right>$ induces a ${\cal D}^+_{{\mu_+ \over F}}$ representation. With the analogous identifications as for the vectorial and spinorial representations we get $\hat {\cal D}_{{\mu_+ \over F}}$. \subsubsection*{Application to $F-$Lie algebras with $so(6)$} Having constructed the representations of $V(so(6))$, one can easily prove that $\hat {\cal D}^+_{\mu_+}$ is included in ${\cal S}^F\left(\hat {\cal D}^+_{\mu_+ \over F} \right)_{\mathrm red}$ (${\cal S}^F\left(\hat {\cal D}^+_{\mu_+ \over F} \right)_{\mathrm red}$ is defined as in section 3.2). We just have to notice that $$\hat \mu_F \left( {\cal S}^F \left( \left|p_1+{1 \over 2F},p_1+{1 \over 2F},p_1+{1 \over 2F} \right>, \cdots, \left|p_F+{1 \over 2F},p_F+{1 \over 2F},p_F+{1 \over 2F} \right> \right) \right)$$ \noindent with $\sum p_i = p$ is a primitive vector which induces the spinorial representations (\ref{eq:spinind}) obtained with $\left| \mu_p \right>$ in ${\cal D}_{\mu_+}$. This leads to the $F-$Lie algebra $\Big(V(so(6)) \oplus ({\cal D}^+_{{\mu_+ \over F}})_{\mathrm red} \Big) \oplus \hat {\cal D}^+_{{\mu_+ \over F}}$ \\ \vskip.5truecm To conclude, the results of this sub-section probably extend, along the same lines, to any Lie algebra $g$. \mysection{Conclusion} Supersymmetry is a well established mathematical structure ($\hbox{$Z$_2$ graded algebras) and beyond its purely formal aspects it has found a wide range of applications in field theories and particle physics. In this article, we have defined a mathematical structure generalizing the concept of super-Lie algebras ($F-$Lie algebras). These algebras seems to be appropriate if one wants to generalize supersymmetry in the sense of $F^{{\mathrm th}}-$roots of representations. Indeed, we have shown that, within the framework of $F-$Lie algebras, it is possible to take the $F^{{\mathrm th}}-$root of {\it any representation of any (complex semi-simple) Lie algebra}. In addition, $F-$Lie algebras naturally lead to the infinite dimensional Lie algebra $V(g)$ containing $g$ as a sub-algebra. For the special case $g=sl(2,\hbox{\it I\hskip -2.pt R })$, $V(g)$ is the centerless Virasoro algebra. As a consequence, one may wonder whether or not central extensions of $V(g)$ exist. Furthermore, one has a geometrical interpretation of the Virasoro algebra (as vector fields on the circle) so is there a similar interpretation for $V(g)$ ? Unitary representations of FSUSY, for $g=so(1,2)$ have also been constructed. It has also been checked that it is a symmetry acting on relativistic anyons \cite{fsusy3d}. In the same way, since the Lorentz group in higher dimensions is just $SO(1,d-1)$, what is the interpretation of FSUSY for $g=so(1,d-1)$ when the $F-$Lie algebra induces the $F-$root of the spin or the vector representations ? \vskip1truecm {\bf Acknowledgements:} We are grateful to A. Neveu and J. Thierry-Mieg for useful discussions and critical comments. \vskip .3 in \baselineskip=1.6pt
\section{Introduction} During the last years there has been a considerable resurgance of mathematical studies of topics related to the black hole equilibrium states. Not only restricted to the pure vacuum Einstein or Einstein-Maxwell theory but also including nonlinear matter models, general sigma models or fields occuring in the low energy limits of the superstring theories. In his pioneering investigations \cite{iso,is1} Israel established the uniqueness of the Schwarzschild metric and its Reissner-Nordstr\"om generalization as static asymptotically flat solutions of the Einstein and Einstein-Maxwell vacuum field equations subject to the condition of regularity outside a well behaved {\it ergosurface} where the static Killing vector becomes null. In view of the subsequent demonstration \cite{caj,ca0} that in the static case (though not more generally) the {\it ergosurface} will be an event horizon, it follows that the Schwarzschild and Reissner-Nordstr\"om solutions are the only Einstein or Einstein-Maxwell (non-extreme) solutions that satisfy the condition of being static black hole metrics in the strict modern sense of the term. \par Israel's article \cite{iso} was the foundation of the next works of M\"uller zum Hagen {\it et al.} \cite{mi} and Robinson \cite{ro} establishing the generalization of Israel's theorem of the uniqueness of the Schwarzschild black hole solution. \par The uniqueness results for rotating configurations, i.e., for stationary and axisymmetric black hole spacetimes were achieved by Carter in \cite{cak} and completed by Hawking and Ellis \cite{haw} and the next works of Carter \cite{ca0,ca}, Robinson \cite{rok} and Wald \cite{wak}. These works were devoted to the vacuum case. Bunting \cite{bun} and Mazur \cite{ma} generalize these results to the electromagnetic case. Bunting's approach was based on using a general class of harmonic mappinng between Riemannian manifolds. In Mazur's proof the key point was the observation that the Ernst Eqs. described a nonlinear sigma model on a symmetric space. The review of the new methods presented by Bunting and Mazur was discussed in \cite{car}. \par Bunting and Masood-ul-Alam \cite{bu} used the positive mass theorem \cite{ya,wi} to prove the uniqueness, the spherical symmetry or the nonexistence for several static black hole solutions of the Einstein's Eqs. In Ref.\cite{bu} the afore mentioned technique was explored to prove the nonexistence of multiple black holes in the asymptotically Euclidean static vacuum spacetime. They found the conformal transformation which caused that the mass of the spatial part of the metric was equal to zero, but the scalar curvature tensor was non-negative. Ruback \cite{ru} applied this technique to prove the uniqueness theorem for charged black holes in static Einstein-Maxwell spacetime. Masood-ul-Alam \cite{ma1} gave an alternative, much simpler, a rigorous proof that the unique non-degenerate electrovac static black hole metrics are the Reissner-Nordstr\"om family. It was done without assuming the connectedness of the event horizon. The further generalization of the uniqueness proof for the static electrovac black holes including the case of a non-vanishing magnetic charge was proposed by Heusler \cite{he1}. He used a generalization of the conformal factor and established the nonexistence of multiple black hole solutions of Einstein-Maxwell system with electric and magnetic fields in a static, asymptotically flat spacetime. Heusler \cite{he2} demonstrated also the uniqueness of multiple black hole solutions for any self-coupled, stationary scalar mapping (sigma-model) with nonrotating horizon. \par Using the fact that Einstein-Abelian gauge field Eqs. can be formulated as a sigma model on the adequate K\"ahler manifold, G\"urses \cite{gu} found that $(n-1)$ Abelian gauge charged Kerr black hole is a unique stationary black hole solution of Einstein-Abelian gauge field Eqs. In Ref.\cite{gu1}, he also proved that, the boundary value problems of some sigma models in a non-Riemannian background have unique solutions. \par Quite recently a uniqueness theorem was extended to the case of the Ernst solution and C-metric \cite{we}. \par Uniqueness theorems for black holes are closely related to the problem of staticity. Lichnerowicz \cite{li} was the first who considered the idea of staticity for a stationary perfect fluid, locally static in the sence that its flow vector was connected with the Killing vector. The next extensions was atributed to Hawking \cite{haw} (vacuum case) and the Hawking's {\it strong rigidity theorem} \cite{hel} emphasized that the event horizon of a stationary black hole had to be a Killing horizon. Recently, Sudarsky and Wald \cite{wa}, by means of the notion of an asymptotically flat maximal slice with a compact interior, obtained the conditions that the solution of Einstein-Yang-Mills Eqs. is static when it had a vanishing Yang-Mills electric field on static hypersurfaces. They also reached to the conclusion that nonrotating Einstein-Maxwell black hole must be static when it has a vanishing magnetic field on static slices \cite{wa1}. \par For a recent review concerning various aspects of uniqueness theorems for nonrotating and rotating black holes see \cite{book}, while the mathematical rigor of the afore mentioned problems has been studied in the review articles provided by Chru\'sciel \cite{chr1,chr2}. \medskip \noindent Recently, there has been an active period for constructing black hole solutions in the string theory which seems the most promising for a theory of quantum gravity (see \cite{hu} for a recent review of the subject). \par The uniqueness of static, charged dilaton black hole solutions in the low energy string theory was certified by Masood-ul-Alam \cite{ma2}. He found a conformal spatial metric which had the sufficient properties for existing suitable Dirac spinors. By means of the Lichnerowicz Eq. it was shown that these spinors are constant. \par The alternative proof of a uniqueness of a static charged dilaton black hole was provided by G\"urses and Sermutlu \cite{gs}. They used a sigma model formulation of equations of motion. The problem of black hole solutions and their uniqueness in axion-dilaton gravity was studied by Bowick {\it et al.} \cite{bow}. They managed to find uniqueness theorem in the case of the minimal coupling of axion field to gravity. Cambell {\it et al.} \cite{cam} showed the existence of axion {\it hair} for a Kerr black hole and calculated it explicitily in the case of a slow motion. They considered axion fields with a Lorentz Chern-Simons coupling to gravity. A dilaton coupling to axion fields strengths were considered in \cite{mig}, where the authors calculated dilaton {\it hair} arising from the specific axion source. \par Wells \cite{we1} wrote down the analogue of the Robinson's identity for dilatonic black holes which allowed him to prove uniqueness theorem for a class of accelerating stringy black holes. \par The problem of staticity theorems in Einstein-Maxwell axion-dilaton gravity was studied by Rogatko in Ref.\cite{r1}, where the modified Carter arguments were used to find staticity conditions for fields and the metric. It was found, in Ref.\cite{r2}, that static black hole solutions in the above theory had vanishing electric and axion-electric fields on static slices. \par Our paper is organized as follows. In Sec.II we present the equations of motion for the bosonic sector of $N = 4, d= 4$ supergravity in a static axially symmetric spacetime. Introducing the adequate forms of the pseudopotentials and complex scalars enables us to write Eqs. of motion as two complex Ernst-like Eqs. Then, using the matrix formulation of Ernst Eqs., conceived by G\"urses and Xanthopoulos \cite{gx}, we reached the conclusion that two metrics satisfying the Eqs. of motion and having the same boundary conditions must be equal to each other in all points of the region of the two-dimensional manifold. Which implies in turns, that all black hole solutions in the theory under consideration subject to the same boundary conditions are the same everywhere in the spacetime. We considered the $SU(4)$ and $SO(4)$ versions of the underlying theory. We conclude in Sec.III with a brief summary of our researches and their implications. \section{Uniqueness of Black Hole Solutions} \subsection{Doubly Charged Black Holes in $SU(4)$ version of $N = 4, d = 4$ supergravity} Superstring theories provide interesting generalizations of the Einstein-Maxwell theory in the so-called low energy limit. A dimensionally reduced superstring theory can be described in terms of $N=4, d = 4$ supergravity. It turned out, that one can refer to the $SO(4)$ version \cite{so} or $SU(4)$ one \cite{su}. In our paper, we shall consider bosonic sectors of these theories, taking into account two $U(1)$ gauge fields and a dilaton field $\phi$, called $U(1)^2$ theories. We begin with the $SU(4)$ version of $N = 4, d = 4$ supergravity, which the action is of the form \cite{so,kal} \begin{equation} I_{SU(4)} = \int d^4 x \sqrt{-g} \left [ R - 2(\nabla \phi)^{2} - e^{-2\phi} \left ( F_{\alpha \beta} F^{\alpha \beta} + G_{\gamma \delta} G^{\gamma \delta} \right ) \right ], \label{act} \end{equation} where the strengths of the gauge fields are descibed by $F_{\mu \nu} = 2\nabla_{[\mu} A_{\nu]}$ and $G_{\alpha \beta} = 2\nabla_{[\alpha} B_{\beta] }$. The resulting equations of motion, derived from the variational principle, are as follows: \begin{eqnarray R_{\mu \nu} = e^{-2 \phi} \left ( 2 F_{\mu \rho} F_{\nu}{}{}^{\rho} - {1 \over 2} g_{\mu \nu}F^2 \right ) + e^{-2 \phi} \left ( 2 G_{\mu \rho} G_{\nu}{}{}^{\rho} - {1 \over 2} g_{\mu \nu}G^2 \right ) + 2 \nabla_{\mu} \phi \nabla_{\nu} \phi, \\ \nabla_{\mu} \nabla^{\mu} \phi + {1 \over 2} e^{-2\phi} F^2 + {1 \over 2} e^{-2\phi} G^2 = 0, \\ \nabla_{\mu} \left ( e^{-2 \phi} F^{\mu \nu} \right ) = 0, \\ \nabla_{\mu} \left ( e^{-2 \phi} G^{\mu \nu} \right ) = 0. \end{eqnarray} The black hole solutions in the theory under consideration were widely discussed by Kallosh {\it et al.} in Ref.\cite{kal}. \par Our main task will be to prove the uniqueness of the obtained results. We want to provide some continuity with the researches of G\"urses \cite{gu1,gs}, in some respects to generalize them, we shall present our analysis of the problem in a form and notation similar to theirs. First, we shall formulate the corresponding Eqs. of motion as a two-dimensional sigma model and prove the uniqueness of the static solution under the same boundary conditions. In order to do so we introduce the static axially symmetric line element expressed as \begin{equation} ds^2 = - e^{2 \psi} dt^2 + e^{-2 \psi} \left [ e^{2 \gamma} \left ( dr^2 + dz^2 \right ) + r^2 d\phi^2 \right ], \label{gij} \end{equation} where $\psi $ and $\gamma$ depended only on $r$ and $z$ coordinates. The components of the $U(1)$ gauge strength tensors will be in the direction of time $F_{\mu \nu} = (A_{0}, 0, 0, 0)$ and in the azimuthal angle $G_{\alpha \beta} = (0, 0, 0, B_{\phi})$. In our further considerations we assume that the components of the $U(1)$ gauge fields are functions of $r$ and $ z$. Then, the resulting Eqs. of motion are as follows: \begin{equation} \nabla^2 \phi - e^{-2\psi - 2\phi} \left ( A_{0, r}^2 + A_{0, z}^2 \right ) + {e^{2\psi - 2 \phi} \over r^2} \left ( B_{\phi, r}^2 + B_{\phi, z}^2 \right ) = 0, \label{em1} \end{equation} \begin{equation} \nabla^2 \psi - e^{-2\psi - 2\phi} \left ( A_{0, r}^2 + A_{0, z}^2 \right ) - {e^{2\psi - 2 \phi} \over r^2} \left ( B_{\phi, r}^2 + B_{\phi, z}^2 \right ) = 0, \label{em2} \end{equation} \begin{equation} \nabla^2 A_{0} - 2 \left ( \psi_{, r} + \phi_{, r} \right ) A_{0, r} - 2 \left ( \psi_{, z} + \phi_{, z} \right ) A_{0, z} = 0, \label{em3} \end{equation} \begin{equation} \triangle B_{\phi} + 2 \left ( \psi_{, r} - \phi_{, r} \right ) B_{\phi, r} + 2 \left ( \psi_{, z} - \phi_{, z} \right ) B_{\phi, z} = 0, \label{em4} \end{equation} \begin{equation} e^{-2\psi - 2\phi} \left ( A_{0, r}^2 - A_{0, z}^2 \right ) + {1 \over r^2} e^{2\psi - 2 \phi} \left ( B_{\phi, r}^2 - B_{\phi, z}^2 \right ) + \left ( \phi_{,z}^2 - \phi_{, r}^2 \right ) = \psi_{, r}^2 - \psi_{, z}^2 - {\gamma_{, r} \over r}, \label{em5} \end{equation} \begin{equation} {\gamma_{, z} \over r} - 2 \psi_{, r} \psi_{, z} = - 2 e^{-2\psi - 2\phi} A_{0, r} A_{0, z} + {2 \over r^2} e^{2\psi - 2 \phi} B_{\phi, r} B_{\phi, z} + 2 \phi_{, r} \phi_{, z}, \label{em6} \end{equation} where $\nabla^2$ is the Laplacian operator in the $(r, z)$ coordinates, namely, $\nabla^2 = \partial_{r}^2 + \partial_{z}^2 + {1 \over r} \partial_{r}$ and $\triangle = \partial_{r}^2 + \partial_{z}^2 - {1 \over r}\partial_{r}$.\\ From Eq.(\ref{em5}) or (\ref{em6}) one can determine the function $\gamma$ if $\psi, \phi, A_{0}, B_{\phi}$ are known. Hence, the essential part of the Eqs. of motion consists of Eqs.(\ref{em1}-\ref{em4}). \par Now, let us define the quantities \begin{equation} E = - \phi - \psi, \qquad M = \psi - \phi, \end{equation} and \begin{equation} {\tilde A}_{0} = {A_{0} \over \sqrt{2}}, \qquad {\tilde B}_{\phi} = {i B_{\phi} \over \sqrt{2}}. \end{equation} Consistently with the above definitions, Eqs.(\ref{em1}-\ref{em4}) can be rewritten in the forms \begin{equation} \nabla^2 E + e^{2 E} \nabla {\tilde A}_{0} \nabla {\tilde A}_{0} = 0, \label{a0} \end{equation} \begin{equation} \nabla^2 {\tilde A}_{0} + 2 \nabla E \nabla {\tilde A}_{0} = 0, \label{a1} \end{equation} \begin{equation} \nabla^2 M + e^{2 M} \nabla {\tilde B}_{\phi} \nabla {\tilde B}_{\phi} = 0, \label{a2} \end{equation} \begin{equation} \triangle {\tilde B}_{\phi} + 2 \nabla M \nabla {\tilde B}_{\phi} = 0. \label{a3} \end{equation} Thus, we obtain two pairs of Eqs., one described in terms of ${\tilde A}_{0}, E$ and the other for ${\tilde B}_{\phi}$ and $M$.\\ We observe that Eq.(\ref{a1}) allows us to define a pseudopotential $\omega_{(A)}$, given by \begin{equation} \omega_{(A) r} = r e^{2 E} {\tilde A}_{0, z}, \qquad \omega_{(A) z} = - r e^{2 E} {\tilde A}_{0, r}, \label{wa} \end{equation} while from Eq.(\ref{a3}), one can establish the following $\omega_{(B)}$ pseudopotential: \begin{equation} \omega_{(B) r} = { - e^{2 M} {\tilde B}_{\phi, z} \over r}, \qquad \omega_{(B) z} = {e^{2 M} {\tilde B}_{\phi, r} \over r}. \label{wb} \end{equation} Then, we want to rewrite the Eqs.(\ref{a0}-\ref{a3}) in the forms similar to the Ernst ones. In order to do this, we introduce two complex scalars, determined by \begin{eqnarray \label{e1} \epsilon_{1} &=& r e^E + i \omega_{(A)}, \\ \epsilon_{2} &=& e^{M} + i \omega_{(B)}. \end{eqnarray} Now, Eqs.(\ref{a0}-\ref{a1}) and (\ref{a2}-\ref{a3}) can be arranged into the following two complex Eqs.: \begin{eqnarray \label{ee1} \left ( \bar \epsilon_{1} + \epsilon_{1} \right ) \nabla^2 \epsilon_{1} = 2 \nabla \epsilon_{1} \nabla \epsilon_{1}, \\ \left ( \bar \epsilon_{2} + \epsilon_{2} \right ) \nabla^2 \epsilon_{2} = 2 \nabla \epsilon_{2} \nabla \epsilon_{2}, \label{ee2} \end{eqnarray} where a bar denotes complex conjugation.\\ Eqs.(\ref{ee1}) and (\ref{ee2}) are two Ernst Eqs., which each of them combine in a convenient and a symmetric fashion the two Eqs. governing $\phi, \psi$ and the adequate gauge field. Each of them defined a sigma model on $SU(2)/U(1)$.\\ It was shown in Ref.\cite{gx}, that the various combinations of Ernst's Eqs. were included in the single matrix Eq. In our case, the matrix Eq. is determined by \begin{equation} \partial_{r} \left [ P_{(i)}^{-1} \partial_{r} P_{(i)} \right ] + \partial_{z} \left [ P_{(i)}^{-1} \partial_{z} P_{(i)} \right ] = 0, \label{ma} \end{equation} where the subscript $(i)$ in $P$ matrix refers respectively to ${\tilde A}, {\tilde B}$ gauge fields. The explicit form of the matrices are given by \begin{equation} P_{(A)} = {1 \over r e^E} \pmatrix{1 & \omega_{(A)} \cr \omega_{(A)} & r^2 e^{2 E} + \omega_{(A)}^2 \cr}, \qquad P_{(B)} = {1 \over e^M} \pmatrix{1 & \omega_{(B)} \cr \omega_{(B)} & e^{2 M} + \omega_{(B)}^2 \cr}. \end{equation} One can check that, the above matrix Eqs. when written out explicitly in terms of its elements constitute four Eqs., all of them are various combinations of Eqs. (\ref{ee1}-\ref{ee2}) and Eqs. the complex conjugate of them. In order to prove a uniqueness theorem we shall follow the line described by G\"urses \cite{gu1}. To begin with, one should assume enough differentiability for the matrices components in a region ${\cal D}$ of the two-dimensional manifold ${\cal M}$ with boundary $\partial {\cal D}$. Let $P_{(i) 1}$ and $P_{(i) 2}$ will be two different solutions of Eqs. (\ref{ma}) respectively for the cases of ${\tilde A}$ and ${\tilde B}$ gauge fields, than the difference of their Eqs. will have be given by \begin{equation} \nabla \left ( P_{(i) 1}^{-1} \left ( \nabla Q_{(i)} \right ) P_{(i) 2} \right ) = 0, \label{dif} \end{equation} where $Q_{(i)} = P_{(i) 1} P_{(i) 2}^{-1}$. Multiplying Eq.(\ref{dif}) by $Q_{(i)}^{\dagger}$, one arives at the expression \begin{equation} \nabla^2 q_{(i)} = Tr \left [ \left ( \nabla Q_{(i)}^{\dagger} \right ) P_{(i) 1}^{-1} \left ( \nabla Q_{(i)} \right ) P_{(i) 2} \right ], \label{g} \end{equation} where $q = Tr Q$. Taking into account hermicity and positive definiteness of the matrices $P_{(i) 1}$ and $P_{(i) 2}$, we can postulate the form of the above matrices, satisfying \begin{equation} P_{(A) \alpha} = A_{\alpha} A_{\alpha}^{\dagger}, \qquad P_{(B) \alpha} = B_{\alpha} B_{\alpha}^{\dagger}, \label{pa} \end{equation} where $\alpha = 1, 2$. The explicit forms of the matrices $A_{\alpha}$ and $B_{\alpha}$ can be established as follows: \begin{equation} A_{\alpha} = {1 \over \sqrt{r} e^{E_{\alpha}/2}} \pmatrix{ 1 & 0 \cr \omega_{(A) \alpha} & r e^{E_{\alpha}} \cr}, \qquad B_{\alpha} = {1 \over e^{M_{\alpha}/2}} \pmatrix{ 1 & 0 \cr \omega_{(B) \alpha} & e^{M_{\alpha}} \cr}. \label{aa} \end{equation} Using relation (\ref{pa}) one can rewrite Eq.(\ref{g}) in the form \begin{equation} \nabla^2 q_{(i)} = Tr \left ( {\cal J}_{(i)}^{\dagger} {\cal J}_{(i)} \right ), \label{jj} \end{equation} where $ {\cal J}_{(A)} = A_{1}^{-1} (\nabla Q_{(A)}) A_{2}$ and ${\cal J}_{(B)} = B_{1}^{-1} (\nabla Q_{(B)}) B_{2}$. Thus, the explicit versions of $q_{(i)}$ are given by \begin{eqnarray q_{(A)} &=& 2 + {1 \over r^2 e^{E_{1} + E_{2}}} \left [ \left ( \omega_{(A)1} - \omega_{(A)2} \right )^2 + r^2 \left ( e^{E_{1}} - e^{E_{2}} \right )^2 \right ], \\ q_{(B)} &=& 2 + {1 \over e^{M_{1} + M_{2}}} \left [ \left ( \omega_{(B)1} - \omega_{(B)2} \right )^2 + \left ( e^{M_{1}} - e^{M_{2}} \right )^2 \right ]. \end{eqnarray} It is evident that on the boundary $q_{(A)} = q_{(B)} = 2 $ and their first derivatives disappear there. Taking into account the boundary conditions on $\partial {\cal D}$, one can integrate Eq.(\ref{jj}) to obtain the relation \begin{equation} \int_{\partial {\cal D}} Tr \left ( {\cal J}_{(i)}^{\dagger} {\cal J}_{(i)} \right ) = 0. \label{jj1} \end{equation} The expression (\ref{jj1}) implies vanishing of ${\cal J}_{(A)}$ and ${\cal J}_{(B)}$, which in turns causes that $Q_{(i)} = const$ in all region ${\cal D}$. Because of this fact, $Q_{(i)} = I$ matrix on $\partial {\cal D}$, then $Q_{(i)}= I$ in ${\cal D}$. Therefore we reach to the conclusion that, $P_{(i) 1} = P_{(i) 2}$ at all points of the region ${\cal D}$ of the two-dimensional manifold ${\cal M}$. As was mentioned in Ref.\cite{gs}, the other way of reaching these conclusions is to observe that, vanishing of (\ref{jj1}) implies the harmonicity of $q$ in ${\cal D}$ region. Since $q_{(i)} = 2$ on the boundary $\partial {\cal D}$, then it must be equal to the same constant value in the region ${\cal D}$. Thus, $P_{(i) 1} = P_{(i) 2}$ in ${\cal D}$. \vspace{0.2cm} \noindent The above considerations enables us to formulate the main result, the following. \noindent {\it Theorem:} Consider a two-dimensional manifold ${\cal M}$ equipped with a local coordinates $(r, z)$. Suppose that, ${\cal D}$ is a region in ${\cal M}$ with boundary $\partial {\cal D}$. Let $P_{(i)}$ be hermitian, positive definite two-dimensional matrices with unit determinants, respectively for $i = {\tilde A}, {\tilde B} $ gauge fields. Suppose further that, matrices $P_{(i) 1}$ and $P_{(i)2}$ satisfy Eq.(\ref{dif}), namely \begin{equation} \nabla \left ( P_{(i) 1}^{-1} \left ( \nabla Q_{(i)} \right ) P_{(i) 2} \right ) = 0, \end{equation} in the region ${\cal D}$ and have the same boundary conditions on $\partial {\cal D}$. Then, in all points of the region ${\cal D}$, one has that $P_{(i) 1} = P_{(i) 2}$. \par The doubly charged dilaton black holes in $SU(4)$ version of $N = 4, d = 4$ supergravity are characterized by mass $M$, the $F$-field electric charge, the $G$-field magnetic charge and the dilaton charge $\Sigma$. The above theorem envisages that all black holes subject to the same boundary conditions, as the solution obtained by Kallosh {\it et al.} in Ref.\cite{kal}, are the same everywhere in the spacetime. \subsection{Doubly Charged Black Holes in $SO(4)$ version of $N = 4, d = 4$ supergravity} The bosonic part of the $SO(4)$ version of $N = 4$ supergravity in four dimensions can be described by the action \cite{so,kal} \begin{equation} I_{SO(4)} = \int d^4 x \sqrt{-g} \left [ R - 2(\nabla \phi)^{2} - \left ( e^{-2\phi} F_{\alpha \beta} F^{\alpha \beta} + e^{2 \phi} {\tilde G}_{\gamma \delta} {\tilde G}^{\gamma \delta} \right ) \right ],\label{so4} \end{equation} where the strengthes of the gauge fields are descibed by $F_{\mu \nu} = 2\nabla_{[\mu} A_{\nu]}$ and ${\tilde G}_{\alpha \beta} = 2\nabla_{[\alpha} G_{\beta] }$. The equations derived from the variational principle are as follows: \begin{eqnarray R_{\mu \nu} = e^{-2 \phi} \left ( 2 F_{\mu \rho} F_{\nu}{}{}^{\rho} - {1 \over 2} g_{\mu \nu}F^2 \right ) + e^{2 \phi} \left ( 2 {\tilde G}_{\mu \rho} {\tilde G}_{\nu}{}{}^{\rho} - {1 \over 2} g_{\mu \nu} {\tilde G}^2 \right ) + 2 \nabla_{\mu} \phi \nabla_{\nu} \phi, \\ \nabla_{\mu} \nabla^{\mu} \phi + {1 \over 2} e^{-2\phi} F^2 - {1 \over 2} e^{2\phi} {\tilde G}^2 = 0, \\ \nabla_{\mu} \left ( e^{-2 \phi} F^{\mu \nu} \right ) = 0, \\ \nabla_{\mu} \left ( e^{2 \phi} {\tilde G}^{\mu \nu} \right ) = 0. \end{eqnarray} The components of the $U(1)$ gauge fields are in the time direction \cite{kal}, namely $F_{\mu \nu} = (A_{0}, 0, 0, 0)$ and ${\tilde G}_{\mu \nu} = (G_{0}, 0 , 0, 0)$. As in the proceding paragraph the components of the $U(1)$ gauge fields are functions depending only on $(r, z)$ coordinates. The Eqs. of motion in metric (\ref{gij}) satisfy \begin{equation} \nabla^2 \phi - e^{-2\psi - 2\phi} \left ( A_{0, r}^2 + A_{0, z}^2 \right ) + {e^{- 2\psi + 2 \phi} \over r^2} \left ( G_{0, r}^2 + G_{0, z}^2 \right ) = 0, \end{equation} \begin{equation} \nabla^2 \psi - e^{-2\psi - 2\phi} \left ( A_{0, r}^2 + A_{0, z}^2 \right ) - e^{- 2\psi + 2 \phi} \left ( G_{0, r}^2 + G_{0, z}^2 \right ) = 0, \end{equation} \begin{equation} \nabla^2 A_{0} - 2 \left ( \psi_{, r} + \phi_{, r} \right ) A_{0, r} - 2 \left ( \psi_{, z} + \phi_{, z} \right ) A_{0, z} = 0, \end{equation} \begin{equation} \nabla^2 G_{0} + 2 \left (- \psi_{, r} + \phi_{, r} \right ) G_{0, r} + 2 \left ( - \psi_{, z} + \phi_{, z} \right ) G_{0, z} = 0, \end{equation} \begin{equation} e^{-2\psi - 2\phi} \left ( A_{0, r}^2 - A_{0, z}^2 \right ) + e^{- 2\psi + 2 \phi} \left ( G_{0, r}^2 - G_{0, z}^2 \right ) + \left ( \phi_{,z}^2 - \phi_{, r}^2 \right ) = \psi_{, r}^2 - \psi_{, z}^2 - {\gamma_{, r} \over r}, \end{equation} \begin{equation} {\gamma_{, z} \over r} - 2 \psi_{, r} \psi_{, z} = - 2 e^{-2\psi - 2\phi} A_{0, r} A_{0, z} - 2 e^{ - 2\psi + 2 \phi} G_{0, r} G_{0, z} + 2 \phi_{, r} \phi_{, z}. \end{equation} Thus, with the substitution \begin{equation} E = - \phi - \psi, \qquad N = \phi - \psi, \end{equation} and \begin{equation} {\tilde A}_{0} = {A_{0} \over \sqrt{2}}, \qquad {\tilde G}_{0} = {G_{0} \over \sqrt{2}}, \end{equation} we find that, Eqs. of motion can be rewritten as follows: \begin{eqnarray \label{so1} \nabla^2 E + e^{2 E} \nabla {\tilde A}_{0} \nabla {\tilde A}_{0} = 0, \\ \nabla^2 {\tilde A}_{0} + 2 \nabla E \nabla {\tilde A}_{0} = 0, \\ \nabla^2 N + e^{2 N} \nabla {\tilde G}_{0} \nabla {\tilde G}_{0} = 0, \\ \nabla^2 {\tilde G}_{0} + 2 \nabla N \nabla {\tilde G}_{0} = 0. \label{so} \end{eqnarray} Now, the pseudopotentials for gauge fields ${\tilde A}$ and ${\tilde G}$ have the same form as in Eq.(\ref{wa}), where one should substitute the adequate values of $E, (N)$ and derivatives of the gauge fields under consideration. Then, using Eq.(\ref{e1}) to introduce the complex scalars, enables to rewrite the system of Eqs.(\ref{so1}-\ref{so}) as the decoupled two Ernst's like Eqs. Then, the proof follows the same line as that of the proceding subsection. Finally one can reach the conclusion, that all points of the region ${\cal D}$ equipped with the boundary $\partial {\cal D}$ on two-dimensional manifold ${\cal M}$, $P_{(i)1} = P_{(i)2}$. \\ In view of the foregoing uniqueness theorem, we conclude that all doubly charged black hole solutions characterized by mass $M$, the $F$-field electric charge, the $G$-field electric charge and the dilaton charge $\Sigma$, with the same boundary conditions as found in Ref.\cite{kal} are the same everywhere in the spacetime. \section{Conclusions} In our work we were studying the doubly charged dilaton black holes in the bosonic sector of $N = 4, d = 4$ supergravity, being the low energy limit of the superstring theories. We were interested in the uniqueness of these solutions. Using the method proposed in Ref.\cite{gu1} and finding the adequate forms of pseudopotentials and complex scalars for $SU(4)$ and $SO(4)$ versions of the theory, we were able to find that Eqs. of motion could be arranged in the two Ernst's Eqs. for each of the gauge fields appearing in the theory. These Eqs. give the Ernst's formulation of the generalized Einstein Eqs. for the bosonic sector of $N = 4, d =4$ supergravity. Then, using the idea of G\"urses and Xantopoulous \cite{gx} that Ernst Eqs. are included in the single matrix Eq., we prove the uniqueness of the previously \cite{kal} obtained black hole solutions. \\ It will be interesting to find the exact form of the Ernst Eqs. for more complicated version of the low energy string theory, for instance Einstein-Maxwell axion-dilaton gravity or $N = 8$ black holes now intensively studied \cite{ort}. We hope to return to these problems elsewhere.
\section{Introduction} After standard recombination at $z \simeq 1100$, the universe is considered to be reionized at some redshift $z \gtrsim 5$ from the negative results of Gunn-Peterson experiments in quasar spectra. It has been pointed out that first stars can play an important role in the reionization of the universe (e.g., Couchman \& Rees 1986). This scenario has been investigated in detail using numerical simulations (e.g., Gnedin \& Ostriker 1997) or semi-analytical models (e.g., Fukugita \& Kawasaki 1994; Haiman \& Loeb 1997). In the latter models, it is crucial to know whether a cloud of mass $M$ that virializes at redshift $z$ can cool or not, i.e., if star formation occurs or not. Haiman, Thoul, \& Loeb (1996) investigated this problem (see also Tegmark et al. 1997). They found that molecular cooling plays a crucial role for clouds with $T_{\rm vir} \lesssim 10^{4} {\rm K}$ (``small'' pregalactic clouds, hereafter) at $10 \lesssim z_{\rm vir} \lesssim 100$, where $T_{\rm vir}$ and $z_{\rm vir}$ are the virial temperature and the redshift at virialization respectively, and determined the minimum mass of the virialized cloud must have in order to cool in a Hubble time. On the other hand, Haiman, Rees, \& Loeb (1997) pointed out that in the presence of ultraviolet (UV) background radiation at the level needed to ionize the universe, molecular hydrogen is photodissociated by far ultraviolet (FUV) photons, whose radiation energy is less than the Lyman limit, in small pregalactic clouds. Thus, they asserted that molecular hydrogen in small pregalactic clouds is universally destroyed long before the reionization of the universe. In their reionization model, Haiman \& Loeb (1997) assumed that only objects with virial temperature above $10^{4}$ K can cool in a Hubble time, owing to atomic cooling and star formation occurs subsequently. Recently, Ciardi, Ferrara, \& Abel (1999) found that the photodissociated regions are not large enough to overlap at $z \simeq 20-30$. In the same redshift range, the flux of FUV background is well below the threshold required by Haiman et al.(1997) to prevent the collapse of the clouds. However, molecular hydrogen in a virialized cloud is photodissociated not only by external FUV background radiation, but also by FUV photons produced by massive stars within the cloud. Here, we assess the negative feedback of massive star formation on molecular hydrogen formation in a primordial cloud. \section{Region of Influence of an OB star} Around an OB star, hydrogen is photoionized, and an HII region is formed. Ionizing photons hardly escape from the HII region, but photons whose radiation energy are below the Lyman limit can get away. Such FUV photons photodissociate molecular hydrogen, and a photodissociation region (PDR) is formed just outside the HII region. In this section, we study how much mass in a primordial cloud is affected by such FUV photons from an OB star and, as a result, becomes unable to cool in a free-fall time owing to the lack of the coolant. We consider a small pregalactic cloud of primordial composition. In such an object, H$_{2}$ is formed mainly by the H$^{-}$ process at $z \lesssim 100$: \begin{eqnarray} \label{eq:Hm1} {\rm H}+e^{-} &\rightarrow & {\rm H^{-}}+\gamma ; \\ \label{eq:Hm2} {\rm H}+{\rm H^{-}} &\rightarrow & {\rm H_{2}}+e^{-}. \end{eqnarray} At $z \gtrsim 100$, H$^{-}$ is predominantly photodissociated by CMB photons before the reaction (\ref{eq:Hm2}) proceeds. On the other hand, in a PDR, photodissociation of H$^{-}$ by UV radiation from an OB star does not dominate the reaction (\ref{eq:Hm2}) except in the vicinity of the star. Then we neglect the photodissociation of H$^{-}$. The rate-determining stage of the H$^{-}$ process is the reaction (\ref{eq:Hm1}), whose rate coefficient $k_{\rm H^{-}}$ is (de Jong 1972) \begin{equation} k_{\rm H^{-}}=1.0 \times 10^{-18} T~{\rm s^{-1} cm^{3}}. \end{equation} In a PDR, H$_{2}$ is dissociated mainly via the two-step photodissociation process: \begin{equation} {\rm H_2}+\gamma \rightarrow {\rm H_{2}^*} \rightarrow 2{\rm H}, \end{equation} whose rate coefficient $k_{\rm 2step}$ is given by (Kepner, Babul, \& Spergel 1997; Draine \& Bertoldi 1996) \begin{equation} k_{\rm 2step}=1.13 \times 10^{8} F_{\rm LW}~{\rm s^{-1}}. \end{equation} Here $F_{\rm LW}~({\rm ergs~s^{-1}cm^{-2}Hz^{-1}})$ is the averaged radiation flux in the Lyman and Werner (LW) bands and can be written as \begin{equation} F_{\rm LW}=F_{\rm LW,ex} f_{\rm shield}, \end{equation} where $F_{\rm LW,ex}$ is the incident flux into the PDR at 12.4 eV and the shielding factor $f_{\rm shield}$ is given by (Draine \& Bertoldi 1996) \begin{equation} \label{eq:fsh} f_{\rm shield}={\rm min} \left[ 1,(\frac{N_{\rm H_2}}{10^{14}})^{-0.75} \right]. \end{equation} The timescale in which the H$_{2}$ fraction reaches the equilibrium value is given by \begin{equation} t_{\rm dis}=k_{\rm 2step}^{-1}. \end{equation} In the presence of FUV radiation, if the temperature and density were fixed, the H$_{2}$ fraction $f$ initially would increase and reach the equilibrium value for a temporal ionization fraction after $\sim t_{\rm dis}$, and then it would decline as ionization fraction decreased as a result of recombination. Actually, if the pregalactic cloud can once produce a sufficient amount of molecular hydrogen to cool in a free-fall time, the cloud can collapse and star formation occurs subsequently. Note that because of their low ionization degree, inverse Compton cooling by CMB photons is not effective in small objects with $T_{\rm vir} \lesssim 10^{4}$ K. We investigate here how much mass around an OB star is affected by the photodissociating FUV radiation from the star and becomes unable to cool in a free-fall time. In particular, we seek the lower bound of such mass. The equilibrium number density of H$_{2}$ under ionization degree $x$ is \begin{eqnarray} \label{eq:neq} n_{\rm H_2} &=& \frac{k_{\rm H^{-}}}{k_{\rm 2step}}x n^{2} \\ &=& 0.88 \times 10^{-26} x F_{\rm LW}^{-1} T n^{2}. \end{eqnarray} Near the star, $F_{\rm LW}$ is so large that the dissociation time $t_{\rm dis}$ is smaller than the recombination time $t_{\rm rec}=(k_{\rm rec} x_{\rm i} n)^{-1}$, where $k_{\rm rec}$ is the recombination coefficient, $x_{\rm i}$ is the ionization degree at virialization, and $n$ is the number density of hydrogen nuclei. Then the chemical equilibrium between above processes is reached before significant recombination proceeds. Far distant from the star, the recombination proceeds before the molecular fraction reaches the equilibrium value and the ionization degree significantly diminishes. However, since we are seeking how much mass is at least affected by the photodissociating FUV radiation from the star, we use the equilibrium value (\ref{eq:neq}) with initial ionization degree $x=x_{\rm i}$ as the H$_{2}$ number density. Consider an OB star, which radiates at the rate of $L_{\rm LW}$ [ergs s$^{-1}$ Hz$^{-1}$] in the LW bands. For an O5 star, whose mass is 40 $M_{\sun}$, $L_{\rm LW} \simeq 10^{24}$ ergs s$^{-1}$ Hz$^{-1}$. At the point whose distance from the star is $r$, the averaged flux in the Lyman and Werner bands is approximately given by \begin{equation} F_{\rm LW}=\frac{L_{\rm LW}}{4 \pi r^{2}} f_{\rm shield}. \end{equation} Using above relations, we obtain the H$_2$ column density between the star and the point whose distance from the star is $r$: \begin{equation} \label{eq:column} N_{\rm H_2}= \left\{ \begin{array}{ll} C V & \mbox{($N_{\rm H_2}<10^{14} {\rm cm^{-3}}$)} \\ 10^{14}[0.25 (C V/10^{14}) + 0.75]^{4} & \mbox{($N_{\rm H_2}>10^{14}{\rm cm^{-3}}$)} \end{array} \right. \end{equation} where \begin{eqnarray} C &=& 0.88 \times 10^{-26} x_{\rm i}L_{\rm LW}^{-1} T n^{2},\\ V &=& \frac{4 \pi}{3} r^{3}. \end{eqnarray} Here we have assumed $n=$const. in space, for simplicity. We define here the region of influence around a star as that where the cooling time $t_{\rm cool}=\frac{(3/2)kT}{fn \Lambda_{\rm H_2}}$ becomes larger than the free-fall time $t_{\rm ff}=(\frac{3 \pi \Omega_{\rm b}} {32G m_{\rm p} n})^{1/2}$ as a result of the photodissociation of molecular hydrogen, where $f=n_{\rm H_{2}}/n$ is the H$_{2}$ concentration, $\Lambda_{\rm H_2}$ is the cooling function of molecular hydrogen, and $\Omega_{\rm b}$ is the baryon mass fraction. The condition $t_{\rm cool} > t_{\rm ff}$ is satisfied as long as the H$_2$ fraction \begin{eqnarray} f < f^{\rm (cool)} &=& (\frac{24 G m_{\rm p}}{\pi})^{1/2} \frac{kT}{\Omega_{\rm b}^{1/2} n^{1/2} \Lambda_{\rm H_2}}\\ &=& 1 \times 10^{-3} (\frac{n}{1 {\rm cm^{-3}}})^{-1/2} (\frac{T}{10^3 {\rm K}})^{-3} (\frac{\Omega_{\rm b}}{0.05})^{-1/2}, \end{eqnarray} where we used our fit to the Martin, Schwarz,\& Mandy (1996) H$_{2}$ cooling function \begin{equation} \Lambda_{\rm H_2} \simeq 4 \times 10^{-25} (\frac{T}{1000{\rm K}})^{4} {\rm ergs~s^{-1}cm^{3}} \end{equation} for the low temperature ($600{\rm K} \lesssim T \lesssim 3000{\rm K}$) and low density ($n \lesssim 10^{4} {\rm cm^{-3}}$) regime. The same condition leads to the condition on the averaged flux in the LW bands with equation (\ref{eq:neq}): \begin{eqnarray} F_{\rm LW} > F_{\rm LW}^{\rm (cool)} &=& \frac{k_{\rm H^{-}} x_{\rm i}n }{1.13 \times 10^{8} f^{(\rm cool)}}\\ &=& 0.7 \times 10^{-24} {\rm ergs~s^{-1} cm^{-2} Hz^{-1}} (\frac{x_{\rm i}}{10^{-4}}) (\frac{n}{1 {\rm cm^{-3}}})^{3/2} (\frac{T}{10^3 {\rm K}})^{4} (\frac{\Omega_{\rm b}}{0.05})^{1/2}. \end{eqnarray} Corresponding to this critical LW flux $F_{\rm LW}^{\rm (cool)}$, a critical radius $r^{\rm (cool)}$ is determined by the relation \begin{equation} F_{\rm LW}[r^{\rm (cool)}]=F_{\rm LW}^{\rm (cool)}. \end{equation} Actually, the timescale for molecular hydrogen to reach the equilibrium value at $r=r^{\rm (cool)}$, \begin{equation} t_{\rm dis}[r^{\rm (cool)}]=4.0 \times 10^{8} {\rm yr} (\frac{x_{\rm i}}{10^{-4}})^{-1} (\frac{n}{1 {\rm cm^{-3}}})^{-3/2} (\frac{T}{10^3 {\rm K}})^{-4} (\frac{\Omega_{\rm b}}{0.05})^{-1/2}, \end{equation} is longer than the lifetime of an OB star $t_{\rm OB} \sim 3 \times 10^{6} $ yr. This means that as far region as $r=r^{\rm (cool)}$ is rarely affected within the lifetime of a single massive star. We define here another critical LW flux $F_{\rm LW}^{(\rm eq)}$ and its corresponding radius $r^{\rm (eq)}$ where the timescale for molecular hydrogen to reach the equilibrium value $t_{\rm dis}$ becomes equal to the lifetime of an OB star; \begin{equation} F_{\rm LW}^{(\rm eq)}=0.93 \times 10^{-22} {\rm ergs~s^{-1} cm^{-2} Hz^{-1}} (\frac{t_{\rm OB}}{3 \times 10^6 {\rm yr}})^{-1}, \end{equation} \begin{equation} F_{\rm LW}[r^{\rm (eq)}]=F_{\rm LW}^{\rm (eq)}. \end{equation} In the region of influence, we require that two conditions be met: (1) the cooling time for the equilibrium H$_{2}$ fraction is longer than the free-fall time (i.e., $t_{\rm cool}[f=f^{\rm (eq)}]>t_{\rm ff}$,where $f^{\rm (eq)}$ is the equilibrium H$_{2}$ fraction) and (2) the equilibrium H$_{2}$ fraction is reached within the lifetime of the central star (i.e., $t_{\rm dis}< t_{\rm OB}$). Hence the radius of influence $r^{\rm (inf)}$ is determined by the smaller one of either $r^{\rm (cool)}$ or $r^{\rm (eq)}$; \begin{equation} r^{\rm (inf)}={\rm min}[r^{\rm (cool)},r^{\rm (eq)}]. \end{equation} Note the LW flux $F_{\rm LW}$ at which $t_{\rm dis}$ becomes equal to $t_{\rm rec}$ is $2 \times 10^{-24} (x_{\rm i}/10^{-4})(n/1 {\rm cm^{-3}})(T/10^{3} {\rm K})^{-0.64} {\rm ergs~s^{-1} cm^{-2} Hz^{-1}}$. Here we used the recombination coefficient $k_{\rm rec}=1.88 \times 10^{-10} T^{-0.64}~{\rm s^{-1} cm^{3}}$ (Hutchins 1976). Therefore, even at the edge of the region of influence, the condition $t_{\rm dis}<t_{\rm rec}$ is usually satisfied. In this case, the chemical equilibrium value for H$_2$ is reached before the ionization degree significantly decreases from the initial value $x_{\rm i}$. If the self-shielding of LW band photons could be neglected, the radius $r^{\rm (cool)}$ would be given by \begin{eqnarray} \label{eq:rinf} r^{({\rm cool})}=r^{({\rm cool})}_{\rm no-sh} &=&[\frac{L_{\rm LW}}{4 \pi F_{\rm LW}^{\rm (cool)}}]^{1/2} \\ &=&3.4 \times 10^{23} {\rm cm} (\frac{x_{\rm i}}{10^{-4}})^{-1/2} (\frac{L_{\rm LW}}{10^{24} {\rm ergs~s^{-1} Hz^{-1}}})^{1/2} (\frac{T}{10^3 {\rm K}})^{-2} (\frac{n}{1 {\rm cm^{-3}}})^{-3/4} (\frac{\Omega_{\rm b}}{0.05})^{-1/4}. \end{eqnarray} This expression is valid only when $N_{\rm H_{2}}<10^{14} {\rm cm^{-2}}$. When the H$_2$ column density $N_{\rm H_2}$ becomes larger than $10^{14} {\rm cm^{-2}}$, self-shielding of LW band photons begins as can be seen from equation (\ref{eq:fsh}). Here, we define the shielding radius $r_{\rm sh}$ as the radius where $N_{\rm H_2}(r_{\rm sh})=10^{14} {\rm cm^{-2}}$. Using equation (\ref{eq:column}), the shielding radius is \begin{eqnarray} \label{eq:rsh} r_{\rm sh}&=&[\frac{10^{14}}{(4 \pi/3) C}]^{1/3} \\ &=&3.0 \times 10^{21} {\rm cm} (\frac{x_{\rm i}}{10^{-4}})^{-1/3} (\frac{L_{\rm LW}}{10^{24} {\rm ergs~s^{-1} Hz^{-1}}})^{1/3} (\frac{T}{10^3 {\rm K}})^{-1/3} (\frac{n}{1 {\rm cm^{-3}}})^{-2/3}. \end{eqnarray} When the self-shielding becomes important, $N_{\rm H_2}$ increases and $F_{\rm LW}$ decreases rapidly with $r$. Then $r^{\rm (cool)}$ is not much larger than $r_{\rm sh}$. In such a case, we put $r^{\rm (cool)}=r_{\rm sh}$ as a lower bound. Then $r^{\rm (cool)}$ is given by the lesser one of those given by equations (\ref{eq:rinf}) or (\ref{eq:rsh}). In all the same way as above, we can obtain the value of $r^{({\rm eq})}$; \begin{equation} r^{({\rm eq})}={\rm min}[r^{({\rm eq})}_{\rm no-sh},r_{\rm sh}], \end{equation} where \begin{eqnarray} \label{eq:req} r^{({\rm eq})}_{\rm no-sh} &=&[\frac{L_{\rm LW}}{4 \pi F_{\rm LW}^{\rm (eq)}}]^{1/2} \\ &=&2.9 \times 10^{22} (\frac{L_{\rm LW}}{10^{24} {\rm ergs~s^{-1} Hz^{-1}}})^{1/2} (\frac{t_{\rm OB}}{3 \times 10^{6} {\rm yr}})^{1/2}. \end{eqnarray} The baryonic mass within the region of influence $M_{\rm b}^{\rm (inf)}$ is then \begin{eqnarray} M_{\rm b}^{\rm (inf)} &=& \frac{4 \pi}{3} n m_{\rm p} {r^{\rm (inf)}}^{3}\\ \label{eq:Minf} &=& {\rm min} \left\{ \begin{array}{ll} 1.4 \times 10^{14} M_{\sun} (\frac{x_{\rm i}}{10^{-4}})^{-3/2} (\frac{L_{\rm LW}}{10^{24} {\rm ergs~s^{-1} Hz^{-1}}})^{3/2} (\frac{T}{10^3 {\rm K}})^{-6} (\frac{n}{1 {\rm cm^{-3}}})^{-5/4} (\frac{\Omega_{\rm b}}{0.05})^{-3/4} \\ 0.85 \times 10^{11} M_{\sun} (\frac{L_{\rm LW}}{10^{24} {\rm ergs~s^{-1} Hz^{-1}}})^{3/2} (\frac{t_{\rm OB}}{3 \times 10^{6} {\rm yr}})^{3/2} (\frac{n}{1 {\rm cm^{-3}}}) \\ 1.0 \times 10^{8} M_{\sun} (\frac{x_{\rm i}}{10^{-4}})^{-1} (\frac{L_{\rm LW}}{10^{24} {\rm ergs~s^{-1} Hz^{-1}}}) (\frac{T}{10^3 {\rm K}})^{-1} (\frac{n}{1 {\rm cm^{-3}}})^{-1} \end{array} \right\} \end{eqnarray} In equation (\ref{eq:Minf}), each expression corresponds to the case of $r^{\rm (inf)}=r_{\rm no-sh}^{\rm (cool)}$, $r_{\rm no-sh}^{\rm (eq)}$, and $r_{\rm sh}$ from the top to bottom, respectively. We shall keep this order hereafter. At first glance, the first expression in equation (\ref{eq:Minf}) seems to be always larger than the others, but its stronger dependence on temperature makes it important for higher temperature (i.e., more massive) objects than the normalized value. From equation (\ref{eq:Minf}), we can see that the mass within a region of influence of an O star already exceeds the scale of the small pregalactic object. We have considered in this letter the regulation of star formation by photodissociation of molecular hydrogen in a pregalactic cloud. On the other hand, Lin \& Murray (1992) considered only the regulation by photoionization. In such a case, the mass affected by an OB star, namely the baryonic mass within a Str\"{o}mgren sphere, is \begin{eqnarray} M_{\rm b}^{\rm (St)}&=&\frac{m_{\rm p} n Q_{\ast}}{k_{\rm rec} n^{2}}\\ &=& 3.7 \times 10^{3} M_{\sun} (\frac{T}{10^3 {\rm K}})^{0.64} (\frac{n}{1 {\rm cm^{-3}}})^{-1} (\frac{Q_{\ast}}{10^{49} {\rm s^{-1}}}), \end{eqnarray} where $Q_{\ast}$ is the flux of ionizing photons by a OB star and $Q_{\ast} \simeq 10^{49} {\rm s^{-1}}$ for an O5 star. This is by far smaller than our estimated mass of photodissociative influence $M_{\rm b}^{\rm (inf)}$. To be more specific in the cosmological context, we consider here pregalactic clouds at virialization. The number density at virialization is \begin{equation} \label{eq:nvir} n_{\rm vir}=0.68 {\rm cm^{-3}} h_{50}^{2} (\frac{\Omega_{\rm b}}{0.05})(\frac{1+z_{\rm vir}}{30})^{3}, \end{equation} and the virial temperature is \begin{equation} \label{eq:Tvir} T_{\rm vir}=6.8 \times 10^{2} {\rm K} h_{50}^{2/3} (\frac{\Omega_{\rm b}}{0.05})^{-2/3} (\frac{M_{\rm b}}{10^{4} M_{\sun}})^{2/3} ( \frac{1+z_{\rm vir}}{30}). \end{equation} Substituting equations (\ref{eq:nvir}) and (\ref{eq:Tvir}) into equation (\ref{eq:Minf}), we obtain \begin{equation} \label{eq:Minf2} M_{\rm b}^{\rm (inf)}= {\rm min} \left\{ \begin{array}{ll} 2.3 \times 10^{15} M_{\sun} (\frac{x_{\rm i}}{10^{-4}})^{-3/2} (\frac{L_{\rm LW}}{10^{24} {\rm ergs~s^{-1} Hz^{-1}}})^{3/2} h_{50}^{-13/2} (\frac{\Omega_{\rm b}}{0.05})^{2} (\frac{M_{\rm b}}{10^{4} M_{\sun}})^{-4} ( \frac{1+z_{\rm vir}}{30})^{-39/4} \\ 0.58 \times 10^{11} M_{\sun} (\frac{L_{\rm LW}}{10^{24} {\rm ergs~s^{-1} Hz^{-1}}})^{3/2} (\frac{t_{\rm OB}}{3 \times 10^{6} {\rm yr}})^{3/2} h_{50}^{2} (\frac{\Omega_{\rm b}}{0.05})( \frac{1+z_{\rm vir}}{30})^{3} \\ 2.2 \times 10^{8} M_{\sun} (\frac{x_{\rm i}}{10^{-4}})^{-1} (\frac{L_{\rm LW}}{10^{24} {\rm ergs~s^{-1} Hz^{-1}}}) h_{50}^{-8/3} (\frac{\Omega_{\rm b}}{0.05})^{-1/3} (\frac{M_{\rm b}}{10^{4} M_{\sun}})^{-2/3} ( \frac{1+z_{\rm vir}}{30})^{-4} \end{array} \right\} \end{equation} In order for the star formation to continue after a massive star forms, the region of influence must be smaller than the original pregalactic cloud. Then $M_{\rm b}^{\rm (inf)}<M_{\rm b}$ is the necessary condition, which leads to \begin{equation} \label{eq:cond} M_{\rm b}> {\rm min} \left\{ \begin{array}{ll} 1.9 \times 10^{6} M_{\sun} (\frac{x_{\rm i}}{10^{-4}})^{-3/10} (\frac{L_{\rm LW}}{10^{24} {\rm ergs~s^{-1} Hz^{-1}}})^{3/10} h_{50}^{-13/10} (\frac{\Omega_{\rm b}}{0.05})^{2/5} ( \frac{1+z_{\rm vir}}{30})^{-39/20}\\ 0.58 \times 10^{11} M_{\sun} (\frac{L_{\rm LW}}{10^{24} {\rm ergs~s^{-1} Hz^{-1}}})^{3/2} (\frac{t_{\rm OB}}{3 \times 10^{6} {\rm yr}})^{3/2} h_{50}^{2} (\frac{\Omega_{\rm b}}{0.05})( \frac{1+z_{\rm vir}}{30})^{3}\\ 4.0 \times 10^{6} M_{\sun} (\frac{x_{\rm i}}{10^{-4}})^{-3/5} (\frac{L_{\rm LW}}{10^{24} {\rm ergs~s^{-1} Hz^{-1}}})^{3/5} h_{50}^{-8/5} (\frac{\Omega_{\rm b}}{0.05})^{-1/5} ( \frac{1+z_{\rm vir}}{30})^{-12/5} \end{array} \right\}. \end{equation} On the other hand, the baryonic mass of a pregalactic cloud that has virial temperature $T_{\rm vir}$ is \begin{equation} \label{eq:Mb} M_{\rm b}=1.8 \times 10^{4} M_{\sun} (\frac{T_{\rm vir}}{1000 {\rm K}})^{3/2} h_{50}^{-1} (\frac{\Omega_{\rm b}}{0.05}) ( \frac{1+z_{\rm vir}}{30})^{-3/2}. \end{equation} Comparing equations (\ref{eq:cond}) and (\ref{eq:Mb}), we can see that for a small pregalactic cloud ($T_{\rm vir}<10^{4}$K) the condition $M_{\rm b}^{\rm (inf)}<M_{\rm b}$ is hardly satisfied in the redshift range of $10 \lesssim z_{\rm vir} \lesssim 100$. This indicates that FUV radiation from one or a few OB stars prohibits the whole small pregalactic cloud from H$_{2}$ cooling and quenches subsequent star formation in it. After the death of the first OB star, star formation could occur somewhere in the cloud, and another OB star could form successively. Thereafter, some massive stars might form one after another, but only a few could co-exist simultaneously, as we have shown above. The timescale for reformation of OB stars depends on that for ${\rm H_2}$ replenishment after the death of the dissociating OB star, which is \begin{eqnarray} t_{\rm rep}&=&\frac{f^{(\rm cool)}}{k_{\rm H^{-}} x_{\rm e} n} \\ &=&3 \times 10^{8} {\rm yr} (\frac{n}{1 {\rm cm^{-3}}})^{-3/2} (\frac{T}{10^3 {\rm K}})^{-4} (\frac{x_{\rm e}}{10^{-4}})^{-1} (\frac{\Omega_{\rm b}}{0.05})^{-1/2}. \end{eqnarray} If the ionization degree is as high as unity, which is typical in the HII region, this timescale can be very short, namely, $t_{\rm rep} \simeq 3 \times 10^{4}$ yr. If the timescale for reformation of OB stars is smaller than the lifetime of OB star, it would be possible for a few OB stars to form every few million years, and, as a result, a considerable amount of stars would form in a Hubble time. However, we do not expect that the star formation continues as long as a Hubble time, because the gravitational binding energy of baryonic gas in such a small pregalactic cloud, \begin{eqnarray} E_{\rm gr} & \simeq & \frac{G M M_{\rm b}}{R_{\rm vir}} \\ & \simeq & 3.3 \times 10^{48} {\rm ergs} (\frac{M_{\rm b}}{10^{4} M_{\sun}})^{5/3} (\frac{\Omega_{\rm b}}{0.05})^{-2/3} h_{50}^{2/3} ( \frac{1+z_{\rm vir}}{30}), \end{eqnarray} where $M$ is the total mass, and $R_{\rm vir}$ is the virial radius of the cloud, is so small that a few supernova explosions would blow out such a small pregalactic cloud (see, e.g., Dekel \& Silk 1986). Thus, our probable scenario is as follows: in a small pregalactic cloud, star formation occurs only in a photodissociatively regulated fashion until several supernovae explode, and it stops thereafter. If numerous small stars were formed per a massive star, a substantial proportion of the original cloud would be converted to stars at last. However, in the primordial circumstance, the stellar initial mass function could be strongly biased toward the formation of massive stars because of the higher temperatures relative to the present-day counterpart. If this was the case and OB stars were formed selectively, the amount of gas mass that was converted to stars would be extremely small. \section{Summary} We have studied the H$_{2}$ photodissociation region around an OB star in a primordial gas cloud. A region as large as the whole small pregalactic cloud is affected by only one or a few OB stars and becomes unable to cool in a free-fall time under the condition appropriate to virialization. Therefore, in those clouds which have virial temperatures less than $10^{4}$ K, star formation does not occur efficiently, unless the primordial initial mass function is extremely weighted toward low mass stars. If the reionization of the universe is caused by stellar UV radiation, some OB stars must form. However as we have shown, an OB star formed in a small pregalactic cloud would inevitably photodissociate the whole cloud and subsequent star formation would be strongly suppressed. Therefore, stellar UV radiation from small pregalactic clouds cannot play a significant role in the reionization of the universe. \acknowledgements The authors would like to thank Toru Tsuribe for fruitful discussions, Humitaka Sato and Naoshi Sugiyama for continuous encouragement, Evan Scannapieco for checking the English, and the referee, Zolt\'{a}n Haiman, for a careful reading of the manuscript and for useful comments. This work is supported in part by the Grant-in-Aid for Scientific Research on Priority Areas (No. 10147105) (R.N.), and the Grant-in-Aid for Scientific Research from the Ministry of Education, Science, Sports and Culture, No. 08740170 (R.N.). \newpage
\section{Abstract} I review several topics pertaining to Weak decays of $b$ and $c$ quarks, including measurements of $|V_{cb}|$, $|V_{ub}/V_{cb}|$, $f_{D_s}$ and $b\to s\gamma$. \section{Introduction} Leptons and quarks, along with gluons, photons and gauge bosons are the fundamental objects in nature described by the Standard Model of electroweak interactions. Although the model has been successful at describing the interactions between these objects, many important questions remain. \begin{tabbing} \=$\bullet$Why are there so many fundamental constants?\\ \>$\bullet$What is the relationship of these constants to quark masses?\\ \>$\bullet$Are quarks and leptons really pointlike?\\ \>$\bullet$Is the Standard Model description correct, especially of CP violation?\\ \>$\bullet$What is the connection between CP and matter-antimatter asymmetry?\\ \end{tabbing} In weak interactions of quarks, we are interested in the couplings of quarks to each other and leptons, but have to deal with the ``brown muck'' of hadrons. The basic weak $V-A$ structure has been verified with purely leptonic decays, for example, $\mu\to e\nu_e\nu_{\mu}$, $\tau\to e\nu_e\nu_{\tau}$. I do not have enough space to report on all interesting aspects of weak decays here, so I will report on a few, but miss others, even ones which I covered in my presentation. \subsection{The CKM Matrix and CP Violation} \label{sec:Intro} The physical point-like states of nature that have both strong and electroweak interactions, the quarks, are mixtures of base states described by the Cabibbo-Kobayashi-Maskawa matrix,\cite{ckm} \begin{displaymath} \left(\begin{array}{c}d'\\s'\\b'\\\end{array} \right) = \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}{c}d\\s\\b\\\end{array}\right) \end{displaymath} The unprimed states are the mass eigenstates, while the primed states denote the weak eigenstates. There are nine complex CKM elements. These 18 numbers can be reduced to four independent quantities by applying unitarity and the fact that the phases of the quark wave functions are arbitrary. These four remaining numbers are fundamental constants of nature that need to be determined from experiment, like any other fundamental constant such as $\alpha$ or $G$. In the Wolfenstein approximation the matrix is written as\cite{wolf} \begin{displaymath} V_{CKM} = \left(\begin{array}{ccc} 1-\lambda^2/2 & \lambda & A\lambda^3(\rho-i\eta(1-\lambda^2/2)) \\ -\lambda & 1-\lambda^2/2-i\eta A^2\lambda^4 & A\lambda^2(1+i\eta\lambda^2) \\ A\lambda^3(1-\rho-i\eta) & -A\lambda^2& 1 \end{array}\right). \end{displaymath} This expression is accurate to order $\lambda^3$ in the real part and $\lambda^5$ in the imaginary part. It is necessary to express the matrix to this order to have a complete formulation of the physics we wish to pursue. The constants $\lambda$ and $A$ have been measured using semileptonic $s$ and $b$ decays;\cite{virgin} $\lambda\approx 0.22$, and $A\approx 0.8$. The phase $\eta$ allows for CP violation. CP violation thus far has only been seen in the neutral kaon system. If we can find CP violation in the $B$ system we could see if the CKM model works or perhaps discover new physics that goes beyond the model, if it does not. It is also of great interest to measure the magnitudes of each of the matrix elements. Techniques used have included: $V_{ud}$ from $0^+\to 0^+$ nuclear $\beta$-decay, $V_{us}$ from $K\to\pi\ell\nu$ and hyperon semileptonic decays, $V_{ub}$ from charmless semileponic $b$ decays, $V_{cd}$ from neutrino interactions and charm semileptonic decay, $V_{cs}$ from direct $W^{\pm}$ decays at LEP II, $V_{cb}$ from charmed semileptonic $b$ decays, $V_{td}$ from $B_d^o$ mixing, limits on $V_{ts}$ from $B_s$ mixing, and limits on $V_{tb}$ from $t$ decays. The measurements of $V_{cb}$ and $V_{ub}$ will be discussed here. \subsection{Measurement Of $|V_{cb}|$ Using $B\to D^*\ell\nu$} Currently, the most favored technique is to measure the decay rate of $B\to D^{*}\ell^-\bar{\nu}$ at the kinematic point where the $D^{*+}$ is at rest in the $B$ rest frame (this is often referred to as maximum $q^2$ or $\omega =1$). Here, according to Heavy Quark Effective Theory, the theoretical uncertainties are at a minimum. There are results from several groups using this technique for the decay sequence $D^{*+}\to \pi^+ D^o$; $D^o\to K^-\pi^+$, or similar decays of the $D^{*o}$. The ALEPH results\cite{aleph_pi} are shown in Fig.~\ref{aleph_vcb}. \begin{figure} \vspace{-19mm} \centerline{\epsfig{figure=aleph_vcb_2.eps,width=4.2in}} \vspace{-50mm} \caption{\label{aleph_vcb} $\overline{B}^o\to D^{+}\ell^-\bar{\nu}$ from ALEPH. The data have been fit to a functional form suggested by Caprini {\it et al.} ~The abcissa gives the value of the product $|F(1)*V_{cb}|^2$.} \end{figure} In a recent analysis, DELPHI detects only the slow $\pi^+$ from the $D^{*+}$ decay and does not explicitly reconstruct the $D^o$ decay.\cite{delphi_pi} Table~\ref{tab:Vcb} summaries determinations of $|V_{cb}|$; here, the first error is statistical, the second systematic and the third, an estimate of the theoretical accuracy in predicting the form-factor $F(\omega=1)=0.91\pm 0.003$.\cite{formfactor} Currently, DELPHI has the smallest error, however, CLEO has only used 1/6 of their current data. The quoted average $|V_{cb}|=0.0381\pm 0.0021$ combines the averaged statistical and systematic errors with the theoretical error in quadrature and takes into account the common systematic errors, such as the $D^*$ branching ratios. \begin{table}[th] \vspace{-2mm} \begin{center} \caption{Modern Determinations of $|V_{cb}|$ using $B\to D^*\ell^-\overline{\nu}$ decays at $\omega = 1$ \label{tab:Vcb}} \vspace*{2mm} \begin{tabular}{lc}\hline\hline Experiment & $V_{cb}$ $(\times 10^{-3})$\\\hline ALEPH\cite{aleph_pi} & $34.4\pm 1.6 \pm 2.3 \pm 1.4$ \\ DELPHI\cite{delphi_pi} & $41.2\pm 1.5 \pm 1.8 \pm 1.4$ \\ OPAL\cite{opal_pi} & $36.0\pm 2.1 \pm 2.1 \pm 1.2$ \\ CLEO\cite{cleo_pi} & $39.4\pm 2.1 \pm 2.0 \pm 1.4$ \\\hline Average & $38.1\pm 2.1$\\ \hline\hline \end{tabular} \end{center} \end{table} There are other ways of determining $V_{cb}$. One new method based on QCD sum rules uses the operator product expansion and the heavy quark expansion, in terms of the parameters $\alpha_s (m_b)$, $\overline{\Lambda}$, and the matrix elements $\lambda_1$ and $\lambda_2$. The latter quantities arise from the differences \begin{displaymath} m_B-m_b=\overline{\Lambda}-{{\lambda_1+3\lambda_2}\over{2m_b}}~~~ m_B^*-m_b=\overline{\Lambda}-{{\lambda_1+-\lambda_2}\over{2m_b}}~~. \end{displaymath} The $B^*-B$ mass difference determines $\lambda_2 = 0.12$ GeV$^2$. The total semileptonic decay width is then related to above parameters as \begin{eqnarray*} &&\Gamma_{sl}={{G_F^2\left|V_{cb}\right|^2m_B^5}\over {192\pi^3}}0.369\times\nonumber\\ &&\left[1-1.54{\alpha_s \over \pi}-1.65{\overline{\Lambda}\over m_B}\left( 1-.087{\alpha_s\over\pi}\right)-0.95{\overline{\Lambda}^2\over m_B^2} -3.18{\lambda_1\over m_B^2}+0.02{\lambda_2\over m_B^2}\right]\nonumber\\ \end{eqnarray*} CLEO has measured the semileptonic branching ratio using lepton tags as (10.49$\pm$0.17$\pm$0.43)\% and using the world average lifetime for an equal mixture of $B^o$ and $B^-$ mesons of 1.613$\pm$0.020 ps, CLEO finds $\Gamma_{sl} = 65.0\pm 3.0$ ns$^{-1}$. (Note that LEP has a somewhat larger value of 68.6$\pm$1.6 ns$^{-1}$.) CLEO then attempts to measure the remaining unknown parameters $\lambda_1$ and $\overline{\Lambda}$ by using moments of the either the hadronic mass or the lepton energy.\cite{moment} The results are shown in Fig.~\ref{moments}. Here the measurements are shown as bands reflecting the experimental errors. Unfortunately, this preliminary CLEO result shows a contradiction. The overlap of the mass moment bands gives different values than the lepton energy moments! The mass moments are theoretically favored and give the values $\lambda_1$= (0.13$\pm$0.01$\pm$0.06) GeV$^2$, and $\overline{\Lambda}$ = (0.33$\pm$0.02$\pm$0.08) GeV. The discrepancy between the two methods is serious. It either means that there is something wrong with the CLEO analysis or there is something wrong in the theory. If the latter is true it would shed doubt on the method used by the LEP experiments to extract a value of $|V_{ub}|$ using the same theoretical framework. \begin{figure}[b] \vspace{-5mm} \centerline{\epsfig{figure=moments.eps,width=2.1in}} \vspace{5mm} \caption{\label{moments}Bands in $\overline{\Lambda}-\lambda_1$ space found by CLEO in analyzing first and second moments of hadronic mass squared and lepton energy. The intersections of the two moments for each set determines the two parameters. The one standard deviation error ellipses are shown.} \end{figure} \subsection{Measurement Of $|V_{ub}|$} Another important CKM element that can be measured using semileptonic decays is $V_{ub}$. The first measurement of $V_{ub}$ done by CLEO and subsequently confirmed by ARGUS, used only leptons which were more energetic than those that could come from $b\to c\ell^- \bar{\nu}$ decays.\cite{first_vub} These ``endpoint leptons'' can occur, $b\to c$ background free, at the $\Upsilon (4S)$ because the $B$'s are almost at rest. Unfortunately, there is only a small fraction of the $b\to u \ell^-\bar{\nu}$ lepton spectrum that can be seen this way, leading to model dependent errors. ALEPH\cite{aleph_vub} L3\cite{L3_vub} and DELPHI\cite{delphi_vub} try to isolate a class of events where the hadron system associated with the lepton is enriched in $b\to u$ and thus depleted in $b\to c$. They define a likelihood that hadron tracks come from $b$ decay by using a large number of variables including, vertex information, transverse momentum, not being a kaon. Then they require the hadronic mass to be less than 1.6 GeV, which greatly reduces $b\to c$, since a completely reconstructed $b\to c$ decay has a mass greater than that of the $D$ (1.83 GeV). They then examine the lepton energy distribution for this set of events, shown in Fig.~\ref{delphi_vub} for DELPHI. \begin{figure}[bt] \vspace{-9mm} \centerline{\epsfig{figure=delphi_vub.eps,width=2.in}} \vspace{-4mm} \caption{\label{delphi_vub}The lepton energy distribution in the $B$ rest frame from DELPHI. The data have been enriched in $b \to u$ events, and the mass of the recoiling hadronic system is required to be below 1.6 GeV. The points indicate data, the light shaded region, the fitted background and the dark shaded region, the fitted $b \to u \ell \nu$ signal. } \end{figure} I have averaged all three LEP results and show them in Fig.~\ref{vub} without any theoretical error, which is estimated at $\pm$8\% by Uraltsev.\cite{vub_thy_inc} However, another calculation using the same type of model by Jin\cite{jin} gives a $\pm$14\% lower value, with a quoted error of $\pm$10\%. My best estimate of $\left|V_{ub}/V_{cb}\right|$ using this technique includes a $\pm$14\% theoretical error added in quadrature with a common systematic error of $\pm$14\%, since the Monte Carlo calculations at LEP are known to be strongly correlated. Also shown in Fig.~\ref{vub} are results from CLEO using the measured the decay rates for the exclusive final states $\pi\ell\nu$ and $\rho\ell\nu$,\cite{cleo_pirho} and results from endpoint leptons, dominated by CLEO II.\cite{cleo_vub} Several theoretical models are used.\cite{vub_thy} From the exclusive results, the model of Korner and Schuler (KS) is ruled out by the measured ratio of $\rho/\pi$. This model deviated the most from the others used to get values of $|V_{ub}|$ from endpoint leptons. Thus the main use of the exclusive final states has been to restrict the models. The endpoint lepton results are statistically the most precise. Assigning a model dependent error is quite difficult. I somewhat arbitrarily have assigned a $\pm$14\% irreducible systematic error to these models and used the average among them to derive a value. My best overall estimate is that $|V_{ub}/V_{cb}|=0.087\pm 0.012$. \begin{figure}[htb] \centerline{\epsfig{figure=Vub_cb.eps,width=3.6in}} \caption{\label{vub}Measurements of $|V_{ub}/V_{cb}|$ using different techniques and theoretical models. (The KS model has been ruled out.) } \end{figure} This estimate must be treated as highly suspect. The value and error depends on uncertain theoretical estimates. We can use this estimate, along with other measurements. To get some idea of what the values of $\rho$ and $\eta$ are. There is a constraint on $\rho$ and $\eta$ given by the $K_L^o$ CP violation measurement ($\epsilon$), given by\cite{buras} \begin{displaymath} \eta\left[(1-\rho)A^2(1.4\pm 0.2)+0.35\right]A^2{B_K \over 0.75}=(0.30\pm 0.06), \end{displaymath} where the errors arise mostly from uncertainties on $|V_{cb}|$ and $B_K$. Here $B_K$ is taken as 0.75$\pm$0.15 according to Buras.\cite{Buras_bk} The constraints on $\rho$ versus $\eta$ from the $|V_{ub}/V_{cb}|$ determination, $\epsilon$ and $B$ mixing are shown in Fig.~\ref{ckm_tri_6}. The bands represent $\pm1\sigma$ errors, for the measurements and a 95\% confidence level upper limit on $B_s$ mixing. The width of the $B_d$ mixing band is caused mainly by the uncertainty on $f_B$, taken here as $240> f_B > 160$ MeV. Other parameters include $|V_{cb}|=0.381\pm 0.0021$, $|V_{ub}/V_{cb}| = 0.087\pm 0.012$, limit on $\Delta M_s > 12.4$ ps$^{-1}$, and the ratio $f_{B_s}\sqrt{B_{B_s}}/f_{B_d}\sqrt{B_{B_d}}\leq 1.25$.\cite{Jonnew} \begin{figure}[htbp] \centerline{\epsfig{figure=ckm_tri_6.eps,height=2.5in}} \vspace{-.4cm} \caption{\label{ckm_tri_6}The regions in $\rho-\eta$ space (shaded) consistent with measurements of CP violation in $K_L^o$ decay ($\epsilon$), $V_{ub}/V_{cb}$ in semileptonic $B$ decay, $B_d^o$ mixing, and the excluded region from limits on $B_s^o$ mixing. The allowed region is defined by the overlap of the 3 permitted areas, and is where the apex of the CKM triangle sits.} \end{figure} \section{The decays $B^-\to\ell^-\overline{\nu}$ and $D_s^+\to\mu^+\nu$} This reaction proceeds via the annihilation of the $b$ quark with the $\overline{u}$ into a virtual $W^-$ which materializes as $\ell^-\overline{\nu}$ pair as illustrated in Fig.~\ref{btolnu}. The decay rate for this process can be written as \begin{displaymath} \Gamma(B^-\to \ell^- \overline{\nu})={{G_F^2}\over 8\pi}f_{B}^2m_{\ell}^2M_{B} \left(1-{m_{\ell}^2\over M_{B}^2}\right)^2 \left|V_{ub}\right|^2~~~, \label{eq:equ_rate} \end{displaymath} where $f_B$ is the so called ``decay constant," a parameter that can be calculated theoretically or determined by measuring the decay rate. This formula is the same for all pseudoscalar mesons using the appropriate CKM matrix element and decay constant. \begin{figure}[htb] \vspace{-.2cm} \centerline{\epsfig{figure=btolnu.eps,height=1.2in}} \vspace{-1.8cm} \caption{\label{btolnu} Diagram for a $B^-\to \ell^-\overline{\nu}$ decay.} \end{figure} Knowledge of $f_B$ is important because it is used to determine constraints on CKM matrix elements from measurements of neutral $B$ mixing. Since the decay is helicity suppressed, the heavier the lepton the larger the expected rate. Thus looking for the $\tau^-\overline{\nu}$ has its advantages. The big disadvantage is that there are least two missing neutrinos in the final state. The most stringent limit has been set by L3 of $<5.7\times 10^{-4}$ at 90\% confidence level, using a missing energy technique.\cite{L3taunu} This is still one order of magnitude higher than what is expected. Other limits are poorer.\cite{othertaunu} Since $f_B$ is so difficult to measure, models, especially lattice gauge models, are used.\cite{Bernard} However, it is prudent to test these models. $D_s^+\to\mu^+\nu$ can be used; it is Cabibbo favored and the predicted branching ratio is close to 1\%. CLEO has made the highest statistics measurement to date of ${\cal B}(D^+_s\to \mu^+\nu)$, by searching for the decay sequence $D_s^{*+}\to\gamma D_s^+$, $D^+_s\to \mu^+\nu$. Since the decay $D_s\to e\nu$ is suppressed by four orders of magnitude due to helicity, they use this mode to measure the physics backgrounds due to real muons. Then they need correct only for differences in muon and electron efficiencies and fake rates. They use missing energy and momentum to define the $\nu$ direction. The mass difference $\Delta M$ is calculated as difference in $D_s^*$ and $D_s$ invariant mass. The $\Delta M$ distributions for the muon and electron data and the calculated effective excess of muon fakes over electron fakes are shown in Fig.~\ref{Fdata}(a). The histogram is the result of a $\chi^2$ fit of the muon spectrum to the sum of three contributions: the signal, the scaled electrons, and the excess of muon over electron fakes. Here, the sizes of the electron and fake contributions are fixed and only the signal normalization is allowed to vary. The signal consists of two components, whose relative normalization is fixed. These two components are the decay $D_s^{*+}\to\gamma D_s^+$, $D_s^+\to\mu^+\nu$ and the direct decay $D_s^+\to\mu\nu$ and $D^+\to\mu^+\nu$ combined with a random photon. CLEO finds a signal of 182$\pm$22 events in the peak which are attributed to the process $D_s^{*+}\to\gamma D_s^+$, $D_s^+\to\mu^+\nu$. They also find 250$\pm$38 events in the flat part of the distribution corresponding to $D_s^+\to\mu^+\nu$ or $D^+\to\mu^+\nu$ decays coupled with a random photon. The contribution of a real $D^+\to\mu^+\nu$ decay with random photons is not entirely negligible since the $D^{*+}\to\gamma D^+$ branching ratio does not enter. The $D^+$ fraction is estimated to be about (18$\pm$8)\% relative to the total $D_s^+\to\mu^+\nu$ plus random photon contribution. \begin{figure}[b] \vspace{-0.8cm} \centerline{\epsfig{figure=dsmunu_dat.eps,width=4.1in}} \caption{\label{Fdata} (a) The $\Delta M$ mass difference distribution for $D_s^{*+}$ candidates for both the muon data (solid points), the electron data (dashed histogram) and the excess of muon fakes over electron fakes (shaded). The histogram is the result of the fit described in the text. (b) The $\Delta M$ mass difference distribution for $D_s^{*+}$ candidates with electrons and excess muon fakes subtracted. The curve is a fit to the signal shape described in the text.} \end{figure} Several other groups have made measurements. The results are shown in Table~\ref{tab:dsmunu}. I have changed the values of $f_{D_s}$ according to the updated PDG $D_s$ decay branching fractions for the normalization modes,\cite{PDG} and have corrected the old CLEO result by using the new fake rates determined in their updated analysis. In addition, there are new results using the $D_s^+\to \tau^+\nu$ decay from the L3 collaboration\cite{L3taunu} of ($309\pm 58\pm 33 \pm 38$) MeV, and ($330 \pm 95$) MeV from the DELPHI collaboration.\cite{othertaunu} The world average value for $f_{D_s}$ is ($255\pm 21\pm 28$) MeV, where the common systematic error is due the error on the absolute branching ratio for $D_s^+\to \phi\pi^+$. These numbers are consistent with C. Bernard's world average for lattice theories of (221$\pm$25) MeV.\cite{Bernard} \begin{table}[hbt] \vspace{-2mm} \caption{Measured values of $f_{D_s}$ from experimental values of $\Gamma(D_s^+\to\mu^+\nu)$\label{tab:dsmunu}} \begin{center} \footnotesize \begin{tabular}{|lccc|}\hline Collaboration & Observed & Published $f_{D_s}$ & Corrected $f_{D_s}$ \\ & Events & value (MeV) & value (MeV) \\ \hline CLEO (old) \cite{cleo} & 39$\pm$8 & $344\pm 37 \pm 52 \pm 42$ & $282 \pm 30 \pm 43 \pm 34$ \\ WA75 \cite{Bullshit} & 6 & $232 \pm 45 \pm 20 \pm 48$ & $ 213 \pm 41 \pm 18 \pm 26$\\ BES \cite{Bes} & 3 & $430 ^{+150}_{-130} \pm 40$ & Same\\ E653 \cite{E653} & $23.2\pm 6.0 ^{+1.0}_{-0.9}$ & $194\pm 35\pm 20\pm 14$ & $200\pm 35\pm 20\pm 26$\\ CLEO \cite{chanda} & 182$\pm$22 & - & $280 \pm 19 \pm 28 \pm 34$ \\ \hline \end{tabular} \end{center} \end{table} \section{Rare Decays as Probes beyond the Standard Model} Rare decays have loops in the decay diagrams so they are sensitive to high mass gauge bosons and fermions. Thus, they are sensitive to new physics. However, it must be kept in mind that any new effect must be consistent with already measured phenomena such as $B_d^o$ mixing and $b\to s\gamma$. These processes are often called ``Penguin" processes, for unscientific reasons. A Feynman loop diagram is shown in Fig.~\ref{loop} that describes the transition of a $b$ quark into a charged -1/3 $s$ or $d$ quark, which is effectively a neutral current transition. The dominant charged current decays change the $b$ quark into a charged +2/3 quark, either $c$ or $u$. \begin{figure}[htb] \vspace{-.6cm} \centerline{\epsfig{figure=loop_dia.eps,height=0.7in}} \caption{\label{loop}Loop or ``Penguin" diagram for a $b\to s$ or $b\to d$ transition.} \vspace{-1mm} \end{figure} The intermediate quark inside the loop can be any charge +2/3 quark. The relative size of the different contributions arises from different quark masses and CKM elements. In terms of the Cabibbo angle ($\lambda$=0.22), we have for $t$:$c$:$u$ - $\lambda^2$:$\lambda^2$:$\lambda^4$. The mass dependence favors the $t$ loop, but the amplitude for $c$ processes can be quite large $\approx$30\%. Moreover, as pointed out by Bander, Silverman and Soni,\cite{BSS} interference can occur between $t$, $c$ and $u$ diagrams and lead to CP violation. In the standard model it is not expected to occur when $b\to s$, due to the lack of a CKM phase difference, but could occur when $b \to d$. In any case, it is always worth looking for this effect; all that needs to be done, for example, is to compare the number of $K^{*-}\gamma$ events with the number of $K^{*+}\gamma$ events. There are other possibilities for physics beyond the standard model to appear. For example, the $W^-$ in the loop can be replaced by some other charged object such as a Higgs; it is also possible for a new object to replace the $t$. \subsection{$b \to s\gamma$} This process occurs when any of the charged particles in Fig.~\ref{loop} emits a photon. CLEO first measured the inclusive rate\cite{oldbsg} as well as the exclusive rate into $K^*(890)\gamma$.\cite{fbsg} There is an updated CLEO measurement\cite{CLEObsg} using 1.5 times the original data sample and a new measurement from ALEPH.\cite{ALEPHbsg} To remove background CLEO used two techniques originally, one based on ``event shapes" and the other on summing exclusively reconstructed $B$ samples. CLEO uses eight different shape variables,\cite{oldbsg} and defines a variable $r$ using a neural network to distinguish signal from background. The idea of the $B$ reconstruction analysis is to find the inclusive branching ratio by summing over exclusive modes. The allowed hadronic system is comprised of either a $K_s\to\pi^+\pi^-$ candidate or a $K^{\mp}$ combined with 1-4 pions, only one of which can be neutral. The restriction on the number and kind of pions maximizes efficiency while minimizing background. It does however lead to a model dependent error. Then both analysis techniques are combined. Currently, most of the statistical power of the analysis ($\sim$80\%) comes from summing over the exclusive modes. Fig.~\ref{bsg_van} shows the photon energy spectrum of the inclusive signal, compared with the model of Ali and Greub.\cite{Ali} A fit to the model over the photon energy range from 2.1 to 2.7 GeV/c gives the branching ratio result shown in Table~\ref{btosgresults}, where the first error is statistical and the second systematic. \begin{figure}[bh] \centerline{\epsfig{figure=bsg_van.eps,height=1.4in}} \caption{\label{bsg_van}The background subtracted photon energy spectrum from CLEO. The dashed curve is a spectator model prediction from Ali and Greub.} \end{figure} \begin{table}[hbt] \vspace{-5mm} \caption{Experimental results for $b\to s\gamma$\label{btosgresults}} \begin{center} \footnotesize \begin{tabular}{|lc|} \hline Sample & branching ratio\\\hline CLEO & $(3.15\pm 0.35\pm 0.41)\times 10^{-4}$\\ ALEPH & $(3.11\pm 0.80\pm 0.72)\times 10^{-4}$ \\ Average & $(3.14\pm 0.48)\times 10^{-4}$\\ Theory\cite{bsgthy}&$(3.28\pm 0.30)\times 10^{-4}$\\ \hline \end{tabular} \end{center} \end{table} ALEPH reduces the backgrounds by weighting candidate decay tracks in a $b \to s\gamma$ event by a combination of their momentum, impact parameter with respect to the main vertex and rapidity with respect to the $b$-hadron direction.\cite{ALEPHbsg} There result is shown in Table~\ref{btosgresults}. The world average value experimental value is also given, as well as the theoretical prediction. The consistency with standard model expectation has ruled out many models. Hewett has given a good review of the many minimal supergravity models which are excluded by the data.\cite{Hewett} Triple gauge boson couplings are of great interest in checking the standard model. If there were an anomalous $WW\gamma$ coupling it would serve to change the standard model rate. $p\overline{p}$ collider experiments have also published results limiting such couplings.\cite{D0} In a two-dimensional space defined by $\Delta\kappa$ and $\lambda$, the D0 constraint appears as a tilted ellipse and and the $b\to s\gamma$ as nearly vertical bands. In the standard model both parameters are zero. \section*{Acknowledgments} I thank Marina Artuso, B. Kayser, R. Peccei, Jon Rosner, and Tomasz Skwarnicki for interesting discussions. This work was supported by the U. S. National Science Foundation.
\section{Introduction} Threshold resummations have recently been studied for a variety of hadronic processes, in particular heavy quark and jet production (for a review see Ref.~\cite{NK}). Near threshold for the production of the final state in these processes one finds large logarithmic corrections which originate from soft gluon emission off the partons in the hard scattering. These logarithms can be resummed to all orders in perturbative QCD. The resummation of the leading theshold logarithms in heavy quark and dijet production cross sections follows the earlier study of the Drell-Yan process \cite{St87,CT} since these logarithms originate from the incoming partons and are thus universal. The resummation of next-to-leading loga-\break rithms (NLL) for QCD hard scattering and heavy quark production, in particular, was presented in Refs. \cite{Thesis,KS}. At NLL accuracy one has to take into account the color exchange in the hard scattering. The resummation is formulated in terms of soft anomalous dimension matrices which describe the factorization of soft gluons from the hard scattering. Applications to top and bottom quark production at a fixed center-of-mass scattering angle, $\theta=90^{\circ}$ (where the relevant anomalous dimension matrices are diagonal), were discussed in Ref. \cite{NKJSRV}. More recently the total cross section was calculated for top quark production at the Tevatron \cite {NKRV}. Resummation for single-particle inclusive cross sections has been considered in Ref.~\cite{LOS}. The NLL resummation formalism for the \break hadroproduction of dijets has been presented in Refs. \cite{KOS1,KOS2}. Jet production involves additional complications relative to heavy quark production because of the presence of the final-state jets. \section{Resummation for heavy quark \break production} We begin with the resummation formalism for heavy quark production. First, we write the factorized form of the cross section and identify singular distributions in it near threshold. Then we refactorize the cross section into functions associated with gluons collinear to the incoming quarks, non-collinear soft gluons, and the hard scattering. Resummation follows from the renormalization properties of these functions. The resummed cross section is written in terms of exponentials of anomalous dimension matrices for each partonic subprocess involved. \subsection{Factorized cross section} We study the production of a pair of heavy quarks of momenta $p_1$, $p_2$, in collisions of hadrons $h_a$ and $h_b$ with momenta $p_a$ and $p_b$, \begin{equation} h_a(p_a)+h_b(p_b) \rightarrow {\bar Q}(p_1) +Q(p_2) + X \, , \end{equation} with total rapidity $y$ and scattering angle $\theta$ in the pair center-of-mass frame. The heavy quark production cross section can be written in a factorized form as a convolution of the perturbatively calculable hard scattering $H_{f \bar f}$ with parton distributions $\phi_{f_i/h}$, at factorization scale $\mu$, for parton $f_i$ carrying a momentum fraction $x_i$ of hadron $h$: \begin{eqnarray} \frac{d\sigma_{h_a h_b\rightarrow Q{\bar Q}}} {dQ^2 \, dy \, d\cos\theta}&=&\sum_{f \bar f} \int \frac{dx_a}{x_a} \frac{dx_b}{x_b} \, \phi_{f/h_a}(x_a,\mu^2) \nonumber \\ &&\hspace{-27mm}\times \, \phi_{{\bar f}/h_b}(x_b,\mu^2) \; H_{f{\bar f}}\left(\frac{Q^2}{x_a x_b S},y,\theta,{Q\over \mu}\right), \label{convolutionHQ} \end{eqnarray} where $S=(p_a+p_b)^2$, $Q^2=(p_1+p_2)^2$, and we sum over the two main production partonic subprocesses, $q{\bar q} \rightarrow Q {\bar Q}$ and $gg \rightarrow Q {\bar Q}$. We note that the short-distance hard scattering $H_{f {\bar f}}$ is a smooth function only away from the edges of partonic phase space. By replacing the incoming hadrons by partons in Eq.~(\ref{convolutionHQ}), we may write the infrared regularized partonic scattering cross section, after integrating over the total rapidity of the heavy quark pair, as \begin{eqnarray} \frac{d\sigma_{f{\bar f}\rightarrow Q{\bar Q}}} {dQ^2 \; d\cos \theta} &=& \int_\tau^1 dz \, \int \frac {dx_a}{x_a} \, \frac{dx_b}{x_b} \, \phi_{f/f}(x_a,\mu^2,\epsilon) \, \nonumber \\ && \hspace{-5mm} \times \; \phi_{{\bar f}/{\bar f}}(x_b,\mu^2,\epsilon) \quad \delta\left (z-{Q^2\over x_ax_bS}\right) \, \nonumber \\ && \hspace{-10mm} \times \, {\hat \sigma}_{f{\bar f}\rightarrow Q{\bar Q}}\left(1-z, \frac{Q}{\mu},\theta,\alpha_s(\mu^2)\right), \label{convpart} \end{eqnarray} where the argument $\epsilon$ represents the universal collinear singularities, and we have introduced a simplified hard-scattering function, ${\hat \sigma}_{f{\bar f}\rightarrow Q{\bar Q}}$. The threshold for the partonic subprocess is given in terms of the variable $z$, \begin{equation} z\equiv \frac{Q^2}{s} \, , \end{equation} with $s=x_a x_b S$ the invariant mass squared of the incoming partons. At partonic threshold, $z_{\rm max}=1$, there is just enough partonic energy to produce the observed final state. We also define a variable $\tau\equiv z_{\rm min}=Q^2/S$. In general, ${\hat \sigma}$ includes ``plus'' distributions with respect to $1-z$, which arise from incomplete cancellations near threshold between diagrams with real gluon emission and with virtual gluon corrections, with singularities at $n$th order in $\alpha_s$ of the type \begin{equation} -\frac{\alpha_s^n}{n!}\, \left[\frac{\ln^{m}(1-z)}{1-z} \right]_{+}, \; m\le 2n-1\, . \end{equation} If we take Mellin transforms (moments with respect to the variable $\tau$) of Eq.~(\ref{convpart}), the convolution becomes a simple product of the moments of the parton distributions and the hard scattering function ${\hat \sigma}$: \begin{eqnarray} && \hspace{-10mm}\int_0^1 d\tau \, \tau^{N-1} \frac{d\sigma_{f{\bar f}\rightarrow Q{\bar Q}}} {dQ^2 \; d\cos \theta}={\tilde \phi}_{f/f}(N,\mu^2,\epsilon) \nonumber \\ && \hspace{-10mm}\times {\tilde \phi}_{{\bar f}/{\bar f}}(N,\mu^2,\epsilon) \; {\hat \sigma}_{f{\bar f}\rightarrow Q{\bar Q}}(N, Q/\mu,\theta,\alpha_s(\mu^2)) \label{sigmom} \end{eqnarray} with the moments of the hard-scattering function and the parton distributions defined, respectively, by \begin{eqnarray} \tilde{\sigma}(N)=\int_0^1dz\; z^{N-1}\hat{\sigma}(z) \, , \nonumber \\ \tilde{\phi}(N)=\int_0^1dx\; x^{N-1}\phi(x) \, . \label{momphisigma} \end{eqnarray} We then factorize the initial-state collinear divergences into the parton distribution functions, expanded to the same order in $\alpha_s$ as the partonic cross section, and thus obtain the perturbative expansion for the infrared-safe hard scattering function, ${\hat \sigma}$. We note that under moments divergent distributions in $1-z$ produce powers of $\ln N$: \begin{eqnarray} \int_0^1 dz\, z^{N-1}\left[{\ln^m(1-z)\over 1-z}\right]_+ &=&{(-1)^{m+1}\over m+1}\ln^{m+1}N \nonumber \\ && \hspace{-10mm} +{\cal O}\left(\ln^{m-1}N\right)\, . \end{eqnarray} The hard scattering function ${\hat \sigma}$ is sensitive to soft-gluon dynamics through its $1-z$ dependence (or the $N$ dependence of its moments). We may now refactorize (moments of) the cross section into $N$-independent hard components $H_{IL}$, which describe the truly short-distance hard scattering, center-of-mass distributions $\psi$, associated with gluons collinear to the incoming partons, and a soft gluon function $S_{LI}$ associated with non-collinear soft gluons. Note that the mass of the heavy quarks eliminates final state collinear singularities in heavy quark production. This factorization is shown in Fig.~1 where $f{\bar f}$ represents a pair of light quarks or gluons that produce a heavy quark pair. Fig. 2 defines the soft gluon function, $S_{LI}$, which describes the coupling of soft gluons to the partons, represented by eikonal lines, in the hard scattering. $I$ and $L$ are color indices that describe the color structure of the hard scattering. The hard-scattering function takes contributions from the amplitude and its complex conjugate, $H_{IL} ={h^*}_{L}\;h_{I}$. \EPSFIGURE{fig1.ps,width=4.05in} {Factorization for heavy quark production near partonic threshold.} \EPSFIGURE{fig2.ps,width=4.05in} {The soft-gluon function $S_{LI}$ in which the vertices $c_I,c_L^*$ link eikonal lines that represent the incoming and outgoing partons.} Then we write the refactorized cross section as~\cite{KS} \begin{eqnarray} && \hspace{-4mm} \int_0^1 d\tau \tau^{N-1} \frac{d\sigma_{f{\bar f}\rightarrow Q{\bar Q}}} {dQ^2 \, d\cos \theta} =\sum_{IL} H_{IL}\left({Q\over\mu},\theta, \alpha_s(\mu^2)\right) \nonumber \\ && \quad \quad \times \; {\tilde S}_{LI} \left({Q\over N \mu },\theta, \alpha_s(\mu^2) \right) \nonumber \\ && \quad \quad \times \; {\tilde\psi}_{f/f}\left ( N,{Q\over \mu },\epsilon \right) \; {\tilde\psi}_{{\bar f}/{\bar f}}\left (N,{Q\over \mu },\epsilon \right ) . \label{sigref} \end{eqnarray} The center-of-mass distribution functions $\psi$ absorb the universal collinear singularities of the initial-state partons in the refactorized cross section. They differ from the standard light-cone parton distributions $\phi$ by being defined at fixed energy rather than light-like momentum fraction. The moments of $\psi$ can then be written as products of moments of $\phi$ and an infrared safe function. \subsection{Resummation} Now, comparing Eqs.\ (\ref{sigmom}) and (\ref{sigref}), we see that the moments of the partonic heavy-quark production cross section are given by \begin{eqnarray} &&{\tilde \sigma}_{f{\bar f}\rightarrow Q{\bar Q}}(N)= \left[{{\tilde\psi}_{f/f}(N,Q/\mu,\epsilon) \over{\tilde \phi}_{f/f}(N,\mu^2,\epsilon)}\right]^2\, \nonumber \\ && \hspace{-2mm} \times \sum_{IL} H_{IL}\left(\frac{Q}{\mu},\theta,\alpha_s(\mu^2)\right) {\tilde S}_{LI}\left(\frac{Q}{N\mu},\theta,\alpha_s(\mu^2)\right) \nonumber \\ && \label{psiphiHS} \end{eqnarray} where $f{\bar f}$ denotes $q{\bar q}$ or $gg$, and we have used the relations ${\psi}_{q/q}={\psi}_{{\bar q}/{\bar q}}$ and ${\phi}_{q/q}={\phi}_{{\bar q}/{\bar q}}$. We will now discuss the exponentiation of logarithms of $N$ for each factor in the above equation. The first factor, $(\tilde \psi_{f/f}/\tilde \phi_{f/f})^2$, in Eq.~(\ref{psiphiHS}) is universal between electroweak and QCD hard processes, and was computed first for the Drell-Yan cross section \cite{St87}. The resummed expression for this factor is \begin{eqnarray} &&\frac{{\tilde{\psi}}_{f/f}(N,Q/\mu,\epsilon)} {{\tilde{\phi}}_{f/f}(N,\mu^2,\epsilon)} = R_{(f)}\, \exp \left[E^{(f)}(N,Q^2)\right] \nonumber\\ && \quad \times \exp \left\{-2\int_{\mu}^Q \frac{d\mu'}{\mu'}\; \left [\gamma_f(\alpha_s(\mu'{}^2))\right. \right. \nonumber\\ && \hspace{25mm} \left. \left. -\gamma_{ff}(N,\alpha_s(\mu'{}^2)) \right]\right\}\, , \label{psiphimu} \end{eqnarray} where $\gamma_f$ is the anomalous dimension of the field of flavor $f$, which is independent of $N$, and $\gamma_{ff}$ is the anomalous dimension of the color-diagonal splitting function for flavor $f$. $R_{(f)}(\alpha_s)$ is an $N$-independent function of the coupling, which can be normalized to unity at zeroth order. The exponent $E^{(f)}$ is given by \begin{eqnarray} && \hspace{-8mm} E^{(f)}\left(N,Q^2\right)= -\int^1_0 dz \frac{z^{N-1}-1}{1-z} \nonumber \\ && \hspace{-3mm} \times \left \{\int^{(1-z)^{m_S}}_{(1-z)^2} \frac{d\lambda}{\lambda} A^{(f)}\left[\alpha_s(\lambda Q^2)\right] \right. \nonumber\\ && \quad {}+B^{(f)}\left[\alpha_s((1-z)^{m_s} Q^2)\right] \nonumber\\ && \quad \left. {}+\frac{1}{2}\nu^{(f)}\left[\alpha_s((1-z)^2 Q^2)\right] \right\} \, , \label{Eexp} \end{eqnarray} where the parameter $m_S$ and the resummed coefficients $B^{(f)}$ depend on the factorization scheme. In the DIS and $\overline{\rm MS}$ factorization schemes we have $m_S=1$ and $m_S=0$, respectively. The function $A^{(f)}$ is given by \begin{equation} A^{(f)}(\alpha_s) = C_f\left ( {\alpha_s\over \pi} +\frac{1}{2} K \left({\alpha_s\over \pi}\right)^2\right )\, , \label{Aexp} \end{equation} with $C_f=C_F=(N_c^2-1)/(2N_c)$ for an incoming quark, and $C_f=C_A=N_c$ for an incoming gluon, with $N_c$ the number of colors, and $K=C_A(67/18-\pi^2/6)-5n_f/9$, where $n_f$ is the number of quark flavors. $B^{(f)}$ is given for quarks in the DIS scheme by $B^{(q)}(\alpha_s)=-3C_F\alpha_s/(4\pi)$, while it vanishes in the $\overline {\rm MS}$ scheme for quarks and gluons. Finally, the lowest-order approximation to the scheme-independent $\nu^{(f)}$ is $\nu^{(f)}=2C_f \alpha_s/\pi$. Next, we discuss resummation for the soft function. The soft matrix $S_{LI}$ depends on $N$ through the ratio $Q/(N\mu)$; its $N$-dependence can then be resummed by renormalization group analysis. The product $H_{IL}S_{LI}$ of the soft function and the hard factors needs no overall renormalization, because the UV divergences of $S_{LI}$ are balanced by construction by those of $H_{IL}$. Thus, we have~\cite{KS} \begin{eqnarray} H^{(0)}_{IL}&=& \prod_{i=f,\bar f} Z_i^{-1}\; \left(Z_S^{-1}\right)_{IC} H_{CD} \left[\left(Z_S^\dagger \right)^{-1}\right]_{DL} \, , \nonumber \\ S^{(0)}_{LI}&=&(Z_S^\dagger)_{LB}S_{BA}Z_{S,AI}, \label{HSren} \end{eqnarray} where $H^{(0)}$ and $S^{(0)}$ denote the unrenormalized quantities, $Z_i$ is the renormalization constant of the $i$th incoming partonic field external to $h_I$, and $Z_{S,LI}$ is a matrix of renormalization constants which describes the renormalization of the soft function, including mixing of color structures, and which is defined to include the wave function renormalization for the heavy quarks. From Eq.\ (\ref{HSren}), the soft function $S_{LI}$ satisfies the renormalization group equation \begin{eqnarray} \left(\mu {\partial \over \partial \mu}+\beta(g){\partial \over \partial g} \right)\,S_{LI}&=&-(\Gamma^\dagger_S)_{LB}S_{BI} \nonumber \\ && \hspace{-8mm} -S_{LA}(\Gamma_S)_{AI}\, , \label{RGE} \end{eqnarray} where $\Gamma_S$ is an anomalous dimension matrix that is calculated by explicit renormalization of the soft function. In a minimal subtraction renormalization scheme and with $\epsilon=4-n$, where $n$ is the number of space-time dimensions, the matrix of anomalous dimensions at one loop is given by \begin{equation} \Gamma_S (g)=-\frac{g}{2} \frac {\partial}{\partial g}{\rm Res}_{\epsilon \rightarrow 0} Z_S (g, \epsilon) \, . \end{equation} Using Eqs.~(\ref{psiphiHS}), (\ref{psiphimu}), and the solution of the renormalization group equation for the soft function, we can write the resummed heavy quark cross section in moment space as \begin{eqnarray} && \hspace{-3mm}\tilde{{\sigma}}_{f{\bar f}\rightarrow Q{\bar Q}}(N) =R_{(f)}^2 \; \exp \left \{2 \left[ E^{(f_i)}(N,Q^2) \right. \right. \nonumber\\&& \hspace{-5mm} \left. \left. {}-2\int_\mu^Q{d\mu'\over\mu'} \left[\gamma_{f_i}(\alpha_s(\mu'{}^2))-\gamma_{f_if_i}(N,\alpha_s(\mu'{}^2)) \right] \right] \right\} \nonumber\\ && \hspace{-3mm} \quad \times \; {\rm Tr} \left \{ H\left({Q\over\mu},\theta,\alpha_s(\mu^2)\right) \right. \nonumber\\ && \hspace{-3mm} \quad \times \; \bar{P} \exp \left[\int_\mu^{Q/N} {d\mu' \over \mu'} \; \Gamma_S^\dagger\left(\alpha_s(\mu'^2)\right)\right] \nonumber\\ && \hspace{-3mm} \quad \times \; {\tilde S} \left(1,\theta,\alpha_s(Q^2/N^2) \right) \; \nonumber\\ && \hspace{-3mm} \quad \left. \times \; P \exp \left[\int_\mu^{Q/N} {d\mu' \over \mu'}\; \Gamma_S \left(\alpha_s(\mu'^2)\right)\right] \right\}\, , \label{resHQ} \end{eqnarray} where the symbols $P$ and $\bar{P}$ refer to path ordering and the trace is taken in color space. We may simplify this result by choosing a color basis in which the anomalous dimension matrix $\Gamma_S$ is diagonal, with eigenvalues $\lambda_I$ for each basis color tensor labelled by $I$. Then, we have \begin{eqnarray} && \hspace{-5mm} {\tilde S}_{LI}\left(\frac{Q}{N\mu}, \theta, \alpha_s(\mu^2)\right)= {\tilde S}_{LI}\left(1,\theta, \alpha_s\left(\frac{Q^2}{N^2}\right)\right) \nonumber \\ && \hspace{-5mm} \times \, \exp\left[-\int^{\mu}_{Q/N}\frac{d \bar{\mu}}{\bar{\mu}} [\lambda_I(\alpha_s(\bar{\mu}^2)) +\lambda^*_L(\alpha_s(\bar{\mu}^2))]\right]\, . \nonumber \\ \label{rgedgsol} \end{eqnarray} Thus, in a diagonal basis, and with $\mu=Q$ and $R_{(f)}$ normalized to unity, we can rewrite the resummed cross section in a simplified form as \begin{eqnarray} && \hspace{-3mm}\tilde{{\sigma}}_{f{\bar f}\rightarrow Q{\bar Q}}(N) = H_{IL}\left(1,\theta,\alpha_s(\mu^2)\right) \nonumber \\ && \hspace{-3mm} \times {\tilde S}_{LI} \left(1,\theta,\alpha_s(Q^2/N^2) \right) \, \exp \left[E_{LI}^{(f {\bar f})}(N,\theta,Q^2)\right] \, , \nonumber \\ && \label{crossdiag} \end{eqnarray} where the exponent is \begin{eqnarray} && \hspace{-5mm} E_{LI}^{(f {\bar f})}(N,\theta,Q^2)=-\int_0^1 dz \frac{z^{N-1}-1}{1-z} \nonumber \\ && \times \, \left\{\int^{(1-z)^{m_S}}_{(1-z)^2} \frac{d\lambda}{\lambda} g_1^{(f{\bar f})}[\alpha_s(\lambda Q^2)] \right. \nonumber \\ && \quad \quad {}+g_2^{(f{\bar f})}[\alpha_s((1-z)^{m_S} Q^2)] \nonumber \\ && \quad \quad \left. {}+g_3^{(IL,f{\bar f})}[\alpha_s((1-z)^2 Q^2),\theta] \right\}\, , \nonumber \\ \label{ELI} \end{eqnarray} with the functions $g_1$, $g_2$, and $g_3$ defined by \begin{eqnarray} &&g_1^{(f\bar f)}=A^{(f)}+A^{(\bar f)} \, , \quad g_2^{(f \bar f)}=B^{(f)}+B^{(\bar f)} \, , \nonumber \\ && g_3^{(IL,f {\bar f})}=-\lambda_I-\lambda_L^*+\frac{1}{2}\nu^{(f)} +\frac{1}{2}\nu^{(\bar f)} \, . \label{g1g2g3} \end{eqnarray} In the next section we present the soft anomalous dimension matrices for heavy quark production through light quark annihilation and gluon fusion; also we give NLO expansions of the resummed cross section in both partonic production channels. \section{Soft anomalous dimension matrices for heavy quark production} \subsection{Soft anomalous dimension for $q {\bar q} \rightarrow Q {\bar Q}$} First, we present the soft anomalous dimension matrix for heavy quark production through light quark annihilation, \begin{equation} q(p_a,r_a)+{\bar q}(p_b,r_b) \rightarrow {\bar Q}(p_1,r_1) + Q(p_2,r_2)\, , \end{equation} where the $p_i$'s and $r_i$'s denote momenta and colors of the partons in the process. We introduce the Mandelstam invariants \begin{eqnarray} s&=&(p_a+p_b)^2 \, , \nonumber \\ t_1&=&(p_a-p_1)^2-m^2 \, , \nonumber \\ u_1&=&(p_b-p_1)^2-m^2 \, , \label{Mandelstam} \end{eqnarray} with $m$ the heavy quark mass, which satisfy $s+t_1+u_1=0$ at partonic threshold. \EPSFIGURE{fig3.ps,width=4.05in} {Eikonal vertex corrections to $S_{LI}$ for partonic subprocesses in heavy quark or dijet production.} \EPSFIGURE{fig4.ps,width=4.05in} {Eikonal self-energy graphs for heavy quark production.} For the determination of $\Gamma_S$ we evaluate the one-loop graphs in Figs. 3 and 4. For the $q {\bar q} \rightarrow Q {\bar Q}$ partonic subprocess $\Gamma_S$ is a $2 \times 2$ matrix, since the process can be described by a basis of two color tensors. Here we calculate the soft matrix in the $s$-channel singlet-octet basis: \begin{eqnarray} c_1 &=& c_{\rm singlet}=\delta_{r_ar_b}\delta_{r_1r_2}\, , \quad \quad \nonumber \\ c_2 &=& c_{\rm octet}=(T^c_F)_{r_b r_a}(T^c_F)_{r_2 r_1}\, , \label{18basis} \end{eqnarray} where the $T_F^c$ are the generators of $SU(3)$ in the fundamental representation. The result of our calculation in a general axial gauge is~\cite{NK,Thesis,KS} \begin{equation} \Gamma_S=\left[\begin{array}{cc} \Gamma_{11} & \Gamma_{12} \\ \Gamma_{21} & \Gamma_{22} \end{array} \right] \, , \label{matrixqqQQ} \end{equation} with \begin{eqnarray} \Gamma_{11}&=&-\frac{\alpha_s}{\pi}C_F \, [L_{\beta} +\ln(2\sqrt{\nu_a\nu_b})+\pi i] \, , \nonumber \\ \Gamma_{21}&=&\frac{2\alpha_s}{\pi} \ln\left(\frac{u_1}{t_1}\right) \, , \nonumber \\ \Gamma_{12}&=&\frac{\alpha_s}{\pi} \frac{C_F}{C_A} \ln\left(\frac{u_1}{t_1}\right) \, , \nonumber \\ \Gamma_{22}&=&\frac{\alpha_s}{\pi} \nonumber \\ && \hspace{-10mm} \times \left\{C_F\left[4\ln\left(\frac{u_1}{t_1}\right) -\ln(2\sqrt{\nu_a\nu_b}) -L_{\beta}-\pi i\right]\right. \nonumber \\ && \hspace{-12mm} \left.{}+\frac{C_A}{2}\left[-3\ln\left(\frac{u_1}{t_1}\right) -\ln\left(\frac{m^2s}{t_1u_1}\right)+L_{\beta}+\pi i \right]\right\}\, , \nonumber \\ \label{GammaqqQQ} \end{eqnarray} where $L_\beta$ is the velocity-dependent eikonal function \begin{equation} L_{\beta}=\frac{1-2m^2/s}{\beta}\left(\ln\frac{1-\beta}{1+\beta} +\pi i \right)\, , \end{equation} with $\beta=\sqrt{1-4m^2/s}$. The gauge dependence of the incoming eikonal lines is given in terms of $\nu_i \equiv (v_i \cdot n)^2/|n|^2$, with $n^{\mu}$ the gauge vector, which cancels against corresponding terms in the parton distributions. Here the dimensionless velocity vectors $v_i^{\mu}$ are defined by $p_i^{\mu}=Q v_i^{\mu}/\sqrt{2}$ and obey $v_i^2=0$ for the light incoming quarks and $v_i^2=2m^2/Q^2$ for the outgoing heavy quarks. The gauge dependence of the outgoing heavy quarks is cancelled by the inclusion of the self-energy diagrams in Fig. 4. $\Gamma_S$ is diagonalized in this singlet-octet color basis at absolute threshold, $\beta=0$. It is also diagonalized for arbitrary $\beta$ when the parton-parton c.m. scattering angle is $\theta=90^\circ$ (where $u_1=t_1$), with eigenvalues that may be read off from Eq.\ (\ref{GammaqqQQ}). We can expand the resummed heavy quark production cross section to any fixed order in perturbation theory without having to diagonalize the soft anomalous dimension matrix. The NLO expansion for $q {\bar q} \rightarrow Q {\bar Q}$ in the $\overline{\rm MS}$ scheme is \begin{eqnarray} && \hspace{-3mm} {\hat \sigma}^{\overline{\rm MS} \, (1)}_{q{\bar q}\rightarrow Q{\bar Q}} (1-z,m^2,s,t_1,u_1)=\sigma^B_{q{\bar q}\rightarrow Q{\bar Q}} \frac{\alpha_s}{\pi} \nonumber \\ && \hspace{-3mm} \times \left\{4C_F\left[\frac{\ln(1-z)}{1-z}\right]_{+} +\left[\frac{1}{1-z}\right]_{+} \right. \nonumber \\ && \hspace{-3mm} \times \left[C_F\left(8\ln\left(\frac{u_1}{t_1}\right) -2-2 \, {\rm Re} \, L_{\beta}+2\ln\left(\frac{s}{\mu^2}\right)\right)\right. \nonumber \\ && \hspace{-5mm} \left.\left. {}+C_A\left(-3\ln\left(\frac{u_1}{t_1}\right)+{\rm Re} \, L_{\beta} -\ln\left(\frac{m^2s}{t_1 u_1}\right)\right)\right]\right\}\, , \nonumber \\ \end{eqnarray} where $\sigma^B_{q{\bar q}\rightarrow Q{\bar Q}}$ is the Born cross section. Similar results have been obtained in the DIS scheme. These results are consistent with the approximate one-loop results in Ref.~\cite{mengetal}. Analytical and numerical results have also been obtained for the expansions at NNLO \cite{NK,KLMV} in both pair inclusive and single-particle inclusive kinematics. \subsection{Soft anomalous dimension for $gg \rightarrow Q {\bar Q}$} Next, we present the soft anomalous dimension matrix for heavy quark production through gluon fusion, \begin{equation} g(p_a,r_a)+g(p_b,r_b) \rightarrow {\bar Q}(p_1,r_1) + Q(p_2,r_2)\, , \end{equation} with momenta and colors labelled by the $p_i$'s and $r_i$'s, respectively. Here $\Gamma_S$ is a $3 \times 3$ matrix since the process can be described by a basis of three color tensors. The relevant graphs for the calculation of $\Gamma_S$ are the same as in Figs.~3 and 4, where now the incoming eikonal lines represent gluons. We make the following choice for the color basis: \begin{eqnarray} c_1&=&\delta_{r_ar_b}\,\delta_{r_2r_1} \, , \nonumber \\ c_2&=&d^{r_ar_bc}\,(T^c_F)_{r_2r_1} \, , \nonumber \\ c_3&=&i f^{r_ar_bc}\,(T^c_F)_{r_2r_1} \, , \end{eqnarray} where $d^{abc}$ and $f^{abc}$ are the totally symmetric and antisymmetric $SU(3)$ invariant tensors, respectively. Then the anomalous dimension matrix in a general axial gauge is given by~\cite{NK,Thesis,KS} \begin{equation} \Gamma_S=\left[\begin{array}{ccc} \Gamma_{11} & 0 & \frac{\Gamma_{31}}{2} \vspace{2mm} \\ 0 & \Gamma_{22} & \frac{N_c}{4} \Gamma_{31} \vspace{2mm} \\ \Gamma_{31} & \frac{N_c^2-4}{4N_c}\Gamma_{31} & \Gamma_{22} \end{array} \right] \, , \label{GammaggQQ33} \end{equation} where \begin{eqnarray} && \Gamma_{11}=\frac{\alpha_s}{\pi}\left [-C_F(L_{\beta}+1) \right. \nonumber \\ && \hspace{15mm} \left. {}+C_A\left(-\frac{1}{2}\ln\left({4\nu_a\nu_b}\right) +1-\pi i\right)\right ], \nonumber \\ && \Gamma_{31}=\frac{\alpha_s}{\pi}\ln\left(\frac{u_1^2}{t_1^2}\right) \, , \nonumber \\ && \Gamma_{22}=\frac{\alpha_s}{\pi}\left\{-C_F(L_{\beta}+1) +\frac{C_A}{2} \right. \nonumber \\ && \left. \times \left[-\ln\left(4\nu_a\nu_b\right) +2+\ln\left(\frac{t_1 u_1}{m^2 s}\right)+L_{\beta}-\pi i \right]\right\}. \nonumber \\ \label{GammaggQQ} \end{eqnarray} We note that $\Gamma_S$ is diagonalized in this basis at absolute threshold, $\beta=0$, and also for arbitrary $\beta$ when the parton-parton c.m. scattering angle is $\theta=90^\circ$ , with eigenvalues that may be read off from Eq.\ (\ref{GammaggQQ}). Again, we may expand the resummed cross section for $gg \rightarrow Q {\bar Q}$ at NLO and NNLO or higher. The NLO expansion in the ${\overline{\rm MS}}$ scheme is \begin{eqnarray} && \hspace{-3mm} {\hat \sigma}^{\overline {\rm MS} \, (1)}_{gg \rightarrow Q{\bar Q}} (1-z,m^2,s,t_1,u_1)= \sigma^B_{gg\rightarrow Q{\bar Q}} \frac{\alpha_s}{\pi} \nonumber \\ && \hspace{-4mm}\times \left\{4C_A\left[\frac{\ln(1-z)}{1-z}\right]_{+} -2C_A \ln\left(\frac{\mu^2}{s}\right) \left[\frac{1}{1-z}\right]_{+}\right\} \nonumber \\ && \hspace{-4mm} {}+\alpha_s^3 K_{gg} B_{QED} \left[\frac{1}{1-z}\right]_{+} \left\{N_c(N_c^2-1)\frac{(t_1^2+u_1^2)}{s^2} \right. \nonumber \\ && \hspace{-4mm} \times \left[\left(-C_F+\frac{C_A}{2}\right) {\rm Re} \, L_{\beta} +\frac{C_A}{2}\ln\left(\frac{t_1u_1}{m^2s}\right) -C_F\right] \nonumber \\ && \hspace{-4mm} {}+\frac{N_c^2-1}{N_c}(C_F-C_A) {\rm Re} \, L_{\beta} -(N_c^2-1)\ln\left(\frac{t_1u_1}{m^2s}\right) \nonumber \\ && \hspace{-4mm} \left. {}+C_F \frac{N_c^2-1}{N_c} +\frac{N_c^2}{2}(N_c^2-1) \ln\left(\frac{u_1}{t_1}\right)\frac{(t_1^2-u_1^2)}{s^2} \right\} \nonumber\\&& \label{gg1loop} \end{eqnarray} where $\sigma^B_{gg\rightarrow Q{\bar Q}}$ is the Born cross section, \begin{equation} B_{\rm QED}=\frac{t_1}{u_1}+\frac{u_1}{t_1}+\frac{4m^2s}{t_1u_1} \left(1-\frac{m^2s}{t_1u_1}\right), \end{equation} and $K_{gg}=(N^2-1)^{-2}$ is a color average factor. \section{Diagonalization and numerical results for $q {\bar q} \rightarrow t{\bar t}$} In this section we give some numerical results for top quark production at the Tevatron. As we noted in the previous section, the soft anomalous dimension matrices, $\Gamma_S$, are in general not diagonal. They are only diagonal at specific kinematical regions, i.e. at absolute threshold, $\beta=0$, and at a scattering angle $\theta=90^{\circ}$. In these cases the exponentiated cross section has a simpler form, Eq.~(\ref{crossdiag}), and numerical studies are easier to perform. Thus, the resummed cross section at $\theta=90^{\circ}$ for top quark production at the Fermilab Tevatron and bottom quark production at HERA-B were presented in Ref.~\cite{NKJSRV}. In general, however, we must diagonalize $\Gamma_S$, i.e. we must find new color bases where $\Gamma_S$ is diagonal so that the resummed cross section can take the simpler form of Eq. (\ref{crossdiag}). The diagonal basis is given by $C'=CR$ where the columns of the matrix $R$ are the eigenvectors of $\Gamma_S$. A detailed discussion of the diagonalization procedure has been given in Refs.~\cite{NK,NKRV}. The resummed partonic cross section for the process $q {\bar q} \rightarrow Q {\bar Q}$ can be written as ~\cite{NKRV} \begin{eqnarray} && \hspace{-4mm} \sigma^{\rm res}_{q \bar q}(s,m^2)=\sum_{i,j=1}^2\int_{-1}^1 d\cos \theta \, \nonumber \\ && \hspace{-4mm} \times \left[-\int^{s-2ms^{1/2}}_{s_{\rm cut}} ds_4 \, f_{q \bar q, ij}(s_4, \theta) \frac{d{\overline \sigma}_{q \bar q, ij}^{(0)}(s,s_4,\theta)}{ds_4}\right] \nonumber \\ \label{respart} \end{eqnarray} where $s_4=s+t_1+u_1$. The $d{\overline \sigma}_{q \bar q, ij}^{(0)}(s,s_4,\theta)/ds_4$ are color components of the differential of the Born cross section. The function $f_{q \bar q, ij}$ is given at NLL by the exponential \begin{equation} f_{q \bar q, ij}=\exp[E_{q \bar q, ij}] =\exp[E_{q \bar q}+E_{q \bar q}(\lambda_i, \lambda_j)] \, , \end{equation} where in the DIS scheme \begin{eqnarray} && E_{q\overline q}^{\rm DIS} = \int_{\omega_0}^1\frac{d\omega'}{\omega'} \Big\{\int_{\omega'^2 Q^2/\Lambda^2}^{\omega' Q^2/\Lambda^2} \frac{d\xi}{\xi}\, \Big[ \frac{2 C_F}{\pi} \Big( \alpha_s(\xi) \nonumber \\ && \quad {}+ \frac{1}{2\pi} \alpha^2_s(\xi) K\Big) \Big] - \frac{3}{2} \frac{C_F}{\pi} \alpha_s \Big( \frac{\omega' Q^2}{\Lambda^2}\Big) \, \Big\} \, , \label{Eofm} \end{eqnarray} with $\omega_0=s_4/(2m^2)$ and $\Lambda$ the QCD scale parameter. The color-dependent contribution in the exponent is \begin{eqnarray} E_{q\overline q}(\lambda_i, \lambda_j)&=& -\int_{\omega_0}^1\frac{d\omega'}{\omega'} \left\{\lambda_i' \left[\alpha_s \left(\frac{\omega'^2 Q^2} {\Lambda^2}\right), \theta \right] \right. \nonumber \\ && \hspace{-2mm} \left. +{\lambda_j'}^* \left[\alpha_s \left(\frac{\omega'^2 Q^2}{\Lambda^2}\right), \theta \right] \right\} \, , \end{eqnarray} in both mass factorization schemes, where $i,j= 1,2$. Here $\lambda'=\lambda-\nu^{(f)}/2$ (see Eq.~(\ref{g1g2g3})) where the $\lambda$'s are the eigenvalues of the soft anomalous dimension matrix. The cutoff $s_{\rm cut}$ in Eq.~(\ref{respart}) regulates the divergence of $\alpha_s$ at low $s_4$: $s_{\rm cut}>s_{4,{\rm min}}=2m^2\Lambda/Q$. We choose a value of the cutoff consistent with the sum of the first few terms in the perturbative expansion \cite{NKJSRV,LSN,HERAB}, in the range $30 s_{4, {\rm min}} < s_{\rm cut} < 40 s_{4, {\rm min}}$, which corresponds to a soft gluon energy cutoff of the order of the decay width of the top, thus giving a natural boundary of the nonperturbative region (see also the discussion in \cite{BC}). The central value in our cutoff range, $s_{\rm cut} = 35 s_{4, {\rm min}}$, corresponds to $s_4/(2m^2)=0.04$ for $m=175$ GeV/$c^2$ and $\Lambda=0.2$ GeV. After the diagonalization of the soft anomalous dimension matrix, the resummed partonic cross section is given at NLL accuracy by \cite{NKRV} \begin{eqnarray} && \sigma^{\rm res}_{q \overline q}(s,m^2) = -\int_{-1}^1 d\cos \theta \, \int^{s-2ms^{1/2}}_{s_{\rm cut}} ds_4 \; \nonumber \\ && \times \frac{1}{|\lambda_1-\lambda_2|^2} \frac{d{\overline \sigma}_{q \overline q}^{(0)}(s,s_4,\theta)}{ds_4} \nonumber \\ && \times \left[\left(\frac{4N_c^2}{N_c^2-1}\Gamma_{12}^2 +|\lambda_1-\Gamma_{11}|^2 \right) e^{E_{q \overline q, 11}} \right. \nonumber \\ && {}+\left(\frac{4N_c^2}{N_c^2-1}\Gamma_{12}^2 +|\lambda_2-\Gamma_{11}|^2 \right) e^{E_{q \overline q, 22}} \nonumber \\ && {}-\frac{8N_c^2}{N_c^2-1}\Gamma_{12}^2 {\rm Re}\left(e^{E_{q \overline q, 12}}\right) \nonumber \\ && \left. {}-2{\rm Re} \left((\lambda_1-\Gamma_{11})(\lambda_2-\Gamma_{11})^* e^{E_{q \overline q, 12}}\right)\right]. \label{partres} \end{eqnarray} To calculate the NLL resummed hadronic cross section we convolute the parton distributions $\phi_{i/h}$, for parton $i$ in hadron $h$, with the partonic cross section: \begin{eqnarray} && \sigma^{\rm res}_{q \overline q, {\rm had}}(S,m^2)=\sum_{q=u}^b \int_{\tau_0}^1 d\tau \int_\tau^1 \frac{dx}{x} \phi_{q/h_1}(x,\mu^2) \nonumber \\ && \quad \quad \times \, \phi_{{\bar q}/h_2}(\frac{\tau}{x},\mu^2) \; \sigma_{q \overline q}^{\rm res}(\tau S, m^2) \, , \label{hadres} \end{eqnarray} where $\sigma_{q \overline q}^{\rm res}(\tau S, m^2)$ is defined in Eq.~(\ref{partres}) and $\tau_0 = (m + \sqrt{m^2 + s_{\rm cut}})^2/S$. In Fig. 5 we show numerical results for the $t \overline t$ production cross section at the Fermilab Tevatron with $\sqrt S=1.8$ TeV. We use the CTEQ 4D DIS parton densities \cite{CTEQ}. The NLO exact cross sections, including the factorization scale dependence, are shown as functions of the top quark mass. We also plot the NLO approximate cross section, i.e. the one-loop expansion of the resummed cross section, calculated with $s_{\rm cut}=0$ and $\mu^2=m^2$. We note the excellent agreement between the NLO exact and approximate cross sections. \EPSFIGURE{fig5.ps,width=5.05in} {The NLO exact and approximate and the NLL improved hadronic $t \overline t$ production cross sections in the $q \overline q$ channel and the DIS scheme are given as functions of top quark mass for $p \overline p$ collisions at the Tevatron energy, $\sqrt{S} = 1.8$ TeV. The NLO exact cross section is given for $\mu^2 = m^2$ (solid curve), $4m^2$ (lower-dashed) and $m^2/4$ (upper-dashed). The NLO approximate cross section with $s_{\rm cut} = 0$ is shown for $\mu^2 = m^2$ (lower dot-dashed). The NLL improved cross section, Eq.~(\ref{improved}), is given for $s_{\rm cut} = 35 s_{4, {\rm min}}$ (upper dot-dashed), $30 s_{4, {\rm min}}$ (upper-dotted) and $40s_{4, {\rm min}}$ (lower-dotted).} The scale dependence of the NLL resummed cross section is significantly reduced relative to that of the NLO cross section. In order to match our results to the exact NLO cross section we define the NLL improved cross section \begin{equation} \sigma_{q \overline q, {\rm had}}^{\rm imp} = \sigma_{q \overline q, {\rm had}}^{\rm res} - \sigma_{q \overline q, {\rm had}}^{\rm NLO, approx} + \sigma_{q \overline q, {\rm had}}^{\rm NLO, exact} \, \, , \label{improved} \end{equation} with the same cut applied to the NLO approximate and the NLL resummed cross sections. We plot the hadronic improved cross section in Fig. 5 for $\mu^2 = m^2$ along with the variation with $s_{\rm cut}$. The variation of the improved cross section with change of cutoff is small over the range $30 s_{4, {\rm min}}< s_{\rm cut} <40 s_{4, {\rm min}}$. At $m = 175$ GeV/$c^2$ and $\sqrt{S} = 1.8$ TeV, the value of the improved cross section for $q \overline q \rightarrow t \overline t$ with $s_{\rm cut}/(2m^2)=0.04$ is 6.0 pb compared to a NLO cross section of 4.5 pb at $\mu=m$. The contribution of the $gg \rightarrow Q {\bar Q}$ channel to top quark production at the Tevatron is relatively small. With its inclusion the total $t {\bar t}$ cross section at the Tevatron is 7 pb. Gluon fusion is much more important for $b$-quark production at HERA-B \cite{HERAB}. \section{Resummation for dijet production} Here we review the resummation formalism for the hadronic production of a pair of jets. Dijet production entails additional complications due to the presence of final state jets; otherwise, the formalism is analogous to that for heavy quark production. \subsection{Factorized dijet cross section} We consider dijet production in hadronic processes \begin{equation} h_a(p_a)+h_b(p_b) \rightarrow J_1(p_1,\delta_1)+J_2(p_2,\delta_2)+X \, , \label{dijet} \end{equation} where $\delta_1$ and $\delta_2$ are the cone angles of the jets. The partonic cross section is infrared safe since the initial-state collinear singularities are factorized into universal parton distribution functions, and the use of cones removes all the final-state collinear singularities. We study inclusive dijet cross sections at fixed rapidity interval \begin{equation} {\Delta}y= (1/2)\ln [(p_1^+\; p_2^-) /(p_1^-\; p_2^+)] \, , \end{equation} and total rapidity \begin{equation} y_{JJ}= (1/2) \ln[(p_1^++p_2^+)/(p_1^-+p_2^-)] \, . \end{equation} To construct the dijet cross sections, we define a large invariant, $M_{JJ}$, which is held fixed. A natural choice is the dijet invariant mass, $M^2_{JJ}=(p_1+p_2)^2$, analogous to $Q^2$ for heavy quark production, but other choices are possible, such as the scalar product of the two jet momenta, $M^2_{JJ}=2p_1 \cdot p_2$. We note that for large $M_{JJ}$ at fixed $\Delta y$ we have a large momentum transfer in the partonic subprocess. The resummed cross section depends critically on the choice for $M_{JJ}$. We shall see that the leading behavior of the resummed cross section is the same as for heavy quark production when $M_{JJ}$ is the dijet invariant mass, while it is different for the other choice. The dijet cross section can be written in factorized form as \begin{eqnarray} &&\frac{d\sigma_{h_ah_b{\rightarrow}J_1J_2}} {dM^2_{JJ}\, dy_{JJ} \, d{\Delta}y}= \sum_{f_a,f_b=q,\overline{q},g} \int \frac{dx_a}{x_a} \, \frac{dx_b}{x_b} \nonumber \\ && \times \, \phi_{f_a/h_a}(x_a,\mu^2) \; \phi_{f_b/h_b}(x_b,\mu^2) \nonumber \\ && \times \, H_{f_af_b}\left(\frac{M_{JJ}^2}{x_ax_bS},y,\Delta y,\frac{M_{JJ}}{\mu}, \alpha_s(\mu^2),\delta_1,\delta_2\right) , \nonumber \\ \label{dijetfact} \end{eqnarray} where $H$ is the hard scattering and the $\phi$'s are parton distribution functions. As for heavy quark production, we may replace the incoming hadrons by partons in the above equation, and integrate over the total rapidity of the jet pair. Then, we can write the infrared regularized partonic cross section, which factorizes as the hadronic cross section, as \begin{eqnarray} && \hspace{-3mm} \frac{d\sigma_{f_af_b{\rightarrow}J_1J_2}}{dM^2_{JJ} \; d{\Delta}y}=\int_{\tau}^1 dz \int \frac{dx_a}{x_a} \, \frac{dx_b}{x_b}\; \nonumber \\ && \hspace{-3mm} \times \, \phi_{f_a/f_a}(x_a,\mu^2,\epsilon) \, \phi_{f_b/f_b}(x_b,\mu^2,\epsilon) \, \delta\left(z-\frac{M^2_{JJ}}{s}\right) \nonumber \\ && \hspace{-3mm} \times \sum_{{\rm f}} \, \hat{\sigma}_{{\rm f}} \left(1-z,\frac{M_{JJ}}{\mu},{\Delta}y,\alpha_s(\mu^2), \delta_1,\delta_2\right) \, , \label{sigpart} \end{eqnarray} where the argument $\epsilon$ represents the universal collinear singularities, and we have introduced a simplified hard-scattering function, ${\hat \sigma}_{{\rm f}}$, where ${{\rm f}}$ denotes the partonic subprocesses. The threshold region is given in terms of the variable $z$, \begin{equation} z\equiv \frac{M^2_{JJ}}{s}\, , \end{equation} where, as before, $s=x_ax_bS$ with $S=(p_a+p_b)^2$. Partonic threshold is at $z_{\rm max}=1$ while the lower limit of $z$ is $z_{\rm min}\equiv\tau=M^2_{JJ}/S$. By taking a Mellin transform of the rapidity-integrated partonic cross section, Eq. (\ref{sigpart}), we reduce the convolution to a simple product of moments, \begin{eqnarray} && \hspace{-5mm} \int_0^1 d\tau\; \tau^{N-1}\; \frac{d\sigma_{f_af_b{\rightarrow}J_1J_2}}{dM^2_{JJ} \; d{\Delta}y} \nonumber \\ && = \sum_{{\rm f}}{\tilde \phi}_{f_a/f_a}(N,\mu^2,\epsilon) \, {\tilde \phi}_{f_b/f_b}(N,\mu^2,\epsilon) \nonumber \\ && \quad \times \, {{\tilde{\sigma}}}_{{\rm f}}(N,M_{JJ}/\mu,\alpha_s(\mu^2),\delta_1, \delta_2)\, , \label{moment} \end{eqnarray} with $\tilde{\sigma}_{{\rm f}}(N)$ and $\tilde{\phi}(N)$ defined as in Eq.~(\ref{momphisigma}). We then factorize the initial-state collinear divergences into the light-cone distribution functions $\phi_{f/f}$, expanded to the same order in $\alpha_s$ as the partonic cross section, and thus obtain the perturbative expansion for the infrared-safe hard scattering function, ${\tilde {\sigma}}_{{\rm f}}$. \EPSFIGURE{fig6.ps,width=5.05in} {Refactorization for dijet production. The soft gluon function is as in Fig. 2.} We can refactorize the cross section, as shown in Fig.~6, into hard components $H_{IL}$, which describe the truly short-distance hard-scattering, center-of-mass distributions $\psi$, associated with gluons collinear to the incoming partons, a soft gluon function $S_{LI}$, associated with non-collinear soft gluons, and jet functions $J_i$, associated with gluons collinear to the outgoing jets. As for heavy quarks, here $I$ and $L$ denote color indices that describe the color structure of the hard scattering. The refactorized cross section may then be written as \begin{eqnarray} && \int_0^1 d\tau\; \tau^{N-1}\; \frac{d\sigma_{f_af_b{\rightarrow}J_1J_2}}{dM^2_{JJ} \; d{\Delta}y} \nonumber\\ && =\sum_{{\rm f}}\sum_{IL} H_{IL}^{({\rm f})}\left({M_{JJ}\over\mu},\Delta y \right) \; {\tilde S}_{LI}^{({\rm f})} \left({M_{JJ}\over N \mu}, \Delta y \right) \nonumber\\ && \quad\times\; {\tilde\psi}_{f_a/f_a}\left(N,{M_{JJ}\over \mu}, \epsilon \right) \; {\tilde\psi}_{f_b/f_b}\left(N,{M_{JJ}\over \mu}, \epsilon \right) \nonumber \\ && \quad \times \; {\tilde J}_{(f_1)}\left(N,{M_{JJ}\over \mu }, \delta_1\right)\; {\tilde J}_{(f_2)}\left(N,{M_{JJ}\over \mu }, \delta_2\right). \nonumber \\ . \label{refact} \end{eqnarray} This refactorization is similar to the heavy quark case except that we now include in addition outgoing jet functions in order to absorb the final state collinear singularities. \subsection{Resummed dijet cross section} By comparing Eqs.\ (\ref{moment}) and (\ref{refact}), we can write the refactorized expression for the Mellin transform of the hard-scattering function ${\tilde {\sigma}}_{{\rm f}}$ as \begin{eqnarray} \tilde{{\sigma}}_{{\rm f}}(N)&=& \left[\frac{ {\tilde{\psi}}_{f_a/f_a}(N,M_{JJ}/\mu) {\tilde{\psi}}_{f_b/f_b}(N,M_{JJ}/\mu )} {{\tilde{\phi}}_{f_a/f_a}(N,\mu^2 ) \, {\tilde{\phi}}_{f_b/f_b}(N,\mu^2)} \right]\nonumber \\ && \hspace{-15mm} \times \, \sum_{IL}H_{IL}^{({\rm f})}\left({M_{JJ}\over\mu},\Delta y \right)\; {\tilde S}_{LI}^{({\rm f})} \left( {M_{JJ}\over N \mu},\Delta y \right) \nonumber \\ && \hspace{-15mm} \times \, {\tilde J}_{(f_1)}\left(N,{M_{JJ}\over \mu}, \delta_1\right)\; {\tilde J}_{(f_2)}\left(N,{M_{JJ}\over \mu}, \delta_2\right). \label{sigfactjet} \end{eqnarray} The results for the resummation of the universal ratio $\psi/\phi$ and the soft gluon function were given in the context of heavy quark production in Section 2. Explicit expressions for the soft anomalous dimension matrices for dijet production will be given in the next section. Thus, the only thing remaining to write down the resummed dijet cross section are expressions for the resummation of the final state jets. The moments of the final-state jet with $M_{JJ}^2=(p_1+p_2)^2$ are given by~\cite{KOS1} \begin{equation} \tilde{J}_{(f_i)}\left(N,{M_{JJ}\over \mu}, \delta_i\right) =\exp \left[E'_{(f_i)}(N,M_{JJ})\right]\, , \label{finaljet} \end{equation} with \begin{equation} E'_{(f_i)}(N,M_{JJ})= \int_\mu^{M_{JJ}/N} {d\mu' \over \mu'}\; \; C'_{(f_i)}(\alpha_s(\mu'{}^2))\, , \label{Epr2exp} \end{equation} where the first term in the series for $C'_{(f_i)}(\alpha_s)$ may be read off from a one-loop calculation. The leading logarithmic behavior of the cross section in this case is not affected by the final state jets, so we always have an enhancement of the cross section at leading logarithm, as is the case for Drell-Yan and heavy quark cross sections. The moments of the final-state jet with $M_{JJ}^2=2p_1\cdot p_2$ are given by Eq.~(\ref{finaljet}) with~\cite{KOS1} \begin{eqnarray} \hspace{-5mm} E'_{(f)}\left(N,M_{JJ}\right) &=& \int^1_0 dz \frac{z^{N-1}-1}{1-z}\; \nonumber\\ && \hspace{-25mm} \times \left \{\int^{(1-z)}_{(1-z)^2} \frac{d\lambda}{\lambda} A^{(f)}\left[\alpha_s(\lambda M_{JJ}^2)\right] \right. \nonumber\\ && \hspace{-25mm} \quad \quad \left. {}+B'_{(f)}\left[\alpha_s((1-z) M_{JJ}^2) \right] \right\}\, , \label{Eprexp} \end{eqnarray} where the function $A^{(f)}$ is the same as in Eq.\ (\ref{Aexp}) and the lowest-order term in $B'_{(f)}$ may be read off from the one-loop jet function. The results include a gauge dependence, which cancels against a corresponding dependence in the soft anomalous dimension matrix. The leading logarithms for final-state jets with $M_{JJ}^2=2p_1\cdot p_2$ are negative and give a suppression to the cross section, in contrast to the initial-state leading-log contributions. Using Eqs.~(\ref{sigfactjet}), (\ref{psiphimu}), and (\ref{finaljet}), together with the evolution of the soft function, we can write the resummed dijet cross section in moment space as \begin{eqnarray} && \hspace{-3mm} \tilde{{\sigma}}_{{\rm f}}(N) = R_{(f)}^2\; \exp \left \{ \sum_{i=a,b} \left[ E^{(f_i)}(N,M_{JJ}) \right. \right. \nonumber\\ && \hspace{-6mm}\left. \left. -2\int_\mu^{M_{JJ}}{d\mu'\over\mu'} \left[\gamma_{f_i}(\alpha_s(\mu'{}^2))-\gamma_{f_if_i}(N,\alpha_s(\mu'{}^2)) \right]\right]\right\} \nonumber \\ && \times\; \exp \left \{\sum_{j=1,2}E'_{(f_j)}(N,M_{JJ}) \right\} \nonumber\\ && \times\; {\rm Tr} \left\{ H^{({\rm f})}\left({M_{JJ}\over\mu},\Delta y,\alpha_s(\mu^2)\right) \right. \nonumber\\ && \times\; \bar{P} \exp \left[\int_\mu^{M_{JJ}/N} {d\mu' \over \mu'}\; \Gamma_S^{({\rm f})}{}^\dagger\left(\alpha_s(\mu'^2)\right)\right] \nonumber\\ && \times\; {\tilde S}^{({\rm f})} \left(1,\Delta y,\alpha_s\left(M_{JJ}^2/N^2\right) \right) \nonumber\\ && \left. \times\; P \exp \left[\int_\mu^{M_{JJ}/N} {d\mu' \over \mu'} \; \Gamma_S^{({\rm f})}\left(\alpha_s(\mu'^2)\right)\right] \right\}. \end{eqnarray} This expression is analogous to Eq.~(\ref{resHQ}) for heavy quark production except for the addition of the exponents for the final-state jets. We give explicit expressions for the soft anomalous dimension matrices, $\Gamma_S$, for the partonic subprocesses in dijet production in the next section. \section{Soft anomalous dimension matrices for dijet production} We consider partonic subprocesses \begin{equation} f_a\left(p_a, r_a \right) + f_b\left(p_b, r_b \right) \rightarrow f_1\left(p_1, r_1 \right) + f_2\left(p_2, r_2 \right) , \end{equation} where the $p_i$'s and $r_i$'s denote momenta and colors of the partons in the process. To facilitate the presentation of the results for $\Gamma_S$ we use the notation \begin{equation} T\equiv \ln\left(\frac{-t}{s}\right)+\pi i, \, \U\equiv \ln\left(\frac{-u}{s}\right)+\pi i, \label{eq:new2form} \end{equation} where \begin{equation} s=\left(p_a+p_b \right)^2, \, t=\left(p_a-p_1 \right)^2, \, u=\left(p_a-p_2 \right)^2, \label{Mandlst} \end{equation} are the usual Mandelstam invariants. We note that for the definition of the color basis we can choose any physical channel $s$, $t$, or $u$. We will use $t$-channel bases for the partonic processes in dijet production except for the processes $q {\bar q}\rightarrow gg$ and $gg \rightarrow q{\bar q}$ which are better described in terms of $s$-channel color structures. Since the full cross section is gauge independent, the gauge dependence in the product of the hard and soft functions, $H^{({\rm f})} \, S^{({\rm f})}$, must cancel the gauge dependence of the incoming jets, $\psi$, and outgoing jets, $J_{(f_i)}$. Since the jets are incoherent relative to the hard and soft functions, the gauge dependence of the anomalous dimension matrices $\Gamma_S^{({\rm f})}$ must be proportional to the identity matrix. Then we can rewrite the anomalous dimension matrices as \begin{eqnarray} && \hspace{-10mm} (\Gamma^{({\rm f})}_S)_{KL}=(\Gamma^{({\rm f})}_{S'})_{KL} + \delta_{KL} \frac{\alpha_s}{\pi} \nonumber \\ && \hspace{-10mm} \times \sum_{i=a,b,1,2}C_{(f_i)} \, \frac{1}{2} \, (-\ln\nu_i-\ln 2+1-\pi i) \, , \label{gammagaug} \end{eqnarray} with $C_{f_i}=C_F\ (C_A)$ for a quark (gluon), and with $\nu_i \equiv (v_i \cdot n)^2/|n|^2$. Here the dimensionless and lightlike velocity vectors $v_i^{\mu}$ are defined by $p_i^{\mu}=M_{JJ}v_i^{\mu}/{\sqrt{2}}$ and satisfy $v_i^2=0$. In the following subsections we will present the explicit expressions for $\Gamma^{({\rm f})}_{S'}$ for all the partonic subprocesses in dijet production, by evaluating one-loop diagrams as in Fig.~3. The full anomalous dimension matrices can be retrieved from Eq.~(\ref{gammagaug}). \subsection{Soft anomalous dimension for $q \bar{q}\rightarrow q \bar{q}$} First, we present the soft anomalous dimension matrix for the process \begin{equation} q\left(p_a, r_a \right)+\bar{q}\left(p_b, r_b \right) \rightarrow q\left(p_1, r_1 \right)+\bar{q}\left(p_2, r_2 \right) \, , \end{equation} in the $t$-channel singlet-octet color basis \begin{eqnarray} c_1&=&\delta_{r_a r_1}\delta_{r_b r_2} \, , \nonumber\\ c_2&=&(T_F^c)_{r_1 r_a}(T_F^c)_{r_b r_2}. \end{eqnarray} We find~\cite{KOS2,BottsSt} \begin{equation} \Gamma_{S'}=\frac{\alpha_s}{\pi}\left[ \begin{array}{cc} 2{C_F}T & -\frac{C_F}{N_c} \U \vspace{2mm} \\ -2\U &-\frac{1}{N_c}(T-2\U) \end{array} \right]\, . \end{equation} We note that the dependence on $T$ is diagonal in this $t$-channel color basis and in the forward region of the partonic scattering, $T\rightarrow -\infty$, where $\Gamma_{S}$ becomes diagonal, color singlet exchange is exponentially enhanced relatively to color octet. \subsection{Soft anomalous dimension for $q q\rightarrow q q$ and ${\bar q}{\bar q}\rightarrow {\bar q}{\bar q}$} Next, we consider the process \begin{equation} q\left(p_a, r_a \right)+q\left(p_b, r_b \right) \rightarrow q\left(p_1, r_1 \right)+q\left(p_2, r_2 \right) \, , \end{equation} in the $t$-channel singlet-octet color basis \begin{eqnarray} c_1&=&(T_F^c)_{r_1 r_a}(T_F^c)_{r_2 r_b} \, , \nonumber\\ c_2&=&\delta_{r_a r_1} \delta_{r_b r_2}. \end{eqnarray} The anomalous dimension matrix is~\cite{KOS2,BottsSt} \begin{equation} \Gamma_{S'}=\frac{\alpha_s}{\pi}\left[ \begin{array}{cc} -\frac{1}{N_c}(T+\U)+2C_F \U & 2\U \vspace{2mm} \\ \frac{C_F}{N_c} \U & 2{C_F}T \end{array} \right] \end{equation} which also applies to the process \begin{equation} {\bar q}\left(p_1, r_1 \right)+{\bar q}\left(p_2, r_2 \right) \rightarrow {\bar q}\left(p_a, r_a \right)+{\bar q}\left(p_b, r_b \right) \, . \end{equation} Again, we note that the dependence on $T$ is diagonal in this $t$-channel color basis, and the color singlet dominates in the forward region of the partonic scattering. \subsection{Soft anomalous dimension for $q \bar{q}\rightarrow g g$ and $g g \rightarrow q \bar{q}$} Here, we present the soft anomalous dimension matrix for the process \begin{equation} q\left(p_a, r_a \right)+\bar{q}\left(p_b, r_b \right) \rightarrow g\left(p_1, r_1 \right)+g\left(p_2, r_2 \right) \, , \end{equation} in the $s$-channel color basis \begin{eqnarray} c_1&=&\delta_{r_a r_b}\delta_{r_1 r_2} \, , \nonumber \\ c_2&=&d^{r_1 r_2 c}{\left( T_F^c \right)}_{r_b r_a} \, , \nonumber \\ c_3&=&if^{r_1 r_2 c}{\left( T_F^c \right)}_{r_b r_a} \, . \end{eqnarray} We find~\cite{KOS2} \begin{eqnarray} &&\Gamma_{S'}=\frac{\alpha_s}{\pi} \\ && \nonumber \times \left[ \begin{array}{ccc} 0 & 0 & \U-T \vspace{2mm} \\ 0 & \frac{C_A}{2}\left(T+\U \right) & \frac{C_A}{2} \left(\U-T\right) \vspace{2mm} \\ 2\left(\U-T \right) & \frac{N_c^2-4}{2N_c}\left(\U-T \right) & \frac{C_A}{2}\left(T+\U \right) \end{array} \right]. \label{Gammaqqgg} \end{eqnarray} The same anomalous dimension describes also the time-reversed process \cite{Thesis,KS} \begin{equation} g\left(p_1, r_1 \right)+g\left(p_2, r_2 \right) \rightarrow \bar{q}\left(p_a, r_a \right)+q\left(p_b, r_b \right) . \end{equation} \subsection{Soft anomalous dimension for $qg \rightarrow qg$ and $\bar{q} g \rightarrow \bar{q} g$} Next, we consider the ``Compton'' process \begin{equation} q\left(p_a, r_a \right)+g\left(p_b, r_b \right) \rightarrow q\left(p_1, r_1 \right)+g\left(p_2, r_2 \right) \, , \end{equation} in the $t$-channel color basis \begin{eqnarray} c_1&=&\delta_{r_a r_1}\delta_{r_b r_2} \, , \nonumber \\ c_2&=&d^{r_b r_2 c}{\left( T_F^c \right)}_{r_1 r_a} \, , \nonumber \\ c_3&=&if^{r_b r_2 c}{\left( T_F^c \right)}_{r_1 r_a} \, . \label{eq:basqgqg} \end{eqnarray} The soft anomalous dimension matrix is \cite{KOS2} \begin{eqnarray} &&\Gamma_{S'}=\frac{\alpha_s}{\pi} \\ && \nonumber \times \left[ \begin{array}{ccc} \left( C_F+C_A \right) T & 0 & \U \vspace{2mm} \\ 0 & C_F T+ \frac{C_A}{2} \U & \frac{C_A}{2} \U \vspace{2mm} \\ 2\U & \frac{N_c^2-4}{2N_c}\U & C_F T+ \frac{C_A}{2}\U \end{array} \right] \label{Gammaqgqg} \end{eqnarray} which also applies to the process \begin{equation} \bar{q}\left(p_1, r_1 \right)+g\left(p_2, r_2 \right) \rightarrow \bar{q}\left( p_a, r_a \right)+g\left(p_b, r_b \right) \, . \end{equation} We note that $T$ appears only in the diagonal of the anomalous dimension matrix and that in the forward region ($T \rightarrow -\infty$) the color singlet dominates. \subsection{Soft anomalous dimension for $gg \rightarrow gg$} Finally, we consider the much more complicated process \begin{equation} g\left(p_a, r_a \right)+g\left(p_b, r_b \right) \rightarrow g\left(p_1, r_1 \right)+g\left(p_2, r_2 \right) \, . \end{equation} A complete color basis for this process is given by the eight color structures~\cite{KOS2} \begin{eqnarray} c_1&=&\frac{i}{4}\left[f^{r_a r_b l} d^{r_1 r_2 l} - d^{r_a r_b l}f^{r_1 r_2 l}\right] \, , \nonumber \\ c_2&=&\frac{i}{4}\left[f^{r_a r_b l} d^{r_1 r_2 l} + d^{r_a r_b l}f^{r_1 r_2 l}\right] \, , \nonumber \\ c_3&=&\frac{i}{4}\left[f^{r_a r_1 l} d^{r_b r_2 l}+d^{r_a r_1 l}f^{r_b r_2 l}\right] \, , \nonumber \\ c_4&=&P_1(r_a,r_b;r_1,r_2)=\frac{1}{8}\delta_{r_a r_1} \delta_{r_b r_2} \, , \nonumber \\ c_5&=&P_{8_S}(r_a,r_b;r_1,r_2) =\frac{3}{5} d^{r_ar_1c} d^{r_br_2c} \, , \nonumber \\ c_6&=&P_{8_A}(r_a,r_b;r_1,r_2)=\frac{1}{3} f^{r_ar_1c} f^{r_br_2c} \, , \nonumber \\ c_7&=&P_{10+{\overline {10}}}(r_a,r_b;r_1,r_2)= \frac{1}{2}(\delta_{r_a r_b} \delta_{r_1 r_2} \nonumber \\ && \quad {}-\delta_{r_a r_2} \delta_{r_b r_1}) -\frac{1}{3} f^{r_ar_1c} f^{r_br_2c} \, , \nonumber \\ c_8&=&P_{27}(r_a,r_b;r_1,r_2)=\frac{1}{2}(\delta_{r_a r_b} \delta_{r_1 r_2} \nonumber \\ && \hspace{-3mm} {}+\delta_{r_a r_2} \delta_{r_b r_1}) -\frac{1}{8}\delta_{r_a r_1} \delta_{r_b r_2} -\frac{3}{5} d^{r_ar_1c} d^{r_br_2c} \, , \nonumber \\ \label{8x8basis} \end{eqnarray} where the $P$'s are $t$-channel projectors of irreducible representations of $SU(3)$ \cite{Bart}, and we use explicitly $N_c=3$. The soft anomalous dimension matrix in this basis is~\cite{KOS2} \begin{equation} \Gamma_{S'}=\left[\begin{array}{cc} \Gamma_{3 \times 3} & 0_{3 \times 5} \\ 0_{5 \times 3} & \Gamma_{5 \times 5} \end{array} \right] \, , \label{gammagggg} \end{equation} with \begin{equation} \Gamma_{3 \times 3}=\frac{\alpha_s}{\pi} \left[ \begin{array}{ccc} 3T & 0 & 0 \\ 0 & 3U & 0 \\ 0 & 0 & 3\left(T+\U \right) \end{array} \right] \end{equation} and \begin{eqnarray} &&\Gamma_{5 \times 5}=\frac{\alpha_s}{\pi} \\ && \nonumber \hspace{-3mm} \times \left[\begin{array}{ccccc} 6T & 0 & -6\U & 0 & 0 \vspace{2mm} \\ 0 & 3T+\frac{3\U}{2} & -\frac{3\U}{2} & -3\U & 0 \vspace{2mm} \\ -\frac{3\U}{4} & -\frac{3\U}{2} &3T+\frac{3\U}{2} & 0 & -\frac{9\U}{4} \vspace{2mm} \\ 0 & -\frac{6\U}{5} & 0 & 3\U & -\frac{9\U}{5} \vspace{2mm} \\ 0 & 0 &-\frac{2\U}{3} &-\frac{4\U}{3} & -2T+4\U \end{array} \right] \, . \end{eqnarray} We note that the dependence on $T$ is diagonal and that color singlet exchange dominates in the forward region of the partonic scattering, $T \rightarrow -\infty$, where $\Gamma_{S}$ becomes diagonal. This has been the case for all the processes we analyzed in $t$-channel bases; suppression increases with the dimension of the exchanged color representation. The calculation of the eigenvalues and eigenvectors of $\Gamma_{S'}$ for the $gg \rightarrow gg$ process is more difficult than for the other partonic processes since we are now dealing with a $8 \times 8$ matrix; however, explicit results have been obtained \cite{KOS2}. The eigenvalues of the anomalous dimension matrix, Eq.~(\ref{gammagggg}), are \begin{eqnarray} \lambda_1&=&\lambda_4=3 \, \frac{\alpha_s}{\pi} \, T \, , \nonumber \\ \lambda_2&=&\lambda_5=3 \, \frac{\alpha_s}{\pi} \, \U \, , \nonumber \\ \lambda_3&=&\lambda_6=3 \, \frac{\alpha_s}{\pi} \, (T+\U) \, , \nonumber \\ \lambda_{7,8}&=&2 \, \frac{\alpha_s}{\pi} \left[T+\U \right. \nonumber \\ && \left. \, \mp 2\sqrt{T^2-T\U+\U^2} \right] \, . \end{eqnarray} The eigenvectors have the general form \begin{eqnarray} e_i&=&\left[\begin{array}{c} e_i^{(3)} \\ 0^{(5)} \end{array}\right], \; i=1,2,3 \, , \nonumber \\ e_i&=&\left[\begin{array}{c} 0^{(3)} \\ e_i^{(5)} \end{array}\right], \; i=4 \ldots 8 \, , \end{eqnarray} where the superscripts refer to the dimension. The three-dimensional vectors $e_i^{(3)}$ are given by \begin{equation} e_i^{(3)}=\left[\begin{array}{c} \delta_{i1} \\ \delta_{i2} \\ \delta_{i3} \end{array}\right], \; i=1,2,3 \, . \label{ei3} \end{equation} The five-dimensional vectors $e_4^{(5)}$, $e_5^{(5)}$ and $e_6^{(5)}$ are given by~\cite{KOS2} \begin{eqnarray} e_4^{(5)}&=&\left[\begin{array}{c} -15 \vspace{2mm} \\ 6-\frac{15}{2}\frac{T}{\U} \vspace{2mm} \\ -\frac{15}{2}\frac{T}{\U} \vspace{2mm} \\ 3 \vspace{2mm} \\ 1 \end{array} \right], \; \; e_5^{(5)}=\left[\begin{array}{c} 0 \vspace{2mm} \\ -\frac{3}{2} \vspace{2mm} \\ 0 \vspace{2mm} \\ \frac{3}{4}-\frac{3}{2}\frac{T}{\U} \vspace{2mm} \\ 1 \end{array} \right], \nonumber \\ e_6^{(5)}&=&\left[\begin{array}{c} -15 \vspace{2mm} \\ -\frac{3}{2}+\frac{15}{2}\frac{T}{\U} \vspace{2mm} \\ \frac{15}{2}-\frac{15}{2}\frac{T}{\U} \vspace{2mm} \\ -3 \vspace{2mm} \\ 1 \end{array} \right] \, . \end{eqnarray} The expressions for $e_7^{(5)}$ and $e_8^{(5)}$ are long but can be given succinctly by~\cite{KOS2} \begin{eqnarray} e_i^{(5)}&=& \left[\begin{array}{c} b_1(\lambda_i') \vspace{2mm} \\ b_2(\lambda_i') \vspace{2mm} \\ b_3(\lambda_i') \vspace{2mm} \\ b_4(\lambda_i') \vspace{2mm} \\ 1 \end{array} \right], \; i=7,8 \, , \end{eqnarray} where \begin{equation} \lambda_i'=\frac{\pi}{\alpha_s} \lambda_i \, , \end{equation} and where the $b_i$'s are given by \begin{eqnarray} b_1(\lambda_i')&=&\frac{3}{\U^2 K'} [80 T^4+103 \U^4-280 \U T^3 \nonumber \\ && \quad \quad \; \; {}-300 T \U^3 +404 T^2 \U^2 \nonumber \\ && \hspace{-12mm} {}+(40 T^3-16 \U^3 -60 T^2 \U +52 T\U^2)\lambda_i'] \, , \nonumber \\ b_2(\lambda_i')&=&\frac{3}{2K'}[20 T^2-50 \UT +44 \U^2 \nonumber \\ && \quad \quad {}+(10 T-5 \U)\lambda_i'] \, , \nonumber \\ b_3(\lambda_i')&=&-\frac{3}{2 \U K'}[40T^3-64\U^3-120 T^2 \U \nonumber \\ && \hspace{-12mm} {}+130 T \U^2+(20 T^2+13 \U^2-20 T\U) \lambda_i'] \, , \nonumber \\ b_4(\lambda_i')&=&\frac{3\U}{K'}(2T+5\U-2\lambda_i') \, , \end{eqnarray} with \begin{equation} K'=20 T^2-20 \UT+21 \U^2 \, . \end{equation} \section{Conclusion} We have discussed soft gluon resummation at next-to-leading logarithmic accuracy for heavy quark and dijet production in hadronic collisions. We have constructed the resummed cross sections in terms of exponentials of soft anomalous dimension matrices which describe the factorization of noncollinear soft gluons from the hard scattering. We have presented explicit results for the soft anomalous dimension matrices for the relevant partonic subprocesses. Numerical results have been presented for top quark production at the Fermilab Tevatron. Similar resummations have been recently applied to direct photon \cite{NK,LOS,NKJO} and $W$ + jet \cite{NK,NKVD} production, as well as to calculations of transverse momentum and rapidity distributions \cite{KLMV,pty} for heavy quark production.
\section{Introduction} One of the central problems of particle physics is to understand how the electroweak scale associated with the mass of the W boson is generated, and why it is so small as compared to the Planck scale associated with the Newton's constant. In the Standard Model (SM) the electroweak scale is generated through the vacuum expectation value (VEV) of the neutral component of a Higgs doublet \cite{Gunion:1989we}. Apart from the fact that this VEV is an arbitrary parameter in the SM, the mass parameter of the Higgs field suffers from quadratic divergences, making the weak scale unstable under radiative corrections. Supersymmetry is at present the only known framework in which the weak scale is stable under radiative corrections \cite{Haber:1985rc}, although it does not explain how such a small scale arises in the first place. As such, considerable importance attaches to the study of supersymmetric models, especially the Minimal Supersymmetric Standard Model (MSSM), based on the gauge group $SU(2)_L \times U(1)_Y$, with two Higgs doublet superfields. It is well known that, because of underlying gauge invariance and supersymmetry (SUSY), the lightest Higgs boson in MSSM has a tree level upper bound of $M_Z$ (the mass of Z boson) on its mass \cite{Inoue:1982ej}. Although radiative corrections \cite{Okada:1991vk} to the tree level result can be appreciable, these depend only logarithmically on the SUSY breaking scale, and are, therefore, under control. This results in an upper bound of about $125-135$ GeV on the one-loop radiatively corrected mass \cite{Rosiek:1995dy} of the lightest Higgs boson in MSSM\footnote{The two-loop corrections to the Higgs boson mass matrix in the MSSM are significant, and can reduce the lightest Higgs boson mass by up to $\sim 20$ GeV as compared to its one-loop value \cite{twoloop}.}. Because of the presence of the additional trilinear Yukawa couplings, such a tight constraint on the mass of the lightest Higgs boson need not {\em a priori} hold in extensions of MSSM based on the gauge group $SU(2)_L \times U(1)_Y$ with an extended Higgs sector. Nevertheless, it has been shown that the upper bound on the lightest Higgs boson mass in these models depends only on the weak scale, and dimensionless coupling constants (and only logarithmically on the SUSY breaking scale), and is calculable if all the couplings remain perturbative below some scale $\Lambda$ \cite{Ellis:1989er,Binetruy:1992mk,Pandita:1993hx,Elliott:1993ex,Ellwanger:1993hn,Kane:1993kq,EQ}. This upper bound can vary between $150$ GeV to $200$ GeV depending on the Higgs structure of the underlying supersymmetric model. Thus, nonobservation of such a light Higgs boson below this upper bound will rule out an entire class of supersymmetric models based on the gauge group $SU(2)_L \times U(1)_Y$. The existence of the upper bound on the lightest Higgs boson mass in MSSM (with arbitrary Higgs sectors) has been investigated in a situation where the underlying supersymmetric model respects baryon ($B$) and lepton ($L$) number conservation. However, it is well known that gauge invariance, supersymmetry and renormalizibility allow $B$ and $L$ violating terms in the superpotentioal of the MSSM \cite{Weinberg:1982wj}. The strength of these lepton and baryon number violating terms is, however, severely limited by phenomenological \cite{Zwirner:1984is,Smirnov:1996ey}, and cosmological \cite{Campbell:1991fa} constraints. Indeed, unless the strength of baryon-number violating term is less than $10^{-13}$, it will lead to contradiction with the present lower limits on the lifetime of the proton. The usual strategy to prevent the appearance of $B$ and $L$ violating couplings in MSSM is to invoke a discrete $Z_2$ symmetry \cite{Farrar:1978xj} known as matter parity, or R-parity. The matter parity of each superfield may be defined as \begin{equation} (\rm{matter\; parity}) \equiv (-1)^{3 (B-L)} . \label{eq:parity} \end{equation} The multiplicative conservation of matter parity forbids all the renormalizable $B$ and $L$ violating terms in the superpotential of MSSM. Equivalently, the R-parity of any component {\em field} is defined by $R_p = (-1)^{3(B-L)+2S}$, where $S$ is the spin of the field. Since $(-1)^{2S}$ is conserved in any Lorentz-invariant interaction, matter parity conservation and R-parity conservation are equivalent. Conservation of R-parity then immediately implies that superpartners can be produced only in pairs, and that the lightest supersymmetric particle (LSP) is absolutely stable. Although the Minimal Supersymmetric Standard Model with R-parity conservation can provide a description of nature which is consistent with all known observations, the assumption of $R_p$ conservation appears to be {\em ad hoc}, since it is not required for the internal consistency of MSSM. Furthermore, all global symmetries, discrete or continuous, could be violated by the Planck scale physics effects \cite{Giddings:1988cx}. The problem becomes acute for low energy supersymmetric models, because $B$ and $L$ are no longer automatic symmetries of the Lagrangian, as they are in the Standard Model. It is, therefore, more appealing to have a supersymmetric theory where R-parity is related to a gauge symmetry, and its conservation is automatic because of the invariance of the underlying theory under this gauge symmetry. Fortunately, there is a compelling scenario which does provide for exact R-parity conservation due to a deeper principle. Indeed, $R_p$ conservation follows automatically in certain theories with gauged $(B-L)$, as is suggested by the fact that matter parity is simply a $Z_2$ subgroup of $(B-L)$. It has been noted by several authors \cite{Mohapatra:1986su,Martin:1992mq} that if the gauge symmetry of MSSM is extended to $SU(2)_L \times U(1)_{I_{3R}} \times U(1)_{B-L}$, or $SU(2)_L \times SU(2)_R \times U(1)_{B-L}$, the theory becomes automatically R-parity conserving. Such a left-right supersymmetric theory (SLRM) solves the problems of explicit $B$ and $L$ violation of MSSM, and has received much attention recently \cite{Cvetic:1984su,Francis:1991pi,Kuchimanchi:1993jg,Huitu:1994gf,Huitu:1995zm,Huitu:1997iy,Aulakh:1997ba}. Of course left-right symmetric theories are also interesting in their own right, for among other appealing features, they offer a simple and natural explanation for the smallness of neutrino mass through the so called see-saw mechanism \cite{Gell-Mann:1979}. Such a naturally R-parity conserving theory necessarily involves the extension of the Standard Model gauge group, and since the extended gauge symmetry has to be broken, it involves a ``new scale'', the scale of left-right symmetry breaking, beyond the SUSY and $SU(2)_L \times U(1)_Y$ breaking scales of MSSM. In \cite{Huitu:1997rr} we showed that in the SLRM with minimal particle content the upper bound on the mass of the lightest neutral Higgs boson depends only on gauge couplings and those VEVs which break the $SU(2)_L\times U(1)_Y$ symmetry. Here we will present a detailed analysis of the Higgs sector of the left-right supersymmetric models, and consider some of the distinguishing features of the lightest Higgs boson in these models. In the SLRMs there are typically also other light Higgs particles. The light doubly charged Higgs boson, which we will consider in detail as well, should provide a clear signal in experiments. The plan of the paper is as follows. In section \ref{sec:tree}, we review the various left-right supersymmetric models that we consider in this paper. In section \ref{sec:higgs}, we obtain the tree level upper bound on the mass of the lightest CP-even Higgs boson for the various SLRMs considered in section \ref{sec:tree}. We use a general procedure to obtain this (tree-level) upper bound on the mass of the lightest CP-even Higgs boson in models with extended gauge groups, such as SLRMs. We show that the upper bound so obtained in the renormalizable models is independent of the supersymmetry breaking scale, as well as the left-right breaking scale. In the case of models containing non-renormalizable terms, although the upper bound depends on the left-right breaking scale, the dependence is extremely weak, being suppressed by powers of Planck mass. In section \ref{sec:rad} we calculate the radiative corrections to the upper bound on the mass of the lightest Higgs boson, and show that the most important radiative corrections arising from quark-squark loops are of the same type as in the MSSM based on $SU(2)_L \times U(1)_Y$. The radiatively corrected upper bound so obtained is numerically considerably larger than the corresponding bound in the MSSM, but for most of the parameter space is below 200 GeV. In section \ref{sec:decoup} we consider the branching ratios of the lightest neutral Higgs boson, and find that its couplings are similar to the corresponding Higgs couplings in the Standard Model in the decoupling limit. In section \ref{sec:double} we discuss the mass of the lightest doubly charged Higgs boson and the possibility of its detection at colliders. In section \ref{sec:conclusion}, we present our conclusions. The full scalar potential of the minimal supersymmetric left-right model is presented in the Appendix A. \section{The Higgs sector of the left-right supersymmetric models} \label{sec:tree} In this section we briefly review the minimal supersymmetric left-right model, and then discuss models which have an extended Higgs sector, and finally discuss models with non-renormalizable interaction terms in the superpotential. The minimal SLRM is based on the gauge group $SU(3)_C \times SU(2)_L \times SU(2)_R \times U(1)_{B-L}$. The matter fields of this model consist of the three families of quark and lepton chiral superfields with the following transformation properties under the gauge group: \begin{eqnarray} Q=\left( \begin{array}{c} U \\ D \end{array} \right) \sim (3,2,1,\frac{1}{3}) & ,\; & Q^c = \left( \begin{array}{c} D^c \\ U^c \end{array} \right) \sim (3^*,1,2,-\frac{1}{3}) , \nonumber \\ L=\left( \begin{array}{c} \nu \\ E \end{array} \right) \sim (1,2,1,-1) & ,\; & L^c =\left( \begin{array}{c} E^c \\ \nu^c \end{array} \right) \sim (1,1,2,1) , \label{eq:fields1} \end{eqnarray} where the numbers in the brackets denote the quantum numbers under $SU(3)_C \times SU(2)_L \times SU(2)_R \times U(1)_{B-L}$. The Higgs sector consists of the bidoublet and triplet Higgs superfields: \begin{eqnarray} \Phi = \left( \begin{array}{cc} \Phi^0_1 & \Phi^+_1 \\ \Phi^-_2 & \Phi^0_2 \end{array} \right) \sim (1,2,2,0) & ,\; & \chi = \left( \begin{array}{cc} \chi^0_1 & \chi^+_1 \\ \chi^-_2 & \chi^0_2 \end{array} \right) \sim (1,2,2,0) , \nonumber \\ \Delta_R = \left( \begin{array}{cc} \frac{1}{\sqrt{2}} \Delta_R^- & \Delta_R^0 \\ \Delta_R^{--} & -\frac{1}{\sqrt{2}} \Delta^-_R \end{array} \right) \sim (1,1,3,-2) & ,\; & \delta_R = \left( \begin{array}{cc} \frac{1}{\sqrt{2}} \delta_R^+ & \delta_R^{++} \\ \delta_R^0 & -\frac{1}{\sqrt{2}} \delta^+_R \end{array} \right) \sim (1,1,3,2) , \nonumber \\ \Delta_L = \left( \begin{array}{cc} \frac{1}{\sqrt{2}} \Delta_L^- & \Delta_L^0 \\ \Delta_L^{--} & -\frac{1}{\sqrt{2}} \Delta^-_L \end{array} \right) \sim (1,3,1,-2) & ,\; & \delta_L = \left( \begin{array}{cc} \frac{1}{\sqrt{2}} \delta_L^+ & \delta_L^{++} \\ \delta_L^0 & -\frac{1}{\sqrt{2}} \delta^+_L \end{array} \right) \sim (1,3,1,2) . \label{eq:fields2} \end{eqnarray} There are two bidoublet superfields in order to implement the $SU(2)_L \times U(1)_Y$ breaking, and to generate a nontrivial Kobayashi-Maskawa matrix. Furthermore, two $SU(2)_R$ Higgs triplet superfields $\Delta_R$ and $\delta_R$ with opposite $(B-L)$ are necessary to break the left-right symmetry spontaneously, and to cancel triangle gauge anomalies due to the fermionic superpartners. The gauge symmetry is supplemented by a discrete left-right symmetry under which the fields can be chosen to transform as \begin{eqnarray} Q \leftrightarrow Q^c , \; L \leftrightarrow L^c , \; \Phi \leftrightarrow -\tau_2 \Phi^T \tau_2 , \; \chi \leftrightarrow -\tau_2 \chi^T \tau_2 , \; \Delta_R \leftrightarrow \delta_L , \; \delta_R \leftrightarrow \Delta_L . \end{eqnarray} Thus, the $SU(2)_L$ triplets $\Delta_L$ and $\delta_L$ are needed in order to make the Lagrangian fully symmetric under $L \leftrightarrow R$ transformation, although these are not needed phenomenologically for symmetry breaking, or the see-saw mechanism. The most general gauge invariant superpotential involving these superfields can be written as (generation indices suppressed) \begin{eqnarray} W_{min}&=& h_{\Phi Q}Q^T i\tau_2 \Phi Q^c + h_{\chi Q}Q^T i\tau_2 \chi Q^c + h_{\Phi L}L^T i\tau_2 \Phi L^c + h_{\chi L}L^T i\tau_2 \chi L^c \nonumber\\ &&+h_{\delta_L} L^T i\tau_2 \delta_L L + h_{\Delta_R} L^{cT} i\tau_2 \Delta_R L^c+ \mu_1 \rm{Tr} (i\tau_2\Phi^T i\tau_2 \chi) + \mu_1' \rm{Tr} (i\tau_2\Phi^T i\tau_2 \Phi) \nonumber\\ &&+ \mu_1'' \rm{Tr} (i\tau_2\chi^T i\tau_2 \chi) +\rm{Tr} (\mu_{2L}\Delta_L \delta_L + \mu_{2R}\Delta_R\delta_R). \label{eq:superpotential} \end{eqnarray} The scalar potential can be calculated from \begin{equation} V=V_F+V_D+V_{soft}, \label{eq:scalarpotential} \end{equation} where $V_F$, $V_D$, and $V_{soft}$ represent the contribution of $F$-terms, the $D$-terms, and the soft SUSY breaking terms, respectively. The full scalar potential can be found in Appendix A. The general form of the vacuum expectation values of the various scalar fields which preserve the $U(1)_{\rm{em}}$ gauge invariance can be written as \begin{eqnarray} &&\langle \Phi \rangle = \left( \begin{array}{cc} \kappa_1 & 0 \\ 0 & e^{i \phi_1}\kappa_1' \end{array} \right) ,\; \langle \chi \rangle =\left(\begin{array}{cc} e^{i \phi_2} \kappa_2' & 0 \\ 0 & \kappa_2\end{array} \right) , \; \langle \Delta_R \rangle = \left(\begin{array}{cc} 0 & v_{\Delta_R} \\ 0 & 0 \end{array} \right) ,\; \langle \delta_R \rangle = \left(\begin{array}{cc} 0 & 0 \\ v_{\delta_R} & 0 \end{array} \right) , \nonumber \\ && \langle \Delta_L \rangle = \left(\begin{array}{cc} 0 & v_{\Delta_L} \\ 0 & 0 \end{array} \right) ,\; \langle \delta_L \rangle = \left(\begin{array}{cc} 0 & 0 \\ v_{\delta_L} & 0\end{array} \right) , \; \langle L \rangle =\left(\begin{array}{c} \sigma_L \\ 0 \end{array} \right) ,\; \langle L^c\rangle = \left( \begin{array}{c} 0 \\ \sigma_R \end{array} \right) . \label{eq:vevs} \end{eqnarray} We note that the triplet vacuum expectation values $v_{\Delta_R}$ and $v_{\delta_R}$ represent the scale of $SU(2)_R$ breaking and are, according to the lower bounds \cite{Caso} on heavy W- and Z-boson masses, in the range $v_{\Delta_R},v_{\delta_R} \mathrel{\vcenter{\hbox{$>$}\nointerlineskip\hbox{$\sim$}}} 1$ TeV. These represent a new scale, the right-handed breaking scale. We note that $\kappa_1'$ and $\kappa_2'$ contribute to the mixing of the charged gauge bosons and to the flavour changing neutral currents, and are usually assumed to vanish. Furthermore, since the electroweak $\rho$ parameter is close to unity, $\rho = 0.9998 \pm 0.0008$ \cite{Caso}, the triplet vacuum expectation values $v_{\Delta_L}$ and $v_{\delta_L}$ must be small. The Yukawa coupling $h_{\chi L}$ is proportional to the neutrino Dirac mass $m_D$. The light neutrino mass in the see-saw mechanism is given by $\sim m_D^2 /m_M$, where $m_M = h_{\Delta_R} v_{\Delta_R}$ is the Majorana mass. The magnitude of $h_{\chi L}$ is not accurately determined given the present upper limit on the light neutrino masses. On the other hand the Yukawa coupling $h_{\Phi L}$ is proportional to the electron mass and is, thus, small. In the minimal model described above, parity cannot be spontaneously broken at the renormalizable level without spontaneous breaking of $R$-parity. This may be cured by adding more fields to the theory. In \cite{Kuchimanchi:1993jg,Kuchimanchi:1995vk} it was suggested that a parity-odd singlet, coupled appropriately to triplet fields, be introduced so as to ensure proper symmetry breaking. This leads to a set of degenerate minima connected by a flat direction, all of them breaking parity. When soft SUSY breaking terms are switched on, the degeneracy is lifted, but the global minimum that results breaks $U(1)_{\rm{em}}$. Because of the flat direction connecting the minima, there is no hope that the fields remain in the phenomenologically acceptable vacuum, which rolls down to global minimum after SUSY is softly broken. The only option that is left is to have a relatively low $SU(2)_R$ breaking scale, with spontaneously broken $R$-parity ($\langle \nu^c \rangle \equiv \sigma_R$ is non-zero). We note that present experiments allow for a low $SU(2)_R$ breaking scale. There is an alternative to the minimal left-right supersymmetric model which involves the addition of a couple of triplet fields, $\Omega_L(1,3,1,0)$ and $\Omega_R(1,1,3,0)$, instead of a singlet Higgs superfield, to the minimal model \cite{Aulakh:1997ba}. In these extended models the breaking of $SU(2)_R$ is achieved in two stages. In the first stage the gauge group $SU(2)_L \times SU(2)_R \times U(1)_{B-L}$ is broken to an intermediate symmetry group $SU(2)_L \times U(1)_R \times U(1)_{B-L}$, and at the second stage $U(1)_R \times U(1)_{B-L}$ is broken to $U(1)_Y$ at a lower scale. In this theory there is only one parity-breaking minimum, in contrast to the minimal model, which respects the electromagnetic gauge invariance. The superpotential for this class of models obtains additional terms involving the triplet fields $\Omega_L$ and $\Omega_R$: \begin{eqnarray} W_{\Omega}&=& W_{min} + \frac 12 \mu_{\Omega_L} {\rm{Tr}}\,\Omega_L^2 +\frac 12 \mu_{\Omega_R} {\rm Tr}\,\Omega_R^{2} +a_L {\rm Tr}\,\Delta_L \Omega_L \delta_L +a_R {\rm Tr}\,\Delta_R \Omega_R \delta_R \nonumber \\ & & + {\rm Tr}\, \Omega_L \left( \alpha_L \Phi i\tau_2 \chi^T i\tau_2+{\alpha_L}' \Phi i\tau_2 \Phi^T i\tau_2 +{\alpha_L}'' \chi i\tau_2 \chi^T i\tau_2 \right) \nonumber \\ & & + {\rm Tr}\, \Omega_R \left( \alpha_R i \tau_2 \Phi^T i\tau_2 \chi+{\alpha_R}' i\tau_2 \Phi^T i\tau_2 \Phi +{\alpha_R}'' i\tau_2\chi^T i\tau_2 \chi \right), \label{superpot} \end{eqnarray} where $W_{min}$ is the superpotential (\ref{eq:superpotential}) of the minimal left-right model. In these models the see-saw mechanism takes its canonical form with $m_{\nu} \simeq m_D^2/M_{BL}$, where $m_D$ is the neutrino Dirac mass. In this case the low-energy effective theory is the MSSM with unbroken $R$-parity, and contains besides the usual MSSM states, a triplet of Higgs scalars much lighter than the $B-L$ breaking scale. A second option is to add non-renormalizable terms to the Lagrangian of the minimal left-right supersymmetric model, while retaining the minimal Higgs content \cite{Martin:1992mq,AMRS,AMS} The superpotential for this class of models can be written as \begin{eqnarray} W_{NR} &=& W_{min} + {a_L \over 2 M} ({\rm Tr}\, \Delta_L \delta_L)^2 + {a_R \over 2 M} ({\rm Tr}\, \Delta_R \delta_R)^2+ {c \over M}{\rm Tr}\, \Delta_L \delta_L {\rm Tr}\, \Delta_R \delta_R \nonumber \\ & & + {b_L \over 2 M} {\rm Tr}\, \Delta_L^2 {\rm Tr}\, \delta_L^2 + {b_R \over 2 M} {\rm Tr}\, \Delta_R^2 {\rm Tr}\, \delta_R^2 + {1 \over M} \left[ d_1{\rm Tr}\, \Delta_L^2 {\rm Tr}\, \delta_R^2 + d_2 {\rm Tr}\, \delta_L^2 {\rm Tr}\, \Delta_R^2 \right] \nonumber \\ & & + {\lambda_{ijkl}\over M}{\rm Tr}\,i\tau_2 \Phi_i^T i\tau_2 \Phi_j{\rm Tr}\,i\tau_2 \Phi_k^T i\tau_2 \Phi_l + {\alpha_{ijL}\over M} {\rm Tr}\, \Delta_L \delta_L \Phi_i i\tau_2 \Phi_j^T i\tau_2 \nonumber\\ &&+ {\alpha_{ijR}\over M} {\rm Tr}\, \Delta_R \delta_R i\tau_2 \Phi_i^T i\tau_2 \Phi_j + {1 \over M}{\rm Tr}\,\tau_2 \Phi_i^T \tau_2 \Phi_j [\beta_{ijL} {\rm Tr}\, \Delta_L \delta_L + \beta_{ijR} {\rm Tr}\, \Delta_R \delta_R ] \nonumber \\ & &+ {\eta_{ij} \over M } {\rm Tr}\,\Phi_i \Delta_R i\tau_2 \Phi_j^T i \tau_2 \delta_L + {\overline \eta_{ij} \over M } {\rm Tr}\,\Phi_i \delta_R i\tau_2 \Phi_j^T i\tau_2 \Delta_L \nonumber \\ & & + { k_{ql} \over M}Q^T i\tau_2 L\,Q^{cT} i\tau_2 L^c + {k_{qq} \over M}Q^T i\tau_2 Q\,Q^{cT} i\tau_2 Q^c + { k_{ll} \over M}L^T i\tau_2 L\,L^{cT} i\tau_2 L^c \nonumber \\ & &+ {1\over M} [j_L Q^T i\tau_2 Q\,Q^T i\tau_2 L + j_R Q^{cT} i\tau_2 Q^c \,Q^{cT} i\tau_2 L^c]. \label{nonsuperpot} \end{eqnarray} It has been shown that the addition of non-renormalizable terms suppressed by a high scale such as Planck mass, $M \sim M_{Planck} \sim 10^{19}$ GeV, with the minimal field content ensures the correct pattern of symmetry breaking in the supersymmetric left-right model. In particular the scale of parity breakdown is predicted to be in the intermediate region $M_R \mathrel{\vcenter{\hbox{$>$}\nointerlineskip\hbox{$\sim$}}} 10^{10} - 10^{11}$ GeV, and $R$-parity remains exact. This theory contains singly charged and doubly charged Higgs scalars with a mass of order $M_R^2/M_{Planck}$, which may be experimentally accessible. However, what is different is the nature of see-saw mechanism. Whereas in the renormalizable version the see-saw mechanism takes its canonical form, in the non-renormalizable case it takes a form similar to what occurs in the non-supersymmetric left-right models, with the neutrino mass depending on the unknown parameters of the Higgs potential. This in general leads to different neutrino mass spectra, which can be experimentally distinguished. \section{The tree-level upper bound on the lightest Higgs mass} \label{sec:higgs} Given the fact that the Higgs sector of SLRM models contain a large number of Higgs multiplets, and the VEVs of some of the Higgs fields involve possibly large mass scales compared to the electroweak and SUSY breaking scales, it is important to ask what is the mass of the lightest Higgs boson in these models. The upper bound on the lightest Higgs boson mass in the minimal model was derived in \cite{Huitu:1997rr} in the limit when $\kappa_1',\kappa_2',\sigma_L,v_{\Delta_L},v_{\delta_L} \rightarrow 0$, using the fact that for any Hermitean matrix the smallest eigenvalue must be smaller than that of its upper left corner $2 \times 2$ submatrix. In the basis in which the first two indices correspond to ($\Phi^0_1$,$\chi^0_2$), we find for matrix elements $m_{11}^2,m_{22}^2,m_{12}^2$ (see Appendix \ref{sec:1}) \begin{eqnarray} m_{11}^2&=& -m_{\Phi\chi }^2\frac{\kappa_2}{\kappa_1} +\frac 12 (g_L^2+g_R^2)\kappa_1^2,\nonumber\\ m_{22}^2&=& -m_{\Phi\chi }^2\frac{\kappa_1}{\kappa_2} +\frac 12 (g_L^2+g_R^2)\kappa_2^2,\nonumber\\ m_{12}^2&=& m_{\Phi\chi }^2 -\frac 12 (g_L^2+g_R^2)\kappa_1\kappa_2. \end{eqnarray} It follows that the upper bound on the lightest Higgs boson mass in the minimal supersymmetric left-right model can be written as \cite{Huitu:1997rr}: \begin{equation} m_{h}^2\leq \frac 12 (g_L^2+g_R^2)(\kappa_1^2 +\kappa_2^2)\cos^2 2\beta = \left(1+\frac{g_R^2}{g_L^2} \right) m_{W_L}^2 \cos^2 2 \beta , \label{eq:treeupper1} \end{equation} where $\tan\beta = \kappa_2/\kappa_1 $. The upper bound (\ref{eq:treeupper1}) is not only independent of the supersymmetry breaking parameters (as in the case of supersymmetric models based on $SU(2)_L \times U(1)_Y$), but it is also independent of the $SU(2)_R$ breaking scale, which, {\em a priori}, can be large. The upper bound is controlled by the weak scale vacuum expectation value, $\kappa_1^2+\kappa_2^2$, and the dimensionless gauge couplings ($g_L$ and $g_R$) only. Since the former is essentially fixed by the electroweak scale, the gauge couplings $g_L$ and $g_R$ determine the bound. We will see below that even when $\kappa_1',\kappa_2',\sigma_L,v_{\Delta_L},v_{\delta_L}$ are non-zero, the upper bound on the lightest Higgs mass at the tree level does not depend on either the right-handed breaking scale or the SUSY breaking scale. A general method to find an upper limit for the lightest Higgs mass in models based on $SU(2)_L\times U(1)_Y$ and maximally quartic potentials was presented in \cite{Comelli:1996xg}. We will apply this method to the case of SLRMs, with possible nonrenormalizable terms, in an appropriate manner. Consider a set of scalar fields $\Phi_j$ transforming under $SU(2)_L$, and define a discrete transformation $P:\, \Phi_j\rightarrow (-1)^{2T_j}\Phi_j$ of these fields, where $2T_j+1$ is the dimension of $SU(2)_L$ representation\footnote{One could as well choose any other spontaneously broken $U(1)$ \cite{Comelli:1996xg}, but $P$ provides the best limit in the case of SLRM.}. By setting the $P$-even fields to their VEVs, a normalized field can be defined in the direction of $SU(2)_L$ breaking, $\phi=\frac 1{v_0}\sum_{odd} v_i\Phi_i$, where $v_0^2=\sum_iv_i^2$, and all the fields orthogonal to $\phi$ have zero VEVs. Since the original lagrangian is invariant under the transformatiom $P$, the potential must be invariant under $P$, so that the potential contains only even powers of $\phi$, \begin{eqnarray} V(\phi )=V(0)-\frac 12 m^2\phi^2 +\frac 18 \lambda_\phi \phi^4 +\frac 16 A\phi^6+\dots . \end{eqnarray} We take into account only the leading nonrenormalizable terms and thus $\phi^6$ is the largest power in the potential $V(\phi)$. If $\phi$ were the mass eigenstate, then by using the minimization condition the Higgs mass would be $\lambda_\phi v_0^2 + 4Av_0^4$. In the general case, this expression provides an upper bound on the mass of the lightest Higgs boson, \begin{eqnarray} m_h^2\le \lambda_\phi v_0^2 + 4Av_0^4. \end{eqnarray} Since only the $SU(2)_L$ doublet fields are relevant for obtaining this bound, adding extra singlets or triplets in the model has no effect on the bound. Thus, there are three separate cases to be considered for deriving an upper bound on the mass of the lightest Higgs boson in the models discussed in the last section, namely: (A) $R$-parity is broken (sneutrinos get VEVs), (B) $R$-parity is conserved because there are additional triplets, and (C) $R$-parity is conserved because there are nonrenormalizable terms. Let us first consider the situation in the minimal model, case (A). We define a new neutral scalar field in the direction of the breaking in the $SU(2)_L$ doublet space: \begin{equation} \Phi^0 = \frac 1v \left( \kappa_1 \rm{Re} \,\Phi^0_1 +\kappa_1' \rm{Re} \,\Phi^0_2+ \kappa_2' \rm{Re} \,\chi^0_1 + \kappa_2 \rm{Re} \,\chi^0_2 + \sigma_L \rm{Re} \,\tilde\nu \right) , \end{equation} where \begin{equation} v^2 = \kappa_1^2+ \kappa_1'^2+\kappa_2^2+\kappa_2'^2+\sigma_L^2 . \end{equation} All the $SU(2)_L$ doublet fields which are orthogonal to $\Phi^0$ have vanishing VEVs. We calculate next the quartic term $(\Phi^0)^4$ in the potential. It has contributions from both the F-terms and the D-terms, and we find that the upper bound on the mass of the lightest Higgs boson in the minimal SLRM is \begin{equation} m_h^2 \leq \frac 1{2 v^2} \left[ g_L^2 (\omega_\kappa^2+\sigma_L^2)^2+ g_R^2 \omega_\kappa^4+g_{B-L}^2 \sigma_L^4 + 8 (h_{\Phi L} \kappa_1'+h_{\chi L} \kappa_2 )^2 \sigma_L^2 + 8 h_{\Delta_L}^2 \sigma_L^4 \right] , \label{eq:treeupper2} \end{equation} where \begin{equation} \omega_\kappa^2 = \kappa_1^2-\kappa_2^2-\kappa_1'^2+\kappa_2'^2 . \end{equation} As is evident, the upper bound (\ref{eq:treeupper2}) is independent of SUSY and right-handed breaking scales, and depends only on the dimensionless gauge and Yukawa couplings, and vacuum expectation values which are determined by the weak scale: \begin{equation} m_{W_L}^2 = \frac 12 g_L^2 \left( \kappa_1^2+\kappa_2^2+\kappa_1'^2+\kappa_2'^2 +\sigma_L^2+ 2 v_{\Delta_L}^2+2 v_{\delta_L}^2 \right) +{\cal{O}}\left(\frac{\kappa^{'2} m_{W_L}^2}{m_{W_R}^2} \right). \end{equation} The triplet VEVs $v_{\Delta_L}$ and $v_{\delta_L}$ must be small in order to maintain $\rho \simeq 1$. In the limit when $\kappa_1',\kappa_2',\sigma_L,v_{\Delta_L}, v_{\delta_L} \rightarrow 0$, the bound (\ref{eq:treeupper2}) reduces to the upper bound (\ref{eq:treeupper1}). It is obvious that the addition of extra triplets does not change this bound. Thus, the bound for the case (B), the SLRM with additional triplets to ensure that $R$-parity is not spontaneously broken, can be obtained from (\ref{eq:treeupper2}) by taking the limit $\sigma_L\rightarrow 0$. The total number of nonrenormalizable terms in case (C) is rather large. All the coefficients in nonrenormalizable terms are proportional to inverse powers of a large scale. Thus the largest contribution comes from those terms which have the smallest number of large scales and the largest number of potentially large VEVs from $SU(2)_R$ triplets. We recall that the terms of the form $(\Phi^0)^4$ and $(\Phi^0)^6$ in the potential are needed to determine the contributions to the mass bound. The nonrenormalizable terms have no effect on the contribution from $D$-terms. The leading terms in the superpotential which can give a $(\Phi^0)^4$ and $(\Phi^0)^6$ type F-term contribution are of the type \begin{eqnarray} W_{NR}=A{\rm Tr}(i\tau_2\Phi^Ti\tau_2\chi)\,{\rm Tr}(\Delta_R\delta_R)+ B{\rm Tr}(i\tau_2\Phi^Ti\tau_2\chi)^2, \end{eqnarray} i.e. one term with two and another with four bidoublet fields. Here $A,\,B\sim 1/M_{\rm Planck}$. With $SU(2)_L$ singlets fixed to their VEVs, the corresponding $F$-terms are \begin{eqnarray} V_{NR}={\cal{O}}(ABv_{\Delta_R}v_{\delta_R})(\Phi^0)^4 +{\cal{O}}(B^2)(\Phi^0)^6. \end{eqnarray} The contribution to the Higgs mass bound from these nonrenormalizable terms is \begin{equation} {\cal{O}}(v_R^2/M_{\rm Planck}^2)\langle \Phi^0\rangle ^2 + {\cal{O}}(1/M_{\rm Planck}^2)\langle \Phi^0\rangle ^4. \end{equation} If the VEV $v_R\sim 10^{10}$ GeV in these models, the contribution is numerically negligible. Therefore the upper bound for this class of models is essentially the same as in the case (B). \section{Radiative corrections} \label{sec:rad} Since it is known that the radiative corrections to the lightest Higgs mass are significant in the MSSM, as well as its extensions based on the $SU(2)_L \times U(1)_Y$, it is important to consider the radiative corrections to the upper bound on the lightest Higgs boson mass obtained in the previous section. In this section we discuss the one-loop radiative corrections to the upper bound on the lightest Higgs mass in the minimal supersymmetric left-right model, which was obtained in last section. We shall use the method of one-loop effective potential \cite{Coleman:1973tz} for the calculation of radiative corrections, where the effective potential may be expressed as the sum of the tree-level potential plus a correction coming from the sum of one-loop diagrams with external lines having zero momenta, \begin{equation} V_{\rm{1-loop}} = V_{\rm{tree}} + \Delta V_1 , \end{equation} where $ V_{\rm{tree}}$ is the tree level potential (\ref{eq:scalarpotential}) evaluated at the appropriate running scale $Q$, and $\Delta V_1$ is the one loop correction given by \begin{equation} \Delta V_1 = \frac 1{64 \pi^2} \sum_i (-1)^{2 J_i} (2 J_i+1) m_i^4 \left( \ln \frac{m_i^2}{Q^2} - \frac 3 2 \right) , \label{eq:deltav1} \end{equation} where $m_i$ is the field dependent mass eigenvalue of the $i$th particle of spin $J_i$. The dominant contribution to (\ref{eq:deltav1}) comes from top-stop ($t-\tilde t$) system. However, under certain conditions the contribution of bottom-sbottom ($b-\tilde b$) can be nonneglible. We shall include both these contributions in our calculations of the radiative corrections. In order to evaluate the contributions of top-stop and bottom-sbottom to (\ref{eq:deltav1}), we need the stop and sbottom mass matrices for the SLRM. From (\ref{eq:fterms}), (\ref{eq:dterms}) and (\ref{eq:softterms}), it is straightforward to calculate squark mass matrices \cite{Huitu:1997iy}. Ignoring the interfamily mixing, the part of the potential containing the stop and sbottom mass terms can be written as \begin{equation} V_{squark}=\left( \begin{array}{cc} U_L^* & U_R^* \end{array} \right) \tilde M_U \left( \begin{array} {c} U_L \\ U_R \end{array} \right) + \left( \begin{array} {cc} D_L^* & D_R^* \end{array} \right) \tilde M_D \left( \begin{array} {c} D_L \\ D_R \end{array} \right) , \end{equation} where the mass matrix elements for the stop are \begin{eqnarray} (\tilde M_U)_{ U_L^* U_L} &=& \tilde m_Q^2+ m_u^2 + \frac 14 g_L^2 \omega^2_\kappa - \frac 16 g_{B-L}^2 \omega^2_R , \nonumber\\ (\tilde M_U)_{ U_R^* U_L} &=& h_{\Phi Q} A_{\Phi Q} \kappa '_1+h_{\chi Q} A_{\chi Q}\kappa_2 - \mu_1 (h_{\phi Q}\kappa '_2+h_{\chi Q}\kappa_1)- 2h_{\phi Q}\mu '_1 \kappa_1- 2 h_{\chi Q}\mu''_1\kappa '_2\nonumber\\ &&+ (h_{\phi L }h_{\phi Q } +h_{\chi L }h_{\chi Q }) \sigma_L\sigma_R \nonumber\\ &=&\left[(\tilde M_U)_{ U_L^* U_R }\right]^*,\nonumber\\ (\tilde M_U)_{ U_R^* U_R} &=& m_{Q^c}^2+ m_u^2 + \frac 14 g_R^2(\omega^2_\kappa -2\omega^2_R ) + \frac 16 g_{B-L}^2 \omega^2_R , \end{eqnarray} while for the sbottom these are \begin{eqnarray} (\tilde M_D)_{ D_L^* D_L} &=& m_Q^2+ m_d^2 -\frac 14 g_L^2 \omega^2_\kappa - \frac 16 g_{B-L}^2 \omega^2_R ,\nonumber\\ (\tilde M_D)_{ D_R^* D_L} &=& -h_{\Phi Q} A_{\Phi Q} \kappa_1-h_{\chi Q} A_{\chi Q}\kappa '_2 + \mu_1 (h_{\phi Q}\kappa_2+h_{\chi Q}\kappa '_1) + 2h_{\phi Q}\mu '_1 \kappa '_1+ 2 h_{\chi Q}\mu''_1\kappa_2 \nonumber\\ & = &\left[(\tilde M_D)_{ D_L^* D_R }\right]^*,\nonumber\\ (\tilde M_D)_{ D_R^* D_R } &=& m_{Q^c}^2+ m_d^2 -\frac 14 g_R^2(\omega^2_\kappa -2\omega^2_R ) + \frac 16 g_{B-L}^2 \omega^2_R , \end{eqnarray} where top and sbottom squared masses are given by $(h_{\chi Q} \kappa_2)^2$ and $(h_{\Phi Q} \kappa_1)^2$, respectively, $m ^2_Q$, $m ^2_{Q^c}$, $A_{\Phi Q}$ and $A_{\chi Q}$ are soft supersymmetry breaking parameters (see eq. (\ref{eq:softterms})), and \begin{equation} \omega^2_R=v_{\Delta_R}^2-v_{\delta_R}^2-\frac 12\sigma_R^2, \; \omega^2_\kappa = \kappa_1^2+{\kappa '_2}^2-\kappa_2^2-{\kappa '_1}^2. \end{equation} In order that $SU(3)_C \times U(1)_{\rm{em}}$ is unbroken, none of the physical squared masses of squarks can be negative. Necessarily then all the diagonal elements of the squark mass matrices should be non-negative. Combining the diagonal elements of the stop and sbottom mass matrices leads to the inequality \begin{equation} m_Q^2+m_{Q^c}^2 \geq |\frac 12 g_R^2\omega^2_R|= \frac 12g_R^2|v_{\Delta_R}^2- v_{\delta_R}^2 -\frac 12 \sigma_R^2 |. \label{eq:susyineq} \end{equation} where we have ignored terms which are of the order of the weak scale or less. The eigenvalues $m^2_{\tilde t _{1,2}},m^2_{\tilde b _{1,2}}$ of the stop and sbottom mass squared matrices are given by ($m^2_{\tilde t _1} > m^2_{\tilde t _2}, m^2_{\tilde b _1} > m^2_{\tilde b _2}$) \begin{eqnarray} m^2_{\tilde t_{1,2}} = m^2_{\tilde t} \pm \Delta^2_{\tilde t} , \nonumber \\ m^2_{\tilde b_{1,2}} = m^2_{\tilde b} \pm \Delta^2_{\tilde b} , \label{eq:squark1} \end{eqnarray} where \begin{eqnarray} m^2_{\tilde t}& = &\frac 12 \left[ m ^2_Q+ m ^2_{Q^c} + 2 m^2_t + \frac 14 (g_L^2+g_R^2) \omega_\kappa^2-\frac 12 g_R^2 \omega_R^2 \right] , \nonumber \\ \Delta^2_{\tilde t} &= &\frac 12 \left\{ \left[ m ^2_Q- m ^2_{Q^c} + \frac 14 (g_L^2-g_R^2) \omega_\kappa^2 + \frac 12 g_R^2 \omega_R^2- \frac 13 g_{B-L}^2 \omega_R^2 \right] ^2 \right. \nonumber \\ && \left. + 4 \left[ h_{\chi Q} A_t \kappa_2 - h_{\chi Q}\tilde\mu_t \kappa_1 \right]^2 \right\} ^{\frac 12} , \label{eq:squark2} \end{eqnarray} and \begin{eqnarray} m^2_{\tilde b} &= &\frac 12 \left[ m ^2_Q+ m ^2_{Q^c} + 2 m^2_b - \frac 14 (g_L^2+g_R^2) \omega_\kappa^2+\frac 12 g_R^2 \omega_R^2 \right] , \nonumber \\ \Delta^2_{\tilde b} &= &\frac 12 \left\{ \left[ m ^2_Q- m ^2_{Q^c} - \frac 14 (g_L^2-g_R^2) \omega_\kappa^2 - \frac 12 g_R^2 \omega_R^2- \frac 13 g_{B-L}^2 \omega_R^2 \right] ^2 \right. \nonumber \\ && \left. + 4 \left[ h_{\Phi Q} A_b \kappa_1 - h_{\Phi Q}\tilde\mu_b \kappa_2 \right]^2 \right\} ^{\frac 12} . \label{eq:squark3} \end{eqnarray} Here we have defined \begin{eqnarray} &&\tilde\mu_t\equiv \mu_1+2 \mu_1' \frac{m_b}{m_t} \tan \beta,\;\; \tilde\mu_b\equiv \mu_1+2 \mu_1''\frac{m_t}{m_b} \cot \beta,\;\; A_t\equiv A_{\chi Q},\;\; A_b\equiv A_{\phi Q}. \end{eqnarray} Using eqs.\ (\ref{eq:squark1}), (\ref{eq:squark2}) and (\ref{eq:squark3}) in (\ref{eq:deltav1}), we have calculated the radiatively-corrected expressions for the matrix elements of the upper left corner $2 \times 2$ submatrix of the $10 \times 10$ CP-even Higgs mass matrix. After imposing the appropriate one-loop minimization conditions, we find the following form for the radiatively corrected upper left corner $2 \times 2$ submatrix of CP-even Higgs mass matrix: \begin{eqnarray} \left( \begin{array}{cc} \frac 12 (g_L^2+g_R^2) \kappa_1^2 & - \frac 12 (g_L^2+g_R^2) \kappa_1 \kappa_2 \\ - \frac 12 (g_L^2+g_R^2) \kappa_1 \kappa_2 & \frac 12 (g_L^2+g_R^2) \kappa_2^2 \end{array} \right) + \left( \begin{array}{cc} \tan \beta & -1 \\ -1 & \cot \beta \end{array} \right) \left( \frac \Delta 2 \right) + \frac{3 g_L^2}{16 \pi^2 m_{W_L}^2} \left( \begin{array}{cc} \Delta_{11} & \Delta_{12} \\ \Delta_{12} & \Delta_{22} \end{array} \right) ,\nonumber\\ \label{eq:squarkmass} \end{eqnarray} where \begin{eqnarray} \label{eq:radelements} \Delta &=& \left[ -2 m^2_{\Phi \chi} + \frac 3 {32 \pi^2} \left( \frac{g_L^2}{\sin^2 \beta} \frac{m_t^2}{m_W^2} \right) A_t \tilde\mu_t \frac{f(m_{\tilde t_1}^2)-f(m_{\tilde t_2}^2)} {m_{\tilde t_1}^2-m_{\tilde t_2}^2} \right. \nonumber\\ && \left. + \frac 3 {32 \pi^2} \left( \frac{g_L^2}{\cos ^2 \beta} \frac{m_b^2}{m_W^2} \right) A_b \tilde\mu_b \frac{f(m_{\tilde b_1}^2)-f(m_{\tilde b_2}^2)} {m_{\tilde b_1}^2-m_{\tilde b_2}^2} \right] , \label{eq:radelementsa} \end{eqnarray} \begin{eqnarray} &&f(x^2)=2x^2(\ln(x^2/Q^2)-1), \label{def:f} \end{eqnarray} and \begin{eqnarray} \Delta_{11} &= & \frac{m_b^4}{\cos^2 \beta} \left[ \ln \left( \frac{ m^2_{\tilde b_1} m^2_{\tilde b_2}}{m_b^4} \right) + \frac{2 A_b \left( A_b - \tilde\mu_b \tan \beta \right) }{m^2_{\tilde b _1}-m^2_{\tilde b _2}} \ln \left( \frac{m^2_{\tilde b_1}}{m^2_{\tilde b _2}} \right) \right] \nonumber \\ && + \frac{m_b^4}{\cos^2 \beta} \frac{A_b^2 \left( A_b - \tilde \mu_b \tan \beta \right) ^2 }{\left( m^2_{\tilde b _1} - m^2_{\tilde b _2} \right) ^2} g ( m^2_{\tilde b _1} , m^2_{\tilde b _2} ) + \frac{m_t^4}{\sin^2 \beta} \frac{\left( A_t - \tilde\mu_t \cot \beta \right) ^2 \tilde\mu_t ^2 }{ (m^2_{\tilde t _1} - m^2_{\tilde t _2})^2 } g ( m^2_{\tilde t _1} , m^2_{\tilde t _2} ) ,\nonumber\\ \label{eq:radelementsb} \end{eqnarray} \begin{eqnarray} \Delta_{22}& = & \frac{m_t^4}{\sin^2 \beta} \left[ \ln \left( \frac{m^2_{\tilde t_1 } m^2_{\tilde t _2}}{m_t^4} \right)+ \frac{2 A_t \left( A_t- \tilde\mu_t \cot \beta \right) }{m^2_{\tilde t _1}-m^2_{\tilde t _2}} \ln \left( \frac{m^2_{\tilde t _1}}{m^2_{\tilde t _2}} \right) \right] \nonumber \\ && + \frac{m_t^4}{\sin^2 \beta} \frac{A_t^2 \left( A_t - \tilde \mu_t \cot \beta \right) ^2}{\left( m^2_{\tilde t _1}-m^2_{\tilde t _2} \right) ^2 } g( m^2_{\tilde t _1},m^2_{\tilde t _2} ) + \frac{m_b^4}{\cos ^2 \beta} \frac{ \left( A_b - \tilde \mu_b \tan \beta \right) ^2 \tilde\mu_b ^2 }{ \left( m^2_{\tilde b _1}- m^2_{\tilde b _2} \right) ^2 } g(m^2_{\tilde b _1}, m^2_{\tilde b _2}) ,\nonumber\\ \label{eq:radelementsc} \end{eqnarray} \begin{eqnarray} \Delta_{12} &= & \frac{m_t^4}{\sin^2 \beta} \frac{ \left( A_t - \tilde \mu_t \cot \beta \right) \left( -\tilde\mu_t \right) }{m^2_{\tilde t _1}-m^2_{\tilde t_2}} \times \left[ \ln \left( \frac{m^2_{\tilde t_1}}{m^2_{\tilde t _2}} \right) + \frac{ A_t \left( A_t - \tilde\mu_t \cot \beta \right) }{m^2_{\tilde t _1}-m^2_{\tilde t _2}} g( m^2_{\tilde t _1}, m^2_{\tilde t _2} ) \right] \nonumber \\ && + \frac{m_b^4}{\cos^2 \beta} \frac{ \left( A_b - \tilde\mu_b \tan \beta \right) \left( -\tilde\mu_b \right) } {m^2_{\tilde b _1}-m^2_{\tilde b _2}} \times \left[ \ln \left( \frac{m^2_{\tilde b _1}}{m^2_{\tilde b _2}} \right) + \frac{A_b \left( A_b- \tilde\mu_b \tan \beta \right) }{m^2_{\tilde b _1}-m^2_{\tilde b _2}} g(m^2_{\tilde b _1},m^2_{\tilde b _2} ) \right] ,\nonumber\\ \label{eq:radelementsd} \end{eqnarray} with \begin{equation} g(m_1^2,m_2^2) = 2- \frac{m_1^2 + m_2^2}{m_1^2-m_2^2} \ln \left( \frac{m_1^2}{m_2^2} \right) . \label{eq:radelementse} \end{equation} We have neglected D-terms in the squark masses, because these are small, and, since we are including only the quark-squark contributions to $\Delta V_1$, in order to gain approximate independence of the renormalization scale $Q$ (see also the inequality (\ref{eq:susyineq})). Using eqs.\ (\ref{eq:squarkmass}) and (\ref{eq:radelements}), we obtain the one-loop radiatively corrected upper bound on the lightest Higgs boson mass in the SLRM: \begin{eqnarray} m_h^2 &\leq& \frac 12 \left[ (g_L^2+g_R^2) \left( \kappa_1^2+\kappa_2^2 \right) \cos^2 2 \beta + \frac{3 g_L^2}{8 \pi^2 m^2_{W_L}} \left( \Delta_{11} \cos^2 \beta + \Delta_{22} \sin^2 \beta + \Delta_{12} \sin 2 \beta \right) \right]\nonumber\\ \label{eq:upperbound1} \end{eqnarray} For $\tan\beta \mathrel{\vcenter{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}} 20 $, one can neglect the b-quark contribution in the radiative corrections. Then, in the approximation \cite{Carena:1995bx,Casas:1995us} \begin{equation} |m_{\tilde t_1}^2 - m_{\tilde t_2}^2| \ll |m_{\tilde t_1}^2 + m_{\tilde t_2}^2|, \end{equation} the upper bound (\ref{eq:upperbound1}) on the lightest Higgs mass reduces to \begin{eqnarray} m_h^2 &\leq& \frac 12 \left[ (g_L^2+g_R^2) \left( \kappa_1^2+\kappa_2^2 \right) \cos^2 2 \beta\right.\nonumber\\ &&\left. + \frac{3 g_L^2 m_t^4}{8 \pi^2 m^2_{W_L}} \left( \ln (\frac{m_{\tilde t_1}^2 m_{\tilde t_2}^2}{m_t^4}) +2 \frac{\tilde A_t^2}{M_s^2} (1 - \frac{\tilde A_t^2}{12M_s^2}) - 8 \frac{\mu_1''^4}{3M_s^4} \right ) \right ] \label{eq:upperbound2} \end{eqnarray} where $\tilde A_t = A_t - \mu_1 \cot\beta$, and $M_s$ is the supersymmetry breaking scale ($2M_s^2 = m_{\tilde t_1}^2 + m_{\tilde t_2}^2$). In this limit the upper bound eq.\ (\ref{eq:upperbound2}) on the lightest Higgs mass in the supersymmetric left-right model differs in form from the corresponding MSSM upper bound only because of $\mu_1''$ being nonzero. The upper bound is maximised by \begin{equation} |\tilde A_t| = (\sqrt 6) M_s \end{equation} for a given value of $\mu_1''$. \begin{figure}[t] \leavevmode \begin{center} \mbox{\epsfxsize=8.truecm\epsfysize=8.truecm\epsffile{mh2.eps} \epsfxsize=8.truecm\epsfysize=8.truecm\epsffile{mh20.eps}} \end{center} \caption{\label{mh1}The upper bound on the radiatively corrected mass of the lightest neutral Higgs boson for two different values of $\tan\beta$. The right-handed scale $M_R$ is indicated in the Figure. The bi- and trilinear soft supersymmetry breaking parameters are 1 TeV (solid line) and 10 TeV (dashed line). Supersymmetric Higgs mixing parameters are assumed to vanish, and $m_{top}=175$ GeV.} \end{figure} The radiatively corrected upper bound (\ref{eq:upperbound1}) on the mass of the lightest Higgs boson is plotted in Fig.\ref{mh1} as a function of the large scale $\Lambda $ up to which the supersymmetric left-right model remains perturbative. The upper bound comes from the requirement that all the gauge couplings of the SLRM remain perturbative below the scale $\Lambda $. We have taken into account the dominant radiative corrections coming from the quark and squark loops in our calculations. In Fig.\ref{mh1} we have taken two values of $\tan\beta=2$ and $\tan\beta=20$. In the figure the upper bound is shown for two different values of the $SU(2)_R$ breaking scale, $M_R=10$ TeV and $M_R=10^{10}$ GeV, respectively, and for two values of soft supersymmetry breaking mass parameter, $M_s=1$ TeV and $M_s=10$ TeV. It is seen from this figure that if the difference between the $SU(2)_R$ breaking scale and the large scale $\Lambda$ is more than two orders of magnitude, the radiatively corrected upper bound on the mass of the lightest Higgs boson remains below 250 GeV. For large values of $\Lambda $ the upper bound is below 200 GeV. The upper bound increases with increasing $M_R$ and with increasing soft supersymmetry breaking parameters. It is considerably larger than the corresponding upper bound in the MSSM. \section{Couplings of the lightest neutral Higgs to fermions and the decoupling limit} \label{sec:decoup} In order to study the phenomenology of the lightest Higgs boson in the SLRM, we must obtain its couplings to the fermions and see how these differ from the corresponding MSSM couplings. The major difference between the two models arises due to the triplet Higgs couplings to leptons. As discussed earlier, the lightest neutral Higgs is composed mainly of the bidoublet fields, and we do not expect the triplet couplings to have a large effect on the lightest neutral Higgs branching ratios. Another difference in the Higgs couplings in the SLRM and the MSSM arises because of the mixing between the Higgs(ino) and lepton sectors in models where $R$-parity is spontaneously broken. Since the lightest chargino contains gaugino and higgsino admixture, one might expect that in some region of the parameter space the couplings are significantly different from the couplings in the MSSM. We will take the lightest chargino to correspond to the $\tau $-lepton. It will be denoted by $\tau$ in the following. Thus, it is important to study in detail the couplings of the lightest Higgs boson to the $\tau$. \begin{figure}[t] \leavevmode \begin{center} \mbox{\epsfxsize=8.truecm\epsfysize=8.truecm\epsffile{kuva1a.epsi}} \end{center} \caption{\label{tau2} The branching ratio of the lightest neutral CP even Higgs boson $h$ to fermion pairs in SLRM (solid curves) as a function of the second lightest Higgs $H$ mass. The tree level mass of the lightest Higgs is $\sim$78 GeV and $\tan\beta=50$. The dotted curves correspond to MSSM.} \end{figure} In Figure \ref{tau2} we have plotted the branching ratio of the lightest neutral Higgs boson to $b\bar b$, $c\bar c$, and $\tau^+\tau^-$ pair both in the MSSM and in the SLRM for large $\tan\beta$ and a tree-level mass $m_{h}\sim 78$ GeV as a function of the second lightest neutral Higgs boson mass. We note that for the second lightest neutral Higgs with mass $\sim 160$ GeV, the branching ratio to bottom pair almost vanishes while the branching ratio to a $\tau$ pair is enhanced. This behaviour occurs in a transition region where the composition of the light Higgs bosons change rapidly, and the lightest Higgs coupling to $b\bar b$ becomes negligible. Similar phenomenon occurs for small $\tan\beta$. It is well known that in the left-right symmetric models problems with FCNC are expected if several light Higgs bosons exist \cite{EGN}. In the supersymmetric case, the problem could be expected to be more severe, since the potential is more constrained. Furthermore, in the four Higgs-doublet model \cite{MR}, which is also the number of doublets in the supersymmetric left-right model, it was found that the potential is always real, thus preventing spontaneous violation of CP. Strong limits on the FCNC Higgs were obtained in \cite{pospelov} assuming that the only source for CP violation is the complex phase in the Kobayashi-Maskawa matrix and that the model is manifestly left-right symmetric. If these conditions are relaxed, for example by defining the symmetry transformation by Eq. (4), a weakened lower limit for second lightest neutral Higgs is obtained from the neutral meson mass difference\footnote{When R-parity is broken, the ground state can violate CP via the soft breaking terms involving sneutrinos \cite{mohapatra}.}. The limit is less severe for large values of $\tan\beta$, but remains of the same order than in the nonsupersymmetric left-right model. Taking into account the uncertainties in calculating the mass difference, $m_{H_{FCNC}}\mathrel{\vcenter{\hbox{$>$}\nointerlineskip\hbox{$\sim$}}} {\cal{O}}(1$ TeV). Thus the relevant limit to discuss is the one in which all the neutral Higgs bosons, except the lightest one, are heavy. In order to consider in detail the neutral Higgs couplings to the $\tau$-leptons, we'll need the interaction Lagrangian. We have seen that at the tree level the lightest Higgs can be written as \begin{eqnarray} h=\frac{1}{v}\sum_k\left(1+{\cal{O}}\left( \frac{m_h^2}{m_{H_2}^2}\right)\right) \langle\phi_k\rangle \phi_k + \sum_k{\cal{O}}\left( \frac{m_h}{m_{H_2}}\right)\psi_k , \end{eqnarray} where $\phi_k$ are the scalar bidoublet fields and $\psi_k$ are all the other fields. In the decoupling limit $m_{H_2}\gg m_h$, the relevant Higgs-fermion interaction Lagrangian can be written as \begin{eqnarray} L=-\frac 1v\psi_i^+ v_ky_{ijk}\psi_j^-h +h.c., \end{eqnarray} where $\psi^{-T}=(-i\lambda_L^-,-i\lambda_R^-,\phi_2^-,\chi_2^-,\Delta_R^-, \tau )^T$ and $\psi^{+T}=(-i\lambda_L^+,-i\lambda_R^+,\phi_1^+,\chi_1^+,\delta_R^+, \tau^c )^T$. The mass eigenstates $\chi^\pm$ are found by diagonalizing the mass matrices by unitary matrices $U,\,V$ \cite{Haber:1985rc}, \begin{eqnarray} \chi^+=V\psi^+,\;\;\; \chi^-=U\psi^-. \end{eqnarray} The diagonal chargino mass matrix is \begin{eqnarray} M_{\chi^\pm}^2=VX^\dagger XV^\dagger=U^*XX^\dagger U^T. \end{eqnarray} We decompose\footnote{For a more general treatment, see \cite{kai}.} the mass matrix $X$ and the diagonalizing matrices $V$ and $U$ as \begin{eqnarray} X=X_0+\epsilon X_1,\;\;\; V=(1+\epsilon Y)V_0,\;\;\; U=(1+\epsilon Z)U_0 , \end{eqnarray} where the contribution of the terms proportional to $\epsilon $ in determining the $\tau $ lepton eigenstate is small. One can solve for the matrix elements to the order $O(\epsilon^2)$, \begin{eqnarray} Y_{jk}&=&\left\{\begin{array}{l}0,\;j=k,\\ \frac{[V_0(X_0^\dagger X_1+X_1X_0)V_0^\dagger]_{jk}} {(M_0^2)_{jj}-(M_0^2)_{kk}}, \;j\neq k \end{array},\right . \\ Z_{jk}&=&\left\{\begin{array}{l}0,\;j=k,\\ \frac{[U_0^*(X_0X_1^\dagger+X_1X_0^\dagger )U_0^T]_{jk}} {(M_0^2)_{jj}-(M_0^2)_{kk}}, \;j\neq k \end{array},\right . \end{eqnarray} where we have denoted $M_0^2\equiv U_0^*X_0X_0^\dagger U_0^T$. In terms of the mass eigenstates, the interaction Lagrangian can be written as \begin{eqnarray} L_{int}=-\frac 1v(\chi^{-T}U^*)_iv_ky_{ijk}(V^\dagger\chi^+)_j h +h.c. \end{eqnarray} Here we denote $C\equiv \frac 1v U^*v_ky_{ijk}V^\dagger $, where $C$ is the coupling matrix for charginos. Thus the diagonal couplings to $\tau$ are given by \begin{eqnarray} C_{11}\sim\frac 1v \sum_{jk}U_{01j} X_{jk}V_{01k}^*. \end{eqnarray} The chargino mass matrix can be written as $X=X_0+\epsilon X_1$, where \begin{eqnarray} X_0=\left(\begin{array}{cccccc} m_L&0&0&0&0&0\\ 0&m_R&0&0&\sqrt{2}g_R v_{\delta_R} &g_R\sigma_R\\ 0&0&0&\mu_1&0&0\\ 0&0&\mu_1&0&0&0\\ 0&-\sqrt{2} g_Rv_{\Delta_R} &0&0&\mu_2&-\sqrt{2}h_\Delta\sigma_R\\ 0&0&-h_{\Phi L}\sigma_R&-h_{\chi L}\sigma_R&0&0 \end{array}\right) \end{eqnarray} and \begin{eqnarray} \epsilon X_1=\left(\begin{array}{cccccc} 0&0&0&g_L\kappa_2&0&0\\ 0&0&-g_R\kappa_1&0&0&0\\ g_L\kappa_1&0&0&0&0&0\\ 0&-g_R\kappa_2&0&0&0&0\\ 0&0&0&0&0&0\\ 0&0&0&0&0&-h_{\Phi L}\kappa_1 \end{array}\right). \end{eqnarray} We have here assumed for simplicity that $\mu_1'=\mu_1''=0$, and $\sigma_L=v_{\delta_L}=v_{\Delta_L}=0$. The matrix $X_0$ has one zero eigenvalue corresponding to the $\tau $ mass, $(M_0)^2_{11}$=0. The $\tau$ mass is \begin{eqnarray} m_\tau &=&(U^*XV^\dagger)_{11}\nonumber\\ &=&(Z_1^*U_0^*X_0V_0^\dagger +U_0^*X_0V_0^\dagger Y_1+U_0^*X_1V_0^\dagger)_{11}. \end{eqnarray} For the SM one has $v C_{11}=(U_0^*X_1V_0^\dagger )_{11}=m_\tau $. Thus, in the decoupling limit, we obtain the following result for the ratio of the Yukawa couplings $y_{SM}$ and $y_{SLRM}$: \begin{eqnarray} \frac{y_{SLRM}}{y_{SM}}=1+\frac{(Z_1^*M_0 -M_0 Y_1)_{11}} {(U_0^*X_1V_0^\dagger)_{11}}=1, \end{eqnarray} since the matrix $M_0$ is diagonal. Even if the $\tau$'s contained large fraction of gauginos or higgsinos, the couplings responsible for Higgs decays to the charged leptons or quarks are the same as in the Standard Model, since the physically relevant parameter region is close to the decoupling limit. Contrary to the MSSM, one might have the lightest Higgs decaying to a $\tau$-lepton and a heavier chargino. Whether this decay mode is kinematically possible is much more model dependent than the decay to leptons, and we cannot say anything general about it. \section{The lightest doubly charged Higgs boson} \label{sec:double} We have seen that the neutral Higgs sector contains one relatively light Higgs boson, which however may be heavier as compared to the lightest Higgs in the MSSM, and which has MSSM like couplings to the fermions of the model. Thus, one cannot tell from the properties of the lightest Higgs boson, about the nature of the model. If a light Higgs is found before any supersymmetric particles are observed, one would not even know whether it is the Higgs boson of a supersymmetric model. Altogether there are four doubly charged Higgs bosons in the SLRM, of which two are right-handed and two left-handed. The mass matrices of the left-handed triplets depend on the soft SUSY breaking parameters, left-triplet VEVs, and the weak scale. Thus, their masses are expected to be of the same order as the soft terms. The mass matrix for the right-handed scalars depends on the right-triplet VEV instead of the left-triplet VEV. Nevertheless, it was noticed in \cite{Huitu:1995zm} that in the SLRM with broken R-parity the lightest doubly charged scalar tends to be light. Also, it was shown in \cite{CM58} that in the nonrenormalizable case it is possible to have light doubly charged Higgs bosons. On the other hand, in the nonsupersymmetric left-right model all the doubly charged scalars tend to have a mass of the order of the right-handed scale \cite{GGMKO}. This is also true in the SLRM with enlarged particle content \cite{AMRS}. Thus a light doubly charged Higgs would be a strong indication of a supersymmetric left-right model with minimal particle content. \begin{figure}[t] \leavevmode \begin{center} \mbox{\epsfxsize=8.truecm\epsfysize=8.truecm\epsffile{mpp_ad.eps}} \end{center} \caption{\label{mpp_ad}The mass $m^{++}$ of the lightest doubly charged Higgs boson as a function of the soft trilinear coupling $A_\Delta$ for different values of the right-handed sneutrino VEVs $\sigma_R$. We have varied $\sigma_R$ in its allowed range of 100 GeV to 8.45 TeV as indicated in the figure.} \end{figure} In the case of broken R-parity, the mass matrix of the right-handed doubly charged scalars can be obtained from the potential in appendix A. The allowed parameter space is strongly constrained by demanding that all the eigenvalues of the mass squared scalar matrices remain positive. In order to find the relevant parameter region, we have studied the bounds on some of the masses in the model. Although these may not be important as actual bounds, they restrict the parameter space. {}In the following we shall assume that $\tan\beta > 1$. {}From the bound obtained from submatrix $(\Phi_2^{0r},\chi_1^{0r})$, we find the following constraint \begin{eqnarray} -\frac 12 g_L^2(\kappa_2^2-\kappa_1^2)-\frac 12 g_R^2D -(h_{\chi L}^2\kappa_2^2-h_{\phi L}^2\kappa_1^2)\sigma_R^2\geq 0, \label{constr1} \end{eqnarray} where $D=2v_{\Delta_R}^2-2v_{\delta_R}^2-\sigma_R^2+\kappa_2^2-\kappa_1^2$. If the terms with bidoublet lepton Yukawa couplings are ignored, and $\tan\beta> 1$, Eq. (\ref{constr1}) indicates that $D<0$ and thus $2v_{\delta_R}^2+\sigma_R^2>2v_{\delta_R}^2$. Other useful constraints follow from $(\Delta_R^{--},\delta_R^{++})$ and $(\Delta_R^{0i},\delta_R^{0i})$ mass matrices, \begin{eqnarray} (A_\Delta v_{\Delta_R}-4h_{\Delta_R}^2v_{\Delta_R}^2+h_{\Delta_R}\mu_{2r} v_{\delta_R})\sigma_R^2+g_R^2D(v_{\delta_R}^2-v_{\Delta_R}^2)&\geq& 0, \nonumber\\ A_\Delta v_{\Delta_R}+h_{\Delta_R}\mu_{2r}v_{\delta_R}&\geq & 0. \end{eqnarray} We have studied in Figure \ref{mpp_ad} an example with the soft masses and right-handed breaking scale, as well as the $\mu_{2R}$ parameter, of the order of 10 TeV. The maximum Majorana Yukawa coupling allowed by positivity of the mass eigenvalues in this case is $h_\Delta\sim 0.4$. For $h_\Delta = 0.4$ we have plotted the allowed doubly charged Higgs mass $m^{++}$ as a function of $A_\Delta $ for fixed $\sigma_R= 100$ GeV$,\dots,8.45$ TeV. It is seen that relatively narrow bands of $A_\Delta$ and $\sigma_R$ are allowed. Even in the maximal case the mass of the doubly charged scalar $m^{++}\sim 1 $ TeV and from the figure we see that the lightest of the doubly charged scalars can be as light or even lighter than the lightest neutral Higgs boson. The mass of the dangerous flavour changing Higgs boson is $\sim \frac{1}{\sqrt{2}} g_R \sqrt{D}$. The mass of the lightest doubly charged Higgs, $m_{H_1^{++}}$, in the case of nonrenormalizable couplings has been considered in \cite{CM58,DM,AMRS}. A general statement in these works is that the mass is $m_{H_1^{++}} \sim v_R^2/M $, where $M$ is the scale of the nonrenormalizable terms. Relevant constraints on the parameters follow again from the positivity of the mass bounds. {}From the submatrix $(\Delta_R^-,\delta_R^+)$, a bound on $m_{H_1^+}$ can be obtained and we get \begin{eqnarray} 0 &<&\frac 12 g_R^2 (v_{\delta_R}^2-v_{\Delta_R}^2) (\kappa_2^2-\kappa_1^2). \end{eqnarray} This gives a condition $v_{\delta_R}/v_{\Delta_R}>1$. {}From the bound on $m_{h}$ from $(\phi_2^0,\chi_1^0)$ submatrix, we find \begin{eqnarray} 0 < \frac 12 [-g_L^2(\kappa_2^2-\kappa_1^2) -g_R^2(2v_{\Delta_R}^2-2v_{\delta_R}^2 -\kappa_2^2+\kappa_1^2)] \end{eqnarray} Thus the $D$-term, $D=2v_{\Delta_R}^2-2v_{\delta_R}^2+\kappa_2^2-\kappa_1^2$ has to be negative. {}From the mass bound on the doubly charged Higgses, we get \begin{eqnarray} 0&<&m_{H_1^{++}}^2\nonumber\\ &<&[-g_R^2(v_{\Delta_R}^2-v_{\delta_R}^2) D]/ (v_{\Delta_R}^2+v_{\delta_R}^2) +\frac 1M 8b_Rv_{\Delta_R}v_{\delta_R}\mu_{2R} +\frac1{M^2}4b_R(2a_R+b_R)v_{\Delta_R}^2v_{\delta_R}^2.\nonumber\\ \label{mppbound} \end{eqnarray} We see that the $b_R$-parameter must necessarily be nonvanishing. \begin{figure}[t] \leavevmode \begin{center} \mbox{\epsfxsize=8.truecm\epsfysize=8.truecm\epsffile{mppnonren1.eps} \epsfxsize=8.truecm\epsfysize=8.truecm\epsffile{mppnonren2.eps}} \end{center} \caption{\label{nonren} Mass of the doubly charged Higgs as a function of the nonrenormalizable $b_R$-parameter. In a) $v_R^2 / M = 10^4 $ GeV and $D =m_{ soft }^2 $, while in b) $v_R^2/M=10^{2}$ GeV and $D =(3$ TeV)$^2$ (solid line), except for $m_{soft} =10$ TeV also $D=(10$ TeV$^2$ is shown (dashed line). The soft supersymmetry breaking parameters are marked in the figure, as well as the $b_R$ parameters. The curves are brought closer to each other for convenience. In both figures $\tan\beta=50$, $M=10^{10}$ GeV, $\mu_{2R}=1$ TeV and $\mu_1=\mu_1'=\mu_1''=500$ GeV. In a) $M_R=10^7$ GeV and in b) $M_R=10^6$ GeV.} \end{figure} In Fig. \ref{nonren} we have plotted the mass of the lightest doubly charged Higgs boson as a function of the nonrenormalizable $b_R$-parameter for two values of the ratios, $v_R^2/M=10^2$ GeV and $v_R^2/M=10^4$ GeV. One should note that at low energies $b_R$ is the only nonrenormalizable parameter, which is necessarily nonvanishing. In particular one can find solutions with $a_R=0$. In Appendix B we give examples of mass spectra using the tree level potential for the renormalizable and nonrenormalizable models. One can expect that radiative corrections will change the spectrum somewhat, but qualitative features should remain unchanged. The value of $b_R$ needed to produce a proper minimum depends on the other parameters in the model. In Fig. \ref{nonren} a) $v_R^2/M=10^4$ GeV and $\sqrt{D}$ is the same as the soft supersymmetry breaking parameters. The mass of the flavour changing Higgs is roughly proportional to the value of $\sqrt{D}$. For larger values of the $D$-term, larger $b_R$'s are needed to retain positive doubly charged Higgs mass, as seen from Eq. (\ref{mppbound}). It is seen that the mass of the doubly charged Higgs rises rapidly with $b_R$ to multi-TeV region as is expected for large $v_R^2/M$. In Fig.\ref{nonren} b) $v_R^2/M=10^2$ GeV, and $\sqrt{D}=3$ TeV (solid lines) for all the values of the soft supersymmetry breaking parameters that are plotted. This shows the dependence on the soft mass parameters. For the soft mass parameter of 10 TeV we have plotted for comparison also the dependence on $b_R$ when $\sqrt{D}=10$ TeV (dashed line). It is seen that increasing the soft parameter, while the $D$-term remains constant, smaller $b_R$'s are needed to produce positive masses. In Fig. (\ref{nonren}) we have taken $\tan\beta$ to be 50. Large $\tan\beta$ somewhat increases the masses of the neutral and pseudoscalar Higgses and decreases the mass of the lightest doubly charged Higgs. With both values of $v_R^2/M$ it is possible to obtain a doubly charged Higgs, which is light enough to be detected in future experiments, but $m_{H^{++}}$ increases with $b_R$ much faster for large $v_R^2/M$. The collider phenomenology of the doubly charged scalars has been actively studied, since they appear in several extensions of the Standard Model, can be relatively light and have clear signatures. Here we will discuss the production of the doubly charged scalar in colliders \cite{prod,HMPR}. If the light doubly charged Higgs $H^{++}$ is lighter than any of the supersymmetric partners, the dominant decay mode of $H^{++}$ is to like sign leptons. Kinematically, production of a single doubly charged scalar would be favoured. If the couplings between electrons and triplet Higgses are large enough, the doubly charged particles can be detected in $e^+e^-$ linear colliders in single production almost up to the kinematical limit. The $l^-l^-$ colliders are especially useful in studying the doubly charged Higgs, since for nonzero triplet couplings it can be produced as an s-channel resonance. The possibilities for single production in hadron colliders in $WW$ fusion does not depend on the coupling to leptons, but it does depend, in addition to $m_{H^{--}}$, on the VEV of the triplet to which the doubly charged scalar belongs. If $m_{W_R}\sim 1$ TeV, one can detect doubly charged Higgs bosons with $m_{H^{++}}\sim 1$ TeV at the LHC \cite{HMPR}. The advantage of pair production of the doubly charged scalars compared to the single production is that it is relatively model independent. It can occur even if $W_R$ is very heavy, as in the nonrenormalizable case, or the triplet Yukawa couplings are very small. The doubly charged Higgses can be produced in $f\bar f\rightarrow\gamma^*,Z^*\rightarrow H^{++}H^{--}$ both at lepton and hadron colliders, if kinematically allowed. At LHC this cross section falls off rapidly close to $m_{H^{++}}\sim 500$ GeV \cite{HMPR}. The pair production cross section at Tevatron and at LEP II are given in \cite{DM}. For $e^+e^-$ linear colliders the pair production cross section for $\sqrt{s}=500$ GeV is $106\, (78)$ fb for $m_{H^{++}}\mathrel{\vcenter{\hbox{$<$}\nointerlineskip\hbox{$\sim$}}} 230\, (240)$ GeV. Thus the detection of doubly charged scalars in pair production is possible close to the kinematical limit. \section{Conclusions} \label{sec:conclusion} Supersymmetric left-right models are well motivated extensions of the MSSM, since they conserve $R$-parity as a consequence of gauge invariance. We have made a detailed study of the Higgs sector in these models. The lightest CP even Higgs boson can be considerably heavier than in the MSSM, but its couplings to fermions remain similar to the couplings of the Standard Model Higgs boson. If a Higgs, which nevertheless is too heavy to be the Higgs boson of the MSSM, is detected, one should consider extended supersymmetric models such as the ones studied in this paper. In the SLRM with the minimal particle content one has typically also a light doubly charged Higgs boson. If this particle is found, it is a strong indication of the SLRM with minimal particle content. We wish to emphasize the importance of studying the full mass matrices, including the soft mass parameters in determining the low energy mass spectrum, especially the masses of the light scalar particles. \section{Acknowledgments} One of us (PNP) would like to thank the Helsinki Institute of Physics, where this work was started, and to DESY, where it was completed, for hospitality. The work of PNP is supported by the University Grants Commission, India under the project 10-26/98(SR-I). The work of KH and KP is partially supported by the Academy of Finland (no. 44129).
\section{Introduction} It is one of the most intriguing properties of black holes in general relativity that one can derive a set of laws, called the laws of black hole mechanics, which are formally equivalent to the laws of thermodynamics \cite{Haw}. For instance, the first law of thermodynamics, which relates variations of the internal energy and the work done to the variation of the entropy, has its counterpart in black hole mechanics. The first law for black holes in general relativity relates the variation of the mass and angular momentum of the black hole to the change in the area of its event horizon. This then leads to the celebrated Bekenstein-Hawking area law \cite{Bek}, which expresses the black hole entropy in terms of the horizon area. The first law is quite a remarkable result as it relates global quantities of the black hole, such as its energy or mass and its angular momentum, which can be entirely determined from the behaviour of the fields at spatial infinity, to the horizon area which is defined at the inner boundary of the black hole solution. Also, the connection with thermodynamics suggests a possible interpretation of the entropy in terms of microstates. Such an interpretation has recently been provided in the context of string theory \cite{StromVafa}. {From} the point of view of the field theory, be it fundamental or effective, it is rather surprising that variations near the outer boundary at spatial infinity are related to variations near the inner boundary at the horizon. Moreover, an effective field theory action will contain more than just the standard Einstein-Hilbert term and will depend on higher derivatives of the fields. So one may wonder whether there could be a natural principle that explains the behaviour implied by the first law of black hole mechanics for more generic field theories. Such a principle can be provided by making use of the concept of a surface charge, and a specific proposal for this charge was put forward by Wald \cite{Wald}. The surface charge is related to the conventional Noether current. In a certain sense there is no Noether current associated with a local gauge symmetry, but there exists a current associated with the residual invariance of a background configuration. When evaluating this current subject to the field equations, current conservation becomes trivial and the current takes the form of an improvement term, i.e., the derivative of an antisymmetric tensor. This antisymmetric tensor, sometimes called the Noether potential, can be written down for arbitrary gauge parameters. Integration of this potential over the boundary of some (spatial) hypersurface leads to a surface charge, which, when restricting the gauge transformation parameters to those that leave a certain background invariant, is equal to the Noether charge in the usual sense. Variations of this surface charge that continuously connect solutions of the equations of motion thus relate the various surface contributions and this is how eventually the first law of black hole mechanics follows (this is, for example, reviewed in \cite{JacKaMy,IWald,WaldRev}). In this approach the entropy is related to the integral of the Noether potential over the horizon. The entropy defined in this way is a local geometric quantity and equals the Bekenstein-Hawking entropy (i.e., one fourth of the horizon area) for Einstein gravity. But for more general actions containing terms of higher order in derivatives the entropy formula contains additional terms. Since the low-energy effective action constitutes the macroscopic or `thermodynamic' level of description of quantum gravity, the laws of black hole mechanics have to hold for such effective actions if they are more than accidental analogies. Naturally, since string theory is a candidate for a theory of quantum gravity, one expects to be able to describe black holes also at the microscopic or `statistical mechanics' level. Progress into this direction has been made after nonperturbative dualities, D-branes and the M-theory description of string theory were discovered (see, for example, \cite{Pol,Mal} for a review). In particular, quantitative agreement has been found between macroscopic black hole entropies extracted from supergravity solutions and microscopic entropies calculated by counting excitations of D-branes and M-branes. A crucial ingredient in establishing agreement is extended supersymmetry, because one needs to interpolate between different regimes of the theory, such as the low-energy regime and the regime of small string coupling. This interpolation is possible for extremal black holes which are BPS solitons of extended supergravity. The resulting expressions for the entropy concern extremal charged black holes and depend exclusively on the electric and magnetic charges. These results are obtained in cases where (some of) these charges are large, but sometimes it is also possible to evaluate some of the subleading corrections. The comparisons between macroscopic and microscopic entropy have mostly been made in situations where the distinction between the Bekenstein-Hawking area law and more general definitions of macroscopic black hole entropy are immaterial. Only recently has it become apparent that microscopic entropy formulae contain subleading corrections which on the macroscopic level are due to higher-derivative interactions in the field theory \cite{MalStrWit,Vaf,CarDeWMoh}. In this case the correct macroscopic definition of entropy is crucial for finding agreement, as we will review below. The structure of this paper is as follows. First we will review the derivation of the Noether potential for Yang-Mills theories and for gravity in the context of field theories with higher derivatives. Then we will discuss the laws of black hole mechanics and certain key notions of black hole physics in the framework of general effective Lagrangians. Particular emphasis will be put on Wald's derivation of the first law and the related definition of black hole entropy as a Noether charge. Then we recall the special features of extremal black holes and their relation to supersymmetry. We briefly describe how special geometry encodes the couplings of vector multiplets to $N=2$ supergravity in the presence of a certain class of higher-order derivative terms proportional to the square of the Weyl tensor. This part of the discussion is based on the effective $N=2$ Wilsonian action, which is defined in terms of a holomorphic quantity. Then we review our work on the entropy of extremal $N=2$ black holes in the presence of these higher-derivative curvature terms, which uses a definition of the macroscopic entropy that deviates from the area law. We conclude with a few examples of black hole solutions arising in string theory compactifications. By recalling results of microscopic entropy calculations in the context of string and M-theory compactifications, we show that they perfectly agree with macroscopic entropy calculations based on Wald's definition of macroscopic entropy. We also briefly discuss how to incorporate non-holomorphic corrections to the effective Lagrangian into macroscopic entropy formulae. \section{Noether potential and charge} For any local symmetry there exists a globally-defined Noether potential, denoted here by an antisymmetric tensor ${\cal Q}^{\mu\nu}(\phi, \xi)$, which is a local function of the fields and of the gauge transformation parameters, here generically denoted by $\phi$ and $\xi$ respectively \cite{Wald}. To derive this potential, one starts from a gauge-invariant action and one derives the Noether current as if one is dealing with a (rigid) residual symmetry associated with a certain background field configuration. To illustrate this we will first present the case of a Yang-Mills theory as a pedagogical example, with a gauge-invariant Lagrangian ${\cal L}(F_{\mu\nu},\nabla_{\!\rho} F_{\mu\nu}, \psi, \nabla_{\!\mu}\psi)$ depending on the field strenghts $F_{\mu\nu}$, matter fields $\psi$ and first derivatives thereof. We begin by multiplying the gauge transformations with a test function $\epsilon(x)$. This test function as well as its derivatives will satisfy certain boundary conditions, such that the variation of the action is proportional to the field equations, in accord with a modified (in view of the higher-derivative interactions) version of Hamilton's variational principle. Hence we have \begin{equation} \begin{array}{rcl} \delta A_\mu &=& \epsilon(x) \,\nabla_{\!\mu\,} \xi(x)\,,\\[1mm] \d\psi &=& \epsilon \, \xi\, \psi\,, \end{array} \qquad \begin{array}{rcl} \delta F_{\mu\nu} &=& 2\, \partial_{[\mu}\epsilon \,\nabla_{\!\nu]}\xi + \epsilon\,[\xi, F_{\mu\nu}]\, , \\[1mm] \qquad \d \nabla_{\!\mu} \psi &=& \partial_\mu\epsilon \,\xi\, \psi + \epsilon \,\xi\,\nabla_{\!\mu} \psi\,. \end{array} \end{equation} Here the gauge field, the field strength and the transformation parameters $\xi$ are written as Lie-algebra valued quantities in the representation relevant for $\psi$. The explicit variation of the action now leads directly to the current (using that $\epsilon$ and its first derivative vanish at the boundary) and one finds (we suppress the gauge-invariant inner product notation), \begin{equation} J^\mu = 2\, {\cal L}^{\mu\nu}\,\nabla_{\!\nu\,}\xi - 2\nabla_{\!\rho\,} {\cal L}^{(\rho,\mu)\nu}\, \nabla_{\!\nu\,} \xi +2 \,{\cal L}^{[\rho,\mu]\nu}\,\nabla_{\!\rho} \nabla_{\!\nu\,}\xi - {\cal L}^{\mu,\rho\sigma}\, [F_{\rho\sigma}, \xi] + {\cal L}^{\mu}\, \xi\psi\,, \label{currentg} \end{equation} where\footnote Note that the variations are defined such that we do not differentiate with respect to {\it independent} field components. Specifically, the variation of the Lagrangian takes the form $\d{\cal L} = {\cal L}^{\mu\nu}\d F_{\mu\nu} + \cdots$. } \begin{equation} {\cal L}^{\mu\nu}= {\partial{\cal L}\over \partial F_{\mu\nu}} \,,\qquad {\cal L}^{\rho, \mu\nu} = {\partial{\cal L}\over \partial \nabla_{\!\rho} F_{\mu\nu}} \,, \qquad {\cal L}^{\mu} = {\partial {\cal L}\over \partial \nabla_{\!\mu} \psi}\,. \end{equation} Observe that the Bianchi identity implies ${\cal L}^{[\rho,\mu\nu]}=0$. The equations of motion for the gauge fields take the form \begin{equation} -2 \,\nabla_{\!\mu\,} {\cal L}^{\mu\nu}+2 \nabla_{\!\mu} \nabla_{\!\rho\,} {\cal L}^{\rho, \mu\nu}+ {\cal L}^{\nu,\rho\sigma} [F_{\rho\sigma},\cdot] -{\cal L}^{\nu}\,\psi= 0 \,. \label{gauge-field-equation} \end{equation} We refrain from giving the matter field equation for $\psi$. By virtue of the combined field equations one can verify that this current is conserved. However, one can also use the field equation \eqn{gauge-field-equation} directly on the current so as to obtain \begin{equation} J^\mu = \partial_\nu \,{\cal Q}^{\mu\nu}\,, \end{equation} where ${\cal Q}^{\mu\nu}$ is the Noether potential, which in the case at hand takes the form \begin{equation} {\cal Q}^{\mu\nu} = 2 \,{\cal L}^{\mu\nu}\, \xi - 2\,\nabla_{\!\rho\,}{\cal L}^{\rho, \mu\nu} \, \xi + {\cal L}^{\rho, \mu\nu}\, \nabla_{\!\rho\,} \xi \,, \label{chargeg} \end{equation} and is thus a local function of the fields and of the transformation parameter. Obviously the fact that the Noether current is conserved is now trivial. Nevertheless the Noether potential is still tied directly to the invariant action, up to terms that are exact, i.e., that can be written as $\partial_\rho X^{\mu\nu\rho}$, with $X^{\mu\nu\rho}$ a totally antisymmetric tensor, and up to improvement terms that vanish in the symmetric background (see below). {From} the Noether potential one can determine the charge by integrating over a closed hypersurface of co-dimension two, in which case the ambiguity drops out. For a spacelike hypersurface we then determine the Noether charge in the usual sense, but written as a surface integral. This charge is associated with the residual gauge symmetry corresponding to parameters that satisfy $\nabla_{\!\mu\,}\xi =0$ in the background. Here we observe that we have used a somewhat different algorithm for computing (\ref{currentg}) than the more conventional one used, for instance, in \cite{Wald,JacKaMy,IWald}. Compared with the latter, this reflects itself in the presence of an additional improvement term in (\ref{currentg}) equal to $ \partial_{\nu} ( {\cal L}^{\rho,\mu\nu} \nabla_{\!\rho\,} \xi )$, and hence in the presence of the last term in (\ref{chargeg}). We thus see that both approaches yield the same Noether potential up to a term which vanishes in the symmetric background. Here we have assumed that the Lagrangian is gauge invariant, which implies that the Noether current is proportional to the symmetry variations of the fields (as one can for instance verify directly for \eqn{currentg}). Consider now the situation where we are dealing with a continuous variety of solutions of the field equations which are left invariant under a corresponding continuous variety of residual gauge transformations. The Noether current associated with these residual transformations is then vanishing for every one of these solutions, from which it follows that, under changes of the solution (and the corresponding symmetry parameters $\xi$), the Noether potential must vary into a term that is exact, i.e., \begin{equation} \delta Q^{\mu\nu} = \partial_{\rho\,}\omega^{\mu\nu\rho}\,, \end{equation} where $\omega$ is a totally antisymmetric tensor. This implies that the value of the integral of the Noether potential over the boundary of a hypersurface of co-dimension one must be constant and equal to zero for all these solutions, provided that the fields are regular on the hypersurface. While this example is therefore not so interesting, the situation changes when the Lagrangian is only invariant modulo a total derivative, because then the current does in general not vanish in symmetric background solutions. For Yang-Mills theory this happens, for example, when the Lagrangian contains Chern-Simons terms. For the case of gravity with general coordinate invariance, the Lagrangian is never invariant under general coordinate transformations, so here the current will not necessarily vanish in a symmetric background. So let us follow the same procedure for gravity and construct the Noether potential, starting from an invariant action depending on the Riemann curvature, and on a matter field $\psi_{\mu\nu}$ (with no particular symmetry) and its first derivative. After multiplying the diffeomorphisms with a test function $\epsilon(x)$ with appropriate boundary conditions, the transformation rules read \begin{eqnarray} \delta g_{\mu\nu} &=& - \epsilon(x) \Big(\nabla_{\!\mu\,} \xi_\nu(x) + \nabla_{\!\nu\,} \xi_\mu(x)\Big) \,, \nonumber\\ \d \psi_{\mu\nu} &=& -\epsilon(x) \Big(\nabla_{\!\mu\,} \xi^\sigma \, \psi_{\sigma\nu} + \nabla_{\!\nu\,} \xi^\sigma \,\psi_{\mu\sigma} + \xi^\sigma\,\nabla_{\!\sigma} \psi_{\mu\nu}\Big) \,. \end{eqnarray} We note the following useful equations for variations of connections and curvatures, \begin{eqnarray} \delta\Gamma_{\nu\rho}{}^\sigma &=& {\textstyle{1\over 2}} g^{\sigma\lambda} \Big[ \nabla_{\!\nu\,} \delta g_{\rho\lambda} + \nabla_{\!\rho\,} \delta g_{\nu\lambda} - \nabla_{\!\lambda\,} \delta g_{\nu\rho} \Big]\,, \nonumber\\ \delta R_{\mu\nu\rho\sigma} &=& R_{\mu\nu[\rho}{}^\lambda \,\delta g_{\sigma]\lambda} + 2\,\nabla_{[\mu} \nabla_{[\rho} \,\delta g_{\sigma]\nu]} \,. \end{eqnarray} The variation of the action, assuming the boundary conditions for $\epsilon$, now yields the current \begin{eqnarray} J^\mu &=& \xi^\mu \,{\cal L} - 2 \,{\cal L}^{\mu\nu\rho\sigma}\Big[ R_{\lambda \nu\rho\sigma} \,\xi^\lambda + \nabla_{\!\nu} \nabla_{\!\rho\,} \xi_\sigma \Big] +4 \,\nabla_\rho {\cal L}^{\mu\nu\rho\sigma}\, \nabla_{(\nu\,}\xi _{\sigma)} \nonumber \\ && - {\cal L}_\psi^{\mu,\rho\sigma} \, \Big[ \nabla_{\!\rho\,}\xi^\lambda \,\psi_{\lambda\sigma} + \nabla_{\!\sigma\,} \xi^\lambda\, \psi_{\rho\lambda} + \xi^\lambda \,\nabla_{\!\lambda} \psi_{\rho\sigma} \Big] \nonumber\\ && + \ft{1}{2} (\nabla_{\!\lambda\,} \xi_\rho + \nabla_{\!\rho\,} \xi_\lambda) \Big[ {\cal L}_\psi^{\mu,\rho\sigma} \psi^{\lambda}{}_{\sigma} + {\cal L}_\psi^{\mu,\sigma \rho} \psi_{\sigma}{}^{\lambda} + {\cal L}_\psi^{\rho,\mu\sigma} \psi^{\lambda}{}_{\sigma} \nonumber\\ && \hspace{35mm} + {\cal L}_\psi^{\rho,\sigma \mu} \psi_{\sigma}{}^{\lambda} - {\cal L}_\psi^{\rho,\lambda \sigma} \psi^{\mu}{}_{\sigma} - {\cal L}_\psi^{\rho,\sigma \lambda} \psi_{\sigma}{}^{\mu} \Big] \,, \label{gravNcurrent} \end{eqnarray} where \begin{equation} {\cal L}^{\mu\nu}= {\partial{\cal L}\over \partial g_{\mu\nu}} \,,\qquad {\cal L}^{\mu\nu\rho\sigma} = {\partial{\cal L}\over \partial R_{\mu\nu\rho\sigma}} \,, \qquad {\cal L}_\psi^{\rho,\mu\nu} = {\partial {\cal L}\over \partial \nabla_{\!\rho} \psi_{\mu\nu}}\,. \end{equation} Observe that ${\cal L}^{\mu\nu\rho\sigma}$ is antisymmetric in $[\mu\nu]$ and in $[\rho\sigma]$; furthermore it is symmetric under pair exchange, ${\cal L}^{\mu\nu\rho\sigma} ={\cal L}^{\rho\sigma\mu\nu}$, and satisfies the cyclicity property ${\cal L}^{[\mu\nu\rho]\sigma}=0$. The current \eqn{gravNcurrent} is conserved by virtue of the equations of motion for the metric, \begin{eqnarray} &&{\textstyle{1\over 2}} g^{\mu\nu} \,{\cal L} + {\cal L}^{\mu\nu} + {\cal L}^{\rho\sigma\lambda(\mu}\,R_{\rho\sigma\lambda}{}^{\nu)} - 2 \,\nabla_{\!(\rho} \nabla_{\!\sigma)} {\cal L}^{\rho\mu\nu\sigma} \nonumber\\ &&+\ft14 \nabla_{\!\lambda}\Big[ {\cal L}_\psi^{\lambda,\mu\rho} \, \psi^\nu{}_\rho + {\cal L}_\psi^{\lambda,\rho\mu} \,\psi_\rho{}^\nu + {\cal L}_\psi^{\mu,\lambda\rho} \,\psi^\nu{}_\rho \nonumber \\ && \hspace{15mm} +{\cal L}_\psi^{\mu,\rho\lambda} \,\psi_\rho{}^\nu -{\cal L}_\psi^{\mu,\nu\rho} \,\psi^\lambda{}_\rho -{\cal L}_\psi^{\mu,\rho\nu} \,\psi_\rho{}^\lambda + (\mu\leftrightarrow\nu)\Big] =0 \,, \label{metric-eq-motion} \end{eqnarray} and of the equation of motion for $\psi_{\mu\nu}$, \begin{equation} \nabla_{\!\rho\,}{\cal L}_\psi^{\rho,\mu\nu} - {\cal L}_\psi^{\mu\nu} = 0\,, \label{psi-eq-motion} \end{equation} where \begin{equation} {\cal L}^{\mu\nu}_\psi = {\partial{\cal L}\over \partial \psi_{\mu\nu}}\,, \end{equation} as well as of the general covariance of the Lagrangian. The latter implies \begin{eqnarray} &&2\,{\cal L}^{\mu\nu} \,\nabla_{\!\mu\,} \xi_\nu +4\, {\cal L}^{\mu\nu\rho\sigma}\,R_{\mu\nu\rho}{}^\lambda \, \nabla_{\!\sigma\,}\xi_{\lambda} + {\cal L}_\psi^{\mu\nu}(\nabla_{\!\mu\,} \xi^\rho \,\psi_{\rho\nu} +\nabla_{\!\nu\,}\xi^\rho \,\psi_{\mu\rho} ) \nonumber\\ &&+ {\cal L}_\psi^{\rho,\mu\nu} (\nabla_{\!\rho\,}\xi^\sigma \,\nabla_{\!\sigma\,}\psi_{\mu\nu} + \nabla_{\!\mu\,}\xi^\sigma \,\nabla_{\!\rho\,}\psi_{\sigma\nu}+ \nabla_{\!\nu\,}\xi^\sigma \,\nabla_{\!\rho\,} \psi_{\mu\sigma} ) =0 \,,\; \end{eqnarray} which in turn gives rise to the identity \begin{eqnarray} &&2{\cal L}^{\mu\nu} \,\xi_\nu -4\, {\cal L}^{\mu\nu\rho\sigma}\,R_{\rho\sigma\nu}{}^{\lambda} \, \xi_{\lambda} +{\cal L}_\psi^{\mu\nu}\xi^\sigma \,\psi_{\sigma\nu} + {\cal L}_\psi^{\nu\mu}\xi^\sigma \,\psi_{\nu\sigma} \nonumber\\ && + {\cal L}_\psi^{\mu,\nu\rho} \xi^\sigma \,\nabla_{\!\sigma\,}\psi_{\nu\rho} + {\cal L}_\psi^{\rho,\mu\nu} \xi^\sigma \,\nabla_{\!\rho\,}\psi_{\sigma\nu} + {\cal L}_\psi^{\rho,\nu\mu} \xi^\sigma \,\nabla_{\!\rho\,}\psi_{\nu\sigma} =0 \,. \label{covariance} \end{eqnarray} Imposing the equations of motion \eqn{metric-eq-motion} and \eqn{psi-eq-motion} and the relation \eqn{covariance}, the current takes the form $J^\mu= \nabla_\nu {\cal Q}^{\mu\nu}$ with the Noether potential equal to \begin{eqnarray} {\cal Q}^{\mu\nu} &=& - 2\, {\cal L}^{\mu\nu\rho\sigma} \,\nabla_{\!\rho\,}\xi_\sigma + 4 \,\nabla_{\!\rho\,}{\cal L}^{\mu\nu\rho\sigma} \, \xi_\sigma \nonumber\\ && +\ft12 \Big[-{\cal L}_\psi^{\mu,\nu\rho} \,\psi^\sigma{}_{\rho} -{\cal L}_\psi^{\mu,\rho\nu} \,\psi_{\rho}{}^\sigma +{\cal L}_\psi^{\mu,\sigma\rho} \,\psi^\nu{}_{\rho} \nonumber \\ &&\hspace{9mm} +{\cal L}_\psi^{\sigma,\mu\rho} \,\psi^\nu{}_{\rho} +{\cal L}_\psi^{\mu,\rho\sigma} \,\psi_{\rho}{}^\nu +{\cal L}_\psi^{\sigma,\rho\mu} \,\psi_{\rho}{}^\nu - (\mu\leftrightarrow\nu) \Big] \,\xi_\sigma \,. \label{noetgrav} \end{eqnarray} When spacetime exhibits an isometry with a corresponding Killing vector $\xi^{\mu}$, so that \begin{equation} \nabla_{\!\mu\,} \xi_{\nu} + \nabla_{\!\nu\,} \xi_{\mu} = 0\,, \label{Killing} \end{equation} the corresponding Noether potential is only proportional to $\xi^\mu$ and to its curl $\nabla_{[\mu\,}\xi_{\nu]}$ in view of the identity \begin{equation} \nabla_{\!\mu} \nabla_{\!\nu\,} \xi_{\rho} = R_{\nu \rho \mu}{}^\sigma \xi_{\sigma}\,. \label{KillingA} \end{equation} The Noether charge associated with the isometry is given as the integral of the Noether potential over the boundary of a spatial hypersurface. The variation of the charge under infinitesimal variations of the background solution will be discussed in the following section. \section{Black hole entropy} As was shown by Wald one can employ the Noether charge in order to find generalized definitions of the black hole entropy that are consistent with the first law of black hole mechanics. Although most of the following applies to black holes in any number of spacetime dimensions we will restrict ourselves to four dimensions. As we discussed already in the introduction, the description in terms of a surface charge can in principle explain how variations of the fields subject to the equations of motion at the horizon can be expressed in terms of the variations of quantities that are defined at spatial infinity. The crucial observation \cite{Wald} is that there exists a Hamiltonian, whose change under a variation of the fields can be expressed in terms of the corresponding change of the Noether charge and takes the following form \begin{equation} \d H = \d\Big (\int_C d\Omega_\mu \;J^\mu \Big ) - \int_C d\Omega_\mu \: \nabla_{\!\nu\,} ( \xi^\mu\,\theta^\nu - \xi^\nu \,\theta^\mu) \,, \label{variation1} \end{equation} where we integrate over a Cauchy surface $C$ with volume element $d\Omega_\mu$, $J^\mu$ is the Noether current associated with a particular Killing vector $\xi^\mu$ (to be discussed below) and $\theta^\mu$ is defined by the surface integral that one obtains when considering the change of the action under the field variation, \begin{equation} \d S = \int d^4x\; \partial_\mu \Big(\sqrt{-g} \,\theta^\mu \Big)\,, \end{equation} after subsequently imposing the field equations. We assume that the Killing vector field is timelike so that $H$ can be associated with a Hamiltonian that governs the evolution along the integral timelike lines of $\xi^\mu$. For black holes, $\xi^\mu$ is a horizon-generating Killing field, meaning that it is the normal vector of a null hypersurface, called the Killing horizon. The Cauchy surface in \eqn{variation1} is chosen to extend from spatial infinity down to the Killing horizon where the Killing field turns lightlike. In Einstein gravity it can be shown that under certain assumptions, such as that the dominant-energy condition holds and that the matter field equations of motion have a well defined Cauchy problem, all event horizons are Killing horizons \cite{HawEll}, but in more general theories this is not obvious. For static black holes, the event horizon is always a Killing horizon \cite{Carter}, irrespective of the precise form of the action. In that case the relevant Killing field is just the static Killing vector field. In the following we shall always assume that we are dealing with a Killing horizon. When imposing the equations of motion, \eqn{variation1} takes the form of surface integrals over the boundary $\partial C$ with surface element $d\Sigma_{\mu\nu}$, \begin{equation} \d H= \int_{\partial C} d\Sigma_{\mu\nu} \, \Big(\delta {\cal Q}^{\mu\nu} - \xi^\mu\,\theta^\nu + \xi^\nu \,\theta^\mu\Big) \,. \label{deltah} \end{equation} Here it is important that the last two terms are proportional to the Killing vector and not to its curl. Furthermore one assumes that these terms can be rewritten as variations of some (not necessarily globally defined) quantity. Because one can prove that $\d H=0$ (but not that $H$ itself vanishes) whenever $\xi^\mu$ is a Killing vector characterizing an invariance of the full background, it follows that the sum of the surface integrals in \eqn{deltah} has to vanish! If one identifies, up to a certain proportionality factor, the resulting variations of the surface integrals at infinity with the mass and angular momentum variations that one has in the first law, then the surface integral at the horizon defines the variation of the entropy. Hence the mass, the angular momentum and the entropy are all surface charges derived from the same current, and the entropy takes the form of an integral over the Noether potential \cite{Wald}, \begin{equation} {\cal S} = -\pi \, \int_{\Sigma_{\rm hor}} {\cal Q}^{\mu \nu} \; \epsilon_{\mu \nu}\; \Big\vert_{\xi^\mu=0,\; \nabla_{[\mu\,}\xi_{\nu]}= \epsilon_{\mu\nu}}\;. \label{entronoet} \end{equation} Here $\Sigma_{\rm hor}$ denotes a spacelike cross section of the Killing horizon (which usually has the topology of $S^2$) and we have used $d\Sigma_{\mu\nu} = \epsilon_{\mu\nu}\,\sqrt{h} \,d^2x$. Here $\epsilon_{\mu \nu}$ denotes the binormal, whose definition we will review below and which is normalized according to $\epsilon_{\mu\nu}\epsilon^{\mu\nu}=-2$; $\sqrt{h} \,d^2x$ gives the surface element induced on $\Sigma_{\rm hor}$. We already mentioned that the Noether potential ${\cal Q}^{\mu\nu}$ can be decomposed according to ${\cal Q}^{\mu \nu} = Y^{\mu \nu \rho \sigma} \nabla_{[\rho\,}\xi_{\sigma]}+N^{\mu \nu \rho} \, \xi_{\rho} $. We shall discuss in due course how one is led to impose the conditions $\xi^\mu=0$ and $\nabla_{[\mu\,}\xi_{\nu]}= \epsilon_{\mu\nu}$ on the Killing vector field at $\Sigma_{\rm hor}$ associated with the Noether potential in \eqn{entronoet}. Let us already point out that the contributions we are suppressing in (\ref{entronoet}) by imposing these conditions have actually been shown to vanish for non-extremal black holes \cite{JacKaMy}; this is more subtle for extremal black holes, where the surface gravity is vanishing. Nevertheless we simply adopt \eqn{entronoet} as the definition for the entropy in both the extremal and non-extremal case. Note that in both cases \eqn{entronoet} leads in principle to a well-defined result. Finally, the normalization in \eqn{entronoet} has been chosen such that we reproduce the Bekenstein-Hawking area law for static black holes in general relativity. To see this, we note that the Lagrangian associated to Einstein gravity, with the conventions of \cite{CarDeWMoh}, reads $8 \pi {\cal L} = - \ft{1}{2} R$, so that $8\pi\,{\cal L}^{\mu\nu\rho\sigma}= -\ft12 g^{\mu[\rho}\,g^{\sigma]\nu}$. Consequently we have, subject to the conditions in \eqn{entronoet}, that ${\cal Q}^{\mu \nu} = - 2 \, {\cal L}^{\mu\nu\rho\sigma} \, \epsilon_{\rho \sigma}$ which equals $(8\pi)^{-1}\,\epsilon^{\mu\nu}$ for Einstein gravity. Substitution of this result into \eqn{entronoet} immediately yields one-fourth of the area, expressed in Planck units. In the presence of higher-derivative curvature terms, the entropy of a static black hole solution will in general not any longer be simply given by the Bekenstein-Hawking area law. As shown in the previous section (c.f. \eqn{noetgrav}), when the Lagrangian depends on the Riemann curvature tensor but not on derivatives thereof, as well as on matter fields with at most second derivatives, then ${\cal Q}^{\mu \nu} = - 2 \, {\cal L}^{\mu\nu\rho\sigma} \, \epsilon_{\rho \sigma}$ when we impose the conditions on the Killing vector at the horizon, specified in \eqn{entronoet}. The entropy of the static black hole is then given by \cite{Visser,JacKaMy,IWald} \begin{equation} {\cal S} = 2\pi \int_{\Sigma_{\rm hor}} \; {\cal L}^{\mu \nu \rho \sigma} \; \epsilon_{\mu \nu} \; \epsilon_{\rho \sigma} \;. \label{graventro} \end{equation} This is the result that we will need in the next section. We will now review some of the concepts involved in the definition (\ref{entronoet}) of the black hole entropy \cite{Wald,JacKaMy,IWald,Waldbook} in somewhat more detail. Wald's construction of the entropy as a surface charge applies to stationary and to some extent also to non-stationary black holes in arbitrary spacetime dimensions, but in the following we will for concreteness restrict the discussion to four-dimensional static black hole solutions. A spacetime containing a black hole consists of an asymptotically flat region, which is separated by a (future) event horizon from an interior region, such that after crossing the horizon one is trapped in the interior region. Usually, the term `horizon' designates either the `horizon at a given instant of time', i.e. a spacelike two-surface $\Sigma_{\rm hor}$, or the corresponding worldvolume $\Delta$ swept out in time by $\Sigma_{\rm hor}$, which is a hypersurface in spacetime. In order to avoid any confusion, we will consistently distinguish between these two cases and use the symbols $\Sigma_{\rm hor}$ and $\Delta$ throughout. A hypersurface can be defined by an equation $f(x^{\mu})=0$. Since the gradient $\nabla_{\mu} f$ is automatically normal to the surface (i.e. $t^{\mu}\, \nabla_{\mu} f = 0$ for all tangent vectors $t^{\mu}$) one can equally well specify this surface in terms of the normal vector field $n_{\mu} = \nabla_{\mu} f$. According to the Frobenius theorem, which formulates necessary and sufficient conditions for vector fields to define smoothly embedded submanifolds, a vector field $n^{\mu}$ is hypersurface orthogonal in the above sense if and only if \begin{equation} n_{[\mu} \nabla_{\nu} n_{\rho]} = 0 \;. \label{Frobenius} \end{equation} If the normal vector field of a hypersurface is a null vector field, $n_{\mu} n^{\mu} =0$, then the hypersurface is called a null hypersurface. Note that the normal vector field of a null hypersurface is also tangential to it. The (future) event horizon $\Delta$ is such a null hypersurface. Since we assumed that it is a Killing horizon we know that its normal vector is in fact a Killing vector field. By taking a spacelike cross section of $\Delta$ one obtains a spacelike surface $\Sigma_{\rm hor}$. In the two-dimensional space normal to $\Sigma_{\rm hor}$ there exists one linearly independent antisymmetric tensor of second rank, $\epsilon_{\mu\nu}$, which is called the normal bivector or simply the binormal. We will normalize it according to $\epsilon_{\mu \nu} \, \epsilon^{\mu \nu} = -2$. The binormal can be explicitly written as a bivector, as follows. First we note that the space normal to $\Sigma_{\rm hor}$ has signature $(-+)$, and therefore it can be spanned by two null vectors. Since $\Sigma_{\rm hor}$ is contained in $\Delta$, one of the null vectors can be taken to be the normal $n^{\mu}$ of $\Delta$, that is, one of the null vectors is proportional to the Killing vector $\xi^{\mu}$. The other null vector, which we denote by $N^{\mu}$, is chosen such that $N^{\mu} n_{\mu} = -1$. Then, the bivector \begin{equation} \epsilon_{\mu \nu} = N_{\mu} n_{\nu} - N_{\nu} n_{\mu}\, \label{bivector} \end{equation} is non-vanishing in the normal directions and has the required normalization. The spacetime metric describing a static and spherically symmetric black hole solution can be written as \begin{equation} ds^2 = - e^{2g(r)}dt^2 + e^{2 f(r)} ( dr^2 + r^2 d \Omega^2 ) \label{StaticMetric} \end{equation} in isotropic coordinates $(t,r,\phi,\theta)$. In such an adapted coordinate system, the time coordinate $t$ and the radius $r$ denote the directions normal to the two-sphere $\Sigma_{\rm hor}$, whose coordinates are $\phi$ and $\theta$. Thus, the associated binormal $\epsilon_{\mu \nu}$ has the non-vanishing components (up to an overall sign) $\epsilon_{tr} = - \epsilon_{rt}= \exp[ g(r)+f(r)]$. In order to formulate the laws of black hole mechanics, we need to define the notion of surface gravity $\kappa_{\rm S}$ associated with a Killing horizon. Since the Killing vector field is null on the horizon (but in general $\xi^\mu\not=0$), we can take $f=\xi^{\mu} \xi_{\mu} = 0$ as the defining equation of the horizon $\Delta$. Since $\nabla_{\mu} f$ is normal to $\Delta$, it must be proportional to $\xi_{\mu}$ itself. The coefficient of proportionality defines the surface gravity $\kappa_{\rm S}$ of the black hole: \begin{equation} \nabla_{\!\mu\,}(\xi^{\nu} \xi_{\nu}) = -2 \kappa_{\rm S}\, \xi_{\mu}\,. \label{kappa} \end{equation} Observe that this presupposes a certain intrinsic normalization for the Killing vector field, which is usually specified at spatial infinity. Subsequently one shows that \begin{equation} \kappa^{2}_{\rm S} = - \ft{1}{2} (\nabla^{\mu} \xi^{\nu})( \nabla_{\mu} \xi_{\nu}) = \ft{1}{2} R_{\mu \nu} \, \xi^{\mu} \xi^{\nu} \;\;\;\; {\rm and} \;\;\;\; \xi^{\mu}\, \partial_{\mu} \kappa_{\rm S} = 0 \,. \label{kappasquared} \end{equation} The first equation follows by multiplying \eqn{kappa} with $\nabla^\mu\xi^\rho$ and by making use of the Killing equation (\ref{Killing}) together with the Frobenius theorem (\ref{Frobenius}) with $n^\mu=\xi^\mu$. On the other hand, it follows from \eqn{KillingA} that $\xi^\rho\,\nabla_{\!\rho\,} \nabla_{\!\mu\,}\xi_\nu=0$, so that $\kappa_{\rm S}$ is constant along the integral curves of $\xi^{\mu}$ on $\Delta$. This is the last equation. The second equality is then obtained by applying a covariant derivative $\nabla^{\mu}$ on (\ref{kappa}) and by using (\ref{KillingA}). For a static and spherically symmetric black hole, the surface gravity $\kappa_{\rm S}$ is thus constant over all of $\Delta$. The constancy of $\kappa_{\rm S}$ over $\Delta$ is known as the zero-th law of black hole mechanics. We should point out here that the above considerations do not quite apply to extremal black holes, because they have zero surface gravity. We return to this point shortly. The comparison with the laws of thermodynamics suggests to identify the surface gravity with the temperature of the black hole, up to a multiplicative constant. Then, the zero-th law is reinterpreted by stating that the Killing horizon is in thermodynamical equilibrium and that it radiates like a black body. This interpretation is confirmed by the phenomenon of Hawking radiation which is found when quantizing matter fields in a classical black hole background \cite{Bek}. In this way the proportionality constant is fixed according to $T = {\kappa_{\rm S}}/{2 \pi}$, where $T$ denotes the Hawking temperature. It is instructive to calculate the surface gravity for a Reissner-Nordstr{\o}m black hole in general relativity, which provides an example of a static and charged black hole. In a coordinate system where the curvature singularity is located at $r=0$, the associated spherically symmetric spacetime line element is given by \begin{equation} ds^2 = - e^{2h(r)} dt^2 + e^{-2h(r)} dr^2 + r^2 d \Omega^2 \;\;,\;\; e^{2h(r)} = 1 - \frac{2M}{r} + \frac{Q^2}{r^2} \;, \label{rn} \end{equation} where $M$ and $Q$ denote the mass and the charge of the black hole, respectively. The cases $Q=0$, $M > |Q|$ and $M=|Q|$ yield the Schwarzschild, the non-extremal and the extremal Reissner-Nordstr{\o}m black hole, respectively. The case $M<|Q|$ is excluded by the generalized positivity theorem for the ADM mass when assuming the so-called dominant-energy condition, i.e., a non-negative energy density and a non-spacelike energy flow for every observer \cite{GibHul}. The surface gravity is computed to be \cite{Waldbook} \begin{equation} \kappa_{\rm S} = \frac{ \sqrt{M^{2}-Q^{2}}}{2M(M + \sqrt{M^{2}-Q^{2}}) -Q^{2}} \;\;, \end{equation} which shows that charged and neutral black holes behave very differently when loosing energy by Hawking radiation. The Schwarzschild black hole ($Q=0$) will heat up in the process, because $\kappa_{\rm S} = (4M)^{-1}$, which suggests that it will completely evaporate into radiation. On the other hand a charged black hole will cool down when approaching the extremal limit, namely $\kappa_{\rm S} \rightarrow 0 $ for $M \rightarrow Q$. The final state, an extremal black hole with $\kappa_{\rm S} =0$, has vanishing Hawking temperature and could therefore be a stable object. In the following, we will always call black hole solutions extremal if they have $\kappa_{\rm S}=0$. Note that this does not necessarily imply that they are extremal in the sense of being on the edge of developing a naked singularity. We now turn to the first law of black hole mechanics which relates changes of the black hole mass to changes in the entropy as well as to changes of other quantities which characterize the black hole, such as its charges. In the following we will omit these latter changes for the sake of clarity. For a static black hole, the first law follows from (\ref{deltah}) with $\d H=0$ by taking the Cauchy surface $C$ to have a boundary $\Sigma_{\rm hor} \cup \Sigma_{\infty}$, where $\Sigma_{\rm hor}$ denotes a spacelike cross section of the event horizon, and where $\Sigma_{\infty}$ denotes a two-sphere at spatial infinity. Inspection of (\ref{deltah}) and \eqn{entronoet} (where in (\ref{entronoet}) we have adopted a different normalization for the Killing vector, see below) then suggests to identify the change in the mass and in the entropy of the black hole with \begin{eqnarray} \d M&=& - \ft12 \int_{\Sigma_{\infty}} d\Sigma_{\mu\nu} \, \Big( \delta {\cal Q}^{\mu\nu} - \xi^\mu\,\theta^\nu + \xi^\nu \,\theta^\mu\Big) \;,\nonumber\\ \frac{\kappa_{\rm S}}{2\pi} \, \d {\cal S}&=& -\ft12 \int_{\Sigma_{\rm hor}} d\Sigma_{\mu\nu} \, \Big(\delta {\cal Q}^{\mu\nu} - \xi^\mu\,\theta^\nu + \xi^\nu \,\theta^\mu\Big) \,,\; \label{first} \end{eqnarray} in accordance with the first law of thermodynamics, $\delta E = T \delta {\cal S} + \cdots$, with $E = M$ and $T= \kappa_{\rm S}/{2 \pi}$. We note that the first law applies to non-extremal black holes, which have a non-vanishing surface gravity. We now proceed to express the entropy as a surface integral of a local geometrical quantity. As already mentioned, the Noether potential has the generic form ${\cal Q}^{\mu \nu} = Y^{\mu \nu \rho \sigma} \nabla_{[\rho\,}\xi_{\sigma]} + N^{\mu \nu \rho} \xi_{\rho}$, where the tensors $N^{\mu \nu \rho}$ and $Y^{\mu \nu \rho \sigma}$ are local quantities constructed out of the Riemann tensor and its derivatives. The antisymmetric tensor $\nabla_{\!\mu\,} \xi_{\nu}$ can be decomposed as follows, \begin{equation} \nabla_{\!\mu\,} \xi_{\nu} = \kappa_{\rm S} \, \epsilon_{\mu \nu} + t_{[\mu} \,\xi_{\nu]} \,, \label{decomp} \end{equation} where $t_{\mu}$ is tangential to $\Sigma$. This decomposition expresses the fact that according to the Frobenius theorem the non-vanishing components of $\nabla_{\mu} \xi_{\nu}$ are those where at least one of the indices is not tangential to $\Sigma$. The coefficient of $\epsilon_{\mu \nu}$ is determined by contracting (\ref{decomp}) with $\nabla^{\mu} \xi^{\nu}$ and by comparing with (\ref{kappasquared}), where one also uses the explicit realization (\ref{bivector}) of $\epsilon_{\mu \nu}$ as a bivector. By substituting the decomposition (\ref{decomp}) into ${\cal Q}^{\mu \nu}$ we obtain ${\cal Q}^{\mu \nu} = \kappa_{\rm S} \, Y^{\mu \nu \rho \sigma} \epsilon_{\rho\,\sigma} + [ N^{\mu \nu \lambda} + Y^{\mu \nu \rho \sigma} \, t_{[\rho}\,\d^\lambda{}_{\!\sigma]\,}]\, \xi_{\lambda}$. Observe that the last two terms of the integrand in \eqn{first} are proportional to the Killing vector and can thus be absorbed into $N^{\mu\nu\rho}$, so they don't need to be discussed separately. For non-extremal black holes there is a theorem \cite{RacWal} which states that the horizon hypersurface $\Delta$ contains, or can be analytically extended to contain, a spacelike cross section, called the bifurcation surface $\Sigma_{0}$, where the timelike Killing vector field has a zero, $\xi^{\mu} = 0$, so that $\nabla_{\mu} \xi_{\nu} = \kappa_{\rm S} \epsilon_{\mu \nu}$. In theories with matter one also has to assume that the matter fields can likewise be analytically continued and that all the fields are regular at the bifurcation surface. Thus, by evaluating the Noether potential on $\Sigma_{0}$ one can get rid of the terms in ${\cal Q}^{\mu \nu}$ proportional to the Killing vector field, so that we are left with $\kappa_{\rm S}/2\pi$ times the variation of the entropy defined in \eqn{entronoet} but with $\Sigma_{\rm hor}$ replaced by $\Sigma_0$. As already mentioned, it can be shown \cite{JacKaMy} that, when replacing $\Sigma_{0}$ with any other spacelike cross section $\Sigma_{\rm hor}$ of the horizon, the resulting expression for the entropy is given by a similar expression, namely by (\ref{entronoet}), which indeed expresses the entropy as a surface integral of a local geometrical quantity over an arbitrary cross section of the horizon. As stressed above, the above procedure for deriving the first law of black hole mechanics applies to non-extremal black holes. Nevertheless, the resulting expression for the entropy (\ref{entronoet}) remains well behaved in the extremal limit $\kappa_{\rm S} \rightarrow 0$, as it is independent of $\kappa_{\rm S}$ and of the Killing field $\xi^{\mu}$. Since extremal black holes do not possess a bifurcation surface, it is important that the entropy can be evaluated on any spacelike cross section $\Sigma_{\rm hor}$, as is the case with (\ref{entronoet}). Thus, we expect that the entropy of an extremal black hole, if computed from (\ref{entronoet}), will be non-vanishing in general. It should be pointed out though that the question whether or not an extremal black hole has a non-vanishing entropy, is a somewhat subtle issue that depends on the approach used to compute its entropy. For instance, in the context of semiclassical quantization of matter coupled to Einstein gravity, one finds that the entropy of an extremal Reissner-Nordstr{\o}m black hole is non-vanishing and given by the Bekenstein-Hawking area law, provided the extremal limit is taken after quantization. If, on the other hand, the quantization is performed after extremalization (that is, if the quantization is applied to the part of phase space that only contains extremal configurations), then the resulting entropy is vanishing. This has been shown both in the Euclidean path integral framework \cite{GhoMit} and in the Minkowskian canonical framework \cite{KieLou}. In the context of string theory, various microscopic state countings yield a non-vanishing entropy in all cases where extremal black holes have a non-vanishing horizon area. Thus, string theory favours to treat extremal black holes as limiting cases of non-extremal ones. A comparison of Wald's Noether charge approach with various other approaches can be, for instance, found in \cite{Visser,IyeWal2} and references therein. Within their domain of applicability, all of these other approaches yield results in agreement with the Noether charge approach. \section{Supersymmetric black holes} The electrically charged static extremal Reissner-Nordstr{\o}m solution of Einstein-Maxwell theory, which we briefly described in (\ref{rn}), enjoys several remarkable properties, as follows. It interpolates between two maximally symmetric spaces, namely Minkowski flat spacetime at spatial infinity and Bertotti-Robinson spacetime at the horizon. There exists a dyonic version of it, whose mass $M$ and entropy ${\cal S}$ (the latter computed from the area law) are determined in terms of its electric and magnetic charges $q$ and $p$ as $M = |q + ip |$ and ${\cal S}= \pi |q + i p|^2$. Since its temperature vanishes, this suggests that such a dyonic black hole is quantum mechanically stable. Moreover, its mass saturates the Bogomolnyi bound which follows from the generalized positivity theorem for the ADM mass \cite{GibHul}. There also exists a static multi-center version of it, which is described by the Majumdar-Papapetrou metric \cite{MajPap} and which resembles the static multi-monopole solutions of Yang-Mills-Higgs theories in the Prasad-Sommerfield limit. All these properties can be explained in terms of a symmetry principle, namely supersymmetry, by embedding Einstein-Maxwell theory into $N=2$ supergravity \cite{GibHul}. Then the extremal Reissner-Nordstr{\o}m black hole solution can be interpreted as a supersymmetric soliton which interpolates between two maximally supersymmetric vacua of $N=2$ supergravity. Globally the solution is invariant under 4 of the 8 supersymmetries. Its mass formula follows from the $N=2$ supersymmetry algebra and takes the form $M=|z|$, where $z$ denotes the central charge of the supersymmetry algebra. It is determined in terms of the electric and magnetic charges $q$ and $p$ associated with the gauge field, namely by $z=q+ip$. More recently, following the work of \cite{FerKalStr}, this has been extended to the study of static supersymmetric black hole solutions in four-dimensional theories describing the coupling of $n$ abelian vector multiplets to $N=2$ supergravity in the presence of a certain class of $R^2$-terms \cite{BCDWLMS,CarDeWMoh}. Such theories arise as low-energy effective field theories of string and M-theory compactifications on suitable compact manifolds. These effective Lagrangians contain, in general, various other terms describing the couplings of additional sectors, such as matter associated with hypermultiplets, to supergravity. These other sectors, however, play only a limited role in the following and we will therefore omit them. Let us first review how the couplings of vector multiplets to $N=2$ supergravity can be described in terms of special geometry. Since we will allow for the presence of certain $R^2$-terms, we will utilize the so-called superconformal framework, which provides a systematic and powerful approach for constructing these couplings \cite{deWvPr}. It makes use of the fact that a conformal theory of $N=2$ supergravity with suitable couplings to compensating multiplets is gauge equivalent to $N=2$ Poincar\'e supergravity. In the superconformal framework, there is a multiplet, the so-called Weyl multiplet, which comprises the gravitational degrees of freedom, namely the graviton, two gravitini as well as various other superconformal gauge fields and also some auxiliary fields. One of these auxiliary fields is an anti-selfdual Lorentz tensor field $T^{ab\, i j}$, where $i,j=1,2$ denote chiral $SU(2)$ indices (we recall that the associated $SU(2)$ algebra is part of the $N=2$ superconformal algebra). The field strengths corresponding to the various gauge fields in the Weyl multiplet reside in a so-called reduced chiral multiplet, denoted by $W^{ab\, i j }$, from which one then constructs the unreduced chiral multiplet $W^2 = (W^{ab\,i j } \varepsilon_{ij})^2$ \cite{BDRDW}. The lowest component field of $W^2$ is ${\hat A} = (T^{ab\, i j } \varepsilon_{ij})^2$. In addition, there are $n + 1$ abelian vector multiplets labelled by an index $I = 0, \dots, n$. Each vector multiplet contains a complex scalar field $X^I$, a vector gauge field $W_{\mu}^I$ with field strength $F_{\mu \nu}^I$, as well as two gaugini and a set of auxiliary scalar fields. The couplings of these $n+1$ vector multiplets to the Weyl multiplet are encoded in a holomorphic function $F(X^I, {\hat A})$, which is homogenous of degree two and consequently satisfies $X^I F_I + 2 {\hat A} F_{\hat A} = 2 F$, where $F_I = \partial F/\partial {X^I}$, $F_{\hat A} = \partial F/\partial {\hat A}$. The field equations of the vector multiplets are subject to equivalence transformations corresponding to electric-magnetic duality, which do not involve the fields of the Weyl multiplet. These equivalence transformations are symplectic ${\rm SP}(2n+2;{\rm \bf Z})$ transformations. Two complex $(2n+2)$-component vectors can now be defined which transform linearly under ${\rm SP}(2n+2;{\rm \bf Z})$ transformations, namely \begin{equation} V = \left( \begin{array}{c} X^I \\ F_I(X,{\hat A}) \\ \end{array} \right) \;\;\;\;\; {\rm and} \;\;\;\;\; \left( \begin{array}{c} F_{\mu \nu}^{\pm I} \\ G^\pm_{\mu \nu I} \\ \end{array} \right)\,, \end{equation} where $(F_{\mu \nu}^{\pm I},G^\pm_{\mu \nu I})$ denotes the (anti-)selfdual part of $(F_{\mu \nu}^{I},G_{\mu \nu I})$. The field strength $G^\pm_{\mu \nu I}$ is defined by the variation of the action with respect to $F^{\pm I}_{\mu \nu}$. The precise definition reads $G^{\pm}_{\mu \nu I} = - 4 i \,(-g)^{-1/2}\, \partial S / \partial F^{\pm I}_{\mu \nu}$. By integrating the gauge fields over two-dimensional surfaces enclosing their sources, it is possible to associate to $(F_{\mu \nu}^{\pm I},G^{\pm}_{\mu \nu I})$ a symplectic vector $(p^I,q_I)$ comprising the magnetic and electric charges. It is then possible to construct a complex quantity $Z$ out of the charges and out of $V$ which is invariant under symplectic transformations, as follows, \begin{equation} Z = e^{ {\cal K}/2} \, (p^I F_I (X,{\hat A}) - q_I X^I) \,, \label{z} \end{equation} where $e^{-\cal K} = i [\bar{X}^I F_I (X,{\hat A}) - \bar{F}_I ({\bar X}, {\bar {\hat A}}) X^I ]$. This quantity will play a role in the following. The associated (Wilsonian) Lagrangian describing the coupling of these vector multiplets to supergravity is quite complicated \cite{deWit1}. Here we only display those terms which will be relevant for the computation of the entropy of a static supersymmetric black hole, \begin{equation} {8 \pi \cal L} = - \ft{1}{2} e^{-{\cal K}}R + \ft12( {i} F_{\hat{A}} \,\hat{C} + \mbox{h.c.}) + \cdots \,, \label{Paction} \end{equation} where ${\hat C} = 64 \,C^{- \mu \nu \rho \sigma } C^-_{ \mu \nu \rho \sigma} + 16 \,\varepsilon_{ij}\, T^{\mu \nu i j} f_{\mu}{}^{\rho\,} T_{\rho \nu k l} \,\varepsilon^{kl} + \cdots \,$. Here $C^-_{ \mu \nu \rho \sigma}$ denotes the anti-selfdual part of the Weyl tensor $C_{ \mu \nu \rho \sigma}$, and $f_{\mu}{}^{\nu} = \ft{1}{2} R_{\mu}{}^{\nu} - \ft{1}{12} R \, \delta_{\mu}{}^{\nu}+ \cdots$. Eventually we will set $e^{-{\cal K}}$ to unity in order to obtain a properly normalized Einstein-Hilbert term. In this way we fix the local scale invariance which is present in a superconformal formulation of the theory. We note that the Lagrangian contains $C_{ \mu \nu \rho \sigma}^2$-terms, but no terms involving derivatives of the Riemann curvature tensor. By expanding the holomorphic function $F(X, {\hat A})$ in powers of ${\hat A}$, $F(X, {\hat A}) =\sum_{g=0}^{\infty} F^{(g)}(X) {\hat A}^g$, we see that the Lagrangian (\ref{Paction}) contains an infinite set of higher-derivative curvature terms of the type $C^2{}_{\!\!\!\!\mu \nu \rho \sigma} ( T^{ab\, i j} \varepsilon_{ij}){}^{2g-2}$ ($g\geq 1$) with field-dependent coupling functions $F^{(g)} (X)$. As alluded to above, we will view a static supersymmetric black hole solution of the Lagrangian (\ref{Paction}) as a solitonic interpolation between two $N=2$ supersymmetric groundstates, namely flat spacetime at spatial infinity and the horizon, whose geometry we now proceed to determine. The spacetime line element associated with a static spherically symmetric solution is of the form (\ref{StaticMetric}) in isotropic coordinates. The near-horizon solution can be specified by imposing full $N=2$ supersymmetry on the solution. This is achieved by requiring that the supersymmetry variations of all the fermions present in the theory vanish in the bosonic black hole background. We stress here that we do not analyze the equations of motion, as their validity is implied by full $N=2$ supersymmetry. A careful analysis \cite{CarDeWMoh} of the resulting restrictions on the bosonic background then shows that the $X^I$ and ${\hat A}$ are constant at the horizon. Since the black hole is charged and can carry both magnetic and electric charges $(p^I, q_I)$, the associated quantity $Z$ (\ref{z}) is therefore generically non-vanishing and constant at the horizon. Moreover, we also find that $T^{01\,ij} = - i\, T^{23\,ij} = 2 \varepsilon^{ij} \, e^{-{\cal K}/2}\, \bar Z^{-1}$, while all other components of $T^{ab\,ij}$ vanish. Therefore we have $\hat A =- 64\,e^{-{\cal K}}\,\bar Z^{-2}$. And finally, the near-horizon spacetime geometry is determined to be of the Bertotti-Robinson type, that is of the form (\ref{StaticMetric}) with $e^{2g(r)} = e^{-2f(r)}= e^{-{\cal K}} \, \vert Z\vert^{-2} \,r^2$. The requirement of $N=2$ supersymmetry at the horizon does not by itself fix the actual values of the constants $X^I$. To do so, we have to invoke the so-called fixed-point behaviour \cite{FerKalStr} for the scalar fields $X^I$ at the horizon (for a recent reference on the fixed-point behaviour, see \cite{Moore}). The fixed-point behaviour implies that regardless of what the values of the scalar fields are at spatial infinity, they always take the same values at the horizon. In the absence of higher-derivative terms it has been shown \cite{FerKalStr} that the supersymmetric black hole solutions do indeed always exhibit fixed-point behaviour as a result of their residual $N=1$ supersymmetry. In the presence of higher-derivative terms this has not yet been shown to be the case. In the following we will assume that such a fixed-point behaviour holds for black hole solutions of the Lagrangian (\ref{Paction}). {From} electric-magnetic duality one may then deduce that the values of the scalar fields $X^I$ at the horizon are determined from a set of equations, called the stabilization equations, which take the following form \cite{BCDWLMS,CarDeWMoh} \begin{equation} \bar{Z} \left( \begin{array}{c} X^I \\ F_I (X,{\hat A}) \\ \end{array} \right) - Z \left( \begin{array}{c} \bar{X}^I \\ \bar{F}_I ({\bar X}, {\bar {\hat A}}) \\ \end{array} \right) = i e^{-{\cal K}/2} \left( \begin{array}{c} p^I \\ q_I \\ \end{array} \right) \,. \label{stab} \end{equation} At this point it is convenient to introduce rescaled variables $Y^I = e^{{\cal K}/2} \bar{Z} X^I$ and $\Upsilon = e^{ {\cal K} } \bar{Z}^2 \hat{A} = -64$. Using the homogeneity property of $F$ mentioned earlier, it follows that the stabilization equations (\ref{stab}) now simply read $Y^I - \bar{Y}^I = ip^I$ and $F_I(Y,\Upsilon) - \bar{F}_I(\bar{Y}, \bar{\Upsilon}) = i q_I$. These equations then determine the value of the rescaled fields $Y^I$ in terms of the charges carried by the black hole. On the other hand, it follows from (\ref{z}) that $|Z|^2 = p^I F_I(Y,\Upsilon) - q_I Y^I$, which determines the value of $|Z|$ in terms of $(p^I, q_I)$. The entropy of the static black hole solution described above can now be computed from (\ref{graventro}), using (\ref{Paction}). The result takes the remarkably concise form \cite{CarDeWMoh} \begin{equation} {\cal S} = \pi \left[ |Z|^2 - 256\, \mbox{Im} \,F_{\hat A} \right] \,. \label{entropia} \end{equation} The first term denotes the Bekenstein-Hawking entropy contribution, whereas the second term is due to Wald's modification of the definition of the entropy in the presence of higher-derivative terms. Here we point out that this contribution does not actually originate from the $C^2{}_{\!\!\!\!\mu \nu \rho \sigma}$-terms in the action, because the Weyl tensor vanishes at the horizon, but from the term in ${\hat C}$ (see below (\ref{Paction})) proportional to the product of the Ricci tensor with the tensor field $T^{ab\,ij} T_{cd \,kl}$! Note that when switching on higher-derivative interactions the value of $|Z|$ changes and hence also the horizon area changes. There are thus two ways in which the presence of higher-derivative interactions modifies the black hole entropy, namely by a change of the near-horizon geometry and by an explicit deviation from the Bekenstein-Hawking area law. Also note that the entropy (\ref{entropia}) is entirely determined in terms of the charges carried by the black hole, ${\cal S} = {\cal S} (q,p)$. Let us now exhibit the generic dependence of the macroscopic entropy (\ref{entropia}) on the charges. Let $Q$ denote a generic electric or magnetic charge carried by the black hole. For large charges $Q$, the stabilization equations (\ref{stab}) and the homogeneity property of $F$ imply that the generic dependence of the entropy (\ref{entropia}) on the charges is given by \begin{eqnarray} {\cal S} = \pi \; \sum_{g=0}^{\infty} a_g \; Q^{2-2g} \;\;\; \label{sf} \end{eqnarray} with constant coefficients $a_g$. We recall that (\ref{sf}) encodes the contributions from a particular set of $R^2$-terms, namely terms proportional to $C_{\mu \nu \rho \sigma}^2$. In general, however, the effective Lagrangian will not only contain these particular terms, but also many other (even higher-derivative) curvature terms which, in principle, will lead to further contributions to the entropy. One might then worry that the inclusion of such contributions could wash out the contributions appearing in (\ref{sf}). In the context of string theory this appears to be unlikely, given that the coefficients multiplying the various powers of the charges in (\ref{sf}) encode topological information about the $N=2$ compactification. \section{Examples} Let us now briefly discuss various classes of black hole solutions arising in string theory compactifications. We will use the rescaled variables $Y^I = e^{{\cal K}/2}{\bar Z} X^I, \Upsilon = e^{{\cal K}}{\bar Z}^2 {\hat A} = - 64$ throughout. Let us first consider type-IIA string theory compactified on a Calabi-Yau threefold, in the limit where the volume of the Calabi-Yau threefold is taken to be large. For the associated homogenous function $F(Y,\Upsilon)$ we take (with $I = 0, \ldots, n$ and $A = 1, \ldots, n$) \begin{equation} F(Y,\Upsilon) = \frac{D_{ABC} Y^A Y^B Y^C}{Y^0} + d_{A}\, \frac{Y^A}{Y^0} \; \Upsilon \,\,\;,\,\,\; D_{ABC} = - \ft16 \, C_{ABC} \,\,\;,\,\,\; d_{A} = - \ft{1}{24} \, \ft{1}{64}\; c_{2A} \,, \label{prep2a} \end{equation} where the coefficients $C_{ABC}$ denote the intersection numbers of the four-cycles of the Calabi-Yau threefold, whereas the coefficients ${c_{2A}}$ denote its second Chern-class numbers \cite{Bershadskyetal}. The Lagrangian (\ref{Paction}) associated with this homogenous function thus contains a term proportional to $c_{2A} \, {\rm Im }\, z^A \,C_{\mu \nu \rho \sigma}^2$, where $z^A = {Y^A}/{Y^0}$. The associated stabilization equations can be solved \cite{CarDeWMoh} for black holes with $p^0 = 0$, yielding $Y^I=Y^I(q,p)$. The result for the macroscopic entropy, which is computed from (\ref{entropia}), reads \cite{CarDeWMoh} \begin{eqnarray} {\cal S}= 2 \pi \sqrt{\ft16 \, |{\hat q}_0| (C_{ABC} \,p^A p^B p^C + {c}_{2A} \, p^A) }\,, \label{entrot2} \end{eqnarray} where ${\hat q}_0 = q_0 + \ft{1}{12} D^{AB} q_A q_B \,,\, D_{AB} = D_{ABC} p^C \,,\, D_{AB} D^{BC} = \delta_A^C$. The expression (\ref{entrot2}) for the macroscopic entropy is in exact agreement with the microscopic entropy formula computed in \cite{MalStrWit,Vaf} via state counting. Next, let us consider black hole solutions arising in heterotic string compactifications on $K_3 \times T_2$. The associated tree-level function is given by \begin{eqnarray} F(Y,\Upsilon) = - \frac{Y^1 Y^a \eta_{ab} Y^b}{Y^0} + c \; \frac{Y^1}{Y^0} \; \Upsilon \;\;\;,\;\; a = 2, \dots, n \;\;\;, \label{hetprep} \end{eqnarray} where the real constants $\eta_{ab}$ and $c$ are related to the intersection numbers of two-cycles and to the second Chern-class number of $K_3$, respectively. For a function $F(Y, \Upsilon)$ of the form (\ref{hetprep}), the stabilization equations can be solved in full generality and the resulting expression for the entropy reads \cite{CDWM} \begin{eqnarray} {\cal S} = \pi \; \sqrt{ \langle M,M \rangle \langle N,N \rangle - ( M \cdot N )^2} \; \sqrt{1 - \frac{ 512 \, c \,}{\langle N,N \rangle}} \;, \label{hetentro} \end{eqnarray} where the $M_I = (q_0, - p^1, q_2, q_3, \dots, q_n)$ and the $N^I = (p^0, q_1, p^2, p^3, \dots, p^n)$ now denote the electric and magnetic charges carried by the heterotic black hole. The bilinears $\langle M,M \rangle$ and $\langle N,N \rangle$ are given by \begin{eqnarray} \langle M,M \rangle = 2 \left( M_0 M_1 + \ft{1}{4} M_a \eta^{ab} M_b \right) \;\;\;,\;\;\; \langle N,N \rangle = 2 \left( N^0 N^1 + N^a \eta_{ab} N^b \right) \;, \end{eqnarray} whereas $M \cdot N = M_I N^I$. These bilinears are invariant under tree-level target-space duality transformations of the charges \cite{clm}. We thus see that turning on a term proportional to ${\rm Im }\, z^1 \, C_{\mu \nu \rho \sigma}^2$ ($z^1 = {Y^1}/{Y^0}$) in the Lagrangian leads again to a modification of the entropy. There will be various corrections to (\ref{entrot2}) and to (\ref{hetentro}) from terms proportional to $C_{\mu \nu \rho \sigma}^2 \, {\hat A}^{g-1} (g \geq 2)$ in the effective Lagrangian (\ref{Paction}). These corrections are such that they no longer preserve the square-root feature of the entropy formulae (\ref{entrot2}) and (\ref{hetentro}) \cite{CDWM}. The examples of black hole solutions discussed so far occur in $N=2$ compactifications of string theory. Let us now consider black hole solutions occuring in heterotic $N=4$ compactifications. At tree-level there is a term proportional to ${\rm Re} \,S \,C_{\mu \nu \rho \sigma}^2$ in the effective Lagrangian of heterotic string theory compactified on a six-torus, where $S$ here denotes the heterotic dilaton field. The entropy of a black hole solution in the presence of such an $C_{\mu \nu \rho \sigma}^2$-term is given by an expression analogous to (\ref{hetentro}). As can be seen from (\ref{hetentro}), the analogous expression is not invariant under the exchange of the electric and the magnetic charges. The heterotic $N=4$ theory is, however, expected to be invariant under strong-weak coupling duality. Here we recall that it is essential to distinguish between the Wilsonian couplings of an effective theory and the physical (in general momentum-dependent) effective couplings. In theories with massless fields, the effective couplings are different from the Wilsonian couplings and do not share their analytic properties. In the context of heterotic $N=4$ compactifications, the effective coupling function multiplying $C_{\mu \nu \rho \sigma}^2$ is non-holomorphic and invariant under strong-weak coupling duality transformations \cite{HarMoo}, whereas the Wilsonian coupling function, which at tree-level is proportional to $S$, is holomorphic although not invariant under strong-weak coupling duality. Thus, in order to arrive at a manifestly strong-weak coupling duality invariant expression for the entropy of a heterotic $N=4$ black hole, its computation should be based on the effective rather than the Wilsonian coupling function of $C^2_{\mu \nu \rho \sigma}$. We will now restrict ourselves to black hole solutions in an $N=2$ subsector of the heterotic $N=4$ theory. Both the stabilization equations (\ref{stab}) and the entropy formula (\ref{entropia}) were derived in the Wilsonian context and as such they are defined in terms of a holomorphic function $F$. In the effective coupling approach, on the other hand, we expect that both (\ref{stab}) and (\ref{entropia}) will receive non-holomorphic corrections so as to be consistent with strong-weak coupling duality. The function $F$, in particular, will not any longer be holomorphic: \begin{eqnarray} F(Y, \bar Y, \Upsilon) = - \frac{Y^1 Y^a \eta_{ab} Y^b}{Y^0} + \; F^{(1)}( z^1, {\bar z}^1 ) \; \Upsilon \;\;\;, \label{nonholof} \end{eqnarray} where we require that (\ref{nonholof}) turns into the tree-level function (\ref{hetprep}) at weak coupling, that is $F^{(1)}( z^1, {\bar z}^1 ) \rightarrow c \,z^1$ as $ S + {\bar S} \rightarrow \infty$, where $S = -i z^1 = -i \, Y^1/Y^0$. The associated stabilization equations now read \begin{eqnarray} Y^I - {\bar Y}^I = i p^I \;\;\;,\;\;\; F_I(Y, \bar Y, \Upsilon) - {\bar F}_I (Y, \bar Y, \bar \Upsilon) = i q_I \;\;\;,\;\;\; \Upsilon = -64 \;\;\;. \label{nonhstab} \end{eqnarray} The form of $F^{(1)}( z^1, {\bar z}^1 )$ can then be determined by requiring (\ref{nonhstab}) to transform in a consistent way under strong-weak coupling duality transformations and also by requiring $F^{(1)}( z^1, {\bar z}^1 )$ to have the weak-coupling behaviour specified above \cite{CDWM}: \begin{eqnarray} F^{(1)} (S, {\bar S}) = - 6 \,i \, \frac{c }{\pi} \, \Big( \log \eta^2 (S) + \log (S + {\bar S}) \Big) \,, \end{eqnarray} where $\eta (S)$ denotes the Dedekind function. The associated non-holomorphic quantity $|Z|^2 = p^I F_I(Y, {\bar Y}, \Upsilon) - q_I Y^I $ is then invariant under strong-weak coupling duality. In analogy to (\ref{entropia}) we thus propose the following strong-weak coupling duality invariant expression for the entropy of a black hole in an $N=2$ subsector of the heterotic $N=4$ theory \cite{CDWM}: \begin{eqnarray} {\cal S} = \pi \Big[ \; |Z|^2 - 256\, {\rm Im} \,\Big( \, F^{(1)} (S, {\bar S}) + 3 \, i \, \frac{c }{\pi} \, \log (S + {\bar S}) \Big)\; \Big] \,. \label{entropiaynonholo} \end{eqnarray} The additional non-holomorphic piece in (\ref{entropiaynonholo}) is there to render the expression invariant under strong-weak coupling duality. We note that the value of the dilaton $S$ at the horizon is, in principle, determined in terms of the charges $M_I$ and $N^I$ carried by the black hole through the stabilization equations (\ref{nonhstab}), although in practice they are hard to solve. We refer to \cite{CDWM} for a more detailed discussion of these and related issues. \vskip0.5cm \leftline{\bf Acknowledgements} \noindent We thank Soo-Jong Rey for valuable discussions. T.M. thanks the Institute for Theoretical Physics of Utrecht University for hospitality during the final stages of this work.
\section{Introduction} \label{sec:intro} The geometrical phase transition known as percolation (see, for a review, Stauffer and Aharony \cite{stauffer94}) is appreciated by many to be an elegant and simply defined yet fully featured example of a second order phase transition. A number of variations of the original percolation problem were proposed as better models of some physical phenomena in the past. This includes the {\it backbone} percolation for studying electrical conduction through random media, {\it polychromatic} percolation for multi-component composites, and {\it four-coordinated} bond percolation for hydrogen-bonded water molecules. In particular, Blumberg et al \cite{blumberg80} and Gonzalez and Reynolds \cite{gonzales80} studied a random bond, site-correlated percolation problem they call four-coordinated percolation on the square lattice. They conclude that this problem belongs to the same universality class as the ordinary random percolation with the same set of (static) exponents. In this paper, we revisit a problem in this realm, though not exactly the same one. We define {\it fully coordinated percolation} as the site percolation problem where only the occupied sites all of whose neighboring sites are also occupied can transmit connectivity. Since the random element is the site, this problem is slightly different from the bond problem referred to above. Thus, after generating a random site configuration with the independent site occupation probability {\it p}, we only select those occupied sites with all 4 neighbors also occupied on the square lattice and study the clusters formed by nearest neighbor connections among those sites. It should be noted that this problem is distinct from the so-called bootstrap percolation (see, e.g., \cite{privman91}) where sites of less connectivity are iteratively removed. In our problem, no iterative procedures are involved; rather, sites of less than full connectivity are marked first and then all of them removed at one time. This problem arose in the context of studying the vibrational properties of fractal structures tethered at their boundaries \cite{mukherjee96,mukherjee98}. In that problem, scaling was observed in the normal mode spectrum whose origin may lie in the ratio of 2 length scales, one of which is the size of highly connected regions of a cluster. In this context, we have embarked on revisiting the characteristics of randomly generated, but highly connected geometrical structures. In the next section, we summarize the Monte Carlo and finite size scaling analyses of the static critical properties of fully coordinated percolation. In Section 3, we discuss the normal modes of the {\it transition probability matrix} for tracer diffusion on the structure using the methods of Arnoldi and Saad (see, e.g., \cite{nakanishi94}). Then in Section 4, we describe the classification of the cluster sites into external boundary, internal boundary, and interior ones and using these to show the major distinctions between the critical clusters of ordinary and fully coordinated percolation. We summarize the results in the final section. \narrowtext \section{STATIC CRITICAL BEHAVIOR} \label{sec:static} To determine the static critical behavior of fully coordinated percolation we first performed Monte Carlo simulations on a square lattice in two dimensions. Each site is occupied with probability $p$ independently and subsequent fully coordinated sites are marked and their connectivity searched. Lattice sizes of $L^{2}$ where $L = 256$, $512$, $1024$, and $2048$ were constructed. For each lattice size we further made a thousand realizations wherein a different random number seed was used on every run. The unnormalized susceptibilities, i.e., $\Xi(L) = \sum_{s}^{'} s^{2} \hat{n}_{s}$ where $\hat{n}_{s}$ is the number of clusters of size $s$, are calculated on each run and are then summed at the end of the thousand realizations. The average susceptibilities $\chi$ are calculated by dividing the sum by the number of realizations and the lattice size. The prime on the summation indicates the fact that the contribution of the largest cluster to $\chi$ near and above what we perceived to be the critical probability $p_{c}$ has been subtracted as usual \cite{stauffer94}. In Fig. 1 we plot the average susceptibilities against the probability $p$ for the corresponding lattice sizes. The data correspond to the values of $L = 256$, $512$, $1024$, and $2048$ from the lowest to highest. We can see that the effects due to the finite sizes of the lattices are exhibited clearly. In particular, there are well-defined peaks which scale with lattice sizes as \begin{equation} \chi(p_{max}, L)~\sim~L^{\gamma/\nu} \label{equ:peaks_scaling} \end{equation} where the known exact value of $\gamma/\nu$ for the ordinary percolation is $\frac{43}{24} \sim 1.7917$. To demonstrate the precision of our calculations, we plot $\chi(p_{max}, L)$ against the corresponding lattice sizes in the inset of Fig. 1. Notice that the data follow an excellent power law, leading to a least squares fit of $\chi(p_{max}, L) \sim L^{1.7911}$. The value of $\gamma$ found is identical with the ordinary percolation value to within about $0.03\%$. This result confirms previous work\cite{blumberg80,gonzales80} stating that fully coordinated percolation and ordinary percolation belong to the same {\em static} universality class. The critical behavior of susceptibility is known to scale as \cite{stauffer94} \begin{equation} \chi(p, \infty)~\sim~|p~-~p_{c}|^{-\gamma} \label{equ:chi_scaling} \end{equation} where for ordinary percolation $\gamma_o = \frac{43}{18} \sim 2.3889$. Notice however that in Fig. 1 the peaks are very near $p = 1.0$. This would provide data to the right of the peaks in only a small probability interval. In our simulations, we would therefore use $\chi$ only to the left of the peaks. Since the scaling relation in Eq.\ (\ref{equ:chi_scaling}) is expected only for infinite lattices, we use only the data taken from $L = 2048$ to test it. Since there are two unknowns in Eq.\ (\ref{equ:chi_scaling}), we first choose a particular $p_c$ and make a fit to see what value of $\gamma_{exp}$ is obtained. If we choose $p_c = 0.886$ we get $\gamma_{exp} = 2.4004$. The correlation coefficient, $|R|$, for this fit is $0.99999$. The discrepancy between $\gamma_{exp}$ and $\gamma_o$ is around $0.481\%$. Choosing $p_c = 0.8858$ we obtain $\gamma_{exp} = 2.3864$. The discrepancy this time is around $0.10\%$ and $|R| = 1$. So we have an exact fit for this value of $p_c$ and the $\gamma_{exp}$ found is very close to the $\gamma_o$. Choosing $p_c = 0.885$ we obtain $\gamma_{exp} = 2.3302$ with an $|R| = 0.99999$. The discrepancy for this value of $p_c$ is $2.46\%$. Fits done with $p_c$ between $0.885$ and $0.886$ gave $|R| = 1$; however, the $\gamma_{exp}$ found when $p_c = 0.8858$ gave the closest value to $\gamma_o$. This allows us to conclude that $\gamma$ for fully coordinated percolation is the same as that for ordinary percolation while also giving an estimate for the value of $p_c$ close to $0.8858$. (We will state the experimental uncertainty for $p_{c}$ after all our analyses are presented.) From the fit done to examine the scaling in Eq.\ (\ref{equ:peaks_scaling}) we could further conclude that $\nu$ for fully coordinated percolation should be the same with that for ordinary percolation. This again confirms the statement that fully coordinated percolation is in the same {\em static} universality class as ordinary percolation. Another universal constant often used to characterize ordinary percolation is the amplitude ratio $C_{+}/C_{-}$ of susceptibility $\chi$ (whose value is about 200 in $d=2$ \cite{stauffer94}). In fully coordinated percolation, this quantity is unfortunately difficult to calculate accurately because the critical region for $p > p_c$ is very small (see below). When we constrain the exponent $\gamma$ to be close to $\gamma_o$ and use the $p_c$ estimated in this work, however, we find that $C_{+}/C_{-}$ is of $\vartheta(10^2)$, which is consistent with the above observation as well. The contribution of the largest cluster to the susceptibility is not significant when $p < p_{c}$. However when $p \sim p_{c}$~a significant number of sites will belong to the largest cluster and when $p > p_{c}$ the largest cluster is dominant in the whole lattice. The average susceptibility contribution due to this largest cluster is $\chi_{1} = \sum s^{2}_{max}/(L^{2} N)$, where the summation is over $N = 1000$ realizations and $s_{max}$ is the size of the largest cluster. The fractal dimension, $d_f$, can be obtained from $s_{max}$ by \begin{equation} s_{max}(p_c)~\sim~L^{d_f} \label{equ:frac_dim} \end{equation} where $\overline{s}_{max}(p_c)$ is the mean size of the largest cluster at $p_c$. $\chi_{1}$ should therefore scale as \begin{equation} \chi_{1}~\sim~L^{2 d_f~-~2}. \label{equ:large_scaling} \end{equation} For ordinary percolation on a two dimensional lattice (see, e.g., \cite{stauffer94}), it is known that $d_f = 91/48$ and $y \equiv 2 d_f - 2 \sim 1.7917$. For fully coordinated percolation, the scaling in Eq.\ (\ref{equ:large_scaling}) have two unknowns, $p_c$ and $d_f$. Similar to what we have done when examining the scaling in Eq.\ (\ref{equ:chi_scaling}), we choose trial values for $p_c$ and then perform a least squares fit to obtain the corresponding $y = 2 d_f - 2$. By looking for the range of trial $p_c$ that maximizes the regression coefficient $|R|$, we arrive at an estimate of $p_c$ to be close to $0.8845$ where $|R| = 1$ and $y = 1.7855$. The variation of $|R|$ is about 2 parts in $10^5$ if $p_c$ is varied by 0.0002, always with less than 1\% deviation from the ordinary percolation value of $y$. From these results we conclude that $d_f$ for fully coordinated percolation is the same as that for ordinary percolation as well as the estimate of about $0.8845$ for $p_c$. In addition to the above, we have also performed the scaling analysis of the quantity $\chi_1 (p,L)$, as both $p$ and $L$ are varied, in the form of \begin{equation} \chi_{i1}~=~L^{2 d_f~-~2}~g(|p~-~p_c|^{\nu}~L) \label{equ:large_outside_scaling} \end{equation} where $g(|p - p_c|^{\nu}~L)$ is a scaling function. Using the exactly known ordinary percolation values of the exponents $d_f$ and $\nu$ (as they have been shown to be the same for fully coordinated percolation above), we obtained the maximum data collapsing in the range of $0.884 < p_c < 0.885$. Independent of the above analyses based on the fully coordinated clusters obtained by Monte Carlo simulations of fixed-sized square grids, we have also performed Monte Carlo simulations by growing fully coordinated clusters starting from a seed site using a variant of the {\em breadth-first search} algorithm \cite{nakanishi94}. This latter approach has an advantage that there is no obvious finite-size effects and that statistics taken while a cluster is still growing represents a {\em partial sum} automatically. That is, we start growing such clusters 10,000 times at each of $p = 0.880$, 0.884, 0.885, and 0.890, and keep track of how many of them are still growing at predetermined intervals of size ($2^n$ where $n = 1$, 2, ... , 15 in our case). This number, say, $N_s$ represents the partial sum \begin{equation} N_s /N_1 = \frac{1}{p} \sum_{s' \geq s} s' n_{s'} \end{equation} Since the normalized number of size-$s$ clusters, $n_s$ scales as $s^{-\tau} f(\epsilon s^{\sigma})$ where $\tau =187/91$ and $\sigma = 36/91$, we expect that $N_s$ scales as \begin{equation} N_s s^{\tau -2} \sim \hat{f}(\epsilon s^{\sigma}) \end{equation} near $p_c$ and for large $s$. In particular, at $p_c$, this quantity should be constant independent of (large) $s$. The numerical results are shown in Fig.~2, where the data correspond, from highest to lowest, to $p=0.890$, 0.885, 0.884, and 0.880. The horizontal dashed line drawn to guide the eye makes it clear that data for $p=0.885$ best approximates a horizontal line as $s \rightarrow \infty$, suggesting that a good estimate of $p_c$ would be 0.885. We now consider all the above results together. The results from scaling in Eq.\ (\ref{equ:chi_scaling}) indicate a range $0.885 - 0.886$, and those from scaling in Eq.\ (\ref{equ:large_scaling}) indicate $0.8844 - 0.8847$, while those from scaling in Eq.\ (\ref{equ:large_outside_scaling}) hints $p_c$ to be in the $0.884 - 0.885$ interval. Another result that could also be used are the values of $p$ for the peaks in Fig.~1, which vary from $0.8841$ to $0.8844$ (with the peak for the largest grid $L = 2048$ occurring at $p_{peak} = 0.8844$). Combining all these results, our final estimate is $p_c = 0.885 \pm 0.001$. \narrowtext \section{DYNAMIC CRITICAL BEHAVIOR} \label{sec:dynamic} By dynamic critical behavior here we simply mean the asymptotic long-time behavior of diffusion taking place on an incipient infinite cluster of fully coordinated percolation, or its equivalent {\em scalar} elastic behavior. This represents the simplest kind of dynamics associated with these complicated geometrical objects and is mainly reflected in the two {\em dynamic} critical exponents called $d_s$ (spectral dimension) and $d_w$ (walk dimension). It is well known that the return-to-the starting point probability of the random walk, $P(t)$, in the long-time limit obeys the power law, \begin{equation} P(t) \sim t^{-d_s/2} , \end{equation} where $d_s$ is the spectral dimension of the walk. In a fractal medium, $d_s$ is less than the space dimension $d$, because the progressive displacement of the random walker further from the starting point is hampered by its encounter with the irregularities of the medium at all scales. Thus, $d_s$ is expected to be greater for environments that provide higher connectivity at large length scales, independently of the fractal dimension itself which is mainly the measure of the overall {\em size} scales or {\em how many} sites are connected, not {\em how well} those sites are connected to each other. For media with long-range loops, $d_f$ (fractal dimension), $d_s$, and $d_w$ are not independent but are expected to obey the well-known Alexander and Orbach scaling law \cite {alexander} \begin{equation} \label{eq:ao} d_s = 2d_f/d_w \;. \end{equation} For this reason, we only calculate $d_s$ here though both of $d_s$ and $d_w$ can be conveniently calculated by numerically studying the {\em transition probability matrix} {\bf W} which represent the random walk on a specific fractal medium. Our calculation in this work is only one aspect of such an analysis: finite size scaling of the dominant non-trivial eigenvalue which describes the longest finite time scale of the Brownian process. This approach has already been described in detail elsewhere \cite{mukherjee94}, and thus we merely state the main feature and then immediately report our specific numerical results. The matrix {\bf W} is constructed from the elements $W_{ij}$ being equal to a hopping probability per step (equal to $\frac{1}{4}$ here) for available nearest neighbor sites $i$ and $j$. For each neighbor site which is not present, a probability of $\frac{1}{4}$ is added to the probability for not taking a step for one time period - this is called the {\em blind ant} rule. Many large matrices {\bf W} are obtained by Monte Carlo simulation (by growing a fully coordinated percolation cluster from a seed site and stopping the growth when a predetermined desired size is reached) and their largest eigenvalues are numerically obtained by the so-called {\em Arnoldi-Saad} method. The dominant non-trivial eigenvalue $\lambda_1$ is the largest eigenvalue just below the stationary eigenvalue 1 and it is known to satisfy the following finite size scaling law: \begin{equation} \label{eq:fsfords} | \ln \lambda_1 | \approx 1-\lambda_1 \sim S^{-2/d_s} \;. \end{equation} Shown in Fig.~3 are our results from such an analysis. We have generated at least 1000 independent realizations of the underlying fully coordinated percolation clusters for sizes $S = 1250$, 2500, 5000, 10000, and 20000 at each of the three nominal probabilities $p=0.883$, 0.885, and 0.887, and numerically obtained $\lambda_1$ for each cluster. The main part of Fig.~3 shows the data from $p=0.885$, the value shown to be closest to $p_c$ in this work. The figure shows an excellent power law fit (regression coefficient of -0.99993) to Eq.(\ref{eq:fsfords}) with the exponent $2/d_s = 1.486 \pm 0.01$. This power translates to $d_s = 1.346 \pm 0.011$ which is close to but definitely larger than the corresponding ordinary percolation value of $d_s^{(O)} = 1.30 \pm 0.02$ estimated by many independent calculations. (See, e.g., \cite{nakanishi94}.) For comparison, a {\em loopless} variant of percolation has exactly the same static exponents as ordinary percolation but has $d_s \approx 1.22$ in two dimensions, about twice as much deviation from ordinary percolation in the opposite direction as the present fully coordinated percolation problem. \cite{loopless} In the inset for Fig.~3, we show a normalized $\lambda_1$ by plotting $(1-\lambda_1) S^{2/d_s^{(o)}}$ where the circles are for $p=0.887$, squares for $p=0.885$, diamonds for $p=0.883$ and the solid line is a horizontal line (for ordinary percolation) to guide the eye. In all cases, the standard errors of the mean for each set of data are substantially smaller than the size of the symbols used in the figure. The distribution of $\lambda_1$ in each case appears to be Gaussian with the standard deviations scaling in the same way as the means. >From these results, we conclude that, though the numerical differences are small, there is a high likelihood that the fully coordinated percolation clusters are significantly different from the ordinary percolation counterparts even at long length scales. In the next section, we show that this analysis is vindicated by exposing one dramatic difference in the cluster morphology which will not be obvious to an uncritical observer. \narrowtext \section{CLUSTER GEOMETRY} \label{sec:geometry} In this section we examine the geometry of fully coordinated percolation clusters more closely. First, we present Fig.~4 which show in grey scale the sites of (a) fully coordinated percolation cluster and (b) ordinary percolation cluster at respective $p_c$. The overall visual impression is that they are very similarly shaped even down to the details of the boundaries and internal holes. Their shapes are also essentially independent of the underlying lattice anisotropy (as Fig.~4(a) has actually been rotated by 45 degrees with respect to the coordinate axes of the square lattice). However, the number and distributions of the especially dark points are evidently quite distinct in (a) and (b). They cluster more and are much more abundant in (a) than in (b). These sites are actually the {\em interior} or fully coordinated sites in the internal part of the cluster. The remaining sites (shaded grey) are either the external {\em hull} sites or internal boundary sites. In Fig.~5 quantitative examination is made on the different classes of sites of the two kinds of clusters. In the main part of the figure, the average numbers of two kinds of sites are shown, {\em interior} (diamonds for fully coordinated percolation, crosses for ordinary percolation), and {\em external} (squares for fully coordinated percolation, plusses for ordinary percolation). It is clear that the interior sites are more than 3 times abundant in fully coordinated percolation as is visually suggested by Fig.~4. Though this is primarily a local effect due to the full coordination rule, they do have a multiplicative effect at long-range connectivity and thus may well be the source of the small difference in the value of $d_s$. Of course just the fact that there are more than 3 times as many interior sites (and correspondingly, much fewer {\em hull} sites) in fully coordinated percolation must have quantitative consequences (even if not qualitative) for any process on the cluster which depend on degrees of connectivity rather than just on the number of connected sites. An example of the effect of the different numbers of the hull sites may be in oxidation or catalysis of a material through the external embedding phase or even a irregularly shaped breakwater in the form of the external boundary of a percolation cluster. The {\em external} sites are those which are sometimes called {\em hull} sites and they are known to scale with an exact exponent in ordinary percolation as \begin{equation} N_{hull} \sim S^{d_h /d_f} \end{equation} where $d_h^{(o)} = 7/4$ and thus the exponent is $84/91 = 0.923...$ for ordinary percolation. Since this is less than 1, these sites comprise less and less fraction of the cluster as $s \rightarrow \infty$ and the remaining sites (i.e., {\em interior} and {\em internal boundary} sites) eventually dominate the whole cluster. This is already evident from the greater slopes close to 1 for the {\em interior} sites in Fig.~5. The linear regression fits for the {\em hull} sites in Fig.~5 indicate the slopes of about 0.913 for ordinary percolation and 0.922 for fully coordinated percolation with essentially perfect fits, again reinforcing the conclusion that they show the same {\em static} critical behavior. \narrowtext \section{SUMMARY AND CONCLUSION} \label{sec:summary} In summary, we have studied both static and dynamic critical behaviors associated with a model of the highly connected regions of a disordered cluster. The model is a {\em site} variant of the four-coordinated percolation \cite{blumberg80,gonzales80} on the square lattice we call {\em fully coordinated percolation}. While the bond version was studied for {\em static} critical behavior, neither bond nor site version was previously studied for the dynamic behavior to the best of our knowledge. We have used various methods such as Monte Carlo simulations, finite-size scaling and Arnoldi-Saad approximate diagonalization of large random matrices for this purpose. Though all indications are that the static behavior of this model is exactly the same as the ordinary percolation (as previous work suggested), the dynamic behavior shows a small but significant difference in the values of the universal critical exponents. We have looked for the cause of this difference and found a three-fold increase in the number and significantly enhanced clustering of the interior sites (i.e., those not on the exterior or internal boundaries) and the associated decrease in the number of boundary sites. Thus, although the deviations from ordinary percolation in terms of the values of the dynamic critical exponents are not large, there will be rather significant differences in any processes that depend sensitively on those numbers. Possible examples of such processes include the oxidation of a material through the external embedding phase and the vibrational normal modes with boundary conditions such as {\em clamping} or {\em tethering} of the external boundaries (through the contrast in elastic constants of embedding and embedded materials, for example). \acknowledgements One of us (JHK) wishes to thank the Purdue University Department of Physics for the hospitality during his visit there when part of the work was done. We are also grateful to D. Stauffer and R. Ziff for insightful remarks.
\section{Introduction} There are two reasons for studying the data from precision electroweak measurements. First, we would like to know how well the Standard Model (SM) agrees with the experimental data: any significant deviation between the two would be a signal of new physics beyond the SM. Also, since the SM predictions depend on the mass of the still unobserved Higgs boson, the value of $m_H$ which best reproduces the experimental data serves as its prediction. Second, we would like to know if any of the proposed theories of new physics beyond the SM is viable, i.e. does not make theoretical predictions inconsistent with the experimental data (or perhaps makes the agreement even better than with the SM). The way to kill these two birds with one stone is to constrain the sizes of radiative corrections coming from new physics using the precision electroweak data. If the data prefers a non--zero value for the size of these extra corrections for some choice of $m_H$, it would signal that: (1) the agreement between the SM and data is not perfect for that particular choice of Higgs mass, and (2) any new physics which predicts such corrections would make the agreement between theory and experiment better. \section{Constraints on Oblique Electroweak Corrections} In order to constrain the size of radiative corrections coming from new physics, we must first make some assumptions about the new physics giving rise to them (since it is impossible to consider all possible corrections from all possible new physics all at once) and express them in terms of a few model independent parameters. \begin{figure}[t] \setlength{\unitlength}{1mm} \begin{picture}(130,55)(-65,-10) \put(-40,0){\line(1,0){80}} \put(0,-10){\vector(0,1){50}} \put(-5,42){Energy} \put(-32,-5){Charged Channel} \put(5,-5){Neutral Channel} \put(-22,4){$G_\mu$} \put(18.8,4.2){$\alpha$} \put(-22,28){$M_W$} \put(17.3,33.2){$M_Z$} \put(0,5){\vector(-1,0){15}} \put(0,5){\vector(1,0){15}} \put(20,20){\vector(0,-1){10}} \put(20,19){\vector(0,1){10}} \put(-20,20){\vector(0,-1){10}} \put(-20,20){\vector(0,1){5}} \put(3,6.5){$T$} \put(21.5,19){$S$} \put(-30.5,15){$S+U$} \put(-20,5){\circle{7}} \put(20,5){\circle{7}} \put(20,34){\circle{7}} \end{picture} \caption{The $S$, $T$, and $U$ variables quantify the extra vacuum polarization corrections due to new physics when relating physics in different channels or physics at different energy scales. Low energy and neutral current observables depend only on $S$ and $T$ because $\alpha$, $G_\mu$, and $M_Z$ are used as inputs to make the SM predictions. \label{FIG1}} \end{figure} The simplest assumptions that would sufficiently constrain the number of corrections one must consider while still encompassing a large class of models are the following: \begin{enumerate} \item{The electroweak gauge group is the standard $SU(2)_L \times U(1)_Y$. The only electroweak gauge bosons are the photon, the $W^\pm$, and the $Z$.} \item{The couplings of new physics to light fermions are highly suppressed. Since all precision electroweak measurements only involve four fermion processes with light external fermions, this means that vertex and box corrections from new physics can be neglected. Only vacuum polarization (i.e. oblique) corrections need to be considered. } \item{The mass scale of new physics is large compared to the $W$ and $Z$ masses.} \end{enumerate} The first two assumptions let us focus our attention on just four vacuum polarization corrections: $\AQ{WW}$, $\AQ{ZZ}$, $\AQ{Z\gamma}$, and $\AQ{\gamma\gamma}$. Here, $\AQ{XY}$ is the transverse part of the vacuum polarization function between gauge bosons $X$ and $Y$. The third assumption lets us expand these vacuum polarization functions around $q^2=0$ and neglect the higher order terms since they are suppressed by powers of $q^2/M_\mathrm{new}^2$ where $q^2 \le M_Z^2$ for the processes under consideration: \begin{eqnarray*} \AQ{WW} & = & \AZERO{WW} + q^2 \APZERO{WW} + \cdots \cr \AQ{ZZ} & = & \AZERO{ZZ} + q^2 \APZERO{ZZ} + \cdots \cr \AQ{Z\gamma} & = & q^2 \APZERO{Z\gamma} + \cdots \cr \AQ{\gamma\gamma} & = & q^2 \APZERO{\gamma\gamma} + \cdots \end{eqnarray*} Therefore, we can express the radiative corrections from new physics in terms of just six parameters: $\AZERO{WW}$, $\APZERO{WW}$, $\AZERO{ZZ}$, $\APZERO{ZZ}$, $\APZERO{Z\gamma}$, and $\APZERO{\gamma\gamma}$. Of these six, three will be absorbed into the renormalization of the three input parameters $\alpha$, $G_\mu$ and $M_Z$, leaving us with three observables parameters which can be taken to be: \cite{PESKIN:90} \begin{eqnarray*} \alpha S & = & 4s^2 c^2 \left[ \APZERO{ZZ} -\frac{c^2-s^2}{sc}\APZERO{Z\gamma} -\APZERO{\gamma\gamma} \right]\,, \nonumber \\ \alpha T & = & \frac{\AZERO{WW}}{M_W^2} - \frac{\AZERO{ZZ}}{M_Z^2}\,, \\ \alpha U & = & 4s^2 \left[ \APZERO{WW} - c^2\APZERO{ZZ} - 2sc\APZERO{Z\gamma} - s^2\APZERO{\gamma\gamma} \right]\,. \nonumber \phantom{\frac{e^2}{s^2}} \end{eqnarray*} Here, $\alpha$ is the fine structure constant and $s$ and $c$ are the sine and cosine of the weak mixing angle. Only the contribution of new physics to these functions are to be included. The parameters $T$ and $U$ are defined so that they vanish if new physics does not break custodial $SU(2)$ symmetry. See Ref.~2 for a discussion on the symmetry properties of $S$. \begin{table}[t] \begin{center} \begin{tabular}{|l|c|c|c|} \hline Observable & SM prediction & Measured Value & Reference\\ \hline\hline \underline{$\nu_\mu e$ and $\bar{\nu}_\mu e$ scattering} & & & \\ $g_V^{\nu e}$ & $-0.0365$ & $-0.041 \pm 0.015$ & \cite{PDG:98} \\ $g_A^{\nu e}$ & $-0.5065$ & $-0.507 \pm 0.014$ & \cite{PDG:98} \\ \hline \underline{Atomic Parity Violation} & & & \\ $Q_W({}^{133}_{\phantom{1}55}{\rm Cs})$ & $-73.19$ & $-72.41 \pm 0.84$ & \cite{PDG:98} \\ $Q_W({}^{205}_{\phantom{1}81}{\rm Tl})$ & $-116.8$ & $-114.8 \pm 3.6$ & \cite{PDG:98} \\ \hline \underline{$\nu_\mu N$ and $\bar{\nu}_\mu N$ DIS} & & & \\ $g_L^2$ & $0.3031$ & $0.3009 \pm 0.0028$ & \cite{PDG:98} \\ $g_R^2$ & $0.0304$ & $0.0328 \pm 0.0030$ & \cite{PDG:98} \\ NuTeV & $0.2289$ & $0.2277 \pm 0.0022$ & \cite{NUTEV:98} \\ \hline \underline{LEP/SLD} & & & \\ $\Gamma_{\ell^+\ell^-}$ & $0.08392$ GeV & $0.08390 \pm 0.00010$ GeV & \cite{LEP:98} \\ $\sin^2\theta_\mathrm{eff}^\mathrm{lept}$ (LEP) & $0.23200$ & $0.23153 \pm 0.00034$ & \cite{LEP:98} \\ $\sin^2\theta_\mathrm{eff}^\mathrm{lept}$ (SLD) & $0.23200$ & $0.23109 \pm 0.00029$ & \cite{LEP:98} \\ \hline \underline{$W$ mass} & & & \\ $M_W$ ($p\bar{p}$ + LEP2) & $80.315$ GeV & $80.39 \pm 0.06$ GeV & \cite{LEP:98} \\ \hline \end{tabular} \caption{ The data used for the oblique correction analysis. The value of $\sin^2\theta_\mathrm{eff}^\mathrm{lept}$ for LEP is from leptonic asymmetries only. The SM predictions for the $W$ mass and the LEP/SLD observables were obtained using the program ZFITTER 4.9 \protect\cite{ZFITTER:92}. The predictions for the low energy observables were calculated from the formulae given in Ref.~7. The parameter choice for the reference SM was $M_Z = 91.1867$ GeV\protect\cite{LEP:98}, $m_t = 173.9$ GeV\protect\cite{TOPMASS:98}, $m_H = 300$ GeV, $\alpha^{-1}(M_Z) = 128.9$\protect\cite{ALEMANY:98}, and $\alpha_s(M_Z) = 0.120$. \label{DATA1}} \end{center} \end{table} The fact that three parameters are necessary to describe the effects of new physics can also be understood as follows: Since vacuum polarizations modify the gauge boson propagators, their presence can be seen when comparing the exchange of different electroweak gauge bosons, or when comparing the exchange of the same boson at different energy scales. In order to make predictions based on the three input parameters $\alpha$, $G_\mu$, and $M_Z$, which are neutral current--low energy, charged current--low energy, and neutral current--high energy observables, respectively, one must compare the theory in the charged ($W$ exchange) and the neutral ($Z$ and photon exchange) channels, as well as in the same channel at different energy scales as shown by arrows in Fig.~\ref{FIG1}. New physics effects will manifest themselves in each of the three arrows shown in Fig.~\ref{FIG1}. The parameter $S$ quantifies the extra correction from new physics one must include when comparing neutral current processes at different energy scales while the parameter $T$ quantifies the extra corrections that must be included when comparing charged and neutral current processes at low energy. A third parameter, in this case $S+U$, is necessary to quantify the correction due to new physics when comparing charged channel processes at different energy scales. This discussion also shows that all low energy and neutral current observables will receive extra corrections from only $S$ and $T$. Of all the precision electroweak measurements, the only quantity which receives a correction from $U$ is the $W$ mass. Therefore, the majority of the precision data can be used to constrain just two parameters $S$ and $T$, which in turn can be calculated for each model of new physics beyond the SM. The details of how to calculate the corrections to various observables from $S$, $T$, and $U$ can be found elsewhere\cite{PESKIN:90}. One obtains expressions such as \[ M_W/\SM{M_W} = 1 + \frac{\alpha}{2(c^2-s^2)} \left[ -\frac{1}{2}S + c^2 T + \frac{c^2-s^2}{4s^2} U \right], \] where $\SM{M_W}$ is the SM prediction of $M_W$. These expressions can be compared directly with the experimental data to place constraints on $S$, $T$, and $U$. \begin{figure}[t] \setlength{\unitlength}{1mm} \begin{picture}(130,100)(0,0) \put(0,50){ \epsfxsize=2.5in \epsfbox{st-apv.ps} } \put(40.5,85){$Q_w(\mathrm{Tl})$} \put(37.5,66){$Q_w(\mathrm{Cs})$} \put(65,50){ \epsfxsize=2.5in \epsfbox{st-enu.ps} } \put(86,72){$g_V^{\nu e}$} \put(118.5,72){$g_A^{\nu e}$} \put(0,0){ \epsfxsize=2.5in \epsfbox{st-dis.ps} } \put(57,30.5){$g_L^2$} \put(50.5,12.5){$g_R^2$} \put(19,9.5){NuTeV} \put(65,0){ \epsfxsize=2.5in \epsfbox{st-lep.ps} } \put(92,9){LEP leptonic asym.} \put(93.5,40.5){SLD} \put(116.5,36.5){$M_W$} \put(115,21){$\Gamma_{\ell^+\ell^-}$} \put(12,93){(a)} \put(77,93){(b)} \put(12,43){(c)} \put(77,43){(d)} \end{picture} \begin{picture}(130,50)(0,0) \put(0,0){ \epsfxsize=2.5in \epsfbox{st-all90.ps} } \put(16,25){$e\nu$} \put(44,35){$eN$} \put(20.5,40){$Q_w$} \put(65,0){ \epsfxsize=2.5in \epsfbox{st-all.ps} } \put(101.5,29){$m_t$} \put(102.5,18.5){$m_H$} \put(104.5,25){$\alpha^{-1}$} \put(12,43){(e)} \put(77,43){(f)} \end{picture} \caption{The 1--$\sigma$ limits on $S$ and $T$ from (a) atomic parity violation, (b) $e\nu_\mu$ and $e\bar{\nu}_\mu$ scattering, (c) $\nu_\mu$ and $\bar{\nu}_\mu$ DIS, and (d) LEP/SLD + $M_W$. (e) The 90\% confidence contours for the four classes of experiments. The LEP/SLD + $M_W$ contour is the small shaded area in the middle. (f) The 68\% and 90\% confidence limits on $S$ and $T$, all experiments combined. The arrows indicate how the SM point will move relative to the contours for $m_t = 173.9 \pm 5$ GeV, $m_H = 300^{+700}_{-210}$ GeV, $\alpha^{-1}(M_Z) = 128.9 \pm 0.1$. \label{FIG2}} \end{figure} In Table~\ref{DATA1} we list the data we used in our analysis. The definitions of the parameters $g_{V/A}^{e\nu}$ for $e\nu_\mu$ and $e\bar{\nu}_\mu$ scattering, the weak charge $Q_w$ for atomic parity violation, and $g_{L/R}^2$ for $\nu_\mu$ and $\bar{\nu}_\mu$ deep inelastic scattering (DIS) can be found in the Review of Particle Physics\cite{PDG:98} from which the data were taken. The quantity measured by the NuTeV collaboration\cite{NUTEV:98} is a linear combination of $g_L^2$ and $g_R^2$ for which the uncertainty due to the charm threshold cancels. The rest of the data is from Ref.~5. Comparing the experimental data to SM predictions with $m_t = 173.9$~GeV\cite{TOPMASS:98}, $m_H = 300$~GeV, and $\alpha^{-1}(M_Z) = 128.9$\cite{ALEMANY:98}, we obtain the constraints shown in Fig.~\ref{FIG2}. Note that Figs.~\ref{FIG2}d and \ref{FIG2}f are drawn at a different scale from the other four. Fig.~\ref{FIG2}e shows the 90\% confidence limits on $S$ and $T$ due to the four classes of experiments separately, and Fig.~\ref{FIG2}f shows the 68\% and 90\% confidence limits from all experiments combined. As is evident from Fig.~\ref{FIG2}, the LEP/SLD measurements provide the tightest constraints on $S$ and $T$. All of the other observables combined have little effect on the final result, which is \begin{eqnarray*} S & = & -0.30 \pm 0.13, \cr T & = & -0.14 \pm 0.15, \cr U & = & \phantom{-}0.15 \pm 0.21. \end{eqnarray*} These limits of course depend on the values of $m_t$, $m_H$, and $\alpha^{-1}(M_Z)$ used as input to calculate the SM predictions. The dependence of the limits on these input parameters is shown by arrows in Fig.~\ref{FIG2}f. We can see that the current data favor either a small value of the Higgs mass or a larger value of $\alpha^{-1}(M_Z)$. \section{Limits on Topcolor Assisted Technicolor} \begin{table}[t] \begin{center} \begin{tabular}{|c||c|c|c|c|c|} \hline & $SU(3)_s$ & $SU(3)_w$ & $U(1)_s$ & $U(1)_w$ & $SU(2)_L$ \\ \hline\hline $(t,b)_L$ & 3 & 1 & $\dfrac{1}{3}$ & 0 & 2 \\ \hline $(t,b)_R$ & 3 & 1 & $\left(\dfrac{4}{3},-\dfrac{2}{3}\right)$ & 0 & 1 \\ \hline $(\nu_\tau,\tau^-)_L$ & 1 & 1 & $-1$ & 0 & 2 \\ \hline $\tau^-_R$ & 1 & 1 & $-2$ & 0 & 1 \\ \hline $(c,s)_L$, $(u,d)_L$ & 1 & 3 & 0 & $\dfrac{1}{3}$ & 2 \\ \hline $(c,s)_R$, $(u,d)_R$ & 1 & 3 & 0 & $\left(\dfrac{4}{3},-\dfrac{2}{3}\right)$ & 1 \\ \hline $(\nu_\mu,\mu^-)_L$, $(\nu_e,e^-)_L$ & 1 & 1 & 0 & $-1$ & 2 \\ \hline $\mu^-_R$, $e^-_R$ & 1 & 1 & 0 & $-2$ & 1 \\ \hline \end{tabular} \end{center} \caption{Charge assignments of the ordinary fermions in topcolor assisted technicolor.\label{CHARGES}} \end{table} The limits on $S$ and $T$ are useful in constraining new physics models which satisfy the three initial assumptions but are less useful for other theories. As an example of such a theory, let us consider topcolor assisted technicolor\cite{HILL:95} with the gauge group \[ SU(3)_s \times SU(3)_w \times U(1)_s \times U(1)_w \times SU(2)_L \] The coupling constants for the two $SU(3)$'s and the two $U(1)$'s are assumed to satisfy $g_{3s}\gg g_{3w}$ and $g_{1s}\gg g_{1w}$. The charges of the ordinary fermions under these groups are shown in Table~\ref{CHARGES}. At a scale of about one TeV, it is assumed that technicolor breaks the two $SU(3)$'s and the two $U(1)$'s to their diagonal subgroups: \[ SU(3)_s \times SU(3)_w \rightarrow SU(3)_c,\qquad U(1)_s \times U(1)_w \rightarrow U(1)_Y, \] which are identified with the SM color and hypercharge groups. Because of the assumption $g_{3s}\gg g_{3w}$ and $g_{1s}\gg g_{1w}$, the broken $SU(3)$ gauge bosons (the \textit{colorons}) and the broken $U(1)$ gauge boson (the $Z'$) couple strongly to the third generation fermions but only weakly to the first and second generation fermions. Coloron exchange is attractive in both the $t\bar{t}$ and $b\bar{b}$ channels, while $Z'$ exchange is attractive for $t\bar{t}$ but repulsive for $b\bar{b}$. The combined strength of the coloron and $Z'$ interactions is assumed to be strong enough to condense the top but not so strong as to condense the bottom. As a result, only the top quark becomes heavy. It is easy to see that this model does not fit into the $STU$ framework since (1) it has an extra electroweak gauge boson, the $Z'$, and (2) coloron and $Z'$ exchange can lead to large vertex corrections for the third generation fermions ($b$, $\tau$, and $\nu_\tau$). How would we place constraints on such a model? A naive extension of the $STU$ formalism to include the $Z'$ vacuum polarization functions turns out to be too complicated to be illuminating. A much better way is to concentrate our attention on the vertex corrections at the $Z$ mass scale, where we have a wealth of data from LEP and SLD. Recall from our previous discussion that $S$ and $T$ are relevant only when comparing processes at different energy scales or in different channels. If we only look at the LEP/SLD data, which come from neutral current processes at the $Z$ mass scale, we can make our analysis completely blind to the vacuum polarization corrections and obtain limits on the vertex corrections only. Another way to see this is to notice that most of the observables at LEP/SLD are asymmetries and branching fractions which are just \textit{ratios of coupling constants} at the $Z$ mass scale. The SM predictions for these observables can be fixed by using only one of them as input to predict all the others, and any deviations must come from the vertex corrections due to new physics.\footnote{% We have used a similar technique in Ref.~11.} Since the details of our analysis has been presented elsewhere\cite{LOINAZ:98}, we give only an outline here. In the topcolor assisted technicolor model considered above, vertex corrections come in two classes: (1) gauge boson mixing terms, and (2) proper vertex corrections. Gauge boson mixing modifies the current to which the $Z$ couples to from $J_Z^0 = J_{I_3} - s^2 J_Q$ to \[ J_Z = J_{I_3} - (s^2 + \delta s^2) J_Q + \epsilon J_{1s}. \] The parameters $\delta s^2$ and $\epsilon$ quantify the amount of $Z$--photon and $Z$--$Z'$ mixing, respectively. The relevant proper vertex corrections are the coloron and $Z'$ corrections to the third generation fermion vertices\cite{HILL:95A}, the sizes of which we parametrize by \[ \kappa_i = \frac{g_{is}^2}{4\pi} \left( \frac{g_{is}^2}{g_{is}^2+g_{iw}^2} \right), \qquad (i=1,3), \] and a correction to the left--handed coupling of the $b$ to the $Z$ from the top--pion loop\cite{BURDMAN:97} which we denote $\Delta$. The corrections to various LEP/SLD obserables from $\delta s^2$, $\epsilon$, $\kappa_1$, $\kappa_3$, and $\Delta$ were calculated and compared to the experimental data. In performing the fit, we kept the value of $\Delta$ fixed and let the four other parameters and the QCD coupling constant $\alpha_s(M_Z)$ float. In Fig.~\ref{FIG3}, we show only the results in the $\kappa_1$--$\kappa_3$ plane for two choices of the value of $\Delta$: $0.003$ and $0.006$. These correspond to top--pion masses of $m_+ = 1000$~GeV and $m_+ = 600$~GeV, respectively. $\kappa_1$ and $\kappa_3$ must fall into the shaded region in order for the coloron and $Z'$ interactions to condense the top while not condensing the bottom. Clearly the $\Delta = 0.003$ case is viable while the $\Delta = 0.006$ case is ruled out. \begin{figure}[t] \setlength{\unitlength}{1mm} \begin{picture}(130,50)(0,0) \put(0,0){ \epsfxsize=2.5in \epsfbox{tc2-1.ps} } \put(39,40){$R_b$} \put(54,17){$R_\tau$} \put(10,44){$\Delta=0.003$} \put(65,0){ \epsfxsize=2.5in \epsfbox{tc2-2.ps} } \put(121,40){$R_b$} \put(95,10.5){$R_\tau$} \put(75,44){$\Delta=0.006$} \end{picture} \caption{Limits on $\kappa_1$ and $\kappa_3$ from LEP/SLD observables. $\Delta$ parametrizes the size of the top--pion correction to the $Zb\bar{b}$ vertex. The shaded region is where the combined coloron and $Z'$ interaction can condense the top without condensing the bottom or the $\tau$. \label{FIG3}} \end{figure} \section{Conclusions} Precision electroweak measurements provide stringent constraints on new physics beyond the SM. For models which satisfy the three conditions listed in section~2, the limits can be described in a model independent way using the $STU$--formalism of Ref.~1. Currently, the tightest limits come from the LEP/SLD observables, with all other observables having little effect. Current data also favors a smaller Higgs mass or a larger $\alpha^{-1}(M_Z)$. For models which are not encompassed within the $STU$ framework, in particular, those with extra electroweak gauge bosons and/or large vertex corrections, one can still obtain stringent limits, albeit in a model dependent way, by focussing only on the vertex corrections at the $Z$ mass scale and using the LEP/SLD observables to constrain them. \section*{Acknowledgments} We would like to thank M.~W.~Gr\"unewald for providing us with the data used in this analysis. This work was supported in part by the U. S. Department of Energy, grant DE-FG05-92-ER40709, Task A (W.L.), and the National Science Foundation, grant NSF~PHY-9802709 (A.K.G.).
\section{Introduction} Galactic bars are the sites of highly diversified star formation activities. Phillips (1993, 1996), Garc\'{\i}a-Barreto {\rm et al.}\ (1996) and Martin \& Friedli (1997, hereafter MF97) have all showed that {\it along} certain bars, mostly found in late-type spirals, star formation (SF) can be quite intense. For instance, the bar of the SBcd galaxy NGC\,4731 has a total star formation rate (SFR) of about 1.5 M$_\odot$ yr$^{-1}$ (MF97). However, in other cases, SF in the bar can be very weak or completely absent. This is the case for most of the bars in early-type barred spirals (e.g. NGC\,1300, NGC\,1512, NGC\,3351). The origin of these differences is not yet completely understood. Numerical simulations (Friedli \& Benz 1995, MF97, Martinet \& Friedli 1997) and observations (Martin \& Roy 1995) suggest that the existence of massive star formation in certain bars is a relatively brief event ($\sim$0.5--1.0\,Gyr) in the evolution of a barred system and that it mostly takes place during the formation of the bar itself. The amplitude and duration of the event is, however, controlled by a complicated combination of different physical processes. For example, the time-dependent bar evolution, initial gas content, amplitude of gas flows, and mechanical energy injected in the interstellar medium (ISM) by supernovae ejecta are all factors that can influence the level of SF activity in bars (MF97). Hence, it is essential to acquire more data on the properties of the {H\,\sc ii}\ regions formed in such environments to better constrain the relative importance of these factors, and consequently significantly improve SF recipes used in numerical simulations. The morphology, SFRs and other properties of SF along the bars of a sample of eleven spiral galaxies were studied in the first paper in this series (MF97). Including the large diversity in SF activity, MF97 found that the distribution of {H\,\sc ii}\ regions can be highly asymmetrical in the bar. For some SBc spirals, large regions can be present outside the bar major axis. This morphology suggests that the SF process along the bar of late-type spirals is a chaotic process, not strictly confined along the major stellar/gas orbits defined by the barred potential. In such a case, one could expect similar properties for these {H\,\sc ii}\ regions when compared to the disc star forming regions. The situation could be different for bars in earlier types of galaxies for which the star forming regions are generally located next to dust lanes (MF97). For one galaxy of their sample (NGC\,7479), MF97 also estimated the amount of gas flowing in the bar and falling into the galaxy center. They found that possibly as much as 75\% of the gas in the bar is not transformed into stars. Numerical simulations suggest that this number is very dependent on the presence of bar-induced shocks in the star forming ISM. It is then critical to determine whether there is some signature of these shocks in the {H\,\sc ii}\ regions along the bars. In this paper, spectrophotometric data are used to derive the physical properties of a sub-sample of {H\,\sc ii}\ regions found along the bars of the galaxies studied by MF97. As a comparison sample, a few regions located in the discs of these galaxies have also been analysed. Using different diagnostic line ratio diagrams as defined by Baldwin {\rm et al.}\ (1981) and Veilleux \& Osterbrock (1987), the excitation of regions located in bar environments is compared to that of ``normal'' disc regions (Sect.~3.1). A similar analysis is performed for the electronic density using the appropriate sulfur line ratio (Sect.~3.2). The distributions of the visual extinction for both populations are also studied in Sect.~3.3. Using the H$\alpha$\ equivalent width indicator (e.g. Leitherer {\rm et al.}\ 1999), we also infer the approximate age of {H\,\sc ii}\ regions (Sect.~3.4). Since large-scale mixing of the chemical composition by bar-driven gas radial flows occurs in galaxy discs (Martin \& Roy 1994; Friedli {\rm et al.}\ 1994), the O/H distribution in bar environments is also investigated (Sect.~3.5). \section{Database} \subsection{Observations} The long slit observations were conducted during two runs in April and October 1994 at the equivalent 4.5 meter Multi-Mirror Telescope on Mt Hopkins, Arizona. The ``Blue Channel'' spectrograph was used with a Loral 3k$\times$1k CCD. A grating with 500 grooves/mm and a blaze at 5410\,{\AA} was employed; the spectral range covered in the reduced spectra is about 3500\,{\AA} ($\sim$3500 $\rightarrow$ $\sim$7000) with a dispersion of 1.17\,{\AA}/pixel. An order sorting filter was used (UV-36) and all the data were obtained with the CCD binned by a factor of two in the spatial direction (0.6\arcsec/pixel). All these observations were carried out with an unvignetted 2\arcsec$\times$150\arcsec\ slit positioned close to the parallactic angle to avoid any light loss. Three exposures of 15 or 20 minutes were obtained for each slit position in the bar. Between exposures, the alignment of the MMT six mirrors was verified and corrected, if necessary. Numerous standard stars were observed with a larger slit (5\arcsec) during the night for the flux calibration procedure. Table~1 presents the journal of observations for the {H\,\sc ii}\ regions of ten of the eleven objects studied by MF97 (NGC\,5068 was not observed). For some objects, a few slit positions were required to optimize the number of {H\,\sc ii}\ regions observed in the bar and the disc. Average seeing during these observations was about 1.2\arcsec\ and conditions were photometric. Note that the only galaxy in the sample considered as an AGN (LINER) is NGC\,7479. \begin{table*}[t] \caption{Journal of observations.} \begin{tabular}{llccccccc} \hline\hline Galaxy & Epoch & \# Slits & Slit PA [$^\circ$] & Exposure Times [s] & $<$Airmass$>$ & $N_{\rm bar}$$^a$ & $N_{\rm disc}$$^a$ & L$_{bar}$$^b$ [kpc] \smallskip \\ \hline NGC\,1073 & 1994 Oct 9 & 1 & 68 & 3 $\times$ 1200 & 1.2 & 2 & 2 & 5.0\\ NGC\,1087 & 1994 Oct 9 & 1 & 139 & 3 $\times$ 1200 & 1.4 & 2 & 2 & 1.6\\ NGC\,3319 & 1994 Apr 17 & 1 & 40 & 1 $\times$ 1200 & 1.0 & 2 & 3 & 4.0\\ NGC\,3359 & 1994 Apr 16 & 1 & 26 & 3 $\times$ 1200 & 1.2 & 8 & 5 & 5.8\\ NGC\,3504 & 1994 Apr 16 & 2 & 140; 175 & 2 $\times$ 3 $\times$ 900 & 1.1 & 6 & 7 & 6.8\\ NGC\,4731 & 1994 Apr 16 & 1 & 126 & 3 $\times$ 900 & 1.3 & 6 & 4 & 15.2\\ NGC\,4900 & 1994 Apr 16 & 1 & 145 & 3 $\times$ 900 & 1.3 & 4 & 6 & 3.4 \\ NGC\,5921 & 1994 Apr 16 & 1 & 159 & 3 $\times$ 1200 & 1.2 & 5 & 4 & 8.1\\ NGC\,7479 & 1994 Oct 4 & 2 & 2; 11 & 2 $\times$ 3 $\times$ 1200 & 1.3 & 7 & 9 & 16.6\\ NGC\,7741 & 1994 Oct 4 & 2 & 92; 98 & 2 $\times$ 3 $\times$ 1200 & 1.1 & 7 & 1 & 3.5\\ \hline \end{tabular} \begin{minipage}{160mm} \smallskip $^a$ Number of bar and disc {H\,\sc ii}\ regions with detected H$\beta$.\\ $^b$ Total length of the bar based on distance and measurements given in MF97. \end{minipage} \end{table*} \subsection{Data reduction and analysis} The long slit spectra were reduced following standard procedures available in the {\sc longslit} package in {\sc iraf} {\footnote{IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation}. First, a bias and flat-field correction was applied. The illumination pattern along the slit was corrected using a set of sky flat-fields taken in the same optical configuration. Wavelength calibration was done using a Helium--Neon--Argon exposure taken immediately after the science observation. Geometric distortion and alignment of the spectra were corrected by 2D-mapping of the spectral lines from the calibration sources using the {\sc fitcoords} and {\sc transform} algorithms. The next step, sky subtraction, was done by extracting a sky background using sections along the long slit outside the galaxy. To make sure that the signal from the galaxy was minimized, the average sky backgrounds were compared between the different slit positions since most of the slits used for the disc {H\,\sc ii}\ regions were less contaminated by the disc emission. The spectra were then flux calibrated using a sensitivity function obtained from a set of standard stars observed through the nights. Extraction of the spectrum for each {H\,\sc ii}\ region was done using the spatial profile at H$\alpha$ seen along slit. In general, no continuum trace was found for these {H\,\sc ii}\ regions. The extraction trace was forced across the spectral range using the positions of the main nebular lines along the spectral domain. Finally, the individual spectra were combined. The integrated fluxes of the main nebular lines H$\beta$, {\rm \mbox{[O\,\sc ii]}}\,$\lambda$3727, {\rm \mbox{[O\,\sc iii]}}\,$\lambda\lambda$4959,\,5007, {\rm \mbox{[O\,\sc i]}}\,$\lambda$6300, {\rm \mbox{[N\,\sc ii]}}\,$\lambda$6584, H$\alpha$, and {\rm \mbox{[S\,\sc ii]}}\,$\lambda\lambda$6717,\,6731 were measured using a Gaussian fitting algorithm available with {\sc splot} in {\sc iraf}. A background continuum estimated from each side of the lines was automatically subtracted. Many {H\,\sc ii}\ regions in bars show a strong underlying Balmer absorption at H$\beta$. McCall {\rm et al.}\ (1985) have shown that adding about 2\,{\AA} of equivalent width for normal disc regions constitutes an appropriate correction. This correction was applied for all the H$\beta$\ fluxes for the {H\,\sc ii}\ regions in our control sample and the bar regions with shallow underlying absorption. However, for about 20\% of the bar sample, this correction was not sufficient. For these regions, mostly located in the bars with a strong continuum (e.g. NGC 3504, NGC 5921, NGC 7479), the amplitude of the underlying absorption was estimated from a Gaussian fit. In general, we found that about 5\,{\AA} of equivalent width were necessary to assure a good correction. This value was applied for all bar regions with absorption higher than normal. All of the line fluxes were corrected for interstellar reddening by comparing the H$\alpha$/H$\beta$\ ratio to the theoretical Balmer decrement (2.86) for Case B recombination (that is, for nebulae with large optical depths for the H\,{\sc i} resonance lines) at $10^4$\,K. In reality, the temperature of bar {H\,\sc ii}\ regions is probably around 7000--8000\,K due to their high O/H abundances (see below) so that our extinction based on the Balmer decrement might be overestimated by about 0.1\,mag. Since the Balmer ratio represents only an approximation of the real extinction, we did not take this difference into account. The reddening law formulated by Savage \& Mathis (1979) was assumed for reddening correction. It is generally difficult to evaluate the accuracy of the absolute spectrophotometric fluxes for individual objects. The main uncertainties include the contamination by the bright galactic continuum, the underlying absorption at H$\beta$, the accuracy of the photometric calibration, the spectrum extraction in a crowded field of {H\,\sc ii}\ regions, the positioning of the slit, the correction for the interstellar extinction and, of course, the intensity of the lines. We evaluate the accuracy of the fluxes to be at about 20--30\%. Since we will only compare the global behavior of {H\,\sc ii}\ regions, these high uncertainties should not influence our conclusions. \section{Nebular properties} \subsection{Excitation} The basic source of ionization in an {H\,\sc ii}\ region is the UV radiation field from young massive stars. The emission line spectra resulting from a pure photoionization field in a Str\"omgren sphere can be studied using diagnostic diagrams (Baldwin {\rm et al.}\ 1981; Evans \& Dopita 1985; Veilleux \& Osterbrock 1987; Osterbrock 1989; Dopita \& Sutherland 1995; Rola et al. 1997). These standard diagrams are based on nebular line ratios like {\rm \mbox{[O\,\sc iii]/{H$\beta$}}}, {\rm \mbox{[O\,\sc i]/{H$\alpha$}}}, {\rm \mbox{[N\,\sc ii]/H$\alpha$}}, and {\rm \mbox{[S\,\sc ii]/{H$\alpha$}}}. For high redshift galaxies, the red part of the spectrum is shifted to the near-infrared. If optical studies are needed, other diagnostic lines like {\rm \mbox{[O\,\sc ii]}}\,$\lambda$3727 or {\rm \mbox{[Ne\,\sc iii]}}\,$\lambda$3869 can also be relied upon (Rola {\rm et al.}\ 1997). These diagnostic diagrams are extremely useful in distinguishing between normal photoionized regions and regions with another ionization mechanism (e.g. high-velocity shocks, hard UV fields). The line ratios above are also independent of the reddening correction and depend only on the accuracy achieved for the spectrophotometry. However, despite the usefulness of these diagnostic diagrams for determining whether another ionization mechanism is present or not, it is very difficult to identify this mechanism. For this, sophisticated nebular models must be used (e.g. Stasi\'nska 1990; Dopita \& Sutherland 1995). \begin{figure}[t] \vskip -0.3truecm \centerline{ \psfig{figure=figure1.eps,width=8.8cm,clip=} } \vskip -0.5truecm \caption[]{Diagnostic diagram of our sample of bar and disc {H\,\sc ii}\ regions based on Osterbrock (1989). The sequence of normal regions is represented by the dotted lines. The full line indicates the separation between normal regions and regions with another ionization mechanism. The squared and circular symbols correspond to the average location of Seyfert~2 and LINER galaxies from Rola {\rm et al.}\ (1997). The arrow lines indicate the effect of high-velocity shocks and magnetic fields on the line ratios from models by Dopita \& Sutherland (1995). From the beginning to the end of the arrow, the velocity for the shock varies from 150 to 500\,km\,s$^{-1}$. The magnetic field parameter is zero for the bottom arrow and 4\,$\mu$G\,cm$^{3/2}$ for the top arrow} \end{figure} Figure~1 illustrates the first diagnostic diagram with a correlation between {\rm \mbox{[O\,\sc iii]/{H$\beta$}}}\ and {\rm \mbox{[N\,\sc ii]/H$\alpha$}}\ for the bar regions (dark symbols) and the disc regions of our control sample (open symbols). The dotted lines define the limits of the sequence of normal {H\,\sc ii}\ regions that can be found in Osterbrock (1989) and Kennicutt {\rm et al.}\ (1989). The full line is the separation between normal photoionized regions and regions with another ionization mechanism. We have also indicated the effect of high-velocity shocks and magnetic fields on these line ratios from models by Dopita \& Sutherland (1995). The average location of Seyfert 2 and LINER galaxies in this diagram are also displayed. At this time, it is still not entirely clear that the spectral characteristics (mainly the high {\rm \mbox{[N\,\sc ii]/H$\alpha$}}\ ratio) of these last objects are due to high UV radiation or high-velocity shocks (see discussion by Dopita \& Sutherland 1995). However, Fig. 1 shows that excepting for four regions, {\it all the {H\,\sc ii}\ regions in our sample are located inside the sequence of normal {H\,\sc ii}\ regions}. These four regions are located close to the nucleus in NGC\,3504 (a starburst galaxy) and NGC\,7479 (LINER). Thus, from this diagram alone, bar {H\,\sc ii}\ regions do not exhibit any sign of high-velocity shocks or hard-UV radiation. Apart from these central peculiarities, there is no dependence on radius. Although {\rm \mbox{[N\,\sc ii]/H$\alpha$}}\ is a good diagnostic ratio for high-velocity shocks or very hard ultraviolet radiation (Veilleux \& Osterbrock 1987; Dopita \& Sutherland 1995), other ratios like {\rm \mbox{[S\,\sc ii]/{H$\alpha$}}}\ or {\rm \mbox{[O\,\sc i]/{H$\alpha$}}}\ are more sensitive to these ionization mechanisms. Figure~2 illustrates another diagnostic diagram: the correlation between {\rm \mbox{[O\,\sc iii]/{H$\beta$}}}\ and {\rm \mbox{[S\,\sc ii]/{H$\alpha$}}}. The dashed and full lines are as in Fig.~1. Arrows also show the effect of the high-velocity shocks and magnetic fields. Although most of the {H\,\sc ii}\ regions fall within the normal region sequence, there is a small number of bar regions outside the sequence. Figure~3 shows the sequence between {\rm \mbox{[O\,\sc iii]/{H$\beta$}}}\ and {\rm \mbox{[O\,\sc i]/{H$\alpha$}}}. The latter ratio was detected in about 67\% of the sample of bar regions but only in about 40\% of the disc region sample. From these diagrams, it is clear that {\it most bar {H\,\sc ii}\ regions do not exhibit any systematic evidence of high-velocity shocks ($>$150\,km\,s$^{-1}$) or very hard UV radiation}. Only circumnuclear regions exhibit obvious signs of another ionization mechanism (fast shocks and/or hard-UV radiation), as expected from the work of Kennicutt et al. (1989). Nevertheless, if these conditions do not appear to be present in bar regions, Figs. 2 and 3 suggest that shocks with lower velocity or an abnormal UV photoionization field cannot be excluded. This possibility is discussed in Sect.~4. \begin{figure}[t] \vskip -0.3truecm \centerline{ \psfig{figure=figure2.eps,width=8.8cm,clip=} } \vskip -0.5truecm \caption[]{Diagnostic diagram comparing bar and disc {H\,\sc ii}\ regions. The dotted lines are the sequence of normal regions defined by Osterbrock (1989) and the full line separates the normal regions from regions which are not exclusively photoionized. The square and circular symbols correspond to the average locations of the Seyfert~2 and LINER galaxies from Rola {\rm et al.}\ (1997). The arrows are as in Fig.~1 except that the magnetic field for the top one is 2\,$\mu$G\,cm$^{3/2}$} \end{figure} \begin{figure}[h] \vskip -0.3truecm \centerline{ \psfig{figure=figure3.eps,width=8.8cm,clip=} } \vskip -0.5truecm \caption[]{Diagnostic diagram of bar and disc {H\,\sc ii}\ regions. The dotted lines indicate the sequence of normal {H\,\sc ii}\ regions. The full line is the separation between normal and abnormally excited regions (Osterbrock 1989). The square and circular symbols correspond to the average location of Seyfert~2 and LINER galaxies from Rola {\rm et al.}\ (1997). The arrows are as in Fig.~1} \end{figure} \subsection{Electronic density} The electronic density $N_e$ in {H\,\sc ii}\ regions can be accessed with the line ratio of the {\rm \mbox{[S\,\sc ii]}}\ doublet at 6717--6731\,{\AA}, i.e. $\gamma \!=\! {{\rm \mbox{[S\,\sc ii]}}\,\lambda 6717 \over {\rm \mbox{[S\,\sc ii]}}\,\lambda 6731}$ (Osterbrock 1989). The density can be derived from nebular models published by Blair \& Kirshner (1985). Since the calibration depends on the nebular temperature ($N_e \propto T_e^{0.5}$), we will assume $T_e \!=\! 10^4$\,K. This value is probably too high for the {H\,\sc ii}\ regions in our sample with abundances higher than the solar value (see Sect.~3.5). However, for comparison purposes, this approximation is appropriate. The electronic density distributions for the bar and disc {H\,\sc ii}\ regions are illustrated in Fig.~4. In both samples, several regions show line ratios that are very close or larger than the low-density limit ($\gamma \ga 1.4$). The electronic densities derived in this regime are very uncertain because the doublet ratio becomes only weakly dependent on $N_e$ for values larger than about 1.3. Kennicutt {\rm et al.}\ (1989) found that nuclei {H\,\sc ii}\ regions tend to possess higher electronic densities on average than disc regions, with both classes showing a large range of densities. As seen in Fig.~4, the case of {H\,\sc ii}\ regions located in bars is different. {\it No significant difference is observed between the distributions of electronic densities of both populations}. On average, $\gamma \approx 1.33 \pm 0.02$ (bar regions) and $\gamma \approx 1.31 \pm 0.03$ (disc regions). Bar {H\,\sc ii}\ regions do not show any compactness with respect to disc regions. \begin{figure}[t] \centerline{ \psfig{figure=figure4.eps,height=8.8cm,clip=} } \caption[]{Electronic density distributions for bar (top) and disc (bottom) {H\,\sc ii}\ regions. The density conversion derived from the {\rm \mbox{[S\,\sc ii]}}\ doublet ratio is from Blair \& Kirshner (1985). Vertical dotted lines indicate the mean values} \end{figure} \subsection{Extinction} As discussed in the Sect.~2.2, the visual interstellar extinction of individual {H\,\sc ii}\ regions can be derived from the H$\alpha$/H$\beta$\ line ratio. The values derived, however, are approximate since in reality, the real extinction is probably not distributed uniformly but is patchy. Figure~5 presents the distribution of the visual interstellar extinction from both the bar and disc {H\,\sc ii}\ regions. There is a considerable extinction in both populations of {H\,\sc ii}\ regions. The mean values are $A_V \approx 1.3$ and $A_V \approx 1.0$ visual magnitudes for the bar and disc regions, respectively. This difference is probably not significant since the extinction derived in bar regions, based on the assumption that the nebular temperature is $10^4$\,K, might be overestimated by about 0.1 to 0.2 magnitude. \begin{figure}[t] \centerline{ \psfig{figure=figure5.eps,height=8.8cm,clip=} } \caption[]{Distributions of the interstellar extinction $A_V$ derived from the H$\alpha$/H$\beta$\ line ratio for bar (top) and disc (bottom) {H\,\sc ii}\ regions. Vertical dotted lines indicate the mean values} \end{figure} The extinction values for the bar regions show a large dispersion. Most of the {H\,\sc ii}\ regions contributing to the highest values are located in the bars of NGC\,7479 and NGC\,5921. As noticed in MF97, these regions are located close to the strong dust lanes seen in these bars. Nevertheless, the overall behavior shows that the interstellar extinction as derived from the Balmer decrement is similar for bar and disc regions. However, using IRAS observations, Phillips (1993) has shown that for circumnuclear regions the extinction derived from the H$\alpha$/H$\beta$\ ratio can be underestimated by as much as 2 magnitudes. The ``uniform screen'' model assumed for the extinction is probably over-simplistic. Any line ratios (e.g. $R_{23}$) or other quantitative properties (e.g. integrated fluxes) severely affected by the interstellar extinction should be interpreted with caution for regions located in the inner parts of galaxies. \subsection{H$\alpha$\ equivalent widths} The equivalent widths (EW) of the Balmer emission-lines provide a measure between the number of ionizing and continuum photons emitted in the {H\,\sc ii}\ region. As such, the EWs depend strongly on the stage of evolution of the ionizing stars, the initial mass function (IMF), and the metallicity (Dottori 1981; McCall {\rm et al.}\ 1985; Copetti {\rm et al.}\ 1986; Bresolin \& Kennicutt 1997; Bresolin {\rm et al.}\ 1999; Leitherer {\rm et al.}\ 1999). As shown by Copetti {\rm et al.}\ (1986) and more recently by Leitherer {\rm et al.}\ (1999), EW(H$\alpha$) and EW(H$\beta$) can both be used as age indicators for {H\,\sc ii}\ regions. In disc galaxies, the distribution of EW(H$\alpha$) extends from about 100\,{\AA} to 1500\,{\AA} with a median value around 400\,{\AA}. No obvious correlation with the Hubble type is found (Bresolin \& Kennicutt 1997). Assuming an instantaneous burst of star formation and a solar metallicity, these values correspond to an age range between 1\,Myr to 7\,Myr (Leitherer {\rm et al.}\ 1999). The accuracy of the Balmer line EWs is mostly determined by the uncertainty in the level of the nebular continuum which is severely contaminated by the galactic continuum. Because we could directly measure the contribution from the galactic continuum on our ``off'' band images used in MF97, we have only measured the EWs for the H$\alpha$\ line. The EW(H$\alpha$) is also less affected by the interstellar extinction and the underlying absorption. The fraction of the galaxy light contributing to the nebular continuum was estimated from two photometric apertures: one covering the integrated light of the {H\,\sc ii}\ regions, and the other located on nearest area devoid of any H$\alpha$\ emission (determined from the H$\alpha$\ images). A correction factor was then applied to the EW values measured directly from the spectra. These correction factors vary from 1.1 to 20 depending on the location of the {H\,\sc ii}\ region. \begin{figure}[t] \centerline{ \psfig{figure=figure6.eps,height=8.8cm,clip=} } \caption[]{Distribution of the H$\alpha$\ equivalent widths EW(H$\alpha$) for the bar (top) and disc (bottom) {H\,\sc ii}\ regions. The scale in the bottom panel indicates the approximate age of the {H\,\sc ii}\ regions from models with a Salpeter IMF, instantaneous burst of star formation and solar metallicity (see Leitherer {\rm et al.}\ 1999 for details). Vertical dotted lines indicate the mean values} \end{figure} Figure~6 illustrates the EW(H$\alpha$) distributions of the bar and disc {H\,\sc ii}\ regions. The mean values for the distributions differ by about a factor of two: EW(H$\alpha$) $\approx$ 250\,{\AA} (bar regions) and EW(H$\alpha$) $\approx$ 560\,{\AA} (disc regions). Following the models of Leitherer {\rm et al.}\ (1999) for an instantaneous burst of star formation with a Miller-Scalo mass function and solar metallicity, the mean age of bar regions is about 5.3\,Myr while disc regions are about 4.0\,Myr old. However, the difference in age is less ($<1 \times 10^6$\,yr) when a Salpeter function is used to describe the IMF. Also, no age gradient seems to exist along the sequence of bar HII regions. This is an indication that HII regions should be ignified all along the bar and not only at bar ends with a subsequent migration towards the center. In their study of nuclear {H\,\sc ii}\ regions, Kennicutt {\rm et al.}\ (1989) found that the EW(H$\alpha$) of the {H\,\sc ii}\ region nuclei, with a median value around 25\,{\AA}, are approximately 20 times lower on average than that of normal disc regions. Such a difference cannot easily be explained by assuming that the correction for galactic continuum was underestimated. The authors rather favor the idea that the stellar continuum is high due to continuous star formation in the same region or an unusual stellar mass spectrum in the ionizing clusters. In the present case, however, it is difficult to completely discard the effect of the contamination of the galactic continuum to the nebular continuum to explain the discrepancy observed between bar and disc regions. The location of the aperture used to measure the galaxy continuum has a strong effect on the correction performed. The light distribution in bars is not uniform and some bars have very patchy dust features. A systematic error of a factor 2 cannot be ruled out. In any case, no firm conclusion based on the difference observed in the EW({H$\alpha$}) of both populations of {H\,\sc ii}\ regions can be drawn from the actual sample. \subsection{O/H abundance (within bars)} The oxygen abundance in {H\,\sc ii}\ regions can be derived either through semi-empirical calibrations (e.g. Edmunds \& Pagel 1984; McGaugh 1991; Pagel 1997) or directly when the temperature can be measured from the nebular lines {\rm \mbox{[O\,\sc i]}}\,$\lambda$4363 or {\rm \mbox{[N\,\sc ii]}}\,$\lambda$5755. The latter, however, are generally detectable only for {H\,\sc ii}\ regions with low oxygen abundance. In our case, almost all the regions have solar or above-solar oxygen abundances; semi-empirical techniques have to be used. Even if the uncertainties related to these methods are generally quite large ($\pm$0.2\,dex), it is worthwhile to derive the O/H values to address the important question of mixing of the ISM in bars. It is now well established that bars induce large-scale mixing of the chemical composition in the disc of spirals (Martin \& Roy 1994; Zaritsky {\rm et al.}\ 1994; Friedli {\rm et al.}\ 1994). The radial flows of gas formed by bars flatten the strong (negative) abundance gradients generally observed in unbarred late-type disc galaxies. The importance of the homogenization effect is related to the bar strength as shown by Martin \& Roy (1994). Very strong gas flows ($v \ga 100$\,km\,s$^{-1}$) are taking place along bars; efficient mixing should be also observed and the O/H scatter between the bar {H\,\sc ii}\ regions should be smaller than what is observed in normal galaxy discs (See Sect.~4). \begin{figure}[t] \centerline{ \psfig{figure=figure7.eps,height=9.6cm,clip=} } \caption[]{Comparison between the O/H abundances values given by three different line ratios: {\rm \mbox{[N\,\sc ii]}}/{{\rm \mbox{[O\,\sc iii]}}}, {\rm \mbox{[O\,\sc iii]/{H$\beta$}}}, and $R_{23}$} \end{figure} The oxygen abundances for our sample of bar {H\,\sc ii}\ regions were determined using three line ratios: {\rm \mbox{[N\,\sc ii]}}/{{\rm \mbox{[O\,\sc iii]}}}, {\rm \mbox{[O\,\sc iii]/{H$\beta$}}}, and $R_{23} \!=\! {{\rm \mbox{[O\,\sc ii]}}\,\lambda 3727 + {\rm \mbox{[O\,\sc iii]}}\,\lambda\lambda 4959,\,5007 \over \rm H\beta}$. The conversion to relative oxygen abundances was done using the calibration of Edmunds \& Pagel (1984). Much has been written on the accuracy of these line ratios as abundance indicators (e.g. McGaugh 1991; Martin \& Roy 1995; Stasi\'nska 1998). For our sample, Fig.~7 compares the different O/H values derived with all three indicators. The O/H values derived from {\rm \mbox{[N\,\sc ii]}}/{{\rm \mbox{[O\,\sc iii]}}}\ are slightly higher ($\sim$0.1\,dex) than the values derived from {\rm \mbox{[O\,\sc iii]/{H$\beta$}}}\ and $R_{23}$ for $12+\log(\rm O/H) < 8.9$. For $12 + \log(\rm O/H) > 8.9$, that is, the abundance regime for most of the bars in our sample, the abundances from {\rm \mbox{[O\,\sc iii]/{H$\beta$}}}\ and {\rm \mbox{[N\,\sc ii]}}/{{\rm \mbox{[O\,\sc iii]}}}\ are slightly below the values given by $R_{23}$. These results were previously discussed by Martin \& Roy (1994) and are due to discrepancies in the semi-empirical calibrations. For our purposes, we use the O/H values derived from the {\rm \mbox{[N\,\sc ii]}}/{{\rm \mbox{[O\,\sc iii]}}}\ indicator; our conclusions are not affected by this choice. The distribution of the oxygen abundance along the bars of our sample is illustrated in Fig.~8. It is clear that {\it the O/H scatter observed in these bars is well smaller than $\pm$0.1\,dex or even less}. Martin \& Belley (1996, 1997) have shown that the azimuthal O/H dispersion observed in the discs of normal and barred galaxies is generally equal or larger than $\pm$0.2\,dex. These abundance variations are probably not intrinsic but are in fact a combination of real inhomogeneities and uncertainties associated with empirical techniques (Roy \& Kunth 1995). However, a comparative analysis remains possible when the same method is used to derive the O/H values (Martin \& Belley 1997). Clearly, the chemical composition of {H\,\sc ii}\ regions in a bar is well-homogenized. This indicates that an efficient mixing of the chemical composition is taking place in the bar region (see next section). \begin{figure*} \vskip -2truecm \centerline{ \psfig{figure=figure8.eps,width=16cm,clip=} } \vskip -1.5truecm \caption[]{O/H abundance distributions along the bars of the galaxies in our sample. The O/H values were derived using the {\rm \mbox{[N\,\sc ii]}}/{{\rm \mbox{[O\,\sc iii]}}}\ line ratio. The horizontal axis is the position of the regions normalized to the bar radius given in MF97. The open and black symbols differentiate the regions with regard to what side of the bar they are located. The abundance interval is 0.06\,dex for galaxies displayed in the left column and 0.2\,dex for those in the right column} \end{figure*} \section{Discussion} {\it Degree of ionization.} As discussed in Sect.~3.1, bar {H\,\sc ii}\ region spectra do not exhibit any obvious signs of high-velocity shocks or hard UV radiation. However, there is marginal evidence from Figs.~2 and 3 that the ionization might be different for some bar regions. The degree of ionization at a specific position in a nebula can be accessed through the ionization parameter: \begin{equation} U = {Q(H^0) \over 4 \pi r^2 c N_e} \end{equation} where $Q(H^{0})$ is the number of ionizing photons per unit time by the central source, $r$ is the position in the nebula, and $c$ the speed of light (Osterbrock 1989). The most obvious evidence that bar regions are different from disc regions is seen in the {\rm \mbox{[O\,\sc i]}}\,$\lambda$6300 line. As described in Evans \& Dopita (1985), the {\rm \mbox{[O\,\sc i]}}\ line is emitted in the transition zone of an {H\,\sc ii}\ region which contains a significant fraction of neutral hydrogen. The line is then stronger when $U$ is lower (or the ionizing stellar temperature is lower). Since we have detected the {\rm \mbox{[O\,\sc i]}}\ line in a much larger fraction of bar regions than in disc regions, this suggests that the ionization parameter could be lower in the former population. Figure~9 shows the correlation between {\rm \mbox{[O\,\sc i]}}\,$\lambda$6300/{\rm \mbox{[O\,\sc iii]}}\,$\lambda$5007 and {\rm \mbox{[O\,\sc ii]}}\,$\lambda$3727/{\rm \mbox{[O\,\sc iii]}}\,$\lambda$5007 which is particularly dependent on $U$ (Evans \& Dopita 1985). A correction for the interstellar extinction has been applied to these line ratios. The bulk of bar {H\,\sc ii}\ regions is located at $U$$\sim$0.0005. For the disc regions, the scatter is quite large but on average $U$$\sim$0.001, larger than the value for bar regions. Unfortunately, our sample is not large enough to firmly confirm that $U$ is indeed different for both populations of {H\,\sc ii}\ regions. \begin{figure}[t] \centerline{ \psfig{figure=figure9.eps,height=9.6cm,clip=} } \caption[]{Diagnostic diagram showing the relation between the {\rm \mbox{[O\,\sc i]}}\,$\lambda$6300/{\rm \mbox{[O\,\sc iii]}}\,$\lambda$5007 and {\rm \mbox{[O\,\sc ii]}}\,$\lambda$3727/{\rm \mbox{[O\,\sc iii]}}\,$\lambda$5007 nebular line ratios. Both ratios have been corrected for the extinction. The curves represent models from Evans \& Dopita (1985). The horizontal dashed lines show three values for $U$, 0.03, 0.003, and 0.0003. The vertical lines indicate the stellar temperatures used in the models, 50\,000\,K, 40\,000\,K and 37\,000\,K.} \end{figure} If a difference in the ionization parameter is really present, this could be due to many factors: differences in the initial mass function, age, richness of the OB associations, or spatial distribution of the ionized material (Evans \& Dopita 1985). As recently shown by Rozas et al. (1999), the luminosity function (LF) of the bar regions is much less regular than the LF of the disc regions in the strongly barred spiral NGC\,7479. Their result, combined with our study on the nebular excitation, suggest strongly that the properties of the OB associations formed in bars differ from the normal associations of the disc. More work comparing LFs and nebular properties of a larger sample of bar and disc regions would allow us to investigate the origin of this difference. In the final paper in this series (Friedli \& Martin, in preparation), we will also examine the properties of the clusters formed in diverse bar environments with high-spatial numerical simulations. \medskip \noindent {\it Mixing and element production.} The estimated timescale given by numerical simulations for which the star formation activity phase lasts in bars is $\tau_{\rm SF} \sim 5 \times 10^8$\,yr (Martin \& Friedli 1997). How does this compare with mixing timescale? It is possible to roughly quantitatively evaluate the timescale of mixing of the ISM due to radial flows. Roy \& Kunth (1995) have discussed the diverse mixing mechanisms of the oxygen abundance in the ISM in galaxy discs. Assuming a pure radial mixing due to gas flows funnelled in the bar, the upper limit for the time for gas to diffuse a length scale, $\Delta x_{\rm rad}$, in the radial direction is: \begin{equation} \tau_{\rm rad}={\Delta x_{\rm rad}^2 \over v l} \, , \end{equation} where $v$ is the radial flow velocity, and $l$ is the mean free path for molecular clouds. In a typical bar, $\Delta x_{\rm rad} \!=\! 5$\,kpc (see Table 1) and $v\!=\!100$\,km\,s$^{-1}$. The value of the mean free path for the gas clouds is not a well-defined quantity in bars. In galaxy discs, $l\!=\!300-1000$\,pc (Roberts \& Hausman 1984; Roy \& Kunth 1995). If we assume $l\!=\!500$\,pc for bars, we find $\tau_{\rm rad} \approx 5 \times 10^8$\,yr. However, since radial flows in bars are not stationary, the real mixing timescale could be even of the order of $\tau_{\rm rad} \!=\! \Delta x_{\rm rad}/v$, i.e. $\sim 5 \times 10^7$\,yr. Putting all this together yields the following reasonable interval for the mixing timescale: $5 \times 10^7 \la \tau_{\rm rad} \la 5 \times 10^8$\,yr. Thus, $\tau_{\rm rad}$ is shorter than $\tau_{\rm SF}$ meaning that the abundance content in {H\,\sc ii}\ regions formed during this phase must be homogenized. It is also instructive to make rough (i.e. close-box) estimates of the global abundance increase during $\tau_{\rm SF}$ as well as of the abundance fluctuations in {H\,\sc ii}\ regions which should result from their age spread. If $Z^i$ is the initial mean gaseous abundance in the bar region, $\epsilon$ the global star formation efficiency, and $Y_Z$ the net yield for the species considered, then the final mean abundance is given by: \begin{equation} Z^f = Z^i + \epsilon (1-\epsilon)^{-1} Y_Z \, . \end{equation} Interestingly enough, $Z^f$ does not depend on the initial gas mass fraction. For instance, for the oxygen with an yield $Y_O \approx 0.006$, an initial solar abundance $Z^i \approx 0.01$, and a typical SF efficiency $\epsilon \approx 0.25$, then $Z^f \approx 0.012$. The global increase of oxygen abundance is thus only about 0.1\,dex. Equation~3 can in fact also be applied to each individual {H\,\sc ii}\ regions with exactly the same numbers; the fluctuations in the oxygen abundance are thus expected to be of the order 0.1\,dex, which is indeed what is observed (Sect.~3.5). However, we do not observe any clear trend between the age and metallicity for bar {H\,\sc ii}\ regions. In dwarf galaxies, the metal enrichment of {H\,\sc ii}\ is not observed and metals are probably locked in the hot phase. (see e.g. Tenorio-Tagle 1996; Kobulnicky 1998). The situation could be similar for star forming regions in bars. \section{Summary} The main results concerning the properties of {H\,\sc ii}\ regions located within bars can be expressed as follows: \smallskip \noindent {\it 1) From standard diagnostic diagrams, the excitation of most {H\,\sc ii}\ regions appears normal and similar to the one of disc regions.} There are some exceptions for nuclear regions where an ionization mechanism other than photoionization seems to be present. However, there is marginal evidence that the ionization parameter in bar regions is lower than in disc regions, suggesting that the properties of the OB associations might be different in bar environments. \smallskip \noindent {\it 2) The electronic density distribution as derived from the {\rm \mbox{[S\,\sc ii]}}\ line ratio is similar to that observed for normal disc regions.} The mean density is $N_e \approx 80$\,cm$^{-3}$. Star formation regions in bars have the same ``compactness'' as disc regions. \smallskip \noindent {\it 3) The H$\alpha$/H$\beta$\ extinction indicator reveals that, on average, bar regions have a visual extinction $A_V \sim 1.3$\,mag, 0.3\,mag more than disc regions.} This difference is mainly due to the fact that some regions are located near the bar dust lanes of the earlier types of galaxies in our sample. \smallskip \noindent {\it 4) The average H$\alpha$\ equivalent width for the bar regions is about 250\,\AA, half that of disc regions.} While this could indicate an older population for bar {H\,\sc ii}\ regions, the corresponding age difference is probably too small to be significant owing to the large uncertainties introduced by the galactic continuum correction. \smallskip \noindent {\it 5) The O/H abundance distribution of these {H\,\sc ii}\ regions is remarkably homogeneous.} This is the result of the gaseous radial flows in bars inducing mixing of the ISM, as seen on a larger scale in the discs of barred spirals. \begin{acknowledgements} Discussions with P.~Ferruit, L. ~Binette, R.~Kennicutt, and J.-R.~Roy were most appreciated throughout this work. We also thank the referee, Fran\c{c}oise Combes, for her helpful comments. The efficient support offered by the technical staff of the Multiple Mirror Telescope during the observations was much appreciated. This work was supported by NSERC (Canada), FCAR (Qu\'ebec) and in part by the NSF through grant AST-94-21145. D.F. acknowledges the kind hospitality of CFHT. \end{acknowledgements}
\section*{Introduction} The problem of the crossover from BCS superconducting state to a Bose-Einstein condensate (BEC) of local pairs \cite{1,2,3} becomed very important in the context of high temperature superconductors (HTSC). While at the present time there is no quantitative microscopic theory for the occurrence of the superconducting state in the doped antiferromagnetic materials, it is generally accepted that the superconducting state can be described in therms of a pairing picture. The short coherence length ($\xi\sim$ 10-20 A) increased the interest for the problem \cite{4,5,6,7,8,9,10,11,12,13,14} because it showed that the BCS equations of highly overlapping pairs, or the description in terms of composite bosons cannot describe the whole regime between weak and strong coupling. The mean field method developed by different authors \cite{3,4,6,14} and solved analitically in two and three dimension, and the Ginzburg-Landau description \cite{7,8,13} showed that the evolution between the two limits is continous, no singularities during this evolution appearing. The zero temperature coherence length in the framework of field-theoretical method has been in Ref. \onlinecite{12}. The problem of the BCS-BEC crossover in arbitrary dimension $d$ using the field-theoretical method has been extesively discussed in Refs. \onlinecite{15,16,17,18}, where the chemical potential, the number of condensed pairs and the repulsive interaction between pairs have been calculating using the analogy with the field-theoretical description of superfluidity. In this paper we apply this method to study the crossover problem for a non-Fermi superconductor described by the Anderson model \cite{19,20} (See also Refs. \onlinecite{20,21,22,23,24,25,26,27,28}), to study the crossover between weak coupling and strong coupling. The paper is organized as follows. In Section II we study the weak coupling model for a $d=2$ non-Fermi superconductor. Section III contains the strong coupling limit. To make the paper self-contained we present in Appendix the Lagrangian formalism for the superfluid phase following Refs. \onlinecite{15,16}. The results are discussed in Section IV. \section*{Weak coupling limit} The BCS-like model for the non-Fermi system is described by the Lagrangian \begin{equation} {\cal L}=\psi^\dagger_{\ua} G_0^{-1}\psi_{\ua} + \psi^\dagger_{\da} (G_o^{-1})^*\psi_{\da} - \l_0\psi^\dagger_{\ua}\psi^\dagger_{\da}\psi_{\da}\psi_{\ua} \label{e1} \end{equation} where the normal state is described by the Green function \begin{equation} G_0(\bp,\omega} \def\G{\Gamma} \def\D{\Delta)=\frac{g(\a)}{\omega} \def\G{\Gamma} \def\D{\Delta_c^\a(\omega} \def\G{\Gamma} \def\D{\Delta-\xi(\bp)+i\d)^{1-\a}} \label{e2} \end{equation} where $\omega} \def\G{\Gamma} \def\D{\Delta_c\leq\omega} \def\G{\Gamma} \def\D{\Delta\leq\omega} \def\G{\Gamma} \def\D{\Delta_c$, $g(\a)=\pi\a/(2\sin{(\pi\a/2)})$ and $\l_0<0$ is the coupling constant describing the attraction between electrons. The Green function given by Eq. (\ref{e2}) contain a cut-off $\omega} \def\G{\Gamma} \def\D{\Delta_c$ and the exponent $\a$. This form has been proposed first by Anderson \cite{19} to describe the 2D non-Fermi properties of the superconducting state. If we introduce the two-component fermionic field \begin{equation} \Psi=\left(\begin{array}{c} \psi_\ua\\ \psi^\dagger_\da \end{array}\right) \hspace{2cm} \Psi^\dagger=\left(\begin{array}{cc} \psi^\dagger_\ua & \psi_\da \end{array}\right) \label{e4} \end{equation} the non-interacting part of the Lagrangian (\ref{e1}) is \begin{equation} {\cal L}_0=\Psi^\dagger\left(\begin{array}{cc}G_0^{-1} & 0 \\ 0 & (G_0^{-1})^*\end{array}\right) \Psi \label{e5} \end{equation} In order to calculate the partition function \begin{equation} Z=\int D\Psi^\dagger D\Psi \exp{\left[i\int_x {\cal L}\right]} \label{e6} \end{equation} we will transform the interaction contribution from the Lagrangian (\ref{e1}) as \begin{equation} \exp{\left[-i\l_0\int_x \psi_\ua^\dagger \psi_\da^\dagger \psi_\da\psi_\ua\right]}= \int D\D^\dagger D\D \exp{\left[-i\int_x \left(\D^\dagger\psi_\da\psi_\ua+\psi_\ua^\dagger\psi_\da^\dagger\D- \frac{1}{\l_0}\D^\dagger\D\right)\right]} \label{e7} \end{equation} where $\int_x=\int dt\int d^dx$ as in Ref. \onlinecite{16,17,18}, and $\D=\l_0\psi_\da\psi_\ua$ is a bosonic field. The partition function defined by Eq. (\ref{e6}) will be expressed using Eq. (\ref{e7}) in a bilinear form as \begin{equation} Z=\int D\Psi^\dagger D\Psi\int D\D^\dagger D\D \exp{\left(\frac{i}{\l_0}\int_x \D^\dagger\D\right)} \exp{\left[i\int_x\Psi^\dagger\left(\begin{array}{cc} G_0^{-1} & -\D \\ -\D^\dagger & (G_0^{-1})^*\end{array}\right)\Psi\right]} \label{e8} \end{equation} Performing the integral over the Grassmann fields the partition function becomes \begin{equation} Z=\int D\D^\dagger D\D \exp{\left(i S_{eff}[\D^\dagger,\D]+\frac{1}{\l_0}\int_x \D^\dagger \D\right)} \label{e9} \end{equation} where $S_{eff}[\D^\dagger,\D]$ is the one loop effective action, which can be written as \begin{equation} S_{eff}[\D^\dagger,\D]=-i Tr \ln{\left(\begin{array}{cc} f(\a) (p_0-\xi(\bp))^{1-\a} & -\D\\ -\D^\dagger & f(\a) (p_0+\xi(\bp))^{1-\a}\end{array}\right)} \label{e10} \end{equation} where $f(\a)=\omega} \def\G{\Gamma} \def\D{\Delta_c^\a g^{-1}(\a)$ and the trace $Tr$ has been used according to the meaning from Ref. \onlinecite{16}. In the mean field approximation the integral from Eq. (\ref{e9}) can be performed using the solution given by the saddle point and for $T\neq 0$ the critical temperature $T_c$ will be obtained as \begin{equation} T_c(\a)=\omega} \def\G{\Gamma} \def\D{\Delta_D\left[\frac{D(\a)}{C(\a)}\right]^{1/2\a}\left[1-\frac{1}{A(\a)D(\a)}\frac{1}{|\l_0|} \left(\frac{\omega} \def\G{\Gamma} \def\D{\Delta_c}{\omega} \def\G{\Gamma} \def\D{\Delta_D}\right)^{2\a}\right]^{1/2\a} \label{e11} \end{equation} where $A(\a)=g^2(\a) 2^{2\a} \sin{(\pi(1-\a))}/\pi$, $C(\a)=\G^2(\a) [1-2^{1-2\a}] \zeta(\a)$, $D(\a)=\G(1-2\a) \G(\a)/(2\a \G(1-\a))$, $\G(x)$ being the Euler's gamma function. This expression is valid only in the limit $0<\a<0.5$ and a positive critical temperature implies for the coupling constant the condition $|\l_0|>\l_c$, with $\l_c=(\omega} \def\G{\Gamma} \def\D{\Delta_c/\omega} \def\G{\Gamma} \def\D{\Delta_D)^\a/(A(\a)D(\a))$. We have to mention that the critical temperature obtained in Eq. (\ref{e11}), calculated also in Ref. \onlinecite{29} is different from the one obtained in Ref. \onlinecite{22,24,25}, and it is easy to show that it gives the exact BCS result in the limit $\a\rightarrow} \def\la{\leftarrow 0$. If we consider the effective action as \begin{eqnarray} S_{eff}[\D^\dagger,\D]&=&-i Tr\ln{\left(\begin{array}{cc} f(\a)(p_0-\xi(\bp))^{1-\a} & 0\\ 0 & f(\a)(p_0+\xi(\bp))^{1-\a}\end{array}\right)}\nonumber\\ &-&i Tr \ln{\left[1-\frac{|\bar{\D}|^2}{f(\a)(p_0^2-\xi^2(\bp))^{1-\a}}\right]} \label{e13} \end{eqnarray} and the system as space time independent the partition function can be written as \begin{equation} Z=Z_0\exp{\left[\frac{i}{\l_0}\bar{\D}^\dagger\bar{\D}\right]} \label{e14} \end{equation} $Z_0$ containing the non-interacting contribution, and we get for the renormalized coupling constant $\l$ the expression \begin{equation} \frac{1}{\l}=\frac{1}{\l_0}+\frac{i}{f^2(\a)}\int\frac{d^2\bp}{(2\pi)^2}\int\frac{dp_0}{2\pi} \frac{1}{(p_0^2-\xi^2(\bp))^{1-\a}} \label{e15} \end{equation} Using the integral \begin{displaymath} \int_{k_0}\frac{1}{(k_0^2-E^2+i\eta)^l}=i(-1)^l\sqrt{\pi}\frac{\G(l-1/2)}{\G(l)}\frac{1}{E^{2l-1}} \end{displaymath} we calculated $\l$ as \begin{equation} \frac{1}{\l}=\frac{1}{\l_0}+\frac{1}{\l_1} \label{e17} \end{equation} where \begin{equation} \l_1=-\frac{4\pi\a g^{-2}(\a)}{\cos{(\pi(\a-1))}}\frac{1}{B(1/2,1/2-\a)} \left(\frac{\omega} \def\G{\Gamma} \def\D{\Delta_c}{\omega} \def\G{\Gamma} \def\D{\Delta_D}\right)^{2\a} \label{e18} \end{equation} $B(x,y)$ being the Euler beta function $B(x,y)=\G(x)\G(y)/\G(x+y)$. The expression given by Eq. (\ref{e18}) is positive for $\a<1/2$. The new coupling constant $\l$ has to be also negative in order to have superconductivity ($\l<0$) and this condition is satisfied if $|\l_0|<\l_1$. If we consider also the condition $\l_c<|\l_0|$ we get the general condition for the bare coupling constant $\l_0$, $\l_c<|\l_0|<\l_1$, which is satisfied for $0<\a<0.5$. We mention that for the weak coupling limit $\l_0\rightarrow} \def\la{\leftarrow 0^-$ the BCS limit studied in Ref. \onlinecite{24} is reobtained, but we also showed that the critical constant calculated from the critical temperature is smaller than $\l_1 (\a)$, which also satisfies condition $\l_1(\a\ra0)=0$. In the limit $\l_0\rightarrow} \def\la{\leftarrow -\infty$, called the strong coupling limit, we expect an important effect of the non-Fermi character of the electrons in the coupling constant. \section*{Strong coupling limit} In this limit we consider $\D(x)=\bar{\D}+\tilde{\D}(x)$ and consider the action $S_{eff}[\tilde{\D}^\dagger,\tilde{\D}]$ obtained from Eq. (\ref{e10}) as \begin{equation} S_{eff}[\tilde{\D}^\dagger,\tilde{\D}]=-i Tr \ln{\left[1+\hat{G}_0\hat{\tilde{\D}}\right]} \label{e23} \end{equation} where \begin{equation} \hat{G}_0^{-1}=\left(\begin{array}{cc} f(\a)(p_0-\xi(\bp))^{1-\a} & -\bar{\D}\\ -\bar{\D}^\dagger & f(\a)(p-0+\xi(\bp))^{1-\a} \end{array}\right) \label{e24} \end{equation} \begin{equation} \hat{\tilde{\D}}= \left(\begin{array}{cc}0 & -\tilde{\D}\\ -\tilde{\D}^\dagger & 0 \end{array}\right) \label{e25} \end{equation} which can be written as \begin{equation} S_{eff}[\tilde{\D}^\dagger,\tilde{\D}]=-i Tr \sum_{l=1}^\infty \frac{1}{l} \left[\hat{G}_0\hat{\tilde{\D}}\right]^l \label{e26} \end{equation} with \begin{equation} \hat{G}_0(p_0,\bp)=\frac{1}{f^2(\a)(p_0^2-\xi^2(\bp))^{1-\a}-|\bar{\D}|^2} \left(\begin{array}{cc}0 & -\tilde{\D}\\ -\tilde{\D}^\dagger & 0 \end{array}\right) \label{e27} \end{equation} We are interested in quadratic terms in $\tilde{\D}$ and we will take the approximation \begin{equation} S_{eff}[\tilde{\D}^\dagger,\tilde{\D}]=S_{eff}^{(2)}(0)+S_{eff}^{(2)}(\bq) \label{e28} \end{equation} which contains the quadratic contributions. The first term in Eq. (\ref{e28}) has the form \begin{eqnarray} S_{eff}^{(2)}(0)&=&\frac{1}{2}i Tr \frac{1}{f^2(\a)(p_0^2-\xi^2(\bp))^{1-\a}-|\bar{\D}|^2} \left(\bar{\D}^2\tilde{\D}^\dagger\tilde{\D}^\dagger+\bar{\D}^{\dagger^2}\tilde{\D}\tilde{\D} +2|\tilde{\D}|^2|\bar{\D}|^2\right)\\ &+&\frac{1}{2}i Tr \frac{1}{f^2(\a)(p_0^2-\xi^2(\bp))^{1-\a}-|\bar{\D}|^2} 2|\tilde{\D}|^2 \label{e29} \end{eqnarray} which will be approximated, taking in the dominator $\bar{\D}\approx 0$ as \begin{equation} S_{eff}^{(2)}(0)\cong \frac{1}{2} i Tr \frac{1}{f^4(\a)(p_0^2-\xi^2(\bp))^{2(1-\a)}} \left[\bar{\D}^2\tilde{\D}^\dagger\tilde{\D}^\dagger+ \bar{\D}^{\dagger^2}\tilde{\D}\tilde{\D}+2|\tilde{\D}|^2|\bar{\D}|^2\right] \label{e30} \end{equation} the last term giving no contribution to the renormalized coupling constant. Following the same approximation we calculated $S_{eff}^{(2)}(\bq)$ as \begin{eqnarray} S_{eff}^{(2)}(\bq)&=&\frac{1}{2}i Tr \frac{1}{f^2(\a)(p_0-\xi(\bp))^{1-a}(p_0+q_0+\xi(\bp+\bq))^{1-\a}} \tilde{\D}\tilde{\D}^\dagger\nonumber\\ &+& \frac{1}{2}i Tr \frac{1}{f^2(\a)(p_0+\xi(\bp))^{1-\a}(p_0+q_0-\xi(\bp+\bq))^{1-\a}} \label{e31} \end{eqnarray} From Eqs. (\ref{e30}) and (\ref{e31}) we have \begin{eqnarray} {\cal L}^{(2)}(0)&=&-\frac{B(1/2,3/2-2\a)}{4\pi f^4(\a)}(2m)^{3-4\a}\nonumber\\ &\times&\int\frac{d^2\bp}{(2\pi)^2}\frac{1}{(p^2+m\ve_a)^{3-4\a}} \left[\bar{\D}^2\tilde{\D}^\dagger\tilde{\D}^\dagger+ \bar{\D}^{\dagger^2}\tilde{\D}\tilde{\D}+2|\tilde{\D}|^2|\bar{\D}|^2\right] \label{e32} \end{eqnarray} and \begin{eqnarray} {\cal L}^{(2)}(\bq)&=&-\frac{\sin{(\pi(1-\a))B(\a,\a)}}{4\pi f^2(\a)}m^{1-2\a} \int\frac{d^2\bp}{(2\pi)^2}\frac{1}{(p^2+m\ve_a+q_0m+q^2/4)^{1-2\a}}\nonumber\\ &+&-\frac{\sin{(\pi(1-\a))B(\a,\a)}}{4\pi f^2(\a)}m^{1-2\a} \int\frac{d^2\bp}{(2\pi)^2}\frac{1}{(p^2+m\ve_a-q_0m+q^2/4)^{1-2\a}} \label{e33} \end{eqnarray} The integrals from Eqs. (\ref{e32}) and (\ref{e33}) can be performed using the formula \begin{displaymath} \int_\bp\frac{1}{(p^2+A^2)^N}=\frac{\G(N-d/2)}{(4\pi)^{d/2}\G(N)}\frac{1}{(A^2)^{N-d/2}} \end{displaymath} and we obtain \begin{eqnarray} {\cal L}^{(2)}=&-&\frac{m}{16\pi^2f^2(\a)}\frac{2^{2-4\a}}{1-2\a}\frac{B(1/2,3/2-2\a)}{\ve_a^{2-4\a}} \left[\bar{\D}^2\tilde{\D}^\dagger\tilde{\D}^\dagger+ \bar{\D}^{\dagger^2}\tilde{\D}\tilde{\D}+2|\tilde{\D}|^2|\bar{\D}|^2\right]\nonumber\\ &+&\frac{m}{16\pi^2f^2(\a)}\frac{\sin{(\pi(\a-1))}}{2\a}\frac{B(\a,\a)}{(\ve_a+q_0+q^2/4m)^{-2\a}} \tilde{\D}\tilde{\D}^\dagger\nonumber\\ &+&\frac{m}{16\pi^2f^2(\a)}\frac{\sin{(\pi(\a-1))}}{2\a}\frac{B(\a,\a)}{(\ve_a-q_0+q^2/4m)^{-2\a}} \tilde{\D}^\dagger\tilde{\D}\nonumber\\ \label{e34} \end{eqnarray} Using the approximation \begin{displaymath} \left(\ve_a\pm q_0+\frac{q^2}{4m}\right)^{2\a}\cong \ve_a^{2\a}+2\a\ve_a^{2\a-1} \left(\pm q_0+\frac{q^2}{4m}\right) \end{displaymath} and using the notation \begin{equation} \tilde{\Psi}=\left(\begin{array}{c} \tilde{\D} \\ \tilde{\D}^\dagger \end{array}\right) \label{e36} \end{equation} we obtain from Eq. (\ref{e34}) \begin{equation} {\cal L}^{(2)}=\frac{m}{16\pi^2f^2(\a)}\sin{(\pi(1-\a))}B(\a,\a)\ve_a^{2\a-1}\frac{1}{2} \tilde{\Psi}^\dagger M \tilde{\Psi} \label{e37} \end{equation} where \begin{equation} M=\left(\begin{array}{cc} q_0-\frac{q^2}{2m_b}-\mu_0 & -\mu_0\\ -\mu_0 & -q_0-\frac{q^2}{2m_b}-\mu_0 \end{array}\right) \label{e38} \end{equation} $m_b=2m$ being the boson mass and $\mu_0$ the chemical potential \begin{equation} \mu_0=\frac{1}{f^2(\a)}\frac{2^{2-4\a}B(1/2,3/2-2\a)}{(1-2\a)\sin{(\pi(1-\a))}B(\a,\a)} |\bar{\D}|^2\ve_a^{2\a-1} \label{e39} \end{equation} The velocity $c_0$ of the sound mode is \begin{equation} c_0^2=\frac{\mu_0}{m_b}=\frac{1}{f^2(\a)}\frac{2^{2-4\a}B(1/2,3/2-2\a)}{(1-2\a)\sin{(\pi(1-\a))}B(\a,\a)m} |\bar{\D}|^2\ve_a^{2\a-1} \label{e40} \end{equation} and the repulsive interaction $\l_{0b}$ between pairs is \begin{equation} \l_{0b}(\a)=\frac{\pi^2}{m}\frac{2^{4-4\a}B(1/2,3/2-2\a)}{(1-2\a)[\sin{(\pi(1-\a))}B(\a,\a)]^2} \label{e41} \end{equation} We mention that $\lim_{\a\rightarrow} \def\la{\leftarrow 0} \l_{0b}(\a)=2\pi/m$ a result identical to the result obtained in Ref. \onlinecite{16} for the two dimensional case. \section{Results and discussions} Using the field-theoretical methods we studied the crossover between BCS and BEC in a non-Fermi liquid. The weak coupling case leads to the same results as in the mean field like models \cite{25,26,27,28}. In the strong coupling limit we showed that the pairs form a Bose gas with a repullsive coupling constant which is controled by $\a$. \section{Acknowledgements} The authors are grateful to Adriaan M. J. Schakel for the enlightening (e-mail) discussions on the field theoretical method.
\section{ Introduction } At present we know of three quark-lepton chiral families in the standard model (SM). Their mixing within the present experimental accuracy is well known to be described by the $3\times 3$ unitary matrix~\cite{KM}. But beyond it, whether there are extra families and, if so, what their masses and mixings are --- this is yet unsolved problem. A recent two-loop renormalization group analysis~\cite{pir} of the SM shows that subject to the precision experiment restriction on the Higgs mass, $M_H\le 215$~GeV at 95\%~C.L.~\cite{tournefier}, the forth chiral family, if alone, is excluded.\footnote{The recent more conservative restrictions $m_H\le 262$ GeV or $M_H\le 300$ GeV at $95\%$ C.L., respectively, from the first and second papers of Ref.~\cite{higgs} render the fourth chiral family only marginally possible.} In fact, it does not depend on whether this extra family has the normal chiral structure or the mirror one. But as it is noted in Ref.~\cite{pir}, a pair of the opposite chirality families with the relatively low Yukawa couplings evades the SM self-consistency restrictions and could still exist. In order to conform to observations these extra families, which otherwise can be considered as the vectorial ones, should get large direct masses and drop out of the light particle spectrum of the SM in the decoupling limit. Nevertheless, at the not too high masses, say, in the TeV region, such families could result in observable corrections to the SM interactions through mixing with the light fermions. Various vector-like fermions are generic in many extensions of the SM like the superstring and grand unified theories, composite models, etc. Many issues concerning those fermions, both the electroweak doublets and singlets, the latter ones of the up and down types, were considered in the literature~\cite{vlf},~\cite{lavoura}. On the other hand there are numerous studies of the $n>3$ chiral family extensions of the SM~\cite{n_families}, \cite{santa}. Some topics concerning the SM extensions with the vector-like families are studied in Ref.~\cite{fuji}. In a previous letter \cite{VLF2} we presented the results for the SM light quark masses and mixings in the presence of the extra vector-like families. In the current paper we give the complete results including those for the heavy quarks. In Section 2 we carry out the model independent analysis for the general case. In Section 3 an explicit realization for the case with a pair of the heavy vector-like families is presented. In Appendix we give the technical details of the diagonalization procedure and the explicit form of the mixing matrices through the elements of the general mass matrices. \section{Model independent analysis} The most general content of the SM families consisting of the $\mathrm SU(2)_W \times U(1)_Y$ doublets and singlets is illustrated in Table~1. The notations with a hat sign designate quarks in the symmetry/electroweak basis where, by definition, the SM symmetry structure is well stated. ``Normal'' in the row means the $n\ge 3$ chiral families, similar in their chiral and quantum number structure to three ordinary families of the minimal SM. ``Mirror'' means the $m\ge 0$ mirror conjugate families with the normal quantum numbers, or in other terms, the charge conjugate families with the normal chiral structure. We suppose for definiteness that $n>m$. ``Chiral'' in the column means the chiral notations, and ``mixed'' corresponds to the more traditional left-right notations.\footnote{To be as clear as possible, what we are talking about, say, in terms of the 15-plets of the GUT $SU(5)$ ($15 = 10\oplus \overline{5}$) is $n 15_L\oplus m 15_R$, or $n 15_L\oplus m \overline{15}_L$. Nevertheless, the scales we have in mind are much lower than those of the GUT's, typically ${\cal O} (1 - 100$) TeV, i.e.\ rather those of the composite models.} \begin{table}[htbp] \paragraph{Table 1} The general content of the SM families. \vspace{1ex} \begin{tabular}{|c|c|c|c|} \hline &$\#$&Chiral&Mixed\\ \hline Normal&$n$&$Q_L=({\hat q}_L,{\hat u}_L^c, {\hat d}_L^c)$&$({\hat q}_L, {\hat u}_R, {\hat d}_R)$\\ Mirror&$m$&$Q'_R=({\hat q}_R', {\hat u}'^c_R, {\hat d}'^c_R)$&$({\hat q}'_R, {\hat u}'_L, {\hat d}'_L)$\\ \hline \end{tabular} \end{table} In general, quarks gain masses from two different physical mechanisms: that of the SM Yukawa interactions and that of a New Physics resulting in the SM invariant direct mass terms. Being chirally unprotected the latter ones should naturally be characterized by a high mass scale $M$, $M\gg v$, with $v$ being the SM Higgs vacuum expectation value. In the symmetry basis the kinetic, Yukawa and direct mass Lagrangian has the following most general form: \defD\hspace{-0.28cm}/\hspace{0.08cm}{D\hspace{-0.28cm}/\hspace{0.08cm}} \begin{eqnarray} \label{eq:lagrangian} {\cal L}&=& ~~~i \overline{{\hat q}_L}D\hspace{-0.28cm}/\hspace{0.08cm} {\hat q}_L + i \overline{{\hat u}_R}D\hspace{-0.28cm}/\hspace{0.08cm} {\hat u}_R + i \overline{{\hat d}_R}D\hspace{-0.28cm}/\hspace{0.08cm} {\hat d}_R\nonumber\\ &&+\, i\overline{{\hat q}'_R}D\hspace{-0.28cm}/\hspace{0.08cm} {\hat q}'_R + i \overline{{\hat u}'_L}D\hspace{-0.28cm}/\hspace{0.08cm} {\hat u}'_L + i \overline{{\hat d}'_L}D\hspace{-0.28cm}/\hspace{0.08cm} {\hat d}'_L\nonumber\\ &&-\Big{(}\overline{{\hat q}_L} Y^u {\hat u}_R \phi^c + \overline{{\hat q}_L} Y^d {\hat d}_R\phi +\overline{{\hat u}'_L} {Y^u}' {\hat q}'_R {\phi^c}^{\dagger} + \overline{{\hat d}'_L} {Y^d}' {\hat q}'_R \phi^{\dagger} + \mbox{h.c.}\Big{)} \nonumber\\ && - \Big{(} \overline{{\hat q}_L} M {\hat q}'_R + \overline{{\hat u}'_L}{M^u}' {\hat u}_R + \overline{{\hat d}'_L} {M^d}' {\hat d}_R + \mbox{h.c.}\Big{)}~, \end{eqnarray} where $D\hspace{-0.28cm}/\hspace{0.08cm}\equiv \gamma^\mu D_\mu$ is the SM covariant derivative, $\phi$ is the Higgs doublet and $\phi^c$ is the charged conjugate one. In Eq.~(\ref{eq:lagrangian}), $Y$ and $Y'$ are, respectively, the square $n\times n$ and $m\times m$ Yukawa matrices; $M$ and $M'$ are, respectively, the rectangular $n\times m$ and $m\times n$ direct mass matrices. Without loss of generality, the matrices $M$ and $M'$ can always be brought to the $m\times m$ triangular form with the rest being zero. Now, one can rewrite the Lagrangian~(\ref{eq:lagrangian}) in terms of the $m$ pairs of the Dirac families $Q= (Q_L, Q'_R)$, constituting the vector-like representations of the SM, and the $n-m$ chiral families $Q_L$. In neglect of the Yukawa couplings, the Lagrangian of the Dirac families is explicitly $P$ invariant. Hence, of those initial $n+m$ chiral families, the $2m$ ones transform after mass diagonalization to $m$ pairs of the heavy vector-like families (VLF's).\footnote{To be precise we call as VLF the family mass eigenstate which possesses the (approximate) left-right symmetric SM interactions.} This is to be expected according to the survival hypothesis~\cite{georgi} because the chirally conjugate families lose their chiral protection. The unbalanced $n-m$ families can be considered as the (approximate) pure chiral ones. In practice, we suppose that the net number of the chiral families is three and hence $n=3+m$. We generalize the parameter counting for the chiral families of Ref.~\cite{santa} to the case with extra VLF's. It goes as is shown in Table 2. Here $G$ is the global symmetry of the kinetic part of the Lagrangian~(\ref{eq:lagrangian}). It is broken explicitly by the mass terms, only the residual symmetry $H = U(1)$ of the baryon number being left in the general case we consider.\footnote{The degenerate cases leave more residual symmetries and require special consideration.} Hence, the transformations of $G/H$ can be used to absorb the spurious parameters in Eq.~(\ref{eq:lagrangian}) leaving only the physical set ${\cal M}_{phys}$ of them. The last four lines in Table~2 present the physical parameters for the minimal SM and for the three its simplest extensions: the traditional one with a normal family, the one with a mirror family and the one with of a pair of the normal and mirror families.\footnote{The first two cases are practically excluded by the SM self-consistency requirements~\cite{pir}.} The last case will be considered in detail in the next section. \begin{table}[htbp] \paragraph{Table 2}Parameter counting in the symmetry/electroweak basis. \vspace{1ex} \begin{tabular}{|c|c|c|} \hline Couplings&Moduli&Phases\\ and symmetries&&\\ \hline $Y^u, Y^d, {Y^u}', {Y^d}',$&$2(n^2 + m^2)$&$2(n^2 + m^2)$\\ $M, {M^u}', {M^d}'$&$+ 3n m$&$+ 3n m$\\ \hline $G = U(n)^3\times U(m)^3$&$-\frac{3}{2} [n(n-1) + m(m-1)]$& $-\frac{3}{2} [n(n+1) + m(m+1)]$\\ \hline $H = U(1)$&$0$&$1$\\ \hline ${\cal M}_{phys}(n, m)$&$\frac{1}{2} (n+m)(n+m-1) $& $\frac{1}{2} (n+m-2)(n+m-1)$\\ &$ + 2n m +2(n + m)$&$ +2n m$\\ \hline ${\cal M}_{phys}^{\mbox{\scriptsize SM}}(3, 0)$&$9 = 3 + 6$&$1$\\ \hline ${\cal M}_{phys}(4, 0)$&$ 8 + 6 = 14 $&$3$\\ \hline ${\cal M}_{phys}(3, 1)$&$ 8 + 12 = 20 $&$9$\\ \hline ${\cal M}_{phys}(4, 1)$&$10 + 18 = 28$&$14$\\ \hline \end{tabular} \end{table} Further, the kinetic part of the effective Lagrangian with the $W$, $Z$ and Higgs bosons being integrated out is \begin{eqnarray} {\cal L}_{eff}&=& ~~i \overline{u_L}D\hspace{-0.28cm}/\hspace{0.08cm} u_L + i \overline{d_L}D\hspace{-0.28cm}/\hspace{0.08cm} d_L + i \overline{u_R}D\hspace{-0.28cm}/\hspace{0.08cm} u_R + i \overline{d_R}D\hspace{-0.28cm}/\hspace{0.08cm} d_R\nonumber\\ &&-\big(\overline{u_L}{\cal M}^{u}_{diag} u_R + \overline{d_L}{\cal M}^{d}_{diag} d_R +\mbox{h.c.} \big)~, \end{eqnarray} where $D\hspace{-0.28cm}/\hspace{0.08cm}$ means the covariant derivatives w.r.t.\ the QED and QCD only; $ u_\chi$ and $d_\chi$ ($\chi = L$, $R$) generically mean the quarks in the mass/flavour basis, and ${\cal M}^{u, d}_{diag}$ are the diagonal mass matrices defining the basis. The corresponding parameter counting is presented in Table~3. Due to the absence of mutual quark transitions, the total residual symmetry of the mass matrices ${\cal M}^{u,d}_{diag}$ is here $H = U(1)^{2(n+m)}$. Table~3 clearly shows the breakdown of the moduli of ${\cal M}_{phys}$ in Table~2 on the physical masses and mixing angles. \begin{table}[htbp] \paragraph{Table 3}Parameter counting for the effective Lagrangian. \vspace{1ex} \begin{tabular}{|c|c|c|} \hline Couplings&Moduli&Phases\\ and symmetries&&\\ \hline ${\cal M}^{u}, {\cal M}^{d}$&$2(n + m)^2$&$2(n + m)^2$\\ \hline $G = U(n + m)^4$&$-2 (n + m)(n + m - 1)$&$-2 (n + m)(n + m + 1)$\\ \hline $H = U(1)^{2(n + m)}$&$0$&$2 (n + m)$\\ \hline ${\cal M}^{u}_{diag}, ~{\cal M}^{d}_{diag}$&$2 (n + m)$&$0$\\ \hline \end{tabular} \end{table} Let us now redefine collectively quarks in the symmetry basis as \def\kappa{\kappa} ${\hat \kappa}_\chi = {\hat u}_\chi$, ${\hat d}_\chi$ and these in the mass basis, i.e.\ the quark eigenstates with ${\cal M}_{phys}$ being diagonal, as $\kappa_\chi = u_\chi$, $d_\chi$ ($\chi = L$, $R$). The bases are related by the unitary $(n+m)\times(n+m)$ transformations \begin{equation} {\hat\kappa}_\chi{}_{A} = ({U^\kappa_\chi})^{ F}_{ A} \,{\kappa_\chi}_{F}~, \end{equation} with the ensuing bi-unitary mass diagonalization \begin{equation} \label{4} {U^{\kappa}_L}^\dagger {\cal M}^{k} U^{\kappa}_R = {\cal M}^{\kappa}_{diag} = \mbox{diag\,}(\overline{ m}^{\kappa}{}_f , \overline{ M}^{\kappa}{}_4, \dots ,\overline{ M}^{\kappa}{}_{n+m})~. \end{equation} In the equations above, the indices $A = A_L,A_R$; $A_L = 1,\dots, n$; $A_R = n+1,\dots, n+m$ are those in the symmetry basis, and $F = f,4, \dots ,n+m$; $f = 1,2,3$ are indices in the mass basis. It is assumed that $\overline{m}^\kappa{}_f \ll \overline{M}^\kappa{}_4, \dots , \overline{ M}^\kappa{}_{n+m}$. The matrices $U^\kappa_\chi$ satisfy the unitarity relations \begin{equation} \label{6} {U^{\kappa}_\chi}\,U^{\kappa}_\chi{}^{\dagger} =I \end{equation} and \begin{equation}\label{5} {U^{\kappa}_\chi}^{\dagger} I_L {U^{\kappa}_\chi} + {U^{\kappa}_\chi}^{\dagger}I_R {U^{\kappa}_\chi} = I~, \end{equation} were $I_L$, $I_R$ are the projectors onto the normal and mirror subspaces in the symmetry basis: \begin{eqnarray} I_L&=&\mbox{diag}\, (\,\underbrace{1,\dots,1}_{n}\,;\underbrace{0,\dots,0}_{m}\,)~, \nonumber\\ I_R&=&\mbox{diag}\, (\,\underbrace{0,\dots,0}_{n}\,;\underbrace{1,\dots,1}_{m}\,) \end{eqnarray} with $I_L+I_R=I$ and $I_\chi^2=I_\chi$. Let us also introduce their transformation to the mass basis \begin{equation} \label{eq:projector} X^\kappa_\chi = U^\kappa_\chi{}^{\dagger} I_\chi U^\kappa_\chi~. \end{equation} ($\kappa = u$, $d$ and $\chi = L$, $R$). Clearly, $X^\kappa_\chi$ are Hermitian and satisfy the projector condition: $X^\kappa_\chi{}^2 = X^\kappa_\chi{}$ (but note that $X_L^\kappa + X_R^\kappa\neq I$ in the notations adopted). Now, the charged current Lagrangian is \begin{equation}\label{eq:L_W} - {\cal L}_W = \frac{g}{\surd\overline{2}} W^+_\mu \sum_\chi \overline{u_\chi} \gamma^\mu V_\chi d_\chi + \mbox{h.c.} \end{equation} and the neutral current one is \begin{equation}\label{eq:L_Z} - {\cal L}_Z = \frac{g}{c} Z_\mu \sum_{\kappa,\chi} \overline{\kappa_\chi} \gamma^\mu N^{\kappa}_\chi\, \kappa_\chi~, \end{equation} where $c\equiv\cos\theta_W$, with $\theta_W$ being the Weinberg mixing angle. The corresponding quark mixing matrices for the charged currents are \begin{equation}\label{V_chi} {V_\chi} = {U^{u}_\chi}^{\dagger}I_\chi U^{d}_\chi~, \end{equation} and for the neutral currents with the operator $T_3-s^2 Q$ \begin{equation}\label{eq:N_chi} N^\kappa_\chi=T^\kappa_3 X^\kappa_\chi-s^2 Q^\kappa_\chi~. \end{equation} Here one has for the electroweak isospin: $T_3^\kappa = 1/2$ at $\kappa=u$ and $-1/2$ at $\kappa=d$; for the electric charge: $Q_{L,R}^\kappa\equiv Q^\kappa I$ with $Q^\kappa =2/3$ at $\kappa=u$ and $-1/3$ at $\kappa=d$; $s\equiv\sin\theta_W$. The charged current mixing matrices $V_L$ and $V_R$ play the role of the generalized CKM matrices. But contrary to the minimal SM case, they as well as the neutral current mixing matrices $N^{\kappa}_\chi$ are non-unitary. Namely, one gets by the unitarity relations~(\ref{6}) \begin{eqnarray}\label{14} V_\chi V_\chi^\dagger &=& X^u_\chi ~,\nonumber\\ V_\chi^\dagger V_\chi &=& X^d_\chi~, \end{eqnarray} where $X_\chi^\kappa$ ($X_\chi^\kappa\neq I$ in general) are given by Eq.~(\ref{eq:projector}). From the considerations above, the representations for the $V_\chi$ follow \begin{equation}\label{eq:representation} V_\chi=X^u_\chi S_\chi=S_\chi X^d_\chi \end{equation} with the unitary matrices $S_\chi={U^u_\chi}^\dagger U^d_\chi$ and the positive definite Hermitian matrices $X^\kappa_\chi$, only one in a pair with fixed $\chi$ being independent, say, $X_\chi^d\equiv S_\chi^\dagger X_\chi^u S_\chi$. The decomposition (\ref{eq:representation}) is known to be unique. In a case where there are only the normal families, one gets $X_L^\kappa=I$ and $X_R^\kappa=0$, so that $V_L$ is unitary, $V_L=S_L$, and $V_R=0$. It is seen that the neutral current matrices $N^\kappa_\chi$ are not independent of the charged current ones $V_\chi$. In fact, one can convince oneself that $V_\chi$ and the diagonal mass matrices ${\cal M}^{\kappa}_{diag}$ suffice to parametrize all the fermion interactions in a general class of the SM extensions by means of the arbitrary numbers of the vector-like isodoublets and isosinglets~\cite{lavoura}. Indeed, in the case at hand, using the unitarity relations (\ref{5}), one gets for the Yukawa Lagrangian in the unitary gauge \begin{eqnarray}\label{L_Y} - {\cal L}_Y &=& \frac{H}{v} \sum_\kappa \overline{\kappa_L} \Big{(} X^\kappa_L {\cal M}^\kappa_{diag} - 2 X^\kappa_L {\cal M}^\kappa_{diag} X^\kappa_R + {\cal M}^{\kappa}_{diag} X^\kappa_R \Big{)} \kappa_R \nonumber\\ && + \sum_\kappa \overline{\kappa_L} {\cal M}^{\kappa}_{diag} \kappa_R + \mbox{h.c.}~, \end{eqnarray} $H$ being the physical Higgs boson. It follows from the above expression and Eqs.~(\ref{eq:L_Z}), (\ref{eq:N_chi}) that all the flavour changing neutral currents are induced entirely by the lack of unitarity of the charged current mixing matrices $V_\chi$. In the case with only the normal families ($X^\kappa_L=I$, $X^\kappa_R=0$) the usual SM expressions for ${\cal L}_W$, ${\cal L}_Z$ and ${\cal L}_Y$ are recovered, the two latter ones being flavour conserving. We propose the following prescription for the model independent paramet\-ri\-za\-tion of the $V_\chi$. The problem is that they are non-unitary and thus are difficult to parametrize directly. So, the idea is to express them in terms of a set of the auxiliary unitary matrices. First of all, note that in the absence of any restrictions on the Lagrangian the unitary matrices $U^\kappa_\chi$ in Eq.~(3) would be arbitrary. Now, an arbitrary $(n + m)\times (n + m)$ unitary matrix $U$ can always be uniquely decomposed as $U = U{\vert}_{n\times n} ~U{\vert}_{m\times m} ~U{\vert}_{n\times m}$. Here $U{\vert}_{n\times n}$ is a unitary matrix in the $n\times n$ subspace. It is built of the $n^2$ generators. Similarly, $U{\vert}_{m\times m}$ is the restriction of $U$ onto the $m\times m$ subspace, and it is built of the $m^2$ generators. And finally, $U{\vert}_{n\times m}$ means a unitary $(n + m)\times (n + m)$ matrix built of the $2nm$ generators which mix the two subspaces. Now, by means of the symmetry basis transformations $G$ of Table~2 one can always put, without loss of generality, the matrices $U^\kappa_\chi$ to the form \begin{eqnarray}\label{eq:U_repr} U^u_L &=& U^u_L{\vert}_{n\times m}~, \nonumber\\ U^u_R &=& U^u_R{\vert}_{n\times m}~, \nonumber\\ U^d_L &=& U^d_L{\vert}_{n\times n} ~U^d_L{\vert}_{n\times m}~, \nonumber\\ U^d_R &=& U^d_R{\vert}_{m\times m} ~U^d_R{\vert}_{n\times m}~. \end{eqnarray} This representation includes six auxiliary unitary matrices. Clearly, they depend on the $[n(n-1)/2 + m(m-1)/2 + 4mn]$ moduli and $[n(n+1)/2 + m(m+1)/2 + 4mn]$ phases, and these numbers are redundant. But the $nm$ moduli and the same number of phases can be eliminated through the $n\times m$ matrix constraint \begin{equation}\label{eq:constraint} I_L U^u_L {\cal M}^u_{diag} {U^u_R}^{\dagger} I_R = I_L U^d_L {\cal M}^d_{diag} {U^d_R}^{\dagger} I_R~. \end{equation} The latter one follows from the equality of the direct mass matrices $M$ in Eq.~(1) for the up and down quarks, and it includes additionally the $2(n+m)$ independent moduli which enter ${\cal M}^u_{diag}$ and ${\cal M}^d_{diag}$. By means of Eq.~(\ref{eq:constraint}) one can express, e.g., one of the $U^\kappa_\chi{\vert}_{n\times m}$ in terms of all other matrices. And finally, the $2(n+m)-1$ phases can be removed via the residual phase redefinition for the quarks in the mass basis. Putting all together, one can easily verify that the total number of the independent parameters is precisely as expected from Table 2. Having parametrized the auxiliary unitary matrices, one gets for the $V_\chi$ \begin{eqnarray}\label{eq:V_repr} V_L &=& {U^u_L}^\dagger{\vert}_{n\times m}~ I_L~ {U^d_L}{\vert}_{n\times n} ~{U^d_L}{\vert}_{n\times m}~, \nonumber\\ V_R &=& {U^u_R}^\dagger{\vert}_{n\times m}~ I_R~ {U^d_R}{\vert}_{m\times m} ~{U^d_R}{\vert}_{n\times m} \end{eqnarray} and for the $X^\kappa_\chi$ \begin{equation}\label{eq:X_repr} X^\kappa_\chi = {U^\kappa_\chi}^\dagger{\vert}_{n\times m}~ I_\chi~ {U^\kappa_\chi}{\vert}_{n\times m}~. \end{equation} When eliminating the $2(n+m)-1$ redundant phases one can always take such a choice as to render the diagonal and above-the-diagonal elements of the $V_L$ (or $V_R$) to be real and positive. This gives a principal solution to the problem. When there are only the normal families ($m = 0$) the usual parametrization in terms of just one unitary matrix $U^d_L{\vert}_{n\times n}$ is readily recovered. For the case with a pair of VLF's ($n=4$, $m=1$) we got also the explicit expressions of all the relevant quantities in terms of a minimal common set of the independent arguments parametrizing the mass matrices (see the next section). It is of much use at the model independent parametrization to estimate the relative magnitudes of the various mixing elements in terms of a small quantity $\epsilon = v^2/M^2\ll 1$. Otherwise, one has a priori no idea of this. Finally, under small mixing it is useful to decompose \begin{equation} V_\chi = V_{0 \chi} + \Delta V_{\chi}~, \end{equation} with the decoupling limit taken as the zeroth order approximation $V_{0 \chi}$, and with corrections $\Delta V_{\chi}$ vanishing at $M\gg v$. To illustrate the behavior in the limit, let us consider the aforementioned case with a pair of the VLF's. One gets here \begin{equation}\label{eq:V_0L} V_{0L}= \left({ \begin{array}{ccc} V_{C}&0&0\\ 0&1&0\\ 0&0&0 \end{array} }\right) \end{equation} and \begin{equation}\label{eq:V_0R} V_{0R} = \mbox{diag\,}(0,0,0,1,0)~, \end{equation} $V_{C}$ being the usual $3\times 3$ charged current matrix of the SM. Hence, for the $X^\kappa_\chi$ as given by Eq.~(\ref{14}) one has in the zeroth order \begin{eqnarray}\label{eq:declim} X^{u,d}_{0L}&=&\mbox{diag\,}(1,1,1,1,0)~,\nonumber\\ X^{u,d}_{0R}&=&\mbox{diag\,}(0,0,0,1,0)~. \end{eqnarray} It follows from Eqs.~(\ref{eq:V_0L})--(\ref{eq:declim}) that in the limit $M\gg v$ there are indeed two VLF's, the forth and the fifth ones, that interact in the left-right symmetric manner, one of the VLF's, the fifth one, being singlet under interactions with the $W$ boson. Besides, as it follows from Eq.~(\ref{L_Y}), both these families decouple from the Higgs boson in the leading order of ${\cal O}(M/v)$, only the Yukawa terms ${\cal O}(M^0)$ being left at most. \section{Explicit realization} The mass/flavour basis parameters, ${\cal M}^{u,d}_{diag}$ and $V_{L,R}$, are phenomenological by their very nature. They reflect an obscure mixture of contributions of quite a different physical origin. In particular, they shed no light on the mixing magnitudes. On the contrary, the parameters in the symmetry basis, i.e.\ Yukawa couplings and the direct mass terms $M$ and ${M^u}'$, ${M^d}'$, have the straightforward theoretical meaning. So, we express the former ones in terms of the latter ones. This permits us to expand upon the idea of the relative magnitude of the various mixing elements in terms of the small quantity $v/M$. The asymptotic freedom requirement for the $\mathrm SU(2)_W$ electroweak interactions results in the restriction that the total number of the electroweak doublets should not exceed 21. The number of doublets in a chiral family being 4, this is equivalent to the restriction that the total number of the families is $(n+m)\le 5$. Hence the maximum number of the extra VLF's allowed by the asymptotic freedom is two, the case we stick to in what follows.\footnote{This might be a landmark for the number of the extra families.} Using here the global symmetries $G$ of the Table~2 one can bring, without loss of generality, the quark mass matrices in the symmetry basis to the following canonical form \begin{equation}\label{20} {\cal M}^\kappa = \left( \begin{array}{ccc} \vspace{0.5ex} {m^\kappa}^g_f&{\mu^\kappa}'_f&0\\ \vspace{0.5ex} {\mu^\kappa}^g&m^\kappa{}_4&M\\ 0&{M^\kappa}'&m^\kappa{}_5 \end{array} \right)~, \end{equation} where $M$, ${M^\kappa}'$ are the real scalars and $\mu^\kappa{}^f$, ${\mu^\kappa}'_f$, $m^\kappa{}_4$, $m^\kappa{}_5$ are in general complex. Here the lower case characters generically mean the masses of the Yukawa origin ($\sim Yv$). Let us remind that $M$ in Eq.~(\ref{20}) is common for both ${\cal M}^u$ and ${\cal M}^d$. The three-dimensional matrices $m^\kappa$ are Hermitian and positive definite, and one of them, e.g. $m^u$, can always be chosen diagonal. Under such a choice one can simplify further: \begin{equation}\label{21} {\cal M}^\kappa_0 = {U^\kappa_0}^\dagger {\cal M}^\kappa U^\kappa_0, \end{equation} where \begin{equation}\label{eq:M_0_kappa} {\cal M}_0^{\kappa} = \left(\begin{array}{ccccc} \vspace{0.5ex} m^\kappa{}_1&0&0&{\mu^\kappa}'_1&0\\ \vspace{0.8ex} 0&m^\kappa{}_2&0&{\mu^\kappa}'_2&0\\ \vspace{0.5ex} 0&0&m^\kappa{}_3&{\mu^\kappa}'_3&0\\ \vspace{0.5ex} {\mu^\kappa}^1&{\mu^\kappa}^2&{\mu^\kappa}^3&m^\kappa{}_4&M\\ \vspace{0.5ex} 0&0&0&{M^\kappa}'&m^\kappa{}_5 \end{array}\right) \end{equation} \noindent with a redefinition of $\mu^\kappa{}^f$ and ${\mu^\kappa}'_f$, and with the diagonal elements $m^\kappa{}_f$ being real and positive. The matrices ${\cal M}_0^{\kappa}$ have a lot of texture zeros and are easiest to operate. The corresponding unitary $U^\kappa_0$ are given by \begin{eqnarray}\label{22a} U^{u}_0 &=& I~,\nonumber\\ U^d_{0 } &=& \left({ \begin{array}{cc} V_{C}&0\\ 0&I_2 \end{array} }\right)~, \end{eqnarray} $V_C$ being the $3\times 3$ CKM matrix and $I_2$ the $2\times 2$ identity matrix. The mass matrices of Eq.~(\ref{eq:M_0_kappa}) possess the residual symmetry $U(1)^6$ which is reduced to $U(1)^5$ by the baryon number conservation. So, one can use the phase redefinitions for two of the light $d$ quarks which leave just one complex phase in $V_C$ in accordance with the decoupling limit requirement. It is seen from Eqs.~(\ref{eq:M_0_kappa}) and (\ref{22a}) that in this parametrization the total number of physical moduli is $10 + 15 + 3 = 28$ as it should be according to Table~2. As for the phases, their number is in general $16 + 1 = 17$, i.e.\ three of them are spurious and can be removed. For example, by means of the residual phase redefinition for the three light $u$ quarks one can make $\mu^u{}^f$ or ${\mu^u}'_f$ to be real, or put some other three relations on their phases. This exhausts the freedom of the phase redefinitions, leaving only the physical parameters. The characteristic equations (see Appendix) \begin{equation}\label{23} \det\,({\cal M}^\kappa_0 {{\cal M}^\kappa_0}^\dagger - \lambda^\kappa I) = 0 \end{equation} give for the roots in the first order (i.e. up to the relative corrections ${\cal O}(v^2/M^2)$ to the leading order): \begin{eqnarray}\label{eq:lambda} \lambda_f\equiv \overline{ m}_f^2 &=&m_f^2\bigg{(}1 - \Big(\frac{\vert\mu^f\vert^2}{M^2} + \frac{\vert\mu'_f\vert^2}{M^{'2}}\Big)\bigg{)} + \frac{m_f}{M M'} (m_5\mu^f\mu'_f + \mbox{h.c.})~,\nonumber\\ \lambda_4\equiv \overline{ M}_4^2&=&M^2 + \Sigma \vert{\mu^f}\vert^2+ \vert m_4\vert^2 + \vert{m_5}\vert^2\nonumber\\ &&+ \frac{M'{}^2}{M^2 - M'{}^{2}} \Big{(} (\vert{m_4}\vert^2 + \vert{m_5}\vert^2) + \frac{M}{M'}(m_4 m_5 + \mbox{h.c.})\Big{)}~,\nonumber\\ \lambda_5\equiv \overline{ M}_5^2&=&M^{'2} + \Sigma\vert\mu'_f\vert^2 + \vert m_4\vert^2 + \vert m_5\vert^2\nonumber\\ &&+ \frac{M^2}{M'{}^{2} - M^2} \Big{(} (\vert m_4\vert^2 + \vert m_5\vert^2) + \frac{M'}{M} (m_4 m_5 + \mbox{h.c.})\Big{)} \end{eqnarray} with the superscripts $\kappa = u$, $d$ being suppressed.\footnote{Hence, the up and down quarks of the fourth family are always (almost) degenerate, whereas those of the fifth family are in general not. Nevertheless, because the fifth family does not couple to the $W$ boson in the zeroth order (see Eqs.~(21), (22)) this does not result in the strong coupling $\sim ({M^u}' - {M^d}')$ of the longitudinal $W$ with the fifth heavy family, as well as with the fourth one. } Here it is supposed that one has, in general, $M\sim M'$ but $M\neq M'$.\footnote{The degenerate case $M = {M^\kappa}'$ (for one or both $\kappa = u,d$) is to be studied separately. It modifies the results for heavy families, but fortunately does not influence the validity of those concerning the light quarks exclusively.} It is seen that corrections to $m_f^2$ are proportional to $m_f$ themselves, i.e. the light quarks are still chirally protected. This property drastically reduces the otherwise dangerous corrections to the masses of the lightest $u$ and $d$ quarks at the moderate $M$. In the limit $m_f\to 0$ it naturally happens without any fine tuning beyond that of the SM. On the other hand, it means that within the perturbation theory the masses of the lightest quarks cannot entirely be induced by an admixture of the vector-like families: if $m_f = 0$ then $\overline{m}_f = 0$, too. But at the finite $m_f$ one finds for the masses of the light quarks \begin{equation}\label{eq:m_light} \overline{m}_f = m_f \Bigg{(} 1 - \frac{1}{2} \Big{(} \frac{\vert\mu^f\vert^2}{M^2} + \frac{\vert\mu'_f\vert^2}{{M'}^2} \Big{)}\Bigg{)} + \frac{1}{2} \Big{(} \frac{m_5\mu^f \mu'_f}{MM'} + \mbox{h.c.} \Big{)}~, \end{equation} and for the validity of perturbative expansion it could require some fine tuning for $m_5$ at the moderate $M$. Once the physical masses are known, one can obtain the matrices $U^\kappa_{1L}$ and $U^\kappa_{1R}$ of the bi-unitary transformation \begin{equation}\label{26} U^{\kappa\dagger}_{1L} {\cal M}^\kappa_0 U^\kappa_{1R} = {\cal M}^\kappa_{diag}~. \end{equation} Obviously, they satisfy the relations \begin{eqnarray}\label{27} {{\cal M}^\kappa_0}^\dagger {\cal M}^\kappa_0 U^\kappa_{1R} &=& U^\kappa_{1R} {{\cal M}^\kappa_{diag}}^2~,\nonumber\\ {\cal M}^\kappa_0 {{\cal M}^\kappa_0}^\dagger U^\kappa_{1L} &=& U^\kappa_{1L} {{\cal M}^\kappa_{diag}}^2 \end{eqnarray} \noindent which are to be considered as the sets of the independent linear equations for their columns. Having solved the equations, one can find the elements of $U^\kappa_{1\chi}$ which are given in Appendix. Finally, one has for the total matrices of the bi-unitary transformations of Eq.~(\ref{4}) \begin{equation}\label{27a} U^\kappa_\chi = {U^\kappa_0}\, U^\kappa_{1\chi}~, \end{equation} where $U^\kappa_0$ are given by Eq.~(\ref{22a}). Hereof one gets the mixing matrices $V_\chi$ given by Eqs.~(\ref{A9}), (\ref{A10}) of Appendix and then the charged current Lagrangian ${\cal L}_W$ given by Eq.~(\ref{eq:L_W}). The $Z$-mediated neutral current Lagrangian ${\cal L}_Z$ is given by Eqs.~(\ref{eq:L_Z}), (\ref{eq:N_chi}) with $X^\kappa_\chi$ from Eqs.~(\ref{A11}), (\ref{A12}). The neutral scalar current Lagrangian takes the general form \begin{equation}\label{29} - {\cal L}_H = \frac{H}{v} \sum_\kappa \overline{\kappa_L}\, {U^\kappa_L}^\dagger ({\cal M}^\kappa - {\cal M}^\kappa_{dir}) U^\kappa_R \, \kappa_R + \mbox{h.c.} \end{equation} with the direct mass matrices \begin{equation}\label{30} {\cal M}^\kappa_{dir} = \left({ \begin{array}{ccc} O_3&0&0\\ 0&0&M\\ 0&{M^\kappa}'&0 \end{array} }\right)~, \end{equation} where $O_3$ is the $3\times 3$ zero matrix. As a consequence of the substraction of the direct mass terms, the total mass and Yukawa matrices are not diagonalizable simultaneously in the same basis, at variance with the SM case. In the mass basis, the Higgs interaction Lagrangian is non-diagonal \begin{eqnarray}\label{eq:Higgs_Lagr} - {\cal L}_H &=& \frac{H}{v} \sum_\kappa {\overline{\kappa_L}}~ {\cal H}^\kappa \kappa_R + \mbox{h.c.} \end{eqnarray} with the explicit form of ${\cal H}^\kappa$ given by Eqs.~(\ref{A13}), (\ref{A14}) of Appendix. One should stress that for the light quarks all the off-diagonal components of the Lagrangian ${\cal L}_W$ (beyond that of the minimal SM), as well as those of the ${\cal L}_Z$ and ${\cal L}_H$ are suppressed by the ratio $v^2/M^2$, and it does not depend on the details of the mass matrices. Besides, it follows from the above that, among the off-diagonal interactions, the Higgs mediated interactions are the only ones that do not vanish in the decoupling limit. Hence, the heavy quarks are expected to decay mainly into the light ones and the Higgs boson with the natural decay width $\Gamma\sim \vert{}Y\vert^2/4\pi~M$. As a result, all the leading loop corrections to the light quark processes with the internal heavy vector-like quarks are expected to be mediated by the Higgs boson exchanges. So, the modern SM physics, i.e. predominantly that of the light fermions and the gauge bosons, may be succeeded by that of the heavy vector-like fermions and the Higgs boson. \section{Conclusions} We have shown that the mere addition of a pair of the VLF's drastically changes all the characteristic features of the minimal SM. First of all, the generalized CKM matrix for the left-handed charged currents ceases to be unitary. Moreover, this non-unitarity takes place in the whole flavour space but not only in the light quark sector which would occur for adding only the normal families. Further, there appear the right-handed charged currents, the flavour changing neutral currents, both the vector and scalar ones, all with the non-unitary mixing matrices and with a number of $CP$ violating phases. Due to decoupling relative to the large direct mass terms $M$, the extended SM definitely does not contradict experiment in the limit $M\gg v$. But at the moderate $M>v$, the addition of a pair of the VLF's would make the model phenomenology, especially that of the flavour and $CP$ violation, extremely diverse. So, the extension opens new prospects for studying the deviations from the SM in the future experiments at high energies. \section*{ Appendix } \setcounter{equation}{0} \defA.\arabic{equation}{A.\arabic{equation}} One has generically (with the indices $\kappa = u$, $d$ being omitted) \begin{equation}\label{A1} \begin{array}{l} {\cal M}_0 {\cal M}_0^\dagger = \\ \\ \left({ \begin{array}{ccccc} \vspace{1ex} (m_1^2+\vert\mu'_1\vert^2)&\mu'_1 {\mu'_2}^*&\mu'_1 {\mu'_3}^*& (m_1\mu_1^*+\mu'_1 m_4^*)&\mu'_1 {M'}\\ \vspace{1ex} \mu'_2 {\mu'_1}^*&(m_2^2+\vert\mu'_2\vert^2)&\mu'_2 {\mu'_3}^*& (m_2\mu_2^*+\mu'_2 m_4^*)&\mu'_2 {M'}\\ \vspace{1ex} \mu'_3 {\mu'_1}^*&\mu'_3 {\mu'_2}^*&(m_3^2+\vert\mu'_3\vert^2)& (m_3\mu_3^* + \mu'_3 m_4^*)&\mu'_3 {M'}\\ \vspace{1ex} (m_1\mu_1\phantom{+}&(m_2\mu_2\phantom{+}&(m_3\mu_3\phantom{+}&({ M}^2+{\vert m_4\vert}^2&(m_4 {M'}\phantom{+}\\ \vspace{1ex} ~+{\mu'_1}^* m_4)&+{\mu'_2}^* m_4)&+{\mu'_3}^* m_4)& +\Sigma{\vert\mu^f\vert}^2)& + M m_5^*)\\ \vspace{1ex} {\mu'_1}^* M'&{\mu'_2}^* M'&{\mu'_3}^* M'&(m_4^* M'+M m_5)&({ M'}^2 + {\vert m_5\vert}^2) \\ \end{array} }\right)\,. \end{array} \end{equation} \noindent The characteristic equation \begin{equation}\label{A2} \det\,({\cal M}_0 {\cal M}_0^{\dagger} - \lambda I) = 0 \end{equation} in the explicit form is \begin{eqnarray}\label{A3} && \lambda^5 - \lambda^4 \bigg[M^2 + {M'}^2 + \Sigma \Big{(} m_f^2 + \vert\mu^f\vert^2 + \vert\mu_f'\vert^2 \Big{)} + \vert m_4\vert^2 + \vert m_5\vert^2\bigg] \nonumber\\&& + \lambda^3 \bigg{[} M^2 {M'}^2 + M^2 \Sigma \Big{(} m_f^2 + \vert\mu_f'\vert^2\Big{)} + {M'}^2 \Sigma \Big{(} m_f^2 + \vert\mu^f\vert^2\Big{)} \nonumber\\&& - M {M'} (m_4 m_5 + \mbox{h.c.}) + m_1^2 m_2^2 + m_1^2 m_3^2 + m_2^2 m_3^2\bigg{]} \nonumber\\&& - \lambda^2 \bigg{[}M^2 {M'}^2 \Sigma m_f^2 + M^2 \Big{(}m_1^2 m_2^2 + m_1^2 m_3^2 + m_2^2 m_3^2 \nonumber\\&& + m_1^2 (\vert\mu_2'\vert^2 + \vert\mu_3'\vert^2) + m_2^2 (\vert\mu_1'\vert^2 + \vert\mu_3'\vert^2) + m_3^2 (\vert\mu_1'\vert^2 + \vert\mu_2'\vert^2) \Big{)} \nonumber\\&& + {M'}^2 \Big{(}m_1^2 m_2^2 + m_1^2 m_3^2 + m_2^2 m_3^2 \nonumber\\&& + m_1^2 (\vert{\mu}_2\vert^2 + \vert{\mu}_3\vert^2) + m_2^2 (\vert{\mu}_1\vert^2 + \vert{\mu}_3\vert^2) + m_3^2 (\vert{\mu}_1\vert^2 + \vert{\mu}_2\vert^2) \Big{)} \nonumber\\&& + M {M'} \Big{(}( - m_4 \Sigma m_f^2 + \Sigma m_f\mu^f\mu'_f) m_5 + \mbox{h.c.} \Big{)} + m_1^2 m_2^2 m_3^2 \bigg{]} \nonumber\\&& + \lambda \bigg{[} M^2 {M'}^2 \Big{(} m_1^2 m_2^2 + m_1^2 m_3^2 + m_2^2 m_3^2 \Big{)} \nonumber\\&& + M^2 \Big{(} m_1^2 m_2^2 m_3^2 + m_2^2 m_3^2 \vert\mu_1'\vert^2 + m_1^2 m_3^2 \vert\mu_2'\vert^2 + m_1^2 m_2^2 \vert\mu_3'\vert^2 \Big{)} \nonumber\\&& + {M'}^2 \Big{(} m_1^2 m_2^2 m_3^2 + m_2^2 m_3^2 \vert\mu_1\vert^2 + m_1^2 m_3^2 \vert\mu_2\vert^2 + m_1^2 m_2^2 \vert\mu_3\vert^2 \Big{)} \bigg{]} \nonumber\\&& - \bigg{[}M^2 {M'}^2 m_1^2 m_2^2 m_3^2 + M {M'} \Big{(}\big{(} - m_1^2 m_2^2 m_3^2 m_4 + m_1 m_2^2 m_3^2 \mu_1\mu'_1 \nonumber\\&& + m_1^2 m_2 m_3^2 \mu_2\mu'_2 + m_1^2 m_2^2 m_3 \mu_3\mu'_3 \big{)} m_5 + \mbox{h.c.} \Big{)}\bigg{]} + \dots = 0~. \end{eqnarray} \noindent Let us rewrite it in terms of the dimensionless quantity $x\equiv\lambda/M^2$. Then, one can transform Eq.~(\ref{A3}) as \begin{equation}\label{A4} \Big{[}\prod_f (x - x^{(0)}_f)\Big{]} (x - x^{(0)}_4)(x - x^{(0)}_5) = \epsilon P_4(x)~, \end{equation} \noindent where \begin{equation}\label{A5} \epsilon = \frac{1}{M^2} \Big{(}\Sigma\vert\mu^f\vert^2 + \Sigma\vert\mu'_f\vert^2 + \vert m_4\vert^2 + \vert m_5\vert^2 \Big{)} \end{equation} is the small paremeter ($\epsilon = {\cal O}(v^2/M^2)$) and $x^{(0)}_f\equiv m_f^2/M^2 = {\cal O}(\epsilon)$, $x^{(0)}_4 = 1$, $x^{(0)}_5 = {M'}^2/M^2$ are the zeroth order roots. The fourth power polynomial $P_4(x) = (x^4 + \dots)$ has coefficients ${\cal O}(1)$ or less. The dropped out terms corresponding to dots in Eq.~(\ref{A3}) result in the relative corrections ${\cal O}(\epsilon^2)$, and hence they can be omitted in our approximation. Iterating Eq.~(\ref{A4}) one arrives at the roots of Eq.~(\ref{eq:lambda}). The elements of the $U_{1L}$ matrix (with the indices $\kappa = u$, $d$ being suppressed) are as follows \begin{eqnarray}\label{A6} &&U_{1L}{}^f_g = \delta^f_g\bigg{(} 1 - \frac{1}{2 M^2}n^f_f\bigg{)} + (\delta^f_g - 1) \frac{1}{M^2} p^f_g~,\nonumber\\ \vspace{0.5ex} &&U_{1L}{}^f_4 = \frac{1}{M^2} p^f_4~~,~~ U_{1L}{}^f_5 = \frac{1}{M} p^f_5~,\nonumber\\ \vspace{0.5ex} &&U_{1L}{}_f^4 = \frac{1}{M^2} p^4_f~~,~~ U_{1L}{}_f^5 = \frac{1}{M} p^5_f~,\nonumber\\ \vspace{0.5ex} &&U_{1L}{}_5^4 = \frac{1}{M} p^4_5~~,~~ U_{1L}{}_4^5 = \frac{1}{M} p^5_4~,\nonumber\\ \vspace{0.5ex} &&U_{1L}{}_4^4 = 1 - \frac{1}{2M^2} n_4^4~~,~~ U_{1L}{}_5^5 = 1 - \frac{1}{2M^2} n_5^5~,\nonumber\\ \end{eqnarray} and \begin{eqnarray}\label{A7} &&U_{1R}{}^f_g = \delta^f_g\bigg{(} 1 - \frac{1}{2{M'}^2} {n'}^f_f\bigg{)} + (\delta^f_g - 1) \frac{1}{{M'}^2} {p'}^f_g~,\nonumber\\ \vspace{0.5ex} &&U_{1R}{}^f_4 = \frac{1}{{M'}^2} {p'}^f_4~~,~~ U_{1R}{}^f_5 = \frac{1}{M'} {p'}^f_5~,\nonumber\\ \vspace{0.5ex} &&U_{1R}{}_f^4 = \frac{1}{{M'}^2} {p'}^5_f~~,~~ U_{1R}{}_f^5 = \frac{1}{M'} {p'}^4_f~,\nonumber\\ \vspace{0.5ex} &&U_{1R}{}_4^4 = \frac{1}{M'} {p'}^5_4~~,~~ U_{1R}{}_5^5 = \frac{1}{M'} {p'}^4_5~,\nonumber\\ &&U_{1R}{}_5^4 = 1 - \frac{1}{2{M'}^2} {n'}_5^5~~,~~ U_{1R}{}_4^5 = 1 - \frac{1}{2{M'}^2} {n'}_4^4~, \end{eqnarray} where \begin{eqnarray}\label{A8} p^f_g &=& \frac{\mu^f (m_f^2-\vert m_5\vert^2)(m_f {\mu^f}^*\mu'_g - m_g {\mu^g}^*\mu'_f) + k_f (m_f\mu'_g - \frac{m_g}{m_f}\frac{M'}{M} {\mu^g}^* m_5^*)} {(m_g^2-m_f^2)(m_f\mu'_f - \frac{M'}{M} m_5^* {\mu^f}^*)}~,\nonumber\\ \vspace{0.5ex}\nonumber\\ p^f_4 &=& - k_f ~\frac{\Big(k_f+\vert\mu^f\vert^2 (m_f^2-\vert m_5\vert^2) \Big)\Big(\frac{M'}{M} m_5^* + \frac{1}{k_f} m_f\mu^f\mu'_f (m_f^2-\vert m_5\vert^2)\Big)} {m_f(m_f\mu'_f - \frac{M'}{M} m_5^* {\mu^f}^*)}~,\nonumber\\ \vspace{0.5ex}\nonumber\\ p_f^4 &=& m_f {\mu^f}^* - \frac{\mu'_f (\rho + \vert m_5\vert^2)} {m_4 + \frac{M'}{M} m_5^*} ~,\nonumber\\ \vspace{0.5ex}\nonumber\\ p^f_5 &=& \frac{\frac{M'}{M} (k_f+m_f^2\vert\mu^f\vert^2) - m_f m_5\mu^f\mu'_f} {m_f(m_f\mu'_f - \frac{M'}{M} m_5^* {\mu^f}^*)}~~~,~~~ p_f^5 = \frac{M}{M'} \mu'_f~,\nonumber\\ \vspace{0.5ex}\nonumber\\ p_5^4 &=& \frac{m_4 m_5 - \frac{M'}{M}\rho}{m_4 + \frac{M'}{M} m_5^*}~~~,~~~ p_4^5 = \frac{MM'}{{M'}^2-M^2} \Big{(} m_4 + \frac{M}{M'} m_5^*\Big{)} ~,\nonumber\\ \vspace{0.5ex}\nonumber\\ n^f_f &=& \bigg{\vert} \frac{\frac{M'}{M} (k_f+m_f^2\vert\mu^f \vert^2) - m_f m_5\mu^f\mu'_f} {m_f(m_f\mu'_f - \frac{M'}{M} m_5^* {\mu^f}^*)} \bigg{\vert}^2~,\nonumber\\ \vspace{0.5ex}\nonumber\\ n^4_4 &=& \bigg{\vert} \frac{m_4 m_5 - \frac{M'}{M} \rho} {m_4 + \frac{M'}{M} m_5^*} \bigg{\vert}^2~,\nonumber\\ \vspace{0.5ex}\nonumber\\ n^5_5 &=& \Big{\vert} \frac{{M'}^2}{{M'}^2-M^2} (m_4 + \frac{M}{M'} m_5^*) \Big{\vert}^2 + \Sigma \vert\mu'_f\vert^2~ \end{eqnarray} and $k_f = M^2 (\overline{m}_f^2 - m_f^2)$, $\rho = M^2 + \Sigma{\vert\mu^f\vert}^2 - \overline{ M}_4^2$. The $p'$, $n'$ are obtained from $p$, $n$, respectively, by substituting $\mu^f\leftrightarrow {\mu'_f}^*$, $m_4\leftrightarrow m_4^*$, $m_5\leftrightarrow m_5^*$, $M\leftrightarrow M'$. All these auxiliary parameters are in general of order ${\cal O}(M^0)$. The elements of the matrix $U_{1R}$ are obtained from those for $U_{1L}$ by the same substitution followed by changing column indices $4\leftrightarrow 5$ for the matrix elements $(U_{1L})^4_A$ and $(U_{1L})^5_A$. Hereof one gets for the charged current matrix $V_L = {V_0}_L + \Delta V_L$ \def{p^u}'{}{{p^u}{}} \def{p^d}'{}{{p^d}{}} \begin{equation}\label{A9} \begin{array}{l} \Delta V_L = \\ \\ \left({ \begin{array}{c|cc} - \frac{1}{M^2} \sum \Big{(} {{p^u}'{}^f_h}^* {V_C}^g_h + {V_C}_f^h {p^d}'{}^g_h \Big{)}& \frac{1}{M^2} \Big{(} \sum {V_C}_f^h {p^d}'{}^4_h + {{p^u}'{}^f_4}^* \Big{)}& \frac{1}{M} \sum {V_C}_f^h {p^d}'{}^5_h \\ - \frac{1}{2M^2} ({n^u}{}^f_f + {n^d}{}^g_g) {V_C}_f^g& &\\ \cline{1-3} &&\\ \frac{1}{M^2} \Big{(} \sum {{p^u}'{}^4_h}^* {V_C}_h^g + {p^d}'{}^g_4 \Big{)}& - \frac{1}{2M^2} ({n^d}^4_4 + {n^u}^4_4)& \frac{1}{M} {p^d}'{}^5_4 \\ &&\\ \frac{1}{M} \sum {{p^u}'{}^5_h}^* {V_C}_h^g& \frac{1}{M} {{p^u}'{}^5_4}^*& \frac{1}{M^2} \Big{(} \sum {{p^u}'{}^5_h}^* {p^d}'{}^5_k {V_C}^k_h\\ & & + {{p^u}'{}^5_4}^* {p^d}'{}^5_4 \Big{)} \end{array} }\right)\,, \\ \\ \end{array} \end{equation} with ${V_0}_L$ from Eq.~(\ref{eq:V_0L}) and similarly for $V_R = {V_0}_R + \Delta V_R$ with ${V_0}_R = \mbox{diag}~ (0,0,0,1,0)$ and \def{p^u}'{}{{p^u}'{}} \def{p^d}'{}{{p^d}'{}} \begin{equation}\label{A10} \begin{array}{l} \Delta V_R = \left({ \begin{array}{c|cc} \frac{1}{{M^u}' {M^d}'} {{{p^u}'{}}^f_5}^* {{p^d}'{}}^g_5& \frac{1}{{M^u}'} {{p^u}'{}^f_5}^*& \frac{1}{{M^u}' {M^d}'} {{{p^u}'{}}^f_5}^* {{p^d}'{}}^4_5\\ &&\\ \cline{1-3} &&\\ \frac{1}{{M^d}'} {{p^d}'{}}^g_5& - \frac{1}{2{{M^u}'}^2} {{n'}^u}{}^5_5 - \frac{1}{2{{M^d}'}^2} {{n'}^d}{}^5_5& \frac{1}{{M^d}'} {{p^d}'{}}^4_5 \\ &&\\ \frac{1}{{M^u}' {M^d}'} {{{p^u}'{}}^4_5}^* {{p^d}'{}}^g_5& \frac{1}{{M^u}'} {{{p^u}'{}}^4_5}^*& \frac{1}{{M^u}' {M^d}'} {{{p^u}'{}}^4_5}^* {{p^d}'{}}^4_5 \end{array} }\right)~~. \\ \end{array} \end{equation} For the neutral current matrices ($\kappa = u, d$ being suppressed everywhere below) one gets \defp{p} \begin{equation}\label{A11} \begin{array}{ll} X_L = X_{0L} - \left({ \begin{array}{ccc} \frac{1}{M^2} {p^f_5}^* p^g_5& \frac{1}{M^2} {p^f_5}^* p^4_5& \frac{1}{M} {p^f_5}^* \\ &&\\ \frac{1}{M^2} {p^4_5}^* p^g_5& \frac{1}{M^2} \vertp^4_5\vert^2& \frac{1}{M} {p^4_5}^*\\ &&\\ \frac{1}{M} p^g_5& \frac{1}{M} p^4_5& - \frac{1}{M^2} n^5_5 \end{array} }\right)~~, \\ \end{array} \end{equation} and \def{p'}{}{{p'}{}} \begin{equation}\label{A12} \begin{array}{l} X_R = X_{0R} + \left({ \begin{array}{ccc} \frac{1}{{M'}^2} {{p'}{}^f_5}^* {p'}{}^g_5& \frac{1}{M'} {{p'}{}^f_5}^*& \frac{1}{{M'}^2} {{p'}{}^f_5}^* {p'}{}^4_5\\ &&\\ \frac{1}{M'} {p'}{}^g_5& - \frac{1}{{M'}^2} {n'}^5_5& \frac{1}{M'} {p'}{}^4_5 \\ &&\\ \frac{1}{{M'}^2} {{p'}{}^4_5}^* {p'}{}^g_5& \frac{1}{M'} {{p'}{}^4_5}^*& \frac{1}{{M'}^2} \vert{p'}{}^4_5\vert^2 \end{array} }\right)~~, \\ \end{array} \end{equation} with $X_{0L} = I_L$ and $X_{0R} = \mbox{diag}~ (0,0,0,1,0)$. Finally, for the Higgs mediated neutral current matrix ${\cal H} = {\cal H}_0 + \Delta {\cal H}$ one has \def\kappa{{}} \begin{equation}\label{A13} {\cal H}_0 = ~ \left( { \begin{array}{ccc} \overline{m}^\kappa_f\delta^g_f& 0& - \frac{M'}{M} {p^f_5}^*\\ &&\\ - \frac{M}{M'} {p'}^g_5& 0& - \frac{M'}{M} {p^4_5}^* - \frac{M}{M'} {p'}^4_5\\ &&\\ 0& - ({p^5_4}^* + {p'}^5_4)& 0\\ \end{array} }\right) \end{equation} and \begin{equation}\label{A14} \begin{array}{l} \Delta {\cal H} = \\ \\ \left( { \begin{array}{c|cc} - \frac{1}{MM'} \bigg{(} {p^f_4}^* {p'}^g_5 + {p^f_5}^* {p'}^g_4\bigg{)}& - \frac{1}{M} \bigg{(} {p^f_5}^* {p'}^5_4 + {p^f_4}^*\bigg{)}& \frac{1}{MM'} \bigg{(} \frac{1}{2} {p^f_5}^* {{n'}^\kappa}^4_4 - {p^f_4}^* {p'}^4_5 \bigg{)}\\ \cline{1-3} &&\\ \frac{1}{MM'} \bigg{(} \frac{1}{2} n_4^4 {p'}^g_5 - {p^4_5}^* {p'}^g_4 \bigg{)}& - \frac{1}{2M} (\rho - \Sigma\vert\mu^f\vert^2)& \frac{1}{2MM'} \bigg{(} n_4^4 {p'}^4_5 + {n'}^4_4 {p^4_5}^* \bigg{)}\\ & + \frac{1}{2M} n^4_4 + \frac{M}{2{M'}^2} {n'}_5^5& \\ & - \frac{1}{M} {p^4_5}^* {p'}^5_4& \\ - \frac{1}{M'} \bigg{(} {p^5_4}^* {p'}^g_5 + {p'}^g_4 \bigg{)}& \frac{1}{2{M'}^2} {n'}^5_5 {p^5_4}^*& - \frac{1}{2M'} (\rho' - \Sigma\vert\mu'_f\vert^2)\\ & + \frac{1}{2M^2} n^5_5 {p'}^5_4 & + \frac{M'}{2M^2} n^5_5 + \frac{1}{2M'} {n'}^4_4\\ & & - \frac{1}{M'} {p^5_4}^* {p'}^4_5 \end{array} }\right)\,, \\ \\ \end{array} \end{equation} where $\rho$ is defined above in Appendix, and $\rho'$ can be obtained from $\rho$ by the usual substitutions $\mu^f\leftrightarrow {\mu'_f}^*$, $m_4\leftrightarrow m_4^*$, $m_5\leftrightarrow m_5^*$, $M\leftrightarrow M'$.
\section{Baryogenesis and lepton number violation} The cosmological matter antimatter asymmetry, the ratio of the baryon density to the entropy density of the universe, \begin{equation} Y_B = {n_B-n_{\Bar{B}}\over s} = (0.6 - 1)\cdot 10^{-10}\;, \label{blub} \end{equation} can in principle be understood in theories where baryon number, C and CP are not conserved \cite{sac}. The presently observed value of the baryon asymmetry is then explained as a consequence of the spectrum and interactions of elementary particles, together with the cosmological evolution. A crucial ingredient of baryogenesis is the connection between baryon number ($B$) and lepton number ($L$) in the high-temperature, symmetric phase of the standard model. Due to the chiral nature of the weak interactions $B$ and $L$ are not conserved\cite{thoo}. At zero temperature this has no observable effect due to the smallness of the weak coupling. However, as the temperature approaches the critical temperature $T_{EW}$ of the electroweak phase transition, $B$ and $L$ violating processes come into thermal equilibrium\cite{krs}. These `sphaleron processes' violate baryon and lepton number by three units, \begin{equation} \Delta B = \Delta L = 3\;. \label{sphal1} \end{equation} It is generally believed that $B$ and $L$ changing processes are in thermal equilibrium for temperatures in the range \begin{equation} T_{EW} \sim 100\ \mbox{GeV} < T < T_{SPH} \sim 10^{12}\ \mbox{GeV}\;. \end{equation} The non-conservation of baryon and lepton number has a profound effect on the generation of the cosmological baryon asymmetry. Eq.~\ref{sphal1} suggests that any $B+L$ asymmetry generated before the electroweak phase transition, i.e., at temperatures $T>T_{EW}$, will be washed out. However, since only left-handed fields couple to sphalerons, a non-zero value of $B+L$ can persist in the high-temperature, symmetric phase if there exists a non-vanishing $B-L$ asymmetry. An analysis of the chemical potentials of all particle species in the high-temperature phase yields the following relation between the baryon asymmetry $Y_B$ and the corresponding $L$ and $B-L$ asymmetries $Y_L$ and $Y_{B-L}$, respectively\cite{chem}, \begin{equation}\label{basic} Y_B\ =\ a\ Y_{B-L}\ =\ {a\over a-1}\ Y_L\;, \end{equation} where $a$ is a number ${\mathcal O}(1)$. In the standard model with three generations and two Higgs doublets one has $a=8/23$. We conclude that $B-L$ violation is needed if the baryon asymmetry is generated before the electroweak transition, i.e. at temperatures $T > T_{EW} \sim 100$~GeV. In the standard model, as well as its supersymmetric version and its unified extensions based on the gauge group SU(5), $B-L$ is a conserved quantity. Hence, no baryon asymmetry can be generated dynamically in these models. The remnant of lepton number violation at low energies is an effective $\Delta L=2$ interaction between lepton and Higgs fields, \begin{equation}\label{dl2} {\mathcal L}_{\Delta L=2} = {1\over 2} f_{ij}\ l_{Li}^T H_2\ C\ l_{Lj} H_2 +\mbox{ h.c.}\;.\label{intl2} \end{equation} Such an interaction arises in particular from the exchange of heavy Majorana neutrinos. In the Higgs phase of the standard model, where the Higgs field acquires a vacuum expectation value $\VEV{H_2}=v_2$, it gives rise to Majorana masses of the light neutrinos $\nu_e$, $\nu_\mu$ and $\nu_\tau$. At finite temperature the $\Delta L=2$ processes described by (\ref{dl2}) take place with the rate\cite{fy1} \begin{equation} \Gamma_{\Delta L=2} (T) = {1\over \pi^3}\,{T^3\over v_2^4}\, \sum_{i=e,\mu,\tau} m_{\nu_i}^2\; . \end{equation} In thermal equilibrium this yields an additional relation between the chemical potentials which implies \begin{equation} Y_B\ =\ Y_{B-L}\ =\ Y_L\ =\ 0 \; . \end{equation} To avoid this conclusion, the $\Delta L=2$ interaction (\ref{intl2}) must not reach thermal equilibrium. For baryogenesis at a temperature $T_B < T_{SPH} \sim 10^{12}$ GeV, one has to require $\Gamma_{\Delta L=2} < H|_{T_B}$, where $H$ is the Hubble parameter. This yields a stringent upper bound on Majorana neutrino masses, \begin{equation}\label{nbound} \sum_{i=e,\mu,\tau} m_{\nu_i}^2 < \left(0.2\;\mbox{eV}\, \left({T_{SPH}\over T_B}\right)^{1/2}\,\right)^2\;. \end{equation} For $T_B \sim T_{SPH}$, this bound would be comparable to the upper bound on the electron neutrino mass obtained from neutrinoless double beta decay. However, eq.~(\ref{nbound}) also applies to the $\tau$-neutrino mass. Note, that the bound can be evaded if appropriate asymmetries are present for particles which reach thermal equilibrium only at temperatures below $T_B$ \cite{cli93}. The connection between lepton number and the baryon asymmetry is lost if baryogenesis takes place at or below the Fermi scale\cite{dol}. However, detailed studies of the thermodynamics of the electroweak transition have shown that, at least in the standard model, the deviation from thermal equilibrium is not sufficient for baryogenesis\cite{jansen}. In the minimal supersymmetric extension of the standard model (MSSM) such a scenario appears still possible for a limited range of parameters\cite{dol}. \section{Decays of heavy Majorana neutrinos} Baryogenesis above the Fermi scale requires $B-L$ violation, and therefore $L$ violation. Lepton number violation is most simply realized by adding right-handed Majorana neutrinos to the standard model. Heavy right-handed Majorana neutrinos, whose existence is predicted by all extensions of the standard model containing $B-L$ as a local symmetry, can also explain the smallness of the light neutrino masses via the see-saw mechanism\cite{seesaw}. The most general Lagrangian for couplings and masses of charged leptons and neutrinos reads \begin{equation}\label{yuk} {\mathcal L}_Y = -h_{e ij}\Bar{e_R}_i l_{Lj} H_1 -h_{\nu ij}\Bar{\nu_R}_i l_{Lj} H_2 -{1\over2}h_{r ij} \Bar{\nu^c_R}_i \nu_{Rj} R +\mbox{ h.c.}\;. \end{equation} The vacuum expectation values of the Higgs field $\VEV{H_1}=v_1$ and $\VEV{H_2}=v_2=\tan{\beta}\ v_1$ generate Dirac masses $m_l$ and $m_D$ for charged leptons and neutrinos, $m_e=h_e v_1$ and $m_D=h_{\nu}v_2$, respectively, which are assumed to be much smaller than the Majorana masses $M = h_r\VEV{R}$. This yields light and heavy neutrino mass eigenstates \begin{equation} \nu\simeq K^{\dg}\nu_L+\nu_L^c K\quad,\qquad N\simeq\nu_R+\nu_R^c\, , \end{equation} with masses \begin{equation} m_{\nu}\simeq- K^{\dg}m_D{1\over M}m_D^T K^*\, \quad,\quad m_N\simeq M\, . \label{seesaw} \end{equation} Here $K$ is a unitary matrix which relates weak and mass eigenstates. The right-handed neutrinos, whose exchange may erase any lepton asymmetry, can also generate a lepton asymmetry by means of out-of-equilibrium decays. This lepton asymmetry is then partially transformed into a baryon asymmetry by sphaleron processes\cite{fy}. The decay width of the heavy neutrino $N_i$ reads at tree level, \begin{equation} \Gamma_{Di}=\Gamma\left(N_i\to H_2+l\right)+\Gamma\left(N_i\to H_2^c+l^c\right) ={1\over8\pi}(h_\nu h_\nu^\dg)_{ii} M_i\;. \label{decay} \end{equation} From the decay width one obtains an upper bound on the light neutrino masses via the out-of-equilibrium condition\cite{fisch}. Requiring $\Gamma_{D1}< H|_{T=M_1}$ yields the constraint \begin{equation}\label{ooeb} \wt{m}_1\ =\ (h_\nu h_\nu^\dg)_{11} {v_2^2\over M_1}\ < \ 10^{-3}\, \mbox{eV}\;. \end{equation} More direct bounds on the light neutrino masses depend on the structure of the Dirac neutrino mass matrix. Interference between the tree-level amplitude and the one-loop self-energy and vertex corrections yields $CP$ asymmetries in the heavy Majorana neutrino decays. In a basis, where the right-handed neutrino mass matrix $M = h_r\VEV{R}$ is diagonal, one obtains \cite{cov,bp2} \begin{eqnarray} \varepsilon_1&=&{\Gamma(N_1\rightarrow l \, H_2)-\Gamma(N_1\rightarrow l^c \, H_2^c) \over \Gamma(N_1\rightarrow l \, H_2)+\Gamma(N_1\rightarrow l^c \, H_2^c)} \nonumber\\[1ex] &\simeq&-{3\over16\pi}\;{1\over\left(h_\nu h_\nu^\dg\right)_{11}} \sum_{i=2,3}\mbox{Im}\left[\left(h_\nu h_\nu^\dg\right)_{1i}^2\right] {M_1\over M_i}\label{cpa}\;. \end{eqnarray} Here we have assumed $M_1 < M_2,M_3$, which is satisfied in the applications considered in the following sections. In the early universe at temperatures $T \sim M_1$ the CP asymmetry (\ref{cpa}) leads to a lepton asymmetry\cite{kw}, \begin{equation}\label{basym} Y_L\ =\ {n_L-n_{\Bar{L}}\over s}\ =\ \kappa\ {\varepsilon_1\over g_*}\;. \end{equation} Here the factor $\kappa<1$ represents the effect of washout processes. In order to determine $\kappa$ one has to solve the full Boltzmann equations \cite{lut92,plu97}. In the examples discussed below one has $\kappa\simeq 10^{-1}\ldots 10^{-3}$. \section{Neutrino masses and mixings} The CP asymmetry (\ref{cpa}) is given in terms of the Dirac and the Majorana neutrino mass matrices. Depending on the neutrino mass hierarchy and the size of the mixing angles the CP asymmetry can vary over many orders of magnitude. It is therefore interesting to see whether a pattern of neutrino masses motivated by other considerations is consistent with leptogenesis. An attractive framework to explain the observed mass hierarchies of quarks and charged leptons is the Froggatt-Nielsen mechanism \cite{fro79} based on a spontaneously broken U(1)$_F$ generation symmetry. The Yukawa couplings arise from non-renormalizable interactions after a gauge singlet field $\Phi$ acquires a vacuum expectation value, \begin{equation} h_{ij} = g_{ij} \left({\VEV\Phi\over \Lambda}\right)^{Q_i + Q_j}\;. \end{equation} Here $g_{ij}$ are couplings ${\mathcal O}(1)$ and $Q_i$ are the U(1) charges of the various fermions with $Q_{\Phi}=-1$. The interaction scale $\Lambda$ is expected to be very large, $\Lambda > \Lambda_{GUT}$. In the following we shall discuss two different realizations of this idea which are motivated by the recently reported atmospheric neutrino anomaly \cite{atm98}. Both scenarios have a large $\nu_\mu -\nu_\tau$ mixing angle. They differ, however, by the symmetry structure and by the size of the parameter $\epsilon$ which characterizes the flavour mixing. \subsection{$SU(5)\times U(1)_F$} This symmetry has been considered by a number of authors \cite{lol99}. Particularly interesting is the case with a nonparallel family structure where the chiral $U(1)_F$ charges are different for the $\bf 5^*$-plets and the $\bf 10$-plets of the same family \cite{sat98}-\cite{bij87}. An example of possible charges $Q_i$ is given in table~1. \begin{table}[b] \begin{center} \begin{tabular}{c|ccccccccc}\hline \hline $\psi_i$ & $ e^c_{R3}$ & $ e^c_{R2}$ & $ e^c_{R1}$ & $ l_{L3}$ & $ l_{L2}$ & $ l_{L1}$ & $ \nu^c_{R3}$ & $ \nu^c_{R2}$ & $ \nu^c_{R1}$ \\\hline $Q_i$ & 0 & 1 & 2 & $0$ & $0$ & $1$ & 0 & $1$ & $2$ \\ \hline\hline \end{tabular} \end{center} \caption{{\it Chiral charges of charged and neutral leptons with $SU(5)\times U(1)_F$ symmetry} \cite{buc99}.} \end{table} The assignment of the same charge to the lepton doublets of the second and third generation leads to a neutrino mass matrix of the form \cite{sat98,ram98}, \begin{equation}\label{matrix} m_{\nu_{ij}} \sim \left(\begin{array}{ccc} \epsilon^2 & \epsilon & \epsilon \\[-1ex] \epsilon & \; 1 \; & 1 \\[-1ex] \epsilon & 1 & 1 \end{array}\right) {v_2^2\over \VEV R}\;. \end{equation} This structure immediately yields a large $\nu_\mu -\nu_\tau$ mixing angle. The phenomenology of neutrino oscillations depends on the unspecified coefficients ${\mathcal O}(1)$. The parameter $\epsilon$ which gives the flavour mixing is chosen to be \begin{equation}\label{exp1} {\VEV\Phi\over\Lambda} = \epsilon \sim {1\over 17}\;. \end{equation} The three Yukawa matrices for the leptons have the structure, \begin{equation}\label{yuk1} h_e \sim\ \left(\begin{array}{ccc} \epsilon^3 & \epsilon^2 & \epsilon^2 \\[-1ex] \epsilon^2 &\; \epsilon \; & \epsilon \\[-1ex] \epsilon & 1 & 1 \end{array}\right) \;, \quad h_{\nu} \sim\ \left(\begin{array}{ccc} \epsilon^3 & \epsilon^2 & \epsilon^2 \\[-1ex] \epsilon^2 &\; \epsilon \; & \epsilon \\[-1ex] \epsilon & 1 & 1 \end{array}\right) \;, \quad h_{r} \sim\ \left(\begin{array}{ccc} \epsilon^4 & \epsilon^3 & \epsilon^2 \\[-1ex] \epsilon^3 &\; \epsilon^2 \; & \epsilon \\[-1ex] \epsilon^2 & \epsilon & 1 \end{array}\right) \;. \end{equation} Note, that $h_e$ and $h_\nu$ have the same, non-symmetric structure. One easily verifies that the mass ratios for charged leptons, heavy and light Majorana neutrinos are given by \begin{eqnarray} \qquad\quad m_e : m_\mu : m_\tau \sim \epsilon^3 : \epsilon : 1\;, &\quad& M_1 : M_2 : M_3 \sim \epsilon^4 : \epsilon^2 : 1\;,\\ m_1 : m_2 : m_3 &\sim& \epsilon^2 : 1 : 1\;. \end{eqnarray} The masses of the two eigenstates $\nu_\mu$ and $\nu_\tau$ depend on unspecified factors of order one, and may easily differ by an order of magnitude \cite{irg98}. They can therefore be consistent with the mass differences $\Delta m^2_{\nu_e \nu_\mu}\simeq 4\cdot 10^{-6} - 1\cdot 10^{-5}$~eV$^2$ \cite{sol98} inferred from the MSW solution of the solar neutrino problem \cite{msw86} and $\Delta m^2_{\nu_\mu \nu_\tau}\simeq (5\cdot 10^{-4}-6\cdot 10^{-3})$~eV$^2$ associated with the atmospheric neutrino deficit \cite{atm98}. For numerical estimates we shall use the average of the neutrino masses of the second and third family, $\Bar{m}_\nu=(m_{\nu_\mu}m_{\nu_\tau})^{1/2} \sim 10^{-2}$~eV. Note, that for a different choice of U(1) charges the coefficients in eq.~(\ref{matrix}) automatically yield the hierarchy $m_2/m_3 \sim \epsilon^{2/3}$ \cite{alt98}. \begin{table}[b] \begin{center} \begin{tabular}{c|ccccccccc} \hline \hline $\psi_i$ & $ e^c_{R3}$ & $ e^c_{R2}$ & $ e^c_{R1}$ & $ l_{L3}$ & $ l_{L2}$ & $ l_{L1}$ & $ \nu^c_{R3}$ & $ \nu^c_{R2}$ & $ \nu^c_{R1}$ \\\hline $Q_i$ & 0 & ${1\over 2}$ & ${5\over 2}$ & $0$ & ${1\over 2}$ & ${5\over 2}$ & 0 & ${1\over 2}$ & ${5\over 2}$ \\ \hline\hline \end{tabular} \end{center} \caption{{\it Chiral charges of charged and neutral leptons with $SU(3)_c \times SU(3)_L \times SU(3)_R \times U(1)_F$ symmetry} \cite{lol99}.} \end{table} The choice of the charges in table~1 corresponds to large Yukawa couplings of the third generation. For the mass of the heaviest Majorana neutrino one finds \begin{equation} M_3\ \sim\ {v_2^2\over\Bar{m}_\nu}\ \sim\ 10^{15}\ \mbox{GeV}\;. \end{equation} This implies that $B-L$ is broken at the unification scale $\Lambda_{GUT}$. \subsection{$SU(3)_c \times SU(3)_L \times SU(3)_R \times U(1)_F$} This symmetry arises in unified theories based on the gauge group $E_6$. The leptons $e_R^c$, $l_L$ and $\nu_R^c$ are contained in a single $(1,3,\bar{3})$ representation. Hence, all leptons of the same generation have the same $U(1)_F$ charge and all leptonic Yukawa matrices are symmetric. Masses and mixings of quarks and charged leptons can be successfully described by using the charges given in table~2 \cite{lol99}. Clearly, the three Yukawa matrices have the same structure\footnote{Note, that with respect to ref.~\cite{lol99}, $\epsilon$ and $\Bar{\epsilon}$ have been interchanged.}, \begin{equation}\label{yuk2} h_e,\ h_r \sim\ \left(\begin{array}{ccc} \epsilon^5 & \epsilon^3 & \epsilon^{5/2} \\ \epsilon^3 & \epsilon & \epsilon^{1/2} \\ \epsilon^{5/2} \; & \epsilon^{1/2} \; & 1 \end{array}\right) \;, \quad h_{\nu} \sim\ \left(\begin{array}{ccc} \bar\epsilon^5 &\bar\epsilon^3 & \bar\epsilon^{5/2} \\ \bar\epsilon^3 &\bar\epsilon & \bar\epsilon^{1/2} \\ \bar\epsilon^{5/2} \; &\bar\epsilon^{1/2} \; & 1 \end{array}\right) \;. \end{equation} Note, that the expansion parameter in $h_{\nu}$ is different from the one in $h_e$ and $h_r$. From the quark masses, which also contain $\epsilon$ and $\bar{\epsilon}$, one infers $\bar{\epsilon} \simeq \epsilon^2$ \cite{lol99}. From eq.~(\ref{yuk2}) one obtains for the masses of charged leptons, light and heavy Majorana neutrinos, \begin{equation} m_e : m_\mu : m_\tau\ \sim\ M_1 : M_2 : M_3\ \sim \epsilon^5 : \epsilon : 1\;, \end{equation} \begin{equation} m_1 : m_2 : m_3 \ \sim\ \epsilon^{15} : \epsilon^3 : 1\;. \end{equation} Like in the example with $SU(5)\times U(1)_F$ symmetry, the mass of the heaviest Majorana neutrino, \begin{equation} M_3 \sim {v_2^2\over m_3} \sim 10^{15}\;\mbox{GeV} \;, \end{equation} implies that $B-L$ is broken at the unification scale $\Lambda_{GUT}$. The $\nu_\mu-\nu_\tau$ mixing angle is mostly given by the mixing of the charged leptons of the second and third generation \cite{lol99}, \begin{equation} \sin{\Theta_{\mu\tau}} \sim \sqrt{\epsilon} + \epsilon \;. \end{equation} This requires large flavour mixing, \begin{equation}\label{exp2} \left({\VEV\Phi\over\Lambda}\right)^{1/2} = \sqrt{\epsilon} \sim {1\over 2}\;. \end{equation} In view of the unknown coefficients ${\mathcal O}(1)$ the corresponding mixing angle $\sin{\Theta_{\mu\tau}} \sim 0.7$ is consistent with the interpretation of the atmospheric neutrino anomaly as $\nu_\mu-\nu_\tau$ oscillation. It is very instructive to compare the two scenarios of lepton masses and mixings described above. In the first case, the large $\nu_\mu-\nu_\tau$ mixing angle follows from a nonparallel flavour symmetry. The parameter $\epsilon$, which characterizes the flavour mixing, is small. In the second case, the large $\nu_\mu-\nu_\tau$ mixing angle is a consequence of the large flavour mixing $\epsilon$. The $U(1)_F$ charges of all leptons are the same, i.e., one has a parallel family structure. Also the mass hierarchies, given in terms of $\epsilon$, are rather different. This illustrates that the separation into a flavour mixing parameter $\epsilon$ and coefficients ${\mathcal O}(1)$ is far from unique. It is therefore important to study other observables which depend on the lepton mass matrices. A particular example is the baryon asymmetry. \section{Matter antimatter asymmetry} We can now evaluate the baryon asymmetry for the two patterns of neutrino mass matrices discussed in the previous section. A rough estimate of the baryon asymmetry can be obtained from the CP asymmetry $\varepsilon_1$ of the heavy Majorana neutrino $N_1$. A quantitative determination requires a numerical study of the full Boltzmann equations \cite{plu97}. \subsection{$SU(5)\times U(1)_F$} In this case one obtains from eqs.~(\ref{cpa}) and (\ref{yuk1}), \begin{equation} \varepsilon_1\ \sim\ {3\over 16\pi}\ \epsilon^4\;. \end{equation} From eq.~(\ref{basym}), $\epsilon^2 \sim 1/300$ (\ref{exp1}) and $g_* \sim 100$ one then obtains the baryon asymmetry, \begin{equation}\label{est1} Y_B \sim \kappa\ 10^{-8}\;. \end{equation} For $\kappa \sim 0.1\ldots 0.01$ this is indeed the correct order of magnitude. The baryogenesis temperature is given by the mass of the lightest of the heavy Majorana neutrinos, \begin{equation} T_B \sim M_1 \sim \epsilon^4 M_3 \sim 10^{10}\ \mbox{GeV}\;. \end{equation} This set of parameters, where the CP asymmetry is given in terms of the mass hierarchy of the heavy neutrinos, has been studied in detail \cite{buc96}. The generated baryon asymmetry does not depend on the flavour mixing of the light neutrinos. The $\nu_\mu-\nu_\tau$ mixing angle is large in the scenario described in the previous section whereas it was assumed to be small in \cite{buc96}. \begin{figure}[t] \mbox{ }\hfill \epsfig{file=SimplestChoice.eps,width=8.2cm} \hfill\mbox{ } \caption{\it Time evolution of the neutrino number density and the lepton asymmetry in the case of the $SU(5)\times U(1)_F$ symmetry. The solid line shows the solution of the Boltzmann equation for the right-handed neutrinos, while the corresponding equilibrium distribution is represented by the dashed line. The absolute value of the lepton asymmetry $Y_L$ is given by the dotted line and the hatched area shows the lepton asymmetry corresponding to the observed baryon asymmetry. \label{asyB}} \end{figure} The solution of the full Boltzmann equations is shown in fig.~\ref{asyB} for the non-supersymmetric case \cite{buc96}. The initial condition at a temperature $T \sim 10 M_1$ is chosen to be a state without heavy neutrinos. The Yukawa interactions are sufficient to bring the heavy neutrinos into thermal equilibrium. At temperatures $T\sim M_1$ this is followed by the usual out-of-equilibrium decays which lead to a non-vanishing baryon asymmetry. The final asymmetry agrees with the estimate (\ref{est1}) for $\kappa \sim 0.1$. The change of sign in the lepton asymmetry is due to the fact that inverse decay processes, which take part in producing the neutrinos, are $CP$ violating, i.e.\ they generate a lepton asymmetry at high temperatures. Due to the interplay of inverse decay processes and lepton number violating scattering processes this asymmetry has a different sign than the one produced by neutrino decays at lower temperatures. \subsection{$SU(3)_c\times SU(3)_L\times SU(3)_R\times U(1)_F$} \begin{figure}[t] \mbox{ }\hfill \epsfig{file=LolaRoss1.eps,width=8.2cm} \hfill\mbox{ } \caption{\it Solution of the Boltzmann equations in the case of the $SU(3)_c\times SU(3)_L\times SU(3)_R\times$ $U(1)_F$ symmetry. \label{asyLR}} \end{figure} In this case the neutrino Yukawa couplings (\ref{yuk2}) yield the CP asymmetry \begin{equation} \varepsilon_1\ \sim\ {3\over 16\pi}\ \epsilon^5\;, \end{equation} which correspond to the baryon asymmetry (cf.~(\ref{basym})) \begin{equation}\label{est2} Y_B \sim \kappa\ 10^{-6}\;. \end{equation} Due to the large value of $\epsilon$ the CP asymmetry is two orders of magnitude larger than in the case with $SU(5)\times U(1)_F$ symmetry. However, washout processes are now also stronger. The solution of the Boltzmann equations is shown in fig.~\ref{asyLR}. The final asymmetry is again $Y_B \sim 10^{-9}$ which now corresponds to $\kappa \sim 10^{-3}$. The baryogenesis temperature is considerably larger than in the first case, \begin{equation} T_B \sim M_1 \sim \epsilon^5 M_3 \sim 10^{12}\ \mbox{GeV}\;. \end{equation} The baryon asymmetry is largely determined by the parameter $\wt m_1$ defined in eq.~(\ref{ooeb}) \cite{plu97}. In the first example, one has $\wt m_1 \sim \Bar m_\nu$. In the second case one finds $\wt m_1 \sim m_3$. Since $\Bar m_\nu$ and $m_3$ are rather similar it is not too surprizing that the generated baryon asymmetry is about the same in both cases. \section{Conclusions} Detailed studies of the thermodynamics of the electroweak interactions at high temperatures have shown that in the standard model and most of its extensions the electroweak transition is too weak to affect the cosmological baryon asymmetry. Hence, one has to search for baryogenesis mechanisms above the Fermi scale. Due to sphaleron processes baryon number and lepton number are related in the high-temperature, symmetric phase of the standard model. As a consequence, the cosmological baryon asymmetry is related to neutrino properties. Baryogenesis requires lepton number violation, which occurs in extensions of the standard model with right-handed neutrinos and Majorana neutrino masses. Although lepton number violation is needed in order to obtain a baryon asymmetry, it must not be too strong since otherwise any baryon and lepton asymmetry would be washed out. This leads to stringent upper bounds on neutrino masses which depend on the particle content of the theory. The solar and atmospheric neutrino deficits can be interpreted as a result of neutrino oscillations. For hierarchical neutrinos the corresponding neutrino masses are very small. Assuming the see-saw mechanism, this suggests the existence of very heavy right-handed neutrinos and a large scale of $B-L$ breaking. It is remarkable that these hints on the nature of lepton number violation fit very well together with the idea of leptogenesis. For hierarchical neutrino masses, with $B-L$ broken at the unification scale $\Lambda_{\mbox{\scriptsize GUT}}\sim 10^{16}\;$GeV, the observed baryon asymmetry $Y_B \sim 10^{-10}$ is naturally explained by the decay of heavy Majorana neutrinos. Although the observed baryon asymmetry imposes important constraints on neutrino properties, other observables are needed to discriminate between different models. The two examples considered in this paper predict different baryogenesis temperatures. Correspondingly, in supersymmetric models the predictions for the gravitino abundance are different \cite{khl84}-\cite{bol98}. In the case with $SU(5)\times U(1)_F$ symmetry, stable gravitinos can be the dominant component of cold dark matter \cite{bol98}. The models make also different predictions for the rate of lepton flavour changing radiative corrections.
\section{\@startsection{section}{1}{\z@}% {-3.25ex\@plus -1ex \@minus -.2ex}% {0.8ex \@plus .2ex}{\reset@font\large\bfseries}} \@addtoreset{equation}{section} \makeatother \renewcommand{\theequation}{\thesection.\arabic{equation}} \def\begin{equation}\label} \def\ba{\begin{eqnarray}{\begin{equation}\label} \def\ba{\begin{eqnarray}} \def\end{equation}} \def\ea{\end{eqnarray}{\end{equation}} \def\ea{\end{eqnarray}} \def\ar{\begin{array}} \def\er{\end{array}} \def\nn{\nonumber} \def\ts{\textstyle} \def\wt{\widetilde} \def\wh{\widehat} \def\rightarrow} \def\={\,={\rightarrow} \def\={\,=} \textwidth 15.4cm \hoffset -1.5cm \textheight 23.5cm \topmargin -1.9cm \hfuzz=2pt \begin{document} \setcounter{page}{0} \thispagestyle{empty} {\small\sc April~~1999 \hfill hep-th/9904059 } \vspace*{1.7cm} \begin{center} {\large\bf Fermionic representations for characters of \\ ${\cal M}(3,t)$, ${\cal M}(4,5)$, ${\cal M}(5,6)$ and ${\cal M}(6,7)$ minimal models and \\ related Rogers-Ramanujan type and dilogarithm identities} \\ [1.1cm] {\sc Andrei G. Bytsko} \\ [4mm] Steklov Mathematics Institute, Fontanka 27,\\ St.Petersburg~~191011, Russia \\ [20mm] \hfill \begin{tabular}{c} Dedicated to Professor L.D.Faddeev \\ on his 65th birthday. \end{tabular} \\ [40mm] {\bf Abstract} \\ [9mm] \parbox{13cm}{ Characters and linear combinations of characters that admit a fermionic sum representation as well as a factorized form are considered for some minimal Virasoro models. As a consequence, various Rogers-Ramanujan type identities are obtained. Dilogarithm identities producing corresponding effective central charges and secondary effective central charges are derived. Several ways of constructing more general fermionic representations are discussed. } \end{center} \vspace*{2cm} \vfill{ \hspace*{-9mm} \begin{tabular}{l} \rule{6 cm}{0.05 mm} \\ <EMAIL> \end{tabular} } \setcounter{section}{0} \newpage \section{Introduction} A minimal Virasoro model \cite{BPZ} ${\cal M}(s,t)$ is parameterized by two positive integers $s$ and $t$ such that $\langle s, t \rangle =1$ (i.e.~they are co-prime numbers). It has the central charge $c(s,t) = 1 - \frac{6(s-t)^{2}}{st}$. The characters of its irreducible representations with highest weights $h_{n,m}^{s,t}=\frac{(nt-ms)^2-(s-t)^2}{4st}$ are given by \cite{FF} \begin{equation}\label} \def\ba{\begin{eqnarray}{char} \chi_{n,m}^{s,t}(q) \= \frac{ q^{\eta_{n,m}^{s,t}} }{(q)_\infty } \sum_{k=-\infty }^{\infty }q^{st k^2} \left(q^{k(nt-ms)}-q^{k(nt+ms)+nm}\right) \, , \end{equation}} \def\ea{\end{eqnarray} where $1{\leq} n{\leq} s{-}1$, $1{\leq} m{\leq} t{-}1$,\, $\eta_{n,m}^{s,t}:={h_{n,m}^{s,t} - \frac{c(s,t)}{24}}$ and $(q)_m:=\prod_{k=1}^m (1-q^k)$. The characters possess the following symmetries \begin{equation}\label} \def\ba{\begin{eqnarray}{hsym} \chi^{s,t}_{n,m}(q) \= \chi^{t,s}_{m,n}(q) \= \chi^{s,t}_{s-n,t-m}(q) \= \chi^{t,s}_{t-m,s-n}(q) \, . \end{equation}} \def\ea{\end{eqnarray} In addition, (\ref{char}) allows us to relate some characters of different models \begin{equation}\label} \def\ba{\begin{eqnarray}{alp} \chi^{\alpha s,t}_{\alpha n,m}(q) \= \chi^{s, \alpha t}_{n, \alpha m}(q) \, , \end{equation}} \def\ea{\end{eqnarray} where $\alpha$ is a positive number such that $\langle s, \alpha t \rangle = \langle t, \alpha s \rangle =1$. For instance, $ \chi^{5,6}_{m,2}(q) = \chi^{3,10}_{1,2m}(q)$, $m=1,2$. Below we will also use the identity proven in \cite{BF2}: \begin{equation}\label} \def\ba{\begin{eqnarray}{36} \chi^{3n,2t}_{n,t-2m}(q) - \chi^{3n,2t}_{n,t+2m}(q) \= \chi^{6n,t}_{n,m}(q) - \chi^{6n,t}_{n,t-m}(q) \,, \end{equation}} \def\ea{\end{eqnarray} where $m< t/2$ and $\langle t,6 \rangle = \langle n,2\rangle = \langle t,n\rangle =1$. In some cases characters (\ref{char}) admit the form named ``fermionic representation" \begin{equation}\label} \def\ba{\begin{eqnarray}{F1} \chi(q) \= q^{\rm{const}} \sum\limits_{\vec{m}=\vec{0}}^\infty \frac{q^{\vec{m}^{\,t} A \vec{m} + \vec{m}\cdot\vec{B} }} {(q)_{m_{1}}\ldots (q)_{m_{r}}}\,. \end{equation}} \def\ea{\end{eqnarray} Hitherto examples of such representation were obtained for two large classes of characters. For $A$ being related to the inverse Cartan matrix for some Lie algebra they were studied in \cite{KKMM}. In this case certain restrictions are often imposed on the summation over $\vec{m}$. Examples of (\ref{F1}) for $A=0$ or, at most, being a diagonal matrix (and $(q)_{m_i}$ being replaced with $(q^b)_{m_i}$, $b>0$) were obtained in \cite{BF2,BF1} as consequences of representation of characters in the ``factorized form" \begin{equation}\label} \def\ba{\begin{eqnarray}{fact} \chi(q) \= q^{\rm{const}} \prod_{m=1}^M \left(\{x_m\}^+_y\right)^{\gamma^+_m} \prod_{n=1}^N \left(\{x_n\}^-_y\right)^{\gamma^-_n} \,, \end{equation}} \def\ea{\end{eqnarray} where the multiplicities $\gamma^\pm_i$ are integer. Here and below we use the notations of \cite{BF2} \begin{equation}\label} \def\ba{\begin{eqnarray}{bl} \{x\}_{y}^{\pm }:=\prod\limits_{k=0}^{\infty }\left( 1\pm q^{x+ky}\right) \,,\quad0<x<y\,;\quad\quad \left\{ x_{1};\ldots ;x_{n}\right\}_{y}^{\pm}:= \prod\limits_{a=1}^{n}\{x_{a}\}_{y}^{\pm }\,. \end{equation}} \def\ea{\end{eqnarray} Equivalence of fermionic and factorized forms for characters which admit both types of representation gives rise to non-trivial identities. We will refer to them as to Rogers-Ramanujan type identities. The paper is organized as follows. Section 2 contains fermionic sum representations for certain families of characters and linear combinations of characters for ${\cal M}(3,t)$. These results are extensions of some previously known examples. Here are also given several fermionic sum representations for ${\cal M}(4,5)$, ${\cal M}(5,6)$ and ${\cal M}(6,7)$ that seem to be new. In Section 3 we observe that all the considered (combinations of) characters admit also the factorized form (\ref{fact}) and present the corresponding Rogers-Ramanujan type identities. In Section 4 we apply the saddle point analysis to the fermionic representations given in Section 2 and derive Bethe ansatz type equations that yield the corresponding effective central charge or (for the differences of characters) the secondary effective central charge as a sum of dilogarithms. For the latter case we show how to modify the saddle point analysis if we are dealing with fermionic sum with alternated signs in summation. We solve the Bethe ansatz type equations explicitly and obtain four infinite dilogarithm sum rules corresponding to ${\cal M}(3,t)$ and one non-trivial identity for ${\cal M}(4,5)$. In Section 5 we employ certain identities for Virasoro characters and expand the list of fermionic representations for ${\cal M}(4,5)$, ${\cal M}(5,6)$ and ${\cal M}(6,7)$. Section 6 contains a brief discussion and conclusion. \section{Fermionic representations for ${\cal M}(3,t)$, ${\cal M}(4,5)$ ${\cal M}(5,6)$ and ${\cal M}(6,7)$.} It was observed in \cite{KKMM} that the characters $\chi^{3,3k+1}_{1,k}(q)$ and $\chi^{3,3k+2}_{1,k}(q)$ as well as all the four characters of ${\cal M}(3,5)$ admit the fermionic form (\ref{F1}) with $A_k$ and ${\tilde A}_k$ given by \ba && \bigl(A_k\bigr)_{ij} = \bigl(A_k\bigr)_{ji} = \min (i,j) \,, \quad {\rm for}\ 1\leq i,j \leq k-1 \,, \label{A3a} \\ && \bigl(A_k\bigr)_{kj} = \bigl(A_k\bigr)_{jk} = {\ts \frac j2} + {\ts \frac{1-k}{4}} \delta_{jk} \,, \label{A3b} \\ && \bigl({\tilde A}_k\bigr)_{ij} \= \bigl(A_k\bigr)_{ij} - {\ts \frac 14 } \delta_{ik} \delta_{jk} \,. \label{A3c} \ea It turns out that these results can be extended as follows: \ba \chi^{3,3k+1}_{1,n}(q) \pm \chi^{3,3k+1}_{1,3k+1-n}(q) &=& q^{\eta^{3,3k+1}_{1,n}} \sum\limits_{\vec{m}=\vec{0}}^\infty \frac{(\pm 1)^{m_k} q^{\vec{m}^{\,t} A_k \vec{m} + \vec{m}\cdot\vec{B}^{3k+1}_{n}}}{ (q)_{m_{1}} \ldots (q)_{m_{k}}} \, , \label{M31} \\ \chi^{3,3k+2}_{1,n}(q) \pm \chi^{3,3k+2}_{1,3k+2-n}(q) &=& q^{\eta^{3,3k+2}_{1,n}} \sum\limits_{\vec{m}=\vec{0}}^\infty \frac{(\pm 1)^{m_k} q^{\vec{m}^{\,t} {\tilde A}_k \vec{m} + \vec{m}\cdot\vec{B}^{3k+2}_{n} }} {(q)_{m_{1}} \ldots (q)_{m_{k}}} \, , \label{M32} \ea where $1{\leq} n {\leq} (k{+}1)$, the matrices $A_k$ and ${\tilde A}_k$ are defined by (\ref{A3a})-(\ref{A3c}) and the corresponding $k$-component vectors ${\vec B}^{3k+1}_{n}$ and ${\vec B}^{3k+2}_{n}$ are such that \ba && \label{B3} \bigl( {\vec B}^{3k+1}_{n} \bigr)_j \= \max (j-n+1,0) + {\ts \frac{n-k-2}{2}} \delta_{jk} \,, \\ && \label{B4} \bigl( {\vec B}^{3k+2}_{n} \bigr)_j \= \bigl({\vec B}^{3k+1}_{n} \bigr)_j +{\ts \frac 12}\delta_{jk} \,. \ea For example, ${\vec B}^{3k+1}_{1}=(1,2,3,\ldots,k-1,{\ts \frac{k-1}{2}})$, ${\vec B}^{3k+1}_{k}={\vec 0}$ and ${\vec B}^{3k+1}_{k+1} = (0,\ldots,0,{\ts -\frac 12})$. It turns out that besides the infinite families (\ref{M31}) and (\ref{M32}) of fermionic sums for the combinations of characters there exist similar ones for ``single" characters: \ba \label{M33} && \chi^{3,6t-2}_{1,3t-1}(q) = q^{\eta^{3,6t-2}_{1,3t-1}} \sum\limits_{\vec{m}=\vec{0}}^\infty \frac{ q^{\vec{m}^{\,t} A_{2t-1} \vec{m} + \vec{m}\cdot\vec{C}^{2t-1}}}{(q)_{m_{1}} \ldots (q)_{m_{2t-1}}} \,,\\ \label{M34} && \chi^{3,6t+2}_{1,3t+1}(q) = q^{\eta^{3,6t+2}_{1,3t+1}} \sum\limits_{\vec{m}=\vec{0}}^\infty \frac{ q^{\vec{m}^{\,t} {\tilde A}_{2t} \vec{m} + \vec{m}\cdot\vec{C}^{2t} }} {(q)_{m_{1}} \ldots (q)_{m_{2t}}} \, , \ea where $t=1,2,3\ldots$, $A$ and ${\tilde A}$ are defined in (\ref{A3a})-(\ref{A3c}), and \begin{equation}\label} \def\ba{\begin{eqnarray}{M35} \bigl(\vec{C}^{2t-1}\bigr)_j \= \max (j-t+1,0) - {\ts \frac t2} \delta_{j,2t-1} \,,\quad \quad \bigl(\vec{C}^{2t}\bigr)_j \= \max (j-t+1,0) - {\ts \frac{t+2}{2} } \delta_{j,2t} \,. \end{equation}} \def\ea{\end{eqnarray} For example, $\vec{C}^1=(\frac 12)$, $\vec{C}^2=(1,\frac 12)$, $\vec{C}^2=(0,1,1)$, $\vec{C}^3=(0,1,2,1)$. Although formal derivation of (\ref{M31})-(\ref{M32}) and (\ref{M33})-(\ref{M34}) is beyond the scope of the present paper, I have verified these identities using Mathematica for $k\leq 7$, $t\leq 5$ expanding them typically up to $q^{100}$. A rigorous proof can presumably be achieved with the help of the machinery of Bailey pairs (see e.g.~\cite{BP}). In contrast to a generic fermionic representation, eqs.~(\ref{M31}) and (\ref{M32}) do not have restrictions on the summation. However, if we combine them to obtain an expression for a single character, the summation over $m_k$ will be restricted to either odd or even numbers. In particular, eq.~(\ref{M31}) in the case of $k=1$, $n=2$ gives two expressions for $\chi^{3,4}_{1,2}(q)$ (as noted also in \cite{KKMM}). Moreover, eq.~(\ref{M33}) yields one more expression for the same character and hence we get \begin{equation}\label} \def\ba{\begin{eqnarray}{34} q^{-\eta^{3,4}_{1,2}} \chi^{3,4}_{1,2}(q) \= \sum_{m\geq 1 \atop m-{\rm odd}} \frac{ q^{\frac 12 m^2-\frac 12 m} } {(q)_m} = \sum_{m\geq 0 \atop m-{\rm even}} \frac{ q^{\frac 12 m^2-\frac 12 m} } {(q)_m} =\sum_{m=0}^\infty \frac{q^{\frac 12 m^2+\frac 12 m}}{(q)_m} \,. \end{equation}} \def\ea{\end{eqnarray} While (\ref{M31}) and (\ref{M32}) extend previously known examples, the following fermionic representations for ${\cal M}(4,5)$, ${\cal M}(5,6)$ and ${\cal M}(6,7)$ to my knowledge have not been discussed in the literature so far: \noindent \underline{${\cal M}(4,5)$:}\ two characters are representable in the form (\ref{F1}) with \begin{equation}\label} \def\ba{\begin{eqnarray}{M45a} A \= {\ts \frac 12} \left( \ar{cc} 4 & 1 \\ 1 & 1 \er\right)\, , \end{equation}} \def\ea{\end{eqnarray} namely, for $n=1,2$ the following equality holds \begin{equation}\label} \def\ba{\begin{eqnarray}{M45b} \chi^{4,5}_{2,n}(q) \= q^{\eta^{4,5}_{1,n}} \sum_{\vec{m}=\vec{0}}^\infty \frac{ q^{2m_1^2+\frac 12 m_2^2+m_1m_2+(4-2n)m_1+ \frac 12 m_2} } { (q)_{m_1} (q)_{m_2} } \,. \end{equation}} \def\ea{\end{eqnarray} \noindent \underline{${\cal M}(5,6)$:}\ certain linear combinations of characters admit the form (\ref{F1}) with \begin{equation}\label} \def\ba{\begin{eqnarray}{M56a} A \={\ts \frac 12} \left( \ar{cc} 1 & 1 \\ 1 & 1 \er\right)\,, \end{equation}} \def\ea{\end{eqnarray} namely, for $n=1,2$, we have \ba && \chi^{5,6}_{n,2}(q) \pm \chi^{5,6}_{n,4}(q) \=q^{\eta^{5,6}_{n,2}} \sum_{\vec{m}=\vec{0}}^\infty \frac{ (\pm 1)^{m_2} q^{\frac 12 (m_1^2 + m_2^2) +m_1m_2+ \frac 12 m_1+(2- n)m_2} } { (q)_{m_1} (q)_{m_2} } \,, \label{M56b} \\ && \chi^{5,6}_{n,1}(q) - \chi^{5,6}_{n,5}(q) \=q^{\eta^{5,6}_{n,1}} \sum_{\vec{m}=\vec{0}}^\infty \frac{ (-1)^{m_2} q^{\frac 12 (m_1^2 + m_2^2) +m_1m_2+ (n-1)(m_1+m_2) } } { (q)_{m_1} (q)_{m_2} } \,. \label{M56c} \ea \underline{${\cal M}(5,6)$:}\ Another fermionic representation for the characters of the ${\cal M}(5,6)$ model can be obtained if we notice that eqs.~(\ref{alp}) and (\ref{36}) allow us to relate these characters to those of the ${\cal M}(3,10)$ model: $ \chi^{5,6}_{n,2}(q) \pm \chi^{5,6}_{n,4}(q)= \chi^{3,10}_{1,2n}(q) \pm \chi^{3,10}_{1,10-2n}(q)$ and $ \chi^{5,6}_{n,1}(q) - \chi^{5,6}_{n,5}(q)= \chi^{3,10}_{1,5-2n}(q) - \chi^{3,10}_{1,5+2n}(q)$, $n=1,2$. Combining these relations with formulae (\ref{M31}) for $k=3$, we find \ba && \chi^{5,6}_{n,2}(q) \pm \chi^{5,6}_{n,4}(q) \=q^{\eta^{5,6}_{n,2}} \sum_{\vec{m}=\vec{0}}^\infty\frac{ (\pm 1)^{m_3} q^{\vec{m}^t A \vec{m} + (2-n)m_2 + (\frac 32 -n)m_3} } { (q)_{m_1} (q)_{m_2} (q)_{m_3}} \,, \label{M56d} \\ && \chi^{5,6}_{n,1}(q) - \chi^{5,6}_{n,5}(q) \=q^{\eta^{5,6}_{n,1}} \sum_{\vec{m}=\vec{0}}^\infty \frac{ (-1)^{m_3} q^{ \vec{m}^t A \vec{m}+ (n-1)(m_1+2m_2 +m_3)} } { (q)_{m_1} (q)_{m_2} (q)_{m_3}} \,, \label{M56e} \ea where $n=1,2$ and \begin{equation}\label} \def\ba{\begin{eqnarray}{M56f} A \= \left( \ar{ccc} 1 & 1 & {\ts \frac 12} \\ 1 & 2 & 1 \\ {\ts \frac 12} & 1 & 1 \er\right)\, . \end{equation}} \def\ea{\end{eqnarray} Furthermore, eq.~(\ref{alp}) also implies that $ \chi^{5,6}_{n,3}(q) = \chi^{2,15}_{1,3n}(q)$, $n=1,2$. For the ${\cal M}(2,2k+1)$ model a fermionic representation is well known \cite{And,Kac,NRT}: \begin{equation}\label} \def\ba{\begin{eqnarray}{2t} \chi^{2,2k+1}_{1,n}(q) \= q^{\eta^{2,2k+1}_{1,n}} \sum\limits_{\vec{m} =\vec{0}}^\infty \frac{q^{\vec{m}^{\,t} {\bar A}_k \vec{m} + \vec{m} \cdot \vec{F}^k_n }}{(q)_{m_{1}}\ldots (q)_{m_{k-1}}}\,, \end{equation}} \def\ea{\end{eqnarray} where $1{\leq}n{\leq}k$,\, $\bigl(\vec{F}^k_n\bigr)_j=\max(j-n+1,0)$, and ${\bar A}_k$ coincides with the $\textstyle (k{-}1){\times}(k{-}1)$ minor (\ref{A3a}) which is the inverse Cartan matrix of the tadpole graph with $(k{-}1)$ nodes. This gives us yet another fermionic representation for ${\cal M}(5,6)$ with $A$ being 6$\times$6 matrix. \noindent \underline{${\cal M}(6,7)$:}\ Employing again eqs.~(\ref{alp}) and (\ref{36}), we observe that $ \chi^{6,7}_{2,n}(q) \pm \chi^{6,7}_{4,n}(q)= \chi^{3,14}_{1,2n}(q) \pm \chi^{3,14}_{1,14-2n}(q)$ and $ \chi^{6,7}_{1,n}(q) - \chi^{6,7}_{5,n}(q)= \chi^{3,14}_{1,7-2n}(q) - \chi^{3,14}_{1,7+2n}(q)$, where $n=1,2,3$. Combining these relations with formulae (\ref{M32}) for $k=4$, we find \ba && \!\!\!\!\!\!\!\!\!\!\! \chi^{6,7}_{2,n}(q) \pm \chi^{6,7}_{4,n}(q) = q^{\eta^{6,7}_{2,n}} \sum_{\vec{m}=\vec{0}}^\infty \frac{ (\pm 1)^{m_4} q^{\vec{m}^t A \vec{m} + (2-n)(m_2 +2m_3) + (\frac 52-n)m_4} }{(q)_{m_1} (q)_{m_2} (q)_{m_3} (q)_{m_4}} ,\quad n=1,2 , \label{M67a} \\ && \!\!\!\!\!\!\!\!\!\!\! \chi^{6,7}_{1,n}(q) - \chi^{6,7}_{5,n}(q) = q^{\eta^{6,7}_{1,n}} \sum_{\vec{m}=\vec{0}}^\infty \frac{ (-1)^{m_4} q^{\vec{m}^t A \vec{m} + \vec{m}\cdot\vec{D}^n} }{ (q)_{m_1} (q)_{m_2} (q)_{m_3} (q)_{m_4} } ,\quad n=1,2,3 ,\label{M67c} \ea where $\vec{D}^1=(0,0,0,0)$, $\vec{D}^2=(0,0,1,1)$, $\vec{D}^3=(1,2,3,2)$ and \begin{equation}\label} \def\ba{\begin{eqnarray}{M67d} A \= \left( \ar{cccc} 1& 1 & 1 & {\ts \frac 12} \\ 1 & 2 & 2 & 1 \\ 1& 2& 3 &{\ts \frac 32}\\ {\ts \frac 12} & 1 & {\ts \frac 32}& 1 \er\right)\, . \end{equation}} \def\ea{\end{eqnarray} Furthermore, due to eq.~(\ref{alp}) we can identify $ \chi^{6,7}_{3,n}(q) = \chi^{2,21}_{1,3n}(q)$, $n=1,2,3$. Therefore, these three characters of ${\cal M}(6,7)$ admit fermionic form of the type (\ref{2t}) with $A$ being 9$\times$9 matrix. \section{Rogers-Ramanujan type identities.} In the previous section we considered some characters and combinations of characters which possess fermionic representations. It turns out that all of them have another common feature -- they are factorizable, that is they admit also the form (\ref{fact}). For the characters of the ${\cal M}(2,2k+1)$ models this is the well-known representation \begin{equation}\label} \def\ba{\begin{eqnarray}{AG} \chi^{2,2k+1}_{1,n}(q) \= q^{\eta^{2,2k+1}_{1,n}} \prod_{{j=1} \atop{j\neq n}}^{k} \frac{1}{ \{j;2k+1-j\}^-_{2k+1} } \,, \end{equation}} \def\ea{\end{eqnarray} where we use the notations (\ref{bl}). Equality of the r.h.s.~of eqs.~(\ref{AG}) and (\ref{2t}) yields a family of identities known as the Andrews-Gordon identities \cite{And}. For $k=2$ these are the famous Rogers-Ramanujan identities \cite{RSR} \begin{equation}\label} \def\ba{\begin{eqnarray}{RR} \sum_{m=0}^\infty \frac{q^{m^2+m}}{(q)_m} \= \frac{1}{ \{2;3\}_{5}^{-} } \,, \quad\quad \sum_{m=0}^\infty \frac{q^{m^2}}{(q)_m} \= \frac{1}{ \{1;4\}_{5}^{-} } \,. \end{equation}} \def\ea{\end{eqnarray} Actually, (\ref{AG}) is only a particular case of a more general formula \begin{equation}\label} \def\ba{\begin{eqnarray}{2s} \chi_{n,m}^{2n,t}(q) \= \frac{q^{\eta^{2n,t}_{n,m} }}{\{1\}^-_1} \{nm;nt-nm;nt\}_{nt}^- \, , \end{equation}} \def\ea{\end{eqnarray} which together with \begin{equation}\label} \def\ba{\begin{eqnarray}{3s} \chi_{n,m}^{3n,t}(q) \= \frac{q^{ \eta^{3n,t}_{n,m} }}{\{1\}^-_1} \{nm;2nt-nm;2nt\}_{2nt}^- \{2nt-2nm;2nt+2nm\}_{4nt}^{-} \end{equation}} \def\ea{\end{eqnarray} exhausts the possibility for single characters to be factorizable on the base of the $A_1^{(1)}$ and $A_2^{(2)}$ Macdonald identities \cite{BF2,Chr,CIZ}. Furthermore, it was shown in \cite{BF2} that in certain cases the following combinations \begin{equation}\label} \def\ba{\begin{eqnarray}{pm} \chi_{n,m}^{s,t}(q) \pm \chi_{n,t-m}^{s,t}(q) \end{equation}} \def\ea{\end{eqnarray} also are factorizable on the base of the same Macdonald identities. The explicit formulae found in \cite{BF2} read \begin{eqnarray} \chi_{n,m}^{3n,t}(q) \pm \chi_{2n,m}^{3n,t}(q) &=& \frac{q^{\eta^{3n,t}_{n,m} }}{ \{1\}^-_1} \{ nm;nt-nm\} _{nt}^- {\ts \left\{ \frac{nt}{2}\right\}_{\frac{nt}{2}}^- \left\{ \frac{nt-2nm}{4}; \frac{nt+2nm}{4}\right\}_{\frac{nt}{2}}^{\pm} } \,, \label{f3} \\ \chi_{n,m}^{4n,t}(q) \pm \chi_{3n,m}^{4n,t}(q) &=& \frac{q^{\eta^{4n,t}_{n,m} }}{ \{1\}^-_1} \{nm;nt-nm;nt\} _{nt}^{-} {\ts \left\{ \frac{nt}{2}-nm;\frac{nt}{2}+nm; \frac{nt}{2}\right\}_{nt}^\pm } \, , \label{f4} \\ \chi_{n,m}^{6n,t}(q) - \chi_{5n,m}^{6n,t}(q) &=& \frac{q^{\eta^{6n,t}_{n,m} }}{ \{1\}^-_1} \, \{nm;nt-nm;nt\}_{nt}^- \{ nt-2nm;nt+2nm\}_{2nt}^- \,. \label{f6} \end{eqnarray} Combining the fermionic representations given in the previous section with these factorized representations, we obtain various identities of the ``sum--product" type. They can be regarded as generalizations of the Rogers-Ramanujan identities and it seems that many of them (especially those with a multivariable summation) have not appeared in the literature so far. \noindent \underline{${\cal M}(3,t)$:}\ According to (\ref{f3}) the fermionic sums on the r.h.s.~of (\ref{M31})-(\ref{M32}) are equal, respectively, to \ba \frac{q^{\eta^{3,3k+1}_{1,n}}}{ \{1\}^-_1} \{ n;3k+1-n\} _{3k+1}^- {\ts \left\{ \frac{3k+1}{2}\right\}_{\frac{3k+1}{2}}^- \left\{\frac{3k+1-2n}{4}; \frac{3k+1+2n}{4}\right\}_{\frac{3k+1}{2}}^{\pm} } \,,\label{rr1} \\ \frac{q^{\eta^{3,3k+2}_{1,n}}}{ \{1\}^-_1} \{ n;3k+2-n\} _{3k+2}^- {\ts \left\{ \frac{3k+2}{2}\right\}_{\frac{3k+2}{2}}^- \left\{\frac{3k+2-2n}{4}; \frac{3k+2+2n}{4}\right\}_{\frac{3k+2}{2}}^{\pm} } \,.\label{rr2} \ea For instance, for $k=1,2$ we get the following identities \ba && \!\!\!\!\!\!\!\!\!\!\!\! \sum_{m=0}^\infty \frac{ (\pm 1)^{m}\, q^{ \frac 12 m^2}} {(q)_m} = \bigl\{{\ts\frac 12}\bigr\}^\pm_{1} \,, \quad\quad\ \label{k1a} \sum_{m=0}^\infty \frac{(\pm 1)^m q^{\frac 12 m^2-\frac 12 m }} {(q)_m} = (1\pm 1)\, \{1\}^+_1 \,, \\ && \!\!\!\!\!\!\!\!\!\!\!\! \sum_{m=0}^\infty \frac{(\pm 1)^{m}\,q^{\frac 14 m^2+(1-\frac n2)m }} {(q)_m} \= \frac{ \bigl\{ \frac{5-2n}{4};\frac{5+2n}{4} \bigr\}^\pm_{\frac 52} }{ \{ \frac 52\}^+_{\frac 52}\{3-n;2+n\}^-_5 } \,,\quad n=1,2 \,, \label{k1b} \\ && \!\!\!\!\!\!\!\!\!\!\!\! \sum\limits_{\vec{m}=\vec{0}}^\infty \frac{(\pm 1)^{m_2} q^{ m_1^2 + \frac 34 m_2 + m_1 m_2+ \delta_{1n}m_1 +(1-\frac n2)m_2 }}{ (q)_{m_{1}} (q)_{m_{2}}} = \frac{ \bigl\{\frac{7-2n}{4};\frac{7+2n}{4}\bigr\}^\pm_{\frac 72} } {\bigl\{\frac 72\bigr\}^+_{\frac 72} \!\! \prod\limits_{j=1 \atop j\neq n}^3 \{j;7-j\}^-_7 } \,,\quad 1\leq n \leq 3 \,, \label{k2a} \\ && \!\!\!\!\!\!\!\!\!\!\!\! \sum\limits_{\vec{m}=\vec{0}}^\infty \frac{(\pm 1)^{m_2} q^{ m_1^2 + \frac 12 m_2 + m_1 m_2+ \delta_{1n}m_1 +(\frac 32 -\frac n2)m_2 }}{ (q)_{m_{1}} (q)_{m_{2}}} =\frac{ \bigl\{2-\frac{n}{2};2+\frac{n}{2}\bigr\}^\pm_{4} } { \prod\limits_{j=1 \atop j\neq n}^3 \{j;8-j\}^-_8 } \,,\quad 1\leq n \leq 3 \,. \label{k2b} \ea Here we simplified the product sides exploiting the Euler identity $\{x\}_{x}^+ \{x\}_{2x}^- =1$ and other transformations (see \cite{BF2}). Eqs.~(\ref{k1a}) are well known (see e.g.~\cite{Sl}), eqs.~(\ref{k1b}) were presented in \cite{BF2}. It should be remarked that in some cases the combinations on the l.h.s.~of (\ref{M31})-(\ref{M32}) belong both to (\ref{f3}) and (\ref{f4}). In this case the product side acquires more compact form \cite{BF2}: \begin{equation}\label} \def\ba{\begin{eqnarray}{f34} \chi_{m,n}^{3m,4n}(q) \pm \chi_{m,3n}^{3m,4n}(q) \= \frac{q^{\,\frac{nm-2}{48}}}{ \{1\}^-_1 } \{nm\}^-_{nm} \{ \frac{nm}{2} \}^\pm_{nm} \, . \end{equation}} \def\ea{\end{eqnarray} Therefore we obtain \ba \sum\limits_{\vec{m}=\vec{0}}^\infty \frac{(\pm 1)^{m_{4t+1}} q^{\vec{m}^{\,t} A_{4t+1} \vec{m} + \vec{m}\cdot\vec{B}^{12t+4}_{3t+1}}}{(q)_{m_{1}} \ldots (q)_{m_{4t+1}}} \=\frac{ \{3t+1\}^-_{3t+1} \{\frac{3t+1}{2}\}^\pm_{3t+1} }{ \{1\}^-_1}\,, \label{rr3} \\ \sum\limits_{\vec{m}=\vec{0}}^\infty \frac{(\pm 1)^{m_{4t+2}} q^{\vec{m}^{\,t} {\tilde A}_{4t+2} \vec{m} + \vec{m}\cdot\vec{B}^{12t+8}_{3t+2}}}{(q)_{m_{1}} \ldots (q)_{m_{4t+2}}} \=\frac{ \{3t+2\}^-_{3t+2} \{\frac{3t+2}{2}\}^\pm_{3t+2} }{ \{1\}^-_1} \,, \label{rr4} \ea where $t=0,1,2,\ldots$, and $A$, ${\tilde A}$ and $\vec{B}$ are defined in (\ref{A3a})-(\ref{A3c}) and (\ref{B3})-(\ref{B4}). For instance, (\ref{rr4}) yields for $t=0$ \begin{equation}\label} \def\ba{\begin{eqnarray}{38} \sum_{\vec{m}=\vec{0}}^\infty \frac{q^{m_1^2+ \frac 12 m_2^2+ m_1m_2+ \frac 12 m_2}}{ (q)_{m_1} (q)_{m_2}} = \{1\}^+_1 \{1\}^+_2 \,, \quad \sum_{\vec{m}=\vec{0}}^\infty \frac{(-1)^{m_2} q^{m_1^2+ \frac 12 m_2^2+ m_1m_2+ \frac 12 m_2}}{ (q)_{m_1} (q)_{m_2}} = 1 \, . \end{equation}} \def\ea{\end{eqnarray} The last equality is due to the fact that $\chi^{3,8}_{1,2}(q)-\chi^{3,8}_{1,6}(q)=1$ (see \cite{BF2,CIZ}). For the fermionic sums (\ref{M33})-(\ref{M34}) the product side also simplifies since these characters belong both to (\ref{2s}) and (\ref{3s}). Namely, we obtain for $t=1,2,3\ldots$ \begin{equation}\label} \def\ba{\begin{eqnarray}{rr5} \sum\limits_{\vec{m}=\vec{0}}^\infty \frac{ q^{\vec{m}^{\,t} A_{2t-1} \vec{m} + \vec{m}\cdot\vec{C}^{2t-1}}}{(q)_{m_{1}} \ldots (q)_{m_{2t-1}}} \= \frac{ \{3t-1\}^-_{3t-1} }{ \{1\}^-_1 } \,,\quad\quad \sum\limits_{\vec{m}=\vec{0}}^\infty \frac{ q^{\vec{m}^{\,t} {\tilde A}_{2t} \vec{m} + \vec{m}\cdot\vec{C}^{2t} }} {(q)_{m_{1}} \ldots (q)_{m_{2t}}} \= \frac{ \{3t+1\}^-_{3t+1} }{ \{1\}^-_1 } \,, \end{equation}} \def\ea{\end{eqnarray} where $A$, ${\tilde A}$ and $\vec{C}$ are defined in (\ref{A3a})-(\ref{A3c}) and (\ref{M35}). For instance, for $t=1$ we get \begin{equation}\label} \def\ba{\begin{eqnarray}{rr6} \sum_{m=0}^\infty \frac{ q^{\frac 12 m^2+\frac 12 m} }{(q)_m} \= \{1\}^+_1 \,, \quad\quad \sum_{\vec{m}=\vec{0}}^\infty \frac{q^{m_1^2+ \frac 12 m_2^2+ m_1m_2+ m_1+\frac 12 m_2}}{ (q)_{m_1} (q)_{m_2}} = \{1\}^+_1 \{2\}^+_2 \,. \end{equation}} \def\ea{\end{eqnarray} To complete the discussion of the ${\cal M}(3,t)$ case, we recall that since each of eqs.~(\ref{M31})-(\ref{M32}) comprises two equalities, we can express the involved characters as sums without alternation of the sign but with summation over $m_k$ restricted to odd or even numbers. Then, according to (\ref{3s}), each character obtained in this way admits also the product representation. Thus, we get for $t=3k{+}1$,\, $1{\leq} n {\leq} k{+}1$ \ba && \sum\limits_{\vec{m}=\vec{0} \atop m_k{\rm -even}}^\infty \!\! \frac{q^{\vec{m}^{\,t} A_k \vec{m} + \vec{m}\cdot\vec{B}^{t}_{n}}}{ (q)_{m_{1}} \ldots (q)_{m_{k}}} \= \frac{1}{\{1\}^-_1} \{n;2t- n;2t\}_{2t}^- \{2t-2n;2t+2n\}_{4t}^- \,, \\ && \sum\limits_{\vec{m}=\vec{0} \atop m_k{\rm -odd}}^\infty \!\! \frac{q^{\vec{m}^{\,t} A_k \vec{m}+\vec{m}\cdot\vec{B}^{t}_{n} }} {(q)_{m_{1}} \ldots (q)_{m_{k}}} \= \frac{1}{\{1\}^-_1} \{t-n;t+n;2t\}_{2t}^- \{2n;4t-2n\}_{4t}^- \,, \ea and analogous formulae for $t=3k{+}2$ if $A_k$ is replaced by $\tilde{A}_k$. \vspace*{3pt} \noindent \underline{${\cal M}(4,5)$:}\ From (\ref{M45b}) and (\ref{2s}) we obtain for $n=1,2$ \begin{equation}\label} \def\ba{\begin{eqnarray}{RR5} \sum_{\vec{m}=\vec{0} }^\infty \frac{ q^{2m_1^2+\frac 12 m_2^2+ m_1m_2+(4-2n)m_1+ \frac 12 m_2} } { (q)_{m_1} (q)_{m_2} } \= \frac{\{5\}_5^+}{ \{n;5-n\}_{5}^{-} \{5-2n;5+2n\}_{10}^{-} } \,. \end{equation}} \def\ea{\end{eqnarray} \noindent \underline{${\cal M}(5,6)$:}\ Combining (\ref{M56b})-(\ref{M56e}) with (\ref{f3}) and (\ref{f6}), we obtain for $n=1,2$ \ba && \!\!\!\!\!\!\!\!\!\! \sum_{\vec{m}=\vec{0}}^\infty \frac{ (\pm 1)^{m_2} q^{\frac 12 (m_1^2 +m_2^2) +m_1m_2+ \frac 12 m_1+(2-n)m_2} }{ (q)_{m_1} (q)_{m_2}} \= \frac{1}{ \{n;5-n\}_{5}^{-} \bigl\{\frac 52-n;\frac 52 +n \bigr\}_{5}^{\mp} } = \label{RR6} \\ && = \sum_{\vec{m}=\vec{0}}^\infty\frac{ (\pm 1)^{m_3} q^{ m_1^2 +2m_2^2+ m_3^2 + 2m_1m_2 +m_1m_3 +2m_2m_3 +(2-n)m_2 + (\frac 32-n)m_3} }{ (q)_{m_1} (q)_{m_2} (q)_{m_3}} \,, \nn\\ && \!\!\!\!\!\!\!\!\!\! \sum_{\vec{m}=\vec{0}}^\infty \frac{ (-1)^{m_2} q^{\frac 12 (m_1^2 + m_2^2) +m_1m_2+ (n-1)(m_1+m_2)} }{ (q)_{m_1} (q)_{m_2} } = \frac{1}{ \{2n;10-2n\}_{10}^{-} } = \label{RR7} \\ && = \sum_{\vec{m}=\vec{0}}^\infty \frac{ (-1)^{m_3} q^{ m_1^2 +2m_2^2+ m_3^2 + 2m_1m_2 +m_1m_3 +2m_2m_3 + (n-1)(m_1+2m_2 +m_3)} }{ (q)_{m_1} (q)_{m_2} (q)_{m_3}} \,. \nn \ea \noindent \underline{${\cal M}(6,7)$:}\ The fermionic sums on the r.h.s.~of (\ref{M67a}) and (\ref{M67c}) are equal, respectively, to \begin{equation}\label} \def\ba{\begin{eqnarray}{RR8} \frac{q^{\eta^{6,7}_{2,n}}}{ \{1;4-n;3+n;6\}_{7}^- \{\frac 72 -n;\frac 72 +n \}^\mp_7 } \,, \quad\quad \frac{q^{\eta^{6,7}_{1,n}}}{ \{d_n;7-d_n\}_{7}^{-} \{2n;14-2n\}_{14}^{-} } \,, \end{equation}} \def\ea{\end{eqnarray} where $d_1=3$, $d_2=1$, $d_3=2$. \section{Dilogarithm identities.} If a $q$-series $\chi(q)$ (not necessarily identified in terms of characters) admits both fermionic representation (\ref{F1}) and product representation (\ref{fact}), it implies not only existence of a Rogers-Ramanujan type identity but also leads to a certain identity involving the dilogarithm function, $L(x)=\sum_{n=1}^{\infty} \frac{x^n}{n^2} + \frac 12 \ln{x}\ln(1-x)$. Indeed, the product side allows us to find easily the number $c_{\rm eff}$ that governs the asymptotics of $\chi(q)$ in the $q\rightarrow 1$ limit (see e.g.~\cite{BF2,BF1}). On the other hand, the same number can be obtained from the fermionic sum by the saddle point analysis (see e.g.~\cite{KKMM,NRT}). Equivalence of the two expressions for $c_{\rm eff}$ is typically a nontrivial identity. Of course, if it is known a-priory that $\chi(q)$ is a character, then its fermionic form alone leads to a dilogarithm identity since $c_{\rm eff}$ (effective central charge) is fixed by the properties of $\chi(q)$ with respect to the modular transformations. Namely, let $q=e^{2\pi i \tau}$ and $\hat{q}=e^{-2\pi i /\tau}$, then for the minimal Virasoro model ${\cal M}(s,t)$ we have $\chi^{s,t}_{n,m}(q)\sim \hat{q}^{-\frac{c_{\rm eff}(s,t)}{24}}$ as $q\rightarrow 1$, where \begin{equation}\label} \def\ba{\begin{eqnarray}{ce} c_{\rm eff}(s,t) \= c(s,t) - 24 h^\prime \= 1- \frac{6}{st} \,. \end{equation}} \def\ea{\end{eqnarray} Here $h^\prime$ denotes the lowest conformal weight in the model. Furthermore, as it was shown in \cite{BF2}, a difference of characters of the type (\ref{pm}) for all minimal models but ${\cal M}(2,t)$ has the asymptotics $\hat{q}^{-\frac{\tilde{c}(s,t)}{24}}$ when $q\rightarrow 1$. Here $\tilde{c}$ (secondary effective central charge) is given by \begin{equation}\label} \def\ba{\begin{eqnarray}{ct} \tilde{c}(s,t) \= c(s,t) - 24 h^{\prime\prime} \= 1- \frac{24}{st} \,, \end{equation}} \def\ea{\end{eqnarray} where $h^{\prime\prime}$ stands for the second lowest conformal weight in the model. For our purposes we need to consider a slightly generalized version of (\ref{F1}): \begin{equation}\label} \def\ba{\begin{eqnarray}{F2} \chi(q) \= q^{\rm{const}} \sum\limits_{\vec{m}} \frac{q^{\vec{m}^{\,t} A\vec{m}+ \vec{m}\cdot \vec{B} }}{(q^{b_1})_{m_{1}}\ldots (q^{b_r})_{m_{r}}} \,, \end{equation}} \def\ea{\end{eqnarray} where $b_i$ are some positive numbers. Modifying properly the standard saddle point analysis of a fermionic sum (see \cite{KKMM,NRT} for the case $b_i=1$ and \cite{BF1} for $b_i=b\neq 1$), we find that (\ref{F2}) has the asymptotics $\hat{q}^{-\frac{c_{\rm eff}}{24}}$ as $q\rightarrow 1$ with \begin{equation}\label} \def\ba{\begin{eqnarray}{cb} c_{\rm eff} \= \frac{6}{\pi^2} \sum_{i=1}^r \frac{1}{b_i} L(x_i) \,. \end{equation}} \def\ea{\end{eqnarray} Here the set of numbers $0<x_i<1$ satisfies the following equations \begin{equation}\label} \def\ba{\begin{eqnarray}{xb} x_i \= \prod_{j=1}^r (1-x_j)^{\frac{1}{b_j}(A_{ij}+A_{ji})} \,, \quad\ i=1,\ldots,r \,. \end{equation}} \def\ea{\end{eqnarray} Let us define $(\widehat{A})_{ij}=\frac{1}{2b_j}(A_{ij}+A_{ji})$. If matrix $\widehat{A}$ is invertible, it is convenient to introduce $I=2-(\widehat{A})^{-1}$ (generalized incidence matrix) and make the substitution $x_i=1/\mu^2_i$. Then (\ref{cb}) and (\ref{xb}) turn into \begin{equation}\label} \def\ba{\begin{eqnarray}{cm} c_{\rm eff} \=\frac{6}{\pi^2} \sum_{i=1}^r \frac{1}{b_i} L(\frac{1}{\mu^2_i}) \,,\quad\quad \mu_i^2 \= 1 + \prod_{j=1}^r (\mu_j)^{I_{ij}} \,. \end{equation}} \def\ea{\end{eqnarray} As we have seen above, certain differences of characters of the type (\ref{pm}) admit the fermionic form with alternated summation over the last variable, \begin{equation}\label} \def\ba{\begin{eqnarray}{cha} \chi(q) \= q^{\rm{const}} \sum\limits_{\vec{m}} \frac{(-1)^{m_r} q^{\vec{m}^{\,t}A\vec{m}+ \vec{m}\cdot \vec{B} }}{(q)_{m_{1}} \ldots (q)_{m_{r}}} \, . \end{equation}} \def\ea{\end{eqnarray} Let us find equations describing the $q\rightarrow 1$ limit of such series. To this end we notice that \begin{equation}\label} \def\ba{\begin{eqnarray}{tr} \frac{1}{(q)_m} \=(-1)^m q^{-\,\frac 12 m(m+1)}\frac{1}{(q^{-1})_m } \,. \end{equation}} \def\ea{\end{eqnarray} Therefore we can rewrite (\ref{cha}) as follows \begin{equation}\label} \def\ba{\begin{eqnarray}{cha2} \chi(q) \= q^{\rm{const}} \sum\limits_{\vec{m}} \frac{q^{\vec{m}^{\,t} A^{\prime} \vec{m}+ \vec{m}\cdot \vec{B}^{\prime} }}{(q)_{m_{1}} \ldots (q)_{m_{r-1}}(q^{-1})_{m_{r}}} \, , \end{equation}} \def\ea{\end{eqnarray} where \begin{equation}\label} \def\ba{\begin{eqnarray}{Ap} (A^{\prime})_{ij} \= A_{ij}- {\ts \frac 12} \delta_{ir}\delta_{jr} \,,\quad\quad (\vec{B}^{\prime})_j \= (\vec{B})_j- {\ts \frac 12} \delta_{jr} \,. \end{equation}} \def\ea{\end{eqnarray} Eq.~(\ref{cha2}) is a particular case of (\ref{F2}) with $b_1=\ldots=b_{r-1}=-b_r=1$ and thus we can apply eqs.~(\ref{cb})-(\ref{xb}). We conclude that (\ref{cha}) has the asymptotics $\hat{q}^{-\frac{\tilde{c}}{24}}$ as $q\rightarrow 1$ with \begin{equation}\label} \def\ba{\begin{eqnarray}{ca} \tilde{c} \=\frac{6}{\pi^2} \Bigr(\sum_{i=1}^{r-1} L(y_i) - L(y_r) \Bigl) \, , \end{equation}} \def\ea{\end{eqnarray} where the set of numbers $0< y_i <1$ satisfies the following equations \begin{equation}\label} \def\ba{\begin{eqnarray}{ya} y_i \= \prod_{j=1}^r (1-y_j)^{ (-1)^{\delta_{jr}} (A^{\prime}_{ij}+A^{\prime}_{ji})} \,, \quad i=1,\ldots,r \, . \end{equation}} \def\ea{\end{eqnarray} It is again convenient to introduce $I^\prime=2-({\widehat{A}}^\prime)^{-1}$, where $({\widehat{A}}^\prime)_{ij}=\frac 12 (-1)^{\delta_{jr}}(A^\prime_{ij}+ A^\prime_{ji})$. Then, making the substitution $y_i=1/\nu^2_i$, we transform (\ref{ca})-(\ref{ya}) to \begin{equation}\label} \def\ba{\begin{eqnarray}{cn} \tilde{c} \=\frac{6}{\pi^2} \Bigr(\sum_{i=1}^{r-1} L(\frac{1}{\nu^2_i})- L(\frac{1}{\nu^2_r}) \Bigl) \,,\quad\quad \nu_i^2 \= 1 + \prod_{j=1}^r (\nu_j)^{ I^\prime_{ij}} \,. \end{equation}} \def\ea{\end{eqnarray} Let us remark that performing the following change of variables in (\ref{ca})-(\ref{ya}): $z_i=y_i$, $i<r$ and $z_r=\frac{y_r}{y_r-1}$, we can transform these equations to the form almost coinciding with (\ref{cb})-(\ref{xb}) (for $b_i=1$) and involving the initial matrix $A$: \begin{equation}\label} \def\ba{\begin{eqnarray}{za} \tilde{c} \=\frac{6}{\pi^2} \sum_{i=1}^{r} L(z_i) \,, \quad\quad (-1)^{\delta_{ir}} z_i \= \prod_{j=1}^r (1-z_j)^{A_{ij}+A_{ji}} \,, \quad\ i=1,\ldots,r \,. \end{equation}} \def\ea{\end{eqnarray} In contrast with (\ref{xb}), now $z_r<0$. Deriving (\ref{za}) we used the definition \cite{Lew,Kir}: $L(x)=L(\frac{1}{1-x})-L(1)$ for $x<0$, and the property $L(x)=-L(\frac{x}{x-1})$ for $x<1$. Generalization of (\ref{ca})-(\ref{za}) for the case of a fermionic sum involving alternated summation over several variables is obvious. Now let us list dilogarithm identities that follow from the formulae (\ref{cm}), (\ref{cn}) and (\ref{ce})-(\ref{ct}) for the (combinations of) characters considered in the previous sections. \noindent \underline{${\cal M}(3,3k+1)$ and ${\cal M}(3,3k+2)$}: \ The explicit expressions (\ref{A3a})-(\ref{A3c}) for the matrices $A_k$ and ${\tilde A}_k$ allow us to compute the corresponding matrices $I_k$ and $\tilde{I}_k$ \ba && \bigl(I_k\bigr)_{ij} \= \delta_{i,j+1} + \delta_{i+1,j} + {\ts \frac 12} \delta_{i,k-1} \delta_{j,k-1} \,, \label{I1} \\ && \bigl(\tilde{I}_k\bigr)_{ij} \= \delta_{i,j+1} + \delta_{i+1,j} + \delta_{i,k} \delta_{j,k-1} + \delta_{i,k-1} \delta_{j,k} - 2 \delta_{i,k} \delta_{j,k} \,. \label{I2} \ea These generalized incidence matrices differ from those of the Lie algebra $A_k$ only by a few entries in the lower-right corner. This hints on a possibility to solve the corresponding sets of equations (\ref{cm}) in a uniform manner, similar to that known for the ${\cal M}(2,t)$ models \cite{NRT,KM}. Indeed, we find the following solutions of (\ref{cm}) (they can be verified by a straightforward substitution): \begin{equation}\label} \def\ba{\begin{eqnarray}{m1} \mu_i \= \frac{\sin\frac{(i+1)\pi}{3k+1}}{\sin\frac{\pi}{3k+1}} \,, \quad \ 1\leq i \leq k-1 \,; \quad\quad\ \mu_k^2 \= 1+ \frac{\sin\frac{k\pi}{3k+1}}{\sin\frac{\pi}{3k+1}} \end{equation}} \def\ea{\end{eqnarray} for $I_k$ given by (\ref{I1}) (that is in the case of ${\cal M}(3,3k+1)$ model) and \begin{equation}\label} \def\ba{\begin{eqnarray}{m2} \mu_i \= \frac{\sin\frac{(i+1)\pi}{3k+2}}{\sin\frac{\pi}{3k+2}} \,, \quad \ 1\leq i \leq k-1 \,; \quad\quad\ \mu_k^2 \= \frac{\sin\frac{(k+1)\pi}{3k+2}}{\sin\frac{\pi}{3k+2}} \end{equation}} \def\ea{\end{eqnarray} for $\tilde{I}_k$ given by (\ref{I2}) (${\cal M}(3,3k+2)$ model). Combining these results with (\ref{ce}), we derive the following identities \ba \label{di1} && \sum_{i=1}^{k-1} L\Bigl(\frac{\sin^2\frac{\pi}{3k+1}} {\sin^2\frac{(i+1)\pi}{3k+1}}\Bigr) + L\Bigl(\frac{\sin\frac{\pi}{3k+1}} {\sin\frac{\pi}{3k+1}+\sin\frac{k\pi}{3k+1}}\Bigr) \= \frac{\pi^2}{6} \,\frac{3k-1}{3k+1} \,,\\ \label{di2} && \sum_{i=1}^{k-1} L\Bigl(\frac{\sin^2\frac{\pi}{3k+2}} {\sin^2\frac{(i+1)\pi}{3k+2}}\Bigl) + L\Bigr(\frac{\sin\frac{\pi}{3k+2}} {\sin\frac{(k+1)\pi}{3k+2}}\Bigl) \=\frac{\pi^2}{6} \,\frac{3k}{3k+2} \,. \ea Let us remark that although these identities were derived here exploiting the modular properties of characters and the saddle point analysis, they resemble the ``general $A_1$-type" dilogarithm identities \cite{Kir} and probably can be proved in more direct way based on the functional relations for the dilogarithm. For instance, eq.~(\ref{di2}) for $k=2$ yields the equality $\bigl(L(1-\frac{1}{\sqrt{2}})+ L(\sqrt{2}-1)\bigr)=\frac{\pi^2}{8}$. It can be proved with the help of the Abel duplication formula \cite{Lew,Kir}. It should be mentioned that eq.~(\ref{di1}) for $k$ odd was encountered in \cite{Ahn} in the context of the thermodynamic Bethe ansatz. Next we consider the differences of characters in (\ref{M31})-(\ref{M32}). First, we compute the matrices $I^\prime_k$ and $\tilde{I}^\prime_k$. It turns out that for ${\cal M}(3,3k+1)$\, $\det({\widehat{A}_k}^\prime)=0$ that is $I^\prime_k$ does not exist and we have to solve the equations (\ref{ya}). For ${\cal M}(3,3k+2)$ the matrix $\tilde{I}^\prime_k$ exists and is given by \begin{equation}\label} \def\ba{\begin{eqnarray}{I3} \bigl(\tilde{I}_k^\prime\bigr)_{ij} \= \delta_{i,j+1} + \delta_{i+1,j} + \delta_{i,k} \delta_{j,k-1} - 3\delta_{i,k-1} \delta_{j,k} + 2\delta_{i,k-1} \delta_{j,k-1} -2\delta_{i,k} \delta_{j,k} \,. \end{equation}} \def\ea{\end{eqnarray} We obtain the following solution of (\ref{ya}) for ${\cal M}(3,3k+1)$ (written in terms of $\nu_i=1/\sqrt{y_i}$ for the sake of uniformness) \begin{equation}\label} \def\ba{\begin{eqnarray}{m3} \nu_i \= \frac{\sin\frac{(2i+2)\pi}{3k+1}}{\sin\frac{2\pi}{3k+1}} \,, \quad \ 1\leq i \leq k-1 \,; \quad\quad\ \nu_k^2 \= \frac{\sin\frac{2k\pi}{3k+1}}{\sin\frac{2\pi}{3k+1}} \,. \end{equation}} \def\ea{\end{eqnarray} and of (\ref{cn}) for ${\cal M}(3,3k+2)$: \begin{equation}\label} \def\ba{\begin{eqnarray}{m4} \nu_i \= \frac{\sin\frac{(2i+2)\pi}{3k+2}}{\sin\frac{2\pi}{3k+2}} \,, \quad \ 1\leq i \leq k-1 \,; \quad\quad\ \nu_k^2 \= 1+\frac{\sin\frac{(2k+2)\pi}{3k+2}}{\sin\frac{2\pi}{3k+2}}\,. \end{equation}} \def\ea{\end{eqnarray} Combining these results with (\ref{ct}), we obtain the following dilogarithm identities \ba \label{di3} && \sum_{i=1}^{k-1} L\Bigl(\frac{\sin^2\frac{2\pi}{3k+1}} {\sin^2\frac{(2i+2)\pi}{3k+1}}\Bigr) - L\Bigl(\frac{\sin\frac{2\pi}{3k+1}} {\sin\frac{2k\pi}{3k+1}}\Bigr) \=\frac{\pi^2}{6}\,\frac{3k-7}{3k+1} \,,\\ \label{di4} && \sum_{i=1}^{k-1} L\Bigl(\frac{\sin^2\frac{2\pi}{3k+2}} {\sin^2\frac{(2i+2)\pi}{3k+2}}\Bigl) -L\Bigr(\frac{\sin\frac{2\pi}{3k+2}} {\sin\frac{2\pi}{3k+2}+\sin\frac{(2k+2)\pi}{3k+2}} \Bigl) \= \frac{\pi^2}{6} \,\frac{3k-6}{3k+2} \, . \ea \noindent \underline{${\cal M}(4,5)$}:\ Eq.~(\ref{xb}) for $A$ given by (\ref{M45a}) can be reduced to a bi-quadratic equation. Solving it we obtain $x_1=1-\sqrt{\rho}$, $x_2=1-\frac{1}{1+\sqrt{\rho}}$, where $\rho:=\frac{\sqrt{5}-1}{2}$. According to (\ref{ce}) and (\ref{cb}) this leads to the identity \begin{equation}\label} \def\ba{\begin{eqnarray}{di5} {\ts L\Bigl(1-\sqrt{\frac{\sqrt{5}-1}{2}}\Bigr)+ L\Bigl(1- \frac{\sqrt{2}}{\sqrt{2}+\sqrt{\sqrt{5}-1}}\Bigr) } \= \frac{7}{10} \frac{\pi^2}{6} \,. \end{equation}} \def\ea{\end{eqnarray} Although the summation in the fermionic representation (\ref{M45b}) is not alternated, we can nevertheless introduce matrix $A^\prime$ according to (\ref{Ap}) and solve eqs.~(\ref{ca})-(\ref{ya}). The solution is $y_1=1-\rho$, $y_2=\rho$ and thus we get $\tilde{c} = \frac{6}{\pi^2} \bigl(L\bigl(1-\rho) - L\bigl(\rho)\bigr) = -\frac 15 $. This identity is rather trivial mathematically (since it is well known that $L(\rho)=\frac{\pi^2}{10}$ and $L(1-\rho)=\frac{\pi^2}{15}$) but it is remarkable that the value of $\tilde{c}$ is in agreement with (\ref{ct}). \noindent \underline{${\cal M}(5,6)$}:\ It is easy to see that for this model any matrix of the form \begin{equation}\label} \def\ba{\begin{eqnarray}{Aa} A \=\left(\ar{cc} 1-\alpha & \alpha \\ \alpha & 1-\alpha \er\right)\,, \qquad \alpha \leq 1/2 \,, \end{equation}} \def\ea{\end{eqnarray} yields $x_1=x_2=1-\rho$ and gives the correct central charge: $c_{\rm eff}=\frac{6}{\pi^2} 2L(1-\rho)=\frac 45$. Besides our example (\ref{M56a}) corresponding to $\alpha=\frac 12$ the case of $\alpha=\frac 13$ is known \cite{KKMM}. However, matrix $A^\prime$ constructed from (\ref{Aa}) according to (\ref{Ap}) leads to the correct value of $\tilde{c}$ only for $\alpha=\frac 12$. Namely, in this case we get $y_1=\rho$, $y_2=1-\rho$ and $\tilde{c} =\frac{6}{\pi^2} \bigl(L(\rho) - L(1-\rho)\bigr) =\frac 15$. It would be interesting to see if there are fermionic representations for characters of ${\cal M}(5,6)$ corresponding to other values of $\alpha$. Below we will show that $\alpha=0$ appears not in ${\cal M}(5,6)$ but in closely related ${\cal M}(3,10)$ model. \section{Further fermionic representations and identities.} So far we considered only ``irreducible" fermionic representations of characters, i.e.~those that are not decomposable into a product of other fermionic sums. However, there are many ways to construct ``reducible" fermionic representations. One of them was discussed in \cite{BF2,BF1} and based on the fact that any factorizable character can be brought to the form (\ref{F2}) (typically with $b_i\neq 1$) with the help of the following formulae \begin{equation}\label} \def\ba{\begin{eqnarray}{PF} \{x\}^\pm_y \= \sum_{m=0}^\infty \frac{(\pm 1)^m q^{\frac y2 (m^2-m) +mx}}{(q^y)_m} \,, \quad\quad\ \frac{1}{\{x\}^\pm_y} \= \sum_{m=0}^\infty \frac{ (\mp 1)^m q^{mx}}{(q^y)_m} \,. \end{equation}} \def\ea{\end{eqnarray} Another possibility is to use relations expressing a character as a product of other characters for which fermionic representation is known. To derive and prove such relations it is often convenient to exploit the factorized forms of characters. For instance, with the help of (\ref{2s}), (\ref{3s}) and (\ref{f3}) it is straightforward to check the following identities (see \cite{BF2} for a similar derivation) \ba && \!\!\!\! \chi^{2n,3m}_{n,m}(q) \chi^{3n,10m}_{n,5m}(q) \= \chi^{2n,5m}_{n,m}(q) \,\chi^{2n,5m}_{n,2m}(q) \,, \label{cc1}\\ && \!\!\!\! \chi^{2n,3m}_{n,m}(q) \, \Bigl( \chi^{3n,10m}_{n,m(2k-1)}(q) + \chi^{3n,10m}_{n,m(11-2k)}(q) \Bigr) \= \chi^{2n,5m}_{n,mk}(q) \,\chi^{2n,5m}_{n,mk}(q) \,, \quad k=1,2\,. \label{cc2} \ea Choosing here $n=m=1$ (recall that $\chi^{2,3}_{1,1}(q)=1$) and using the sum side of the Rogers-Ramanujan identities (\ref{RR}), we obtain the following formulae \ba && \!\!\!\!\!\!\!\!\!\!\!\!\!\!\! \chi^{3,10}_{1,5}(q) =\chi^{2,5}_{1,1}(q) \chi^{2,5}_{1,2}(q) = q^{\frac 16} \sum_{\vec{m}=\vec{0}}^\infty \frac{ q^{m_1^2 +m_2^2 + m_1}}{(q)_{m_1} (q)_{m_2}} \,, \label{cc3} \\ && \!\!\!\!\!\!\!\!\!\!\!\!\!\!\! \chi^{3,10}_{1,2k-1}(q)+\chi^{3,10}_{1,11-2k}(q) = \chi^{2,5}_{1,k}(q) \chi^{2,5}_{1,k}(q) = q^{\frac{23-12k}{30}} \sum_{\vec{m}=\vec{0}}^\infty \frac{q^{m_1^2 +m_2^2 + (4-2k)m_1}} {(q)_{m_1} (q)_{m_2}} , \label{cc4} \ea where $k=1,2$. These reducible fermionic representations correspond to the choice $\alpha=0$ for the matrix $A$ given by (\ref{Aa}). Using in the similar way the following identities for $k=1,2$ (they were found in \cite{Tao} and generalized to the form similar to (\ref{cc1})-(\ref{cc2}) in \cite{BF2}) \ba && \label{cc5} \chi_{2,k}^{4,5}(q) \= \chi_{1,2}^{3,4}(q) \, \Bigl(\chi_{1,(3-k)}^{6,5}(q) - \chi_{1,(2+k)}^{6,5}(q) \Bigr) \,, \\ && \label{cc6} \chi_{1,k}^{4,5}(q) \pm \chi_{3,k}^{4,5}(q) \= \Bigl(\chi_{1,1}^{3,4}(q) \pm \chi_{1,3}^{3,4}(q) \Bigr) \, \Bigl(\chi_{2,(3-k)}^{6,5}(q) \mp \chi_{2,(2+k)}^{6,5}(q) \Bigr) \ea and employing any of the fermionic representations for ${\cal M}(3,4)$ and ${\cal M}(5,6)$ discussed above, we get different reducible fermionic representations for ${\cal M}(4,5)$. For instance, substituting (\ref{34}) and (\ref{M56c}) into (\ref{cc5}) we obtain \begin{equation}\label} \def\ba{\begin{eqnarray}{cc7} \chi_{2,k}^{4,5}(q) \=q^{\eta^{4,5}_{2,k}} \sum_{\vec{m}=\vec{0}}^\infty \frac{ (-1)^{m_3} q^{\frac 12 (m_1^2 +m_2^2 +m_3^2)+m_2m_3+\frac 12 m_1 +(2-k)(m_2+m_3) } }{ (q)_{m_1} (q)_{m_2} (q)_{m_3} } \,,\quad k=1,2 \,. \end{equation}} \def\ea{\end{eqnarray} One more possibility to extend the list of fermionic representations is to consider relations between characters with rescaled argument $q$. For instance, we have \cite{BF2} \begin{equation}\label} \def\ba{\begin{eqnarray}{q1} \chi_{n,2}^{5,6}(q) + \chi_{n,4}^{5,6}(q) \= \chi_{1,n}^{2,5}(q^{\frac 12})\,, \quad\quad \chi_{n,1}^{5,6}(q) - \chi_{n,5}^{5,6}(q) \= \chi_{1,3-n}^{2,5}(q^2)\,, \quad n=1,2\,. \end{equation}} \def\ea{\end{eqnarray} Together with (\ref{RR}) this gives yet another fermionic representation for ${\cal M}(5,6)$ \begin{equation}\label} \def\ba{\begin{eqnarray}{q2} \chi_{n,2}^{5,6}(q) + \chi_{n,4}^{5,6}(q) \= \sum_{m=0}^\infty \frac{ q^{\frac 12 m^2 + (1-\frac n2)m}} {(q^{\frac 12})_m}\,, \quad\quad \chi_{n,1}^{5,6}(q) - \chi_{n,5}^{5,6}(q) \= \sum_{m=0}^\infty \frac{q^{2 m^2+(2n-2)m}}{(q^2)_m} \,. \end{equation}} \def\ea{\end{eqnarray} On the other hand, reading eqs.~(\ref{q1}) from the right to the left and using (\ref{M56b})-(\ref{M56c}), we obtain alternative sum sides for the Rogers-Ramanujan identities (\ref{RR}): \begin{equation}\label} \def\ba{\begin{eqnarray}{q3} \sum_{m=0}^\infty \frac{q^{m^2+(2-n)m}}{(q)_m} \= \sum_{\vec{m}=\vec{0}}^\infty \frac{ q^{(m_1 +m_2)^2 + m_1+(4-2n)m_2} }{ (q^2)_{m_1} (q^2)_{m_2}} \= \sum_{\vec{m}=\vec{0}}^\infty \frac{ (-1)^{m_2} q^{\frac 14 (m_1 + m_2)^2 + (1-\frac n2 )(m_1+m_2)} }{ (q^{\frac 12})_{m_1} (q^{\frac 12})_{m_2} } \,, \end{equation}} \def\ea{\end{eqnarray} where $n=1,2$. This sequence of identities can be continued further employing the fermionic representations (\ref{M56d})-(\ref{M56e}) and also those found in \cite{KKMM} (corresponding to (\ref{Aa}) with $\alpha=\frac 13$). Another set of relations observed in \cite{BF2} \begin{equation}\label} \def\ba{\begin{eqnarray}{q4} \chi_{2,n}^{6,7}(q^2) + \chi_{4,n}^{6,7}(q^2) \= \chi_{1,d_n}^{6,7}(q) - \chi_{5,d_n}^{6,7}(q) \,, \end{equation}} \def\ea{\end{eqnarray} where $d_1=3$, $d_2=1$, $d_3=2$, together with eqs.~(\ref{M67a})-(\ref{M67c}) can be used to get fermionic representations of the type (\ref{F2}) for the ${\cal M}(6,7)$ model. It is also possible to combine rescaling of $q$ and constructing of reducible fermionic sums. For example, let $p$ be a prime number such that $\langle p,t \rangle =1$. Then, using (\ref{char}), it is easy to derive the following relation \begin{equation}\label} \def\ba{\begin{eqnarray}{qp} \chi_{1,p}^{3,2p}(q) \, \chi_{n,m}^{s,t}(q^p) \= \chi_{pn,m}^{ps,t}(q) \,. \end{equation}} \def\ea{\end{eqnarray} Here $\chi_{1,p}^{3,2p}(q)$ should be replaced with $\chi_{1,3}^{2,9}(q)$ if $p=3$. If $p$ is not a prime number, generalization of (\ref{qp}) can be achieved by decomposing $p$ into proper factors. Now choosing $p=2$, $t=5$ and $s=2$ or $s=3$ in (\ref{qp}), we get for $k=1,2$ \begin{equation}\label} \def\ba{\begin{eqnarray}{qq2} \chi_{2,k}^{4,5}(q) =\chi_{1,k}^{2,5}(q^2) \chi_{1,2}^{3,4}(q) \,, \quad\quad \chi_{k,2}^{5,6}(q) \pm \chi_{k,4}^{5,6}(q) = \Bigl( \chi_{1,k}^{3,5}(q^2) \pm \chi_{1,5-k}^{3,5}(q^2) \Bigr) \chi_{1,2}^{3,4}(q) \,. \end{equation}} \def\ea{\end{eqnarray} Substituting here (\ref{2t}), (\ref{34}) and (\ref{k1b}), we obtain more fermionic representations of the type (\ref{F2}) for the ${\cal M}(4,5)$ and ${\cal M}(5,6)$ models: \ba && \chi_{2,k}^{4,5}(q) \= \sum_{\vec{m}=0}^\infty \frac{q^{2 m_1^2+\frac 12 m_2^2 + (4-2k)m_1 + \frac 12 m_2}} {(q^2)_{m_1} (q)_{m_2}} \,, \quad k=1,2 \,, \label{qq3} \\ && \chi_{k,2}^{5,6}(q) \pm \chi_{k,4}^{5,6}(q) \= \sum_{\vec{m}=0}^\infty \frac{(\pm 1)^{m_1}\, q^{\frac 12 (m_1^2 + m_2^2) +(2-k)m_1 + \frac 12 m_2}} {(q^2)_{m_1} (q)_{m_2}} \,, \quad k=1,2\,. \label{qq4} \ea \section{Discussion.} Having a character (linear combination of characters) in the fermionic form (\ref{F2}), we can rewrite it as a series $\chi(q)=\sum_{k=0}^{\infty} \mu_k q^k$, where the level $k$ admits partitioning, $k=\sum_{a=1}^r \sum_{i_a} p_a^{i_a}$, into parts of a specific form. The interpretation of the $p_a^{i_a}$ as momenta of massless particles gives rise to the quasi-particle picture, where a character is regarded as a partition function, $\chi(q)=\sum_k \mu_k e^{-\beta E_k}$. Here $q=e^{-2\pi \beta v/L}$, with $v$ being the speed of sound, and $L$ -- the size of the system. This quasi-particle representation was developed originally in \cite{KKMM} (for $b_i=1$) and became a standard technique by now. It is also applicable to factorized characters \cite{BF2,BF1,BF3} (in this case $b_i \neq 1$). For the fermionic form (\ref{F1}) or (\ref{F2}) of a character the quasi-particle representation involves $r$ quasi-particles. They are naturally interpreted as a conformal limit of particles presented in a massive theory related to the given conformal model. Moreover, it was suggested in \cite{BM} that different non-equivalent fermionic representations of the same character correspond to different integrable perturbations of the conformal model in question. It would be interesting to understand if the representations for ${\cal M}(4,5)$, ${\cal M}(5,6)$ and ${\cal M}(6,7)$ discussed above in Section 2 agree with this picture. Our results demonstrate that the number of non-equivalent fermionic representations for ${\cal M}(s,t)$ increases if $st$ can be represented as a product of two other co-prime numbers. For instance, we have encountered above three representations of the type (\ref{F1}) for ${\cal M}(5,6)$ besides the one considered in \cite{KKMM}. Furthermore, we can considerably expand the list of non-equivalent fermionic representations, if we are looking for representations of the type (\ref{F2}), including those that are reducible. For instance, in this way one obtains representations with one (\ref{q2}) and two quasi-particles (\ref{qq4}) for ${\cal M}(5,6)$ (another two-particle representation of the type (\ref{F2}) follows from the factorized characters \cite{BF2,BF3}). To summarize, in the present paper we have extended the list of fermionic representations for some minimal Virasoro models. Physical content of these representations, in particular, their connection with massive integrable models remains to be investigated. For all considered fermionic representations we have established Rogers-Ramanujan type identities and the corresponding dilogarithm identities. The Rogers-Ramanujan type identities possibly can be employed to construct various quasi-particle representations for certain physical entities arising in the lattice models of statistical mechanics. \vspace{0.5mm} After submission of this manuscript I was informed by the referee that connection of the matrix (\ref{M45a}) to the ${\cal M}(4,5)$ model was found earlier by M.~Terhoeven (unpublished). \vspace*{3mm} \noindent {\bf Acknowledgement:} \ I am grateful to A.~Fring for useful discussions. \vspace*{-3mm} \newcommand{\sbibitem}[1]{ \vspace*{-1.5ex} \bibitem{#1} }
\section*{Introduction} Five years ago ``precision cosmology'' would have been an oxymoron. That is not true today. Our knowledge of the cosmological parameters reflects that change. Once an embarrassment, they are quickly becoming a source of pride. The COBE FIRAS team led by John Mather determined the temperature of the cosmic microwave background radiation to four significant figures, $T_0 = (2.7277\pm 0.00001\pm 0.002)\,$K. Here, the first error is the error relative to the internal reference black body and the second error, which dominates the error budget, is the uncertainty in determining the temperature of the internal reference black body (Fixsen et al, 1996). Had the on-board thermometers performed better, they might have had another couple of significant figures (which boggles the mind!). In any case, this is a truly impressive measurement, cosmological or otherwise. For the first time in thirty years, there is agreement (within the uncertainties) on the value of the Hubble constant and a reliable estimate of the uncertainty. {\it My summary} of the recent measurements is:\footnote{I use standard conventions for the meaning of error bars: $1\sigma$ bet your graduate student's reputation; $2\sigma$ bet the family pet; $3\sigma$ bet a child (other than first born male).} $H_0 = (65\pm 5)\,{\rm km\,s^{-1}\,Mpc^{-1}}$ (see Freedman, 1999 for a more expert assessment of the situation). The improvement is due in large measure to the Hubble Key Project, which has established Cepheid calibrations for the secondary distance indicators and just as importantly stimulated renewed interest in the examining the distance scale. Systematic uncertainties remain; foremost are the distance to the LMC and reddening and metallicity effects on the Cepheid period-luminosity relation. Similar progress has been made in dating the age of the oldest objects in the Universe: globular clusters, the heavy elements and white dwarfs (see e.g., Chaboyer, 1998). Allowing for a time of a 2\,Gyr until the formation of the first generation of stars, the age of the Universe is $t_0=(14\pm 1.5)\,$Gyr. While the hard lower limit to the age of 10\,Gyr derives from all three methods, the central value and error is driven by age dating of globular clusters. Here progress has occurred due to better stellar modelling, improvement in the distance scale and better microphysics. Regarding the distance scale; the parallax distance measurements made by Hipparcos played a role, though, probably more important was the fact that it spurred scrutiny of the entire enterprise. I devote the rest of this article to a summary of our knowledge of the matter and energy content of the Universe. \begin{figure} \centerline{\psfig{figure=omega_sum.eps,width=5in}} \caption{Summary of matter/energy in the Universe. The right side refers to an overall accounting of matter and energy; the left refers to the composition of the matter component. The contribution of relativistic particles, CBR photons and neutrinos, $\Omega_{\rm rel}h^2 = 4.170\times 10^{-5}$, is not shown. The upper limit to mass density contributed by neutrinos is based upon the failure of the hot dark matter model of structure formation (White, Frenk \& Davis, 1983; and Dodelson et al, 1996) and the lower limit follows from the evidence for neutrino oscillations (Fukuda et al, 1998). Here $H_0$ is taken to be $65\,{\rm km\,s^{-1}\,Mpc^{-1}}$. } \label{fig:omega} \end{figure} \section*{An Inventory of Matter and Energy} \subsection*{Preliminaries} The quantity and composition of matter and energy in the Universe is of fundamental importance in cosmology. The fraction of the critical energy density contributed by all forms of matter and energy, \begin{equation} \Omega_0 \equiv {\rho_{\rm tot} \over \rho_{\rm crit}} = \sum_i \Omega_i \,, \end{equation} determines the geometry of the Universe: \begin{equation} R_{\rm curv}^2 = {H_0^{-2} \over \Omega_0 - 1}\,. \end{equation} Here, subscript `0' denotes the value at the present epoch, $\rho_{\rm crit} = 3H_0^2/8\pi G \simeq 1.88h^2\times 10^{-29}\,{\rm g\ cm^{-3}}$, $\Omega_i$ is the fraction of critical density contributed by component $i$ (e.g., baryons, photons, stars, etc) and $H_0 = 100h\,{\rm km\,s^{-1}\, Mpc^{-1}}$. The sign of $R_{\rm curv}^2$ specifies the spatial geometry: positive for a 3-sphere, negative for a 3-saddle and 0 for the flat space. The present rate of deceleration of the expansion depends upon $\Omega_0$ as well as the composition of matter and energy, \begin{equation} q_0 \equiv {(\ddot R /R)_0 \over H_0^2} = {1 \over 2}\Omega_0 + {3\over 2} \sum_i \Omega_i w_i \,. \end{equation} The pressure of component $i$, $p_i \equiv w_i \rho_i$; e.g., for baryons $w_i = 0$, for radiation $w_i = 1/3$, and for vacuum energy $w_i = -1$. The fate of the Universe -- expansion forever or recollapse -- is not directly determined by $\Omega_0$ and/or $q_0$. It depends upon precise knowledge of the composition of {\em all} components of matter and energy, for all times in the future. Recollapse occurs only if there is a future turning point, that is an epoch when the expansion rate, \begin{equation} H^2 = {8\pi G \over 3} \sum_i \rho_i - {1\over R_{\rm curv}^2}\,, \end{equation} vanishes and \begin{equation} {\ddot R \over R} = -{4\pi G\over 3} \sum_i\,\rho_i[1+w_i] \end{equation} is less than zero. In a universe comprised of matter alone, a positively curved universe ($\Omega_0 > 1$) eventually recollapses and a negatively curved universe ($\Omega_0 < 1$) expands forever. However, exotic components complicate matters: a positively curved universe with positive vacuum energy can expand forever, and a negatively curved universe with negative vacuum energy can recollapse. The quantity and composition of matter and energy in the Universe is also crucial for understanding the past. It determines the relationship between age of the Universe and redshift, when the Universe ended its early radiation dominated era, and the growth of small inhomogeneities in the matter. Ultimately, the formation and evolution of large-scale structure and even individual galaxies depends upon the composition of the dark matter and energy. Measuring the quantity and composition of matter and energy in the Universe is a challenging task. Not just because the scale of inhomogeneity is so large, around $10\,$Mpc; but also, because there may be components that remain exactly or relatively smooth (e.g., vacuum energy or relativistic particles) and only reveal their presence by their influence on the evolution of the Universe itself. \subsubsection*{Radiation} Because the cosmic background radiation (CBR) is known to be black-body radiation to very high precision (better than $0.005\%$) and its temperature is known to four significant figures, $T_0 = 2.7277\pm 0.002\,$K, its contribution is very precisely known, $\Omega_\gamma h^2 = 2.480\times 10^{-5}$. If neutrinos are massless or very light, $m_\nu \ll 10^{-4}\,$eV, their energy density is equally well known because it is directly related to that of the photons, $\Omega_\nu = {7\over 8}(4/11)^{4/3} \Omega_\gamma$ (per species) (there is a small 1\% positive correction to this number; see Dodelson \& Turner, 1992). It is possible that additional relativistic species contribute significantly to the energy density, though big-bang nucleosynthesis (BBN) severely constrains the amount (the equivalent of less than 0.4 of a neutrino species; see e.g., Burles et al, 1999) unless they were produced by the decay of a massive particle after the epoch of BBN. In any case, we can be confident that the radiation component of the energy density today is small. The matter contribution (denoted by $\Omega_M$), consisting of particles that have negligible pressure, is the easiest to determine because matter clumps and its gravitational effects are thereby enhanced (e.g., in rich clusters the matter density typically exceeds the mean density by a factor of 1000 or more). With this in mind, I will decompose the present matter/energy density into two components, matter and vacuum energy, \begin{equation} \Omega_0 = \Omega_M + \Omega_\Lambda\,, \end{equation} ignoring the contribution of the CBR and ultrarelativistic neutrinos. I will use vacuum energy as a stand in for a smooth component (more later). Vacuum energy and a cosmological constant are indistinguishable: a cosmological constant corresponds to a uniform energy density of magnitude $\rho_{\rm vac} = \Lambda /8\pi G$. \subsection*{Total Matter/Energy Density: $\Omega_0 = 1\pm 0.2$} There is a growing consensus that the anisotropy of the CBR offers the best means of determining the curvature of the Universe and thereby $\Omega_0$, cf., Eq. (2). This is because the method is intrinsically geometric -- a standard ruler on the last-scattering surface -- and involves straightforward physics at a simpler time (see e.g., Kamionkowski et al, 1994). \begin{figure} \centerline{\psfig{figure=knox_better.eps,width=5in}} \caption{Summary of CBR anisotropy measurements, binned to reduce error bars and visual confusion. The temperature variations across the sky have been expanded in spherical harmonics, $\delta T(\theta , \phi ) = \sum_i a_{lm}Y_{lm} (\theta ,\phi )$ and $C_l \equiv \langle |a_{lm}|^2\rangle$. In plain language, this plot shows the size of the temperature variation between two points on the sky separated by angle $\theta$ (ordinate) vs. multipole number $l=200^\circ / \theta$ ($l=2$ corresponds to $100^\circ$, $l=200$ corresponds to $\theta = 1^\circ$, and so on). The theoretical curve is for the $\Lambda$CDM model with $H_0=65\,{\rm km\, s^{-1}\,Mpc^{-1}}$ and $\Omega_M =0.4$, which is a good fit to the data. Note also the evidence for the first of a series of ``acoustic peaks.'' The presence of these acoustic peaks is a key signature of the density perturbations of quantum origin predicted by inflation (Figure courtesy of L. Knox). } \label{fig:cbr_knox} \end{figure} At last scattering baryonic matter (ions and electrons) was still tightly coupled to photons; as the baryons fell into the dark-matter potential wells the pressure of photons acted as a restoring force, and gravity-driven acoustic oscillations resulted. These oscillations can be decomposed into their Fourier modes; Fourier modes with $k\sim l H_0/2$ determine the multipole amplitudes $a_{lm}$ of CBR anisotropy. Last scattering occurs over a short time, making the CBR is a snapshot of the Universe at $t_{\rm ls} \sim 300,000\,$yrs. Each mode is ``seen'' in a well defined phase of its oscillation. (For the density perturbations predicted by inflation, all modes the have same initial phase because all are growing-mode perturbations.) Modes caught at maximum compression or rarefaction lead to the largest temperature anisotropy; this results in a series of acoustic peaks beginning at $l\sim 200$ (see Fig.~\ref{fig:cbr_knox}). The wavelength of the lowest frequency acoustic mode that has reached maximum compression, $\lambda_{\rm max} \sim v_s t_{\rm ls}$, is the standard ruler on the last-scattering surface. Both $\lambda_{\rm max}$ and the distance to the last-scattering surface depend upon $\Omega_0$, and the position of the first peak $l\simeq 200/\sqrt{\Omega_0}$. This relationship is insensitive to the composition of matter and energy in the Universe. CBR anisotropy measurements, shown in Fig.~\ref{fig:cbr_knox}, now cover three orders of magnitude in multipole number and involve more than twenty experiments. COBE is the most precise and covers multipoles $l=2-20$; the other measurements come from balloon-borne, Antarctica-based and ground-based experiments using both low-frequency ($f<100\,$GHz) HEMT receivers and high-frequency ($f>100\,$GHz) bolometers. Taken together, all the measurements are beginning to define the position of the first acoustic peak, at a value that is consistent with a flat Universe. Various analyses of the extant data have been carried out, indicating $\Omega_0 \sim 1\pm 0.2$ (see e.g., Lineweaver, 1998). It is certainly too early to draw definite conclusions or put too much weigh in the error estimate. However, a strong case is developing for a flat Universe and more data is on the way (Python V, Viper, MAT, Maxima, Boomerang, DASI, and others). Ultimately, the issue will be settled by NASA's MAP (launch late 2000) and ESA's Planck (launch 2007) satellites which will map the entire CBR sky with 30 times the resolution of COBE (around $0.1^\circ$) (see Page \& Wilkinson, 1999). \subsection*{Matter} \subsubsection*{Baryons} For more than twenty years big-bang nucleosynthesis has provided a key test of the hot big-bang cosmology as well as the most precise determination of the baryon density. Careful comparison of the primeval abundances of D, $^3$He, $^4$He and $^7$Li with their big-bang predictions defined a concordance interval, $\Omega_Bh^2 = 0.007 - 0.024$ (see e.g., Copi et al, 1995; for another view, see Hata et al, 1995). Of the four light elements produced in the big bang, deuterium is the most powerful ``baryometer'' -- its primeval abundance depends strongly on the baryon density ($\propto 1/\rho_B^{1.7}$) -- and the evolution of its abundance since the big bang is simple -- astrophysical processes only destroy deuterium. Until recently deuterium could not be exploited as a baryometer because its abundance was only known locally, where roughly half of the material has been through stars with a similar amount of the primordial deuterium destroyed. In 1998, the situation changed dramatically, launching BBN into the precision era of cosmology. Over the past four years there have been claims of upper limits, lower limits, and determinations of the primeval deuterium abundance, ranging from (D/H)\,$=10^{-5}$ to (D/H)\,$=3\times 10^{-4}$. In short, the situation was confusing. In 1998 Burles and Tytler clarified matters and established a strong case for (D/H)$_P = (3.4\pm 0.3)\times 10^{-5}$ (Burles \& Tytler, 1998a,b). That case is based upon the deuterium abundance measured in four high-redshift hydrogen clouds seen in absorption against distant QSOs, and the remeasurement and reanalysis of other putative deuterium systems. In this enterprise, the Keck I 10-meter telescope and its HiRes Echelle Spectrograph have played the crucial role. The Burles -- Tytler measurement turns the previous factor of three concordance range for the baryon density into a 10\% determination of the baryon density, $\rho_B = (3.8\pm 0.4)\times 10^{-31}\,{\rm g\,cm^{-3}}$ or $\Omega_Bh^2 = 0.02 \pm 0.002$ (see Fig.~\ref{fig:bbn}), with about half the error in $\rho_B$ coming from the theoretical error in predicting the BBN yield of deuterium. [A very recent analysis has reduced the theoretical error significantly, and improved the accuracy of the determination of the baryon density, $\Omega_Bh^2 = 0.019\pm 0.0012$ (Burles et al, 1999).] \begin{figure} \centerline{\psfig{figure=bbn_new.eps,width=5in}} \caption{Predicted abundances of $^4$He (mass fraction), D, $^3$He, and $^7$Li (number relative to hydrogen) as a function of the baryon density; widths of the curves indicate ``$2\sigma$'' theoretical uncertainty. The dark band highlights the determination of the baryon density based upon the recent measurement of the primordial abundance of deuterium (Burles \& Tytler, 1998a,b), $\Omega_Bh^2 = 0.019 \pm 0.0024$ (95\% cl); the baryon density is related to the baryon-to-photon ratio by $\rho_B = 6.88\eta \times 10^{-22}\,{\rm g\,cm^{-3}}$ (from Burles et al, 1999). } \label{fig:bbn} \end{figure} This precise determination of the baryon density, based upon the early Universe physics of BBN, is consistent with two other measures of the baryon density, based upon entirely different physics. By comparing measurements of the opacity of the Lyman-$\alpha$ forest toward high-redshift quasars with high-resolution, hydrodynamical simulations of structure formation, several groups (Meiksin \& Madau, 1993; Rauch et al, 1997; Weinberg et al, 1997) have inferred a lower limit to the baryon density, $\Omega_Bh^2 > 0.015$ (it is a lower limit because it depends upon the baryon density squared divided by the intensity of the ionizing radiation field). The second test involves the height of the first acoustic peak: it rises with the baryon density (the higher the baryon density, the stronger the gravitational force driving the acoustic oscillations). Current CBR measurements are consistent with the Burles -- Tytler baryon density; the MAP and Planck satellites should ultimately provide a 5\% or better determination of the baryon density, based upon the physics of gravity-driven acoustic oscillations when the Universe was 300,000\,yrs old. This will be an important cross check of the BBN determination. \subsubsection*{Weighing the dark matter: $\Omega_M = 0.4\pm 0.1$} Since the pioneering work of Fritz Zwicky and Vera Rubin, it has been known that there is far too little material in the form of stars (and related material) to hold galaxies and clusters together, and thus, that most of the matter in the Universe is dark (see e.g. Trimble, 1987). Weighing the dark matter has been the challenge. At present, I believe that clusters provide the most reliable means of estimating the total matter density. Rich clusters are relatively rare objects -- only about 1 in 10 galaxies is found in a rich cluster -- which formed from density perturbations of (comoving) size around 10\,Mpc. However, because they gather together material from such a large region of space, they can provide a ``fair sample'' of matter in the Universe. Using clusters as such, the precise BBN baryon density can be used to infer the total matter density (White et al, 1993). (Baryons and dark matter need not be well mixed for this method to work provided that the baryonic and total mass are determined over a large enough portion of the cluster.) Most of the baryons in clusters reside in the hot, x-ray emitting intracluster gas and not in the galaxies themselves, and so the problem essentially reduces to determining the gas-to-total mass ratio. The gas mass can be determined by two methods: 1) measuring the x-ray flux from the intracluster gas and 2) mapping the Sunyaev - Zel'dovich CBR distortion caused by CBR photons scattering off hot electrons in the intracluster gas. The total cluster mass can be determined three independent ways: 1) using the motions of clusters galaxies and the virial theorem; 2) assuming that the gas is in hydrostatic equilibrium and using it to infer the underlying mass distribution; and 3) mapping the cluster mass directly by gravitational lensing (Tyson, 1999). Within their uncertainties, and where comparisons can be made, the three methods for determining the total mass agree (see e.g., Tyson, 1999); likewise, the two methods for determining the gas mass are consistent. \begin{figure} \centerline{\psfig{figure=cbf2.eps,width=5in}} \caption{Cluster gas fraction as a function of cluster gas temperature for a sample of 45 galaxy clusters (Mohr et al, 1998). While there is some indication that the gas fraction decreases with temperature for $T< 5\,$keV, perhaps because these lower-mass clusters lose some of their hot gas, the data indicate that the gas fraction reaches a plateau at high temperatures, $f_{\rm gas} =0.212 \pm 0.006$ for $h=0.5$ (Figure courtesy of Joe Mohr). } \label{fig:gas} \end{figure} Mohr et al (1998) have compiled the gas to total mass ratios determined from x-ray measurements for a sample of 45 clusters; they find $f_{\rm gas} = (0.075\pm 0.002)h^{-3/2}$. Carlstrom (1999), using his S-Z gas measurements and x-ray measurements for the total mass for 27 clusters, finds $f_{\rm gas} =(0.06\pm 0.006)h^{-1}$. (The agreement of these two numbers means that clumping of the gas, which could lead to an overestimate of the gas fraction based upon the x-ray flux, is not a problem.) Invoking the ``fair-sample assumption,'' the mean matter density in the Universe can be inferred: \begin{eqnarray} \Omega_M = \Omega_B/f_{\rm gas} & = & (0.3\pm 0.05)h^{-1/2}\ ({\rm X ray})\nonumber\\ & = & (0.25\pm 0.04)h^{-1}\ ({\rm S-Z}) \nonumber \\ & = & 0.4\pm 0.1\ ({\rm my\ summary})\,. \end{eqnarray} I believe this to be the most reliable and precise determination of the matter density. It involves few assumptions, and most of them have now been tested (clumping, hydrostatic equilibrium, variation of gas fraction with cluster mass). \subsubsection*{Supporting evidence for $\Omega_M=0.4\pm 0.1$} This result is consistent with a variety of other methods that involve different physics. For example, based upon the evolution of the abundance of rich clusters with redshift, Henry (1999) finds $\Omega_M = 0.45 \pm 0.1$ (also see, Bahcall \& Fan, 1998 and N. Bahcall, 1999). Dekel and Rees (1994) place a low limit $\Omega_M > 0.3$ (95\% cl) derived from the outflow of material from voids (a void effectively acts as a negative mass proportional to the mean matter density). The analysis of the peculiar velocities of galaxies provides an important probe of the mass density averaged over very large scales (of order several hundred Mpc). By comparing measured peculiar velocities with those predicted from the distribution of matter revealed by redshift surveys such as the IRAS survey of infrared galaxies, one can infer the quantity $\beta = \Omega_M^{0.6}/b_I$ where $b_I$ is the linear bias factor that relates the inhomogeneity in the distribution of IRAS galaxies to that in the distribution of matter (in general, the bias factor is expected to be in the range 0.7 to 1.5; IRAS galaxies are expected to be less biased, $b_I\approx 1$.). Recent work by Willick \& Strauss (1998) finds $\beta = 0.5\pm 0.05$, while Sigad et al (1998) find $\beta = 0.9\pm 0.1$. The apparent inconsistency of these two results and the ambiguity introduced by bias preclude a definitive determination of $\Omega_M$ by this method. However, Dekel (1994) quotes a 95\% confidence lower bound, $\Omega_M >0.3$, and the work of Willick \& Strauss seems to strongly indicate that $\Omega_M$ is much less than 1; both are consistent with $\Omega_M\sim 1$. Finally, there is strong, but circumstantial, evidence from structure formation that $\Omega_M$ is around 0.4 and significantly greater than $\Omega_B$. It is based upon three different lines of reasoning. First, there is no viable model for structure formation without a significant amount of nonbaryonic dark matter. The reason is simple: in a baryons-only model, density perturbations grow only from the time of decoupling, $z\sim 1000$, until the Universe becomes curvature dominated, $z\sim \Omega_B^{-1} \sim 20$; this is not enough growth to produce all the structure seen today with the size of density perturbations inferred from CBR anisotropy. With nonbaryonic dark matter and $\Omega_M \gg \Omega_B$, dark-matter perturbations begin growing at matter -- radiation equality and continue to grow until the present epoch, or nearly so, leading to significantly more growth and making the observed large-scale structure consistent with the size of the density perturbations inferred from CBR anisotropy. Second, the transition from radiation domination at early times to matter domination at around 10,000\,yrs leaves its mark on the shape of the present power spectrum of density perturbations, and the redshift of matter -- radiation equality depends upon $\Omega_M$. Measurements of the shape of the present power spectrum based upon redshift surveys indicate that the shape parameter $\Gamma = \Omega_M h \simeq 0.25\pm 0.05$ (see e.g., Peacock \& Dodds, 1994). For $h\sim 2/3$, this implies $\Omega_M \sim 0.4$. (If there are relativistic particles beyond the CBR photons and relic neutrinos, the formula for the shape parameter changes and $\Omega_M\gg 0.4$ can be accommodated; see Dodelson et al, 1996). Third, the Ly-$\alpha$ forest, seen in the absorption spectra of high-redshift quasars, gives us a unique view of the Universe at a time when structure was just becoming nonlinear. Based upon a comparison of high-resolution $N$-body simulations with 28 Keck HIRES spectra, Weinberg et al (1998) infer $\Omega_M = 0.34\pm 0.1$. \subsection*{Mass-to-light ratios and $\Omega_M$} The most mature approach to estimating the matter density involves the use of mass-to-light ratios, the measured luminosity density, and the deceptively simple equation, \begin{equation} \langle \rho_M \rangle = \langle M/L \rangle \, {\cal L}\,, \end{equation} where ${\cal L}= 2.4h\times 10^8\,L_{B\odot}\, {\rm Mpc^{-3}}$ is the measured (B-band) luminosity density of the Universe. Once the average mass-to-light ratio for the Universe is determined, $\Omega_M$ follows by dividing it by the critical mass-to-light ratio, $(M/L)_{\rm crit} = 1200h$ (in solar units). Though it is tantalizingly simple -- and it is far too easy to take any measured mass-to-light ratio and divide it by $1200h$ -- this method does not provide an easy and reliable method of determining $\Omega_M$. Based upon the mass-to-light ratios of the bright, inner regions of galaxies, $(M/L)_*\sim $\ few, the fraction of critical density in stars (and closely related material) has been determined, $\Omega_* \simeq (0.003\pm 0.001)h^{-1}$ (see e.g., Faber \& Gallagher, 1979). Persic \& Salucci (1992) derive a similar value based upon the observed stellar-mass function. Luminous matter accounts for only a tiny fraction of the total mass density and only about a tenth of the baryons. CNOC (Carlberg et al, 1996, 1997) have done a very careful job of determining a mean cluster mass-to-light ratio, $(M/L)_{\rm cluster} = 240\pm 50$, which translates to an estimate of the mean matter density, $\Omega_{\rm cluster} = 0.20 \pm 0.04$. Because clusters contain thousands of galaxies and cluster galaxies do not seem {\em radically} different from field galaxies, one is tempted to take this estimate of the mean matter density seriously. However, it is significantly smaller than the value I advocated earlier, $\Omega_M = 0.4\pm 0.1$. Which estimate is right? I believe the higher number, based upon the cluster baryon fraction, is more reliable and that we should be surprised that the CNOC number is so close, closer than we had any right to expect! After all, only a small fraction of galaxies are found in clusters and the luminosity density ${\cal L}$ itself evolves strongly with redshift and corrections for this effect are large and uncertain. (We are on the tail end of star formation in the Universe: 80\% of star formation took place at a redshift greater than unity.) While the value for $\Omega_M$ derived from the cluster baryon fraction also relies upon clusters, the underlying assumption is far weaker and much more justified, namely that clusters provide a fair sample of matter in the Universe. Even if mass-to-light ratios were measured in the red (they typically are not), where the light is dominated by low-mass stars and reflects the integrated history of star formation rather than the current star-formation rate as it does in the blue, one would still require the fraction of baryons converted into stars in clusters to be identical to that in the field to have agreement between the CNOC estimate and that based upon the cluster baryon fraction. Apparently, the fraction of baryons converted into stars in the field and in clusters is similar, but not identical. To put this in perspective and to emphasize the shortcomings of the mass-to-light technique, had one used the cluster mass-to-x-ray ratio and the x-ray luminosity density, one would have inferred $\Omega_M \sim 0.05$. A factor of two discrepancy based upon optical mass-to-light ratios does not seem so bad. Enough said. \subsection*{Missing energy?} The results $\Omega_0 = 1\pm 0.2$ and $\Omega_M = 0.4\pm 0.1$ are in apparent contradiction, suggesting that one or both are wrong. However, prompted by a strong belief in a flat Universe, theorists have explored the remaining logical possibility: a dark, exotic form of energy that is smoothly distributed and contributes 60\% of the critical density (Turner et al, 1984; Peebles, 1984). Being smoothly distributed its presence would not have been detected in measurements of the matter density. The properties of this missing energy are severely constrained by other cosmological facts, including structure formation, the age of the Universe, and CBR anisotropy. So much so, that a smoking-gun signature for the missing energy was predicted (see e.g., Turner, 1991). To begin, let me parameterize the bulk equation of state of this unknown component: $w = p_X/\rho_X$. This implies that its energy density evolves as $\rho_X \propto R^{-n}$ where $n=3(1+w)$. The development of the structure observed today from density perturbations of the size inferred from measurements of the anisotropy of the CBR requires that the Universe be matter dominated from the epoch of matter -- radiation equality until very recently. Thus, to avoid interfering with structure formation, the dark-energy component must be less important in the past than it is today. This implies that $n$ must be less than $3$ or $w< 0$; the more negative $w$ is, the faster this component gets out of the way (see Fig.~\ref{fig:xmatter}). More careful consideration of the growth of structure implies that $w$ must be less than about $-{1\over 3}$ (Turner \& White, 1997). Next, consider the constraint provided by the age of the Universe and the Hubble constant. Their product, $H_0t_0$, depends the equation of state of the Universe; in particular, $H_0t_0$ increases with decreasing $w$ (see Fig.~\ref{fig:wage}). To be definite, I will take $t_0 =14\pm 1.5\,$Gyr and $H_0=65\pm 5\,{\rm km\,s^{-1}\,Mpc^{-1}}$ (see e.g., Chaboyer, 1998 and Freedman, 1999); this implies that $H_0t_0 = 0.93 \pm 0.13$. Fig.~\ref{fig:wage} shows that $w<-{1\over 2}$ is preferred by age/Hubble constant considerations. \begin{figure} \centerline{\psfig{figure=xmatter.eps,width=5in}} \caption{Evolution of the energy density in matter, radiation (heavy lines), and different possibilities for the dark-energy component ($w=-1,-{1\over 3},{1\over 3}$) vs. scale factor. The matter-dominated era begins when the scale factor was $\sim 10^{-4}$ of its present size (off the figure) and ends when the dark-energy component begins to dominate, which depends upon the value of $w$: the more negative $w$ is, the longer the matter-dominated era in which density perturbations can go into the large-scale structure seen today. These considerations require $w<-{1\over 3}$ (Turner \& White, 1997). } \label{fig:xmatter} \end{figure} \begin{figure} \centerline{\psfig{figure=wage.eps,width=5in}} \caption{$H_0t_0$ vs. the equation of state for the dark-energy component. As can be seen, an added benefit of a component with negative pressure is an older Universe for a given Hubble constant. The broken horizontal lines denote the $1\sigma$ range for $H_0=65\pm 5\,{\rm km\,s^{-1}\,Mpc^{-1}}$ and $t_0=14\pm 1.5\,$Gyr, and indicate that $w<-{1\over 2}$ is preferred. } \label{fig:wage} \end{figure} To summarize, consistency between $\Omega_M \sim 0.4$ and $\Omega_0 \sim 1$ along with other cosmological considerations implies the existence of a dark-energy component with bulk pressure more negative than about $-\rho_X /2$. The simplest example of such is vacuum energy (Einstein's cosmological constant), for which $w=-1$. The smoking-gun signature of a smooth, dark-energy component is accelerated expansion since $q_0 = 0.5 + 1.5w\Omega_X \simeq 0.5 + 0.9w < 0$ for $w<-{5\over 9}$. \subsection*{Missing energy found!} In 1998 evidence for accelerated expansion was presented in the form of the magnitude -- redshift (Hubble) diagram for fifty-some type Ia supernovae (SNe Ia) out to redshifts of nearly 1. Two groups, the Supernova Cosmology Project (Perlmutter et al, 1998) and the High-z Supernova Search Team (Riess et al, 1998; Schmidt et al, 1998), working independently and using different methods of analysis, each found evidence for accelerated expansion. Perlmutter et al (1998) summarize their results as a constraint to a cosmological constant (see Fig.~\ref{fig:omegalambda}), \begin{equation} \Omega_\Lambda = {4\over 3}\Omega_M +{1\over 3} \pm {1\over 6}\,. \end{equation} For $\Omega_M\sim 0.4 \pm 0.1$, this implies $\Omega_\Lambda = 0.85 \pm 0.2$, or just what is needed to account for the missing energy! (A simple explanation of the SN Ia results may be useful. If galactic distances and velocities were measured today they would obey a perfect Hubble law, $v_0=H_0d$, because the expansion of the Universe is simply a rescaling. Because we see distant galaxies at an earlier time, their velocities should be larger than predicted by the Hubble law, provided the expansion is slowing due to the attractive force of gravity. Using SNe Ia as standard candles to determine the distances to faraway galaxies, the two groups in effect found the opposite: distant galaxies are moving slower than predicted by the Hubble law, implying the expansion is speeding up!) Recently, two other studies, one based upon the x-ray properties of rich clusters of galaxies (Mohr et al, 1999) and the other based upon the properties of double-lobe radio galaxies (Guerra et al, 1998), have reported evidence for a cosmological constant (or similar dark-energy component) that is consistent with the SN Ia results (i.e., $\Omega_\Lambda \sim 0.7$). There is another powerful test of an accelerating Universe whose results are more ambiguous. It is based upon the fact that the frequency of multiply lensed QSOs is expected to be significantly higher in an accelerating universe (Turner, 1990). Kochanek (1996) has used gravitational lensing of QSOs to place a 95\% cl upper limit, $\Omega_\Lambda < 0.66$; and Waga and Miceli (1998) have generalized it to a dark-energy component with negative pressure: $\Omega_X < 1.3 + 0.55w$ (95\% cl), both results for a flat Universe. On the other hand, Chiba and Yoshii (1998) claim evidence for a cosmological constant, $\Omega_\Lambda = 0.7^{+0.1}_{-0.2}$, based upon the same data. From this I conclude: 1) lensing excludes $\Omega_\Lambda$ larger than 0.8, and 2) when larger objective surveys of gravitational-lensed QSOs are carried out (e.g., the Sloan Digital Sky Survey), there is the possibility of uncovering another smoking-gun for accelerated expansion. By far, the strongest evidence for dark energy is the SN Ia data. The statistical errors reported by the two groups are smaller than possible systematic errors. Thus, the believability of the results turn on the reliability of SNe Ia as one-parameter standard candles. SNe Ia are thought to be associated with the nuclear detonation of Chandrasekhar-mass white dwarfs. The one parameter is the rate of decline of the light curve: The brighter ones decline more slowly (the so-called Phillips relation; see Phillips, 1993). The lack of a good theoretical understanding of this (e.g., what is the physical parameter?) is offset by strong empirical evidence for the relationship between peak brightness and rate of decline, based upon a sample of nearby SNe Ia. It is reassuring that in all respects studied, the distant sample of SNe Ia appear to be similar to the nearby sample. For example, distribution of decline rates and dispersion about the Phillips relationship. The local sample spans a range of metallicity, likely spanning that of the distant sample, and further, suggesting that metallicity is not an important second parameter. At this point, it is fair to say that if there is a problem with SNe Ia as standard candles, it must be subtle. Cosmologists are even more inclined to believe the SN Ia results because of the preexisting evidence for a ``missing-energy component'' that led to the prediction of accelerated expansion. \subsection*{Cosmic concordance} With the SN Ia results we have for the first time a complete and self-consistent accounting of mass and energy in the Universe (see Fig.~\ref{fig:omega}). The consistency of the matter/energy accounting is illustrated in Fig.~\ref{fig:omegalambda}. Let me explain this exciting figure. The SN Ia results are sensitive to the acceleration (or deceleration) of the expansion and constrain the combination ${4\over 3}\Omega_M -\Omega_\Lambda$. (Note, $q_0 = {1\over 2}\Omega_M - \Omega_\Lambda$; ${4\over 3}\Omega_M - \Omega_\Lambda$ corresponds to the deceleration parameter at redshift $z\sim 0.4$, the median redshift of these samples). The (approximately) orthogonal combination, $\Omega_0 = \Omega_M + \Omega_\Lambda$ is constrained by CBR anisotropy. Together, they define a concordance region around $\Omega_0\sim 1$, $\Omega_M \sim 1/3$, and $\Omega_\Lambda \sim 2/3$. The constraint to the matter density alone, $\Omega_M = 0.4\pm 0.1$, provides a cross check, and it is consistent with these numbers. Cosmic concordance! \begin{figure} \centerline{\psfig{figure=omegalambda.eps,width=5in}} \caption{Two-$\sigma$ constraints to $\Omega_M$ and $\Omega_\Lambda$ from CBR anisotropy, SNe Ia, and measurements of clustered matter. Lines of constant $\Omega_0$ are diagonal, with a flat Universe shown by the broken line. The concordance region is shown in bold: $\Omega_M\sim 1/3$, $\Omega_\Lambda \sim 2/3$, and $\Omega_0 \sim 1$. (Particle physicists who rotate the figure by $90^\circ$ will recognize the similarity to the convergence of the gauge coupling constants.) } \label{fig:omegalambda} \end{figure} \begin{figure} \centerline{\psfig{figure=kraus.eps,width=5in}} \caption{Constraints used to determine the best-fit CDM model: PS = large-scale structure + CBR anisotropy; AGE = age of the Universe; CBF = cluster-baryon fraction; and $H_0$= Hubble constant measurements. The best-fit model, indicated by the darkest region, has $H_0\simeq 60-65\,{\rm km\,s^{-1} \,Mpc^{-1}}$ and $\Omega_\Lambda \simeq 0.55 - 0.65$. Evidence for its smoking-gun signature -- accelerated expansion -- was presented in 1998 (adapted from Krauss \& Turner, 1995 and Turner, 1997).} \label{fig:best_fit} \end{figure} But there is more. We also have a consistent and well motivated picture for the formation of structure in the Universe, $\Lambda$CDM. The $\Lambda$CDM model, which is the cold dark matter model with $\Omega_B \sim 0.05$, $\Omega_{\rm CDM}\sim 0.35$ and $\Omega_\Lambda \sim 0.6$, is a very good fit to all cosmological constraints: large-scale structure, CBR anisotropy, age of the Universe, Hubble constant and the constraints to the matter density and cosmological constant; see Fig.~\ref{fig:best_fit} (Krauss \& Turner, 1995; Ostriker \& Steinhardt, 1995; Turner, 1997). Further, as can be seen in Fig.~\ref{fig:cbr_knox}, CBR anisotropy measurements are beginning to show evidence for the acoustic peaks characteristic of the Gaussian, curvature perturbations predicted by inflation. Until 1998, $\Lambda$CDM's only major flaw was the absence of evidence for accelerated expansion. Not now. \section*{Three Dark Matter Problems} While stars are very interesting and pretty to look at -- and without them, astronomy wouldn't be astronomy and we wouldn't exist -- they represent a tiny fraction of the cosmic mass budget, only about 0.5\% of the critical density. As we have known for several decades, the bulk of the matter and energy in the Universe is dark. The present accounting defines three dark matter/energy problems; none is yet fully addressed. \subsection*{Dark Baryons} By a ten to one margin, the bulk of the baryons are dark and not in the form of stars. With the exception of clusters, where the ``dark'' baryons exist as hot, x-ray emitting intracluster gas, the nature of the dark baryons is not known. Clusters only account for around 10\% or so of the baryons in the Universe (Persic \& Salucci, 1992) and the (optically) dark baryons elsewhere, which account for 90\% or more of all the baryons, could take on a different form. The two most promising possibilities for the dark baryons are diffuse hot gas and ``dark stars'' (white dwarfs, neutron stars, black holes or objects of mass around or below the hydrogen-burning limit). I favor the former possibility for a number of reasons. First, that's where the dark baryons in clusters are. Second, the cluster baryon fraction argument can be turned around to infer $\Omega_{\rm gas}$ at the time clusters formed, redshifts $z\sim 0 -1$, \begin{equation} \Omega_{\rm gas}h^2 = f_{\rm gas} \Omega_Mh^2 = 0.023\,(\Omega_M/0.4)(h/0.65)^{1/2}\,. \end{equation} That is, at the time clusters formed, the mean gas density was essentially equal to the baryon density (unless $\Omega_Mh^{1/2}$ is very small), thereby accounting for the bulk of baryons in gaseous form. Third, numerical simulations suggest that most of the baryons should still be in gaseous form today (Rauch et al, 1997). There are two arguments for dark stars as the baryonic dark matter. First, the gaseous baryons not associated with clusters have not been detected. Second, the results of the microlensing surveys toward the LMC and SMC (Spiro, 1999) are consistent with about one-third of our halo being in the form of half-solar mass white dwarfs. I find neither argument compelling; gas outside clusters will be cooler ($T\sim 10^5 - 10^6$\,K) and difficult to detect, either in absorption or emission. There are equally attractive explanations for the Magellanic Cloud microlensing events (e.g., self lensing by the Magellanic Clouds, lensing by stars in the spheroid, or lensing due to disk material that, due to flaring and warping of the disk, falls along the line of sight to the LMC; see Sahu, 1994; Evans et al, 1998; Gates et al, 1998; Zaritsky \& Lin, 1997; Zhao, 1998). The white-dwarf interpretation for the halo has a host of troubles: Why haven't the white dwarfs been seen (Graff et al, 1998)? The star formation rate required to produce these white dwarfs -- close to $100\,{\rm yr^{-1}\,Mpc^{-3} }$ -- far exceeds that measured at any time in the past or present (see Madau, 1999). Where are the lower-main-sequence stars associated with this stellar population and the gas, expected to be 6 to 10 times that of the white dwarfs, that didn't form into stars (Fields et al, 1997)? Finally, there is evidence that the lenses for both SMC events are stars within the SMC (Alcock et al, 1998; EROS Collaboration, 1998a,b) and at least one of the LMC events is explained by an LMC lens. The SMC/LMC microlensing puzzle can be stated another way. The lenses have all the characteristics of ordinary, low-mass stars (e.g., mass and binary frequency). If this is so, they cannot be in the halo (they would have been seen); the puzzle is to figure out where they are located. \subsection*{Cold Dark Matter} The second dark-matter problem follows from the inequality $\Omega_M\simeq 0.4 \gg \Omega_B\simeq 0.05$: There is much more matter than there are baryons, and thus, nonbaryonic dark matter is the dominant form of matter. The evidence for this very profound conclusion has been mounting for almost two decades, and this past year, the Burles -- Tytler deuterium measurement anchored the baryon density and allowed the cleanest determination of the matter density, through the cluster baryon fraction. Particle physics provides an attractive solution to the nonbaryonic dark matter problem: relic elementary particles left over from the big bang (see Ellis, 1999). Long-lived or stable particles with very weak interactions can remain from the earliest moments of particle democracy in sufficient numbers to account for a significant fraction of critical density (very weak interactions are needed so that their annihilations cease before their numbers are too small). The three most promising candidates are a neutrino of mass 30\,eV or so (or $\sum_i\,m_{\nu_i} \sim 30\,$eV), an axion of mass $10^{-5\pm 1}\,$eV, and a neutralino of mass between $50\,$GeV and $500\,$GeV. All three are motivated by particle-physics theories that attempt to unify the forces and particles of Nature. The fact that such particles can also account for the nonbaryonic dark matter is either a big coincidence or a big hint. Further, the fact that these particles interact with each other and ordinary matter very weakly, provides a simple and natural explanation for dark matter being more diffusely distributed. At the moment, there is significant circumstantial evidence against neutrinos as the bulk of the dark matter. Because they behave as hot dark matter, structure forms from the top down, with superclusters fragmenting into clusters and galaxies (White, Frenk \& Davis, 1983), in stark contrast to the observational evidence that indicates structure formed from the bottom up. (Hot + cold dark matter is still an outside possibility, with $\Omega_\nu\sim 0.15$ or less; see Dodelson et al, 1996 and Gawiser \& Silk, 1998.) Second, the evidence for neutrino mass based upon the atmospheric (Totsuka, 1999) and solar-neutrino (Kirsten, 1999 and Bahcall, 1999) data suggests a neutrino mass pattern with the tau neutrino at $0.1\,$eV, the muon neutrino at $0.001\,$eV to $0.01$\,eV and the electron neutrino with an even smaller mass. In particular, the factor-of-two deficit of atmospheric muons neutrinos with its dependence upon zenith angle is very strong evidence for a neutrino mass difference squared between two of the neutrinos of around $10^{-2}$\,eV$^2$ (Fukuda et al, 1998). This sets a {\em lower} bound to neutrino mass of about $0.1\,$eV, implying neutrinos contribute at least as much mass as bright stars. WOW! Both the axion and neutralino behave as cold dark matter; the success of the cold dark matter model of structure formation makes them the leading particle dark-matter candidates. Because they behave as cold dark matter, they are expected to be the dark matter in our own halo; in fact, there is nothing that can keep them out (Gates \& Turner, 1994). As discussed above, 2/3 of the dark matter in our halo -- and probably all the halo dark matter -- cannot be explained by baryons in any form. The local density of halo material is estimated to be $10^{-24}\,{\rm g\, cm^{-3}}$, with an uncertainty of slightly less than a factor of 2 (Gates et al, 1995). This makes the halo of our galaxy an ideal place to look for cold dark matter particles! An experiment at Livermore National Laboratory with sufficient sensitivity to detect halo axions is currently taking data (van Bibber et al, 1998; Rosenberg and van Bibber, 1999) and experiments at several laboratories around the world are beginning to search for halo neutralinos with sufficient sensitivity to detect them (Sadoulet, 1999). The particle dark-matter hypothesis is compelling, but very bold, and most importantly, it is now being tested. Finally, while the axion and the neutralino are the most promising particle dark-matter candidates, neither one is a ``sure thing.'' Moreover, any sufficiently heavy particle relic (mass greater than a GeV or so) will behave as cold dark matter. A host of more exotic possibilities have been suggested, from solar-mass primordial black holes produced at the quark/hadron transition (see e.g., Jedamzik, 1998 and Jedamzik \& Niemeyer, 1998) that masquerade as MACHOs in our halo to supermassive (mass greater than $10^{10}\,$GeV) particles produced by nonthermal processes at the end of inflation (see e.g., Kolb, 1999). Lest we become overconfident, we should remember that Nature has many options for the particle dark matter. \subsection*{Dark Energy} I have often used the term exotic to refer to particle dark matter. That term will now have to be reserved for the dark energy that is causing the accelerated expansion of the Universe -- by any standard, it is more exotic and more poorly understood. Here is what we do know: it contributes about 60\% of the critical density; it has pressure more negative than about $-\rho /2$; and it does not clump (otherwise it would have contributed to estimates of the mass density). The simplest possibility is the energy associated with the virtual particles that populate the quantum vacuum; in this case $p=-\rho$ and the dark energy is absolutely spatially and temporally uniform. This ``simple'' interpretation has its difficulties. Einstein ``invented'' the cosmological constant to make a static model of the Universe and then he discarded it; we now know that the concept is not optional. The cosmological constant corresponds to the energy associated with the vacuum. However, there is no sensible calculation of that energy (see e.g., Zel'dovich, 1967; Bludman and Ruderman, 1977; and Weinberg, 1989), with estimates ranging from $10^{122}$ to $10^{55}$ times the critical density. Some particle physicists believe that when the problem is understood, the answer will be zero. Spurred in part by the possibility that cosmologists may have actually weighed the vacuum (!), particle theorists are taking a fresh look at the problem (see e.g., Harvey, 1998; Sundrum, 1997). Sundrum's proposal, that the energy of the vacuum is close to the present critical density because the graviton is a composite particle with size of order 1\,cm, is indicative of the profound consequences that a cosmological constant has for fundamental physics. Because of the theoretical problems mentioned above, as well as the checkered history of the cosmological constant, theorists have explored other possibilities for a smooth, component to the dark energy (see e.g., Turner \& White, 1997). Wilczek and I pointed out that even if the energy of the true vacuum is zero, as the Universe as cooled and went through a series of phase transitions, it could have become hung up in a metastable vacuum with nonzero vacuum energy (Turner \& Wilczek, 1982). In the context of string theory, where there are a very large number of energy-equivalent vacua, this becomes a more interesting possibility: perhaps the degeneracy of vacuum states is broken by very small effects, so small that we were not steered into the lowest energy vacuum during the earliest moments. Vilenkin (1984) has suggested a tangled network of very light cosmic strings (also see, Spergel \& Pen, 1997) produced at the electroweak phase transition; networks of other frustrated defects (e.g., walls) are also possible. In general, the bulk equation-of-state of frustrated defects is characterized by $w=-N/3$ where $N$ is the dimension of the defect ($N=1$ for strings, $=2$ for walls, etc.). The SN Ia data almost exclude strings, but still allow walls. An alternative that has received a lot of attention is the idea of a ``decaying cosmological constant'', a termed coined by the Soviet cosmologist Matvei Petrovich Bronstein in 1933 (Bronstein, 1933). (Bronstein was executed on Stalin's orders in 1938, presumably for reasons not directly related to the cosmological constant; see Kragh, 1996.) The term is, of course, an oxymoron; what people have in mind is making vacuum energy dynamical. The simplest realization is a dynamical, evolving scalar field. If it is spatially homogeneous, then its energy density and pressure are given by \begin{eqnarray} \rho & = & {1\over 2}{\dot\phi}^2 + V(\phi ) \nonumber \\ p & = & {1\over 2}{\dot\phi}^2 - V(\phi ) \end{eqnarray} and its equation of motion by (see e.g., Turner, 1983) \begin{equation} \ddot \phi + 3H\dot\phi + V^\prime (\phi ) = 0 \end{equation} The basic idea is that energy of the true vacuum is zero, but not all fields have evolved to their state of minimum energy. This is qualitatively different from that of a metastable vacuum, which is a local minimum of the potential and is classically stable. Here, the field is classically unstable and is rolling toward its lowest energy state. Two features of the ``rolling-scalar-field scenario'' are worth noting. First, the effective equation of state, $w=({1\over 2}\dot\phi^2 - V)/({1\over 2}\dot\phi^2 +V)$, can take on any value from 1 to $-1$. Second, $w$ can vary with time. These are key features that may allow it to be distinguished from the other possibilities. In fact, there is some hope that more SNe Ia will be able to do this and perhaps even permit the reconstruction of the scalar-field potential (Huterer \& Turner, 1998). At moment, taking the SNe Ia data together with large-scale structure data we can already constrain the average value of $w$ to be less than $-.60$ (95\% cl) (Perlmutter, White \& Turner, 1999). The rolling scalar field scenario (aka mini-inflation or quintessence) has received a lot of attention over the past decade (Freese et al, 1987; Ozer \& Taha, 1987; Ratra \& Peebles, 1988; Frieman et al, 1995; Coble et al, 1996; Turner \& White, 1997; Caldwell et al, 1998). It is an interesting idea, but not without its own difficulties. First, one must {\em assume} that the energy of the true vacuum state ($\phi$ at the minimum of its potential) is zero; i.e., it does not address the cosmological constant problem. Second, as Carroll (1998) has emphasized, the scalar field is very light and can mediate long-range forces. This places severe constraints on it. Finally, with the possible exception of one model (Frieman et al, 1995), none of the scalar-field models address how $\phi$ fits into the grander scheme of things and why it is so light ($m\sim 10^{-33}\,$eV). \section*{Concluding Remarks} The cosmological parameters $T_0$, $H_0$, $t_0$, and $\Omega_i$ are now all measured with reliable and respectable error estimates. This is a remarkable achievement for a science about which it used to be said, ``the errors are always in the exponents.'' Already, the consistency between $H_0$, $t_0$ and $\Omega_i$ provides an important crosscheck on the consistency of the hot big-bang framework, which it passes with flying colors. Over the next two decades we can expect to see even more progress. After the MAP and Planck missions and the completion of the Sloan Digital Sky Survey, it is not unreasonable to expect a 1\% determination of $H_0$ and $\Omega_0$, and a 5\% determination of $\Omega_M$ and $\Omega_B$, with significant crosschecks. In addition, we should have a much better understanding of the mysterious dark energy. The COBE FIRAS measurement of $T_0$ has set a standard that is unlikely to be exceeded for a long time. These precision measurements of the cosmological parameters will test the big-bang framework and general relativity as well as testing the paradigm that aspires to extend it, Inflation + Cold Dark Matter. These are exciting times in cosmology!
\section{Introduction} As a close, luminous active galactic nucleus (AGN) [14.4 Mpc (\cite{Tully88}) so that 1 arcsec=72 pc), NGC 1068 has been studied at nearly every available spatial resolution and wavelength for thirty years. While classified as a Seyfert 2 based on the presence of narrow emission lines and the absence of broad ones, polarization studies have detected broad wings on its narrow emission lines (\cite{Ant85}). These observations suggest that NGC 1068 harbors an obscured Seyfert 1 nucleus whose broad lines are scattered into our line of sight. Significant modeling of the spectrum and spectral energy distribution has been done by Pier \& Krolik (1993) and Granato et al. (1997) for the presence of a dusty torus which conceals the nucleus, and these models can reproduce the observed emission from the X-ray through the near-infrared. At scales of a few hundred parsecs around the nucleus, HST narrow band (\cite{Macch94}) and continuum (\cite{Ly91}) imaging show a non-uniform conical narrow-line region. These observations suggest that clumps of gas have been ionized by a partially collimated nuclear source. Mid-infrared measurements of this area have revealed the presence of warm gas (\cite{Cam93}). Previous near infrared one$-$dimensional speckle measurements of the nucleus (\cite{McCarthy82}; \cite{Chelli87}) showed extended emission on the 100 pc scale, and more recent two$-$dimensional speckle work finds extended emission closer to the nucleus (\cite{Wittkowski98}). Radio VLBI measurements (\cite{Green96}) of water maser emission demonstrate the presence of a thick torus, and high-resolution radio maps of the nucleus (\cite{Galli96b}) suggest that the obscuring material takes the form of a warped disk. The combination of these results have made the nucleus of NGC 1068 the prototypical obscured Seyfert 1. In addition, the nucleus resides in an SB host galaxy with a 3 kpc bar (\cite{Sco88}; \cite{Th89}) and active star formation in the inner 10 kpc (\cite{Teles88}). Authors have speculated on the relationship between this star formation and the activity of the nucleus (\cite{Normsco88}). Near-infrared measurements trace the distribution of hot dust and stars in NGC 1068, and thus characterize the physical condition of the material near the nucleus. The ability to do speckle imaging with the W. M. Keck Telescope allows a resolution of 0.$''$05, or 3.6 pc, at 2.2 $\mu$m which for the first time provides a direct comparison between near infrared and visual (HST) measurements. We use these speckle measurements, as well as complimentary 1.6 $\mu$m speckle imaging from the 200-inch Telescope and direct imaging at both wavelengths from the Keck Telescope to investigate the physical conditions in the near nuclear region of NGC 1068 on hitherto unavailable scales. \section{Observations} Speckle observations of NGC 1068 were made on four nights, 1994 October 18, 1994 December 19, and 1995 November 4$-$5, with the 200-inch Hale Telescope. A 64$\times$64 subsection of a 256$\times$256 Santa Barbara Research Center InSb array was used in order to allow continuous readout of speckle frames every 0.07 or 0.10 s. Speckle frames were collected in sets of $\sim$400 images on the AGN and on the two nearby unresolved SAO catalog stars. The sources were observed at both {\it H}$-$band~ ($\lambda_0$=1.65 $\mu$m, $\Delta\lambda$=0.32) and {\it K}$-$band~ ($\lambda_0$=2.2 $\mu$m, $\Delta\lambda$=0.4). Additional observations were made on three nights, 1995 December 18-20, at the W. M. Keck Observatory. Images from the full 256$\times$256 InSb array of the facility's Near Infrared Camera (NIRC; \cite{MS94}), were taken at a rate of one 0.118~s image every 1.5~s in sets of 100 images on the AGN and on the two nearby unresolved stars SAO 130046 and SAO 110692. A {\it K}$-$band~ ($\lambda_0$=2.21 $\mu$m, $\Delta\lambda$=0.43) filter was used for all the observations. The basic observing strategy was reported in Matthews et al. (1996). To reduce the noise contributed by phase discontinuities between the 36 segments of the Keck Telescope, the object and calibrators were observed at 12 different pupil orientations. A summary of observations is provided in Table \ref{tablejournal}. \begin{table}[h] \small \caption{Journal of Observations of NGC 1068} \medskip \begin{center} \begin{tabular}{cllcccccc} \hline Date &Cal 1 &Cal 2 &integ. &\#frames &\# K &\# H &pixel scale &seeing \\ &(SAO) &(SAO) &time (s) &per set &sets &sets &(arcsec &(arcsec) \\ & & & & & & &$\rm pixel^{-1}$) &\\ \hline \hline \multicolumn{9}{c}{200-Inch Telescope}\\ & & & & & & & & \\ 18 Oct 1994 & 130057 & 110709 &0.1 &400 &11 &10 &0.036 &0.7 \\ 19 Dec 1994 & 130057 & 110709 &0.1 &400 &20 &10 &0.036 &0.7 \\ 4 Nov 1995 & 130057 & 110692 &0.07 &480 &14 &20 &0.034 &0.7 \\ 5 Nov 1995 & 130057 & 110692 &0.1 &480 &16 &14 &0.034 &0.9 \\ & & & & & & & & \\ \multicolumn{9}{c}{Keck Telescope}\\ & & & & & & & & \\ 18 Dec 1995 & 130046 & 110692 &0.118 &100 &11 &$-$ &0.021 &0.5 \\ 19 Dec 1995 & 130046 & 110692 &0.118 &100 &12 &$-$ &0.021 &0.5 \\ 20 Dec 1995 & 130046 & 110692 &0.118 &100 &16 &$-$ &0.021 &0.4 \\ 16 Jan 1998 &$-$ &$-$ &10.0 &1 &3 &4 &0.021 &0.4 \\ 16 Jan 1998 &$-$ &$-$ &5.0 &1 &1 &$-$ &0.15 &0.4 \\ \hline \end{tabular} \end{center} \label{tablejournal} \end{table} At both telescopes, reimaging optics were used to convert the standard detector plate scales to scales appropriate for diffraction limited imaging. At the 200-inch Hale Telescope, detector pixel scales of 0.$''$034 and 0.$''$036 pixel${-1}$ were used and at the Keck Telescope, a detector pixel scale of 0.$''$021 pixel${-1}$ was used. The pixel scale at the 200-inch Telescope was chosen to oversample the {\it K}$-$band~ while still allowing diffraction limited imaging in $H-$band. The pixel scale at the Keck Telescope was chosen to sample optimally the aperture in the $K-$band. Several long exposure images of the nuclear region were also taken at the Keck Telescope under photometric conditions in both the {\it H}$-$band~ and $K-$band. In order to avoid saturating the detector, the speckle plate scale of 0.$''$021 pixel${-1}$ was used for these images. The total integration times were 30 seconds at {\it K}$-$band~ and 40 seconds at $H-$band, and the seeing for these images was 0.$''$45. A single 5 second {\it K}$-$band~ image of NGC 1068 was also obtained at the Keck Telescope with a pixel scale of 0.$''$15/pixel and seeing of 0.$''$45. Although the center is saturated, this image captured the distribution of galactic {\it K}$-$band~ flux at distances greater than 0.$''$75 from the nucleus. HST infrared standard stars of (\cite{Persson}) were observed both with and without the reimaging optics. \section{Data Analysis} In preliminary processing, each image was sky subtracted and flat fielded, and bad pixels were corrected by interpolation. The full NIRC frame of 256$\times$256 pixels was clipped around the centroid of each speckle frame to 128$\times$128 pixels (2.$''$6 on a side). Outside this smaller field, the signal-to-noise-ratio in each pixel is less than one-fifth, so clipping cut out pixels which would add only noise to the Fourier analysis. In the 200-inch data no clipping was necessary since a smaller field was used in collecting the images. In the second stage of analysis, the object's Fourier amplitudes and phases were recovered via classical speckle analysis (\cite{Labey70}) and bispectral analysis (\cite{Weig87}) respectively. For the data from the Keck Telescope, both processes were modified from the standard procedure to incorporate the field rotation that occurred during the observations (\cite{Matth96}). The observations were made over a spread of 103$^\circ$ in parallactic angle, although the change in parallactic angle over a single stack of 100 frames was always less than $\sim$2$^\circ$. Linear interpolation was used to find the rotated pixel values in each frame. The bispectral analysis was sufficiently computationally intensive as to require the use of the Caltech Concurrent Supercomputing Facility's nCUBE2 and Intel Delta computers. In order to go from the Fourier components calculated above to a final image, it is necessary to include a smoothing function, or effectively a telescope transfer function; a Gaussian of FWHM equal to the $\lambda / D$ where $\lambda$ is the wavelength of the observations and $D$ is the diameter of the telescope, was multiplied by the Fourier amplitudes. Then, the amplitudes and phases were combined directly in an inverse transform to produce the final image. \section{Results} \label{results} \subsection{$K-$band Results from the Keck Telescope} \label{kfit} Figure \ref{fig_Keckimage} presents the 0.$''$05 (or 3.6 pc) resolution {\it K}$-$band~ image produced with speckle imaging from using all of the nearly 4000 frames obtained in 1995 December. Each pixel is 0.$''$021 $\times$ 0.$''$021 and the field of view has been clipped to 0.$''$67 on a side. The nuclear emission is seen to be comprised of two components, an unresolved point source and an extended region symmetric about the nucleus with a major axis of $\sim$0.$''$3 (22 pc) and a minor axis of $\sim$0.$''$18 (13 pc). The calculated Fourier phases were consistent with zero, so the extended component is symmetric about the nucleus. \placefigure{fig_Keckimage} For comparison with previous results and with models of the nucleus, we estimated the fractions of the total nuclear flux density arising in the point source and extended components. These quantities, along with the orientation and size of the extended emission, were found by fitting a two component model to the two-dimensional average object visibility. The fit was performed in the spatial frequency, i.e. Fourier, domain rather than the image domain so that no tapering function had to be applied to the high spatial frequencies. This model was intended not to reproduce the exact distribution of flux but to provide a robust estimate of the magnitude of the contributions of the two components. The extended emission was modeled as a smoothly falling exponential of the form $I e^{-kd^n}$. The parameter $k$ measures the size of the extended emission, and the parameter $d$ measures the shape and orientation of the extended emission; $d$ was parameterized by an ellipticity and angle. The power $n$ determines how quickly the visibility falls with spatial frequency and hence the overall shape of the extended emission. The power $n$ was assumed, by ad-hoc phenomenological inspection, to be $3$ for the Keck Telescope {\it K}$-$band~ visibility. Since this model has only the explicit purpose of measuring the contributions from the two components, it is important for it to fit the overall shape of the visibility but unimportant whether it reproduced the details of the visibility at low spatial frequency. In particular, the lowest spatial frequencies were not used in the fit both because they are the most corrupted by global changes in the seeing in the time between when the object and calibrator were measured and because the extended flux from the galaxy causes the visibility to drop at the low frequencies. The point source was included as a constant visibility offset, i.e. the same visibility at all spatial frequencies. The fit was performed as a $\chi^2$ minimization of the two-dimensional object visibility, where each frequency was weighted by its statistical uncertainty as calculated from the ensemble of power spectra, subject to the constraints that the parameters be greater than or equal to zero. All of the points from 0$-$20 cycles arcsec$^{-1}$ were included in the fit. The salient results of this model are given in Table \ref{tablefit}. Figure \ref{fig_1068_visib} shows the radial (i.e. azimuthally averaged) profiles of the measured two-dimensional visibilities and of the fit are shown. The differences between the data and model, i.e. the residuals, are also shown in the figure. \placefigure{fig_1068_visib} The result of the fitting demonstrates that 49\% of the {\it K}$-$band~ flux density in the speckle image is contained in the unresolved core, i.e. in a diffraction limited beam, and 51\% is in the extended region. The uncertainty in this fit is taken to be the uncertainty in the normalization of the power spectra, or 8\%. \S~\ref{color} describes how the total flux in the speckle image was determined. \subsection{$H-$band Results from the Palomar 200-inch Telescope} \label{hfit} Of all the {\it H}$-$band~ data taken at the 200-inch Telescope, the observations from 19 December 1994 had the best signal to noise ratio at high spatial frequencies, so they were used for the analysis which follows. The observations from other nights are consistent with that of 19 December, but of lower quality. The final {\it H}$-$band~ image from the 200-inch Telescope, with half the resolution of the {\it K}$-$band~ image from the Keck Telescope (0.$''$1 or 7.2 pc), similarly consists of both a point source and extended emission. The same fitting procedure described in the previous section was used to fit the two-dimensional {\it H}$-$band~ visibility, but the power $n$ was taken to be $1$. Since the value of $n$ primarily affects the shape of the visibility at low spatial frequencies (where, as noted above, the measurements are sensitive to seeing variation and the galactic emission) we make no comparison between the {\it H}$-$band~ and {\it K}$-$band~ data on these scales. In the {\it H}$-$band~ data from the 200-inch Telescope, 63\% of the flux density from the speckle image is contained in the unresolved point source (i.e. within a diffraction limited beam) and 37\% is in the extended region. The uncertainty in this fit is taken to be the uncertainty in the normalization of the visibilities, or 8\%. A discussion of how the total flux in the {\it H}$-$band~ image is determined is given in \S~\ref{color}. \begin{table}[ht] \small \caption{Results of Model Fits to 2-D Object Visibility} \medskip \begin{center} \begin{tabular}{lccc} \hline measurement &eccentricity &angle &fraction unresolved \\ \hline \hline Keck {\it K}$-$band~ (at 200-inch Resolution) &0.81 &158.2 &0.64 \\ 200-Inch {\it H}$-$band~ &0.81 &159.4 &0.63 \\ Keck {\it K}$-$band~ (at full Resolution) &0.81 &158.6 &0.49 \\ \hline \end{tabular} \end{center} \label{tablefit} \end{table} \subsection{{\it H~}$-${\it K~} Color} The data from the Keck Telescope, at higher spatial resolution than that from the 200-inch Telescope, resolves more of the nuclear {\it K}$-$band~ flux into extended emission. In order to compute the {\it H~}$-${\it K~} colors of the unresolved point source and the extended emission, however, the {\it K}$-$band~ data from the Keck Telescope must be smoothed to the 200-inch Telescope resolution. Instead of smoothing the reconstructed image, the object visibility from the Keck Telescope data was fit out to a spatial frequency of 11.3 cycles arcsec${-1}$, i.e. the same resolution obtainable at the 200-inch Telescope, using the procedure described above in \S~\ref{kfit}. In this case the {\it K}$-$band~ shows the same distribution of flux density as the $H-$band, i.e. 64\% in an unresolved core and 36\% in an extended region. Since both the {\it K}$-$band~ and {\it H}$-$band~ data show the same fraction of their respective total fluxes in the point source, the H$-$K color of the point source is the same as that of the extended emission. The uncertainty in this color is 11\%, the combination of the uncertainties in each of the fits to the visibilities. The actual value of this color is computed below in the discussion section. \subsection{Upper Limit to the Size of the Point-like Nucleus} \label{ptsrc} The speckle data can be used to place an upper limit on the size of the nuclear point source. If the core were actually extended, it would have the effect of reducing the visibilities at high spatial frequencies, but instead we find the visibilities flatten out at high spatial frequencies. The fit to the Keck Telescope data, shown in a radial profile plot in Figure \ref{fig_1068_visib}, leaves residuals of less than 10\% at frequencies above 19 cycles~arcsec${-1}$. So, the presence of another undetected component is therefore constrained by the uncertainty in the residuals at the highest frequencies. Without making an a priori assumption as to the shape of the extension, should it be present, the upper limit to its size can be taken as the highest spatial frequency at which high S/N information was obtained in the data, in this case 19.7 cycles~arcsec${-1}$ or 0.$''$051. A more stringent upper limit to the point source size can be placed under the assumption that the true nucleus has a Gaussian shape. The width of the largest Gaussian which could be hidden in the {\it K}$-$band~ visibility data (i.e. which will not differ from the data by more than 3$\sigma$ at the highest spatial frequency) is 0.$''$02 or 1.4 pc. \subsection{Photometry} \label{keckphot} Because of limitations deriving from the small field size, low S/N, and image wander in speckle images, it is difficult to make photometric measurements from speckle data. Therefore, we measured the total flux density in a beam the size of the speckle frames from 0.$''$45 resolution long exposure images made at the Keck Telescope in both the {\it H~} and $K-$bands. In a beam radius of 1.$''$25, the {\it K}$-$band~ magnitude is 7.5 and the {\it H}$-$band~ magnitude is 9.26, resulting in an {\it H}$-${\it K} color of 1.76~mag. Aperture photometry from these images (in both magnitudes and Janskys) at a variety of other beam sizes is reported in Table \ref{table_photometry} and shown in Figure \ref{fig_1068cog}. \placefigure{fig_1068cog} The contribution of the galaxy to the photometry at small beam sizes was estimated by fitting the galaxy surface brightness at radii between 1.$''$8 and 27$''$ with a deVaucouleurs function. Extrapolating the fit to a beam of radius 1.$''$25, approximately the same size as the speckle field of view, implies that only 15\% percent of the flux arises in the galaxy. Furthermore, the shape of the surface brightness profile within 1.$''$25 is consistent with being the sum of the deVaucouleurs profile and a point source. \begin{table}[ht] \small \caption{Aperture Photometry and Colors} \medskip \begin{center} \begin{tabular}{cccccc} \hline beam diameter &K &K &H &H &H$-$K \\ ($''$) &(mag) &(mJy) &(mag) &(mJy) &(mag) \\ \hline \hline 0.4 &8.75 &204 &11.06 &39 &2.31 \\ 0.5 &8.50 &257 &10.78 &51 &2.28 \\ 0.8 &7.98 &415 &10.13 &92 &2.14 \\ 1.0 &7.87 &459 &9.95 &109 &2.08 \\ 1.5 &7.68 &547 &9.63 &146 &1.95 \\ 2.0 &7.58 &600 &9.42 &178 &1.84 \\ 2.5 &7.50 &646 &9.26 &206 &1.76 \\ 2.9 &7.45 &676 &9.14 &230 &1.69 \\ 3.5 &7.39 &715 &9.00 &261 &1.61 \\ 3.75 &7.36 &735 &8.94 &276 &1.58 \\ \hline \end{tabular} \end{center} \label{table_photometry} Uncertainties are 3\% in both the K and H magnitudes. Within a diameter of 1$''$, the uncertainties are systematically larger due to the effects of the seeing disk. \end{table} \subsection{Comparison to Previous Measurements} The size of the near-infrared core of NGC 1068 has been the subject of investigation by many previous researchers. One-dimensional speckle by McCarthy et al. (1982) at the 3.8 m Mayall Telescope placed an upper limit of 0.$''$2 on the size of the unresolved core, and found extended emission that was 25\% of the total 2.2 $\mu$m flux in the beam. As shown in Figure \ref{fig_1068_visib}, the {\it K}$-$band~ visibility measured in this work is consistent with 1.0 out to 1 cycle~arcsec${-1}$, but in the McCarthy et al. data, the visibility decreased to 0.8 by 0.5 cycles~arcsec${-1}$. This is probably a consequence of their large, 5$''$ by 10$''$, beam, which captured more light from the galactic stars, which have large spatial extent, than our smaller 2.$''$6 square beam. Similarly, 1-D speckle at 3.6 $\mu$m by Chelli et al. (1987) at the 3.6 m ESO Telescope found an unresolved core, large scale (100 pc) emission, and a third component of extended yet compact emission 0.$''$2 around the nucleus. The visibility obtained by Chelli et al. agrees very well with what we report. Their data also suggested that, of the two position angles they measured, the compact extended emission was larger along an angle of 135$^\circ$ than 45$^\circ$ which is consistent with the emission we measure shown in Figure \ref{fig_Keckimage}. In more recent imaging with an aperture mask, Thatte et al. (1997) find that 94\% of the {\it K}$-$band~ flux in a 1$''$ diameter aperture comes from a point source smaller than 0.$''$03. They report the flux from this source as 190 mJy. Both of these measurements disagree with what is reported above in \S~\ref{kfit} and Table~\ref{table_photometry}, respectively. Recent two dimensional speckle imaging by Wittkowski et al. (1998) at the 6 m Special Astrophysical Observatory Telescope finds extended emission which is 20\% of the total {\it K}$-$band~ flux and in addition places a limit on the core size of 0.$''$03. These estimates came from assuming a uniform disk model for the reconstructed emission, but they admit that an alternative explanation for their data would be an unresolved central object and extended emission. Fitting their data with this model, as described in \S\S~\ref{kfit} and \ref{hfit} above, would increase the fraction of the flux density attributable to the extended emission. While of insufficient sensitivity to show the extended emission reported in this work, their results are consistent with what is described here. \section{Discussion} Discussed in this section are three components of the emission from the central 1.$''$25 radius of the nucleus: (1) the central point source, (2) the newly imaged extended nuclear emission reported in \S~\ref{results} (and accounting for approximately 50\% of the flux at 2$\mu$m previously attributed to the point source), and (3) the stars of the underlying host galaxy. \subsection{{\it H~}$-${\it K~} Color of the Nucleus} \label{color} While it is possible to tell from the speckle measurements alone that the color of the point source and the color of the newly imaged extended emission are the same, it is not possible to determine the color itself. This is because speckle, as an interferometric technique, resolves out (i.e. is not sensitive to) smooth large-scale extended emission which fills the field of view. Thus, all of the flux measured in a beam the size of the speckle frames, as reported in Table \ref{table_photometry}, cannot be automatically attributed to the features observed in the speckle image. However, the color of the emission in the speckle image can be deduced by subtracting the contribution from large scale galactic emission from the total flux measured in the speckle beam. We assume that the only such contribution comes from the distribution of stars in the host galaxy. From the deVaucouleurs model fitting described in \S~\ref{keckphot} the galactic stellar contribution to the 1.$''$25 radius speckle beam was determined to be 97 mJy at $K-$band, or 15\% of the total {\it K}$-$band~ emission. Since no large (38$''$ square) image of NGC 1068 at {\it H}$-$band~ such as the one taken at {\it K}$-$band~ was available, it was assumed that the $H-K$ color of the galactic stellar population was 0.3 mag in the nuclear region, i.e. the same color as measured in aperture photometry off the nucleus (\cite{Th89}). Combining this color with the {\it K}$-$band~ measurement, it was calculated that 118 mJy, or 57\%, of the {\it H}$-$band~ emission in the 1.$''$25 radius beam is due to stars. All flux in excess of the galactic stellar contribution, was assumed to come from the nucleus plus extended emission reported in \S~\ref{results}, and this flux then has an $H-K$ color of 2.5 mag. If there is another population of stars in excess of the assumed galactic contribution, this estimate of the $H-K$ color would be low. If there is substantial reddening of the stars in the nucleus compared with far from the nucleus, this estimate would be high. However, $H-K$~=~2.1~mag can safely be considered a lower limit to the color based on the aperture photometry reported in Table \ref{table_photometry}. The statistical uncertainty in this color is a combination of the uncertainties in the photometric calibration, the aperture photometry, and the fit to the photometry, for a total of 9\%. Combined with the uncertainty in the visibility fitting from \S\S~\ref{kfit} and \ref{hfit}, our best estimate of the color of the extended emission is 2.5 $\pm$ 0.2 mag. \subsection{Possible Mechanisms for Extended Emission} \label{extem} It is of interest to consider the origin of the observed extended nuclear emission. There are several possibilities: it could be emission from stars, from nuclear light reflected off dust or electrons, or emission from hot dust, either in equilibrium or from single photon heating. The $H-K$ color of the extended emission, 2.5 mag, is significantly redder than that of any stellar population. Thus, if the emission is from stars, it must be highly extincted. A color excess of 2.2 mag, obtained from assuming the stars have the same intrinsic color as the galactic stars far from the nucleus, (i.e., $H-K$=0.3 mag) necessitates $A_v$=34 mag. This is similar to the extinction found in the center parsec of our own Galaxy and produced in models of thermal emission from dust in a thick torus in NGC 1068 (\cite{Efstat95}; \cite{Young95}). It is, however, much higher than the extinction of $A_v$=0$-$2 mags suggested in the data of Thatte et al. (1997) for the central stellar cluster. It is known, however, that the extinction is quite patchy near the nucleus (\cite{Blietz94}). The extended emission reported in this work is smooth over a length of 20 pc, and the substantial reddening required for the extended emission to be from stars seems unlikely to be similarly smooth over this scale. A bigger problem with this hypothesis comes from the luminosity of the extended emission. {\it K}$-$band~ imaging spectroscopy of the nuclear region (\cite{Thatte97}) has revealed the presence of a dense stellar cluster, which, based on the equivalent width of the CO band-head, accounts for 7\% of the total nuclear luminosity. However, 50\% of the nuclear flux is resolved by speckle. The {\it K}$-$band~ luminosity of the extended region, if it is emitted isotropically, is 4.7$\times$10$^{8}$~L$_\odot$, a factor of ten larger than the stellar luminosity given in Thatte et al. Finally, in this scenario the fact that the point source and the extended emission have the same color would be purely accidental unless the point source is also composed of stars. If the extended near-infrared structure is comprised of stars, it is also unlikely to be a continuation of a larger scale structure in the host galaxy. NGC 1068 does have a well known large scale ($\approx$ 1 kpc) stellar bar (\cite{Sco88}; \cite{Th89}), but it is also oriented at a position angle of approximately 45$^\circ$. Light from the point source reflected off dust or electrons in the narrow-line region would, on the other hand, provide a natural explanation for the similar colors of the point source and extended emission. The nucleus is highly (4$-$5\%) polarized at 2.2$\mu$m in a 4$''$ beam suggesting that there is extensive scattering of the nuclear radiation (\cite{Lebofsky78}). There are three possible sources of scattering: the warm electron gas which scatters the broad line emission, another population of electrons, or dust. The warm electrons modeled by Miller, Goodrich, \& Matthews (1991) are located at least 30 pc from the nucleus, i.e. outside the extended emission reported here. Therefore scattering from these electrons is unlikely be the source of the extended region. Scattering from other electrons which are within a few parsecs of the point source would also tend to reflect the broad line region, so it is reasonable to conclude that there is no second population of electrons beyond that found by Miller et al. The albedo, and wavelength dependence thereof, of an ensemble of dust grains varies widely with their size distribution (e.g. \cite{Lehtinen96}). For the small grains that are to be expected in high UV-radiation field locations such as that around NGC 1068, observations and theory (\cite{Draine84}) predict that the albedo at 2.2 $\mu$m is approximately 20\% lower than at 1.6 $\mu$m. Therefore, if the observed extended emission were reflected light from the point source of NGC 1068, it would be significantly bluer than the point source itself, whereas we observe the same color in the two sources. On the optimistic assumption that the albedo at 2.2 $\mu$m has a value of 0.8 and that the dust scatters isotropically, the central source would have a true 2.2 $\mu$m luminosity which is 15 times greater than observed. Observations by Glass (1997) have shown that the nucleus of NGC 1068 became steadily brighter at {\it K}$-$band~ over a twenty year span from 1976 to 1994 before leveling off. He did not detect a concomitant rise in the {\it H}$-$band~ emission, but this is understandable given the galactic stellar contamination of his 12$''$ beam. If the nuclear emission comes from dust on the inner edge of the torus which is heated to just below its sublimation point, an increase in luminosity will push the inner edge of the torus further away from the nucleus but not change the intrinsic color of the emission. However, the time constant for destroying grains may be long enough, on the order of years (\cite{Voit92}) so that an increase in luminosity produces temporarily higher temperatures and therefore bluer colors. The light crossing time of the extended emission reported in \S~\ref{results} is approximately 10~yr, so if it is reflected point source light, it should show a 10~yr lag in color compared to the point source. We do not know what the color of the point source was 10 years ago, but for it to have been just red enough to offset the tendency of dust reflection to make the emission bluer would be quite a conspiracy. The final possibility is that the extended emission comes from hot dust. The color temperature implied by an H$-$K color of 2.5 mags is 800 K. Taking the central luminosity of the AGN as $1.5 \times 10^{11} L_{\sun}$, we can calculate that the dust grains which would be heated to this temperature in equilibrium would lie at most 1 pc or 0.$''$01 (for silicate grains) from the point source. The extended emission reported in this paper is a factor of 10 larger. It has been suggested by other authors (\cite{Baldwin87}; \cite{Braatz93}; \cite{Bock98}) that extended 10 $\mu$m emission on 200 pc scales may be caused by heating of grains by the central source if the luminosity is beamed along the direction of the radio jets rather than emitted isotropically. The good spatial correspondence between the mid-infrared and radio jet (\cite{Cam93}; \cite{Bock98}) also lends credence to this idea. The component of the radio jet thought to lie at the infrared point source (\cite{Galli96a}), S1, sits in a region of extended radio continuum emission which lies at a position angle of 175$^\circ$. A beaming factor of 200 which would be sufficient to explain the extended mid-infrared emission would also be sufficient to produce 800~K grains at 10 pc from the point source. By contrast, Efstathiou et al. (1995) derive a beaming factor of $\gtrsim$~6 based on their fit to the near-infrared spectrum of the nucleus, and this would be insufficient to heat grains to 800~K, 10 pc from the point source. It is possible that the extended near-infrared emission is from hot dust which is heated not externally (i.e. by the central AGN), but internally, for example, by an interaction with the jet. The jets are observed to drive the motion of the emission line gas (\cite{Axon97}), so it is reasonable to assume that they are dumping energy into the circumnuclear gas. More complete models would have to be made to examine this hypothesis. A natural explanation for the high color temperature which is reconcilable with an isotropically emitting central source would be single photon heating of small grains (\cite{Sellgren84}). The rate of UV photons necessary to produce the total {\it K}$-$band~ luminosity of the extended nuclear emission can be calculated if one knows the mass of dust present in the region. If the hydrogen density is 10$^5$ cm$^{-3}$ (\cite{Tacconi94}), the dust to gas mass ratio is 10$^{-2}$, and the grains radiate with unity efficiency in the infrared, the rate of photons needed per grain is $\sim$10$^{-5}$~s$^{-1}$. This rate is well below that expected from the intrinsic UV/X-ray spectrum of NGC 1068 (\cite{Pier94}). However, polycyclic aromatic hydrocarbons (PAHs) are thought to be destroyed by intense X-ray/UV radiation fields (\cite{Voit92}) and the 3.3 $\mu$m emission feature associated with PAHs has not been unambiguously detected at the nucleus (\cite{BlandHaw97}). However, if the dust along the edges of the extended torus, as in Figure \ref{fig_1068model}, were illuminated by UV photons which were reflected off of the high lying electron cloud and yet protected from the nucleus by the bulk of the torus, the X-ray flux they intercept would be substantially reduced. Miller et al. (1991) predict that if the optical depth to electron scattering is about 0.1 and there is a dusty disk of dimension 10$^{20}$ cm surrounding the central region, then about 10\% of the central UV luminosity would be back scattered onto the disk. If the disk is not uniform, as is likely considering the lumpy high resolution radio maps, some regions would have high enough column density to stop the grain destroying X-rays, yet see a reflected UV flux sufficient to create transient grain heating. \subsection{Comparison to Models and Line Emission} In the model of infrared emission from NGC 1068 (e.g. \cite{pnk93}; \cite{Efstat95}; \cite{Gran97}) commonly found in the literature, the central source is surrounded by an optically thick torus. The torus is heated by the ultraviolet and X-ray photons from the accretion disk plus black hole system to which it is optically thick, but ionizing photons escape along the axis of the torus. The inner radius of the torus, $\sim$0.2 pc, is set by the sublimation point of the dust, and its outer radius, $\leq$ 40 pc (\cite{Gran97}), is set by models of its infrared emission. The line of sight to this torus is nearly edge on, passing through 70-1000 magnitudes of visual extinction, depending on the model, and therefore does not permit a direct view of the central source. The 1$-$2 $\mu$m emission observed is produced by the thermal radiation from hot dust on the inner edge of the torus, which, because it is on the edge, escapes through a region of moderate extinction. The geometry of the torus is constrained by the conical shape of the narrow-line region to have an opening angle of approximately 45$^\circ$. A cartoon of this model is shown in Figure \ref{fig_1068model}. The 2 pc upper limit placed on the size of the point source in \S~\ref{ptsrc} is consistent with this model, but the fact that the extended emission we observe is much larger than the size of the inner edge of the torus means that we must add to this picture. The emission we observe at 10 pc from the nucleus could come from a larger scale dusty structure, perhaps an extension of the torus, if the emission can be produced by one of the mechanisms outlined in \S~\ref{extem}. \placefigure{fig_1068model} Light scattering off of dust is observed in the narrow-line region much further from the central source (\cite{MGM91}) than predicted by models of the spectral energy distribution (\cite{Efstat95}). The narrow-line emission comes from clouds excited by the central source (\cite{Macch94}), and it extends to hundreds of parsecs, at a position angle of approximately 45$^\circ$. The placement of the 2 $\mu$m point source and extended region, shown superposed on the HST narrow-line image in Figure \ref{fig_hst}, shows that the extended 2 $\mu$m emission lies alongside the bright emission knots in the visual ionization cone. Of course, the registration of the infrared and optical images is not known to exquisite precision. The best estimates from Thatte et al. (1997) have a 0.$''$1 uncertainty in the registration and this uncertainty nearly encompasses the size of the extended near-infrared emission. \placefigure{fig_hst} \section{Conclusions} Two components of the nuclear 2.2 and 1.6 $\mu$m emission of NGC 1068, in addition to its galactic stellar population, have been detected with speckle imaging on the Keck and 200-Inch Telescopes. The observations reveal an extended region of emission that accounts for nearly 50\% of the nuclear flux at $K-$band. This region extends 10 pc along its major axis and 6 pc along its minor axis on either side of an unresolved point source nucleus which is at most, 0.$''$02 or 1.4 pc in size. Both the point source and the newly imaged extended emission are very red, with identical $H-K$ colors corresponding to a color temperature of 800~K. While the point source is of a size to be consistent with grains in thermal equilibrium with the nuclear source, the extended emission is not. The current data do not allow us to unambiguously determine the origin of the extended emission, but it is most likely either scattered nuclear radiation from an extended dusty disk or emission from thermally fluctuating small grains heated by reflected nuclear UV photons. \acknowledgements We thank Andrea Ghez for her help with the observations and data analysis, Tom Soifer for many helpful conversations on models of NGC 1068, and the telescope operators at Palomar and Keck Observatories for their efforts during time consuming speckle observing. Infrared astronomy at Caltech is supported by the NSF.
\section{Introduction} During the past few years, neutrino oscillations have been used to explore exotic properties of neutrinos such as possible violations of Lorentz invariance (VLI)\cite{cg1,cg2,ghklp} and/or violations of the equivalence principle (VEP)\cite{fly,hlp,imy,hl,g1,g2}. Since neutrinolesss double beta decay has served as a window into neutrino masses for the last two decades \cite{hmpr,dkt,m}, it is natural to enquire if this rare decay can tell us anything about VLI and VEP processes. We begin with the observation that the properties of neutrinos enter into the neutrino exchange diagram for the neutrinoless double beta decay amplitude (without right handed currents) in the form of the factor \cite{k} \begin{equation} A^{\nu}=(1-\gamma_5)P(1-\gamma_5), \end{equation} where P is a linear combination of the propagators for each of the Majorana neutrino fields that constitute the $\nu_e$ field, \begin{equation} P=U_{ea}^2 P_a. \end{equation} The $U_{ea}$ are elements of the unitary matrix connecting mass and weak eigenstate neutrinos. In the absence of external fields, and in the absence of Lorentz invariance violation, the Majorana fields have definite masses, $m_a$, and all neutrinos have the same limiting velocity, $c$. In that case the $ P_a$ are, of course, given by \begin{equation} P_a=(\gamma^0E-\gamma^k p_k c +m_a c^2)^{-1}. \end{equation} \section{Modifications Due to Violation of Lorentz Invariance.} In the Lorentz invariance violating scheme of Coleman and Glashow, the limiting velocity of each neutrino may be distinct, so that $c\rightarrow c_a$. Our conclusions are unaltered by the introduction of distinct mass and velocity bases, and for the sake of clarity such a complication will be ignored. To first order in $m_a$, the VLI modified $A^{\nu}$ is then given by \begin{equation} A^{\nu}_{VLI} \simeq - \frac{2 U^2_{ea}(1 - \gamma_5) m_a c_a^2}{E^2- (p c_a)^2} \end{equation} where the chirality factors have been used to eliminate contributions from factors of $E\gamma^0$ or $\gamma^k p_k$ in the numerator. For neutrinoless double beta decay in nuclei we usually make the approximation of ignoring nuclear recoil, so the energy of the exchanged neutrino is set equal to zero, that is $E=0$. With this standard approximation, \begin{equation} A^{\nu}_{VLI} \simeq 2 U_{ea}^2 (1 - \gamma_5) m_a/p^2. \end{equation} Since this expession is independent of limiting velocities, we conclude that VLI cannot enter into neutrinoless double beta decay in any significant way. \section{Modifications Due to Violation of the Equivalence Principle.} Following the formalism of Ref.\cite{imy}, the Dirac equation governing neutrino $a$ is modified by the presence of an external gravitational field which couples to neutrinos with strength $f_a$ relative to the usual universal Newtonian coupling. For the sake of clarity, we take the mass and gravitational coupling bases to be the same. In the presence of a constant Newtonian potential, $\Phi$, the neutrino propagator becomes \begin{equation} P_a = [(1+ f_a \Phi)E\gamma^0 -(1 - f_a\Phi)\gamma^k p_k + m_a]^{-1} \end{equation} where we have set the common limiting vacuum (i.e. $\Phi = 0$) velocity equal to 1. To first order in $m_a$ we then have, making use of the chirality factors as before, \begin{equation} A^{\nu}_{VEP} \simeq - \frac{2 U_{ea}^2 (1 - \gamma_5) m_a}{(1 + f_a \Phi)^2 E^2 - (1 - f_a\Phi)^2 p^2}. \end{equation} Making the zero recoil approximation as above and retaining only to first order in $\Phi$ for consistency, we then have \begin{equation} A^{\nu}_{VEP} \simeq 2 U_{ea}^2 (1 - \gamma_5)m_a(1 + 2f_a\Phi)/p^2. \end{equation} The expression above includes only the modification to the neutrino propagator due to the presence of $\Phi$. To this we must also add modifications to the W- boson, quark and electron lines due to $\Phi$. Assuming that only neutrinos have anomalous gravitational couplings, restoration of gravitational gauge invariance in the limit that all $f_a$ are equal guarantees that the final $\Phi$-dependent contribution depends only upon $\Delta f_a = f_a - f_0$, where $f_0 = 1$ in Einsteinian gravity. The $\Phi$-dependent contribution to the total neutrinoless double beta decay rate will then be proportional to $U_{ea}^2 m_a \Delta f_a$. Thus, the VEP effect is proportional to both $m_a$ and $\Delta f_a$ and is therefore extremely small. \section{Conclusions} We have examined the modifications to the usual neutrino exchange diagram for the neutrinoless double beta decay amplitude arising from violations of Lorentz invariance and/or the equivalence principle in the neutrino sector. We find that the VLI parameters disappear from the decay amplitude in the usual zero recoil approximation and that the VEP parameters enter the amplitude only in combination with with neutrino mass factors. We therefore conclude that neutrinolesss double beta decay cannot provide a significant window into VLI and VEP neutrino processes. This result appears to contradict the conclusions of a recent paper \cite{kps}. \acknowledgments{A.H. would like to thank the particle theory group at The University of Melbourne and the Korean Institute for Advanced Study for their hospitality during a portion of this work. He also thanks C.N. Leung for discussions during the early stages of this work. This work was supported in part by the US Department of Energy grant DE-FG02-84ER40163. R.R.V. was supported by the Australian Research Council.}
\section{\protect\large LQP Principles and some Consequences} If one thinks about the fundamental physical principles of this century which have stood their grounds in the transition from classical into quantum physics, relativistic causality as well as the closely related locality of quantum operators (together with the localization of quantum states) will certainly be the most prominent one. This principle entered physics through Einstein's 1905 special relativity, which in turn resulted from bringing the Galilei relativity principle of classical mechanics into tune with Maxwell's theory of electromagnetism. Therefore it incorporated Faraday's ``action at a neighborhood'' principle which revolutionized 19$^{th}$ century physics. The two different aspects of Einstein's special relativity, namely Poincar% \'{e} covariance and the locally causal propagation of waves (in Minkowski space) were kept together in the classical setting. In the adaptation of relativity to \textbf{LQP} (local quantum physics\footnote{% We use this terminology, whenever we want to emphasize that we relate the principles of QFT not with necessarily with the standard text-book formalism that is based on quantization through Lagrangian formalism.}) on the other hand \cite{Haag}, it is appropriate to keep them at least initially apart in the form of positive energy representations of the Poincar\'{e} group (leading to Wigner's concept of particles) and Einstein causality of local observables (leading to observable local fields and local generalized ``charges''). Here a synthesis is also possible, but it happens on a deeper level than in the classical setting and results in LQP as a new physical realm which is conceptually very different from both classical field theory and general QT (quantum theory). The elaboration of some of these differences, in particular as they may be relevant with respect to the measurement process, constitutes one of the aims of these notes. For material which already entered textbooks or review articles, we have preferred to quote the latter. A more detailed account of the consequences of causality in a much broader context can be found in \cite{Cau}\cite {schroer}. As a result of this added locality, LQP acquires a different framework than the kind of general quantum theory setting \cite{Landsman} in which the basics of quantum theory and measurement (including those ideas, which in the fashionable language of the day, are referred to as ``quantum computation'') are presented . Those concepts, which originate from the quantum adaptation of Einstein causality, lead in the presence of interactions to real particle creation (which artificially could be incorporated into a multichannel version quantum theory of particles) and, what has more importance within our presentation, to \textit{virtual particle structure} (related to the phenomenon of vacuum polarization) which has no counterpart in global general quantum theory as quantum mechanics and cannot be incorporated into it at all. The latter remark preempts already the greater significance of superselected charges and their fusion, as opposed to particles and their quantum mechanical bound states. Thus the hierarchy of particles in QM is replaced by the hierarchy of charges and consequently we obtain ``nuclear democracy'' between particles. This is closely related to an almost anthropological principle which LQP realizes in a perfect way in laboratory particle physics: whenever energy-momentum and (generalized) charge conservation allow for particle creation channels to be opened, nature will maximally use this possibility. To be sure there are theoretical models of LQP (integrable/factorizing models in d=1+1 spacetime dimensions) which do not follow this dictum, but even in those cases at least its theoretical ``virtual'' version is realized: a vector state created by the application of an interacting field to the vacuum which has a one-particle component, is inexorably accompanied by a ``polarization cloud'' of particles/antiparticles (the hallmark of LQP). As already emphasized the \textit{only exception} \textit{are free bosonic/fermionic fields} and in a somewhat pointed (against history), but nevertheless correct manner, one may say that this very exception is the reason why QM as a nonrelativistic limit of LQP has a physical reality at all. More general braid group statistics, as it can occur together with exotic spin in low dimensional QFT, requires these polarization clouds already in the ``freest'' realization of anyons/plektons and they are not fading away in the nonrelativistic limit because they are needed to uphold braid group statistics in that limit. This is the reason why the attempts of Leinaas- Myrheim, Wilscek and many others, which draw on the analogy with the Aharanov-Bohm quantum mechanics may catch some aspects of plektons but miss the spin-statistics connection which is their most important property (i.e. their LQP characterization). This aspect of virtuality, which at first sight seems to complicate life since it activates the coupling between infinitely many degrees of freedom/channels, is counterbalanced by some very desirable and useful features: whereas general quantum theory needs an outside interpretative support, LQP carries this already within itself. It was emphasized already at the end of the 50$^{ies}$ (notably by Rudolf Haag \cite{Haag}), that e.g. for a particle interpretation one does not need to resolve the distinction between the various local observables which are localized in the same space-time region (laboratory extension and time duration of measurement), the knowledge of the space-time affiliation of a generic observable from a region $\mathcal{O}$ is enough. The experimenter does not know more than the geometric spacetime placement of his counters and their sensitivity; the latter he usually has to determine by monitoring experiments. The basic nature of locality in interpreting the particle aspect of a theory is underlined by the fact that despite intense efforts nobody has succeeded to construct a viable \textit{nonlocal} theory. Here ``viable'' is meant in the sense of conceptual completeness, namely that a theory is required to \textit{contain its own physical interpretation} i.e. that one does not have to invent or impose formulas from outside this theory. Although physical reality may unfold itself like an onion or an infinite Russian ``matrushka'' with infinitely many layers of ever more general physical principles towards higher energies (smaller distances), it should still continue to be possible to have a mathematically consistent theory in each layer which is faithful to the principles valid in that layer. This has been fully achieved for quantum mechanics, but this goal was not yet reached in QFT. As a result of lack of nontrivial d=1+3 models or structural arguments which could demonstrate that the physical locality and spectral requirements allow for nontrivial solutions, the theory is still far from conceptual maturity, despite its impressive perturbation successes in QED, the Standard Model and in the area of Statistical Mechanics/Condensed Matter physics. Causality and locality are in a profound way related to the foundations of quantum theory in the spirit of von Neumann, which brings me a little closer to the topic of this symposium. In von Neumann's formulation, observables are represented by selfadjoint operators and measurements are compatible if the operators commute. The totality of all measurements which are relatively compatible with a given set (i.e. noncommutativity within each set is allowed) generate a subalgebra: the commutant $L^{\prime }$ of the given set of operators $L$. In particular in LQP, a conceptual framework which was not yet available to von Neumann, one is dealing with an isotonic ``net'' of subalgebras (in most physically interesting cases von Neumann factors, i.e. weakly closed operator algebras with a trivial center) $\mathcal{O}% \rightarrow \mathcal{A}(\mathcal{O}).$ Therefore unlike quantum mechanics, the spatial localization and the time duration of observables becomes an integral part of the formalism. \textit{Causality gives an a-priori information about the size} \textit{of spacetime }$\mathcal{O}$\textit{\ -affiliated operator (von Neumann) algebras:} \begin{equation} \mathcal{A}(\mathcal{O})^{\prime }\supset \mathcal{A}(\mathcal{O}^{\prime }) \label{Einstein} \end{equation} in words: the commutant $\mathcal{A}(\mathcal{O})^{\prime }$ of the totality of local observables $\mathcal{A}(\mathcal{O})$ localized in the spacetime region $\mathcal{O}$ contains the observables localized in its spacelike complement (disjoint) $\mathcal{O}^{\prime }.$ In fact in most of the cases the equality sign will hold in which case one calls this strengthened (maximal) form of causality ``Haag duality'' \cite{Haag}: \begin{equation} \mathcal{A}(\mathcal{O})^{\prime }=\mathcal{A}(\mathcal{O}^{\prime }) \end{equation} In words, the spacelike localized measurements are not only commensurable with the given observables in $\mathcal{O}$, but every measurement which is commensurable with all observables in $\mathcal{O},$ is necessarily localized in the causal complement $\mathcal{O}^{\prime }.$ Here we extended for algebraic convenience von Neumann's notion of observables to the whole complex von Neumann algebra generated by hermitian operators localized in $% \mathcal{O}.$ If one starts the theory from a net indexed by compact regions $\mathcal{O}$ as double cones, then algebras associated with unbounded regions $\mathcal{O}^{\prime }$ are defined as the von Neumann algebra generated by all $\mathcal{A}(\mathcal{O}_{1})$ if $\mathcal{O}_{1}$ ranges over all net indices $\mathcal{O}_{1}\subset \mathcal{O}^{\prime }.$ Whereas the Einstein causality (\ref{Einstein}) allows a traditional formulation in terms of pointlike fields $A(x)$ as \begin{equation} \left[ A(x),A(y)\right] =0,\,\,\,\left( x-y\right) ^{2}<0, \end{equation} Haag duality can only be formulated in the algebraic net setting of LQP, since it is not a property which can be expressed in terms of individual operators. This aspect is shared by many other important properties and results \cite{Haag}. One can prove that Haag duality always holds after a suitable extension of the net to the so-called dual net $\mathcal{A}(\mathcal{O})^{d}.$ The latter may be defined independent of locality in terms of relative commutation properties as \begin{equation} \mathcal{A}(\mathcal{O})^{d}:=\bigcap_{\mathcal{O}_{1},\mathcal{O}% _{1}^{\prime }\subset \mathcal{O}}\mathcal{A}(\mathcal{O}_{1})^{\prime } \end{equation} The relative commutance with respect to the observables is called (algebraic) ``localizability''. These considerations show that causality, locality and localization in LQP have a natural and deep relation to the notion of compatibility of measurements. In addition there are subtle modifications with respect to the basic quantum structure with possible changes of environmental and other aspects of quantum measuring. The fundamental reason for all such modifications in the interpretation of LQP versus QM is the different structure of local algebras: the vacuum is not a pure state with respect to any algebra which is equal to or contained in an $% \mathcal{A}(\mathcal{O})$ with $\mathcal{O}^{\prime }$ nonempty, and the sharply localized algebras $\mathcal{A}(\mathcal{O})$ themselves do not admit pure states at all\footnote{% In order to find local algebras which are anywhere near quantum mechanical algebras and admit pure states and tensor products with entanglement similar to the inside/outside quantization box situation in Schr\"{o}dinger theory, one has to allow for a ``fuzzy'' transition ``collar'' between a double cone and its causal disjoint outside, in more precise terms one has to consider a so-called split inclusion \cite{Haag}.}! They possess an algebraic structure which has not been taken into account in the present day presentation of quantum basics including quantum computation. Since these fine points can only be appreciated with some more preparation, I will postpone their presentation. If the vacuum net (i.e. the vacuum representation of the observable net) is Haag dual, then all associated ``charged'' nets share this property, unless the charges are nonabelian (in which case the deviation from Haag duality is measured by the Jones index of the above inclusion, or in physical terms the statistics- or quantum-dimension \cite{S-W}). If on the other hand even the vacuum representation of the observable net violates Haag duality, then this indicates spontaneous symmetry breaking \cite{Roberts} i.e. not all internal symmetry algebraic automorphisms are spatially implementable. As already mentioned, in that case one can always maximize the algebra without destroying causality and without changing the Hilbert space, such that Haag duality is restored. This turns out to be related to the descend to the unbroken part of the symmetry which allows (since it is a subgroup) more invariants i.e. more observables. Since QM and what is usually referred to as the basics of quantum theory do not know these concepts at all, I am presenting in some sense a contrasting program to the (global) QT orientation of this symposium. But often one only penetrates the foundations of a framework more profoundly, if one looks at a contrasting structure even if the difference is (presently) not measurable. For an analogy we may refer to the Hawking effect which has attracted ever increasing attention as a matter of principle, even though there is hardly any experimental chance. In connection with this main theme of this symposium, it is interesting to ask if LQP could add something to our understanding of classical versus quantum reality (the ERP, Bell issue) or the measurement process i.e. production of ``Schr\"{o}dinger cat states'' and observation of their subsequent decoherence. For the first issue I refer to \cite{Summers}. Apart from some speculative remarks \cite{Landsman}, there exists no investigation of the measurement process which takes into consideration the characteristic properties of the local algebras in LQP. I tend to believe that, whereas most of the present ideas on coherent states of Schr\"{o}dinger cats and their transition to von Neumann mixtures will remain or at least not suffer measurable quantitative modifications, LQP could be expected to lead to significant conceptual changes. Certainly it will add a universal aspect to the issue of decoherence through environments. Contrary to QM where the environment is introduced by extending the system, localized systems in LQP are always open subsystems for which the ``causal disjoint'' defines a kind of universal environment which is build into its formalism. Another structurally significant deviation which was already alluded to results from the fact that the vacuum becomes a thermal state with respect to the local algebras $\mathcal{A}(\mathcal{O}).$ There are two different mechanisms to generate thermal states: the standard coupling with a heat bath and the thermal aspect through restriction or localization and the creation of horizons \cite{Sew}\cite{Verch}. The latter is in one class with the Hawking-Unruh mechanism; the difference being that in the localization situation the horizon is not classical i.e. is not defined in terms of a differential geometric Killing generator of a symmetry transformation of the metric. The fact that algebras of the type $\mathcal{A}(\mathcal{O})$ have no pure states is related to the different behavior of the pair inside/outside with respect to factorization: whereas in QM the boxed system factorizes with the system outside the box, the total algebra $B(H)$ in LQP is generated by $% A(O) $ and its commutant $B(H)=A(O)\vee A(O)^{\prime },$ but \textit{it is not the tensor product} of the two factor algebras $\mathcal{A}(\mathcal{O})$ and $\mathcal{A}(\mathcal{O})^{\prime }=\mathcal{A}(\mathcal{O}^{\prime }).$ In order to get back to a tensor product situation and be able to apply the concepts of entanglement and entropy, one has to do a sophisticated split which is only possible if one allows for a ``collar'' (see later) between $% \mathcal{O}$ and $\mathcal{O}^{\prime }$ \cite{Haag}. Since the thermal aspects of localization are analogous to black holes% \footnote{% The analogy is especially tight for the wedge localization since the boundary of wedges define bifurcated classical ``Killing horizons'' (Unruh), whereas the boundary of e.g. a double cone in a massive theory defines a ``quantum horizon''. This concept has a cood meaning with respect to the nongeometrically acting modular group associated with the latter situation, and it has no classical analogon (it is in fact a ``hidden symmetry'').}, there is no chance to directly measure such tiny effects. However in conceptual problems, e.g. the question if and how not only classical relativistic field theory, but also QFT excludes superluminal velocities, these subtle differences play a crucial role. Because of an unusual property of the vacuum in QFT (the later mentioned Reeh-Schlieder property), the exclusion of superluminal velocities requires more conceptual and mathematical understanding than in the classical case. Imposing the usual algebraic structure of QM (i.e. assuming tacitly that the local observables allow pure states) onto the local photon observables will lead to nonsensical results. Most sensational theoretical observations on causality violations which entered the press and in one case even Phys. Rev. Letters, suffer from incorrect tacit assumptions (if they are not already caused by a misunderstanding of the classical theory). We urge the reader to look at the fascinating reference \cite{Yng-Buch} and the conceptually wrong preceding article. Historically the first conceptually clear definition of localization of relativistic wave function was given by Newton and Wigner \cite{New Wig} who adapted Born's x-space probability interpretation to the Wigner relativistic particle theory. Apparently the result that there is no exact satisfactory relativistic localization (but only one sufficient for all practical purposes) disappointed Wigner so much, that he became distrustful of the usefulness of QFT in particle physics altogether (private communication by R. Haag). Whereas we know that this distrust was unjustified, we should at the same time acknowledge his stubborn insistence in the importance of the locality concept which he thought of as an indispensable requirement in addition the positive energy property and irreducibility of the Wigner representations. Without explanation we state that modular localization of state vectors is different from the Born probability interpretation. Rather subspaces of modular localized wave functions preempt the existence of causally localized observables already on the level of the Hilbert space of relativistic wave functons and have no counterpart at all in N-particle quantum mechanics. As will be explained later, modular localization may serve as a starting point for the construction of interacting nonperturbative LQP's \cite{S-W}\footnote{% In fact the good modular localization properties are guarantied in finite component positive energy representations, with the Wigner infinite component ``continuous spin'' representations being the only exception.. In this infinite component finite energy representation it is not possible to come from the wedge localization down to the spacelike cone localization which is the coarsest localization which one needs for a particle interpretation.}. It is worthwhile to emphasize that sharper localization of local algebras in LQP is not defined in terms of support properties of classical smearing functions but via the rather unusual formation of intersection of localized algebras; although in some cases as CCR- or CAR-algebras (or more generally Wightman fields) the algebraic formulation (% \ref{Einstein}) can be reduced to this more classical concept. Since the modular structure is related to the so-called KMS property \cite {Haag}, it is not surprising that the modular localization has thermal aspects. In fact as mentioned before, there are two manifestations of thermality, the standard heat bath thermal behavior which is described by Gibbs formula or, after having performed the thermodynamic limit, by the KMS condition, and thermality caused by localization either with classical bifurcated Killing-horizons as in black holes \cite{Sew}\cite{Verch} curved spacetime and (Rindler, Unruh, Bisognano-Wichmann) wedge regions, or in a purely quantum manner as the boundary of the Minkowski space double cones. In the latter case the KMS state has no natural limiting description in terms of a Gibbs formula (which only applies to type $I$ and $II$, but not to type $III$ von Neumann algebras), a fact which is also related to the boundedness from below of the Hamiltonian, whereas the e.g. Lorentz boost (the modular operator of the wedge) does not share this property. In \cite {Jaekel} the reader also finds an discussion of localization and cluster properties in a heat bath thermal state. Although in these notes we will not enter these interesting thermal aspects, it should be emphasized that thermality (similar to the concept of virtual particle clouds) is an inexorable aspect of localization in LQP and does not need the Hawking type of Killing vector horizons. The close relation of particle and thermal physics (KMS thermal property$\simeq $crossing symmetry of S-matrix and formfactors \cite{S-W}) is a generic property of LQP and should not be counted as a characteristic success of string theory. Already in the very early development of algebraic QFT \cite{Haag-Sch} the nature of the local von Neumann algebras became an interesting issue. Although it was fairly easy (and expected) to see that i.e. wedge- or double cone- localized algebras are von Neumann factors (in analogy to the tensor product factorization of standard QT under formation of subsystems, it took the ingenuity of Araki to realize that these factors were of type $III$ (more precisely hyperfinite type $III_{1},$ as we know nowadays thanks to the profound contributions of Connes and Haagerup), at that time still an exotic mathematical structure. Hyperfiniteness was expected from a physical point of view, since approximatability as limits of finite systems (matrix algebras) harmonizes very well with the idea of thermodynamic+scaling limits of lattice approximations. A surprise was the type $III_{1}$ nature which,as already mentioned, implies the absence of pure states (in fact all projectors are Murray von Neumann equivalent to 1) on such algebras; this property in some way anticipated the thermal aspect (Hawking-Unruh) of localization. Overlooking this fact (which makes local algebras significantly different from QM), it is easy to make conceptual mistakes which could e.g. suggest an apparent breakdown of causal propagation \cite {Yng-Buch} as already mentioned before. If one simply grafts concepts of QM onto the causality structure of LQP (e.g. quantum mechanical tunnelling, structure of states) without deriving them in LQP , one runs the risk of wrong conclusions about e.g. the possibility of superluminal velocities. A very interesting question is: what is the influence of the always present causally disjoint environment on the measurement process, given the fact that in the modern treatment the coupling to the environment and the associated decoherence relaxation are very important. Only certain aspects of classical versus quantum reality, as expressed in terms of Bell's inequalities, have been discussed in the causal context of LQP \cite{Summers}% . In the following we will sketch some more properties which set apart QM from LQP and whose conceptual impacts on decoherence of Schr\"{o}dinger cats, entanglement etc. still is in need of understanding. Let me mention two more structural properties, intimately linked to causality, which distinguish LQP rather sharply from QM. One is the Reeh-Schlieder property: \begin{eqnarray} \overline{\mathcal{P}(\mathcal{O})\Omega } &=&H,\,\,i.e.\,\,cyclicity\,\,of% \,\,\Omega \label{cyc} \\ A\in \mathcal{P}(\mathcal{O}),\,\,A\Omega &=&0\Longrightarrow A=0\,\,i.e.\,\,\Omega \,\,separating \nonumber \end{eqnarray} which either holds for the polynomial algebras of fields or for operator algebras $\mathcal{A}(\mathcal{O}).$ The first property, namely the denseness of states created from the vacuum by operators from arbitrarily small localization regions (a state describing a particle behind the moon% \footnote{% This weird aspect should not be held against QFT but rather be taken as indicating that localization by a piece of hardware in a laboratory is also limited by an arbitrary large but finite energy, i.e. is a ``phase space localization'' (see subsequent discussion). In QM one obtains genuine localized subspaces without energy limitations.} and an antiparticle on the earth can be approximated inside a laboratory of arbitrary small size and duration) is totally unexpected from the global viewpoint of general QT and has even attracted the interest of philosophers of natural sciences. If the naive interpretation of cyclicity/separability in the Reeh-Schlieder theorem leaves us with a feeling of science fiction, the way out is to ask: which among the dense set of localized states can be really produced with a controllable expenditure (of energy)? In QM to ask this question is not necessary since, as already mentioned, the localization at a given time via support properties of wave functions leads to a tensor product factorization of inside/outside so that the ground state factorizes and the application of the inside observables never leads to a dense set in the whole space. It turns out that most of the very important physical and geometrical informations are encoded into features of dense domains and in fact the aforementioned modular theory is explaining such relations. For the case at hand, the reconciliation of the Reeh-Schlieder theorem with common sense has led to the discovery of the physical relevance of \textit{localization with respect to phase space in LQP}, i.e. the understanding of the size of degrees of freedom in the set: \begin{eqnarray} P_{E}\mathcal{A}(\mathcal{O})\Omega \,\,is\;compact &&\,\,\, \\ \,\,e^{-\beta \mathbf{H}}\mathcal{A}(\mathcal{O})\Omega \;is\,\,nuclear,\,\,\,\,\mathbf{H} &=&\int EdP_{E}\,\, \nonumber \end{eqnarray} The first property was introduces way back by Haag and Swieca \cite{Haag}, whereas the second statement (and similar nuclearity statements involving modular operators of local regions instead of the global Hamiltonian) which is more informative and easier to use, is a later result of Buchholz and Wichmann \cite{Haag}. It should be emphasized that the LQP degrees of freedom counting of Haag-Swieca, which gives an infinite (but still nuclear) number of localized states is different from the finiteness in QM, a fact often overlooked in present day's string theoretic degree of freedom counting. The difference to the case of QM decreases if one uses instead of a strict energy cutoff a Gibbs damping factor $\ e^{-\beta H}.$ In this case the map $\mathcal{A}(\mathcal{O})\rightarrow e^{-\beta H}\mathcal{A}(% \mathcal{O})\Omega $ is ``nuclear'' if the degrees of freedom are not too much accumulative in order to prevent the existence of a maximal (Hagedorn) temperature. The nuclearity assures that a QFT, which was given in terms of its vacuum representation, also exists in a thermal state. An associated nuclearity index turns out to be the counterpart of the quantum mechanical Gibbs partition function \cite{Haag} and behaves in an entirely analogous way. The peculiarities of the above Haag-Swieca degrees of freedom counting are very much related to one of the oldest ``exotic'' and at the same time characteristic aspects of QFT: vacuum polarization. As discovered by Heisenberg, the partial charge: \begin{equation} Q_{V}=\int_{V}j_{0}(x)d^{3}x=\infty \end{equation} diverges as a result of uncontrolled vacuum fluctuations near the boundary. For the free field current it is easy to see that a better definition involving test functions, which takes into account the fact that the current is a 4-dim distribution and has no restriction to equal times, leads to a finite expression. The algebraic counterpart is the already mentioned so called ``split property'' namely \cite{Haag} that if one leaves between say the double cone (``relativistic box'') observable algebra $\mathcal{A}(% \mathcal{O})\,$and its causal disjoint $\mathcal{A}(\mathcal{O}^{\prime })$ a ``collar'' region, then it is possible to construct in a canonical way a type $I$ tensor factor $\mathcal{N}$ which extends into the collar and one obtains inside/outside factorization if one leaves out the collar region (a fuzzy box). This is then the algebraic analog of Heisenberg's smoothening of the boundary to control vacuum fluctuations. It is this ``split inclusion'' which allows to bring back some of the familiar structure of QM, since type I factors allow for pure states, tensor product factorization, entanglement and all the other properties at the heart of quantum theory and the measurement process. Although there is no time to explain this, let us nevertheless mention that the most adequate formalism for LQP which substitutes quantization and is most characteristic of LQP in contradistinction to QT, is the formalism of modular localization related to the Tomita modular theory of von Neumann algebras. The interaction enters through wedge algebras, thus giving wedges a similar fundamental role as they already had in the Unruh illustration of the thermal aspects of the Hawking effect. Modular localization also leads to a vast enlargement of the symmetry concepts in QFT \cite{Hidden}\cite{Buch} beyond those geometric symmetries which enter the theory through quantized Noether currents. If by these remarks I have created the impression that local quantum physics is one of the conceptually most fertile and spiritually (not historically) young areas of future basic research with relevance to the basics of measurement and quantum computation, I have accomplished the purpose of these notes. Indeed I know of no other framework which brings together such seemingly different ideas as Spin \& Statistics, TCP and crossing symmetry of particle physics on the one hand together with thermal and entropical aspects of (modular) localization \& black hole physics on the other hand.
\section{Introduction} Some kinds of bacterial colonies present interesting structures during their growth \cite{Mat,ben1,ben2,budrene,shapiro,rauprich,williams,rafols,Matsus,Ohg,fujikawa}. Depending on the bacterial species and the culture conditions, colonies can exhibit a great diversity of forms. In general, the complexity of the growth pattern increases as the environmental conditions become less favorable. Bacteria respond to adverse growth conditions by developing sophisticated strategies and higher micro-level organization in order to cooperate more efficiently. Examples of these strategies are: the differentiation into longer-motile bacteria, the production of extracellular wetting fluid, the secretion of surfactants which change the surface tension or the chemotactic response to chemical agents produced by bacteria \cite{Mat,ben1,ben2,budrene,shapiro,rauprich,williams,rafols}. The experiments are usually made in a petri dish, which contains a solution of nutrient and agar. A drop of bacterial solution is then inoculated in the center of the dish. The growth conditions are controlled by the initial concentration of the medium components. The agar concentration determines the consistency of the medium, which becomes harder as the amount of agar increases, and the nutrient concentration controls the bacterial reproduction. Depending on these two factors, the colony grows at a higher or lower rate, developing different kinds of patterns. In particular, colonies of the bacterium {\it Bacillus subtilis} OG-01 present a rich variety of structures \cite{rafols,Matsus,Ohg,fujikawa}. Fig.\ \ref{fig1} shows the morphological diagram obtained by Matsushita and co-workers \cite{Ohg}. They classified the colony patterns into five types, from A to E, whose main features can be summarized as follows. If the medium is very hard, i.e., with a high concentration of agar, bacteria can hardly move and the colony essentially grows due to the consumption of nutrient and subsequent reproduction. If the level of nutrient is also low (region $A$), the growth is controlled by the diffusion of the nutrient up to the bacteria placed at the interface. The colony develops a ramified structure very similar to the patterns obtained with the diffusion-limited aggregation model (DLA)\cite{Matsus,witten}. It takes approximately one month to cover the dish. If the initial agar concentration remains high and the nutrient concentration is increased, the growth is faster than in region $A$. The branches grow thicker until they fuse into a dense disk with rough interface (region $B$), similar to the patterns obtained with an Eden model\cite{vicsek}. This structure needs 5-7 days to cover the disk. When the level of agar is decreased, which produces a medium a little softer than in region B, and the level of nutrient remains high, the colony forms concentric rings (region $C$). This region is characterized by periodic dynamics: for 2-3 hours the colony expands while the bacteria move actively ("migration phase") and then they almost stop for 2-3 hours ("consolidation phase"), during which the colony does not grow appreciably and the bacterial density increases due to reproduction. The crossover between the two phases is sharp. The periodic cycles of subsequent migration and consolidation phases create the pattern of concentric rings \cite{rafols,fujikawa}. Accurate measurements show that in the growth phase there is a high concentration of longer and more motile bacteria, as a consequence of a differentiation process \cite{rafols}. When a high level of nutrient is maintained, and the agar concentration is decreased further, the colony spreads over the agar plate, and after less than 8 hours homogeneous disk of low bacterial density is formed covering all the dish (region $D$). In this thin surface, bacteria are always short and can move easily by swimming. By decreasing the nutrient concentrations for a semi-solid medium, the colony develops a densely branched pattern (region $E$) similar to the dense branching morphology (DBM) found in other systems \cite{ben3}. The ratio of the width of the branches to the gap between them is constant over the whole colony. The colony grows quite fast, showing its main activity at the tips of the fingers, and covering the dish in less than 24 hours. The dynamics is related to both the consumption of nutrients and the bacterial motility. In general, when environmental conditions are adverse (low nutrient or hard surface), a higher level of cooperation is observed. The existence of a cooperative behavior seems to be determinant in the formation of the rings patterns of region $C$. The same kind of concentric rings has been found in experiments with other bacterial species. In the case of the bacterium {\it Proteus mirabilis}, the migration phases clearly involve the movement of differentiated swarmer bacteria (elongated and hyperflagellated)\cite{shapiro,rauprich,williams}. Similar ring patterns have also been observed in other non-living systems, like the Liesegang rings produced by precipitation in the wake of a moving reaction front \cite{chopard}, or some experiments of interfacial electrodeposition \cite{zeiri}. In the case of the Liesegang patterns, it is well known that the distance between rings increases as $t^{1/2}$, whereas in bacterial and electrodeposition it is constant. \begin{figure} \begin{center} \def\epsfsize#1#2{0.50\textwidth} \leavevmode \epsffile{fig1.ps} \end{center} \caption{ \label{fig1} Morphological diagram of patterns observed in colonies of {\it Bacilus subtilis} OG-01 as a function of nutrient concentration ($C_n$) and the hardness of agar surface ($1/C_a$ being $C_a$ the agar concentration). Experiments were performed in a petri dish with a diameter of 88 mm. Taken with permission from \protect\cite{rafols}.} \end{figure} Several models have been proposed to explain the variety of patterns exhibited by {\it Bacillus subtilis}, as shown in Fig.\ \ref{fig1}. \cite{ben1,ben2,Matsus,Murray,Wakita,kawasaki,mimura,esipov}. DLA-like patterns (region $A$ in Fig.\ \ref{fig1}) have been interpreted \cite{Matsus} as growth controlled by the diffusion of nutrients in the context of the DLA model. Ben-Jacob and co-workers proposed a communicating walkers model to describe some of the morphologies \cite{ben1,ben2}. This model reproduces the crossover between regions A and B by coupling random walkers to fields representing the nutrients. DBM-like patterns are also obtained by introducing a chemotactic agent. Other kinds of models are based on reaction-diffusion equations for bacterial density. The Fisher equation \cite{Murray} can be used for reproducing the homogeneous circular morphology (region $D$) \cite{Wakita}. Further developments were achieved by introducing new elements to the Fisher model, such as a field for nutrient and nonlinear diffusion coefficients \cite{kawasaki,mimura}. Depending on the new elements introduced, these models reproduce some of the patterns of Fig.\ \ref{fig1}. However, the ring patterns (region $C$) have so far eluded a satisfactory modelization. Although the model suggested in Ref.\ \cite{mimura} can generate concentric ring patterns, they are rather different from those observed in experiments \cite{rafols}. In fact, dynamical cycles of consolidation and growth phases are not found. Finally, we must mention a model proposed by Esipov {\it et al.} \cite{esipov} for the study of {\it Proteus mirabilis} colonies, which introduces a life-time for the differentiated swarmer bacteria. This model reproduces concentric ring patterns but does not explain why no periodicity is observed in other regions of the morphological diagram. In this paper, we propose a model consisting of two coupled diffusion-reaction equations for bacteria and nutrient concentrations, where the bacterial diffusion coefficient can adopt two different expressions, corresponding to two possible mechanisms of motion. The first is the usual random swimming performed by bacteria in liquid medium. The second is developed by bacteria in response to adverse growth conditions, and depends on their concentration. Bacterial response is modeled as a global variable that can present hysteresis. Our model reproduces the five morphologies observed in the experiments, including the ring patterns. \section{Mathematical Model} We consider a two-dimensional system containing bacteria and nutrients. Both diffuse, while bacteria proliferate by feeding on nutrient. Let us denote by ${\bar b}({\bf r}',\tau)$ the density of bacteria at time $\tau$ and spatial position ${\bf r}'$, and by ${\bar n}({\bf r}',\tau)$ the concentration of nutrient. Then, ${\bar b}$ and ${\bar n}$ are in general governed by the following equations\cite{Murray,kawasaki}: \begin{eqnarray} \frac{\partial {\bar b}}{\partial \tau}&=& \nabla D_b \nabla {\bar b} + \theta f({\bar b},{\bar n}), \nonumber \\ \label{mod1} \\ \frac{\partial {\bar n}}{\partial \tau}&=& D_n \nabla^2 {\bar n} - f({\bar b},{\bar n}). \nonumber \end{eqnarray} The function $f({\bar b},{\bar n})$ denotes the consumption term of nutrient by bacteria, and can be described by Michaelis-Menten kinetics\cite{Murray} \begin{equation} f({\bar b},{\bar n})=\frac{k {\bar n}{\bar b}}{(1+\gamma' {\bar n})}, \label{growth} \end{equation} where $k$ is the intrinsic consumption rate. For small ${\bar n}$, the consumption rate is approximately linear in ${\bar n}$ and it saturates at the value $k/\gamma'$ as ${\bar n}$ increases. $D_b$ and $D_n$ are the diffusion coefficients of bacteria and nutrient respectively. We assume that $D_n$ is constant, but $D_b$ can depend on nutrient and bacterium concentrations. As explained above, experiments show that in adverse conditions, bacteria can adapt themselves in order to improve their motility. In a soft medium and high nutrient concentration (region $D$ of Fig.\ \ref{fig1}), short bacteria can swim randomly without difficulty, but in an adverse environment (regions $A, B, C$ and $E$) they need to develop mechanisms to become more motile. For intermediate conditions of semi-solid medium and sufficient nutrient, there are periods of fast growth (migration phase) and slow growth (consolidation phase) that lead to the concentric ring patterns. The analysis of periodic rings suggests a dynamical scheme with hysteresis that can be outlined in the following way: during the consolidation phase, the population of longer-motile bacteria increases in order to overcome the opposition to the movement. When this population exceeds a certain value, enhanced movement becomes possible and a migration phase begins. Then, however, a progressive decrease in long-bacterial population ensues, until it reaches a minimum at which the "enhanced-movement mechanism" does not work. Then a new consolidation phase begins. Within this scheme, region $D$ corresponds to a case where the maximum value is never reached (and therefore bacteria always move by usual diffusion), whereas in regions $A, B$ and $E$ long-bacterial population does not fall below the minimum (and therefore always moves by the enhanced-movement mechanism). All these ideas can be introduced in our model by means of two basic points: (a) The diffusion coefficient $D_b({\bar n},{\bar b})$ can take two different expressions depending on the long-bacterial population. (b) The net production of long bacteria depends on the environmental conditions and also on the colony phase of growth. According to these ideas, we propose the following function for $D_b$: \begin{equation} D_b=D ({\bar d}_1+{\bar d}_2 {\bar b}) {\bar n}, \label{dif} \end{equation} where $D$ depends on the concentration of agar, which is lower for harder medium. To take into account the inhomogeneities of the medium, we introduce a quenched disorder in $D$, which is written as $D=D_0(1+\xi ({\bf r}'))$, $\xi ({\bf r}')$ being a random term defined on a square lattice. From now on, $D_0$ will be referred to as the diffusion parameter. The first term of Eq.\ (\ref{dif}) describes the usual diffusion of bacteria in a liquid medium. The second describes the cooperative enhanced-movement mechanism promoted by long bacteria. This second mechanism can be modeled by a diffusion coefficient that depends on the bacterial concentration. We multiply both terms by nutrient concentration to take into account the fact that bacteria are inactive in the region where nutrient has been depleted. This dependence on ${\bar n}$ would not have been necessary if we had considered a "death" term in the equation for ${\bar b}$. The coefficients $d_1$ and $d_2$ can adopt two different values (one of them zero) depending on the concentration of the long bacteria, as will be specifed below. Equations (\ref{mod1}) with Eqs.\ (\ref{growth})-(\ref{dif}) can be written in a simpler form as \begin{eqnarray} \frac{\partial b}{\partial t}&=& \nabla \left \{ D \left( d_1 +d_2 b \right ) n \nabla b \right \} + \frac{nb}{1+\gamma n}, \nonumber \\ \label{model} \\ \frac{\partial n}{\partial t}&=& \nabla^2 n -\frac{nb}{1+\gamma n}, \nonumber \end{eqnarray} with \begin{eqnarray} {\bf r}&=&\left( \frac{\theta k^2}{D_n} \right )^{1/4} {\bf r}' ~,~~~ t = k \left ( \theta D_n \right)^{1/2} \tau ~, \nonumber \\ n&=&\left (\frac{\theta}{D_n}\right )^{1/2}{\bar n}~,~~~ b=\left (\frac{1}{\theta D_n}\right )^{1/2}{\bar b}~, \label{dimen} \\ \gamma& =& \left ( \frac{D_n}{\theta} \right )^{1/2} \gamma'~,~~~ d_1=\left (\frac{1}{\theta D_n}\right )^{1/2}{\bar d}_1~, ~~~ d_2={\bar d}_2. \nonumber \end{eqnarray} At this point, we need to specify how to choose $d_1$ and $d_2$ depending on the population of long bacteria. In order to do this, we introduce a global phenomenological quantity $W(t)$ that measures the amount of long bacteria. The evolution of this quantity should have a "creation term" that represents the transformation of short bacteria into long ones, and an "annihilation term" that represents the opposite transformation (septation). It seems reasonable to assume that the creation term is directly dependent on the mean bacterial concentration, and inversely dependent on the level of nutrient ($n_0$) and on the diffusion parameter (adverse conditions, {\it i.e.} $D_0$ and $n_0$ small, means a faster differentiation process). With regard to the annihilation term, it can adopt two possible values depending on the growth phase. The simplest equation that includes all these considerations can be written as \begin{eqnarray} \frac{\partial W}{\partial t}&=&\lambda \frac{B}{n_0 D_0}-c_i, \label{population} \end{eqnarray} where $\lambda$ is a constant and $c_i$ can have two different values ($c_g$ or $c_s$, $c_g>c_s$). The quantity $B$, defined as $B=\sum b^2 / \sum b$, is a measure of the mean concentration of bacteria inside the colony. We introduce the hysteresis previously pointed out by assuming that there are two limit values $W_{MAX}$ and $W_{MIN}$ for which: \begin{eqnarray} {\rm when}~~~ W &\ge& W_{MAX} ~~~ {\rm then}~~~ c_i=c_g~,~~~ d_2=D_2~,~~~d_1=0, \nonumber \\ \label{cond} \\ {\rm when}~~~ W &\le& W_{MIN} ~~~ {\rm then}~~~ c_i=c_s~,~~~ d_2=0~,~~~d_1=1. \nonumber \end{eqnarray} With a suitable choice of parameters $\lambda$, $c_s$ and $c_g$, and by changing only $n_0$ and $D_0$, we can obtain colonies that always move with one of the two types of diffusion, or colonies that periodically change from one type to the other. This will occur if $c_s < \lambda B / (n_0 D_0) < c_g$, which will give rise to the ring patterns. Although the bacterial response has been expressed in terms of the population of long-bacteria, other possible kinds of responses admit identical modelization. In this sense, our model is quite general. \section{Numerical results} We have numerically integrated Eqs.\ (\ref{model}) with (\ref{population})-(\ref{cond}) in a square lattice of lateral size $L=600$ using a 4-th order Runge-Kutta's method with mesh-size $\Delta x=0.5$ and time step $\Delta t=0.005$. The system was initially prepared by assigning to each point a nutrient concentration $n({\bf r},0)=n_0+\eta({\bf r},0)$, $\eta$ being a uniform random number in the interval $(-0.1,0.1)$, and a bacterial concentration $b({\bf r},0)=0$, except in a small central square where $b({\bf r},0)=b_0$. The random term of the diffusion, $\xi({\bf r})$, takes a different and uncorrelated value in each box of side $4\Delta x$. The random values are assumed to be uniformly distributed in the interval $(-\epsilon,\epsilon)$. The box size and the intensity $\epsilon$ do not essentially affect the results. In all our simulations, we used the parameters $b_0=0.7$, $\gamma=0.5$, $D_2=30$, $\epsilon=0.4$, $\lambda=0.18$, $c_g=2$, $c_s=1.6$, $W_{MAX}=3$ and $W_{MIN}=2$. We reproduce the different morphologies observed in experiments by changing the values of the initial concentration of nutrients $n_0$ and the softness of the media, related to $D_0$. \begin{figure} \begin{center} \def\epsfsize#1#2{0.50\textwidth} \leavevmode \epsffile{fig2.ps} \end{center} \caption{ \label{fig2} Bacterial colonies (top) and nutrient patterns (bottom) for a fixed value of $D_0=0.01$ and two values of initial nutrient: $n_0=1 (a)$ and $n_0=5 (b)$. They correspond to times $t=2500$ and $75$ respectively.} \end{figure} In Figs. \ref{fig2}-\ref{fig3} we present the results obtained for $D_0=0.005$ and different values of $n_0$. By increasing $n_0$ we reproduce the crossover between regions $A$ and $B$ of Fig.\ \ref{fig1}, from DLA-like patterns (Fig.\ \ref{fig2}(a)) to a dense rough structure similar to that found with an Eden model (Fig.\ \ref{fig2}(b)). All of them correspond to a situation in which, due to the small value of $D_0$, the creation term of Eq.\ (\ref{population}) is greater than $c_g$, except at the very beginning. The response $W$ can never decrease below the value $W_{MIN}$, and therefore the colony will always grow with the enhanced-movement mechanism. In spite of this cooperative mechanism, and because of the hardness of the medium, the effective diffusion $D_0 D_2 b n$ is still small. The growth is mostly due to reproduction by feeding on the nutrient. For low level of nutrient, {\it i.e.} small $n_0$, the colony growth is limited by the diffusion of these nutrients. It develops branches, which are thicker as $n_0$ increases. The prototype model that reproduces this kind of structure is the diffusion-limited aggregation (DLA) \cite{witten}, which is known to form a fractal pattern with a fractal dimension of $d_F=1.71$\cite{Mat,barabasi}. Experiments performed by Matsushita et al. \cite{Matsus} in region $A$ of Fig.1 also show a fractal growth with dimension $d_F=1.73$ \cite{Matsus}. We have analyzed the fractal nature of the patterns obtained with our model, for $D_0=0.005$ and several values of the initial nutrient, from $n_0=1$ (DLA-like) to $n_0=5$ (rough structure). We have calculated their fractal dimensions by using the box-counting method \cite{Mat,barabasi}. In Fig.\ \ref{fig3} we show, in a log-log plot, the number $N$ of boxes of size $e$ that contains any part of the pattern, versus the size of the boxes. The slopes of the lines represent the fractal dimensions. We observe that the cases that correspond to low nutrient have a fractal dimension of about $d_F=1.73$, showing good agreement with experiments. On the other hand, there is an abrupt change between these patterns and those that are not fractal ($d_F=1$). These last cases can be analyzed in terms of the roughness of their interfaces. \begin{figure} \begin{center} \def\epsfsize#1#2{0.50\textwidth} \leavevmode \epsffile{fig3.ps} \end{center} \caption{ \label{fig3} Fractal dimensions determined by the box--counting method, for $D_0=0.01$ and $n_0=1 (\bigcirc),2 (\Diamond), 3 (\Box)$ and $5 (\triangle)$. Two lines of slopes $1$ and $1.73$ are also plotted for comparison with experiments.} \end{figure} It is well known that Eden structures are not themselves fractal, but their surfaces exhibit a self-affine scaling \cite{barabasi,krug}. This implies that, for a long enough time, the width of the rough interface $\sigma$ scales with an exponent $\alpha$ as a function of the length of the interface $l$ ($\sigma \sim l^\alpha$). The roughness exponent for the Eden model is $\alpha=0.5$. Vicsek et al. \cite{vicsek} analyzed experimental data corresponding to the region $B$ of Fig.\ \ref{fig1}. They concluded that these colony surfaces are self-affine with a roughness exponent $\alpha=0.74$. We have checked this point for our dense rough pattern (Fig.\ \ref{fig2}(b)) by measuring the width $\sigma$ for intervals of interface of length $l$. The results, as a function of $l$, are presented in Fig.\ \ref{fig4}. In order to avoid additional effects derived from the radial growth of the colony, we have also performed a complementary simulation for the same parameters as Fig.\ \ref{fig2}(b) but with a strip geometry. To do this, we have used a rectangular lattice of horizontal lateral size $L_x=600$, with periodic boundary conditions in $x$ direction, and taken as an initial condition for bacteria a horizontal line of length $L_x$. The results for this case are also plotted in Fig.\ \ref{fig4}. For both circular and strip cases, we observe analogous behavior to that observed in experiments \cite{vicsek}. Our results show a linear region with a slope compatible with the experimental value $\alpha=0.74$. \begin{figure}[h] \begin{center} \def\epsfsize#1#2{0.50\textwidth} \leavevmode \epsffile{fig4.ps} \end{center} \caption{ \label{fig4} Width of the rough interface as a function of the length of the interval in which it is measured. They correspond to the pattern of Fig.\ \ref{fig2}(b) ($\Box$) and to a case with the same parameters but with strip geometry ($\triangle$). A line of slope $0.74$ is also plotted for comparison with experimental results.} \end{figure} With the aim of reproducing other morphologies of Fig.\ \ref{fig1}, we now keep the initial nutrient fixed at the value $n_0=1$ and increase the diffusion parameter $D_0$. Results are shown in Fig.\ \ref{fig5}(a)-(b). We observe a crossover from the DLA-like structure (Fig.\ \ref{fig2}(a)) to a dense branching morphology analogous to that represented in region $E$ of Fig.\ \ref{fig1}. In Fig.\ \ref{fig5}(c)-(d), we present two snapshots obtained for a fixed value of the initial nutrient $n_0=5$. They show how different kinds of patterns are obtained when $D_0$ is increased: from the dense rough structure (Fig.\ \ref{fig2}(b)), to concentric rings (Fig.\ \ref{fig5}(c)) and homogeneous disk (Fig.\ \ref{fig5}(d)). They correspond to the regions $B, C$ and $D$ respectively. Homogeneous disks are obtained when $D_0$ and $n_0$ are so high that the creation term of Eq.\ (\ref{population}) is always smaller than $c_s$. This means that the value $W_{MAX}$, above which the enhanced-movement mechanism begins, is never reached, and bacteria move with the usual diffusion coefficient $D_0 n$. \begin{figure}[h] \begin{center} \def\epsfsize#1#2{0.50\textwidth} \leavevmode \epsffile{fig5.ps} \end{center} \caption{ \label{fig5} Patterns obtained for $n_0=1$ with $D_0=0.05 (a)$ and $0.1 (b)$ and for $n_0=5$ with $D_0=0.1 (c)$ and $2 (d)$. They correspond to times $t=500, 300, 50$ and $25$ respectively.} \end{figure} Ring patterns correspond to a narrow region of parameters $D_0$ and $n_0$ for which the creation term of Eq.\ (\ref{population}) takes a value between $c_s$ and $c_g$. As explained in Section II, this leads to dynamics in which bacteria move alternatively by usual diffusion $D_0 n$ (consolidation phase) or by the enhanced-movement mechanism $D_0 D_2 b n$ (migration phase). The two phases are clearly manifested in Fig. \ref{fig6}(a), where we represent the radius of the colony as a function of time. The pattern of concentric rings is a consequence of this dynamic behavior. In Fig.\ref{fig6}(b) we plot the radial density profile, circularly-averaged, corresponding to the ring pattern shown in Fig.\ \ref{fig5}(c). The maxima are formed in the positions where a consolidation phase began. To illustrate this point, we have pointed out in Fig.\ \ref{fig6} the positions corresponding to the colony radius at the beginning of each consolidation phase. \begin{figure}[h] \begin{center} \def\epsfsize#1#2{0.50\textwidth} \leavevmode \epsffile{fig6.ps} \end{center} \caption{ \label{fig6} $(a)$ Time evolution of the colony radius, for parameters $D_0=0.1$ and $n_0=5$; $(b)$ Radial density profile corresponding to time $t=50$ (pattern of Fig.\ \ref{fig5}(c)). The positions where each consolidation phase started are pointed out.} \end{figure} Numerically, our model also reproduces the experimentally observed robustness of the growth-plus-consolidation period, which is barely dependent on changes in either nutrient or agar concentrations over a wide range. For high enough $n_0$, the value of the global quantity $B$ approaches $n_0$. In this limit, as can be derived from Eq.\ (\ref{population}) the period is given by \begin{eqnarray} T=D_0 (W_{MAX}-W_{MIN})\left (\frac{1}{\lambda-D_0 c_s}- \frac{1}{\lambda-D_0 c_g}\right), \label{period} \end{eqnarray} which does not depend on $n_0$. Moreover, as a function of $D_0$, the period also maintains a rather constant value within a certain range (determined by parameters $\lambda$, $c_g$ and $c_s$) to increase sharply in the boundaries of the ring patterns region ($D_0 \longrightarrow {\lambda / c_s}$, $D_0 \longrightarrow {\lambda / c_g}$). For equal period, the width of the rings increases with $n_0$. \section{Conclusions} We have proposed a reaction-diffusion model for the study of bacterial colony growth on agar plates, which consists of two coupled equations for nutrient and bacterial concentrations. The most important feature, which introduces differences from previous models, is the fact that here we consider two mechanisms for the bacterial movement: the random swimming in a liquid medium, and a cooperative enhanced movement developed by bacteria when the growth conditions are adverse. The two mechanisms are introduced in our model by means of a diffusion term with two different expressions which depend on the bacterial response to the environmental conditions. This response is modeled as a global variable that presents hysteresis depending on the conditions of the medium. The inhomogeneities of the agar plate have been taken into account as a quenched disorder in the diffusion parameter. We have shown that, simply by changing the parameters related to the hardness of the medium and the initial nutrient, our model reproduces all the patterns obtained experimentally with the bacterium {\it Bacillus subtilis}: DLA-like, dense-rough disk, DBM-like, ring patterns and homogeneous disk. We have calculated the fractal dimension of the DLA-like structures and the roughness exponent of the rough disk surface, obtaining results in good agreement with experiments. The ring patterns have been obtained for intermediate values of agar and high nutrient. In this region, the bacterial response presents hysteresis and the two mechanisms of motion work alternatively, leading to cycles of migration and consolidation phases. The duration of these cycles is roughly constant for different values of nutrient and agar concentration over a wide range. This periodical dynamics generates patterns of concentric rings. In summary, the model proposed satisfactorily reproduces the whole experimental morphological diagram. It represents a first attempt at describing the response of bacteria to adverse growth conditions and, in certain conditions, their ability to improve their motility. Further refinements could be made. The bacterial response, here described as a global variable $W(t)$, could be considered in a more realistic way by introducing a coupling term in a local version of Eq.\ (\ref{population}) for a field $W({\bf r},t)$. However, preliminar studies with such a model \cite{lacasta} shows the same essential features previously described. \acknowledgements We thank I. R\`afols and J.M. Sancho for helpful discussions. This research was supported by Direcci\'on General de Investigaci\'on Cient\'{\i}fica y T\'ecnica (Spain) (PB96-0241-02), by Comissionat per Universitats i Recerca de la Generalitat de Catalunya (SGR97-439) and by Universitat Polit\`ecnica de Catalunya (PR-9608). We also acknowledge computing support from Fundaci\'o Catalana per a la Recerca and Centre Catal\`a de Computaci\'o i Comunicacions.
\section{Introduction} \indent The 30 micron emission was first observed in the bright AGB star, IRC$+$10216 (Low et al 1973). Observations of several carbon stars by Goebel \& Moseley (1985) found that the emission also occurs in AFGL 3068. Recent ISO observations of some carbon stars turned up a few more candidates: AFGL\,2256, AFGL\,2155 and IRC$+$40540 (Yamamura et al. 1998). This feature is seen not only in extreme C--rich AGB stars, but also in C--rich proto--planetary nebulae (PPNe; called also post--AGB objects) and planetary nebulae (PNe). Omont et al. (1995) observed five C--rich PPN objects with an unidentified emission feature at 21 micron in their IRAS LRS spectra (Kwok et al. 1989) and detected the 30 micron feature in all of them. In the case of PNe a similar feature is observed for IC\,418 and NGC\,6752 (Forrest et al. 1981). A few suggestions for the carrier of this emission have been proposed. Because the feature has never been detected in O--rich objects, and is the broadest known feature in the mid--infrared, solid species that form in the absence of oxides are taken into consideration. Omont et al. (1995) suggested iron atoms bound to PAH molecules as a possible emitter for the 30 micron feature. But Chan et al. (1997) raise doubt about this suggestion after the detection of this feature in direction of the Galactic center, where the existence of PAH molecules is questionable. A more acceptable candidate is solid magnesium sulfide (MgS), first suggested by Goebel \& Moseley (1985). The laboratory spectra of MgS samples showed very good agreement of band turn--on and cut--off wavelengths, as well as the overall band shape, with the observed feature seen in AFGL\,3068 and IRC$+$10216. A reasonable fit was achieved to the 30 micron feature of the PPN object IRAS\,22272$+$5435 (Szczerba et al. 1997) by means of MgS grains with a distribution of shapes. In addition, MgS is one of the molecules which condensate at low temperature when no oxides are present. This is consistent with the detection of the 30 micron feature only in objects with cold dust shells. IRAS\,03313$+$6058 was classified as a candidate extreme carbon star based on the similarity of its IRAS LRS spectrum to those of AFGL\,3068 and IRC$+$10216 (Volk et al. 1992). It has no optical counterpart in the POSS plates (Jiang \& Hu 1992). In the near infrared, the source was detected at K$=$15.6 mag and was not detected in J and H bands with upper limit of magnitude 17 and 16, respectively (Jiang et al. 1997). The color index between 12 and 25 micron, based on the IRAS PSC catalogue, indicates a color temperature of about 250\,K. Therefore, this is an object with a cold and optically thick circumstellar envelope. Though the IRAS LRS type of this object is 22, no clear silicate feature is seen (note that Kwok et al. 1997 classified the LRS spectrum of this object as unusual). The detection of HCN\,(1--0) line (Omont et al. 1993) indicated the possibility of a C--rich nature. O--rich maser lines such as OH (Le Squeren et al. 1992, Galt et al. 1989), H$_{2}$O (Wouterloot et al. 1993) or SiO (Jiang et al. 1996) have not been detected at all and this is an indirect indication of a C--rich circumstellar envelope. The CO (2--1) line profile suggests that this object is a late--type star rather than a young stellar object, and gives an expansion velocity of the shell of 13.9\,km/s (Volk et al. 1993). \section{Observation and data reduction} \indent The spectroscopic observation was carried out by using the SWS spectrometer (de Graauw et al. 1996) of the Infrared Space Observatory (ISO) satellite on 31 July 1997 with the fastest scan speed covering full wavelength range from 2.3 $\mu$m to 45 $\mu$m (AOT\,01, speed 1). The achieved resolution is about 200 $\sim$ 300 with S/N higher than 100 at wavelength longer than 5\,$\mu$m. The original pipeline data were corrected for dark current, up--down scan difference, flat--field and flux calibration by using the SWS Interactive Analysis (IA) at MPE Garching \footnote{We acknowledge support from the ISO Spectrometer Data Centre at MPE Garching, funded by DARA under grant 50 QI 9402 3}. The deglitching and averaging to equidistant spectral point in wavelength across the scans and detectors was done using the ISAP package. The ISOCAM imaging was performed on 31 July 1997 with the CAM camera in the mode AOT\,01 with the filter LW3, centered at 15 micron. The data were reduced by using CIA (a joint development by the ESA Astrophysics Division and the ISOCAM Consortium led by the ISOCAM PI, C. Cesarsky, Direction des Sciences de la Matiere, C.E.A., France; Cesarsky et al 1996). The object is still point-like at this angular resolution of 1.5 arcsec/pixel. The flux through the filter LW3 is 59.03 Jy as measured by aperture photometry method, about twice of the 12 $\mu$m flux (30.87\,Jy) given by the IRAS PSC catalogue. This difference could be caused by the strong emission in the mid--infrared range of the spectrum or by the object being variable. However, the flux ($\sim$ 39 Jy) at 15 micron of the SWS spectrum, which was taken at the same day as the ISOCAM image, does not match the photometric result from the ISOCAM image while it is in rough agreement with the 12 $\mu$m IRAS flux. By calculating the IRAS fluxes at 12 and 25 micron from SWS, correction factors of 1.11 and 1.06 should be applied to the corresponding SWS bands, respectively, to agree with the photometric data from IRAS PSC. After correction, the spectrum is smooth and the shape is similar to its IRAS LRS spectrum. In the same time, the IRAS LRS spectrum should be multiplied by a factor of 1.27 to agree with the IRAS PSC data. These factors are within the calibration uncertainties of the ISO--SWS and the IRAS LRS. Such agreement may show that the flux calibration of ISO--SWS is reliable, and that the object is non--variable (the variability index from IRAS observation is 50, at the border between variable and nonvariable indices). Therefore, we suspect that the discrepancy between ISO--SWS and ISO--CAM results is related to the uncertainties in the flux calibration from the image at LW3 band, but the reason for this is not clear to us. Anyway this ISO-CAM result is shown in Fig. 1 as an open circle. \section{Modeling} \indent \subsection{Spectral Energy Distribution} The overall spectral energy distribution of this object is observationally determined from the near infrared to the far infrared. The K band magnitude at 2.2 micron is 15.6 (Jiang et al. 1997), i.e., the flux is 3.7$\times$10$^{-4}$\,Jy. The fluxes at 12, 25, 60 and 100\,$\mu$m are 30.87, 43.44, 15.08 and 6.40\,Jy, respectively, from the IRAS photometric observations ( quality index = 3 in all four IRAS bands). Combination of these photometric results with the ISO SWS\,01 spectrum defines the observed spectral energy distribution. In Fig. 1, the observational data are shown where the dots represent the photometric results and the thin solid line shows the ISO SWS\,01 spectrum. The spectrum of ISO--SWS from 2.3 micron to 3.52 micron (band 1A, 1B and 1D of ISO--SWS) is not shown because it is too noisy. \begin{figure} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{h1307.f1}} \caption{Observed and modelled spectral energy distribution. The dot represents the near infrared and IRAS photometric results and the thin solid line is the ISO--SWS spectrum starting from band 1E. The flux at 15 micron measured from the ISO--CAM LW3 image is given by an open circle. The thick line is calculated from the model with the values of the parameters listed in Table 1 and the long--dashed line represents the radiation spectrum from the central star which radiates as a blackbody of temperature 2500\,K. } \end{figure} The model we used to fit the observational data is described in detail by Szczerba et al. (1997). In brief, the frequency--dependent radiative transfer equations are solved under the assumption of spherically--symmetric geometry simultaneously with the thermal balance equation for a dusty envelope. The radiation of the central star is assumed to be a blackbody. The mass loss rate is taken to be constant and the envelope is assumed to expand at the velocity derived from the CO\,(2--1) line observation. The dust opacity is represented by amorphous carbon grains of AC type (Bussoletti et al. 1987, Rouleau \& Martin 1991). The appropriate values of important parameters for fitting the observational data are listed in Table 1. The symbols have their usualmeanings, but more details can be found in Szczerba et al. (1997). Dynamical time t$_{dyn}$ means the time required for the matter to reach the outer radius of the envelope. \begin{table} \begin{flushleft} \label{tablines} \caption[]{Model parameters} \begin{tabular}{ll} \hline parameter & value \\ \hline T$_{eff}$ & 2500\,K \\ L$_{star}$ & 8000\,L$_{\odot}$ \\ d & 4.25\,kpc \\ T$_{inner}$ & 760\,K \\ R$_{inner}$ & 5.8 $\times$ 10$^{14}$cm \\ R$_{out}$ & 7.0 $\times$ 10$^{17}$cm \\ \.{M} & 8.0 $\times$ 10$^{-5}$ M$_{\odot}$/yr \\ t$_{dyn}$ & 1.6 $\times$ 10$^{4}$ yr\\ Dust--to--gas ratio & 5.0 $\times$ 10$^{-3}$ \\ a$_{-}$ & 0.005\,$\mu$m \\ a$_{+}$ & 0.25$\,\mu$m \\ p (power--law index of size dist.) & 3.5\\ \hline \end{tabular} \end{flushleft} \end{table} The result from model calculation with the values listed in Table 1 is plotted in Fig. 1 by a heavy solid line, while the long--dashed line represents the radiation from the central star, which radiates as a blackbody of effective temperature 2500\,K. The luminosity and distance of the star depend on each other, as well as the dust--to--gas mass ratio and mass loss rate assumed. We adopted the value of 8000 L$_{\odot}$ for the luminosity. The corresponding distance is 4.25 kpc. Since the outer radius R$_{out}$ is 7.0 $\times$ 10$^{17}$cm, the angular diameter of the circumstellar envelope is predicted to be 22 arcseconds at this distance. This size is big enough to be resolved by the ISOCAM at the resolution mode of 1.5$''$. However, the ISOCAM observations centered at 15\,$\mu$m are the most sensitive for the peak temperature of about 200\,K which, according to modeling results, corresponds to radius of 4.1$\times$10$^{15}$\,cm and reflects a much smaller angular size of about 0.13 arcseconds. So the result of ISOCAM that no extension is seen may be attributed to a large distance of the object. The value of R$_{out}$ is determined from the model--fitting to the observational results and its choice affects mainly the flux intensity in the mid-- and far--infrared. A reasoanable fit can be achieved with R$_{out}$ values ranging from 5.0 $\times$ 10$^{17}$cm to 9.0 $\times$ 10$^{17}$cm assuming a constant mass loss rate. For increasing mass loss rate (which means density distribution steeper than r$^{-2}$), the outer radius should be larger to compensate for the smaller far infrared emission of the outer enevelope layers, but range of the allowed changes in R$_{out}$ is quite similar. Note, however, such density distribution cannot be much different from that corresponding to the constant mass loss rate due to strong constraints from the 60 and 100 $\mu$m flux densities unless we assume the dust optical properties at far infrared wavelengths have a less steep slope than in the case of amorphous carbon used here. The adopted dust--to--gas mass ratio of 5.0$\times$ 10$^{-3}$ corresponds to mass loss rate of 8.0 $\times$ 10$^{-5}$ M$_{\odot}$/yr. This mass loss rate is higher than 5.8 $\times$ 10$^{-6}$ M$_{\odot}$/yr deduced from CO\,(1--0) line (Loup et al. 1993) or 1.4 $\times$ 10$^{-5}$ M$_{\odot}$/yr inferred from CO\,(2--1) line (Omont et al. 1993). Note, in addition, that to get velocity of the shell around 14\,km/s dynamical considerations would suggest even smaller value of dust--to--gas mass ratio (see Steffen et al. 1998), and in consequence a larger mass loss rate and a larger total mass of the envelope. However, the assumption of a density distribution slightly more steep than r$^{-2}$ would cancel the increase in total mass of the envelope. The calculation of the mass loss rate from the CO line suffers mainly from the uncertainty of the distance and mass fraction factor for CO molecules. For example, the mass loss rate of W Hya derived from infrared water lines lies a factor about 30 above the estimates based on the CO line observation (Neufeld et al. 1996). The case of Y CVn is similar in that the mass loss rate derived from interpretation of far infrared ISOPHOT images is 2 orders of magnitude higher than that found from the CO line (Izumiura et al. 1996). Izumiura et al.(1996) explained the result for Y CVn by suggesting that the far infrared and CO observations represent different epoches of mass loss. There may be another possibility that, the mass fraction of CO molecules is overestimated so that the mass loss rate is underestimated. The mass loss rate of IRAS\,03313$+$6058 from modeling lies a little above that estimated from the flux at 60 micron, 4.8 $\times$ 10$^{-5}$ M$_{\odot}$/yr (Omont et al. 1993). Derivation of mass loss rate from the flux at IRAS bands depends on the distance and on the bolometric correction factor which would induce some uncertainty. Though the mass loss rate derived from our modeling depends on the value of dust--to--gas ratio and density distribution, the result is relatively stable (probably better than a factor of two) because of the constraints required to fit the spectral energy distribution over the wide--range of wavelengths. The object then experiences quite strong wind, and has a circumstellar envelope perhaps as massive as 1.282\,M$_{\odot}$. By considering that the outer radius R$_{out}$ can vary in the range from 5.0 $\times$ 10$^{17}$ to 9.0 $\times$ 10$^{17}$cm, the mass of the circumstellar envelope may be in the range of 0.92\,M$_{\odot}$ to 1.65\,M$_{\odot}$ under the assumed dust--to--gas ratio of 0.005. Since the mass of the cirsumstellar envelope appears to be above one solar mass, the star could very possibly be an intermediate--mass AGB star. \subsection{The 30 micron emission} As can be seen in Fig. 1, the object shows emission around 30 micron which is superimposed on the continuum radiation from the star and the circumstellar envelope. The emission starts at 20 micron, peaks at about 30 micron and extends to longward of 40 micron. Because of another unidentified emission around 43.6 micron (see Discussion), it is difficult to define the cut--off position of the 30 micron emission band from this spectrum, though the emission may extend to the long--wavelength limit of ISO--SWS. As described in the Introduction, MgS is regarded as a reasonable candidate to be the carrier of the 30 micron emission. We tried to model this with the optical constants taken from the tables based on laboratory measurements of a MgS(90\%)--FeS(10\%) mixture (Begemann et al. 1994). For our computations, we used two shapes for the grains, i.e. the CDE (Continuous Distribution of Ellipsoids) and Mie theory (spherical grains). Assumptions and method used are described in detail by Szczerba et al. (1997). Because the temperature structure of the circumstellar envelope is determined from modeling of the spectral energy distribution, and because there is little difference in the dust temperature structure between the largest and smallest AC grains (a$_{-}$\,=\,0.005 and a$_{+}$\,=\,0.25\,$\mu$m) used in the calculations, the only free parameter after taking MgS into account is the number ratio between Sulfur atoms in MgS and total Hydrogen atoms n(S)/n(H). Under the CDE approximation a value of 3.0$\,\times$\,10$^{-6}$ for n(S)/n(H) gives a fit to the observed feature at the long--wavelength wing, though there is a little inadequacy in the short--wavelength wing of the band. On the other hand, the Mie theory (spherical) grains can make up the emission at the short--wavelength wing. This means that a combination of MgS grains shapes may account for the observed emission. In Fig. 2, the observation and modeling of this band are shown. \begin{figure} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{h1307.f2}} \caption{Fitting to the 30 micron emission. The thin solid line is the ISO--SWS spectrum and the thick line stands for the model fitting result to spectral energy distribution, the same as in Fig.1. The dot and dash lines represent the fitting by assuming the CDE and Mie approximation for the computation of optical properties of the MgS grains, respectively; n(S)/n(H) = 3.0\,$\times$\,10$^{-6}$. } \end{figure} \section{Discussion} \indent It is interesting to compare the 30 micron feature of this object with those of other AGB and post--AGB stars. Fig. 3 exhibits the spectrum of IRAS 03313$+$6058 together with the profiles of this feature for another AGB star, AFGL\,3068, and for the post--AGB object IRAS 22272$+$5435. The data for AFGL\,3068 are taken from Yamamura et al. (1998) while the data for the IRAS\,22272$+$5435 are from Omont et al. (1995). Because these two objects are much brighter than IRAS\,03313$+$6058, their spectra are scaled downward to agree at about 20 $\mu$m with the spectrum of IRAS\,03313$+$6058. Besides that IRAS\,22272$+$5435 has another feature at 21 micron, the most evident difference is that the emission of IRAS\,03313$+$6058 is much weaker than that of IRAS 22272$+$5435. On the other hand, AFGL\,3068 exhibits a similar strength of the 30 $\mu$m band as does IRAS\,03313$+$6058 which seems, however, to show more fine structures in this wavelength range. \begin{figure} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{h1307.f3}} \caption[short title]{Comparison of the 30 micron band in IRAS\,03313$+$6058 (thin solid line) with the post--AGB object IRAS\,22272$+$5435 (dotted line) and AGB star AFGL\,3068 (dashed line). The modeling continuum spectrum at this band to IRAS 03313$+$6058 is represented by the thick solid line which lies below the emission features. } \end{figure} The 30 micron emission for IRAS\,22272$+$5435 was previously modeled by wing of MgS grains (Szczerba et al. 1997). That fit resulted in an estimate of n(S)/n(H) = 4\,$\times$\,10$^{-6}$ for the highest dust temperature distribution and 1.6\,$\times$\,10$^{-5}$ for the lowest dust temperature distribution considered. The value of 3.0\,$\times$\,10$^{-6}$ for IRAS\,03313$+$6058 is very close to the case of highest dust temperature distribution of IRAS\,22272$+$5435. Then there is probably little difference in the sulfur abundance between these two objects and the strength of the 30 micron emission band may be influenced more strongly by other factors. The temperature may be one of the important factors through the way that lower temperature favors the excitation of this band. As the dust temperature of post--AGB envelopes is generally lower than that of AGB ones, this emission is stronger in post--AGB stars. For example IRAS\,22272$+$5425 has colder dust than IRAS\,03313$+$6058 and much stronger emission at this band. From laboratory experiment, MgS is one of the molecules which condenses at low temperature in the environment without oxides. Up to now, this band emission is detected in the C--rich AGB stars only with cold and high optical depth dust shells. This may indicate the existence of a critical temperature to form the 30 micron emission band carrier, and that such low temperatures together with approporiate chemical conditions are only possible in the extreme carbon stars. On the other hand, it is still not clear if carbon stars with smaller mass loss rates could form a carrier of this band. The existence of the 30 $\mu$m emission in AFGL\,2155 and IRC$+$40540 (Yamamura et al. 1998), neither of them was classified as extreme C--stars by Volk et al. (1992), suggests that formation of the approporiate chemical material is much more common and that higher envelope temperature in other carbon stars (not extreme ones) does not allow us to detect this emission feature. But it could be as well that some special chemical reactions responsible for the carrier formation of the 30 $\mu$m band are efficient only when temperature is enough low and/or the chemical composition is approporiate. Unfortunately, since without optical spectra the determination of the chemical abbundances is rather impossible other methods of investigation should be elaborated, and especially better statistics created by the ISO data could help solve the problem of the 30 $\mu$m carrier formation. Besides this feature around 30 micron, some other features, e.g. absorptions around 7.5 and 14 micron, emissions around 41 and 43.5 $\mu$m in the SWS spectrum of this object are not discussed here; they are currently under investigation. We note only that the absorptions are probably related to the C$_2$H$_2$ and/or HCN molecular bands, while emissions could be related to the crystaline silicates (especially as the enstatite mass absorption coefficient matches these emissions well - see J{\"a}ger et al. 1998). If crystalline silicate emissions are confirmed then it will allow to deduce an evolutionary status of IRAS\,03313$+$6058 which could bring some more information on the exciting transition phase between oxygen and carbon rich parts of the AGB evolution. \acknowledgements{B.W.J. thanks the people in N. Copernicus Astronomical Center, Torun for their help and support. We express also our gratitudes to Dr. Kevin Volk for his careful reading of the manuscript and useful suggestions. This work has been partly supported by grant 2.P03D.002.13 of the Polish State Committee for Scientific Research.}
\section{Introduction} Squeezed states of light are nonclassical \cite{Wal86}. The foremost consequence of this is that they can produce a homodyne photocurrent having a noise level below the shot-noise limit. The shot-noise limit is what is predicted by a theory in which the light is classical, with no noise, but the process of photo-electron emission is treated quantum-mechanically. There is, however, a simple way to produce a sub-shot-noise photocurrent without squeezed light: modulating the light incident on the photodetector by a current originating from that very detector. This was first observed \cite{WalJak85,MacYam86} around the same time as the first incontestable observation of squeezing \cite{Slu85}. The sub-shot noise spectrum of an in-loop photocurrent is not regarded as evidence for squeezing for a number of reasons. First, the two-time commutation relations for an in-loop field are not those of a free field \cite{Sha87}. This means that it is possible to reduce the fluctuations in the measured (amplitude) quadrature without increasing those in the other (phase) quadrature. Second, attempts to remove some of the supposedly low-noise light by a beam splitter yields only above shot-noise light, as verified experimentally \cite{WalJak85,YamImoMac86}. Because of these differences, the presence of a sub-shot noise photocurrent spectrum for in-loop light has by and large been omitted from discussions of squeezing \cite{Gar91,WalMil94,Bac98}. Nevertheless, it has been argued \cite{Tau95} that the in-loop field can justifiably be called ``sub-Poissonian'' (even if not squeezed), because a perfect quantum-non-demolition (QND) intensity meter for the in-loop light would register the same sub-shot-noise statistics as the (perfect) in-loop detector. Furthermore, it was proposed in Ref.\cite{Tau95} that the apparent in-loop noise reduction could be used to improve the signal to noise ratio for a measurement of a modulation in the coupling coefficient of such a QND intensity meter. Following on from Ref.~\cite{Tau95}, it has been shown that in-loop optical noise suppression may have other, more practical, applications. Buchler {\em et al.} showed \cite{Buc99} that such in-loop light can suppress radiation pressure noise in a gravitational wave detector by a factor of two. Even more strikingly, I recently showed \cite{Wis98b} that a two-level atom coupled to the in-loop field can exhibit linewidth narrowing exactly analogous to that produced by squeezed light \cite{Gar86}. In this paper I will discuss the properties of ``squashed'' states of light, as the in-loop analogues of squeezed states of light are called in Ref.~\cite{Buc99}. Section 3 covers the general theory of squashed states of light, including states which are squashed in both quadratures and states which are simultaneously squashed and squeezed. In Section 4, I generalize the analysis of Ref.~\cite{Wis98b} by considering the effects of these more general state of light on a two-level atom. But to begin, I review the properties of squeezed states of light in the following section. \section{Squeezed States} \subsection{Single-mode squeezing} The noise in the quadratures of a single-mode light field of annihilation operator $a$, \beq x = a+a^\dagger \;,\;\; y = -ia +ia^\dagger, \end{equation} is limited by the \hei uncertainty relation \beq V_{x} V_{y} \geq \left|\half[x,y]\right|^{2}=1, \end{equation} where $V$ denotes variance. Squeezed states of light are states such that one of the quadrature variances, say $V_{x}$, is less than one \cite{Wal86,Cav81,Yue76}. Clearly the other quadrature variance, say $V_{y}$, must be greater than one. If the equality is attained in the uncertainty relation then these are called minimum uncertainty squeezed states. The only other sort of minimum uncertainty state is the coherent state with $V_{x}=V_{y}=1$. \subsection{Continuum Squeezing} Single-mode squeezing was generalized early to multimode squeezing \cite{CavSch85}. In this work I wish to consider the limit of an infinite number of modes: the electromagnetic continuum. Considering polarized light propagating in one direction, only a single real-valued index is needed for the modes, and we take that to be the mode frequency $\omega = k$ (using units such that the speed of light is one). Then the continuum field operators $b(k)$ obey \beq [b(k),b^\dagger(k')] = \delta(k-k'). \end{equation} For light restricted in frequency to a relatively narrow bandwidth $B$ around a carrier frequency $\Omega$ it is possible to convert from the frequency domain to the time or distance domain by defining \beq b(t) = \int_{\Omega-B}^{\Omega+B} dk b(k)e^{-i(k-\Omega)t}. \end{equation} These obey \beq [b(t),b^\dagger(t')] = \delta(t-t'), \label{comrel2} \end{equation} where $b^\dagger(t) \equiv [b(t)]^\dagger$, and the $\delta$ function is actually a narrow function with width of order $B^{-1}$. The operator $b^\dagger(t)b(t)$ can be interpreted as the photon flux operator. Defining continuum quadrature operators \beq X(t)= b(t)+b^\dagger(t) \;,\;\; Y(t) = -ib(t)+ib^\dagger(t), \end{equation} one obtains \beq [X(t),Y(t')] = 2i\delta(t-t'). \end{equation} The singularity in the associated uncertainty relations can be avoided by quantifying the uncertainty by the {\em spectrum} \cite{WalMil94} \beq S^{X}(\omega) = \ip{\tilde{X}(\omega)X(0)}_{\rm ss} - \ip{\tilde{X}(\omega)}_{\rm ss}\ip{X(0)}_{\rm ss}. \end{equation} Then one can derive the finite uncertainty relations \cite{Sha87} \beq \label{uncrel1} S^{X}(\omega)S^{Y}(\omega) \geq 1. \end{equation} For very high frequencies the spectra always go to unity. This represents the shot-noise or vacuum noise level. However for finite frequencies it is possible to have for example $S_{X}(\omega)<1$. This indicates squeezing of the $X$ quadrature. The uncertainty in the conjugate quadrature would of course be increased. The quadrature operators $X(t),Y(t)$ can be directly measured (one at a time) using homodyne detection \cite{YueSha80}. For detection of efficiency $\epsilon \leq 1$ the photocurrent $I_{\rm hom}^{X}$ has the same statistics as (and therefore can be represented by) the operator \beq I_{\rm hom}^{X}(t) = \rt{\epsilon}X(t) + \rt{1-\epsilon}\xi_{X}(t), \end{equation} where $\xi_{X}(t)$ is a Gaussian white noise term. Here the normalization has been chosen so that the photocurrent spectrum \beq S^{X}_{\rm hom}(\omega) = \epsilon S^{X}(\omega) + (1-\epsilon), \end{equation} remains equal to unity (the shot noise limit) at high frequencies, where $S^{X}(\omega)=1$. \subsection{An Atom in a Squeezed Bath} \label{atomsec1} Now consider the situation where a two-level atom is immersed in a beam of squeezed light with annihilation operator $b_{0}(t)$. If the degree of mode matching of the squeezed light to the atom's dipole radiation mode is $\eta$, then the dipole coupling Hamiltonian in the rotating-wave approximation is \beq \label{vac} H_{0}(t) = -i[\rt\eta b_{0}(t) + \rt{1-\eta}\nu(t)]\sigma^\dagger(t) + {\rm H.c.} \end{equation} Here $\sigma$ is the atomic lowering operator and the atomic linewidth has been set to unity. The operator $b_0(t)$ represents the squeezed field with spectra $S^{X}_0(\omega)$ and $S^{Y}_0(\omega)$, and $\nu(t)$ represents the vacuum field interacting with the atom, satisfying $\ip{\nu(t)\nu^\dagger(t')}=\delta(t-t')$. Now if the quadrature spectra of the squeezed light are much broader than the atomic linewidth then one can make the white noise approximation that they are constant. For minimum-uncertainty squeezing, one has \beq S^{X}_{0}(\omega) = L = 1/S^{Y}_{0}(\omega), \end{equation} and the atom will obey the master equation \cite{Gar86} \beq \dot\rho = (1-\eta){\cal D}[\sigma]\rho + \frac{\eta}{4L} {\cal D}[(L+1)\sigma-(L-1)\sigma^\dagger]\rho, \end{equation} where ${\cal D}[A]B \equiv A^\dagger B A - \half \{A^\dagger A ,B\}$ as usual. This leads to the following atomic dynamics: \bqa {\rm Tr}[\dot\rho \sigma_{x}] &=& -\gamma_{x}{\rm Tr}[\rho \sigma_{x}], \label{onex}\\ {\rm Tr}[\dot\rho \sigma_{y}] &=& -\gamma_{y}{\rm Tr}[\rho \sigma_{y}],\\ {\rm Tr}[\dot\rho \sigma_{z}] &=& -\gamma_{z}{\rm Tr}[\rho \sigma_{z}]-C \label{threez}, \end{eqnarray} where \bqa \gamma_{x} &=& \half\left[ (1-\eta)+\eta L\right] \label{gx2},\\ \gamma_{y} &=& \half\left[ (1-\eta)+\eta L^{-1}\right], \label{gy2} \\ \gamma_{z} &=& \gamma_{x}+\gamma_{y} \;,\;\;C=1. \end{eqnarray} Note that for $L< 1$ the decay rate of the $x$ component of the atomic dipole is reduced below the vacuum level of $\half$, while the decay rate of the other component is increased. The reduction or increase in the decay rates are directly attributable to the reduction or increase in the fluctuations of the respective quadrature of the input continuum field. The prediction of this effect by Gardiner \cite{Gar86} began the study of quantum spectroscopy (that is, the interaction of nonclassical light with matter) \cite{FicDru97}. For sufficiently large $\eta$ this effect would be easily detectable experimentally in the fluorescence power spectrum of the atom (into the vacuum modes): \bqa \label{genps} P(\omega) &=& \frac{1-\eta}{2\pi}\ip{\tilde{\sigma}^\dagger(-\omega) \sigma(0)}_{\rm ss} \\ &=& \frac{(1-\eta)(\gamma_{z}-C)}{8\pi\gamma_{z}}\left[ \frac{\gamma_{x}}{\gamma_{x}^{2}+\omega^{2}} + \frac{\gamma_{y}}{\gamma_{y}^{2}+\omega^{2}}\right]. \end{eqnarray} For $L<1$ the spectrum consists of two Lorentzians, one with a sub-natural linewidth and one with a super-natural linewidth. The overall linewidth (defined as the full-width at half-maximum height) is reduced. This line-narrowing is only noticeable if the degree of mode matching $\eta$ of the squeezed light to the atom is significant, which is hard to do with a squeezed beam. One way around this is to make the atom strongly coupled to a microcavity, which can be driven by a squeezed beam. The microcavity enhances the atomic decay rate, but a squeezed input should suppress this enhancement in one quadrature. Unfortunately, experiments to date have failed to see this suppression, due to imperfections of various kinds \cite{Tur98}. \section{Theory of Squashed States} \subsection{Generation of Squashed States} Consider the feedback loop shown in Fig.~1. The field entering the modulator is $b_{0}(t)= \half[X_{0}(t)+iY_{0}(t)]$. The quadrature operators are assumed to have independent statistics defined by the spectra $S^{X}_{0}(\omega)$, $S^{Y}_{0}(\omega)$. The modulator simply adds a coherent amplitude to this field. There are various ways of achieving this, one of which is discussed in Ref.~\cite{Wis98b}. The field exiting is in any case given by \beq \label{b1} b_{1}(t) = \half\left[X_{0}(t)+iY_{0}(t) + \chi(t) + i\upsilon(t)\right], \end{equation} where $\chi(t),\upsilon(t)$ are real functions of time. This field now enters a homodyne detection device, set up so as to measure the $X$ quadrature. If the efficiency of the measurement is $\epsilon_{X}$ then the photocurrent is given by \beq \label{ihom} I_{\rm hom}^{X}(t) = \rt{\epsilon_{X}} \left[X_{0}(t)+ \chi(t) \right] + \rt{1-\epsilon_{X}}\xi_{X}(t). \end{equation} Now, through the feedback loop, this current may determine the classical field amplitudes $\chi,\upsilon$. Obviously in the case of measuring the $X$ quadrature, the only interesting results will come from controlling $\chi(t)$, and we set \beq \chi(t) = \int_{0}^{\infty} g_{X} h(s) I_{\rm hom}^{X}(t-\tau-s)/\rt{\epsilon_{X}}ds. \end{equation} Here $\tau$ is the minimum delay in the feedback loop, $h(s)$ is the feedback loop response function normalized to $\int_{0}^{\infty}h(s)ds = 1$ and $g_{X}$ is the round-loop gain \cite{Ste90}. Taking the fourier transform of this expression and substituting into Eqs.~(\ref{b1}) and (\ref{ihom}) yields \beq \label{b1p} \tilde{b}_{1}(\omega) = \half\left[ \frac{\tilde{X}_{0}(\omega)+ \rt{\theta_{X}}g_{X} e^{i\omega\tau}\tilde{h}(\omega) \tilde{\xi}_{X}(\omega)} {1-g_{X} e^{i\omega\tau}\tilde{h}(\omega)}+i\tilde{Y}_{0}(\omega) + i\upsilon(t)\right]. \end{equation} where \beq \theta_{X} \equiv \epsilon_{X}^{-1}-1. \end{equation} The $X$ quadrature spectrum of this light is \beq S_{1}^{X}(\omega) = \frac{S_{0}^{X}(\omega)+\theta_{X}g_{X}^{2}|\tilde{h}(\omega)|^{2}} {|1-g_{X} e^{i\omega\tau}\tilde{h}(\omega)|^{2}}. \end{equation} Say we are only interested in frequencies well inside the bandwidth of $\tilde{h}(\omega)$, much less than $\tau^{-1}$, and much less than the bandwidth of $S_{0}^{X}(\omega)$. Then we can replace $e^{i\omega\tau}\tilde{h}(\omega)$ by unity and $S_{0}^{X}(\omega)$ by a constant $L$. This gives \beq \label{conspec} S_{1}^{X} = \frac{L+g_{X}^{2}\theta_{X}}{(1-g_{X})^{2}} \geq \frac{L}{1+L/\theta_{X}}, \end{equation} where the minimum is achieved for negative feedback $g_{X} = -L/\theta_{X}$. Evidently this minimum is less than $L$. This means that even starting with shot-noise limited light ($L=1$) it is possible to produce sub-shot-noise light. However, it is important to note that this is not squeezed light in the ordinary sense. For example, it is impossible to remove any of the squeezed light by putting a beam splitter in the path of the in-loop beam. Under the above conditions the resulting out-of-loop beam actually has a noise level above the shot noise \cite{WalJak85,YamImoMac86,Sha87,Tau95}. Nevertheless, the fluctuations of the in-loop light do produce genuine physical effects in other circumstances, as investigated in Sec.~4. \subsection{Violation of the Uncertainty Relations} A curious point in the apparent squeezing of the $X$ quadrature is that the feedback has no effect on the $Y$ quadrature of $b_{1}$. From \erf{b1p}, $S_{1}^{Y} = L^{-1}$, assuming a minimum uncertainty input $b_{0}$ and $\upsilon=0$. Thus the uncertainty relation (\ref{uncrel1}) is violated for this in-loop light. For this reason, the in-loop light exhibiting sub-shot-noise fluctuations has been called ``squashed light'' \cite{Buc99}. By feedback, the noise in one quadrature can be squashed (reduced), but there is no ``squeezing'' of phase-space area into an increased noise in the other quadrature. The reason that the uncertainty relation (\ref{uncrel1}) is violated is that the commutation relations (\ref{comrel2}) are no longer valid for time differences $|t-t'| > \tau$, the minimum feedback loop delay. This is a direct consequence of the feedback loop. It is important to realize that the parts of the in-loop field separated in time by greater than $\tau$ never actually exist together. That is because the propagation time from the modulator to the detector is necessarily less than $\tau$. Thus the fundamental commutation relations \cite{Gar91} between parts of the field at different points in {\em space} at the same time are never violated. For freely propagating fields there is no real distinction between space and time separations, but for an in-loop field it is a crucial distinction. The temporal anticorrelations in the in-loop squeezed light only exist for time separations greater than $\tau$, and hence greater than the time for which any part of the in-loop light exists. There is never any anticorrelation between parts of the in-loop field in existence at any given time. By contrast, conventional squeezed light can propagate for an arbitrarily long time before detection, so the anticorrelations are between parts of the field which can exist simultaneously (even if they may not actually do so in a given experiment). \subsection{Simultaneous Squashing in Both Quadratures} It is interesting now to consider feedback in both the $X$ and $Y$ quadratures. Obviously one cannot simultaneously measure both of these quadratures with unit efficiency, but one can measure $Y$ with efficiency $\epsilon_{Y} \leq 1- \epsilon_{X}$. Carrying through the same sort of analysis as above shows that the $Y$ quadrature spectrum can be simultaneously reduced to \beq S_{1}^{Y} = \frac{L^{-1}}{1+L^{-1}/\theta_{Y}}. \end{equation} For the special case of $L=1$ (a vacuum input), one finds \beq S_{1}^{Y} + S_{1}^{X} = 2 - \epsilon_{X} - \epsilon_{Y} \geq 1. \end{equation} For feedback based on heterodyne detection (which is equivalent to homodyne detection on both quadratures with equal efficiency), one can have $S_{1}^{Y} = S_{1}^{X} = 1 - \half \epsilon$, which goes to one half in the limit of perfect detectors. \subsection{Simultaneous Squeezing and Squashing} \label{eezash} In the general case of $L \neq 1$, the sum of the quadrature spectra need not even be greater than one. Rather, for fixed $L<1$ and fixed $\epsilon = \epsilon_{X}+\epsilon_{Y}$ one finds, for $\epsilon_{X}=0$ and $\epsilon_{Y} = \epsilon$ \beq S_{1}^{Y} + S_{1}^{X} = L + \frac{\epsilon^{-1}-1}{L(\epsilon^{-1}-1) + 1} \geq 0, \end{equation} where the limit of zero noise in both quadratures is approached for $L\to 0$ and $\epsilon \to 1$. For example, with experimentally realisable parameters of 6dB squeezing \cite{Tur98} and detection efficiency $\epsilon = 0.95$ \cite{Sch96}, one could obtain $S_{1}^{X} = 0.25$ and $S_{1}^{Y} = 0.05$, giving $S_{1}^{Y}+S_{1}^{X} = 0.30$, compared to the limit of $2$ implied by the uncertainty relation (\ref{uncrel1}). \section{Application to Quantum Spectroscopy} Since the quadrature spectra calculated above apply to an in-loop field, which cannot be extracted using a beam splitter, it might seem that they have no physical significance. However, this is not the case. As I showed recently \cite{Wis98b}, placing a two-level atom in an squashed bath leads to linewidth narrowing of one atomic dipole quadrature, entirely analogous to that produced by a squeezed bath. The master equation is not identical, however, because the non-squashed quadrature is still shot-noise limited, so there is no line broadening of the other atomic dipole quadrature. In this work I generalize the results of Ref.~\cite{Wis98b} to include light which is simultaneously squashed in both quadratures, or simultaneously squeezed and squashed. If the atom is coupled to the in-loop beam $b_{1}(t)$ with mode-matching $\eta$ then the atomic Hamiltonian is \beq \label{totalHam} H(t)= -i[\rt\eta b_{1}(t) + \rt{1-\eta}\nu(t)]\sigma^\dagger(t) + {\rm H.c.} \end{equation} Following the methods of Ref.~\cite{Wis98b}, the expression for $b_{1}$ is modified from (\ref{b1p}) by the addition of the atom's radiated field in the direction of the beam, $\rt{\eta}\tilde{\sigma_{x}}(\omega)$, to the input operator $\tilde{X}_{0}(\omega)$. Thus the total Hamiltonian can be written \beq H(t)= H_{\rm fb}(t) + H_{0}(t), \end{equation} where $H_{0}(t)$ is as given in \erf{vac} and the Hamiltonian due to the feedback is \bqa H_{\rm fb}(t) &=& \lambda_{X} \half \sigma_{y}(t)\left\{ \sigma_{x}(t^{-}) + [X_{0}(t^{-})+\rt{\theta_{X}}\xi_{X}(t^{-})]\sqrt{\eta}\right\} \nn\\ && \,+ \lambda_{Y} \half\sigma_{x}(t)\left\{ \sigma_{y}(t^{-}) + [Y_{0}(t^{-})+\rt{\theta_{Y}}\xi_{Y}(t^{-})]\sqrt{\eta}\right\} .\label{Hfb} \end{eqnarray} The feedback parameters are defined as \beq \lambda_{X} = \frac{g_{X}\eta}{1-g_{X}}\;;\;\; \lambda_{Y} = \frac{g_{Y}\eta}{1-g_{Y}}. \end{equation} for the feedback of the homodyne current $I_{\rm hom}^{Y}$ are defined analogously to those from the feedback of $I_{\rm hom}^{X}$. In \erf{Hfb} the limit of broad-band feedback has been taken, with $\tilde{h}(\omega)e^{i\omega\tau}$ in \erf{b1p} set to unity. This Markov approximation is justified provided the bandwidth of the feedback is very large compared to the characteristic rates of response of the system \cite{Wis94a}. In the present context the rate of atomic decay is unity so we require, for instance, $\tau \ll 1$. For a typical electro-optic feedback loop with a bandwidth in the MHz range, the atom would have to have be metastable to satisfy this inequality. The precise requirements for the validity of the Markov approximation will be investigated in a future publication. Of course even in the broad-band limit the feedback from the measurement of a particular part of the field must act after that part of the field has interacted with the atom. This is the reason for the use of the time argument $t^{-}$ rather than $t$ in \erf{Hfb}. Now to describe the evolution generated by the total Hamiltonian (\ref{totalHam}), the theory of homodyne detection and feedback in the presence of white noise is required. This was first detailed in Ref.~\cite{WisMil94b}, generalizing the earlier work in Refs.~\cite{WisMil93b} and \cite{Wis94a}. The basic equation is \beq \label{cke} d\rho(t) = \ip{\exp[-iH_{\rm fb}(t)dt]\exp[-iH_{0}(t)dt] \rho(t)\exp[iH_{0}(t)dt]\exp[iH_{\rm fb}(t)dt]}-\rho(t). \end{equation} Here the ordering of the unitary operators has been chosen such that the time delay of the feedback has been taken into account and one can replace $t^{-}$ in \erf{Hfb} by $t$. In \erf{cke} the expectation value indicates an average over the bath operators $b_{0}(t)$ and $\nu(t)$, and the detector noise terms $\xi_{X}(t),\xi_{Y}(t)$. This is effected by making replacements such as \beq [X_{0}(t)dt]^{2} \to Ldt\;;\;\; \nu(t)dt\, \nu^\dagger(t)dt \to dt. \end{equation} The result is \bqa \dot{\rho} &=& (1-\eta){\cal D}[\sigma]\rho + \frac{\eta}{4L} {\cal D}[(L+1)\sigma-(L-1)\sigma^\dagger]\rho \nn \\ && -i{\lambda_{X}}\left[ \half\sigma_{y},\half\{(L+1)\sigma-(L-1)\sigma^\dagger\} \rho+\rho\half\{(L+1)\sigma^\dagger-(L-1)\sigma\}\right] \nn \\ && + i{\lambda_{Y}}\left[ \half\sigma_{x},-i\half\{(L^{-1}+1)\sigma+(L^{-1}-1)\sigma^\dagger\} \rho+i\rho\half\{(L^{-1}+1)\sigma^\dagger-(L^{-1}-1)\sigma\}\right] \nn \\ && + \frac{\lambda_{X}^{2}(L+\theta_{X})}{\eta}{\cal D}\left[\frac{\sigma_{y}}{2}\right]\rho + \frac{\lambda_{Y}^{2}(L^{-1}+\theta_{Y})}{\eta} {\cal D}\left[\frac{\sigma_{x}}{2}\right]\rho. \end{eqnarray} This again produces the atomic dynamics of Eqs.~(\ref{onex})--(\ref{threez}), but with \bqa \gamma_{x} &=& \half\left[ (1-\eta)+\eta L(1+\lambda_{X}/\eta)^{2}+\lambda_{X}^{2}\theta_{X}/\eta\right] = \half \left[ (1-\eta)+\eta S_{X}\right],\\ \gamma_{y} &=& \half\left[ (1-\eta)+\eta L^{-1}(1+\lambda_{Y}/\eta)^{2}+\lambda_{Y}^{2}\theta_{Y}/\eta\right] = \half \left[ (1-\eta)+\eta S_{Y}\right], \\ \gamma_{z} &=& \gamma_{x}+\gamma_{y} \;,\;\;C=1+\lambda_{X}+\lambda_{Y}. \end{eqnarray} In the above equations, the introduction of the spectrum $S^{X}$ is based on \erf{conspec}, with the identification $\lambda_{X} = \eta g_{X} /(1-g_{X})$, and similarly for $Y$. Note that the expressions for $\gamma_{x}$ and $\gamma_{y}$ depend upon the quadrature spectra (in the absence of the atom) in precisely the same way as those for pure squeezed (not squashed) light in Eqs.~(\ref{gx2}) and (\ref{gy2}). This suggests that the natural explanation for the change in the decay rates is again the reduced fluctuations of the input light. It seems that squashed fluctuations are much the same as squeezed fluctuations as far as the atom is concerned. The one difference is that the constant $C$ in the equation of motion for $\ip{\sigma_{z}}$ is also altered by the feedback. Consider now the case as in Sec.~\ref{eezash} where $L<1$, $\epsilon_{X}=0$ and $\epsilon_{Y}=\epsilon$. Then choosing $\lambda_{Y}=-\eta/(1+L\theta)$, so as to minimize the in-loop $Y$ quadrature spectrum, gives \bqa \gamma_{x} &=& \half\left[ (1-\eta)+\eta L\right] ,\\ \gamma_{y} &=& \half\left[ (1-\eta)+\eta{\theta}/({1+L\theta})\right], \\ \gamma_{z} &=& \gamma_{x}+\gamma_{y} \;,\;\;C=1 - {\eta}/({1+\theta L}), \end{eqnarray} where $\theta = \epsilon^{-1}-1$ as usual. Choosing $\epsilon=0.95$ and $L\approx 0.25$, as before, gives $\gamma_{z} \approx 1 - 0.85 \eta$. That is, in the limit of $\eta \to 1$, the rate of decay of the atomic population would be slowed by $85\%$. In the ideal limit of $L \to 0$, and $\epsilon,\eta \to 1$, the atom would be frozen in its initial state and would not decay at all. \section{Conclusion} In this work I have presented for the first time the theory for a new class of in-loop light, namely light which may be both squashed (in either or both quadratures) and squeezed. Squeezing here refers to conventional quantum noise reduction, whereas squashing refers to the noise reduction produced by the feedback loop. Even without a squeezed input it is possible to reduce the noise in {\em both} quadratures of the in-loop field below the shot-noise limit. With a squeezed input it is possible, in principle, to reduce the noise in both quadratures to zero (by squeezing one and squashing the other). I next derived the effect of this arbitrarily squeezed and squashed light on an in-loop atom. The calculated in-loop spectra precisely reflect the noise to which the atom responds, provided the bandwidth of the squeezing and the bandwidth of the feedback are much greater than the atomic linewidth. As the quantum fluctuations seen by the atom are reduced, the decay rates for the quadratures of the atom's dipole are reduced. In the limit that the atom is coupled only to in-loop light which is perfectly squeezed in one quadrature and perfectly squashed in the other, the atomic decay rates vanish and the atom's dynamics are frozen. As discussed above, the main experimental difficulty with seeing squeezing-induced line-narrowing is related to efficiently coupling the squeezed light onto the atom. Using squashed light rather than squeezed light would not overcome this difficulty. However, highly squashed light should be easier to generate than highly squeezed light, because it is limited only by the homodyne detector efficiency. Also, it can be produced at any frequency for which a coherent source and the appropriate electro-optic equipment is available. These factors, plus the intriguing possibility of observing simultaneous linewidth narrowing on both atomic quadratures, suggest an important role for squashed light in experimental quantum spectroscopy.
\section{INTRODUCTION} The next generation of X--ray instruments, such as AXAF or XMM, will provide deep X--ray images with very high source density (up to $1000/{\rm deg{^2}}$). To fully exploit the scientific content of these data, new and more refined detection techniques have to be considered. Algorithms based on the wavelet transform provide one of the best analysis tools, which has been already used in astronomy over the last decade (Coupinot et al. 1992; Slezak et al. 1993, 1994; Rosati 1995; Rosati et al. 1995; Grebenev et al. 1995; Damiani et al. 1997a; Vikhlinin et al. 1998). We have fully implemented a Wavelet Detection Algorithm (WDA) in order to meet the confidence requirements needed to deal with a large set of data (see the accompanying paper by Lazzati el at. 1999; Paper I hereafter). Here we focus on the application of this WDA to the X--ray images taken with the High Resolution Imager (HRI) on board the ROSAT satellite, detailing the HRI-specific features of the algorithm and presenting the first results of an on-going automatic analysis on all available data. Catalogs of X--ray sources with more than thousands of objects have been produced in the last few years (WGA: White, Giommi \& Angelini 1994; ROSATSRC: Zimmermann 1994; RASS: Voges et al. 1996; ASCASIS: Gotthelf \& White 1997; ROSAT Results Archive: Arida et al. 1998). These catalogs are mainly based on the Position Sensitive Proportional Counter (PSPC) on board the ROSAT satellite and have been heavily used in many research projects on different class of X--ray sources (e.g. Padovani \& Giommi 1996 on blazars; Fiore et al. 1998 on quasars; Angelini, Giommi \& White 1996 on X--ray variable sources; Israel 1996 on X--ray pulsars). The most appealing feature in using the ROSAT HRI data rather than the PSPC one is provided by the sharp core of the Point Spread Function (PSF), of the order of few arcseconds FWHM on-axis. This allows to detect and disentangle sources in very crowded fields and to detect extended emission on small angular sizes. Moreover, the search for counterparts at different wavelengths will be greatly simplified by the reduced error circles. On the other hand, the ROSAT HRI instrument has a very crude spectral resolution, thus a spectral analysis can not be carried out. It is less efficient than the PSPC by a factor of $\sim 4$ (3--8 for a plausible range of incident spectra) and, finally, it has a higher instrumental background. More details on the performances of the HRI are given in Section 2. In Section 3 we describe the application of the WDA to HRI data. Section 4 is devoted to the illustration of the simulations carried out to test the detection pipeline. In Section 5 we show the WDA results on ``difficult'' fields (e.g. very crowded fields, clusters of galaxies and of stars, supernova remnants). In Section 6 we summarise our conclusions. \section{HIGH RESOLUTION IMAGER} The HRI on board the ROSAT satellite is a position sensitive detector based on microchannel plates, that reveals single X--ray photons and determines their positions and arrival times (for more details see David et al. 1998). The ROSAT HRI is very similar to the Einstein HRI. The nominal pixel size of $0.5''$ has been reduced to $0.4986''$ after detailed observations on the Lockman hole field (Hasinger et al. 1998). The HRI field of view is given by the intersection of the circular detector and a square readout, resulting in octagon-like shape with $\sim 19'$ radius. The HRI PSF on-axis is $\sim 5''$ FWHM, well modeled with two Gaussians plus an exponential function. Due to random errors in the aspect solution however, images may be occasionally ellipsoidal and the PSF parameters may vary up to $\sim 15\%$. The HRI PSF degrades rapidly for off-axis angles beyond $\sim 5'$ whereas it becomes very asymmetric beyond $\sim 12'$. The PSF off-axis is not known with good accuracy and, up to now, only a description of the Gaussian widths with the off-axis angle has been published (David et al. 1998). Here we adopted a different approach. We make use of a ray-tracing simulator in order to extrapolate the well known on-axis PSF to any off-axis angle; in this way we get rid of aspect solution and photon statistic problems (see Appendix A for more detail). The on-axis effective area is 83 cm$^2$ at 1 keV, while the vignetting is less than 10\% within $10'$ at all energies. The effective area has not varied significantly since launch. Systematic uncertainties amount to $\lesssim 3\%$. The HRI covers an energy range of 0.1--2.4 keV, divided in 16 Pulse Height Analyzer (PHA), which provide very crude spectral information. Hardness ratios can give some qualitative information, but the gain variations lead to the definition of the PHA boundaries on a case by case basis (Prestwich et al. 1998). The HRI background is made of several components: the internal background due to the residual radioactivity of the detector (1--2 cts s$^{-1}$), the externally-induced background from charged particles (1--10 cts s$^{-1}$) and the X--ray background (0.5--2 cts s$^{-1}$). The background is the highest in the first few (1--3) PHA channels and it is dominated by the first two components. As shown by sky calibration sources, most of the source photons instead fall between PHA channel 3 and 8, approximately (David et al. 1998). \section{Wavelet Detection Algorithm for the HRI} The analysis and source detection of HRI images takes place in several steps, here we briefly describe the analysis of HRI images (see Fig.~\ref{block}). The detection algorithm is presented and described in detail in Paper I (see also Fig.~1 therein). \subsection{Image extraction} Due to computer limitations it is not efficient to analyse the entire HRI image in one single step, preserving the original angular resolution (it would result in a $\sim 5000\times 5000$ pixels image). In order to maintain the superb HRI angular resolution, we analyse the images in more steps. We extract a $512\times512$ pixels image rebinned by a factor of 1 (pixel size $0.5''$), as well as concentric images rebinned by a factor of 3 (pixel size $1.5''$), 6 (pixel size $3''$) and 10 (whole image with pixel size $5''$), respectively. These images are extracted from the relevant event files using the XIMAGE program (Giommi et al. 1992). Each image is then analysed separately. \subsection{Background estimate} For each image a mean background value is estimated using a $\sigma-$clipping algorithm (see Paper I): the mean and standard deviation of an image are calculated and pixels at more than $3\,\sigma$ above the mean value are discarded. The procedure works iteratively, rebinning the image and discarding outstanding pixels. We carried out various simulations on HRI fields and found that even in crowded fields the background value is recovered within $\sim 10\%$. We checked these values using the background estimator within XIMAGE obtaining the same results and accuracy. \subsection{Exposure map} X--ray mirrors, like optical mirrors, are vignetted. This generates non-flat fields where the detection of X--ray sources is made difficult by the second-order derivative of the background component (the adopted WT is insensitive to constant or first-order background components, see Paper I). In addition to this component, there may be obscuring structures (like the ROSAT PSPC rib) or hot regions in the detector (like in the EXOSAT Channel Multiplier Array) or real sky background variations (e.g. in the presence of extended sources). For these reasons it is usually better to perform the source search on top of a background map, rather than on a flat background. In the particular case of the ROSAT HRI, this allows us to search also for sources which lie in the edge region, where the detector efficiency rapidly drops to zero. In this case the linearity of the wavelet transform helps us: since an image can be thought as the sum of a background component plus the sources, the transform of the source component can be obtained subtracting the transform of the background map from that of the whole image. Background maps are provided together with the relevant data (as in the case of ROSAT PSPC) or can be generated through dedicated software (as the ESAS software for both ROSAT PSPC and HRI instruments; Snowden et al. 1994). A different approach is to properly smooth the source image, filtering out the brightest sources, and using this image as the background map (Vikhlinin et al. 1995b; Damiani et al. 1997a). This approach is also useful in the presence of extended emission which cannot be modeled analytically (e.g. supernova remnants) even if the smearing of extended and faint sources tends to reduce source significance and hence the completeness of the catalog. At variance with the ROSAT PSPC, the HRI background is dominated by the unvignetted particle background. In order to minimise the impact of this background and more generally to increase the signal-to-noise (S/N) ratio of X--ray sources, we restrict our analysis to PHA 2--9 (see also David et al. 1998; Hasinger et al. 1998). This range reduces the detector background by about 40\% with a minimum loss of cosmic X--ray photons ($\lesssim 10\%$; David et al. 1998). To build the exposure map we adopt the ESAS software (Snowden 1994) that makes use of the bright Earth and dark Earth data sets to produce a vignetted sky background map and a background detector map, respectively. The mean image background, as estimated with the $\sigma-$clipping algorithm (see above), is used to normalise the sky background map provided by the ESAS software. The total background map is then obtained by summing up the detector and sky maps. The ESAS maps are produced at rebin 10 (pixel scale $5''$), thus we interpolate them to obtain the maps at rebin 1, 3 and 6. These three maps are then smoothed with a Gaussian filter with a size twice the mean PSF, in order to get rid of interpolation inhomogeneities. \subsection{Image analysis} The four images and relative exposure maps are searched for significant enhancements using the WDA. On the first three images the detection is performed within an annulus of 255 pixels excluding a border at the detector edge of the local PSF width ($\sim 5''$, $\sim 7''$ and $\sim 12''$ at rebin 1, 3 and 6, respectively). This strategy has been adopted in order to preserve the original circular symmetry of the image and to avoid the occurrence of azimuthally dependent detection thresholds. In fact, detecting sources on the whole $512\times 512$ pixels image would result in a bias at the corners between 255 pixels (i.e. the image radius) and $255\,\sqrt{2}$ (i.e. half the square diagonal). Sources in the region left over are recovered at the successive rebin. At rebin 10 the whole image is handled. The detection threshold choice has deep impact on the characteristics of the catalog to be produced. The use for statistical purposes (such as $\log(N)-\log(S)$) of the catalog requires a high threshold (i.e. low contamination), whereas a low threshold is needed for the detection of a large number of sources, even if plagued by higher contamination. For these aims, we consider two different detection thresholds: a contamination of 0.1 spurious source per field is allowed for the Brera Multi-scale Wavelet Bright Source Catalog (BMW-BSC); a contamination of 1 spurious source per field is used for the Faint Source Catalog (BMW-FSC). The equivalent thresholds applied to BMW-BSC and BMW-FSC correspond to a source significance of $\sim 4.5\,\sigma$ and $\sim 3.5\,\sigma$, respectively. These detection thresholds are given for each field as a whole and must therefore be shared among the four rebinned images, holding the relative contamination constant over the full HRI field. The detection threshold for each rebinned image cannot be just one forth of the whole threshold (e.g. 0.25 in the case of 1 spurious source per field), due to the different area analysed at each rebin. A proper weighting factor is given by the ratio between the analysed area and the mean PSF width at each rebin. \subsection{Corrections} The HRI data are pre-processed with the ROSAT Standard Analysis System (Voges 1992) which provides images corrected for detector non-linearities and attitude control. We also apply the vignetting and gain correction, deriving them from the final exposure map. A deadtime correction in the case of bright sources is also applied (David et al. 1998). To these we have to add other small corrections related to the detection algorithm we use. To fit the sources in the wavelet space we approximate the PSF with a single Gaussian. However, the PSF has an extended tail and becomes increasingly asymmetric with the off-axis angle, thus we systematically loose counts. Since the HRI PSF is not well known at large off-axis angles, we performed ray-tracing simulations all over the field of view (see Appendix A). We then fit these ray-tracing simulated images with the WDA and derived the PSF correction needed to obtain the real flux (the PSF correction is a common characteristic of all detection algorithms). This PSF correction depends on the off-axis angles and varies from $\sim 18\%$ for sources at off-axis angles $\theta\lesssim 2'$ (i.e for all sources in the image at rebin 1) to $\sim 13\%$ for $6'\lesssim \theta\lesssim 13'$ (i.e. at rebin 6; see Table \ref{cts_off}). At rebin 10 the PSF degradation makes this correction a steep function of the off-axis angle (see Table \ref{cts_off}). The reduction of the PSF correction with the off-axis angle is likely due to the vanishing importance of the second Gaussian in the PSF. The HRI PSF becomes increasingly asymmetric with the off-axis angle, developing a bright spot (the center of which coincides with the source position) on top of an extended emission (as shown also by our ray-tracing simulations). The center of this extended emission is shifted by a few arcsec from the spot in the direction opposed to the field center. Fitting the counts distribution of an X--ray source with a Gaussian, the WDA finds its position in between the bright spot and the center of the extended emission (even if much nearer to the spot). In order to correct for this (small) effect which is mainly due to the larger support of the wavelet functions, we selected 4 HRI exposures with a large number of sources (Trapezium ROR 200500; P1905 ROR 200006; IC348 ROR 201674 and NGC2547 ROR 202298). We first performed a boresight correction on the central sources, then we selected X--ray sources associated with Guide Star Catalogue objects within $10''$. We measured the distance between the X--ray sources and the optical counterparts as a function of the off-axis angle. Even if a (small) number of spurious identifications can take place, we note that a systematic effect is clearly evident in Fig.~\ref{off}. In particular the radial shifts are linearly correlated with the off-axis angle and the best fit line provides a correction of about $7''$ for a source at $18'$ off-axis (this value includes also the $\sim 3''$ correction for the smaller pixel size, cf. Section 2). The scatter in the source off-sets is rather large so we decided to conservatively compute the $3\,\sigma$ uncertainty on the best fit line by individuating the region that includes the 99.7\% of sources (cf. Fig.~\ref{off}). This is achieved by considering the two dashed lines in Fig.~\ref{off}. The derived error is of about $7''$ for a source at $18'$ off-axis at a $3\,\sigma$ level. This shift error has been summed in quadrature to the position error derived from the fit and it provides the total error for the sources in our catalog. We point out that this result should not be regarded as a different HRI plate scale as reported in David et al. 1998 and Hasinger et al. 1998. The radial shift found in this analysis is mainly caused by the larger support of the wavelet functions that probe the source PSF on a larger scale than the sliding box techniques. \subsection{Creation of the catalog} For each observation we derive a catalog of sources with position, count rate, extension along with the relative errors, as well as ancillary information about the observation itself and source fitting. The count rate has been computed by fitting the image transform at all scales simultaneously in the wavelet space (see Paper I). We provide also a second count rate estimate following the standard approach of counting the source photons, however this method fails e.g. in the case of crowded fields or extended emission. The counts to flux conversion factor is determined based on a Crab spectrum. In the case of high latitude fields ($|b| > 30^\circ$) the galactic column density is assumed, whereas for lower latitude fields we consider either a null column density and the galactic value, therefore providing a range of fluxes. Together with the information relative to the X--ray data, we cross correlate the detected sources with databases at different wavelengths to give a first identification. One of the most interesting feature of the wavelet analysis is the possibility of characterising the source extension, however this cannot be assessed simply by comparing the source width ($\sigma$, as derived by the fitting procedure) with the HRI PSF (e.g. as derived by the ray-tracing simulator) at a given off-axis angle, due to the energy dependence of the PSF width as well as to errors in the aspect reconstruction (near on-axis). Thus, to assess the source extension, we considered a version of the catalog consisting of sources detected in the observations that have a star(s) as a target (ROR number beginning with 2). We considered 756 HRI fields and we detected 6013 sources in the BMW-BSC catalog (Fig.~\ref{extend}). The distribution of source extensions has been divided into bins of $1'$ each, as a function of the source off-axis angles. In each $1'$ bin, we applied a $\sigma-$clipping algorithm on the source extension: the mean and standard deviation in each bin are calculated and sources with widths at more than $3\,\sigma$ above the mean value are discarded. This method iteratively discards truly extended sources and provides the mean value of the source extension ($\sigma$) for each bin along with its error. We then determine the $3\,\sigma$ dispersion on the mean extension for each bin. The mean value plus the $3\,\sigma$ dispersion provide the line demarking source extension (cf. the dashed line in Fig.~\ref{extend}; see also Rosati et al. 1995). We conservatively classify a source as extended if its error on the extension parameter is such that it lays more than $2\,\sigma$ from this limit (see Fig.~\ref{extend}). Combining this threshold with the $3\,\sigma$ on the intrinsic dispersion we obtain a $\sim 4.5\,\sigma$ confidence level for the extension classification. 254 sources have been classified as extended, which makes up $\sim 4\%$ of the total. Note that no source has been classified as extended on-axis, as should be expected being stars the targets. A word of caution has to be spent for the flux estimate of extended sources. The source flux is computed by fitting a Gaussian to the surface brightness profile and, in many cases, this provides a poor approximation. Therefore, fluxes of extended sources are usually underestimated. A solution in the case of extended sources with well-defined surface brightness profiles (i.e. clusters of galaxies) has been presented by Vikhlinin et al. 1998. \medskip \section{SIMULATIONS} The WDA presented in paper I has provided very good results on ideal fields, with flat background, Gaussian sources and no crowding. To test the whole pipeline also on realistic images, we simulated sets of HRI observations using real instrumental background maps generated with the Snowden's software and superposing X--ray sources following the soft X--ray $\log(N)-\log(S)$ by Hasinger et al. 1993 (see also Vikhlinin et al. 1995b). We took the exposure map of one of the longest observations in the ROSAT public archive (NGC 6633 - ROR 202056a01 - $\sim 120$ ks), and superpose X--ray sources down to a flux of $\sim 5\times 10^{-16}\rm \ erg \, s^{-1}\rm \ cm^{-2}$ (a conversion factor of $1\,{\rm cts\,s^{-1}} = 1.71\times 10^{-11}\rm \ erg \, s^{-1}\rm \ cm^{-2}$ in the 0.5--2.0 keV energy band has been adopted; Hasinger et al. 1998). Each simulated image contains 500 sources, the faintest of which have $\sim 3$ counts and enhance the sky background being well below the detection threshold. The sources were distributed homogeneously all over the field of view and the appropriate PSF obtained with the ray-tracing simulator was used to spread their photons. A total of 100 fields were simulated and analysed in the same automatic fashion as the real data. Every detected output source has been identified with a simulated input source within the $3\,\sigma$ error box. These simulations allow to probe the WDA algorithm behaviour on ``real'' HRI fields, revealing the presence and the influence of biases and selection effects (e.g. Hasinger et al. 1993; Vikhlinin et al. 1995a). The great majority (more than 90\%) of source fluxes are recovered within a factor of 2 of their input values. The tail of the counts distribution starts enlarging at $\sim 120$ counts over the entire field of view (i.e. $\sim 2\times 10^{-14}\rm \ erg \, s^{-1}\rm \ cm^{-2}$, see Fig.~\ref{cts}). This effect is produced by the combination of source confusion (e.g. Hasinger et al. 1998) and the bias in the source intensity determination extensively discussed by Vikhlinin et al. 1995a. The latter occurs due to the preferential selection of sources coincident with positive background fluctuations, near the detection threshold and it is more severe for surveys with low S/N ratios. Comparing the results on the simulated HRI fields with the ones with equally spaced sources discussed in Paper I (cf. Figure 5), we conclude that source confusion is more important. We remark that source confusion affects only a small fraction of the sources ($\lower.5ex\hbox{\ltsima} 10\%$). If, in fact, we quantify the confusion following Hasinger et al. 1998 (cf. Equation 5 therein), we are far below the strong confusion regime. In order to explore the influence of biases and selection effects we simulated also 50 fields at five different exposure times (1, 7, 15, 30 and 60 ks) each. Rather than adopting the Hasinger's $\log(N)-\log(S)$ distribution (which implies about 3 sources per field at 7 ks), we considered in these cases a simpler one, i.e. a power law $\log(N)-\log(S)$ with an index of --2.5 and with arbitrary normalisation set to have about 200 sources per frame. This approach has been adopted in order to reduce the number of simulations and therefore computing time (which is mainly spent in the calculation of the image and background transforms) and it results in a more stringent test on the WDA due to the heavier crowding. The relative completeness functions are plotted in Fig.~\ref{compl}. As one can see the 95\% limiting flux moves from $\sim 4\times 10^{-13}\rm \ erg \, s^{-1}\rm \ cm^{-2}$ to $\sim 3\times 10^{-14}\rm \ erg \, s^{-1}\rm \ cm^{-2}$ (see Fig.~\ref{compl_fit}). These numbers refer to a completeness achieved over the full field of view. Lower values can be obtained reducing the area of interest. The total number of counts needed to achieve the 95\% completeness level as a function of the exposure time can be well approximated by a constant plus a square root function (see Fig. \ref{compl}). \subsection{General properties of the survey} The sensitivity of the HRI instrument is not uniform over the field of view but decreases rapidly with the off-axis angle as a consequence of the PSF broadening and mirror vignetting. For this reason the surveyed region at a given limiting flux should not necessarily coincide with the HRI detector area but is in general a smaller circular area. We compute the sky coverage of a single observation by calculating the detection thresholds over the entire field of view. With the help of the ray-tracing program we simulated sources from $0'$ to $18'$ off-axis with a $1'$ step. We then properly rebin these simulated images and calculate the minimum number of counts needed to reveal the source at the selected off-axis angle as a function of the image background. A sky survey of $\sim 400$ square degrees is expected down to a limiting flux of $\sim 10^{-13}\rm \ erg \, s^{-1}\rm \ cm^{-2}$, and of 0.3 square degrees down to $\sim 3\times 10^{-15}\rm \ erg \, s^{-1}\rm \ cm^{-2}$. A preliminary and conservative estimate based on the extragalactic soft X--ray $\log(N)-\log(S)$ indicates that about 16000 sources will be revealed in the complete analysis of the whole set of HRI data (BMW-BSC). In Fig.~\ref{sky} we show the differential and integral distributions of exposure times and galactic absorptions. We also computed the $\log(N)-\log(S)$ distribution based on the dataset of 120 ks simulations (see above). The recovered distribution is complete down to a flux of $\sim 10^{-14}\rm \ erg \, s^{-1}\rm \ cm^{-2}$ over the entire ($18'$) field of view. The knowledge of the sky coverage allows us to correct for the loss of sources, enabling us to recover input $\log(N)-\log(S)$ down to a flux of $\sim 3\times 10^{-15}\rm \ erg \, s^{-1}\rm \ cm^{-2}$ (Fig.~\ref{lognlogs}). \section{FIRST RESULTS ON SELECTED HRI FIELDS} To test our HRI pipeline, we examined a sample of HRI fields (see Table \ref{fields}) selected for their ``difficulty'' in terms of source confusion and extended emission, for which sliding box techniques face serious problems. We compare the results obtained with our WDA with the detections made with XIMAGE/XANADU (Giommi et al. 1992) and EXSAS/MIDAS (Zimmermann et al. 1993). We have to point out that the detection algorithm in XIMAGE is not optimised for the HRI so that, especially at large off-axis angles where the PSF degrades, strong sources are usually detected as multiple; on the other hand EXSAS/MIDAS has been specifically developed to deal with the ROSAT data. We ran our WDA on these fields automatically without particular settings. For XIMAGE we set the probability threshold to $7\times 10^{-6}$ and for EXSAS the maximum likelihood (ML) to 8 but kept sources with a final ML$\,\ge 12$, which correspond to a statistical significance of $\sim 4.5\,\sigma$. In Table \ref{fields} we report the number of sources detected with the different algorithms as well as the number of sources for each rebin image (in the case of WDA we report the number of sources for each image, rather than in a circle as described above, in order to allow the comparison). As one can see the WDA is more efficient in detecting faint sources. Even for fields where the number of sources detected with the other two algorithms is higher than for the WDA, a closer inspection reveals that this is typically due to the presence of a strong or extended source, that has been splitted into multiple sources. To better compare the algorithms we plot in Fig.~\ref{m31} the inner part of the M31 and Trapezium images: of the 59 sources detected (21 in M31 and 38 in Trapezium) at rebin 1 by our WDA, 35 (13+22) are in common with the other two algorithms, 11 (3+8) are found by WDA and EXSAS and 10 (4+6) by WDA and XIMAGE. Two sources (0+2) are found by XIMAGE and EXSAS and not by the WDA. 3 (1+2) sources are found by WDA only, 9 (5+4) by EXSAS only and 12 (0+12) by XIMAGE only. If we retain sources detected by at least two algorithms as ``real'' and sources with only one detection as ``non-real'', we have that the WDA is characterised by the smallest number of ``missed'' and ``spurious'' sources. \section{CONCLUSIONS} The general theory underlying the WT-based algorithm we developed has been described in Paper I, together with the extensive testing we carried out. Here we focused on the application of this WDA to ROSAT HRI images, describing the HRI-dependent features, the major problems encountered (e.g. the sharp drop in the background at the detector edges and the PSF broadening with the off-axis angle) as well as the extensive testing on simulated images. The use of the Snowden's background maps and the modeling of the HRI PSF with a ray-tracing simulator allowed us to overcome them and to perform the source search over the entire HRI field of view. In particular, we were able to optimise our WDA such that more than 90\% of sources have output fluxes within a factor of two of their input values in the deepest images we simulated (120 ks). The use of a wavelet-based algorithm allows to flag extended sources in a complete catalog of X--ray sources. The completeness functions for different exposure times have been computed with simulated images. For an exposure time of 120 ks we reach a completeness level of 95\% at a limiting flux of $\sim 10^{-14}\rm \ erg \, s^{-1}\rm \ cm^{-2}$ over the full field of view. Correcting for the sky coverage the $\log(N) - \log(S)$ distribution can be extended down to $\sim 3\times 10^{-15}\rm \ erg \, s^{-1}\rm \ cm^{-2}$ over the entire field of view. The analysis with the WDA of the large set of HRI data will allow a sky survey of $\sim 400$ square degrees down to a limiting flux of $\sim 10^{-13}\rm \ erg \, s^{-1}\rm \ cm^{-2}$, and of $\sim 0.3$ square degrees down to $\sim 3\times 10^{-15}\rm \ erg \, s^{-1}\rm \ cm^{-2}$. A conservative estimate based on the extragalactic $\log(N)-\log(S)$ indicates that at least 16000 sources will be revealed in the complete analysis of the whole set of HRI data (BMW-BSC). The WDA we developed is also tested on difficult fields and it compares favorably with other detection algorithms such as XIMAGE and EXSAS, both for what concerns the sensitivity to blended and/or weak sources and the reliability of the detected sources. A complete and public catalog will be the outcome of our analysis: it will consist of a BMW-BSC with sources detected with a significance $\gtrsim 4.5\,\sigma$ and a fainter BMW-FSC with sources at $\gtrsim 3.5\,\sigma$. All the detected HRI sources will be characterised in flux, size and position and will be cross-correlated with other catalogs at different wavelengths (e.g. Guide Star Catalog, NRAO/VLA Sky Survey etc.), providing a first identification. The BSC can and will be used for systematic studies on different class of sources as well as for statistical studies on source number counts. These images and related information would be available through a multi-wavelength Interactive Archive via WWW developed in collaboration with BeppoSAX-Science Data Center, Palermo and Rome Observatories. The layout of a typical image is displayed in Fig.~\ref{tra_tot}. The WDA used for the analysis of HRI sources can be adapted for future X--ray missions, such as JET-X, XMM and AXAF. The application of wavelet-based detection algorithms to these new generation of X--ray missions will provide an accurate, fast and user friendly source detection software. \acknowledgments{ We thank P. Conconi for his help with the ray-tracing simulator. We acknowledge S. Snowden for providing a beta version of the ESAS software in advance of release. We are grateful to S. De Grandi, P. Giommi and P. Rosati for useful comments. This research has made use of data obtained through the High Energy Astrophysics Science Archive Research Center (HEASARC) provided by NASA's Goddard Space Flight Center and through the ROSAT data archive maintained at the Max Planck Institute f\"ur Extraterrestrische Physik. This work was supported through CNAA and ASI grants.}
\section{Selected Results from Axiomatic Vertex Algebras}\label{a:axiom_facts} In this appendix we gather a few facts about ordinary axiomatic vertex algebras which we have used throughout this paper. For more details see \cite{kac}. Let $(V, Y(\cdot,x)\cdot, T, \mbox{$|0\rangle$})$ be a vertex algebra. This first lemma is a consequence of the vacuum axioms: \begin{lemma} For any $a \in V$, $Y(a,x)\mbox{$|0\rangle$} = \sum_{i\geq 0} \frac{T^i}{i!}a x^i = e^{xT}a.$ \end{lemma} \begin{proof} Since $Y(a,x)\mbox{$|0\rangle$}$ can be evaluated at zero, we know that it has no singularities. By the first vacuum axiom we know that $T\mbox{$|0\rangle$} =0$, so the translation covariance axiom says that for any $n \geq 0$, $(\partial_x)^n Y(a,x)\mbox{$|0\rangle$} = \frac{T^n}{n!}Y(a,x)\mbox{$|0\rangle$}$. Letting $Y(a,x)\mbox{$|0\rangle$} = \sum_{i\geq 0} a_ix^i$, and evaluating this at zero we have $a_n = \frac{T^n}{n!}a$ and the lemma is proved. \end{proof} The next lemma uses this result, and is a consequence of the locality axiom. \begin{lemma}[Quasisymmetry] For any $a,b \in V$ we have $Y(a,x)b = e^{xT}Y(b,-x)a.$ \end{lemma} \begin{proof} Recall that the translation covariance axiom says that $Y(a,x)T = (T-\partial_x)Y(a,x)$. So we have \begin{eqnarray*} Y(b,y)e^{xT}a &=& \sum_{i\geq 0}Y(b,y)\frac{T^i}{i!}a\\ &=& \sum_{p,q\geq 0}(-1)^q\frac{T^p}{p!}\frac{\partial_y^q}{q!}x^{p+q} Y(b,y)a \\ &=& \sum_{p\geq 0}\frac{T^p}{p!}x^p Y(b,y-x)a \\ &=& e^{xT}Y(b,y-x)a. \end{eqnarray*} From the locality axiom we know that for some $N \gg 0$ the following holds \[(x-y)^{N}Y(a,x)Y(b,y)\mbox{$|0\rangle$} = (x-y)^{N}Y(b,y)Y(a,x)\mbox{$|0\rangle$}. \] Combining this with the previous lemma, it just says that \begin{eqnarray*} (x-y)^{N}Y(a,x)e^{yT}b &=& (x-y)^{N}Y(b,y)e^{xT}a\\ &=& (x-y)^{N}e^{xT}Y(b,y-x)a. \end{eqnarray*} Setting $y$ equal to zero we have $x^{N}Y(a,x)b = x^{N}e^{xT}Y(b,-x)a$, and since this is an equality of Laurent series, we may divide by $x^{N}$ to give our result. \end{proof} \section{Borcherds' Singular maps parameterised by trees}\label{ss:bor_sing_maps} This appendix provides a review of the singular multilinear maps defined in \cite{bor}. In that paper, the definitions which follow were intended only for trees of constant height (i.e., trees whose root is separated from each leaf by the same number of edges), which are equivalent to the \defn{sieves} defined in that paper. Here we extend the definition of that paper to all trees. Our definition for multilinear singular maps uses these singular maps as a base over which we pullback to get a more specific collection of multilinear singular maps. \begin{dfn} If $p$ is a tree with $n$ leaves and height $d$, then we define the \defn{space of singular functions of type $p$ from $\G{n}$ to $B$}, written as $\ensuremath{\mathit{Fun}}_p(\G{n},B)$, recursively on height. If the height of $p$ is one, then $p$ is the flat tree with zero internal vertices, and we define $\ensuremath{\mathit{Fun}}_{p}(\G{n},B)$ to be $\ensuremath{\mathit{Fun}}(\G{n},B)$ as in definition \ref{d:fun}. If $d > 1$, then let $q$ be the tree of height $d-1$ with $m$ leaves, formed by removing the $s$ edges which are separated from the root by $d-1$ edges. We define $\ensuremath{\mathit{Fun}}_{p}(\G{n},B)$ as follows: \begin{enumerate} \item\label{step1}Begin by forming the collection of $\G{m}$\mbox{-} invariant maps $\ensuremath{\mathrm{Hom}}_{\G{m}}(\G{s},\ensuremath{\mathit{Fun}}_{q}(\G{m},B))$. We make $\G{s}$ into a $\G{m}$\mbox{-} module by first noticing that each leaf $0 \leq l \leq m$ of $q$ is joined to $k_{l}$ edges of $p$. So the $l$th entry of $\G{m}$ acts diagonally on the corresponding $k_{l}$ entries of $\G{n}$. \item Localise $\ensuremath{\mathrm{Hom}}_{\G{m}}(\G{n},\ensuremath{\mathit{Fun}}_{q}(\G{m},B))$ at all pairs $(i,j)$ of the $k_{l}$ entries of $\G{n}$ which are joined to each leaf $1 \leq l \leq m$. This collection of singular maps is $\ensuremath{\mathit{Fun}}_{p}(\G{n},B)$. \end{enumerate} \end{dfn} \begin{note} This difference between this definition and the one given in \cite{bor} is that in step \ref{step1} above, we require the maps from $\G{n}$ to $\ensuremath{\mathit{Fun}}_q(\G{m},B)$ to be $G$\mbox{-} invariant at each leaf of $q$. We will see in the following examples that this definition gives the results put forward in that paper. It is our belief that this assumption of $\G{m}$\mbox{-} invariance is presumed in the definition given in \cite{bor}. \end{note} The definition is made clear by working out a number of examples. \begin{example} Let $G$ be a vertex group and $B$ be a $G$\mbox{-} module. We begin by looking at the space of singular functions from $G$ to $B$, parameterised by non-branching trees. For the tree, $p_1$ consisting of just one edge, we know that $\ensuremath{\mathit{Fun}}_{p_1}(G,B) = \ensuremath{\mathit{Fun}}(G,B) \cong \ensuremath{\mathrm{Hom}}(G,B)$. Using the given recursion definition, if we let $p_i$ be the tree of height $i$, then $\ensuremath{\mathit{Fun}}_{p_2}(G,B) = \ensuremath{\mathrm{Hom}}_G(G,\ensuremath{\mathrm{Hom}}(G,B)) \cong \ensuremath{\mathrm{Hom}}(G,B)$, and so for any $i \geq 1$, $\ensuremath{\mathit{Fun}}_{p_i}(G,B) \cong \ensuremath{\mathrm{Hom}}(G,B)$. \end{example} \begin{example} Similarly, if we have an $n$\mbox{-} leafed tree $p$, and if $q$ is the tree formed from pasting the root of $p$ onto the end of the tree, \epsfig{file=Images/edge.eps,height=3mm}, then $\ensuremath{\mathit{Fun}}_{q}(\G{n},B)$ is the module $\ensuremath{\mathrm{Hom}}_G(\G{n},\ensuremath{\mathrm{Hom}}(G,B))$ localised at all pairs $(i,j)$. But we know $\ensuremath{\mathrm{Hom}}_G(\G{n},\ensuremath{\mathrm{Hom}}(G,B)) \cong \ensuremath{\mathrm{Hom}}_R(\G{n},B)$, and so $\ensuremath{\mathit{Fun}}_{q}(\G{n},B) \cong \ensuremath{\mathit{Fun}}_{p}(\G{n},B)$. \end{example} \begin{example} Let $t_3$ be the unique 3\mbox{-} leafed tree of height 1. We saw in example \ref{ex:G3} that for the classical vertex group, \begin{eqnarray*} \ensuremath{\mathit{Fun}}_{t_3}(\G{3},B) &\cong & \ensuremath{\mathit{Fun}}(\G{3},B) \\ &\cong & B\ps{x_1,x_2,x_3}[(x_1-x_2)^{-1},(x_1-x_3)^{-1}(x_2-x_3)^{-1}]. \end{eqnarray*} \noindent If we let the following tree be denoted $l$: \[\flatthreeleafedleft{}{}{}{}\] then we have that $\ensuremath{\mathit{Fun}}_{l}(\G{3},B)$ is the following module localised at $(1,2)$: \[\ensuremath{\mathrm{Hom}}_{\G{2}}(\G{3},Fun_{t_2}(\G{2},B)) \cong \ensuremath{\mathrm{Hom}}_{\G{2}}(\G{3},B\ps{z_1,z_2}[(z_1-z_2)^{-1}]).\] Localising, we can identify this module with a subset of the collection of formal power series \begin{equation}\label{e:3leafed_labeled} B\ps{z_1,z_2}[(z_1-z_2)^{-1}]\ps{y_1,y_2,y_3} [(y_1-y_2)^{-1}]. \end{equation} The $\G{2}$\mbox{-} invariance can be realised as the restriction to power series which satisfy $\partial_{z_1} = \partial_{y_1}+\partial_{y_2}$ and $\partial_{z_2} = \partial_{y_3}$. We represent this module pictorially as the labelled tree, \[\flatthreeleafedleftsides{y_1}{y_2}{y_3}{z_1}{z_2}{B}\] where at each internal node we have equalised over an action of $G$ on the incoming and outgoing edges. These differential equations give conditions which allow us to rewrite this power series module as a power series of only three variables. Either of the changes of variables given in equations \eqref{e:varchange1}-\eqref{e:varchange3} work, and so we may made the identification \begin{eqnarray*} \ensuremath{\mathit{Fun}}_l(\G{3},B) &\cong& B\ps{X_1,X_3}[(X_1-X_3)^{-1}] \ps{(X_1-X_2)}[(X_1-X_2)^{-1}]\\ &\cong& B\ps{X_1,X_3}[(X_1-X_3)^{-1}]\ps{X_2}[(X_1-X_2)^{-1}]. \end{eqnarray*} \end{example} \section{Classical Vertex Group and its Representations} \label{s:classical_vg} In the introduction we concentrated on examining the products of vertex operators. We purposely avoided the question of what types of homomorphisms we were considering, because we were mainly concerned with demonstrating the dependence of singularities on products of vertex operators. In this section we return to the vertex operator itself, examining it closely and paying special attention to its interaction with the infinitesimal translation operator. Before beginning we recall the definition of a vertex algebra (following the definition in \cite{kac}). \begin{dfn}\label{d:vetex_algebra} A \defn{vertex algebra} consists of a complex vector space $V$ (the \defn{state space}) together with an endomorphism, $T:V \rightarrow V$ (the \defn{infinitesimal translation operator}), a distinguished vector denoted $\mbox{$|0\rangle$} \in V$ (the \defn{vacuum vector}), and a linear map (the \defn{vertex operator}): \[Y(\cdot, x)\cdot:V \mbox{$\otimes$} V \rightarrow V\ps{x}[x^{-1}],\] taking any $a\mbox{$\otimes$} b \in V\mbox{$\otimes$} V$ to $Y(a,x)b \in V\ps{x}$. The axioms for a vertex algebra say that \begin{description} \item[Vacuum axioms:] For any $a \in V$, the vacuum satisfies: \begin{eqnarray} T\mbox{$|0\rangle$} &=& 0 \label{e:vac_triv_axiom}\\ Y(\mbox{$|0\rangle$}, x)a &=& a\label{e:vac_id_axiom} \\ Y(a,x)\mbox{$|0\rangle$}|_{x=0} &=& a\label{e:vac_ps_axiom}. \end{eqnarray} \item[Translation covariance axiom:] The infinitesimal translation operator interacts with a vertex operator according to: \begin{equation} [T,Y(a,x)] = \partial_x Y(a,x):V \rightarrow V\ps{x}[x^{-1}].\label{e:inv_axiom} \end{equation} \item[Locality axiom:] And, for any $a, b \in V$, there exists some $N \gg 0$ such that the following holds: \begin{equation}\label{e:locality} (x-y)^N[Y(a,x), Y(b, y)] = 0. \end{equation} \end{description} \end{dfn} If we consider the free group ring generated by the infinitesimal translation operator, denoted $G$, then $V$ is a $G$\mbox{-} module simply by virtue of the fact that $T$ is an endomorphism of $V$. From a simple vertex algebra calculation, one can show that $\partial_x Y(a,x) = Y(Ta,x)$ (see \cite[Prop. 4.8]{kac}). So, from the translation covariance axiom, we see that given any $b \in V$, we have an equality of Laurent series, \[Y(a,x)Tb + Y(Ta,x)b = TY(a,x)b.\] Now, considering the vertex operator as a map from $V\mbox{$\otimes$} V$ to $V\ps{x}[x^{-1}]$, this presentation of the translation covariance axiom just says that a vertex operator is a $G$\mbox{-} invariant map, where $G$ acts on $V\mbox{$\otimes$} V$ by \begin{eqnarray*} G \mbox{$\otimes$} (V \mbox{$\otimes$} V) &\longrightarrow &V \mbox{$\otimes$} V\\ T\mbox{$\otimes$} a \mbox{$\otimes$} b &\longmapsto &Ta \mbox{$\otimes$} b + a \mbox{$\otimes$} Tb. \end{eqnarray*} From this account, it is clear that a general vertex algebra is going to be some type of module. In the next section we will define a vertex group, which will be the algebra over which we will want to work. We show that $G$ is a nontrivial example of a vertex group. We then move on to consider modules over a vertex group, and modules over $G$ in particular. We finish this section by considering ways of expressing maps between $G$\mbox{-} modules and we show how holomorphic vertex algebras arise naturally from this presentation. \subsection{Definition of a Vertex Group} This section reviews some important definitions from \cite{bor}. In particular, we define an elementary vertex structure on a cocommutative Hopf algebra which we use to introduce the notion of a vertex group as a Hopf algebra together with a specific choice of elementary vertex structure. Recall that a Hopf algebra is a module $H$ over a commutative ring $R$ (with unit) that has both the structure of an algebra and a coalgebra. It also possesses an antipode map $S:H \rightarrow H$ which is both a map of algebras and a map of coalgebras, and serves to connect the algebra and coalgebra structures. \begin{dfn} Let $H$ be a Hopf algebra over a commutative ring $R$, and let $H^* = \ensuremath{\mathrm{Hom}}_R(H,R)$ be the collection of $R$\mbox{-} linear maps from $H$ to $R$. We say that an $R$\mbox{-} module, $K$, gives an \defn{elementary vertex structure} on $H$ if it is an associative algebra over the algebra $H^*$ satisfying the following properties: \begin{enumerate} \item \defn{(Closure under left and right translation)} $K$ is a 2\mbox{-} sided $H$\mbox{-} module such that the product on $K$ is invariant under the left and right actions of $H$. By this we mean that for $h \in H$ and $k,l \in K$, we have: \begin{align*} h\mbox{$\cdot$} (kl)=\sum_{(h)}(h_{(1)}\mbox{$\cdot$} k)(h_{(2)}\mbox{$\cdot$} l) && (kl)\mbox{$\cdot$} h=\sum_{(h)}(k\mbox{$\cdot$} h_{(1)})(l\mbox{$\cdot$} h_{(2)}). \end{align*} We also require that the unit map for the algebra, $\eta:H^* \rightarrow K$, be a homomorphism of 2\mbox{-} sided $H$\mbox{-} modules. \item \defn{(Closure under Inversion)} The antipode on $H$ gives rise to a map $S^*$ on $H^*$ that can be extended to an $R$\mbox{-} linear map on $K$. By abuse of notation, we will refer to this map as $S:K \rightarrow K$, and because we are extending the the dual map it is not quite an $H^*$\mbox{-} algebra map, but instead satisfies $S(kl)=S(l)S(k)$ and $S(1)=1$ for $k,l,1 \in K$. \end{enumerate} \end{dfn} \begin{dfn}\label{d:K} If $H$ is a cocommutative Hopf algebra and $K$ is an elementary vertex structure on $H$ which is commutative and satisfies $S^2=1_K$, then we say that we have an \defn{elementary vertex group}, $G$. $H$ will be referred to as the \defn{group ring} of the vertex group $G$ or the \defn{underlying Hopf algebra}, and $K$ will be called the \defn{ring of singular functions on $G$}. \end{dfn} \begin{note} Since there are examples of vertex groups that are more general than this definition allows, we have chosen the name elementary vertex groups. Throughout the rest of this paper we will only be working with elementary vertex groups, so for simplicity, we will refer to them simply as vertex groups. \end{note} \begin{dfn} For any $H$\mbox{-} module, $B$, we abuse terminology by calling $\ensuremath{\mathrm{Hom}}_R(H^n,B)$ the \defn{collection of nonsingular functions from $\G{n}$ to $B$}. This will also be written $\ensuremath{\mathrm{Hom}}_R(\G{n},B)$. \end{dfn} Notice that for any cocommutative Hopf algebra, $H$, letting $K=H^*$ gives $H$ the structure of a trivial vertex group. This is a 2\mbox{-} sided $H$\mbox{-} module with left and right actions given as above for $g,h \in H$ and $f \in H^*$, \begin{align}\label{2sideaction} (h\mbox{$\cdot$} f)(g)=\sum_{(h)}h_{(1)}f(S(h_{(2)})g) && (f\mbox{$\cdot$} h)(g)=f(hg). \end{align} Since we are working with the dual of a Hopf algebra this is closed under left and right translation and under inversion. \subsection{Classical Vertex Group}\label{ss:classicalvg} The Hopf algebra which will prove most important for our later work is the complex polynomial algebra in one variable, $H = {\mathbb C}[T]$. For simplicity, we shall denote powers of $T$ by $\D{i} = T^i/i!$. Thus we have an associative algebra with multiplication $\mu (\D{i} \mbox{$\otimes$} \D{j}) = \binom{i+j}{i} \D{i+j}$ for any $i,j \geq 0$ and unit $\D{0}$. We shall denote multiplication as $\D{i} \mbox{$\cdot$} \D{j}$. This module has the additional structure of a Hopf algebra from the following three maps: \begin{align*} \text{Coalgebra Structure:} && &\begin{aligned}[t]\Delta: H &\rightarrow H \mbox{$\otimes$} H\\\D{i} &\mapsto \sum_{p+q = i} \D{p}\mbox{$\otimes$}\D{q} \end{aligned} & &\begin{aligned}[t] \epsilon:H &\rightarrow {\mathbb C}\\ \D{i} &\mapsto \begin{cases} 1 & i=0 \\ 0 & \text{otherwise} \end{cases} \end{aligned}\\[5mm] \text{Antipode:} && &\begin{aligned}[t]S: H &\rightarrow H.\\ \D{i} &\mapsto (-1)^{i}\D{i}. \end{aligned} \end{align*} Note that in particular this structure is cocommutative (i.e., $(1 \mbox{$\otimes$} \Delta)\Delta = (\Delta \mbox{$\otimes$} 1)\Delta$). It is easy to see that the linear dual of $H$, $H^* = \ensuremath{\mathrm{Hom}}_{\mathbb C}(H,{\mathbb C})$, is the ring of power series in one variable, $H^* = {\mathbb C}\ps{x}$. The dual pairing is given by letting $\D{1}$ act as differentiation on the $x$ variable evaluated at $x=0$. We can extend this action of $H$ on its dual to give an $H$\mbox{-} module structure on $H^*$ by defining the obvious action of derivation $\D{j} \mbox{$\cdot$} x^i = \binom{i}{j}x^{i-j}$, so that $H$ is just a group ring of derivations on ${\mathbb C}\ps{x}$. It is easy to see that more generally, $\ensuremath{\mathrm{Hom}}_{\mathbb C}(\ensuremath{H^{\otimes n}},{\mathbb C})$ can be identified with ${\mathbb C}\ps{x_1,\ldots,x_n}$, where a map $f$ is paired with $\sum r_{i_1, \ldots, i_n} x_1^{i_1}, \ldots, x_n^{i_n}$ when $f(\D{i_1} \mbox{$\otimes$} \ldots \mbox{$\otimes$} \D{i_n}) = r_{i_1, \ldots, i_n}$. If we choose our elementary vertex structure to be $K={\mathbb C}\ps{x}[x^{-1}]$, the space of Laurent series in one variable, we see immediately that it is an associative algebra over $H^*$ by construction. In other words, it is an infinite dimensional vector space over $H^*$ spanned by $\set{x^{-i}}{i \geq 0}$, with multiplication and unit maps giving it the structure of an algebra. The action of $H$ is the obvious extension of the action of derivation on $H^*$. So for any $j \geq 0$ and $i \in {\mathbb Z}$ we have $\D{j} \mbox{$\cdot$} x^i = \binom{i}{j}x^{i-j}$. Recall that for negative $i$, the binomial coefficient is given by \[\binom{i}{j} = \frac{i(i-1) \cdots (i-j+1)}{j(j-1) \cdots 1} = (-1)^j \binom{j-i-1}{j} .\] If $i=j$ then this gives back our dual pairing between $x^i$ and $\D{i}$. We extend the antipode to $K$ by defining $S(x^i)=(-1)^ix^i$ for all $i\in {\mathbb Z}$. By the product rule for derivation, we also see that the action of $H$ on the product of two elements in $K$ commutes with the multiplication by using the diagonal map. In other words, for any $j \geq 0$, $i_1,i_2 \in {\mathbb Z}$; \begin{eqnarray*} \D{j} \mbox{$\cdot$} (x^{i_1} x^{i_2}) &=& \mu\bigl(\Delta(\D{j}) \mbox{$\cdot$} (x^{i_1} \mbox{$\otimes$} x^{i_2})\bigr)\\ &=& \sum_{p+q=j} \mu\bigl((\D{p} \mbox{$\cdot$} x^{i_1}) \mbox{$\otimes$} (\D{q} \mbox{$\cdot$} x^{i_2})\bigr)\\ &=& \sum_{p+q=j} \binom{i_1}{p}\binom{i_2}{q} x^{i_1 + i_2 - p - q}\\ &=& \binom{i_1 + i_2}{j} x^{i_1 + i_2 - j}. \end{eqnarray*} and so this vertex algebra is closed under left and right translation. We shall call this the classical vertex group since it will give rise to the vertex algebras of \cite{kac}. It shall be denoted $G$ and will be the most important example for the sections that follow. \subsection{Representations of the Classical Vertex Group} We now return our attention to the case of a general vertex group, $G$, with underlying Hopf algebra, $H$. An $R$\mbox{-} module, $B$, is a \defn{representation} of the vertex group if it is a representation of the the underlying Hopf algebra. Hence the category of modules for a vertex group, $G$, is exactly the category of $H$\mbox{-} modules. When we refer to the action of $G$ on a module, we mean that its group ring is acting on the module. The reason for introducing the category of $G$\mbox{-} modules is because we will use the elementary vertex structure of the vertex group to provide this category with some additional singular structure. This additional structure will make it into a relaxed multilinear category (see definition \ref{d:relmultcat} below). Before we introduce this additional structure we look more closely at the underlying category of modules for a vertex group. \begin{note} We will want to consider the collection of all representations of a vertex group. We shall not restrict our attention to the full subcategory of finite dimensional representations because we want to allow ourselves the freedom to work with representations possessing a singular structure freely generated as an algebra. \end{note} We begin by pointing out that since the underlying Hopf algebra for any vertex group is cocommutative, the category of $G$\mbox{-} modules possesses a symmetric tensor product, which is just the tensor product of objects considered as $R$\mbox{-} modules. Given $G$\mbox{-} modules $A$ and $B$, the action of $G$ on $A \mbox{$\otimes$} B$ is given by the diagonal map. So we can think of $A \mbox{$\otimes$} B$ as having both a $G$\mbox{-} action and a $(G \mbox{$\otimes$} G)$\mbox{-} action. More generally, given the tensor product of a collection of $G$\mbox{-} modules, $A_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} A_n$, for every $1 \leq i \leq n$ there exist actions of $\G{i}$ corresponding to the different ways of comultiplying $\G{i} \rightarrow \G{n}$. Next we notice that the collection of $R$\mbox{-} linear maps between any two $G$\mbox{-} modules $A$ and $B$, denoted $\ensuremath{\mathrm{Hom}}_R(A,B)$, can be given the structure of a module in a number of ways. Given any such map, $f:A \rightarrow B$, we first define an action of $G$ on $f$ for any $a \in A$ and $g \in G$ by \begin{equation} \label{e:hom_action} (g \mbox{$\cdot$} f)(a) = \sum_{(g)}g_{(1)}\mbox{$\cdot$} f (S(g_{(2)})\mbox{$\cdot$} a). \end{equation} We say that $f$ is \defn{invariant} under the action of $G$ if $g \mbox{$\cdot$} f = \epsilon(g) f$ for all $g \in G$. This is equivalent to saying $f(g \mbox{$\cdot$} a) = g \mbox{$\cdot$} f(a)$, or in other words the $G$\mbox{-} invariant ring maps are exactly the $G$\mbox{-} module morphisms. There also exist two additional actions of $G$ on $\ensuremath{\mathrm{Hom}}_R(A,B)$; one where $G$ acts on the domain, the other where it acts on the codomain. The $G$\mbox{-} action on the domain of a map $f:A \rightarrow B$ is a right action on $f$, while the $G$\mbox{-} action on the codomain of $f$ is a left action. In fact, considering $\ensuremath{\mathrm{Hom}}_R(A_1 \mbox{$\otimes$} \ldots \mbox{$\otimes$} A_n,B)$ we see that in addition to being a map of $G$\mbox{-} modules, the domain possesses the additional structure of a module for $\G{n}$. So $\ensuremath{\mathrm{Hom}}_R(A_1 \mbox{$\otimes$} \ldots \mbox{$\otimes$} A_n,B)$ possesses the structure of a (right) $\G{n}$\mbox{-} module by action on the domain of $f$ (and similarly, a (left) $G$\mbox{-} module for action on the codomain, $B$). The fact that $\ensuremath{\mathrm{Hom}}_R(A_1 \mbox{$\otimes$} \ldots \mbox{$\otimes$} A_n,B)$ can be a module for both $G$ and $\G{n}$ will be important when we look at the composition of singular maps below (see section \ref{ss:sing_mult_maps}). Returning to the particular case of the classical vertex group, we have shown that $\ensuremath{\mathrm{Hom}}_{\mathbb C}(G,{\mathbb C}) \cong {\mathbb C}\ps{x}$. We see immediately that for any $G$\mbox{-} module, $B$, the linear dual, $\ensuremath{\mathrm{Hom}}_{\mathbb C}(G,B)$, is isomorphic as a $G$\mbox{-} module to $B\ps{x}$, where $B\ps{x}$ denotes the collection of power series in $x$ with coefficients in $B$. This isomorphism between maps linear maps and power series can be extended to maps from $n$\mbox{-} fold products of $G$ for any $n \geq 0$. In such a case we have $\ensuremath{\mathrm{Hom}}_{\mathbb C}(\G{n},B) \cong B \ps{x_1, \ldots, x_n}$ (where we have tacitly included the case where $n=0$ and $\ensuremath{\mathrm{Hom}}_{\mathbb C}(\G{0},B) = \ensuremath{\mathrm{Hom}}_{\mathbb C}({\mathbb C},B) \cong B$). From the discussion above we know that $\G{n}$ has the structure of a $G$\mbox{-} module using the $n$\mbox{-} fold diagonal map, so the collection of linear maps $\ensuremath{\mathrm{Hom}}_{\mathbb C}(\G{n},B)$ also has the structure of a $G$\mbox{-} module as described above. Under the isomorphism $\ensuremath{\mathrm{Hom}}_{\mathbb C}(\G{n},B) \cong B \ps{x_1, \ldots, x_n}$, the action of $\D{i} \in G$ on any $b(x_1, \ldots, x_n) \in B \ps{x_1, \ldots, x_n}$ is given by: \begin{equation}\label{e:action} \D{i} \mbox{$\cdot$} b(x_1, \ldots, x_n) = \sum_{j_0 + \cdots + j_n = i} (-1)^{j_0}\bigl(\D{j_0}_B + \D{j_1}_{x_1} \cdots + \D{j_n}_{x_n}\bigr) b(x_1,\ldots,x_n). \end{equation} In the previous expression, $D_B$ denotes the action of $G$ on the coefficients of the power series $b(x_1, \ldots, x_n)$, and where $\D{j_k}_{x_k}$ denotes the action of differentiation on the variable $x_k$. As before, we say that $b(x_1, \ldots, x_n)$ is invariant under the action of $H$ (or $G$) if $\D{i} \mbox{$\cdot$} b(x_1, \ldots, x_n)$ is zero. This is the same as requiring that $b(x_1, \ldots, x_n)$ satisfy the following for all $i \geq 0$: \begin{equation}\label{e:invariant} \D{i}_B b(x_1, \ldots, x_n) = \sum_{j_1 + \cdots + j_n = i} \bigl(\D{j_1}_{x_1} + \cdots + \D{j_n}_{x_n}\bigr) b(x_1, \ldots, x_n). \end{equation} The collection of all such invariant elements is denoted $\ensuremath{\mathrm{Hom}}_G(\G{n},B)$. Working out equation \eqref{e:action} explicitly for the case of $\ensuremath{\mathrm{Hom}}_{\mathbb C}(\G{2},B) \cong B \ps{x,y}$, we have \begin{eqnarray*} \D{1}\mbox{$\cdot$}(\sum_{i,j \geq 0} b_{i,j} x^{i} y^{j}) &=& \sum_{i,j \geq 0} (\D{1}\mbox{$\cdot$} b_{i,j}) x^{i} y^{j} - (\D{1}_{x} + \D{1}_{y})\sum_{i,j \geq 0} b_{i,j} x^{i} y^{j} \\ &=& \sum_{i,j \geq 0} (\D{1}b_{i,j}) x^{i} y^{j} - \sum_{i,j \geq 0} i b_{i,j} x^{i-1} y^{j} - \sum_{i,j \geq 0} j b_{i,j} x^{i} y^{j-1}. \end{eqnarray*} And from equation \eqref{e:invariant}, we see that the $G$\mbox{-} invariant elements of $B\ps{x,y}$ are exactly those which satisfy $\D{1} \mbox{$\cdot$} b_{i,j} = (i+1)b_{i+1,j} + (j+1)b_{i,j+1}$, or more generally \begin{equation} \D{k}\mbox{$\cdot$} b_{i,j} = \sum_{p+q=k} \binom{p+i}{i} \binom{q+j}{j}b_{p+i,q+j}. \end{equation} For any $n$, this equation also has an ``inverted form'' in which we can write any $b_{i_1,\ldots,i_n}$ as a sum of $b_{j_1,\ldots,j_n}$ where one of the indices $(j_1,\ldots,j_n)$ is set to zero. For $n=2$ we can write $b_{i,j}$ as: \begin{eqnarray} b_{i,j} &=& \sum_{p+q=j} (-1)^p \binom{i+p}{i}\preup{\D{q}}{b_{i+p,0}}\label{eq:d_i0} \\ &=& \sum_{p+q=i} (-1)^q \binom{j+q}{j}\preup{\D{p}}{b_{0,j+q}}.\label{eq:d_0i} \end{eqnarray} This ability to represent $G$\mbox{-} invariant power series in various way will be important for encapsulating duality for vertex algebras. \subsection{Extended Representation of Morphisms - Holomorphic Vertex Algebras}\label{ss:extended} Now that we have a characterisation of modules for a vertex group, we shall look more closely at maps between them. We would like to describe module maps in a way which will allow us to naturally add the elementary vertex structure. This treatment will avoid a very general abstract treatment, instead aiming to provide the flavour of theory. The details can be found in \cite{cts_valgdetails}. Given any linear map between $G$\mbox{-} modules, $f:A \rightarrow B$, we define a unique $G$\mbox{-} linear map $\widehat{f}:A \rightarrow \ensuremath{\mathrm{Hom}}_R(G,B)$ by: \[ \widehat{f}(a)(g) = f(g \mbox{$\cdot$} a). \] This pairing between linear maps and $G$\mbox{-} linear maps is actually an isomorphism in the category of $G$\mbox{-} modules, so we may easily pass from one description to the other. We shall call this an \defn{extended representation} of $f$. If the map $f$ was also $G$\mbox{-} invariant, then the extended representation of $f$ would be a $G$\mbox{-} linear map from $A$ to $G$\mbox{-} invariant elements of $\ensuremath{\mathrm{Hom}}_R(G,B)$. These extended representations can also be used to reexpress linear maps of the form $f:A_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} A_n \rightarrow B$ as $\G{n}$\mbox{-} module maps $\widehat{f}: A_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} A_n \rightarrow \ensuremath{\mathrm{Hom}}_R(\G{n},B)$, where the action of $\G{n}$ on the domain, $A_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} A_n$, passes to an action on the domain of the maps $\ensuremath{\mathrm{Hom}}_R(\G{n},B)$. \begin{example} In the case of the classical vertex group, this means that linear maps between $G$\mbox{-} modules $f:A_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} A_n \rightarrow B$, are the same as $\G{n}$\mbox{-} linear maps \[\widehat{f}: A_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} A_n \rightarrow B\ps{x_1, \ldots, x_n}.\] In particular, we have that the action of $G$ on any of the $A_i$ carries over to differentiation of $x_i$ as \[ \widehat{f}(a_1\mbox{$\otimes$} \cdots \mbox{$\otimes$} \D{1}a_i \mbox{$\otimes$} \cdots \mbox{$\otimes$} a_n) = \partial_{x_i} \widehat{f}(a_1\mbox{$\otimes$} \cdots \mbox{$\otimes$} a_i \mbox{$\otimes$} \cdots \mbox{$\otimes$} a_n).\] \end{example} The composition of maps in this extended representation is slightly delicate. Given ordinary linear maps of $G$\mbox{-} modules, say $f:A \rightarrow B$ and $g:B \rightarrow C$, then clearly these compose to give a map $g \circ f:A \rightarrow C$. Now $f$ and $g$ can be considered in their extended representation, i.e., $G$\mbox{-} linear maps $\widehat{f}:A \rightarrow \ensuremath{\mathrm{Hom}}_R(G,B)$ and $\widehat{g}:B \rightarrow \ensuremath{\mathrm{Hom}}_R(G,C)$, which compose to give a $G$\mbox{-} linear map $\widehat{g} \circ \widehat{f}:A \rightarrow \ensuremath{\mathrm{Hom}}_R(\G{2}, C)$. But the ordinary composite $g \circ f$ has an extended representation $\widehat{g \circ f}:A \rightarrow \ensuremath{\mathrm{Hom}}_R(G, C)$. We would expect these to be related, but in fact, we need to require that $f$ be $G$\mbox{-} invariant in order to relate/reduce the extra factor of $G$ in the codomain of $\widehat{g} \circ \widehat{f}$. Intuitively, this problem appears because the action of $G$ on $B$ appears in $\ensuremath{\mathrm{Hom}}_R(G,C)$, but the composite $g \circ f$ ``hides'' that action. So by taking $f$ to be $G$ \mbox{-} invariant we eliminate the action of $G$ on $f$ explicitly. More generally, maps in the extended representation compose pointwise as multilinear functions in the usual way, provided that all maps (except possibly the bottom map) are $G$\mbox{-} invariant. \begin{example}\label{ex:nonsing_comp} Again we consider the case of the classical vertex group. Let $A_i, B_j, C_l$ and $D$ be $G$\mbox{-} modules. Given a $\G{n}$\mbox{-} linear map $g:C_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} C_n \rightarrow D\ps{z_1, \ldots, z_n}$, and $G$\mbox{-} invariant maps $f:A_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} A_p \rightarrow C_1\ps{x_1, \ldots, x_p}$, $k:B_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} B_q \rightarrow C_2\ps{y_1, \ldots, y_q}$ ($\G{p}$\mbox{-} linear and $\G{q}$\mbox{-} linear respectively), then the composite is a $\G{(p+q+n-2)}$\mbox{-} linear map \[g \circ (f \mbox{$\otimes$} k):A_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} A_p \mbox{$\otimes$} B_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} B_q \mbox{$\otimes$} C_3 \mbox{$\otimes$} \cdots \mbox{$\otimes$} C_n \rightarrow D\ps{x_1, \ldots, x_p, y_1, \ldots, y_q, z_1, \ldots, z_n}\] which satisfies $\partial_{z_1} = \partial_{x_1}+ \cdots + \partial_{x_p}$ and $\partial_{z_2} = \partial_{y_1}+ \cdots + \partial_{y_q}$. In other words, the map $g \circ (f \mbox{$\otimes$} k)$ factors through $D\ps{x_1+z_1, \ldots, x_p+z_1, y_1+z_2, \ldots, y_q+z_2, z_3, \ldots, z_n}$. Also, it makes sense to consider compositions in a particular order, say $(g \circ k) \circ f$ or $(g \circ f) \circ k$, and these are equal. \end{example} \begin{claim}\label{c:holo_equivalence} Let {\boldmath $H \mbox{-} \mathit{Mod}$} be the category of representations of the underlying Hopf algebra, $H$, of the classical vertex group. Then category of holomorphic vertex algebra is isomorphic to the category of $H$\mbox{-} invariant commutative algebras in {\boldmath $H \mbox{-} \mathit{Mod}$}. \end{claim} \begin{proof} In order to make sense of this statement, we will first review both the definition of a holomorphic vertex algebra, and the definition of an algebra in a category. \begin{dfn} A \defn{holomorphic vertex algebra} is a vertex algebra without singularities. In other words, the vertex operator of a holomorphic vertex algebra is a map: \[Y(\cdot, x)\cdot:V \mbox{$\otimes$} V \rightarrow V\ps{x}.\] It follows that the locality axiom reduces to the statement that products of vertex operators commute. The collection of holomorphic vertex algebras is given the structure of a category by defining a morphism in the category of holomorphic vertex algebras to be a map of complex vector spaces taking vacuum to vacuum, commuting with multiplication, and respecting the actions of the infinitesimal translation operators. \end{dfn} \begin{dfn}\label{d:algebra_defn} A \defn{commutative (associative) algebra} in a symmetric tensor category consists of an object $A$, a multiplication map $\mu: A \mbox{$\otimes$} A \rightarrow A$ which is invariant under the symmetry action for the tensor product, and a unit for the multiplication $\eta:I \rightarrow A$ satisfying the usual axioms for associativity and unit (where $I$ is the unit for the tensor product). \end{dfn} For the category of representations of the Hopf algebra $H$, the unit for multiplication is ${\mathbb C}$. and the multiplication map $\mu$ is $H$\mbox{-} invariant. Given any such algebra in this category of representations, where the multiplication map $\mu$ is $H$\mbox{-} invariant, we form a holomorphic vertex algebra as follows. We begin by taking $T = \D{1}$ as the infinitesimal translation operator, and take the vacuum vector to be $\eta(1)$. Then by considering $\mu$ in the extended representation, we have a map \[ \widehat{\mu}:A\mbox{$\otimes$} A \rightarrow A\ps{x,y},\] which satisfies \begin{eqnarray} \widehat{\mu}(Ta\mbox{$\otimes$} b) &=& \partial_x \widehat{\mu}(a \mbox{$\otimes$} b)\\ \widehat{\mu}(a\mbox{$\otimes$} Tb) &=& \partial_y \widehat{\mu}(a \mbox{$\otimes$} b)\\ T\widehat{\mu}(a \mbox{$\otimes$} b)&=& \widehat{\mu}(Ta\mbox{$\otimes$} b)+\widehat{\mu}(a\mbox{$\otimes$} Tb). \end{eqnarray} From this we define a vertex operator on $A$ to be \begin{equation}\label{e:operator_from_alg} Y(\cdot, x)\cdot = \widehat{\mu}(\cdot \mbox{$\otimes$} \cdot)|_{y=0}:A \mbox{$\otimes$} A \rightarrow A\ps{x}. \end{equation} Checking that this satisfies the axioms for a holomorphic vertex algebra, we see immediately that the vacuum axioms are satisfied because $H$ acts trivially on ${\mathbb C}$, and because $\mu(\eta(1)\mbox{$\otimes$} a) = a$ for all $a \in V$. Writing out the translation covariance axiom, we use the $H$\mbox{-} invariance of $\widehat{\mu}$ to give: \begin{eqnarray} T\widehat{\mu}(a \mbox{$\otimes$} \cdot)|_{y=0} - \widehat{\mu}(a \mbox{$\otimes$} T\cdot)|_{y=0} &=& \widehat{\mu}(Ta\mbox{$\otimes$} \cdot)|_{y=0} +\widehat{\mu}(a\mbox{$\otimes$} T\cdot)|_{y=0} - \widehat{\mu}(a \mbox{$\otimes$} T\cdot)|_{y=0} \\ &=& \widehat{\mu}(Ta\mbox{$\otimes$} \cdot)|_{y=0} \\ &=& \partial_x \widehat{\mu}(a\mbox{$\otimes$} \cdot)|_{y=0}. \end{eqnarray} And the locality axiom follows from the commutativity of $\widehat{\mu}$. Notice that $\widehat{\mu}(a \mbox{$\otimes$} b) = Y(a,x)Y(b,y)\mbox{$|0\rangle$}$. Similarly, given any vertex algebra, we can easily define an algebra in the category of of representations of the Hopf algebra $H$ by taking $\eta(r) = r\mbox{$|0\rangle$} \in V$ for any $r \in R$, and $\mu(\cdot \mbox{$\otimes$} \cdot) = Y(\cdot, x)\cdot|_{x=0}$. The unit axioms follow from the properties of the vacuum, and associativity follows from locality. The translation covariance axiom says that the multiplication is $H$\mbox{-} invariant. And, locality axiom acting on the vacuum says that this multiplication is commutative. If $(V, Y(\cdot,x)\cdot, T, \mbox{$|0\rangle$})$ is a vertex algebra, and the corresponding $H$\mbox{-} invariant algebra is $(V, \mu, \eta)$, then from this algebra, we get back the vertex algebra, $(V, Y^\prime(\cdot,z)\cdot, T, \mbox{$|0\rangle$})$. If $Y(a,x)b = \sum_{i \geq 0} c_i x^i$ for $a\mbox{$\otimes$} b \in V\mbox{$\otimes$} V$, then \[\widehat{\mu}(a\mbox{$\otimes$} b) = \sum_{i, j \geq 0}\biggl[Y(\D{i}a,x)\D{j}b\biggr]_{x=0} z^i y^j \] so setting $y=0$, we have \begin{eqnarray*} \widehat{\mu}(a\mbox{$\otimes$} b)|_{y=0} &=& \sum_{i \geq 0}\biggl[Y(\D{i}a,x)b\biggr]_{x=0} z^i \\ &=& \sum_{i \geq 0}\biggl[\sum_{j \geq 0} \binom{j+i}{i}c_{j+i} x^j\biggr]_{x=0} z^i \\ &=& \sum_{i \geq 0}c_i z^i, \end{eqnarray*} and so $Y = Y^\prime$. Similarly, starting with an $H$\mbox{-} invariant algebra, $(A,\mu, \eta)$, the process of creating a holomorphic vertex algebra and then mapping back to an $H$\mbox{-} invariant algebra clearly maps $(A,\mu, \eta)$ to itself. Finally, a morphism in the category of holomorphic can be seen to correspond exactly to a morphism of algebras in the category of $H$\mbox{-} modules. Because this pairing of objects in each category is completely natural we have an isomorphism of categories and so our claim is proved. \end{proof} \section{Evaluation Maps}\label{a:evaluation} This appendix is concerned with making clear what types of evaluation maps arise naturally when dealing with collections of multilinear maps of $G$\mbox{-} modules for an arbritrary vertex group $G$. \begin{example} If $A$ and $B$ are $G$\mbox{-} modules, then we know that there exists a natural evaluation map \[\ensuremath{\mathrm{Hom}}(A,B) \mbox{$\otimes$} B \longrightarrow B.\] This of course holds for ordinary $R$\mbox{-} linear maps, $\ensuremath{\mathrm{Hom}}_R(A,B)$, as well as for $G$\mbox{-} invariant maps $\ensuremath{\mathrm{Hom}}_G(A,B)$. In fact, given $G$ modules, $A_1, \ldots, A_n, B$ we have partial evaluation maps \begin{equation}\label{e:partial_eval} \ensuremath{\mathrm{Hom}}(A_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} A_n, B) \mbox{$\otimes$} A_1 \mbox{$\otimes$} A_2 \longrightarrow \ensuremath{\mathrm{Hom}}(A_3 \mbox{$\otimes$} \cdots \mbox{$\otimes$} A_n, B). \end{equation} \end{example} The problem with evaluation arises when we consider singular maps. Recall that for a $G$\mbox{-} module, $B$, the collection of singular maps $\ensuremath{\mathit{Fun}}(\G{2},B)$ was defined to be $\ensuremath{\mathrm{Hom}}(\G{2},B)\mbox{$\otimes$}_{f_{1,2}} K$, where the tensor product is taken over $H^*$. We are tempted to evaluate this map for some $\G{2}$, \[\ensuremath{\mathrm{Hom}}(\G{2},B)\mbox{$\otimes$}_{f_{1,2}} K \mbox{$\otimes$} \G{2} \longrightarrow B \mbox{$\otimes$} K,\] but no such natural map exists because $B$ does not have the structure of an $H^*$\mbox{-} module. (Recall that the action of $H^*$ on $\ensuremath{\mathrm{Hom}}(\G{2},B)$ is on its domain.) The following example makes the problem more explicit: \begin{example} When $G$ is the classical vertex group, $\ensuremath{\mathit{Fun}}(\G{2},B) \cong B\ps{x,y}[(x-y)^{-1}]$, and so an arbritrary element of this collection of maps is, \[(x-y)^{-k}\sum_{i,j \geq 0} b_{i,j} x^i y^j, \] for some $b_{i,j} \in B$ and some $k \geq 0$. Naively evaluating this at $\D{p}\mbox{$\otimes$}\D{q} \in \G{2}$ we have $b_{p,q}(x-y)^{-k} \in B \mbox{$\otimes$} K$. But if we now rewrite our power series as \[(x-y)^{-k-1}\sum_{i,j \geq 0} b_{i,j} (x^{i+1} y^{j} + x^{i} y^{j+1}, \] we see that this evaluates to $(x-y)^{-k-1}(b_{p-1,q} + b_{p,q-1})$ for $p,q \geq 1$. Thus we see that the same power series seems to have evaluated to two different solutions. We still might think that the situation could be reconciled by setting the two evaluation equal to one another, giving \[b_{p-1,q} + b_{p,q-1} = b_{p,q}(x-y).\] But this is just a restriction on the form of our power series, it does not define an action of $H^*$ on $B$ as we might have hoped. \end{example} Now that we see that there are evaluations which we can and can not make, we would like to know how to evaluate multilinear singular maps, $\ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/generic.eps,height=3mm}}(A_1, \ldots, A_n; B)$, for an arbritrary tree, p. Since these are defined as pullbacks over $\ensuremath{\mathrm{Hom}}_{\G{n}}\biggl(A_1\mbox{$\otimes$} \cdots \mbox{$\otimes$} A_n, \ensuremath{\mathit{Fun}}_p(\G{n}, B)\biggr)$, we know that they can be evaluated partially or completely for all $A_1, \ldots, A_n$ as in equation \eqref{e:partial_eval}. In fact, the singular maps, $\ensuremath{\mathrm{Sing}}_\sigma(A_1, \ldots, A_n; \ensuremath{\mathrm{Hom}}(\G{n},B))$ of equation \eqref{e:nary_general_pbk} can also be evaluated, because although there are many copies of the singularity $K$ among the $A_i$, the tensoring over $H^*$ is with the inner copy of $\G{n}$ (this can be realised as a suitable quotient), and so we could for example evaluate the following at $A_{\sigma(2)}, A_{\sigma(3)}$: \begin{multline}\label{e:sing_eval} \ensuremath{\mathrm{Hom}}\biggl(A_{\sigma(1)}\mbox{$\otimes$} A_{\sigma(2)}, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl(A_{\sigma(3)}, K^{\otimes 2}\mbox{$\otimes$} \bigl(\cdots K^{\otimes n-1} \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{n},B)\cdots\bigr)\Bigr)\biggr) \\ \longrightarrow \ensuremath{\mathrm{Hom}}\biggl(A_{\sigma(1)}, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl(A_{\sigma(4)}, K^{\otimes 2}\mbox{$\otimes$} K^{\otimes 3}\mbox{$\otimes$} \bigl(\cdots K^{\otimes n-1} \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{n},B)\cdots\bigr)\Bigr)\biggr). \end{multline} \section{Relaxed Multi Categories} So far in this paper, we have introduced the notion of a vertex group. We then looked at representations of vertex groups, paying special attention to representation of the classical vertex group. We then introduced the notion of multilinear singular maps, and saw that in order for them to compose properly, we needed to parameterise them by trees. Finally we described the canonical maps between singular maps which arise when we refine one tree to another. In this section we shall be concerned with general properties satisfied by our multilinear singular maps. The work which we have already done in defining, composing and refining the singular maps will be enough to make it clear that we have been working with examples of the following structure. \begin{dfn}\label{d:relmultcat} A \bemph{relaxed multilinear category} is an ordinary category of $R$\mbox{-} modules for some ring $R$ with the following additional structure: \begin{enumerate} \item \bemph{Multi Maps:} For any $n+1$ objects $A_1, \ldots, A_n, B$ and any $n$ leafed tree, $p$, there is a collection of multi maps from $A_1, \ldots, A_n$ to $B$ denoted $\ensuremath{\mathit{Multi}}_p(A_1, \ldots, A_n, B)$. In particular, if $\bullet$ is the unique tree with no internal vertices, then for all objects $A$, $\ensuremath{\mathit{Multi}}_\bullet(A,A)$ contains the identity morphism. \item\label{d:relaxed_composition} \bemph{Composition:} For any $(n+1)$\mbox{-} tuple of multi maps, $\ensuremath{\mathit{Multi}}_{p_i}(A_{i1}, \ldots, A_{im_1}, B_i)$, $\ensuremath{\mathit{Multi}}_q(B_1, \ldots, B_n, C)$, $1 \leq i \leq n$ there is a map that defines composition: \begin{multline*} \ensuremath{\mathit{Multi}}_{p_1}(A_{11},\ldots, A_{1m_1}, B_1)\mbox{$\otimes$} \ldots \mbox{$\otimes$} \ensuremath{\mathit{Multi}}_{p_n}(A_{n1},\ldots, A_{nm_n}, B_n) \\ \mbox{$\otimes$} \ensuremath{\mathit{Multi}}_q(B_1, \ldots, B_n, C) \longrightarrow \ensuremath{\mathit{Multi}}_{q(p_1, \ldots, p_n)}(A_{11},\ldots, A_{nm_n}, C), \end{multline*} where $q(p_1, \ldots, p_n)$ is the tree with $\sum m_i$ leaves formed by gluing the root of each tree $p_i$ to the $i$th external edge. Composition of maps of this type is associative. \item\label{d:relaxed_refinement} \bemph{Refinement:} If $p$ and $q$ are $n$ leaved trees and $p$ is a refinement of $q$, then for every collection of $n+1$ objects $A_1, \ldots, A_n, B$ there is a map \[r_{p,q}: \ensuremath{\mathit{Multi}}_p(A_1, \ldots, A_n, B) \longrightarrow \ensuremath{\mathit{Multi}}_q(A_1, \ldots, A_n, B).\] \item \bemph{Linearity:} The collections of multi maps, $\ensuremath{\mathit{Multi}}_p(A_1, \ldots, A_n, B)$ are $R$\mbox{-} modules and composition is multilinear. \end{enumerate} \end{dfn} The category of representations of a vertex group possesses the structure of a relaxed multilinear category using the $G$\mbox{-} invariant collections of singular functions. Notice that we need to restrict to $G$\mbox{-} invariant multi maps in order for the composition condition to hold. With this categorical structure defined, the next natural step is to define an algebra in a relaxed multilinear category. Recall that we defined an (associative) algebra in a symmetric tensor category in definition \ref{d:algebra_defn}. In a relaxed multilinear category we define an (associative) algebra as follows: \begin{dfn} An \bemph{(associative) algebra} in a relaxed multilinear category, ${\cal B}$, consists of an object $B \in {\cal B}$ and a collection of maps $\{f_{p}\} = \set{f_{p} \in \ensuremath{\mathit{Multi}}^K_{G,p}(B, \ldots, B, B)}{\text{$p$ is an $n$ leafed tree},n \in {\mathbb N}}$ . These maps must satisfy the following axioms: \begin{enumerate} \item\label{d:algebra_composition} \bemph{Composition:} If $q(p_1, \ldots, p_n)$ is the tree formed by gluing the root of each tree $p_i$ to the $i$th external edge of an $n$ leafed tree, $q$, ($p_i$ possibly empty), then \begin{equation} f_{q(p_1, \ldots, p_n)} = f_q \circ (f_{p_1}, \ldots, f_{p_n}). \end{equation} \item \bemph{Unit:} The map $f_\circ:R \rightarrow B$ (where $\circ$ is the empty tree) defines a unit for the algebra in the sense that for any $n$ leafed tree, $p$, and any $1 \leq k \leq n$, \[f_p \circ_k f_\circ = f_{p^\prime}\] where $\circ_k$ denotes composition at the $k$th leaf of $p$, and $p^\prime$ is the $n-1$ leaved tree arrived at by removing the $k$th leaf from $p$. \item \bemph{Refinement:} If $p,q \in \T{n}$ and $p$ is a refinement of $q$, then $r_{p,q} (f_p) = f_q$ where $r_{p,q}$ is the refinement map given by the refinement axiom for a relaxed multilinear category. \end{enumerate} \end{dfn} This is an algebra in the sense that each map $f_p$ defines an ``$n$\mbox{-} fold multiplication'' for elements of $B$. For all $n \in {\mathbb N}$ we denote the multilinear map associated to the flat tree with $n$ leaves by $f_n$. Since composition of multi maps in ${\cal B}$ is associative, the associativity of $(B,\{f_{p}\})$ is a consequence of the composition axiom. Considering $\bullet$, the 1 leafed tree with zero edges, then since $f_p \circ_k f_{\bullet} = f_p$ and $f_{\bullet} \circ f_p = f_p$, we see that $f_{\bullet} = 1_B$. The algebra defined by $(B,\{f_{p}\})$ is said to be \defn{commutative} if there exists an action of the symmetric group on each of the multilinear maps in $\{f_{p}\}$. \begin{example}\label{ex:last_part_of_vacuum} Since $f_{\bullet}$ refines to $f_{\ \epsfig{file=Images/1_flat.eps,height=3mm}}$, we know that for any $b \in B$, and $g \in G$, \[f_{\ \epsfig{file=Images/1_flat.eps,height=3mm}}(b)(g) = gb.\] \end{example} \begin{note} This definition of an algebra is just a functor from the opposite of the category or trees (see appendix \ref{a:cat_of_trees}) to ${\cal B}$ where each object $p \in \mathcal{T}$ is mapped to an element of $\ensuremath{\mathit{Multi}}^K_{G,p}$. \end{note} \begin{thm} \label{t:from_alg_to_valg} Given any commutative algebra, $(B,\{f_p\})$, for the classical vertex group, it gives rise to a vertex algebra as in definition \ref{d:vetex_algebra} where $B$ is the space of fields and $\mbox{$|0\rangle$} = f_0(1)$. The infinitesimal translation operator is $T=\D{1}$, and the state-field correspondence is given by $Y(\cdot,x)\cdot = f_2(\cdot,\cdot)|_{y=0} : B \rightarrow B\ps{x}[x^{-1}]$. \end{thm} \begin{proof} In order to show that $(B,\{f_p\})$ gives a vertex algebra, we check the axioms for the vertex algebra. The vacuum axioms follow naturally from the discussion of section \ref{ss:vacuum}, together with previous example, \ref{ex:last_part_of_vacuum}, giving \[f_{\epsfig{file=Images/2_leaf.eps,height=3mm}}(\mbox{$|0\rangle$} \mbox{$\otimes$} b) = \sum_{i\geq 0} \frac{T^i}{i!}b y^i = e^{Ty}b.\] Translation covariance is automatically satisfied because our functions, $f_p$, are $G$\mbox{-} invariant. And, we saw that the locality condition is satisfied in section \ref{ss:locality}. \end{proof} \begin{thm} A vertex algebra (as in definition \ref{d:vetex_algebra}) with state space $V$, defines an algebra in the relaxed multilinear category of representations of the classical vertex group. \end{thm} \begin{proof} As in the proof for holomorphic vertex algebras (claim \ref{c:holo_equivalence}), the vacuum defines a map $f_\circ$ for the empty tree. We described in example \ref{ex:getting_singmaps} how to construct $f_{\epsfig{file=Images/2_leaf.eps,height=3mm}}$, and section \ref{ss:locality} showed how how to construct the singular function associated to the flat tree with three leaves. Carrying out a similar process leads to the construction of singular functions associated to all flat trees, and the singular functions associated to trees with internal nodes arise from composition of singular functions associated to flat trees. Closure under refinement follows by construction. \end{proof} \section{Introduction} \subsection{Motivation} \label{ss:motivation} The theory of (non-supersymmetric) axiomatic vertex algebras has as its data, a complex vector space, $V$, together with a \defn{vertex operator}: \[Y(\cdot,x)\cdot:V\mbox{$\otimes$} V \longrightarrow V\ps{x}[x^{-1}],\] as well as a distinguished vector $\mbox{$|0\rangle$}$ and an automorphism $T:V\rightarrow V$. So for vectors $a,b \in V$, the vertex operator can be written as \[Y(a,x)b = x^{-k} \sum_{i \geq 0} c_i x^i, \] where $c_i \in V$ and $k \geq 0$. We would like to describe carefully the products of vertex operators, and to give a precise description of what types of maps arise from such products. We represent the collection of all such vertex operators by the labelled tree \[\flattwoleafed{V}{V}{V}{x}{}\] By extending the domain of the vertex operator pointwise to $V\ps{x}[x^{-1}]$, we have two ways to take the product of a pair of vertex operators, giving maps: \begin{eqnarray} Y(\cdot,x)Y(\cdot,y)\cdot:&V\mbox{$\otimes$} V\mbox{$\otimes$} V \longrightarrow& V\ps{x}[x^{-1}]\ps{y}[y^{-1}] \label{e:wrong_comp1}\\ Y(Y(\cdot,y)\cdot,x)\cdot:&V\mbox{$\otimes$} V\mbox{$\otimes$} V \longrightarrow& V\ps{x}[x^{-1}]\ps{y}[y^{-1}].\label{e:wrong_comp2} \end{eqnarray} Immediately we notice that the collection of all maps from $V^{\otimes 3}$ to $V\ps{x}[x^{-1}]\ps{y}[y^{-1}]$ contains many maps which can not be realised as the composite of two vertex operators. For example, the singularity in the variable $y$ only depends on two of the copies of $V$ in $V^{\otimes 3}$. Composing two copies of the labelled tree, we can represent the collection of all maps arising as the composite of two vertex operators by \[\threeleafedleft{V}{V}{V}{V}{y}{}{x}{} \threeleafedright{V}{V}{V}{V}{y}{}{x}{} \] This labeled tree notation makes clear this dependency of singularities on inputs. Looking more closely at the product, $Y(\cdot,x)Y(\cdot,y)\cdot$, we see that a more accurate description of the space of maps defined by this product is given by the collection, \[\ensuremath{\mathrm{Hom}}\Bigl(V\mbox{$\otimes$} V,\ensuremath{\mathrm{Hom}}(V, V\ps{x}[x^{-1}])\ps{y}[y^{-1}]\Bigr).\] We can repeat this process, so that the collection of $n$\mbox{-} fold products of vertex operators, $Y(\cdot,x_n)\cdots Y(\cdot,x_1)\cdot$ can be described exactly as the product, \[\ensuremath{\mathrm{Hom}}\biggl(V\mbox{$\otimes$} V,\ensuremath{\mathrm{Hom}}\Bigl(V, \cdots \ensuremath{\mathrm{Hom}}(V, V\ps{x_n}[x_n^{-1}])\cdots \Bigr)\ps{x_1}[x_1^{-1}]\biggr).\] Similarly, this space describes the other type of $n$\mbox{-} fold product of vertex operators, \[Y\Bigl(Y\bigl(\cdots Y(Y(\cdot,x_1)\cdot,x_{2})\cdots\bigr)\cdot,x_n\Bigr)\cdot.\] The difficulty arises when we begin to consider the space of products of vertex operators which contain both types of composition. Take for example the composite, \[Y(Y(\cdot,x)\cdot,z)Y(\cdot,y)\cdot.\] This product can be achieved either by composing three vertex operators, $Y(\cdot, x)\cdot$, $Y(\cdot, y)\cdot$, and $Y(\cdot, z)\cdot$, or by appropriately composing a vertex operator $Y(\cdot, y)\cdot$ with an element of \[\ensuremath{\mathrm{Hom}}(V\mbox{$\otimes$} V,\ensuremath{\mathrm{Hom}}(V, V\ps{z}[z^{-1}])\ps{x}[x^{-1}])= \threeleafedleft{V}{V}{V}{V}{x}{}{z}{}\] or even by composing $Y(\cdot, x)\cdot$ in the other way with an element of \[\ensuremath{\mathrm{Hom}}(V\mbox{$\otimes$} V,\ensuremath{\mathrm{Hom}}(V, V\ps{z}[z^{-1}])\ps{y}[y^{-1}])= \threeleafedright{V}{V}{V}{V}{y}{}{z}{}\] Either way, we would like to understand the space of all maps which can be realized as any of the composites given above. Using the tree notation, we shall denote the space of all such maps by \[\fourleafdoubleA \] but this space can not be described as explicitly as the previous spaces of products of vertex operators. In addition to the desire to simply describe these spaces of products of vertex operators, we would like to be able to relate them in such a way as to take account of the axioms of the vertex algebra. The usual way of working with products of vertex operators is to consider them inside very large spaces of formal distributions. While this has the benefit of leaving plenty of room to manipulate power series, it tends to obscure the important features of the vertex operators. In this paper we shall take the opposite tack, opting instead to use this minimal description of the spaces in which these products of vertex operators live. In this way, we will be able to make clear the essential features of composition of vertex operators. \subsection{Outline} The idea behind this paper is the extraction the essential features of axiomatic vertex algebras, with the ultimate goal of showing how they can be seen to arise naturally in a certain context. By demonstrating this naturality, the generalisation of vertex algebras to more general settings becomes simply a matter of choosing the correct categories in which to work, with the long term goal of applying them to higher dimensional field theories. We shall approach the problem of giving an abstract account of vertex algebras by first considering carefully the vertex operator. After reviewing the definition of a vertex algebra, we define a vertex group which ties together the power series and the infinitesimal translation operator. Temporarily setting aside the singularities of a vertex operator, we show how holomorphic vertex algebras arise naturally when considering representations of the vertex group. Next we look at the role of singularities in the definition of a vertex operator. We show how to use the elementary vertex structure of our vertex group to define a space of singular functions, and after considering composition of these singular functions we define spaces of singular functions parameterised by binary trees. Our definitions, while motivated by the Borcherds definition in \cite{bor}, are different because they emphasise the relation of singularities on inputs. For the classical vertex group, we demonstrate the correspondence between a binary singular functions and vertex operators. In the next section we consider the axioms for our vertex algebra. After discussing the types of multi maps which arise for one leafed trees, we see that the vacuum axioms lead us to consider maps associated to trees with zero leaves. We show how the locality condition for vertex operators suggests the notion of singular maps associated to a three leafed tree with no internal vertices. We finish this section by defining singular maps associated to an arbritrary tree, and that natural maps between $n$ leafed trees give rise to natural maps between corresponding singular functions. We finish this paper by describing the general categorical structure satisfied by these singular maps. They form a relaxed multilinear category, and we show that a vertex algebra is just an algebra in this category. For clarity, our presentation does not give the supersymmetric details, but the generalisation can be made easily. \subsection{Notation} Throughout this paper we will often refer to modules of polynomials, power series, and Laurent series. Given any $R$\mbox{-} module, $B$ (where $R$ is a commutative ring with 1), our convention will be to denote the collection of power series with coefficients in $B$ by $B\ps{x}$, and the collection of Laurent series by $B\ps{x}[x^{-1}]$. The notation can be combined to form larger modules such as $B\ps{x}[x^{-1}]\ps{y}$, the collection of power series in the variable $y$, whose coefficients are Laurent series in the variable $x$. Note that $B\ps{x}[x^{-1}]\ps{y} \supsetneq B\ps{x,y}[x^{-1}]$ because, for example, $\sum_{i \geq 0} x^{-i} y^i$ is contained in the first module but not the second. We also adopt Sweedler's notation \cite{sweedler} to denote comultiplication by $\Delta(h) = \sum_{(h)} h_{(1)} \mbox{$\otimes$} h_{(2)}$ for any coalgebra element $h$. \subsection*{Acknowledgements} I am very grateful to Martin Hyland for his time and patience with this project. I would also like to thank Richard Borcherds for frequently fielding my detailed questions about his idea of a vertex algebra. And, a special thanks goes to Tom Leinster for the many discussions related to the completion of this paper, and especially for taking the time to show me his definition of the category of trees. Also a great thanks to the creators of \Xy-pic who made the tree diagrams possible. \section{Vacuum and Locality - Extending from Binary Trees} Now that we have defined spaces of (generalised) vertex operators and their composites, we shall return our attention to the remaining axioms for a vertex algebra. In particular, we will be interested in incorporating the vacuum axioms of equations \eqref{e:vac_triv_axiom}, \eqref{e:vac_id_axiom}, and \eqref{e:vac_ps_axiom} into our description, and making sense of the locality axiom of equation \eqref{e:locality}. But first, we return to the case of singular maps between $G$\mbox{-} modules $A$ and $B$, and parameterise them by non-branching trees. \subsection{Non-branching Trees}\label{ss:non_branch_trees} When we first considered singular multilinear maps in section \ref{ss:sing_mult_maps}, we defined $\ensuremath{\mathit{Multi}}^K(A;B)$ to be $\ensuremath{\mathrm{Hom}}_G(A,\ensuremath{\mathrm{Hom}}(G,B))$ (see definition \ref{d:sing_1and2}). Now that we have singular multilinear maps parameterised by trees, we shall think of this collection of maps as being associated to the flat tree with one leaf, \epsfig{file=Images/1_flat.eps,height=3mm}. (For a review of definitions related to trees, see appendix \ref{a:cat_of_trees} Thus we have \[\ensuremath{\mathit{Multi}}^K_{\ \epsfig{file=Images/1_flat.eps,height=3mm}}(A;B) = \ensuremath{\mathit{Multi}}^K(A;B) = \ensuremath{\mathrm{Hom}}_G(A,\ensuremath{\mathrm{Hom}}(G,B)).\] In fact, we can extend the construction of $\ensuremath{\mathit{Multi}}^K_{p}$, described in definition \ref{d:bin_sing_maps} for binary trees, to trees with non-branching subtrees. We do this by simply extending the definition of $\ensuremath{\mathrm{Sing}}_{q}$ to those internal nodes $q$ with only one incoming edge. \begin{dfn} \label{d:1_sing_maps} If the tree $p$ in definition \ref{d:bin_sing_maps} is allowed to also have non-branching subtrees, then $\ensuremath{\mathit{Multi}}^K_p(A_1, \ldots, A_n;C)$ is defined exactly as in that definition except that when an internal vertex, $p_i$ has only one incoming edge, we define an operator to act on $G$\mbox{-} modules, \[\ensuremath{\mathrm{Sing}}_{p_i} = \ensuremath{\mathrm{Hom}}_G(X_{p_i},\cdot),\] where $X_{p_i}$ is the label of the incoming node as in section \ref{ss:multimaps_binary}. \end{dfn} \begin{lemma} The definition of $\ensuremath{\mathit{Multi}}^K_{\ \epsfig{file=Images/1_flat.eps,height=3mm}}(A;B)$ coincides with definition \ref{d:1_sing_maps} applied to the tree \epsfig{file=Images/1_flat.eps,height=3mm}. \end{lemma} \begin{proof} Repeating the construction of definition \ref{d:1_sing_maps} for the tree \epsfig{file=Images/1_flat.eps,height=3mm}, we have \[\ensuremath{\mathit{Multi}}^K_{\ \epsfig{file=Images/1_flat.eps,height=3mm}}(A;B) = \ensuremath{\mathrm{Hom}}_G(A, \ensuremath{\mathrm{Hom}}(G, B)) \cong \ensuremath{\mathrm{Hom}}(A,B) \] as defined above. We also see that the $G$\mbox{-} invariant maps are the collection, \[\ensuremath{\mathit{Multi}}^K_{G,\epsfig{file=Images/1_flat.eps,height=3mm}}(A;B) = \ensuremath{\mathrm{Hom}}_G(A,\ensuremath{\mathrm{Hom}}_G(G,B)) \cong \ensuremath{\mathrm{Hom}}_G(A,B).\]\end{proof} \begin{example} If $p$ is the non-branching tree with $n$ internal nodes (including the root) then \[\ensuremath{\mathit{Multi}}^K_p(A;B) = \ensuremath{\mathrm{Hom}}_G(A,\bigl(\ensuremath{\mathrm{Sing}}_{\ \epsfig{file=Images/1_flat.eps,height=3mm}}\bigr)^{(n-1)} \ensuremath{\mathrm{Hom}}(G,B)).\] Taking advantage of the $G$\mbox{-} invariance of $\ensuremath{\mathrm{Sing}}_{\ \epsfig{file=Images/1_flat.eps,height=3mm}}$, we see that this collection is isomorphic to $\ensuremath{\mathrm{Hom}}_G(A,\ensuremath{\mathrm{Hom}}(G,B))$ in the usual way. \end{example} \begin{example} If we take $G$ to be the classical vertex group, then $\ensuremath{\mathit{Multi}}^K_{\ \epsfig{file=Images/1_flat.eps,height=2mm}}(A;B) \cong \ensuremath{\mathrm{Hom}}_{G}(A, B\ps{x}).$ (c.f., example \ref{ex:sing_1_hom}.) For the non-branching tree with $n$ internal nodes, $p$, we have $\ensuremath{\mathit{Multi}}^K_p(A;B) \cong \ensuremath{\mathrm{Hom}}(A, B\ps{x_1, \ldots, x_n}),$ where the $G$\mbox{-} invariance of $\ensuremath{\mathrm{Sing}}_{\ \epsfig{file=Images/1_flat.eps,height=3mm}}$ means that any such map is an element of the subcollection, $\ensuremath{\mathrm{Hom}}(A, B\ps{x_1+ \cdots + x_n})$. \end{example} \begin{example}\label{ex:binary_with_tail} Consider the singular maps associated to the tree, \epsfig{file=Images/2_tail.eps,height=3mm}. From the definition we have \begin{eqnarray*} \ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/2_tail.eps,height=3mm}}(A_1,A_2;B) &=& \ensuremath{\mathrm{Hom}}_{G,\G{2}}\Bigl(A_1\mbox{$\otimes$} A_2, K\mbox{$\otimes$}\ensuremath{\mathrm{Hom}}_G\bigl(\G{2},\ensuremath{\mathrm{Hom}}(G,B)\bigr)\Bigr)\\ &\cong& \ensuremath{\mathrm{Hom}}_{G,\G{2}}\Bigl(A_1\mbox{$\otimes$} A_2, K\mbox{$\otimes$}\ensuremath{\mathrm{Hom}}\bigl(\G{2},B\bigr)\Bigr). \end{eqnarray*} When $G$ is the classical vertex group this says that there is a bijection between the collection of multilinear singular maps associated to the following trees: \begin{equation}\label{e:binary_with_tail} \twoleafedtail{A_1}{A_2}{B}{x}{y}{z} \cong \flattwoleafed{A_1}{A_2}{B}{x+z}{y+z} \end{equation} This isomorphism follows immediately from the $G$\mbox{-} invariance at the internal node, where it provides the relation $\partial_x +\partial_y = \partial_z$. The map between these spaces is just given by binomial expansion as in example \ref{ex:getting_singmaps}, and the singularity remains unaffected. When these maps are fully $G$\mbox{-} invariant, the single edge attached to the bottom of the tree corresponds to operating on each singular map with \[e^{Tz} = \sum_{i\geq 0} \D{i}z^i:B \longrightarrow B\ps{z}.\] \end{example} We finish this section by noting in passing that composing any tree with the tree consisting of a single node, $\bullet$, leaves the tree unchanged. So we would like to define $\ensuremath{\mathit{Multi}}^K_\bullet$ so that it composes with a singular map of type $p$ (for some tree, $p$) to give a singular map of type $p$. This suggests the following definition: \begin{dfn}\label{d:id_sing_maps} For any $G$\mbox{-} module, $A$, the singular multilinear maps associated to the tree $\bullet$ are just the endomorphisms of $A$. \[\ensuremath{\mathit{Multi}}^K_\bullet(A;A) = \ensuremath{\mathrm{Hom}}(A,A).\] \end{dfn} \begin{lemma} The definition of $\ensuremath{\mathit{Multi}}^K_\bullet(A;A)$ coincides with definition \ref{d:1_sing_maps} applied to the tree $\bullet$. \end{lemma} \begin{proof} Definition \ref{d:1_sing_maps} applied to $\bullet$ gives $\ensuremath{\mathit{Multi}}^K_\bullet(A;A) = \ensuremath{\mathrm{Hom}}(A,A))$ as desired. \end{proof} \subsection{Vacuum - Trees with no Leaves}\label{ss:vacuum} Now that we have a description of unary singular maps, it is natural to consider the nullary type multi maps. By way of motivation, recall that the vacuum was defined to be a distinguished vector denoted $\mbox{$|0\rangle$} \in V$ such that for any $a \in V$, the vacuum satisfies: \begin{eqnarray*} T\mbox{$|0\rangle$} &=& 0 \\ Y(\mbox{$|0\rangle$}, x)a &=& a \\ Y(a,x)\mbox{$|0\rangle$}|_{x=0} &=& a. \end{eqnarray*} We saw in section \ref{ss:extended} that for vertex algebras without singularities, the vacuum acted as a unit object. Later we shall show that even with singularities, the vacuum will continue to act as a unit. But in this section we are concerned primarily with showing how the collection of all possible vacuum vectors arises naturally when considering singular maps parameterised by trees. They arise naturally as the multilinear singular map associated to the empty tree (which we denote $\circ$). Applying definition \ref{d:1_sing_maps}, to the case of the empty tree, we construct the collection of multilinear singular maps by first augmenting the empty tree, giving \epsfig{file=Images/0_flat.eps,height=3mm}. As with the tree $\bullet$, the only internal node is $\bot$. Thus we have $X_\bot = R$ since it is not connected to any other leaf or internal node (the empty node is not counted as a leaf), and so we have: \begin{dfn}\label{d:0_sing_maps} The multilinear singular maps parameterised by the empty tree are given by: \[ \ensuremath{\mathit{Multi}}^K_{\circ}(R;V) = \ensuremath{\mathrm{Hom}}(R,V) \cong V.\] \end{dfn} The $G$\mbox{-} invariant elements of this collection are the vectors in $V$ with trivial $G$\mbox{-} action, so we see that a $G$\mbox{-} invariant singular multilinear map of type $\circ$ is just a vector $v \in V$ satisfying the first vacuum axiom. What happens when the vacuum vector is used as an input for a binary singular multi map? We begin with an example for the classical vertex group. \begin{example}\label{ex:compose_with_vac} Let $G$ be the classical vertex group and $f$ be a binary singular map in \[\ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/2_leaf.eps,height=2mm}}(A_1,A_2;B) \cong \ensuremath{\mathrm{Hom}}_{\G{2}}(A_1\mbox{$\otimes$} A_2, B\ps{x,y}[(x-y)^{-1}]).\] We know that for a vector $v \in A_1$ arising as above, we have $Tv=0$, so given any $a \in A_2$ we have, \[\partial_x f(v\mbox{$\otimes$} a)= f(Tv\mbox{$\otimes$} a)=0.\] So $f(v\mbox{$\otimes$} a) \in B\ps{y}$. We see further what if we assume $f$ to be $G$\mbox{-} invariant, we have that $Tf(v\mbox{$\otimes$} a) = \partial_y f(v\mbox{$\otimes$} a)$, and so writing $f(v\mbox{$\otimes$} a) = \sum_{i\geq 0} b_i y^i$, we see that $Tb_i=(i+1)b_{i+1}$, and so we have \begin{equation}\label{e:exp_T} f(v\mbox{$\otimes$} a) = \sum_{i\geq 0} \frac{T^i}{i!}b_0 y^i = e^{Ty}b_0. \end{equation} But we saw in the previous section that $\ensuremath{\mathit{Multi}}^K_{\ \epsfig{file=Images/1_flat.eps,height=2mm}}(A_2;B) \cong \ensuremath{\mathrm{Hom}}_{G}(A_2, B\ps{y}),$ and the $G$\mbox{-} invariant maps in this collection were of the form of equation \eqref{e:exp_T}. Thus we see that we have a composition map \begin{equation}\label{e:null_comp} \ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/2_leaf.eps,height=2mm}}(A_1,A_2;B) \mbox{$\otimes$} \ensuremath{\mathit{Multi}}^K_{G,\circ}(R;A_1) \longrightarrow \ensuremath{\mathit{Multi}}^K_{\ \epsfig{file=Images/1_flat.eps,height=2mm}}(A_2;B), \end{equation} and $G$\mbox{-} invariant binary singular maps compose to give $G$\mbox{-} invariant maps $\ensuremath{\mathit{Multi}}^K_{G,\epsfig{file=Images/1_flat.eps,height=2mm}}(A_2;B)$. Notice that if $A_1 = A_2 = B = V$ for some complex vector space $V$, then equation \eqref{e:exp_T} gives the remaining vacuum axioms when $b_0 = a$. We shall see later that this will hold for a suitable algebra. \end{example} This is a very satisfactory notion for composition because we have taken a binary tree and composed it with the empty node to give a tree with only one leaf. In fact, this holds equally well for an arbritrary vertex group, with equation \eqref{e:null_comp} holding for any $G$\mbox{-} modules $A_1, A_2$ and $B$. For the general theory see \cite{cts_valgdetails}. \subsection{Locality}\label{ss:locality} We have attempted to be quite clear throughout this paper as to the ramifications of our description for the classical vertex group. We would now like to apply the locality axiom to the classical vertex group story so far, and interpret it in the abstract presentation. Recall that the locality axiom says that for any $a, b \in V$, there exists some $N \gg 0$ such that the following holds: \[(x-y)^N[Y(a,x), Y(b, y)] = 0.\] In other words, the following two maps are equal: \begin{eqnarray*} (x-y)^NY(a,x)Y(b, y)\cdot:V &\longrightarrow& V\ps{x}[x^{-1}]\ps{y}[y^{-1}]\label{e:loc_xy}\\ (x-y)^NY(b,y)Y(a, x)\cdot:V &\longrightarrow& V\ps{y}[y^{-1}]\ps{x}[x^{-1}]\label{e:loc_yx}. \end{eqnarray*} In light of our previous discussion showing how $Y(a,x)Y(b, y)\cdot$ and $Y(b,y)Y(a, x)\cdot$ are actually elements of different spaces of singular maps, this presents us with the question of how to interpret such an equality of maps. The usual way to interpret them is by considering them inside the space of formal power series $V\ps{x, x^{-1},y, y^{-1}}$. But since both of the spaces of power series in equations \eqref{e:loc_yx} and \eqref{e:loc_xy} are properly contained in the larger space of formal power series, they can only be equal if they map to the intersection $V\ps{y}[y^{-1}]\ps{x}[x^{-1}] \cap V\ps{x}[x^{-1}]\ps{y}[y^{-1}]$. \begin{lemma} $V\ps{y}[y^{-1}]\ps{x}[x^{-1}] \cap V\ps{x}[x^{-1}]\ps{y}[y^{-1}] = V\ps{x,y}[x^{-1}, y^{-1}].$ \end{lemma} \begin{proof} It is clear that the right hand side is contained in the intersection, so we need only to prove that an arbitrary element of the intersection is contained in the power series on the right. But the only difference between the power series $V\ps{x}[x^{-1}]\ps{y}[y^{-1}]$, and $V\ps{x,y}[x^{-1}, y^{-1}]$, is that in the former, polynomials in the variable $x^{-1}$ can exist as coefficients of the power series in the variable $y$. But this can not occur in $V\ps{y}[y^{-1}]\ps{x}[x^{-1}]$, so the intersection is as given. \end{proof} So we see that for some $N \gg 0$, $(x-y)^NY(a,x)Y(b, y)\cdot$ and $(x-y)^NY(b,y)Y(a, x)\cdot$ are equal as maps from $V$ to $V\ps{x,y}[x^{-1}, y^{-1}]$. Fixing $N$, we denote this map $g(a\mbox{$\otimes$} b\mbox{$\otimes$} \cdot)$. Considering $g$ as a map to the larger space, $V\ps{x,y}[x^{-1}, y^{-1}, (x-y)^{-1}],$ of power series with the inverse of $(x-y)$ adjoined. Then the map, $(x-y)^{-N}g(a\mbox{$\otimes$} b\mbox{$\otimes$} \cdot)$, is well defined on $V$. Notice first that this map is not necessarily equal to $Y(a,x)Y(b, y)\cdot$ or $Y(b,y)Y(a, x)\cdot$, because $(x-y)^{-N}$ does not exist in the codomain of either of these maps. But, if we expand $(x-y)^{-N}$ as a power series in either the variable $x$ or $y$, we have maps \begin{equation}\label{e:refine_from_three} \xymatrix{ & V\ps{x,y}[x^{-1},y^{-1}, (x-y)^{-1}] \ar[ld]_{i_{x,y}} \ar[rd]^{i_{y,x}}&\\ V\ps{x}[x^{-1}]\ps{y}[y^{-1}] & &V\ps{y}[y^{-1}]\ps{x}[x^{-1}]} \end{equation} which are identities except on $(x-y)^{-1}$, where they are given by \begin{eqnarray*} i_{x,y}\bigl((x-y)^{-j-1}\bigr) & = & \sum_{n \in {\mathbb N}} \binom{n+j}{j} x^{-n-j-1} y^n \\ i_{y,x}\bigl((x-y)^{-j-1}\bigr) & = & (-1)^{j+1}\sum_{n \in {\mathbb N}} \binom{n+j}{j} x^{n} y^{-n-j-1}, \end{eqnarray*} then we see that under these maps, the singular map $(x-y)^{-N}g(a\mbox{$\otimes$} b\mbox{$\otimes$} \cdot)$ canonically maps down to the two composites of vertex operators: \begin{eqnarray*} i_{x,y}\bigl((x-y)^{-N}g(a\mbox{$\otimes$} b\mbox{$\otimes$} \cdot)\bigr)&=& Y(a,x)Y(b, y)\cdot\\ i_{y,x}\bigl((x-y)^{-N}g(a\mbox{$\otimes$} b\mbox{$\otimes$} \cdot)\bigr)&=& Y(b,y)Y(a, x)\cdot. \end{eqnarray*} So we have constructed an element of the following collection of singular maps (with appropriate $G$\mbox{-} invariance taken): \begin{equation*} \ensuremath{\mathrm{Hom}}\biggl(V\mbox{$\otimes$} V, \ensuremath{\mathrm{Hom}}\Bigl(V, V\ps{x,y}[x^{-1}, y^{-1}]\Bigr)[(x-y)^{-1}]\biggr), \end{equation*} where we have taken special care to emphasise that the singularity at $(x-y)$ depends only upon the outer two copies of $V$. We shall see shortly that this dependence of singularities on inputs is a crucial feature of the more general theory. But first we examine the effect of the vacuum on these composites. \begin{example} Consider the action of $(x-y)^{-N}g(a\mbox{$\otimes$} b\mbox{$\otimes$} \cdot)$ on the vacuum, $\mbox{$|0\rangle$}$. From its definition we have \begin{eqnarray*} (x-y)^{-N}g(a\mbox{$\otimes$} b\mbox{$\otimes$} \mbox{$|0\rangle$}) &=& (x-y)^{-N}(x-y)^NY(a,x)Y(b, y)\mbox{$|0\rangle$}\\ &=& (x-y)^{-N}(x-y)^NY(b,y)Y(a, x)\mbox{$|0\rangle$}, \end{eqnarray*} and we know from the vacuum axioms that $Y(a, x)\mbox{$|0\rangle$}$ and $Y(b, y)\mbox{$|0\rangle$}$ have no singularities, therefore $(x-y)^{-N}g(a\mbox{$\otimes$} b\mbox{$\otimes$} \mbox{$|0\rangle$})$ is an element of $V\ps{x,y}[(x-y)^{-1}]$. From our previous discussion we know that it maps down to the composites \begin{eqnarray*} i_{x,y}\bigl((x-y)^{-N}g(a\mbox{$\otimes$} b\mbox{$\otimes$} \mbox{$|0\rangle$})\bigr)&=& Y(a,x)e^{Ty}b\\ i_{y,x}\bigl((x-y)^{-N}g(a\mbox{$\otimes$} b\mbox{$\otimes$} \mbox{$|0\rangle$})\bigr)&=& Y(b,y)e^{Tx}a, \end{eqnarray*} which are elements of $V\ps{x}[x^{-1}]\ps{y}$ and $V\ps{y}[y^{-1}]\ps{x}$ respectively. Recall that in example \ref{ex:getting_singmaps} we showed how to construct a singular map from a vertex operator. \begin{lemma} The binary singular power series given by regarding $e^{Ty}Y(a,x-y)b$ as an element of $V\ps{x,y}[(x-y)^{-1}]$ is equal to $(x-y)^{-N}g(a\mbox{$\otimes$} b\mbox{$\otimes$} \mbox{$|0\rangle$})$. \end{lemma} \begin{proof} Each of these is an element of $V\ps{x,y}[(x-y)^{-1}]$ which agree when $x=0$ (and when $y=0$). Because each satisfies $\partial_x +\partial_y = \partial_V$ they are completely determined by their values at $x=0$ (or $y=0$), and so are equal. \end{proof} \end{example} Just as we looked at the action of $(x-y)^{-N}g(a\mbox{$\otimes$} b\mbox{$\otimes$} \cdot)$ on the vacuum, we consider its action on an arbritrary $c \in V$. From the vacuum axioms, we know that $c = Y(c,z)\mbox{$|0\rangle$}|_{z=0}$, so in fact we shall consider \[(x-y)^{-N}g\Bigl(a\mbox{$\otimes$} b\mbox{$\otimes$} Y(c,z)\mbox{$|0\rangle$}\Bigr)\Bigr|_{z=0} = (x-y)^{-N}(x-y)^NY(a,x)Y(b, y)Y(c,z)\mbox{$|0\rangle$}|_{z=0}.\] Using the locality axiom, we know that for some $M\gg 0$ the following equality holds for any $a \in V$: \[(y-z)^MY(a,x)Y(b, y)Y(c,z)\mbox{$|0\rangle$} = (y-z)^MY(a,x)Y(c,z)Y(b,y)\mbox{$|0\rangle$}.\] Setting $z=0$, we see that as an element of $V\ps{x,y}[(x-y)^{-1},x^{-1},y^{-1}]$, the singularity $y^{-1}$ of $(x-y)^{-N}g(a\mbox{$\otimes$} b\mbox{$\otimes$} c)$ depends only on $b$ and $c$. Similarly, the singularity $x^{-1}$ depends only on $a$ and $c$. Recalling how when we defined the binary singular maps, we defined maps $f(a\mbox{$\otimes$} b)$ such that setting $y=0$ gave $Y(a,x)b$ and setting $y=0$ gave $Y(b,y)a$, we see that a more natural definition for $g$ would be as a map to $V\ps{x,y,z}[(x-y)^{-1}, (y-z)^{-1},(x-z)^{-1}]$. For any $a,b,c \in V$ we may define it (with $N\gg 0$ as before): \[(x-y)^{-N}(x-y)^{N}\Bigl(e^{Tz}Y(a,x-z)Y(b, y-z)c\Bigr).\] Thus we have a ternary function which, according to the dependence of singularities on inputs discussed above, can we regarded as an element of: \begin{eqnarray} \ensuremath{\mathrm{Hom}}\biggl(V\mbox{$\otimes$} V, \ensuremath{\mathrm{Hom}}\Bigl(V, V\ps{x,y,z}[(x-z)^{-1}, (y-z)^{-1}]\Bigr)[(x-y)^{-1}]\biggr),\label{e:ternary_1} \\ \ensuremath{\mathrm{Hom}}\biggl(V\mbox{$\otimes$} V, \ensuremath{\mathrm{Hom}}\Bigl(V, V\ps{x,y,z}[(x-z)^{-1}, (x-y)^{-1}]\Bigr)[(y-z)^{-1}]\biggr),\label{e:ternary_2} \\ \ensuremath{\mathrm{Hom}}\biggl(V\mbox{$\otimes$} V, \ensuremath{\mathrm{Hom}}\Bigl(V, V\ps{x,y,z}[(y-z)^{-1}, (x-y)^{-1}]\Bigr)[(x-z)^{-1}]\biggr).\label{e:ternary_3} \end{eqnarray} All three of these collections of singular maps are properly contained in the larger collection, \begin{equation} \ensuremath{\mathrm{Hom}}\biggl(V\mbox{$\otimes$} V \mbox{$\otimes$} V, V\ps{x,y, z}[(x-z)^{-1}, (y-z)^{-1}, (x-y)^{-1}]\biggr),\label{e:ternary_4} \end{equation} where the dependence of each singularity on input is ignored. So we are led to define a new collection of multi maps which we associate to the flat three leafed tree, \epsfig{file=Images/flat_3.eps,height=3mm}, as the pullback of the spaces in equations \eqref{e:ternary_1}, \eqref{e:ternary_2}, and \eqref{e:ternary_3} over the space in equation \eqref{e:ternary_4}. We shall denote it $\ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/flat_3.eps,height=3mm}}(V,V,V;V)$. In fact, this collection formalises the idea of the \defn{operator product expansion} (see \cite[section 4.6]{kac}). \begin{note} We have tacitly been using the isomorphism of trees of equation \ref{e:binary_with_tail}. \end{note} \subsection{Flat Trees and Singular Maps} In the previous section we used our knowledge of axiomatic vertex algebras to define a collection of multi maps associated to a three leaved flat tree. Following that example, we now provide a more general description of singular maps for an arbritrary vertex group associated to any $n$\mbox{-} leafed flat tree. This definition arises naturally from the previous discussion, and incorporates the definitions provided for trees with $0,1$ and $2$ leaves. \begin{dfn}\label{d:n_flat_sing_maps} If $G$ is a vertex group and $A_1, \ldots, A_n, B$ are $G$\mbox{-} modules, then the collection of singular maps associated to the flat tree with $n$ leaves, \epsfig{file=Images/generic.eps,height=3mm}, is denoted $\ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/generic.eps,height=3mm}}(A_1, \ldots, A_n; B)$, and is defined to be the pullback of the following singular maps for $\sigma \in A_n$ (the alternating group): \begin{multline}\label{e:nary_general_pbk} \ensuremath{\mathrm{Sing}}_\sigma(A_1, \ldots, A_n; \ensuremath{\mathrm{Hom}}(\G{n},B)) =\\ \ensuremath{\mathrm{Hom}}\biggl(A_{\sigma(1)}\mbox{$\otimes$} A_{\sigma(2)}, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl(A_{\sigma(3)}, K^{\mbox{$\otimes$} 2}\mbox{$\otimes$} \bigl(\cdots K^{\otimes n-1} \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{n},B)\cdots\bigr)\Bigr)\biggr), \end{multline} over the collection, \begin{equation}\label{e:nary_general} \ensuremath{\mathrm{Hom}}_{\G{n}}\biggl(A_1\mbox{$\otimes$} \cdots \mbox{$\otimes$} A_n, \ensuremath{\mathit{Fun}}(\G{n}, B)\biggr) = \ensuremath{\mathrm{Hom}}_{\G{n}}\biggl(A_1\mbox{$\otimes$} \cdots \mbox{$\otimes$} A_n, K^{\mbox{$\otimes$} \binom{n}{2}} \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{n},B)\biggr), \end{equation} where equation \eqref{e:nary_general_pbk} is taken to be invariant under the action of $\G{n}$ inferred from its action on equation \eqref{e:nary_general}. The singularities of equation \eqref{e:nary_general_pbk} are the singularities are tensored over $f_{i,j}$ as in definition \ref{d:fun}, of $\ensuremath{\mathit{Fun}}(\G{n}, B)$. We have \defn{$G$\mbox{-} invariant multilinear singular maps} exactly when $\ensuremath{\mathrm{Hom}}(\G{n},B)$ is $G$\mbox{-} invariant. \end{dfn} \begin{example} When $n=0$, this definition says that $\ensuremath{\mathit{Multi}}^K_{\circ}(R;B) = \ensuremath{\mathrm{Hom}}(R,B)$ as expected, and when $n=1$, we have that $\ensuremath{\mathit{Multi}}^K_{\ \epsfig{file=Images/1_flat.eps,height=3mm}}(A;B) = \ensuremath{\mathrm{Hom}}_G(A,\ensuremath{\mathrm{Hom}}(G,B))$ as expected. \end{example} \begin{example} When $n=2$, we see that $\ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/2_leaf.eps,height=3mm}}(A_1,A_2;B) = \ensuremath{\mathrm{Hom}}_{\G{2}}(A_1 \mbox{$\otimes$} A_2, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2},B))$ as before. \end{example} \begin{example}\label{d:ternary_multimaps} If $G$ is a vertex group and $A_1, A_2, A_3, B$ are $G$\mbox{-} modules, then the collection of singular maps, $\ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/flat_3.eps,height=3mm}}(A_1, A_2, A_3; B)$, is the pullback of the following three collections of singular maps: \begin{eqnarray} \ensuremath{\mathrm{Hom}}\biggl(A_1\mbox{$\otimes$} A_2, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl(A_3, K^{\mbox{$\otimes$} 2}\mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{3},B)\Bigr)\biggr),\label{e:ternary_general_1} \\ \ensuremath{\mathrm{Hom}}\biggl(A_2\mbox{$\otimes$} A_3, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl(A_1, K^{\mbox{$\otimes$} 2}\mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{3},B)\Bigr)\biggr),\label{e:ternary_general_2} \\ \ensuremath{\mathrm{Hom}}\biggl(A_3\mbox{$\otimes$} A_1, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl(A_2, K^{\mbox{$\otimes$} 2}\mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{3},B)\Bigr)\biggr),\label{e:ternary_general_3} \end{eqnarray} over the collection, \begin{equation} \ensuremath{\mathrm{Hom}}_{\G{3}}\biggl(A_1\mbox{$\otimes$} A_2 \mbox{$\otimes$} A_3, \ensuremath{\mathit{Fun}}(\G{3}, B)\biggr) = \ensuremath{\mathrm{Hom}}_{\G{3}}\biggl(A_1\mbox{$\otimes$} A_2 \mbox{$\otimes$} A_3, K^{\mbox{$\otimes$} 3} \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{3},B)\biggr),\label{e:ternary_general_4} \end{equation} with appropriate $G$\mbox{-} invariance taken. We see immediately that the definition of $\ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/flat_3.eps,height=3mm}}(V,V,V;V)$ for the classical vertex group follows immediately. \end{example} \subsection{Multi Maps Parameterised by All Trees} After extending our definition of singular maps to trees with many leaves, we are ready to define the collection of singular maps associated to an arbitrary tree. In this section we give the general definition for an arbritrary vertex group, and show that singular maps compose as desired for the classical vertex group. This definition generalises definitions \ref{d:bin_sing_maps}, \ref{d:1_sing_maps} and \ref{d:n_flat_sing_maps}. For clarity we give it in its complete form here. Recall from section \ref{ss:multimaps_binary} that a tree is said to be \defn{augmented} when we have added an additional vertex and edge to the root of the tree, such that the new vertex inherits the former root label, and the old vertex is labelled $\G{n}$ where $n$ is the number of incoming nodes. We denote the new vertex $\bot$. Also recall that to each internal node, $q$, of a tree, we can associate the tensor product of the labels of the incoming nodes, and denote it $X_q = X_{q,1} \mbox{$\otimes$} \cdots \mbox{$\otimes$} X_{q,n}$. Thus we have $X_\bot = \G{n}$. \begin{dfn} If $G$ is a vertex group and $A_1, \ldots, A_n, B$ are $G$\mbox{-} modules, then the collection of singular maps associated to an arbritrary tree, $p$, with $n$ leaves is denoted, $\ensuremath{\mathit{Multi}}^K_p(A_1, \ldots, A_n; B)$, and is defined as follows: \begin{itemize} \item Let $t$ denote a total ordering, $\bot < \ensuremath{\mathit{root}} < p_1 < \cdots < p_l$, of the internal vertices of augmented $p$, compatible with the the partial ordering inherited from the tree structure of $p$. For each internal vertex, $p_i$, denote the number of incoming edges $n_i$. Let $\sigma \in A_{n_i}$ and define an operator taking $\G{n_i}$\mbox{-} modules to the suitably $\G{n_i}$\mbox{-} invariant maps: \begin{equation} \ensuremath{\mathrm{Sing}}_{\sigma,p_i}\cdot= \ensuremath{\mathrm{Hom}}\biggl(X_{p_i,\sigma(1)}\mbox{$\otimes$} X_{p_i,\sigma(2)}, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl(X_{p_i,\sigma(3)}, K^{\mbox{$\otimes$} 2}\mbox{$\otimes$} \bigl(\cdots K^{\otimes n_i-1} \mbox{$\otimes$} (\cdot)\cdots\bigr)\Bigr)\biggr). \end{equation} The tensor products of the singularities, $K$, are taken over appropriate copies of $f_{i,j}$ as in definition \ref{d:fun}. \item Let $\ensuremath{\mathrm{Sing}}_{p_i}$ denote the pullback of $\ensuremath{\mathrm{Sing}}_{\sigma,p_i}$ for all $\sigma \in A_{n_i}$ over $\ensuremath{\mathrm{Hom}}\Bigl(X_{p_i},K^{\otimes \binom{n_i}{2}} \mbox{$\otimes$} \cdot\Bigr)$. This can be thought of as giving the multilinear singular maps associated to the flat subtree with root $p_i$. \item Iterating this operator for all internal vertices of $p$, we have \[ \mathrm{Ord}_t(A_1, \ldots, A_n, B) = \ensuremath{\mathrm{Sing}}_{p_l}\cdots \ensuremath{\mathrm{Sing}}_{p_1} \ensuremath{\mathrm{Sing}}_{\ensuremath{\mathit{root}}} \ensuremath{\mathrm{Hom}}(X_\bot,B).\] \end{itemize} Then $\ensuremath{\mathit{Multi}}^K_p(A_1, \ldots, A_n; B)$ is defined to be the wide pullback of each $\mathrm{Ord}_t$ for all possible total orderings, $t$, of of the internal vertices of augmented $p$, over \begin{equation}\label{e:pull_over_me} \ensuremath{\mathrm{Hom}}_{\G{n}}\Bigl(A_1\mbox{$\otimes$}\cdots\mbox{$\otimes$} A_n, \ensuremath{\mathit{Fun}}_{p}(\G{n},B)\Bigr) \end{equation} where $\ensuremath{\mathit{Fun}}_{p}(\G{n},B)$ is the collection of singular functions as defined in \cite{bor}. See appendix \ref{ss:bor_sing_maps} for more details. We have \defn{$G$\mbox{-} invariant multilinear singular maps}, $\ensuremath{\mathit{Multi}}^K_{G,p}$, exactly when $\ensuremath{\mathrm{Hom}}(X_\bot,B)$ is $G$\mbox{-} invariant. \end{dfn} \begin{note} Because limits commute with one another, we could avoid forming the pullback $\ensuremath{\mathrm{Sing}}_{p_i}$ by instead defining \begin{equation}\label{e:ord_plus_perms} \mathrm{Ord}_{t,\sigma_I}(A_1, \ldots, A_n, B) = \ensuremath{\mathrm{Sing}}_{p_l, \sigma_l}\cdots \ensuremath{\mathrm{Sing}}_{p_1, \sigma_1}\ensuremath{\mathrm{Sing}}_{\ensuremath{\mathit{root}}, \sigma_0} \ensuremath{\mathrm{Hom}}(X_\bot,B) \end{equation} for each $\sigma_i \in A_{n_i}$, and taking the wide pullback for all possible total orderings, $t$, and permutations $\sigma_i \in A_{n_i}$ for $1\leq i \leq l$. \end{note} \begin{example} We see automatically that if $p$ is a flat tree, then there is only one total ordering, and so $\ensuremath{\mathit{Multi}}^K_p(A_1, \ldots, A_n; B)$ reduces to the pullback of $\ensuremath{\mathrm{Sing}}_{\ensuremath{\mathit{root}}} \ensuremath{\mathrm{Hom}}(X_\bot,B)$, over the maps in equation \eqref{e:pull_over_me}, which is just definition \ref{d:n_flat_sing_maps}. \end{example} \begin{thm} If $G$ is a vertex group over a field and $\ensuremath{\mathit{Multi}}^K_p(A_1, \ldots, A_n; B_1)$, $\ensuremath{\mathit{Multi}}^K_q(B_1, \ldots, B_m; C)$ are collections of multilinear singular maps, then they compose pointwise to give an element of \[\ensuremath{\mathit{Multi}}^K_{p\circ q}(A_1,\ldots, A_n, B_2 \ldots, B_m; C).\] \end{thm} \begin{note} Keep in mind that we are composing the trees $p$ and $q$ and not the augmented trees. We only use augmented trees for the purpose of describing their associated singular multi maps. \end{note} \begin{proof} Let $f \in \ensuremath{\mathit{Multi}}^K_p$ and $g \in \ensuremath{\mathit{Multi}}^K_q$. We shall show that $f$ and $g$ compose to give an element of $\ensuremath{\mathit{Multi}}^K_{p\circ q}$. To do so, we select a pullback object $\mathrm{Ord}_{t,\sigma_I}(A_1, \ldots, A_n, B_2 \ldots, B_m; C)$ as in equation \eqref{e:ord_plus_perms}. The total ordering of the internal vertices of the tree $p \circ q$ provides a total ordering of both $p$ and $q$ as $\ensuremath{\mathit{root}}_p < p_1 < \cdots < p_k$ and $\ensuremath{\mathit{root}}_q < q_1 < \cdots < q_l$. Such a pullback object also consists of a choice of permutation of labels for each $X_{p_i}, X_{q_j}$, and so we have uniquely determined objects, \begin{eqnarray*} \mathrm{Ord}_{t_p}(A_1, \ldots, A_n, B_1) &=& \ensuremath{\mathrm{Sing}}_{\sigma_k, p_k}\cdots \ensuremath{\mathrm{Sing}}_{\sigma_1, p_1} \ensuremath{\mathrm{Sing}}_{\sigma_0, \ensuremath{\mathit{root}}_p} \ensuremath{\mathrm{Hom}}(X_{\bot_p},B_1)\\ \mathrm{Ord}_{t_q}(B_1, \ldots, B_m, C) &=& \ensuremath{\mathrm{Sing}}_{\delta_l, q_l}\cdots \ensuremath{\mathrm{Sing}}_{\delta_1, q_1} \ensuremath{\mathrm{Sing}}_{\delta_0, \ensuremath{\mathit{root}}_q} \ensuremath{\mathrm{Hom}}(X_{\bot_q},C). \end{eqnarray*} Because $f$ and $g$ are contained in the pullbacks, we can regard them as elements of $\mathrm{Ord}_{t_p}$ and $\mathrm{Ord}_{t_q}$ respectively. We know that the internal vertices of the composed tree, $p \circ q$ consist exactly of the internal vertices of $p$, the internal vertices of $q$, and the root of $p$. We shall therefore prove that $f$ and $g$ compose to give an element of $\mathrm{Ord}_{t,\sigma_I}$ by induction on the number of internal vertices of $p \circ q$, $k+l+1$. If the composed tree $p \circ q$ has zero internal vertices, then we are composing with a tree of the form $\bullet$ or $\circ$, and it is clear from the discussions of sections \ref{ss:non_branch_trees} and \ref{ss:vacuum} that maps compose as desired. Now, without loss of generality we may assume that in the total ordering $t$, $p_k > q_l$. We need only to prove that there exists a natural map, \begin{equation}\label{e:composition_proof}\begin{split} \biggl(\ensuremath{\mathrm{Sing}}_{\sigma_k, p_k}&\cdots \ensuremath{\mathrm{Sing}}_{\sigma_1, p_1} \ensuremath{\mathrm{Sing}}_{\sigma_0,\ensuremath{\mathit{root}}_p} \ensuremath{\mathrm{Hom}}(X_{\bot_p},B_1)\biggr) \mbox{$\otimes$} \biggl(\ensuremath{\mathrm{Sing}}_{\delta_l, q_l}\cdots \ensuremath{\mathrm{Sing}}_{\delta_1, q_1} \ensuremath{\mathrm{Sing}}_{\delta_0, \ensuremath{\mathit{root}}_q} \ensuremath{\mathrm{Hom}}(X_{\bot_q},C)\biggr)\\ &\begin{split}\longrightarrow \ensuremath{\mathrm{Sing}}_{\sigma_k, p_k}\biggl(\Bigl(\ensuremath{\mathrm{Sing}}_{\sigma_{k-1}, p_{k-1}}&\cdots \ensuremath{\mathrm{Sing}}_{\sigma_1, p_1} \ensuremath{\mathrm{Sing}}_{\sigma_0,\ensuremath{\mathit{root}}_p} \ensuremath{\mathrm{Hom}}(X_{\bot_p},B_1)\Bigr) \mbox{$\otimes$} \\ &\Bigl(\ensuremath{\mathrm{Sing}}_{\delta_l, q_l}\cdots \ensuremath{\mathrm{Sing}}_{\delta_1, q_1} \ensuremath{\mathrm{Sing}}_{\delta_0, \ensuremath{\mathit{root}}_q} \ensuremath{\mathrm{Hom}}(X_{\bot_q},C)\Bigr)\biggr), \end{split}\end{split}\end{equation} because on the right hand side, the operator $\ensuremath{\mathrm{Sing}}_{\sigma_k, p_k}$ is acting on a collection of singular maps associated to a pair of trees with $k+l$ internal vertices, which we know compose by induction. We know that $\Bigl(\ensuremath{\mathrm{Sing}}_{\sigma_k, p_k}\cdots \ensuremath{\mathrm{Sing}}_{\sigma_1, p_1} \ensuremath{\mathrm{Sing}}_{\sigma_0,\ensuremath{\mathit{root}}_p} \ensuremath{\mathrm{Hom}}(X_{\bot_p},B_1)\Bigr)$ can be evaluated at $X_{p_k}$ (see appendix \ref{a:evaluation}), so the left hand side of equation \eqref{e:composition_proof} can be evaluated at $X_{p_k}$, giving the inner collection of singular maps on the right hand side. The transpose of this evaluation map provides us with the desired map. Now that we have seen that the maps $f$ and $g$ compose naturally into each pullback object, $\mathrm{Ord}_{t}$, our proof will be complete if we show that given two pullback objects, $\mathrm{Ord}_{t_1,\sigma_I}$ and $\mathrm{Ord}_{t_2,\sigma_J}$, the composite of $f$ and $g$ maps through each of them to the same element of \[\ensuremath{\mathrm{Hom}}\Bigl(A_1\mbox{$\otimes$} \cdots \mbox{$\otimes$} A_n \mbox{$\otimes$} B_2 \mbox{$\otimes$} \cdots \mbox{$\otimes$} B_m, \ensuremath{\mathit{Fun}}_{p\circ q}(\G{n+m-1},C)\Bigr),\] the collection over which we are pulling back. Denote this collection $Z$. Since both $f$ and $g$ are elements of a pullback over objects \begin{eqnarray*} X &=& \ensuremath{\mathrm{Hom}}\Bigl(A_1\mbox{$\otimes$} \cdots \mbox{$\otimes$} A_n,\ensuremath{\mathit{Fun}}^G_{p}(\G{n},B_1)\Bigr)\\ Y &=& \ensuremath{\mathrm{Hom}}\Bigl(B_1\mbox{$\otimes$} \cdots \mbox{$\otimes$} B_m,\ensuremath{\mathit{Fun}}_{q}(\G{m},C)\Bigr) \end{eqnarray*} respectively, then if there exists an injective maps from $Z$ to the composite of $X$ and $Y$, then we know that the composite of $f$ and $g$ maps to the same element of $Z$, and hence is an element of the pullback. Clearly $X$ and $Y$ compose to give an element of the collection, \[W = \ensuremath{\mathrm{Hom}}\Bigl(A_1\mbox{$\otimes$} \cdots \mbox{$\otimes$} A_n \mbox{$\otimes$} B_2 \mbox{$\otimes$} \cdots \mbox{$\otimes$} B_m, \ensuremath{\mathit{Fun}}_{p}\bigr(\G{n},\ensuremath{\mathit{Fun}}_{q}(\G{m},C)\bigr)\Bigr),\] and because we are working with a vertex group over a field, there is a natural injection from $Z$ to $W$, so our proof is complete. \end{proof} \begin{remark} Composition can be checked to be associative. \end{remark} \subsection{Refinement and Maps Between Singular Functions}\label{ss:refinement} Now that we have a definition of multilinear singular maps for every possible tree, we consider how these collections are related. Recall that in equation \eqref{e:refine_from_three} we saw that we could map from the collection of singular maps associated to the tree, \epsfig{file=Images/flat_3.eps,height=3mm}, to the collections associated to the trees \epsfig{file=Images/left_3.eps,height=3mm} and \epsfig{file=Images/right_3.eps,height=3mm} by expanding the singularity, $(x-y)^{-1}$, as a power series in either $x$ or $y$. We shall now generalise this expansion for an arbritrary vertex group and see that this is an example of a more general phenomenon where our collections of multilinear singular maps are related to one another through canonical maps. \begin{dfn} Given any ring of singular functions, $K$ (see definition \ref{d:K}), a \defn{refinement for a singularity} is the map, \[K \longrightarrow \ensuremath{\mathrm{Hom}}_G(G,K)\] which takes any $k \in K$ to the map $f \in \ensuremath{\mathrm{Hom}}_G(G,K)$ defined by $f(g) = g \mbox{$\cdot$} k$ for any $g \in G$. For any other $G$\mbox{-} module, $A$, and $G$\mbox{-} invariant map $\alpha:A \rightarrow G$, we define a refinement for $K$ by composition: \[K \longrightarrow \ensuremath{\mathrm{Hom}}_G(G,K) \xrightarrow{\alpha} \ensuremath{\mathrm{Hom}}_G(A,K). \] \end{dfn} We know that this map is well defined since $K$ is a $G$\mbox{-} module. It is also obviously a $G$\mbox{-} linear map. The following examples will show that this is just a generalisation of the idea of power series expansions. \begin{example} We saw in section \ref{ss:classicalvg} that $K = {\mathbb C} \ps{x}[x^{-1}]$. So a refinement for this singularity is a map: \begin{eqnarray*} {\mathbb C} \ps{x}[x^{-1}] &\longrightarrow &{\mathbb C} \ps{x}[x^{-1}]\ps{y} \\ x^k &\mapsto & \sum_{i \geq 0} \binom{k}{i}x^{k-i}y^i. \end{eqnarray*} In other words, refinement for this singularity is a map from $x^k$ to $(x+y)^k$ expanded as a power series in the variable $y$. \end{example} \begin{example} If we compose the refinement map with the antipode map, $S:G \rightarrow G$, we have: \begin{eqnarray*} {\mathbb C} \ps{x}[x^{-1}] &\longrightarrow &{\mathbb C} \ps{x}[x^{-1}]\ps{y} \\ x^k &\mapsto & \sum_{i \geq 0} \binom{k}{i}(-1)^i x^{k-i} y^i, \end{eqnarray*} which is just $(x-y)^k$ expanded as a power series in the variable $y$. \end{example} \begin{example} If we compose the refinement map with the multiplication map, $\mu:\G{2} \rightarrow G$, we have: \begin{eqnarray*} {\mathbb C} \ps{x}[x^{-1}] &\longrightarrow &{\mathbb C} \ps{x}[x^{-1}]\ps{y,z} \\ x^k &\mapsto & \sum_{p,q \geq 0} \binom{k}{p+q} \binom{p+q}{p} x^{k-p-q} y^p z^q, \end{eqnarray*} which is just $(x+y+z)^k$ expanded as a power series in the variables $y$ and $z$. \end{example} In order to apply this notion of refinement to our singular multilinear maps, we shall associate a refinement for a singularity to every map between trees in the category of trees. \begin{thm}\label{t:refinement_works} For $G$\mbox{-} module, $A_1, \ldots, A_n, B$, and any tree $p$ with $n$ leaves, there exists a canonical map $\ensuremath{\mathit{Multi}}^K_p(A_1, \ldots, A_n; B) \rightarrow \ensuremath{\mathit{Multi}}^K_q(A_1, \ldots, A_n; B)$ for all trees $q$ which refine to $p$. This map is given by appropriate refinements of singularities. \end{thm} Before we prove this theorem, we give some examples for the classical vertex group. \begin{example} We saw in example \ref{ex:binary_with_tail} that for the classical vertex group, there is an isomorphism \[\twoleafedtail{A_1}{A_2}{B}{x_1}{x_2}{y} \cong \flattwoleafed{A_1}{A_2}{B}{z_1}{z_2}\] \end{example} \begin{example} Again for the classical vertex group, consider the refinement map \[\flattwoleafed{A_1}{A_2}{B}{z_1}{z_2} \longrightarrow \twoleavedrighttwo{A_1}{A_2}{B}{x_1}{x_2}{y} \] At the level of power series, this is a map, \[\ensuremath{\mathrm{Hom}}\Bigl(A_1\mbox{$\otimes$} A_2, B\ps{z_1, z_2}[(z_1-z_2)^{-1}]\Bigr) \longrightarrow \ensuremath{\mathrm{Hom}}\Bigl(A_2, \ensuremath{\mathrm{Hom}}\bigl(A_1, B\ps{x_1, x_2}[(x_1-x_2)^{-1}]\bigr)\ps{y}\Bigr) \] where $z_1$ is mapped to $x_1$, $z_2$ is mapped to $x_2+y$, and $(z_1-z_2)^{-1}$ is mapped to $(x_1-x_2-y)^{-1}$ and expanded as power series in the variable $y$. Also we see that automatically $\partial_y = \partial_{x_2}$, which is the same as the $G$\mbox{-} invariance requirement at the internal node. Abstractly we have a refinement map \begin{eqnarray*} \ensuremath{\mathrm{Hom}}\Bigl(A_1\mbox{$\otimes$} A_2, \ensuremath{\mathrm{Hom}}(\G{2},B)\mbox{$\otimes$} K\Bigr) &\longrightarrow& \ensuremath{\mathrm{Hom}}\Bigl(A_2, \ensuremath{\mathrm{Hom}}\bigl(A_1 \mbox{$\otimes$} G, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2},B)\bigr)\Bigr) \\ &&\cong \ensuremath{\mathrm{Hom}}\Bigl(A_1 \mbox{$\otimes$} A_2, \ensuremath{\mathrm{Hom}}\bigl(G, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2},B)\bigr)\Bigr), \end{eqnarray*} and so this map reduces to a map \begin{eqnarray*} \ensuremath{\mathrm{Hom}}(\G{2},B)\mbox{$\otimes$} K &\longrightarrow &\ensuremath{\mathrm{Hom}}\bigl(G, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2},B)\bigr)\\ f\mbox{$\otimes$} k &\mapsto & F \end{eqnarray*} where $F(g) = \sum_{(g)} f(\cdot \mbox{$\otimes$} g_{(1)}\cdot) \mbox{$\otimes$} (S(g_{(2)})k)$. \end{example} \begin{example}\label{ex:expand_three} If we again let $G$ be the classical vertex group and consider the refinement map, \[\flatthreeleafedlabelled{A_1}{A_2}{A_3}{B}{z_1}{z_2}{z_3} \longrightarrow \threeleafedleft{A_1}{A_2}{A_3}{B}{x_1}{x_2}{y_1}{y_2}\] At the level of power series, we have a map of variables, \[B\ps{z_1,z_2,z_3}[(z_1-z_2)^{-1}, (z_1-z_3)^{-1}, (z_2-z_3)^{-1}] \longrightarrow B\ps{y_1, y_2}[(y_1-y_2)^{-1}]\ps{x_1, x_2}[(x_1-x_2)^{-1}]\] where the following are expanded as power series in the variables $x_1, x_2$. \begin{align*} z_1 &\mapsto x_1+y_1 & (z_1-z_2)^{-1} &\mapsto (x_1-x_2)^{-1}\\ z_2 &\mapsto x_2+y_1 & (z_1-z_3)^{-1} &\mapsto (x_1+y_1-y_2)^{-1}\\ z_3 &\mapsto y_2 & (z_2-z_3)^{-1} &\mapsto (x_2+y_1-y_2)^{-1}. \end{align*} Abstractly, we need to define a map from $\ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/flat_3.eps,height=3mm}}$ into \[\ensuremath{\mathrm{Hom}}\biggl(A_1 \mbox{$\otimes$} A_2, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl(\G{2} \mbox{$\otimes$} A_3, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2}, B)\Bigr)\biggr).\] Of the possible pullback objects corresponding to the flat 3 leafed tree, we shall show that there exists a canonical map from \[\ensuremath{\mathrm{Hom}}\biggl(A_1 \mbox{$\otimes$} A_2, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl(A_3, K^{\otimes 2} \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{3}, B)\Bigr)\biggr).\] Since the outer terms are the same we need only define a map \begin{eqnarray*} K^{\otimes 2} \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{3}, B) &\longrightarrow &\ensuremath{\mathrm{Hom}}\Bigl(\G{2}, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2}, B)\Bigr)\\ k_1 \mbox{$\otimes$} k_2 \mbox{$\otimes$} f &\mapsto & F, \end{eqnarray*} where $F(g\mbox{$\otimes$} h) = \sum_{(g)}\sum_{(h)} \biggl[\Bigl(S(g_{(1)})k_1\Bigr) \Bigl(S(h_{(1)})k_2\Bigr)\biggr] \mbox{$\otimes$} \biggl[f\Bigl(g_{(2)}\cdot \mbox{$\otimes$} h_{(2)}\cdot \mbox{$\otimes$} \cdot \Bigr)\circ(\Delta \mbox{$\otimes$} 1)\biggr].$ \end{example} \begin{example} From the discussion of locality in section \ref{ss:locality}, we see that by setting $z_3$, $x_2$ and $y_2$ equal to zero, the refinement map between $G$\mbox{-} invariant singular functions, \[\flatthreeleafedlabelled{V}{V}{V}{}{z_1}{z_2}{0} \longrightarrow \threeleafedright{V}{V}{V}{V}{x_1}{0}{y_1}{0}\] maps an element $f \in \ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/flat_3.eps,height=3mm}}(V,V,V;V)$ to the product of vertex operators, $Y(cdot, y_1)Y(\cdot, x_1)\cdot$. Then the locality condition reduces to the requirement that $f$ be symmetric (this is equivalent to the rationality and commutativity of products of Frenkel, Huang, and Lepowsky, \cite{FHL}). In fact, the refinement, \[\flatthreeleafedlabelled{V}{V}{V}{}{z_1}{z_2}{0} \longrightarrow \threeleafedleft{V}{V}{V}{V}{x_1}{0}{y_1}{0}\] is interpreted as giving a map from $f$ to $Y(Y(\cdot, x_1)\cdot, y_1)\cdot$, and thus formalises the notion of associativity for a vertex algebra as in \cite{FHL} or \cite[section 4.6]{kac}. \end{example} \begin{proof}[Proof of theorem \ref{t:refinement_works}] The idea of the proof is that we get a map between pullbacks by showing that there is a map between the spaces over which we are pulling back. There is also a map between the pullback objects, and these maps form a commuting diagram, inducing a map between pullbacks. For more of the abstract details see \cite{cts_valgdetails}. Explicitly, a tree $p$ can arise as a refinement of a tree $q$ by a sequence of either replacing nodes of $q$ with edges, or shrinking internal edges of $q$ down to nodes (see appendix \ref{a:cat_of_trees}). These two moves can be interchanged, and can be carried out one move at a time. Thus we shall consider them separately. And because we have defined $\ensuremath{\mathit{Multi}}^K_p$ and $\ensuremath{\mathit{Multi}}^K_q$ as an iteration of operators associated to each internal node, we need only consider refinements involving flat trees. \begin{description} \item[Case 1:] We consider the case where $p$ arises as a refinement of $q$ by replacing a node of $q$ with an edge. Taking $q$ to be the flat $n$ leafed tree, we may either replace the root or a leaf to give $p$. If we replace the root, then we are defining a canonical map \[\generictailA{A}{B}{n} \longrightarrow \genericA{A}{B}{n}\] Labelling the interior node of the tree on the left hand side, $p_1$, we see that an arbritrary pullback object in the definition of $\ensuremath{\mathit{Multi}}^K_p$, is \[\ensuremath{\mathrm{Sing}}_{\sigma,p_1}\ensuremath{\mathrm{Hom}}_G\bigl(\G{n},\ensuremath{\mathrm{Hom}}(G,B)\bigr),\] which is manifestly isomorphic to the pullback object, $\ensuremath{\mathrm{Sing}}_{\sigma,\ensuremath{\mathit{root}}}\ensuremath{\mathrm{Hom}}(\G{n},B)$, of the right hand side. The same holds for the space over which we pullback, and so we see that in fact this map is an isomorphism. If we instead replace a leaf, we are defining a canonical map \[\generictopA{A}{B}{n} \longrightarrow \genericA{A}{B}{n}\] A similar proof shows that this is also an isomorphism. \item[Case 2:] We next consider the case where $p$ arises by shrinking an internal edge of $q$ down to a node. In this case we are defining a canonical map \[\genericA{A}{B}{n} \longrightarrow \genericcompositionB\] where, without loss of generality, we have chosen to consider refinement of the first $m$ leaves ($m \leq n$). Labelling the internal node of the tree on the right hand side, $q_1$, we see that an arbritrary pullback object is of the form \[\ensuremath{\mathrm{Sing}}_{\sigma,q_1}\ensuremath{\mathrm{Sing}}_{\delta,\ensuremath{\mathit{root}}_q}\ensuremath{\mathrm{Hom}}(\G{n-m+1},B),\] where $\sigma \in A_m$ and $\delta \in A_{n-m+1}$. We need to show that there exists a natural map from $\ensuremath{\mathit{Multi}}^K_p$ to this object, so considering an arbritrary element of $\ensuremath{\mathit{Multi}}^K_p$, we consider it as an element of the pullback object $\ensuremath{\mathrm{Sing}}_{(\sigma,\delta),\ensuremath{\mathit{root}}_p} \ensuremath{\mathrm{Hom}}(\G{n},B)$, where we define a permutation, $(\sigma,\delta) \in A_n$: \begin{eqnarray*} (1, \ldots, m) &\mapsto & (\sigma(1), \ldots, \sigma(m))\\ (m+1, \ldots, n) &\mapsto & (\delta(x), \delta(m+1),\ldots, \widehat{\delta(i)}, \ldots, \delta(n)) \end{eqnarray*} where $x$ is the placeholder for the internal edge, and where we exclude $\widehat{\delta(i)}$ when $i$ is mapped to $x$. Notice that we have shifted the domain of $\delta$ so that it acts on $(m,\ldots, n)$, and that the permutation, $(\sigma,\delta)$, will arise for $n-m+1$ different choices of $\delta$. Consider the canonical map we are trying to exhibit: \[\ensuremath{\mathrm{Sing}}_{(\sigma,\delta),\ensuremath{\mathit{root}}_p} \ensuremath{\mathrm{Hom}}(\G{n},B)\longrightarrow \ensuremath{\mathrm{Sing}}_{\sigma,q_1}\ensuremath{\mathrm{Sing}}_{\delta,\ensuremath{\mathit{root}}_q}\ensuremath{\mathrm{Hom}}(\G{n-m+1},B).\] It will be an identity on the outer terms corresponding to $\ensuremath{\mathrm{Sing}}_{\sigma,q_1}$. So our problem reduces to showing that there exists a map \begin{multline}\label{e:refine_problem} \ensuremath{\mathrm{Hom}}\biggl(A_{(\sigma,\delta)(m+1)}, K^{\otimes m} \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl( A_{(\sigma,\delta)(m+2)}, K^{\otimes m+1} \mbox{$\otimes$} \cdots \ensuremath{\mathrm{Hom}}(\G{n},B)\Bigr)\biggr) \\ \longrightarrow \ensuremath{\mathrm{Hom}}\Biggl(A_{\delta(x)} \mbox{$\otimes$} A_{\delta(m+1)}, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\biggl(A_{\delta(m+1)}, K^{\otimes 2} \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl( A_{\delta(m+2)}, K^{\otimes 3} \mbox{$\otimes$} \cdots \ensuremath{\mathrm{Hom}}(\G{n-m+1},B)\Bigr)\biggr)\Biggr), \end{multline} where $A_x = \G{m}$. We immediately notice that there are fewer copies of the singularity $K$ on the right hand side. This is because when we expand the singularities, we will have maps of the form: \[K\mbox{$\otimes$} K \longrightarrow \ensuremath{\mathrm{Hom}}_G(G,K)\mbox{$\otimes$} \ensuremath{\mathrm{Hom}}_G(G,K) \longrightarrow \ensuremath{\mathrm{Hom}}_G(\G{2},K),\] where the second arrow is achieved by multiplication in $K$ (e.g., example \ref{ex:expand_three}). A complicated but routine calculation shows that a succession of maps of this form give the desired map. \end{description} \end{proof} \section{Singularities} Up to this point, we have concentrated on the underlying Hopf algebra structure of our vertex group. We now use the elementary vertex structure of the vertex group to add singularities to the extended representations of morphisms of $G$\mbox{-} modules, paying special attention to the case for the classical vertex group. It is in this chapter that this paper departs from Richard Borcherds' paper \cite{bor}. The ideas in this section were inspired by that paper, but the collections of singular maps we shall define are much smaller than those defined in his paper. In fact, they will be defined in such a way as to make their relationship to his singular maps clear. \subsection{Localisation}\label{localisation} The following definitions illustrate how we will put together the space of nonsingular functions from $\G{n}$ to $B$ with the ring of singular functions on $G$ in order to arrive at a notion of singular functions from $\G{n}$ to $B$. These will also be referred to as \emph{singular functions of type $K$}. \begin{dfn} For $1 \leq i < j \leq n$ and any (right) $\ensuremath{\mathrm{Hom}}_R(\G{n},R)=(\G{n})^*$\mbox{-} module, $M$, the \defn{localisation of $M$ at $(i,j)$} is defined to be $M \mbox{$\otimes$}_{G^*} K$, where $M$ is given the structure of an $G^*$\mbox{-} module by the $R$\mbox{-} module dual of the map \begin{equation}\label{e:f_ij} \begin{aligned} \widehat{f}_{ij}&: && \ \G{n} \ \ \longrightarrow \ \ && \ \ G. \\ &g_1 \mbox{$\otimes$} &&\cdots \mbox{$\otimes$} g_n \mapsto &&g_iS(g_j) \end{aligned} \end{equation} \end{dfn} We have already seen that the space of nonsingular functions from $\G{n}$ to $B$ has the structure of both an $G$\mbox{-} module and an $\G{n}$\mbox{-} module. Because of the coassociativity of $G$, it can also be considered as an $(\G{n})^*$\mbox{-} module where the action of any $\alpha \in (\G{n})^*$ on $f:\G{n} \rightarrow B$ is \begin{equation} \label{e:dual_act} (\alpha\cdot f)(h_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} h_n) = \alpha(h_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} h_n)\cdot f(h_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} h_n) \end{equation} for all $h_1 \mbox{$\otimes$} \cdots \mbox{$\otimes$} h_n \in \G{n}$. With this action defined, it makes sense to localise the space of nonsingular functions from $\G{n}$ to $B$. \begin{lemma}\label{l:sequentail_locs} The localisation of $\ensuremath{\mathrm{Hom}}_R(\G{n},B)$ is also a $(\G{n})^*$\mbox{-} module, and sequential localisations commute. \end{lemma} \begin{proof} This is clear from the action defined in equation \eqref{e:dual_act} and the coassociativity of $G$. \end{proof} \begin{dfn}\label{d:fun} The \defn{space of singular functions} from $\G{n}$ to $B$, denoted $\ensuremath{\mathit{Fun}}(\G{n},B)$, is defined to be the localisation of the space of nonsingular functions from $\G{n}$ to $B$ at all $(i,j)$ for $1 \leq i < j \leq n$. It is the space \[\ensuremath{\mathrm{Hom}}_R(\G{n},B) {\bigotimes}_{\substack{f_{i,j}\\1 \leq i < j \leq n}} K \] \end{dfn} The space of singular functions from $\G{n}$ to $B$ is a module for a number of different actions. Firstly, from lemma \ref{l:sequentail_locs}, we have that $\ensuremath{\mathit{Fun}}(\G{n},B)$ is a $(\G{n})^*$\mbox{-} module. Next we know that there is a left action of $G$ on $B$ which carries over to $\ensuremath{\mathit{Fun}}(\G{n},B)$. We also know that there is a (left) diagonal action of $G$ on the domain of the nonsingular maps $\ensuremath{\mathrm{Hom}}_R(\G{n},B)$. An easy check shows that this diagonal action carries over to a trivial action on $K$ via the maps $f_{i,j}$ defined in equation \eqref{e:f_ij}. Hence the standard action of $G$ on any singular function is the same as the standard action of $G$ on the nonsingular functions given by \eqref{e:hom_action}. This gives us the following lemma: \begin{lemma} If we define the \defn{space of singular $G$\mbox{-} invariant functions} from $\G{n}$ to $B$, denoted $\ensuremath{\mathit{Fun}}_G(\G{n},B)$, to be the localisation of $\ensuremath{\mathrm{Hom}}_G(\G{n},B)$ at all $(i,j)$ for $1 \leq i < j \leq n$, then this coincides with the $G$\mbox{-} invariant elements of $\ensuremath{\mathit{Fun}}(\G{n},B)$ under the standard action of $G$. \end{lemma} Finally we know that the nonsingular maps, $\ensuremath{\mathrm{Hom}}(\G{n},B)$, possesses a (right) action of $\G{n}$ on their domain. This action can be extended to singular maps by using comultiplication to act on each of the $\otimes_{f_{i,j}} K$ terms through that map $f_{i,j}$. \begin{example} If we are working with a trivial vertex group (i.e., $K = H^*$) then the localisation of the nonsingular functions from $\G{n}$ to $B$ has no effect, and we just have $\ensuremath{\mathit{Fun}}(\G{n},B) = \ensuremath{\mathrm{Hom}}_R(\H{n},B)$. \end{example} \begin{example}\label{ex:G1} If we let $G$ be the classical vertex group then we showed in section \ref{ss:classicalvg} that the collection of nonsingular functions, $\ensuremath{\mathrm{Hom}}_{\mathbb C}(G,B)$, corresponds to $B\ps{x}$. Since this can not be localised, we have $\ensuremath{\mathit{Fun}}(G,B)= \ensuremath{\mathrm{Hom}}_{\mathbb C}(G,B) \cong B\ps{x}$. The $G$\mbox{-} invariant maps are given by $\ensuremath{\mathit{Fun}}_G(G,B)=\ensuremath{\mathrm{Hom}}_G(G,B) \cong B$. \end{example} \begin{example}\label{ex:G2} Again letting $G$ be the classical vertex group, we have also shown that the collection of nonsingular functions from $\G{2}$ to $B$ corresponds to $B\ps{x_{1},x_{2}}$, which we consider as a module over ${\mathbb C}\ps{x_{1},x_{2}}$. The ring of singular functions on $G$ is isomorphic to ${\mathbb C}\ps{z}[z^{-1}]$. We can only localise at $(1,2)$ so consider the dual of the map $\widehat{f}_{12}:\H{2} \rightarrow H$. For any $\D{a},\D{b} \in H$, this gives $\widehat{f}_{12}(\D{a}\mbox{$\otimes$} \D{b}) = \binom{a+b}{a}(-1)^b \D{a+b}$, so the dual map $f_{12}:{\mathbb C}\ps{z} \rightarrow {\mathbb C}\ps{x_{1},x_{2}}$ takes $z^k \in H^*$ to \begin{eqnarray*} f_{12}(z^k) & = & \sum_{p+q=k} \binom{p+q}{p} (-1)^q x_{1}^p x_{2}^q \\ & = & (x_{1}-x_{2})^k. \end{eqnarray*} This map can be extended from $H^*$ to include negative values of $k$, in which case we will have infinitely many nonzero summands for negative values of $q$, and will still have $f_{12}(z^k)=(x_{1}-x_{2})^k$. Therefore our space of singular functions from $\G{2}$ to $B$ is given by \begin{eqnarray*} \ensuremath{\mathit{Fun}}(\G{2},B) & = & B\ps{x_{1},x_{2}} \mbox{$\otimes$}_{H^*} {\mathbb C}\ps{z}[z^{-1}] \\ &\cong& B\ps{x_{1},x_{2}}\ps{(x_{1}-x_{2})}[(x_{1}-x_{2})^{-1}] \\ &=& B\ps{x_{1},x_{2}}[(x_{1}-x_{2})^{-1}]. \end{eqnarray*} The $G$\mbox{-} invariant subcollection of maps in $\ensuremath{\mathit{Fun}}(\G{2},B)$ is the quotient of $B\ps{x_{1},x_{2}}[(x_{1}-x_{2})^{-1}]$ by the relation $\D{1}_B -\Partial{x_{1}} - \Partial{x_{2}}$, where $\D{1}_B$ is the action of $\D{1}$ on the module $B$. We thus have $\ensuremath{\mathit{Fun}}_G(\G{2},B) \cong B\ps{x_{1}-x_{2}}[(x_{1}-x_{2})^{-1}]$. \end{example} \begin{example}\label{ex:G3} Continuing the previous example for the case of $\G{3}$, we know that the space of nonsingular functions from $\G{3}$ to $B$ corresponds to $B\ps{x_{1},x_{2},x_{3}}$ and we need to localise at the three ordered pairs (i,j) for $1 \leq i < j \leq 3$. The dual maps $f_{12}$, $f_{13}$, $f_{23}$, behave exactly as in the previous example, and so we have for our space of singular functions from $\G{3}$ to $B$ the collection \begin{eqnarray*} \ensuremath{\mathit{Fun}}(\G{3},B) &=& \begin{aligned}[t]B\ps{x_{1},x_{2},x_{3}}\mbox{$\otimes$} {\mathbb C} \ps{(x_{1}-x_{2})}&[(x_{1}-x_{2})^{-1}] \mbox{$\otimes$} {\mathbb C} \ps{(x_{1}-x_{3})}[(x_{1}-x_{3})^{-1}] \\&\mbox{$\otimes$} {\mathbb C} \ps{(x_{2}-x_{3})}[(x_{2}-x_{3})^{-1}] \end{aligned}\\ &\cong&B\ps{x_{1},x_{2},x_{3}}[(x_{1}-x_{2})^{-1},(x_{1}-x_{3})^{-1}, (x_{2}-x_{3})^{-1}]. \end{eqnarray*} This can be extended in the obvious way to functions from $\G{n}$ to $B$ giving: \begin{equation} \ensuremath{\mathit{Fun}}(\G{n}, B) \cong B\ps{x_1, \ldots, x_n}[(x_i-x_j)^{-1}_{1 \leq i < j \leq n}]. \end{equation} \end{example} \subsection{Singular Multilinear Maps}\label{ss:sing_mult_maps} Now that we know how to use the elementary vertex structure of our vertex groups to add singularities to the picture, we can easily see how to generalise the extended representations of multilinear maps to define singular multilinear maps. But following the discussion given in section \ref{ss:motivation}, we want to define these maps very delicately. Consider first the collection of linear maps between $G$\mbox{-} modules $A$ and $B$, $\ensuremath{\mathrm{Hom}}_R(A,B)$. We saw in section \ref{ss:extended} that this collection is isomorphic to the collection of $G$\mbox{-} linear maps from $A$ to $\ensuremath{\mathrm{Hom}}_R(G,B)$. Replacing the collection of nonsingular maps $\ensuremath{\mathrm{Hom}}_R(G,B)$ with the collection of singular maps $\ensuremath{\mathit{Fun}}(G,B)$ has no effect since they are equal, so we have: \begin{dfn}\label{d:sing_1and2} The singular multilinear maps from $A$ to $B$ are just the multilinear maps in the extended representation: \[\ensuremath{\mathit{Multi}}^K(A;B) = \ensuremath{\mathrm{Hom}}_{G}(A, \ensuremath{\mathrm{Hom}}(G,B)).\] Repeating the same process for the collection of maps between $G$\mbox{-} modules $A_1\mbox{$\otimes$} A_2$ and $B$, we have the following as our singular multi maps: \[\ensuremath{\mathit{Multi}}^K(A_1,A_2;B) = \ensuremath{\mathrm{Hom}}_{\G{2}}(A_1\mbox{$\otimes$} A_2, \ensuremath{\mathit{Fun}}(\G{2},B)).\] \end{dfn} \begin{lemma}\label{l:singular_inv} The $G$\mbox{-} invariant singular multilinear maps, $\ensuremath{\mathit{Multi}}_G^K(A_1, A_2;B)$ are the $\G{2}$\mbox{-} invariant maps from $A_1\mbox{$\otimes$} A_2$ to $\ensuremath{\mathit{Fun}}_G(\G{2},B)$. \end{lemma} \begin{proof} The $\G{2}$\mbox{-} invariance of the singular multilinear maps allows the action of $G$ on the domain, $A_1\mbox{$\otimes$} A_2$, to carry over to an action on the domain of $\ensuremath{\mathit{Fun}}(\G{2},B)$ as described in section \ref{ss:extended}. Hence the action of $G$ on $A_1\mbox{$\otimes$} A_2$ passes through to an action on $B$ exactly when $\ensuremath{\mathit{Fun}}(\G{2},B)$ is $G$\mbox{-} invariant. \end{proof} \begin{example}\label{ex:sing_1_hom} Take $G$ to be the classical vertex group, and let $f$ be a singular multilinear maps in $\ensuremath{\mathit{Multi}}^K(A;B) = \ensuremath{\mathrm{Hom}}_{G}(A,B\ps{x}).$ Then for any $a \in A$ we can write $f(a) = \sum_{i\geq 0} b_i x^i$ for $b_i \in B$. If $f$ is $G$\mbox{-} invariant, then $Tf(a) = \partial_xf(a)$, and so $Tb_i=(i+1)b_{i+1}$. Thus \[f(a) = \sum_{i\geq 0} \frac{T^i}{i!}b_0 x^i = e^{Ty}b_0,\] and so we have an isomorphism of $G$\mbox{-} modules, \[\ensuremath{\mathit{Multi}}^K_G(A;B) \cong \ensuremath{\mathrm{Hom}}(A,B).\] \end{example} \begin{example}\label{ex:getting_vertexops} With $G$ taken to be the classical vertex group, the singular multilinear maps from $A_1\mbox{$\otimes$} A_2$ to $B$ are the $\G{2}$\mbox{-} invariant maps from $A_1\mbox{$\otimes$} A_2$ to $B\ps{x,y}[(x-y)^{-1}]$. Given any such map, say $f$, the $\G{2}$\mbox{-} invariance just says that for any $a \mbox{$\otimes$} b \in A \mbox{$\otimes$} B$, \begin{eqnarray*} f(Ta\mbox{$\otimes$} b) &=& \partial_x f(a\mbox{$\otimes$} b)\\ f(a\mbox{$\otimes$} Tb) &=& \partial_y f(a\mbox{$\otimes$} b). \end{eqnarray*} If we add the additional requirement that $f$ be $G$\mbox{-} invariant, then we have \begin{eqnarray*} Tf(a\mbox{$\otimes$} b) &=& f(Ta\mbox{$\otimes$} b) + f(a\mbox{$\otimes$} Tb)\\ && \partial_x f(a\mbox{$\otimes$} b) + \partial_y f(a\mbox{$\otimes$} b), \end{eqnarray*} so $G$ acts on $B\ps{x,y}[(x-y)^{-1}]$ as $T_B = \partial_x + \partial_y$. Taking $V = A_1 = A_2 = B$, and setting \[Y(\cdot, x)\cdot = f(\cdot \mbox{$\otimes$} \cdot)|_{y=0}:V\mbox{$\otimes$} V \longrightarrow V\ps{x}[x^{-1}],\] we have a vertex operator that satisfies the translation covariance axiom of equation \eqref{e:inv_axiom}. \end{example} \begin{example}\label{ex:getting_singmaps} Having seen how to compute a vertex operator from a singular map, we now show how to compute a binary singular map from a vertex operator. Let $Y(\cdot,x)\cdot$ be a vertex operator on a complex vector space $V$. We claim that the map, \[f(\cdot \mbox{$\otimes$} \cdot)= e^{Ty}Y(\cdot,x-y)\cdot,\] gives the desired $G$\mbox{-} invariant binary singular map. (Here the operator $e^{Ty}=\sum_{i\geq 0} y^i \D{i}$ provides a linear map from $V$ to $V\ps{y}$, and we are regarding $Y(\cdot,x-y) \in V\ps{x-y}[(x-y)^{-1}]$ as an element of $V\ps{x,y}[(x-y)^{-1}]$ under binomial expansion. We shall see in example \ref{ex:binary_with_tail} that this inclusion is important for the general description that will follow.) For any $a\mbox{$\otimes$} b \in V\mbox{$\otimes$} V$, clearly we see that $f(a\mbox{$\otimes$} b)|_{y=0} = Y(a,x)b$, and \begin{eqnarray*} f(a\mbox{$\otimes$} b)|_{x=0} &=& e^{Ty}Y(a, -y)b \\ &=& Y(b,y)a \end{eqnarray*} where the second equality follows by quasisymmetry (see appendix \ref{a:axiom_facts}). Notice also that we could have just as easily reconstructed $f(a\mbox{$\otimes$} b)$ from $e^{Tx}Y(b,y-x)a$. Thus we see that a binary singular map incorporates the quasisymmetry of a vertex operator. This example provides an interesting contrast with the case for holomorphic vertex algebras in section \ref{ss:extended}. There we saw that the entire theory was determined by a vertex operator $Y(\cdot,x)\cdot$ evaluated at $x=0$. Because we have singularities here, we can not make such an evaluation, but we still can use the actions of $G$ to reconstruct a singular map from the vertex operator. \end{example} The next obvious question is what happens when we compose two singular multilinear maps? We begin with an example using the classical vertex group: \begin{example} Following the discussion of composition in section \ref{ss:extended}, we shall assume that all maps are $G$\mbox{-} invariant, except possibly the bottom map. With an eye to emulating the behaviour of axiomatic vertex algebras, we shall also consider pointwise composition of these maps. Taking the classical vertex group, $G$, the simplest nontrivial example of composition is the composition of $f \in \ensuremath{\mathit{Multi}}^K_G(A_1, A_2;B_1)$ with $h \in \ensuremath{\mathit{Multi}}^K(B_1, B_2; C)$. Writing these out explicitly, we have: \begin{align} \begin{aligned} A_1 \mbox{$\otimes$} A_2 & \xrightarrow{f} B_1\ps{x_1, x_2}[(x_1-x_2)^{-1}] \end{aligned} && &\begin{aligned} B_1 \mbox{$\otimes$} B_2 & \xrightarrow{h} C\ps{z_1, z_2}[(z_1-z_2)^{-1}]. \end{aligned} \end{align} These compose to give a map \[A_1 \mbox{$\otimes$} A_2 \mbox{$\otimes$} B_2 \xrightarrow{h \circ f} C\ps{z_1, z_2}[(z_1-z_2)^{-1}]\ps{x_1, x_2}[(x_1-x_2)^{-1}].\] By the $G$\mbox{-} invariance of $f$, we know that this $\G{2}$\mbox{-} linear map satisfies $\partial_{z_1} = \partial_{x_1} + \partial_{x_2}$. Using these differential equations, the composite $h \circ f$ can be seen to factor through $C\ps{X_1,X_3}[(X_1-X_3)^{-1}]\ps{X_1-X_2}[(X_1-X_2)^{-1}]$ under either of the following changes of variables: \begin{align} X_1 & = x_1 + z_1 & X_1 & = x_2 + z_1 \label{e:varchange1}\\ X_2 & = x_2 + z_1 & X_2 & = x_1 + z_1 \label{e:varchange2}\\ X_3 & = z_2 & X_3 & = z_2. \label{e:varchange3} \end{align} \end{example} We notice immediately that even though the composite factors through a collection which has only 3 variables, it does not factor through $\ensuremath{\mathit{Fun}}(\G{3},B) \cong B\ps{x_1, x_2, x_3}[(x_1-x_2)^{-1}, (x_1-x_3)^{-1}, (x_2-x_3)^{-1}]$. In fact it is simple to find examples of such composites which do not fit inside $\ensuremath{\mathit{Fun}}(\G{3},B)$. Thus our singular multilinear maps do not compose in the way we might have initially hoped. Since we haven't yet defined singular multilinear maps for three input modules, $\ensuremath{\mathit{Multi}}^K(A_1, A_2, A_3;B)$, we could think of defining it to be a space large enough to contain the composite described in the previous example, and all other possible composites. But resulting spaces would end up simply being the spaces of formal distributions which we had initially set out to avoid. Instead, we shall define a collection of singular multilinear maps for every way of composing singular multilinear maps. In other words, if we denote the collection $\ensuremath{\mathit{Multi}}^K(A_1, A_2;B)$ by the labelled tree \[\flattwoleafed{A_1}{A_2}{B}{}{}\] then the composite would be an element of a space associated to the labelled tree \[\threeleafedleft{A_1}{A_2}{B_2}{C}{}{}{}{}\] and so to every binary tree $p$, we would associate a collection of singular multilinear maps, $\ensuremath{\mathit{Multi}}_{p}^K(A_1, \ldots, A_n;C)$. How then shall we define $\ensuremath{\mathit{Multi}}_{\epsfig{file=Images/left_3.eps,height=3mm}}^K(A_1, A_2, A_3;C)$? Following the discussion of section \ref{ss:motivation} we shall take the following definition: \begin{dfn} The collection of multilinear singular maps from $A_1, A_2, A_3$ to $C$ associated to the tree \epsfig{file=Images/left_3.eps,height=4mm} is given by the following equation: \[\ensuremath{\mathit{Multi}}_{\epsfig{file=Images/left_3.eps,height=3mm}}^K(A_1, A_2, A_3;C) := \ensuremath{\mathrm{Hom}}\Bigl(A_1\mbox{$\otimes$} A_2, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\bigl(\G{2}\mbox{$\otimes$} A_3, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2}, C)\bigr)\Bigr)/\sim,\] where we are quotienting by four actions of $G$. These will be clear by regarding this collection of multi maps inside the larger collection: \[\ensuremath{\mathrm{Hom}}_{\G{3}}\Bigl(A_1\mbox{$\otimes$} A_2\mbox{$\otimes$} A_3, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}_{G}\bigl(\G{2}, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2}, C)\bigr)\Bigr).\] Then as with the binary singular maps described above, the action on each of the $A_i$ carries over to an action on its respective occurrence of $G$, and the additional action of $G$ acts on the one hand on the remaining occurrence of $G$, and on the other hand it acts diagonally on the outer $\G{2}$ term. \end{dfn} \begin{note} Would it be better to use something like this: \begin{eqnarray*} \ensuremath{\mathit{Multi}}_{\epsfig{file=Images/left_3.eps,height=3mm}}^K(A_1, A_2, A_3;C) &:=& \ensuremath{\mathrm{Hom}}_{\G{3}}(A_1\mbox{$\otimes$} A_2\mbox{$\otimes$} R, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}_{\G{2}}(\G{2}\mbox{$\otimes$} A_3, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2}, C)))\\ &=& \ensuremath{\mathrm{Hom}}_{\G{2}}(A_1\mbox{$\otimes$} A_2, \ensuremath{\mathit{Fun}}_{\G{2}}(\G{2}\mbox{$\otimes$} A_3, \ensuremath{\mathit{Fun}}(\G{2}, C))). \end{eqnarray*} \end{note} \begin{example} Taking $G$ to be the classical vertex group, this just says that the collection of singular multilinear maps, $\ensuremath{\mathit{Multi}}_{\epsfig{file=Images/left_3.eps,height=3mm}}^K(A_1, A_2, A_3;C)$ is the subcollection of \[\ensuremath{\mathrm{Hom}}\Bigl(A_1\mbox{$\otimes$} A_2, \ensuremath{\mathrm{Hom}}\bigl(A_3, C\ps{x_1, x_2}[(x_1-x_2)^{-1}]\bigr) \ps{y_1, y_2}[(y_1-y_2)^{-1}]\Bigr),\] satisfying the equations \begin{eqnarray*} T_{A_1} &=& \partial_{y_1}, \\ T_{A_2} &=& \partial_{y_2},\\ T_{A_3} &=& \partial_{x_2}\text{, and}\\ \partial_{y_1}+\partial_{y_2} &=& \partial_{x_1}, \end{eqnarray*} where $T_{A_i}$ denotes the action of $G$ on the module, $A_i$. We can represent this subcollection pictorially by the labelled tree \[\threeleafedleft{A_1}{A_2}{A_3}{C}{y_1}{y_2}{x_1}{x_2}\] or consider it to be the appropriate subcollection of \[\ensuremath{\mathrm{Hom}}\Bigl(A_1\mbox{$\otimes$} A_2\mbox{$\otimes$} A_3, C\ps{x_1,x_2}[(x_1-x_2)^{-1}] \ps{y_1, y_2}[(y_1-y_2)^{-1}]\Bigr),\] where the singularity $(y_1-y_2)^{-1}$ is ``independent'' of $A_3$. \end{example} We can extend this definition naturally to include all binary trees that have exactly one splitting at each level. A difficulty arises when we consider two compositions at a single level. For example, the collection of singular maps associated to the binary labelled tree: \[\fourleafdoubleB{A_1}{A_2}{A_3}{A_4}{C}\] We could arrive at a singular multilinear map of type \epsfig{file=Images/4ht2.eps,height=3mm} via composition in three ways: for $G$\mbox{-} modules $B_1$ and $B_2$, we could compose three binary singular maps: \begin{equation} \label{e:sing1} \ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/2_leaf.eps,height=2mm}}(A_1,A_2;B_1) \mbox{$\otimes$} \ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/2_leaf.eps,height=2mm}}(A_3,A_4;B_2) \mbox{$\otimes$} \ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/2_leaf.eps,height=2mm}}(B_1,B_2;C) \longrightarrow \ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/4ht2.eps,height=3mm}}(A_1, A_2, A_3, A_4;C), \end{equation} or we could compose an appropriate binary singular map with a singular map of type \epsfig{file=Images/left_3.eps,height=3mm} or \epsfig{file=Images/right_3.eps,height=3mm} as: \begin{eqnarray} \ensuremath{\mathit{Multi}}_{\epsfig{file=Images/left_3.eps,height=3mm}}^K(A_1, A_2, B_2;C) \mbox{$\otimes$} \ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/2_leaf.eps,height=2mm}}(A_3,A_3;B_2) &\longrightarrow & \ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/4ht2.eps,height=3mm}}(A_1, A_2, A_3, A_4;C) \label{e:sing2}\\ \ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/2_leaf.eps,height=2mm}}(A_1,A_2;B_1) \mbox{$\otimes$} \ensuremath{\mathit{Multi}}_{\epsfig{file=Images/right_3.eps,height=3mm}}^K(B_1, A_3, A_4;C) &\longrightarrow & \ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/4ht2.eps,height=3mm}}(A_1, A_2, A_3, A_4;C).\label{e:sing3} \end{eqnarray} We would like to end up with a space which consists exactly of maps that could arise as composites. The important feature of this definition is the restriction of the dependence of each singularity to the relevant controlling modules. This suggests the following definition: \begin{dfn}\label{d:4leaf_height2} The collection of singular multilinear maps of type \epsfig{file=Images/4ht2.eps,height=3mm} from $A_1, A_2, A_3, A_4$ to $C$ is defined to be the pullback of the singular multilinear maps, \begin{eqnarray} \ensuremath{\mathrm{Hom}}_{G,\G{2}}\Bigl(A_1\mbox{$\otimes$} A_2,K\mbox{$\otimes$} \ensuremath{\mathrm{Hom}}_{G,\G{2}}\bigl(A_3\mbox{$\otimes$} A_4, K\mbox{$\otimes$} \ensuremath{\mathrm{Hom}}_{G,\G{2}}(\G{4}, K\mbox{$\otimes$}\ensuremath{\mathrm{Hom}}(\G{2},C))\bigr)\Bigr) \label{e:4multi_1}\\ \ensuremath{\mathrm{Hom}}_{G,\G{2}}\Bigl(A_3\mbox{$\otimes$} A_4,K\mbox{$\otimes$} \ensuremath{\mathrm{Hom}}_{G,\G{2}}\bigl(A_1\mbox{$\otimes$} A_2, K\mbox{$\otimes$} \ensuremath{\mathrm{Hom}}_{G,\G{2}}(\G{4}, K\mbox{$\otimes$}\ensuremath{\mathrm{Hom}}(\G{2},C))\bigr)\Bigr) \label{e:4multi_2} \end{eqnarray} over \begin{equation} \ensuremath{\mathrm{Hom}}_{\G{4}}\Bigl(A_1\mbox{$\otimes$} A_2\mbox{$\otimes$} A_3\mbox{$\otimes$} A_4, K\mbox{$\otimes$} K\mbox{$\otimes$} \ensuremath{\mathrm{Hom}}_{\G{2}}\bigl(\G{4}, K\mbox{$\otimes$}\ensuremath{\mathrm{Hom}}(\G{2},C)\bigr)\Bigr) \end{equation} where the appropriate $G$\mbox{-} invariance is taken into account. \end{dfn} \begin{note} Because invariance just appears as an equaliser, it commutes with pullbacks. \end{note} \begin{thm}\label{t:4compose} The composites of singular maps in equations \eqref{e:sing1}, \eqref{e:sing2}, and \eqref{e:sing3} compose as desired. \end{thm} Before we prove this theorem, we need the following lemma. \begin{lemma} In any symmetric monoidal category, ${\cal C}$, the following diagram commutes: \[\xymatrix{ \ensuremath{\mathrm{Hom}}(A_1, B_1\mbox{$\otimes$} C_1)\mbox{$\otimes$}\ensuremath{\mathrm{Hom}}(A_2,B_2\mbox{$\otimes$} C_2)\ar[r]\ar[d]& \ensuremath{\mathrm{Hom}}(A_1, C_1 \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(A_2,B_1\mbox{$\otimes$} B_2\mbox{$\otimes$} C_2))\ar[d] \\ \ensuremath{\mathrm{Hom}}(A_2, C_2 \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(A_1,B_1\mbox{$\otimes$} C_1\mbox{$\otimes$} B_2))\ar[r]& \ensuremath{\mathrm{Hom}}(A_1\mbox{$\otimes$} A_2, B_1\mbox{$\otimes$} C_1\mbox{$\otimes$} B_2\mbox{$\otimes$} C_2)}\] \end{lemma} \begin{proof} The proof follows immediately from the fact that the evaluation of \[\ensuremath{\mathrm{Hom}}(A_1, B_1\mbox{$\otimes$} C_1)\mbox{$\otimes$}\ensuremath{\mathrm{Hom}}(A_2,B_2\mbox{$\otimes$} C_2)\] on $A_1\mbox{$\otimes$} A_2$ gives the same result when carried out by either first evaluating $A_1$, or by first evaluating $A_2$ or by evaluating both together. \end{proof} \begin{proof}[Proof of theorem \ref{t:4compose}] We shall only prove that the composition works for the classical vertex group. For the general theory, see \cite{cts_valgdetails}. Since composition is pointwise, the composite \eqref{e:sing1} factors through both \eqref{e:sing2} and \eqref{e:sing3}, and so we can focus on those compositions. In order to prove that the composites map into the pullback, we shall first prove that they map into each of the pullback objects (the singular maps given in \eqref{e:4multi_1} and \eqref{e:4multi_2}). Taking the composite in equation \eqref{e:sing2}, we see that it can be considered a multi map as in equation \eqref{e:4multi_1} through the following natural map: \begin{equation*} \begin{split} \ensuremath{\mathit{Multi}}&_{\epsfig{file=Images/left_3.eps,height=3mm}}^K(A_1, A_2, B_2;C) \mbox{$\otimes$} \ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/2_leaf.eps,height=2mm}}(A_3,A_3;B_2)=\\ \ensuremath{\mathrm{Hom}}&\Bigl(A_1\mbox{$\otimes$} A_2, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\bigl(\G{2}\mbox{$\otimes$} B_2, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2}, C)\bigr)\Bigr) \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl(A_3\mbox{$\otimes$} A_4, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2},B_2)\Bigr)\\ \longrightarrow& \ensuremath{\mathrm{Hom}}\biggl(A_1\mbox{$\otimes$} A_2, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\bigl(\G{2}\mbox{$\otimes$} B_2, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2}, C)\bigr) \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl(A_3\mbox{$\otimes$} A_4, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2},B_2)\Bigr)\biggr) \\ \longrightarrow& \ensuremath{\mathrm{Hom}}\biggl(A_1\mbox{$\otimes$} A_2, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl(A_3\mbox{$\otimes$} A_4, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2},\ensuremath{\mathrm{Hom}}\bigl(\G{2}, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2}, C)\bigr) )\Bigr)\biggr) \\ \cong & \ensuremath{\mathrm{Hom}}\biggl(A_1\mbox{$\otimes$} A_2, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\Bigl(A_3\mbox{$\otimes$} A_4, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{4}, K \mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{2}, C))\Bigr)\biggr). \end{split} \end{equation*} Similarly we can can see that it gives a multi map as in equation \eqref{e:4multi_2} through evaluating $A_3 \mbox{$\otimes$} A_4$ first. From the lemma, we know that these two ways of evaluating are equal, so the composite must factor through the pullback. \end{proof} In the next section we give a general definition for singular multilinear maps parameterised by any binary tree. But before we do so, we finish this section with a description of $\ensuremath{\mathit{Multi}}^K_{\epsfig{file=Images/4ht2.eps,height=3mm}}(A_1, A_2, A_3, A_4;C)$ for the classical vertex group. \begin{example} Letting $G$ be the classical vertex group, the collection of singular multilinear maps of type \epsfig{file=Images/4ht2.eps,height=3mm} from $A_1, A_2, A_3, A_4$ to $C$ is the submodule of \[\ensuremath{\mathrm{Hom}}\Bigl(A_1\mbox{$\otimes$} A_2\mbox{$\otimes$} A_3\mbox{$\otimes$} A_4,C\ps{x_1,x_2}[(x_1-x_2)^{-1}] \ps{y_1,y_2, z_1, z_2}[(y_1-y_2)^{-1},(z_1-z_2)^{-1}]\Bigr),\] where the singularity $(y_1-y_2)^{-1}$ is independent of $A_3\mbox{$\otimes$} A_4$, the singularity $(z_1-z_2)^{-1}$ is independent of $A_1 \mbox{$\otimes$} A_2$, and the maps satisfy the following equations: \begin{align*} T_{A_1}&=\partial_{y_1} & T_{A_2}&=\partial_{y_2}\\ T_{A_3}&=\partial_{z_1} & T_{A_4}&=\partial_{z_2}\\ \partial_{y_1}+\partial_{y_2} &= \partial_{x_1} & \partial_{z_1}+\partial_{z_2} &= \partial_{x_2} \end{align*} We represent this collection pictorially by the following labelled tree: \[\fourleafdoubleC{A}{C}{y}{z}{x}\] The $G$\mbox{-} invariant such maps are the further subcollection satisfying $T_C=\partial_{x_1}+\partial_{x_2}$. \end{example} \subsection{Multi Maps Parameterised by Binary Trees}\label{ss:multimaps_binary} In this section we are concerned with giving the definition of a multilinear singular map associated to an arbritrary binary tree. Given any binary tree $p$, we may consider its collection of internal vertices. For our purposes, we shall assume that these include the root, but they do not include the leaves. Considering them as a set, this set inherits a partial order from the tree, where the root is the least element. We know that any partial order can be extended to at least one total ordering, possibly many. We have already been working with trees whose leaves and root are labelled by $G$\mbox{-} modules. It will be useful for the explanations to follow to assume that every internal node is also labelled. For any internal node, $q$, connected to $n$ \defn{incoming} nodes (i.e., non-empty nodes whose height is equal to the height of $q$ plus one and connected to $q$ by a single edge), we shall label $q$ by $\G{n}$. We can also associate to $q$ the tensor product of the labels of the incoming nodes, and denote it $X_q$. Thus the following labelled tree has two internal nodes, \[\threeleafedleft{A_1}{A_2}{A_3}{C}{}{}{}{}\] and we have $X_{\text{root}}=\G{2}\mbox{$\otimes$} A_3$, and $X_{\text{internal}}=A_1\mbox{$\otimes$} A_2$. We say that our tree is \defn{augmented} when we have added an additional vertex and edge to the root of the tree, such that the new vertex inherits the former root label. The former root is labelled $\G{2}$ as expected. We denote the new vertex $\bot$, and we automatically have $X_\bot = \G{2}$. \begin{dfn}\label{d:bin_sing_maps} Let $p$ be a binary $n$\mbox{-} leafed tree, and let $t$ denote a total ordering, $\bot < root < p_1 < \cdots < p_l$, of the internal vertices of augmented $p$, compatible with the the partial ordering inherited from the tree structure of $p$. We define an operator on $\G{2}$\mbox{-} modules: \[\ensuremath{\mathrm{Sing}}_{p_i} = \ensuremath{\mathrm{Hom}}_{G,\G{2}}(X_{p_i}, K\mbox{$\otimes$}\cdot).\] Iterating this operator we have \[ \mathrm{Ord}_t(A_1, \ldots, A_n, C) = \ensuremath{\mathrm{Sing}}_{p_l}\cdots \ensuremath{\mathrm{Sing}}_{p_1} \ensuremath{\mathrm{Sing}}_{\text{root}} \ensuremath{\mathrm{Hom}}(X_\bot,C).\] Then $\ensuremath{\mathit{Multi}}^K_p(A_1, \ldots, A_n, C)$ is defined to be the wide pullback of each $\mathrm{Ord}_t$ for all possible total orderings, $t$, of of the internal vertices of augmented $p$, over \[\ensuremath{\mathrm{Hom}}_{\G{n}}(A_1\mbox{$\otimes$}\cdots\mbox{$\otimes$} A_n, \ensuremath{\mathit{Fun}}_{p}(\G{n},C))\] where $\ensuremath{\mathit{Fun}}_{p}(\G{n},C))$ is the collection of singular functions as defined in \cite{bor}. See appendix \ref{ss:bor_sing_maps} for more details. We have \defn{$G$\mbox{-} invariant multilinear singular maps}, $\ensuremath{\mathit{Multi}}^K_{G,p}$ exactly when $\ensuremath{\mathrm{Hom}}(X_\bot,C)$ is $G$\mbox{-} invariant. \end{dfn} \begin{example} If $p$ is a tree with only one binary tree at each level, then there is only one total ordering, $t$, of internal vertices of the tree, and so \[\ensuremath{\mathit{Multi}}^K_p(A_1,\ldots, A_n, C) = \mathrm{Ord}_t(A_1, \ldots, A_n, C).\] \end{example} \begin{example} When $p = \epsfig{file=Images/4ht2.eps,height=3mm}$, there are exactly two total orderings of internal vertices of this tree, and the corresponding $\mathrm{Ord}_t(A_1, \ldots, A_4, C)$ functions are given by \begin{eqnarray} \mathrm{Ord}_{t_1} &=& \ensuremath{\mathrm{Hom}}\Bigl(A_1\mbox{$\otimes$} A_2,K\mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\bigl(A_3\mbox{$\otimes$} A_4,K\mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{4}, K\mbox{$\otimes$}\ensuremath{\mathrm{Hom}}(\G{2},C))\bigr)\Bigr)\\ \mathrm{Ord}_{t_2} &=& \ensuremath{\mathrm{Hom}}\Bigl(A_3\mbox{$\otimes$} A_4,K\mbox{$\otimes$} \ensuremath{\mathrm{Hom}}\bigl(A_1\mbox{$\otimes$} A_2, K\mbox{$\otimes$} \ensuremath{\mathrm{Hom}}(\G{4}, K\mbox{$\otimes$}\ensuremath{\mathrm{Hom}}(\G{2},C))\bigr)\Bigr), \end{eqnarray} where the $\ensuremath{\mathrm{Hom}}$ functions are both $G$\mbox{-} invariant and $\G{2}$\mbox{-} invariant. Since we are pulling back over \begin{multline} \ensuremath{\mathrm{Hom}}_{\G{4}}(A_1\mbox{$\otimes$}\cdots\mbox{$\otimes$} A_4, \ensuremath{\mathit{Fun}}_{p}(\G{4},C)) = \\ \ensuremath{\mathrm{Hom}}_{\G{4}}(A_1\mbox{$\otimes$}\cdots\mbox{$\otimes$} A_4, K\mbox{$\otimes$} K\mbox{$\otimes$} \ensuremath{\mathrm{Hom}}_{\G{2}}(\G{4}, K\mbox{$\otimes$}\ensuremath{\mathrm{Hom}}(\G{2}, C))), \end{multline} this definition reduces to definition \ref{d:4leaf_height2}. \end{example} \section{Trees}\label{a:cat_of_trees} Formally, a \defn{tree} is defined to be a connected oriented planar graph with a finite number of vertices and no circuits. The empty tree is denoted by a circle, $\circ$. All other trees have at least one vertex, and given any such graph we choose a particular vertex which we call the \defn{root} of the tree. For convenience we will always draw the root of the tree at the bottom of the graph. The vertices which are joined to exactly one edge (excluding the root) are called the external vertices or the \defn{leaves} of the tree. All other vertices are called internal vertices. For $n \geq 0$, we let \T{n} denote the collection of all $n$ leafed trees. We say that a tree has \defn{height} $d$, if $d$ is the greatest number of edges between the root vertex and a leaf. In \eqref{e:3leafs}, the first tree has height 1 while the second and third trees have height 2. The unique tree with zero internal vertices is a refinement for every tree in \T{n} and is the only tree with height 1 in that collection. We call trees of height 1 \defn{flat} tress. Each collection of trees \T{n} has a subcollection in which the leaves of each tree are separated from the root vertex by the same number of edges. We denote these \defn{the trees of constant height} \TT{n}. \begin{equation} \label{e:3leafs} \flatthreeleafed{}{}{}{} \flatthreeleafedleft{}{}{}{} \flatthreeleafeddown \end{equation} For any two trees $p,q$ with the same number of leaves, we say that $p$ is a \defn{refinement} of $q$ if $p$ arises after a succession (possibly zero) of the following moves: \begin{itemize} \item an internal edge is shrunk down to a vertex, \item a vertex is replaced by an edge. \end{itemize} By internal edges, we mean those edges that do not end in a leaf. Two trees are considered to be the same exactly when they have the same oriented graph. In \eqref{e:3leafs} the first tree is a refinement of the second, and the first and third trees are refinements of one another. There are a number of (possibly inequivalent) ways of giving the collection of all trees the structure of a category. For our purposes we take the definition due to Tom Leinster \cite{leinster_enrichment} which seems to arise most naturally when dealing with higher dimensional categories. (For possibly different definitions see \cite{soibelman} or \cite{kreimer_trees}.) In this categorical structure, we define a single morphism from a tree $q$ to a tree $p$ exactly when $p$ is a refinement of $q$. So in equation \eqref{e:3leafs} there exists an arrow from the second tree to the first, and there exists an arrow in each direction between the first tree and the third. Refinement is transitive so morphisms compose and each tree is a refinement of itself so we have identity morphisms. For each $n \geq 0$, \T{n} is a category and so $\mathcal{T}$ is a category with countably many disconnected components. Under this definition, the subcollection of trees of constant height forms a full subcategory of the category of trees, and each \TT{n} is a full subcategory of \T{n}. The category of trees naturally forms an operad, which is the usual multi category of composable trees.
\section{Introduction} In a recent letter \cite{apjl1} we have claimed that the 4 Year COBE DMR data exhibits evidence of non Gaussianity; our quantitative claim was that the hypotheses that this data set is due to a Gaussian random process can be ruled out at the $98 \%$ confidence level. This came somewhat as a surprise given that this data set was lauded as strong evidence for the inflationary paradigm: its statistics were thought to be {\it consistent} with Gaussianity \cite{kog96a}. We have performed an extensive analysis of the 4 year COBE DMR data set and have been unable to find a non cosmological origin for the non-Gaussian signal. The resuls of this analysis are presented in \cite{bigpaper}. In this report I will only address a few often raised questions: \begin{itemize} \item Why does this statistic get a different result from all the ones that were previously used? \item Isn't the effect we are seeing just some systematic effect of the analysis? \item If it is signal what could it be? \end{itemize} \section{The Statistic and Result} In our analysis we propose, and work with, an estimator for the {\it normalized bispectrum}. To construct such an estimator we work in the spherical harmonic representation $a_{\ell m}$ and consider the tensor product of $3$ $\Delta T_\ell=\sum_{m}a_{\ell m}$s One is interested in rotationally invariant quantities. These can be trivially obtained if one rewrites the tensor product in terms of the total angular momentum basis. The coefficient of the singlet will be the higher order invariant we are looking for. This procedure leads to an estimator of the bispectrum \begin{eqnarray} {\hat B}_\ell&=&\alpha_\ell\sum_{m_1m_2m_3}{\cal W}^{\ell \ell \ell}_{m_1m_2m_3} a_{\ell m_1}a_{\ell m_2} a_{\ell m_3} \nonumber \\ \alpha_\ell&=&\frac{1}{(2\ell+1)^{\frac{3}{2}}} \left ({\cal W}^{\ell \ell \ell}_{000} \right )^{-1} \label{bispec} \end{eqnarray} where ${\cal W}^{\ell_1\ell_2\ell_3}_{m_1m_2m_3}$ are the Wigner 3J coefficients. ${\hat B}_\ell$ may then be divided by the appropriate power of an estimator for $C_\ell$ in order to make it dimensionless, and suitably normalised. \begin{eqnarray} I^3_\ell &=&\left| { {\hat B}_\ell\over ({\hat C}_\ell)^{3/2}} \right| \label{defI} \end{eqnarray} Let us comment on a number of features of the estimator. Firstly this statistic is global on pixel space, i.e. all the estimators are a function of all the pixel values. This means that it will be very good at identifying the scale dependence of non-Gaussianity (in the same way as the $C_\ell$s are very discrimative of the scale dependence of the variance of the fluctuations); however it performs very poorly at identifying the location of any signal on the map. In this it contrasts with statistics which have been applied to the COBE data until now: the three point correlation function, the statistics of peaks and topological measures are all defined on pixel space \cite{kog96a}. Secondly we have assumed statistical isotropy in constructing this estimator. We know, however, that the COBE data has a number of anisotropic features which violate this assumption, in particular the presence of the galaxy (which must be removed) and the anisotropic sky coverage of the observation pattern. The only way we can truly assess the significance of the $I^3_\ell$ is by comparing them to an ensemble of $I^3_\ell$ measured on Gaussian sky maps generated with exactly the same characteristics as that of the 4 year COBE DMR data. The assumptions of the Monte Carlo must be carefully checked. We have tested the inverse noise variance weighted, average maps of the 53A, 53B, 90A and 90B {\it COBE}-DMR channels, with monopole and dipole removed, at resolution 6, in ecliptic pixelization. We use the extended galactic cut of \cite{banday97}, and \cite{benn96} to remove most of the emission from the plane of the Galaxy. We apply our statistics to the DMR maps before and after correction for the plausible diffuse foreground emission outside the galactic plane as described in \cite{kog96b}, and \cite{COBE}. To estimate the $I^3_\ell$s we set the value of the pixels within the galactic cut to 0 and the average temperature {\it of the cut map} to zero. We then integrate the map multiplied with spherical harmonics to obtain the estimates of the $a_{\ell m}$s and apply equations \ref{bispec} and \ref{defI}. We then compare the estimates to the distributions generated from Monte Carlos simulations of Gaussian maps. \begin{figure} \centering \leavevmode\epsfysize=7cm \epsfbox{fig1.ps}\\ \caption[flatness]{\label{fig1}The vertical thick dashed line represents the value of the observed $I^3_\ell$. The solid line is the probability distribution function of $I^3_\ell$ for a Gaussian sky with extended galactic cut and DMR noise.} \end{figure} Since the $I^3_\ell$ distributions are non Gaussian we generalize the $\chi^2$ for a set of probability functions $P_\ell(I^3_\ell)$ associated with observations $\{I^3_\ell\}$ by defining the following functional \begin{equation}\label{presc} X^2={1\over N}{\sum_\ell X_\ell^2}= {1\over N}{\sum_\ell (-2\log P_\ell(I^3_\ell) + \beta_\ell),} \end{equation} where the constants $\beta_\ell$ are defined so that for each term of the sum $\langle X_\ell^2\rangle=1$. The definition reduces to the usual $X^2$ for Gaussian $P_\ell$. Again, we build a $X^2$ for the {\it COBE}-DMR data by means of Monte Carlo simulations. We proceed as follows. First we compute the distributions $P(I^3_\ell)$, for $\ell=2,\dots,18$, for a Gaussian process as measured subject to our galactic cut, and pixel noises. These $P(I^3_\ell)$ were inferred from 25000 realizations (see Fig.~\ref{fig1}). From these distributions we then build the $X^2$ defined in ($\ref{presc}$), taking special care with the numerical evaluation of the constants $\beta_\ell$. We call this function $X^2_{COBE}$. We then find its distribution $F(X^2_{COBE})$ from 10000 random realizations. This is very well approximated by a $\chi^2$ distribution with 12 degrees of freedom (Fig.~\ref{chi2}). We then compute $X^2_{COBE}$ with the actual observations and find $X^2_{COBE}=1.81$. One can compute $P(X^2_{COBE}<1.81)= 0.98$. Hence, it would appear that we can reject Gaussianity at the $98\%$ confidence level. \begin{figure} \centering \leavevmode\epsfysize=7cm \epsfbox{chi2.ps}\\ \caption[chi2]{\label{chi2}The distribution $P(X^2_{COBE})$ and a fit to it, which is a chi squared with 12 degress of freedom. The observed chi squared, with COBE, is represented as a vertical line: $X^2_{COBE}=1.81$. One can compute $P(X^2_{COBE}<1.81)= 0.98$. Hence, it would appear that we can reject Gaussianity at the $98\%$ confidence level.} \end{figure} \section{Is it a systematic effect?} The non-Gaussianity we are finding in the 4 year COBE DMR maps is surprising and fascinating enough that we have gone through an exhaustive test of all possible systematic effects. We summarize them in the following list: \begin{quote} \begin{enumerate} \item Foregrounds contamination: \begin{itemize} \item Dust (using the DIRBE sky maps and also the Schlegel {\it et al} dust model) \item Synchrotron (with the Haslam template) \item Foreground corrected maps \end{itemize} \item Noise model: \begin{itemize} \item Anisotropic sky coverage \item Noise correlations between different pixels \item Analysis of noise templates \end{itemize} \item Galactic cut: \begin{itemize} \item Dependence on shape (``custom'' versus constant elevation) \item Dependence on elevation \item Dependence on monopole and dipole subtraction, before or after the cut, with or with out galaxy. \end{itemize} \item Systematic templates \begin{itemize} \item Spurious offsets induced by the cut. \item Instrument susceptibility to the Earth magnetic field. \item Callibration errors . \item Errors due to incorrect removal of the COBE Doppler and Earth Doppler signals. \item Errors in correcting for emissions from the Earth, and eclipse effects. \item Artifacts due to uncertainty in the correction for the correlation created by the low-pass filter on the lock-in amplifiers (LIA) on each radiometer \item Errors due to emissions from the moon, and the planets. \end{itemize} \item Assumptions in Monte Carlos: \begin{itemize} \item Dependence on tilt \item Dependence on smooth versus discontinuous power spectrum \item Dependence on beam shape \item Dependence on pixelization. \end{itemize} \end{enumerate} \end{quote} Let us just highlight the foreground tests. In figure \ref{foreg} (left panel) we plot the $I^3_\ell$s estimated directly from the dust maps produced by the DIRBE $100\mu m$ and $240\mu m$ channels and from the dust template constructed in \cite{sfd98} using the DIRBE and IRAS maps. The feature we identify in the DMR data is not there. Furthemore, if we use the dust templates to subtract any foreground contribution (see figure \ref{foreg} right panel) the value of $I^3_{16}$ actually increases! \begin{figure} \centering \leavevmode\epsfysize=7cm \epsfbox{dustrawBW.ps} \leavevmode\epsfysize=7cm \epsfbox{forcorrBW.ps} \caption[foreg]{\label{foreg}Left panel: The measured $I_\ell$ for dust galactic foregrounds. We have plotted values concerning the DIRBE 100$\mu$m and DIRBE 240$\mu$m as well as the Schlegel, Finkbeiner, and Davis maps. Right panel: The measured $I_\ell$ for DMR foreground corrected data in ecliptic coordinates, galactic coordinates, and ecliptic coordinates with galactic frame coupling.} \end{figure} We have found that none of the effects listed above change our result by much, in fact for most corrections the confidence level rises to over $99\%$. We therefore claim that the non-Gaussianity we find in the COBE DMR data is not due to any of the {\it known} systematics. \section{The nature of the signal in $I^3_{16}$} The structure of the non-Gaussian signal is truly intriguing. It manifests itself as a spectral ``spike'' at $\ell=16$; it is difficult to associate such a pattern to some known or speculated source of fluctuations. If the signal is cosmological, the minimal inflationary models cannot be right. On the other hand it is not obvious that the main competitor to inflation, topological defects, could explain this type of non-Gaussianity. Topological defects are non-Gaussian, but in ways which are often more subtle than commonly thought. An interesting possibility was recently proposed by Peebles \cite{peeb}. This is an isocurvature model in which the underlying fluctuations are not a Gaussian random field, but the square of a Gaussian random field. The model is based on non minimal inflation, but produces fluctuations radically different from minimal inflationary fluctuations. A more useful exercise is to try and understand, given the characteristics of the experiment, what the power spectrum of such a non-Gaussian signal might be. The simplest thing to consider is combination of three signals. On very large scales (upto $\ell=12-14$) we have a Gaussian sky signal. It may be fundamentally Gaussian or merely a manifestation of the law of large numbers. For $\ell\ge14$ the sky signal is non-Gaussian, and this should manifest itself all the way upto much higher $\ell$s. However the instrument noise of the 4 Year COBE DMR data sets starts to dominate the $I^3_{\ell}$s at $\ell=18-20$, where the signal to noise drops below unity. Given that the noise has been shown to be extremely well characterized by a Gaussian, the sky map will manifest itself as Gaussian for these higher $\ell$s. Following they publications of our result, a number of groups have reported similar results: Pando {\it et al} \cite{pando} have applied a wavelet based technique and find evidence that the non-Gaussian signal is localized in the northern hemisphere (a result we tentatively confirm \cite{bigpaper}) while Novikov {\it et al} \cite{nov} have applied topological tests to detect non-Gaussianity in the COBE DMR data. According to skeptics, this may merely reflect a change in the psychological prior, triggered by our work. More seriously one should remember that the work performed by us and by these groups makes use of the same data set. Therefore this work provides an independent confirmation of our analysis of the DMR maps. In the very least this may mean a revision of the data analysis techniques which are currently in vogue for power spectrum estimation from CMB data sets: the underlying assumption in these Bayesian techniques is that the data set is Gaussian (greatly simplifying the estimation algorithms) \cite{gorski97}. However the fact that COBE data set seems to be non-Gaussian raises the question weather the current estimates of the $C_\ell$s of the COBE data set are as accurate as they are claimed to be. If the sky is truly non-Gaussian in the way we describe above, then we are at the threshold of uncovering its statistical nature with the higher resolution experiments that are coming online in particular with the BOOMERANG LDB experiment (described elsewhere in these proceedings) and ultimately with the MAP and Planck Surveyor satellite experiments. \section*{Acknowledgments} We thank the organizers for an excellent meeting. We thank JNICT, NASA-ADP, NASA-COMBAT, NSF, RS, Starlink, TAC for support.
\section{Introduction} One of the most fundamental models of quantum optical resonance is the interaction of a single two-level atom with a single quantised mode of radiation, described by the Jaynes-Cummings Hamiltonian \cite {Jaynes63,Shore93}. Despite being simple enough to be analytically soluble in the rotating-wave approximation, this model has been a long-lasting source of insight into the nuances of the interaction between light and matter. It has led to nontrivial predictions, such as the existence of `collapses' and `revivals' in the atomic excitation \cite{Eberly80}, and has also allowed a deeper understanding of the dynamical entangling and disentangling of the atom-field system in the course of time \cite {Phoenix88}. Further interest in the Jaynes-Cummings model (JCM) comes from the fact that these predictions are directly accessible to experimental verification. A JCM interaction can be experimentally realized in cavity-QED setups \cite{Haroche94} and also, as an effective interaction, in laser-cooled trapped ions \cite{Blockley92} (in which case the ionic harmonic motion assumes the role normally played by the field mode). For example, revivals in the atomic excitation have recently been observed in a cavity-QED experiment \cite{Brune96}, providing direct evidence for the discreteness of photons. In spite of these successes, a closed analytical description of the collapse-revival pattern has so far proved to be elusive; however, an elegant approximation scheme valid for a number of initial conditions has been presented by Fleischhauer and Schleich \cite{Fleischhauer93}, improving the earlier work of Eberly and co-workers \cite{Eberly80}. Among other things, they have demonstrated that, when the atom is initially completely excited or de-excited, and the initial photon-number distribution of the field ($P_{n}$) is sufficiently smooth (as is the case in a coherent state) then the shape of each revival is a direct reflection of the shape of $P_{n}$% . This direct relationship can be affected by the presence of initial atomic coherence. It has been noticed \cite{Zaheer89} that if the atom is initially prepared in a coherent superposition of its energy eigenstates, then the revivals can be largely suppressed, effectively freezing the value of the atomic state populations. Appropriately, these initial atomic states have been referred to as ``atomic trapping states'' \cite {Zaheer89,GeaBanacloche91}{\em . }This population trapping has been connected to the existence of a specific phase difference between the relative phase of the atomic superposition and the phase of the initial field state \cite{Zaheer89}. In particular, in the resonant case exact trapping has been shown to exist \cite{Cirac90} if the field is initially prepared in a phase coherent state \cite{Shapiro91} (eigenstate of the Susskind-Glogower phase operator \cite{Susskind64}). As has been pointed out by Cirac and S\'{a}nchez-Soto \cite{Cirac91}, this can be understood by the fact that in these cases only one state out of each pair of dressed eigenstates of the JCM is ever populated. We now note that, since the atomic excitation dynamics can be modified merely by altering the initial atomic state, then it is apparent that the initial photon distribution cannot be the sole responsible for the existence and shape of revivals. In fact, as we demonstrate in this paper, the key to understand the collapse-revival pattern under more general initial conditions is to consider not only the initial coherence of the atom and field by themselves, but also their joint properties as a single quantum system. This is so even if the system is initially disentangled. Building on Cirac and S\'{a}nchez-Soto's observation, we find that the revival structure depends{\em \ }essentially{\em \ }on the {\it relative weight} of each dressed eigenstate in the initial atom-field state. We{\em \ }are able to identify an atom-field variable which plays, in the general case, the same role as that of the photon distribution when the atom is initially excited or de-excited. This allows us to estimate for any initial condition the degree of suppression of the revivals (that is, the amount of trapping which occurs). In addition, in some particular cases we are able to describe analytically to a good approximation the shape of the revivals, even when they are partially suppressed. To illustrate the dressed-state distribution dependence of the revival patterns, we show that essentially identical examples can be found both in the presence of initial atomic coherence, and in cases where this coherence is not only atomic, but where atom and field are already entangled. Our work is organised as follows: in section II, we rewrite the dynamics of the Jaynes-Cummings model from a point of view based on the entangled dressed-state eigenbasis. With the help of an appropriate coordinate system, the expression for the atomic inversion assumes a simpler form; as a consequence, we are able to calculate bounds for the evolution of this quantity for arbitrary initial conditions. In order to illustrate our technique, specific cases are treated in section III. In section IV, we generalise Fleischhauer and Schleich's scheme to include the cases where trapping occurs. Our conclusions are summarised in section V. \section{Dressed-State Coordinates for the JCM} \subsection{The Model\label{modelsec}} The Jaynes-Cummings Hamiltonian, on resonance and in the rotating-wave approximation is given by \cite{Jaynes63,Shore93}: \begin{equation} H=\frac{1}{2}\hbar \omega \hat{\sigma}_{z}+\hbar \omega \hat{a}^{\dagger }\hat{a}+\hbar \lambda (\hat{\sigma}_{+}\hat{a}+\hat{a}^{\dagger }\hat{% \sigma}_{-}) \end{equation} where $\omega $ is the atomic transition frequency, $\lambda $ is the atom-field dipole coupling constant, $\hat{a}^{\dagger }(\hat{a})$ is the field photon creation (annihilation) operator and $\hat{\sigma}_{z},\hat{% \sigma}_{\pm }$ are the atomic inversion, rising and lowering operators. The first two terms in this expression describe the internal energy levels of the spin-1/2-like 2-level atom and the harmonic oscillator-like field mode, and the interaction term can be straightforwardly understood as a one-excitation exchange between these two systems.{\it \ } An important feature of this fully-quantised matter-radiation interaction is that its steady states (known as `dressed states') are{\em \ }{\it % entangled. }Switching to an interaction-picture representation for convenience, it can be shown that these states and their corresponding energy levels are \cite{Meystre91}: \begin{mathletters} \begin{eqnarray} \left| n\pm \right\rangle &=&\frac{1}{\sqrt{2}}\left[ \left| e,n\right\rangle \pm \left| g,n+1\right\rangle \right] \label{dressedstates} \\ E_{n\pm } &=&\pm \hbar \lambda \sqrt{n+1}\equiv \pm \frac{\hbar }{2}% \Omega _{n} \label{energies} \end{eqnarray} ($\Omega _{n}$ is the Rabi frequency). It is apparent that in all these cases the atomic state is completely undetermined, being an equal mixture of the excited and ground states. Thus, if we assume that at $t=0$ the atom and field are independently prepared in the states \end{mathletters} \begin{eqnarray} \left| \psi \left( 0\right) \right\rangle _{A} &=&p\left| e\right\rangle +q\left| g\right\rangle \\ \left| \Phi \left( 0\right) \right\rangle _{F} &=&\sum_{n=0}^{\infty }c_{n}\left| n\right\rangle \end{eqnarray} then the initial state $\left| \Psi \left( 0\right) \right\rangle _{AF}=\left| \psi \left( 0\right) \right\rangle _{A}\otimes \left| \Phi \left( 0\right) \right\rangle _{F}$ will evolve into \begin{eqnarray} \left| \Psi \left( t\right) \right\rangle _{AF} &=&qc_{0}\left| g,0\right\rangle + \nonumber \\ &&+\sum_{n=0}^{\infty }\left[ \begin{array}{c} pc_{n}\cos \left( \lambda t\sqrt{n+1}\right) - \\ -iqc_{n+1}\sin \left( \lambda t\sqrt{n+1}\right) \end{array} \right] \left| e,n\right\rangle + \nonumber \\ &&+\sum_{n=0}^{\infty }\left[ \begin{array}{c} qc_{n+1}\cos \left( \lambda t\sqrt{n+1}\right) - \\ -ipc_{n}\sin \left( \lambda t\sqrt{n+1}\right) \end{array} \right] \left| g,n+1\right\rangle \label{interevol} \end{eqnarray} Despite being straightforwardly solvable in this way, the JCM is well-known for the fact that the time-evolution of most expectation values is usually expressible only in series form. For instance, assume that at time $t=0$ we have: \begin{eqnarray} \left| \Psi \left( 0\right) \right\rangle &=&\left( p\left| e\right\rangle +q\left| g\right\rangle \right) \otimes \left( \sum_{n=0}^{\infty }c_{n}\left| n\right\rangle \right) \nonumber \\ &=&qc_{0}\left| g,0\right\rangle +\sum_{n=0}^{\infty }pc_{n}\left| e,n\right\rangle +qc_{n+1}\left| g,n+1\right\rangle \label{t=0fact} \end{eqnarray} The time evolution of the atomic population inversion is then given by: \begin{eqnarray} \left\langle \hat{\sigma}_{z}\right\rangle \left( t\right) &=&2\sum_{n=0}^{\infty }\left| pc_{n}\cos \left( \lambda t\sqrt{n+1}\right) -iqc_{n+1}\sin \left( \lambda t\sqrt{n+1}\right) \right| ^{2}-1 \nonumber \\ &=&-\left| qc_{0}\right| ^{2}+\sum_{n=0}^{\infty }\left( \left| pc_{n}\right| ^{2}-\left| qc_{n+1}\right| ^{2}\right) \cos \left( 2\lambda t% \sqrt{n+1}\right) + \nonumber \\ &&+2{\it Im} \left( \left( pc_{n}\right) ^{*}qc_{n+1}\right) \sin \left( 2\lambda t\sqrt{n+1}\right) \label{complinver} \end{eqnarray} where \begin{equation} \sum_{n=0}^{\infty }\left| pc_{n}\right| ^{2}+\left| qc_{n+1}\right| ^{2}=1-\left| qc_{0}\right| ^{2} \end{equation} In particular, if at $t=0$ the atom is completely excited $(q=0)$ or in the ground state $(p=0)$, and admitting for simplicity that $c_{0}$ is negligible, then this expression reduces to the simpler form \begin{equation} \left\langle \sigma _{z}\right\rangle \left( t\right) =\pm \sum_{n=0}^{\infty }P_{n}\cos \left( 2\lambda \sqrt{n+1}t\right) \label{invfactcase} \end{equation} where $P_{n}=\left| c_{n}\right| ^{2}$is the initial photon distribution of the field. Due to the $\sqrt{n}$ scaling of the various Rabi frequencies in this sum, no closed form for it is known. Nevertheless, it is apparent that features of this time evolution, such as collapse-revival phenomena, must be directly related to the characteristics of $P_{n}$. The Fleischhauer-Schleich approximation scheme, which we shall discuss in section IV, allows this relationship to be explicitly described in certain situations, such as when $P_{n}$ is sufficiently smooth. Apart from the simple case represented above, in general there is no way to relate the inversion dynamics exclusively to the initial field state. As can be seen from expression (\ref{complinver}), it will usually depend on atomic and field variables in an apparently complicated way. In the next section, we show how this evolution can be recast in a simple form, reminiscent of expression (\ref{invfactcase}), independently of the initial conditions. \subsection{Dressed-State Coordinates} Our discussion will be based on the relative contribution of each dressed state for the global atom-field state. For simplicity, we restrict ourselves to {\it pure }initial conditions. Therefore, we may expand an arbitrary initial state in terms of $\left| n\pm \right\rangle :$ \begin{equation} \left| \Psi (0)\right\rangle =w_{-1}\left| g;0\right\rangle +\sum_{n=0}^{\infty }w_{n}e^{i\chi _{n}}\left| \psi _{n}\right\rangle , \label{inista} \end{equation} where \begin{equation} \left| \psi _{n}\right\rangle =\left[ \cos \left( \frac{\theta _{n}}{2}% \right) \left| n+\right\rangle +e^{-i\phi _{n}}\sin \left( \frac{\theta _{n}% }{2}\right) \left| n-\right\rangle \right] , \label{dresscomp} \end{equation} with the normalisation condition $\sum_{n=-1}^{\infty }w_{n}^{2}=1.$ The parameters $w_{n}\in \left[ 0,1\right] ,$ $\theta _{n}\in \left[ 0,\pi \right] $ and $\chi _{n},\phi _{n}\in \left[ 0,2\pi \right] ,$ will henceforth be referred to as {\it dressed-state coordinates}. The use of this coordinate system allows us to obtain a simple geometrical picture of the manifold of pure atom-field states. For each value of $% n\geq 0$ one can imagine a Bloch-type spherical shell corresponding to the two-level system formed by $\left| n\pm \right\rangle $ (Fig. 1), which are parametrised by the spherical coordinates $\theta _n$ and $\phi _n.$ The poles $\left( \theta _n=0,\pi \right) $ are associated to the dressed states $\left| n\pm \right\rangle $, and points on the equator $% \left( \theta _n=\frac \pi 2\right) $ to the states of the form $\frac 1{% \sqrt{2}}\left[ \left| n+\right\rangle +e^{-i\phi _n}\left| n-\right\rangle \right] $, which include the particular product states $\left| e,n\right\rangle $ and $\left| g,n+1\right\rangle $ (corresponding to $\phi _n=0,\pi $). Therefore the coordinate $\theta _n$ gives a measure of what we shall call the `{\it dressedness}' (or degree of proximity to the nearest dressed state) of the components $\left| \psi _n\right\rangle $ of $\left| \Psi (0)\right\rangle $. We note that, while this property is related to the {\it entanglement }of $\left| \psi _n\right\rangle $ (since the dressed states are in fact maximally entangled), they are two different concepts: states with the same dressedness, such as $\left| e,n\right\rangle $ and $% \cos \mu \left| e,n\right\rangle -i\sin \mu \left| g,n+1\right\rangle $ can have different amounts of entanglement (as measured by the von Neumann entropy of their subsystems). Meanwhile, the weight factors $w_n$ measure the relative importance of each of these components. In physical terms, their squares $% w_n^2$ correspond to the probability distribution for measurements of the {\it total excitation number }operator{\it \ } \begin{equation} \hat{X}_{tot}=\frac 12\hat{\sigma}_z+\hat{N}. \end{equation} {\em \ }on the initial state $\left| \Psi \left( 0\right) \right\rangle $ ( $% w_n^2$ corresponds to the probability for $n+1$ excitations). We note that, in the special case where the atom is prepared in one of the states $\left| e\right\rangle $ or $\left| g\right\rangle $, then $\theta _n\equiv \frac \pi 2$ and $\phi _n\equiv 0\left( \pi \right) $ for all $n$, while $w_n^2$ reduces to the photon number distribution: \begin{equation} w_n^2\rightarrow \left\{ \begin{array}{c} P_n=\left| \left\langle n|\Phi \left( 0\right) \right\rangle _F\right| ^2\text{ (if }\left| e\right\rangle )\\ P_{n+1}=\left| \left\langle n+1|\Phi \left( 0\right) \right\rangle _F\right| ^2\text{ (if }\left| g\right\rangle ) \end{array}\right. . \end{equation} We now show that the dressed-state coordinate system allows a simple description of the time evolution of any initial state. From equations (\ref {dressedstates},b), the evolution of state (\ref{inista}) is given by \begin{equation} \left| \Psi \left( t\right) \right\rangle =w_{-1}\left| g,0\right\rangle +\sum_{n=0}^\infty w_ne^{\frac i2\left( 2\chi _n-\Omega _nt\right) }\left[ \cos \left( \frac{\theta _n}2\right) \left| n+\right\rangle +e^{-i\left( \phi _n-\Omega _nt\right) }\sin \left( \frac{\theta _n}2\right) \left| n-\right\rangle \right] . \end{equation} We can re-express this directly in terms of the dressed-state coordinates: \begin{eqnarray} w_{n}(t) &=&w_{n}(0) \nonumber \\ \theta _{n}\left( t\right) &=&\theta _{n}\left( 0\right) \\ \chi _{n}\left( t\right) &=&\chi _{n}\left( 0\right) -\frac{1}{2}\Omega _{n}t \nonumber \\ \phi _{n}\left( t\right) &=&\phi _{n}\left( 0\right) -\Omega _{n}t \nonumber \end{eqnarray} Thus, $w_{n}$ and $\theta _{n}$ are both constants of motion, while the angles $\chi _{n}$ and $\phi _{n}$ precess with a constant angular velocity proportional to the Rabi frequency $\Omega _{n}$. A way of visualising this is by means of a geometrical image as in Fig. 1. The simplicity of this picture reflects the symmetries of the resonant JCM: the conservation of $% \hat{X}_{tot}$ and the fact that, in the interaction picture, the energies in each 2-dimensional eigensubspace generated by $\left| n\pm \right\rangle $ are always symmetric with respect to zero. \subsubsection{Population Inversion in Dressed-State Coordinates\label% {inversec}} Let us now apply this coordinate system to describe the evolution of atomic variables.{\em \ }In the basis $\left\{ \left| e\right\rangle ,\left| g\right\rangle \right\} ,$ the reduced atomic density operator can be expressed as: \begin{equation} \rho _{A}\left( t\right) =\left[ \begin{array}{ll} \rho _{ee}\left( t\right) & \rho _{eg}\left( t\right) \\ \rho _{eg}^{*}\left( t\right) & \rho _{gg}\left( t\right) \end{array} \right] \label{atomDM} \end{equation} where \begin{mathletters} \begin{eqnarray} \rho _{ee}\left( t\right) &=&\frac{1}{2}\left( 1-w_{-1}^{2}\right) +\frac{1}{% 2}\sum_{n=0}^{\infty }w_{n}^{2}\sin \left( \theta _{n}\right) \cos \left( \phi _{n}\left( t\right) \right) ;\;\;\;\;\;\;\rho _{gg}\left( t\right) =1-\rho _{ee}\left( t\right) \label{rhoee(t)} \\ \rho _{eg}\left( t\right) &=&\frac{w_{-1}w_{0}}{\sqrt{2}}e^{i\left( \chi _{0}\left( t\right) \right) }\left( \cos \left( \frac{\theta _{0}}{2}\right) +e^{-i\phi _{0}\left( t\right) }\sin \left( \frac{\theta _{0}}{2}\right) \right) + \nonumber \\ &&+\frac{1}{2}\sum_{n=0}^{\infty }w_{n}w_{n+1}e^{i\left( \chi _{n+1}\left( t\right) -\chi _{n}\left( t\right) \right) }\left( \cos \left( \frac{\theta _{n+1}}{2}\right) +e^{-i\phi _{n+1}\left( t\right) }\sin \left( \frac{\theta _{n+1}}{2}\right) \right) \times \nonumber \\ &&\times \left( \cos \left( \frac{\theta _{n}}{2}\right) -e^{i\phi _{n}\left( t\right) }\sin \left( \frac{\theta _{n}}{2}\right) \right) \label{rhoeg(t)} \end{eqnarray} We thus see that the evolution of the atomic level populations, and therefore of the population inversion, can be expressed in a simple form for any initial condition: \end{mathletters} \begin{equation} \left\langle \hat{\sigma}_z\left( t\right) \right\rangle =2\rho _{ee}\left( t\right) -1=-w_{-1}^2+\sum_{n=0}^\infty w_n^2\sin \left( \theta _n\right) \cos \left( \phi _n\left( t\right) \right) . \label{popinv} \end{equation} This should be contrasted with the much more complicated expression (\ref {complinver}) for this same quantity, written in terms of separate atomic and field coordinates. In fact, we can see that the present expression is a direct generalisation of eq. (\ref{invfactcase}), obtained by phase-shifting the $n^{th}$ term in the sum by $\phi _n\left( 0\right) $ and by replacing the photon number distribution $P_n$ with \begin{equation} D_n=w_n^2\sin \left( \theta _n\right) . \label{weidist} \end{equation} It is not difficult to understand the reason for this substitution. First of all, as was noted by Cirac and S\'{a}nchez-Soto \cite{Cirac91}, dressed states have no population inversion. Therefore, if a component $\left| \psi _n\right\rangle $ of a given state $\left| \Psi \left( 0\right) \right\rangle $ has a large `dressedness' ($\sin \left( \theta _n\right) \rightarrow 0$), it will contribute very little to the overall inversion$.$ Moreover, even if $\left| \psi _n\right\rangle $ does have a large inversion, this may still be of little consequence to the total average $% \left\langle \hat{\sigma}_z\left( t\right) \right\rangle $ if its relative importance $w_n^2$ is small. Thus, the product $D_n,$ which we shall refer to as the `weighted dressedness distribution' gives the appropriate magnitude of the contribution of $\left| \psi _n\right\rangle $ to the inversion. In other words, for general initial conditions it is this distribution, not $P_n$, that will control the evolution of the inversion. For instance, we can expect the existence of collapse-revival structures if $% D_n$ is `narrow', in the sense that few values of $n$ (and therefore few Rabi frequencies $\Omega _n)$ feature significantly in the sum. \subsubsection{Population Trapping\label{trap}} The main difference between the general evolutions allowed by equation (\ref {popinv}) and those obtained in the special case (\ref{invfactcase}) lies in the fact that the distribution $P_{n}$ is normalised{\it ,} while $D_{n}$ is generally {\it not}. This implies that, in general, there will be an upper bound for the allowed range of variation of the population inversion, given by: \begin{equation} \left| \left\langle \hat{\sigma}_{z}\left( t\right) \right\rangle +w_{-1}^{2}\right| \leq \sum_{n=0}^{\infty }D_{n}\equiv M\leq 1 \label{bound} \end{equation} As we can see, this bound is specified essentially by the {\it average dressedness }of the components of $\left| \Psi \left( 0\right) \right\rangle $. For example: if the most important components (those with largest weights $w_{n}^{2}$) are highly dressed ($\sin \left( \theta _{n}\right) $ is small), then the amplitude of variation of the population inversion will also be small ($M\sim 0$).{\em \ }In other words, these are the conditions for the existence of {\it population trapping }in the (resonant) JCM{\it . }% The steady-state value of the inversion is given simply by $-w_{-1}^{2}$, i.e., by the fixed population in the ground state. We emphasise that, given any initial condition, the bound $M$ can be immediately calculated from the constants of the motion $w_{n},$ $\theta _{n}$, thus giving an estimate of the amount of trapping that can be expected from that state's evolution \cite {Hillery87}. \section{Examples} In order to illustrate our technique, we analyse three classes of initial atom-field states for which population trapping occurs: \subsection{Perfect Trapping States} In \cite{Cirac90} , Cirac and S\'{a}nchez-Soto found a class of factorised initial conditions which exhibit {\it perfect} trapping, that is, for which the atomic populations are strictly constant over time. These were of the form: \begin{equation} \left| \Psi \left( 0\right) \right\rangle =\sqrt{1-\left| z^{2}\right| }% \left( z\left| e\right\rangle +\left| g\right\rangle \right) \otimes \frac{1% }{\sqrt{1+\left| z^{2}\right| }}\sum_{n=0}^{\infty }z^{n}\left| n\right\rangle \label{exact} \end{equation} where $\left| z\right| <1.$ In this case, the field is prepared in an eigenstate of the Susskind-Glogower phase operator $\hat{V}% =\sum_{n=0}^{\infty }\left| n\right\rangle \left\langle n+1\right| $ \cite {Susskind64} (`phase-coherent state'\cite{Shapiro91}), while the atomic superposition is chosen so that the phase of the atomic dipole matches that of the eigenvalue $z$ of the field state \cite{remark}. The resulting population trapping can thus be attributed to a suitable matching of atomic and field parameters. However, as was later noticed by the same authors \cite {Cirac91}, a clearer explanation can be found by using a dressed-state point of view: Rewriting this state in terms of the dressed-state basis, we find \begin{equation} \left| \Psi \left( 0\right) \right\rangle =\sqrt{\frac{1-\left| z^2\right| }{% 1+\left| z^2\right| }}\left[ \left| g,0\right\rangle +\sum_{n=0}^\infty \sqrt{2}z^n\left| n+\right\rangle \right] . \label{exact2} \end{equation} The reason for the existence of trapping now becomes immediately clear: all the components $\left| \psi _n\right\rangle $ of this state are completely dressed ($\sin \left( \theta _n\right) \equiv 0$), and thus do not contribute to the inversion. Cirac and S\'{a}nchez-Soto also stated \cite{Cirac90} that these are the {\it only }factorised states exhibiting perfect trapping. We note that this is not strictly true: the same result is still achieved for all states of the form \begin{equation} \left| \Psi \left( 0\right) \right\rangle =\sqrt{1-\left| z^{2}\right| }% \left( z\left| e\right\rangle +\left| g\right\rangle \right) \otimes \frac{1% }{\sqrt{1+\left| z^{2}\right| }}\sum_{n=0}^{\infty }j\left( n\right) z^{n}\left| n\right\rangle \label{Exact3} \end{equation} where $j\left( n\right) $ can be $\pm 1$ and $\left| z\right| <1$ (in this case, the field state is not in general an eigenstate of $\hat{V}$). Once again, this can be seen by rewriting the state in the dressed-state basis. In this case, we obtain \begin{equation} \left| \Psi \left( 0\right) \right\rangle =\sqrt{\frac{1-\left| z^{2}\right| }{1+\left| z^{2}\right| }}\left[ j\left( 0\right) \left| g,0\right\rangle +\sum_{n=0}^{\infty }\sqrt{2}j\left( n\right) z^{n}\left| n\pm \right\rangle \right] \end{equation} where for each $n$ there is a single dressed state $\left| n\pm \right\rangle $ present, whose sign is the same as that of the ratio $\frac{% j\left( n+1\right) }{j\left( n\right) }.$ The converse statement is also easy to show: suppose $\left| \Psi \left( 0\right) \right\rangle $ is a perfect trapping state, of the form: \begin{eqnarray} \left| \Psi \left( 0\right) \right\rangle &=&k\left( z\left| e\right\rangle +\left| g\right\rangle \right) \otimes \sum_{n=0}^{\infty }a_{n}\left| n\right\rangle \nonumber \\ &=&ka_{0}\left| g,0\right\rangle +\sum_{n=0}^{\infty }za_{n}\left| e,n\right\rangle +a_{n+1}\left| g,n+1\right\rangle \end{eqnarray} where $k$ is a constant and $a_{n},$ $z$ are arbitrary. Then for each $% \left| \psi _{n}\right\rangle $ component of this state to be perfectly dressed it is necessary that: \begin{equation} a_{n+1}=\pm za_{n} \end{equation} and hence \begin{equation} a_{n}\propto \pm z^{n}. \end{equation} We recover expression (\ref{Exact3}) by noting that $\left| z\right| $ must be less than $1$ for $\left| \Psi \left( 0\right) \right\rangle $ to be normalisable. Thus, for the resonant JCM\ there can be no population trapping with positive population inversion. \subsection{Zaheer-Zubairy `atomic trapping states'} While the initial conditions considered above lead to perfect trapping, preparing the appropriate field states requires elaborate state-engineering techniques. It has been shown (numerically) by Zaheer and Zubairy, however, that {\it approximate} population trapping is also possible if the cavity field is initially in a coherent state $\left| \alpha \right\rangle $, which are easily accessible in experiment \cite{Zaheer89}. This is portrayed in Fig. 2: if the atom is initially completely excited, the inversion evolves according to the familiar collapse-revival pattern \cite {Eberly80}; however, by suitably rotating the atomic state, the revival peaks become more and more reduced, practically disappearing when the atom is prepared in state \begin{equation} \left| \psi \left( 0\right) \right\rangle _A=\frac 1{\sqrt{2}}\left( \left| e\right\rangle +e^{-i\nu _a}\left| g\right\rangle \right) \label{zz} \end{equation} (where $\nu _\alpha $ is the phase of the field's coherent amplitude $\alpha $). Close inspection also reveals that, in this limit, the revival peaks become split into doublets. In their work, Zaheer and Zubairy gave no quantitative explanation for this quenching of the revivals, attributing it qualitatively to a ``destructive interference between the atomic dipole and the cavity eigenmode ''. Later, Cirac and S\'{a}nchez-Soto pointed out that the existence of approximate trapping in this case{\em \ }should be expected due to similarities between the properties of coherent states and of the phase coherent field states \cite{Cirac90}. With the help of our dressed-state coordinate system, we can now give a quantitative description of this effect. Let us suppose that the atom-field system is prepared in the state: \begin{equation} \left| \Psi _{ZZ}\left( \alpha ,\gamma ,\xi \right) \right\rangle =\left[ \cos \gamma \left| e\right\rangle +e^{-i\xi }\sin \gamma \left| g\right\rangle \right] \otimes \left| \alpha \right\rangle , \label{zzstate} \end{equation} where the atom is in a coherent superposition of ground and excited states and the field is in a coherent state $\left( \alpha =|\alpha |e^{i\nu _{\alpha }}\right) $. Expanding this state in terms of the dressed-state basis, we find that its dressed-state coordinates have the following values (we have only listed those relevant for the atomic inversion): \begin{eqnarray} w_{n}^{2}\left( \alpha ,\gamma ,\xi \right) &=&\frac{|\alpha |^{2n}e^{-\left| \alpha \right| ^{2}}}{\left( n+1\right) !}\left[ \left( n+1\right) \cos ^{2}(\gamma )+|\alpha |^{2}\sin ^{2}(\gamma )\right] \text{ }% ,n\geq -1 \nonumber \\ \sin \theta _{n}\left( \alpha ,\gamma ,\xi \right) &=&\frac{\left| \left( n+1\right) \cos ^{2}(\gamma )-|\alpha |^{2}e^{i2\left( \nu _{\alpha }-\xi \right) }\sin ^{2}(\gamma )\right| }{\left( n+1\right) \cos ^{2}(\gamma )+|\alpha |^{2}\sin ^{2}(\gamma )}\;,\;\left( \theta _{n}\in \left[ 0,\pi \right] \right) \label{zzdc} \\ \cos \phi _{n}\left( \alpha ,\gamma ,\xi \right) &=&\frac{\left( n+1\right) \cos ^{2}(\gamma )-|\alpha |^{2}\sin ^{2}(\gamma )}{\left| \left( n+1\right) \cos ^{2}(\gamma )-|\alpha |^{2}e^{i2\left( \nu _{\alpha }-\xi \right) }\sin ^{2}(\gamma )\right| } \nonumber \\ \sin \phi _{n}\left( \alpha ,\gamma ,\xi \right) &=&\frac{|\alpha |\sqrt{% \left( n+1\right) }\sin (2\gamma )\sin \left( \nu _{\alpha }-\xi \right) }{% \left| \left( n+1\right) \cos ^{2}(\gamma )-|\alpha |^{2}e^{i2\left( \nu _{\alpha }-\xi \right) }\sin ^{2}(\gamma )\right| },\left( \phi _{n}\in \left[ 0,2\pi \right] \right) . \nonumber \end{eqnarray} The weight factors $w_{n}^{2}$ follow a Poisson distribution when $\gamma $ is equal to zero (they reduce to the photon number distribution of the coherent field) and have small deviations from that distribution for any value of $\gamma $. For moderately large $\left| \alpha \right| $, therefore, $w_{n}^{2}$ is maximised around the integer $n_{\max }$ closest to $\left| \alpha \right| ^{2}$ or $\left| \alpha \right| ^{2}-1.$ Meanwhile, it is not difficult to show that the $\sin \theta _{n}$ distribution has an absolute minimum, located around $n_{\min }\approx $ $% \left| \alpha \right| ^{2}\tan ^{2}\gamma -1,$ with minimum value given by \begin{equation} \sin \left( \theta _{n_{\min }}\right) \simeq \sin \left( \left| \nu _{\alpha }-\xi \right| \right) \end{equation} This indicates the component of $\left| \Psi \left( \alpha ,\gamma ,\xi \right) \right\rangle $ having the largest dressedness. Following the procedure outlined in section \ref{trap}, we can now see that to attain the highest possible degree of population trapping we must maximise the largest dressedness, and match it with the largest weight factor: \begin{mathletters} \begin{eqnarray} n_{\max } &=&n_{\min }\rightarrow \tan \gamma \simeq 1 \label{cond1} \\ \sin \left( \theta _{n_{\min }}\right) &=&0\rightarrow \nu _{\alpha }=\xi \label{cond2} \end{eqnarray} We thus explain the ``atomic trapping state'' (\ref{zz}) found by Zaheer and Zubairy: it is the one for which the total atom-field state $\left| \Psi _{ZZ}\left( \alpha ,\gamma ,\xi \right) \right\rangle $ most closely resembles a ``perfect trapping state'', in the sense that its most important components $\left| \psi _{n}\right\rangle $ closely resemble dressed states. It is clear from expression (\ref{zzdc}) that the degree of trapping experienced from initial state $\left| \Psi _{ZZ}\left( \alpha ,\gamma ,\xi \right) \right\rangle $ depends both on the relative weight $\gamma $ of each atomic state and on the phase difference $\left( \nu _\alpha -\xi \right) $ between the atomic dipole phase and the field's coherent amplitude. In Fig. 3 we illustrate the change in the weighted dressedness distribution $D_n$ as a function of these parameters:\ on the left-hand side, we take equal weights for $\left| e\right\rangle $ and $\left| g\right\rangle $ , satisfying condition (\ref{cond1}), and vary $% \left( \nu _\alpha -\xi \right) $ from $\frac \pi 2$ to $0$. We can see the quenching of the distribution as the dressedness of the component $\left| \psi _{n_{\max }}\right\rangle $ is increased. On the right-hand side, we keep the phase difference null and vary the relative weight from $0$ to $% \frac \pi 4$, obtaining a similar effect (in both cases, we have chosen $% \alpha =7$). Comparing this figure with Fig. 2, it is apparent that there is a striking similarity between the profile of the $D_n$ distribution and the envelope of the revivals in the corresponding time evolution, including the appearance of a doublet structure in the limit of the optimal trapping conditions. As we shall see in section IV, this is no coincidence. \subsection{`Even-odd' entangled states} The achievement of population trapping in the previous examples was seen to be ultimately caused by the quenching of the $D_{n}$ distribution, and therefore by the joint properties of the atom-field system. However, since the initial states considered were factorised into atomic and field states, it could also be argued that the trapping was due to a suitable matching of independent atomic and field parameters (i.e., the phase of the field's coherent amplitude matching the atomic dipole phase in the previous example). In order to stress the underlying importance of the dressed-state point of view, we now show that essentially identical evolution and trapping patterns for the population inversion can be obtained even when the atom and field cannot be considered as independent quantum systems. Consider the following class of states of the atom-field system: \end{mathletters} \begin{equation} \left| \Psi _{EO}\left( \alpha ,\gamma ,\xi \right) \right\rangle =\cos (\gamma )\left| e\right\rangle _A\left| even\right\rangle _f+\sin (\gamma )e^{i\xi }\left| g\right\rangle _A\left| odd\right\rangle _f, \label{evenodd} \end{equation} where $\left| even\right\rangle _f$\ $=\frac 1{\sqrt{2}}\left( \left| \alpha \right\rangle +\left| -\alpha \right\rangle \right) $ and $\left| odd\right\rangle _f=\frac 1{\sqrt{2}}\left( \left| \alpha \right\rangle -\left| -\alpha \right\rangle \right) $ are respectively even and odd coherent states of the cavity field (here $\left| \alpha \right| $ is assumed large enough so that these states can be considered normalised). It can be seen that, apart from the limiting cases where $\gamma \rightarrow 0,\frac \pi 2$, these states are {\it entangled}. Furthermore, since $\left| even\right\rangle _f$\ $\left( \left| odd\right\rangle _f\right) $\ have nonzero amplitudes only for even (odd) photon numbers, an expansion of this state in terms of dressed states $\left| n\pm \right\rangle $ features only those with {\it even }$n$ (in other words, $w_{2n-1}=0$). Physically, this implies that states of this form have {\it zero} average electric and magnetic fields, and {\it zero} average atomic polarisation. \begin{equation} \left\langle \hat{a}+\hat{a}^{\dagger }\right\rangle =\left\langle \hat{a}-% \hat{a}^{\dagger }\right\rangle =\left\langle \hat{\sigma}_x\right\rangle =\left\langle \hat{\sigma}_y\right\rangle =0 \end{equation} Thus, if at t = 0 the atom-filed system is in state $\left| \Psi _{EO}\left( \alpha ,\gamma ,\xi \right) \right\rangle $, the population inversion evolves according to eq. (\ref{popinv}), but with only the even terms present: \begin{equation} \left\langle \hat{\sigma}_z\left( t\right) \right\rangle =\sum_{n=0}^\infty w_{2n}^2\sin (\theta _{2n})\cos (\phi _{2n}(t)). \end{equation} Now, it turns out that all these even-indexed dressed-state coordinates have values virtually identical to the ones listed in expression (\ref{zzdc}) above for the state $\left| \Psi _{ZZ}\left( \alpha ,\gamma ,\xi \right) \right\rangle $ with the same parameters $\alpha ,\gamma ,\xi .$ The only difference is a multiplication of the weights $w_{2n}$ in the present case by a factor of 2, to preserve normalisation in spite of the absence of odd-indexed components. The resulting time-evolution of the atomic inversion, plotted in Fig. 4, shows that the quenching of revivals and appearance of a doublet structure as $\left( \nu _\alpha -\xi \right) \rightarrow 0$ proceed in essentially the same manner as in the previous case, the main difference being the doubling of the frequency of revivals \cite{remark2}.{\em \ }We note, however, that in the present case neither of the parameters $\nu _\alpha ,\xi ${\em \ }can be unambiguously assigned exclusively to the atom or field, so that population trapping in this case must necessarily be understood as a result of the joint atom-field properties of the state in question. It is also worth remarking that, due to the lack of atomic polarisation, the atomic reduced density operators of these states are always diagonal in the $% \left| e\right\rangle ,\left| g\right\rangle $ basis . Therefore, the {\it % reduced entropy } $S_a=-Tr\rho _a\ln \rho _a$, which measures the degree of entanglement between the atom and the field, is entirely determined by the population inversion: \begin{eqnarray} S_a\left( t\right) &=&-\rho _{ee}\left( t\right) \ln \left( \rho _{ee}\left( t\right) \right) -\left( 1-\rho _{ee}\left( t\right) \right) \ln \left( 1-\rho _{ee}\left( t\right) \right) \nonumber \\ &=&-\frac 12\left( 1-\left\langle \hat{\sigma}_z\left( t\right) \right\rangle \right) \ln \left( \frac 12\left( 1-\left\langle \hat{\sigma}% _z\left( t\right) \right\rangle \right) \right) \nonumber \\ &&-\frac 12\left( 1+\left\langle \hat{\sigma}_z\left( t\right) \right\rangle \right) \ln \left( \frac 12\left( 1+\left\langle \hat{\sigma}_z\left( t\right) \right\rangle \right) \right) . \end{eqnarray} Thus, in this case the existence of population trapping is equivalent to the atom and field remaining (nearly) maximally entangled during their entire time-evolution. For any given state $\left| \Psi _{EO}\left( \alpha ,\gamma ,\xi \right) \right\rangle ,$ a lower bound to the value of $S_a\left( t\right) $ at any given instant of the evolution is then given by: \begin{equation} S_{\min }=-\frac 12\left( 1-M\right) \ln \left( \frac 12\left( 1-M\right) \right) -\frac 12\left( 1+M\right) \ln \left( \frac 12\left( 1+M\right) \right) \label{sbound} \end{equation} where $M$ is the bound defined in eq. (\ref{bound}). For instance, in the case of the `trapping state' depicted in Fig. 4, where, $\alpha =7$% , $\gamma =\frac \pi 4,$ $\xi =0$ , we have $S_{\min }=0.69005$, very close to the maximum possible value $\ln 2\simeq 0.69315.$ Finally, we note that despite the existence of this lower bound, it is still possible to devise schemes by which such entangled atom-field states can be constructed \cite {Jonathan97,Jonathan98}. \section{Poisson Summation Formula for revivals in the case of population trapping} In \cite{Fleischhauer93}, Fleischhauer and Schleich obtained approximate analytical expressions for the evolution of the atomic inversion, using a stationary-phase method based on the {\it Poisson Summation Formula} \cite {Courant53}. They showed that in many cases the inversion can be written as a sum in which each term $\omega _k\left( t\right) $ is non-negligible only during a certain extension of time. This is in contrast for instance with the expressions used in section \ref{modelsec} above, where each term in the summation is periodic. Their work was restricted to initial states of the form $\left| g\right\rangle _{A}\otimes \left| \Phi \right\rangle _{F}$, in which the inversion is given by eq. (\ref{invfactcase}). They were able to show that, if the photon distribution $P_{n}$ of $\left| \Phi \right\rangle _{F}$ is sufficiently smooth, then the $k^{th}$ term in the Poisson sum for the inversion is given by \begin{equation} \omega _{k}\left( t\right) =-\frac{\lambda t}{\pi \sqrt{2k^{3}}}P\left( n=% \frac{\lambda ^{2}t^{2}}{4\pi ^{2}k^{2}}\right) \cos \left( \frac{\lambda ^{2}t^{2}}{2\pi k}-\frac{\pi }{4}\right) \end{equation} where $P\left( n\right) $ is a continuous interpolation of $P_{n}$. The main interest of this formulation is the fact that, if $P_n$ is also sufficiently narrow, then the term $\omega _k\left( t\right) $ describes to a high accuracy the evolution of the inversion during the $k^{th}$ revival. This can be readily seen from the formula above: this term describes a rapid oscillation in time, modulated by an envelope centred around $t\simeq \frac{% 2\pi k\sqrt{\left\langle n\right\rangle }}\lambda $ and whose format is essentially that of $P_n.$ Therefore, for narrow enough $P_n$, some of the terms $\omega _k\left( t\right) $ become completely disjoint from the rest, describing an independent revival. Fleischhauer and Schleich were able to use this formulation to derive many interesting results, such as the decrease in amplitude and increase in width of each successive revival, and also the number of revivals that can be resolved for a given state, before they become scrambled due to their increasing width and consequent interference with each other. Finally, they were also able to extend the technique to some cases where $P_n$ is not smooth, such as in squeezed states. Using our dressed-state formulation for the population inversion, eq. (\ref {popinv}), we are able to extend the Poisson Summation Formula method to even more general initial states, including ones with atomic or atom-field coherence. In the Appendix we show that, under suitable conditions, the behaviour of the atomic inversion during its $k^{th}$ revival is described by: \begin{equation} \left\langle \sigma _z\left( t\right) \right\rangle \simeq \left( \frac{% \lambda t}{\pi \sqrt{2k^3}}\right) \left. \left[ D\left( n\right) \cos \left( \phi _n\left( 0\right) +\frac{\lambda ^2t^2}{2\pi k}-\frac \pi 4\right) \right] \right| _{n+1=\frac{\lambda ^2t^2}{4\pi ^2k^2}} \label{kthrevival} \end{equation} where $D\left( n\right) $ is a continuous interpolation of the `weighted dressedness' distribution $D_n$ of the initial state. Thus, in general it is this distribution, not $P_n$, that is reflected in the shape of each revival. In particular, in the case of population trapping, the quenching of the revivals mirrors that of $D_n$. This explains the similarity between the profiles in Figs. 2 and 3, including the doublet structure of the revivals in the limit of maximum trapping. In order to display the accurateness of this expression, we use it to calculate the revivals in the case of the Zaheer-Zubairy states $\left| \Psi _{ZZ}\left( \alpha ,\gamma ,\xi \right) \right\rangle $ introduced above. From expressions (\ref{zzdc}) for the dressed coordinates, we see that in this case the weighted dressedness distribution is given by \begin{equation} D_n\left( \alpha ,\gamma ,\xi \right) =\sqrt{Q_1^2\left( n\right) +\ Q_2^2\left( n\right) -2Q_1\left( n\right) \cdot \ Q_2\left( n\right) \cdot \cos 2\left( \nu _\alpha -\xi \right) } \label{Ddist} \end{equation} where \begin{mathletters} \begin{eqnarray} Q_1\left( n\right) &=&\exp \left( -\left| \alpha \right| ^2\right) \frac{% \left| \alpha \right| ^{2n}}{\left( n\right) !}\cos ^2\left( \gamma \right) \\ \ Q_2\left( n\right) &=&\exp \left( -\left| \alpha \right| ^2\right) \frac{% \left| \alpha \right| ^{2(n+1)}}{\left( n+1\right) !}\sin ^2\left( \gamma \right) \end{eqnarray} For sufficiently large $\alpha $, these Poissonian distributions are well approximated by Gaussians: \end{mathletters} \begin{mathletters} \begin{eqnarray} Q_1\left( n\right) &\simeq &\frac{\cos ^2\left( \gamma \right) }{\sqrt{2\pi }% \left| \alpha \right| }\exp \left( \frac{-\left( n-\left| \alpha \right| ^2\right) ^2}{2\left| \alpha \right| ^2}\right) \\ Q_2\left( n\right) &\simeq &\frac{\sin ^2\left( \gamma \right) }{\sqrt{2\pi }% \left| \alpha \right| }\exp \left( \frac{-\left( n+1-\left| \alpha \right| ^2\right) ^2}{2\left| \alpha \right| ^2}\right) . \end{eqnarray} Extending these to continuous values of $n$ and substituting in (\ref{Ddist}% ) (\ref{kthrevival}) we obtain an analytical expression for the $k^{th}$ revival in the inversion. In Fig. 5 we plot the first two of these $(k=1,2)$ in the cases of maximum and minimum population trapping, alongside the corresponding exact evolutions obtained numerically. Despite a little distortion, agreement is seen to be very good. A similar calculation is also possible in the case of the entangled `even-odd' states $\left| \Psi _{EO}\left( \alpha ,\gamma ,\xi \right) \right\rangle $ presented in the previous section. In this case, formula (% \ref{kthrevival}) given above for the inversion is not valid, due to the strong oscillations in $D_{n}$ [$D_{n}=0$ for odd $n$]. However, it is possible to adapt our calculations to this situation (see section \ref {varisec} in the Appendix), obtaining the expression: \end{mathletters} \begin{equation} \left\langle \sigma _z\left( t\right) \right\rangle \simeq \left( \frac{% \lambda t}{\pi \sqrt{k^3}}\right) \left[ D\left( m\right) \cos \left( \phi _{2m}(0)+\frac{\lambda ^2t^2}{2\pi k}+\pi k-\frac \pi 4\right) \right] \left| _{m+1=\frac{\lambda ^2t^2}{\pi ^2k^2}}\right. \end{equation} for the inversion during the $k^{th}$ revival. Here $D\left( m\right) $ is a continuous interpolation of the even-indexed terms of $D_n$, renumbered with the new index $m$ $(D\left( m=3\right) =D_6$, for instance). Comparing with expression (\ref{kthrevival}) above for the Zaheer-Zubairy states, we can see that the main difference in the present case is that the frequency of revivals is {\it doubled}: the $k^{th}$ revival occurs around $% t_k^{EO}\simeq \frac{\pi k\sqrt{\left\langle m\right\rangle +1}}\lambda $ , compared to $t_k^{ZZ}=\frac{2\pi k\sqrt{\left\langle n\right\rangle +1}}% \lambda $ in the previous case. This difference is due essentially to the doubling of the separation between adjacent Rabi frequencies in the present case \cite{Antonio}, and is illustrated by comparing Figs. 2 and 4. \section{Conclusion} Intuitive pictures of the interaction between a two-level atom and an electric field commonly involve the expectation that the atomic level populations must change as both systems exchange excitations over the course of time \cite{Eberly85}. This is due to the absence of further atomic levels, which precludes the existence of destructive interference between different atomic transitions. However, in a fully-quantised interaction model such as the Jaynes-Cummings model, it is indeed possible to have states in which the atomic populations are completely or nearly completely trapped. This can be ultimately traced to the fact that the eigenstates of this model are entangled. We have shown that, by giving up the traditional point of view based on the individuality of each subsystem and assuming instead one based on these entangled dressed states, it is possible to obtain a quantitative understanding of population trapping in this model. This is achieved via the introduction of a set of joint atom-field state variables, the `weighted dressedness' distribution $D_{n}$ (eq. \ref{weidist}). We have shown that, for general initial conditions, this distribution governs the evolution of the atomic inversion, assuming a role commonly attributed to the photon number distribution $P_{n}$ (to which it reduces in particular cases).\ Using $D_{n}$, we have obtained an upper bound to the amplitude of population oscillations that can be expected from a given state at any instant of its evolution. We have also been able to obtain an approximate analytical description of the behaviour of the atomic inversion during partially suppressed revivals. We have found that in general the shape of revival envelopes is a direct reflection of the form of (a continuous interpolation of ) $D_{n}$. In the particular case of a field initially in a coherent state, this explains the appearance of a doublet structure in the revivals in the limit of greatest population trapping. \section{Acknowledgements} We would like to thank Dr. S.M. Dutra and Prof. P.L. Knight for helpful comments and discussions. This work was supported by Conselho Nacional de Desenvolvimento Cient\'{i}fico e Tecnol\'{o}gico (CNPq) and Funda\c{c}\~{a}o de Amparo \`{a} Pesquisa do Estado de S\~{a}o Paulo (FAPESP), Brazil. \begin{appendix} \section{Approximate expressions for revivals in the atomic population inversion for generalised initial conditions\label{AppendixA}} In this Appendix, we calculate approximate expressions for the revivals in the atomic population inversion, which are valid for a wide variety of {\it % pure }initial conditions of the atom-field system. These calculations generalise the method presented by Fleischhauer and Schleich \cite {Fleischhauer93}, who assumed an initially factorised state of the form $% \left| g\right\rangle _A\otimes \left| \psi \right\rangle _F.$ The result they obtained for the $k^{th}$ revival is: \begin{equation} \left\langle \hat{\sigma}_z\left( t\right) \right\rangle \simeq -P\left( n=% \frac{\lambda ^2t^2}{4\pi ^2k^2}\right) \frac{\lambda t}{\pi \sqrt{2k^3}}% \cos \left( \frac{\lambda ^2t^2}{2\pi k}-\frac \pi 4\right) \label{Fleisresult} \end{equation} where $P_n$ is the photon distribution of state $\left| \psi \right\rangle _C.$ In the case of more general initial conditions, including ones with atomic and atom-field coherence, we shall see that the form of the revival envelope depends not on the state's photon-number distribution, but on the `weighted dressedness' distribution $D_n$ given in equation (\ref{weidist}). As was shown in section \ref{inversec}, the inversion, written in terms of dressed-state coordinates, has the form: \begin{equation} \left\langle \hat{\sigma}_{z}\left( t\right) \right\rangle =-w_{-1}+\sum_{n=0}^{\infty }D_{n}\cos \left( \phi _{n}\left( t\right) \right) \label{inversion0} \end{equation} where \begin{equation} \phi _{n}\left( t\right) =\phi _{n}\left( 0\right) -\Omega _{n}t \end{equation} This expression can be rewritten according to the Poisson Summation Formula (see Courant and Hilbert \cite{Courant53}, p.. 76 ) \begin{equation} \left\langle \hat{\sigma}_z\left( t\right) \right\rangle =\sum_{k=-\infty }^\infty \omega _k\left( t\right) +\tau _0\left( t\right) \label{inversion} \end{equation} where \begin{mathletters} \begin{eqnarray} \omega _k\left( t\right) &=&\int_0^\infty dnD\left( n\right) \cos \left( \phi _n\left( 0\right) -\Omega _nt\right) e^{2i\pi kn} \label{wn} \\ \tau _0\left( t\right) &=&\frac 12D\left( 0\right) \cos \left( \phi _0\left( 0\right) -2\lambda t\right) -w_{-1} \end{eqnarray} and where $D\left( n\right) $ is {\it any }`reasonable' function of a continuous variable $n$ (continuous, differentiable, etc.), that interpolates between the values of $D_n$ at the points where $n$ integer. Noting that the sum in $k$ in $\left( \ref{inversion}\right)$ extends to $% \pm \infty $, so that the expression is invariant under $k\leftrightarrow -k$% , it is possible to substitute eq. $\left( \ref{wn}\right)$ by: \end{mathletters} \begin{eqnarray} \omega _k\left( t\right) &=&\int_0^\infty dnD\left( n\right) \cos \left( \phi _n\left( 0\right) -2S_k\left( n,t\right) \right) = \\ &=&\int_0^\infty dnD\left( n\right) \cos \left( \phi _n\left( 0\right) \right) \cos \left( 2S_k\left( n,t\right) \right) + \nonumber \\ &&+\int_0^\infty dn\sin \left( \phi _n\left( 0\right) \right) \sin \left( 2S_k\left( n,t\right) \right) \\ &=&{Re}\int_0^\infty dnD_1\left( n\right) \exp \left( 2iS_k\left( n,t\right) \right) +{Im}\int_0^\infty dnD_2\left( n\right) \exp \left( 2iS_k\left( n,t\right) \right) \end{eqnarray} where we have defined \begin{mathletters} \begin{eqnarray} S_k\left( n,t\right) &=&\pi kn-\lambda t\sqrt{n+1} \label{Phase} \\ D_1\left( n\right) &=&D\left( n\right) \cos \left( \phi _n\left( 0\right) \right) \label{E1} \\ D_2\left( n\right) &=&D\left( n\right) \sin \left( \phi _n\left( 0\right) \right) . \label{E2} \end{eqnarray} In this way, the inversion may be rewritten (without any approximation) as \end{mathletters} \begin{equation} \left\langle \hat{\sigma}_{z}\left( t\right) \right\rangle =\sum_{k=-\infty }^{\infty }\omega _{k}^{1}\left( t\right) +\omega _{k}^{2}\left( t\right) +\tau _{0}\left( t\right) \label{inversionexp} \end{equation} where \begin{mathletters} \begin{eqnarray} \omega _{k}^{1}\left( t\right) &\equiv &{Re} \int_{0}^{\infty }dnD_{1}\left( n\right) \exp \left( 2iS_{k}\left( n,t\right) \right) \\ \omega _{k}^{2}\left( t\right) &\equiv &{Im} \int_{0}^{\infty }dnD_{2}\left( n\right) \exp \left( 2iS_{k}\left( n,t\right) \right) \end{eqnarray} Now, assuming the envelopes $D_1\left( n\right) $ and $D_2\left( n\right) $ are sufficiently smooth if compared to the oscillating functions $\cos \left( 2S_k\left( n,t\right) \right) ,\sin \left( 2S_k\left( n,t\right) \right) ,$ we may apply the method of stationary phases, approximating these expressions by: \end{mathletters} \begin{mathletters} \begin{eqnarray} \omega _k^1\left( t\right) &\simeq &D_1\left( n=n_k\right) {Re}\left\{ \begin{array}{c} \exp \left( 2iS_k\left( n=n_k\right) \right) \times \\ \times \int_0^\infty dn\exp \left[ i\left. \frac{\partial ^2S_k}{\partial n^2}\right| _{n=n_k}\left( n-n_k\right) ^2\right] \end{array} \right\} \label{statphaprox} \\ \omega _k^2\left( t\right) &\simeq &D_2\left( n=n_k\right) {Im}\left\{ \begin{array}{c} \exp \left( 2iS_k\left( n=n_k\right) \right) \times \\ \times \int_0^\infty dn\exp \left[ i\left. \frac{\partial ^2S_k}{\partial n^2% }\right| _{n=n_k}\left( n-n_k\right) ^2\right] \end{array} \right\} \end{eqnarray} where $n_{k}$ is the point at which $\frac{\partial S_{k}}{\partial n}=0:$% \end{mathletters} \begin{equation} n_{k}+1=\frac{\lambda ^{2}t^{2}}{4\pi ^{2}k^{2}} \label{statpoint} \end{equation} We note that these expressions are invalid for $k=0,$ since in this case the phase is always stationary (=0). Substituting in $\left( \ref{Phase}\right) , $we obtain: \begin{eqnarray} S_{k}\left( n=n_{k}\right) &=&-\left( \pi k+\frac{\lambda ^{2}t^{2}}{4\pi k}% \right) \label{S} \\ \left. \frac{\partial ^{2}S_{k}}{\partial n^{2}}\right| _{n=n_{k}} &=&2\frac{% \pi ^{3}k^{3}}{\lambda ^{2}t^{2}}\equiv F \label{K} \end{eqnarray} Now, the integral in $\left( \ref{statphaprox},b\right) $may be written in terms of the Fresnel integrals \cite{Gradsteyn65}: \begin{mathletters} \begin{eqnarray} C\left( x\right) &=&\sqrt{\frac{2}{\pi }}\int_{0}^{x}dy\cos \left( y^{2}\right) dy \\ S\left( x\right) &=&\sqrt{\frac{2}{\pi }}\int_{0}^{x}dy\sin \left( y^{2}\right) dy \end{eqnarray} For instance, for $F>0:$% \end{mathletters} \begin{equation} \int_{0}^{\infty }dn\exp \left( iF\left( n-n_{k}\right) ^{2}\right) =\sqrt{% \frac{\pi }{2F}}\left[ C\left( x\rightarrow \infty \right) +C\left( \sqrt{F}% n_{k}\right) +i\left( S\left( x\rightarrow \infty \right) +S\left( \sqrt{F}% n_{k}\right) \right) \right] \end{equation} The asymptotic form of the Fresnel integrals for $x\rightarrow \infty $ \cite {Gradsteyn65} is: \begin{mathletters} \begin{eqnarray} C\left( x\right) &\simeq &\frac{1}{2}+\sqrt{\frac{1}{2\pi }}\frac{\sin \left( x^{2}\right) }{x}+O\left( \frac{1}{x^{2}}\right) \\ S\left( x\right) &\simeq &\frac{1}{2}+\sqrt{\frac{1}{2\pi }}\frac{\cos \left( x^{2}\right) }{x}+O\left( \frac{1}{x^{2}}\right) \end{eqnarray} Assuming $\sqrt{\left| F\right| }n_{k}\gg 1$ and taking these approximations to zeroth order, we have: \end{mathletters} \begin{equation} \int_{0}^{\infty }dn\exp \left( iF\left( n-n_{k}\right) ^{2}\right) \simeq \sqrt{\frac{\pi }{2F}}\left( 1+i\right) \text{ }(\text{valid for }F>0) \label{asympt1} \end{equation} Similarly, for $F<0$% \begin{equation} \int_{0}^{\infty }dn\exp \left( iF\left( n-n_{k}\right) ^{2}\right) \simeq \sqrt{\frac{\pi }{2\left| F\right| }}\left( 1-i\right) \label{asympt2} \end{equation} Using $\left( \ref{K}\right) $ and $\left( \ref{statpoint}\right) ,$% condition $\sqrt{\left| F\right| }n_{k}\gg 1$ becomes: \begin{mathletters} \begin{eqnarray} \lambda t &\gg &2\left( \sqrt{\pi \left| k\right| }+\sqrt{2\pi \left| k\right| +4\pi ^{2}k^{2}}\right) \label{tcond} \\ \text{ou} &\ll &2\left( \sqrt{\pi \left| k\right| }-\sqrt{2\pi \left| k\right| +4\pi ^{2}k^{2}}\right) \end{eqnarray} for $k=1,$ for example, this requires $\lambda t\gg 17$ Substituting the asymptotic expressions $\left( \ref{asympt1}\right) ,\left( \ref{asympt2}\right) $in $\left( \ref{statphaprox},b\right) ,$ we obtain for $k>0$ (or $k<0$):% \end{mathletters} \begin{mathletters} \begin{eqnarray} \omega _{k}^{1}\left( t\right) &\simeq &D_{1}\left( n=n_{k}\right) \sqrt{% \frac{\pi }{2F}}\left[ \cos \left( \left. 2S_{k}\right| _{n=n_{k}}\right) \mp \sin \left( \left. 2S_{k}\right| _{n=n_{k}}\right) \right] \\ \omega _{k}^{2}\left( t\right) &\simeq &D_{2}\left( n=n_{k}\right) \sqrt{% \frac{\pi }{2F}}\left[ \cos \left( \left. 2S_{k}\right| _{n=n_{k}}\right) \pm \sin \left( \left. 2S_{k}\right| _{n=n_{k}}\right) \right] \end{eqnarray} Thus, using $\left( \ref{K}\right) $ and $\left( \ref{inversionexp}\right) ,$% the inversion may be written: \end{mathletters} \begin{equation} \left\langle \hat{\sigma}_{z}\left( t\right) \right\rangle = \sum_{{k=-\infty, }{k\neq 0} } ^{\infty }\left( \frac{\lambda t}{2\pi k^{\frac{3}{2}}}% \right) \left[ \begin{array}{c} D_{1}\left( n=n_{k}\right) \left[ \cos \left( \left. 2S_{k}\right| _{n=n_{k}}\right) \mp \sin \left( \left. 2S_{k}\right| _{n=n_{k}}\right) \right] + \\ D_{2}\left( n=n_{k}\right) \left[ \cos \left( \left. 2S_{k}\right| _{n=n_{k}}\right) \pm \sin \left( \left. 2S_{k}\right| _{n=n_{k}}\right) \right] \end{array} \right] +\omega _{0}^{1}+\omega _{0}^{2}+\tau _{0} \end{equation} (where the upper(lower) sign is valid for the terms with $k>0(<0)$. Finally, substituting the value $\left( \ref{S}\right) $ for $\left. S_{k}\right| _{n=n_{k}}:$ \begin{eqnarray} \left\langle \hat{\sigma}_{z}\left( t\right) \right\rangle &=&\sum_{{% k=-\infty, }{k\neq 0} }^{\infty }\left( \frac{\lambda t}{2\pi k^{\frac{3}{2}}}% \right) \left[ \begin{array}{c} \left( \left. D_{1}+D_{2}\right| _{n=n_{k}}\right) \cos \left( 2\pi k+\frac{% \lambda ^{2}t^{2}}{2\pi k}\right) \pm \\ \pm \left( \left. D_{1}-D_{2}\right| _{n=n_{k}}\right) \sin \left( 2\pi k+% \frac{\lambda ^{2}t^{2}}{2\pi k}\right) \end{array} \right] +\omega _{0}^{1}+\omega _{0}^{2}+\tau _{0} \nonumber \\ &=&\sum_{{k=-\infty, }{k\neq 0} }^{\infty }\left( \frac{\lambda t}{% 2\pi k^{\frac{3}{2}}}\right) \left. D\left( n\right) \right| _{n=n_{k}}\left[ \begin{array}{c} \left( \left. \cos \left( \phi _{n}\left( 0\right) \right) +\sin \left( \phi _{n}\left( 0\right) \right) \right| _{n=n_{k}}\right) \cos \left( \frac{% \lambda ^{2}t^{2}}{2\pi k}\right) \pm \\ \pm \left( \left. \cos \left( \phi _{n}\left( 0\right) \right) -\sin \left( \phi _{n}\left( 0\right) \right) \right| _{n=n_{k}}\right) \sin \left( \frac{% \lambda ^{2}t^{2}}{2\pi k}\right) \end{array} \right] + \nonumber \\ &&+\omega _{0}^{1}+\omega _{0}^{2}+\tau _{0} \end{eqnarray} Finally, since: \begin{equation} \left( \cos \left( x\right) +\sin \left( x\right) \right) \cos \left( y\right) \pm \left( \cos \left( x\right) -\sin \left( x\right) \right) \cos \left( y\right) =\sqrt{2}\cos \left( x\pm y-\frac \pi 4\right) \end{equation} then the approximate expression for the inversion is: \begin{equation} \left\langle \hat{\sigma}_z\left( t\right) \right\rangle =\sum_{{% k=-\infty, }{k\neq 0}}^\infty \left( \frac{\lambda t}{\sqrt{2}\pi k^{\frac 32}% }\right) \left. \left[ D\left( n\right) \cos \left( \phi _n\left( 0\right) \pm \frac{\lambda ^2t^2}{2\pi k}-\frac \pi 4\right) \right] \right| _{n+1=% \frac{\lambda ^2t^2}{4\pi ^2k^2}}+\omega _0^1+\omega _0^2+\tau _0 \label{approxinversion} \end{equation} where the upper (lower) sign is valid for the terms with $k>0$ $(<0),$and where $\tau _0$ and $\omega _0^1+\omega _0^2$ are given by \begin{mathletters} \begin{eqnarray} \tau _0\left( t\right) &=&\frac 12D\left( 0\right) \cos \left( \phi _0\left( 0\right) -2\lambda t\right) -w_{-1} \label{aux} \\ \omega _0^1+\omega _0^2\left( t\right) &=&\omega _0\left( t\right) =\int_0^\infty dnD\left( n\right) \cos \left( \phi _n\left( 0\right) -2\lambda t\sqrt{n+1}\right) \end{eqnarray} The first of these two last terms represents the contribution to the inversion of the states in which the field is in a vacuum state. For the initial conditions which satisfy the approximations that have been made above, this term will normally be negligible (see below). The second term assumes non-negligible values only in the vicinity of $t=0$. This is due to the fact that it is an integral over oscillating functions with different frequencies, which rapidly cancel each other out; also, since these frequencies form a continuum, they cannot re-phase substantially at subsequent times. Fleischhauer and Schleich have thus conjectured that zero-order terms of the Poisson Formula such as this always describe the first {\it collapse }of the inversion shortly after $t=0$. The remaining terms ( $k\neq 0$ ) each describe a modulated oscillation{\it % \ }in the inversion, with an envelope given essentially by the shape of $% D\left( n\right) $ and assuming non-negligible values only in an interval around $t\simeq \frac{2\pi k\sqrt{\left\langle n\right\rangle +1}}\lambda $ (here $\left\langle n\right\rangle $ represents a value of $n$ around the peak of $D\left( n\right) $). If $D\left( n\right) $ is sufficiently narrow, the first few of these modulated oscillations will be well-separated in time, thus constituting an independent {\it revival} during which the inversion is described by \end{mathletters} \begin{equation} \left\langle \hat{\sigma}_z\left( t\right) \right\rangle \simeq \left( \frac{% \lambda t}{\pi \sqrt{2k^3}}\right) \left. \left[ D\left( n\right) \cos \left( \phi _n\left( 0\right) \pm \frac{\lambda ^2t^2}{2\pi k}-\frac \pi 4\right) \right] \right| _{n+1=\frac{\lambda ^2t^2}{4\pi ^2k^2}} \end{equation} When the initial state is of the form$\left| g\right\rangle _A\otimes \left| \psi \right\rangle _C$ , we have \begin{equation} \phi _n\left( 0\right) \rightarrow \pi \text{ },\text{ }D\left( n\right) \rightarrow P_{n+1} \end{equation} ( $P_n$ being the photon distribution of $\left| \psi \right\rangle _C)$, so that Fleischhauer and Schleich's result (\ref{Fleisresult}) is recovered (eq. $\left( 2.8b\right) $ in \cite{Fleischhauer93} ). This expression is valid as long as: \begin{enumerate} \item[a) ] The distributions $D_{n}\cos \left( \phi _{n}\left( 0\right) \right) $ and $D_{n}\sin \left( \phi _{n}\left( 0\right) \right) $ vary slowly with $n$ if compared with $\cos S_{k}\left( n,t\right) =\cos (\pi kn-\lambda t\sqrt{n+1})$. \item[b) ] The value of $t$ obeys conditions $\left( \ref{tcond},b\right) .$ This implies that, for the $k^{th}$ term of the sum above to describe well the $k^{th}$ revival, $D\left( n\right) $ must assume its largest values in the region of $n$ where: \begin{equation} n+1\gg \frac{4\left( 3\pi k+4\pi ^{2}k^{2}+2\pi k\sqrt{2+4\pi k}\right) }{% 4\pi ^{2}k^{2}}=4+\frac{1}{2\pi k}\left( \frac{3}{2}+\sqrt{2+4\pi k}\right) \end{equation} Thus, the approximation can be expected to be good for all revivals if the initial state has at least 10 or so photons on average in the field. This will usually also ensure that the component $\tau _{0}$, which depends on $% D\left( 0\right) $, (eq. (\ref{aux})), can be ignored. \end{enumerate} It is straightforward to show that in the examples of section IV, where the dressed-state coordinates are given by expressions (\ref{zzdc}) and where $\left| \alpha \right| =7$, both of these conditions are satisfied. \subsection{Variation for `even-odd' states [where $w_{2n-1}=0$]\label% {varisec}} In the case of `even-odd'-type states such as $\left| \Psi _{EO}\left( \alpha ,\gamma ,\xi \right) \right\rangle $, where only the dressed-state coordinates $w_{n}$ with {\it even }index are non-null $(w_{2n-1}=0),$ condition (a) above is violated and expression (\ref{approxinversion}) is thus invalid. Nevertheless, a similar analytical expression for the inversion can still be derived if eq. $\left( \ref{inversion0}\right) $ is rewritten considering only the terms with even $n.$ In this case, it is straightforward to show that, as long as the `continuous versions' of distributions $D_{2n}\cos \left( \phi _{2n}\left( 0\right) \right) $ and $% D_{2n}\sin \left( \phi _{2n}\left( 0\right) \right) $ are sufficiently smooth {\it as a function of }${\it n}$, then the same stationary-phase method as was used above can be applied, resulting in: \begin{equation} \left\langle \hat{\sigma}_{z}\left( t\right) \right\rangle \simeq \sum_{% {k=-\infty, }{k\neq 0}}^{\infty }\left( \frac{\lambda t}{\pi k^{\frac{3% }{2}}}\right) \left. \left[ D\left( m\right) \cos \left( \phi _{2m}\left( 0\right) \pm \frac{\lambda ^{2}t^{2}}{2\pi k}+\pi k-\frac{\pi }{4}\right) \right] \right| _{m+1=\frac{\lambda ^{2}t^{2}}{\pi ^{2}k^{2}}}\text{ }% +\omega _{0}^{1}+\omega _{0}^{2}+\tau _{0} \end{equation} Here $D\left( m\right) $ is a continuous interpolation of the even-indexed terms of $D_{n}$, renumbered with the new index $m$ $(D\left( m=3\right) =D_{6}$, for instance). $\tau _{0}$ and $\omega _{0}^{1}+\omega _{0}^{2}$ are still given by $\left( \ref{aux},b\right) ,$except one must substitute $% n\rightarrow m.$ Once again, it is possible to show that the approximations realised in the course of obtaining this formula remain valid as long as $D\left( m\right) $ assumes significant values only for $2m\gtrsim 10.$ (Thus, whenever the formula is applicable the term $\tau _0$ can be ignored). \end{appendix}
\section{Introduction} In 1997 Franco Selleri \cite{selleri}, in his long quest for inconsistencies in the special relativity theory (SRT), pointed out a paradox concerning the speed of light as measured on board a rotating disk. Actually his point treats the speed of light along a closed circuit encircling the rotation axis: when the platform is moving, the speed obtained dividing the length of the contour by the time of flight, as measured in the ''relative space'' of the disk (defined in Sec. 3) by an observer at rest on the platform, is different whether measured in the rotation sense or in the opposite sense. More explicitly, suppose a light beam is sent along the rim in the rotation sense and another one in the opposite sense; then measure the average velocities of both beams for a complete round trip, i.e. the ratio between the length of the path and the times of flight read on a clock at rest on the rim, and call them $c_{+}$ and $c_{-}$; then the ratio $\rho =c_{+}/c_{-} $ differs from 1. Since a rotating reference frame is not inertial, this anisotropy of the light propagation is not, on itself, a surprising result; however Selleri notices that when letting the platform's radius $R$ go to infinity and the angular speed $\omega $ go to $0$ in such a way that the peripheral speed $% \omega R$ of the turntable remains constant, the ratio $\rho $ too keeps a constant value. However in the limit of infinite radius the uniform rotation becomes a uniform translation, i.e. the local reference frame becomes inertial. Here, according to SRT, the speed of light is assumed to be exactly the same in any direction (since all inertial frames are assumed to be optically isotropic); hence $\rho $ must be strictly $1$. This alleged discontinuity in the behavior of $\rho $, under such limit process, is the core of what we could call Selleri's paradox. This issue has already been discussed elsewhere \cite{rita}, showing that a full 4-dimensional relativistic treatment of the problem of the rotating platforms avoids any discontinuity or inconsistency whatsoever, since the speed of light, consistently defined, turns out to be exactly the same both clockwise and counterclockwise, just as in an inertial reference frame. However, though the 4-dimensional geometric point of view is clear and consistent, nothing prevents from considering the problem from a different viewpoint, rather natural for an observer living on the rotating disk. Then some doubt is apparently allowed \cite{selleri}, \cite{selleri2}, since the ratio $\rho $, when measured by means of meter rods and clocks (or rather a single clock) at rest on the platform, actually differs from 1. The root of Selleri's paradox can be identified in the basic assumption - founded on the homogeneity of the disk along the rim - that the ''global ratio'' $\rho =c_{+}/c_{-}$ of the average light velocities for complete round trips, coincides with the ''local ratio'' $\rho _{o}$ of the forward and backward light velocities. We shall however show, analyzing the actual measurement procedures of the velocities in both cases, that $\rho $ cannot in general be assumed to equal $\rho _{o}$, contrary to the claim by Selleri. In fact, $\rho $ does not depend on the criterium for simultaneity adopted along the rim, as Selleri correctly points out; $\rho _{o}$ does instead strictly depend on the local simultaneity criterion. The two ratios, which Selleri labels by the same letter $\rho $, refer to two different kinds of measurements, so that there is no point in comparing them: they are and remain different, whatever the size of $R$ is, be it finite or infinite, with no harm for SRT. That this was the weak point of Selleri's argument has been already remarked also by Budden \cite{budden}. In sect. 2 the four-dimensional approach considered in ref. [2] is breafly reexamined. Sect. 3 discusses two possible alternative definitions of space of the platform along the rim, and compares the Minkowskian and the operational approach to the interpretation of the measurements of space and time intervals on board the rotating disk. Sect. 4 draws the general conclusions. \section{Constancy of the speed of light in Minkowski spacetime} On a formal point of view, the SRT is the description of a four-dimensional manifold, whose geometrical structure is uniquely determined by two principles: the Einstein relativity principle and the principle of constancy of the (one way) velocity of light in vacuum\footnote{% Of course, the axiomatic basis of the SRT is not completely established by the two principles mentioned above, but also embodies the so-called ''principle of locality'', which states the local equivalence of any accelerated reference frame with a momentarily comoving inertial frame.}. As well known, such manifold is the familiar Minkowski spacetime, in which the time evolution of any massive particle is described in terms of a world line $\gamma _{m}$ which lies everywhere inside the light cone associated with any point of $\gamma _{m}$. This can be visualized in a standard spacetime diagram (in which space and time are measured by the same unities and the coordinate lines are drawn orthogonal to each other) as a world line whose slope, although variable, is everywhere greater than 45$^{o}$. Only massless particles, particularly photons, are described by null world lines, i.e. by world lines whose slope, in the graphic representation, is always 45$^{o}$: any light beam in free spacetime is described by a 45$^{o}$ slanting straight line, which can be regarded as a generator of the light cone. This is a geometrical expression of the principle of costancy of the one way velocity of light in minkowskian spacetime. The interaction of the light beam with a mirror may change the space direction of propagation of the beam, curving the trajectory in space and the world line in spacetime, {\it without affecting its slope}. As a consequence, when a light beam is lead to move along the rim of a rotating disk, grazing a cylindrical mirror, its world line in $2+1$ dimensions turns out to be a ''null helix'' wrapped around the world tube of the disk and keeping everywhere a 45$^{o}$ slope. A $2+1$ geometrical analysis of the Sagnac effect (see \cite{rita}) shows how and why the times of flight for co-rotating and counter-rotating beams are different, {\it although their world lines are helixes of constant (45}$% ^{o}${\it ) slope}; or, frasing it differently, {\it although their velocities are the same - namely c - in any inertial frame, in particular in the local inertial comoving frame at any point of the rim.} To sum up, the special relativistic assumption of the constancy of the slope of the world lines of light does not lead to inconsistencies or unphysical discontinuities. On the operational point of view, this means that, in the framework of SRT, the apparent global anisotropy of the propagation of light along the rim is perfectly compatible with the local isotropy{\it , }% contrary to Selleri's assumption. \section{Actual measurements of the speed of light} Once the internal consistency of the geometry of Minkowskian spacetime has been established again, it still remains to confront it with the operational procedures an observer at rest on the rotating disk uses, in order to attribute actual values to the physical quantities of interest. The problem is that any measurement concerning the geometry of the disk and the synchronization of clocks on it is a well defined set of physical and mathematical operations on an extended region of space (in particular along the rim), whereas in a rotating frame special relativistic formulae are merely local: any result obtained by extrapolating them globally cannot be considered as a pure consequence of SRT, but depends on some (usually hidden) further assumptions. In our opinion, the presence of recurrent contradictions and paradoxes simply underlines the arbitrariness of such extrapolations, from local to global. Now, the obvious operational way to define and determine the (one way) speed of a (massive or not) moving object is to measure the length of a given travel and the time it takes, then divide the former by the latter. Of course this procedure determines the slope of the world line of the moving object only when the measurements are local (infinitesimal extension of the space and time intervals); finite measurements can yield the slope only in very special cases (constant slope world lines). In the case of uniform rotation and of light travelling along the rim of the rotating disk, the slope of the light world line is constant as well as that of the observer's one; as a consequence, it can be determined not only by (a sequence of) local measurements of space and time intervals, performed in the local comoving frames, but also by global measurements of space and time intervals referred to a complete (either co-rotating or counter-rotating) round trip. However, in the second case the operational procedure should be carefully defined, because of the presence of some unavoidable conventional extrapolations, as pointed out before. More precisely: (i) the measure of the length of a complete round trip depends on the definition of ''space on the platform'', at least along the rim; (ii) the time taken by the light beam for a complete round trip is an observable quantity (it is the proper time lapse of a single clock), but the impossibility of a global synchronization along the rim could require a suitable correction. The form of the correction is imposed by the space-time geometry, according to the particular definition chosen for the space of the platform (see later). Among the many possible definitions of ''space on the platform along the rim'', we consider in particular the following two: (i) the ''space of locally Einstein simultaneous events'', defined as the set of events along the rim such that any nearby pair of them are simultaneous according to the Einstein criterium; (ii) the ''relative space'' $S:=T/I$, defined as the quotient of the world tube $T$ of the disk by the congruence $I$ of the word lines of the points of the disk. The former is obtained extrapolating the local Einstein synchronization procedure to the whole rim of the disk. This space coincides with the space-like helix $\gamma _{S}$ considered at the beginning of Sec. 4 of ref. 2, and is everywhere Minkowski-orthogonal to the time-like helixes corresponding to the world lines of the points of the rim. The latter turns out to be the space of locations on the disk, regardless of any kind of synchronization: ''two points of spacetime which lie on the same disk word line ... are identified in the relative space'' \cite{tim}. This space seems rather artificial on a Minkowskian point of view, but it should appear quite natural for the observer on the platform, since the space spanned by meter sticks arranged on the platform by this observer is precisely the ''relative space'' of the disk. Notice that we used both spaces, namely:\ the ''space of Einstein locally simultaneous events'' when we adopt a Minkowskian approach, like in the main part of \cite{rita}; and the ''relative space'' when we adopt an operational approach, like in sect. 5 of \cite{rita} and everywhere in \cite{tartaglia}. \subsection{{\bf \ }{\it Minkowskian approach}} In this approach, the ''space of locally Einstein simultaneous events'' along the rim coincides with the space-like helix $\gamma _S$ considered before; the important point is that $\gamma _S$ is not a circumference (this is true only in the absence of rotation), but an open line whose slope depends on the rotation velocity. Now, the proper length of an open line is not a uniquely defined entity; we showed in particular in \cite{rita} that the geometry of Minkowskian spacetime imposes different lengths for the portion of $\gamma _S$ covered by the co-rotating and by the counter-rotating light beams in a complete round trip. The difference in these two lengths turns out to be \begin{equation} \delta s_{\gamma _{S}}=\frac{4\pi \left( \omega R\right) }{c\sqrt{1-\left( \omega R\right) ^{2}/c^{2}}}R \label{1} \end{equation} which exactly coincides, dividing by $c$, with the difference in time of flight along the two round trips (see eq. (\ref{2}) later, which is consistent with the Sagnac effect). As a consequence, this definition of space ensures the equality of the global speed of light both for the co-rotating and the counter-rotating light beams, restoring the isotropy of light propagation. We point out that this definition of space is the only one which can insure the equality between global and local velocities, i.e. between the ''global ratio'' $\rho $ and the ''local ratio'' $\rho _{o}$: this agrees with Selleri's assumption, but both ratios equal exactly 1, with no harm for the SRT. \subsection{\it Operational approach} In this approach, the ''relative space'' $S$ along the rim allows the observer at rest on the platform to consider a unique length for the rim of the disk (see sect. 5 of \cite{rita} and \cite{tartaglia}). The measure of the two round trip times is obtained by one single clock (no need for special synchronization procedures), and gives two different results. In particular, the difference in time between the two round trips is \begin{equation} \delta \tau =\frac{4\pi \left( \omega R\right) }{c^{2}\sqrt{1-\left( \omega R\right) ^{2}/c^{2}}}R \label{2} \end{equation} which is an expression of the Sagnac effect \cite{sagnac}, \cite{stedman} . In this case, the observer can draw the following conclusions, on the basis of his measurements of space and time on the platform and without any knowledge of Minkowskian spacetime structure (see \cite{rita}, sect.5): (i) the platform on which he lives is rotating, and the desynchronization $% \delta \tau $ of a pair of clocks, after slow round trips in opposite directions, is a measure of the speed of this rotation; (ii) the durations of travels along the closed path are not uniquely defined and, to obtain reliable measures of them, the readings of clocks must be corrected by a quantity $\pm \delta \tau /2$ to account for the desynchronization effect, which is the same result obtained by Bergia and Guidone \cite{bergia}; (iii) as a consequence of this correction, the speed of light is actually the same both forward and backward. On the other hand, if the readings are used without any theoretical correction, the global measurement actually gives an anisotropic result at all radii (as far as $\omega R\neq 0$); but this procedure, which is the one proposed by Selleri, cannot prove his basic assumption, only founded on the homogeneity of the disk along the rim, that the ''global ratio'' $\rho $ coincides with the ''local'' one $\rho _{o}$. In fact, the measurements performed by the observer on the platform in order to calculate the two ratios are completely different. The value of the ''global ratio'' turns out to be \begin{equation} \rho =\frac{c_{+}}{c_{-}}=\frac{c-\omega R}{c+\omega R}\;\; \label{3} \end{equation} and depends on the measurement of a difference of proper times read on a single clock. This measurement is independent from any assumption about synchronization. On the contrary, the ''local ratio'' $\rho _{o}$ depends: (i) on the measurement of two infinitesimal lengths (forward and backward) in the local comoving frame; (ii) on the readings of three clocks (placed at the starting point of the light beams and at the arrival points, in opposite directions), Einstein synchronized in the local comoving frame. If Einstein synchronization is used, the ''local ratio'' $\rho _{o}$ is exactly 1, and cannot be identified with the ''global ratio'' $\rho $, which differs from 1. One could object that also a local measurement of the light velocity can be performed by means of a single clock, when the light beam is reflected by a mirror placed at an infinitesimal distance from the source (two ways average light speed). But in this case - that is what is usually made in actual experiments, like e.g. Michelson-like experiments - the difference of measurement procedures is still more evident: the global method, which measures two one-way velocities of two light beams performing two complete round trips along a closed path in opposite directions, cannot be used in the local inertial frame, in which only the two ways light speed is measurable. So the ''global ratio'' $\rho $ only is an observable; the ''local ratio'' $\rho _{o}$ is not. Selleri's assumption is \begin{equation} \rho =\rho _{o}=\frac{c-\omega R}{c+\omega R}\neq 1\;\;\;\;\;\;% \forall % \omega \neq 0 \label{quattro} \end{equation} This assumption is equivalent to assuming a suitable non Einstein synchronization in the local comoving frame, which could be called ''Selleri synchronization'', consistent with the condition \cite{selleri2}, \cite {budden}: \begin{equation} c_{+}=c\left( 1+\frac{\omega R}{c}\right) ^{-1}\;;\;\;\;c_{-}=c\left( 1-% \frac{\omega R}{c}\right) ^{-1} \label{5} \end{equation} Such a synchronization requires of course a suitable non Lorentz coordinate transformation, which in turn implies the existence of a privileged frame and the absolute character of synchronization, see \cite{selleri}, \cite {selleri2}, \cite{selleri3}. An obvious consequence of eqs. (\ref{5}) is that light propagates anisotropically in any local comoving frame along the rim, but the observable two ways light speed is again $c$. As a consequence, the ''Selleri synchronization'' does not conflict with known experiments, but conflicts with the standard Einstein synchronization (which assumes that light propagates isotropically in any inertial frame: $c_{+}=c_{-}$ $=c$). If Selleri's synchronization is used, the SRT is violated; in this case Selleri's paradox only shows that, starting from an assumption violating the SRT, a result violating the SRT follows. We cannot treat, in the limits of this letter, the question of which synchronization (Einstein or Selleri) is more adequate to the whole experimental and theoretical context: we limit ourselves to claiming that both are consistent and compatible with experiments, but the ''serious logical problem in the SRT'' declared by Selleri does not exist. \section{Conclusion} To sum up our line of thought, we have seen that the direct measurement of the speed of light along a closed path, free of any theoretical corrections, does indeed reveal an anisotropy when the observer is rotating along the contour. It would continue to be so also for a contour of infinitely great curvature radius, though it is impossible to actually perform the experiment. On the other side local measurements of the speed of light cannot evidence any anisotropy. The global and local ratios between forward and backward light velocities, which Selleri labels by the same letter $\rho $, refer to two different kinds of measurements, which cannot be reduced one to the other: they are and remain different, whatever the size of the platform radius is, be it finite or infinite.The two classes of measurements do not overlap and do not reveal, in the framework of the SRT, any internal contradiction.
\section{Introduction} Hole-doped perovskite-type manganese oxides have attracted considerable interest in recent years, motivated by the observation of colossal magnetoresistance (CMR) in numerous related compounds, and the great variety of magnetic and transport properties in this class of materials.\cite{Review} Among the La-based systems, the ground state of the stoichiometric parent compound LaMnO$_{3}$ is insulating A-type antiferromagnetic (AF), which is attributed to a cooperative effect of orbital ordering and superexchange interactions.\cite{SuExch} Substitution of a fraction $x$ of La$^{3+}$ by divalent cations such as Sr$^{2+}$, Ca$^{2+}$ or Ba$^{2+}$ causes the conversion of a proportional number of Mn$^{3+}$ to Mn$^{4+}$. At certain doping ranges ($0.2 \lesssim x \lesssim 0.5$) this induces a metal-insulator transition and the appearance of a ferromagnetic (FM) state. The simultaneous FM and metallic transitions have been qualitatively explained by the double-exchange (DE) model,\cite{Zen} which considers the magnetic coupling between Mn$^{3+}$ and Mn$^{4+}$ resulting from the motion of an electron between the two partially filled $d$ shells. Nevertheless, this DE mechanism does not account for several experimental results, and it has been claimed\cite{Millis} that a Jahn-Teller type electron-phonon coupling plays an important role in explaining the large magnetoresistive effects. Conversion of Mn$^{3+}$ to Mn$^{4+}$ can also be achieved by the presence of non-stoichiometric oxygen in undoped LaMnO$_{3+\delta}$, with a nominal Mn$ ^{4+}$ content of $2\delta$. For simplicity this is the crystallographic representation used in the present work and in most other studies in this system. However, it does not reflect the fact that the system contains randomly distributed La and Mn vacancies rather than oxygen excess, which can not be accommodated interstitially in the lattice.\cite{Oxy} The actual crystallographic formula is better written as La$_{1-x}$Mn$_{1-y}$O$_3$. By varying the oxygen stoichiometry the resulting compounds display a wide variety of structural and magnetic phases, previously studied by x-rays and neutron scattering,\cite{Ritter,Alon2} as well as magnetic and transport measurements.\cite{Ritter,Les} It is well known that the low-temperature magnetic phase of non-stoichiometric LaMnO$_{3+\delta}$ changes from AF to FM for small values of $\delta$ due to the DE interaction caused by the presence of Mn$^{4+}$ ions in the sample. However, unlike the cation-doped systems, the material remains insulating at all temperatures, and the FM transition temperature decreases for increasing content of Mn$^{4+}$. The relevant fact to be considered appears to be the competing effect between La vacancies, enhancing the Mn$^{3+}$-Mn$^{4+}$ DE interaction, and Mn vacancies which introduce considerable disorder in the lattice. For large values of $\delta$ the FM order is suppressed, and the low-temperature phase is better described by a spin-glass-like state.\cite{Ritter,Les} The competing effect between cation and manganese vacancies makes LaMnO$ _{3+\delta }$ a model system for studying magnetic interactions and disorder effects in mixed-valence manganites. In order to achieve a better understanding of the low temperature properties of this system we have performed magnetic and specific-heat measurements in three different samples of non-stoichiometric LaMnO$_{3+\delta}$. Magnetic data show signatures of a double transition: as the temperature is lowered, the system first orders ferromagnetically in small weakly-connected clusters, and then changes to a cluster-glass phase. Results of low-temperature specific-heat measurements show an unexpectedly large linear coefficient and a spin-wave contribution. This is interpreted in terms of the existence of disorder-induced charge-localization in these compounds. \section{Experiments} The bulk samples of LaMnO$_{3+\delta}$ investigated in the present study were thoroughly characterized in Refs.\ \onlinecite{Alon2,Les,Alon3}. They were prepared in polycrystalline form by a citrate technique, as described elsewhere.\cite{Alon2} The products were annealed at 1100 $^\circ$C in air (Sample 1), 1000 $^\circ$C in air (Sample 2) and 1000 $^\circ$C under 200 bar of O$_2$ (Sample 3). The determination of $\delta$ was initially performed by thermogravimetric analysis. The final materials were characterized by x-ray diffraction. Neutron powder diffraction diagrams were also collected in the temperature range 2-250$\;$K. The Rietveld method was used to refine the crystal and magnetic structures. The neutron-diffraction refinements showed that all investigated samples have stoichiometric oxygen content of $3.00\pm 0.05$. The Mn$^{4+}$ content was calculated from the vacancy concentration of La and Mn determined from the neutron data, and found to be in good agreement with the thermogravimetric analysis. Sample 1, with $\delta =0.11$ and 23\% of Mn$ ^{4+}$, consists of a mixture of a main orthorhombic phase (64\%) and a minor rhombohedral phase (36\%). Sample 2, with $\delta =0.15$ and 33\% of Mn $^{4+}$, and Sample 3, with $\delta =0.26$ and 52\% of Mn$^{4+}$, both have rhombohedral symmetry. Samples 1 and 2 showed a FM ordered structure at low temperatures (with some canting observed in Sample 2), whereas Sample 3 showed spin-glass-like signatures. Transport measurements\cite{Les} revealed that all the studied compounds are insulating down to low temperatures, with a typical semiconductor-like behavior. Selected sample parameters are summarized in Table I. \bigskip \vbox{\narrowtext \begin{table} \caption{Selected physical parameters and preparation conditions of the LaMnO $_{3+\delta}$ samples. The actual crystallographic formula is better written as La$_{1-x}$Mn$_{1-y}$O$_3$. The FM transition temperature is given by $T_c$ .} \label{tab:1} \begin{tabular}{c|ccc} Sample & 1 & 2 & 3 \\ $\delta $ & 0.11 & 0.15 & 0.26 \\ $x$, $y$ & 0.022, 0.054 & 0.029, 0.069 & 0.029, 0.128 \\ Mn$^{4+}$ (\%) & 23 & 33 & 52 \\ Prep.\ Conditions & 1100 $^\circ$C/air & 1000 $^\circ$C/air & 1000 $^\circ$C/O$_{2}$ \\ Cryst.\ Structure & Ortho./Rhomb. & Rhomb. & Rhomb. \\ $T_c$ (K) & 154 & 142 & -- \end{tabular} \end{table} } \narrowtext \begin{figure} \begin{center} \includegraphics[width=7.50cm,angle=0,clip]{ghiv-fig1.ps} \end{center} \caption{Real and imaginary parts of the AC susceptibility of LaMnO$ _{3+\delta }$, measured in an alternating field $h_{ac}$ = 1$\;$Oe and frequencies $f = 25$, 125 and 1000$\;$Hz. For samples with $\delta = 0.11$ and 0.15 (Samples 1 and 2) the data is shifted vertically for clarity. For the sample with $\delta = 0.26$ (Sample 3) the results are multiplied by 3.} \label{fig:1} \end{figure} In the present study, specific-heat results were obtained from 4.5 to 200 K with an automated quasi-adiabatic pulse technique. The absolute accuracy of the data, checked against a copper sample, is better than 3\%. The measured samples had masses of approximately 50$\;$mg. Detailed AC susceptibility and DC magnetization measurements where performed in a commercial magnetometer (Quantum Design PPMS). The FM transition temperatures of Samples 1 and 2, obtained from AC susceptibility data, are also shown in Table I. \section{Magnetic Measurements} Figure\ 1 shows the AC susceptibility of LaMnO$_{3+\delta }$. Real and imaginary parts, respectively $\chi ^{\prime }$ and $\chi ^{\prime \prime }$ , were measured in zero DC field, with an alternating field $h_{ac}=1\;$Oe, and frequencies of 25, 125 and 1000 Hz. Results for Sample 3 are multiplied by a factor of 3. Part \end{multicols} \widetext \begin{figure} \begin{center} \includegraphics[width=16.0cm,angle=0,clip]{ghiv-fig2.ps} \end{center} \caption{Field-cooled (FC) and zero-field-cooled (ZFC) DC magnetization of LaMnO$_{3+\delta }$. The applied field, from bottom to top in the figures, is 50, 200, 500 and 5000$\;$Oe.} \label{fig:2} \end{figure} \begin{multicols}{2} \noindent of the data was shifted vertically for clarity. The first point to note is a pronounced FM transition, observed at 154 and 142 K for Samples 1 ($\delta =0.11$) and 2 ($\delta =0.15$), respectively. The values of T$_{c}$ were determined from the maximum derivative in $\chi ^{\prime }$. For Sample 3, with the higher vacancy content ($\delta =0.26$), at 48$\;$K we observe a much lower cusp-like anomaly in $\chi ^{\prime }$, typical of a spin glass behavior. As mentioned in the introduction, the evolution from FM to spin glass features for increasing oxygen content in LaMnO$_{3+\delta }$ was previously observed in the literature.\cite {Ritter,Les} Moreover, it is most interesting to note in Fig.\ 1 that the results for the two FM samples show a double-peak structure and a frequency dependence of the imaginary component, $\chi^{\prime\prime}$. The high-temperature peak is frequency independent, whereas the position of the low-temperature peak strongly depends on the measuring frequency. The maximum in $ \chi^{\prime\prime}$ shifts to higher temperatures as the frequency increases. These are clear signatures of a cluster-glass behavior, as previously reported for other manganite\cite{doubleX} and cobaltite\cite {Co1,Co2} systems. The high-temperature peak signals the onset of FM order, whereas the low-temperature frequency-dependent peak is associated with freezing of the cluster magnetic moments. In connection with the low-temperature peak in $\chi^{\prime\prime}$, a frequency-dependent shoulder can be observed in the real component $\chi^{\prime}$. Results for Sample 3 also show a distinct frequency dependence in $\chi^{\prime}$, not visible in the scale of the figure. In order to probe disorder-induced features in the system, we have measured the field-cooled (FC) and zero-field-cooled (ZFC) magnetization of the studied samples. Figures\ 2(a) and 2(b) display the results for Samples 1 and 2 respectively. The low-field data was taken with $H=50\;$Oe. A pronounced irreversibility is observed, again indicative of a disordered state. In our results the irreversibility starts just below T$_{c}$, which is the typical behavior of a cluster-glass phase, whereas in reentrant-spin-glass systems irreversibility occurs far below T$_{c}$. As the field increases the irreversible behavior is reduced, and is no longer present at $H=5000\;$Oe. Measurements of AC susceptibility with an applied DC field (not shown) confirm that the frequency dependence in $\chi ^{\prime }$ disappears with increasing fields. These results show that the application of a DC field tends to align the cluster moments, and stabilizes a reversible FM ordered state. In the magnetization results for Sample 3, shown in Fig.\ 2(c), the behavior is quite different. The magnetization peak is more than two orders of magnitude lower than in the other samples, and the irreversibility persists with higher applied DC field, which confirms the standard spin-glass features in the high-vacancy sample. The difference between the ZFC and FC magnetizations is much higher in the cluster-glass phase (Samples 1 and 2) compared to the spin-glass phase (Sample 3), reflecting the presence of FM order within the clusters.\cite{Co1} Isothermal $M$ vs.\ $H$ curves measured at 10$\;$K are plotted in Fig.\ 3. For the FM samples (1 and 2) the magnetization saturates at fields of the order of 1--2$\;$T. The saturation values are $3.70\,\mu _{B}$ and $ 3.57\,\mu _{B}$ for Samples 1 and 2, respectively. The magnetic moment expected from the spin contribution is $gS\,\mu _{B}$, where $S$ is the spin of the ion, which is 3/2 for Mn$^{4+}$ and 2 for Mn$^{3+}$, and the gyromagnetic factor $g=2$ in both cases. Taking into account the relative concentrations of Mn$^{4+}$ and Mn$^{3+}$ in the compounds, we get an effective moment of $3.77\,\mu _{B}$ for Sample 1 (23\% Mn$^{4+}$), and $ 3.67\,\mu _{B}$ for Sample 2 (33\% \narrowtext \begin{figure} \begin{center} \includegraphics[width=7.50cm,angle=0,clip]{ghiv-fig3.ps} \end{center} \caption{Magnetization as a function of field of LaMnO$_{3+\delta }$, measured at 10$\;$K. The inset shows low-field data for Sample 1. The arrows indicate measurements increasing and decreasing the field.} \label{fig:3} \end{figure} \noindent Mn$^{4+}$). This prediction virtually coincides with the values observed experimentally, indicating that the applied field fully polarizes the FM clusters. Hysteresis is observed at very low fields, up to about 400 Oe, as shown for Sample 1 in the inset of Fig.\ 3. This hysteresis is consistent with the $M$ vs.\ $T$ data of Fig.\ 2, and is attributed to the low field cluster-glass nature of the samples. For Sample 3 (52\% Mn$^{4+}$), the low-temperature magnetization does not saturate at our highest field, and a large hysteretic behavior is observed. At 9$\;$T the measured magnetic moment is $2.15\,\mu _{B}$, much smaller than the predicted value of $3.48\,\mu _{B}$. This is an additional indication of the spin-glass-like properties of this sample. In order to verify the consistency of our magnetic results, we have performed the same measurements on another similar series of LaMnO$ _{3+\delta}$ samples. The cluster-glass behavior of the intermediate-vacancy FM samples, i.e., the frequency-dependent AC susceptibility and the irreversibility in low-field magnetization, were confirmed to exist in this second series of samples. \section{Specific Heat Measurements} Figure\ 4(a) shows the specific heat of the investigated samples plotted as $C/T$ vs. $T^{2}$, in the temperature range of 4.5--15$\;$K. For comparison, measurements on La$_{0.90}$Ca$_{0.10}$MnO$_{3}$, a ferromagnetic insulator, and \nopagebreak on La$_{0.67}$Ca$_{0.33}$MnO$_{3}$, ~~a ~ferromagnetic ~metal, ~~are ~also \narrowtext \begin{figure} \begin{center} \includegraphics[width=7.50cm,angle=0,clip]{ghiv-fig4.ps} \end{center} \caption{(a) From top to bottom, low-temperature specific heat, plotted as $ C/T$ vs.\ $T^{2}$, for LaMnO$_{3+\delta}$ with $\delta = 0.26$ (circles), 0.15 (down triangles), and 0.11 (up triangles), and for La$_{1-x}$Ca$_x$MnO$ _3$ with $x = 0.11$ (squares) and 0.33 (diamonds). (b) Plot of $C$ vs.\ $T$ for LaMnO$_{3+\delta }$; the solid lines are fitted curves, as discussed in the text.} \label{fig:4} \end{figure} \noindent displayed.\cite{Ghiv} The latter has the same Mn$^{4+}$ content as in LaMnO$_{3+\delta }$ with $\delta =0.15$. However, it is clear from the figure that the heat capacity is considerably higher in LaMnO$_{3+\delta }$ as compared to the Ca-doped compounds. In order to interpret these results and evaluate the different contributions to the specific heat, the low-temperature data of each studied sample were fitted to the expression \begin{equation} C=\gamma T+\beta T^{3}+BT^{3/2}. \end{equation} The linear coefficient $\gamma $ is usually attributed to charge carriers, and is proportional to the density of states at the Fermi level. However, transport measurements showed that all the investigated samples of LaMnO$ _{3+\delta }$ are insulating, and the appearance of a linear term in the specific heat must be more carefully interpreted. The lattice contribution is given by $\beta T^{3}$. A higher-order lattice term proportional to $ T^{5} $ was not needed to fit the data in the temperature range up to 10$\;$ K. The term $BT^{3/2}$ is associated with FM spin-wave excitations. The coefficient \end{multicols} \widetext \begin{table}[tbp] \caption{Fitting results of the low-temperature specific heat of LaMnO$% _{3+\delta }$. The linear coefficient is given by $\gamma$. The Debye temperature $\theta_D$ is obtained from the cubic coefficient $\beta $, and the spin-wave stiffness constant $D$ is obtained from the magnetic term $% BT^{3/2}$.} \label{tab:2} \begin{tabular}{cccccc} Sample & $\gamma$ (mJ/mol$\,$K$^{2}$) & $\beta$ (mJ/mol$\,$K$^{4}$) & $% \theta_{D}$ (K) & $B$ (mJ/mol$\,$K$^{5/2}$) & $D$ (meV\AA$^{2}$) \\ \hline 1 ($\delta = 0.11$) & $23 \pm 3$ & $0.193 \pm 0.02$ & $369 \pm 13$ & $2.1 \pm 0.6$ & $75 \pm 15$ \\ 2 ($\delta = 0.15$) & $19 \pm 2$ & $0.168 \pm 0.02$ & $387 \pm 15$ & $7.5 \pm 1.0$ & $32 \pm 3$ \\ 3 ($\delta = 0.26$) & 0 & $0.0786 \pm 0.007$ & $498 \pm 15$ & $21.2 \pm 0.2$ & $16.1 \pm 0.1$% \end{tabular} \end{table} \begin{multicols}{2} \noindent $\beta $ is related to the Debye temperature $\theta _{D}$, and the coefficient $B$ to the spin-wave stiffness constant $D$.\cite{book} The fitting parameters obtained for all samples are given in Table II, and the fitted curves can be seen in Fig.\ 4(b) in a plot of $C$ vs.\ $T$. For Samples 1 and 2 ($\delta $ = 0.11 and 0.15) although the plot of $C/T$ vs.\ $ T^{2}$ gives approximately straight lines, a careful fitting procedure confirms the existence of a magnetic $BT^{3/2}$ term. The uncertainty in the coefficients is estimated mostly by varying the fitted temperature range. All fitted curves fall within the experimental data with a maximum dispersion smaller than $\pm$0.7\% in more than 90\% of the points, and no systematic departures from the fitted curves are observed. In Sample 3 ($ \delta = 0.26$) we found no contribution arising from a linear term $\gamma T $. An upper estimate gives $\gamma < 0.8\;$ mJ/mol$\,$K$^{2}$, obtained using a maximum fitting temperature above 9$\;$K. Below this range, the inclusion of a linear term in the fitted expression yields negative values of $\gamma$. By ~allowing ~the ~magnetic contribution ~to vary as ~$BT^n$ \narrowtext \begin{figure} \begin{center} \includegraphics[width=8.0cm,angle=0,clip]{ghiv-fig5.ps} \end{center} \caption{High-temperature (30--200$\;$K) specific heat of LaMnO$_{3+\delta}$ , plotted as $C/T$ vs.\ $T$. Results for Samples 2 and 3 are shifted downward for clarity, as indicated by the arrows. The inset shows the temperature derivative $d(C/T)/dT$ vs.\ $T$ for Samples 1 and 2, with the FM transition temperatures indicated.} \label{fig:5} \end{figure} \noindent we find a best fit with $n$ very close to the assumed value of 3/2. One of the most important and unexpected results obtained from our low-temperature specific-heat data is the observation of a very high linear coefficient $ \gamma$ in Samples 1 and 2. In this case, by fitting the data only with linear and cubic terms we obtain even higher values of $\gamma$. Possible origins of this contribution will be discussed below. The Debye temperature $\theta_{D}$ significantly increases with the increase of vacancy content in LaMnO$_{3+\delta}$. The values of $\theta _{D}$, in the range 370-500$\;$K, are comparable to those previously reported in manganite perovskites.\cite{Ghiv,Ham,Coey,Wood,Tok} The tendency to an increase of $\theta_{D}$ with higher hole doping has been previously observed.\cite{Ghiv,Wood,Tok} It has been argued\cite{Tok} that the reduction of lattice stiffness at low doping values could be related to dynamic Jahn-Teller distortion in the compounds. The large value of $ \theta_{D}$ in Sample 3, with the highest content of Mn$^{4+}$, is close to that observed\cite{Ghiv} in the AF insulator La$_{0.38}$Ca$_{0.62}$MnO$_3$. This suggests that AF interactions, also present in Sample 3, may contribute to a hardening of the lattice vibrations. The magnitude of the $BT^{3/2}$ term is also of relevance, providing information on the spin-wave excitations in the compounds. The value of the spin-wave stiffness constant determined for Sample 1, $D = 75 \;$meV\AA$^{2}$ , is approximately half of that obtained for La$_{0.7}$Ca$_{0.3}$MnO$_3$ ($D = 170 \;$meV\AA$^{2}$)\cite{Lynn} and La$_{0.7}$Sr$_{0.3}$MnO$_3$ ($D = 154 \;$meV\AA$^{2}$),\cite{Smol} both in the FM metallic phase. The value of $D = 32\;$meV\AA$^{2}$ in Sample 2 is of the same order as in the FM insulator La$_{0.9}$Ca$_{0.1}$MnO$_3$ ($D = 40 \;$meV\AA$^{2}$),\cite{Ghiv} whose insulating character is also interpreted as a disorder effect. This is consistent with the fact that increasing disorder should give rise to lower values of $D$, i.e.,``softer'' spin waves, as it is expected to reduce the strength of the ferromagnetic coupling. The observation of a magnetic $ BT^{3/2}$ contribution in Sample 3, for which a spin-glass phase is observed, will be addressed in the next section. For completeness, Fig.\ 5 displays the high-temperature (30--200$\;$K) specific heat of the investigated LaMnO$_{3+\delta }$ samples, plotted as $ C/T$ vs.\ $T$. Results for Samples 2 and 3 are shifted downward for clarity. Sample 1 shows a small anomaly associated with the FM transition at 152$\;$ K, coinciding with the transition temperature obtained from the AC susceptibility. No anomaly is observed in the results for Samples 2 and 3. The inset shows the temperature derivative, $d(C/T)/dT$, for Samples 1 and 2. The FM transition in Sample 2 is visible in the derivative plot at 143$\;$ K, again coinciding with the susceptibility measurements. Phase transitions with a large temperature width often show no specific-heat anomaly, as reported\cite{Ghiv} in the FM insulator La$_{0.90}$Ca$_{0.10}$MnO$_3$. For Sample 3, a specific-heat anomaly is not observed even in the derivative plot (not shown), as expected for a spin glass. For Sample 1, the entropy associated with the FM transition, which can be obtained from $\Delta S=\int (C/T)dT$, is $\Delta S = 0.21 \pm 0.02\;$J/mol$\,$K. The subtracted lattice contribution is estimated by excluding the peak region from the data, and fitting the remaining data with a sum of three Einstein optical modes. The value of $\Delta S$ is about an order of magnitude smaller than reported on Ca-doped samples,\cite{Ghiv,Tanaka} and on other manganite compounds,\cite {Nd,PrCa} where in turn the $\Delta S$ values are also smaller than expected from the ordering of the spin system. A thorough discussion related to this ``missing'' entropy can be found elsewhere.\cite{PrCa} \section{Discussion} From our susceptibility and magnetization data we have established that the FM phase of hole doped LaMnO$_{3+\delta }$ samples evolves to a cluster-glass-like state. Several other manganite compounds present similar behavior when substitution occurs in the manganese site. If one takes, for instance, the standard CMR compound La$_{0.7}$Ca$_{0.3}$MnO$_{3}$, substitution of Mn by Co\ (Ref. \onlinecite{Gay}) or In (Ref.\ \onlinecite{Sanchez}) also gives rise to an insulating cluster-glass phase. This suggests that the DE interaction, mostly responsible for the metallic FM state of doped manganites, is inhibited by random disorder in the system. The formation of FM clusters is accompanied by strong charge-localizing effects which yield an insulating state. Nevertheless, the size of the clusters must be large enough for the $e_{g}$ electrons to extend over several sites, and provide the observed FM interaction. The most striking feature of the specific-heat data for Samples 1 and 2 is the appearance of an unexpectedly large linear term, in excess of 19$\;$ mJ/mol$\,$K$^{2}$, although the system as whole is an insulator with respect to transport properties. It is most important to understand the origin of this anomalous contribution. Compared with the increasing number of publications on doped manganite perovskite samples, relatively few reports on heat capacity have been presented. Low-temperature data for LaMnO$_{3}$ doped with Ca,\cite{Ghiv,Ham,Coey} Sr,\cite{Coey,Wood} and Ba,\cite{Ham,Coey} all in the metallic FM phase, observed a specific-heat linear term $\gamma $ in the range of 5--7$\;$mJ/mol$\,$K$^{2}$, associated with conduction electrons. However, few previous investigations have reported high $\gamma $ values in insulating manganite samples: in the electron doped system La$ _{2.3}$Ca$_{0.7}$Mn$_{2}$O$_{7}$,\cite{LaY} the authors found $\gamma =41\;$ mJ/mol$\,$K$^{2}$, and in Nd$_{0.67}$Sr$_{0.33}$MnO$_{3}$\cite{Nd} a value of $\gamma =25\;$mJ/mol$\,$K$^{2}$ was observed. A detailed explanation for this contribution has not been put forward. As already mentioned, our magnetic results clearly allow us to infer that ferromagnetic order in LaMnO$ _{3+\delta }$ develops in regions of limited size (clusters), whose magnetic moments undergo a spin-glass-like transition. We will argue now that our heat capacity results are consistent with this picture. The stoichiometric compound LaMnO$_{3}$ ($\delta =0$) has an orthorhombic crystal structure,\cite{ortho71} which is a distorted form of the cubic perovskite structure. The ideal cubic system would have a FM metallic character, with the Fermi energy lying in the middle of the $e_{g}$ band. \cite{Sat} The splitting of the $e_{g}$ bands, due to the Jahn-Teller distortion, leads to a small gap (1.5$\;$eV) between the Mn $e_{g}^{1}$ and $ e_{g}^{2}$ bands. This stabilizes the A-type AF order and makes the system a Mott insulator. As we dope with holes, Mn$^{4+}$ are created, the perovskite distortion decreases, and the Fermi level drops down into the lowest half of the split band. Thus the system becomes metallic, with the DE mechanism being responsible for charge transfer among Mn ions, and the consequent polarization of the $t_{2g}$ spins that yield ferromagnetic order. However, in non-stoichiometric LaMnO$_{3+\delta}$, the disorder introduced by random La and Mn vacancies may cause Anderson-like localization of the electron states close to the band edges. In contrast to what happens in the cation-substituted compounds, the disordering effect of Mn vacancies is strong enough for localization to be effective even at high concentrations of Mn$^{4+}$. Previous theoretical investigations\cite{Allub1,Allub2,Varma} confirmed that disorder leads to charge localization in doped manganites. As is well known since Anderson's original paper,\cite{Ander} a distribution of site-dependent diagonal energies produces localization of the electronic states from the edges of the bands to an energy within them which is called the {\it mobility edge\/}. Allub and Alascio\cite{Allub1,Allub2} have shown that, for La$_{1-x}$Sr$_x$MnO$_3$, according to the amount of disorder and concentration of carriers, the Fermi level can cross the mobility edge to produce a metal-insulator transition. Disorder is quantified by a distribution width $\Gamma$. If $\Gamma$ is large, charge localization is enhanced, and the system remains insulating, as observed in our LaMnO$ _{3+\delta}$ samples. It is worth mentioning that electron localization also occurs in metallic manganite compounds. It has been argued\cite{Coey} that the $e_{g}$ electrons may be localized in large wave packets due to potential fluctuations arising from cation substitution, and additionally by spin-dependent fluctuations due to local deviations from FM order. In LaMnO$ _{3+\delta}$ the missing Mn ions enhance these random fluctuations, favoring charge localization.\cite{Ranno} On the other hand, it is reasonable to assume that at low vacancy concentration the Fermi level does not fall too far above the mobility edge, which implies that the localization length may be fairly large. Charge carriers can thus hop between a number of Mn ions, which actually defines the FM clusters. The electron mobility inside the clusters ensures the effectiveness of the DE interaction, giving rise to the observed FM behavior. Furthermore, if the FM regions are not too small, regular spin waves can be excited inside them, yielding the observed $T^{3/2}$ contribution to the specific heat. This is consistent with the values obtained for the spin wave stiffness in these compounds. The electron levels, although localized, are not largely spaced in energy, allowing for thermal excitations that contribute with a linear term to the specific heat as a function of temperature. It remains to be understood why the coefficient of the specific-heat linear term is so large in our results, as compared to other perovskite systems. A number of mass enhancement mechanisms may be envisaged, like magnetic polarons, lattice polarons related to the dynamical Jan-Teller effect, or Coulomb interaction effects. However, it is not straightforward to understand why these effects would not be equally noticeable in most doped manganite compounds. We suggest that the explanation lies in the fact that localization has changed the Fermi level to a region of high density of states. For instance, it is possible that the disorder yields an enhancement of the two-dimensional character of the bands, giving rise to a high density of states. Indeed, band structure calculations\cite{Sat,Sing} on stoichiometric LaMnO$_3$ revealed sharp features resembling the typical logarithmic van Hove singularities of two-dimensional tight-binding bands. Sample 3, with the largest vacancy content, shows qualitatively distinct characteristics. Nevertheless, its behavior can be interpreted with the same arguments discussed above. The localization length is now very small due to the high degree of disorder. Thus, FM clusters are no longer formed, which is consistent with the observed magnetic response of the compound. The absence of a linear term in the specific heat reflects the higher degree of localization of the charge carriers, which effectively prevents the DE mechanism. The system behavior closely resembles that of a regular spin glass, with short range FM interaction competing with AF coupling, the latter arising from the high Mn$^{4+}$ content. It is somewhat puzzling, though, that the dominant contribution to the specific heat is a term proportional to $T^{3/2}$, which is usually attributed to spin waves in a long range FM system. However, according to computer simulations by Walker and Walstedt\cite{ww} on a model spin glass, the low-energy excitations are collective modes, even though the local magnetic moments do not show long range order. Thus, some power-law behavior of the specific heat with temperature can be expected. Linear and quadratic terms are obtained in Ref.\ \onlinecite{ww} for a model metallic spin glass with RKKY interactions, but the actual value of the exponent depends on details of the distribution of low-lying excitations, and a value of 3/2 cannot be ruled out. \section{Conclusions} In this work we have presented measurements of AC susceptibility, DC magnetization, and specific heat in a series of LaMnO$_{3+\delta}$ samples with large $\delta$ values, and therefore high degree of disorder. The aim is to provide a better understanding of the role of La and Mn vacancies in the properties of mixed-valence manganites. From our analysis we may draw two main conclusions: (i) magnetic measurements showed that the previously known FM insulating phase of these compounds displays a disorder-induced cluster-glass-like behavior; (ii) the anomalous high specific-heat linear coefficient $\gamma$ in this case gives evidence of a high density of localized states around the Fermi level, even though the latter falls in a region of Anderson-localized states. Hence, charge localization is enhanced due to disorder in the system, and the low temperature FM insulating state consists of randomly oriented FM clusters, which align in small applied fields. The carriers, though localized, may hop between several Mn sites to ensure the DE interaction responsible for the FM order. In the sample with the highest vacancy content the increased random disorder, and the competition between FM and AF interactions give rise to a spin-glass state. \section{Acknowledgments} We thank Mucio Continentino and Gerardo Mart\'{\i }nez for helpful discussions. This research was financed by the Brazilian Ministry of Science and Technology under the contract PRONEX/FINEP/CNPq no 41.96.0907.00. Additional support was also given by FUJB and FAPERJ. J.A.A. thanks the Spanish CICyT for funds to the project PB97-1181. L.F.C. was supported by the EPSRC grant number GR/K 73862 and by the Royal Society, U.K.
\section{Introduction} Since the papers by Tsallis \cite{first,second}, non-extensive statistical formalism has been shown to be not only {\it robust} --it allows generalizations of all necessary fundamental concepts of thermostatistics \cite{concepts}--, but also {\it useful} --it provides a suitable theoretical tool to explain some of the experimental situations where standard thermostatistics has shortcomings, due to the presence of long-range interactions, or long-range memory effects, or (multi)-fractal space-time constraints--. See Ref. \cite{biblio} for a periodically updated bibliography. \\ The core of this generalized formalism is defined through a generalized entropy \begin{equation} S_q = k \frac{1- \sum_{i=1}^{W} p_i^q}{q-1}, \;\;\;\;\;\; (q \in \Re), \end{equation} where $k$ is a positive constant, $\{p_i\}$ is a set of probabilities and $W$ is the total number of microscopic configurations. It is easy to verify that the $q\rightarrow 1$ limit immediately recovers the usual (extensive) Boltzmann-Gibbs entropy. Moreover, if a composed system $A+B$ has probabilities which factorize into those corresponding to the subsystems $A$ and $B$, then $S_q(A+B)/k = S_q(A)/k + S_q(B)/k + (1-q) S_q(A)S_q(B)/k^2$. This property clearly exhibits the fact that the parameter $q$ characterizes the degree of non-extensivity of any physical system.\\ The generalization of quantum statistics for non-extensive systems was only accomplished, up to recent days, in an {\it approximate fashion}, by using two different schemes. One of them is the Asymptotic Approach (AA), of Tsallis et al. \cite{bbAAtsallis}, the other one is the Factorization Approach (FA), of B\"{u}y\"{u}kk{\i}l{\i}\c{c} et al. \cite{FA}. The physical applications studied so far within these two approximations include the blackbody radiation \cite{bbAAtsallis,bbAAroditi,bbFAugur}, the Stefan-Boltzmann constant \cite{SB-AAplastino,SB-AAwang,SB-FAugur}, and some aspects of the early universe physics \cite{early-AA,early-FA}. Moreover, the AA has also been used in some other works such as the Bose-Einstein condensation \cite{curilef}, the specific heat of $^4$He \cite{curilef-helyum}, thermalization of an electron-phonon system \cite{koponen} and cosmology \cite{torres1,torres2}. Although some detailed analysis on these approximate schemes \cite{wang-FA} suggest that both schemes could be helpful in physical applications --at least, for $(1-q)$-order corrections--, this was still doubtful. A complete verification needed a comparison between the results of these approximate schemes and the exact ones. But an exact treatment of non-extensive quantum distributions was not available up to the recent papers of Rajagopal et al. \cite{raj,ervin}. Just after this work, Lenzi and Mendes have also given an exact treatment of blackbody radiation \cite{bb-exact}. All these recent efforts enable us to make a comparison between the approximate and exact schemes, which will ultimately show whether the AA and the FA are useful or not. This will be the main purpose of this paper.\\ In Section II, we review the approximate and exact results and develop an analytical method to derive the explicit form of any measurable quantity within the AA. We compare the approximate and exact results in Section III using (i) the predictions of one of the experimental tests suggested in \cite{raj} and (ii) the blackbody radiation. Finally, we give our final comments in Section IV. \section{Non-extensive quantum statistics} \subsection{Asymptotic approach} Within the AA, namely in the $\beta(1-q)\rightarrow 1$ limit, the generalized partition function is given by \cite{bbAAtsallis} \begin{equation} Z_q\simeq Z_{1} \left\{1-\frac 12 (1-q) \beta^2 \left<\hat {\cal H}^2\right>_{1}\right\}, \end{equation} from where the generalized distribution function of non-interacting bosons can be found, up to $(1-q)$-order, as \begin{equation} \label{nAA} \left<n\right>_q =\left<n\right>_1 + (1-q) \left<n\right>_1 \left\{\ln(1/Z_1) + (x-\psi) \left[\frac{\left<n^2\right>_1} {\left<n\right>_1}+(x-\psi)\left(\left<n^2\right>_1 - \frac{\left<n^3\right>_1}{\left<n\right>_1}\right)\right]\right\}, \end{equation} where $x\equiv\beta\epsilon$ , $\psi\equiv\beta\mu$, and \begin{equation} \left<n\right>_1 = \frac{1}{e^{x-\psi}-1},\hspace{0.4cm} \left<n^2\right>_1 = \frac{e^{-(x-\psi)}+e^{-2(x-\psi)}} {\left[1-e^{-(x-\psi)}\right]^2},\hspace{0.4cm} \left<n^3\right>_1 = \frac{e^{-(x-\psi)}+ 4e^{-2(x-\psi)}+e^{-3(x-\psi)}} {\left[1-e^{-(x-\psi)}\right]^3}. \end{equation} The standard ($q=1$) partition function is given by \begin{equation} Z_1 = \frac{1}{1-e^{-(x-\psi)}}. \end{equation} This approximation has found a wide range of applications up to now, however, no attempt has been made for deriving some of the thermodynamical quantities within this approach, directly using Eq. (\ref{nAA}). \\ One aim of this paper is to provide a technique for computing, in a closed form, the kind of integrals needed to find the average number of particles within the AA. To do this, let us start by writing down the definition of the average number of particles: \begin{equation} \label{bosN} \left<N\right>_q = \frac{2\pi V(2mk)^{3/2} T^{3/2}}{h^3} \int_0^{\infty} \epsilon^{1/2} \left<n\right>_q d\epsilon , \end{equation} where all variables have the usual meaning. Using Eq. (\ref{nAA}) and the definitions of $x$ and $\psi$, this expression turns out to be \begin{equation} \left<N\right>_q = \frac{2\pi V(2mk)^{3/2} T^{3/2}}{h^3} \left[I_{st} + (1-q)(I_2 + I_3) \right], \end{equation} where \begin{equation} I_{st} = \int_0^{\infty}\frac{x^{1/2} dx}{e^{x-\psi}-1},\hspace{0.4cm} I_{2} = \int_0^{\infty}\frac{(x-\psi) x^{1/2} dx} {e^{x-\psi}-1}, \end{equation} and \begin{equation} I_{3} = \int_0^{\infty}\frac{(x-\psi) x^{1/2} dx} {e^{x-\psi}-1} \left[\frac{\left<n^2\right>_1} {\left<n\right>_1}+(x-\psi)\left(\left<n^2\right>_1 - \frac{\left<n^3\right>_1}{\left<n\right>_1}\right)\right]. \end{equation} $I_2$ and $I_3$ are the $(q-1)$ order correction to the standard ($q=1$) result and here $I_{st}$ stands for the standard integral appearing in the solution of the extensive case \cite{pathria}. $I_{st}$ and $I_2$ have standard forms, and could easily be solved as: \begin{equation} I_{st} = \Gamma(3/2) g_{3/2}(z), \end{equation} \begin{equation} \label{2RESUL} I_2 = \int_0^{\infty} \frac{x^{5/2-1} dx}{e^{x-\psi}-1} - \psi \int_0^{\infty} \frac{x^{3/2-1} dx}{e^{x-\psi}-1} = \Gamma(5/2) g_{3/2}(z) - \psi \Gamma(3/2) g_{3/2}(z) , \end{equation} where $z$ is the fugacity and is defined as $z\equiv e^{\beta\mu}$. On the other hand, $I_3$ is more involved, and it takes the form: \begin{equation} \label{3RESUL} I_3 = a+b-3c-d, \end{equation} where \begin{equation} a= \int_0^{\infty} \frac{(x-\psi)x^{1/2} dx} {\left[e^{x-\psi}-1\right]^2}, \hspace{0.4cm} b= \int_0^{\infty} \frac{(x-\psi)x^{1/2} e^{x-\psi} dx} {\left[e^{x-\psi}-1\right]^2}, \end{equation} \begin{equation} c=\int_0^{\infty} \frac{(x-\psi)^2 x^{1/2} e^{x-\psi} dx} {\left[e^{x-\psi}-1\right]^3}, \hspace{0.4cm} d=\int_0^{\infty} \frac{(x-\psi)^2 x^{1/2} e^{2(x-\psi)} dx} {\left[e^{x-\psi}-1\right]^3}. \end{equation} In an Appendix, we provide an analytical technique (maybe there are others) to compute each one of these integrals. Using this technique, we obtain the average number of particles as: \begin{eqnarray} \left<N_e\right>_q \frac {h^3}{2\pi V(2mkT)^{3/2}} = \Gamma(3/2) g_{3/2}(z) + (q-1) \sqrt{\pi} \times \hspace{7cm} \mbox{} \nonumber\\ \hspace{3cm} \times \left[ \frac 32 g_{3/2}(z) -\frac 98 g_{5/2}(z) + \frac {7}{4} \psi g_{3/2}(z) - 2 \psi g_{1/2}(z) - \frac 12 \psi^2 g_{1/2}(z) + \frac 98 \psi^2 g_{-1/2}(z) \right]. \end{eqnarray} Here, $\left<N_e\right>_q$ stands for the number of particles in the excited states ($\epsilon\neq 0$). As in the standard case, we have separated the contribution of the state given by $\epsilon =0$, which has zero weight in the integrals. For this level of energy, we found, \begin{equation} \left<N\right>_q (\epsilon = 0) = \frac{z}{1-z} \left\{ 1 + (q-1) \left[ \ln z + \frac{z \ln z}{1-z} - \frac{ \ln z}{1-z} + 3 \frac{z (\ln z)^2}{(1-z)^2} + \frac{(\ln z)^2}{(1-z)^2}\right] \right\} . \end{equation} When $z\ll 1$, all the correction terms go to zero. When $z\rightarrow 1$, some of the terms are divergent but the usual shape is unchanged. This can be seen in Fig. 1.\\ Numerical analysis, which we show in Fig. 2, illustrates that the maximum correction is attained for $z=1$. Then, the number of particles in all excited states is bounded by, \begin{equation} \label{neAA} \left<N_e\right>_q \le \frac {2\pi V(2mk)^{3/2} T^{3/2}} {h^3} \left[2.315 + (q-1) 4.27 \right] . \end{equation} It is worth noticing that the AA is such that not all terms in the $(1-q)$ correction are positive (or negative, depending on the choice of $q$) definite. Moreover, their maximum values are not always attained at middle points of the interval of interest, and although they do have bounded expressions, the maximum correction is obtained only for $z=1$. This differs from what happened in the FA, where each term had a maximum value within the interval of interest \cite{tor-tir}. The order of magnitude of the maximum correction is, however, the same in both approximations.\\ Any interested reader could easily apply the same technique, which we introduced in the Appendix, to compute any other thermodynamical quantity, whenever it is needed. \subsection{Factorization approach} Within the FA \cite{FA}, the generalized distribution function of bosons is given, at $(1-q)$ order, by \cite{tor-tir} \begin{equation} \label{nFA} \left<n\right>_q = \left<n\right>_1 + (q-1) \frac{(x-\psi)^2 e^{x-\psi}}{2 \left(e^{x-\psi}-1\right)^2}, \end{equation} where $\left<n\right>_1$, $x$ and $\psi$ have the same definitions as before. At this point, the remarkably simpler form of this result, when compared to the result of the AA [Eq. (\ref{nAA})], is worth emphasizing.\\ In this context, we have found general expressions for some thermodynamical quantities of bosons and fermions \cite{tor-tir}; here we quote only the average number of particles for bosons, since it will be adequate for our proposed comparison\footnote{We take advantage here to signal out a mistake in the last equation of Ref. \cite{tor-tir}, where the correction appears to be proportional to 0.886 $(q-1)$ and should have a minus sign in front of it \cite{WANG-PRIVA}.}: \begin{equation} \label{neFA} \left<N_e\right>_q \le \frac {2\pi V(2mk)^{3/2} T^{3/2}} {h^3} \left[2.315 + (q-1) 3.079 \right] . \end{equation} \subsection{The exact result} Although the results of the AA and the FA have been successfully used in a wide range of physical applications, an exact treatment of non-extensive quantum statistics was lacking until the recent work of Rajagopal, Mendes and Lenzi \cite{raj,ervin}. In their analysis, they have given the many-particle $q$-Green function in terms of a parametric contour integral over a kernel, multiplied by the usual grand canonical one particle Green function which now depends on $q$. They managed to obtain exact expressions for thermodynamical quantities, such as $\left<N\right>_q$.\\ To proceed further, let us quote here some of the results of \cite{raj,ervin}. Rajagopal et al. have used the general contour integral of the form, \begin{equation} b^{1-z} \frac{i}{2\pi} \int_C du \exp(-bu) (-u)^{-z} = \frac{1}{\Gamma(z)} , \end{equation} with $b>0$ and Re $z>0$, and where the contour $C$ starts from $+\infty$ on the real axis, encircles the origin once counterclockwise and returns to $+\infty$. Using the $q$-Green functions, and after some cumbersome algebra, they finally obtain (for bosons) \begin{equation} \left<N\right>_q = V \int_C du K_q^{(2)}(u) \int_{-\infty}^{\infty} \frac{d\omega}{2\pi} \int \frac{d^D p}{(2\pi)^D} \frac{Z_1(-\beta(1-q)u,\mu)} {\left[e^{-\beta(1-q)u(\omega-\mu)}-1\right]} A({\vec p};\omega), \end{equation} where $D$ is the dimension of space, $ A({\vec p};\omega)$ is the spectral weight function and \begin{equation} K_q^{(2)}(u) = i \frac{\Gamma(1/(1-q))}{2\pi (Z_q)^q} \exp(-u)(-u)^{-1/(1-q)}, \end{equation} and \begin{equation} Z_q(\beta,\mu) = \int_C du K_q^{(1)}(u) Z_1(-\beta u(1-q),\mu). \end{equation} This exact expression for the average number of particles finally gives us the opportunity to make a comparison between the exact and the approximate results. \section{Exact and approximate results} \subsection{Bose-Einstein condensate} One of possible experimental tests of the validity of the $q$-framework is based on a recent work on Bose-Einstein condensation of a small number of atoms (of the order of 100 to 170), confined to a small region of space by magnetic trapping \cite{BE}. By taking free particle spectral weight function, namely $A({\vec p};\omega)=2\pi\delta(\omega-{\vec p}^2/2m)$, near the Bose-Einstein condensation, they have found \begin{eqnarray} \label{NqN1exact} \frac{\left<N\right>_q}{\left<N\right>_1} \simeq \left[\frac{(T_c)_q}{(T_c)_1}\right]^{3/2} \frac{\Gamma\left(\frac{2-q}{1-q}\right)} {(1-q)^{1/2} \Gamma\left(\frac{2-q}{1-q} + \frac{1}{2}\right)} \left\{1 + \frac{\left<N\right>_1}{(1-q)^{3/2}} \frac{\zeta(5/2)}{\zeta(3/2)} \left[\frac{(T_c)_q}{(T_c)_1}\right]^{3/2} \right. \times \nonumber\\ \times \left. \left[\frac{\Gamma\left(\frac{2-q}{1-q} + \frac{1}{2}\right)} {\Gamma\left(\frac{2-q}{1-q} + 2 \right)} - q \frac{\Gamma\left(\frac{2-q}{1-q} \right)} {\Gamma\left(\frac{2-q}{1-q} + \frac{3}{2}\right)} \right] \right\}. \end{eqnarray} Here, the $\simeq$ sign reflects that this equation is valid near the Bose-Einstein condensation, which does not change the fact that it is an exact result. We may now expand this expression in powers of $(q-1)$. Up to first order, \begin{equation} \label{NqN1exactapprox} \frac{\left<N\right>_q}{\left<N\right>_1} \simeq \left[\frac{(T_c)_q}{(T_c)_1}\right]^{3/2} \left\{ 1 + (q-1) \left( 0.456 - 0.023 \left<N\right>_1 \left[\frac{(T_c)_q}{(T_c)_1}\right]^{3/2} \right) \right\}. \end{equation} The two previous equations deviate from each other very soon when $(q-1)^2$ is not negligible.\\ Let us now derive similar expressions for the AA and FA in order to compare them with Eqs. (\ref{NqN1exact}) and (\ref{NqN1exactapprox}). The condition for the appearance of Bose-Einstein condensation can be expressed as \begin{equation} \left<N\right>_q > \left<N_e\right>_q . \end{equation} Alternatively, with constant $\left<N\right>_q$ and $V$, using Eq. (\ref{neAA}) for the AA and Eq. (\ref{neFA}) for the FA, this condition can be recast in the form \begin{equation} T < \left(T_c\right)_q = \frac{h^2}{(2\pi)^{3/2}2mk} \left\{\frac{\left<N\right>_q}{V\left[2.315+(q-1)\kappa\right]} \right\}^{2/3}, \end{equation} where $\kappa=4.27$ for the AA and $\kappa=3.078$ for the FA. Then, we can organize these expressions to give \begin{equation} \label{NqN1appr} \frac{\left<N\right>_q}{\left<N\right>_1} = \left[\frac{\left(T_c\right)_q}{\left(T_c\right)_1}\right]^{3/2} \frac{[2.315+(q-1)\kappa]}{2.315} . \end{equation} Eq.(17) and (19) have corrections which are trivial (not depending on $z$) just because we have approximated them: the actual complete results are Eqs.(15) and (16) for the AA, while those for the FA can be found in our previous paper [24]. We managed the dependence on $z$ in order to obtain an upper bound for the corrections and simplify the analysis that follows. Differences between Eqs. (\ref{NqN1appr}) and (\ref{NqN1exactapprox}) are worth noticing: the later depends on $(T_c)_q$ and $\left<N\right>_1$ in a much stronger way. However, as we shall see, for $q$ close to 1 these differences are not important.\\ We would now like to choose physically suitable $q$ values. An early Universe test based on the FA has shown \cite{early-FA} to produce a bound $|q-1|\le 4.01 \times 10^{-3}$, thus we have $q=0.996$. The other $q$ value which we use comes from a very recent work on pion transverse-momentum correlations in Pb-Pb high-energy nuclear collisions \cite{pion}. In that work, a deviation of $|q-1|=0.015$ from the standard statistics is found to be sufficient for eliminating the puzzling discrepancy between theoretical calculations and experimental data \cite{pion}. Thus, we shall use $q=0.985$ (in fact, in \cite{pion}, $q=1.015$ has been used, but since the exact result is given for $q<1$ values, we must take $q=0.985$, which has the same $|q-1|$ deviation). In Fig. 3 we plot $\left<N\right>_q / \left<N\right>_1$ versus $\left(T_c\right)_q / \left(T_c\right)_1$ for two representative values of $\left<N\right>_1$, and the two quoted values of $q$. However, note again that in our approximated schemes, $\left<N\right>_q / \left<N\right>_1$ as a function of $\left(T_c\right)_q / \left(T_c\right)_1$ is in fact independent of the particular value of $ \left<N\right>_1$. From Fig. 3, the following conclusions can be drawn: At the order of such $q$ values, the AA and the FA are almost the same, and in $(1-q)$-order correction, any of them could be used with the same confidence (maybe the FA would be preferable due to its remarkably simpler form). Only in those situations of extremely high experimental precision one could distinguish between the exact and approximate results. \subsection{Blackbody radiation} Very recently, an exact analysis of the blackbody radiation within the $q$-framework has been given \cite{bb-exact}. This exact analysis gives the generalization of the Stefan-Boltzmann law as \begin{equation} \label{Uexact} U_q =\frac{3kT\xi_3}{Z_q^q} \sum_{m=0}^{\infty} \frac{\xi_3^m}{m!} \frac{\Gamma[(2-q)/(1-q)]}{\Gamma[(2-q)/(1-q)+3(m+1)]}, \end{equation} where \begin{equation} Z_q = \sum_{m=0}^{\infty} \frac{\xi_3^m}{m!} \frac{\Gamma[(2-q)/(1-q)]}{\Gamma[(2-q)/(1-q)+3m]} , \end{equation} and \begin{equation} \xi_3 = \frac{4 \Gamma(3) \zeta(4)}{[2\pi^{1/2}(1-q)]^3\Gamma(3/2)} \left(\frac{2\pi V^{1/3}kT}{hc}\right)^3 . \end{equation} Let us now recall the Stefan-Boltzmann law derived by using the AA \cite{SB-AAplastino,SB-AAwang} and the FA \cite{SB-FAugur}: \begin{equation} \label{Uappr} U_q = \frac{8 \pi k^4 T^4 V}{c^3 h^3} \left[6.4939 - (1-q)\; \theta\; \right] \end{equation} where $\theta = 40.018$ for the AA and $\theta = 62.215$ for the FA.\\ For the comparison of the exact and approximate Stefan-Boltzmann laws, we again try to choose a value of $q$ which is in accordance with the blackbody radiation. The possible $q$-correction could be at the order of $10^{-4}$ or $10^{-5}$. Thus, here we shall use again the largest deviation predicted for the $q$-correction \cite{SB-AAplastino}, namely $|q-1|\le 5.3 \times 10^{-4}$, which gives $q=0.99947\;$. In Fig. 4 we present the behaviour of the exact [Eq. (\ref{Uexact})] and the approximate results [Eq. (\ref{Uappr})] for $q=0.99947$. It is seen from the figure that for such order of $q$-correction the approximate results are very close to the standard ($q=1$) case without exhibiting any curvature, contrary to the exact result. \section{Final remarks} We have managed to develop an analytical technique to express thermodynamical quantities for the asymptotic approach of quantum distribution functions. We have shown that, for simple boson systems, and for all $q$-values admitted by the existing bounds, both approximate schemes (the AA and the FA) are in agreement with the exact result (see figures). The magnitude of the deviation is quantified in previous formulae and could be seen if there is enough experimental precision. Otherwise, the simpler form that the factorization approach exhibits makes a case for its use as a standard and safe procedure for $(1-q)$-order corrections. \section*{Acknowledgments} U.T. acknowledges the partial support of BAYG-C program of TUBITAK (Turkish Agency) and CNPq (Brazilian agency) as well as the support from Ege University Research Fund under the project number 98FEN025. D.F.T. acknowledges support from CONICET and Fundaci\'on Antorchas and wishes to thank A. Lavagno for sending him his valuable Ph. D. Thesis. We also thank A. K. Rajagopal, A. Wang, A. Erzan and an anonymous referee for useful comments.
\section{Introduction.} Being heavy the top quark undergoes fast weak decays. The relatively large width $\Gamma_t$ of the top quark is mainly saturated by the decay channel $t\rightarrow Wb$ and keeps the effective energy of top-antitop system in the complex plane far enough from the cut along the positive semiaxis. Thus it serves as a sufficient infrared cutoff for long distance effects avoiding the problem of strong coupling. This allows one to bypass possible nonperturbative regions and is the key observation for the theoretical study of the top-antitop pair production near the two-particle threshold \cite{FK}. Because the relevant scale $\sqrt{\Gamma_tm_t}$, with $m_t$ being the top quark mass, is much larger than $\Lambda_{\rm QCD}$, the QCD perturbation theory expansion is applicable for the theoretical description of physical phenomena near the top quark threshold if singular Coulomb effects are properly taken into account \cite{FK,ttgg,ttee}. This feature turns the processes involving the top quarks into a unique laboratory for perturbative investigation of threshold effects. Experimental study of the top-antitop pair threshold production is planned to be performed at the Next Linear Collider both in high energy $e^+e^-$ annihilation and $\gamma\gamma$ collision \cite{exp}. High quality experimental data that can be obtained in such experiments along with a very accurate theoretical description of them make the processes of top-antitop pair threshold production a promising place for investigating quark-gluon interactions. This investigation concerns both general features of interaction and precise quantitative properties such as the determination of numerical values of the strong coupling constant $\alpha_s$, the top quark mass, and the top quark width. Though the main features in both $e^+e^-$ and $\gamma\gamma$ processes of top quark pair threshold production are rather similar the strong interaction corrections and relativistic corrections are different for them. Therefore a simultaneous analysis of these two processes extends possibilities of studying fine details of the top quark threshold dynamics. Besides the total cross sections which are mainly saturated by the $S$ wave final state of the top quark-antiquark pair, there is a set of observables sensitive to the $P$ wave component. For example, the $S$ and $P$ partial waves of the final state top quark-antiquark pair produced in $\gamma\gamma$ collisions can be separated by choosing the same or opposite helicities of the colliding photons \cite{ttgg}. This gives an opportunity of direct measurement of the $P$ wave amplitude which is strongly suppressed in the threshold region in comparison with the $S$ wave one. On the other hand, the forward-backward asymmetry of the quark-antiquark pair production in $e^+e^-$ annihilation \cite{Sum,Har} and top quark polarization \cite{Har,FKK} are determined by the $S-P$ wave interference in both processes. This provides us with two additional independent probes of the top quark interactions. The finite order perturbation theory of QCD breaks down in the threshold region of particle production due to the presence of singular $(\alpha_s/\beta)^n$ Coulomb terms. Here $\beta$ is a velocity of the heavy quark. However the resummation of these Coulomb contributions which are most important quantitatively in the threshold region is possible and can be systematically done in the framework of nonrelativistic QCD (NRQCD) \cite{CasLep} (for the recent development of the NRQCD effective theory approach see \cite{Man,LukMan,GriRot,LukSav,PinSot,BenSmi,Lab,Gr}). Note that the characteristic scale of the Coulomb effects for the top quark production $\alpha_s m_t$ is comparable numerically with the cutoff scale for infrared effects $\sqrt{\Gamma_t m_t}$ so the Coulomb effects are not suppressed by the top quark width. The determination of higher order corrections in the QCD coupling constant and of relativistic corrections in the case when Coulomb effects have to be taken into account beyond the finite order perturbation theory requires the perturbative expansion for the complete correlator to be performed near the Coulomb approximation rather than near Green functions of the free theory which is the standard pattern of perturbation theory for the infrared safe high energy processes. Recently has been made a rather essential progress in the theoretical description of heavy quark-antiquark threshold dynamics within NRQCD. The evaluation of next-to-leading order (NLO) and next-to-next-to-leading order (NNLO) corrections to the heavy quark threshold production in $e^+e^-$ annihilation has been done both within the analytical approach \cite{H1,KPP,PP,MY,PP1,BSS,BenSin,H2,KniPen} and numerically \cite{Hoang,Mel,KuhTeu,Nag,HoaTeu} while the NLO corrections to the heavy quark threshold production in $\gamma\gamma$ collision where computed analytically \cite{PP2}. The analysis of NNLO corrections in the last case is still absent. However, this analysis is necessary for the accurate quantitative study of the process since the NNLO contribution is found to be relatively large in the case of the top quark production in $e^+e^-$ annihilation \cite{Hoang,Mel} and one can expect that some large corrections emerge also in the case of the top quark threshold production in $\gamma\gamma$ collision. Moreover, a semianalytical analysis of the high order corrections to the top quark threshold production cross section in $e^+e^-$ annihilation has been performed so far \cite{Hoang,Mel,Nag,HoaTeu} while the essential part of corrections has been accounted for numerically \cite{ttee}. Therefore the complete analytical description of the process is also desirable\footnote{When this work was in its final stage a letter \cite{BSS} appeared where the photon mediated top quark production in $e^+e^-$ annihilation was analyzed analytically.}. Furthermore, the forward-backward asymmetry and top quark polarization has been analyzed in NLO only numerically \cite{Har} as well as the axial contribution in the $e^+e^-\rightarrow t\bar t$ process \cite{KuhTeu}. In this case the numerical study is more involved because of necessity to construct the Green function for the $P$ wave which leads to the more singular differential equations in comparison with the $S$ wave. The case of $P$ wave production in $\gamma\gamma$ collision \cite{ttgg,PP2} clearly demonstrates that the numerical analysis \cite{KuhTeu} with an explicit cutoff of the hard momentum contribution is not sufficient for an accurate account for the finite top quark width for these quantities because the relativistic effects are not taken into account properly. In the present paper we give a simultaneous analysis of several observables relevant to $e^+e^-\rightarrow t\bar t$ annihilation and $\gamma\g\rightarrow t\bar t$ collisions near the top quark production threshold in high orders of NRQCD. The total cross sections are computed in NNLO of NRQCD which includes $\alpha_s^2$, $\alpha_s\beta$ and $\beta^2$ corrections in the coupling constant $\alpha_s$ and the heavy quark velocity $\beta$ to the nonrelativistic Coulomb approximation. Explicit analytical expressions for the soft part of corrections are obtained. The threshold cross section of the $t \bar t$ production in $e^+e^-$ annihilation is obtained in the closed form including the contribution due to the top quark axial coupling. The hard part of the correction to the $\gamma\g\rightarrow t\bar t$ threshold cross section is found with the logarithmic accuracy. The forward-backward asymmetry of the top quark-antiquark pair production in $e^+e^-$ annihilation and top quark polarization in both processes of $e^+e^-$ annihilation and $\gamma\g$ collisions are computed up to NLO. The paper is organized as follows. In the next Section the nonrelativistic approximation for the basic observables of top quark-antiquark pair production near the threshold is formulated. In Section 3 the threshold effects are described by three universal functions related to the $S$, $P$ wave production and $S-P$ wave interference which have been computed analytically within NRQCD. In Section 4 we present our numerical analysis and the discussion of the obtained results. The last Section is devoted to our conclusions. Some explicit analytical formulae are given in Appendix. \section{The nonrelativistic approximation near the production threshold.} In this section we describe the set of observable which will be analyzed: the total cross sections, the forward-backward asymmetry, the polarization of top quark. We formulate the nonrelativistic approximation for these observables setting the stage for the complete NRQCD analysis. In the last subsection we dwell upon the peculiarities of introducing the finite width of the top quark. \subsection{The effective theory description of the heavy quark threshold dynamics.} Near the threshold the heavy quarks are nonrelativistic so that one may consider both the strong coupling constant and heavy quark velocity as small parameters. The threshold expansion of the QCD loop integrals has been developed in \cite{BenSmi}. However, to take into account the singular threshold effects properly one has to go beyond the finite order QCD perturbation theory. For this purpose the expansion in $\beta$ should be performed directly in the QCD Lagrangian within the effective field theory framework. The first step to construct the effective theory is to identify all the scales present in the problem. The threshold dynamics is characterized by four different scales \cite{BenSmi}:\\ (i) the hard scale (energy and momentum scale like $m_q$);\\ (ii) the soft scale (energy and momentum scale like $\beta m_q$);\\ (iii) the potential scale (the energy scales like $\beta^2 m_q$, while the momentum scales like $\beta m_q$); \\ (iv) the ultrasoft scale (both energy and momentum scale like $\beta^2 m_q$). The ultrasoft scale is only relevant for gluons.\\ By integrating out the hard scale of QCD one arrives at the effective theory of NRQCD \cite{CasLep}. Because the NRQCD Lagrangian does not contain explicitly the heavy quark velocity the power counting rules are necessary to construct the regular expansion in this parameter. The list of the power counting rules for dimensionally regularized NRQCD and their relation to the threshold expansion \cite{BenSmi} can be found in \cite{Gr} Integrating out the soft modes and the potential gluons of NRQCD one obtains the effective theory of potential NRQCD \cite{PinSot} which contains potential quarks and ultrasoft gluons as active particles and is relevant for the analysis of the threshold effects. In potential NRQCD the dynamics of the quarks is governed by the effective nonrelativistic Schr{\"o}dinger equation and by their interaction with the ultrasoft gluons. To obtain a regular perturbative expansion in $\beta$ this interaction should be expanded in multipoles. Note that in the process of scale separation some spurious infrared and ultraviolet divergences may appear at intermediate steps of calculation which cancel each other in the final results for physical observables. The dimensional regularization has been recognized as a powerful tool to deal with these divergences \cite{Man,LukSav,PinSot,BenSmi,Lab,Gr,c1,c2,PinSot1,CMY}. If the ultrasoft effects are neglected the propagation of a quark-antiquark pair in the color singlet state is described in the potential NRQCD by the Green function $G({\bf x},{\bf y},E)$ of the Schr{\"o}dinger equation \begin{equation} \left({\cal H}-E\right)G({\bf x},{\bf y},E)=\delta({\bf x}-{\bf y}) \label{Schr} \end{equation} where ${\cal H}$ is the effective nonrelativistic Hamiltonian. Near the threshold the singular $(\alpha_s/\beta)^n$ Coulomb terms should be summed up in all orders in $\alpha_s$. Thus, in threshold region one has to develop the expansion in $\beta$ and $\alpha_s$ around some solution which incorporates properly the threshold effects, for example, around the nonrelativistic Coulomb solution. In this case the leading order approximation for the nonrelativistic Green function is obtained with the Coulomb Hamiltonian \[ {\cal H}_C=-{{\bf \Delta}_{\bf x}\over m_t}+V_C(x) \] where ${\bf \Delta_x} = \partial_{\bf x}^2$ is the kinetic energy operator and $V_C(x)=-C_F\alpha_s/x$ is the Coulomb potential, $x=|{\bf x}|$. The harder scales contributions are represented by the higher-dimension operators in ${\cal H}$ and by the Wilson coefficients of the operators of the nonrelativistic Hamiltonian leading to an expansion in $\beta$ and $\alpha_s$. On the other hand the radiation/absorption of the ultrasoft gluons by the interacting quark-antiquark pair, the effect of retardation, does not contribute in NLO and NNLO (the leading ultrasoft effects in heavy quarkonium have been recently computed in ref.~\cite{KniPen}). Thus, the nonrelativistic Green function of eq.~(\ref{Schr}) is the basic object in NRQCD analysis of the threshold effects up to NNLO. In Sections~2.2-2.4 we relate the observables of $e^+e^-\rightarrow t\bar t$ annihilation and $\gamma\g\rightarrow t\bar t$ collisions in the threshold region to this Green function. \subsection{Cross sections.} We study the normalized cross sections of the top quark-antiquark pair production in $e^+e^-$ annihilation \[ R^e(s)={\sigma(e^+e^-\rightarrow t\bar t)\over \sigma(e^+e^-\rightarrow\mu^+\mu^-)}, \] and in $\gamma\g$ collisions \[ R^\gamma(s)={\sigma(\gamma\gamma\rightarrow t\bar t)\over \sigma(e^+e^-\rightarrow\mu^+\mu^-)} \] where the lepton cross section \[ \sigma(e^+e^-\rightarrow\mu^+\mu^-)={4\pi\alpha_{QED}^2\over 3s} \] is the standard normalization factor with $\alpha_{QED}$ being the fine structure constant. Here $\sqrt{s}$ is the total energy of the colliding particles (electrons or photons) in the center of mass frame. For unpolarized initial states the following decomposition of the total cross sections is useful \begin{equation} R^{e}(s)={D_{V}\over q_t^2}R^{v}(s)+D_{A}R^{a}(s), \label{Ree} \end{equation} \begin{equation} R^{\gamma}(s)={R^{++}(s)+R^{+-}(s)\over 2} \label{Rgg} \end{equation} where $R^v$ ($R^a$) corresponds to the top quark vector (axial) coupling in $e^+e^-$ annihilation while $R^{++}$ ($R^{+-}$) corresponds to the colliding photons of the same (opposite) helicity in $\gamma\g$ collisions. $D_{V,A}$ are the standard combinations of electroweak coupling constants (see below), $q_t$ is the top quark electric charge. The cross section for the polarized electron/positron initial states is discussed, e.g. in ref.~\cite{Kuh}. Near the threshold the cross sections are determined by the imaginary part of the correlators of the nonrelativistic vector/axial quark currents which can be related to the nonrelativistic Green function and its derivatives at the origin. In NNLO the (potential) NRQCD provides one with the following representation of the cross sections \begin{equation} R^{v}(s) ={6\pi q_t^2N_c\over m_t^2}\left( C^{v}(\alpha_s)+B^{v}{k^2\over m_t^2}\right){\rm Im}G(0,0,k), \label{Rv} \end{equation} \begin{equation} R^{a}(s)={4\pi N_c\over m^4}C^a(\alpha_s) \partial^2_{\bf xy} {\rm Im}G({\bf x},{\bf y},k) \label{Ra} \end{equation} \begin{equation} R^{++}(s) ={24\pi q_t^4N_c \over m_t^2} \left(\left(C^{++}(\alpha_s)+B^{++}{k^2\over m_t^2}\right) {\rm Im}G(0,0,k)+\partial^2_{\bf x\bf y} {\rm Im}G({\bf x},{\bf y},k) \right) \label{Rpp} \end{equation} \begin{equation} R^{+-}(s)={32\pi q_t^4N_c\over m^4}C^{+-}(\alpha_s) \partial^2_{\bf x\bf y} {\rm Im}G({\bf x},{\bf y},k) \label{Rpm} \end{equation} where $k^2=-m_tE$, $E=\sqrt{s}-2m_t$ is the energy of a quark pair counted from the threshold $2m_t$. A symbolic notation $\partial^2_{\bf x\bf y}$ is used for the operator \[ \partial^2_{\bf x\bf y}f({\bf x},{\bf y})\equiv\sum_{i=1}^3\left. \partial_{x_i}\left(\partial_{y_i} f({\bf x},{\bf y})|_{y=0}\right)\right|_{x=0} \] that singles out the $P$ partial wave of the Green function. The standard electroweak factors read \[ D_V=q_e^2q_t^2+2q_eq_tv_ev_td+(v_e^2+a_e^2)v_t^2d^2, \] \[ D_A=(v_e^2+a_e^2)a_t^2d^2, \] \[ \begin{array}{lll} q_e=-1, & v_e=-1+4\sin^2\theta_W ,& a_e=-1,\\ q_t={2\over 3}, & v_t=1-8\sin^2\theta_W ,& a_t=1, \end{array} \] \[ d={1\over 16\sin^2\theta_W\cos^2\theta_W}{s\over s-M_Z^2}. \] The coefficients $C^i(\alpha_s)$ and $B^i$ are the parameters of NRQCD which are responsible for matching the effective and full theory cross sections in the limit of weak coupling in NNLO. The coefficients \[ C^{i}(\alpha_s)=1+c^{i}_1C_F{\alpha_s\over \pi}+c^{i}_2C_F \left({\alpha_s\over \pi}\right)^2+\ldots \] account for the hard QCD corrections and are determined by the corresponding amplitudes with on-shell heavy quarks at rest. The numerical values of these hard coefficients in the NLO approximation have been known since long ago \cite{Kar,RRY,HarBr,BarKuh}. They are explicitly \[ c^v_1=-4,\qquad c^a_1=-2, \] \[ c^{++}_1={\displaystyle {\pi^2\over 4}-5},\qquad c^{+-}_1=-4\, . \] The coefficient $C^v$ has recently been computed in NNLO in different schemes \cite{BSS,Hoang,Mel}. Starting from NNLO the hard coefficients acquire the anomalous dimensions and the calculation of the NNLO correction requires an accurate separation of hard and soft contributions. At the same time these coefficients do not depend on the normalization point of the strong coupling constant in NNLO and one can use the different normalization points of $\alpha_s$ entering in the coefficients $C^i$ (the ``hard'' scale $\mu_h$) and the nonrelativistic Green function (the ``soft'' scale $\mu_s$), see Section~3.1. The coefficients $B^i$ in eqs.~(\ref{Rv}) and (\ref{Rpp}) describe the pure relativistic corrections to the cross section which appear when the cross section is evaluated in terms of the correlator of nonrelativistic quark currents. Because the corresponding correction first appears in the order $O(\beta^2)$ the coefficients $B^i$ can be taken in the leading order in $\alpha_s$. The coefficient $B^v$ is related to the nonrelativistic expansion of the vector current and is known \cite{CasLep} to be equal to \[ B^v={4\over 3}. \] The calculation of the coefficient $B^{++}$ necessary for the consistent description of the $\gamma\g$ cross section within NRQCD in NNLO is more involved because the amplitude of $\gamma\g\rightarrow t\bar t$ transition is determined by the nonrelativistic expansion of a $T$ product of two vector currents \cite{PP2,Piv}. This coefficient, however, can be found by direct comparison with the relativistic expression for the cross section expanded in the velocity of the heavy quark (see Section~3.1). For the noninteracting quarks (the Born approximation) one obtains the following results for the cross sections $(\beta=\sqrt{1-4m_t^2/s})$ \[ R^v(\beta)={3\over 2}q_t^2N_c\theta(\beta^2)(\beta+O(\beta^3)),\qquad R^a(\beta)=N_c\theta(\beta^2)(\beta^3+O(\beta^5)), \] \[ R^{++}(\beta)=6q_t^4N_c\theta(\beta^2)(\beta+O(\beta^3)),\qquad R^{+-}(\beta)=8q_t^4N_c\theta(\beta^2)(\beta^3+O(\beta^5)). \] Note that the cross sections $R^v$ and $R^{++}$ are saturated with the $S$ wave contribution and are proportional to the Green function at the origin while $R^a$ and $R^{+-}$ parts are saturated with the $P$ wave contribution and are proportional to the derivative of the Green function at the origin. As a consequence they are suppressed in comparison with $R^v$ and $R^{++}$ by factor $\beta^2$. In the present paper we study the corrections to the total cross sections $R^e$ and $R^\gamma$ up to the NNLO of NRQCD. Thus $R^a$ is a NNLO contribution to the total cross section $R^e$ and only the leading contribution to $R^a$ is important. On the contrary, the $R^{+-}$ part can be separated from $R^\gamma$ by fixing the opposite helicities of the colliding photons. This makes possible the direct study of the $P$ wave production and, therefore, the evaluation of the corrections to $R^{+-}$ cross section is of practical interest. Concluding this subsection we should also mention that the electroweak corrections to the cross sections are known to the one-loop accuracy. For $e^+e^-$ annihilation they have been obtained in ref.~\cite{Holl} and for $\gamma\g$ collisions in ref.~\cite{Denn}. \subsection{Forward-backward asymmetry. } The important parameter related to threshold production is a space asymmetry of the differential cross sections. This parameter gives more detailed information on the process and allows one to obtain independent experimental data for further testing the theory. The forward-backward asymmetry of the top quark production is defined as a difference of cross sections averaged over forward and backward semispheres eith respect to the electron beam direction devided by the total cross section. A nonvanishing asymmetry appears in $e^+e^-$ annihilation due to the axial coupling of the top quark to the $Z$-boson. The expression of this parameter for energies near the threshold is given by \cite{Sum} \begin{equation} A_{FB}={E_{VA}\over D_V}\left(1+{c^a_1-c^v_1\over 2} {C_F\alpha_s\over \pi} \right)\Phi(k) \label{afb} \end{equation} where \[ E_{VA}=q_eq_ta_ea_td+2v_ea_ev_ta_td^2 \] is the electroweak factor. The expression for the asymmetry in eq.~(\ref{afb}) is given in NLO and the explicit correction of order $\alpha_s$ is taken in the linear approximation that leads to the manifest difference of axial and vector hard coefficients in this order. The dynamical quantity is a function \begin{equation} \Phi(k)={1\over m_t}{{\rm Re}\int \tilde G^*(p,k) \tilde F(p,k) p^3dp \over \int \tilde G^*(p,k) \tilde G(p,k)p^2dp} \label{phidef} \end{equation} that describes the overlap of the $S$ and $P$ partial waves. Here ${\bf p}\tilde F(p,k)$ and $\tilde G(p,k)$ are the Fourier transforms of $i\partial_{\bf y}G({\bf x},{\bf y},k)|_{y=0}$ and $G(x,0,k)$ correspondingly. In the Born approximation the expression for the function $\Phi(\beta)$ can be found in the explicit form and is rather simple $\Phi(\beta)={\rm Re}\,\beta$. It vanishes for the real values of energy below threshold. \subsection{Top quark polarization.} The longitudinal top quark polarization in the process $e^+e^-\rightarrow t\bar t$ averaged over the production angle reads \cite{Har,Chi} \[ \langle P_L\rangle =-{4\over 3}{D_{VA}\over D_V} \left(1+{c^a_1-c^v_1\over 2}{C_F\alpha_s\over \pi}\right)\Phi(k) \] where \[ D_{VA}=q_eq_tv_ea_td+(v_e^2+a_e^2)v_ta_td^2 \] and $\Phi(k)$ is given by eq.~(\ref{phidef}). This function enters also the expression for the averaged longitudinal top quark polarization in $\gamma\g\rightarrow t\bar t$ process with the same helicity colliding photons \cite{FKK} \[ \langle P_L\rangle =\pm 2\left(1+{c^{+-}_1-c^{++}_1\over 2}{C_F\alpha_s\over \pi}\right) \Phi(k) \] where $+$ $(-)$ correspond to the positive (negative) helicity photons. The extension of the above expressions to the general electron/positron polarization and photon helicity and to other component of the polarization vector can be found in the literature (e.g. refs.~\cite{Har,FKK}). \subsection{The effects of the finite top quark width.} As has already been mentioned the sufficiently large $t$-quark decay width suppresses the nonperturbative effects of strong interactions at large $(\sim 1/\Lambda_{QCD})$ distances and makes the perturbation theory applicable for the description of the $t$-quark threshold dynamics. The near-threshold dynamics is nonrelativistic and is rather insensitive to the hard momentum details of $t$-quark decays. Therefore as the leading order approximation the instability of the top quark can be parameterized with the constant mass operator. The finite top quark width can then be taken into account by the direct replacement $m_t\rightarrow m_t-i\Gamma_t/2$ in the relevant argument $s-4m_t^2$ describing the functional dependence of physical quantities near the threshold, or $E\rightarrow E+i\Gamma_t$ \cite{FK}. This approximation accounts for the leading imaginary electroweak contribution to the leading order NRQCD Lagrangian. Since the essential features of the physical situation are caught within this approximation we neglect the electroweak effects in higher orders in the strong coupling constant and heavy quark velocity. However, in the case of $P$ wave production and $S-P$ wave interference the above prescription is not sufficient for a proper description of the entire effect of the non-zero top quark width \cite{ttgg} and more thorough analysis is necessary (see Sections~3.2,~3.3). In the context of the top quark finite lifetime we should also mention the unfactorizable corrections due to the top quark interaction with the decay products which are suppressed in the total cross sections \cite{FKM} but should be taken into account as NLO contributions to the angular distribution and top quark polarization \cite{PetSum}. \section{Nonrelativistic Green function beyond the leading order.} The basic quantity in the analysis of the threshold effects is the nonrelativistic Green function of the Schr{\"o}dinger equation~(\ref{Schr}). The Green function has a standard partial wave decomposition \begin{equation} G({\bf x},{\bf y},k)=\sum^\infty_{l=0}(2l+1) (xy)^lP_l({\bf xy}/xy)G_l(x,y,k) \label{lexp} \end{equation} where $P_l(z)$ is a Legendre polynomial. The partial waves of the Green function of the pure Coulomb Schr{\"o}dinger equation $G^C({\bf x},{\bf y},k)$ are known explicitly \begin{equation} G^C_l(x,y,k)={m_tk\over 2\pi}(2k)^{2l}e^{-k(x+y)} \sum_{m=0}^\infty {L_m^{2l+1}(2kx) L_m^{2l+1}(2ky)m!\over (m+l+1-\nu)(m+2l+1)!} \label{Gcl} \end{equation} where $\nu=\lambda / k$, $\lambda =\alpha_s C_F m_t/2$ with $\alpha_s$ is taken at the soft scale $\mu_s$. $L^\alpha_m(z)$ is a Laguerre polynomial which is chosen in the form \[ L_m^\alpha(z)={e^zz^{-\alpha}\over m!}\left({d\over dz}\right)^m (e^{-z}z^{m+\alpha}) . \] We, however, need to know the nonrelativistic Green function for the NNLO Hamiltonian of the following form \[ {\cal H}={\cal H}_C+\Delta{\cal H} . \] The second term of this expression describes the corrections to the Coulomb Hamiltonian \begin{equation} \Delta{\cal H}=-{{\bf \Delta}_{\bf x}^2\over 4m_t^3} +\Delta_1V(x)+\Delta_2V(x) +\Delta_{NA}V(x)+\Delta_{BF}V({\bf x},\partial_{\bf x},{\bf S}) . \label{Ham} \end{equation} The first term of the equation is the standard correction to the kinetic energy operator, $\Delta_{NA}V(x)=-C_AC_F\alpha_s^2/(2m_tx^2)$ is the so called non-Abelian potential of quark-antiquark interaction \cite{Gup} and $\Delta_{BF}V({\bf x}, \partial_{\bf x},{\bf S})$ is a standard Breit-Fermi potential known since long ago (only the overall color factor $C_F$ is new). The Breit-Fermi potential contains the quark spin operator ${\bf S}$, {\it e.g.}~\cite{Landau}. In NNLO the cross section $R^{v}$ is saturated by the final state configuration of $t\bar t$ pair with $l=0$, $S=1$ while $R^{++}$ cross section is saturated by $l=0$, $S=0$ configuration. The Breit-Fermi potential takes the following form when considered on the $l=0$ states \[ \Delta_{BF}V(x)={C_F\alpha_s\over x}{{\bf \Delta_x}\over m_t^2}+ A^i{C_F\alpha_s\pi\over m_t^2}\delta({\bf x}) \] where $A^v=11/3$ corresponds to the spin one final state of $e^+e^-\rightarrow t\bar t$ production and $A^{++}=1$ corresponds to the spin zero final state of $\gamma\g\rightarrow t\bar t$ production. The terms $\Delta_iV$ ($i=1,2$) represent the first and second order perturbative QCD corrections to the Coulomb potential \cite{Fish,Peter} \[ \Delta_1 V(x)={\alpha_s\over 4\pi}V_C(x)(C_0^1+C_1^1\ln(x\mu_s)), \] \[ \Delta_2 V(x) = \left({\alpha_s\over 4\pi}\right)^2V_C(x)(C_0^2+C_1^2\ln(x\mu_s) +C_2^2 \ln^2(x\mu_s)) \] where \[ C_0^1=a_1+2\beta_0\gamma_E,\qquad C_1^1=2\beta_0, \] \[ C_0^2=\left({\pi^2\over 3}+4\gamma_E^2\right)\beta_0^2 +2(\beta_1+2\beta_0a_1)\gamma_E+a_2, \] \[ C_1^2=2(\beta_1+2\beta_0a_1)+8\beta_0^2\gamma_E, \qquad C_2^2=4\beta_0^2, \] \[ a_1={31\over 9}C_A-{20\over 9}T_Fn_f, \] \[ a_2= \left({4343\over 162}+4\pi^2-{\pi^4\over 4} +{22\over3}\zeta(3)\right)C_A^2- \left({1798\over 81} + {56\over 3}\zeta(3)\right)C_AT_Fn_f \] \[ -\left({55\over 3} - 16\zeta(3)\right)C_FT_Fn_f +\left({20\over 9}T_Fn_f\right)^2, \] \[ \beta_1={34\over 3}C_A^2-{20\over 3}C_AT_Fn_f-4C_FT_Fn_f. \] Here $\alpha_s$ is defined in $\overline{\rm MS}$ renormalization scheme. The invariants of the color symmetry $SU(3)$ group have the following numerical values for QCD: $C_A=3$, $C_F=4/3$, $T_F=1/2$, $n_f=5$ is the number of light quark flavors, $\beta_0=11C_A/3-4T_Fn_f/3$ is the first $\beta$-function coefficient, $\gamma_E=0.577216\ldots$ is the Euler constant and $\zeta(z)$ is the Riemann $\zeta$-function. The solution to eq.~(\ref{Schr}) with the Hamiltonian~(\ref{Ham}) can be found within the standard nonrelativistic perturbation theory around the Coulomb Green function as a leading order approximation \[ G({\bf x},{\bf y},k) =G_C({\bf x},{\bf y},k)+\Delta G({\bf x},{\bf y},k), \] \begin{equation} \Delta G({\bf x},{\bf y},k)=-\int G_C({\bf x},{\bf z},k) \,\Delta{\cal H }\,G_C({\bf z},{\bf y},k)d{\bf z}+\ldots \label{totcorr} \end{equation} In the previous section the threshold effects in the basic observables were reduced to three universal functions: the Green function at the origin which is saturated by the $S$ wave contribution, the derivative of the Green function at the origin which is saturated by the $P$ wave contribution and the function $\Phi(k)$ which describes the $S-P$ wave interference. These functions are analyzed in detail in Sections 3.1-3.3 \subsection{$S$ wave production.} Only the $l=0$ component of the Green function (\ref{lexp}) contributes to its value at the origin \[ G(0,0,k)=G_0(0,0,k). \] The explicit expression for the Coulomb part of the Green function has the form \[ G^C_0(x,0,k)|_{x\rightarrow 0}= {m_t\over 4\pi}\left({1\over x}- 2\lambda\ln\left({2x\mu_f}\right) -2\lambda\left({k\over 2\lambda}+ \ln\left({k\over \mu_f}\right) \right.\right. \] \begin{equation} +2\gamma_E-1+ \Psi_1\left(1- \nu\right)\bigg)\bigg) \label{G0} \end{equation} where $\Psi_n(z)=d^n\ln{\Gamma(z)}/dz^n$ and $\Gamma(z)$ is the Euler $\Gamma$-function. The energy independent finite part of this expression is chosen for later convenience. Eq.~(\ref{G0}) can be most easily obtained from the general expression for the Coulomb partial waves \begin{equation} G^C_l(x,0,k)= {m_tk\over 2\pi}(2k)^{2l}e^{-kx}\Gamma(l+1-\nu)U(l+1-\nu,2l+2,2kx) \label{hyp} \end{equation} where $U(a,b,z)$ is the confluent hypergeometric function. In the short distance limit $x\rightarrow 0$ the Coulomb Green function $G^C({\bf x},0,k)=G^C_0({x},0,k)$ has $1/x$ and $\ln(x)$ divergent terms. These terms, however, are energy independent and do not contribute to the cross section. Hence these terms can be subtracted without affecting any physical results. The quantity $\mu_f$ in eq.~(\ref{G0}) is an auxiliary parameter, the factorization scale, which drops out from the physical observables. The NLO correction $\Delta_1G$ to eq.~(\ref{G0}) due to the first iteration of $\Delta_1V$ term of the QCD potential has been found in ref.~\cite{KPP} where the simple and efficient framework for computation of higher orders was formulated. The result of the evaluation of the NLO correction is \[ \Delta_1G_0(0,0,k) ={\alpha_s\beta_0\over 2\pi}{\lambda m_t\over 2\pi}\left( \sum_{m=0}^\infty F(m)^2(m+1) \left(L_1(k)+\Psi_1(m+2)\right)-2\sum_{m=1}^\infty\sum_{n=0}^{m-1} F(m) \right. \] \[ \left. \times F(n){n+1\over m-n} +2\sum_{m=0}^\infty F(m) \left(L_1(k) - 2\gamma_E-\Psi_1(m+1)\right) -\gamma_E L_1(k)+{1\over 2}L_1(k)^2 \right) \] where \[ L_1(k)=\ln\left({\mu_se^{C_0^1/C_1^1}\over 2k }\right) \] and \[ F(m)={\nu\over (m+1)\left(m+1-\nu\right)}. \] The NNLO correction $\Delta_2^{(2)}G$ due to $\Delta_2V$ part of the potential and the correction $\Delta_2^{(1)}G$ due to the second iteration of $\Delta_1V$ part of the correction to the Coulomb static potential have been obtained in refs.~\cite{KPP,PP}. While the technique is rather straightforward the results of the calculations are cumbersome and explicit formulae are relegated to Appendix~A. The method of calculation of the correction to the Green function at the origin due to logarithmic terms in the potential is described in details in ref.~\cite{PP1}. It is based on the representation of the Coulomb Green function as an expansion over the Laguerre polynomials~(\ref{Gcl}). This representation is very close to the standard physical expansion over the eigenfunctions that makes the technique transparent and easily interpretable in physical terms. It is equally suitable for any partial wave contribution as has been shown in ref.~\cite{PP2} where results for the $P$ wave production were found. The results for the $S$ wave part of the corrections were reproduced within a different technical framework based on an integral representation of the Coulomb Green function in ref.~\cite{MY}. The corrections to the Coulomb Green function at the origin due to ${\bf\Delta}^2$, $V_{NA}$ and $V_{BF}$ terms have been presented in \cite{Hoang,Mel}. They are of the following explicit form \[ \Delta_{{\bf \Delta}^2,NA,BF}G= {m_t\over 4\pi}{k^2\over m_t^2}\left( {5\over 8}k+ 4\lambda\left( \ln\left({k\over \mu_f}\right)+\gamma_E +\Psi_1\left(1- \nu\right)\right) -{11\over 8}C_F\alpha_s\nu \Psi_2\left(1-\nu\right)\right) \] \begin{equation} +\pi{C_F\alpha_s\over m_t^2} \left(5-A^i+2{C_A\over C_F}\right)G_C(0,0,k)^2 \, . \label{BFcorr} \end{equation} In the course of evaluation of this correction to the nonrelativistic Green function one encounters the ultraviolet divergence in the imaginary part of the Green function contained in the last term of eq.~(\ref{BFcorr}). This divergence is related to the singular behavior of the Coulomb Green function at the origin. The particular form of this divergence depends on the regularization procedure. The divergence appears in the process of scale separation and is a consequence of the fact that the nonrelativistic approximation is not adequate for the description of the short distance effects. The hard coefficient $C^{v,++}$ computed within the same regularization procedure as the Green function itself must have the infrared singular term which exactly cancels the one appearing in the Green function. The hard coefficient can be evaluated by matching the effective and full theory cross sections in the weak coupling limit \cite{H1,Hoang} or by an explicit splitting of the hard and soft contributions using, for example, the scale factorization in dimensional regularization \cite{BSS,PinSot1,CMY}. Let us consider the cancelation of the divergences and determination of the hard coefficient in the matching scheme. The natural regularization for the analysis of the hard part of the corrections is the dimensional one \cite{c1,c2}. In $4-2\varepsilon$ dimensions the infrared divergence of the hard contribution in NNLO has the form of the first order pole in $\varepsilon$. The Coulomb Green function at the origin in eq.~(\ref{BFcorr}) can be regularized in the same way. The dimensionally regularized Coulomb Green function at the origin can be defined as follows (see Appendix~B) \begin{equation} G^{d.r.}_C(0,0,k)=-{m_t\over 4\pi}\left(k + 2\lambda \left(-{1\over 2\varepsilon}+\ln\left({k\over\mu_f}\right) +\gamma_E+\Psi_1(1-\nu)\right)\right) +O(\varepsilon). \label{dimreg} \end{equation} Note that the Green function regularized in this way differs from that which has been obtained in ref.~\cite{CMY} within another scheme of dimensional regularization. In contrast to eq.~(\ref{G0}) in~eq.~(\ref{dimreg}) there is no divergence in the Born approximation. The Green function in this approximation is a nonrelativistic free propagator and is proportional to $k$. The first order pole in $\varepsilon$ appears only in the first order in $\alpha_s$. The $O(\alpha^2)$ singular $1/\varepsilon$ term in the imaginary part of eq.~(\ref{BFcorr}) is proportional to ${\rm Im}(G_C(0,0,k))$ and, therefore, can be absorbed by the redefinition of the hard coefficient $C^{v,++}$. For the Green function this redefinition results in the substitution $G^{d.r.}_C(0,0,k)\rightarrow G^s_C(0,0,k)$ in eq.~(\ref{BFcorr}) where the ``subtracted'' Green function reads \begin{equation} G^s_C(0,0,k)=G^{d.r.}_C(0,0,k)-{m_t\lambda\over 4\pi} {1\over \varepsilon}. \label{G0r} \end{equation} Within the redefined hard coefficient the $O(\alpha^2)$ ``ultraviolet'' $1/\varepsilon$ term stemming from the corrections to the Green function eq.~(\ref{BFcorr}) exactly cancels the $O(\alpha^2)$ ``infrared'' $1/\varepsilon$ term of the hard part of the corrections regularized in the same way. Then the (finite) coefficient $C^{v,++}$ can be found directly by matching the effective theory expression for the cross sections and the result of perturbative QCD calculation of the spectral density in the formal limit $\alpha_s\ll \beta\ll 1$ up to the order $\alpha_s^2$ for $\mu_h = \mu_s$. Eqs.~(\ref{Ree},~\ref{Rgg}) in the matching limit take the form \[ R^v={3\over 2}N_cq_t^2\beta\left(\left(1+\left(1-B^v\right) \beta^2\right)+ C_F{\alpha_s\over \pi}\left({\pi^2\over 2}{1\over\beta}+ c^v_1+ {\pi^2\over 3}\beta\right)\right. \] \[ +C_F\left({\alpha_s\over \pi}\right)^2 \left(C_F{\pi^4\over 12}{1\over\beta^2}+ \pi^2\left(C_F{c^v_1\over 2}+{1\over 8} \left(-C_1^1\ln{\left({2\beta m_t\over\mu_s}\right)}+C_0^1-2\beta_0\gamma_E\right) \right){1\over\beta}\right. \] \begin{equation} \left. +C_F{5\pi^4\over 36}+c^v_2-C_F\pi^2\left({5-A^v\over 2}+{C_A\over C_F} \right)\ln{\left({\beta m_t\over\mu_f}\right)}\right), \label{re} \end{equation} \[ R^{++}=6N_cq_t^4\beta \left(\left(1+\left(1-B^{++}\right)\beta^2\right)+ C_F{\alpha_s\over \pi}\left({\pi^2\over 2}{1\over\beta}+ c^{++}_1+ {\pi^2\over 3}\beta\right)\right. \] \[ +C_F\left({\alpha_s\over \pi}\right)^2 \left(C_F{\pi^4\over 12}{1\over\beta^2}+\pi^2 \left(C_F{c^{++}_1\over 2}+{1\over 8} \left(-C_1^1\ln{\left({2\beta m_t\over\mu_s}\right)}+C_0^1-2\beta_0\gamma_E\right) \right){1\over\beta}\right. \] \begin{equation} \left. +C_F{5\pi^4\over 36}+c^{++}_2-C_F\pi^2\left({5-A^{++}\over 2}+ {C_A\over C_F} \right)\ln{\left({\beta m_t\over\mu_f}\right)}\right) +R^{++}_P \label{rg} \end{equation} where the terms of the relative order $O(\beta^2)$ are kept. In eq.~(\ref{rg}) the $P$ wave contribution \[ R^{++}_P=6N_cq_t^4\beta^3+\ldots \] is due to the derivative term in eq.~(\ref{Rpp}) to be discussed. By comparing eq.~(\ref{re}) with the NNLO QCD result for the cross section $R^v$ expanded in the velocity of the heavy quark near the threshold \cite{c1,sd} one finds \begin{equation} c^v_2=\tilde c^v_2-c^v_1{\beta_0\over 2} \ln{\left(m_t\over\mu_h\right)}+\pi^2\left({2\over 3}C_F+C_A \right)\ln{\left(m_t\over\mu_f\right)}, \label{ce0} \end{equation} where the coefficient $\tilde c^v_2$ has been obtained in refs.~\cite{Hoang,Mel} \[ \tilde c^v_2=\left({39\over 4}-\zeta(3)+{4\pi^2\over 3} \ln{2}-{35\pi^2\over 18}\right)C_F-\left({151\over 36}+{13\over 2} \zeta(3)+{8\pi^2\over 3} \ln{2}-{179\pi^2\over 72}\right)C_A \] \[ +\left({44\over 9}-{4\pi^2\over 9}+{11\over 9}n_f\right)T_F . \] The first logarithm in eq.~(\ref{ce0}) is determined by the renormalization group running of the strong coupling constant in the hard momentum regime and is proportional to the first coefficient of the $\beta$ function. Thus, both the hard coefficient and the imaginary part of the Green function do not depend on the normalization point of $\alpha_s$ in the fixed order of perturbation theory so one can use different scales for $\alpha_s$ in these quantities. The second logarithm corresponds to the anomalous dimension of the hard coefficient and precisely cancels the factorization scale dependence of the Green function due to eq.~(\ref{BFcorr}) making the total result independent of the factorization scale. Note that the use of different hard and soft normalization scales leads to the incomplete cancelation of the factorization scale dependence which, however, is the higher order ($O(\alpha_s^3)$) effect. As it has been mentioned above one can bypass the direct matching by the consistent use of the same subtraction scheme within dimensional regularization for both hard coefficient and Green function. Then matching is automatic \cite{BenSmi,PinSot1,CMY}. In this approach the hard coefficient is completely determined by the hard renormalization coefficient of the nonrelativistic vector current \cite{BSS}. However, in order to compute the corrections to the Green function in this case one has to define accurately the Breit-Fermi Hamiltonian and the Green function in $3-2\varepsilon$ dimensions \cite{BSS,CMY} (in our analysis we use three dimensional Breit-Fermi Hamiltonian and direct regularization of the Green function therefore the matching is necessary in order to fix the constant relating the two regularization schemes). The NNLO analysis of the $R^{++}$ cross section is still absent and the constant in the hard coefficient is unknown. The logarithmic part of the NNLO contribution to $C^{++}(\alpha_s)$ reads \begin{equation} c^{++}_2=\tilde c^{++}_2- c^{++}_1{\beta_0\over 2} \ln{\left(m_t\over\mu_h\right)}+\pi^2\left(2C_F+C_A \right)\ln{\left(m_t\over\mu_f\right)} \label{cg} \end{equation} where $\tilde c^{++}_2$ is a constant to be determined. The relativistic correction to this cross section, however, can be extracted from the calculations presented earlier in the literature. Comparing the known result \cite{INOK} \begin{equation} R^{++}(\beta)=6q_t^4N_c\beta\left(1+{2\over 3}\beta^2+O(\beta^4)\right) \label{rel} \end{equation} with our expressions from eqs.~(\ref{rg},~\ref{rel}) we find \[ B^{++}={4\over3}. \] The Green function at the origin can be written in the form which includes only single poles in the energy variable. This form seems to be natural for a Green function of the nonrelativistic Schr{\"o}dinger equation \begin{equation} G(0,0,E)=\sum_{m=0}^\infty{|\psi_{0m}(0)|^2\over E_{0m}-E}+ {1\over \pi}\int_0^\infty{|\psi_{0E'}(0)|^2\over E'-E} dE' \label{endenom} \end{equation} where $\psi_{0m,E'}(0)$ is the wave function at the origin, the sum goes over the bound states and the integral goes over the states of a continuous part of the spectrum. In this way the corrections to the Green function stemming from the discrete part of the spectrum reduce to corrections to Coulomb bound state energy levels \[ E_{0m}=E^C_{0m}\left(1+\Delta_1E_{0m}+\Delta_2E_{0m}\right) \] and to the values of Coulomb bound state wave functions at the origin \[ |\psi_{0m}(0)|^2=|\psi^C_{0m}(0)|^2\left(1+\Delta_1\psi^2_{0m} +\Delta_2\psi^2_{0m}\right) \] where \[ E^C_{0m}=-{\lambda^2\over m_t(m+1)^2}, \qquad |\psi^C_{0m}(0)|^2={\lambda^3\over \pi(m+1)^3}, \] \[ \Delta_2E_{0m}=\Delta_{{\bf \Delta}^2,NA,BF}E_{0m} +\Delta^{(2)}_2E_{0m} +\Delta^{(1)}_2E_{0m}, \] \[ \Delta_2\psi^2_{0m}=\Delta_{k^2}\psi^2_{0m}+ \Delta_{{\bf\Delta}^2,NA,BF}\psi^2_{0m}+\Delta^{(2)}_2\psi^2_{0m} +\Delta^{(1)}_2\psi^2_{0m} \] and $\Delta_{k^2}\psi^2_{0m}$ is the correction due to relativistic corrections parameterized by the coefficients $B^i$ which we include into the definition of the wave function. In NLO an explicit analytical expression for the corrections to the bound state parameters has the form \cite{MY,PP1,PY} \[ \Delta_1E_{0m}={\alpha_s\beta_0\over \pi}\left(\bar L_1(m)+\Psi_1(m+2)\right), \] \[ \Delta_1\psi^2_{0m}={\alpha_s\beta_0\over 2\pi} \left(3\bar L_1(m)+\Psi_1(m+2)-2(m+1)\Psi_2(m+1) -1-2\gamma_E+{2\over m+1} \right) \] where $\bar L_1(m)=L_1(\lambda /(m+1))$. The expressions of the NNLO corrections to the energy levels \cite{MY,PP1,PY} and wave functions at the origin \cite{MY,PP1} are rather cumbersome and given in Appendix~C, D. The continuum contributions in eq.~(\ref{endenom}) can be directly found by subtracting the discrete part of these equations expanded around the Coulomb approximation up to NNLO \[ \sum_{m=0}^\infty{|\psi^C_{0m}(0)|^2\over E^C_{0m}-E}\left( 1+\Delta_1\psi^2_{0m}+\Delta_2\psi^2_{0m} +{\left(1+\Delta_1\psi^2_{0m}\right)\Delta_1E_{0m}+\Delta_2E_{0m}\over 1-E/E^C_{0m}} \right. \] \[ \left. +{\Delta_1E_{0m}^2\over (1-E/E^C_{0m})^2}\right) \] from the result obtained within the nonrelativistic perturbation theory for the Green function at the origin~(\ref{totcorr}) multiplied by $(1-B^iE/m_t)$. This procedure removes the double and triple poles from eq.~(\ref{totcorr}) and leaves only the single poles in the expression for the Green function~(\ref{endenom}). An important consequence of the relatively large top quark width is that the most of Coulomb resonances are smoothed out. The numerical analysis shows that only the ground state resonance in the cross sections is distinguishable. Its separation from others is not completely covered by the infrared cutoff provided by the top quark width. Indeed, using the pure Coulomb formulas for estimates within the order of magnitude we find \[ |E^C_{00}-E^C_{01}|={3\lambda^2\over 4m_t}\approx 0.6~{\rm GeV} \] to be compared with the top quark width $\Gamma_t=1.43~{\rm GeV}$. The second spacing between radial excitations for the $l=0$ partial wave and the first spacing for the $l=1$ partial wave are, however, much smaller \[ |E^C_{01}-E^C_{02}|=|E^C_{10}-E^C_{11}|= \frac{5}{36}{\lambda^2\over m_t}\approx 0.11~{\rm GeV} \, . \] Therefore the contributions of higher resonances are completely smeared out with the top quark width. In the limit of vanishing top quark width the NNLO approximation for the energy of the first resonance in $e^+e^-$ annihilation reads \[ E_{00}^v=-{\lambda^2\over m_t}\left(1+{\alpha_s\over 4\pi} 2C_1^1\left(L_1(\lambda)+1-\gamma_E\right) +\left({\alpha_s\over 4\pi}\right)^2\bigg(2C_1^2(L_2(\lambda)+1-\gamma_E) \right. \] \[ \left. +(C_1^1)^2\left((L_1(\lambda)-\gamma_E)^2+1-{\pi^2\over 3}-\Psi_3(1)\right) +2C_2^2\left((L(\lambda)+1-\gamma_E)^2-1+{\pi^2\over 6}\right) \right) \] \begin{equation} \left. +C_F^2\alpha_s^2\left({C_A\over C_F}+{1\over 48}\right) \right) \label{e2} \end{equation} where \[ L(\lambda)=\ln\left({\mu_s\over 2\lambda }\right), \qquad L_2(\lambda)=\ln\left({\mu_se^{C_0^2/C_1^2}\over 2\lambda }\right). \] This value is related to the energy of the resonance of the top quark production in $\gamma\g$ collision by the hyperfine splitting \[ E_{00}^v-E_{00}^{++}= {4\over 3}{\lambda^2\over m_t}C_F^2\alpha_s^2\ . \] The convergence of the perturbation theory series~(\ref{e2}) is not fast. For some typical values of the soft normalization scale the series for the resonance energy reads \[ E^v_{00}= E^{LO}_{00}(1+0.36+0.30),\qquad \mu_s=25~{\rm GeV}, \] \[ E^v_{00}= E^{LO}_{00}(1+0.58+0.38), \qquad \mu_s=50~{\rm GeV}, \] \begin{equation} E^v_{00}= E^{LO}_{00}(1+0.68+0.45),\qquad \mu_s=75~{\rm GeV}, \label{div} \end{equation} \[ E^v_{00}= E^{LO}_{00}(1+0.74+0.50),\qquad \mu_s=100~{\rm GeV}. \] The poor convergence of the series for the resonance energy can be assigned to high infrared sensitivity of the pole mass (see, for example, \cite{BB}). The convergence can be manifestly improved by removing the pole mass from the theoretical expressions in favor of some less infrared sensitive mass parameter, for example, the short-distance \cite{MY}, potential-subtracted \cite{BSS} or $1S$ mass \cite{H2}. Note that in a finite order of the expansion all the mass definitions are perturbatively equivalent. The infrared safe mass parameters, however, are ``closer'' to the physical observables since in contrast to the pole mass the corresponding perturbative series are supposed to be better convergent (less divergent). Due to the finite top quark width the location of the peak (maximum) of the cross section is not given only by the position of the ground state resonance but is also affected by the contribution of the higher (smeared out) resonances and the continuum contribution. Due to this effect the absolute value of the NNLO peak energy~(\ref{e2}) counted from the threshold is less than the absolute value of the energy of the ground state resonance $E^{v,++}_{00}$ by about $200~{MeV}$ {\it i.e.} $\sim 7\%$. This shift is essentially smaller than the one related to the perturbative QCD corrections to Coulomb values but considerably larger than the leading nonperturbative contribution due to the gluon condensate \cite{VL} which is suppressed parametrically as $(\Lambda_{QCD}/\lambda)^4<1\%$. \subsection{$P$ wave production.} The derivative of the Green function at the origin is saturated with its $l=1$ component and explicitly given by the relation \[ \partial^2_{\bf x\bf y} G({\bf x},{\bf y},k)= 9G_1(0,0,k) . \] For the Coulomb Green function from eq.~(\ref{hyp}) we obtain the closed formula for the partial wave $l=1$ Green function at the small space separation of particle \[ G^C_1(x,0,k)|_{x\rightarrow 0}= {m_t\over 36\pi} \left({3\over x^3}+{3\lambda\over x^2}+ {6\lambda^2-3k^2\over 2x} +2\lambda(k^2-\lambda^2)\ln(2x\tilde \mu_f) \right. \] \begin{equation} \left. +\lambda\left(2(k^2-\lambda^2)\left({k\over 2\lambda}+ \ln\left({k\over \tilde\mu_f}\right) +2\gamma_E-{11\over 6}+ \Psi_1\left(1- \nu\right)\right)+{k^2\over 2}\right)\right) \label{dG0} \end{equation} where $\tilde \mu_f$ is the analog of the parameter $\mu_f$ for the $l=0$ partial wave. In the short distance limit $x\rightarrow 0$ the derivative of the Coulomb Green function (or the partial wave with $l=1$) has $1/x^n$ $(n=1,2,3)$ and $\ln(x)$ singularities. In contrast to the case of the $S$ wave production, the value at the origin for the $P$ wave partial Green function contains divergent terms that explicitly depends on the energy (or wave vector) $k$. However, these terms do not contribute to the cross section for the vanishing top quark width $\Gamma_t=0$ because they have no discontinuity across the physical cut in the complex plane of the energy variable in the approximation of top quark zero width. The case of the non-zero top quark width requires to perform a more detailed analysis given below. The correction to the $l=1$ partial wave at the origin due to the first iteration of $\Delta_1V$ term of the QCD potential has been found in ref.~\cite{PP2} \[ \Delta_1G_1(0,0,k) ={\alpha_s\beta_0\over 2\pi}{\lambda m_tk^2\over 18\pi}\left( \sum_{m=0}^\infty \tilde F(m)^2(m+1)(m+2)(m+3) \left(L_1(k)+\Psi_1(m+4)\right) \right. \] \[ -2\sum_{m=1}^\infty\sum_{n=0}^{m-1} \tilde F(m)\tilde F(n){(n+1)(n+2)(n+3)\over m-n} +2\sum_{m=0}^\infty \tilde F(m)\bigg(2\tilde J_0(m)+(m+1)(m+2)L_1(k) \] \[ \left. +(1+\nu)(\tilde J_1(m)+(m+1)L_1(k)) +{\nu(\nu+1)\over 2}(\tilde J_2(m)+ 2L_1(k))\bigg) +\tilde I(k) \right) \] where \[ \tilde F(m)={\nu(\nu^2-1)\over (m+2-\nu)(m+1)(m+2)(m+3)} \] \[ \tilde J_0(m)=-2\Psi_1(m+1)-4\gamma_E +3, \] \[ \tilde J_1(m)=(m+1)(-\Psi_1(m+2)-2\gamma_E +2), \] \[ \tilde J_2(m)={(m+1)(m+2)\over 2}\left(-\Psi_1(m+3)- 2\gamma_E +{3\over2}\right) , \] \[ \tilde I(k)=-{(\gamma_E-1)^2\over 2}-{\pi^2\over 12} -(4-3\gamma_E)\nu +{1-9\gamma_E+6\gamma_E^2+\pi^2\over 4}\nu^2+ {1-3\gamma_E\over 2}\nu^3+{1-\gamma_E\over 4}\nu^4 \] \[ +\left(\gamma_E-1-3\nu+{9-12\gamma_E\over 4}\nu^2 +{3\over 2}\nu^3+{1\over 4}\nu^4\right)L_1(k) +\left(-{1\over 2}+{3\over 2}\nu^2\right)L_1(k)^2 \] For the derivative of the Green function at the origin (or for the $l=1$ partial wave) the analog of eq.~(\ref{endenom}) reads \[ \partial^2_{\bf xy}G({\bf x},{\bf y},E)= \sum_{m=0}^\infty{|\psi'_{1m}(0)|^2\over E_{1m}-E-i0}+ {1\over \pi}\int_0^\infty{|\psi'_{1E'}(0)|^2\over E'-E-i0} dE' \] where, symbolically, \[ |\psi'_{1m,E'}(0)|^2= \partial_{\bf x}\psi^*_{m,E'}({\bf x}) \partial_{\bf y}\psi_{m,E'}({\bf y})|_{x,y=0}. \] Here $E_{1m}$ is the $l=1$ bound state energy. In NLO approximation these quantities read \cite{PP2} \[ E_{1m}=-{\lambda^2\over m_t(m+2)^2}\left(1+ {\alpha_s\over 4\pi} 2C_1^1\left(\bar L_1(m+1)+\Psi_1(m+4)\right)\right)\ , \] and \[ |\psi'_{1m}(0)|^2={\lambda^5\over \pi}{(m+1)(m+3)\over (m+2)^5} \left(1+ {\alpha_s\over 4\pi}C_1^1\left(5\bar L_1(m+1)+5\Psi_1(m+4)- {\pi^2\over 3}(m+2)-1 \right.\right. \] \[ \left.\left. +2\sum^{m-1}_{n=0}{(n+1)(n+2)(n+3)\over (m+1)(m+3)(m-n)^2}\right)\right)\ . \] The continuum contribution is obtained in the same way as it was done in the previous section for the $S$ wave production. In the case of $P$ wave production the simple shift $E\rightarrow E+i\Gamma$ in the nonrelativistic approximation is not sufficient to describe properly the entire effect of the non-zero top quark width \cite{ttgg}. Indeed, eq.~(\ref{dG0}) in the limit $x\rightarrow 0$ with the nonvanishing width has the divergent imaginary part with the leading power singularity $\sim\Gamma_t/x$ related to the free Green function singularity and the logarithmic singularity $\sim\Gamma_t\alpha_s\ln(x)$ produced by the one Coulomb gluon exchange. The presence of these singularities clearly indicates that the coefficient of the constant term linear in $\Gamma_t$ gets a contribution from the large momentum region and cannot be obtained within the pure nonrelativistic approximation. Like the hard coefficients it should be computed in relativistic theory. Parametrically this contribution is not suppressed in comparison to the pure nonrelativistic contribution in the threshold region. At $E=0$, for example, the ratio between the relativistic (proportional to $\Gamma_t$) and nonrelativistic (Coulomb) contributions is of order ${\Gamma_t\over\alpha_s^2m_t}\sim 1$. Since we are interested in the NLO corrections the $O(\Gamma_t\alpha_s)$ term also has to be taken into account. By construction, the nonrelativistic effective theory has to reproduce the perturbation theory predictions in the formal matching limit $\alpha_s,~ \beta \ll \Gamma_t/m_t\ll 1$ where both effective theory and perturbation theory descriptions are valid. Thus one has to compute $O(\Gamma_t)$ and $O(\Gamma_t\alpha_s)$ terms within the relativistic perturbation theory and then to fix the parameters of the effective nonrelativistic theory so that it reproduces the perturbative results in the matching limit. Within the relativistic perturbation theory the relevant contributions can be obtained by inserting the complex momentum-dependent mass operator into the top quark propagator at $\beta =0$ (only the leading terms in $\Gamma_t/m_t$ should be kept). In the leading order in $\alpha_s$ this procedure has been done in refs.~\cite{ttgg}. The result reads \[ G^C_1(0,0,k)|_{\Gamma_t}= {m_t^3\over 36\pi}g_1\Gamma_t \] where $g_1$ is a coefficient coming from the relativistic treatment with numerical value $g_1=0.185\ldots$ For the term of the order $O(\Gamma_t\alpha_s)$ the necessary calculation has been performed in ref.~\cite{PP2}. It has been shown that the proper relativistic analysis leads to fixing the auxiliary parameter of eq.~(\ref{dG0}) $\tilde \mu_f=g_2 m_t$ where $g_2$ is the coefficient coming from the relativistic treatment. Its numerical value is $g_2=0.13\ldots$. Here we should note also the problem of the previous numerical analysis of the $P$ wave contribution \cite{KuhTeu}. While solving the Schr{\"o}dinger equation~(\ref{Schr}) numerically for the finite top quark width one has to introduce an explicit ultraviolet cutoff for the nonrelativistic expressions divergent in the large momentum region. To get rid of the cutoff dependence one has to compute the hard contribution within the relativistic approximation using the similar prescription for the infrared cutoff. This, however, has not been done and as a consequence the $O(\Gamma_t)$ and $O(\Gamma_t\alpha_s)$ contributions to the cross section were not determined within the numerical framework of ref.~\cite{KuhTeu}\footnote{Recently the $O(\Gamma_t)$ contribution has been estimated within the numerical approach \cite{HoaTeu} by using the physical (relativistic) phase space for the unstable top quark to regularize the divergence of the nonrelativistic approximation.}. On the other hand the total $O(\Gamma_t)$ contribution to the cross section is numerically small in comparison with that of the regular completely nonrelativistic terms of eq.~(\ref{dG0}) which saturate the total result for the energies not far below the threshold. \subsection{$S-P$ interference.} In the zero width approximation the function~(\ref{phidef}) allows for the following decomposition \[ \Phi(E) =\Phi_{pol}(E)+\Phi_{con}(E) \] where $\Phi_{con}$ and $\Phi_{pol}$ are the continuum and bound state poles contributions correspondingly. This is known \cite{FKK,Chi} that the continuum contribution is not affected by the Coulomb effects and above the threshold one has the Born approximation result \[ \Phi^C_{con}(E)={\rm Re}\sqrt{E\over m_t} \] even for the Coulomb Green function in eq.~(\ref{phidef}). Below the threshold in the Coulomb approximation one gets \begin{equation} \Phi^C_{pol}(E)=\left(\sum^\infty_{m=0}{\phi_m^C\over (E^C_{1m}-E)^2}\right) \left(\sum^\infty_{m=0}{|\psi_m(0)|^2\over (E^C_{0m}-E)^2}\right)^{-1} \label{ph0} \end{equation} where the quantities \[ \phi_m^C={\lambda^4\over m_t\pi}{(m+1)(m+3)\over (m+2)^5} \] measure the overlap of the $S$ and $P$ wave functions. Note that in the zero width limit the function $\Phi^C_{pol}$ does not vanish due to the Coulomb degeneration of the energy levels with different $l$: $E^C_{0m+1}=E^C_{1m}$. It was pointed in ref.~\cite{Chi} that the continuum contribution gets no soft corrections. Thus in NLO for a finite top quark width we have the simple result \[ \Phi_{con}(E)={\rm Re}\sqrt{E+i\Gamma_t\over m_t} . \] The corrections to the pole contribution are less trivial. They can be computed using the powerful technique developed in refs.~\cite{KPP,PP1,PP2}. The result reads \begin{equation} \Phi_{pol}(E)={\rm Re}\left(\sum^\infty_{m=0}{\phi_m\over (E_{0(m+1)}-E+i\Gamma_t)(E_{1m}-E-i\Gamma_t)}\right) \left(\sum^\infty_{m=0}{|\psi_m(0)|^2\over (E_{0m}-E)^2+\Gamma_t^2}\right)^{-1} \label{ph1} \end{equation} where \[ \phi_m=\phi_m^C\left(1+ {\alpha_s\beta_0\over \pi}\left( 4L(m+1)+4\Psi(m+4)+ {m\over m+3}-2-{\pi\over 6}(m+2) \right.\right. \] \[ \left.\left. +\sum_{n=0}^{m-1}{(n+1)(n+2)(n+3)\over (m+1)(m+3)(m-n)^2}\right)\right) \] In eq.~(\ref{ph1}) we keep the finite top quark width to get a nonvanishing result since the Coulomb degeneration is lifted by the logarithmic corrections to the potential. Strictly speaking, our approach is valid only if the level splitting $E_{0m+1}-E_{1m}$ is much smaller than the top quark width (which is realized for the actual values of these quantities). For $E_{0m+1}-E_{1m}>\Gamma_t$ the nonrelativistic analysis is not applicable for the $S-P$ interference below the threshold because the double poles of eq.~(\ref{ph0}) disappear and the nonrelativistic contribution is not enhanced in comparison to the relativistic one in this case. Note that for the finite top quark width the interference of the free $l=0$ and $l=1$ partial waves results in the logarithmically divergent $O(\Gamma_t)$ term in the numerator of eq.~(\ref{phidef}) (this term does not include the factor $\Gamma_t$ explicitly but it is suppressed in comparison to the leading term which is proportional to $1/\Gamma_t$ as the leading term in the denominator of eq.~(\ref{phidef}) for the free quark Green function). This divergent term is of the same nature as the divergence in the $P$ wave amplitude discussed in Section~3.2. An accurate calculation of this term can be done only within the relativistic approximation. In contrast to the $P$ wave production this term is parametrically suppressed above the threshold in comparison to the nonrelativistic continuum contribution at least by the factor $\sqrt{\Gamma_t/m_t}$ at $E\sim 0$ and can be safely omitted. However, it becomes important below the resonance region when the nonrelativistic contribution becomes small. Moreover, the denominator in the right hand side of eq.~(\ref{phidef}) decreases rapidly below the ground state pole. Therefore a small uncertainty in the numerator would lead to a large uncertainty in the function $\Phi(k)$ and a reliable estimate of its numerical value is not possible in this region within the nonrelativistic approximation. Strictly speaking the accurate determination of the function $\Phi$ below the ground state pole requires the calculation of the relativistic $O(\Gamma_t)$ contribution to the $S$ wave cross section (the denominator of eq.~(\ref{phidef})) which is not usually considered since it does not lead to the divergence in the nonrelativistic expression. \section{Discussion.} The results of the numerical analysis for the physical observables based on the obtained analytical expressions are plotted in Figs.~1-4. The constant $\tilde c^{++}_2$ appearing in the hard coefficient $C^{++}(\alpha_s)$ in the $O(\alpha_s^2)$ order remains unknown. The calculation of this parameter is necessary for the formal completion of the NNLO analysis. To find its numerical value one has to compute the $O(\alpha_s^2)$ perturbative QCD correction to the $\gamma\g$ cross section near the threshold in the formal limit $\alpha_s\ll \beta\ll 1$ and compare it with $O(\alpha_s^2)$ term in eq.~(\ref{rg}). In the case of $e^+e^-$ annihilation, however, the analogous contribution parameterized by $\tilde c_2^v$ is relatively small (about $10\%$ of the total NNLO correction) and the correction to the physical observables in NNLO is saturated with the soft part of the total contribution determined by the corrections to the parameters of the nonrelativistic Green function. Thus one can reasonably hope that the similar situation can also take place for $\gamma\g$ collisions. However the importance of this parameter for physical observables is not crucial, it affects only the overall normalization of the cross sections. For example, it does not shift the position of the resonance which is an important characteristic of the production and does not enter the ratio $R^{++}(E)/R^{++}(0)$. For the numerical analysis of the cross section $R^\gamma$ we set $\tilde c_2^{++}=0$. In our approach we deal with the soft corrections by summing them into the energy denominators of the discrete part of the Green function. In other words we treat the soft corrections as effective corrections to the parameters of the Green function written in a fixed functional form. The same approach has been advocated in refs.~\cite{BSS,Mel,Nag,HoaTeu,Yak} where all the corrections to the Green function have been found (numerically or analytically) in the form (\ref{endenom}). In ref.~\cite{Hoang}, however, a part of the NNLO corrections has not been resummed to the energy denominators of the discrete part of the Green function. On the other hand, in refs.~\cite{Mel,Nag,HoaTeu,Yak} the Schr{\"o}dinger equation~(\ref{Schr}) has been solved numerically, {\it i.e.} the NLO and NNLO correction to the Coulomb Hamiltonian have been taken into account effectively in all orders of the nonrelativistic series~(\ref{totcorr}) for the Green function while we work strictly in NNLO. Our formulae reproduce the numerical result for $R^v$ of the most recent numerical analysis \cite{Nag,HoaTeu,Yak} with $1\%-3\%$ accuracy that can be assigned to the contribution of the higher iterations of the NLO and NNLO corrections to the potential in eq.~(\ref{totcorr}) beyond NNLO. For the total cross sections which are dominated by $S$ wave contribution we find the typical size of the NNLO corrections to be of the order of $20\%$ in the overall normalization of the cross sections and $\sim 40\%$ in the resonance energies expressed in terms of the top quark pole mass, {\it i.e.} of the order of the NLO ones (see Fig.~1). Though the inclusion of the NLO corrections leads to a considerable stabilization of the theoretical results for the cross sections against changing the normalization point, the NNLO corrections do not lead to better stability as compared to NLO. In the overall normalization of the cross sections the NLO and NNLO corrections cancel each other to a large extent while the NLO and NNLO corrections to the resonance energies are of the same sign and shift the resonance farther from the threshold. They also make the peak more distinguishable which is the main difference between the leading Coulomb and NNLO approximations. The leading order approximation for $R^e$ and $R^\gamma$ cross sections are the same up to the normalization factor $2q_t^2$. Up to the overall factor the difference between the cross sections is determined by NNLO QCD and relativistic corrections (see Fig.~2). Above the threshold this difference is determined by the difference between $B^{++}$ and $B^v$ coefficients and between $P$ wave contributions to eqs.~(\ref{Ree},~\ref{Rgg}) {\it i.e.} by the pure relativistic corrections. Below the threshold in the resonance region this difference is determined also by $A^i$ coefficients and is quite sensitive to the value of $\alpha_s$. Though using an infrared safe mass parameter instead of the pole mass improves the convergence of the series for the resonance energies it does not affect the huge NNLO corrections to the cross sections normalization. Moreover, it is not clear if there exist physically motivated mass and strong coupling parameters providing fast uniform convergence of the perturbative expansion for the cross sections in the threshold region. The absence of such a parameterization would mean the unavoidable significance of the high orders terms of the threshold expansion. Some high order effects have been already considered in the literature. The leading logarithmic corrections related to the renormalization group evolution of the hard coefficient $C^{v}$ have been computed \cite{BSS}. The corresponding corrections to the $R^v$ cross section are $\pm 5\%$. In ref.~\cite{Nag} the running of the strong coupling constant has been taken into account by introducing the energy dependent soft normalization point of $\alpha_s$ entering the Coulomb potential in the numerical solution of the Schr{\"o}dinger equation. The resummation of the renormalization group logarithms has an essential (up to $10\%$) effect in the resonance region and reduces the normalization scale dependence of the result. Furthermore, the effect of retardation which introduces a new type of contributions absent in NLO and NNLO has been analyzed for the low lying resonances \cite{KniPen}. The characteristic scale of the leading ultrasoft contribution was found to be about $-5 \%$ for the square of the ground state wave function at the origin and $+100~{\rm MeV}$ for the ground state pole position. The result for the axial coupling contribution to $e^+e^-\rightarrow t\bar t$ cross section is in a good agreement with the numerical analysis of ref.~\cite{KuhTeu}. Up to the trivial normalization this contribution coincides with the cross section $R^{+-}$ (Fig.~3). Numerically it does not exceed $2\%$ of the total cross section and less than the uncertainty due to the normalization scale dependence. The cross section $R^{+-}$ and the function $\Phi(k)$ obtain no contribution from the ground state resonance and, therefore they are rather smooth because the top quark width smears the higher resonance contributions very efficiently (Figs.~3,~4). These quantities is rather insensitive to a variation of the normalization scale. A typical NLO correction to $R^{+-}$ is about $10\%$ while the one to $\Phi(k)$ is about $15\%$ (the corrections to the forward-backward asymmetry and top quark polarization include also the hard normalization factors which have not been included to $\Phi(k)$ itself and the nonfactorizable corrections corrections discussed in Section~2.4). Our result for the function $\Phi(k)$ (Fig.~4) is in a good agreement with the results of numerical analysis \cite{Sum,Har} for the energies above the ground state resonance. There is some discrepancy between the results below the resonance. However the reliable estimate of the function $\Phi$ is not possible in this region with the pure nonrelativistic treatment of the top quark width as has been explained in Section~3.3. The final remark of this section concerns the optimal choice of the normalization and factorization scales. The hard scale appears in the hard coefficients as $\ln{(m_t/\mu_h)}$ {\it i.e.} the typical hard scale of the problem is the top quark mass. Though in a fixed order of the perturbative expansion the hard coefficients do not depend on $\mu_h$ one can put $\mu_h \sim m_t$ to minimize the potentially large logarithmic contributions of the higher order terms. In practice the NNLO results are almost independent of $\mu_h$ when $\mu_h \sim m_t$. On the other hand the requirement of convergence of the perturbative expansion around the Coulomb Green function restricts the allowed range for the choice of a soft normalization point which can be used for reliable estimates. The soft physical scale of the problem is determined by the natural infrared cutoff related to the top quark width $\sqrt{m_t\Gamma_t}$ that measures the distance to the nearest singularity in the complex energy plane and/or by the characteristic scale of the Coulomb problem $\lambda$ {\it i.e.} $\mu_s\sim 15~{\rm GeV}$. Both scales are rather close to each other for the case of top quark that makes possible a uniform description of both perturbative QCD and Coulomb resonance effects. Indeed, for $\mu_s\sim 15~{\rm GeV}$ the soft NLO correction, for example, to the energy level~(\ref{e2}) reaches its minimal magnitude. However, at this scale the NNLO correction exceeds the NLO one and the series for the energy levels seems to diverge. Moreover, for such a low soft normalization point the NNLO corrections to the wave function at the origin which cannot be dumped by the quark mass redefinition become uncontrollable. This is not surprising since the normalization scale is defined in a rather artificial $\overline {\rm MS}$ scheme that has little to do with peculiarities of $t\bar t$ physics and there is no reason for a literal coincidence of $\mu_s$ parameter with any physical scale of the process. The relative weight of the NNLO correction to the Green function as well as the dependence of the cross sections on $\mu_s$ is stabilized at $\mu_s$\raisebox{-3pt}{$\stackrel{>}{\sim}$}$40~{\rm GeV}$ which can be considered as an optimal choice of the soft normalization point. The price one pays for using different soft and hard normalization scales is the incomplete cancelation of the factorization scale dependence but this effect is suppressed by an additional power of $\alpha_s$. Another source of the dependence on the factorization scale is the factorized form~(\ref{Rv},~\ref{Rpp}) of the cross sections where some higher order $\mu_f$-dependent terms are kept. The numerical analysis, however, shows that the results are rather insensitive to the factorization scale chosen in the region $\mu_f\sim m_t$. \section{Conclusion} The basic observables of the top quark pair production in $e^+e^-$ annihilation and $\gamma\gamma$ collision have been considered in the threshold region. The threshold effects are described by three universal functions related to the $S$, $P$ wave production and $S-P$ wave interference which have been computed analytically within (potential) NRQCD. An explicit analytical expression for the soft part of the NNLO corrections to the total cross section has been obtained. The $e^+e^-\rightarrow t\bar t$ threshold cross section has been obtained in NNLO in the closed form including the contribution due to the top quark axial coupling. The forward-backward asymmetry of the quark-antiquark pair production in $e^+e^-$ annihilation and top quark polarization in both processes have been computed analytically up to NLO. The running of the strong coupling constant and the finite top quark width effects in the $P$ wave production and $S-P$ wave interference have been taken into account properly within the analytical approach. In combination, these uncorrelated observables form an efficient tool for investigating quark interactions. As independent sources they can also be used for determination of the theoretical uncertainty in the numerical values of the strong coupling constant $\alpha_s$, the top quark mass, and the top quark width extracted from the experimental date on top-antitop production. The high order corrections turn out to be relatively large for all observables and important for the accurate description of the top quark physics near production threshold. \vspace{5mm} \noindent {\large \bf Acknowledgments}\\[2mm] This work is partially supported by Volkswagen Foundation under contract No.~I/73611 and Russian Fund for Basic Research under contract No.~97-02-17065. The work of A.A.Penin is supported in part by the Russian Academy of Sciences Grant No.~37. The present stay of A.A.Pivovarov in Mainz, where the paper has been completed, was made possible due to Alexander von Humboldt fellowship. We are thankful to K.~Melnikov and A.~Czarnecki \cite{CzaMel} for pointing out an error in the partial wave decomposition of $R^{++}$ in the previous version of this paper.
\section{Introduction} Identical particle correlations have previously been used to determine source sizes and lifetimes of the emission regions formed in heavy-ion collisions \cite{bh_NA49_98,bh_NA44_98,bh_E859_95,bh_e877_97}. In the case of non-central collisions, such measurements usually integrate over the reaction plane of the collision and therefore obscure physics which depends on the relative orientation of the two colliding nuclei. In this report we discuss preliminary results from our investigations of the dependence of HBT source parameters on the reaction plane. This approach should eventually give us another perspective into the source dynamics. Together with other HBT dependencies of the source geometry and dynamics, we may thus be able to form a more complete picture of the nature and evolution of the emission region. \subsection{The HBT Correlation Formalism} Quantum statistics require that the amplitudes for identical particles be added together before interpreting the square modulus as a probability. Specifically, identical bosons, such as pions, are required to have symmetric two-particle amplitudes, and the HBT correlation arises from the cross-term in the square of the symmetrized amplitude. For a pair of identical bosons, emitted by an extended source $\rho({\bf x})$ with four-momenta ${\bf p}_1$ and ${\bf p}_2$, detected at space-time coordinates ${\bf x}_1$ and ${\bf x}_2$, assuming that the bosonic wavefunctions can be described by plane waves, the symmetrized amplitude of this process is: \begin{equation} \Psi_{12}({\bf x}_1,{\bf x}_2,{\bf p}_1,{\bf p}_2) = \frac{1}{\sqrt{2}}(e^{i({\bf p}_1\cdot {\bf x}_1+{\bf p}_2\cdot {\bf x}_2)} + e^{i({\bf p}_1\cdot {\bf x}_2+{\bf p}_2\cdot {\bf x}_1)}) \end{equation} Which leads to the probability of detecting a pair with relative four-momentum ${\bf q} = {\bf p}_1 - {\bf p}_2$: \begin{equation} C_2({\bf q}) \equiv \frac{1}{N({\bf q})} {\int d^4x_1 d^4x_2~ \rho({\bf x}_1) \rho({\bf x}_2){|\Psi_{12}|}^2} = \frac{1 + {|\tilde{\rho}({\bf q})|}^2}{N({\bf q})} \end{equation} \noindent where the two-particle correlation function $C_2({\bf q})$ is simply related to the Fourier transform of the source density. The normalization N({\bf q}) is discussed below in Section \ref{bh_normalization}. The above formulation is only strictly valid for completely incoherent sources. To account for partial coherence effects, as well as contamination from long-lived resonances, an empirical variable $\lambda$ is added to the definition of the correlation function as a coherence scaling parameter: \begin{equation} C_2({\bf q}) \equiv \frac{1 + \lambda{|\tilde{\rho}({\bf q})|}^2} {N({\bf q})} \end{equation} \subsection{HBT Parameterizations} In practice, the source $\rho({\bf r})$ has usually been assumed to be Gaussian in configuration-space. In this case, the momentum-space distribution is also Gaussian: \begin{equation} \rho({\bf r}) \sim e^{\frac{-{|\vec{r}|}^2}{{\bf R}^2}} \Rightarrow \tilde{\rho}({\bf q}) \sim e^{-{|\vec{q}|{\bf R}}^2} \end{equation} For multi-dimensional HBT analyses, the relative momentum variable ${\bf q}$ can be expressed in terms of a variety of orthogonal components, each of which has model-dependent significance \cite{bh_boal90,bh_wu98}. In the preliminary analysis presented here, a simple 3-D Cartesian parameterization is chosen: ($Q_x, Q_y, Q_z$), with conjugate source parameters ($R_x, R_y, R_z$). $R_z$ is taken along the beam axis; $R_x$ is lies in the reaction plane; and $R_y$ is orthogonal to both. In this case, $C_2({\bf q})$ has the following form: \begin{equation} C_2({\bf q}) = \frac{1}{N({\bf q})}[1 + \lambda{e^{-{(q_xR_x)}^2-{(q_yR_y)}^2-{(q_zR_z)}^2}}] \label{bh_Geometrical_HBT} \end{equation} \subsection{Reaction Plane Determination} The reaction plane is determined using the relative orientation of two axes: the direction of the incoming beam particle, $\hat{z}$, and the direction of the impact parameter $\hat{b}$. In our experiment, the beam axis is defined by a beam vertexing detector (BVER). The BVER detector consists of four planes of scintillating fibers each read out by a multi-anode photomultiplier tube. The fiber planes each consist of $\sim$ 150 200 x 200 $\mu m^2$ fibers, situated 5.84 m and 1.72 m upstream from the target \cite{bh_back98}. The position of the projection of $\hat{z}$ onto the hodoscope can be determined with an accuracy of 1.5 mm at 11.4 m downstream from the target. Charged projectile spectator fragments are detected in a hodoscope (HODO), and their charge centroid calculated for each event. HODO consists of two orthogonal planes of 38 plastic scintillator slats with 1 cm$^2$ cross-sections, centered on the beam line, and situated 11.4 m downstream from the target. The response of individual scintillators to deposited charge was calibrated on a run-by-run basis and this information was used to find the charge-weighted centroid $\vec{Q} = Q_x\hat{i} + Q_y\hat{j}$ for each event, where \begin{equation} Q_x = {\sum Q \cdot x \over \sum Q} \end{equation} \label{bh_RP_Defn} The direction of the impact parameter $\hat{b}$ is then $\hat{Q}$, defined with the origin at the projected beam position on HODO, and the reaction plane is defined as the plane spanned by $\hat{Q}$ and $\hat{z}$. $\hat{x}$ is then redefined to lie along $\hat{Q}$. Implicit in this definition of the impact parameter is the assumption that the direction of proton flow -- the deflection of the spectator fragment -- is along the reaction plane. An estimate of the reaction plane resolution is determined by randomly dividing each event into two sub-events and looking at the ($\phi_1~-~\phi_2$) difference distribution for the two reaction planes calculated from each sub-event (see Fig. \ref{bh_rp_resolution}). The actual resolution for the reaction plane determined using the full event statistics is roughly half this value \cite{bh_ahle98}, and in this manner we obtain an estimate of $\delta\phi \approx 32^{\circ}$. \begin{figure}[ht] \centerline{ \epsfig{file=bh_rp_resolution.eps,height=1.8in }} \caption{Distribution of the reaction plane angle difference for angles $\phi_1$ and $\phi_2$ determined by splitting each event into two sub-events. The angle resolution for the full event is roughly half this value ($\delta\phi \approx$ 32$^\circ$).} \label{bh_rp_resolution} \end{figure} \subsection{Normalization} \label{bh_normalization} The correlation function is, by definition, a normalized quantity. To find the proper normalization N({\bf q}), a background is generated which creates two-particle ``events'' out of single tracks from different events. $C_2({\bf q})$ is then simply the ratio between real data and the event-mixed background. \subsection{Corrections to the correlation function} The shape of the measured two-particle correlation function mainly depends on three effects: the Coulomb repulsion between the two particles, the two-particle resolution of the detector, and the HBT correlation itself. For the Coulomb effect, a correction $f_{C}$ is numerically calculated for a simple extended source \cite{bh_pratt86,bh_baker96} of size $R_0$. This correction is applied iteratively to the background until the value of $R_{inv}$ obtained from the one-dimensional correlation function converges to that of $R_0$. The two-particle resolution arises from the finite resolution of the tracking detectors in the spectrometer. Two close tracks are, at some point, indistinguishable from single tracks, and do not, therefore, appear in the measured two-particle correlation. A two-dimensional cut in relative coordinate space, $f_{\rm TPR}$, is applied to both signal and background events. The separation between two tracks, projected onto the first plane in the spectrometer, is shown in figure \ref{bh_TPAC}, together with $f_{\rm TPR}$. \begin{figure}[ht] \centerline{ \epsfig{file=bh_tpac_parkcity_cut.eps,height=150pt }} \caption{Separation between two tracks on T1. The $f_{\rm TPR}$ cut is the solid line.} \label{bh_TPAC} \end{figure} The final correlation function is fitted from a spectrum in {\bf q}, which is generated by dividing signal events by background events. Both signal and background have been corrected by $f_{\rm TPR}$; the background has additionally been corrected by $f_C$. \section{Results} The reader is reminded that the following results and analysis are preliminary, and only contain a limited subset of the E917 data. Thus far, only about 10\% of the data has passed through the HBT analysis. By December 1999, we expect at least another 50\% will have been analyzed. The measured correlation for identical pairs of pions is shown in Fig. \ref{bh_2D}, plotted as a function of $Q_x$ and $Q_y$, where x and y are defined relative to the reaction plane of the event as discussed in Section \ref{bh_RP_Defn}. To improve the statistics, all of $Q_z$ has been integrated over. In addition, since the radii for $\pi^+\pi^+$ and $\pi^-\pi^-$ pairs are similar \cite{bh_baker96}, both datasets were combined in the present analysis. \begin{figure}[ht] \centerline{ \epsfig{file=bh_lego_interesting.eps,height=1.8in}} \caption{$|Q_y|$ vs. $|Q_x|$} \label{bh_2D} \end{figure} \begin{figure}[ht] \epsfig{file=bh_int_qxproj.eps,height=1.8in} \epsfig{file=bh_int_qyproj.eps,height=1.8in} \dblcaption{\label{bh_2D_x}$|Q_x|$, $|Q_y| < 30$ MeV} {\label{bh_2D_y}$|Q_y|$, $|Q_x| < 30$ MeV} \end{figure} Central slices are shown in Figures \ref{bh_2D_x} and \ref{bh_2D_y} along with the projections from the two-dimensional fit. These data have been fit with the function in Eq. \ref{bh_Geometrical_HBT}, and values for $R_x$ and $R_y$ obtained. We find that $R_x = 2.95 \pm 0.26$ fm and $R_y = 3.34 \pm 0.19$ fm, the latter value being larger than the former. The error bars are solely from the fitting procedure and do not include systematic errors. To establish a baseline for comparison, the data were also analyzed in the same fashion, but with $\hat{x}$ chosen along a random direction rather than along the reaction plane (Figs. \ref{bh_2Dcheck}, \ref{bh_2Dcheck_x}, \ref{bh_2Dcheck_y}). In this analysis, the values of $R_x$ and $R_y$ obtained are consistent with each other, indicating that the observed difference relative to the reaction plane is a real effect. \begin{figure}[ht] \centerline{ \epsfig{file=bh_lego_random.eps,height=1.8in}} \caption{$|Q_y|$ vs. $|Q_x|$ -- systematic check} \label{bh_2Dcheck} \end{figure} \begin{figure}[ht] \epsfig{file=bh_rand_qxproj.eps,height=1.8in} \epsfig{file=bh_rand_qyproj.eps,height=1.8in} \dblcaption {\label{bh_2Dcheck_x}$|Q_x|$, $|Q_y| < 30$ MeV} {\label{bh_2Dcheck_y}$|Q_y|$, $|Q_x| < 30$ MeV} \end{figure} \section{Conclusions} In a naive geometrical model of the source, we would expect $R_y$ to be larger than $R_x$. This is consistent with the simple overlap of two non-centrally colliding spheres for which the overlap region resembles an almond in shape, with $R_y > R_x$. Other data and analyses \cite{bh_back99,bh_ollitraut98}, however, indicate that the emission region is not a simple static source at these energies, with $dN/dy$ distributions and flow studies indicating both longitudinal and transverse expansion. The shape of the source may also be obscured by particle absorption by spectator matter. Future analyses with better statistics will allow the investigation of the source asymmetry with respect to the reaction plane and its dependence on collision centrality, pair $m_T$, and rapidity. These, taken together with $dN/dy$ distributions and flow studies, may be able to further disentangle expansion and absorption effects from the source shape and size, and thus assemble a more complete picture of the collision dynamics. This work is in progress. \begin{chapthebibliography}{1} \bibitem{bh_NA49_98} Appelsh\"auer,~H. {\it et al}. (1998) Nucl. Phys. {\bf A638}, 91c. \bibitem{bh_NA44_98} Bearden,~I.~G. {\it et al}. (1998) Phys. Rev. {\bf C58}, 1656. \bibitem{bh_E859_95} Cianciolo,~V. {\it et al}. (1995) Nucl. Phys. {\bf A590}, 459c. \bibitem{bh_e877_97} Barrette,~J. {\it et al}. (1997) Phys. Rev. Let. {\bf 78}, 2916. \bibitem{bh_boal90} Boal,~D.~H., Gelbke,~C., and B.~K.~Jennings (1990) Rev. Mod. Phys. {\bf 62}, 553. \bibitem{bh_wu98} Wu,~Y.-F., Heinz,~U., Tom\'asik,~B., and U.~A.~Wiedemann (1998) Eur. Phys. J. {\bf C1}, 599. \bibitem{bh_back98} Back,~B.~B. {\it et al}. (1998) Nuclear Instruments and Methods {\bf A412}, 191. \bibitem{bh_ahle98} Ahle,~L. {\it et al}. (1998) Phys. Rev. {\bf C57}, 1416 \bibitem{bh_pratt86} Pratt,~S. (1986) Phys. Rev. {\bf D33}, 72. \bibitem{bh_baker96} Baker,~M. {\it et al}. (1996) Nucl. Phys. {\bf A610}, 213c. \bibitem{bh_back99} Back,~B.~B. {\it et al}. (1999) Contribution to these proceedings. \bibitem{bh_ollitraut98} Ollitraut,~J.-Y. (1998) Nucl. Phys. {\bf A638}, 195c. \end{chapthebibliography} \end{document}
\section{Introduction} \par There is considerable current interest in trying to isolate the lightest glueball. Several experiments have been performed using glue-rich production mechanisms. One such mechanism is Double Pomeron Exchange (DPE) where the Pomeron is thought to be a multi-gluonic object. Consequently it has been anticipated that production of glueballs may be especially favoured in this process~\cite{closerev}. \par The WA102 experiment at the CERN Omega Spectrometer studies centrally produced exclusive final states formed in the reaction \noindent \begin{equation} pp \longrightarrow p_{f} X^{0} p_s, \label{eq:1} \end{equation} where the subscripts $f$ and $s$ refer to the fastest and slowest particles in the laboratory frame respectively and $X^0$ represents the central system. \section{A partial wave analysis of the $K \overline K$ system} \par The isolation of the reaction \begin{equation} pp \rightarrow p_{f} (K^+ K^-) p_{s} \label{eq:b} \end{equation} has been described in detail in a previous publication~\cite{re:kkpap}. A Partial Wave Analysis (PWA) of the centrally produced \mbox{$K^+K^-$ } system has been performed, using the reflectivity basis~\cite{chung}, in 40~MeV intervals of the \mbox{$K^+K^-$ } mass spectrum using an event-by-event maximum likelihood method~\cite{re:kkpap}. The $S_0^-$ and $D_0^-$-Waves from the physical solution are shown in fig.~\ref{fi:1}. \begin{figure}[h] \vspace{7.0cm} \begin{center} \special{psfile=kktalk13.eps voffset=-210 hoffset=10 hscale=60 vscale=60 angle=0} \end{center} \caption{ The $S_0^-$ and $D_0^-$-Waves resulting from a partial wave analysis of the $K^+K^-$ system.} \label{fi:1} \end{figure} \par The $S_0^-$-wave shows a threshold enhancement; the peaks at 1.5 GeV and 1.7~GeV are interpreted as being due to the $f_0(1500)$ and $f_J(1710)$ with J~=~0. A fit has been performed to the $S_0^-$ wave using three interfering Breit-Wigners to describe the $f_0(980)$, $f_0(1500)$ and $f_J(1710)$ and a background of the form $a(m-m_{th})^{b}exp(-cm-dm^{2})$, where $m$ is the \mbox{$K^+K^-$ } mass, $m_{th}$ is the \mbox{$K^+K^-$ } threshold mass and a, b, c, d are fit parameters. The resulting fit is shown in fig.~\ref{fi:1} and gives for the $f_0(980)$ M~=~985~$\pm$~10~MeV, $\Gamma$~=~65~$\pm$~20~MeV, for the $f_0(1500)$ M~=~1497~$\pm$~10~MeV, $\Gamma$~=~104~$\pm$~25~MeV and for the $f_0(1710)$ M~=~1730~$\pm$~15~MeV, $\Gamma$~=~100~$\pm$~25~MeV parameters which are consistent with the PDG~\cite{PDG98} values for these resonances. \par The $D_0^-$-wave shows peaks in the 1.3 and 1.5~GeV regions, presumably due to the $f_2(1270)/a_2(1320)$ and $f_2^\prime(1525)$ and a wide structure above 2 GeV. There is no evidence for any significant structure in the D-wave in the region of the $f_J(1710)$. In addition, there are no statistically significant structures in any of the other waves. A fit has been performed to the $D_0^-$ wave above 1.2~GeV using three incoherent relativistic spin 2 Breit-Wigners to describe the $f_2(1270)/a_2(1320)$, $f_2^\prime(1525)$ and the peak at 2.2 GeV and a background of the form described above. The resulting fit is shown in fig.~\ref{fi:1} and gives for the $f_2(1270)/a_2(1320)$ M~=~1305~$\pm$~20~MeV, $\Gamma$~=~132~$\pm$~25~MeV, for the $f_2^\prime(1525)$ M~=~1515~$\pm$~15~MeV, $\Gamma$~=~70~$\pm$~25~MeV and for the $f_2(2150)$ M~=~2130~$\pm$~35~MeV, $\Gamma$~=~270~$\pm$~50~MeV. \section{A partial wave analysis of the $\pi \pi$ system} \begin{figure}[h] \vspace{5.0cm} \begin{center} \special{psfile=pipitalk9.eps voffset=-160 hoffset=0 hscale=80 vscale=60 angle=0} \end{center} \caption{ a), b), c) The $S_0^-$ wave resulting from a partial wave analysis of the $\pi^+\pi^-$ system.} \label{fi:2} \end{figure} \par The isolation of the reaction \begin{equation} pp \rightarrow p_{f} (\pi^+ \pi^-) p_{s} \label{eq:c} \end{equation} has been described in detail in a previous publication~\cite{re:pipipap}. The resulting centrally produced \mbox{$\pi^+\pi^-$ } system consists of 2.87 million events. A PWA of the centrally produced \mbox{$\pi^+\pi^-$ } system has been performed, using the reflectivity basis~\cite{chung}, in 20~MeV intervals of the \mbox{$\pi^+\pi^-$ } mass spectrum using an event-by-event maximum likelihood method~\cite{re:pipipap}. The $S_0^-$-wave from the physical solution is shown in fig.~\ref{fi:2}. and shows a clear threshold enhancement followed by a sharp drop at 1~GeV. \par In order to obtain a satisfactory fit to the $S_0^-$ wave from threshold to 2~GeV it has been found to be necessary to use three interfering Breit-Wigners to describe the $f_0(980)$, $f_0(1300)$ and $f_0(1500)$ and a background of the form $a(m-m_{th})^{b}exp(-cm-dm^{2})$, where $m$ is the \mbox{$\pi^+\pi^-$ } mass, $m_{th}$ is the \mbox{$\pi^+\pi^-$ } threshold mass and a, b, c, d are fit parameters. The fit is shown in fig.~\ref{fi:2}a) for the entire mass range and in fig.~\ref{fi:2}b) for masses above 1 GeV. The resulting parameters are for the $f_0(980)$ M~=~982~$\pm$~3~MeV, $\Gamma$~=~80~$\pm$~10~MeV, for the $f_0(1300)$ M~=~1308~$\pm$~10~MeV, $\Gamma$~=~222~$\pm$~20~MeV and for the $f_0(1500)$ M~=~1502~$\pm$~10~MeV, $\Gamma$~=~131~$\pm$~15~MeV which are consistent with the PDG~\cite{PDG98} values for these resonances. As can be seen, the fit describes the data well for masses below 1~GeV. It was not possible to describe the data above 1~GeV without the addition of both the $f_0(1300)$ and $f_0(1500)$ resonances. However, even with this fit using three Breit-Wigners it can be seen that the fit does not describe well the 1.7 GeV region. This could be due to a \mbox{$\pi^+\pi^-$ } decay mode of the $f_J(1710)$ with J~=~0. Including a fourth Breit-Wigner in this mass region decreases the $\chi^2$ from 256 to 203 and yields for the $f_J(1710)$ M~=~1750~$\pm$~20~MeV and $\Gamma$~=~160~$\pm$~30~MeV parameters which are consistent with the PDG~\cite{PDG98} values for the $f_J(1710)$. The fit is shown in fig.~\ref{fi:2}c) for masses above 1 GeV. \section{A Glueball-$q \overline q$ filter in central production ?} The WA102 experiment studies mesons produced in double exchange processes. However, even in the case of pure DPE the exchanged particles still have to couple to a final state meson. The coupling of the two exchanged particles can either be by gluon exchange or quark exchange. Assuming the Pomeron is a colour singlet gluonic system if a gluon is exchanged then a gluonic state is produced, whereas if a quark is exchanged then a $q \overline q $ state is produced~\cite{closeak}. In order to describe the data in terms of a physical model, Close and Kirk~\cite{closeak}, have proposed that the data be analysed in terms of the difference in transverse momentum ($dP_T$) between the particles exchanged from the fast and slow vertices. The idea being that for small differences in transverse momentum between the two exchanged particles an enhancement in the production of glueballs relative to $q \overline q$ states may occur. \par The ratio of the number of events for $dP_T$ $<$ 0.2 GeV to the number of events for $dP_T$ $>$ 0.5 GeV for each resonance considered has been calculated~\cite{memoriam}. It has been observed that all the undisputed $q \overline q$ states which can be produced in DPE, namely those with positive G parity and $I=0$, have a very small value for this ratio ($\leq 0.1$). Some of the states with $I=1$ or G parity negative, which can not be produced by DPE, have a slightly higher value ($\approx 0.25$). However, all of these states are suppressed relative to the the glueball candidates the $f_0(1500)$, $f_J(1710)$, and $f_2(1930)$, together with the enigmatic $f_0(980)$, which have a large value for this ratio. \begin{figure} \vspace{9.0cm} \begin{center} \special{psfile=phiang.eps voffset=-40 hoffset=60 hscale=50 vscale=50 angle=0} \end{center} \caption{The azimuthal angle between the fast and slow protons ($\phi$) for various final states. } \label{fi:phidep} \end{figure} \section{The azimuthal angle between the outgoing protons} \par The azimuthal angle ($\phi$) is defined as the angle between the $p_T$ vectors of the two protons. Naively it may be expected that this angle would be flat irrespective of the resonances produced. Fig.~\ref{fi:phidep} shows the $\phi$ dependence for two $J^{PC}$~=~$0^{-+}$ final states (the $\eta$ and $\eta^\prime$), two $J^{PC}$~=~$1^{++}$ final states (the $f_1(1285)$ and $f_1(1420)$) and two $J^{PC}$~=~$2^{++}$ final states (the $\phi \phi$ and $K^*(892) \overline K^*(892)$ systems). The $\phi$ dependence is clearly not flat and considerable variation is observed between final states with different $J^{PC}$s. \section{Summary} \par In conclusion, a partial wave analysis of the centrally produced $K \overline K$ system has been performed. The striking feature is the observation of peaks in the $S_0^-$-wave corresponding to the $f_0(1500)$ and $f_J(1710)$ with J~=~0. In addition, a partial wave analysis of a high statistics sample of centrally produced \mbox{$\pi^+\pi^-$ } events shows that the $S_0^-$-wave is composed of a broad enhancement at threshold, a sharp drop at 1 GeV due to the interference between the $f_0(980)$ and the S-wave background, the $f_0(1300)$, the $f_0(1500)$ and the $f_J(1710)$ with J~=~0. \par A study of centrally produced pp interactions show that there is the possibility of a glueball-$q \overline q$ filter mechanism ($dP_T$). All the undisputed $q \overline q $ states are observed to be suppressed at small $dP_T$, but the glueball candidates $f_0(1500)$, $f_J(1710)$, and $f_2(1930)$ , together with the enigmatic $f_0(980)$, survive. In addition, the production cross section for different resonances depends strongly on the azimuthal angle between the two outgoing protons. \section*{References}
\section{Cosmic topology} This workshop is on observational cosmology: how observations confront cosmological theory. Unfortunately, one of the fundamental aspects of Friedmann-Lema\^{\i}tre models of the Universe is weak in theoretical predictions. General relativity says nothing about how big the Universe should be. It describes curvature, which divides up constant curvature 3-manifolds (``spaces'') into three classes corresponding to the three possible signs of curvature. For example, a canonical flat multiply connected model is the hypertorus, $T^3,$ which can be thought of as a cube whose opposite faces are identified. This is a flat 3-manifold without any edges or boundaries, but finite in volume. A $T^3$ universe may be as small as 1{~$h^{-1}$Gpc} or as big as the horizon for the same values of $\Omega_0,$ $\lambda_0,$ $\Omega_b,$ $\sigma_8$ and $H_0$.\footnote{These parameters are defined as usual. The first two correspond to $\Omega_{m}$ and $\Omega_\Lambda$ in the popular Peebles \protect\cite{Peeb93} notation.} Evolution in the luminosity functions of galaxies and quasars, the star formation rate history of the Universe, and similar observational quantities do not distinguish between the different models. They do not constrain the size of the Universe. Although the ``curvature radius'' and $H_0$ have strong effects on the size of the {\em observable} Universe, i.e. on the horizon radius, they only have weak effects on the size of the Universe itself. How can the theory (that spatial sections are 3-manifolds) be confronted with observations? In short, by photons travelling many times across the Universe so that multiple topological images are seen of single objects. In a multiply connected universe, objects (or regions of CMB plasma) would be seen several times in different directions and (in general) at different redshifts. This would be something like gravitational lensing, except that the whole Universe would be the lens and the angular and radial distance differences in multiply imaged objects would be, in general, big fractions of $\pi$ and of the horizon radius respectively, as opposed to arcsecond and sub-parsec differences in the case of gravitational lensing. \section{Recommended reading} Recent reviews of the different observational strategies include \cite{RB98,LR99} (the latter also includes a brief historical and mathematical background). A fuller review including theoretical aspects of cosmic topology and pre-1993 observational work is that of \cite{LaLu95}, but due to exponential growth in the subject, the number of published articles on the subject has roughly doubled since then. Proceedings of the 1997 Cleveland and 1998 Paris workshops on cosmic topology are available as \cite{Stark98} and \cite{BR99} respectively. Mathematical tools, particularly including a ``census'' of a few thousand small compact hyperbolic 3-manifolds are available at \cite{WeeksSP}. \section{A survival kit for the observer: jargon} The minimum concepts and jargon that the workshop participant or reader should retain from the above literature are probably: \begin{list}{(\roman{enumi})}{\usecounter{enumi}} \item {\em ``compact''} essentially means finite in spatial volume \item to avoid confusion, the word ``open'' is dropped in favour of {\em ``hyperbolic'', ``negatively curved'', ``$\Omega_0 < 1$''} or {\em ``$k < 0$''}; and ``closed'' is dropped in favour of {\em ``elliptic'', ``spherical'', ``positively curved'', ``$\Omega_0 > 1$''} or {\em ``$k > 0$} (otherwise, compact hyperbolic models would be referred to as closed open models \dots) \item {\em ``geodesic''} generally means a geodesic in 3-space, but at times is used to mean a geodesic in 3+1 space-time \item the entire ({\em comoving} spatial section of the) Universe can be represented as a polyhedron embedded in $H^3,$ $R^3$ or $S^3$ (for $k < 0, =0, > 0$ respectively) of which faces are identified with one another in some way --- this is the {\em ``fundamental polyhedron''} or {\em ``Dirichlet domain''} \item by pasting together copies of the fundamental domain, an space $H^3,$ $R^3$ or $S^3$ (respectively) can be constructed which corresponds, for the observer, to the apparent space in which objects at high redshift are located under the hypothesis of trivial topology\footnote{``Trivial topology'' refers here to the property of having a trivial $\pi_1$ homotopy group.} --- this is termed the {\em ``universal covering space'',} $\widetilde{M}$ \item in the covering space, the isometries mapping multiple {\em ``topological images''} (or {\em ``topological clones''}) to one another form a group, $\Gamma$, whose elements are linear combinations of a set of {\em ``generators''} \item the 3-manifold can formally be written as $M=\widetilde{M}/\Gamma$ \item for convenience, one often swaps thinking and calculating between the fundamental polyhedron and the covering space. \end{list} \section{An example of a candidate 3-manifold} In the commonly studied case of the rectilinear toroidal models, multiple topological images of an object form a rectilinear grid in comoving space. Among a small selection of the brightest known galaxy clusters, three form a right angle of equal arm lengths to within $2-3\deg$ and $1\%$ accuracy respectively \cite{RE97}. Are the Coma cluster, RX~J1347.5-1145 and CL~09104+4109 three images of a single cluster or is the right angle just a coincidence? A list of arguments for and against this $T^2$ candidate is provided in the discussion section of \cite{RBa99}, and a comparison with COBE data is presented in \cite{Rouk99}. \section{Projects} This is a workshop. The following are ideas suggested for projects. \subsection{Theory} \begin{list}{(\roman{enumi})}{\usecounter{enumi}} \item What should the topology of the Universe be? {\em Can a theory of quantum gravity or of quantum cosmology make any serious predictions about what the topology of the Universe should be at $t\sim t_0$? } \item A group $\Gamma$ relates the covering space $\widetilde{M}$ of a multiply connected universe to the fundamental polyhedron $M=\widetilde{m}/\Gamma.$ The standard model of particle physics relates different particles to one another by a group, e.g. SU(2)$_L \times$ U(1)$_{YW} \times$ SU(3)$^C$. {\em Could the Universe be considered a particle at the quantum epoch and the spatial transformations of $\Gamma$ be related to the gauge bosons?} \end{list} \subsection{Observation} \subsubsection{Methods} \begin{list}{(\roman{enumi})}{\usecounter{enumi}} \addtocounter{enumi}{2} \item The classical magnitude-redshift relation yielded only weak constraints on the curvature parameters ($\Omega_0,$ $\lambda_0$) until an empirical way of improving supernovae of type Ia as standard candles was devised. The results are impressive, even though theoretical understanding of the method of sharpening the standard candle is weak \cite{SCP9812}.\footnote{Cosmic topology could provide high precision estimates of the curvature parameters. Detection of 5-10 multiple topological images of an object up to $z \sim 2-3$ would be sufficient to estimate $\Omega_0$ and $\lambda_0$ to better than 1\% and 10\% respectively \protect\cite{RL99}.} {\em Could some sort of similar ``trick'' improve the presently published methods to the point of extracting a significant topological detection?} \item {\em Realistic simulations including all the observational difficulties could be used to optimise the cosmic crystallography \cite{LLL96,LLU98,ULL99a} and local isometry search methods \cite{Rouk96,ULL99a}.} \item {\em Realistic simulations and analysis should also be used to find the best way to apply the matched circles principle \cite{Corn96,Corn98b,Weeks98,Rouk99}.} \end{list} \subsubsection{Candidates} \begin{list}{(\roman{enumi})}{\usecounter{enumi}} \addtocounter{enumi}{5} \item {\em Generate specific candidates.} \item {\em Observationally refute these in order to understand systematic errors.} \end{list} \subsubsection{New catalogues} \begin{list}{(\roman{enumi})}{\usecounter{enumi}} \addtocounter{enumi}{7} \item radio: GMRT --- $5\,\lower.6ex\hbox{$\buildrel <\over \sim$} \, z \,\lower.6ex\hbox{$\buildrel <\over \sim$} \, 10$ ? proto-clusters \item mm/sub-mm: MMA/LSA --- $5\,\lower.6ex\hbox{$\buildrel <\over \sim$} \, z \,\lower.6ex\hbox{$\buildrel <\over \sim$} \, 10$ ? galaxies \item cm: MAP, Planck ---$ z \approx 1100$ or \dots (integrated Sachs-Wolfe effect) $z \ll 1100$ CMB (plasma); $z \sim 1 - 3$ clusters (SZ effect) \item optical: SDSS, VLT --- $ z \sim 1 - 3$ ? quasars, galaxies \item Xray: XMM --- $z \sim 1 - 2$ clusters, quasars \end{list} \subsubsection{Local (10kpc $-$ 100Mpc)} \begin{list}{(\roman{enumi})}{\usecounter{enumi}} \addtocounter{enumi}{12} \item {\em Understand the Galaxy (or the local unit of large scale structure) well enough to say what its topological image must have looked like at $z\sim 2-5$ and from an ``arbitrary'' angle.} This would be a ``safe'' theme for a thesis project, since the theoretical and/or observational work done in understanding the Galaxy would be valid independently of its use in identifying or refuting high redshift topological images of the Galaxy. \end{list}
\section{\bf Introduction} QCD has a comparatively "simple" basic Lagrangian but offers an enormous variety of physics phenomena. They extend from asymptotically free quarks and gluons at short space-time distances to complex hadronic excitations at larger scales, built on a highly non-trivial vacuum which hosts strong quark and gluon condensates. Given the genuine non-perturbative features of QCD, namely confinement and spontaneous chiral symmetry breaking, the persistent challenge is to identify and explore the active (effective) degrees of freedom at each different scale. The present talk summarizes several topics and recent results related to these issues. We start with an instructive little exercise, elaborating quark and gluon distributions of nucleons and nuclei in coordinate rather than momentum space. Such a description gives a simple geometrical interpretation of the mechanisms at work in deep-inelastic lepton scattering. Turning to lower $Q^2$ we outline the smooth transition from partonic to hadronic degrees of freedom by the example of generalized ($Q^2$-dependent) electromagnetic polarizabilities of the nucleon. Next we move to low-energy QCD with strange quarks. We discuss recent developments and results of Chiral $SU(3)$ Dynamics, our non-perturbative coupled channel approach based on the chiral low-energy effective Lagrangian. In the final part of this survey, the QCD Sum Rule method is unified with aspects of spontaneous chiral symmetry breaking in an attempt to establish model-independent constraints for vector meson spectra, both in vacuum and in nuclear matter. \section{\bf Deep-inelastic lepton scattering in coordinate space} Nucleon structure functions are commonly analyzed in momentum space. In coordinate space, quark and gluon distributions are defined as correlation functions involving two field operators separated by a light-cone distance $y^+ = t + z = 2l$. In deep-inelastic scattering as viewed in the laboratory frame with the target nucleon or nucleus at rest, the longitudinal distance $y^+ /2$ entering in the parton correlation function can be compared with length scales characteristic of nucleons and nuclei and offers new insights into the nature of parton distributions and their interpretation. The quark and gluon distributions are expressed in terms of the squared four-momentum transfer $Q^2$ and the Bjorken variable $x = Q^2 /2 M \nu$, where $M$ is the nucleon mass and $\nu$ is the energy transfer in the lab frame. In the Bjorken limit the dominant contributions to the structure functions at small Bjorken-$x$ come from the light-like separations of order $y^+ \sim 1/Mx$. Consequently, large longitudinal distances $l = y^+/2 = (2 Mx)^{-1}$ are important in the scattering process at small $x$. The space-time pattern of deep-inelastic scattering is then as follows: the virtual photon interacts with partons which propagate a distance $y^+$ along the light cone. The characteristic lab frame correlation length $l$ is half of that distance. These features are naturally implemented in coordinate space (so-called Ioffe time) distribution functions \cite{1}. In accordance with their properties under charge conjugation one introduces these coordinate space distributions as one-dimensional Fourier sine and cosine transforms of the momentum space quark and gluon distributions: \begin{eqnarray} {\cal Q} (y^+, Q^2) & = & \int_0^1 dx \sin (\frac{My^+}{2} x) [q (x, Q^2) + \bar{q} (x, Q^2)],\\ {\cal Q}_{valence} (y^+, Q^2) & = & \int_0^1 dx \cos (\frac{My^+}{2} x) [q (x, Q^2) - \bar{q}(x, Q^2)],\\ {\cal G}(y^+, Q^2) & = & \int_0^1 dx \cos (\frac{My^+}{2}x) x g (x, Q^2). \end{eqnarray} \noindent Consider first the coordinate space parton distributions of a free nucleon. We start from realistic input distributions using the CTEQ4L parametrization \cite{2} at $Q_0^2 = 1.4 \; GeV^2$, perform the QCD (DGLAP) evolution to $Q^2 = 4 \; GeV^2$ and then take the Fourier transforms (1--3) to translate the distributions into coordinate space. The result \cite{3} is shown in Fig. 1. Note that ${\cal Q} (y^+, Q^2)$ and ${\cal G} (y^+, Q^2)$ extend to distances far beyond the diameter of the nucleon. In fact they grow continuously, reflecting the strong rise of the structure function $F_2 (x, Q^2)$ observed at $x < 10^{-3}$. Even the valence quark distribution has a pronounced tail extending beyond the nucleon size. The interpretation in the lab frame is simple: at very small Bjorken-$x$ corresponding to large longitudinal distances the virtual photon converts into a beam of partons which propagate along the light cone and interact with partons of the target nucleon, probing its sea quark and gluon content. This "beam" stretches over length scales much larger than the size of the nucleon itself. The partonic composition of that beam is dominated by a steadily growing number of gluons at larger distances. \begin{figure}[h] \centerline{\input{cteq.tex}} \caption{ Coordinate space quark and gluon distributions at $Q^2 = 4 \: GeV^2$. Dashed curve: Valence quarks; dotted curve: total quark and antiquark distribution; solid curve: gluons. For details see ref. \cite{3}.} \end{figure} It is instructive to examine parton distributions of nuclei using the same picture. A detailed momentum space analysis of these distributions has been performed in refs. \cite{4}, combining data from deep-inelastic lepton-nucleus scattering and Drell-Yan lepton pair production in proton-nucleus collisions. Based on these data the quark and gluon content of the nuclear distributions was extracted using a DGLAP evolution analysis. One can then take the ratio of Fourier transforms of the nuclear and free nucleon distributions: \begin{equation} {\cal R}(y^+, Q^2) = \frac{{\cal Q}^{A} (y^+, Q^2)}{{\cal Q}^N (y^+, Q^2)} \end{equation} for quark distributions, and analogous ratios for gluon distributions. Nuclear effects can now be analysed in coordinate space. The result \cite{3} shown inFig.~2 for $^{40}Ca$ clearly demonstrates the most prominent features. Effects of binding and Fermi motion which modify the structure functions at longitudinal distances $l = y^+ /2$ smaller than the nucleon diameter $(l<2 fm)$ are evidently marginal. The leading effect is shadowing, the reduction of the ratios ${\cal R}$ substantially below one, due to coherent multiple scattering of the parton "beam" from at least two nucleons in the target nucleus. This effect starts as soon as the propagation length of quark and gluon fluctuations of the virtual photon exceeds the average distance between nucleons in the nucleus $(l > d \simeq 2 fm)$. The interesting feature is again the prominent role of gluons in this process. The measured shadowing effect is represented by the ratio ${\cal R}_{F_2}$ of the corresponding $F_2$ structure functions for nuclei and free nucleons. This ratio reflects the shadowing effect on the sum of quark and antiquark distributions seen directly by the virtual photon. The indirect effect of gluon shadowing, expressed in terms of the ratio ${\cal R}_g = {\cal G}^A (y^+, Q^2)/ {\cal G}^N (y^+, Q^2)$, is obviously very strong. While this ratio is not directly observable, it certainly indicates a large effective cross section for gluons interacting with nucleons. \begin{figure}[h] \centerline{\input{calcium.tex}} \caption{ Ratios of nuclear $(^{40}Ca)$ and free nucleon coordinate space distributions for gluons (solid), valence quarks (dashed) and for the structure functions $F_2^{A, N}$ (dotted) at $Q^2 = 4 \: GeV^2$. For further details see ref. \cite{3}.} \end{figure} \section{Partons versus hadrons: generalized nucleon polarizabilities} At intermediate $Q^2\lsim 1 GeV^2$ a detailed analysis of virtual Compton scattering on the nucleon should provide insights into the transition from partonic to hadronic degrees of freedom in the nucleon. Interesting quantities to look at in this context are the generalized ($Q^2$-dependent) electromagnetic polarizabilities. We investigate: \begin{enumerate} \item[a)] the sum of electric and magnetic polarizabilities \begin{equation} \Sigma (Q^2) = (\alpha + \beta)_{Q^2} = \frac{1}{2 \pi^2} \int_{\omega_0}^\infty \frac{d \omega}{\omega^2} \sigma_T (\omega, Q^2), \end{equation} where $\sigma_T (\omega, Q^2)$ denotes the total cross section for scattering of a transverse virtual photon at fixed $Q^2 = \vec{q}\: ^2 - \omega^2 > 0$ from a nucleon, and the integration is taken over the photon energy $\omega$ from (pion-production) threshold to infinity. The $\omega^{-2}$ weighting in this polarizability integral focuses on the low energy part of the excitation spectrum, but probed over a certain range of "resolution" $Q^2$; \item[b)] the spin polarizability \begin{equation} \gamma (Q^2) = \frac{1}{4 \pi^2} \int_{\omega_0}^{\infty} \frac{d \omega}{\omega^3} [\sigma_{1/2} (\omega, Q^2) - \sigma_{3/2} (\omega, Q^2)] \end{equation} which involves the difference of helicity $1/2$ and $3/2$ photon-nucleon cross sections, accessible by polarized electroproduction measurements on the nucleon at fixed $Q^2$. This polarizability has a characteristic $\omega^{-3}$ weighting under the integral. We recall that the corresponding integral with $\omega^{-1}$ gives the Gerasimov-Drell-Hearn sum rule. \end{enumerate} In the limit of large $Q^2$, the polarizabilities $\Sigma (Q^2)$ and $\gamma (Q^2)$ turn into certain moments of the structure functions $F_1$ and $g_1$ measured in unpolarized and polarized deep-inelastic lepton-nucleon scattering: \begin{eqnarray} \Sigma (Q^2) \longrightarrow \frac{2e^2M}{\pi Q^4} \int_0^1 dx \frac{x}{1-x} F_1(x),\\ \gamma (Q^2) \longrightarrow \frac{4e^2 M^2}{\pi Q^6} \int_0^1 dx \frac{x^2}{1-x} g_1 (x). \end{eqnarray} \begin{figure}[bh] \vspace*{-1cm} \centerline{\epsfig{figure=figure3.ps,width=100mm}} \vspace*{-1cm} \caption{Generalized proton electromagnetic polarizability $\Sigma_{\rho} (Q^2)$ (see text and ref. \cite{5}). The calculated full line includes relativistic pion loops and the $\Delta (1232)$ resonance. Data (dashed) from inelastic electron-proton scattering. Dash-dotted curve: partonic description using the structure function $F_1$.} \vspace*{-.3cm} \end{figure} The interesting question is then the following: how does the QCD description of $\Sigma (Q^2)$ and $\gamma (Q^2)$ in terms of quarks and gluons at large $Q^2$ turn into the hadronic low $Q^2$ description of the same quantities? At low $Q^2$, QCD with light (u-, d- and s-) quarks translates into a chiral effective Lagrangian of Goldstone bosons (pions, kaons,...) coupled to baryons and vector mesons. This effective field theory has been used \cite{5} at the level of one-loop chiral perturbation theory in the pion-nucleon sector to evaluate the nucleon polarizabilities $\Sigma (Q^2)$ and $\gamma (Q^2)$ for $Q^2 \stackrel{<}{\sim} 0.5\: GeV^2$. This framework emphasizes the role of the pion cloud and the $N \to \Delta$ excitation of the nucleon in response to the electromagnetic field. At high $Q^2$, on the other hand, nucleon structure is realised in terms of its parton (quark and gluon) content. We have explored whether low-$Q^2$ chiral dynamics matches high-$Q^2$ partonic structure in $\Sigma (Q^2)$ and $\gamma (Q^2)$. The results \cite{5} are presented in Figs. 3, 4. Let us first discuss the $Q^2$-dependent electromagnetic polarizability $\Sigma_{p}$ of the proton (Fig. 3). The chiral dynamics calculation (solid line) summarizes the response of the pion cloud and through resonance (primarily $\Delta (1232)$) excitation and compares very well with the data (dashed curve) derived from inelastic electron scattering in the range $Q^2 \stackrel{<}{\sim} 0.5\: GeV^2$. The downward extrapolation of the parton distribution $F_1$ from deep-inelastic scattering (dash-dotted curve) determines $\Sigma (Q^2)$ at large $Q^2$. Both descriptions evidently meet at intermediate $Q^2$, indicating a smooth crossover from partonic to hadronic degrees of freedom. The result for the $Q^2$-dependent spin polarizability has similar features (Fig.~4). On the low-$Q^2$ side we see the well-known balance between the diamagnetic response of the pion cloud and the spin-paramagnetic effect of the $N \to \Delta$ (spin $1/2$-to-spin $3/2$) transition. Dia- and paramagnetism enter with opposite signs and comparable magnitudes, so that the resulting $\gamma (Q^2)$ is small and changes sign at $Q^2 \simeq 0.4 \: GeV^2$. The downward extrapolation from high $Q^2$ can be made using the spin structure function $g_1$ from polarized deep-inelastic scattering. Again it matches the low-$Q^2$ hadronic calculation remarkably well. \begin{figure}[tbh] \centerline{\epsfig{figure=figure4.ps,width=100mm}} \caption{generalized proton spin polarizability $\gamma (Q^2)$ as calculated in ref. \cite{5}. Pion loop (diamagnetic) and $\Delta$ resonance (paramagnetic) contributions are shown separately. Their sum (solid curve) is compared with the partonic prediction using the spin structure function $g_1$.} \end{figure} \section{Low-energy QCD with strange quarks: Chiral $SU(3)$ Dynamics} QCD with massless u-, d- and s-quarks has a chiral $SU(3)_L \times SU(3)_R$ symmetry. As a consequence of strong QCD forces this symmetry is spontaneously broken. It is also explicitly broken by the quark masses. The symmetry breaking pattern is manifest in the low-energy hadron spectrum, with the light pseudoscalar mesons representing the Goldstone bosons of the spontaneously broken symmetry. The strange quark is special since its mass is intermediate between "light" and "heavy". One of the key questions is then the following: to what extent do the {\it symmetries} of QCD govern strong interaction {\it dynamics}, and what is the role of the strange quark in this context? Progress has recently been made in developing a framework to deal with these problems. The starting point is the chiral effective Lagrangian, a theory of Goldstone bosons (pions, kaons and eta mesons) coupled to the octet of baryons. It is designed according to the rules set by QCD symmetries \cite{6}. This effective Lagrangian introduces a characteristic scale, $4 \pi f_{\pi} \sim 1\: GeV$, where $f_{\pi} = 92.4\: MeV$ is the pion decay constant. Spontaneously broken chiral symmetry also implies that the low-energy interactions of Goldstone bosons are weak: the leading behaviour of any amplitude for scattering of a pseudoscalar meson on a baryon goes like \begin{equation} T_{meson - baryon} = const \cdot \frac{E}{f_{\pi}^2} + ..., \end{equation} \noindent where $E$ is the meson energy in the center-of-mass system. The constant in eq.~(9) is specific for each meson-baryon channel and completely determined by $SU(3)$ symmetry. Pions close to threshold with $E \simeq m_{\pi}$ have small scattering amplitudes. In this case chiral perturbation theory (ChPT), the systematic expansion of observables in powers of energy of momentum (or $m_{\pi}$) is a useful concept. When strange quarks are involved, the scattering amplitudes scale as $E/f^2_{\pi}$ with $E \geq m_K$, and the driving terms in (9) become sizable. One therefore expects that ChPT is only of limited value once strangeness is included. A way to proceed which proves to be quite successful is chiral $SU(3)$ dynamics \cite{7}: a non-perturbative coupled-channel approach based on the chiral $SU(3) \times SU(3)$ effective meson-baryon Lagrangian. The basic strategy is first to generate Born terms $T^{(0)}_{ij}$ of the multi-channel meson-baryon T-matrix from the chiral effective Lagrangian and then to perform a partial loop summation to all orders using a Lippmann-Schwinger equation: \begin{equation} T = [1 - T^{(0)} \cdot G]^{-1} T^{(0)}, \end{equation} \noindent with an appropriate Green function $G$. Introducing a limited number of finite-range parameters, a remarkably good description of a large amount of cross section can be achieved. This method has been successfully applied \cite{7} to $KN$ scattering, the coupled $ \{\bar{K}N, \pi Y, \eta Y\}$ and $\{ \pi N, KY, \eta N \}$ multi-channel systems, and to the photoproduction of $\eta$ and $K$ mesons. Simplified but otherwise very similar calculations are reported elsewhere at this conference \cite{8}. While all previous calculations have dealt with $s$-wave dynamics, the present focus is on the systematic incorporation of $p$-waves as required by a multitude of measured meson-baryon angular distributions and polarization observables. We have now reached the stage where the complete $s$- and $p$-wave chiral $SU(3)$ dynamics is well under control \cite{9}. As an example Fig.~5 shows predicted cross sections for kaon photoproduction once a limited set of parameters has been fixed to reproduce a large variety of $\pi N \to \eta N$, $K \Lambda$ and $K \Sigma$ cross sections and angular distributions. The $l = 1$ partial waves are evidently important in such channels immediately above threshold. With the inclusion of $p$-waves, further detailed tests of the chiral $SU(3)$ effective Lagrangian are now possible. A forthcoming necessary step is to incorporate the axial $U (1)$ anomaly and the dynamics of the $\eta '$. This is where the gluonic sector of QCD should have its impact. \begin{figure}[h] \centerline{\epsfig{figure=tx.8.eps,width=50mm,angle=-90} \epsfig{figure=tx.7.eps,width=50mm,angle=-90}} \caption{Kaon photoproduction cross sections using s- and p-wave chiral $SU(3)$ dynamics \cite{9}, a non-perturbative coupled channel approach based on the chiral meson-baryon Lagrangian \cite{7}. The data are taken from ref. \cite{10}.} \end{figure} \section{QCD Sum Rules: some recent developments} The QCD sum rule method of treating the non-perturbative dynamics of QCD was developed twenty years ago (SVZ sum rules \cite{11}). It connects the expansion of a correlation function in terms of vacuum condensates (the operator product expansion) with the spectrum of this correlation function via dispersion relations. These sum rules have been applied to understand the masses and properties of a variety of hadrons emphasizing their role as excitations of the condensed QCD vacuum. QCD sum rules have also been used in more recent times to arrive at estimates for possible in-medium mass shifts of vector mesons \cite{12}. The validity of such estimates has been under debate, however, for several reasons. Uncertainties exist at the level of factorization assumptions used to approximate four-quark condensates $\langle \bar{q} \Gamma q \bar{q} \Gamma q \rangle$ in terms of $\langle \bar{q} q \rangle^2$, the square of the standard chiral condensate. Furthermore, for broad structures such as the $\rho$ meson, with its large vacuum decay width further magnified by in-medium reactions, the QCD sum rule analysis does not provide a reliable framework to extract a "mass shift" in medium. The situation is more comfortable for the $\omega$ meson which has a vacuum width twenty times smaller than that of the $\rho$ meson and may have a much better chance to survive as a quasi-particle in nuclear matter \cite{13}. We have recently re-examined these questions \cite{14} in search for model-independent sum-rule constraints which do not suffer from the uncertainties introduced by four-quark condensates. We exemplify these constraints for the case of the $\omega$ meson spectral distribution and its changes in the nuclear medium. The starting point is the current-current correlation function \begin{equation} \Pi_{\mu \nu} (q) = i \int d^4 x e^{iq \cdot x} \langle {\cal T}\: j_{\mu} (x) j_{\nu} (0) \rangle = \left( \frac{q_{\mu} q_{\nu}}{q^2} - g_{\mu \nu} \right) \Pi (q^2), \end{equation} and we work with the spectrum \begin{equation} R (s) = \frac{12 \pi}{s} Im \Pi (s). \end{equation} \noindent In vacuum $R (s)$ is directly related to $\sigma (e^+ e^- \to hadrons)$. In dense and hot hadronic matter it enters into the analysis of lepton pair production in high-energy heavy-ion collisions. One proceeds now as follows. First $\Pi (q^2)$ is expanded at large spacelike $q^2$ (i. e. for $Q^2 = -q^2$ positive and large) in powers of $1/Q^2$ using the QCD operator product expansion. At the same time $\Pi (q^2)$ is written in the form of a dispersion relation. Then a Borel transformation is performed which effectively reduces the weight on the uncertain high-energy parts of the spectrum. Let us focus on the $\omega$ meson with the isoscalar current $j_{\mu} = \frac{1}{6} (\bar{u} \gamma_{\mu} u + \bar{d} \gamma_{\mu} d)$ and discuss the vacuum case first. One finds \begin{equation} \int^{\infty}_{0} ds\: R(s)e^{-s / {\cal M}^2} = \frac{1}{6} (1 + \frac{\alpha_s}{\pi}) {\cal M}^2 + \frac{C}{{\cal M}^2} +... \end{equation} \noindent where the Borel mass ${\cal M}$ is a technical scale parameter and $C$ is a combination of quark and gluon condensates: \begin{equation} C = \frac{2 \pi^2}{3} \left( \langle m_u \bar{u} u + m_d \bar{d} d \rangle + \frac{1}{12} \langle \frac{\alpha_s}{\pi} G_{\mu \nu} G^{\mu \nu} \rangle \right) . \end{equation} \noindent A negligibly small quark mass term has been dropped on the r.h.s. of eq. (13). Terms of order ${\cal M}^{-4}$ involve combinations of (uncertain) four-quark condensates. In a nuclear medium the condensate term (14) receives density-dependent corrections, the leading one being proportional to the first moment of the quark distribution in the nucleon. The vacuum spectrum of the $\omega$ meson has the characteristic behaviour shown in Fig. 6: a resonant part below a scale $s_0$ followed by a continuum $R_c (s)$ which approaches the perturbative QCD limit for $s > s_0$: \begin{equation} R_c \simeq \frac{1}{6} (1 + \frac{\alpha_s}{\pi}) \Theta (s - s_0). \end{equation} \noindent Splitting the spectrum into resonance and continuum and choosing ${\cal M} > \sqrt{s_0}$, a term-by-term comparison in eq. (13) gives a set of sum rules for the moments of $R(s)$. The lowest ones (in vacuum) are \begin{eqnarray} \int^{s_0}_0 ds \: R(s) & = & \frac{s_0}{6} (1 + \frac{\alpha_s}{\pi}),\\ \int^{s_0}_0 ds \: s \:R(s) & = & \frac{s^2_0}{12} (1 + \frac{\alpha_s}{\pi})- C, \end{eqnarray} \noindent with the condensate term $C$ of eq. (14). Higher moments of $R(s)$ successively introduce condensates of higher dimensions. \begin{figure}[h] \centerline{\epsfig{figure=Rome2.eps,width=100mm}} \caption{Spectrum R(s) in the $\omega$ meson channel as calculated in ref. \cite{13,14}. The data points refer to $e^+ e^- \to 3 \pi$ and $e^+ e^- \to$ hadrons $(I = 0)$.} \end{figure} Let us now make an attempt to unify QCD sum rules with chiral symmetry and current algebra. Recall that the scale for spontaneous chiral symmetry breaking, $\Delta = 4 \pi\; f_{\pi} \sim 1\: GeV$, is realized as a characteristic gap in the low-mass hadron spectrum. The light vector mesons are the lowest $q \bar{q}$ dipole $(J^{\pi} = 1^-)$ excitations of the QCD vacuum, with their masses located just under the gap $\Delta$. We propose \cite{14} that the scale $\sqrt{s_0}$ which separates the hadronic (resonance) sector from the quark-antiquark continuum in the QCD sum rule analysis, should be identified with the gap $\Delta$, setting \begin{equation} \sqrt{s_0} = \Delta = 4 \pi \;f_{\pi}. \end{equation} \noindent That this makes some sense can be seen instantly by returning to the Vector Meson Dominance (VMD) model for the resonant part of $R (s)$. In this model, taking the zero width limit, we have \begin{equation} R (s) = \frac{4 \pi^2}{3} \frac{m^2_{\omega}}{g^2} \delta (s-m^2_{\omega}) + R_c (s) \end{equation} \noindent with the vector coupling constant $g \simeq 6$. From the sum rule (16) one finds \begin{equation} \frac{8 \pi^2}{g^2} \frac{m^2_{\omega}}{s_0} = 1 + \frac{\alpha_s}{\pi}. \end{equation} \noindent In fact this equation holds both for $\omega$ and $\rho$ meson in the zero width limit, with degenerate masses $m_V = m_{\rho} = m_{\omega}$. Inserting $s_0 = 16 \pi^2 f_{\pi}^2$ one immediately recovers the celebrated current algebra (KSFR) relation \begin{equation} m_V = \sqrt{2} g f_{\pi} , \end{equation} \noindent apart from a small perturbative QCD correction. Note that the second sum rule (17) gives the interesting further constraint $g = 2 \pi$ (up to small $\alpha_s$ and condensate corrections which drive $g$ closer to its empirical value). The in-medium downward shift of the vector meson mass in the zero-width limit found in ref. \cite{12} was about 15 \% at nuclear matter density, $\rho = \rho_0 = 0.17 \: fm^{-3}$, and came primarily through the downward shift of $s_0$ in the QCD sum rule analysis. With the identification (18) in the dipole sum rules (16, 17), this can now be easily understood in terms of the in-medium reduction of the gap $\Delta (\rho) = 4 \pi \: f_{\pi} (\rho) = \sqrt{s_0 (\rho)}$. The pion decay constant (in fact, the one related to the time component of the axial current in hadronic matter) is proportional to the square root of the chiral condensate $\langle \bar{q} q \rangle$. Its leading dependence on baryon density is controlled by the nucleon sigma term $\langle N |m_q \bar{q} q| N \rangle$ which induces an approximate 30 \% reduction of the magnitude of $\langle \bar{q} q \rangle$ at $\rho = \rho_0$. While such relationships are obscured for the $\rho $ meson by its very large in-medium width, they may well be realized for the much narrower $\omega$ meson. In fact explicit calculations \cite{13} of the in-medium $\omega$ meson spectrum using the chiral $SU(3) \times SU(3)$ Lagrangian with inclusion of anomalous couplings from the Wess-Zumino action, turn out to be fully consistent with the QCD sum rule analysis. The suggested in-medium mass shift may even lead to nuclear bound states of $\omega$ mesons \cite{15,16}, an exciting perspective.
\section{Effective Potential Approach } The method of solving the RGE's and the appropriate boundary conditions for the couplings is explained in Ref. 1. In this update, we use the same notation and procedure found in Ref. 1. We also use the following values for $M_{Z}$ and $\alpha_3({M_{Z}})$: $M_{Z} = 91.1867$ GeV and $\alpha_{3}(M_{Z}) = .119$. \section{Bounds on $M_{H}$} We now determine a lower bound on the Higgs boson mass in the SM \cite{5,12}. We first alert the reader to our phenomenologically viable assumption that the physical vacuum corresponds to a global, not merely a local, minimum of the effective potential. This assumption is consistent with our intention to accept the SM as a truly valid theory and compute the consequences, i.e. to zeroth order there is no motivation to consider the physical vacuum to be anything other than the true vacuum. If one considers the possibility that the physical vacuum is a metastable vacuum with a lifetime longer than the age of the universe, that there exist deeper minima of the potential, then the SM lower bounds on the Higgs boson mass become less stringent in general for certain choices of $\Lambda$ and $M_{top}$, where $\Lambda$ is the cutoff beyond which the SM is no longer valid \cite{13}. But, for $M_{top} \sim 177$ GeV and $\Lambda = 10^{19}$ GeV, the SM3 absolute stability lower bound is relaxed by only $\sim O(5)$ GeV when one only imposes metastability requirements, and this small effect only becomes diminished with the inclusion of a fourth generation. We obtain lower limits on the SM Higgs boson mass by requiring stability of this observed vacuum. It is well known that lower values of $\Lambda$ relax the SM lower bounds \cite{15}, but we note that the lower bounds on the SM Higgs boson mass are insensitive to the precise value of $\Lambda$ for large $\Lambda$, i.e. for $10^{11}$ GeV $< \Lambda < 10^{19}$ GeV. Working with the two-loop RGE requires the imposition of one-loop boundary conditions on the running parameters \cite{10}. As pointed out by Casas et al. \cite{5,7}, the necessary condition for vacuum stability is derived from requiring that the effective coupling $\tilde{\lambda}(\mu)>$ 0 rather than $\lambda > 0$ for $\mu(t) < \Lambda$, where $\Lambda$ is the cut-off beyond which the SM is no longer valid. The effective coupling $\tilde{\lambda}$ in the SM4 is defined as: \begin{displaymath} \tilde{\lambda}=\frac{\lambda}{3} -\frac{1}{16 \pi^{2}}\left\{ \sum_{i=1}^{5} 2 \kappa_{i} h_{i}^{4} \left[ \ln \frac{h_{i}^{2}}{2} - 1 \right] \right\} \end{displaymath} where the three generation case is simply the same as the above expression without the fourth generation Yukawa coupling contributions. Choosing $\Lambda = 10^{19}$ GeV and $M_{top} = 172$ GeV, we arrive at a vacuum stability lower bound on $M_{h}$ of $\sim$ 134 GeV for the SM with three generations. Allowing $M_{top}$ to be as large as 179 GeV increases the lower bound on $M_{H}$ to $\sim$ 150 GeV. To compute the MSSM upper bound on $M_{H}$, we assume that all of the sparticles have masses $O(M_{susy})$ or greater and that of the two Higgs isodoublets of the MSSM, one linear combination is massive, also with a mass of $O(M_{susy})$ or greater, while the other linear combination, orthogonal to the first, has a mass of the order of weak-scale symmetry breaking. With these two assumptions, it is clear that below the supersymmetry breaking scale $M_{susy}$, the effective theory is the SM. This fact enables us to use the SM effective potential for the Higgs boson when we treat the lightest Higgs boson in the MSSM. In the MSSM(3,4), the boundary condition for $\lambda$ at $M_{susy}$ is \begin{displaymath} \frac{\lambda}{3}(M_{susy})=\frac{1}{4}\left[g_{1}^{2}(M_{susy})+g_{2}^{2}(M_{susy})\right]\cos^{2}(2\beta)+\frac{\kappa_{i}h_{i}^{4}(M_{susy})}{16 \pi^{2}}\left( 2\frac{X_{i}}{M_{susy}^{2}}-\frac{X_{i}^{4}}{6 M_{susy}^{4}} \right) \end{displaymath} where $\kappa_{i}$ = 3 for $i = (t,T,B)$ and $\kappa_{i}$ = 1 for $i = (N,E)$ and $X_{i}$ is the supersymmetric mixing parameter for the ith fermion. Zero threshold corrections correspond to $X_{i}$ = 0. Maximum threshold corrections occur for $X_{i} = 6 M_{susy}^{2}$. \bigskip \bigskip \vglue -1cm \hglue -2cm \psfig{figure=kang1.eps,height=8cm,angle=-90}\hglue 1cm \psfig{figure=kang2.eps,height=8cm,angle=-90} \hglue 3.5cm (1)\hglue 10cm (2)\\ \begin{quote} \scriptsize Figure 1: The lightest Higgs boson mass $M_{H}$ as a function of $\cos^{2}(2\beta)$. The bottom two curves correspond to MSSM upper bounds with no threshold corrections, for $M_{top}$ = 172 GeV and 179 GeV, respectively. The two upper curves correspond to MSSM upper bounds with maximum threshold corrections, for $M_{top}$ = 172 GeV and 179 GeV, respectively. The two horizontal lines are the $\cos^{2}(2\beta)$-independent SM3 vacuum stability bounds. The lower horizontal line corresponds to $M_{top}$ = 172 GeV, while the other horizontal line was computed with $M_{top}$ = 179 GeV. Figure 2: Same as Figure 1, but now the MSSM bounds correspond to the minimal threshold corrections consistent with the experimental lower limit on $M_{H}$ \end{quote} \normalsize In Fig. (1) we present our numerical two-loop results for the lightest Higgs boson mass bounds in the SM and the MSSM3 as a function of the supersymmetric parameter $\cos^{2}(2\beta)$. The bottom two curves correspond to the MSSM3 upper bound for the two cases $M_{top} = 172$ GeV and the slightly greater upper bound that results when $M_{top}$ = 179 GeV and with no threshold corrections. When the case of maximum threshold corrections is considered, these two curves are translated upwards by $\sim$ 55 GeV - 60 GeV, illustrating the strong dependence of the upper bound on the precise value of the threshold corrections. Yet even with such a dramatic increase in the upper bounds with increasing threshold corrections, we observe that the SM lower bound exceeds the MSSM upper bound for $M_{top} = 172$ GeV and $ 0 < \cos^{2}(2\beta) < .2$ for all values of the threshold correction contribution. Similarly, for $M_{top} = 179$ GeV, the troublesome situation is only exacerbated, as the SM lower bound exceeds the MSSM upper bound for $0 < \cos^{2}(2\beta) < .38$ independent of the threshold corrections. In Fig.(2) we present the problem more clearly. Taking into account the present experimental lower limit on $M_{H}$ of $\sim$ 90 GeV at 95$\%$ CL, we find the value of the threshold correction that gives a smallest upper bound consistent with the experimental lower limit. Clearly, for this phenomenologically determined lower limit of the threshold contributions, there is a large area in $M_{H} \times \cos^{2}(2\beta)$ space that is inconsistent with both the SM and the MSSM. For $M_{top} = 172$ GeV, the region 92 GeV $ < M_{H} < $ 134 GeV invalidates both theories independent of $\cos^{2}(2\beta)$, while for $M_{top} = 179$ GeV, the range of mutual invalidiation is 92 GeV $ < M_{H} < $ 150 GeV. \section{Fourth Generation} To resolve the above conundrum, one would like to either raise the MSSM upper bounds, lower the SM lower bounds, or both. Upon adding a fourth generation, the SM4 lower bounds exceed the SM3 lower bounds and are an increasing function of the fourth generation masses. If a Higgs is detected in the region of mutual invalidation of both the SM and the MSSM, consideration of SM4 vacuum stability lower bounds only exacerbates the problem. It is readily apparent that the way out of the area of inconsistency is to consider the MSSM4 and see if the additional matter of the MSSM4 results in MSSM4 upper bounds that exceed the SM3 lower bounds. We now discuss restrictions on the possible fourth generation fermion masses \cite{2,14,15,16}. The close agreement betweeen the direct measurements of the top quark at the Tevatron and its indirect determination from the global fits of precision electroweak data including radiative corrections within the framework of the SM imply that there is no significant violation of the isospin symmetry for the extra generation. Thus the masses of the fourth generation isopartners must be very close to degenerate \cite{15}; i.e. \begin{displaymath} \frac{\|M_{T}^{2}-M_{B}^{2}\|}{M_{Z}^{2}} \lesssim 1, \frac{\|M_{E}^{2}-M_{N}^{2}\|}{M_{Z}^{2}} \lesssim 1 \end{displaymath} Recently, the limit on the masses of the extra neutral and charged lepton masses, $M_{N}$ and $M_{E}$, has been improved by LEP1.5 to $M_{N} > 59$ GeV and $M_{E} > 62$ GeV. Also, CDF has yielded a lower bound on $M_{B}$ of $\sim$ 140 GeV. In our previous work, we considered a completely degenerate fourth generation of fermions with mass $m_{4}$. We derived an upper bound on $m_{4}$ in the MSSM4 by demanding pertubative validity of all the couplings out to the GUT scale \cite{17}. This constraint led to an upper bound on $m_{4}$ of $\sim$ 110 GeV. The above experimental lower limit on $M_{B}$ naturally forces us to now a consider a fourth generation where degeneracy only holds among the isodoublets seperately. We therefore consider a fourth generation with masses $M_{L}$ and $M_{Q}$. In Fig.(3), we present the SM lower bound, the MSSM4 upper bound with the fourth generation masses at their experimental lower limits and with fourth generation masses large enough to remove the problem area for all values of $\cos^{2}(2\beta)$. The MSSM bounds were calculated with no threshold corrections, and $M_{top}$ is fixed at 172 GeV. Fig.(4) shows the same information for $M_{top}$ = 179 GeV. The MSSM4 upper bounds are much more sensitive to $M_{Q}$ than they are to $M_{L}$. This qualitative behaviour is readily understood from inspection of the equation for $m_{\phi}^{2}$. For this reason, it is necessary to increase $M_{Q}$ appropriately in order to generate a MSSM4 upper bound that is greater than the SM lower bound for all values of $\cos^{2}(2\beta)$. In fact, keeping $M_{Q}$ at 146 GeV and allowing $M_{L}$ to be 110 GeV does not resolve the problem. But increasing both $M_{Q}$ and $M_{L}$ as indicated in the figures does remove the problem. Because all of the bounds increase as $M_{L}$ and $M_{Q}$ increase, and because the upper bounds on $m_{4}$ from the previous work are saturated when the masses of the fourth generation reach some critical values from below, we can conclude that $M_{L}$ must still be $< 110$ GeV. This conclusion follows because it is $h_{N}$ that violates pertubative validity, so in the non-degenerate case, it is $M_{L}$ that must still respect this upper bound if gauge coupling unification is still to be achieved in the MSSM4. \hglue -2cm \psfig{figure=kang3.eps,height=8cm,angle=-90}\hglue 1cm \psfig{figure=kang4.eps,height=8cm,angle=-90} \hglue 3.5cm (3)\hglue 10cm (4)\\ \begin{quote} \scriptsize Figure 3: Plots of the physical Higgs boson mass as a function of $\cos^{2}(2\beta)$. The $\cos^{2}(2\beta)$-independent flat line is the MSSM3 vacuum stability lower bound for $M_{top}$ = 172 GeV. The lower curve is the MSSM4 upper bound for the same value of $M_{top}$, no threshold corrections and the indicated values for $M_{L}$ and $M_{Q}$. Similarly for the upper curve. Figure 4: Same as Figure 3, but with $M_{top}$ = 179 GeV. \end{quote} \normalsize \section{CONCLUDING REMARKS} In conclusion, we have studied the upper bounds on the lightest Higgs boson mass $M_{H}$ in the MSSM with four generations by solving the two-loop RGE's and using the one-loop EP. We find that if the Higgs boson is discovered with a mass $M_{H} < $134 GeV (150 GeV) for $M_{top} =$ 172 GeV (179 GeV), then there is a demand for the introduction of new physics. This mass range for $M_{H}$ will be explored shortly and thus an explanation of what new physics could be consistent with such a Higgs mass measurement is desirable. We propose that such a measurement could be taken as indirect evidence for a fourth generation of fermions. Considering a fourth generation where degeneracy only holds within the isodoublets individually, we find that a measurement of $M_{H}$ in the above range is consistent with the MSSM4 upper bounds on $M_{H}$. In addition, the possibility of gauge coupling unification remains intact for 60 GeV $< M_{L} <$ 110 GeV and $M_{Q} \stackrel{>}{_{\sim}}$ 170 GeV. Therefore, if $M_{H}$ is measured to be below the SM3 lower bound, we suggest a search for fourth generation fermions with 60 GeV $< M_{L} <$ 110 GeV and $M_{Q} \stackrel{>}{_{\sim}}$ 170 GeV. \section{ACKNOWLEDGEMENTS} We wish to thank M. Machacek and M. Vaughn for helpful discussions concerning the RGE used in this investigation. Support for this work was provided in part by U.S. Dept. of Energy Contract DE-FG-02-91ER40688-Task A. \newpage
\section{Introduction} The question why macroscopic superpositions are not observable in everyday life has been raised most strikingly by Schr\"{o}dinger in his famous cat paradox. Recent experiments \cite{MMKW,BrHR}, however, show that at least mesoscopic superpositions can be observed in quantum-optical systems. In quantum optics one usually speaks of a Schr\"{o}dinger cat (SC) state if one has a superposition of two different coherent states of a harmonic oscillator. In one of the experiments \cite{MMKW} a superposition of two different coherent states have been created for an ion oscillating in a harmonic potential. In the other one \cite{BrHR} two coherent states of a cavity mode were superposed, and also the process of the decoherence between these states could be followed by monitoring the field with resonant atoms. The unusual properties of such states have been discussed theoretically in several publications, see e.g. \cite{YS86,JJ90,SPL91,BMKP92}. A different type of Schr\"{o}dinger cat like state can be created in principle in a collection of two-level atoms, as first proposed in \cite{CZ94}. The terminology we may use, is the following: The individual two-level atoms can be regarded as the ``cells'' of the cat, and the cat is definitely alive, if all of its cells are alive, i.e. they are in the $|+\rangle $ state, and it is definitely dead, if all the cells are in the ill, $|-\rangle $ state. In the case of $N$ atoms a prototype of a SC like state is then: \begin{equation} |\Psi _{\text{SC}}\rangle ={\frac{1}{\sqrt{2}}} (|{+,+,\ldots ,+}\rangle +|{-,-,\ldots ,-}\rangle ), \label{SC0} \end{equation} where each of the terms contain $N$ pluses and $N$ minuses. We shall call this state the polar cat state, because the two components are in the farthest possible distance from each other. This state is in the totally symmetric $N+1$ dimensional subspace of the whole $2^{N}$ dimensional Hilbert space, and if such states are manipulated by a resonant electromagnetic field mode with dipole interaction, then the atomic system will remain in this subspace. This is the arena of the collective interaction of the atoms and the electromagnetic field, called superradiance \cite{D54,GH82,BM96}. In this work we present results concerning the properties and dynamics of polar cat states (\ref{SC0}), and also of more general collective atomic states, the generation of which have also been considered recently \cite{AgPS,G98}. Our approach of discussing the properties of quantum states like $|\Psi _{\text{SC}}\rangle $ is based mainly on the method of the Wigner function, which is one of the possible quasi-probability distributions. It has become a customary tool for investigating quantum states of an electromagnetic mode oscillator, or an ion oscillating in an appropriate trapping field \cite{FBLSS97,LMKMIW96}. The method of Wigner function is much less exploited, however, in the description of atomic states like (\ref{SC0}). That is why we first summarize the essentials of this method, and then turn to the determination of the Wigner function for the cat state (\ref{SC0}) in Section II. Next, in Section III. we consider more general cat like states, which we call ``nonpolar cats'' and deteremine their squeezing properties. Finally, in Section IV. we write down and solve the master equation for a cat state in an environment with finite temperature. We define and determine the dissipation and decoherence times of the system, and the characteristic time when the system becomes essentially classical. \section{The Wigner function of the polar cat state} The $N$-atom dipole interaction with the electromagnetic field is equivalent to the dynamics of a spin of $j=N/2$, and the phase space of the atomic subsystem is the surface of a sphere of radius $\sqrt{j(j+1)}$, ($\hbar =1$), which is sometimes called the Bloch sphere. This phase space and quasiprobability distributions corresponding to various operators acting in the $2j+1$ dimensional Hilbert space have been introduced first by Stratonovich \cite{St}. Similar constructions have been considered independently by several authors \cite{ACGT,Ag,G76,VG,SW94}. We use here the construction and notation introduced by Agarwal \cite{Ag}. Similarly to the case of oscillator quasidistributions \cite{CG,AgW}, the quasiprobability functions for angular momentum states are not unique either. Beyond the natural requirements that the possible quasiprobability distribution functions have to satisfy, there is a special property, called the product rule, that distinguishes the most natural choice among the possible quasiprobability distributions. This rule requires that the expectation value of a product of two operators could be calculated by integrating the product of the corresponding quasiprobabilities. This choice is essentially unique, and in accordance with most authors we call it the Wigner function for spin $j$. We note that the construction can be extended to include several values of $j$ \cite{FBC98,Wolf}, and in the same spirit Wigner functions can be defined for arbitrary Lie groups \cite{Brif}. We also note that it is possible to define joint Wigner functions for atom-field interactions, and then a fully phase space description of atom-field dynamics can be considered \cite{CB96}. Here we restrict ourselves to the problem of angular momentum with a fixed value of $j$. Using the procedure proposed in \cite{Ag} we shortly summarize here the method of quasiprobability functions in the $2j+1$ dimensional Hilbert space. One first chooses an operator basis in this space, and the most straightforward set of operators is the set of the spherical tensor operators $T_{KQ}$ which transform among others irreducibly under the action of the rotation operators\cite{BL}. Their explicit expression is: \begin{eqnarray} T_{KQ}=\sum_{m=-j}^{j}(-1)^{j-m}(2K+1)^{1/2}\left( \begin{array}{ccc} j & K & j \\ -m & Q & m-Q \end{array} \right) |j,m\rangle \langle j,m-Q| \end{eqnarray} where $\left( \begin{array}{ccc} j & K & j \\ -m & Q & m-Q \end{array} \right) $ is the Wigner $3j$ symbol. They form a basis in the sense that any operator of the Hilbert space can be expanded in terms of them and they fulfil the Hilbert-Schmidt orthonormality condition $\text{Tr} \left( T^{\dagger}_{KQ} T_{K'Q'} \right)=\delta_{KK'}\delta_{QQ'}$. Introducing the characteristic matrix of the density operator $\rho $ with respect of this operator basis as: \begin{equation} \varrho_{KQ}=\text{Tr} \left( \rho \,T_{KQ}^{\dagger} \right), \label{charf} \end{equation} the Wigner function of the state $\rho$ is defined as: \begin{equation} W_{\rho }(\theta ,\phi )=\sqrt{\frac{2j+1}{4\pi }}\sum_{K=0}^{2j} \sum_{Q=-K}^{K}\varrho _{KQ} \, Y_{KQ}(\theta ,\phi ). \label{WFU} \end{equation} The factor in front of the sum ensures normalization. We note that in a similar way one can associate a Wigner function $W_{A}(\theta ,\phi )$ to any operator $A$, by introducing its characteristic matrix: $A_{KQ}=\text{Tr} \left( A\,T_{KQ}^{\dagger} \right)$, and then forming the sum as in Eq. (\ref{WFU}). It can be easily seen, that this is a very similar procedure according to which one introduces the quasidistributions of oscillator states and operators by the help of characteristic functions of the translation operator basis: $D(\alpha )=\exp (\alpha a^{\dagger }-\alpha ^{*}a)$ \cite{CG,AgW}. The construction of Eq. (\ref{WFU}) can be shown to satisfy the product rule mentioned above, giving the following result for the expectation value of an operator $A$: \begin{eqnarray} \text{Tr} (\rho A)= \sqrt{\frac{4\pi }{2j+1}} \int W_{\rho}(\theta ,\phi ) W_{A}(\theta ,\phi )\sin \theta \,\text{d}\theta \,\text{d}\phi . \end{eqnarray} For other types of quasidistributions of angular momentum, like the analogs of the oscillator $P$ and $Q$ functions see \cite{Ag}. Similarly to the case of the oscillator, the Wigner function allows one to visualize the properties of the state in question. In the work of Dowling \& al. \cite{DAS} graphical representations of the Wigner function of the number, coherent and squeezed atomic states were presented. The Wigner function of a cat state like (\ref{SC0}) has been considered first in \cite {BCAS}. The characteristic matrix of the state given by (\ref{SC0}) can now be calculated according to the definition, Eq. (\ref{charf}), taking into account that the density operator corresponding to $|\Psi _{\text{SC}}\rangle $ is \begin{eqnarray} \rho _{\text{SC}}=\frac{1}{2}\left( \left| j,j\right\rangle \left\langle j,j\right| +\left| j,-j\right\rangle \left\langle j,-j\right| +\left| j,j\right\rangle \left\langle j,-j\right| +\left| j,-j\right\rangle \left\langle j,j\right| \right) \end{eqnarray} in the standard basis, with $j=N/2.$ The characteristic matrix has the form: \begin{eqnarray} (\varrho _{\text{SC}})_{K,Q} = \frac{\sqrt{2K+1}}{2} \left\{ \left( \begin{array}{ccc} j & K & j \\ -j & 0 & j \end{array} \right) (1+(-1)^{K})\delta _{Q,0} + \left( \begin{array}{ccc} j & K & j \\ -j & 2j & -j \end{array} \right) (\delta _{Q,2j}+(-1)^{K}\delta _{Q,-2j})\right\} , \end{eqnarray} and from Eq. (\ref{WFU}) one obtains the following result for the Wigner function: \begin{eqnarray} W_{\text{SC}}(\theta ,\phi )={\frac{1}{2}}\sqrt{\frac{N+1}{4\pi }} \left\{ \sum_{l=0}^{N}{\frac{{\sqrt{2l+1}N!}}{{\sqrt{(N-l)!(N+l+1)!}}}} [Y_{l0}(\theta )+Y_{l0}(\pi -\theta )] +2\sqrt{\frac{{(2N+1)!}}{{4\pi }}}{ \frac{{(\sin \theta )^{N}\cos (N\phi )}}{{2^{N}N!}}}\right\} . \label{WFSC} \end{eqnarray} The first term, containing the sums of two spherical harmonics, corresponds to the individual states $|+,+,\ldots,+\rangle $, and $|-,-,\ldots,-\rangle $, while the last term arises from the interference term between the ``living'' and ``dead'' parts of (\ref{SC0}) (the last two terms of the density operator). Fig. 1 shows the polar diagram of this Wigner function for $N=5$ atoms. The two bumps to the ``north'' and ``south'' correspond to the quasiclassical coherent constituents, while the ripples along the equator -- where the function takes periodically positive and negative values -- are the result of interference between the two kets of Eq. (\ref{SC0}). The factor $\cos (N\phi )$ in (\ref{WFSC}) shows that the number of negative ``wings'' along the equator is equal to the number of atoms. \begin{figure}[htbp] \epsfxsize=3.375in \epsfbox{figure1.ps} \caption{ Wigner function for a polar cat state, Eq. (\ref{WFSC}) for the case of $N=5$ atoms. The absolute value of the function is measured along the radius in the direction $(\theta ,\phi) $, and the surface is shown in light where the function is positive and in dark where it takes on negative values. } \label{fig:polcat} \end{figure} \section{Nonpolar cat states, and their squeezing properties} \subsection{Nonpolar cat states} One can also construct more general SC states by taking the superposition of any two atomic coherent states. An atomic coherent state (a quasiclassical state) \cite{ACGT}, $\left| \tau \right\rangle $ is an eigenstate with the highest eigenvalue $m=j$ of the component of the angular momentum operator `pointing in the direction ${\bf n}$': \begin{eqnarray} \left( {\bf J\cdot n}\right) |\tau \rangle =j|\tau \rangle . \end{eqnarray} The notation $\tau $ refers to a specific parametrization of the unit vector ${\bf n}$ by its stereographic projection to the complex plane. It is connected with the polar angle $\beta $ and the azimuth $\alpha $ of the direction ${\bf n}$ as $\tau =\tan (\beta /2)e^{-i\alpha }$. The atomic coherent state can be expanded in terms of the eigenstates $\left|j,m\right\rangle $ of $J_{z}$ \cite{ACGT}: \begin{eqnarray} \nonumber |\tau \rangle &=&\left( \frac{1}{1+|\tau |^{2}}\right) ^{j}e^{\tau J_{+}}\left| j,-j\right\rangle = \sum_{m=-j}^{j} {2j \choose j+m}^{1/2} \frac{\tau ^{j+m}}{(1+|\tau |^{2})^{j}}\left| j,m\right\rangle \nonumber \\ &=&\sum_{m=-j}^{j} {2j \choose j+m}^{1/2} \sin ^{j+m}(\beta /2)\cos ^{j-m}(\beta /2)e^{-i(j+m)\alpha } \left| j,m \right\rangle . \label{COHS} \end{eqnarray} The superposition of two quasiclassical coherent states is given by the ket: \begin{equation} |\Psi _{12}\rangle ={\frac{{|\tau _{1}\rangle +|\tau _{2}\rangle }} {\sqrt{2(1+\text{Re}\langle \tau _{1}|\tau _{2}\rangle )}}}. \label{NPSC} \end{equation} Recently Agarwal, Puri and Singh \cite{AgPS} and Gerry and Grobe \cite {G98} have proposed methods to generate such states in a cavity, via a dispersive interaction with the cavity mode. \begin{figure}[htbp] \epsfxsize=3.375in \epsfbox{figure2.ps} \caption{ The phase space scheme of a nonpolar cat state, Eq.(\ref{NPSC}) with $\tau _{2}=-\tau _{1}$. } \label{fig:scheme} \end{figure} We choose here $\tau _{1}=\tan (\beta /2)$, $\tau _{2}=-\tau _{1}$. Then $\beta $ is the polar angle of the classical Bloch vector corresponding to the atomic coherent state ${|\tau _{1}\rangle }$ ($\beta $ is measured from the south pole), see Fig. 2. This means that the $x$ component of the expectation value of the dipole moment in these states is proportional to $\pm (N/2)\sin \beta $, respectively, and the $y$ component is zero. Any other equal weight superposition of two atomic coherent states can be obtained from this special choice by an appropriate rotation. The polar cat state of the previous section corresponds to the special case when the two points are the northern and southern poles of the Bloch sphere. If the centres of the two coherent states in question are not in opposite points of the sphere, then we will call their superposition as ``nonpolar'' cat states. The corresponding quasiprobability distribution functions still can be explicitly calculated. For the Wigner function of the cat state $|\Psi _{12}\rangle $ one gets the following expression: \begin{eqnarray} W(\theta ,\phi ) = \sqrt{\frac{N+1}{4\pi}} && \sum_{K=0}^{2j}\sum_{Q=-K}^{K}{\frac{\sqrt{2K+1}{(2j)!}} {{2(1+(\cos \beta )^{2j})}}}{\ } \sum_{m=-j}^{j}{\frac{ (-1)^{j-Q-m}+(-1)^{3j+m}+(-1)^{2j}+(-1)^{2j-Q}}{\sqrt{ (j+m)!(j-m)!(j+Q+m)!(j-Q-m)!}}} \hfill \nonumber \\ && \times \left( \begin{array}{ccc} j & \ K & \ j \\ -m-Q & Q & m \end{array} \right) (\sin {\beta /2})^{2(j+m)+Q}(\cos {\beta /2})^{2(j-m)-Q}\ Y_{KQ}(\theta ,\phi ). \label{WFNP} \end{eqnarray} We present polar plots of this Wigner function in Fig. 3. for $N=5$ atoms and for several values of $\beta$. \begin{figure}[htbp] \epsfxsize=7.in \epsfbox{figure3.ps} \caption{ Wigner functions for the state $|\Psi_{12}\rangle$ for $N=5$ atoms, and for several different values of $\beta$: (a) $\beta=20^{\circ}$, (b) $\beta=45^{\circ}$, (c) $\beta=55^{\circ}$, (d) $\beta=70^{\circ}$. For smaller values of $\beta$ the state goes over into a single coherent state, and then it has essentially only one positive lobe. This graphical presentation shows qualitatively that the $y$ component of the dipole moment is squeezed, the maximal value of the squeezing in the present case ($N=5$) comes about $\beta=43^{\circ}$. } \label{fig:genpolcat} \end{figure} For small $\beta $ values, the interference is weak and the maximum of the Wigner function is around $\theta =0$. For larger $\beta $-s the function has two maxima around $\theta =\pm \beta $, and the interference gets more pronounced. When $\beta =\pi /2$, the two maxima corresponding to the individual coherent states point in the $x$ and $-x$ directions, respectively. In this case we get back the Wigner function of the SC state of Eq. (\ref{SC0}), rotated around the $y$ axis by $\pi /2$. \subsection{Squeezing properties} The expectation values of the dipole operators $J_{x}$ and $J_{y}$ are zero in the state (\ref{NPSC}) with $\tau _{1}=\tan (\beta /2)$, $\tau _{2}=-\tau _{1}$, which is a consequence of the mirror symmetry of this state with respect of both the $x-z$ and the $y-z$ planes. As it is known, the variances of the dipole operators, $J_{x}$ and $J_{y}$ are equal to each other in an atomic coherent state: \begin{equation} (\Delta J_{x})^{2}|_{\left| \tau \right\rangle }=(\Delta J_{y})^{2}|_{\left| \tau \right\rangle }=j/2. \label{VCOH} \end{equation} In order to calculate the variances in the state (\ref{NPSC}), one can use directly expansion (\ref{COHS}) and the known matrix elements of $J_{x}$ and $J_{y}$, but the summations that occur are rather cumbersome to evaluate. A more effective procedure is to apply the method of generating functions \cite{ACGT}. All the necessary expectation values in a cat state can be calculated by the formula : \begin{eqnarray} \left[ \left( \frac{\partial }{\partial \xi }\right) ^{a}\left( \frac{ \partial }{\partial \eta }\right) ^{b}\left( \frac{\partial }{\partial \zeta }\right) ^{c}X_{A}\right] _{\xi =\eta =\zeta =0}=\left\langle \tau _{1}\right| J_{-}^{a}J_{z}^{b}J_{+}^{c}\left| \tau _{2}\right\rangle , \end{eqnarray} where \begin{eqnarray} \nonumber X_{A}(\xi ,\eta ,\zeta ) & \equiv & \left\langle \tau _{1}\right| e^{\xi J_{-}}e^{\eta J_{z}}e^{\zeta J_{+}}\left| \tau _{2}\right\rangle =\frac{ \left\langle -j\right| e^{(\tau _{1}^{*}+\xi )J_{-}}e^{\eta J_{z}}e^{(\tau _{2}+\zeta )J_{+}}\left| -j\right\rangle }{\{(1+|\tau _{1}|^{2})(1+|\tau _{2}|^{2})\}^{j}} \nonumber \\ &=&\frac{\{e^{-\eta /2}+e^{\eta /2}(\tau _{1}^{*}+\xi )(\tau _{2}+\zeta )\}} {\{(1+|\tau _{1}|^{2})(1+|\tau _{2}|^{2})\}^{j}}^{2j} \end{eqnarray} is the (antinormally ordered) generating function. Inserting the necessary operators, we obtain the following expressions for the variances in the state given by (\ref{NPSC}): \begin{eqnarray} (\Delta J_{x})^{2}={\frac{j}{2}}\left( 1+{\frac{(2j-1)\sin ^{2}\beta }{{ 1+(\cos \beta )^{2j}}}}\right) , \end{eqnarray} \begin{eqnarray} (\Delta J_{y})^{2}={\frac{j}{2}}\left( 1-{\frac{(2j-1)(\cos \beta )^{2j-2}\sin ^{2}{\beta }}{{1+(\cos \beta )^{2j}}}}\right) . \label{YC} \end{eqnarray} Comparing these results with Eq. (\ref{VCOH}), we see, that except for some special cases the $J_{y}$ quadrature is squeezed while the $J_{x}$ quadrature is stretched in this state. The reason of this asymmetry lies in the fact, of course, that in the superposition (\ref{NPSC}) we have chosen states that are both centered in points lying in the $x-z$ plane. One of the exceptional cases that is not squeezed is, if there is only one atom: $j=1/2.$ As it is easily seen, for $j=1/2$ any state in the two-dimensional Hilbert space is a coherent state, and therefore it does not show squeezing. The two other exceptions are $\beta =0$ for any $j$, because then the two coherent states coincide, and $\beta =\pi /2$, which is the rotated version of the polar cat state. Writing Eq.(\ref{YC}) in the form $(\Delta J_{y})^{2}={j}(1-{\cal{S}})/2$, we can define the quantity ${\cal{S}}$ as the measure of squeezing. Analysis shows that if $N$ is large enough, then the maximum value of ${\cal{S}}$ is $0.56$ and it is achieved around $\beta _{m}=1.6/\sqrt{N}$. Figure 4 shows the dependence of ${\cal{S}}$ on $\beta $ for several values of $N=2j$. \begin{figure}[htbp] \epsfxsize=3.375in \epsfbox{figure4.ps} \caption{ The $\beta$ dependence of the quantity ${\cal{S}}$ in $(\Delta J_y)^2 = j(1-{\cal{S}}(\beta,j))/2$, for several values of $N=2j$. ${\cal{S}}$ can be considered as the measure of squeezing for a cat state consisting of two atomic coherent states separated by the central angle $2\beta$ on the Bloch sphere.} \label{fig:sq} \end{figure} \section{Decoherence and dissipation} As we mentioned in the Introduction, there have already been realizable methods proposed for the experimental generation of atomic SC states in a collection of two-level atoms \cite{AgPS,G98}. However, such an atomic ensemble can never be perfectly isolated from the surrounding environment. Further, any observation of these states necessarily leads to the interaction of the atomic system with a measuring apparatus. In both of these cases the atomic system interacts with a system containing a large number of degrees of freedom. A possible and successful approach to this problem \cite{Zu} considers that the static environment continously influences the dynamics of the atomic subsystem, which besides exchanging energy with the environment loses the coherence of its quantum superpositions and evolves into a classical statistical mixture. In this section we investigate the decoherence and dissipation of the atomic Schr\"{o}dinger cat states embedded in an environment with many degrees of freedom, by writing down the master equation for the reduced density operator of the atomic subsystem. We provide the solution for the polar cat states (\ref{SC0}). \subsection{Model and solution} We couple our ensemble of two-level atoms to the environment which is supposed to be a multimode electromagnetic radiation with photon annihilation and creation operators $a_k$ and $a_k^\dagger$. Then the interacting system can be described by the following well known model Hamiltonian which considers dipole interaction and uses the rotating wave approximation: \begin{equation} H=\omega _{\text{a}}J_{z}+ \hbar \sum_{k}\omega _{k}a_{k}^{\dagger }a_{k}+ \sum_{k}g_{k}\left( a_{k}^{\dagger }J_{-}+a_{k}J_{+}\right) , \label{model_ham} \end{equation} where $\omega_{\text{a}}$ is the transition frequency between the two atomic energy levels, the $\omega_k$ denote the frequencies of the modes of the environment and $g_k$ are the coupling constants. If we suppose the environment to be in thermal equilibrium at temperature $T$, then the time evolution of the atomic subsystem is determined by a master equation for its reduced density operator $\rho(t)$ \cite{ASp,WM}: \begin{eqnarray} \hbar^2 \frac{\text{d}\rho (t) }{\text{d}t} = -\frac{\gamma }{2} \ (\langle n \rangle+1)\ (J_{+}J_{-}\rho (t) +\rho (t) J_{+}J_{-}-2J_{-}\rho (t) J_{+}) -\frac{\gamma }{2} \ \langle n \rangle\ (J_{-}J_{+}\rho (t) +\rho (t) J_{-}J_{+}-2J_{+}\rho (t) J_{-}) \label{mastereq} \end{eqnarray} which involves the usual Born-Markov approximation and is written in the interaction picture. Here $\langle n \rangle=\left( \exp \left( \hbar \omega _{\text{a}}/ (k_{\text{B}} T)\right) -1\right) ^{-1}$ is the mean number of photons in the environment and $\gamma = (g(\omega_{\text{a}})\sigma(\omega_{\text{a}}))^2$ denotes the damping rate, where $\sigma$ is the mode density of the environment. Eq. (\ref{mastereq}) can be obtained also in a somewhat different context, as described in \cite{BSH}. Then one assumes the atomic subsystem to be placed in a resonant cavity with low quality mirrors causing the damping of the cavity mode at a rate $\kappa$. Under certain reasonable assumptions one can get Eq. (\ref{mastereq}) with $\gamma = 2 \, g(\omega_{\text{a}})^2/\kappa$. From Eq. (\ref{mastereq}) one can easily deduce the following equations for the matrix elements of the density operator $\rho _{m,l}(t)\equiv \left\langle j,m\left| \rho (t)\right| j,l \right\rangle $: \begin{eqnarray} \frac{\text{d}\rho _{m,l}(t)}{\text{d}t}& = & -\frac{\gamma}{2} \, [ \, \langle n \rangle \left( 2j(j+1)-m(m+1)-l(l+1)\right) +(\langle n \rangle+1)\left( 2j(j+1)-m(m-1)-l(l-1) \right) \, ] \, \rho _{m,l} (t) \nonumber \\ & & + \,\gamma \, \langle n \rangle\, \sqrt{\left( j(j+1)-m(m-1)\right) \left( j(j+1)-l(l-1)\right) }\ \rho _{m-1,l-1} (t) \nonumber \\ & & + \,\gamma \, (\langle n \rangle+1)\, \sqrt{\left( j(j+1)-m(m+1)\right) \left( j(j+1)-l(l+1)\right) }\ \rho _{m+1,l+1} (t) . \label{dens_mtrx_dyn} \end{eqnarray} Thus the time evolution of a particular density matrix element is coupled only to the two neighbouring elements in the corresponding diagonal for $\langle n \rangle > 0$, and only to the neighbour with larger index at zero temperature. In the case of a polar cat state (consisting of $N=2j$ atoms), the elements of the density matrix have zero initial values except for $\rho _{-j,-j},\ \rho _{j,j},\ \rho _{-j,j}$ and $\rho _{j,-j}\ (=\rho _{-j,j}^{*}).$ This implies that the density matrix elements, except for those in the main diagonal and for $\rho _{-j,j}$ and $\rho _{j,-j}$, remain identically zero for any time. Setting $\gamma =1$ (i.e. the time unit is $1/\gamma $) the equations for the elements in the main diagonal of the density matrix are the following: \begin{eqnarray} \frac{\text{d}\rho _{m,m} (t)}{\text{d}t} &=& -\left[ \langle n \rangle\, \left( j(j+1)-m(m+1)\right) +(\langle n \rangle+1)\left( j(j+1)-m(m-1)\right) \right] \ \rho _{m,m} (t) \nonumber \\ & & + \ \langle n \rangle \, \left( j(j+1)-m(m-1)\right) \ \rho _{m-1,m-1} (t) \nonumber \\ & & + \ (\langle n \rangle+1)\, \left( j(j+1)-m(m+1)\right) \ \rho _{m+1,m+1} (t) \label{main_diag_dyn} \end{eqnarray} with the initial values $\rho _{m,m}(t=0)=\frac{1}{2}(\delta _{m,j}+\delta _{m,-j})$ (cf. equation (\ref{SC0})). The dynamics of $\rho _{-j,j}$ is governed by the particularly simple equation: \begin{equation} \frac{\text{d}\rho _{-j,j} (t)}{\text{d}t}= -j(2 \langle n \rangle + 1)\rho _{-j,j} (t) , \label{decoh_dyn} \end{equation} yielding immediately the following solution with the initial value $\rho_{-j,j}(0)=1/2$ corresponding to the polar cat state: \begin{equation} \rho _{-j,j}(t)=\frac{1}{2}\exp \left( -j(2 \langle n \rangle + 1)t\right). \label{decoh} \end{equation} As expected, the stationary solution of Eq.~(\ref{main_diag_dyn}) is the Boltzmann distribution of the stationary values $\bar{\rho}_{m,m}$: \begin{eqnarray} \bar{\rho}_{m,m} & = & \exp \left( -(m+j) \frac{\hbar \omega_{\text{a}}}{k_{\text{B}} T} \right) \frac{ 1-\exp \left(-\hbar \omega_{\text{a}}/(k_{\text{B}} T) \right) }{ 1-\exp \left(- (2j+1) \hbar \omega_{\text{a}}/(k_{\text{B}} T) \right) } = \frac{ \left( \langle n \rangle / (\langle n \rangle + 1 ) \right)^{m+j} }{ (\langle n \rangle+1)\ \left(1-\left(\langle n \rangle / (\langle n \rangle + 1)\right)^{2 j+1} \right)}. \label{stac} \end{eqnarray} Approximate analytical time dependent solutions of Eq. (\ref{main_diag_dyn}) can be found especially for the case of superradiance, when $\rho_{j,j}(0)=1$ in \cite{DG}, see also \cite{BM96} and references therein. For the initial conditions corresponding to the polar cat state, the time dependent solution of equations (\ref{main_diag_dyn}) at zero temperature ($\langle n \rangle=0$) can be obtained by the following recursive integration: \begin{eqnarray} \rho _{j,j}(t) &=&\frac{1}{2}\exp \left( -2jt\right) , \nonumber \\ \label{recursion} \\ \rho _{m,m}(t) &=&\exp \left( -b_{m}t\right) \left( \frac{1}{2}\ \delta _{m,-j}+b_{m+1}\int\limits_{0}^{t}\exp \left( b_{m}t^{\prime }\right) \ \rho _{m+1,m+1}(t^{\prime })\ \text{d}t^{\prime }\right) ,\hspace{1cm}-j\leq m<j \nonumber \end{eqnarray} where $b_{m}=j(j+1)-m(m-1)$. These equations show rather explicitly, how does the initial excitation cascade down to the zero temperature stationary state. For non-zero temperatures ($\langle n \rangle>0$) we have solved equations (\ref{main_diag_dyn}) numerically. We are going to analyze the solutions in the next subsection. \subsection{Characteristic times} Figure \ref{fig:dmplot} shows the time evolution of the relevant density matrix elements $\rho_{m,m} (t)$, $m=-j,-j+1,\ldots,j$ (solid lines) and $\rho_{-j,j} (t)$ (dashed line), in the case of initial polar cat states consisting of 5 and 50 atoms, for $\langle n \rangle=0,\ 1,\ 10$. The actual value of $\rho_{-j,j}(t)$ characterizes the coherence of the corresponding state, since $\rho_{-j,j}$ and $\rho_{j,-j}$ ($=\rho_{-j,j}^*$) are the only nonzero matrix elements outside the main diagonal. Their exponential decay (cf. Eq.~(\ref{decoh})) is the decoherence, shown by the dashed lines in the plots of Fig. \ref{fig:dmplot}. Thus it is reasonable to define the characteristic time of the decoherence by \begin{eqnarray} t_{\text{dec}} = \frac{2}{N(2 \langle n \rangle + 1)}, \label{tdec} \end{eqnarray} implying $\rho_{-j,j} (t_{\text{dec}})=\rho_{-j,j}(0)/e$. In contrast to the simple time dependence of $\rho_{-j,j}(t)$, the dynamics of the main diagonal elements $\rho_{m,m} (t)$ depend on the actual value of $\langle n \rangle$ and $N$ rather sensitively. The zero temperature cases, Fig. \ref{fig:dmplot} (a) and (d), clearly show the initial excitation, contained in $\rho_{j,j} (0)= 1/2$, cascading down to $\rho_{-j,-j} (\infty)=1$ as given by Eqs. (\ref{recursion}). At nonzero temperatures ($\langle n \rangle > 0$) the time evolution of the $\rho_{m,m} (t)$-s is more complicated because of the coupling to both neighbours, cf. Eq. (\ref{dens_mtrx_dyn}). More information can be extracted from the time evolution of the $\rho_{m,m} (t)$-s by calculating the energy of the atomic subsystem as the function of time: \begin{eqnarray} E(t) \equiv \langle \omega_{\text{a}} \, J_z \rangle (t) = \omega_{\text{a}} \text{Tr} \left( \rho (t) \, J_z \right) = \hbar \omega_{\text{a}} \sum_{m=-j}^{j} \, m \, \rho_{m,m} (t). \label{energy} \end{eqnarray} \begin{figure}[htbp] \epsfxsize=7.in \epsfbox{figure5.ps} \caption{ Plots of the density matrix elements $\rho_{m,m} (t)$, $m=-j,-j+1,\ldots,j$ (solid lines) and $\rho_{-j,j} (t)$ (dashed line) versus time (the time unit is $1/\gamma$). Plots are given for $N=5$ atoms: (a, b, c), and $N=50$ atoms: (d, e, f), for $\langle n \rangle=0$ (zero temperature): (a, d), for $\langle n \rangle=1$: (b, e), and for $\langle n \rangle=10$: (c, f). The solid lines may be identified as follows: in (a) the smaller $m$ is the later the corresponding $\rho_{m,m}$ has its maximal value; in (b) and (c) the stationary values of the $\rho_{m,m}$-s follow the Boltzmann distribution, see Eq.~(\ref{stac}). In (d, e, f) the $\rho_{m,m}$-s follow each other along the axis labeled by the $m$. The dotted lines (starting from 1 at $t=0$) show the normalized energy of the atomic subsystem $1+E(t)/(-j \hbar \omega_{\text{a}})$, see Eq.~(\ref{energy}). The number $r=t_{\text{diss}}/t_{\text{dec}}$ is the ratio of the characteristic time of dissipation to the characteristic time of decoherence. } \label{fig:dmplot} \end{figure} The process of dissipation (i.e. the change of the energy of the atomic subsystem in time) can be very easily followed by studying $E(t)$. This function, normalized to the zero temperature stationary energy and shifted to vary from 1 to its stationary value: $1+E(t)/(-j \hbar \omega_{\text{a}})$, is shown in the plots of Fig.\ \ref{fig:dmplot} by the dotted lines. Since its asymptotic behavior is exponential-like, it is reasonable to define the characteristic time of dissipation $t_{\text{diss}}$ by requiring \begin{equation} |E(t_{\text{diss}})-E(\infty)|= |E(0)-E(\infty)|/e. \end{equation} In order to ensure that $E(t)$ achieves its stationary value with a good accuracy in the plots of Fig. \ref{fig:dmplot} we have set the time range to $5 \, t_{\text{diss}}$. It is seen that the value of $r=t_{\text{diss}}/t_{\text{dec}}$ grows with both the temperature and the number of atoms. A more detailed analysis of this question follows later in this section. The initial state of the process, the polar cat state, has sharply non-classical features. On the other hand, at non-zero temperature the final stationary state of the present model is a thermal state, which is classical in its nature. (At zero temperature the stationary state is also non-classical, since it is the state $|j,-j\rangle$.) It is natural to ask, when does the transition from the non-classical to the classical stage occur? What is a good measure of non-classicality reflecting the change of non-classical nature of the corresponding state? The spherical Wigner function (\ref{WFU}) provides a good answer to both of these questions. Quantum states are generally considered essentially non-classical if the corresponding Wigner function takes on also negative values. Therefore to answer the second question, for the measure of the degree of non-classicality we propose to use the quantity $\nu = 1 - (I_+ - I_-)/(I_+ + I_-)$, where $I_+$ is the integral of the Wigner function over those domains where it is positive while $I_-$ is the absolute value of the integral of the Wigner function over the domains where it is negative. Since the integral of the Wigner function over the sphere is 1, $I_+ - I_- = 1$, thus $\nu = 2 I_- /(2 I_- + 1)$, and it is easy to see that $0 \leq \nu < 1$. According to this definition, the bigger is the value of $\nu$, the more non-classical is the state, and for all classical states one has $\nu=0$. Regarding now the first question, namely for how long is the state of the atomic system non-classical, we introduce a third kind of characteristic time $t_{\text{ncl}}$. We define $t_{\text{ncl}}$ to be the time instant when the corresponding spherical Wigner function becomes non-negative everywhere on the sphere, i.e. $\nu$ becomes 0. We will return to this question in connection with the time evolution of the Wigner function, which we will present in the next subsection in more detail. Based on the information provided by the three kinds of characteristic times, we consider here the dependence of the process on the number of atoms and on the temperature. In Figure \ref{fig:cht} we plot $t_{\text{diss}}$ (dashed line), $t_{\text{dec}}$ (solid line) and $t_{\text{ncl}}$ (dotted line) as the function of the number of constituent atoms of the polar cat $N$, for several temperatures, on a log-log scale. \begin{figure}[htbp] \epsfxsize=3.375in \epsfbox{figure6.ps} \caption{ Plots of the characteristic times of decoherence, $t_{\text{dec}}$ (solid line), dissipation, $t_{\text{diss}}$ (dashed line) and non-classicality, $t_{\text{ncl}}$ (dotted line) versus the number of atoms, on a log-log scale. The uppermost solid and dashed line are for $\langle n \rangle=0$ (there is no plot for $t_{\text{ncl}}$ at zero temperature, since the state stays non-classical), while the subsequent groups of the three kinds of lines, one under the other, are for $\langle n \rangle=1,\ 10,\ 100$, respectively.} \label{fig:cht} \end{figure} It is seen that the characteristic time of decoherence $t_{\text{dec}}$ is inversely proportional to the number of atoms (the straight solid lines in Fig. \ref{fig:cht}), according to the definition (\ref{tdec}). Compared to this, the characteristic time of non-classicality $t_{\text{ncl}}$ decreases less rapidly with increasing number of atoms. The characteristic time of dissipation $t_{\text{diss}}$ first slightly increases at non-zero temperature then it achieves a maximum which depends on $\langle n \rangle$ and finally it decreases nearly inversely proportional to the number of atoms. The values of $t_{\text{diss}}$ at different temperatures seem to converge slowly beyond a certain number of atoms. It seems however rather surprising that the ratio $t_{\text{diss}}/t_{\text{dec}}$ is not as large as such a quantity is usually expected to be \cite{Zu,large}: in the case of $N=1000$ it is 4.04 for $\langle n \rangle=0$, and it is still just around 350 for $\langle n \rangle=100$. (Note that $\langle n \rangle=100$ corresponds to a temperature of 250~K in the case of typical experiments \cite{BrHR}.) The ratio $t_{\text{diss}}/t_{\text{dec}}$ seems not even to vary considerably with increasing $N$ beyond the maximum of $t_{\text{diss}}$ mentioned above. Thus the process of decoherence is extremely slow in the case of a polar cat state which is coupled to the environment by an interaction leading to the master equation (\ref{mastereq}). Similar effects have already been reported for other physical systems earlier \cite{similar}. In a recent work Braun, Braun and Haake \cite{BBH} investigated the decoherence of an atomic SC state $| \tau_1 \rangle + | \tau_2 \rangle$ based on Eq. (\ref{mastereq}) for zero temperature. By evaluating a certain quantity characterising the decoherence rate at the {\it initial} time, and applying a semiclassical procedure for finite times they concluded that for atomic SC states with $\tau_1 \tau_2^* = 1$ the decoherence slows down. Our initial state, the polar cat state fulfils the former condition. The results presented in Fig. \ref{fig:cht} derive from the solution of the master equation for the whole process. They are in agreement with the statements of Ref.\cite{BBH}, where the initial stage of the decoherence is analyzed for the case of zero temperture. \subsection{Wigner functions} We illustrate now the process of decoherence and dissipation using the spherical Wigner function (\ref{WFU}). \begin{figure}[htbp] \epsfxsize=7.in \epsfbox{figure7.ps} \caption{ Polar plots of the temporal change of the Wigner function representing the decoherence and dissipation of the initial polar cat state (\ref{SC0}), made of 5 atoms and shown in Fig.~\ref{fig:polcat}, at zero temperature. The dynamics of the corresponding density matrix elements is shown in Fig.~\ref{fig:dmplot}(a). The time instants are the following (in units of $1/\gamma$): (a) 0.1, (b) 0.2, (c) 0.4 ($=t_{\text{dec}}$), (d) 0.506 ($=t_{\text{diss}}$), (e) 0.8, (f) 2.5. } \label{fig:wfN5n0} \end{figure} In order to obtain its time dependence we have to calculate first the characteristic matrix $\varrho _{K,Q}(t)=\text{Tr} \left( \rho (t)T_{K,Q}^\dagger \right) $ from the matrix elements $\rho_{m,l}(t)$ according to \begin{equation} \varrho _{K,Q}(t)=\sqrt{2K+1}\sum_{m=-j}^{j}(-1)^{j-m}\left( \begin{array}{ccc} j & K & j \\ -m & Q & m-Q \end{array} \right) \rho _{m,m-Q}(t). \label{rhotrf} \end{equation} From Eq. (\ref{rhotrf}) it can be seen that only $\varrho _{K,0}$ ($K=0,1,\ldots,N$) and $\varrho _{N,N}=(-1)^{N} (\varrho _{N,-N})^{*}$ are nonzero. This fact (which is due to the initial conditions specified by the polar cat state) ensures that the azimuthal dependence of the spherical Wigner function is determined only by the real part of the spherical harmonic $Y_{N,N} (\theta,\phi)$ which is proportional to $\cos (N \phi)$. Therefore the Wigner function keeps its initial azimuthal symmetry during the whole process. Further, since $\varrho _{N,N} (t) =(-1)^{N} \rho _{j,-j} (t)$, the azimuthal modulation of the spherical Wigner function explicitly shows the degree of the coherence of the actual state. Figures \ref{fig:wfN5n0} and \ref{fig:wfN5n10} show the polar plots of the spherical Wigner function transforming its shape in time at $\langle n \rangle = 0$ and $\langle n \rangle = 10$, respectively. The initial state is a polar cat made of $N=5$ atoms and its Wigner function is shown in Fig.~\ref{fig:polcat}. \begin{figure}[htbp] \epsfxsize=7.in \epsfbox{figure8.ps} \caption{ Polar plots of the temporal change of the Wigner function representing the decoherence and dissipation of the initial polar cat state (\ref{SC0}), made of 5 atoms and shown in Fig.~\ref{fig:polcat}, for $\langle n \rangle=10$. The dynamics of the corresponding density matrix elements is shown in Fig.~\ref{fig:dmplot}(c). The time instants are the following (in units of $1/\gamma$): (a) 0.01, (b) 0.019 ($=t_{\text{dec}}$), (c) 0.031 ($=t_{\text{ncl}}$), (d) 0.045, (e) 0.058 ($=t_{\text{diss}}$), (f) 0.25. } \label{fig:wfN5n10} \end{figure} In Figures \ref{fig:wfN5n0} and \ref{fig:wfN5n10} the following main characteristics of the process can be identified. The decoherence is shown by the decreasing and finally disappearing ripples along the equator. The vanishing of non-classicality, i.e. the decrease of the parameter $\nu$, can be easily recognized as the decrease of the negative (dark) wings. At nonzero temperature they disappear exactly at $t_{\text{ncl}}$, as shown in Fig. \ref{fig:wfN5n10}(c). The dissipation is represented by the approach of the initial upper and lower bumps to each other. At zero temperature (Fig. \ref{fig:wfN5n0}) the upper bump disappears and goes over to the lower one. This stationary shape of the Wigner function corresponds to the lowest coherent state $|j,-j\rangle$ \cite{DAS}. For $\langle n \rangle = 10$, when the stationary energy is close to the initial energy, not only the upper bump moves downwards but also the lower one lifts upwards. The stationary Wigner function has nearly a spherical symmetry, although its center is not in the origin. In agreement with Fig. \ref{fig:cht}, the plots of Figures \ref{fig:wfN5n0} and \ref{fig:wfN5n10} show, that the timescales of the decoherence and of the dissipation are very close to each other in the case of 5 atoms, for zero temperature they are practically the same. Then the spherical Wigner function exhibits considerable azimuthal modulation (ripples) also at $t_{\text{diss}}$. We may come back finally to the question of finding the characteristic time of non-classicality $t_{\text{ncl}}$. According to the arguments given after Eq. (\ref{rhotrf}) it is sufficient to study the Wigner function within a $\phi$-range of length $2\pi/N$, e.g. $0\leq \phi \leq 2\pi/N$, because it is invariant with respect of rotations by $\phi=k {2\pi\over N},\ (k=1,2, \dots, N) $ i.e. it has $C_{N}$ symmetry at all times. Therefore the spherical Wigner function of a polar cat state, while subject to dissipation and decoherence, has its minimum value always at $\phi = \pi/N$. Thus in order to calculate $t_{\text{ncl}}$, it is sufficient to follow the time evolution of the section $W(\theta,\phi=\pi/N)$. Further, in connection with the calculation of the measure of non-classicality $\nu$, it is sufficient to consider the above mentioned $\phi$-range when evaluating the integrals $I_+$ and $I_-$. \section{Conclusions} We have considered a class of states in an ensemble of two-level atoms, a superposition of two distinct atomic coherent states which are called atomic Schr\"{o}dinger cat states. According to the relative positions of the constituents we have defined polar and nonpolar cat states. We have investigated their properties based on the spherical Wigner function, which has been proven to be a convenient tool to investigate the quantum interference effects. We have shown that nonpolar cat states generally exhibit squeezing, for which we have introduced the measure ${\cal S}$. The squeezing depends on the separation of the components of the cat and on the number of the atoms the cat is consisting of. By solving the master equation of this system embedded in an external environment we have determined the characteristic times of decoherence, dissipation and non-classicality of an initial polar cat state. We have shown how these depend on the number of the microscopic elements the cat consists of, and on the temperature of the environment. Our results show that the decoherence of the polar cat state is surprisingly slow: $t_{\text{diss}}/t_{\text{dec}}$ is less then a half of an order of magnitude for zero temperature, making these states potentially significant in several areas of quantum physics, e.g. experimental studies of decoherence, quantum computing and cryptography. We have visualised the process, governed by the interaction with the external environment, using the spherical Wigner function. Its transformation in time reflects the characteristics of the behaviour of the atomic subsystem in a suggestive way. \section*{Acknowledgments} The authors thank L.~Di\'{o}si, T.~Geszti, F.~Haake, J.~Janszky and W.~P.~Schleich for enlightening discussions on the subject, and Cs.~Benedek for his help in figure plotting. One of the authors, A.~C. is grateful to the DAAD for financial support. This work was supported by the Hungarian Scientific Research Fund (OTKA) under contracts T022281, F023336 and M028418.
\section{Introduction} Topological field theories \cite{birm1} are mathematically as well as physically interesting \cite{wit}. In four dimensions we have two interesting topological field theories, The topological Yang--Mills model \cite{wit} and the antisymmetric tensor field model \cite{hor1}. The algebraic renormalization\footnote{More details about the subject of algebraic renormalization can be found in the last reference of \cite{pig3}} of the topological Yang--Mills field theory was carried out in \cite{mwo} and then extended to curved space--time in \cite{zer2}. On the other hand, the algebraic renormalization of the antisymmetric tensor field model (in the flat space--time limit) was first done in \cite{sor2} and then generalized to a curved space--time admitting a covariantly constant vector field in \cite{zer3}. In this work we will go a step further and try to generalize the analysis of \cite{zer3} to an arbitrary, curved, Riemannian manifold.\\ The paper is organized as follows, in section 2 we give the gauge fixed action and display the BRS transformations of all the fields appearing in the model. Next, in section 3 we derive the on--shell local supersymmetry--like transformations, i.e. the anticommutator of the BRS operator and of the local susy--like operator leads to Lie derivative . In section 4 we extend the on--shell analysis (of section 3) to the off--shell level. Here we will see that the local susy--like Ward operator, when it acts on the total action describing the model, gives rise to a hard breaking which is quadratic in the quantum fields. In order to eliminate this quadratic breaking we introduce auxiliary fields. The result of this construction is that the anticommutator of the linearized Slavnov operator and the Ward operator of the new symmetry does not close on diffeomorphisms for certain fields. Furthermore, the most important fact is that the total action is invariant under this new, local, nonlinear (and not susy--like) symmetry. \\ On the other hand, one could use this symmetry (first one has to show its validity at all order of perturbation theory) to prove the finiteness of the $4D$ antisymmetric tensor field model in a general class of curved manifolds: generalizing the results of \cite{zer3}. \section{The model} \label{sec1} First let us consider the following classical action in curved space--time \begin{equation} \S_{inv}= - \frac{1}{4}\int_{\cal M} d^4x ~\varepsilon^{\mu\nu\rho\sigma} F^a_{\mu\nu}B^a_{\rho\sigma} \ , \eqn{curv-bf} where ${\cal M}$ is a curved manifold described by the Euclidean metric $g_{\mu\nu}$. The field strength is described by \begin{equation} F_{\mu\nu}^a = \partial_\mu A^a_\nu - \partial_\nu A^a_\mu + f^{abc} A^b_\mu A^c_\nu, \eqn{strengthfga} where $A^a_\mu$ is the gauge field which belongs to the adjoint representation of a compact Lie group whose structure constants are denoted by $f^{abc}$. The antisymmetric tensor field $B^a_{\mu\nu}$ is also Lie algebra valued and $\varepsilon^{\mu\nu\rho\sigma} \renewcommand{\S}{\Sigma}$ is the Levi--Civita tensor density\footnote{In this paper we denote the inverse of the metric by $g^{\mu\nu}$ and its determinant by $g$. Under diffeomorphisms, $\sqrt g$ behaves like a scalar density of weight +1, whereas the volume element density $d^4x$ has weight -1. The Levi--Civita antisymmetric tensor density $\varepsilon^{\mu\nu\rho\sigma} \renewcommand{\S}{\Sigma}$ has weight $+1$ and $\varepsilon_{\mu\nu\rho\sigma} \renewcommand{\S}{\Sigma}$ has weight $-1$. Furthermore, we have the following useful identity \[ \varepsilon_{\mu\nu\rho\sigma} \renewcommand{\S}{\Sigma}= \frac{1}{g} g_{\mu\a}g_{\nu\b}g_{\rho\gamma} \newcommand{\G}{\Gamma}g_{\sigma} \renewcommand{\S}{\Sigma\l} \varepsilon^{\a\b\gamma} \newcommand{\G}{\Gamma\l}. \] } of weight $+1$. On the other hand, the action (\ref{curv-bf}) possesses two kinds of invariance, given by \begin{eqnarray} \delta^{(1)}A^a_\mu \= -(D_\mu \theta)^a = - (\partial_\mu \theta^a + f^{abc}A^b_\mu \theta^c) \ , \nonumber \\ \delta^{(1)}B^a_{\mu\nu} \= f^{abc}\theta^b B^c_{\mu\nu} \ , \eqan{sym1} and \begin{eqnarray} \delta^{(2)}A^a_\mu \= 0 \ , \nonumber \\ \delta^{(2)}B^a_{\mu\nu} \= -(D_\mu \varphi_\nu - D_\nu \varphi_\mu)^a \ , \eqan{sym2} $\theta^a$ and $\varphi^a_\mu$ are local parameters. Now, by choosing a Landau gauge the gauge--fixing part of the action is given by \cite{zer3} \begin{eqnarray} \S_{gf} \= - s \int_{\cal M} d^4 x \sqrt{g} \bigg[g^{\mu\nu}\partial_\mu \bar c^a A^a_\nu + g^{\mu\alpha}g^{\nu\beta}\partial_\a \bar\xi^a_\beta B^a_{\mu\nu} + g^{\mu\nu} \partial_\mu \bar\phi^a \xi^a_\nu - g^{\mu\nu}\partial_\nu e^a \bar\xi^a_\mu - \bar\phi^a \lambda^a \bigg] \nonumber \\ &-& \frac{1}{2}\int_{\cal M} d^4x ~\varepsilon^{\mu\nu\rho\sigma}f^{abc} \partial_\mu \bar \xi} \newcommand{\X}{\Xi_\nu^a \partial_\rho \bar \xi} \newcommand{\X}{\Xi_\sigma} \renewcommand{\S}{\Sigma^b {\phi}} \newcommand{\F}{{\Phi}^c , \eqan{gauge-fix} As already computed in \cite{zer3}, the extended BRS transformations of all fields introduced so far read \begin{eqnarray} sA^a_\mu \= -(D_\mu c)^a = -(\partial_\mu c^a + f^{abc}A^b_\mu c^c) \ , \nonumber \\ sB^a_{\mu\nu} \= -(D_\mu \xi_\nu - D_\nu \xi_\mu)^a + f^{abc} c^b B^c_{\mu\nu} + \varepsilon_{\mu\nu\rho\sigma}f^{abc} \sqrt{g}g^{\rho\a}g^{\sigma\b}(\partial_\a\bar\xi} \newcommand{\X}{\Xi^b_\b){\phi}} \newcommand{\F}{{\Phi}^c \ , \nonumber \\ s\xi^a_\mu \= (D_\mu \phi)^a + f^{abc} c^b \xi^c_\mu \ , \nonumber \\ s\phi^a \= f^{abc} c^b \phi^c \ , \nonumber \\ sc^a \= \frac 1 2 f^{abc} c^b c^c \ , \nonumber \\ s\bar c^a \= b^a \ , \hspace{4.1cm} sb^a = 0 \ , \nonumber \\ s\bar \xi} \newcommand{\X}{\Xi^a_\mu \= h^a_\mu \ , \hspace{4cm} sh^a_\mu = 0 \ , \nonumber \\ s\bar {\phi}} \newcommand{\F}{{\Phi}^a \= \omega^a \ , \hspace{4cm} s\omega^a = 0 \ , \nonumber \\ s e^a \= \lambda^a \ , \hspace{4cm} s\lambda^a = 0 \ , \nonumber \\ s g_{\mu\nu} \= \hat g_{\mu\nu} \ , \hspace{3.8cm} s\hat g_{\mu\nu} = 0 \ . \eqan{brs-set} The vector $\xi} \newcommand{\X}{\Xi^a_\mu$ is the ghost field for the symmetry (\ref{sym2}) whereas $c^a$ is the ghost field for the gauge symmetry (\ref{sym1}). ${\phi}} \newcommand{\F}{{\Phi}^a$ is the ghost for the ghost field $\xi} \newcommand{\X}{\Xi^a_\mu$. Each of the couples of fields $(\bar c^a, b^a)$, $(\bar \xi} \newcommand{\X}{\Xi^a_\mu, h^a_\mu)$, $(\bar {\phi}} \newcommand{\F}{{\Phi}^a, \o^a)$ and $(e^a, \l^a)$ contains an antighost and the corresponding Lagrange multiplier fields.\\ We could extend the BRS transformations (see last line of (\ref{brs-set})) by letting $s$ acting on the metric $g_{\mu\nu}$ because, at the level of the gauge fixed action $\S_{inv} + \S_{gf}$, the metric appears only in a BRS exact expression \cite{zer3}, a fact which guarantee its non physical character. It turned out \cite{zer3} that the BRS operator, constructed above, is nilpotent on--shell. more precisely, \begin{equation} s^2 B^a_{\mu\nu} = - \varepsilon_{\mu\nu\rho\sigma}f^{abc} \frac{\d (\S_{inv}+\S_{gf})}{\d B^b_{\rho\sigma}}\phi^c~~~\hbox{and}~~~ s^2 = 0~~~\hbox{for the other fields} \ . \eqn{inv1} \begin{table}[h] \renewcommand{\arraystretch}{1.2} \begin{center} \begin{tabular}{|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|c|}\hline &$A^a_\mu$ &$B^a_{\mu\nu}$ &$c^a$ &$\bar c^a$ &$b^a$ &$\xi} \newcommand{\X}{\Xi^a_\mu$ &$\bar\xi} \newcommand{\X}{\Xi^a_\mu$ &$h^a_\mu$ &${\phi}} \newcommand{\F}{{\Phi}^a$ &$\bar{\phi}} \newcommand{\F}{{\Phi}^a$ &$\omega^a$ &$e^a$ &$\lambda^a$ &$g_{\mu\nu}$ &$\hat g_{\mu\nu}$ \\ \hline dim &1 &2 &0 &2 &2 &1 &1 &1 &0 &2 &2 &2 &2 &0 &0 \\ \hline $\phi\pi$ &0 &0 &1 &-1 &0 &1 &-1 &0 &2 &-2 &-1 &0 &1 &0 &1 \\ \hline weight &0 &0 &0 &0 &0 &0 &0 &0 &0 &0 &0 &0 &0 &0 &0 \\ \hline \end{tabular} \\ \vspace{0.5cm} \small{Table 1: Dimensions, ghost numbers and weights of the fields.} \end{center} \end{table} \section{The on--shell analysis} In the flat space--time limit the authors of \cite{sor2} constructed, besides the BRS transformations, a further symmetry of the gauge fixed action, the so called vector supersymmetry--like transformations. Their analysis was generalized \cite{zer3} to the case of a curved space--time admitting a covariantly constant vector field. In both cases the supersymmetry--like transformations were rigid transformations. In this paper we make a further step and try to construct a {\it local} supersymmetry--like transformations (at least on--shell), so let us begin by proposing the following transformations \begin{eqnarray} \d_{(\eta)}A^a_\mu \= -\varepsilon_{\mu\nu\rho\sigma}\eta^\nu \sqrt g g^{\rho\a}g^{\sigma\b}\partial_\a\bar\xi^a_\b \ , \nonumber \\ \d_{(\eta)}B^a_{\mu\nu} \= -\varepsilon_{\mu\nu\rho\sigma}\eta^\rho \sqrt g g^{\sigma\a}\partial_\a\bar c^a \ , \nonumber \\ \d_{(\eta)}c^a \= -\eta^\mu A^a_\mu \ , \nonumber \\ \d_{(\eta)}\bar c^a \= 0 \ , \nonumber \\ \d_{(\eta)}b^a \= {\cal L}_\eta\bar c^a \ , \nonumber \\ \d_{(\eta)}\xi} \newcommand{\X}{\Xi^a_\mu \= \eta^\nu B^a_{\mu\nu} \ , \nonumber \\ \d_{(\eta)}\bar\xi} \newcommand{\X}{\Xi^a_\mu \= 0, \nonumber \\ \d_{(\eta)}h^a_\mu \= {\cal L}_\eta\bar\xi} \newcommand{\X}{\Xi^a_\mu , \nonumber \\ \d_{(\eta)}{\phi}} \newcommand{\F}{{\Phi}^a \= \eta^\mu \xi} \newcommand{\X}{\Xi^a_\mu \ , \nonumber \\ \d_{(\eta)}\bar {\phi}} \newcommand{\F}{{\Phi}^a \= 0 \ , \nonumber \\ \d_{(\eta)}\omega^a \= {\cal L}_\eta\bar {\phi}} \newcommand{\F}{{\Phi}^a \ , \nonumber \\ \d_{(\eta)}e^a \= 0 \ , \nonumber \\ \d_{(\eta)}\lambda^a \= {\cal L}_\eta e^a \ , \nonumber \\ \d_{(\eta)}g_{\mu\nu} \= 0 \ , \nonumber \\ \d_{(\eta)}\hat g_{\mu\nu} \= {\cal L}_\eta g_{\mu\nu} \ , \eqan{susy} where ${\cal L}_\eta$ is the Lie derivative and $\eta^\mu$ is the vector parameter of the transformations with ghost number $+2$. It turns out that the on--shell algebra takes the following form \begin{equation} \lbrace s,\ \d_{(\eta)} \rbrace = {\cal L}_\eta + ~~~\hbox{equ. of motion} \eqn{aklsdj} for all fields, except for the antisymmetric tensor field $B^a_{\mu\nu}$ where we get \begin{equation} \left\{}\newcommand{\rac}{\right\} s , \d_{(\eta)} \rac B^a_{\mu\nu} = {\cal L}_\eta B^a_{\mu\nu} +\varepsilon_{\mu\nu\rho\sigma} \renewcommand{\S}{\Sigma} \eta^\rho \frac{\d \S}{\d A^a_\sigma} \renewcommand{\S}{\Sigma} +\varepsilon_{\mu\nu\rho\sigma} \renewcommand{\S}{\Sigma} \eta^\rho f^{abc} \sqrt g g^{\sigma} \renewcommand{\S}{\Sigma\a} \partial_\a \bar{\phi}} \newcommand{\F}{{\Phi}^b {\phi}} \newcommand{\F}{{\Phi}^c \ . \eqn{algebra-B} Now to repair this shortcoming we add to the gauge fixed action the following expression \begin{equation} \S_{K,M}=\int d^4x (K_\mu^a \Xi^{a\mu}- M^a_\mu s \Xi^{a\mu} ) \eqn{sigmakm} with \begin{equation} \Xi^{a\mu}=-f^{abc} \sqrt g g^{\mu\a}\partial_\a \bar {\phi}} \newcommand{\F}{{\Phi}^b{\phi}} \newcommand{\F}{{\Phi}^c. \eqn{Xi} The two auxiliary fields $K^a_\mu$ and $M^a_\mu$ transform under the BRS and $\d_{(\eta)}$ operators according to \begin{equation} \begin{array}{llll} s M^a_\mu &=& K^a_\mu, ~~~&~~~ s K^a_\mu = 0, \\ \d_{(\eta)} M^a_\mu &=& 0, ~~~&~~~ \d_{(\eta)} K^a_\mu = {\cal L}_\eta M^a_\mu. \end{array} \eqn{brskm} In an easy way we can prove that \begin{equation} \left\{}\newcommand{\rac}{\right\} s,\ \d_{(\eta)} \rac = {\cal L}_\eta ~ , \eqn{kusdfgfi95hwj} for the two auxiliary fields. \begin{table}[h] \renewcommand{\arraystretch}{1.2} \begin{center} \begin{tabular}{|c|c|c|}\hline &$K^a_\mu$ &$M^a_\mu$ \\ \hline dim &1 &1 \\ \hline $\phi\pi$ &0 &-1 \\ \hline weight &0 &0 \\ \hline \end{tabular} \\ \vspace{0.5cm} \small{Table 2: Dimensions, ghost numbers and weights.} \end{center} \end{table} The advantage of introducing the auxiliary fields $K^a_\mu$ and $M^a_\mu$ is that (\ref{algebra-B}) will get the same form as (\ref{aklsdj}). Indeed, \begin{equation} \left\{}\newcommand{\rac}{\right\} s , \d_{(\eta)} \rac B^a_{\mu\nu} = {\cal L}_\tau B^a_{\mu\nu} +\varepsilon_{\mu\nu\rho\sigma} \renewcommand{\S}{\Sigma} \eta^\rho \bigg( \frac{\d \S}{\d A^a_\sigma} \renewcommand{\S}{\Sigma} - \frac{\d \S}{\d K^a_\sigma} \renewcommand{\S}{\Sigma} \bigg) \eqn{corralgeb} So, in this way we have constructed an on--shell local supersymmetry--like transformations which anticommute with the BRS operator and lead to Lie derivative (\ref{aklsdj}). Next, we want to know if the transformations, generated by the operator $\d_{(\eta)}$, give rise to a symmetry of the gauge fixed action. This question is investigated in the next section where we also display the off--shell algebra. \section{The off--shell analysis} In order to generalize the above results to the off--shell level we first introduce external sources which couple to the non--linear BRS transformations (\ref{brs-set}) \begin{eqnarray} \S_{ext} \= \int_{\cal M} d^4x \Big[\Omega^{a\mu}(sA^a_\mu) + \gamma^{a\mu\nu}(sB^a_{\mu\nu}) +L^a(sc^a)+D^a(s{\phi}} \newcommand{\F}{{\Phi}^a)+\rho^{a\mu}(s\xi} \newcommand{\X}{\Xi^a_\mu)\Big] \nonumber \\ &+&\in \Big[\frac{1}{2}\varepsilon_{\mu\nu\rho\sigma}f^{abc}\gamma^{a\mu\nu} \gamma^{b\rho\sigma}{\phi}} \newcommand{\F}{{\Phi}^c \Big] \ , \eqan{ext-action} with dimensions, weights and ghost numbers as given in table 3. \begin{table}[h] \renewcommand{\arraystretch}{1.2} \begin{center} \begin{tabular}{|c|c|c|c|c|c|}\hline &$\gamma^{a\mu\nu}$ &$\Omega^{a\mu}$ &$L^a$ &$D^a$ &$\rho^{a\mu}$ \\ \hline dim &2 &3 &4 &4 &3 \\ \hline $\phi\pi$ &-1 &-1 &-2 &-3 &-2 \\ \hline weight &1 &1 &1 &1 &1 \\ \hline \end{tabular} \\ \vspace{0.5cm} \small{Table 3: Dimensions, ghost numbers and weights of the external sources.} \end{center} \end{table} \noindent From the transformations \equ{susy} and the second line of (\ref{brskm}) (and due to presence of the external sources) we get the Ward operator $\bar {\cal W}^S_{(\eta)}$ such that \begin{eqnarray} \bar {\cal W}^S_{(\eta)} \= \in \Bigg[ -\varepsilon_{\mu\nu\rho\sigma}\eta^\nu (\gamma} \newcommand{\G}{\Gamma^{a\rho\sigma} \renewcommand{\S}{\Sigma} + \sqrt g g^{\rho\a}g^{\sigma\b}\partial_\a\bar\xi^a_\b) \dd{A^a_\mu} -\eta^\mu A^a_\mu \dd{c^a} + {\cal L}_\eta\bar c^a \dd{b^a} - \nonumber \\ &-&\varepsilon_{\mu\nu\rho\sigma}\eta^\rho (\O^{a\sigma} \renewcommand{\S}{\Sigma} +\sqrt g g^{\sigma\a}\partial_\a\bar c^a ) \dd{B^a_{\mu\nu}} +\eta^\nu B^a_{\mu\nu} \dd{\xi} \newcommand{\X}{\Xi^a_\mu} +{\cal L}_\eta\bar\xi} \newcommand{\X}{\Xi^a_\mu \dd{h^a_\mu} +\eta^\mu \xi} \newcommand{\X}{\Xi^a_\mu \dd{{\phi}} \newcommand{\F}{{\Phi}^a} + \nonumber \\ &+& {\cal L}_\eta\bar {\phi}} \newcommand{\F}{{\Phi}^a\dd{\omega^a} +{\cal L}_\eta e^a \dd{\lambda^a} + {\cal L}_\eta g_{\mu\nu} \dd{\hat g_{\mu\nu}} -\eta^\mu D^a \dd{\rho^{a\mu}} -\eta^\mu L^a \dd{\Omega^{a\mu}} - \nonumber \\ &-& \eta^\mu \rho^{a\nu} \dd{\gamma^{a\mu\nu}} + ({\cal L}_\eta M^a_\mu - \varepsilon_{\mu\nu\rho\sigma} \renewcommand{\S}{\Sigma} \eta^\nu \gamma} \newcommand{\G}{\Gamma^{a\rho\sigma} \renewcommand{\S}{\Sigma})\dd{K^a_\mu} \Bigg] \ . \eqan{ward-operator-susy} After tedious calculations, the corresponding Ward identity takes the form \begin{equation} \bar{\cal W}^S_{(\eta)} (\S_{inv}+\S_{gf}+\S_{K,M}+\S_{ext}) = \Delta^{cl}_{(\eta)} \ , \eqn{ward-susy} where the classical breaking $\Delta^{cl}_{(\eta)}$ split in linear and quadratic parts \begin{equation} \Delta^{cl}_{(\eta)} = \Delta^L_{(\eta)} + \Delta^Q_{(\eta)}. \eqn{classical-breaking} with, \begin{eqnarray} \Delta^L_{(\eta)} \= \in \Big[ -\gamma^{a\mu\nu}{\cal L}_\eta B^a_{\mu\nu} -\Omega^{a\mu}{\cal L}_\eta A^a_{\mu}+L^a{\cal L}_\eta c^a-D^a{\cal L}_\eta {\phi}} \newcommand{\F}{{\Phi}^a +\rho^{a\mu}{\cal L}_\eta \xi} \newcommand{\X}{\Xi^a_{\mu} - \nonumber \\ &-&\varepsilon_{\mu\nu\rho\sigma}\Omega^{a\mu} \eta^\nu s(\sqrt g g^{\rho\a}g^{\sigma\b}\partial_\a\bar\xi} \newcommand{\X}{\Xi^a_\b) - \varepsilon_{\mu\nu\rho\sigma}\gamma^{a\mu\nu} \eta^\rho s(\sqrt g g^{\sigma\a}\partial_\a\bar c^a) \Big] \eqan{linear-breaking} and \begin{eqnarray} \Delta^Q_{(\eta)} &=& \int d^4 x \Big[ - \varepsilon^{\mu\nu\rho\sigma} \renewcommand{\S}{\Sigma} s (g_{\mu\a}\eta^\a \partial_\nu \bar c^a \partial_\rho \bar \xi} \newcommand{\X}{\Xi^a_\sigma} \renewcommand{\S}{\Sigma) - s (\sqrt g g^{\mu\nu} \partial_\mu \bar {\phi}} \newcommand{\F}{{\Phi}^a \eta^\a B^a_{\a\nu}) - s (\sqrt g \bar {\phi}} \newcommand{\F}{{\Phi}^a {\cal L}_\eta e^a) - \nonumber \\ &-& s (f^{abc} \sqrt g g^{\mu\a} M^a_\mu \partial_\a \bar {\phi}} \newcommand{\F}{{\Phi}^b \eta^\nu \xi} \newcommand{\X}{\Xi^c_\nu) \Big] \ . \eqan{quadratic-breaking} The nonlinear breaking (\ref{quadratic-breaking}) is quadratic in the quantum fields, then it is not harmless in the context of the renormalization procedure.\\ To eliminate the nonlinear expression in (\ref{classical-breaking}) we first add to the action the BRS exact integral \begin{eqnarray} \S_1 &=& \int d^4 x \Big( -L_\mu \Y^\mu + R_\mu s \Y^\mu +W^\mu \Ps_\mu - Z^\mu s \Ps_\mu + I^a_\mu \T^{a\mu} - \nonumber \\ &-& J^a_\mu s \T^{a\mu} + P^a \Lambda^a + Q^a s \Lambda^a \Big) \eqan{lrlrlr} such that \begin{equation} \begin{array}{rcl} \Y^\mu &=& \varepsilon^{\mu\nu\rho\sigma} \renewcommand{\S}{\Sigma} \partial_\nu \bar c^a \partial_\rho \bar \xi} \newcommand{\X}{\Xi^a_\sigma} \renewcommand{\S}{\Sigma , \\ \Ps_\mu &=& \bar {\phi}} \newcommand{\F}{{\Phi}^a \partial_\mu e^a , \\ \T^{a\mu} &=& \sqrt g g^{\mu\rho} \partial_\rho \bar {\phi}} \newcommand{\F}{{\Phi}^a , \\ \Lambda^a &=& f^{abc} \sqrt g g^{\mu\rho} M^b_\mu \partial_\rho \bar {\phi}} \newcommand{\F}{{\Phi}^c . \end{array} \eqn{jkhfglu} All new auxiliary fields introduced in (\ref{lrlrlr}) transform in a BRS doublets \begin{equation} \begin{array}{llll} s R_\mu &=& L_\mu , ~~~~&~~~~ s L_\mu =0, \\ s Z^\mu &=& W^\mu , ~~~~&~~~~ s W^\mu =0, \\ s J^a_\mu &=& I^a_\mu , ~~~~&~~~~ s I^a_\mu =0, \\ s Q^a &=& P^a , ~~~~&~~~~ s P^a =0. \end{array} \eqn{ha98asd} Furthermore, under the $\d_{(\eta)}$ operation, they transform as \begin{equation} \begin{array}{llll} \d_{(\eta)} R_\mu &=& g_{\mu\rho} \eta^\rho , ~~~~&~~~~ \d_{(\eta)} L_\mu = {\cal L}_\eta R_\mu - \hat g_{\mu\rho} \eta^\rho , \\ \d_{(\eta)} Z^\mu &=& - \sqrt g \eta^\mu , ~~~~&~~~~ \d_{(\eta)} W^\mu = {\cal L}_\eta Z^\mu + \eta^\mu s \sqrt g , \\ \d_{(\eta)} J^a_\mu &=& \eta^\nu B_{\mu\nu}^a , ~~~~&~~~~ \d_{(\eta)} I^a_\mu = {\cal L}_\eta J^a_\mu - s (\eta^\nu B_{\mu\nu}^a), \\ \d_{(\eta)} Q^a &=& \eta^\mu \xi} \newcommand{\X}{\Xi^a_\mu , ~~~~&~~~~ \d_{(\eta)} P^a = {\cal L}_\eta Q^a - s (\eta^\mu \xi} \newcommand{\X}{\Xi^a_\mu). \end{array} \eqn{834oihuer} \begin{table}[h] \renewcommand{\arraystretch}{1.2} \begin{center} \begin{tabular}{|c|c|c|c|c|c|c|c|c|}\hline & $L_\mu$ & $R_\mu$ & $W^\mu$ & $Z_\mu$ & $I^a_\mu$ & $J^a_\mu$ & $P^a$ & $Q^a$ \\ \hline dim &-1 &-1 &-1 &-1 &1 &1 &0 &0 \\ \hline $\phi\pi$ &2 &1 &2 &1 &2 &1 &3 &2 \\ \hline weight &0 &0 &1 &1 &0 &0 &0 &0 \\ \hline \end{tabular} \\ \vspace{0.5cm} \small{Table 1: Dimensions, ghost numbers and weights of auxiliary fields.} \end{center} \end{table} \noindent For all auxiliary fields, the anticommutator of the BRS operator $s$ and $\d_{(\eta)}$ closes on the Lie derivative plus equations of motion.\\ On the other hand, the Ward identity (\ref{ward-susy}) gets promoted to \begin{equation} {\cal W}^S_{(\eta)}(\S) = \bar {\cal W}^S_{(\eta)}(\S) + {\cal V}^S_{(\eta)}(\S) = \Delta^L_{(\eta)}, \eqn{ljkafg} where ${\cal V}^S_{(\eta)}(\S)$ is nonlinear, and its expression is \begin{eqnarray} {\cal V}^S_{(\eta)}(\S) &=& \int d^4x \Bigg( g_{\mu\rho}\eta^\rho \ds{R_\mu} + ({\cal L}_\eta R_\mu - \hat g_{\mu\rho} \eta^\rho) \ds{L_\mu} - \sqrt g \eta^\mu \ds{Z^\mu} + \eta^\nu B^a_{\mu\nu} \ds{J^a_\mu} + \nonumber \\ &+& ({\cal L}_\eta Z^\mu + \eta^\mu s \sqrt g) \ds{W^\mu} + ({\cal L}_\eta J^a_\mu - \eta^\nu \ds{\gamma} \newcommand{\G}{\Gamma^{a\mu\nu}})\ds{I^a_\mu} + \eta^\mu \xi} \newcommand{\X}{\Xi^a_\mu \ds{Q^a} + \nonumber \\ &+& ({\cal L}_\eta Q^a - \eta^\mu \ds{\rho^{a\mu}} )\ds{P^a} \Bigg) \eqan{jklerliertkljdfgnmf} The complete gauge fixed action is now given by \begin{equation} \Sigma=\S_{inv}+\S_{gf}+\S_{K,M}+\S_{ext} + \S_1 . \eqn{action-classical} It obeys the Slavnov identity \begin{equation} {\cal S}(\Sigma) = 0 \ , \eqn{slavnov} where \begin{eqnarray} {\cal S}(\Sigma)\= \in \Bigg(\ds{\gamma^{a\mu\nu}}\ds{B^a_{\mu\nu}} +\ds{\Omega^{a\mu}}\ds{A^a_{\mu}}+\ds{L^a}\ds{c^a} +\ds{D^a}\ds{{\phi}} \newcommand{\F}{{\Phi}^a}+\ds{\rho^{a\mu}}\ds{\xi} \newcommand{\X}{\Xi^a_{\mu}}+ \nonumber \\ &+&b^a \ds{\bar c^a}+h^a_\mu \ds{\bar\xi} \newcommand{\X}{\Xi^a_\mu} +\omega^a \ds{\bar {\phi}} \newcommand{\F}{{\Phi}^a}+\lambda^a \ds{e^a} + \hat g_{\mu\nu} \ds{g_{\mu\nu}} + K_\mu^a \ds{M_\mu^a}+ L_\mu \ds{R_\mu} + \nonumber \\ &+& W^\mu \ds{Z^\mu} + I^a_\mu \ds{J^a_\mu} + P^a \ds{Q^a} \Bigg) \ . \eqan{slavnov-identity} It is straightforward to verify that the corresponding linearized Slavnov operator is given by \begin{eqnarray} {\cal S}_\Sigma \= \in \Bigg(\ds{\gamma^{a\mu\nu}}\dd{B^a_{\mu\nu}} +\ds{B^a_{\mu\nu}}\dd{\gamma^{a\mu\nu}} +\ds{\Omega^{a\mu}}\dd{A^a_{\mu}}+\ds{A^a_{\mu}}\dd{\Omega^{a\mu}}+ \nonumber \\ &+&\ds{L^a}\dd{c^a}+\ds{c^a}\dd{L^a} +\ds{D^a}\dd{{\phi}} \newcommand{\F}{{\Phi}^a}+\ds{{\phi}} \newcommand{\F}{{\Phi}^a}\dd{D^a} +\ds{\rho^{a\mu}}\dd{\xi} \newcommand{\X}{\Xi^a_{\mu}}+\ds{\xi} \newcommand{\X}{\Xi^a_{\mu}}\dd{\rho^{a\mu}}+ \nonumber \\ &+&b^a \dd{\bar c^a}+h^a_\mu \dd{\bar\xi} \newcommand{\X}{\Xi^a_\mu} +\omega^a \dd{\bar {\phi}} \newcommand{\F}{{\Phi}^a}+\lambda^a \dd{e^a} + \hat g_{\mu\nu} \dd{g_{\mu\nu}} + K^a_\mu \frac{\d}{\d M^a_\mu} + L_\mu \frac{\d}{\d R_\mu} + \nonumber \\ &+& W^\mu \frac{\d}{\d Z^\mu} + I^a_\mu \frac{\d}{\d J^a_\mu} + P^a \frac{\d}{\d Q^a} \Bigg) \eqan{linear-slavnov-operator} At the functional level, the invariance of the classical action (\ref{action-classical}) under diffeomorphisms can be expressed by an unbroken Ward identity \begin{equation} {\cal W}^D_{(\varepsilon)} \Sigma = 0 \ , \eqn{ward-diff} where ${\cal W}^D_{(\varepsilon)}$ denotes the corresponding Ward operator, \begin{equation} {\cal W}^D_{(\varepsilon)} = \in \sum_{f} \big( {\cal L}_\varepsilon f \big) \dd{f} \ , \eqn{ward-operator-diff} for all fields $f$. The vector parameter of the diffeomorphism transformations is denoted by $\varepsilon^\mu$, and it carries ghost number $+1$.\\ Next, we display the complete nonlinear algebra of the Slavnov operator and the Ward operator ${\cal W}^D_{(\varepsilon)}$. To this end, let $\Gamma$ be an arbitrary functional depending on the fields of the model, then \begin{eqnarray} {\cal S}_\Gamma {\cal S}(\Gamma) \= 0 \ , \nonumber \\ {\cal S}_\Gamma {\cal W}^D_{(\varepsilon)} \Gamma + {\cal W}^D_{(\varepsilon)}{\cal S}(\Gamma) \= 0 \ , \nonumber \\ \left\{}\newcommand{\rac}{\right\} {\cal W}^D_{(\varepsilon)} , {\cal W}^D_{(\varepsilon')} \rac \Gamma \= -{\cal W}^D_{(\left\{}\newcommand{\rac}{\right\}\varepsilon,\varepsilon'\rac)} \Gamma \ . \eqan{nonlinear-algebra1} Now if the functional $\Gamma$ is a solution of the Slavnov identity and of the Ward identity of diffeomorphisms, then the off--shell algebra \equ{nonlinear-algebra1} reduces to the linear algebra \begin{eqnarray} {\cal S}_\Sigma{\cal S}_\Sigma \= 0 \ , \nonumber \\ \left\{}\newcommand{\rac}{\right\} {\cal S}_\Sigma , {\cal W}^D_{(\varepsilon)} \rac \= 0 \ , \nonumber \\ \left\{}\newcommand{\rac}{\right\} {\cal W}^D_{(\varepsilon)} , {\cal W}^D_{(\varepsilon')} \rac \= -{\cal W}^D_{(\left\{}\newcommand{\rac}{\right\}\varepsilon,\varepsilon'\rac)} \ , \nonumber \\ \eqan{off-shell-algebra} with the Lie brackets \begin{equation} \{\varepsilon,\varepsilon'\}^\mu = {\cal L}_\varepsilon\e'^\mu \eqn{lie-def} Contrary to other topological field theories \cite{zer2} \cite{zer1}, the anticommutator of the linearized Slavnov operator (\ref{linear-slavnov-operator}) with the linearized\footnote{First one has to linearize ${\cal W}^S_{(\eta)}$ and then let the anticommutator $\lbrace {\cal S}_\S, {\cal W}^S_{(\eta)} \rbrace$ acting on the different fields.} ${\cal W}^S_{(\eta)}$ Ward operator does not give rise to the diffeomorphisms Ward operator. Indeed, for certain fields we get $\lbrace {\cal S}_\S, {\cal W}^S_{(\eta)} \rbrace = {\cal W}^D_{(\varepsilon)}$ and for other fields the anticommutator does not close on ${\cal W}^D_{(\varepsilon)}$. As a simple example, one can let the above anticommutator acting on the two auxiliary fields $M^a_\mu$ and $K^a_\mu$.\\ The main result of this paper is that we could, after introducing auxiliary fields, construct a local and nonlinear symmetry of the $4D$ antisymmetric tensor field model in a curved manifold. The next natural step to do is to show the renormalizability of the new symmetry. It turns out that if this symmetry is valid at the quantum level then it would be very useful in showing the finiteness of the model in a big class of curved manifolds. This would be the generalisation of the results of \cite{zer3}. \section*{Acknowledgment} I would like to thank the ``Fonds zur F\"orderung der Wissenschaftlichen Forschung'' for the financial support: contract grant number P11582-PHY. \newpage
\section{Cosmological constraints from {\sf JVAS}\dots} The Jodrell Bank-VLA Astrometric Survey ({\sf JVAS}) is a survey for flat-spectrum radio sources with a flux density greater than 200\,mJy at 5\,GHz. Flat-spectrum radio sources are likely to be compact, thus making it easy to recognise the lensing morphology. In addition, they are likely to be variable, making it possible to determine $H_{0}$ by measuring the time delay between the lensed images. (See \cite{ABiggsBHKWP99a} for the description of a time delay measurement in a {\sf JVAS}\ gravitational lens system.) {\sf JVAS}\ is also a survey for MERLIN phase-reference sources and as such is described in \cite{APatnaikBWW92a}, \cite{IBRownePWW98a} and \cite{PWilkinsonBPWS98a}. {\sf JVAS}\ as a gravitational lens survey, the lens candidate selection, followup process, confirmation criteria and a discussion of the {\sf JVAS}\ gravitational lenses is described in detail in \cite{LKingBMPW99a} (see also \cite{LKingIBrowne96a}). In order to have a parent sample which is as large as possible and as cleanly defined as practical, our `{\sf JVAS}\ gravitational lens survey sample' is slightly different than the `{\sf JVAS}\ phase-reference calibrator sample'. For the former, the source must be a point source and must have a good starting position (so that the observation was correctly pointed) while its precise spectral index is not important. For the latter, only the spectral index is important, as the source can be slightly resolved or the observation can be less than perfectly pointed. Thus, the {\sf JVAS}\ astrometric sample \cite{APatnaikBWW92a,IBRownePWW98a,PWilkinsonBPWS98a} contains 2144 sources. To these must be added 103 sources which were too resolved to be used as phase calibrators and 61 sources which had bad starting positions (thus the observations were too badly pointed to be useful for the astrometric sample), bringing the total to 2308. This formed our gravitational lens sample, since these additional sources were also searched for gravitational lenses \cite{LKingBMPW99a} (none were found meeting the {\sf JVAS}\ selection criteria: multiple flat-spectrum point-source components with a separation between 300 mas and 6 arcsec with a flux ratio of $\le$20). We have used the gravitational lens systems in Table~\ref{ta:lenses} in this analysis. \begin{table}[h] \begin{center} \begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}llllll} \hline \hline Name & \# images & $\Delta\theta ['']$ & $z_{\mathrm{l}}$ & $z_{\mathrm{s}}$ & lens galaxy \\ \hline B0218+357 & 2 + ring & 0.334 & 0.6847 & 0.96 & spiral \\ MG0414+054 & 4 & 2.09 & 0.9584 & 2.639 & elliptical \\ B1030+074 & 2 & 1.56 & 0.599 & 1.535 & spiral \\ B1422+231 & 4 & 1.28 & 0.337 & 3.62 & \textbf{?} \\ \hline \end{tabular*} \caption[]{ {\sf JVAS}\ lenses used in this analysis. Of the information in the table, for this analysis we use only the source redshift $z_{\mathrm{s}}$, and the image separation $\Delta\theta$.} \label{ta:lenses} \end{center} \end{table} The {\sf JVAS}\ lens B1938+666 \cite{LKingJBBBdBFKMNW98a} was not included because it is not formally a part of the sample, having a too steep spectral index and having been recognised on the basis of a lensed extended source as opposed to lensed compact components. Also, the {\sf JVAS}\ lens B2114+022 \cite{PAugustoBWJFM99a} was not included because it is not a single-galaxy lens system. For this analysis, due to the paucity of the observational data, we have made rather stark assumptions: the redshift distribution of the sample is assumed to be identical to that of the CJF (Caltech-Jodrell Bank Flat-spectrum) sample \cite{GTaylorVRPHW96a}, independent of flux density, and the number-magnitude relation is assumed to be identical to that of {\sf CLASS}\, the Cosmic Lens All-Sky Survey, \cite{SMyersetal99a}, independent of redshift. Otherwise, we have calculated the likelihood as a function of $\lambda_{0}$ and $\Omega_{0}$ as described in \cite{CKochanek96a}. That is, we use a non-singular isothermal sphere as a lens model, model the lens galaxy population with a Schechter function and use the Faber-Jackson relation to convert between luminosity and velocity dispersion, considering only elliptical galaxies. The results are presented in Fig.~\ref{fi:jvas}. \begin{figure}[t] \centering \epsfig{figure=13_phelbig1_1.ps, width=9cm} \caption[]{ The likelihood function $p(\lambda_0,\Omega_0)$ based on the {\sf JVAS}\ lens sample. The pixel grey level is directly proportional to the likelihood: darker pixels reflect higher likelihoods. The pixel size reflects the resolution of our numerical computations. The contours mark the boundaries of the minimum $0.68$, $0.90$, $0.95$ and $0.99$ confidence regions for the parameters $\lambda_0$ and $\Omega_0$.} \label{fi:jvas} \end{figure} At 95\% confidence, our lower and upper limits on $\lambda_{0}-\Omega_{0}$, using the {\sf JVAS}\ lensing statistics information alone, are respectively $-2.69$ and $0.68$. For a flat universe, these correspond to lower and upper limits on $\lambda_{0}$ of respectively $-0.85$ and $0.84$. (Reducing the constraints in the $\lambda_{0}$-$\Omega_{0}$ plane to $\lambda_{0}-\Omega_{0}$ is, of course, just an approximation, but a reasonably good one when considering upper limits on $\lambda_{0}$ for small $\Omega_{0}$ values. These numbers were derived from the corresponding confidence limits on $\lambda_{0}$ for fixed $\Omega_{0}$ and are thus of course different than the intersection of lines of constant $\lambda_{0}-\Omega_{0}$ with the corresponding contour in Fig.~\ref{fi:jvas}.) \section{\dots and optical gravitational lens surveys\dots} One can improve these constraints by adding those from optical gravitational lens surveys, though one should keep in mind that the systematic errors---for example, lens systems which are missed due to extinction in the lens galaxy or to the fact that the typical seeing is not much better than the typical image separation---are probably less well understood than is the case in the radio (though the statistical properties of the unlensed parent population are better understood). Not only does one have more objects and thus better statistics, but a different redshift range is sampled as well. Essentially repeating the analysis in \cite{CKochanek96a} (but with $\lambda_{0}$ and $\Omega_{0}$ as free parameters, of course) and combining the resulting constraints with those from Fig.~\ref{fi:jvas}, one obtains the better constraints shown in Fig.~\ref{fi:joint}. \begin{figure}[t] \centering \epsfig{figure=13_phelbig1_2.ps, width=9cm} \caption[]{ The same as Fig.~\ref{fi:jvas} but combining {\sf JVAS}\ with optical gravitational lens surveys from the literature.} \label{fi:joint} \end{figure} Using the combination of {\sf JVAS}\ lensing statistics and lensing statistics from the literature as in \cite{CKochanek96a}, the corresponding $\lambda_{0}-\Omega_{0}$ values are $-1.78$ and $0.27$. For a flat universe, these correspond to lower and upper limits on $\lambda_{0}$ of respectively $-0.39$ and $0.64$. \section{\dots and `reasonably well-accepted wisdom'\dots} Gravitational lensings statistics alone cannot usefully constrain $\Omega_{0}$. Thus, it seems sensible to combine the constraints shown in Fig.~\ref{fi:joint} with measurements of $\Omega_{0}$. Fortunately, there seems to be a consensus developing that $\Omega_{0} \approx 0.3$ e.g.~\cite{RCarlbergetal97a,RCarlberg98a,RCarlbergetal98b, NBahcall97a,NBahcallFC97a,XFanBC97a,MBartelmannHCJP98a, SPerlmutteretal98a, ARiessetal98a,BSchmidtetal98a,AKim98a,CLineweaver98a,EGuerraDW98a, RDalyGW98a}. Conservatively, these results can be summarised as \begin{equation} p(\lambda_0,\Omega_0) = L(\Omega_0|0.4,0.2). \label{eq:p2} \end{equation} where the two arguments of $L$ represent the mean and standard deviation of a lognormal distribution. In a similar vein, lensing statistics determines a lower limit on $\lambda_{0}$ much less strongly than an upper limit, so it seems sensible to include some prior information which can give a lower limit on $\lambda_{0}$. To be conservative, we take relatively undisputed estimates for the age of the universe and the Hubble constant, their product setting a (slightly $\Omega_{0}$-dependent) lower limit on $\lambda_{0}$. The best estimate of the absolute age of the oldest galactic globular clusters currently is $t_{\mathrm{gc}}=11.5\pm1.3\,\mbox{Gyr}$ \cite{BChaboyerDKK98a}. We choose to formulate this prior information in the form of a lognormal distribution that meets these statistics \begin{equation} p(t_{\mathrm{gc}}) = L(t_{\mathrm{gc}}|11.5\,\mbox{Gyr}, 1.3\,\mbox{Gyr}). \label{eq:ptgc} \end{equation} Similarly, we roughly estimate $H_0=65\pm10\,\mbox{km}\,\mbox{s}^{-1}\,\mbox{Mpc}^{-1}$ and choose to formulate this prior information in form of a normal distribution \begin{equation} p(H_0) = N(H_0|65\,\mbox{km}\,\mbox{s}^{-1}\,\mbox{Mpc}^{-1}, 10\,\mbox{km}\,\mbox{s}^{-1}\,\mbox{Mpc}^{-1}), \label{eq:ph0} \end{equation} where, again, the notation for $L$ (and $N$) is such that the two arguments correspond to the mean and standard deviation. Fig.~\ref{fi:posterior} shows how inclusion of this prior information, representing a conservative estimate of what we know about the values of the cosmological parameters, tightens the constraints on $\lambda_{0}$ and $\Omega_{0}$ as compared to the constraints from lens statistics alone (Figs.~\ref{fi:joint} and \ref{fi:jvas}). \begin{figure}[t] \centering \epsfig{figure=13_phelbig1_3.ps, width=9cm} \caption[]{ The same as Fig.~\ref{fi:joint} but combining {\sf JVAS}\ and optical gravitational lens surveys from the literature with prior information on the value of $\Omega_{0}$, $H_{0}$ and the age of the universe. This figure thus represents the combination of constraints from lensing statistics and from relatively undisputed knowledge about values of the cosmological parameters.} \label{fi:posterior} \end{figure} \section{\dots and the CMB\dots} It has long been realised, e.g.~\cite{DEisensteinHT98b}, that the direction of degeneracy of constraints from cosmic microwave background anisotropies is roughly orthogonal to that of most other tests, including lensing statistics. Thus, combining the constraints from CMB anisotropies with those from other cosmological tests can give much tighter constraints than either alone. We have performed an analysis similar to that done in \cite{CLineweaver98a}, though including an updated Tenerife data point and calculating two-dimensional joint likelihood constraints as in the calculations done previously in this poster rather than employing Lineweaver's statistical method. Adding the constraints on $\lambda_{0}$ and $\Omega_{0}$ so derived to the previous ones narrows down the region of parameter space further, as is shown in Fig.~\ref{fi:cmb}. \begin{figure}[t] \centering \epsfig{figure=13_phelbig1_4.ps, width=9cm} \caption[]{ The same as Fig.~\ref{fi:posterior} but combining {\sf JVAS}\, optical gravitational lens surveys from the literature and prior information on the value of $\Omega_{0}$, $H_{0}$ and the age of the universe with constraints derived from CMB anisotropies. Since the CMB constraints are more or less orthogonal to the lensing statistics constraints, this reduces the allowed area of the $\lambda_{0}$-$\Omega_{0}$ parameter space significantly. Note that the scale of this plot differs from the previous ones. For technical reasons no models with $k=\mathop{\rm sign}\nolimits(\lambda_{0}+\Omega_{0}-1)=+1$ were calculated; this slightly distorts the contours near the $k=0$ line, which otherwise would extend a bit more into the $k=+1$ region.} \label{fi:cmb} \end{figure} The power spectrum for the best-fit model using the CMB data alone ($\lambda_{0}=0.6$, $\Omega_{0}=0.3$, otherwise not shown here) along with various data points from the literature (see the collection of Max Tegmark at \texttt{http://www.sns.ias.edu/\symbol{126}max/cmb/experiments.html}) is shown in Fig.~\ref{fi:bestfit}. \begin{figure}[t] \centering \epsfig{figure=13_phelbig1_5.ps, width=9cm} \caption[]{ Power spectrum shown with data points from the literature for our best-fit cosmological model, fitting to the CMB data alone (and keeping parameters other than $\lambda_{0}$ and $\Omega_{0}$ fixed at predetermined fiducial values).} \label{fi:bestfit} \end{figure} (Note added for the proceedings. The likelihood based on CMB observations which (combined with other tests) is shown in Fig.~\ref{fi:cmb} is, due to a numerical error, qualitatively but not quantitatively correct. We have since performed the correct computation, which will be presented elsewhere. However, the difference creates a smaller error than that caused by many other approximations made use of in these calculations, so we present the figure as shown in the original poster, in keeping with the concept of providing a record of the conference as opposed to an updated paper on the same subject.) \section{\dots and how this compares to other cosmological tests} Fig.~\ref{fi:cmb} combines constraints based on optical and radio gravitational lensing statistics, `direct' measurements of $H_{0}$, $\Omega_{0}$ and the age of the universe and constraints derived from CMB anisotropies. This restricts $\lambda_{0}$ to a narrow range. If one believes that $\Omega_{0}\approx 0.3$, then it follows that $\lambda_{0}\approx 0.5$. This should be compared to the result of Perlmutter et al.~\cite{SPerlmutteretal99a}: $0.8\Omega_{0} - 0.6 \lambda_{0} \approx -0.2$: inserting $0.3$ for $\Omega_{0}$ one obtains $\lambda_{0}\approx0.43$. Taking the errors into consideration (which are not large enough in either case to allow, for example, $\lambda_{0} = 0$) one obtains perfectly consistent measurements of $\lambda_{0}$ from completely independent methods. Taken together, present measurements of cosmological parameters \emph{definitely} rule out the Einstein-de~Sitter universe ($\lambda_{0}=0$, $\Omega_{0}=1$), \emph{very probably} rule out a universe without a cosmological constant ($\lambda_{0}=0$) and \emph{tentatively} rule out a flat ($\lambda_{0}$ + $\Omega_{0}$ = 1) universe as well. A universe with $\lambda_{0}\approx 0.4$ and $\Omega_{0}\approx 0.3$ seems to be consistent with all observational data, including measurements of the Hubble constant and age of the universe. \section{The future} {\sf CLASS}\ is similar to {\sf JVAS}\ but contains about 4 times as many sources. The definition of both is flat-spectrum between L-band and C-band, i.e.~$\alpha>-0.5$ where $s_{f}\sim f^{\alpha}$, the essential difference being the lower flux density limit of 200\,mJy for {\sf JVAS}\ and 30\,mJy for {\sf CLASS}. However, since {\sf CLASS}\ is defined based on the newer GB6 and NVSS catalogues \cite{PGregorySDJ96a,JCondonCGYPTB98a} than {\sf JVAS}, there will be some essentially random differences due to differing quality of observations and variability of the sources. All the {\sf JVAS}\ lenses mentioned in Table~\ref{ta:lenses} are in the new {\sf CLASS}\ sample, which, having no upper flux density limit, subsumes {\sf JVAS}. The previous samples {\sf CLASS}-I and {\sf CLASS}-II will be similarly subsumed in the same sense as {\sf JVAS}, though the differences here will be slightly larger since bands other than L and C were used in the preliminary definition of these samples. The initial phase of observations is complete; currently lens candidates are being followed up. At present, we have confirmed as gravitational lenses the systems listed in Table~\ref{ta:class} (which for completeness also includes the two {\sf JVAS}\ lens systems not used in the statistical analysis presented here). \begin{table}[h] \begin{center} \begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}llllll} \hline \hline Name & \# images & $\Delta\theta''$ & $z_{\mathrm{l}}$ & $z_{\mathrm{s}}$ & lens galaxy \\ \hline B0712+472 & 4 & 1.27 & 0.406 & 1.34 & spiral \\ B1127+385 & 2 & 0.70 & \textbf{?} & \textbf{?} & \textbf{?} \\ B1600+434 & 2 & 1.39 & 0.414 & 1.589 & spiral \\ B1608+656 & 4 & 2.08 & 0.63 & 1.39 & spiral \\ B1933+507 & $4+4+2$ & 1.17 & 0.755 & \textbf{?} & \textbf{?} \\ B1938+666 & $4+2$ & 0.93 & \textbf{?} & \textbf{?} & \textbf{?} \\ B2045+265 & 4 & 1.86 & 0.867 & 1.28 & \textbf{?} \\ B2114+022 & 2 or 4 & 2.57 & 0.32 \& 0.59 & \textbf{?} & \textbf{?} \\ \hline \end{tabular*} \caption[]{ {\sf CLASS}\ gravitational lenses and the two {\sf JVAS}\ lens systems (1938+666 and 2114+022) not listed in Table~\ref{ta:lenses}.} \label{ta:class} \end{center} \end{table} We hope that the larger size of {\sf CLASS}\ will allow the constraints on cosmological parameters from gravitational lensing statistics to improve. At present, the greatest uncertainty is the redshift-dependent luminosity function (or equivalently the flux-dependent redshift distribution) of the unlensed population (which of course, due to the amplification bias, extends to fainter flux-densities than the survey itself). We are currently taking steps to decrease this uncertainty. \section*{Acknowledgements} {\centering \epsfig{figure=13_phelbig1_6.ps, width=9cm} } \noindent This research was supported in part by the European Commission, TMR Programme, Research Network Contract ERBFMRXCT96-0034 `CERES'. {\sf CERES}, standing for the Consortium for European Research on Extragalactic Surveys, is an EU TMR Network, coordinated by Ian Browne at Jodrell Bank, and involving the Nuffield Radio Astronomy Laboratories of the Universtity of Manchester at Jodrell Bank, the Institute of Astronomy of the University of Cambridge, The Kapteyn Astronomical Institute of the University of Groningen, the Netherlands Foundation for Research in Astronomy at Dwingeloo, the University of Bologna and the University of Portugal. One of the main goals of {\sf CERES}\ is the use of the {\sf JVAS}\ and {\sf CLASS}\ surveys for research in gravitational lensing. We thank our collaborators in the {\sf JVAS}, CJF and {\sf CLASS}\ surveys for useful discussions and for providing data in advance of publication and many colleagues at Jodrell Bank for helpful comments and suggestions. We also thank John Meaburn and Anthony Holloway at the Department of Astronomy in Manchester and the staff at Manchester Computing for providing us with additional computational resources. RQ is grateful to the {\sf CERES}\ collaboration for making possible a visit to Jodrell Bank where part of this work was done. \section*{References}
\section{INTRODUCTION} Although the study of GRBs (gamma-ray bursts) has been revolutionized in the past few years by finding a number of precise positions and distances, the nature of their progenitor objects remains uncertain (see M\'esz\'aros 1999 for a review). The production of a large amount of energy in a short time has naturally led to models involving compact objects (neutron stars and black holes). Fryer, Woosley, \& Hartmann (1999) have summarized possible progenitors involving black hole accretion disks: neutron star - neutron star binary mergers (NS/NS), black hole - neutron star mergers (BH/NS), black hole - white dwarf mergers (BH/WD), massive star core collapses, and black hole - helium star mergers (BH/He). Based on estimated formation rates and on accretion disk models with high viscous forces, Fryer et al. (1999) suggest that NS/NS and BH/NS mergers dominate the population of short-duration GRBs and that massive stars and BH/He mergers dominate the long-duration bursts. The afterglows observed to date would have massive star progenitors in this scenario because they followed long-duration bursts. One way of distinguishing between the progenitor models is to examine the position of the GRB in the parent galaxy. Paczy\'nski (1998) pointed out that NS/NS binaries would be expected to have a significant space velocity, which would carry them many kpc from their birthplaces. The observational evidence for the association of several GRBs with star forming regions then provided weak evidence against the NS/NS merger progenitors and favored massive star progenitors. The population synthesis calculations of Fryer et al. (1999) supported this conclusion. However, Bloom, Sigurdsson, \& Pols (1999a) found a time to NS/NS merger of $\sim 10^8$ years. These objects would then follow the star formation rate, although $\sim 15$\% of them might occur well outside of dwarf galaxy hosts. The connection of GRBs to massive stars became stronger with the discovery of the Type Ic supernova SN 1998bw in the error box of GRB 980425 (Galama et al. 1998). The high energy inferred for the optical supernova, $(2-3)\times 10^{52}$ ergs (Iwamoto et al. 1998; Woosley, Eastman, \& Schmidt 1999), and the high expansion velocity inferred for the radio supernova (Kulkarni et al. 1998) strengthen the GRB connection. Li \& Chevalier (1999) found that the evolution of the radio source indicated non-uniform energy input to the blast wave, as is also inferred in GRBs. They also found evidence that the radio SN 1998bw interacted with the stellar wind expected from the massive star progenitor. In this paper, we emphasize that a stellar wind environment is an unavoidable consequence of a massive star progenitor and that the nature of the GRB afterglow emission can provide a discriminant between massive star and compact binary progenitor models. In \S~2, we discuss the expected afterglow for a massive star progenitor model and in \S~3 place our models in the context of observations. Our discussion concentrates on the cases $s=2$ (stellar wind) and $s=0$ (interstellar medium), where the ambient medium has $\rho\propto r^{-s}$. \section{AFTERGLOWS IN A MASSIVE STAR WIND} In the existing massive star GRB progenitor models, the most likely progenitor is the stripped core of a massive star, i.e., a Wolf-Rayet star. MacFadyen \& Woosley (1999) consider single massive stars whose cores directly collapse to black holes. The stars have an initial mass $\raise0.3ex\hbox{$>$}\kern-0.75em{\lower0.65ex\hbox{$\sim$}} 25~M_\odot$, which is the type of star that is likely to lose its H envelope in winds. Paczy\'nski (1998) noted that the requirement of a rapidly rotating core might necessitate a close binary companion, which again points to Wolf-Rayet stars. The winds from these stars in our Galaxy have velocities of $1,000-2,500\rm ~km~s^{-1}$ and mass loss rates $\dot M\approx 10^{-5}-10^{-4}~\Msun ~\rm yr^{-1}$ (Willis 1991). On evolutionary grounds, Langer (1989) advocated $\dot M \sim 6\times 10^{-8}(M_{\rm WR}/~M_\odot)^{2.5}~\Msun ~\rm yr^{-1}$, where $M_{\rm WR}$ is the mass of the Wolf-Rayet star. If the stellar mass drops to $\sim 3~M_\odot$ at the end of its life because of mass loss, $\dot M\sim 10^{-6}~\Msun ~\rm yr^{-1}$ at that time. The host galaxies of GRBs are likely to be of low metallicity. Willis (1991) notes that there is no evidence for a metallicity dependence of mass loss from a particular type of Wolf-Rayet star, but that metallicity may affect the distribution of Wolf-Rayet types. The effects of Wolf-Rayet winds can be observed in the wind bubbles around some of these objects. In some cases where the bubble is well-studied, the bubble expansion and X-ray emission are consistent with a weaker wind than that inferred from direct observations of the Wolf-Rayet star (e.g., NGC 6888, Garc\'ia-Segura \& Mac Low 1995). The position of the wind termination shock, $r_t$, can be estimated by equating the ram pressure in the wind with the pressure in the bubble. We estimate $r_t/R_b\approx (2R_b/v_wt_w)^{1/2}$, where $R_b$ is the radius of the wind bubble and $t_w$ is its age. For $R_b=3$ pc, $t_w=2\times 10^4$ yr, and $v_w=1,000\rm ~km~s^{-1}$, we have $r_t\approx 5\times 10^{18}$ cm. At $r_t$, the density increases by a factor of 4 and becomes approximately constant at larger radius. The radius $r_t$ is expected to increase as the bubble evolves and its pressure drops. At the end of its life, a Wolf-Rayet star is thus expected to be surrounded by a substantial $\rho\propto r^{-2}$ medium. The radio observations of Type Ib/c supernovae can be interpreted as interactions with such a medium in models with constant efficiencies of production of relativistic electrons and magnetic fields (Weiler et al. 1999; Chevalier 1998). For the wind density, we take $\rho =Ar^{-2}$, where $A=\dot M/4\pi v_w=5\times 10^{11}A_{\star}$ g cm$^{-1}$. The reference value of $A$ corresponds to $\dot M=1\times 10^{-5}~\Msun ~\rm yr^{-1}$ and $v_w=1000\rm ~km~s^{-1}$. If a burst were able to occur in a red supergiant star, the low wind velocity, $v_w\approx 10\rm ~km~s^{-1}$, would lead to a higher circumstellar density because $\dot M\raise0.3ex\hbox{$>$}\kern-0.75em{\lower0.65ex\hbox{$\sim$}} 10^{-7}~\Msun ~\rm yr^{-1}$ up to $\sim 10^{-4}~\Msun ~\rm yr^{-1}$ is expected. The only supernova case where we have inferred a lower circumstellar density is around SN 1987A, which exploded as a B3 I star and was thus too cool to drive a strong wind with its ultraviolet radiation field. In this case, the $r^{-2}$ wind terminated at $\sim (3-4)\times 10^{17}$ cm and the supernova radio flux rose sharply when the shock front encountered denser gas (Ball et al. 1995; Chevalier 1998). The scaling laws that are appropriate for GRB interaction with an $s=2$ medium have been described by M\'esz\'aros, Rees, \& Wijers (1998) and by Panaitescu, M\'esz\'aros, \& Rees (1998). Our aim here is to examine the specific predictions for interaction with a Wolf-Rayet star wind. We calculate the expected emission for a thin shell model (cf. Li \& Chevalier 1999). We adopt a nucleon-to-electron number density ratio of two, which is appropriate for winds of Wolf-Rayet stars that are predominantly helium (and perhaps carbon/oxygen). For the mass loss case, a density of $1.67\times 10^{-24}\rm ~g~cm^{-3}$ is attained at $r\approx 5.5\times 10^{17}A_{\star}^{1/2}$ cm. For an adiabatic blast wave in an $s=2$ medium (Blandford \& McKee 1976), we find $\gamma^2 =R/4ct$ and $R=(9Et/4\pi Ac)^{1/2}= 2.0\times 10^{17} E_{52}^{1/2} A_{\star}^{-1/2}t_{\rm day}^{1/2}$ cm, where $\gamma$ is the Lorentz factor of the gas, $R$ is the observed radius near the line of sight, $E_{52}=E/10^{52}$ ergs is the explosion energy, and $t=t_{\rm day} {\rm ~day}$ is the time in the observer's frame. Over the typical time of observation of a GRB afterglow, the shock front is expected to be within the $s=2$ wind, except for the unusual case of a progenitor like that of SN 1987A. In our model, the synchrotron emission frequency of the lowest energy electrons is \begin{equation} \nu_m= 5\times 10^{12}\left(1+z\over 2\right)^{1/2} (\epsilon_e/0.1)^2(\epsilon_B/0.1)^{1/2}E_{52}^{1/2} t_{\rm day}^{-3/2} {\rm~Hz}, \end{equation} and the flux at this frequency is \begin{equation} F_{\nu_m}= 20\left(\sqrt{1+z}-1\over \sqrt{2}-1\right)^{-2} \left(1+z\over 2\right)^{1/2} (\epsilon_B/0.1)^{1/2} E_{52}^{1/2} A_{\star} t_{\rm day}^{-1/2} {\rm~mJy}, \end{equation} where $z$ is the redshift, and $\epsilon_e$ and $\epsilon_B$ are the electron and magnetic postshock energy fractions. These expressions assume a flat universe with $H_o=65$ km s$^{-1}$ Mpc$^{-1}$. As in discussions of afterglows in the ISM, the magnetic field is assumed to be amplified by processes in the shocked region and is not directly determined by the field in the Wolf-Rayet star wind. For electrons with a power law energy distribution above $\gamma_{em}$, $dN_e/d\gamma_e\propto \gamma_e^{-p}$, the flux above $\nu_m$ is $F_{\nu}=F_{\nu_m}(\nu/\nu_m)^{-(p-1)/2} \propto \epsilon_e^{p-1}\epsilon_B^{(p+1)/4} E^{(p+1)/4} A t^{-(3p-1)/4} \nu^{-(p-1)/2}$, provided the electrons are not radiative. $F_{\nu}=F_{\nu_m}(\nu/\nu_m)^{1/3}$ below $\nu_m$ until the spectrum turns over due to synchrotron self-absorption at \begin{equation} \nu_A\approx 1\times 10^{11} \left(1+z\over 2\right)^{-2/5} (\epsilon_e/0.1)^{-1}(\epsilon_B/0.1)^{1/5}E_{52}^{-2/5}A_{\star}^{6/5} t_{\rm day}^{-3/5} {\rm~Hz}. \end{equation} We estimate the effects of synchrotron cooling following the discussions of Sari, Piran, \& Narayan (1998) and Wijers \& Galama (1999) for the $s=0$ case. Radiative cooling becomes important at a frequency \begin{equation} \nu_c\approx 1\times 10^{12}\left(1+z\over 2\right)^{-3/2} (\epsilon_B/0.1)^{-3/2}E_{52}^{1/2}A_{\star}^{-2} t_{\rm day}^{1/2} {\rm~Hz}, \end{equation} provided $\nu_c>\nu_m$. For $\nu>\nu_c$, $F_{\nu}\propto \epsilon_e^{p-1}\epsilon_B^{(p-2)/4} E^{(p+2)/4} t^{-(3p-2)/4} \nu^{-p/2}$. These results can be compared to similar results for interaction with a constant density interstellar medium with $n_o\approx 1$ cm$^{-3}$ (e.g., Waxman 1997; Sari et al. 1998). In that case $F_{\nu_m}$ is constant with time, $\nu_m$ is comparable to the value given above, $\nu_A$ is independent of time, and $\nu_c$ decreases with time. If $\nu_A$, $\nu_m$, and $\nu_c$ have the same relation to each other in the $s=0$ and $s=2$ cases, the appearance of the spectrum at one time is similar for both cases, but the evolution is different. At high frequency (optical and X-ray), for $s=0$ the flux evolution goes from adiabatic ($\propto t^{-(3p-3)/4}$) to cooling ($\propto t^{-(3p-2)/4}$) while for $s=2$ the flux evolution goes from cooling ($\propto t^{-(3p-2)/4}$) to adiabatic ($\propto t^{-(3p-1)/4}$). While cooling, the two cases have the same spectrum and flux decline. At low frequency (radio) with $\nu<\nu_m$, the flux evolves as $t^{1/2}$ for $s=0$, but can make a transition from $\propto t$ to constant for $s=2$. \section{COMPARISON WITH OBSERVATIONS} We have noted the evidence that SN 1998bw is interacting with a circumstellar wind (Li \& Chevalier 1999), and here discuss the cosmological GRBs. The more detailed models that have been developed for comparison with the best observed GRB afterglows have taken a low, constant density surrounding medium as the starting point. For GRB 970508, Wijers \& Galama (1999) and Granot, Piran, \& Sari (1999) found $n_o=0.030, 5.3$ cm$^{-3}$, $E_{52}= 3.5, 0.53$, $\epsilon_e=0.12, 0.57$, and $\epsilon_B=0.089,0.0082$, respectively. The difference between these results reflects, in part, the difficulty of accurately determining the significant frequencies, such as $\nu_A$. The low density found for this case is indicative of an interstellar density. However, these models primarily depend on the spectrum at one time, which can be the same for the $s=0$ and $s=2$ cases. Panaitescu et al. (1998) have presented a successful $s=0$ model for the evolution of GRB 970508, but a number of deviations from the simple model are included, so the case is not clear. With $s=0$, the optically thin, adiabatic flux evolution can be described by $F_{\nu}\propto t^{\alpha}\nu^{\beta}$ with $\alpha=1.5\beta$. This type of evolution was observed in the afterglow of GRB 970228 (Wijers, Rees, \& M\'esz\'aros 1997). For $s=2$, the expected power law evolution has $\alpha=(3\beta-1)/2$ for a constant energy blast wave, so a mechanism must be found to flatten the time evolution if this GRB evolved in a circumstellar wind. The effects of beaming and a change to nonrelativistic flow steepen the time evolution, but continued power input from ejecta can flatten it. This effect is inferred to occur in radio supernovae (Chevalier 1998). Following Rees \& M\'esz\'aros (1998) for the $s=0$ case with power input from ejecta with a mass gradient $M(\raise0.3ex\hbox{$>$}\kern-0.75em{\lower0.65ex\hbox{$\sim$}} \Gamma_f)\propto \Gamma_f^{-n}$ where $\Gamma_f$ is the Lorentz factor in the freely expanding ejecta, the evolution follows $F_{\nu}\propto t^{-[2-\beta (n+6)]/(n+4)}$ for the $s=2$ case. For $\beta=-0.75$, the property $\alpha=1.5\beta$ is recovered when $n=5.33$. The optically thin, $\nu>\nu_m$ evolution would then be the same as in the constant density ($s=0$) case, but other aspects of the evolution would be different. In particular, the expansion would be much less decelerated and the apparent radius of the blast $r\propto t^{(n+2)/(n+4)}\propto t^{0.79}$ and its Lorentz factor $\gamma\propto t^{-1/(n+4)}\propto t^{-0.11}$. At optically thick wavelengths, the flux would increase as $F_{\nu}\propto t^{1.57}$ as opposed to $F_{\nu}\propto t^{1/2}$ for the constant density case. Radio observations could distinguish this kind of evolution, but GRB 970228 was not detected in the radio. Kulkarni et al. (1999) noted that the afterglow of GRB 990123 was consistent with $p=2.44$ in a $s=0$ model, where $p$ is the electron energy spectral index. The optical emission declined with $\alpha=-1.10\pm 0.03$, suggesting adiabatic evolution, while the X-rays declined with $\alpha=-1.44\pm 0.07$, suggesting cooling evolution. The faster decline at X-ray wavelengths is distinctive of $s=0$ evolution and is the opposite of expectations for $s=2$. Most of the well observed optical afterglows also have $\alpha$ in the range --1.1 to --1.3, which is plausibly modeled by blast wave evolution in a constant density medium. However, the possibilities of flat electron spectra or cooling evolution do not allow $s=2$ models to generally be ruled out. An afterglow with a steep decline was that of GRB 980519. Its optical emission followed $F_{\nu}\propto t^{-(2.05\pm 0.04)}$, which, with the observed $\beta=-1.05\pm 0.10$ for optical and X-ray data, is consistent with expansion in an $s=2$ medium (Halpern et al. 1999). The steep decline might be the result of beaming (Halpern et al. 1999; Sari, Piran, \& Halpern 1999) instead of interaction with a wind. One way of distinguishing the wind interaction case is to observe the optically thick radio flux rise, $F_{\nu}\propto t$ for a wind as opposed to $F_{\nu}\propto t^{1/2}$ for a constant density. The effect of beaming in the $s=0$ case is to decrease the expansion and thus to increase the difference between the models. Radio observations of GRB 980519 were made at ages 0.3, 1.1, and 2.8 days at 8.3 GHz (Frail et al. 1998). Fig. 1 shows that an $s=2$ model is capable of fitting the radio, as well as the optical and X-ray, data. The model light curves in Fig. 1 are obtained using the synchrotron self-absorption model of Li \& Chevalier (1999). The model takes into account the dynamical evolution of a spherical, constant-energy blast wave in a pre-burst $s=2$ stellar wind, relativistic effects on radiation, and synchrotron self-absorption. The distance to GRB 980519 is unknown, so we take a fiducial value of $z=1$. Adopting a relatively low magnetic energy fraction of $\epsilon_B=10^{-5}$, we find that the following combination of parameters fits all available data reasonably well: $E_{52}=0.54$, $\epsilon_e=0.62 $, $p=3$, and $A_{\star}=4.3 $. Rough scalings of these parameters for other choices of $\epsilon_B$ are given in Li \& Chevalier (1999). The value of $p$ is higher than that normally found in GRB afterglows, but is within the range found in radio supernovae (Chevalier 1998 and references therein). The model shown in Fig. 1 does not include synchrotron cooling. The good fit to the X-ray data is then expected because the spectral index joining optical and X-ray emission, $\beta=-1.05\pm 0.10$ (Halpern et al. 1999), is consistent with the separate optical and X-ray indices. From eq. (4) and the parameters for GRB 980519 given above, we have $\nu_c\approx 4\times 10^{16}$ Hz for $t_{\rm day}=1$, which is below the X-ray frequency (3 keV = $7\times 10^{17}$ Hz). We consider the agreement adequate in view of the expected gradual turnover in flux and the observational and theoretical uncertainty in the value of $\nu_c$. The model does require a low value of $\epsilon_B$, as has also been inferred in a number of other afterglows (Galama et al. 1999). The model R-band flux densities on days 60 and 66 of Fig. 1 fall short of the observed values by nearly two orders of magnitude; the emission on these two days is presumed to come from the host galaxy of the burst (Sokolov et al. 1998; Bloom et al. 1998a). However, the observed flux densities on days 60 and 66 are close to those expected of SN 1998bw at a cosmological distance of $z=1$ (see Fig. 1). Despite excellent seeing conditions at Keck II, Bloom et al. (1998a) found little evidence for the extension expected of a host galaxy, so the presence of a SN 1998bw-type supernova is possible. Observations of the source at later times than have been reported should be able to confirm or reject this possibility. Another burst with a relatively steep decline with time is GRB 980326, in which the optical $F_{\nu}\propto t^{-2.10\pm 0.13}$ (Groot et al. 1998). There is no radio data for this object, but the last optical observation is a factor $\sim 10$ above the power law decline; this is not the host galaxy because it is not present at a later time (Bloom \& Kulkarni 1998). A possible source for the emission is a supernova (Bloom et al. 1999b), as in the case of GRB 980519. Such an event would be consistent with expansion into a wind and the explosion of a massive star. Bloom et al. (1999b) found that the spectrum of the late time source is consistent with that of a supernova, which rules out the possibility of a rise in the nonthermal afterglow emission as observed in GRB 970508 (Panaitescu et al. 1998) and the radio emission from SN 1998bw (Kulkarni et al. 1998; Li \& Chevalier 1999). Based on the discovery of the SN 1998bw/GRB 980425 association, Bloom et al. (1998b) proposed a subclass of GRBs produced by supernovae (S-GRBs). We also place SN 1998bw in a supernova class of GRB, but with different properties from those of Bloom et al. (1998b). By supernova, we mean an observed event with an optical light curve and spectrum similar to that of a Type I or Type II supernova. In our picture, the GRB 990123 is an example of a GRB interacting with the ISM (interstellar medium). The similar (slow) rates of decline observed in other GRB afterglows suggests that this type of object is the most common. These objects are not interacting with winds, do not have massive star progenitors, and are not accompanied by supernovae. Plausible progenitors are the mergers of compact objects. The GRBs in the wind type are interacting with winds, have massive star progenitors, and are accompanied by supernovae. SN 1998bw/GRB 980425 is the best example of this class of object. GRB 980519 and GRB 980326 are other possible members. Bloom et al. (1998b) propose that S-GRBs have no long-lived X-ray afterglow because of synchrotron losses. We suggest that X-ray afterglows are possible, although with steeper time evolution than in the ISM case, provided $\epsilon_B$ is small, as is also inferred for some of the ISM type afterglows. Bloom et al. further propose that S-GRBs have single pulse GRB profiles. We have found evidence for nonuniform energy input in SN 1998bw/GRB 980425 (Li \& Chevalier 1999) and suggest that both types of bursts are powered by central engines with irregular power output. The GRB itself may not allow a classification of the event; the two cosmological GRBs mentioned as possible members of the wind class have multiple peak time structure (in 't Zand et al. 1999; Groot et al. 1998). \acknowledgments We are grateful to J. Bloom, S. Kulkarni, and an anonymous referee for useful comments and information. Support for this work was provided in part by NASA grant NAG5-8232. \clearpage
\section{Introduction} Scalar field theory has become the generic playground for building cosmological models related to particle physics, in particular for obtaining inflationary cosmologies. One such class of models involve exponential potentials (exp$(\sqrt{16 \pi/p~m_{\rm Pl}^2} \phi)$), which lead to power law inflation, $a \propto t^p$, with $p>1$, for sufficiently flat potentials \cite{lucchin,halliwell,burd}. A number of related features have also been discovered for such potentials: in a Universe containing a perfect fluid and such a scalar field, then for a wide range of parameters the scalar field `mimics' the perfect fluid, adopting its equation of state \cite{wetterich,wands93,joyce} and leading to attractor scaling solutions at late time \cite{copeland98}. These solutions offer a plausible mechanism for stabilizing the dilaton field in models of gaugino condensation arising in supersymmetry breaking \cite{barreiro}. It is generally assumed that even if there are many scalar fields present, only one of them will dominate the dynamics and roll down the potential slowly. However, recently Liddle, Mazumdar and Schunck (LMS)\cite{lms}, have demonstrated in a particular example that multiple scalar fields, each with an exponential potential, can lead to inflationary solutions, even if the individual field potentials are too steep for inflation. There exists a cumulative effect of all the fields that can give rise to inflationary behavior - a result they termed as ``Assisted Inflation''. Malik and Wands have demonstrated that the associated attractor solution could be identified through a rotation in field space, with a hybrid model where the vacuum energy had an exponential dependence upon the dilaton field \cite{malik}. Multiple exponential potentials do arise in modern Kaluza-Klein theories. Indeed they are a natural outcome of the compactification of higher dimensional theories down to 3+1 dimensions. With this in mind it is worth investigating such potentials in a bit more detail. Indeed, Kanti and Olive have recently proposed a possible realisation of assisted inflation based on the compactification of a five-dimensional Kaluza Klein model \cite{kanti}. It also raises the question, could inflation arise out of the 11 dimensional supergravity models compactified on squashed seven spheres for example? Such models have been investigated by a number of authors as low energy cosmologies from string or M-theory \cite{lukas,stelle,reall,reall1}. Most `realistic' models of dimensional reduction lead to steep potentials, which generally do not lead to inflation. In this paper we consider a more general class of exponential potentials, which can include those generally found in supergravity compactifications, and obtain exact cosmological solutions for them. In particular we demonstrate how difficult it is to obtain assisted inflation when there exists cross-couplings between the scalar fields in the potential, a result also discussed in \cite{reall} and \cite{kanti}. We first recall the model discussed by LMS, before generalizing their potential to exponentials involving cross-coupling terms and demonstrating that the attractor behavior of the scalar-fields still exists, leading to scaling solutions for the generalized potential. We then turn our attention to the case of potentials involving multiple exponential terms containing the same scalar fields, and relate the solutions to those arising in supergravity models. \section{The dynamics of assisted inflation} Liddle et al\cite{lms}, considered $n$ scalar fields, $\phi_i ,\,i=1..n$, each with exponential potentials decoupled from each other: \begin{equation} V_i(\phi_i) = V_0 \exp \left( \alpha_i \, \phi_i \right) \,, \end{equation} where $\alpha_{i}$ is the slope of the individual field with dimensions of inverse Planck mass. Although the fields are not directly coupled through the potential, they are coupled through the Friedmann equation, which implies that the combined role of the fields affect the expansion rate of the Universe: \begin{eqnarray} \label{Motion} H^2 & = & \frac{8\pi}{3m_{{\rm Pl}}^2} \sum_{i=1}^n \left[ V_i(\phi_i) + \frac{1}{2} \dot{\phi}_i^2 \right] \,, \\ \label{eqofmo} \ddot{\phi}_i & = & - 3 H \dot{\phi}_i - \frac{dV_i(\phi_i)}{d\phi_i} \,, \end{eqnarray} where $H=\dot{a}/a$ is Hubble's constant and $a$ is the scale factor of the flat FRW Universe. The solution to this is the modified power law \cite{lms} \begin{equation} \label{power} a(t) \propto t^p \,, \end{equation} where $p$ is given by \begin{equation} \label{slope} p = \frac{16 \pi}{m_{{\rm Pl}}^2} \sum_{i=1}^n \frac{1}{\alpha_i^2} \,. \end{equation} Inflationary solutions exist provided $p > 1$, hence even if each of the $\alpha_i$'s are too steep to individually satisfy the condition for inflation, as long as $n$ is large enough, the inequality $p>1$ can be satisfied. These solutions, and the inflationary ones we shall present below are eternal, they do not possess an exit from the inflationary epoch. Realistic models would of course have to possess such an exit in order to enter the radiation and matter dominated epochs of our Universe. The particular example of Eq.~(\ref{slope}) suggests that it is worth investigating whether or not such assisted inflation exists with more general exponential potentials. An alternative approach with interesting results has been adopted in \cite{kanti}, where they have applied the assistance method to the case of polynomial scalar potentials. \subsection*{Exponential potentials with coupled scalar fields.} To begin with we consider the most natural generalization of the single field exponential, namely the case of two coupled fields: \begin{equation} \label{trial} V(\phi,\psi) = V_0 e^{\alpha \phi + \beta \psi} \,, \end{equation} where $\alpha$ and $\beta$ are the slopes for the fields $\phi$ and $\psi$. we see from Eq.(\ref{Motion}) that dimensionally, the right hand side should decrease as $t^{-2}$, because $H^2 \propto t^{-2} \propto V_i(\phi_{i})$. We further assume that for our potential Eq.(\ref{trial}), \begin{equation} \label{assum1} e^{\alpha \phi} = \frac{k_{\alpha}}{t^c}, ~~~~ e^{\beta \psi} = \frac{k_{\beta}}{t^{2-c}} \,, \end{equation} where $k_{\alpha}$, $k_{\beta}$ are dimensional constants and $c$ is a dimensionless constant. Substituting the power law solution Eq.(\ref{power}) into Eq.(\ref{Motion}) we obtain: \begin{eqnarray} \label{evolv} p^2 & = & H^2 t^2 \, \nonumber \\ & = & \frac{8\pi}{3m_{{\rm Pl}}^2} \left(V_0 k_{\alpha}k_{\beta} + \frac{1}{2} \left(\frac{c}{\alpha}\right)^2 + \frac{1}{2} \left(\frac{2-c}{\beta}\right)^2 \right) \,, \end{eqnarray} which when coupled with Eq.(\ref{eqofmo}) for the $\phi$ and $\psi$ fields leads to, \begin{equation} \label{sol1} V_0 k_{\alpha}k_{\beta} = \frac{(3p-1)c}{\alpha^2} \,, ~~~ V_0 k_{\alpha}k_{\beta} = \frac{(3p-1)(2-c)}{\beta^2} \,, \end{equation} hence \begin{equation} \label{sol3} V_0 k_{\alpha}k_{\beta} = \frac{2(3p-1)}{\alpha^2+\beta^2} \,. \end{equation} Using Eq.(\ref{sol1}--\ref{sol3}) in Eq.~(\ref{evolv}), we obtain a simple scaling solution between the two fields: \begin{eqnarray} \label{scal} p = \frac{16\pi}{m_{{\rm Pl}}^2} \frac{1}{\alpha^2+\beta^2} \,, \\ \left(\frac{\dot{\phi}}{\dot{\psi}} \right)^2 = \left(\frac{\alpha}{\beta} \right)^2. \end{eqnarray} An important problematic feature for inflationary solutions confronts us immediately in Eq.~(\ref{scal}), namely the coupling between the two fields reduces the rate of expansion of the Universe, a point also made in \cite{reall,kanti}. We will return to this point again later. An alternative method which would also lead to Eq.~(\ref{scal}) is described in \cite{malik} in terms of field rotations, which results in the introduction of two orthogonal fields, one of which is massless and the other posseses an exponential potential. Having demonstrated that it is possible to obtain a scaling solution without using slow roll approximations, we now generalize this simple case to include an arbitrary number of fields and exponential terms making up the overall potential. \subsection*{General exponential potentials with coupled scalar fields.} We now consider a potential where we have multiple scalar fields but their corresponding exponential potentials can contain arbitrary combinations of the fields with different slopes. The potential we will consider is: \begin{equation} \label{pot} V = \sum_{s=1}^n V_s = V_0 \sum_{s=1}^n \exp \left(\sum_{j=1}^{m_s} \alpha_{sj} \, \phi_{sj} \right) \,, \\ \end{equation} with the corresponding Friedmann equation : \begin{equation} \label{genevol0} H^2 = \frac{8\pi}{3m_{{\rm Pl}}^2} \left[ V + \sum_{s=1}^n \sum_{j=1}^{m_s} \frac{1}{2} \dot{\phi}_{sj}^2 \right] \,, \end{equation} where, from now on $q,r,s$ stands for index terms in the potential and $i,j,k,l$ stands for field indexes, hence, $\phi_{sj}$ stands for the $j$th field in the $s$th potential term. In other words, there are a total of $\sum_{s=1}^n m_s$ fields distributed in groups of $m_s$ through the terms of the potential. We obtain the solution to this problem by generalising the assumption Eq.(\ref{assum1}). There exists an attractor region with a power law solution, which from Eq.~(\ref{genevol0}), dimensionally satisfies $H^2 \propto t^{-2} \propto V_i$. Hence we write, \begin{eqnarray} \label{genevol} e^{\alpha_{sj} \phi_{sj}} & = & \frac{k_{sj}}{t^{c_{sj}}} \,, \\ \label{csj} \sum_{j=1}^{m_s} c_{sj} & = & 2 \,, \end{eqnarray} where $k_{sj}$ are dimensional and $c_{sj}$ are dimensionless constants respectively. Eq.(\ref{genevol}), coupled with the equations of motion: \begin{equation} \label{gensol1} \ddot{\phi}_{sj} + 3 H \dot{\phi}_{sj} + \frac{\partial V} {\partial \phi_{sj}} = 0 \,, \end{equation} result in: \begin{equation} (3p-1) c_{sj} = \alpha_{sj}^2 V_0 \prod_{k=1}^{m_s} k_{sk} \,, \end{equation} from which we find, using Eq.(\ref{csj}) and Eq.(\ref{gensol1}): \begin{eqnarray} \label{gensol2} V_0 \prod_{k=1}^{m_s} k_{sk} & = & \frac{2 (3p-1)}{\sum_{j=1}^{m_s} \alpha_{sj}^2} \,, \nonumber \\ \left(\frac{c_{sj}}{\alpha_{sj}} \right)^2 & = & \frac{4 \alpha_{sj}^2} {\left(\sum_{k=1}^{m_s} \alpha_{sk}^2 \right)^2} \,. \end{eqnarray} When substituted into Eq.(\ref{genevol0}) this leads to a key result, the generalisation of the original assisted inflation result given by Eq.~(\ref{slope}): \begin{equation} \label{slope1} p = \frac{16 \pi}{m_{{\rm Pl}}^2} \sum_{s=1}^n \frac{1}{\sum_{j=1}^{m_s} \alpha_{sj}^2} \,. \end{equation} We also note that the generalization of the scaling solution found in Eq.(\ref{scal}), quickly follows for the case of any two scalar fields, $\phi_{sj}$ and $\phi_{ql}$: \begin{equation} \label{scal1} \left(\frac{\dot{\phi}_{sj}}{\dot{\phi}_{ql}} \right)^2 = \left(\frac{\alpha_{sj}}{\alpha_{ql}}\right)^2 \left(\frac{\sum_{i=1}^{m_{q}} \alpha_{qi}^2}{\sum_{k=1}^{m_s} \alpha_{sk}^2}\right)^2 \,. \end{equation} It is directly evident that Eqs.(\ref{slope1}) and (\ref{scal1}) reduce to Eq.~(\ref{slope}) for the example of $n$ exponential terms each containing just one field, and Eq.~(\ref{scal}) for the case of one exponential term but containing two fields. We again see the inhibiting affect that multiplicative coupling of the fields (i.e. $m_s>1$) has for obtaining inflationary solutions. However, in this case, as with the original version of assisted inflation, this can be compensated for if there are enough exponential terms present in the potential (i.e. if $n$ is large enough)\cite{lms}. There is another feature of the potentials we have been discussing so far that makes them rather unphysical in general. We have been demanding that any two fields present can not be the same, (i.e. $\phi_{sj} \neq \phi_{ql}$). In other words they can only appear once in the full potential. Nearly all realistic models which emerge from compactifications arising in supergravity models have the same field appearing in at least two separate exponential terms. In the following section, we turn our attention to this case. \section{Exponential potentials inspired by supergravity models.} To set the scene, we generalise Eq.~(\ref{trial}) to the case where the scalar field potential takes the following form: \begin{equation} \label{2x2} V(\phi_1,\phi_2) = z_1 e^{\alpha_{11} \phi_1+ \alpha_{12} \phi_2} + z_2 e^{\alpha_{21} \phi_1 + \alpha_{22} \phi_2} \end{equation} where, $\alpha_{sj} $, are dimensional constants which can take any real value and $z_s>0$. The occurrence of such forms of the potential are quite common in dimensionally reduced supergravity models \cite{stelle,reall}. Remarkably we can solve this system to obtain scaling solutions in a manner analogous to those already presented in Eq.~(\ref{assum1})-(\ref{scal}) obtaining the unique late time scaling solution for the fields $\phi_1$ and $\phi_2$, \begin{equation} \label{scalnew} p = {16\pi \over m_{{\rm Pl}}^2} \left[{(\alpha_{21} - \alpha_{11})^2 + (\alpha_{22} - \alpha_{12})^2 \over (\alpha_{11} \alpha_{22} - \alpha_{21} \alpha_{12})^2 } \right]. \end{equation} This simple result reduces to the particular cases we have already investigated when the appropriate limits are taken. For example, when $\alpha_{21}=\alpha_{12}=0$, we reproduce the result Eq.~(\ref{slope}). The equivalent of the assisted inflation result Eq.~(\ref{scal}) follows by setting $\alpha_{11} \alpha_{21} + \alpha_{12} \alpha_{22} =0$, in which case we find, \begin{equation} \label{scalnew1} p = {16\pi \over m_{{\rm Pl}}^2} \left[{1 \over \alpha_{11}^2 + \alpha_{12}^2} + {1 \over \alpha_{21}^2 + \alpha_{22}^2} \right]. \end{equation} We shall now generalize the potential to $n$ such exponential potentials and $m$ combinations of linear fields in the exponent (explicitly calculating for the simple case of 2 terms $\times$ 2 fields of Eq.~(\ref{2x2})). The generalised Eq.~(\ref{2x2}) is then \begin{equation} V = \sum_{s=1}^n z_s \exp \left( \sum_{j=1}^{m} \alpha_{sj} \phi_j \right) \,. \end{equation} Of course we are allowing here for the possibility that $\alpha_{sj} =0$ for some combination of $sj$. We assume that for late times the fields have an attractor solution, given by \begin{equation} \label{expon} z_s\exp \left( \sum_{j=1}^m \alpha_{sj} \phi_j \right) = \frac{k_s}{t^2} \,, \end{equation} and, following Eq.~(\ref{genevol}) and Eq.~(\ref{csj}) we write \begin{equation} \label{late} \phi_j= a_j - \frac{c_{sj}}{\alpha_{sj}} \ln t \,, \end{equation} where, $a_j$ is a constant depending on the initial conditions and $\sum_{j=1}^m c_{sj} =2, ~~~s=1..n$. Substituting Eq.(\ref{late}) into the equation of motion Eq.(\ref{eqofmo}), we obtain the constraint equation for the $j$th field, which follows from assuming the existence of a power law solution, \begin{equation} \label{cijgeneral} (3p-1) \frac{c_{sj}}{\alpha_{sj}} = \sum_{q=1}^n \alpha_{qj}k_q \,. \end{equation} Again, using $\sum_{j=1}^m c_{ij} =2$ we obtain, \begin{equation} \sum_{j=1}^m \alpha_{sj} \left( \sum_{q=1}^n \alpha_{qj} k_q \right) = 2(3p-1) \,, \end{equation} which is equivalent to writing \begin{equation} \label{seqn} \sum_{q=1}^n A_{sq} k_q = 2(3p-1) \,, \end{equation} where \begin{equation} \label{defa} A_{sq} = \sum_{j=1}^m \alpha_{sj} \alpha_{qj} \,. \end{equation} Since $s$ is a free index, we have a set of $n$ equations that can be witten as \begin{equation} \label{eqk} A {\bf k} = 2(3p-1) \,, \end{equation} where $A$ is the $n \times n$ matrix with elements $A_{sq}$ and ${\bf k}=(k_1,..,k_n)$ a column vector. For the 2 $\times$ 2 case of Eq.~(\ref{2x2}) we obtain \begin{equation} A = \left( \begin{array}{cc} \alpha_{11}^2+\alpha_{12}^2 & \alpha_{11}\alpha_{21}+\alpha_{12}\alpha_{22} \\ \alpha_{21}\alpha_{11}+\alpha_{22}\alpha_{12} & \alpha_{21}^2+\alpha_{22}^2 \end{array} \right) \,. \end{equation} The solution to this system is \begin{equation} {\bf k} = A^{-1} 2(3p-1) \,, \end{equation} with \begin{equation} \label{defb} A^{-1} = \frac{A_{COF}^T}{det A} \,, \end{equation} where $A_{COF}^T$ is the transpose of the cofactor matrix of $A$. To simplify notation we will write $B \equiv A_{COF}^T$ and the sum of the elements in row $s$ of $B$ as $B^s \equiv \sum_{q=1}^n B_{sq}$, hence, each $k_s$ is \begin{equation} \label{defks} k_s = \frac{2(3p-1)}{det A} B^s \,. \end{equation} For the 2 $\times$ 2 case this yields \begin{eqnarray} \label{bkk} B &=& \left( \begin{array}{cc} \alpha_{21}^2+\alpha_{22}^2 & -\alpha_{11}\alpha_{21}-\alpha_{12}\alpha_{22} \\ -\alpha_{21}\alpha_{11}-\alpha_{22}\alpha_{12} & \alpha_{11}^2+\alpha_{12}^2 \end{array} \right) \,, \\ k_1 &=& 2(3p-1) \frac{\alpha_{21}^2+\alpha_{22}^2 -\alpha_{11}\alpha_{21}-\alpha_{12}\alpha_{22}}{(\alpha_{11}\alpha_{22}-\alpha_{ 12}\alpha_{21})^2} \,, \\ k_2 &=& 2(3p-1) \frac{-\alpha_{21}\alpha_{11}-\alpha_{22}\alpha_{12}+\alpha_{11}^2+\alpha_{12}^2 }{(\alpha_{11}\alpha_{22}-\alpha_{12}\alpha_{21})^2} \,. \end{eqnarray} >From Eqs.~(\ref{late}) and (\ref{cijgeneral}) the late time ratio between the kinetic terms of two different fields becomes \begin{equation} \left( \frac{\dot{\phi}_j}{\dot{\phi}_l} \right)^2 = \left( \frac{\sum_{q=1}^n \alpha_{qj}B^q}{\sum_{r=1}^n \alpha_{rl} B^r} \right)^2 \,. \end{equation} Substitution of Eqs.~(\ref{defks}) and (\ref{cijgeneral}) into the Friedmann equation yields, \begin{eqnarray} p^2 &=& \frac{8\pi}{3m_{{\rm Pl}}^2} \left[ \sum_{s=1}^n k_s + \frac{1}{2} \sum_{j=1}^m \left(\frac{c_{sj}}{\alpha_{sj}} \right)^2 \right] \nonumber \, \\ &=& \frac{8\pi}{3m_{{\rm Pl}}^2}\left[ 2(3p-1)\sum_{s=1}^n \frac{B^s}{det A} +2\sum_{j=1}^m \left(\frac{\sum_{q=1}^n \alpha_{qj}B^q}{det A}\right)^2 \right] \nonumber \,. \end{eqnarray} After some algebra, we obtain the simple result for $p$ as the ratio between the sum of all the elements in the cofactor matrix of $A$ and its determinant. \begin{equation} \label{finalp} p = \frac{16\pi}{m_{{\rm Pl}}^2}\frac{\sum_{s}^n \sum_{q}^n B_{sq}}{det A} \,. \end{equation} The reader should have no problem showing that for the 2 $\times$ 2 case this reduces to Eq.(\ref{scalnew}). It is instructive to rewrite Eq.~(\ref{finalp}) in another form. From Eqs.~(\ref{seqn}), (\ref{defks}) and (\ref{finalp}) \begin{equation} p = \frac{16\pi}{m_{{\rm Pl}}^2}\sum_{s=1}^n \frac{1}{\sum_{q=1}^n A_{sq} k_q/k_s} \,, \end{equation} and using Eq.~(\ref{defa}) with $q=s$ and Eq.~(\ref{defks}) we obtain, \begin{equation} \label{presult} p = \frac{16\pi}{m_{{\rm Pl}}^2}\sum_{s=1}^n \frac{1}{\sum_{j=1}^m \alpha_{sj}^2 + \sum_{q \neq s}^n A_{sq} B^q/B^s} \,. \end{equation} A number of points need to be made about Eq.~(\ref{presult}). It is similar in form to Eq.~(\ref{slope1}), which should not be too surprising, the additional terms in the denominator arising from the fact that we have allowed for fields to appear more than once in the potential, hence leading effectively to `self-interaction' type contributions. Indeed if these terms were turned off we would reproduce the result in Eq.~(\ref{slope1}). In the 2 $\times$ 2 case, it is the constraint leading to Eq.~(\ref{scalnew1}). Due to the presence of these `self-interaction' terms, $p$ could increase above the value in Eq.~(\ref{slope1}) if there happened to be a combination of positive and negative slopes in Eq.~(\ref{presult}). An issue emerges when considering these more complicated potentials. For the two field, two term case of Eq.(\ref{2x2}), if $\alpha_{11} > \alpha_{12}$ then, a necessary (but not sufficient) condition for the second term to be comparable to the first term at late times is, $\alpha_{21} < \alpha_{22}$. By comparable we mean that the potential terms reach a constant ratio. If this were not the case, then one of the two terms would quickly dominate the overall dynamics. One way to check if a combination of terms in a given potential will be comparable at late times is to use Eq.~(\ref{defks}) to obtain the ratios $k_s/k_q$ for these terms. From Eq.~(\ref{expon}) it follows that for consistency we require them to be positive, with $p>1/3$. In general the surviving terms (i.e. those which remain comparable) will be the ones with the smallest slopes, corresponding to the largest values of $p$. \section{Conclusions} In this paper we have derived a new class of exact cosmological solutions involving exponential potentials. In doing so we have been able to generalize the assisted inflation solutions discussed by LMS \cite{lms} to the case of multiple exponential terms involving many fields. Such potentials are more likely to arise in realistic models of particle physics where individual fields will occur in a number of separate exponential terms, leading to cross-couplings between the terms. In general, it transpires that it is more difficult to obtain assisted inflation in such models, the fields in any one exponential term tend to conspire to act against one another rather than assist each other, a result also noticed in \cite{reall,kanti}. This is the real reason why such models tend to fail to produce inflationary solutions in supergravity models compactified on squashed seven spheres \cite{reall}. We also investigated the case where a number of exponential terms contained the same scalar fields and demonstrated that a number of novel features emerged, including the possibility of increasing the rate of expansion when there exists a mixture of positive and negative slopes in the potential. \acknowledgements We would like to thank Andrew Liddle, Karim Malik, David Wands and Orfeu Bertolami for useful discussions. EJC was supported by PPARC and is particularly grateful to David Wands for conversations on the nature of the generalised solutions. NJN was supported by FCT (Portugal) under contract PRAXIS XXI BD/15736/98 and AM was supported by the INLAKS and the ORS award.
\section{Introduction} \label{sec1} The polynomials studied in this paper arise independently in graph theory and in statistical mechanics. It is appropriate, therefore, to begin by explaining each of these contexts. Specialists in these fields are warned that they will find at least one (and perhaps both) of these summaries excruciatingly boring; they can skip them. Let $G = (V,E)$ be a finite undirected graph\footnote{ In this paper a ``graph'' is allowed to have loops and/or multiple edges unless explicitly stated otherwise. } with vertex set $V$ and edge set $E$. For each positive integer $q$, let $P_G(q)$ be the number of ways that the vertices of $G$ can be assigned ``colors'' from the set $\{ 1,2,\ldots,q \}$ in such a way that adjacent vertices always receive different colors. It is not hard to show (see below) that $P_G(q)$ is the restriction to ${\mathbb Z}_+$ of a polynomial in $q$. This (obviously unique) polynomial is called the {\bf chromatic polynomial}\/ of $G$, and can be taken as the {\em definition}\/ of $P_G(q)$ for arbitrary real or complex values of $q$.\footnote{ Two excellent reviews on chromatic polynomials are \cite{Read_68,Read_88}. An extensive bibliography on chromatic polynomials is \cite{Chia_97}. } The chromatic polynomial was introduced in 1912 by Birkhoff \cite{Birkhoff_12}. The original hope was that study of the real or complex zeros of $P_G(q)$ might lead to an analytic proof of the Four-Color Conjecture \cite{Ore_67,Saaty_77}, which states that $P_G(4) > 0$ for all loopless planar graphs $G$. To date this hope has not been realized, although combinatoric proofs of the Four-Color Theorem have been found \cite{Appel_77a,Appel_77b,Appel_89,Robertson_97,Thomas_98}. Even so, the zeros of $P_G(q)$ are interesting in their own right and have been extensively studied. Most of the available theorems concern real zeros \cite{Birkhoff_46,Tutte_70,Woodall_77,Woodall_92,Jackson_93,Woodall_97,% Thomassen_97,Edwards_98}, but there has been some study (mostly numerical) of complex zeros as well \cite{Hall_65,Berman_69,Biggs_72,Beraha_79,Beraha_80,Farrell_80,% Baxter_86,Baxter_87,Read_91,Wakelin_92,Brenti_94,Brown_98a,Brown_98b,% Shrock_97a,Shrock_97b,Shrock_97c,Shrock_97d,Shrock_98a,Shrock_98b,% Shrock_98c,Tsai_98,Shrock_98e,Shrock_99a,Shrock_99b,Salas-Sokal_in_prep}. A more general polynomial can be obtained as follows: Assign to each edge $e \in E$ a real or complex weight $v_e$. Then define \begin{equation} Z_G(q, \{v_e\}) \;=\; \sum_{ \{\sigma_x\} } \, \prod_{e \in E} \, \biggl[ 1 + v_e \delta(\sigma_{x_1(e)}, \sigma_{x_2(e)}) \biggr] \;, \label{eq1.1} \end{equation} where the sum runs over all maps $\sigma\colon\, V \to \{ 1,2,\ldots,q \}$, the $\delta$ is the Kronecker delta, and $x_1(e), x_2(e) \in V$ are the two endpoints of the edge $e$ (in arbitrary order). It is not hard to show (see below) that $Z_G(q, \{v_e\})$ is the restriction to $q \in {\mathbb Z}_+$ of a polynomial in $q$ and $\{v_e\}$. If we take $v_e = -1$ for all $e$, this reduces to the chromatic polynomial. If we take $v_e = v$ for all $e$, this defines a two-variable polynomial $Z_G(q,v)$ that was introduced implicitly by Whitney \cite{Whitney_32a,Whitney_32b,Whitney_33} and explicitly by Tutte \cite{Tutte_47,Tutte_54}; it is known variously (modulo trivial changes of variable) as the {\bf dichromatic polynomial}\/, the {\bf dichromate}\/, the {\bf Whitney rank function}\/ or the {\bf Tutte polynomial}\/ \cite{Welsh_93,Biggs_93}.\footnote{ The Tutte polynomial $T_G(x,y)$ is conventionally defined as \protect\cite[p.~45]{Welsh_93} \protect\cite[pp.~73, 101]{Biggs_93} $$ T_G(x,y) \;=\; \sum\limits_{E' \subseteq E} (x-1)^{k(E')-k(E)} \, (y-1)^{|E'| + k(E') - |V|} $$ where $k(E')$ is the number of connected components in the subgraph $(V,E')$. Comparison with (\protect\ref{eq1.2}) below yields $$ T_G(x,y) \;=\; (x-1)^{-k(E)} \, (y-1)^{-|V|} \, Z_G \bigl( (x-1)(y-1), \, y-1 \bigr) \;. $$ } In statistical mechanics, \reff{eq1.1} is known as the partition function of the {\bf \mbox{\boldmath $q$}-state Potts model}\/.\footnote{ The Potts model \cite{Potts_52} was invented in the early 1950's by Domb (see \cite{Domb_74}). The $q=2$ case, known as the Ising model \cite{Ising_25}, was invented in 1920 by Lenz \cite{Lenz_20} (see \cite{Brush_67,Kobe_97,Ising_obit}). The $q=4$ case, which is a special case of the Ashkin--Teller model, was invented in 1943 by Ashkin and Teller \cite{Ashkin-Teller_43}. } In the Potts model \cite{Wu_82,Wu_84}, an ``atom'' (or ``spin'') at site $x \in V$ can exist in any one of $q$ different states (where $q$ is an integer $\ge 1$). The {\bf energy}\/ of a configuration is the sum, over all edges $e \in E$, of $0$ if the spins at the two endpoints of that edge are unequal and $-J_e$ if they are equal. The {\bf Boltzmann weight}\/ of a configuration is then $e^{-\beta H}$, where $H$ is the energy of the configuration and $\beta \ge 0$ is the inverse temperature. The {\bf partition function}\/ is the sum, over all configurations, of their Boltzmann weights. Clearly this is just a rephrasing of \reff{eq1.1}, with $v_e = e^{\beta J_e} - 1$. A coupling $J_e$ (or $v_e$) is called {\bf ferromagnetic}\/ if $J_e \ge 0$ ($v_e \ge 0$) and {\bf antiferromagnetic}\/ if $-\infty \le J_e \le 0$ ($-1 \le v_e \le 0$). To see that $Z_G(q, \{v_e\})$ is indeed a polynomial in its arguments (with coefficients that are in fact 0 or 1), we proceed as follows: In \reff{eq1.1}, expand out the product over $e \in E$, and let $E' \subseteq E$ be the set of edges for which the term $v_e \delta_{\sigma_{x_1(e)}, \sigma_{x_2(e)}}$ is taken. Now perform the sum over configurations $\{ \sigma_x \}$: in each connected component of the subgraph $(V,E')$ the spin value $\sigma_x$ must be constant, and there are no other constraints. Therefore, \begin{equation} Z_G(q, \{v_e\}) \;=\; \sum_{ E' \subseteq E } q^{k(E')} \prod_{e \in E'} v_e \;, \label{eq1.2} \end{equation} where $k(E')$ is the number of connected components (including isolated vertices) in the subgraph $(V,E')$. The expansion \reff{eq1.2} was discovered by Birkhoff \cite{Birkhoff_12} and Whitney \cite{Whitney_32a} for the special case $v_e = -1$ (see also Tutte \cite{Tutte_47,Tutte_54}); in its general form it is due to Fortuin and Kasteleyn \cite{Kasteleyn_69,Fortuin_72} (see also \cite{Edwards-Sokal}). We take \reff{eq1.2} as the {\em definition}\/ of $Z_G(q, \{v_e\})$ for arbitrary complex $q$ and $\{v_e\}$. In statistical mechanics, a very important role is played by the complex zeros of the partition function. This arises as follows \cite{Yang-Lee_52}: Statistical physicists are interested in {\em phase transitions}\/, namely in points where one or more physical quantities (e.g.\ the energy or the magnetization) depend nonanalytically (in many cases even discontinuously) on one or more control parameters (e.g.\ the temperature or the magnetic field). Now, such nonanalyticity is manifestly impossible in \reff{eq1.1}/\reff{eq1.2} for any finite graph $G$. Rather, phase transitions arise only in the {\em infinite-volume limit}\/. That is, we consider some countably infinite graph $G_\infty = (V_\infty, E_\infty)$ --- usually a regular lattice, such as ${\mathbb Z}^d$ with nearest-neighbor edges --- and an increasing sequence of finite subgraphs $G_n = (V_n, E_n)$. It can then be shown (under modest hypotheses on the $G_n$) that the {\bf (limiting) free energy per unit volume}\/ \begin{equation} f_{G_\infty}(q,v) \;=\; \lim_{n \to \infty} |V_n|^{-1} \log Z_{G_n}(q,v) \label{limiting_free_energy} \end{equation} exists for all {\em nondegenerate physical}\/ values of the parameters\footnote{ Here ``physical'' means that the weights are nonnegative, so that the model has a probabilistic interpretation; and ``nondegenerate'' means that we exclude the limiting cases $v=-1$ in (a) and $q=0$ in (b), which cause difficulties due to the existence of configurations having zero weight. }, namely either \begin{quote} \begin{itemize} \item[(a)] $q$ integer $\ge 1$ and $-1 < v < \infty$ \quad [using \reff{eq1.1}: see e.g.\ \cite[Section I.2]{Israel_79}] \item[or (b)] $q$ real $> 0$ and $0 \le v < \infty$ \quad [using \reff{eq1.2}: see \cite[Theorem 4.1]{Grimmett_95} and \cite{Grimmett_78,Seppalainen_98}]. \end{itemize} \end{quote} This limit $f_{G_\infty}(q,v)$ is in general a continuous function of $v$; but it can fail to be a real-analytic function of $v$, because complex singularities of $\log Z_{G_n}(q,v)$ --- namely, complex zeros of $Z_{G_n}(q,v)$ --- can approach the real axis in the limit $n \to\infty$. Therefore, the possible points of physical phase transitions are precisely the real limit points of such complex zeros. As a result, theorems that constrain the possible location of complex zeros of the partition function are of great interest. In particular, theorems guaranteeing that a certain complex domain is free of zeros are often known as {\em Lee-Yang theorems}\/.\footnote{ The first such theorem, concerning the behavior of the ferromagnetic Ising model at complex magnetic field, was proven by Lee and Yang \cite{Lee-Yang_52} in 1952. A partial bibliography (through 1980) of generalizations of this result can be found in \cite{Lieb-Sokal_81}. } The purpose of this paper is to prove an upper bound on the complex $q$-plane zeros of the Potts-model partition function $Z_G(q, \{v_e\})$, valid throughout the ``complex antiferromagnetic regime'' $|1 + v_e| \le 1$, under certain ``local'' conditions on the weights $\{v_e\}$: for example, in terms of the quantity $\max\limits_{x \in V} \sum\limits_{e \ni x} |v_e|$. As a corollary, I obtain upper bounds on the zeros of the chromatic polynomial $P_G(q)$ in terms of the maximum degree of the graph $G$. More precisely, I show that there exist universal constants $C(r) < \infty$ such that, for all loopless graphs $G$ of maximum degree $\le r$, the zeros of $P_G(q)$ lie in the disc $|q| < C(r)$. This answers in the affirmative a question posed by Brenti, Royle and Wagner \cite[Question 6.1]{Brenti_94}, generalizing an earlier conjecture of Biggs, Damerell and Sands \cite{Biggs_72} limited to $r$-regular graphs. The constants $C(r)$ arise as the solution of an explicit minimization problem, and I prove that $C(r) \le 7.963907 r$. This linear dependence on $r$ is best possible, as the example of the complete graph $K_{r+1}$ shows that $C(r) \ge r$. Furthermore, I show that the presence of {\em one}\/ vertex of large degree cannot lead to large chromatic roots. More precisely, if all but one of the vertices of $G$ have degree $\le r$, then the zeros of $P_G(q)$ lie in the disc $|q| < C(r) + 1$. Please note that a result of this kind {\em cannot}\/ hold if ``all but one'' is replaced by ``all but two'', for in this case the chromatic roots can be unbounded, even when $r=2$ and $G$ is planar \cite{Sokal_hierarchical}. The proofs of these results are based on well-known methods of mathematical statistical mechanics. The first step is to transform the Whitney--Tutte--Fortuin--Kasteleyn representation \reff{eq1.2} into a gas of ``polymers'' interacting via a hard-core exclusion (Section \ref{sec2}). I then invoke the Dobrushin condition \cite{Dobrushin_96a,Dobrushin_96b} (or the closely related Koteck\'y--Preiss condition \cite{Kotecky_86,Sokal_Mayer_in_prep}) for the nonvanishing of a polymer-model partition function (Section \ref{sec3}). Lastly, I verify these conditions for our particular polymer model, using a series of simple combinatorial lemmas, some of which may be of independent interest (Section \ref{sec4}); in particular, I give a simple proof of a generalized (multivariate) Brown-Colbourn conjecture on the zeros of the reliability polynomial for the special case of series-parallel graphs (Remark 3 in Section \ref{sec4.1}). The main results of this paper are contained in Section \ref{sec5}; some generalizations and extensions are in Section \ref{sec6}. I conclude with some conjectures and open questions (Section \ref{sec7}). With a little more work, it should be possible to extend the arguments of this paper to prove the existence and analyticity of the limiting free energy per unit volume \reff{limiting_free_energy} for suitable regular lattices $G_\infty$ and translation-invariant edge weights $v_e$, in the same region of complex $q$- and $\{v_e\}$-space where $Z$ will be proven (in Section~\ref{sec5}) to be nonvanishing uniformly in the finite subgraphs $V_n$ (``uniformly in the volume'' in statistical-mechanical lingo). In particular, this would provide a convergent expansion for the limiting free energy in powers of $1/q$. However, I have not worked out the details. This paper would never have seen the light of day without the help and advice of Antti Kupiainen. During my visit to Helsinki in September--October 1997, I told Antti of my conjectures about $P_G(q)$ and $Z_G(q, \{v_e\})$ --- conjectures that I had no good idea how to prove. He immediately saw that they ought to be provable by cluster (or Mayer) expansion. My reaction was, ``Ugh! You know how I {\em detest}\/ the cluster expansion!''; indeed, I had resisted learning it for nearly 20 years and had devoted much of my work in mathematical physics to finding ways of circumventing it \cite{Sokal_82,BFS_83a,BFS_83b,FFS_book}. Antti assured me that the cluster expansion is not so difficult, and he suggested that I study the excellent review article of Brydges \cite{Brydges_86}. We also quickly figured out how to represent $Z_G(q, \{v_e\})$ as a polymer gas. Jean Bricmont then told me about the work of Koteck\'y and Preiss \cite{Kotecky_86}, and Roman Koteck\'y informed me of the work of Dobrushin \cite{Dobrushin_96a}. Here, finally, was a version of the cluster expansion simple enough that even I could understand it! Nine months later, I figured out how to verify the Dobrushin (or Koteck\'y--Preiss) condition and thereby complete the proof. \section{Transformation of the Potts-Model Partition Function to a Polymer Gas} \label{sec2} Let $G=(V,E)$ be a finite undirected graph equipped with complex edge weights $\{ v_e \}_{e \in E}$. If $G$ contains a loop $e$ (i.e.\ an edge connecting a vertex to itself), this simply multiplies $Z_G(q, \{v_e\})$ by a factor $1+v_e$; so we can assume without loss of generality that $G$ is loopless, and we shall do so in this section in order to avoid unnecessary complications. Likewise, if $G$ contains multiple edges $e_1,\ldots,e_n$ connecting the same pair of vertices, they can be replaced, without changing the value of $Z$, by a single edge $e$ with weight $v_e = \prod_{i=1}^n (1+v_{e_i}) - 1$. So we could assume without loss of generality, if we wanted, that $G$ has no multiple edges. But this assumption would not simplify most of our subsequent arguments, so we shall usually refrain from making it. Note, however, that our numerical bounds frequently get better if multiple edges are replaced by a single equivalent edge. So let $G$ be loopless, and consider the Whitney--Tutte--Fortuin--Kasteleyn representation \reff{eq1.2} of the Potts-model partition function $Z_G(q, \{v_e\})$. For each term in \reff{eq1.2} we decompose the subgraph $(V,E')$ into its connected components. Some of these components may consist of a single vertex and no edges; the remaining components are disjoint connected subgraphs $(S_1,E_1), \ldots, (S_N,E_N)$ with $|S_i| \ge 2$. The total number of components is \begin{subeqnarray} k(E') & = & N \,+\, \left( |V| - \sum_{i=1}^N |S_i| \right) \\[2mm] & = & |V| \,-\, \sum_{i=1}^N (|S_i| - 1) \end{subeqnarray} It follows that: \begin{proposition}[jointly with Antti Kupiainen] \label{prop2.1} Let $G=(V,E)$ be a loopless finite undirected graph equipped with edge weights $\{ v_e \}_{e \in E}$. Then \begin{equation} Z_G(q, \{v_e\}) \;=\; q^{|V|} Z_{polymer,G}(q, \{v_e\}) \;, \label{eq2.2} \end{equation} where \begin{equation} Z_{polymer,G}(q, \{v_e\}) \;=\; \sum_{N=0}^\infty {1 \over N!} \; \sum_{S_1,\ldots,S_N \,\hboxscript{disjoint}} \; \prod_{i=1}^N w(S_i) \label{eq2.3} \end{equation} and \begin{equation} w(S) \;=\; \cases{ q^{-(|S|-1)} \!\! \sum\limits_{\begin{scarray} \widetilde{E} \subseteq E \\ (S,\widetilde{E}) \hboxscript{ connected} \end{scarray}} \prod\limits_{e \in \widetilde{E}} v_e & if $|S| \ge 2$ \cr \noalign{\vskip 2mm} 0 & if $|S| \le 1$ } \label{eq2.4} \end{equation} The sum in \reff{eq2.3} runs over pairwise disjoint subsets $S_1,\ldots,S_N$ of $V$, and the term $N=0$ in \reff{eq2.3} is understood to contribute 1. \end{proposition} \smallskip \par\noindent Note, in particular, that $w(S) = 0$ if $S$ is disconnected [i.e.\ if the induced subgraph $(S, E_S)$ is disconnected]. The ``polymer model'' \reff{eq2.3}--\reff{eq2.4} has the form of a grand-canonical gas (see Section~\ref{sec3} for the precise definition) \begin{equation} Z_{polymer,G}(q, \{v_e\}) \;=\; \sum_{N=0}^\infty {1 \over N!} \sum_{S_1,\ldots,S_N} \prod_{i=1}^N w(S_i) \prod_{1 \le i < j \le N} W(S_i,S_j) \label{eq2.5} \end{equation} with single-particle state space ${\cal P}_*(V)$ [the set of all nonempty subsets of $V$], fugacities $w(S)$, and two-particle Boltzmann factor given by a hard-core exclusion \begin{equation} W(S,S') \;=\; \cases{ 1 & if $S \cap S' = \emptyset$ \cr \noalign{\vskip 1mm} 0 & otherwise \cr } \label{eq2.6} \end{equation} Graph theorists will recognize the right-hand side of \reff{eq2.5} as the generating function, in the variables $w(S)$, for independent subsets of vertices of the intersection graph of ${\cal P}_*(V)$. The usefulness of \reff{eq2.2}--\reff{eq2.6} comes from the fact that the fugacities $w(S)$ are all suppressed by powers of $q^{-1}$, hence are small for large $|q|$. Moreover, if the sum over $\widetilde{E}$ in \reff{eq2.4} can be controlled, one expects that $w(S)$ will be exponentially decaying in $|S|$ when $|q|$ is large enough. This raises the hope that the Mayer expansion \cite{Uhlenbeck_62}, which is an expansion of $\log Z_{polymer,G}$ in powers of the fugacities $w(S)$, might converge for sufficiently large $|q|$. If so, this would imply that $Z_{polymer,G} \neq 0$ in the region of convergence. That is what we go about proving in the following sections --- but in the opposite order. \section{Dobrushin and Koteck\'y--Preiss Conditions for\break\hfill the Nonvanishing of $Z$} \label{sec3} In statistical mechanics, a {\bf grand-canonical gas} is defined by a {\em single-particle state space}\/ $X$ (here assumed for simplicity to be finite), a {\em fugacity vector}\/ $w = \{w_x\}_{x \in X} \in {\mathbb C}^X$, and a {\em two-particle Boltzmann factor}\/ $W(x,y)$ [a symmetric function $W \colon\; X \times X \to {\mathbb C}$]. The (grand) partition function $Z(w,W)$ is then defined to be the sum over ways of placing $N \ge 0$ ``particles'' on ``sites'' $x_1, \ldots, x_N \in X$, with each configuration assigned a ``Boltzmann weight'' given by the product of the corresponding factors $w_{x_i}$ and $W(x_i,x_j)$: \begin{equation} Z(w,W) \;=\; \sum_{N=0}^\infty {1 \over N!} \; \sum_{x_1,\ldots,x_N \in X} \; \prod_{i=1}^N w_{x_i} \prod_{1 \le i < j \le N} W(x_i,x_j) \;, \label{eq3.1} \end{equation} where the $N=0$ term is understood to contribute 1. Under very mild conditions on $W$ [e.g.\ $|W(x,y)| \le 1$ for all $x,y$ is more than sufficient], $Z(w,W)$ is an entire analytic function of $w$. Our goal is to find a sufficient condition for $Z(w,W)$ to be nonvanishing in a polydisc $D_R = \{ w \colon\; |w_x| < R_x\}$. This would imply, in particular, that $\log Z(w,W)$ is an analytic function of $w$ in $D_R$. We say that $W$ is \begin{itemize} \item {\em physical}\/ if $0 \le W(x,y) < +\infty$ for all $x,y \in X$ \item {\em repulsive}\/ if $|W(x,y)| \le 1$ for all $x,y \in X$ \item {\em physical and repulsive}\/ if $0 \le W(x,y) \le 1$ for all $x,y \in X$ \item {\em hard-core}\/ if $W(x,y) = 0 \hbox{ or } 1$ for all $x,y \in X$ \item {\em hard-core self-repulsive}\/ if $W(x,x) = 0$ for all $x \in X$ \end{itemize} An important special case is when $W$ is hard-core and hard-core self-repulsive: then $Z(w,W)$ is the generating function for independent sets of vertices of the graph $\widetilde{G} = (X,E)$ defined by placing an edge between each pair of vertices $x \neq y$ for which $W(x,y) = 0$. Dobrushin \cite{Dobrushin_96a,Dobrushin_96b} has given an elegant sufficient condition for the nonvanishing of $Z$ in a polydisc $D_R$, whenever $W$ is hard-core and hard-core self-repulsive. His proof is astoundingly simple, avoiding all the combinatoric complication that has given cluster expansions such a reputation for difficulty. Here I shall present a slight extension of Dobrushin's theorem, in which the condition of hard-core interaction is replaced by the weaker assumption that the interaction is physical and repulsive; moreover, the conclusion of the theorem is slightly strengthened. (We won't really need this extension --- the original Dobrushin theorem would suffice for our purposes --- but the stronger result is no more difficult, and it gives a bit more insight into the method of proof.) The hard-core {\em self-}\/repulsion is, however, essential both in Dobrushin's version and in my own: it guarantees that each ``site'' $x \in X$ can be occupied by at most one ``particle'' $x_i$. It follows that the partition function can be rewritten as a sum over subsets: \begin{equation} Z(w,W) \;=\; \sum_{X' \subseteq X} \; \prod_{x \in X'} w_x \prod_{\< xy \> \in X'} W(x,y) \;, \end{equation} where the second product runs over unordered pairs $x,y \in X'$ ($x \neq y$) with each pair counted once. Let us define, for each subset $\Lambda \subseteq X$, the restricted partition function \begin{equation} Z_\Lambda(w,W) \;=\; \sum_{X' \subseteq \Lambda} \; \prod_{x \in X'} w_x \prod_{\< xy \> \in X'} W(x,y) \;. \label{def_Z_Lambda} \end{equation} Of course this notation is redundant, since the same effect can be obtained by setting $w_x = 0$ for $x \in X \setminus \Lambda$, but it is useful for the purposes of the inductive proof. We have: \begin{theorem} \label{thm3.1} Let $X$ be a finite set, and let $W$ satisfy \begin{itemize} \item[(a)] $0 \le W(x,y) \le 1$ for all $x,y \in X$ \item[(b)] $W(x,x) = 0$ for all $x \in X$ \end{itemize} Suppose there exist constants $R_x \ge 0$ and $0 \le K_x < 1/R_x$ satisfying \begin{equation} K_x \;\ge\; \prod_{y \neq x} {1 - W(x,y) K_y R_y \over 1 - K_y R_y} \label{eq1} \end{equation} for all $x \in X$. Then, for each subset $\Lambda \subseteq X$, $Z_\Lambda(w,W)$ is nonvanishing in the closed polydisc $\bar{D}_R = \{ w \in {\mathbb C}^X \colon\; |w_x| \le R_x \}$ and satisfies there \begin{equation} \left| {\partial\log Z_\Lambda (w,W) \over \partial w_x} \right| \;\le\; \cases{ {\displaystyle {K_x \over 1 - K_x |w_x|} } & for all $x \in \Lambda$ \cr \noalign{\vskip 4mm} 0 & for all $x \in X \setminus \Lambda$ \cr } \label{eq2} \end{equation} Moreover, if $w, w' \in \bar{D}_R$ and $w'_x / w_x \in [0,+\infty]$ for each $x \in \Lambda$, then \begin{equation} \left| \log {Z_\Lambda(w',W) \over Z_\Lambda(w,W)} \right| \;\le\; \sum_{x \in \Lambda} \left| \log {1 - K_x |w'_x| \over 1 - K_x |w_x|} \right| \label{eq3} \end{equation} where on the left-hand side we take the standard branch of the log, i.e.\ $|\mathop{\rm Im}\nolimits\log \cdots| \le \pi$. \end{theorem} \medskip \par\noindent {\bf Remarks.} 1. It follows from \reff{eq1} that $K_x \ge 1$ and hence that $R_x < 1$. 2. The conclusion of Dobrushin's theorem \cite{Dobrushin_96a,Dobrushin_96b} is the special case of \reff{eq3} in which some of the $w'_x$ are equal to $w_x$ and others are equal to 0, and in which only the {\em real part}\/ of the logarithm on the left-hand side is handled. \medskip \par\medskip\noindent{\sc Proof.\ } Note first that \reff{eq2} for any given $\Lambda$ implies \reff{eq3} for the same $\Lambda$, by integration. The proof is by induction on the cardinality of $\Lambda$. If $\Lambda = \emptyset$ the claims are trivial. So let us assume that \reff{eq2} [and hence also \reff{eq3}] holds for all sets of cardinality $< n$, and let a set $\Lambda$ of cardinality $n$ be given. Let $x$ be any element of $\Lambda$, and let $\Lambda' = \Lambda \setminus \{x\}$. It follows from \reff{def_Z_Lambda} that \begin{equation} Z_\Lambda(w,W) \;=\; Z_{\Lambda'}(w,W) \,+\, w_x Z_{\Lambda'}(\widetilde{w},W) \label{eq_basic} \end{equation} where \begin{equation} \widetilde{w}_y \;=\; W(x,y) \, w_y \;; \end{equation} here the first term on the right-hand side of \reff{eq_basic} covers the summands $X' \not\ni x$, while the second covers $X' \ni x$. Note that $\widetilde{w} \in \bar{D}_R$ since $|W(x,y)| \le 1$. {}From \reff{eq_basic} we have \begin{equation} {\partial \over \partial w_x} \log Z_\Lambda(w,W) \;=\; {k(w) \over 1 + k(w) w_x} \end{equation} where \begin{equation} k(w) \;=\; {Z_{\Lambda'}(\widetilde{w},W) \over Z_{\Lambda'}(w,W)} \;. \end{equation} Now by the inductive hypothesis \reff{eq3} for $\Lambda'$, and using the fact that $\widetilde{w}_y/w_y = W(x,y) \ge 0$, we have \begin{equation} |k(w)| \;\le\; \prod_{y \in \Lambda'} {1 - W(x,y) K_y |w_y| \over 1 - K_y |w_y|} \;\le\; \prod_{y \in X \setminus \{x\}} {1 - W(x,y) K_y |w_y| \over 1 - K_y |w_y|} \;, \end{equation} which is $\le K_x$ by the hypothesis \reff{eq1}. This proves \reff{eq2} for $\Lambda$, and hence completes the induction. \hbox{\hskip 6pt\vrule width6pt height7pt depth1pt \hskip1pt}\bigskip Let us now return to the special case of a hard-core interaction. If $W(x,y) = 0$ (resp.\ 1), we say that $x$ and $y$ are {\em incompatible}\/ (resp.\ {\em compatible}\/) and write $x \not\sim y$ (resp.\ $x \sim y$). Note that in our convention $x \not\sim x$, in agreement with some authors' convention \cite{Kotecky_86,Seiler_82,Simon_93} and contrary to others' \cite{Dobrushin_96a,Dobrushin_96b}. The hypothesis \reff{eq1} is then equivalent to the existence of constants $c_x \ge 0$ such that \begin{equation} R_x \;\le\; (e^{c_x} - 1) \, \exp\!\left( - \sum_{y \not\sim x} c_y \right) \label{dob_condition} \end{equation} for all $x \in X$ [set $c_x = -\log(1 - K_x R_x)$]. This is the Dobrushin \cite{Dobrushin_96a,Dobrushin_96b} condition. Slightly stronger, and more convenient to check, is the Koteck\'y--Preiss \cite{Kotecky_86,Sokal_Mayer_in_prep} condition \begin{equation} R_x \;\le\; c_x \, \exp\!\left( - \sum_{y \not\sim x} c_y \right) \;. \label{KP_condition} \end{equation} \bigskip Let us now consider the important special case in which the single-particle state space $X$ can be partitioned as $X = \bigcup\limits_{n=1}^\infty X_n$ in such a way that \begin{equation} \sum\limits_{y \in X_n \colon\; y \not\sim x} R_y \;\le\; A_n m \quad\hbox{for all $x \in X_m$} \label{partition_condition_1} \end{equation} for suitable constants $\{A_n\}_{n=1}^\infty$. [This typically arises when $X$ is some set of nonempty subsets of a finite set $V$, and $x \not\sim y$ means $x \cap y \neq \emptyset$; we will then take $X_n$ to be the sets of cardinality $n$, and will prove \reff{partition_condition_1} by proving \begin{equation} \sum\limits_{y \in X_n \colon\; y \ni i} R_y \;\le\; A_n \quad\hbox{for all $i \in V$} \;, \label{partition_condition_2} \end{equation} which is manifestly stronger than \reff{partition_condition_1}.] Let us take \begin{equation} c_x \;=\; e^{\alpha n} R_x \quad\hbox{for all } x \in X_n \label{partition_choice_cx} \end{equation} with some suitably chosen $\alpha > 0$. Then, for \reff{partition_condition_1} to imply the Koteck\'y--Preiss condition \reff{KP_condition}, it suffices that \begin{equation} \sum\limits_{n=1}^\infty e^{\alpha n} \, A_n \;\le\; \alpha \;. \label{KP_sum_condition} \end{equation} We have therefore proven: \begin{proposition} \label{prop3.2} Suppose that $X = \bigcup\limits_{n=1}^\infty X_n$ (disjoint union) and that there exist constants $\{A_n\}_{n=1}^\infty$ and $\alpha$ such that \begin{itemize} \item[(a)] $\sum\limits_{y \in X_n \colon\; y \not\sim x} R_y \;\le\; A_n m \quad\hbox{for all $m,n$ and all $x \in X_m$}$ \item[(b)] $\sum\limits_{n=1}^\infty e^{\alpha n} \, A_n \;\le\; \alpha \;.$ \end{itemize} Then the Koteck\'y--Preiss condition \reff{KP_condition} holds with the choice $c_x = e^{\alpha n} R_x$ for $x \in X_n$. \end{proposition} \medskip \par\noindent {\bf Remarks.} 1. Suppose we try the more general Ansatz $c_x = b_n R_x$ for $x \in X_n$. Then \reff{partition_condition_1} implies the Koteck\'y--Preiss condition \reff{KP_condition} in case $b_n \ge e^{\alpha n}$ where $\alpha \equiv \sum_{n=1}^\infty b_n A_n$. But in that case $b'_n \equiv e^{\alpha n} \ge e^{\alpha' n}$ where $\alpha' \equiv \sum_{n=1}^\infty b'_n A_n$. So there is no loss of generality in restricting attention to $b_n = e^{\alpha n}$ for some $\alpha$. 2. Since the state space $X$ is finite, only finitely many of the $A_n$ are nonzero. Nevertheless, we often have occasion to consider simultaneously an infinite {\em family}\/ of problems --- e.g.\ in this paper, all loopless graphs $G$ of maximum degree $\le r$ and arbitrarily many vertices --- and it is natural to seek bounds that are {\em uniform}\/ over the family. So it is useful to forget that only finitely many of the $A_n$ are nonzero. (Moreover, similar methods can be applied to problems with an infinite state space $X$, in which case $\{A_n\}$ is a genuinely infinite sequence.) This leads to two further remarks: 3. For the condition \begin{equation} \exists \alpha > 0 \hbox{ such that } \sum\limits_{n=1}^\infty e^{\alpha n} A_n \;\le\; \alpha \label{KP_sum_condition_2} \end{equation} to hold, it is necessary that the sequence $\{ A_n \}_{n=1}^\infty$ have {\em some}\/ exponential decay (i.e.\ $A_n \le C e^{-\epsilon n}$ for some $\epsilon > 0$), but there is no minimum required rate of decay. Indeed, if $\{ A_n \}_{n=1}^\infty$ has any exponential decay at all, then by modifying finitely many of the $A_n$ one can make \reff{KP_sum_condition_2} hold. It can thus be valuable in applications to work hard on estimating the first few coefficients $A_n$ (see \cite{Kennedy_88} for an example).\footnote{ The emphasis in \protect\cite{Cammarota_82,Seiler_82,Brydges_86,Kotecky_86,Simon_93} on the special case $A_n = C e^{-\epsilon n}$ with $C=1$ is thus somewhat misleading, inasmuch as it suggests that there is a minimum allowed rate of decay $\epsilon$. } 4. Let $\delta = \liminf_{n\to\infty} (-\log A_n)/n$. Then $F(\alpha) = \alpha^{-1} \sum_{n=1}^\infty e^{\alpha n} A_n$ is finite-valued and continuous (in fact, real-analytic) on $0 < \alpha < \delta$, left-continuous (as a map into the extended real line) as $\alpha \uparrow \delta$, and identically $+\infty$ for $\alpha > \delta$. In particular, the infimum of $F(\alpha)$ is attained, so \reff{KP_sum_condition_2} is equivalent to \begin{equation} \inf\limits_{\alpha > 0} \alpha^{-1} \sum\limits_{n=1}^\infty e^{\alpha n} A_n \;\le\; 1 \;. \end{equation} \bigskip \par\noindent {\bf Important Final Remark.} The results in this section provide an extraordinarily simple proof of the convergence of the Mayer expansion for a grand-canonical gas with physical and repulsive two-particle interactions. To see what is at issue, let us first trivially rewrite the partition function \reff{eq3.1} as \begin{equation} Z(w,W) \;=\; \sum_{N=0}^\infty {1 \over N!} \; \sum_{x_1,\ldots,x_N \in X} \; \prod_{i=1}^N w_{x_i} \sum_{G \in {\cal G}_N} \prod_{\<ij\> \in G} F(x_i,x_j) \;, \end{equation} where ${\cal G}_N$ is the set of all (simple loopless undirected) graphs on the vertex set $\{1,\ldots,N\}$, and \begin{equation} F(x,y) \;=\; W(x,y) - 1 \end{equation} is called the {\em two-particle Mayer factor}\/. Then standard combinatorial arguments \cite{Uhlenbeck_62} show that \begin{equation} \log Z(w,W) \;=\; \sum_{N=1}^\infty {1 \over N!} \; \sum_{x_1,\ldots,x_N \in X} \; \prod_{i=1}^N w_{x_i} \sum_{G \in {\cal C}_N} \prod_{\<ij\> \in G} F(x_i,x_j) \label{mayer_exp} \end{equation} at least in the sense of formal power series in $w$, where ${\cal C}_N \subseteq {\cal G}_N$ is the set of {\em connected}\/ graphs on $\{1,\ldots,N\}$. This is the Mayer expansion; the principal problem is to prove its convergence in some specified polydisc. The usual approach to proving convergence of the Mayer expansion \cite{Penrose_67,Seiler_82,Cammarota_82,Brydges_86,Brydges_87,Brydges_88,% Simon_93,Brydges_99,Sokal_Mayer_in_prep} is to explicitly bound the terms in \reff{mayer_exp}; this requires some rather nontrivial combinatorics (for example, Proposition~\ref{prop_penrose} below together with the counting of trees). Once this is done, an immediate consequence is that $Z$ is nonvanishing in any polydisc where the series for $\log Z$ is convergent. Dobrushin's brilliant idea \cite{Dobrushin_96a,Dobrushin_96b} was to prove these two results in the opposite order. First one proves, by an elementary induction on the cardinality of the state space, that $Z$ is nonvanishing in some specified polydisc (Theorem~\ref{thm3.1}); it then follows immediately that $\log Z$ is analytic in that polydisc, and hence that its Taylor series \reff{mayer_exp} is convergent there. It is an interesting open question to know whether this approach can be made to work without the assumption of hard-core self-repulsion. \section{Some Combinatorial Lemmas} \label{sec4} \subsection{Reduction to trees} \label{sec4.1} The weight $w(S)$ involves a sum \reff{eq2.4} over connected subgraphs $(S,\widetilde{E})$ of the induced subgraph $(S, E_S)$. The trouble is that there may be ``too many'' connected subgraphs. It is remarkable, therefore, that this sum can sometimes be bounded by a sum over a much smaller set of graphs, namely {\em spanning trees}\/. The following proposition underlines the special role played by the ``complex antiferromagnetic regime'' $A \equiv \{v \in {\mathbb C} \colon\; |1 + v| \le 1 \}$. \begin{proposition}[Penrose \protect\cite{Penrose_67}] \label{prop_penrose} Let $G=(V,E)$ be a finite undirected graph equipped with complex edge weights $\{ v_e \}_{e \in E}$ satisfying $|1 + v_e| \le 1$ for all $e$. Then \begin{equation} \left| \sum\limits_{\begin{scarray} E' \subseteq E \\ (V,E') \hboxscript{ connected} \end{scarray}} \, \prod\limits_{e \in E'} v_e \, \right|\, \;\,\le\;\, \sum\limits_{\begin{scarray} E' \subseteq E \\ (V,E') \hboxscript{ tree} \end{scarray}} \, \prod\limits_{e \in E'} |v_e| \;. \label{penrose_bound} \end{equation} \end{proposition} \par\noindent Penrose \cite{Penrose_67} proved this when $G$ is the complete graph $K_n$; the result then follows for all graphs without loops or multiple edges (it suffices to set $v_e = 0$ on the nonexistent edges). Here I present a minor modification of Penrose's proof that permits loops and multiple edges: \medskip \par\medskip\noindent{\sc Proof.\ } We can assume without loss of generality that $G$ is connected, since otherwise both sides of the inequality are zero. Let ${\cal C}$ (resp.\ ${\cal T}$) be the set of subsets $E' \subseteq E$ such that $(V,E')$ is connected (resp.\ is a tree). Clearly ${\cal C}$ is an increasing family of subsets of $E$ with respect to set-theoretic inclusion, and the minimal elements of ${\cal C}$ are precisely those of ${\cal T}$ (i.e.\ the spanning trees). It is a nontrivial but well-known fact \cite[Sections 7.2 and 7.3]{Bjorner_92} \cite[Section 8.3]{Ziegler_95} that the (anti-)complex ${\cal C}$ is {\em partitionable}\/: that is, there exists a map ${\bf R} \colon\, {\cal T} \to {\cal C}$ such that ${\bf R}(T) \supseteq T$ for all $T \in {\cal T}$ and ${\cal C} = \biguplus \, [T, \, {\bf R}(T)]$ (disjoint union), where $[E_1,E_2]$ denotes the Boolean interval $\{ E' \colon\; E_1 \subseteq E' \subseteq E_2 \}$. In fact, many alternative choices of ${\bf R}$ are available \cite[Sections 7.2 and 7.3]{Bjorner_92} \cite[Sections 2 and 6]{Gessel_96} \cite[Proposition 13.7 et seq.]{Biggs_93}, and none of the subsequent arguments will depend on a specific choice of ${\bf R}$. Nevertheless, for completeness, we shall give at the end of this proof a concrete construction of one possible ${\bf R}$. Given the existence of ${\bf R}$, we have the immediate identity \begin{eqnarray} \sum\limits_{\begin{scarray} E' \subseteq E \\ (V,E') \hboxscript{ connected} \end{scarray}} \; \prod\limits_{e \in E'} v_e & = & \sum\limits_{\begin{scarray} T \subseteq E \\ (V,T) \hboxscript{ tree} \end{scarray}} \; \prod\limits_{e \in T} v_e \sum\limits_{T \subseteq E' \subseteq {\bf R}(T)} \; \prod\limits_{e \in E' \setminus T} v_e \nonumber \\[3mm] & = & \sum\limits_{\begin{scarray} T \subseteq E \\ (V,T) \hboxscript{ tree} \end{scarray}} \; \prod\limits_{e \in T} v_e \; \prod\limits_{e \in {\bf R}(T) \setminus T} (1 + v_e) \;. \label{penrose_identity} \end{eqnarray} In particular, if $|1 + v_e| \le 1$ for all $e$, then \reff{penrose_bound} follows. We now indicate a construction of ${\bf R}$ that is a slight variant of the one used by Penrose \cite{Penrose_67} (he orders the vertices, while I order the edges): Choose (arbitrarily) a vertex $x \in V$ and call it the root; and choose (arbitrarily) a numbering of the edges. For each $E' \in {\cal C}$ and $y \in V$, let $\hbox{\em depth}_{E'}(y)$ be the length of the shortest path in $E'$ connecting $y$ to the root. For each $y \in V \setminus \{x\}$, let $e(y)$ be the lowest-numbered edge in $E'$ connecting $y$ to a vertex $y'$ with $\hbox{\em depth}_{E'}(y') = \hbox{\em depth}_{E'}(y) - 1$. And finally, let ${\bf S}(E') = \{ e(y) \colon\; y \in V \setminus \{x\} \, \}$. Then trivially ${\bf S}(E') \subseteq E'$; moreover, it is easy to see that $(V, {\bf S}(E'))$ is a tree and that $\hbox{\em depth}_{{\bf S}(E')}(y) = \hbox{\em depth}_{E'}(y)$ for all $y \in V$. Conversely, given a spanning tree $(V,T)$, it is not hard to see that ${\bf S}(E') = T$ if and only if $T \subseteq E' \subseteq {\bf R}(T)$, where ${\bf R}(T)$ is obtained from $T$ by adjoining all edges $e \in E$ that \begin{itemize} \item[(a)] connect two vertices of equal $\hbox{\em depth}_{T}$ (this includes loops, if any), or \item[(b)] connect a vertex $y$ to a vertex $y'$ having $\hbox{\em depth}_{T}(y') = \hbox{\em depth}_{T}(y) - 1$ where $e$ is higher-numbered than the edge already in $T$ that connects $y$ to a vertex $y''$ having $\hbox{\em depth}_{T}(y'') = \hbox{\em depth}_{T}(y) - 1$. \end{itemize} This completes the proof. \hbox{\hskip 6pt\vrule width6pt height7pt depth1pt \hskip1pt}\bigskip {\bf Remarks.} 1. The identity \reff{penrose_identity} and the inequality \reff{penrose_bound} generalize to matroids. Indeed, for any matroid $M$, the independent sets of $M$ form a simplicial complex $IN(M)$, called a {\em matroid complex}\/; moreover, every matroid complex is shellable, and every shellable complex is partitionable \cite[Theorem 7.3.3 and Proposition 7.2.2]{Bjorner_92}. For the cographic matroid $M^*(G)$, the independent sets are the complements of connected subgraphs, and the bases are complements of spanning trees, so we recover the situation of Proposition~\ref{prop_penrose}. I thank Criel Merino for teaching me about matroids and for helping me notice a silly error in my original proof of Proposition~\ref{prop_penrose}. Earlier, Dave Wagner had informed me that analogues of \reff{penrose_identity} hold for shellable simplicial complexes (see \cite[sections 0.3 and III.2]{Stanley_96} for the definition). There is a long history of identities related to \reff{penrose_identity}: see, for example, \cite{Rota_64,Malyshev_79,Brydges_86,Brydges_87,Brydges_88,Abdesselam_95,% Brydges_99} and the references cited therein. 2. I conjecture that \reff{penrose_bound} can be strengthened so that on the right-hand side the absolute value is put outside the sum rather than inside. (This would be useful in case the $\{v_e\}$ do not all have the same phase.) In fact, I conjecture more: Let \begin{equation} c(E') \;=\; |E'| \,-\, |V| \,+\, k(E') \end{equation} be the cyclomatic number of the subgraph $(V,E')$, and define the generalized connected sum \begin{subeqnarray} C_G(\lambda, \{v_e\}) & = & \!\!\sum\limits_{\begin{scarray} E' \subseteq E \\ (V,E') \hboxscript{ connected} \end{scarray}} \!\lambda^{c(E')} \, \prod\limits_{e \in E'} v_e \\[2mm] & = & \lambda^{1-|V|} \, C_G(1, \{\lambda v_e\}) \slabel{eq4.4b} \\[2mm] & = & \lim\limits_{q\to 0} \, q^{-1} \lambda^{-|V|} Z_G(\lambda q, \{\lambda v_e\}) \;. \end{subeqnarray} In particular, $\lambda=0$ corresponds to the tree sum and $\lambda=1$ to the connected sum. Then I conjecture that \begin{quote} \begin{itemize} \item[(a)] If $|1 + v_e| \le 1$ for all $e$, then $|C_G(\lambda, \{v_e\})|$ is a decreasing function of $\lambda$ on $0 \le \lambda \le 1$. \end{itemize} \end{quote} I had originally conjectured a stronger result, namely \begin{quote} \begin{itemize} \item[($\hbox{a}'_r$)] If $|r + v_e| \le r$ for all $e$, then $(-1)^n \, (d^n/d\lambda^n) \, |C_G(\lambda, \{v_e\})|^2 \ge 0$ on $0 \le \lambda \le 1$, for all $n \ge 0$ \end{itemize} \end{quote} either for $r=1$ or, failing that, for $r = {1 \over 2}$; but this is in fact false for all $r > 0$, even for the second derivative evaluated at $\lambda = 0$ with equal edge weights $v_e = v$. Indeed, if we write \begin{equation} C_G(\lambda,v) \;=\; v^{|V|-1} \, \bigl[ a_0 + a_1 v \lambda + a_2 v^2 \lambda^2 + \ldots \bigr] \end{equation} where $a_j$ is the number of spanning subgraphs of $G$ having $j$ cycles, then \break $(d^2 / d\lambda^2) \, |C_G(\lambda,v)|^2 \, \Bigr| _{\lambda=0} \ge 0$ holds for all $v$ if $a_1^2 \ge 2 a_0 a_2$, but fails for $v$ in a wedge near the imaginary axis if $a_1^2 < 2 a_0 a_2$. Now the complete bipartite graph $K_{3,4}$ has $a_0 = 432$, $a_1 = 612$, $a_2 = 456$ and hence provides a counterexample. Nevertheless, ($\hbox{a}'_r$) might be true for some interesting subclasses of graphs $G$. For {\em real}\/ $v_e \in [-1,0]$, by contrast, I am able to prove a result even stronger than ($\hbox{a}'_{1/2}$), namely \begin{quote} \begin{itemize} \item[(b)] If $-1 \le v_e \le 0$ for all $e$, then $(-1)^{n+|V|-1} \, (d^n/d\lambda^n) \, C_G(\lambda, \{v_e\}) \ge 0$ on $0 \le \lambda \le 1$, for all integers $n \ge 0$. \end{itemize} \end{quote} Indeed, by \reff{penrose_identity} and \reff{eq4.4b}, we have \begin{equation} C_G(\lambda, \{v_e\}) \;=\; \sum\limits_{\begin{scarray} T \subseteq E \\ (V,T) \hboxscript{ tree} \end{scarray}} \; \prod\limits_{e \in T} v_e \; \prod\limits_{e \in {\bf R}(T) \setminus T} (1 + \lambda v_e) \end{equation} and hence \begin{equation} {d^n \over d\lambda^n} \, C_G(\lambda, \{v_e\}) \;=\; \sum\limits_{\begin{scarray} T \subseteq E \\ (V,T) \hboxscript{ tree} \end{scarray}} \sum\limits_{\begin{scarray} \widetilde{T} \subseteq \hbox{\bf\scriptsize R}(T) \setminus T \\ |\widetilde{T}| = n \end{scarray}} \; \prod\limits_{e \in T \cup \widetilde{T}} v_e \; \prod\limits_{e \in {\bf R}(T) \setminus (T \cup \widetilde{T})} (1 + \lambda v_e) \;, \end{equation} which has the claimed sign whenever $0 \le \lambda \le 1$ and $-1 \le v_e \le 0$ for all $e$. 3. $C_G(1, \{v_e\})$ is equal, up to a prefactor, to the reliability polynomial $R_G(\{p_e\})$ \cite{Colbourn_87}, where $p_e$ is the probability that edge $e$ is operational and $v_e = p_e/(1-p_e)$: \begin{equation} R_G(\{p_e\}) \;=\; \left[ \prod_{e \in E} (1-p_e) \right] C_G(1, \{p_e/(1-p_e)\}) \;. \end{equation} Now the Brown-Colbourn conjecture \cite{Brown_92,Wagner_00} states that for any connected graph $G$ (loops and multiple edges are allowed), $R_G(p) \neq 0$ whenever $|p-1| > 1$. A more general conjecture is that $R_G(\{p_e\}) \neq 0$ whenever $|p_e - 1| > 1$ for all edges $e$, or equivalently, that $C_G(1, \{v_e\}) \neq 0$ whenever $0 < |1 + v_e| < 1$ for all $e$. But this generalized Brown-Colbourn conjecture is an immediate consequence of conjecture (a): for if we had $C_G(1, \{v_e\}) = 0$ with $|1 + v_e| < 1$ for all $e$, then we could choose $\epsilon > 0$ such that $v'_e \equiv (1+\epsilon) v_e$ satisfy $|1 + v'_e| < 1$ for all $e$, and we would have $C_G(\lambda, \{v'_e\}) = 0$ for $\lambda = 1/(1+\epsilon)$ [but not, of course, identically for $1/(1+\epsilon) \le \lambda \le 1$]. Note also that if the generalized Brown-Colbourn conjecture holds for a graph $G$, then it holds also for any graph that can be obtained from $G$ by a sequence of doublings of edges (``parallel expansions'') and/or subdivisions of edges (``series expansions''). This follows from the formulae \cite[p.~35]{Colbourn_87} \begin{eqnarray} R_{G'}(\{p_e, p_1, p_2\}) & = & R_G(\{p_e, 1 - (1-p_1)(1-p_2)\}) \\[2mm] R_{G'}(\{p_e, p_1, p_2\}) & = & [1- (1-p_1)(1-p_2)] \, R_G \Bigl( \Bigl\{ p_e, {\displaystyle p_1 p_2 \over \displaystyle p_1 + p_2 - p_1 p_2 } \Bigr\} \Bigr) \end{eqnarray} where $G'$ is obtained from $G$ by parallel (resp.\ series) expansion of an edge $e_0$ into a pair of edges $e_1, e_2$. It suffices to note that if $|1-p_i| > 1$ for $i=1,2$, then the same inequality holds for $p_{\parallel} \equiv 1- (1-p_1)(1-p_2)$ and for $p_{series} \equiv p_1 p_2/(p_1 + p_2 - p_1 p_2)$; the former is obvious, and the latter follows by observing that the series-expansion formula corresponds to addition of $1/v = 1/p - 1$ and that $|1-p| > 1$ corresponds to $\mathop{\rm Re}\nolimits(1/v) < -1/2$. In particular, since the generalized Brown-Colbourn conjecture manifestly holds for trees, it also holds for all connected graphs without a $K_4$ minor, as these are precisely the graphs that can be obtained from trees by a sequence of series and parallel expansions \cite{Duffin_65,Liu_76,Wald_83,Oxley_86}. The (original) Brown-Colbourn conjecture for series-parallel graphs was first proven by Wagner \cite{Wagner_00}, by a vastly more complicated method. \subsection{Connected subgraphs containing a specified vertex} Let $G=(V,E)$ be a finite or countably infinite undirected graph equipped with edge weights $\{ v_e \}_{e \in E}$, and let $x \in V$. Let us define the weighted sum over connected subgraphs $G' = (V',E') \subseteq G$ containing $n$ vertices, one of which is $x$, and $m$ edges: \begin{equation} C_{n,m}(G, \{ v_e \}, x) \;=\; \sum_{\begin{scarray} G' = (V',E') \subseteq G \\ G' \hboxscript{ connected} \\ V' \ni x \\ |V'| = n \\ |E'| = m \end{scarray}} \prod_{e \in E'} |v_e| \;. \end{equation} Special cases are the tree sum \begin{equation} T_n(G, \{ v_e \}, x) \;=\; C_{n,n-1}(G, \{ v_e \}, x) \end{equation} and the edge-counted sum \begin{equation} C_{\bullet,m}(G, \{ v_e \}, x) \;=\; \sum\limits_{n=1}^{m+1} C_{n,m}(G, \{ v_e \}, x) \;. \end{equation} When the edge weights $v_e$ are all equal to 1, we shall optionally omit them from the notation; note in particular the obvious bound \begin{equation} C_{n,m}(G, \{ v_e \}, x) \;\le\; C_{n,m}(G, x) \, \left( \sup_{e \in E} |v_e| \right) ^{\! m} \;. \end{equation} In this subsection we shall obtain a variety of upper bounds on $C_{n,m}(G, \{ v_e \}, x)$ in terms of ``local'' information about the graph $G$ and the weights $\{ v_e \}$. \begin{proposition} \label{prop4.2} Let $G=(V,E)$ be a finite or countably infinite loopless undirected graph of maximum degree $\le r$, equipped with edge weights $\{ v_e \}_{e \in E}$; and let $x \in V$. Let ${\bf T}_r$ be the infinite $r$-regular tree, and let $y$ be any vertex in ${\bf T}_r$. Then \begin{equation} C_{\bullet,m}(G, x) \;\le\; C_{\bullet,m}({\bf T}_r, y) \;=\; T_{m+1}({\bf T}_r, y) \;\equiv\; t_{m+1}^{(r)} \;=\; r \, {[(r-1)(m+1)]! \over m! \, [(r-2)m + r]!} \label{eq_tnr} \end{equation} and hence \begin{equation} C_{\bullet,m}(G, \{ v_e \}, x) \;\le\; t_{m+1}^{(r)} \, \left( \sup_{e \in E} |v_e| \right) ^{\! m} \;. \label{eq4.16} \end{equation} In particular, \begin{equation} T_n(G, \{ v_e \}, x) \;\le\; t_n^{(r)} \, \left( \sup_{e \in E} |v_e| \right) ^{\! n-1} \;. \label{eq4.17} \end{equation} \end{proposition} \par\medskip\noindent{\sc Proof.\ } We can assume without loss of generality that $G=(V,E)$ is connected. Let $U = (\widetilde{V}, \widetilde{E})$ be the universal covering graph of $G$, with covering map $f \colon\, U \to G$; and let $\widetilde{x}$ be a vertex of $U$ such that $f(\widetilde{x}) = x$. [The universal covering graph of a connected loopless graph $G$ can be constructed as follows: Fix a base vertex $x$ of $G$, and let the vertices of $U$ be the walks in $G$ (of finite length) that begin at $x$ and do not contain any ``doublebacks'' (i.e.\ two consecutive uses of the same edge in opposite directions). Two vertices of $U$ are defined to be adjacent if one of them is a one-step extension of the other, and $f \colon\, U \to G$ maps each walk onto its final vertex. We take $\widetilde{x}$ to be the zero-step walk starting at $x$.] It is easy to see that $U$ is a tree (in general countably infinite even when $G$ is finite). Moreover, since $G$ has maximum degree $\le r$, $U$ is a subtree of ${\bf T}_r$, from which it follows trivially that $C_{\bullet,m}(U, \widetilde{x}) \le C_{\bullet,m}({\bf T}_r, \widetilde{x})$. Let us prove, then, that $C_{\bullet,m}(G, x) \le C_{\bullet,m}(U, \widetilde{x})$. Fix an arbitrary total order on $E$, and choose arbitrarily for each edge $e \in E$ a distinguished direction. Now let $H$ be a connected $m$-edge subgraph of $G$ that contains $x$. Let $S$ be the lexicographically first (with respect to the chosen total order on $E$) spanning tree of $H$. Then $S$ based at $x$ has a unique lifting to a subgraph $\widetilde{S}$ of $U$ based at $\widetilde{x}$: it is defined by mapping each vertex $s$ of $S$ to the unique path in $S$ from $x$ to $s$. Now, for each edge $e$ of $H$ not belonging to $S$, there is a unique edge $\widetilde{e}$ of $U$ such that $f(\widetilde{e}) = e$ and $\widetilde{e}$ is incident with the image in $\widetilde{S}$ of the vertex of $S$ from which $e$ is directed. The addition of these edges to $\widetilde{S}$ produces a connected $m$-edge subgraph $\widetilde{H}$ of $U$ that contains $\widetilde{x}$. Moreover, the map $H \mapsto \widetilde{H}$ is injective, since $H = f(\widetilde{H})$. This completes the proof that $C_{\bullet,m}(G, x) \le C_{\bullet,m}({\bf T}_r, \widetilde{x})$. We conclude by calculating the numbers $t_{m+1}^{(r)} \equiv C_{\bullet,m}({\bf T}_r, \widetilde{x})$.\footnote{ For similar computations, see e.g.\ \protect\cite{Fisher_61}. } Let ${\bf U}_r$ be the infinite tree in which all vertices have degree $r$ except for one vertex $y$ which has degree $r-1$, and let $u_{m+1}^{(r)} = C_{\bullet,m}({\bf U}_r, y)$. Then define, as formal power series, the generating functions \begin{eqnarray} T_r(z) & = & \sum\limits_{n=1}^\infty t_n^{(r)} z^n \label{Tr_def} \\[1mm] U_r(z) & = & \sum\limits_{n=1}^\infty u_n^{(r)} z^n \label{Ur_def} \end{eqnarray} The recursive structure of $r$-regular rooted trees easily implies the functional equations \begin{eqnarray} T_r(z) & = & z [1 + U_r(z)]^r \label{Tr_eqn} \\[1mm] U_r(z) & = & z [1 + U_r(z)]^{r-1} \label{Ur_eqn} \end{eqnarray} We now use the Lagrange Implicit Function Theorem for formal power series \cite[Theorem 1.2.4]{Goulden_83}, which states that for formal power series $f(u) = \sum_{n=0}^\infty f_n u^n$ and $g(u) = \sum_{n=0}^\infty g_n u^n$ with $g_0 \neq 0$, the functional equation $U(z) = z g(U(z))$ has a unique solution $U(z)$, and for all $n \ge 1$ one has \begin{equation} [z^n] f(U(z)) \;=\; {1 \over n} \, [u^{n-1}] (f'(u) g(u)^n) \end{equation} where $[z^n] P(z)$ denotes the coefficient of $z^n$ in the formal power series $P(z)$. Applying this with $f(u) = (1+u)^r$ and $g(u) = (1+u)^{r-1}$ yields \begin{eqnarray} t_n^{(r)} & = & {r \over n-1} \, {(r-1)n \choose n-2} \;=\; r \, {[(r-1)n]! \over (n-1)! \, [(r-2)n + 2]!} \label{tr_result} \\[3mm] u_n^{(r)} & = & {1 \over n} \, {(r-1)n \choose n-1} \;=\; (r-1) \, {[(r-1)n -1]! \over (n-1)! \, [(r-2)n + 1]!} \label{ur_result} \end{eqnarray} \hbox{\hskip 6pt\vrule width6pt height7pt depth1pt \hskip1pt}\bigskip \medskip \par\noindent {\bf Remarks.} 1. The proof presented here is a simplification of my original proof, based on independent suggestions by Paul Seymour, Dave Wagner and an anonymous referee. 2. {\em A posteriori}\/, we learn from Proposition \ref{prop4.3}(c,d) below that the power series \reff{Tr_def}/\reff{Ur_def} in fact define analytic functions in the disc $|z| < (r-2)^{r-2}/(r-1)^{r-1}$. 3. I suspect that Proposition \ref{prop4.2} is known somewhere in the graph-theory literature, but I do not know any reference. A weaker version of Proposition \ref{prop4.2} can be found in \cite[Lemma 5.4]{Dobrushin_96b}. \bigskip Let us also collect some properties of the numbers $t_n^{(r)}$ that arise in Proposition \ref{prop4.2}: \begin{proposition} \label{prop4.3} The quantities \begin{equation} t_n^{(r)} \;=\; r \, {[(r-1)n]! \over (n-1)! \, [(r-2)n + 2]!} \;, \end{equation} defined for integers $n,r \ge 1$, have the following properties: \begin{itemize} \item[(a)] $t_n^{(1)} = \cases{1 & for $n=1,2$ \cr 0 & for $n \ge 3$ \cr }$ \item[(b)] $t_n^{(2)} = n$ \item[(c)] As $n \to \infty$ at fixed $r \ge 3$, \begin{equation} t_n^{(r)} \;=\; {r \, (r-1)^{1/2} \over \sqrt{2\pi} \, (r-2)^{5/2}} \left( {(r-1)^{r-1} \over (r-2)^{r-2}} \right) ^{\! n} \, n^{-3/2} \left[ 1 \,+\, {{1 \over r-1} - {37 \over r-2} - 1 \over 12n} \,+\, O\!\left( {1 \over n^2} \right) \right] \;. \end{equation} \item[(d)] For all $n$ and all $r \ge 3$, \begin{equation} t_n^{(r)} \;\le\; \left( {(r-1)^{r-1} \over (r-2)^{r-2}} \right) ^{\! n-1} \;\le\; [e(r-{\textstyle{3 \over 2}})]^{n-1} \;. \end{equation} \item[(e)] As $r \to \infty$ at fixed $n \ge 1$, \begin{equation} t_n^{(r)} \;=\; {(rn)^{n-1} \over n!} \left[ 1 \,-\, {3(n-1)(n-2) \over 2nr} \,+\, O\!\left( {1 \over r^2} \right) \right] \;. \end{equation} \item[(f)] For all $r,n \ge 1$, \begin{equation} t_n^{(r)} \;\le\; {(rn)^{n-1} \over n!} \;. \end{equation} \end{itemize} \end{proposition} \par\medskip\noindent{\sc Proof.\ } (a) and (b) are trivial, while (c) and (e) follow from Stirling's formula. (f) is trivial for $r=1$, while for $r \ge 2$ it follows immediately from \begin{equation} {[(r-1)n]! \over [(r-2)n + 2]!} \;=\; {(rn-n)! \over (rn-2n+2)!} \;\le\; (rn)^{n-2} \;. \end{equation} The first inequality in (d) is obvious for $n=1$, so assume $n \ge 2$. We have \begin{eqnarray} t_n^{(r)} & = & {rn \over [(r-2)n + 1] [(r-2)n + 2]} \, {(r-1)n \choose n} \nonumber \\[1mm] &\le& {r \over (r-2)^2 n} \, {(r-1)n \choose n} \nonumber \\[1mm] &\le& {r (r-1)^{1/2} \over \sqrt{2\pi} (r-2)^{5/2}} \, n^{-3/2} \left( {(r-1)^{r-1} \over (r-2)^{r-2}} \right) ^{\! n} \nonumber \\[1mm] & = & {r (r-1)^{r - {1 \over 2}} \over \sqrt{2\pi} (r-2)^{r + {1 \over 2}}} \, n^{-3/2} \left( {(r-1)^{r-1} \over (r-2)^{r-2}} \right) ^{\! n-1} \;, \end{eqnarray} where the second inequality uses Lemma \ref{lemma4.4} below. Then straightforward calculus shows that the function $F(r) = r (r-1)^{r - {1 \over 2}} / [\sqrt{2\pi} (r-2)^{r + {1 \over 2}}]$ is decreasing on $r>2$; and we have $F(3) = 12/\sqrt{\pi} < 4^{3/2}$, $F(4) = (27/8) \sqrt{3/\pi} < 3^{3/2}$ and $F(5) = (1280/243) \sqrt{2/(3\pi)} < 2^{3/2}$. So the first inequality in (d) follows except for the cases $(r,n) = (3,2), \, (3,3)$ and $(4,2)$, which can be checked by hand. The final inequality in (d) follows from \begin{eqnarray} \log {\sigma^\sigma \over (\sigma-1)^{\sigma-1}} & = & \log\sigma \,+\, (\sigma-1) \log {\sigma \over \sigma-1} \nonumber \\[2mm] & = & \log\sigma \,+\, 1 \,-\, \sum\limits_{k=1}^\infty {\sigma^{-k} \over k(k+1)} \nonumber \\[2mm] &\le& \log\sigma \,+\, 1 \,-\, \sum\limits_{k=1}^\infty {\sigma^{-k} \over k 2^k} \nonumber \\[2mm] & = & \log\sigma \,+\, 1 \,+\, \log\!\left(1 - {1 \over 2\sigma}\right) \end{eqnarray} where the sums are convergent for $\sigma > 1$, so that $\sigma^\sigma / (\sigma-1)^{\sigma-1} \le e(\sigma - {1 \over 2})$. \hbox{\hskip 6pt\vrule width6pt height7pt depth1pt \hskip1pt}\bigskip \begin{lemma} \label{lemma4.4} Let $n \ge 2$ and $1 \le k \le n-1$ be integers. Then \begin{equation} {n \choose k} \;<\; \left( {n \over k} \right) ^{\! k} \left( {n \over n-k} \right) ^{\! n-k} \sqrt{ {n \over 2\pi k (n-k)} } \;. \end{equation} \end{lemma} \par\medskip\noindent{\sc Proof.\ } We use the following strong form of Stirling's formula \cite[pp.~45--46]{Chow_97}: for integer $n \ge 1$, \begin{equation} \log n! \;=\; (n + \smfrac{1}{2} ) \log n \,-\, n \,+\, \log \sqrt{2\pi} \,+\, \epsilon_n \end{equation} with \begin{equation} {1 \over 12n+1} \;<\; \epsilon_n \;<\; {1 \over 12n} \;\;. \end{equation} (The proof in \cite{Chow_97} is valid only for $n \ge 2$, but $\epsilon_1 = 1 - \log \sqrt{2\pi} \approx 0.08106$ clearly satisfies $1/13 < \epsilon_1 < 1/12$.) Then \begin{eqnarray} \epsilon_n - \epsilon_k - \epsilon_{n-k} & < & {1 \over 12n} \,-\, {1 \over 12k+1} \,-\, {1 \over 12(n-k) + 1} \nonumber \\[1mm] & = & {-144[n^2 - k(n-k)] - 12n + 1 \over 12n \, (12k+1) \, [12(n-k) + 1] } \nonumber \\[1mm] & < & 0 \;. \end{eqnarray} \hbox{\hskip 6pt\vrule width6pt height7pt depth1pt \hskip1pt}\bigskip Proposition \ref{prop4.2} clearly gives the best possible bound for $C_{\bullet,m}(G, x)$ and $T_n(G, x)$ in terms of the maximum degree of $G$, since it is sharp when $G = {\bf T}_r$. On the other hand, Proposition \ref{prop4.2} is somewhat unnatural for general (unequal) edge weights $\{v_e\}$, since adding an edge of small weight $v_e$ makes little change in $C_{n,m}(G, \{ v_e \}, x)$ but can cause the bound to jump (in case it increases the maximum degree). It is of interest, therefore, to find alternative bounds that depend ``smoothly'' on the weights $\{v_e\}$. We shall now give two such bounds (Propositions \ref{prop4.5} and \ref{prop4.6}). Unfortunately, both of them are strictly weaker than Proposition \ref{prop4.2} when the edge weights are equal, and neither one is strictly stronger than the other. \begin{proposition} \label{prop4.5} Let $G=(V,E)$ be a finite or countably infinite loopless undirected graph equipped with edge weights $\{ v_e \}_{e \in E}$. Then for any $x \in V$, \begin{subeqnarray} C_{\bullet,m}(G, \{ v_e \}, x) & \le & {(m+1)^m \over (m+1)!} \, \left( \sup\limits_{i \in V} \sum\limits_{e \ni i} |v_e| \right) ^{\! m} \\[2mm] & \le & \left( e \, \sup\limits_{i \in V} \sum\limits_{e \ni i} |v_e| \right) ^{\! m} \;. \slabel{eq.prop4.5.b} \end{subeqnarray} [The $e$ in front of the sup in \reff{eq.prop4.5.b} denotes, of course, the base of the natural logarithms!] \end{proposition} \par\medskip\noindent{\sc Proof.\ } As in the proof of Proposition \ref{prop4.2}, we pass to the universal covering graph $U = (\widetilde{V}, \widetilde{E})$ of $G$ with covering map $f \colon\, U \to G$; and we define the weight $v_{\widetilde{e}}$ of an edge $\widetilde{e} \in \widetilde{E}$ to be the weight $v_e$ of its image $e = f(\widetilde{e})$. It then follows, as in Proposition \ref{prop4.2}, that $C_{\bullet,m}(G, \{v_e\}, x) \le C_{\bullet,m}(U, \{v_{\widetilde{e}}\}, \widetilde{x})$. Let us now define, for each vertex $x \in \widetilde{V}$, the formal generating function \begin{equation} C_x(z) \;=\; \sum\limits_{m=0}^\infty C_{\bullet,m}(U, \{v_{\widetilde{e}}\}, x) \, z^m \;. \end{equation} Then the recursive structure of rooted trees implies that \begin{subeqnarray} C_x(z) & \preceq & \prod\limits_{y \sim x} [1 + |v_{xy}| z C_y(z)] \\[2mm] & \preceq & \prod\limits_{y \sim x} e^{|v_{xy}| z C_y(z)} \end{subeqnarray} where $y \sim x$ denotes that $y$ is adjacent to $x$, $xy$ denotes the (unique) corresponding edge, and $\preceq$ denotes coefficientwise inequality at all orders in $z$; the second inequality holds because $1 + \alpha z \preceq e^{\alpha z}$ for $\alpha \ge 0$. It then follows, by induction on the power of $z$, that $C_x(z) \preceq \bar{C}(z)$ for all $x$, where $\bar{C}(z)$ is determined by the equation \begin{equation} \bar{C}(z) \;=\; e^{\mu z \bar{C}(z)} \end{equation} with \begin{equation} \mu \;=\; \sup\limits_{x \in \widetilde{V}} \sum\limits_{y \sim x} |v_{xy}| \;=\; \sup\limits_{i \in V} \sum\limits_{e \ni i} |v_e| \;. \end{equation} The coefficients of $\bar{C}(z)$ can be determined by applying the Lagrange Implicit Function Theorem to $\bar{U}(z) = z \bar{C}(z)$, and we have the well-known (e.g. \cite[p.~392]{Knuth_73}) result \begin{equation} \bar{C}(z) \;=\; \sum_{m=0}^\infty {(m+1)^m \over (m+1)!} \mu^m z^m \;. \end{equation} Finally, the inequality $e^m \ge (m+1)^m / (m+1)!$ is trivial for $m=0,1$, and for $m \ge 2$ it follows from $e^x \ge x^n/n!$ by setting $x = n = m+1$. \hbox{\hskip 6pt\vrule width6pt height7pt depth1pt \hskip1pt}\bigskip \medskip \par\noindent {\bf Remarks.} 1. The proof presented here is a simplification of my original proof, based on the suggestions of an anonymous referee. 2. Proposition \ref{prop4.5} holds also for graphs with loops, but they impose slight technical complications. \bigskip An alternative estimate is due to Campanino {\em et al.}\/ \cite[p.~129]{Campanino_79} (see also \cite[p.~522]{Cammarota_82} and \cite[pp.~463--464]{Simon_93}): \begin{proposition} \label{prop4.6} Let $G=(V,E)$ be a finite or countably infinite undirected graph equipped with edge weights $\{ v_e \}_{e \in E}$. Define the matrix $M = (M_{xy})_{x,y \in V}$ by \begin{equation} M_{xy} \;=\; \sum_{e\colon\; \hboxscript{$e$ connects $x$ to $y$}} \! |v_e|^{1/2} \;. \end{equation} Then for any $x \in V$, \begin{equation} C_{\bullet,m}(G, \{ v_e \}, x) \;\le\; (M^{2m})_{xx} \;\le\; \left( \sup\limits_{i \in V} \sum\limits_{e \ni i} |v_e|^{1/2} \right) ^{\! 2m} \;. \end{equation} \end{proposition} \par\medskip\noindent{\sc Proof.\ } Let $G' = (V',E')$ be a connected subgraph of $G$ having $m$ edges. Then for any vertex $x \in V'$, there exists a path on $G'$ starting and ending at $x$ that uses each edge $e \in E'$ exactly twice. (Proof: The multigraph formed by doubling each edge of $G'$ is Eulerian. Alternate proof: By induction on $m$.\footnote{ See e.g.\ \cite[Lemma V.7.A.2]{Simon_93}. }) Conversely, every path on $G$ starting and ending at $x$ corresponds in this way to at most one subgraph $G'$. The claim follows. \hbox{\hskip 6pt\vrule width6pt height7pt depth1pt \hskip1pt}\bigskip \bigskip Let us conclude by examining the relative sharpness of these bounds when $G$ is an $r$-regular graph and the edge weights $v_e$ are equal. Then the tree bound $t_n^{(r)}$ of Proposition \ref{prop4.2} grows as $n\to\infty$ at an exponential rate $(r-1)^{r-1} / (r-2)^{r-2}$: this is less than $e(r-{3 \over 2})$ for all $r$, and behaves as $e [ r - {3 \over 2} - O(1/r)]$ as $r \to\infty$. The bound of Proposition \ref{prop4.5} is slightly weaker: it grows at exponential rate $er$. Finally, the bound of Proposition \ref{prop4.6} grows at exponential rate $r^2$, which is vastly weaker for large $r$ but is slightly better when $r=2$. In particular, when $G$ is a regular lattice, it can be shown by supermultiplicativity arguments \cite{Klarner_67,Klein_81,Whittington_90,Janse_92} that the limits \begin{eqnarray} \lambda_o(G) & = & \lim\limits_{n\to\infty} T_n(G,x)^{1/n} \\[2mm] \lambda_b(G) & = & \lim\limits_{m\to\infty} C_{\bullet,m}(G,x)^{1/m} \end{eqnarray} exist. For the simple hypercubic lattice ${\mathbb Z}^d$ with nearest-neighbor bonds, these growth constants have been computed (non-rigorously) in a large-$d$ asymptotic expansion \cite{Gaunt_94,Peard_95} (see also \cite{Penrose_94,Hara_95,Derbez_98} for related rigorous results): \begin{eqnarray} \log\lambda_o({\mathbb Z}^d) & = & \log\sigma \,+\, 1 \,-\, {1 \over 2} \sigma^{-1} \,-\, {8 \over 3} \sigma^{-2} \,-\, \ldots \\[3mm] \log\lambda_b({\mathbb Z}^d) & = & \log\sigma \,+\, 1 \,-\, {1 \over 2} \sigma^{-1} \,- \left( {8 \over 3} - {1 \over 2e} \right) \sigma^{-2} \,-\, \ldots \end{eqnarray} where $\sigma = r-1 = 2d-1$. Let us compare this with the tree bound of Proposition \ref{prop4.2}: \begin{equation} \log {(r-1)^{r-1} \over (r-2)^{r-2}} \;=\; \log\sigma \,+\, 1 \,-\, {1 \over 2} \sigma^{-1} \,-\, {1 \over 6} \sigma^{-2} \,-\, \ldots \;. \end{equation} Thus, the latter bound is very close to sharp for $G = {\mathbb Z}^d$ in high dimension $d$, confirming the intuition that high-dimensional regular lattices are ``like trees'' to leading order in $1/d$. \section{Application to the Potts-Model Partition Function} \label{sec5} We are now ready for the main theorem of this paper: \begin{theorem} \label{thm5.1} Let $G=(V,E)$ be a loopless finite undirected graph equipped with complex edge weights $\{ v_e \}_{e \in E}$ satisfying $|1 + v_e| \le 1$ for all $e$. Let $Q = Q(G, \{ v_e \}) > 0$ be the smallest number for which \begin{equation} \inf\limits_{\alpha > 0} \alpha^{-1} \sum\limits_{n=2}^\infty e^{\alpha n} \, Q^{-(n-1)} \, \max\limits_{x \in V} T_n(G, \{ v_e \}, x) \;\,\le\;\, 1 \;. \label{eq5.1} \end{equation} [Note that $Q$ is automatically finite, since $T_n(G, \{ v_e \}, x) = 0$ for $n > |V|$.] Then all the zeros of $Z_G(q, \{v_e\})$ lie in the disc $|q| < Q$. \end{theorem} \par\medskip\noindent{\sc Proof.\ } Starting from the polymer-gas representation \reff{eq2.2}--\reff{eq2.4} of $Z_G(q, \{v_e\})$, we apply Theorem \ref{thm3.1} and Proposition \ref{prop3.2} with the choice $R_S = |w(S)|$. We verify hypothesis (a) of Proposition \ref{prop3.2} by verifying \reff{partition_condition_2} with \begin{equation} A_n \;=\; \max_{x \in V} \sum_{\begin{scarray} S \ni x \\ |S| = n \end{scarray}} |w(S)| \;. \label{eq5.2} \end{equation} Now we use Proposition \ref{prop_penrose} to conclude that $w(S)$ can be bounded by a sum over trees: \begin{equation} |w(S)| \;\le\; |q|^{-(|S|-1)} \sum\limits_{\begin{scarray} \widetilde{E} \subseteq E \\ (S,\widetilde{E}) \hboxscript{ tree} \end{scarray}} \; \prod\limits_{e \in \widetilde{E}} |v_e| \;. \end{equation} Inserting this into \reff{eq5.2}, we get \begin{equation} A_n \;\le\; |q|^{-(n-1)} \, \max_{x \in V} T_n(G, \{ v_e \}, x) \;. \end{equation} If $|q| \ge Q$, hypothesis (b) of Proposition \ref{prop3.2} holds (recall Remark 4 following that Proposition) and hence $Z_G(q, \{v_e\}) \neq 0$. \hbox{\hskip 6pt\vrule width6pt height7pt depth1pt \hskip1pt}\bigskip In applying Theorem \ref{thm5.1} we are of course free to use any convenient upper bound on $\max_{x \in V} T_n(G, \{ v_e \}, x)$. In particular, when $G$ has maximum degree $\le r$, Proposition \ref{prop4.2} provides such a bound. Recall that \begin{equation} t_n^{(r)} \;=\; r \, {[(r-1)n]! \over (n-1)! \, [(r-2)n + 2]!} \;, \end{equation} and let $C = C(r) > 0$ be the smallest number for which \begin{equation} \inf\limits_{\alpha > 0} \alpha^{-1} \sum\limits_{n=2}^\infty e^{\alpha n} \, C^{-(n-1)} \, t_n^{(r)} \;\le\; 1 \;. \label{Falpha} \end{equation} The following is then an immediate consequence of Theorem \ref{thm5.1} and Proposition \ref{prop4.2}: \begin{corollary} \label{cor5.2} Let $G=(V,E)$ be a loopless finite undirected graph of maximum degree $\le r$, equipped with complex edge weights $\{ v_e \}_{e \in E}$ satisfying $|1 + v_e| \le 1$ for all $e$. Let $v_{max} = \max\limits_{e \in E} |v_e|$. Then all the zeros of $Z_G(q, \{v_e\})$ lie in the disc $|q| < C(r) v_{max}$. \end{corollary} \noindent And for the chromatic polynomials: \begin{corollary} \label{cor5.3} Let $G=(V,E)$ be a loopless finite undirected graph of maximum degree $\le r$. Then all the zeros of $P_G(q)$ lie in the disc $|q| < C(r)$. \end{corollary} Table \ref{table1} lists rigorous upper bounds on $C(r)$ for $2 \le r \le 20$, proven (with the assistance of {\sc Mathematica}) as follows: After computing numerically an approximate value of $C(r)$,\footnote{ Using the generating function $f(z) = \sum_{n=2}^\infty t_n^{(r)} z^n$ where $z = e^{\alpha}/C$, I solve simultaneously the equations $f(z) = (\log z + \log C) / C$ and $f'(z) = 1/(Cz)$ by solving numerically $f(z) = - z f'(z) \log f'(z)$ and then plugging back in to determine $C = 1/[z f'(z)]$ and $\alpha = - \log f'(z)$. } I added $10^{-6}$ and rounded it upwards to a rational number $p/10^6$. (Thus, the value reported in Table \ref{table1} exceeds my best estimate of $C(r)$ by at most $2 \times 10^{-6}$.) I likewise approximated the numerically-found $\alpha$ by a rational number $p'/10^6$. Thereafter I did all computations in exact rational arithmetic. First I computed a rational upper bound on $e^\alpha$ (differing from the true $e^{\alpha}$ by at most $2 \times 10^{-10}$) by truncating the Taylor series for $e^{-\alpha}$ at odd order (here ninth or eleventh) to obtain a lower bound on $e^{-\alpha}$. Finally, I computed an upper bound on \reff{Falpha} by summing the terms explicitly through $n=$ some $n_0$ and bounding the tail of the series ($n \ge n_0 + 1$) using Proposition \ref{prop4.3}(d); I systematically increased $n_0$ until the inequality \reff{Falpha} was verified. For $r=2$, of course, I just summed the series exactly. As $r \to \infty$ we have the following: \begin{table} \begin{center} \begin{tabular}{rrrrc} \multicolumn{1}{c}{$r$} & \multicolumn{1}{c}{$C(r)$} & \multicolumn{1}{c}{$\alpha$} & \multicolumn{1}{c}{$e^{\alpha}$} & \multicolumn{1}{c}{$n_0$} \\ \hline \\[-4mm] 2 & 13.234367 & 0.453177 & 1.5733026366 & $\infty$ \\ 3 & 21.144294 & 0.436884 & 1.5478765131 & 15 \\ 4 & 29.081607 & 0.428653 & 1.5351882318 & 18 \\ 5 & 37.029702 & 0.423694 & 1.5275940786 & 20 \\ 6 & 44.983130 & 0.420382 & 1.5225430561 & 22 \\ 7 & 52.939585 & 0.418013 & 1.5189404206 & 23 \\ 8 & 60.897921 & 0.416236 & 1.5162436603 & 24 \\ 9 & 68.857505 & 0.414852 & 1.5141466306 & 25 \\ 10 & 76.817961 & 0.413745 & 1.5124713976 & 25 \\ 11 & 84.779049 & 0.412839 & 1.5111017191 & 26 \\ 12 & 92.740610 & 0.412084 & 1.5099612679 & 26 \\ 13 & 100.702534 & 0.411445 & 1.5089967109 & 26 \\ 14 & 108.664743 & 0.410898 & 1.5081715154 & 27 \\ 15 & 116.627179 & 0.410423 & 1.5074553040 & 27 \\ 16 & 124.589800 & 0.410008 & 1.5068298399 & 28 \\ 17 & 132.552573 & 0.409641 & 1.5062769348 & 28 \\ 18 & 140.515473 & 0.409315 & 1.5057859685 & 28 \\ 19 & 148.478479 & 0.409024 & 1.5053478486 & 28 \\ 20 & 156.441575 & 0.408761 & 1.5049519941 & 28 \end{tabular} \end{center} \caption{ Upper bounds on $C(r)$ for $2 \le r \le 20$; they differ from my best estimate of the true $C(r)$ by at most $2 \times 10^{-6}$. The third column gives a value of $\alpha$ for which (\protect\ref{Falpha}) is proven to be $\le 1$. The fourth column gives the upper bound on $e^{\alpha}$ employed in this proof. The fifth column gives the number of terms explicitly summed in the series. } \label{table1} \end{table} \begin{proposition} \label{prop5.4} Let $K \approx 7.963906\ldots$ be the smallest number for which \begin{equation} \inf\limits_{\alpha > 0} \alpha^{-1} \sum\limits_{n=2}^\infty e^{\alpha n} \, K^{-(n-1)} \, {n^{n-1} \over n!} \;\le\; 1 \;. \end{equation} Then $C(r) \le Kr$ for all $r$, and $\lim\limits_{r \to\infty} C(r)/r = K$. Moreover, we have the rigorous bound $K \le 7.963907$. \end{proposition} \par\medskip\noindent{\sc Proof.\ } Clearly $\widetilde{C}(r) \equiv C(r)/r$ is the smallest number for which \begin{equation} \inf\limits_{\alpha > 0} \alpha^{-1} \sum\limits_{n=2}^\infty e^{\alpha n} \, \widetilde{C}^{-(n-1)} \, {t_n^{(r)} \over r^{n-1}} \;\le\; 1 \;. \end{equation} It follows from Proposition \ref{prop4.3}(f) that $\widetilde{C}(r) \le K$ for all $r$. Now suppose that we were to have $\liminf\limits_{r \to \infty} \widetilde{C}(r) \le K-\epsilon < K$. Then there would exist infinite sequences $\{ r_i \} \uparrow \infty$ and $\{ \alpha_i \}$ such that \begin{equation} \alpha_i^{-1} \sum\limits_{n=2}^\infty e^{\alpha_i n} \, (K-\epsilon)^{-(n-1)} \, {t_n^{(r_i)} \over r_i^{n-1}} \;\le\; 1 \label{prop5.4_star1} \end{equation} for all $i$. Now the finiteness of \reff{prop5.4_star1} implies that the $\alpha_i$ are bounded [e.g.\ from Proposition \ref{prop4.3}(c) we have $e^{\alpha_i} \le (K-\epsilon) r_i (r_i -2)^{r_i -2} / (r_i -1)^{r_i -1} \le {3 \over 4} (K-\epsilon)$ whenever $r_i \ge 3$]. So we can extract a subsequence of $\{ \alpha_i \}$ that converges to some value $\alpha_*$. Then Proposition \ref{prop4.3}(e,f) and the dominated convergence theorem imply that \begin{equation} \alpha_*^{-1} \sum\limits_{n=2}^\infty e^{\alpha_* n} \, (K-\epsilon)^{-(n-1)} \, {n^{n-1} \over n!} \;\le\; 1 \;, \end{equation} which contradicts the definition of $K$. Finally, it is easy to prove that $K \le 7.963907$, by a computer-assisted method similar to that used above for $C(r)$. For the tail of the series ($n \ge n_0 + 1$), it suffices to use the crude bound $n^{n-1}/n! \le e^{n-1} \le 3^{n-1}$. The proof succeeds with the choices $\alpha = 0.403774$, $e^\alpha \le 1.4974655$ and $n_0 = 32$. \hbox{\hskip 6pt\vrule width6pt height7pt depth1pt \hskip1pt}\bigskip If we employ Proposition \ref{prop4.5} in place of Proposition \ref{prop4.2}, Theorem \ref{thm5.1} yields the following: \begin{corollary} \label{cor5.5} Let $G=(V,E)$ be a loopless finite undirected graph equipped with complex edge weights $\{ v_e \}_{e \in E}$ satisfying $|1 + v_e| \le 1$ for all $e$. Then all the zeros of $Z_G(q, \{v_e\})$ lie in the disc $|q| < K \max\limits_{i \in V} \sum\limits_{e \ni i} |v_e|$, where $K$ is the constant defined in Proposition \ref{prop5.4}. \end{corollary} \medskip \par\noindent {\bf Remarks.} 1. Since $Z_G(q, \{v_e\})/q$ for any graph $G$ is the product of the same quantity over the blocks of $G$, it is legitimate to apply Theorem \ref{thm5.1} and its corollaries separately to each block. This can lead to large improvements (consider e.g.\ trees). 2. What happens if we drop the assumption that $|1 + v_e| \le 1$? Because we can no longer use Proposition \ref{prop_penrose} to reduce the sum to trees, we need to consider all $n$-vertex connected subgraphs of $G$ containing a given vertex $x$. But the number $m$ of {\em edges}\/ in such a subgraph could be as large as $\lfloor rn/2 \rfloor$ (where $r$ is the maximum degree of $G$). Therefore, the factor $t_n^{(r)} v_{max}^{n-1}$ coming from \reff{eq4.17} has to be replaced by a factor $\sum_{m=n-1}^{\lfloor rn/2 \rfloor} t_{m+1}^{(r)} v_{max}^m$ coming from \reff{eq4.16} [or the analogue from Proposition \ref{prop4.5}]. As a consequence, the radius of the $q$-plane disc containing all the zeros of $Z_G(q, \{v_e\})$ will scale as $\max[v_{max}, v_{max}^{r/2}]$ rather than simply $v_{max}$. And this is not simply an artifact of the method of proof: for the $q$-state Potts {\em ferromagnet}\/ ($v > 0$, $q > 0$) on the the simple hypercubic lattice ${\mathbb Z}^d$ with nearest-neighbor bonds, the first-order phase-transition point $v_t$ indeed behaves as \begin{equation} v_t(q) \;=\; q^{1/d} \, [1 + O(1/q)] \end{equation} as $q \to +\infty$ \cite{Martirosian_86,Laanait_86,Kotecky_90,Laanait_91,Borgs_91}. In the Whitney--Tutte--Fortuin--Kasteleyn polynomial \reff{eq1.2}, this reflects the coexistence at $v=v_t$ (for all $q \gg 1$) between a phase with a low density of occupied edges and a phase with a high density of occupied edges. \section{Some Generalizations} \label{sec6} The following generalization of the Whitney--Tutte--Fortuin--Kasteleyn polynomial \reff{eq1.2} is motivated by some work of Tutte \cite{Tutte_47,Tutte_76}, Farrell \cite{Farrell_79} and Stanley \cite{Stanley_95,Stanley_98} as well as by the statistical-mechanical application to be discussed below. Let us replace the single complex number $q$ by a map ${\sf q}\colon\, {\cal P}_*(V) \to {\mathbb C}$, and define \begin{equation} Z_G({\sf q}, \{v_e\}) \;=\; \sum_{ E' \subseteq E } \left( \prod_{i=1}^{k(E')} {\sf q}(V_i) \right) \left( \prod_{e \in E'} v_e \right) \;, \label{eq6.1} \end{equation} where $(V_1,E_1), \ldots, (V_{k(E')}, E_{k(E')})$ are the connected components of $(V,E')$. We immediately deduce an analogue of Proposition \ref{prop2.1}: the identity \reff{eq2.2} is replaced by \begin{equation} Z_G({\sf q}, \{v_e\}) \;=\; \left( \prod\limits_{x \in V} {\sf q}(\{x\}) \right) Z_{polymer,G}({\sf q}, \{v_e\}) \;, \label{eq6.2} \end{equation} and the fugacities $w(S)$ are now given by \begin{equation} w(S) \;=\; \cases{ {\displaystyle {\sf q}(S) \over \displaystyle \prod\limits_{x \in S} {\sf q}(\{x\}) } \!\! \sum\limits_{\begin{scarray} \widetilde{E} \subseteq E \\ (S,\widetilde{E}) \hboxscript{ connected} \end{scarray}} \prod\limits_{e \in \widetilde{E}} v_e & if $|S| \ge 2$ \cr \noalign{\vskip 2mm} 0 & if $|S| \le 1$ \cr } \label{eq6.3} \end{equation} The proof of Theorem \ref{thm5.1} then goes through without change, and yields: \begin{theorem} \label{thm6.1} Let $G=(V,E)$ be a loopless finite undirected graph equipped with complex edge weights $\{ v_e \}_{e \in E}$ satisfying $|1 + v_e| \le 1$ for all $e$, and let ${\sf q}\colon\, {\cal P}_*(V) \to {\mathbb C}$. Let $\{ Q_n \} _{n=1}^\infty$ be a sequence of positive numbers satisfying \begin{equation} \inf\limits_{\alpha > 0} \alpha^{-1} \sum\limits_{n=2}^\infty e^{\alpha n} \, Q_n^{-1} \, \max\limits_{x \in V} T_n(G, \{ v_e \}, x) \;\,\le\;\, 1 \;, \label{eq6.4} \end{equation} and assume that \begin{equation} \left| {{\sf q}(S) \over \prod\limits_{x \in S} {\sf q}(\{x\})} \right| \;\le\; Q_{|S|}^{-1} \label{eq6.5} \end{equation} for all nonempty subsets $S \subseteq V$. Then $Z_G({\sf q}, \{v_e\}) \neq 0$. \end{theorem} The following special case is of particular interest: Fix an integer $N \ge 0$, and for each $x \in V$ choose a vector $u_x = (u_x^{(1)}, \ldots, u_x^{(N)}) \in {\mathbb C}^N$. Then define \begin{equation} {\sf q}(S) \;=\; q \,-\, N \,+\, \sum_{i=1}^N \prod_{x \in S} (1+ u_x^{(i)}) \label{eq6.star1} \end{equation} where $q$ is a fixed complex number. This corresponds to a $q$-state Potts model in a magnetic field $h_x = (h_x^{(1)}, \ldots, h_x^{(N)})$ in the first $N$ spin directions, where $u_x^{(i)} = \exp(h_x^{(i)}) - 1$. To see this, we first define, for each integer $q \ge N$, the partition function for the $q$-state Potts model in a magnetic field, generalizing \reff{eq1.1}: \begin{equation} Z_G(q, \{v_e\}, \{u_x^{(i)}\}) \;=\; \sum_{ \{\sigma_x\} } \, \prod_{e \in E} \, \biggl[ 1 + v_e \delta(\sigma_{x_1(e)}, \sigma_{x_2(e)}) \biggr] \, \prod_{x \in V} \prod_{i=1}^N \, \biggl[ 1 + u_x^{(i)} \delta(\sigma_x, i) \biggr] \;. \label{eq6.star2} \end{equation} Now expand out the product over $e \in E$, and let $E' \subseteq E$ be the set of edges for which the term $v_e \delta_{\sigma_{x_1(e)}, \sigma_{x_2(e)}}$ is taken. Then perform the sum over configurations $\{ \sigma_x \}$: in each connected component of the subgraph $(V,E')$ the spin value $\sigma_x$ must be constant, and there are no other constraints. The sum over possible spin values in a connected component with vertex set $S$ yields \reff{eq6.star1}. It follows that, for any integer $q \ge N$, the partition function $Z_G(q, \{v_e\}, \{u_x^{(i)}\})$ equals $Z_G({\sf q}, \{v_e\})$ with weights \reff{eq6.star1}. We then take the latter, which is a polynomial in $q$, $\{v_e\}$ and $\{u_x^{(i)}\}$, as the {\em definition}\/ of $Z_G(q, \{v_e\}, \{u_x^{(i)}\})$ for general complex $q$. The following lemma gives a sufficient condition for the applicability of Theorem \ref{thm6.1} to this situation: \begin{lemma} \label{lemma6.2} Let ${\sf q}(S)$ be defined by \reff{eq6.star1}. \begin{itemize} \item[(a)] If $N=1$ and each $u_x^{(i)}$ equals either $0$ or $-1$, then \begin{equation} \left| {{\sf q}(S) \over \prod\limits_{x \in S} {\sf q}(\{x\})} \right| \;\le\; \min(|q|, |q-1|)^{-(|S|-1)} \;. \label{eq_lemma6.2a} \end{equation} \item[(b)] If $-1 \le u_x^{(i)} \le 0$ for all $x,i$ and $|q| > N$, then \begin{equation} \left| {{\sf q}(S) \over \prod\limits_{x \in S} {\sf q}(\{x\})} \right| \;\le\; (|q| - N)^{-(|S|-1)} \;. \label{eq_lemma6.2b} \end{equation} \item[(c)] If $|1 + u_x^{(i)}| \le 1$ for all $x,i$ and $|q-N| > N$, then \begin{equation} \left| {{\sf q}(S) \over \prod\limits_{x \in S} {\sf q}(\{x\})} \right| \;\le\; {|q-N| + N \over (|q-N|-N)^{|S|} } \;. \label{eq_lemma6.2c} \end{equation} \end{itemize} \end{lemma} \noindent The proof of Lemma \ref{lemma6.2} is deferred to the end of this section. We can exploit this example to obtain new results for the ordinary (zero-field) Potts-model partition function $Z_G(q,\{v_e\})$ and in particular for the chromatic polynomial $P_G(q)$, by employing a variant of the ``ghost spin'' trick of Suzuki \cite{Suzuki_65} and Griffiths \cite{Griffiths_67}. Given a finite graph $G_0 = (V_0, E_0)$ and an integer $N \ge 1$, we define $G$ to be the join of $G_0$ with the complete graph on $N$ vertices. Thus, the vertex set of $G$ is $V = V_0 \bigcup \{y_1,\ldots,y_N\}$ (disjoint union) and the edge set is $E = E_0 \bigcup \{ \<x y_i\> \}_{x \in V_0, \, 1 \le i \le N} \bigcup \{ \<y_i y_j\> \}_{ 1 \le i < j \le N}$. We allow the edge weights $\{ v_e \} _{e \in E_0}$ and $\{ v_{\<x y_i\>} \} _{x \in V_0, \, 1 \le i \le N}$ to be arbitrary complex numbers, but we require that $v_{\<y_i y_j\>} = -1$ for $1 \le i < j \le N$ (this condition is crucial). We then have the identity \begin{equation} Z_G(q, \{v_e, v_{\<x y_i\>}, v_{\<y_i y_j\>} \}) \;=\; q_{\<N\>} \, Z_{G_0}(q, \{v_e\}, \{u_x^{(i)}\}) \;, \label{eq6.star3} \end{equation} where $q_{\<N\>} = q (q-1) \cdots (q-N+1)$ is the $N$-th ``falling factorial'' polynomial and $u_x^{(i)} = v_{\<x y_i\>}$. This is most easily proven in the Potts spin representation \reff{eq1.1}/\reff{eq6.star2}: Let $q$ be an integer $\ge N$, and let us compute the left-hand side of \reff{eq6.star3}. There are $q_{\<N\>}$ admissible ways to color the vertices $\{y_1,\ldots,y_N\}$, all of which are equivalent modulo permutations of $\{1,\ldots,q\}$; and with any such coloring fixed, the sum over colorings of $V_0$ yields precisely $Z_{G_0}(q, \{v_e\}, \{u_x^{(i)}\})$ with $u_x^{(i)} = v_{\<x y_i\>}$. Since both sides of \reff{eq6.star3} are polynomials in $q$ and the equality holds for infinitely many values of $q$, it must hold identically. By applying Theorem \ref{thm6.1} to the graph $G_0$, we can obtain new results for the ordinary Potts-model partition function of the graph $G$. In particular, given {\em any}\/ graph $G=(V,E)$ and any vertex $y \in V$, we can interpret $G$ as the join of $G_0 \equiv G \setminus y$ (the graph obtained from $G$ by deleting $y$ and all edges incident on it) and $K_1$. (Any edge $\< xy \>$ that was not originally present in $G$ can be introduced and given $v_{\<xy\>} = 0$.) More generally, given any $N$-clique $y_1,\ldots,y_N$ of $G$, we can interpret $G$ as the join of $G_0 \equiv G \setminus \{y_1,\ldots,y_N\}$ and $K_N$; however, for $N>1$ we must require that $v_{\<y_i y_j\>} = -1$ for each pair $i \neq j$. Theorem \ref{thm6.1}, Lemma \ref{lemma6.2}(b,c) and Proposition \ref{prop4.2} then yield an extension of Corollary \ref{cor5.2}. To state it, we first define $\widetilde{C} = \widetilde{C}(r,N,\bar{v})$ to be the smallest number for which \begin{equation} \inf\limits_{\alpha > 0} \alpha^{-1} \sum\limits_{n=2}^\infty e^{\alpha n} \, {\widetilde{C}+N \over (\widetilde{C}-N)^n} \, \bar{v}^{n-1} \, t_n^{(r)} \;\le\; 1 \;. \end{equation} We then have: \begin{theorem} \label{thm6.3} Let $G=(V,E)$ be a loopless finite undirected graph in which all vertices have degree $\le r$ except perhaps for an $N$-clique $y_1,\ldots,y_N$. Let $G$ be equipped with complex edge weights $\{ v_e \}_{e \in E}$ satisfying $|1 + v_e| \le 1$ for all $e$ and $v_{\<y_i y_j\>} = -1$ for all $i \neq j$. Let $v_{max} = \max\limits_{e \in E_0} |v_e|$, where $E_0$ is the set of edges not incident on any of the vertices $y_1,\ldots,y_N$. Then: \begin{itemize} \item[(a)] All the zeros of $Z_G(q, \{v_e\})$ lie in the disc $|q-N| < \widetilde{C}(r,N,v_{max})$. \item[(b)] If, in addition, all the edges $e$ incident on any of the vertices $y_1,\ldots,y_N$ satisfy $-1 \le v_e \le 0$, then all the zeros of $Z_G(q, \{v_e\})$ lie in the disc $|q| < C(r) v_{max} + N$. \end{itemize} \end{theorem} \noindent And for the chromatic polynomials, we have, using Lemma \ref{lemma6.2}(a): \begin{corollary} \label{cor6.4} Let $G=(V,E)$ be a loopless finite undirected graph in which all vertices, except perhaps one, have degree $\le r$. Then all the zeros of $P_G(q)$ lie in the union of the discs $|q| < C(r)$ and $|q-1| < C(r)$. In particular, they all lie in the disc $|q| < C(r) + 1$. \end{corollary} \noindent Thus, the zeros of $P_G(q)$ can be bounded in terms of the {\em second-largest}\/ degree of a vertex in $G$. Such a result was recently conjectured by Shrock and Tsai \cite{Shrock_99a}; see Section \ref{sec7} for further discussion. Let us note that the phrase ``except perhaps one'' in Corollary \ref{cor6.4} {\em cannot}\/ be replaced here by ``except perhaps two'', not even in the case $r=2$. Indeed, I have elsewhere \cite{Sokal_hierarchical} constructed a family of planar graphs in which all but two vertices have degree 2 and whose chromatic roots are together dense in $\{ q \in {\mathbb C} \colon\; |q-1| \ge 1 \}$. Modifications of these graphs show also \cite{Sokal_hierarchical} that the condition $v_{\<y_i y_j\>} = -1$ for $i \neq j$ in Theorem \ref{thm6.3} (when $N>1$) cannot be relaxed. Let us now give the proof of Lemma \ref{lemma6.2}. We will need the following elementary fact: \begin{lemma} \label{lemma6.5} Let $z$ and $a$ be complex numbers. Then \begin{equation} |z + \lambda a|^2 \;\ge\; |z+a| \, (|z| - |a|) \end{equation} whenever $0 \le \lambda \le 1$. \end{lemma} \par\medskip\noindent{\sc Proof.\ } Simple calculus shows that \begin{equation} \min\limits_{0 \le \lambda \le 1} |z + \lambda a|^2 \;=\; \cases{ |z|^2 & if $\mathop{\rm Re}\nolimits(z^* a) \ge 0$ \cr \noalign{\vskip 2mm} |z|^2 - {\displaystyle \mathop{\rm Re}\nolimits(z^* a)^2 \over \displaystyle |a|^2} & if $-|a|^2 \le \mathop{\rm Re}\nolimits(z^* a) \le 0$ \cr \noalign{\vskip 2mm} |z+a|^2 & if $\mathop{\rm Re}\nolimits(z^* a) \le -|a|^2$ \cr } \end{equation} In the first two cases we clearly have \begin{equation} \min\limits_{0 \le \lambda \le 1} |z + \lambda a|^2 \;\ge\; |z|^2 - |a|^2 \;=\; (|z| + |a|) (|z| - |a|) \;\ge\; |z+a| \, (|z| - |a|) \;, \end{equation} while in the third case we have \begin{equation} \min\limits_{0 \le \lambda \le 1} |z + \lambda a|^2 \;=\; |z+a|^2 \;\ge\; |z+a| \, (|z| - |a|) \;. \end{equation} \hbox{\hskip 6pt\vrule width6pt height7pt depth1pt \hskip1pt}\bigskip \proofof{Lemma \ref{lemma6.2}} We use the shorthand $w(S) = {\sf q}(S) / \prod\limits_{x \in S} {\sf q}(\{x\})$. (a) Let $|S| = n$ and suppose that the sequence $(u_x^{(1)})_{x \in S}$ consists of $m$ $-1$'s and $n-m$ 0's. Then \begin{equation} w(S) \;=\; \cases{ q^{-(n-1)} & if $m=0$ \cr \noalign{\vskip 2mm} q^{-(n-m)} (q-1)^{-(m-1)} & if $1 \le m \le n$ \cr } \end{equation} from which \reff{eq_lemma6.2a} immediately follows. (b) Let $S = \{x_1,\ldots,x_n\}$; we then have ${\sf q}(\{x_j\}) = q + u_j$ and ${\sf q}(S) = q + \bar{u}$ with $-N \le \bar{u} \le u_1,\ldots,u_n \le 0$. Now apply Lemma \ref{lemma6.5} with $z=q$, $a = \bar{u}$ and $\lambda = u_j/\bar{u}$ for $j=1,2$: we have \begin{equation} \left| {{\sf q}(S) \over {\sf q}(\{x_1\}) \, {\sf q}(\{x_2\})} \right| \;\le\; {1 \over |q| - |\bar{u}|} \;\le\; {1 \over |q| - N} \end{equation} and hence \begin{equation} |w(S)| \;\le\; {1 \over |q| - N} \prod\limits_{j=3}^n {1 \over |q + u_j|} \;, \label{eq_lemma6.2b_better} \end{equation} which implies \reff{eq_lemma6.2b}. (c) This bound is trivially obtained by bounding the numerator and denominator separately. \hbox{\hskip 6pt\vrule width6pt height7pt depth1pt \hskip1pt}\bigskip \medskip \par\noindent {\bf Remarks.} 1. For simplicity, I have not bothered to exploit the full strength of \reff{eq_lemma6.2b_better}, which is quite a bit sharper than \reff{eq_lemma6.2b}. 2. I am not entirely happy with Lemma \ref{lemma6.2}, and I suspect that it can be improved. In particular, it is disconcerting that \reff{eq_lemma6.2b} is not uniformly stronger than \reff{eq_lemma6.2c}, even though the corresponding hypothesis on the $u_x^{(i)}$ {\em is}\/ strictly stronger. \section{Some Conjectures and Open Questions} \label{sec7} The bounds in this paper are, of course, far from sharp, and it is of some interest to speculate on what the best-possible results might be. Let us define \begin{equation} C_{opt}(r) \;=\; \max\{|q| \colon\; P_G(q) = 0 \hbox{ for some loopless graph $G$ of maximum degree $r$} \} \;. \end{equation} The example of the complete graph $K_{r+1}$ shows that $C_{opt}(r) \ge r$. It is easy to see that $C_{opt}(1) = 1$ and $C_{opt}(2) = 2$; and there is some evidence that $C_{opt}(3) = 3$.\footnote{ Biggs, Damerell and Sands \cite{Biggs_72} have verified that the chromatic roots of all 3-regular graphs with $\le 10$ vertices, as well as those of ladders (``prisms'') and M\"obius ladders of arbitrary length, lie in $|q| \le 3$. Read and Royle \cite{Read_91} have extended this verification to all 3-regular graphs with $\le 16$ vertices, as well as to some larger graphs. } But, at least for $r \ge 4$, $C_{opt}(r)$ must in fact be strictly larger than $r$, as is shown by numerical computations on the complete bipartite graph $K_{r,r}$ (see Table \ref{table2}).\footnote{ Recall \cite{Swenson_73,Laskar_75} that $$ P_{K_{m,n}}(q) \;=\; \sum_{k=0}^m S(m,k) \, q_{\<k\>} \, (q-k)^n \;, $$ where $S(m,k)$ is the Stirling number of the second kind (the number of ways of partitioning a set of $m$ elements into $k$ nonempty subsets) \cite[pp.~33--38]{Stanley_86} and $q_{\<k\>} = q (q-1) \cdots (q-k+1)$. See Woodall \cite[pp.~219--220]{Woodall_77} and Brown \cite{Brown_98a} for some properties of the chromatic zeros of the $K_{m,n}$. } Indeed, Gordon Royle (private communication) has conjectured that, for $r \ge 4$, $K_{r,r}$ is the graph of maximum degree $r$ having the largest chromatic roots (in modulus). It would be useful to have a better understanding of the chromatic zeros of the complete bipartite graphs $K_{m,n}$. In particular, it would be useful to have a {\em proof}\/ that $K_{r,r}$ has chromatic roots of magnitude $>r$ for all $r \ge 4$; and it would be valuable to understand the asymptotic behavior of the chromatic roots of $K_{m,n}$ as $m,n \to \infty$ in various ways (e.g.\ with $\alpha = m/n$ fixed). \begin{table} \begin{center} \begin{tabular}{rr@{$\:\pm\:$}r@{$\,i\quad$}r} \multicolumn{1}{c}{$r$} & \multicolumn{2}{c}{$q$} & \multicolumn{1}{c}{$|q|$} \\ \hline \\[-4mm] 2 & 1.500000 & 0.866025 & 1.732051 \\ 3 & 2.140640 & 1.948682 & 2.894772 \\ 4 & 2.802489 & 3.097444 & 4.177093 \\ 5 & 3.469365 & 4.291184 & 5.518221 \\ 6 & 4.138450 & 5.516667 & 6.896404 \\ 7 & 4.808805 & 6.765768 & 8.300616 \\ 8 & 5.480007 & 8.033190 & 9.724331 \\ 9 & 6.151830 & 9.315289 & 11.163316 \\ 10 & 6.824136 & 10.609446 & 12.614641 \\ 11 & 7.496833 & 11.913711 & 14.076186 \\ 12 & 8.169855 & 13.226591 & 15.546358 \\ 13 & 8.843156 & 14.546915 & 17.023928 \\ 14 & 9.516697 & 15.873744 & 18.507925 \\ 15 & 10.190450 & 17.206318 & 19.997566 \\ 16 & 10.864391 & 18.544006 & 21.492211 \\ 17 & 11.538501 & 19.886280 & 22.991328 \\ 18 & 12.212764 & 21.232697 & 24.494469 \\ 19 & 12.887165 & 22.582876 & 26.001256 \\ 20 & 13.561693 & 23.936489 & 27.511362 \end{tabular} \end{center} \caption{ The chromatic root of largest modulus for the complete bipartite graphs $K_{r,r}$ for $2 \le r \le 20$. } \label{table2} \end{table} Using the Dobrushin uniqueness theorem \cite{Georgii_88,Simon_93}, it can be proven \cite{Salas_97} that for a countable graph $G$ of maximum degree $r$, the $q$-state Potts-model Gibbs measure on $G$ is unique for all integer $q > 2 r$ whenever $-1 \le v_e \le 0$ for all edges $e$. Uniqueness of the Gibbs measure is one of several (inequivalent) notions of ``absence of phase transition'' \cite{Georgii_88,Simon_93}. It does not imply the analyticity of the free energy, but it does make it plausible.\footnote{ Indeed, it was by meditating on possible extensions of the theorem in \cite{Salas_97} that I was led to conjecture the results in this paper. } Likewise, a result that holds for {\em integer}\/ $q > q_0$ need not hold for all {\em real}\/ $q > q_0$, much less for a complex neighborhood of that real semi-axis; but it does suggest that such a result might be true. It is not unreasonable, therefore, to conjecture that there is a complex domain $D_r$ containing the interval $(2r,\infty)$ of the real axis, such that $Z_G(q, \{v_e\}) \neq 0$ whenever $q \in D_r$, $-1 \le v_e \le 0$ for all edges $e$, and $G$ has maximum degree $\le r$. Indeed, it is quite possible that $D_r = \{ q \colon\; |q| > 2r \}$ works; this would be a slight extension of the conjecture that $C_{opt}(r) \le 2r$. We can pose these questions more generally as follows: Let ${\cal G}$ be a class of finite graphs, and let ${\cal V}$ be a subset of the complex plane. Then we can ask about the sets \begin{eqnarray} S_1({\cal G}, {\cal V}) & = & \bigcup_{G \in {\cal G}} \; \bigcup_{v \in {\cal V}} \; \{q \in {\mathbb C} \colon\; Z_G(q,v) \,=\, 0 \} \\ S_2({\cal G}, {\cal V}) & = & \bigcup_{G \in {\cal G}} \; \bigcup_{\{v_e\} \colon\; v_e \in {\cal V} \; \forall e} \{q \in {\mathbb C} \colon\; Z_G(q, \{v_e\}) \,=\, 0 \} \end{eqnarray} Among the interesting cases are the chromatic polynomials ${\cal V} = \{-1\}$, the antiferromagnetic Potts models ${\cal V} = [-1,0]$, and the complex antiferromagnetic Potts models ${\cal V} = A \equiv \{v \in {\mathbb C} \colon\; |1 + v| \le 1 \}$. Indeed, one moral of this paper is that some questions concerning chromatic polynomials are most naturally studied in the more general context of antiferromagnetic or complex antiferromagnetic Potts models (with not-necessarily-equal edge weights). In Corollary \ref{cor5.2} we have shown that the set $S_2({\cal G}_r, A)$ is bounded, where ${\cal G}_r$ is the set of all loopless graphs of maximum degree $\le r$; and in Theorem \ref{thm6.3} we have extended this to $S_2({\cal G}'_r, A)$, where ${\cal G}'_r$ is the set of all loopless graphs of second-largest degree $\le r$. But it would be interesting to examine in more detail the location of all these sets in the complex plane, and to prove sharper bounds. Another direction in which the results of this paper could be extended is by finding a criterion {\em weaker}\/ than bounded maximum degree (or bounded second-largest degree) under which the zeros of $P_G(q)$ and $Z_G(q, \{v_e\})$ could be shown to be bounded. An interesting idea was suggested very recently by Shrock and Tsai \cite{Shrock_99a}, who studied a variety of families of graphs and arrived at a conjecture that can be rephrased as follows: For $G=(V,E)$ and $x,y \in V$, define \begin{subeqnarray} \lambda(x,y) & = & \hbox{max \# of edge-disjoint paths from $x$ to $y$} \\ & = & \hbox{min \# of edges separating $x$ from $y$} \end{subeqnarray} and \begin{equation} \Lambda(G) \;=\; \max\limits_{x \neq y} \lambda(x,y) \;. \end{equation} Clearly $\lambda(x,y) \le \min[\deg(x), \deg(y)]$ and hence $\Lambda(G) \le $ second-largest degree of $G$. Now let ${\cal G}^\Lambda_r$ be the set of all loopless graphs with $\Lambda(G) \le r$. Then the conjecture is that the set $S_2({\cal G}^\Lambda_r, {\cal V})$ is bounded, where ${\cal V} = \{-1\}$ or $[-1,0]$ or perhaps even $A$.\footnote{ Shrock and Tsai \cite{Shrock_99a} studied only the chromatic-polynomial case ${\cal V} = \{-1\}$, and proposed an even stronger result, based on the quantity $$ \Lambda_{\hboxscript{non-adj}}(G) \;=\; \max\limits_{\begin{scarray} x \neq y \\ x,y \hboxscript{ not adjacent} \end{scarray}} \lambda(x,y) \;. $$ But this cannot work for $v \neq -1$: a counterexample is obtained \cite{Sokal_hierarchical} by gluing together $n$ copies of the cycle $C_k$ (any fixed $k \ge 3$) along a single common edge and then taking $n \to \infty$. } More generally, one could define $\lambda(x,y; \{v_e\})$ to be the maximum flow from $x$ to $y$ when $|v_e|$ is taken to be the capacity of edge $e$, and likewise $\Lambda(G, \{v_e\})$; this {\em might}\/ lead to the appropriate extension of Corollary \ref{cor5.5}. This possible connection of chromatic-polynomial and Potts-model problems with max-flow problems is intriguing. Note that $\Lambda(G)$ and $\Lambda(G, \{v_e\})$ possess a ``naturalness'' property that maximum degree and its relatives lack: namely, for any graph $G$ with blocks $G_1,\ldots,G_b$, we have $\Lambda(G, \{v_e\}) = \max\limits_{1 \le i \le b} \Lambda(G_i, \{v_e\})$; contrast this with Remark 1 after Corollary \ref{cor5.5}. \vspace{2cm} \section*{Acknowledgments} I wish to thank Norman Biggs, Jean Bricmont, Jason Brown, Roberto Fern\'andez, Ira Gessel, David Jackson, Roman Koteck\'y, Antti Kupiainen, Neal Madras, Gordon Royle, Jes\'us Salas, Paul Seymour, Robert Shrock, Gordon Slade, Richard Stanley, Dave Wagner and Stu Whittington for valuable conversations and/or correspondence. I also wish to thank Tony Guttmann, Elliott Lieb and Nick Wormald for putting me in contact with some of these people. Finally, I wish to thank an anonymous referee for many helpful comments, particularly regarding the proofs of Propositions \ref{prop4.2} and \ref{prop4.5}; and to thank Criel Merino for helping me notice a silly error in my original proof of Proposition~\ref{prop_penrose}. This work was begun during a visit to Helsinki sponsored by the University of Helsinki and the Finnish National Board of Education. It was continued during a sabbatical in London where I had the pleasure of affiliation with both Imperial College and the London School of Economics. The final revisions were made during a visiting fellowship at All Souls College, Oxford. I wish to thank all these institutions --- and most especially Antti Kupiainen, Pekka Elo, Chris Isham, Helena Cronin and John Cardy --- for their warm hospitality. This research was supported in part by U.S.\ National Science Foundation grants PHY--9520978 and PHY--9900769 and by U.K.\ Engineering and Physical Sciences Research Council grant GR/M 71626. \clearpage
\section{Introduction} Recently, there has been a great deal of interest in problems involving periodic patterns in the tens of nanometers scale, for example, light conduction by photonic crystals\cite{joanopoulos}, Josephson junctions arrays formed by granular superconducting materials\cite{relevantes}, and lithographic masks for special design chips \cite{park1} among others. In special, nanolithography has brought considerable attention to thin films of microphase separated diblock copolymers (DBCP) as they naturally self organize in periodic structures on that length scale \cite{park1,PhTo,zhong}. The basic technique for the fabrication of those templates is the creation of a well-ordered DBCP film between two flat surfaces and then the transfer of the microdomains to a substrate where one of the components is removed leaving behind an ordered array of stripes or dots in either low or high relief. The desirable ordering, in this case, is one that creates a pattern on the substrate; for example, for even DBCP molecules one must have lamellae perpendicular to the hard boundaries. For that matter it is important to understand the pattern formation in confined films of DBCP since it involves problems not present in bulk systems. This issue has been addressed both theoretically \cite{turner1,kiku,brown-chakra2,pibal1} and experimentally \cite{menelle,lamb,kellogg1} and the basic conclusions are that, when the confining walls are neutral, the equilibrium pattern corresponds to lamellae perpendicular to the walls and, when the substrate prefers one kind of monomer, the pattern may consist of lamellae parallel or perpendicular to the walls, depending on the relation between the film thickness and the bulk lamellar width. The later effect appears because the finiteness of the system brings about frustration and one has to take into account the amount of compression or stretching of the molecules in order to accommodate a certain number of lamellae between the two rigid walls. One important issue that has not been emphasized in the above studies is the possibility of density mismatch between the two parts of the molecules. At the bottom wall, as the denser part of the molecule sinks, lamellae parallel to the walls will form, even if the walls are neutral \cite{canela,bamorg1}. Being a bulk effect, the interaction with the gravitational field is capable of dramatically changing the microphase separation, even for infinite systems, as the lamellae tend to be aligned with the field far from the boundaries \cite{canela}. In finite systems, the lamellae perpendicular to the field present more diffuse interfaces and the gravitational field may completely destroy the microphase separation \cite{bamorg1}. In the present work we consider this problem on two-dimensional films of even DBCP molecules as we analyze the effects of the degree of polymerization and film thickness on frustration by means of a cell dynamical system(CDS) model. We simulate films both in the weak and strong segregation regimes ((WSR)and (SSR)). For the WSR we empirically find the one-dimensional concentration profile and study the distortion of each layer within the lamellae. In Sec. II, we define the model and outline the numerical scheme. Results for neutral and interacting walls in the presence and absence of the gravitational field are discussed in Sec. III. The effects of frustration are analyzed as they affect not only the size and number of lamellae, but also their internal structure. In section V, the main conclusions are summarized. \section{Computational Model} Block copolymers are linear-chain molecules consisting of two subchains $A$ and $B$ grafted covalently to each other. Below some critical temperature $T_c$ these two blocks tend to separate, but due to the covalent bond, they can segregate at best locally to form periodic structures \cite{PhTo}. Here we consider only the case of even molecules corresponding to lamellar equilibrium patterns. CDS models have been successfully used in several problems of phase separation dynamics due to their computational efficiency and versatility \cite{oopu1,enokawa,chakragun1,baoo,mogo1,shinooo2}, so we prefer that method for the simulations. As usual, in this kind of description we assign a scalar variable $\psi(n,t)$ to each lattice site corresponding to the coarse-grained order parameter in the $n$-th cell at time $t$ (time here is defined as the number of iterations). This order parameter represents the difference $\psi_A-\psi_B$, where $\psi_A(\psi_B)$ is the local number density of $A(B)$. The ingredients for the time evolution of $\psi$ are: local dynamics dictated by a function with two symmetric hyperbolic attractive fixed points, diffusive coupling with neighbors, stabilization of the homogeneous solution and conservation of $\psi$. For the present problem, we also add the interaction with the gravitational field and with the confining walls. The conservation, when an external field is present, must be imposed by considering the Kawasaki exchange dynamics explicitly. The detailed explanation of this model is found in \cite{kitaoo} for spinodal decomposition. With this, we come to final equation for a melt of even DBCP molecules: \begin{equation} \psi(n,t+1)=(1-\epsilon)\psi(n,t)+ \left\langle\left\langle C\left(n,j;sgn [I(n,t)-I(j,t)]\right) [I(n,t)-I(j,t)]\right\rangle\right\rangle, \label{eq-model} \end{equation} where \begin{equation} I(n,t) \equiv {\cal A} \tanh\left(\psi(n,t)\right)-\psi(n,t)+ D\left[\langle\langle \psi(n,t)\rangle\rangle-\psi(n,t)\right]+hn_z+ V_s(n) \end{equation} is essentially the chemical potential. $\langle\langle\star\rangle\rangle$ is the isotropic space average of $\star$, ${\cal A}$ is a measure of the quench depth, and $D$ is the diffusion coefficient. The parameter $\epsilon>0$ appears in this model to stabilize the solution $\psi=0$ in the bulk, for $\epsilon=0$ we have a model for spinodal decomposition, in which the domains can grow without bound. Scaling arguments have proved that $\epsilon\sim N^{-2}$, where $N$ is the polymerization index \cite{ooba1}. $h$ is the gravitational field, which we assume is in the $z$ direction, and $n_z$ is the $z$ component of $n$. For molecules with matched densities we just take $h=0$. A possible interaction with the walls appears via the surface term $V_s(n)$. $C$ is the collision coefficient given by: $C(i,j;\alpha)=[\psi_c+\alpha\psi(j)][\psi_c-\alpha\psi(i)]/\psi_c^2$, where $\pm\psi_c$ are the fixed points of ${\cal A}\tanh\psi -\psi$ for ${\cal A}>1$. For all the simulations we used the values ${\cal A}=1.2$ and $D=0.5$, and uniformly distributed random initial conditions. The gravitational field, when present, is parallel to the smaller dimension. The direction normal to the field will be called the $x$ direction. We consider systems with periodic boundary conditions in the $x$ direction and hard walls in the $z$ direction, separated by a distance $L_z$. At the hard walls we impose no flux boundary conditions in the form:$[I(z+1)-I(z)]_{boundaries}= 0$. \section{Results} In order to understand the effect of confinement on the lamellae width, we must first determine its bulk value. For that matter, we ran simulations on $512\times 512$ lattices with periodic boundary conditions for different values of $\epsilon$, and $h=0$. The resulting isotropically striped pattern was then Fourier transformed, and the bulk lamellar width $W_b$ was measured in a standard way. Defining one lamella as $ABBA$ we have: \begin{equation} W_b = \frac{2\pi}{\langle k\rangle_{eq}}, \end{equation} where \begin{equation} \langle k\rangle_{eq} = \frac{\int S(k,\infty)\,k\, {\rm d}k}{\int S(k,\infty)}. \end{equation} $S(k,\infty)$ is the circular average of the structure function $S({\bf k },t)=|\psi({\bf k},t)|^2$, calculated at large times, that is, when the value of $\langle k\rangle$ approaches a constant value. Just to make sure that the bulk lamellar width as measured above was not affected by the interface bending of the disordered pattern, we also measured $W_b$ in $32\times 128$ systems with hard neutral walls and zero field, or matched densities. As expected the equilibrium configuration correspond to lamellae normal to the hard walls \cite{pibal1,kellogg1} as in Fig. \ref{fig-B}(a). $W_b$ was then measured using the one-dimensional structure function for each line and finally averaging along the $z$ direction. The values of $W_b$ found in both determinations agree, so we conclude that the excessive interface curving of the disordered pattern does not affect the lamellar width. Since the disordered patterns are easier to obtain, we will consider the lamellar width obtained from them as our bulk equilibrium value $W_b$. The results below correspond to simulations with $V_s=0$ (neutral walls), $h\neq 0$ (mismatched densities) and $h= 0$ (matched densities), and $V_s\neq 0$ (interacting walls) and $h= 0$. As will be seen, different patterns regarding the lamellae orientation appear: lamellae normal or parallel to the hard walls and a mixture of both. \vspace*{3cm} \begin{figure} \setlength{\unitlength}{1mm} \begin{picture}(90,90)(0,0) \put(50,5){\epsfxsize=8cm\epsfbox{figura1.ps}} \end{picture} \caption{Equilibrium patterns for confined films with $L_z$ = 21 and $L_x$ = 256 (only the first 150 columns are shown). (a) Neutral walls, and matched densities. The lamellae are normal to the walls with the bulk periodicity. (b) Neutral walls and density mismatch, $\epsilon$ in the SSR. 1.5 lamellae are accommodated parallel to the walls. (c) Neutral walls and density mismatch, $\epsilon$ in the WSR. 2.5 lamellae are formed parallel to the walls. The lamellar width is 8.401, smaller then the bulk value of 9.422. (d)Interacting walls and matched densities, $\epsilon$ in the WSR. The lamellae are also parallel to the walls but are more segregated than in (c).} \label{fig-B} \end{figure} \subsection{Neutral walls} We focus now on films with mismatched densities ($h\neq 0$), confined by neutral walls ($V_s=0$). In this situation we observe patterns of lamellae parallel to the hard walls, or a mixture of wetting layers on the hard walls and lamellae normal to the walls in the center part of the film. First we analyze the case of lamellae parallel to the walls only. Due to the density difference of chains $A$ and $B$, the denser part (say $A$) will be at the bottom, and the less dense part (say $B$), at the top. For a blend of two homopolymers $A$ and $B$, the film would have the lower half filled with $A$ and the upper one with $B$. The covalent bond between $A$ and $B$ parts hinders this complete separation and forces the alternation of $A$-rich and $B$-rich microregions that will then have thicker interfaces due to the interpenetration of domains \cite{bamorg1}. In the extreme case, the existence of a density mismatch may completely destroy the segregation of $A$ and $B$. The number of alternating lamellae will depend on both chain size, $\epsilon$, (Fig. \ref{fig-B}) and the separation between walls, $L_z$, (see Figure \ref{fig-L}). Also, the equilibrium patterns will always have $m+1/2$ lamellae, where $m$=0,1,2\dots. \vspace*{4cm} \begin{figure} \setlength{\unitlength}{1mm} \begin{picture}(90,100)(0,0) \put(50,5){\epsfxsize=8cm\epsfbox{figura2.ps}} \end{picture} \caption{Equilibrium patterns for films with $h=0.01$, $\epsilon=0.01$(WSR) and variable width. $L_x=256$ but only the first 150 columns are shown. As the width $L_z$ is decreased, the film goes discontinuously from $m = 3$ to $m=0$ lamellar patterns. (b), (d) and (f) show transition patterns with mixed parallel and perpendicular lamellae.} \label{fig-L} \end{figure} As we vary $L_z$ for a fixed $\epsilon$ we clearly see the effects of frustration. Figure \ref{fig-L} shows the transitions from $m$ = 0, to $m$ = 1, $m=2$ and $m=3$ patterns as $L_z$ is changed from 5 to 29 for $\epsilon$ = 0.01 and $h$ = 0.01. The transition patterns are frustrated and present lamellae normal to the walls in the central region. Since full lamellar patterns are essentially one dimensional, we define the average concentration profile, $\langle \psi(n_z)\rangle_x$, as the average over the $x$ direction of the vertical variation of $\psi$. Figure \ref{fig-profile} shows the behavior of $\langle \psi(n_z)\rangle_x$, for three different situations, for now we are interested in cases (a) and (b) only. Figure \ref{fig-profile}(a) corresponds to the profile for $\epsilon = 0.01$, $h = 0.01$, $V_s=0$ and $L_z=21$. If we try to fit a sine function to that profile, we see that the fitting will miss only the wetting layers. From this we conclude that the system is in the WSR so that the inner layers can be described by just one Fourier component \cite{leib1}. The wetting layers have an enhanced concentration due to gravitational field and the presence of the wall: the bottom and top $AB$ layers are considerably stretched by the effect of buoyancy and do not experiment the penetration of other layers, resulting in an excess of $A$ at the bottom and of $B$ at the top. \begin{figure} \setlength{\unitlength}{1mm} \begin{picture}(90,70)(0,0) \put(40,10){\epsfxsize=8cm\epsfbox{figura3.ps}} \end{picture} \caption{Average concentration profiles for a film with $L_z = 21$ and $L_x = 256$. The continuous line correspond to a fitting using Eq. (\ref{eq-profile}) (a)Neutral walls, mismatched densities. The profile is well fitted by a sine function plus exponential enhancement at the hard walls. (b) Neutral walls, mismatched densities and molecules larger than in (a). From the fitting it is clear that more than one Fourier component must be considered, indicating that the film is already in the SSR; (c) Interacting walls ($V_s(1) = -\sigma = -V_s(L_z)$) and matched densities. Although similar to the profile (a), we notice that Eq. (\ref{eq-profile}) is not adequate to describe this concentration profile.} \label{fig-profile} \end{figure} A correction for this effect led us to the tentative function \begin{equation} \psi(x) = (-1)^{m+1} \eta \sin q x +2Ce^{-\beta L}\sinh 2\beta x\;\;\;\;\mbox{ for $-L/2\leq x \leq L/2$}, \label{eq-profile} \end{equation} which fits very well the profile in Figure \ref{fig-profile}(a). For films in the WSR we found that the fitted value for $q$ is indistinguishable from $2\pi(m+1/2)/L_z$, so we define the average lamellar width $W$, directly from the fitting, as $2\pi/q$. As will be seen below, the interaction with the gravitational field causes a distortion within the lamellae regarding the width of the $A$ and $B$-rich layers, but, if considered as a unit, all the lamellae have a width very close to the average value. The transitions between consecutive values of $m$ as we vary $L_z$ is shown in Figure \ref{fig-m}. From this figure we see that discontinuous transitions occur from a pattern in which the lamellae are stretched, compared to its bulk state, to a compressed state with one more lamella, as $L_z$ is increased. The regions between steps of fixed $m$ correspond to transition patterns in which lamellae normal to the walls form in the center portion of the film. \begin{figure} \setlength{\unitlength}{1mm} \begin{picture}(90,70)(0,0) \put(50,10){\epsfxsize=8cm\epsfbox{figura4.ps}} \end{picture} \caption{Total number of lamellae, $n$, as a function of the film thickness $L_z$ for a film with mismatched densities, confined by neutral walls. The solid points on the steps represent lamellar patterns with $n$ lamellae parallel to the hard walls and between consecutive steps the film is in a mixed configuration with horizontal and vertical lamellae. The vertical dashed lines correspond to the film thickness adequate to accommodate the corresponding number of lamellae but with the bulk width $W_b$. We see that as the film width increases, discontinuous transitions between $n$ stretched lamellae and $n+1$ compressed lamellae occur.} \label{fig-m} \end{figure} We can, alternatively, fix $L_z$ and vary $\epsilon$, which corresponds to fixing the film width and varying the bulk lamellae width. For $0.006<\epsilon<0.018$ we obtain patterns with $m=2$, in the range $0.004 <\epsilon<0.006$ again we observe a transition pattern, and decreasing $\epsilon$ further we find that a $m$ = 1 pattern appears. Figure \ref{fig-trans} shows patterns with $m$=1, $m$=2 and in the transition region. The analysis of the transitions in this case is more complicated since for $\epsilon<0.004$ the system is no longer in the WSR as can be seen from the fitting of Eq. (\ref{eq-profile}) in Fig. \ref{fig-profile}(b). It is clear that other Fourier components need to be included in this case. \begin{figure} \setlength{\unitlength}{1mm} \begin{picture}(90,70)(0,0) \put(50,5){\epsfxsize=8cm\epsfbox{figura5.ps}} \end{picture} \caption{Equilibrium patterns for confined films with $L_z$ = 21 and $L_x$= 256 (only the first 150 columns are shown) three different values of $\epsilon$. As in the case of variable film width, the number of lamellae varies discontinuously and transition patterns with mixed orientation lamellae appear. For $\epsilon=0.01$ 2.5 weakly segregated lamellae are formed, the concentration profile can be well fitted by Eq. \ref{eq-profile}. $\epsilon = 0.005$ produces a transition pattern and $\epsilon = 0.0014$, a more segregated pattern with 1.5.} \label{fig-trans} \end{figure} The accommodation of the lamellae distorts their widths non uniformly as may be easily checked from the plot of the width of each individual $A$-rich and $B$-rich layer. Figure \ref{fig-we} shows the behavior, as a function of $\epsilon$, of the widths of the first $A$-rich layer that wets the bottom wall ($w_1$), the first $B$-rich layer connected to the bottom wetting layer ($w_2$) and one quart of the central lamella ($w_c = W/4$). For the sake of comparison, the variation of the bulk width, $w_b = W_b/4$, of $A$-rich (or $B$-rich) layers is also plotted. Although the variation is small compared to the bulk behavior, we see that $w_2<w_c<w_1$ consistently. This happens because there is a reduction in the number of $A$ and $B$ contacts in the first layer (for the lack of neighboring molecules from below) and an increase in the next one because the gravitational field shifts the $A$ parts downwards. As we separate the regions where the lamellae are compressed and stretched as compared to the bulk, we notice two different situations. In the compressed region, $w_1$ increases as $N$ increases ($\epsilon$ decreases), due to a compression of the internal layers caused by the greater stretching of the surface layers which produces an increase in the internal pressure (for $L_z$ fixed). We expected the inverse effect to happen when $N$ was reduced in the stretched region: the internal layers would shrink producing a tension that would stretch the surface (larger effect) and second layers (smaller effect). But, in fact, we observe a drastic reduction of the second layer acting as a tension center for surface and central layers (Fig. \ref{fig-we}). \begin{figure} \setlength{\unitlength}{1mm} \begin{picture}(90,70)(0,0) \put(50,15){\epsfxsize=10cm\epsfbox{figura6.ps}} \end{picture} \caption{Variation of $A$-rich and $B$-rich layers as a function of $\epsilon$ for a films with 2.5 lamellae. $w_1$, $w_2$ are the widths of the lowest $A$-rich and $B$-rich layers, $w_c$ is 1/4 of the central lamella and $w_b$ is 1/4 of the bulk lamellar width. The vertical line indicates the separation between regions where the film is in compressed and stretched states.} \label{fig-we} \end{figure} It is clear that the above effects are meaningful only for thin films. The transition from this to the bulk behavior may be observed as we analyze $w_1$, $w_2$ and $w_c$ as a function of the film thickness $L_z$. If the bulk behavior prevails, $w_2\approx w_1\approx w_2\approx w_b\approx L_z/m$. As we increase $L_z$ and observe films with increasing number of lamellae, we find $w_1\rightarrow w_2\rightarrow w_c\rightarrow w_b$. On the other hand, the slope $\alpha$ of each group of $w$ values is proportional to $m^{-0.8}$ instead of $m^{-1}$, which reflects the different behavior of each layer of DBCP under stretching or compressions. \begin{figure} \setlength{\unitlength}{1mm} \begin{picture}(90,60)(0,0) \put(50,10){\epsfxsize=8cm\epsfbox{figura7.ps}} \end{picture} \caption{Variation of $A$-rich and $B$-rich layers as a function of $L_z$ for a films with $\epsilon = 0.01$. We use here the same notation of Fig. \ref{fig-we}. Each group of points corresponds to lamellar patterns with $(m+1/2)$ lamellae. As $L_z$, and correspondingly $m$, increases, the film behaves more like a bulk sample in the sense that the distortion of lamellae is less significant. The slope $\alpha$ of each group is proportional to $m^{-0.8}$, for larger values of $L_z$ a crossover to the bulk behavior $\alpha\propto m^{-1}$ is expected} \label{fig-wL} \end{figure} \subsection{Interacting walls} The effect of surface fields in the formation of lamellar patterns has been extensively studied \cite{turner1,kiku,brown-chakra2,pibal1,menelle,lamb,kellogg1}. Our goal here is to compare the effect of the surface and bulk fields, so we consider only a film of DBCP molecules with matched densities confined by interacting walls in such a way that the bottom wall attracts the denser component and the top wall prefers the less dense component, As will seen below, in many ways this choice of walls produces a pattern is similar to the one obtained with neutral walls and a density mismatch, but the two situations are, in fact, different. The above interaction with walls may be simulated by choosing the surface interaction as: $V_s = \sigma$ for $z=1$ and $V_s = -\sigma$ for $z = L_z$. The equilibrium pattern obtained for $\epsilon = 0.01 $, $L_z = 21$, $h = 0$ and $\sigma = 0.01$ is very similar to the one with the same values of $\epsilon$ and $L_z$, $\sigma = 0$ and $h = 0.01$, as both present 2.5 lamellae parallel to the walls (see Fig. \ref{fig-B} (c) and (d), respectively). The first distinction appears in the segregation of domains: it is clear that the pattern in Fig. \ref{fig-B}(c) is less segregated due to effect of interpenetration of domains driven by the gravitational field. As we try to fit the concentration profile with Eq. ({\ref{eq-profile}) we notice that the patterns with density mismatch and surface interaction are also a little different in the surface region, so we conclude that Eq. (\ref{eq-profile}) is a good fit for the concentration profile of lamellar patterns with density mismatch and neutral walls only. A substantial difference appears for larger values of $h$ and $\sigma$. Figure \ref{fig-largefield} shows patterns with the same value of $\epsilon = 0.01$ and $L_z=21$ but one with surface field only and the other with density mismatch only. In this case, the lamellar structure still exists for $\sigma = 0.04$ but here, for $h = \sigma$ the lamellar structure is completely destroyed in the center of the film. \begin{figure} \setlength{\unitlength}{1mm} \begin{picture}(90,50)(0,0) \put(50,5){\epsfxsize=8cm\epsfbox{figura8.ps}} \end{picture} \caption{A comparison between bulk and surface interactions. (a) Interacting walls and matched densities. Although the surface interaction is stronger than the one considered so far, the only noticeable difference is the segregation of the wetting lamellae. (b) Neutral walls and mismatched densities. Increasing the value of the bulk field $h$ by the same amount as the surface field the observed pattern changes dramatically: instead of a lamellae pattern we observe a frustrated mixed orientation pattern.} \label{fig-largefield} \end{figure} \section{Conclusions} In this paper, we study the effects of surface and bulk (gravitational) fields coupled with hard wall restrictions on the lamellar pattern formation of diblock copolymer systems. We find that the two are the predominant factors to determine the final equilibrium pattern: lamellae tend to form normal to the field and their number is determined by the ability of the system to resolve the frustration caused by the confinement. Unresolved patterns present a mixture of wetting lamellae normal to the field and lamellae parallel to the field in the central part of the film. The gravitational field also distorts the periodicity of the lamellar pattern. The bottom $A$ layer is larger than it would be if placed in the central part of the film. On the other hand, the next $B$ layer is narrower, in such a way that the first lamella, defined as the sequence $ABBA$, has a width very close to the central lamellae. We obtained a good fit for the average concentration profile in the WSR by using a trial function which consists of the superposition of a sinusoidal function, characteristic of the WSR, and exponential functions for the enhanced concentration of the wetting layers. \acknowledgements This work was partially supported by CNPq (Brazil) and Faperj (Rio de Janeiro).
\section{Introduction} Reheating in the post-inflationary Universe is a very important process, since it describes the formation of all matter and energy in the Universe today \cite{ref1}. However, it has only relatively recently \cite{ref2} been realised that in certain cases the process of reheating involves a stage of explosive production of particles by parametric resonance which cannot be described by the usual methods of the elementary reheating theory \cite{ref8}. This stage of `preheating' is a non-perturbative and out of equilibrium process and as such is a very involved and little understood proceeding. In fact there have been complex analytical methods of preheating developed for only the very simplest of single-field inflationary potentials \cite{ref7}. It is generally numerically easy to obtain results for the first stages of particle production, yet a complete, self-consistent analysis should include all effects of backreaction of produced particles and their rescattering. At present lattice simulations (such as in \cite{ref31,ref3}) must be performed to fully study these effects. Nevertheless, much progress is being made in developing methods to cope with the rich topic of preheating \cite{ref7}. For a while now, the advantages of using more than one scalar field to implement inflation have been well appreciated. Hybrid inflation models \cite{H} are probably the most popular of all present inflationary models. They succeed in removing the need for fine-tuning of the inflaton potential by providing a natural means to end inflation. Most importantly however, they allow inflation to occur at a relatively low energy density using small scale fields avoiding strong CMB constraints, a task which is virtually impossible to achieve using only one field. Thus, hybrid-type models are attractive from the point of view of particle physics and in particular are naturally well suited to realisations in supersymmetry (SUSY) where there is an abundance of scalar fields and flat directions. Preheating in hybrid inflation has been studied \cite{ref4} using the standard toy hybrid potential of two fields $V(\phi_1,\phi_2)$, coupled to an extra, massless scalar field, $\chi$ by the interaction $\frac{1}{2}(h_1^2\phi_1^2+h_2^2 \phi_2^2)\chi^2$. It is found \cite{ref4} that over most of the parameter space, preheating is rather inefficient in hybrid inflation with little production of $\phi_1,\phi_2$ particles although production of $\chi$ particles is possible for certain values of $h_1, h_2$. However, the results concerning the production of $\phi_1$ and $\phi_2$ particles may change in a supersymmetric version of the model, where the potential is derived from a supersymmetric superpotential of the kind $W \sim \kappa \phi_1 \phi_2^2+\cdots$, such that the same coupling will give the quartic self-interaction for the $\phi_2$ field and the interaction term between $\phi_2$ and the inflaton $\phi_1$. Therefore, supersymmetry will ensure that the masses of both fields at the global minimum of the potential are not only the same order of magnitude but $equal$. This in turn will give rise to a non-chaotic evolution of the classical homogeneous fields at the beginning of the preheating/reheating era. The classical fields will begin to oscillate with relatively large, little damped amplitudes, which will favour production of particles occurring. In addition, in hybrid models inflation ends with a phase transition when the inflaton field reaches a critical value. At this point, the second field $\phi_2$ is sited at a maximum of the potential with vanishing mass. Once the inflaton passes this point, the mass of $\phi_2$ becomes negative allowing both fields to roll down the potential towards the global minimum. It is clear that across the region of the spinodal instability (negative curvature) \cite{refDan1}, production of $\phi_2$ particles can be quite intensive. Therefore, previous to any oscillation of the homogeneous fields, preheating will begin with an initial burst of $\phi_2$ production. If the amplitude of the oscillations is large enough, this will also happen in the successive oscillations every time the classical fields enter the region of the spinodal instability. Although this only directly affects the rate of production of $\phi_2$ particles, the effect is propagated into the production of $\phi_1$ particles through the mixing term in the mass squared matrix. The layout of the paper is as follows: section 2 briefly describes the theory of preheating as applied to many scalar fields and we point out that the coupling between fields should in general be taken into account; in section 3 we study a general model of supersymmetric hybrid inflation, focusing on the issue of production of $\phi_1$ and $\phi_2$ particles; in order to motivate production of other particles, which is model dependent, in section 4 we review a particular model of inflation where the strong-CP problem is solved via a Peccei-Quinn symmetry which leads to the production of axions. It will be important to calculate the effects of preheating in this model since the number of axions produced during reheating must be tightly constrained in order to be cosmologically acceptable. The results of the numerical study of the first stages of preheating in this model are presented, and compared to those results without including mixing terms. The first effects of backreaction of produced particles is studied and the total number of each type of particle calculated. \section{Theory of preheating with multiple fields} We begin with a model of the early Universe described by a potential, $V(\phi_\alpha)$, in an expanding Universe with scale factor $a(t)$. Amongst the fields $\phi_\alpha$ in the potential, there are degrees of freedom corresponding to scalar fields in which all energy is concentrated during inflation. In the usually considered single-field inflation this is the slowly rolling inflaton, $\phi$, with effective potential $V(\phi)$. In the hybrid scenario there are two such fields important to inflation, one in which the potential is relatively flat during inflation which acts as the inflaton field, $\phi_1$, and one which sits in a false vacuum, $\phi_2$, and causes the end of inflation in the phase transition to its true vacuum. All other fields are not important to inflation serving only perhaps to provide the constant vacuum energy during inflation. In order to investigate the reheating of the Universe and determine the extent of the effects of preheating, we must first study the evolution of these fields as inflation ends and beyond into the regime of coherent oscillations. We may write the full quantum fields as: \begin{equation} \mbox{\boldmath $\phi$}_{\alpha}(\textbf{x},t)=\phi_{\alpha}(t) +\delta\phi_{\alpha}(\textbf{x},t), \label{phidelta} \end{equation} where $\delta\phi_{\alpha}(\textbf{x},t)$ represents the quantum fluctuations about the homogeneous zero modes, $\phi_{\alpha}(t)$, and necessarily satisfies $\langle\delta\phi_{\alpha}(\textbf{x},t)\rangle=0$. Neglecting ensuing effects of particle production, we proceed by considering the motion classically, solving the equations of motion for the classical homogeneous fields $\phi_{\alpha}(t)$, but using quantum fluctuations $\sim H_0/2\pi$ (where $H_0$ is the Hubble constant during inflation) in the initial conditions at the end of inflation to initiate the spontaneous symmetry breaking. The equations of motion governing the behaviour of the background fields, $\phi_{\alpha}(t)$, are: \begin{equation} \ddot{\phi}_{\alpha}+3H\dot{\phi_{\alpha}}+\partial V/\partial\phi_{\alpha}=0, \label{cleom} \end{equation} where the Hubble constant is given by the Friedmann equation, \begin{equation} H^2(t) = \frac{8\pi}{3M_P^2}\left[ \frac{1}{2}\sum_{\alpha}\dot{\phi_{\alpha}}^2+V\right], \label{H} \end{equation} and $H=\dot{a}/a$. Here, the subscript $\alpha$ runs over all the classical scalar fields in the theory relevant to inflation and $V=V(\phi_{\alpha})$ is the effective potential of these fields. During inflation all energy is contained in this potential and the Universe is in a vacuum-like state with vanishing temperature, entropy and particle number densities. The inflaton field, $\phi$, is slowly rolling with the friction term dominating in Eq. (\ref{cleom}) so that: \begin{equation} \dot{\phi}\simeq\frac{-1}{3H_0}\frac{\partial V}{\partial\phi}. \label{slowroll} \end{equation} This also describes the relevant situation for each classical field at the beginning of the reheating era. The equations of motion for the quantum fluctuations of the scalar fields in linear perturbation theory are given by: \begin{equation} \ddot{\delta\phi_k}_\alpha + 3H\dot{\delta\phi_k}_\alpha+\frac{k^2}{a^2}{\delta\phi_k}_\alpha + \mathcal{M}^2_{\alpha \beta}{\delta\phi_k}_\beta = 0, \label{queom} \end{equation} where ${\delta\phi_k}_\alpha(t)$ are the time dependent Fourier modes (in comoving momentum space $\textbf{k}$) of the quantum fluctuations $\delta\phi_\alpha$, \begin{equation} \delta\phi_\alpha(\textbf{x},t)=\int\frac{d^3k}{\sqrt{2}(2\pi)^{3/2}} \left[\hat{a}_k{\delta\phi_k}_\alpha(t)e^{-i\textbf{k.x}} +\hat{a}^\dagger_k{\delta\phi_k}_\alpha^*(t)e^{i\textbf{k.x}}\right], \end{equation} and $\hat{a}^\dagger_k$ and $\hat{a}_k$ are creation and annihilation operators. Thus, Eq. (\ref{queom}) describes the interactions between the classical fields, $\phi_{\alpha}$, and the quantum fluctuations, $\delta\phi_\alpha$. When neglecting backreaction effects due to the particles produced, the Hubble constant, $H(t)$, and the mass squared matrix, $\mathcal{M}^2_{\alpha\beta}(t)= \partial^2V/\partial\phi_\alpha\partial\phi_\beta$, are functions of the classical fields only, as given by the classical equations of motion (\ref{cleom}) and (\ref{H}). It should be clear that the equations of motion for the quantum fluctuations will in general be coupled through mixing terms in the mass matrix $\mathcal{M}^2_{\alpha\beta}\not= 0$ (for $\alpha\not=\beta$); in other words the rate of production of each separate particle is dependent on the rate of production of the others, {\it{even}} at the initial stage of preheating when backreaction effects are not yet important. Changing to comoving fields, ${\psi_k}_\alpha=a^{3/2}{{\delta\phi_k}_\alpha}$, these equations can be simplified further to: \begin{equation} \ddot{\psi_k}_\alpha + \left(\frac{k^2}{a^2}-\epsilon\right){\psi_k}_\alpha + \mathcal{M}^2_{\alpha\beta}{\psi_k}_\beta = 0, \label{psieq} \end{equation} where one usually neglects the pressure term, $\epsilon = \frac{3}{4}(H^2+2\ddot{a}/a){\psi_k}_\alpha$ which is a small correction that approaches zero as the Universe approaches matter domination $(a\sim t^{2/3})$. Typically the transition between the vacuum dominated state of inflation and that of matter domination during reheating is fairly rapid, so this term will be negligible. The solutions of (\ref{psieq}), ${\psi_k}_\alpha(t)$, are complex functions whose initial values at the end of inflation at time $t_0$ are taken to be \cite{refDan2}: \begin{equation} {\psi_k}_\alpha(t_0)=1/\sqrt{2{\omega_k}_\alpha(t_0)},\;\;\; \dot{{\psi}_k}_\alpha(t_0)=-i\sqrt{{\omega_k}_\alpha(t_0)/2}, \end{equation} where, \begin{equation} {\omega_k}_\alpha^2 = k^2/a^2 + \mathcal{M}^2_{\alpha\alpha}. \end{equation} The comoving occupation number of particles of wave number $\textbf{k}$ is an adiabatic invariant defined by: \begin{equation} {n_k}_\alpha=\frac{{\omega_k}_\alpha}{2}\left(\frac{|\dot{{\psi}_k}_\alpha|^2}{{\omega_k}_\alpha^2}+|{{\psi}_k}_\alpha|^2\right)-\frac{1}{2}. \label{occno} \end{equation} An instability in the growth of a given mode, ${{\psi}_k}_\alpha$, corresponds to an exponential increase in the occupation number of particles, ${n_k}_\alpha\propto \exp{(2{\mu_k}_\alpha\bar{m}t)}$. $\bar{m}$ denotes the characteristic frequency of oscillations of the classical fields, given by their mass at the global minimum. The efficiency of preheating of a given particle with a given mode $\textbf{k}$ is measured by the growth parameter ${\mu_k}_\alpha$. Of course to be fully self-consistent, the potential and its derivatives in the equations for the classical fields and Hubble constant, (\ref{cleom}), (\ref{H}), should contain the full quantum fields in the theory, $V(\mbox{\boldmath $\phi$}_\alpha)$, taking into account the extra particles being produced. In the Hartree approximation this is done via $\langle\delta\phi_\alpha^2\rangle\ne 0$, where the expectation value contains all modes: \begin{equation} \langle\delta\phi_\alpha^2\rangle = \int\frac{k^2dk}{2\pi^2}|{\delta\phi_k}_\alpha|^2. \end{equation} The backreaction on the classical fields will not be important in the first stage of the evolution when $\langle\delta\phi_\alpha^2\rangle\simeq 0$, but the modes will grow and eventually dominate over the other terms. The classical evolution will be altered and in turn this will affect the quantum equations (\ref{queom}) which also must include further terms corresponding to produced particles. As a first step we may neglect production of particles altogether and study the evolution classically. Once the behaviour of the classical fields in a specific model is given, we may try to rewrite the evolution equation for each quantum fluctuation, Eq. (\ref{psieq}), in the form of a Mathieu equation \cite{mathieu}, \begin{equation} {\psi_k}^{''}_\alpha + (A_{\alpha}(k)-2q_{\alpha}\cos{2z}){\psi_k}_\alpha = 0,\label{mat} \end{equation} where dash denotes differentiation with respect to $z$, and $z\propto \bar{m} t$. The $(A,q)$ plane is divided into instability/stability regions, such that in the unstable regions the solution grows exponentially, $\psi_k \propto e^{\mu_k z}$, and this can be interpreted as production of particles for the associated mode. Therefore, once the parameters $A_{\alpha}(k)$ and $q_\alpha$ are known for a particular model we can get some insight about which particles are more likely to be produced, without the need of numerical integration. In the next section we will apply the above receipt to a general supersymmetric hybrid model. First we analyse the motion of the classical fields when they begin to oscillate, focusing afterwards on the implications this will have for the production of their own quantum fluctuations during preheating. \section{Supersymmetric hybrid inflation.} In the hybrid scenario \cite{H} with just two real fields, the potential is given by, \begin{equation} V(\phi, N)= \frac{\lambda}{4} \left( N^2 - 2 \Lambda^2\right)^2 + \frac{g^2}{2} \phi^2 N^2 + \frac{1}{2} m_\phi^2 \phi^2 \,, \label{Vhyb} \end{equation} where $\phi$ will play the role of the inflaton. During inflation, the field $N$ is trapped in its false vacuum, $N=0$, whilst the $\phi$ field slowly rolls along the potential from above the critical value $\phi_c= (\sqrt{2 \lambda}/g ) \Lambda$ towards the origin. Once $\phi$ reaches the critical value, the $N$ mass changes sign and becomes negative, allowing $N$ to roll away from its zero value towards the global minimum, which in the positive quadrant is given by, \begin{equation} \phi_0=0 \,,\;\;\;\; N_0= \sqrt{2} \Lambda \,. \end{equation} The fields then oscillate about their true vacuum values ending inflation and beginning the preheating/reheating phase. Integrating the classical equations of motion, we obtain the trajectory in the $\phi-N$ plane. In many cases after the initial symmetry breaking, one field may oscillate with an amplitude much larger than that of the other field which is essentially fixed at its VEV. In this case almost all the energy is contained in just one field and reheating is then similar to the single field case. This will greatly simplify the investigation of preheating, although in general it will not be possible. Alternatively both fields may oscillate, and the motion may be chaotic. Having one or another situation will depend on the ratio of the coupling constants, $\lambda/g^2$, and they have been analysed in \cite{ref4}, the chaotic motion being typical of a potential with $\lambda \approx g^2$. However, a different scenario will arise in the supersymmetric hybrid models of inflation for which the exact relation $\lambda= g^2/2$ holds. In supersymmetric hybrid inflation, the potential in Eq. (\ref{Vhyb}) can be derived from the superpotential, \begin{equation} W= \kappa \Phi (\rm{N}^2 - \Lambda^2) \label{supot}\,, \end{equation} where $\Phi$ and $\rm{N}$ are complex superfields, $\phi$ and $N$ in the potential being the real components of the associated canonically normalised complex fields, and $\kappa^2=\lambda= g^2/2$. In addition, the inflaton mass $m_\phi$ will be given either by a soft SUSY-breaking mass or else generated by radiative corrections to the inflaton potential. In either case, it has to be smaller than the combination $\kappa \Lambda$ to ensure a potential flat enough during inflation to allow for the slow rolling scenario \cite{liddle}, that is, \begin{eqnarray} \epsilon &=& \frac{M_P^2}{16 \pi} \left(\frac{V'}{V}\right)^2= \frac{1}{16 \pi}\left( \frac{M_P \phi}{\Lambda^2} \right)^2 \left(\frac{m_\phi}{ \kappa \Lambda} \right)^4 \ll 1\,, \label{epsilon} \\ |\eta| &=& \frac{M_P^2}{8 \pi} \left|\frac{V''}{V}\right| = \frac{1}{8 \pi} \left(\frac{M_P}{\Lambda}\right)^2 \left(\frac{m_\phi}{ \kappa \Lambda} \right)^2 \ll 1 \,, \label{eta} \end{eqnarray} where prime denotes derivative with respect to $\phi$. Moreover, the model should be able to reproduce the correct level of density perturbation, responsible for the large scale structure in the Universe, accordingly to the COBE anisotropy measurements. The spectrum of the density perturbations is given by the quantity \cite{liddle}, \begin{equation} \delta^2_H= \frac{32}{75} \frac{V}{M^4_P} \frac{1}{\epsilon_*} \,, \label{cobe} \end{equation} where $\epsilon_*$ is defined roughly 60 e-folds before the end of inflation, and the COBE value is $\delta_H=1.95 \times 10^{-5}$ \cite{COBE}. Assuming $|\eta| < 1/60 $, then $\phi_*= \phi_c e^{ 60 \eta} \approx \phi_c$, and Eq. (\ref{cobe}) yields the constraint, \begin{equation} \kappa \Lambda \simeq 2.36 \times 10^{10} \left(\kappa m_{\phi}\right)^{2/5}\, GeV\,, \label{cobecon} \end{equation} or similarly, \begin{equation} \kappa \Lambda \simeq 1.27 \times 10^{15} |\eta| \,. \label{cobecon1} \end{equation} The value $|\eta| \simeq 0.01$ is also a reasonable assumption in order not to generate too large a tilt in the spectral index, $n=1 + 2 \eta - 6 \epsilon$, whose present observational constraint \cite{sindex} is $|n-1|< 0.2$. The superpotential in Eq. (\ref{supot}) provides not only the couplings of the inflaton and hybrid field $N$ neccessary to achieve the hybrid mechanism, but also a non-vanishing constant vacuum energy $V(0)$ through the scale $\Lambda$, which is also related to the critical value $\phi_c$. Hybrid inflation models derived from this kind of superpotential are called F-term hybrid inflation. A different version of SUSY hybrid inflation is provided by D-term inflation \cite{DI}, where the critical value and the constant potential energy are related to the Fayet-Illiopoulus term coming from an anomalous $U(1)$ symmetry. Alternatively the constant potential $V(0)$ may owe its origin to the hidden sector of a supergravity (SUGRA) theory \cite{SUGRA}. Nevertheless, the ultimate origin of $V(0)$ and the scale $\phi_c$ is not relevant for the discussion of preheating. We only demand a common coupling giving the quartic interaction term for $N$ and the interaction between $\phi$ and $N$, and the condition that the potential vanishes at the global minimum. Therefore, the SUSY hybrid potential can be written as: \begin{equation} V_0(\phi,N) = \frac{1}{4}{\kappa}^2 (N^2 -2 \phi^2_c)^2 + {\kappa}^2 N^2 \phi^2 + \frac{1}{2} m_\phi^2 \phi^2. \end{equation} where the origin of $\phi_c$ is left unspecified. The slow-rolling and COBE constraints, Eq.s (\ref{epsilon}-\ref{cobecon1}), remain the same with the replacement $\phi_c=\Lambda$. \subsection{Classical field evolution \label{clshyb}} Whatever the values of the parameters $\kappa$, $\phi_c$ and $m_\phi$ which satisfy the above constraints, the behaviour of the classical fields when they begin to oscillate will follow a common pattern, dictated by the form of the potential, and we can make some general observations which are independent of the particular values of the parameters chosen: \noindent \emph{i)} The maximum attainable amplitude of the oscillations of the $\phi$ field, $\phi_{c}=N_0/\sqrt{2}$, is comparable to (more precisely is a factor of $\sqrt{2}$ smaller than) that of the $N$ field, $N_0$. \noindent \emph{ii)} There is only $one$ natural frequency of oscillation, given by the masses of the fields at the global minimum, \begin{equation} \bar{m}_\phi = \bar{m}_N= \sqrt{2} \kappa N_0 \,. \end{equation} This is quite different to the situation in most of the parameter regime of the standard (non-supersymmetric) hybrid potential, and it is due to the exact relation $\lambda= g^2/2$ imposed by SUSY. \noindent \emph{iii)} In the limit $m_\phi \ll \kappa \phi_c$, and due to the fact that both fields have the same mass at the global minimum, there is a particular solution of the classical equations such that, \begin{equation} \frac{\partial V_0}{\partial \phi} \propto \frac{\partial V_0}{\partial N} \,, \end{equation} that is, the fields' trajectory towards the minimum of the potential describes a straight line: \begin{eqnarray} N=\pm\sqrt{2}(\phi_{c}-\phi), \label{traj} \end{eqnarray} in the $\phi-N$ plane. That the fields follow such a trajectory or not will depend on the initial conditions, otherwise fixed by the inflationary dynamics and given by: \begin{eqnarray} & &\phi_{in} = \phi_{c}\pm \frac{H_0}{2\pi},\hspace{0.3in} N_{in} = \frac{H_0}{2\pi}, \nonumber \\ & &\dot{\phi}_{in} =\frac{-1}{3H_0}\frac{\partial V_0}{\partial\phi}(\phi_{in},N_{in}),\hspace{0.2in} \dot{N}_{in} = \frac{-1}{3H_0}\frac{\partial V_0}{\partial N}(\phi_{in},N_{in}). \nonumber \end{eqnarray} Since the initial velocities, $\dot{\phi}_{initial}$ and $\dot{N}_{initial}$, are very small slow roll values, the fields will tend to move following the gradient field of the potential towards the minimum, and will end in the straight line trajectory almost exactly. In effect, we will have one oscillating mode and one stationary mode. The smaller the mass parameter $m_\phi$ with respect to $\bar{m}_\phi$, the better the straight line approximation. \noindent \emph{iv)} If we take $\phi_c \ll M_P$, we then have the relation, $\bar{m}_{\phi}/H_0 = \bar{m}_N/H_0 \simeq M_P/N_0 \gg 1$, with $H_0$ being the value of the Hubble constant at the end of inflation. This allows for many oscillations of each field in a Hubble time, thus the motion is not very suppressed by the expansion of the Universe and we should expect large amplitudes of oscillations. Consequently the large oscillations for a reasonable length of time will allow for the possibility of parametric resonance and subsequent particle production which must be investigated. In Ref. \cite{ref4} they have studied standard hybrid inflation with $\lambda=g^2$ and found chaotic behaviour of the classical fields initially, that eventually becomes more regular. However, the SUGRA hybrid inflation model would correspond to $\lambda= g^2/2$, where the chaotic motion is not present and the trajectory is given by a straight line. This can be seen in Fig. (\ref{sugraplot}), where we have plotted the oscillations of the classical fields and the trajectory in the $N-\phi$ plane, for the choice of parameters: $\phi_c= 10^{16}\,GeV$, $\kappa=10^{-3}$ and $m_\phi=3.7\times 10^{9}\, GeV$, which correspond to the choice $\eta \simeq 0.01$. We have also included for comparison the standard hybrid model with $\sqrt{\lambda} = g = 10^{-3}$, and the same values of $\phi_c$ and $m_\phi$. In both situations, the results will remain qualitatively the same when changing the value of $\phi_c$, and accordingly those of $\kappa$ and $m_\phi$; the main effect of reducing (increasing) the scale $\phi_c$ being a weaker (stronger) effect of the expansion of the Universe during the first oscillations. \begin{figure} \begin{tabular}{cc} \epsfxsize=8cm \epsfysize=8cm \hfil \epsfbox{shyb3.eps} \hfil & \epsfxsize=8cm \epsfysize=8cm \hfil\epsfbox{shyb3t.eps} \hfil \\ \epsfxsize=8cm \epsfysize=8cm \hfil \epsfbox{ghyb3.eps} \hfil & \epsfxsize=8cm \epsfysize=8cm \hfil \epsfbox{ghyb3t.eps} \hfil \end{tabular} \caption{Oscillations of the classical fields (left) and the trajectory in the $N-\phi$ plane (right). The top panels correspond to the supersymmetric hybrid model with $\phi_c=10^{16}\, GeV$, $\kappa=10^{-3}$ and $m_\phi=3.7\times 10^{9}\, GeV$. The time scale is given in terms of the approximate number of oscillations of the fields, $N'=\bar{m}_\phi t /2 \pi$, where $N'=0$ has been defined for convenience as the point where the fields reach the first peak. The lower panels show for comparison the situation in a non-supersymmetric hybrid model with $g= \sqrt{\lambda} = 10^{-3}$, and $\phi_c (=\Lambda)$ and $m_\phi$ as before. \label{sugraplot}} \end{figure} Using the relation Eq. (\ref{traj}), we can effectively replace the two oscillating fields problem by a single field, for example $N(t)$, oscillating along the potential, \begin{equation} V(N)= \frac{1}{4} \kappa^2 N_0^4 + \frac{3}{4} \kappa^2 N^4 - \kappa^2 N^3 N_0. \label{Vsingle} \end{equation} Near the origin the potential is almost flat, with the curvature becoming first negative and then changing sign and increasing as we approach and pass the minimum. Therefore, the field will tend to spend most of its time near the initial values, mainly varying when it approaches the minimum and the evolution speeds up. In a non-expanding Universe, and taking as initial conditions the asymptotic limit $N(-\infty)=\dot{N}(-\infty)=0$, the exact solution of the classical evolution equation is given by: \begin{eqnarray} \frac{N(t)}{N_0} &=& \frac{4/3}{1+2 \bar{m}_\phi^2 t^2/3} \nonumber \\ & = & 1 + \frac{1}{3}\cos (\int^t_0 \sqrt{\frac{3}{2}} \left( \frac{N(t')}{N_0} \right) \bar{m}_\phi dt') - \frac{2}{3} \sin^2 (\int^t_0 \sqrt{\frac{3}{8}} \left( \frac{N(t')}{N_0} \right) \bar{m}_\phi dt') \,, \label{soln} \end{eqnarray} where $t=0$ is defined when $N$ reaches the peak of the oscillation $N(0)/N_0=4/3$. In fact the motion can be decomposed into two different parts, which we label region I/II respectively: \noindent Region I: when $|N/N_0-1|\le 1/3$ the field is near the minimum $N_0$, the motion is more harmonic and the sine squared term in the right hand side of Eq. (\ref{soln}) can be neglected, i.e., \begin{equation} \frac{N(t)}{N_0} \simeq 1 + \frac{1}{3}\cos (\int^t_0 \sqrt{\frac{3}{2}} \left( \frac{N(t')}{N_0} \right) \bar{m}_\phi dt')\,. \label{solnmin} \end{equation} \noindent Region II: when $0\le N/N_0 <2/3$ the field increases/decreases as $N(t) \propto 1/t^2$. In an expanding Universe, the main effect of the friction term in the evolution equation will be to reduce the amplitude of the successive oscillations (and thus the period), such that the amplitude will now decrease as the inverse of time. Then, we can approximate the behaviour of the $N$ field around each peak of the oscillations (region I) by, \begin{equation} \frac{N(t)}{N_0} \simeq 1 + \frac{\Phi(t)}{3}\cos (\int^t_0 \sqrt{\frac{3}{2}} \left( \frac{N(t')}{N_0} \right) \bar{m}_\phi dt')\,, \label{solnminh} \end{equation} where now $\Phi(t) \propto 1/t$. The actual decrease in amplitude over an oscillation period will depend on the ratio $H/\bar{m}_\phi$. However, during the first stages of the oscillations the dominant effect will be to render the total oscillation more harmonic (symmetric) around the minimum, more than to reduce the factor $\Phi(t)$. \subsection{ Particle production: $\delta \phi$, $\delta N$. \label{pp}} Due to the behaviour of the classical fields, we may expect particle production through parametric resonance when the field is $oscillating$ around the peaks, that is, in region I when $N(t)/N_0 \ge 2/3$. Therefore, in that interval of time we can use the approximation given in Eq. (\ref{solnminh}) in order to convert the evolution equations for the quantum fluctuations, Eq. (\ref{queom}), into a Mathieu-like equation. Here we will be mainly concerned about the production of $\phi$ and $N$ particles, which indeed will occur in any SUSY model of hybrid inflation. Production of other species of particle will be model dependent, depending on how they couple to the classical fields. We first analyse the equations neglecting the mixing terms, $\mathcal{M}^2_{\alpha \beta}$ in Eq. (\ref{queom}). As before, we will use the fact that $\phi$ and $N$ follow a straight line trajectory, and work only with the classical field $N(t)$. The effective masses for the quantum fluctuations $\delta \phi$ and $\delta N$ are then given by: \begin{eqnarray} \mathcal{M}^2_{\phi\phi}&=& 2 \kappa^2 N^2 \,,\\ \mathcal{M}^2_{NN} &=& 2 \kappa^2 N ( 2 N - N_0) \,, \label{mnn} \end{eqnarray} which, when the classical fields are near the minimum, become: \begin{eqnarray} \mathcal{M}^2_{\phi\phi}& \simeq & \bar{m}_\phi^2 \left( 1 + \frac{2 \Phi(t)}{3} \cos \int^t_0 \sqrt{\frac{3}{2}} \left( \frac{N(t')}{N_0} \right) \bar{m}_\phi dt' \right) \,,\\ \mathcal{M}^2_{NN} & \simeq & \bar{m}_\phi^2 \left( 1 + \Phi(t)\cos \int^t_0 \sqrt{\frac{3}{2}} \left( \frac{N(t')}{N_0} \right) \bar{m}_\phi dt' \right) \,. \end{eqnarray} We have kept only the contribution linear in the cosine and neglected the sub-dominant cosine squared contributions. Substituting these expressions into Eq. (\ref{psieq}), and defining: \begin{equation} 2 z = \int^t_0 \sqrt{\frac{3}{2}} \left( \frac{N(t')}{N_0} \right) \bar{m}_\phi dt' \,, \end{equation} it is straightforward to get the parameters $A_\alpha$ and $q_\alpha$ entering in the Mathieu equation, that is, \begin{eqnarray} q_{\phi}(t)&=&\frac{8}{9} \left(\frac{N_0}{N(t)}\right)^2 \Phi(t),\;\;\;\; A_{\phi}(t)= \frac{8}{3} \left(\frac{N_0}{N(t)}\right)^2 \frac{k^2}{(a^2 \bar{m}_\phi^2)}+ 3 \frac{q_\phi(t)}{\Phi(t)},\\ q_{N}(t)&=& \frac{4}{3} \left(\frac{N_0}{N(t)}\right)^2 \Phi(t),\;\;\;\; A_{N}(t)= \frac{8}{3} \left(\frac{N_0}{N(t)}\right)^2 \frac{k^2}{(a^2 \bar{m}_\phi^2)}+ 2 \frac{q_N(t)}{\Phi(t)}. \end{eqnarray} We notice that these parameters are functions of time, as is the case in models studied in an expanding Universe. The time dependence purely due to the expansion appears in the scale factor, $a(t)$, such that the momentum $\textbf{k}$ is red-shifted as the system evolves, and in the damping of the amplitude of the classical oscillations, $\Phi(t)$, which will cause $q(t)$ to decrease appreciably in a Hubble time. On top of this, we have an extra time dependence related to the nature of the oscillations, which enters in the factor $N_0/N(t)$, and which would be present even in Minkowski space. \begin{figure} \hfil \scalebox{0.5}{\rotatebox{270}{\includegraphics{{tryagain.ps}}}} \hfil \caption{The trajectories through the Mathieu stability/instability chart for $\textbf{k}=0$ showing how the efficiency of parametric resonance varies over each stage in the classical field oscillation. The contours are incremented in levels of $\mu\sim 0.1$, ranging from the darkly shaded (stable) regions to the light (extremely unstable) regions. Modes $\textbf{k}> 0$ will lie vertically above the zero mode trajectories shown. Particle production occurs mostly in the second instability band. \label{matplot}} \end{figure} Although the Mathieu parameters $A_\alpha$ and $q_\alpha$ depend on time we can infer the behaviour of the solutions using the stability/instability bands analysis of the canonical Mathieu equation, (\ref{mat}). For a given mode $\textbf{k}$, instead of looking to a given point in the chart, we will move along some trajectory in the plane $(A_\alpha(t), q_\alpha(t))$ \cite{ref7,steinhardt}, crossing different stability/instability regions as the fields oscillate. As long as we know the trajectory followed in the Mathieu chart, the standard analysis can be applied. For example, the parameters $q(t)$ are bounded by, \begin{eqnarray} 2 \Phi(t) \ge & q_\phi(t)& \ge \Phi(t)/2 \,, \\ \nonumber 3 \Phi(t) \ge & q_N(t)& \ge \frac{3}{4} \Phi(t) \,, \label{rangeq} \end{eqnarray} so that in each oscillation we will move in the Mathieu plot downward from right to left during the first half period, and back again. This can be seen in Fig. (\ref{matplot}). We first notice that for small values of the momentum $\textbf{k}$ and $\Phi(t) \approx 1$, we have that $A_\phi \simeq 3 q_\phi$ but $A_N \simeq 2 q_N$. The line $A=2 q$ roughly separates the broad (to the right of this line in the Mathieu plot) from the narrow (to the left) resonance regimes, and accordingly we expect much less production of $\phi$ particles than $N$ particles. To see the difference between broad/narrow regimes, we can follow Ref. \cite{ref7} in order to get the maximum $allowed$ growth index in a trajectory of the type $A(k) = d(k) + b q$, with $k$ being the momentum. This is given by, \begin{equation} \mu^{max} = \frac{1}{\pi} \ln \left( \sqrt{ 1 + e^{-\pi c}} + e^{-\pi c} \right) \label{mumax}\,, \end{equation} with:, \begin{equation} c= \frac{A(k)-2 q}{2 \sqrt{q}} \,. \end{equation} The value $\mu^{max}$ should be understood as an upper bound on the effective growth index along the given trajectory. Along the line $A= 2 q$ (i.e. $k=0$) we have a constant maximum possible growth index, $\mu^{max} \approx 0.28$ \cite{ref2,ref7}. Because we are crossing several stability/instability bands in only half a period, it has been argued than the effective $\mu$ will somehow be less than half this value, with the exact value depending on the actual time dependence of the parameters, and how much time is spent in each band. To the left of this trajectory, $A > 2 q$, we have that $c$ grows with $q$ and then $\mu^{max}$ decreases as we move from left to right in the Mathieu plot. In the opposite case, $A < 2 q$, we have the reverse behaviour with $\mu^{max}$ increasing with large $q$, and we are in the broad resonance regime. \begin{figure} \epsfxsize=8cm \epsfysize=8cm \hfil \epsfbox{muaq.eps} \hfil \caption{The growth index $\mu(A,q)$ for the static Mathieu Equation. \label{muindex}} \end{figure} Then, it is clear that $N$ production is in the border line between the broad and narrow resonance, and the resonance can be quite efficient in producing them. With respect to $\phi$ production, we only get a non negligible value of $\mu^{max}_\phi \sim 0.1$ in a small part of the trajectory, as can be seen in Figs. (\ref{matplot}) and (\ref{muindex}). Given the constraints in Eq. (\ref{rangeq}), at the beginning $\phi$ production occurs in the second resonance band, whilst $N$ may partially enter the first resonance band (see Fig. (\ref{muindex})). Moreover, even if the motion is not very suppressed by the Hubble expansion, parametric resonance takes place only during a fraction of the total period of the motion, so the final effective $\mu_k$ parameter averaging over several oscillations will be much smaller. Hence, for the $\phi$ particles we expect a narrow resonance if at all for momenta $k \lsim \bar{m}_\phi$. As the damping factor $\Phi(t)$ gets reduced in time, $N$ production will be shifted further to the left in the Mathieu chart, into the narrow resonance regime. However, $N$ particles with low momenta may be produced when the classical fields are not near the minimum, although this has nothing to do with parametric resonance. Whenever $N(t)/N_0 < 1/2$ the effective squared mass of the $N$ field, Eq. (\ref{mnn}), becomes negative and we enter the spinodal instability region. In this region, for the modes for which $k^2/a^2 + \mathcal{M}^2_{NN} < 0$, we will again have an exponential increasing solution that can be interpreted as production of particles. At the end of the day, it seems that production of $N$ particles is quite efficient for modes with $k \lsim \bar{m}_\phi$, and extremely effective for small modes below roughly $0.35 \bar{m}_\phi$. So far, so good. These would be the qualitative results if the $mixing$, $\mathcal{M}^2_{\phi N}$, were negligible, but this is never the case in SUSY hybrid inflation. The analysis with the mixing term for the $\delta \phi$, $\delta N$ system is greatly simplified given the fact that the background fields follow a straight line trajectory, implying only one oscillating mode. Hence, the coupled system for their quantum fluctuations can also be decomposed into two orthogonal modes, \begin{eqnarray*} \delta \phi_+ &=& \cos \theta \delta \phi + \sin \theta \delta N \,, \\ \delta \phi_- &=& -\sin \theta \delta \phi + \cos \theta \delta N \,, \end{eqnarray*} with $\tan \theta = - \sqrt{2}$. The time dependent frequencies for the modes $\delta \phi_{\pm}$ are now given by: \begin{eqnarray} \mathcal{M}_{+}^2&=& 2 \kappa^2 N(t) ( 3 N(t) - 2 N_0) \,, \\ \mathcal{M}_{-}^2&=& 2 \kappa^2 N(t) N_0 \,, \end{eqnarray} and the corresponding parameters in the Mathieu equation are: \begin{eqnarray} q_{+}(t) &=& \frac{16}{9} \left(\frac{N_0}{N(t)}\right)^2 \Phi(t),\;\;\;\; A_{+}(t)=\frac{8}{3} \left(\frac{N_0}{N(t)}\right)^2 \frac{k^2}{(a^2 \bar{m}_\phi^2)}+ \frac{3}{2} \frac{q_{+}(t)}{\Phi(t)} \,, \\ q_-(t) &=&\frac{4}{9} \left(\frac{N_0}{N(t)}\right)^2 \Phi(t),\;\;\;\; A_-(t)= \frac{8}{3} \left(\frac{N_0}{N(t)}\right)^2 \frac{k^2}{(a^2 \bar{m}_\phi^2)}+ 6 \frac{q_{-}(t)}{\Phi(t)}\,. \end{eqnarray} The Mathieu parameters for the mode $\delta \phi_{-}$ appear well to the left of the line $A=2 q$, i.e. deep in the narrow resonance regime (see Fig. (\ref{matplot}). However, for the mode $\delta \phi_+$ we have the reverse situation, and the amplification of this mode will occur in the efficient broad resonance regime. Moreover, the effective mass $\mathcal{M}^2_{+}$ becomes negative when $N(t)/N_0 < 2/3$, and again we will also have production of particles in region II of the classical oscillation. Finally, given that $\tan \theta \approx O(1) $, both $\delta \phi$ and $\delta N$ are dominated by the production in the $\delta \phi_+$ mode, \begin{equation} \delta \phi \simeq - \sin \theta \delta \phi_+ \,,\;\;\;\;\; \delta N \simeq \cos \theta \delta \phi_+ \,, \end{equation} leading to roughly the same rate of production for both particles. Moreover, in this case the mixing enhances the rate of production of $N$ particles compared with what was expected without it. This enhancement is a two-fold effect: on one hand parametric resonance is more effective due to the mixing, on the other hand we have a greater negative frequency in region II, with $\mathcal{M}^2_+|_{min}= -\bar{m}^2_\phi /3$, which will imply a wider range of momenta, $k/a < 0.58 \bar{m}_\phi$, in which particles can be more efficiently produced. To illustrate these points, we have numerically integrated the equations with $\kappa=10^{-3}$, $\phi_c= 10^{16}\, GeV$ and $m_\phi=3.7 \times 10^{9}\, GeV$, and obtained the occupation number of particles, Eq. (\ref{occno}). The results are plotted in Fig. (\ref{shybln}), and we have included for comparison those obtained when neglecting the mixing (left plot). Without the mixing, $\phi$ particles are hardly produced, whilst production of $N$ particles with low momenta proceeds quite quickly. Moreover, we can distinguish clearly in the plots the two different mechanisms of particle production. The peaks in $\ln n_k$ are due to parametric resonance in an expanding Universe, whereas between peaks $\ln n_k$ grows almost linearly in the first few oscillations reflecting the presence of a negative squared frequency in the evolution equation for these modes. When the Hubble expansion has reduced the amplitude of the oscillation such that $N(t)/N_0 > 1/2$ (without mixing) or further to $N(t)/N_0 > 2/3$ (with mixing), the occupation numbers become constant between the peaks. When the mixing is included (right plot), the occupation numbers for both particles become comparable, and the rate of production of $N$ particles is enhanced compared to the situation without the mixing. \begin{figure} \begin{tabular}{cc} \epsfxsize=8cm \epsfysize=8cm \hfil \epsfbox{shyb3ln.eps} \hfil & \epsfxsize=8cm \epsfysize=8cm \hfil \epsfbox{shyb3mln.eps} \hfil \end{tabular} \caption{Initial rates of production of $\phi$ and $N$ particles for the model with $\kappa=10^{-3}$, $\phi_c= 10^{16}\, GeV$ and $m_\phi= 3.7 \times 10^{9} \, GeV$. In the left side plot the results without taking into account the mixing term in the effective squared mass matrix are displayed; we have also included the oscillation of the classical field $N(t)$ (dashed line) scaled by a factor of 5. The comoving momentum $k$ is given in both plots in units of the natural frequency $\bar{m}_\phi$ and it corresponds to the physical momentum at $N' \approx -17$. In the right side plot, by the time particles are produced the initial value of the physical momentum $k/a=0.7 \bar{m}_\phi$ has been red-shifted below $0.58 \bar{m}_\phi$. \label{shybln}} \end{figure} \subsection{ Comment on backreaction. \label{comment}} Up to now, the analysis of particle production in the SUSY hybrid model has been done neglecting backreaction effects. Even if no other particles couple to the inflaton and $N$ fields, it can be inferred from Fig. (\ref{shybln}) that the production of their own quanta will soon affect the evolution of the system. Due to the presence of the tachyonic mass immediately after the end of inflation, these fluctuations may grow and modify the classical evolution even before the fields reach the minimum for the first time. After the inflaton has crossed the critical value, the classical fields may still move in the slow-rolling regime for a while, and we can consider that the Universe continues inflating for a further fraction of an e-fold $N_e \approx H t$ (for the parameters of Fig. (\ref{sugraplot}) $N_e \approx 0.3$). The first effect of the instabilities during this period would be to speed up the evolution and end the slow roll regime and thus inflation altogether. Their contributions $\langle \delta \phi_\alpha^2 \rangle$ to the classical equations can be computed using the Hartree Fock approximation\footnote{A brief description of this approximation is given later in section \ref{hartree} and Appendix B.} and can be interpreted as corrections to the effective masses of the fields. Thus, even in the slow roll regime with $N(t) \approx 0$, the mass of the inflaton becomes, \begin{equation} m_\phi^2 \rightarrow m_\phi^2 + 2 \kappa^2 \langle \delta N^2 \rangle \,. \end{equation} When the fluctuations grow such that both terms become comparable, the inflaton (and by extension the $N$ field) is pushed away from the slow roll trajectory. These contributions are dominated by the modes with the lower values of the momentum $\textbf{k}$, but their evolution will tend to follow that of the background fields (for small enough values of $\textbf{k}$, the equations of motion are the same). As the classical fields reach the minimum and begin to oscillate, production of particles in these low momentum modes practically stops, and production begins in slightly higher modes. This effect is similar to the ``zero mode assembly'' described in Ref. \cite{refDan1} in the context of new inflation type models. Despite the differences, new inflation and hybrid inflation share the fact that both end with a phase transition. Under certain conditions, it is shown how the spinodal instabilities drive the growth of the quantum fluctuations such that they dominate completely the inflationary dynamics ending inflation. The super-horizon modes together with the inflaton follow the same evolution and assemble into an effective zero-mode field which drives the dynamics of the system. In hybrid inflation, the contribution due to the particles produced in the lower modes during the pre-oscillatory, quasi slow-rolling regime, speeds up the initial evolution towards the minimum, but it does not reach the level of modifying the oscillatory regime. For this to happen, we have to wait until higher modes have been populated and become dominant. On the other hand, other particles can be present and these might be more efficiently produced during the oscillatory phase, such that they would dominate the backreaction effects. Given that in the original SUSY superpotential we have complex fields, their imaginary components, CP-odd singlet scalars, are among the particles that should also be included in the game. However, because the couplings to their real partners is model dependent, we will not attempt to study the production of these particles in a general way here. We will rather focus in the next section in a particular supersymmetric model of inflation motivated by particle physics considerations, where these questions can be addressed. \section{A particular model: $\phi$NMSSM Hybrid Inflation} The $\phi$NMSSM \cite{phiNMSSM} is a simple modified version of a well known particle physics theory, the next-to-minimal supersymmetric model (NMSSM). As such it is a realistic working supersymmetric model which supports inflation. The model has been formulated in the context of a SUGRA theory in \cite{SUGRA}. We will briefly review the global SUSY model here. The model is based on the superpotential: \begin{equation} W = \lambda \rm{N}\rm{H}_1 \rm{H}_2 - \it{\kappa}\rm{N}^2\Phi + ...,\label{SP} \end{equation} where $\rm{H}_1, H_2$ are the standard Higgs doublet superfields, and N and $\Phi$ are singlet superfields. As in the usual NMSSM, ${\kappa}$ and $\lambda$ are two dimensionless coupling constants. Since the singlet, $\Phi$ is linear in the superpotential, the potential will be very flat in its direction enabling it to act as the slowly rolling inflation field. This superpotential admits a global $U(1)_{PQ}$ Peccei-Quinn symmetry which forbids further terms and solves the strong-CP problem. The symmetry is broken at the scale of the singlet VEVs giving rise to an invisible DFSZ axion of that scale. Whilst the axions may provide a Dark Matter candidate, they must be tightly constrained in order not to over close the Universe or destroy the results of nucleosynthesis. Axions may be produced directly during the explosive preheating phase or they may be a decay product of the other massive particles produced during preheating. Previous to the start of inflation it is assumed that the field $N$ is driven to zero. During inflation the standard Higgs fields may be ignored and the scalar potential in terms of the canonically normalised real fields $\phi$ and $N$, becomes: \begin{equation} V_0(\phi,N) = V(0) + \frac{1}{4}{\kappa}^2N^4 + {\kappa}^2(\phi - \phi_{c-})(\phi - \phi_{c+})N^2 + \frac{1}{2}m_\phi^2 \phi^2, \label{Repotl} \end{equation} The constant, $V(0)$ has been added to ensure the necessary vanishing of the potential at its global minimum which corresponds to our vacuum and the critical values of $\phi$ for which the effective $N$ mass vanishes have been written explicitly in the potential and are given below in terms of the soft SUSY parameters, which appear in the soft SUSY-breaking part of the potential, $V_{soft}= \frac{1}{2}m_N^2 N^2 - \frac{1}{\sqrt{2}}\kappa A_\kappa \phi N^2$, \begin{equation} \phi_{c\pm} = \frac{A_{\kappa}}{2\sqrt{2}{\kappa}}\left(1 \pm \sqrt{1-\frac{4m_N^2}{A_{\kappa}^2}}\right). \end{equation} The first derivatives of the potential are, \begin{eqnarray} \partial V_0/\partial\phi &=&2{\kappa}^2(\phi-\phi_0)N^2 + m_\phi^2\phi, \nonumber \\ \partial V_0/\partial N &=&{\kappa}^2N[N^2 + 2(\phi-\phi_{c-})(\phi - \phi_{c+})]. \label{derivs} \end{eqnarray} Taking the fields to be positive, the (tree level) global minimum is found to be, \begin{eqnarray} \phi_0 & = & \frac{\phi_{c+}+\phi_{c-}}{2} = \frac{A_{\kappa}}{2\sqrt{2}{\kappa}}, \nonumber \\ N_0 & = & \frac{\phi_{c+}-\phi_{c-}}{\sqrt{2}} =\frac{A_{\kappa}}{2{\kappa}}\sqrt{1-\frac{4m_N^2}{A_{\kappa}^2}}, \end{eqnarray} and thus the vacuum energy which must be zero at the global minimum is given by, \begin{equation} V(0) \approx -V_0(\phi_0, N_0) = \frac{1}{4}{\kappa}^2 N_0^4. \end{equation} In the SUGRA model \cite{SUGRA} $V(0)$ is provided in the hidden sector by the moduli and dilaton contributions. As usual, during inflation the $N$ field is trapped in its false vacuum, $N=0$, whilst the $\phi$ field slowly rolls along the potential either from above $\phi_{c+}$ towards the origin (hybrid) or from below $\phi_{c-}$ and away from the origin (inverted hybrid \cite{IH}), depending on whether the small inflaton mass squared, $m_{\phi}^2$, is positive or negative respectively. Once at the critical value, the $N$ mass changes sign and both singlets roll down towards the global minimum signaling the end of inflation. The distinctive feature of having two critical values, $\phi_{c\pm}$, relies on the presence of the trilinear parameter $A_k$ in the potential. Since we expect $\lambda N_0 \sim 10^3\, GeV$ in order to motivate the $\mu$ parameter of the MSSM, the inclusion in the potential of the SUSY-breaking soft parameters, masses and trilinears, is mandatory in our model. However, once the potential is written as in Eq. (\ref{Repotl}), it can be seen as a generalisation of the SUSY hybrid potential, independently of the origin of the mass parameters. In the limit $A_k^2 \ll |m^2_N| $,($m_N^2 <0$), and thus $\phi_{c-} = -\phi_{c+}$, we recover indeed the standard SUSY hybrid potential, \begin{equation} V_0(\phi,N) = \frac{1}{4}{\kappa}^2 (N^2 -2 \phi^2_c)^2 + {\kappa}^2 N^2 \phi^2 + \frac{1}{2} m_\phi^2 \phi^2. \label{Repoth} \end{equation} Therefore, the main difference between Eqs. (\ref{Repotl}) and (\ref{Repoth}) lies in the origin of the scale $\phi_c$. In standard SUSY F-term hybrid inflation, this is given in terms of an intermediate mass scale $\Lambda$ coming directly from the superpotential, Eq. (\ref{supot}), whereas in our model no explicit mass scale is included in the superpotential and the critical value, $\phi_c \approx A_k/\kappa$, is generated via soft SUSY-breaking. From the cosmological point of view, this in turn will mark the difference between the models through the value of the Hubble constant and the vacuum energy during inflation. Whatever the origin of the critical scale $\phi_c$, the parameters in the model must satisfy the COBE constraints and the slow rolling conditions, Eqs. (\ref{epsilon}-\ref{cobecon}). Since in our model $\phi_c \sim \phi_0 \sim N_0$ must not exceed the bound on the axion scale, and moreover $\lambda N_0 \sim \kappa \phi_c \sim 1 \, TeV$, rough order of magnitude estimates for the parameters in the potential are found to be: $\phi_0 \sim N_0 \sim \phi_c\sim 10^{13}GeV,\; \lambda\sim {\kappa}\sim 10^{-10}$, hence the vacuum energy and Hubble constant are small, $V(0)^{1/4}\sim 10^8GeV$, $\; H_0 \sim 10^{-3} GeV$. Note that terms involving the Higgs doublets in the potential, (see Appendix A, Eq. (\ref{fullpotl})), will be at least a factor $\mathcal{O}({\kappa}^2)$ below the value of $V(0)$ above so essentially these fields will not contribute at all towards the vacuum energy. We result with a scale independent spectrum, $n=1$ predicted to an accuracy of $10^{-12}$. The inflaton mass, $m_{\phi}\sim eV$ comes from radiative corrections to its zero tree-level mass. The smallness of $\lambda$ and ${\kappa}$, the origin of $V(0)$ and the massless tree-level inflaton can be explained in terms of a no-scale SUGRA, superstring inspired theory in \cite{SUGRA} where it is also shown that the $\eta$-problem common to most ($F$-term) inflation models can be avoided here. In the next subsections we give the results of our investigation of the post-inflationary stage of the $\phi$NMSSM model described in the previous section. Firstly we analyse the behaviour of the classical homogeneous background fields according to the classical potential, neglecting particle production (Figs. \ref{fig:traj}-\ref{fig:Hub}); secondly we study the quantum fluctuations on the classical background of all fields in the theory, relating the growth of these fluctuations to particle production (Figs. \ref{fig:pprod}-\ref{fig:rates}); and lastly we find an estimate of the point at which backreaction of created particles becomes important and show the immediate effect this has on the classical field evolution (Figs. \ref{phiNbr}-\ref{ntbr}). Within the approximate order of magnitude results necessary to realise the model, we have chosen the set of parameters below for our numerical study: \begin{eqnarray} \phi_0 = N_0 = 10^{13}GeV,&&{\kappa} = 10^{-10}, m_\phi = 10^{-9}GeV, \label{params}\\ \Rightarrow&&\frac{\phi_{c\pm}}{\phi_0} = 1 \pm \frac{1}{\sqrt{2}}. \nonumber \end{eqnarray} More explicit details and calculations involving the other fields in the potential needed in the analysis of the post-inflation era are given in the Appendix at the end of the paper. \subsection{Classical results - neglecting particle production}\label{sec4.1} Being no more than a particular version of the SUSY F-term hybrid model, the $\phi$NMSSM model leads to an evolution of the classical homogeneous fields, $N(t)$ and $\phi(t)$ (ignoring particle production) which follows the general pattern given in section \ref{clshyb}. In this case the inflaton field $\phi$ will oscillate around its non-zero VEV, $\phi_0$. The two ratios of scales that fix the behaviour of the oscillations are: \noindent (a) $m_\phi/ (\kappa \phi_c) \approx 10^{-12}$; the motion begins with extremely small slow-rolling values, and moreover the contribution of the tiny inflaton mass $m_\phi$ becomes negligible very soon in the evolution. As a result, the fields will follow the straight line trajectory (Fig. (\ref{fig:traj}) to a very good degree of accuracy, and the actual oscillations of the fields are indeed in phase. \noindent (b) $H_0/ (\kappa \phi_c) \approx 10^{-6}$; the oscillations will begin with very large, lightly damped amplitudes, and oscillate $\sim 10^{5}$ times in a Hubble time. Fig. (\ref{fig:evol}) shows the evolution of the fields from the moment the critical point is passed. The time scale displayed is in terms of the approximate number of oscillations of the fields, given by: \begin{equation} N=\frac{\bar{m}_\phi t}{2\pi}=\frac{\bar{m}_N t}{2\pi}, \end{equation} where one e-fold ($N_e \simeq H t $) corresponds to roughly $\sim 2.7 \times 10^5$ oscillation times. This approximation becomes more exact as the oscillations approach simple harmonic motion at late times in the evolution when the amplitudes are very small and the masses given by (\ref{masses}) become a better approximation to the actual frequency of these late oscillations. In several of the plots we have used the variable $N'=N-N_{start}$, where $N'=0$ is defined to be the moment when the fields reach the global minimum for the first time. We can see in Fig.~\ref{fig:evol} that the fields take a reasonable length of time, $N_{start}\sim$~$2.9 \times 10^4$ oscillation times (about 1/10 of a Hubble time) before they begin to oscillate, so in fact inflation continues for just a fraction of an e-fold after the critical point is passed. The fields then begin to oscillate with amplitudes decreasing due to the expansion of the Universe. In terms of a Hubble time, the total amplitude will decrease rapidly at first and more slowly as time progresses until after about 2 e-folds the oscillations have decreased to $\sim 1/10$th their original amplitude. The frequency of oscillations increases as they progress, and the motion eventually evolves towards simple harmonic. The period of oscillation rapidly decreases to begin with, indicating that in comparison to the Hubble time, simple harmonic motion becomes a good approximation relatively early on in the stages of oscillation. However, particle production will occur during the first few oscillations which are far from harmonic. Due to the low value of the Hubble constant relative to $\bar{m}_\phi$ and hence the little effect it has on the amplitude of the first oscillations, in the top right panel in Fig. (\ref{fig:evol}) we can clearly distinguish the two different parts of each oscillation: the peak, corresponding to more harmonic-like motion around the minimum, Eq. (\ref{solnminh}), and the non-oscillating section where the fields decay as $\sim 1/t^2$ as they approach the initial values. This second part of the motion is rapidly affected by the expansion of the Universe as it causes a slight decrease in the amplitude, and at later stages (top right) it is further suppressed. The behaviour of the Hubble constant is shown in Fig.~\ref{fig:Hub} where we have also plotted the quantity $\varepsilon=-\dot{H}/{H^2}$. $\varepsilon=0$ corresponds to the vacuum dominated era of inflation, $\varepsilon=3/2$ to the reheating phase when matter (specifically the decay products before they have thermalised) dominates, and $\varepsilon=2$ to the final (after the reheat temperature has been reached) radiation dominated Universe. During the oscillations, $\varepsilon$ oscillates between the values 0 and 3 and its average value per oscillation, $\bar{\varepsilon}$ shows the transition between vacuum and matter dominated states. \begin{figure} \hfil \scalebox{0.5}{\rotatebox{270}{\includegraphics{fileC1.ps}}} \hfil \caption{Classical trajectory in field space ($\phi,N$) after inflation, (for the parameters given in (\ref{params})) in the case of both hybrid and inverted hybrid inflation. The fields begin at $\phi=(1 \pm 1/\sqrt{2})\phi_0, N=0$, and oscillate around $\phi=\phi_0, N=N_0$, where, neglecting particle production, they eventually settle.} \label{fig:traj} \end{figure} \begin{figure} \scalebox{0.315}{\rotatebox{270}{\includegraphics{fileosc1.ps}}} \scalebox{0.315}{\rotatebox{270}{\includegraphics{fileosc4.ps}}} \scalebox{0.315}{\rotatebox{270}{\includegraphics{fileshadedamp.ps}}} \scalebox{0.315}{\rotatebox{270}{\includegraphics{filefreq.ps}}} \caption{Top: Oscillations of the classical fields in the hybrid case at the first stage and at a later stage in their evolution. Bottom left: Evolution of the fields and Hubble constant after inflation in the hybrid case over $\sim 2$ e-folds, showing the damping of the amplitudes of oscillations caused by the decreasing Hubble constant. The fields take $\sim 0.1$ e-folds before they begin to oscillate in the shaded regions of the plot. Bottom right: The decreasing period of oscillations in units of $N$ (period of simple harmonic oscillations).} \label{fig:evol} \end{figure} \begin{figure} \scalebox{0.315}{\rotatebox{270}{\includegraphics{filesmallscale2.ps}}} \scalebox{0.315}{\rotatebox{270}{\includegraphics{fileebar.ps}}} \caption{Behaviour of Hubble constant. Left: Small scale behaviour, showing oscillations in $\varepsilon$. $H/H_0$ is shown over the range [0.97814,0.97808], but has been rescaled to make it visible in the plot. Right: Large scale behaviour - Universe approaching matter domination ($\bar{\varepsilon}\rightarrow 3/2)$.} \label{fig:Hub} \end{figure} \subsection{Quantum effects - Preheating} \label{sec4.2} Treating $\delta\phi_{\alpha}$ as fluctuations in the presence of the background fields, $\phi_{\alpha}(t)$ which have the independent dynamics described in the previous section, we solve the equations of motion for the quantum fluctuations of all scalar fields in the theory, \begin{equation} \ddot{\psi_k}_\alpha + \frac{k^2}{a^2}{\psi_k}_\alpha + \mathcal{M}^2_{\alpha\beta}{\psi_k}_\beta = 0, \label{psieq1} \end{equation} where ${\psi_k}=a^{3/2}{\delta\phi_k}$. Here, the term $\epsilon$ in Eq. (\ref{psieq}) is negligible since the Hubble constant is so small relative to $\bar{m}_\phi$ and $\varepsilon \sim O(1)$ (see Fig. (\ref{fig:Hub})) so $\dot{H}$ is also small in this model. In fact, particle production will take place in a tiny fraction of a Hubble time, and we can safely neglect any effect due to the expansion of the Universe in the following. Since we are not considering the quantum fluctuations of the standard Higgs fields, $\delta H_i$, (see Appendix A), the system above reduces to two sets of two coupled equations describing the quantum fluctuations of the $N$, $\phi$, the axion and $X$ fields. The axion and the $X$ field are the massless/massive pseudoscalars in the model, defined in terms of the imaginary components of the complex fields, \begin{eqnarray} a = N_I\cos\alpha_a - \phi_I\sin\alpha_a\,,& &\nonumber\\ X = N_I\sin\alpha_a + \phi_I\cos\alpha_a\,,& & \end{eqnarray} with $\tan \alpha_a = 2 \phi_0/ N_0$. The mass squared matrices, $\mathcal{M}^2_{\phi N}, \mathcal{M}^2_{a X}$, are given in Eqs. (\ref{mphiN}) and (\ref{maX}) in Appendix A. \begin{figure} \scalebox{0.315}{\rotatebox{270}{\includegraphics{fileax.ps}}} \scalebox{0.315}{\rotatebox{270}{\includegraphics{filepap9.ps}}} \caption{Explosive production of particles, displayed here by the exponential increase in the occupation numbers, $n_k$ of axions and massive $X$ particles for $k=0.3\bar{m}_\phi$ (left); and of $\phi$ and $N$ particles for $k=0.6\bar{m}_\phi$ (right) with time (in units of the fundamental period of oscillation of the classical fields, $N$). The results without including the mixing terms are also shown for comparison. Without mixing there is no production of $X$ particles and $\phi$ particle production is negligible. Including the mixing terms results in virtually identical numbers of $\phi$ and $N$ particles; and of axions and $X$ particles being produced.} \label{fig:pprod} \end{figure} The results obtained by solving the equations numerically for the parameters given in (\ref{params}) are summarised in Figs. (\ref{fig:pprod}-\ref{fig:rates}) where the occupation number $n_k$ is defined in Eq. (\ref{occno}). In fact particle production appears to be extremely efficient. We have also plotted for comparison the results obtained treating each particle separately and solving Eqs. (\ref{psieq1}) as if they were four uncoupled equations, ie. without the mixing terms ($\alpha\ne \beta$) in the mass squared matrix $\mathcal{M}^2_{\alpha\beta}$. In particular in Fig. (\ref{fig:pprod}) we can see that with mixing there are less axions produced, but now $X$s are also produced at the same rate, whereas $N$ particle production increases considerably and now $\phi$ particles are also produced at the same rate. The numerical results for the $\phi-N$ system fit quite well with the analytical behaviour obtained in section \ref{pp} using the Mathieu equation . In the axion-$X$ system, the analysis based on the Mathieu equation will depend on the details of the model, for instance the value of $\tan \alpha_a$, and some of the approximations may not be valid, such as neglecting terms proportional to the cosine squared in the effective masses. Bearing this in mind, we still can get some qualitative features studying the equations without the mixing term. The Mathieu parameters are then given by: \begin{eqnarray} q_{a}&\simeq& \frac{4}{9} |\cos^2{\alpha_a}-\sin{2\alpha_a}/\sqrt{2}| \left(\frac{N_0}{N(t)} \right)^2 \simeq 0.16 \left(\frac{N_0}{N(t)} \right)^2 , \\ A_{a}(t)&\simeq& \frac{8}{3} \left(\frac{N_0}{N(t)} \right)^2 \frac{k^2}{(a^2\bar{m}_{\phi}^2)},\\ q_{X} &\simeq & \frac{4}{9} |2+\sin^2{\alpha_a}-\sqrt{2}\sin^2{\alpha_a}\tan{\alpha_a}| \left(\frac{N_0}{N(t)} \right)^2 \simeq 0.24 \left(\frac{N_0}{N(t)} \right)^2, \\ A_{X}(t) &\simeq& \frac{8}{3} \left(\frac{N_0}{N(t)} \right)^2 ( \frac{k^2}{a^2\bar{m}_{\phi}^2}+\tan^2{\alpha_a}+1) \simeq \frac{8}{3} \left(\frac{N_0}{N(t)} \right)^2 \frac{k^2}{(a^2\bar{m}_{\phi}^2)} + 56 q_X \,, \end{eqnarray} where the RHS numerical factors correspond to $\tan \alpha_a =2$. We notice that production of $X$ particles would be practically non-existent, whilst axions may be produced but with $q_a < 0.36$, which leaves us in the narrow resonance regime. However, these are only indications of what could happen if the axion and X were not coupled. Due to the mixing term, we get that both particles are produced at the same rate, but in this case the production of axions will be further suppressed. Fig. (\ref{fig:muk}) plots an estimate of the initial rate of growth of the occupation number of each particle as a function of the wavenumber $k$, ${\mu_k}_\alpha\sim (4\pi)^{-1}d(\ln {n_k})_\alpha/dN.$ Without taking into account the mixing terms, the largest rates correspond to the lowest mode, whilst with the mixing the effective growth index for $\phi$ and $N$ particles is negligible for $ k \ll \bar{m}_\phi$, but it has a maximum around $k \approx 0.15 \bar{m}_\phi$. As mentioned in section \ref{comment}, the lower modes follow the evolution of the homogeneous fields and their production is suppressed during the oscillations. For a particular typical value of $k$, Fig. (\ref{fig:rates}) shows the relative rates of production of particles. For the beginning oscillations, the classical fields spend less than 1/6 of the total period of the oscillation around the minimum, where parametric resonance takes place. In Fig. (\ref{fig:rates}), this would correspond to the spikes in the $N$ rate of production with no mixing. Thus, the relatively large values of $\mu(k)$ for low values of the momentum $k\lsim 0.3 \bar{m}_\phi$ is produced by a simpler mechanism, and is due to the presence of negative squared frequencies in the evolution equation of the quantum fluctuations, as can be seen in Fig. (\ref{fig:wks}). On the other hand, we want to stress the fact that with mixing the rate of production for the axion and the $X$ are brought together, but this now corresponds to a smaller rate of production for the axion. We will see in the next section that indeed the total number of axions produced is several orders of magnitude smaller than that of the inflaton and $N$ particles. The situation for the pseudoscalars may change depending on the model. For example, in models where $\phi_0=0$, ($\phi_{c+}= -\phi_{c-}$) there would be no mixing among the massive and the massless pseudoscalars and the axion would be the imaginary component of the $\rm{N}$ field. When particle production begins, the effective squared mass for the axion would be given by (see Eq. (\ref{mimag}) in Appendix A), \begin{equation} \mathcal{M}^2_{a}= 2 \kappa^2 N ( N - N_0) \,. \end{equation} which again becomes negative whenever $N(t) < N_0$. In this case, we would find that indeed the production of axions may proceed at the same rate as that of the inflaton and $N$ particles in a extremely efficient way. \begin{figure} \scalebox{0.315}{\rotatebox{270}{\includegraphics{filemuk7.ps}}} \scalebox{0.315}{\rotatebox{270}{\includegraphics{filemukphi.ps}}} \caption{Typical values of the growth parameter, $\mu_k$ over the first few ($N'\le 40$) oscillations of the classical fields as a function of the momentum $k$ in units of $\bar{m}_\phi$.} \label{fig:muk} \end{figure} \begin{figure} \hfil \scalebox{0.5}{\rotatebox{270}{\includegraphics{filepap4.ps}}} \hfil \caption{Comparison of initial rates of production of all particles for a typical momentum, $k=0.3\bar{m}_\phi$.} \label{fig:rates} \end{figure} \begin{figure} \hfil \scalebox{0.5}{\rotatebox{270}{\includegraphics{filewks2.ps}}} \hfil \caption{The squared frequencies of the quantum fluctuations, $w_k^2$ in units of $\bar{m}_\phi^2$ over one classical field oscillation for the zero modes ($k=0$) without mixing, ie. ${w_k}_\alpha^2=\mathcal{M}^2_{\alpha\alpha}$.} \label{fig:wks} \end{figure} \subsection{Effects of backreaction. \label{hartree}} Due to the extremely large numbers of $\phi$ and $N$ particles produced during the first few oscillations of the background fields, these particles will quickly give a non-negligible contribution to the evolution equations. Therefore, instead of the tree-level equations considered until now, the full dynamics taking into account the interactions among the particles being produced should be included in the evolution equations. In practice this means that the polarization tensors for the different fields must be computed beyond the tree-level approximation. A consistent treatment of all non-linear effects involved can be obtained by using lattice techniques \cite{ref31,ref3}. However, in order to get some insight on how backreaction affects the process of preheating, the Hartree approximation \cite{hartree,hartree2} for the potential should be good enough. Within the Hartree approximation, the main one-loop contributions to the tadpole and mass terms in the potential are included, not only for the background fields but for any quantum field. A summary of how to expand the potential and the new expressions for the effective masses are given in Appendix B. Including one-loop corrections, the effective mass of $all$ the fields pick up new non-thermal contributions proportional to, \begin{equation} \langle\delta\phi_\alpha^2\rangle = \int\frac{k^2dk}{2\pi^2}|{\delta\phi_k}_\alpha|^2. \label{average} \end{equation} For example, the effective mass for the inflaton field $\phi$ and its quantum fluctuation will become: \begin{equation} \mathcal{M}^2_{\phi \phi}= 2 \kappa^2 (N^2 + \langle\delta N^2\rangle + \cos^2 \alpha_a (\langle \delta a^2 \rangle + t^2 \langle \delta X^2\rangle + 2 t \langle \delta a \delta X \rangle ) \,, \end{equation} where $t=\tan \alpha_a=2 \phi_0/N_0$ is the mixing angle in the axion-$X$ sector defined in Appendix A. The average $\langle \delta \phi_\alpha^2 \rangle$ can be seen as a closed loop of a particle $\delta \phi_\alpha$. The Hartree approximation does not cover all possible types of contributions to the polarization tensors, and the effects of rescattering among the different species are missed. However, by integrating the coupled system of equations together with Eq. (\ref{average}), it succeeds in summing up the infinite set of diagrams of the ``cactus'' or ``daisy'' type. The overall effect of the change in the effective masses will be to shut down the explosive production of particles. On one hand, it will increase the frequency of the oscillation of the background fields, hence decreasing their amplitudes; on the other hand, it will render the squared frequencies of the quantum fluctuations positive, ending the process. Using the results of the previous section we can estimate the time at which backreaction becomes important, that is, when $\langle\delta \phi_\alpha^2\rangle \sim O( \phi_0^2) \sim O (N_0^2)$. The expectation values are related to the total number of particles produced, and the later can be estimated given the growth index function $\mu(k)$. It is then clear from Fig. (6) that we have to worry mainly about $\phi$ and $N$ production, and that the total number of axions and $X$ fields is several orders of magnitude smaller. This simplifies life, because the function $\mu(k)$ for the fields $\phi$ and $N$ can be quite nicely approximated by: \begin{equation} \mu(k) \approx 0.6\left(\frac{k}{\bar{m}_\phi}\right)^{1/3}exp(-5 (k/\bar{m}_\phi)^{3/2}) \,, \end{equation} valid for the first few oscillations. The total number of particles $\phi$ (or $N$) produced at a time $N'=\bar{m}_\phi t/(2 \pi)$ is then: \begin{equation} n_\phi(N') = \frac{1}{4 \pi^2 a^3} \int dk k^2 e^{4 \pi \mu(k) N'}\,, \end{equation} and we get $n_\phi \approx O(\phi_0^2)$ when $N' \approx 20 $. Therefore, backreaction effects will become important just after the third oscillation of the inflaton field. Soon after that, the results given in the previous section are modified and quantum fluctuations cease to be produced at a exponential rate. As a consequence, we will never reach the point where we have produced too many axions (see Fig. (\ref{fig:pprod})). In Fig. (\ref{fig:muk}) it can be seen that only particles with low momenta, $k < \bar{m}_\phi$, are effectively being produced. Then, the value $\bar{m}_\phi$ can be used as a cut-off in the momentum integrals, especially when computing the expectation values $\langle\delta \phi^2_\alpha\rangle$. In principle, these quantities are quadratically divergent and have to be renormalised \cite{refren}. In practice, we avoid this subtle point using the cut-off; the rest of the couplings and mass parameters appearing in the equations are then taken to be the renormalised ones. \begin{figure}[h] \epsfxsize=12cm \epsfysize=10cm \hfil \epsfbox{phiNbr.eps} \hfil \caption{Oscillations of the classical fields with backreaction effects included.} \label{phiNbr} \end{figure} The numerical results confirm our estimation. In Fig. (\ref{phiNbr}) it is shown how the oscillations for the background fields change after $N' \sim 20$. However, they are still in phase and follow the trajectory given in Fig. (\ref{fig:traj}). Recently, it has been stressed that, in models with symmetry breaking, backreaction of the quantum fields could restore the original symmetry \cite{symres}, if they come to dominate the effective mass during the stage of preheating. This is not the case in our particular model. Backreaction itself prevents the quantum fluctuations from reaching the stage where they might restore the symmetry. However, the shape of the potential changes and when particle production has effectively stopped, the classical fields end oscillating around a minimum different to $\phi_0$, $N_0$. We stress that the whole preheating process takes place within much less than a Hubble time. This means the expansion of the Universe can be neglected during this time. The decrease in the amplitude of the oscillating fields in Fig. (\ref{phiNbr}) is purely due to the presence of the fluctuations. \begin{figure}[h] \epsfxsize=12cm \epsfysize=10cm \hfil \epsfbox{nkbr.eps} \hfil \caption{Hartree approximation: Rates of production of all particles for a typical comoving momentum $k= 0.3 \bar{m}_\phi$.} \label{nkbr} \end{figure} In Fig. (\ref{nkbr}) we give the rate of production for all the different species of particle with a typical momentum $k=0.3 \bar{m}_\phi$. Soon after $N' \sim 20$ they reach a plateau, signaling the near end of the preheating period. Fig. (\ref{ntbr}) shows the total number of particles per comoving volume being produced. We end with the ratio of scalars to pseudoscalars being $O(10^{20}$). Preheating is very efficient in producing $\phi$ and $N$ quanta, and in fact it is this effect which precludes other particles from being so efficiently produced. Backreaction due to $\phi$ and $N$ affects the evolution of $every$ field (classical and quantum) in the model. This, however, may not be the end of the story. The Hartree approximation neglects the effects of rescattering among the particles produced, but it is possible that these terms could be the same order of magnitude as the expectation values $\langle\delta \phi_\alpha^2\rangle$ \cite{ref7}. Lattice simulations \cite{ref3} for one-field inflation models have shown that the maximum value of $\langle\delta \phi_\alpha^2\rangle$ is overestimated in the Hartree approximation when the resonance is very efficient. However, once this maximum value is reached the resonance ends qualitatively in the same way as in the Hartree approximation. It is difficult to extrapolate from these studies to our model. We may have overestimated the total number of $\phi$ and $N$ produced, and/or it may happen that rescattering may induce an extra little production of axions and $X$. But it is unlikely this could reduce the large gap among $n_a$ and $n_\phi$. We can venture that, after preheating, the Universe will end with the classical fields still displaced from the global minimum, and fill up with nearly non-relativistic particles, a large fraction of them being $\phi$ and $N$. A fraction of the energy storage in the inflaton field at the end of inflation has been transferred to the particles produced, and this will set the initial conditions for the subsequent evolution of the system, as given by the standard theory of reheating; in the Hartree approximation roughly 60\% of the initial energy density is converted into quantum fluctuations. Moreover, by the time reheating is complete, any density of stable particles (i.e. axions) would be diluted by the entropy released in the decay of the $\phi$ and $N$. \begin{figure}[h] \epsfxsize=12cm \epsfysize=10cm \hfil \epsfbox{ntbr.eps} \hfil \caption{Hartree approximation: total number of particles produced.} \label{ntbr} \end{figure} {\section{Conclusion}} We have analysed in this paper the preheating era which occurs immediately after the end of inflation in a general SUSY model of hybrid inflation. From the point of view of the inflationary potential it is no more than a particular case of the general hybrid model, Eq. (\ref{Vhyb}). Still, preheating in this particular model is found to be quite different when studied in more detail. Supersymmetry leads to equal masses at the global minimum for the real singlets relevant to inflation ($\phi$ the inflaton and $N$), and therefore to the presence of a common frequency of oscillation for the classical fields. That is, the supersymmetric potential allows an approximate solution with only one mode of oscillation for the classical fields, the orthogonal one being stationary. This solution becomes exact as the explicit tiny mass of the inflaton, $m_\phi$, tends to zero. Due to the inflationary dynamics (slow roll conditions) this mass is indeed always negligible, independently of the value of any other parameter. Moreover, the slow roll scenario also fixes the initial conditions at the beginning of the reheating era such that we begin with very small initial velocities. Therefore, the fields will follow almost exactly a straight line trajectory, oscillating with equal frequencies and proportional amplitudes. Hence, for the purpose of studying particle production, the SUSY hybrid model can be simplified to a single field model (the oscillating mode), with the potential given in Eq. (\ref{Vsingle}). Since we are dealing with an effective single field oscillating, we can attempt to convert the evolution equations for the quantum fluctuations into Mathieu-like equations, whose solutions in term of stability/instability regions is well known. We have first studied production of $\phi$ and $N$ particles, which is completely independent of the details of the model. Even in a first approximation neglecting the effects of backreaction, the evolution equations are coupled through mixing terms in the effective mass matrices. The equations for $\delta \phi$ and $\delta N$ taken independently (that is, neglecting the mixing term) result in a higher rate of production of $N$ particles than $\phi$, the latter being almost completely suppressed. However, the effect of the mixing term is to bring together the rate of production for both particles, and overall to become more effective indeed. The fact that we achieve almost the same number of particles is general to any system of coupled particles with non negligible mixing. In other words, during the preheating proccess we can produce particles which in principle are not expected to exist, if they are coupled to other particle species which are efficiently produced. The rate of production of the latter will be increased/decreased depending on the strength of the mixing. Given that hybrid inflation ends by a phase transition, for some values of the fields tachyonic masses will be present, for example that of the $N$ field. As well as an initial production of particles when the fields are first moving towards the minimum, if the amplitude of the oscillations is large enough there will be additional production of particles between oscillations when the classical fields are in the unstable region of the potential with negative curvature. Therefore, it is the combination of parametric resonance and tachyonic mass which causes the production of $\phi$ and $N$ particles to be so efficient. Production of $\phi$ and $N$ fluctuations will rapidly give rise to non-negligible backreaction effects which will modify the subsequent evolution of the system. However, other particles may be efficiently produced in the model which can also contribute to this effect. With this issue being model dependent, we have proceeded to analyse the situation in a particle physics based model of inflation, the $\phi$NMSSM \cite{phiNMSSM}. This model is not only suitable for inflation, but also solves the strong CP problem by an approximate Peccei-Quinn symmetry and leads to an invisible axion, which could be dangerously produced during preheating. However, in this case the mixing term between the axion and the massive pseudoscalar slightly suppresses the axion production rate, which turns out to be much smaller than that of the real scalars. Nevertheless, we obtain the same general result of comparable rates of production for both pseudoscalars. Backreaction effects are then dominated by the production of the real components of the fields, which quite rapidly succeed in preventing the production of any other particle species. Therefore, axions in this model are kept safely under control, thoughtout and after the preheating era. \begin{center} {\bf Acknowledgements} \end{center} We thank Juan Garc\'{\i}a-Bellido both for motivating this work and for discussions. \begin{center}{--------------------------------------------------------------------------} \end{center}
\section{Introduction} Supersymmetry (SUSY) is currently believed to lead to the most attractive scenario of physics beyond the standard model (SM). However, the low-energy spectrum of fermions and bosons does not exhibit this symmetry, so SUSY, if it exists, must somehow be broken to be of any relevance to Nature. SUSY breaking is different from the more familiar symmetry breaking in the SM since the supertrace theorem prevents SUSY breaking by tree-level renormalizable couplings. According to this theorem one must assume that the sector responsible for SUSY breaking is hidden, i.e. it has no renormalizable tree-level couplings with the visible sector. A popular realization of this breaking can be accommodated through gravity, a theory which is altogether non-renormalizable, and this mechanism has been explored for a long time \cite{acn}. Another possibility, in which a different mechanism is used to communicate SUSY breaking, can be provided by keeping the original theory renormalizable, but with a low-energy description in terms of an effective Lagrangian with non-renormalizable terms. Within the first scenario (i.e. theories with gravity-mediated SUSY breaking) soft mass terms are generated at the Planck scale and one cannot produce flavor-invariant SUSY-breaking terms for the sfermion masses. On the other hand, in the second scenario it is possible to generate soft terms at some ``messenger'' scale $\Lambda_M$ below the Planck scale and to break flavor symmetry only through Yukawa couplings. This can be achieved within the framework of theories which take advantage of the general mechanism of gauge-mediated supersymmetry breaking (GMSB) \cite{gmsb} in which, just as in the SM, the Yukawa couplings are the only sources of flavor violation. In GMSB theories one introduces vector-like quarks and leptons whose role is to break supersymmetry. These theories have received considerable attention in recent years \cite{gmsbnew}, largely due to their high predictability and their emphasis on a dynamical origin for SUSY breaking. To carry out the GMSB program, one starts with an observable sector which contains the usual matter and gauge fields and their supersymmetric partners, while leaving the hidden sector unspecified. One then introduces a {\it messenger} sector, formed by the new superfield $X$, whose coupling with the goldstino superfield generates a supersymmetric mass of order $S$ for the messenger fields, and leads to mass splittings of order $F$ (where we denote by $S$, $F$ the vacuum expectation values of the scalar and the auxiliary component of the superfield, respectively), and thus $\sqrt{F}$ is identified with the scale of SUSY breaking in the messenger sector \cite{dineetal}. Apart from the requirement that the messenger fields transform under the SM gauge group, the messenger sector is unknown and is the main source of model-dependence in GMSB theories. Explicit model building in theories which take advantage of the GMSB mechanism has been explored in a number of papers by taking $SU(5)$ to be the unifying gauge group in most cases \cite{gmsbph,borzu,skiba,martin,chanowitz}. Despite serious shortcomings related to problems such as the nucleon decay rates and neutrino masses, as well as $R$-parity breaking, the predictability and simplicity of these models has served as a useful first illustration of the applicability and power of the idea of GMSB. An attempt to go beyond these models was made by the authors of \cite{mohapatra}. They chose to study the electroweak gauge group $SU(2)_L\times U(1)_{I_{3 R}} \times U(1)_{B-L}$ since it automatically guarantees $R$-parity conservation. They also explored messenger fields which can play the role of the Higgs fields and break the chosen gauge group down to the SM. In this work we propose to study the GMSB mechanism and its phenomenological consequences in a large class of supersymmetric $SO(10)$ grand unified theories (GUTs) by taking a different approach. We start with the most general messenger sector and break the $SO(10)$ symmetry to either $SU(2)_L\times U(1)_{I_{3 R}} \times U(1)_{B-L}$ or $SU(2)_L\times SU(2)_R \times U(1)_{B-L}$, with a subsequent breaking to the SM gauge group. We enforce the boundary condition that the gauge couplings must unify at a common scale by using the one-loop renormalization group evolution of the couplings. We use this boundary condition to restrict the messenger field masses and to predict the complete sparticle mass spectrum. Our paper is organized as follows. In section II, we describe the $SO(10)$ model and its breaking. In section III we introduce the $SU(2)_L\times U(1)_{I_{3 R}} \times U(1)_{B-L}$ and $SU(2)_L\times SU(2)_R \times U(1)_{B-L}$ models and give their particle content. In section IV we discuss the restrictions on SUSY breaking and present the resulting solutions. The particle spectrum is discussed in section V. We conclude with some remarks in section VI. \section{$SO(10)$ models} $SO(10)$ GUTs have received a great deal of attention over the years and continue to do so, thanks to the improved measurements of the low-energy gauge couplings which confirm that supersymmetry leads to an extremely accurate (perturbative) unification of couplings \cite{bb}. Also, from a more theoretical point of view, the $SO(10)$ gauge group is a natural candidate for supersymmetric unification since all the quarks and leptons of a single generation are the components of a single spinor representation. Furthermore, the $SO(10)$ gauge group has the particularly attractive feature that it can break to the SM through either $U(1)_{I_{3 R}}$ or $SU(2)_R$, thus providing non-zero neutrino masses through the see-saw mechanism. Due to this, and in view of the recent observations at SuperKamiokande \cite{Kamiokande} which indicate non-vanishing neutrino masses, $SO(10)$ GUTs must be considered even more seriously than before. Additionally, $SO(10)$ GUTs also provide interesting fermion mass relations. For example, depending on the scale of the right-handed symmetry breaking, one can predict lower bounds on neutrino masses. In this context, it was shown that $SO(10)$ is a realistic and natural supersymmetric grand-unified theory \cite{bb}. Yet another important feature of $SO(10)$ SUSY GUTs is that they make it possible to achieve the desired doublet-triplet splitting (i.e. keeping the pair of Higgs doublets of the supersymmetric SM light, while giving their colour triplet partners superheavy masses to avoid proton decay) without fine-tuning \cite{dtsp}. Let us now turn our attention to $SO(10)$ symmetry breaking in $SO(10)$ SUSY GUTs. $SO(10)$ symmetry breaking can proceed in essentially two ways: $SO(10)$ can be broken to $SU(5) \times U(1)$ at scale $\sim 10^{18}$GeV and then further broken down to the minimal supersymmetric standard model (MSSM) at scale $\sim 10^{16}$GeV. Or, alternatively, $SO(10)$ can be broken to some left-right symmetric group, e.g. $G_{224}=SU(2)_L\times SU(2)_R\times SU(4)_C$, at some high energy scale, where the $SU(4)_C$ group contains the subgroup $SU(3)_C\times U(1)_{B-L}$. $G_{224}$ is then broken at some intermediate scale to the MSSM. The latter possibility of breaking $SO(10)$ is of particular interest, since the supersymmetric left-right model can naturally account for parity and comes with the extra bonus of providing possible solutions to both the SUSY CP problem and the R-parity problem \cite{mohras}. In this work we will further assume the D-terms to vanish (they could in principle contribute to the supersymmetry breaking masses). It can be argued that renormalizable see-saw mechanism and spontaneously broken $B-L$ symmetry lead to exact $R$-parity at all energies \cite{amrs}. These are important and attractive features of the left-right models, which is why we wish to study this possibility of breaking through the GMSB mechanism in more detail in this paper. The symmetry breaking chain that we study here is as follows. We assume that the $SO(10)$ gauge group is broken to an intermediate left-right symmetry group $G_{LR}$ at scale $M_G$. $M_G$ should be no less than $10^{16}$~GeV to have stable nuclei and it should be below the Planck scale $10^{19}$~GeV. The intermediate symmetry group $G_{LR}$ is then broken down to the MSSM at scale $M_R$. For simplicity, and to maintain only a minimal number of scales in the theory, we will assume that there are no further symmetry breaking scales between $M_R$ and $M_G$. As possible left-right symmetry groups we will consider $G_{LR}^I=SU(3)_C \times SU(2)_L \times U(1)_{I_{3 R}} \times U(1)_{B-L}$ and $G_{LR}^{II}=SU(3)_C \times SU(2)_L \times SU(2)_R \times U(1)_{B-L}$. Next we must include SUSY breaking into our model. The simplest possibility which avoids the proliferation of scales is to assume that the SUSY breaking scale and the left-right gauge symmetry breaking scale could be somehow related. To realize this we will further assume that the SUSY breaking scale $\Lambda_{\text{SUSY}}$ is the same as the left-right symmetry breaking scale $M_R$. In terms of the scales introduced in the previous section, we relate explicitly the scales as $\Lambda_{\text{SUSY}}=\frac{F}{S}$. This is an attractive choice since it simultaneously connects the scale of the gauge symmetry breaking to the scale of SUSY breaking and fulfills the requirement that the breaking of the electroweak symmetry remain radiative.In order for the sparticle masses to be around 1 TeV, the SUSY breaking scale must be $\Lambda_{\text{SUSY}} \sim M_R \sim 100$~TeV. Our specification of scales leads, among other things, to an interesting scenario which involves a low right-handed scale, effects of which could be observed in precision measurements of low energy observables at LHC or NLC \cite{dmm}. To sum up, in the symmetry breaking chains studied here we restrict the scales as follows: \begin{equation} 10^5{\text{GeV}}<M_R<10^7{\text{GeV}} {\text{ and }} 10^{16}{\text{GeV}}<M_G<10^{19}{\text{GeV}} . \label{eq:scales} \end{equation} \noindent We would like to note that this particular choice of scales is in agreement with general astrophysical/cosmological bounds derived from nucleosynthesis \cite{ggr}. In GMSB models it is the messenger fields that carry the information about SUSY breaking to the visible sector. In our case the messenger sector is strongly limited by the constraint that the gauge couplings must meet at scale $M_G$ and that they remain perturbative up to the Planck scale. We will study these constraints and show that among all the possible choices for messenger-field contents and multiplicities only a few models remain which are consistent with all the constraints that we impose. (In view of the fact that we start out with a minimal number of scales and assumptions, this is a somewhat intriguing result and we are inclined to take these phenomenological implications to be very suggestive.) Once the messenger sector is specified, one can calculate all the SUSY breaking parameters of the MSSM and thus make predictions for the complete mass spectrum in the MSSM. As described below, we have fully carried out this program by taking the one-loop radiative effects into account. \section{Left-right models} As has been mentioned above, we assume that the $SO(10)$ symmetry breaking proceeds through one of the intermediate left-right (product) gauge groups $SU(2)_L\times U(1)_{I_{3R}} \times U(1)_{B-L}$ or $SU(2)_L\times SU(2)_R \times U(1)_{B-L}$, followed by a subsequent breaking down to the SM. Let us now briefly describe the superpotentials and the particle contents in theories with these gauge groups. \subsection*{Model~I: $SU(2)_L\times U(1)_{I_{3R}} \times U(1)_{B-L}$} The model based on the gauge group $SU(2)_L\times U(1)_{I_{3R}} \times U(1)_{B-L}$ is phenomenologically interesting because it contains all the usual matter multiplets {\it plus} the right-handed neutrinos. Furthermore, this model forbids lepton- and baryon-number violating terms in the superpotential, thus guaranteeing automatic $R$-parity conservation \cite{mohras}. The model can therefore lead to a stable lightest supersymmetric particle (LSP) which can be a cold dark matter (CDM) candidate. The superpotential for the matter sector of the theory is: \begin{eqnarray} \label{superpotential1} W & = & h_{u} Q H_u u^{c} + h_d Q H_d d^{c} + h_e L H_d e^{c} + h_{\nu} L H_u \nu^{c} \nonumber \\ & & + \mu H_u H_d + f \delta \nu^c \nu^c + M_R \delta {\bar \delta}+ W_m, \end{eqnarray} where $W_m$ denotes the messenger sector superpotential. The lepton and the quark sectors in this theory consists of the doublets $Q(2,0,1/3)$, $L(2,0,-1)$, and the singlets $u^c(1,-1/2, -1/3)$, $d^c(1, 1/2, -1/3)$, $e^c(1, 1/2, 1)$, $\nu^c(1, -1/3, 1)$, and the corresponding squarks and sleptons. The two Higgs doublets (and their superpartners) in this model are the same as in the MSSM: $H_u(2, 1/2, 0)$ and $H_d( 2, -1/2, 0)$. In addition to these, the model contains two $SU(2)_L$ Higgs singlets $\delta (1, 1, -2)$ and ${\bar \delta} (1, -1, 2)$ which break the $U(1)_{I_{3R}} \times U(1)_{B-L}$ symmetry down to the $U(1)_Y$ of the SM, together with their superpartners. The gauge sector of the the model contains the bosons $W(3, 0, 0)$,~$B(1, 0, 0)$ and $V(1, 0, 0)$, and the corresponding gauginos. \subsection*{Model II: $SU(2)_L\times SU(2)_R \times U(1)_{B-L}$} The left-right supersymmetric model based on the gauge group $SU(2)_L\times SU(2)_R \times U(1)_{B-L}$ has been studied intensively over the years \cite{fh}. In addition to the attractive features of Model~I, this model offers solutions to both the strong and weak CP problems, while at the same time preserving $R$-parity. As opposed to the $SU(2)_L \times U(1)_{I_{3R}}$ model, in the $SU(2)_L \times SU(2)_R$ model $R$-parity conservation is not automatic. However, there exist ways to avoid it being broken spontaneously \cite{mohras,amrs}. The superpotential for the matter sector of this theory is: \begin{eqnarray} \label{superpotential2} W & = & {\bf h}_{q}^{(i)} Q_L^T\tau_{2}\Phi_{i} \tau_{2}Q_R + {\bf h}_{l}^{(i)} L_L^T\tau_{2}\Phi_{i} \tau_{2}L_R + i({\bf h}_{LR}L_L^T\tau_{2} \delta_L L_L + {\bf h}_{LR}L_R^{T}\tau_{2} \Delta_R L_R) \nonumber \\ & & + M_{LR}\left[ Tr(\Delta_L {\bar \delta}_L + Tr(\Delta_R {\bar \delta}_R)\right] + \mu_{ij}Tr(\tau_{2}\Phi^{T}_{i} \tau_{2} \Phi_{j})+W_m, \end{eqnarray} where, as before, $W_m$ denotes the messenger sector superpotential. The particle content in this theory is as follows. The lepton and the quark sectors consist of the doublets $Q_L(2,1,1/3)$, ~$L_L(2,1,-1)$,~$ Q_R(1, 2, -1/3)$,~$L_R( 1, 2 ,1)$, and the corresponding superpartners. This model contains bi-doublet Higgs fields $\Phi_u (2, 2, 0)$ and $\Phi_d( 2, 2, 0)$, the triplet Higgs fields $\Delta_L (3,1,-2)$ and $\Delta_R (1,3,-2)$ which break the left-right model to the SM, as well as $\delta_L (3,1,2)$, $\delta_R (0,3,2)$ (which are required for the cancellation of anomalies in the fermionic sector), and their superpartners. The gauge sector consists of the bosons $W_L(3, 1, 0), W_R(1, 3, 0), B(1, 0, 0)$, and $V(1, 1, 0)$ and their associated superpartners. \section{The messenger sector} The simplest messenger sector consists of $N_f$ flavors of chiral superfields $\Phi_i$ and ${\bar \Phi}_i$, $(i=1,\cdots, N_f)$ which transform in the ${\bf r} +{\bf {\bar r}}$ representation of the gauge group. In order to preserve gauge coupling constant unification, one usually requires that the messengers form complete GUT multiplets. In this case the presence of the messenger fields at an intermediate scale does not modify the value of $M_G$. The gauge coupling strength receives an extra contribution from the messengers, at the unification scale $M_G$ which is given by \cite{dineetal}: \begin{eqnarray} \delta \alpha^{-1}_{GUT}=-\frac{N}{2 \pi}\ln\frac{M_G}{S}, \end{eqnarray} with \begin{eqnarray} N=\sum^{N_f}_{i=1} n_i. \end{eqnarray} Here $n_i$ is twice the Dynkin index of the ${\bf r}$ representation of the gauge group and $i$ is the flavor index. By requiring that the gauge interactions remain perturbative all the way up to the GUT scale $M_G$ one obtains: \begin{eqnarray} N \alt 150/\ln \frac{M_G}{S}. \end{eqnarray} On the one hand, the minimal set of messenger fields can be relaxed and augmented further if one requires a theory sufficiently rich to be viable---for instance by requiring a reasonable set of masses for the scalars and gauginos. On the other hand, the number of messenger fields is restricted by the requirement that the MSSM couplings stay perturbative up to the GUT scale and that gauge couplings unify to a single coupling at $M_G$. The requirement of gauge unification does not provide much information on the messenger sector: indeed it is not necessary that all messenger-scale vector-like superfields obtain their masses by coupling primarily to the chiral superfield $X$. The messengers that do not satisfy this condition have little effect on the MSSM masses, but can still participate in gauge coupling unification. The possible messenger fields in Model~I are: \begin{eqnarray} Q_8 = (8, 1, 0, 0) , \nonumber \\ L_3 = (1, 3, 0, 0) , \nonumber \\ \Delta+\overline \Delta = (1, 3 , 0, -2) +conj. , \nonumber \\ \Delta^c +\overline {\Delta^c} = (1, 1 , -1, 2) +conj. , \nonumber \\ H + \overline H = (1, 2, \frac{1}{2}, 0) + conj. , \nonumber \\ Q + \overline Q = (3, 2, 0, \frac{1}{3}) + conj. ,\nonumber \\ U ^c+\overline{U^c} =({\overline 3}, 1, - \frac{1}{2}, -\frac{1}{3}) +conj. , \nonumber\\ D ^c+\overline{D^c}=({\overline 3}, 1 ,\frac{1}{2},- \frac{1}{3}) +conj. , \nonumber\\ L +\overline L = (1, 2, 0, -1) +conj. , \nonumber\\ e^c +\overline{e^c} = (1, 1, \frac {1}{2} , 1) +conj. , \nonumber \\ \nu^c +\overline{\nu^c}= ( 1, 1, -\frac {1}{2}, 1) +conj. , \end{eqnarray} and they transform under $SU(3)_C \times SU(2)_L \times U(1)_{I_{3 R}} \times U(1)_{B-L}$ as specified by the quantum numbers in brackets. Similarly, the possible messenger fields in Model~II consist of \begin{eqnarray} Q_8=(8,1,1,0) , \nonumber \\ Q_3=(1,3,1,0) , \nonumber \\ Q_3^c=(1,1,3,0) , \nonumber \\ \phi=(1,2,2,0) , \nonumber \\ Q +\overline Q=(3,2,1,\frac {1}{3})+conj. , \nonumber \\ Q^c +\overline {Q^c}=(3,1,2,-\frac {1}{3})+conj. , \nonumber \\ L+\overline L=(1,2,1,-1)+conj. , \nonumber \\ L^c+\overline{L^c}=(1,1,2,1)+conj. , \nonumber \\ \Delta +\overline \Delta =(1,3,1,-2) + conj. , \nonumber \\ \Delta^c+\overline {\Delta^c} =(1,1,3,2) + conj., \end{eqnarray} which transform under $SU(3)_C\times SU(2)_L \times SU(2)_R \times U(1)_{B-L}$ with the quantum numbers specified in brackets. In the following section we shall restrict the messenger sectors by requiring (i) the perturbativity of the gauge couplings ($\alpha_k<1$) up to the Planck scale and (ii) unification at the GUT scale $M_G$. \section{The SO(10) solution} Unification is the requirement that the values of the four gauge couplings be equal to a single value $\alpha_G$ at scale $M_G$. We use subindex ``$V$'', as opposed to ``$B-L$'', to denote the GUT-normalized $B-L$ gauge coupling. The relation between the $U(1)_{B-L}$ gauge coupling $\alpha_{B-L}$ (analogous to the hypercharge $\alpha_Y$ of the SM) and the GUT-normalized gauge coupling $\alpha_V$ (analogous to the $\alpha_1$ of the standard model) is $\alpha_V= \frac{2}{3}~\alpha_{B-L}$. At the left-right breaking scale $M_R$ we match couplings to those of the the MSSM. Denoting by $\beta_L$, $\beta_R$, $\beta_{V}$ and $\beta_C$ the beta-functions corresponding to the $SU(2)_L$, $SU(2)_R$ (or $U(1)_{I_{3R}}$), $U(1)_{V}$, and $SU(3)_C$ gauge groups respectively, the one-loop renormalization group equations at the $M_R$ scale are as follows: \begin{eqnarray} \alpha_L^{-1}(M_R) = \alpha_G^{-1} +\beta_L (t_G-t_R) , \nonumber \\ \alpha_R^{-1}(M_R) = \alpha_G^{-1} +\beta_R (t_G-t_R) , \nonumber \\ \alpha_{V}^{-1}(M_R) = \alpha_G^{-1} +\beta_{V} (t_G-t_R) , \nonumber \\ \alpha_C^{-1}(M_R) = \alpha_G^{-1} +\beta_C (t_G-t_R) , \label{eq:rng1} \end{eqnarray} where we have defined: \begin{equation} t_R = \frac 1{2 \pi} \ln \frac{M_R}{M_Z} ~~{\text{ and }}~~ t_G = \frac 1{2 \pi} \ln \frac{M_G}{M_Z}. \end{equation} The one-loop matching conditions to the MSSM at the scale $M_R$ are: \begin{eqnarray} \frac 53 \alpha_1^{-1}(M_R)&=&\alpha_R^{-1}(M_R)+\frac 23 \alpha_{V}^{-1}(M_R) , \nonumber \\ \alpha_2^{-1}(M_R)&=&\alpha_L^{-1}(M_R) , \nonumber \\ \alpha_3^{-1}(M_R)&=&\alpha_C^{-1}(M_R), \label{eq:rng2} \end{eqnarray} where $\alpha_1$, $\alpha_2$ and $\alpha_3$ are the gauge couplings of $U(1)_Y$, $SU(2)_L$ and $SU(3)_C$ respectively. Combining equations (\ref{eq:rng1}) and (\ref{eq:rng2}) yields \begin{equation} \alpha_k^{-1}(M_Z)=\alpha_G^{-1}+\beta_k^{MSSM} t_R+\beta_k^{LR} (t_G-t _R) , k=1,2,3 , \label{eq:rng3} \end{equation} where we have defined the left-right $\beta$-functions by \begin{equation} \beta^{LR}=\left( \begin{array}{c} \beta_1^{LR} \\ \beta_2^{LR} \\ \beta_3^{LR} \end{array} \right) =\left( \begin{array}{c} \frac 35 \beta_R+\frac 25 \beta_{V} \\ \beta_L \\ \beta_C \end{array} \right) . \label{eq:definebeta} \end{equation} and where the MSSM $\beta$-functions $\beta_k^{MSSM}=(33/5,1,-3)^T$ have been used. We can eliminate the unification coupling $\alpha_G$ from equation (\ref{eq:rng3}) by using constraints from equation (\ref{eq:scales}). We thus obtain the following limits on the differences of LR $\beta$-functions: \begin{equation} 3.1 < \beta_2^{LR}-\beta_3^{LR} < 4.1 {\text{ and }} 7.4 < \beta_1^{LR}-\beta_3^{LR} < 9.9 . \label{eq:constrain1} \end{equation} The choice of messenger fields is determined by the requirement that the gauge couplings remain perturbative ($\alpha_k<1$) up to the Planck scale, as well as by the requirement of obtaining plausible masses for all MSSM particles. The requirement of the perturbativity of the gauge couplings at the scale $M_G$ and between $M_G$ and $M_{P}$, together with (\ref{eq:rng3}), leads to the following constraints on the left-right $\beta$-functions: \begin{equation} \beta_1^{LR}<10.4 {\text{ , }} \beta_2^{LR}<6.1 {\text{ and }} \beta_3^{LR}<3.0 . \label{eq:constrain2} \end{equation} These equations constrain the number of messenger fields, since each new chiral superfield contributes to the $\beta$-functions. The requirement that the gauge couplings unify at a specific energy scale may not always impose severe restrictions on the messenger sector, but it certainly provides for reasonable MSSM masses and agrees well with the apparent experimental confirmation for gauge coupling unification at LEP. Using the $\beta$-functions for Models I and II (see Appendix C), together with the constraints (\ref{eq:constrain1}) and (\ref{eq:constrain2}), we find that there can be no consistent solutions in Model II. In Model~I, on the other hand, there exist consistent solutions with the messenger multiplicities: \begin{equation} n_8=n_3=n_H+n_L=1 {\text{ and }} n_{e^c}+n_{\nu^c}=0,1 . \label{eq:messconstrain} \end{equation} We will study these possibilities in more detail and show that despite apparently having several choices, their consequences are remarkably alike for all the models specified by (\ref{eq:messconstrain}). According to (\ref{eq:messconstrain}) the messenger sector consists of one color octet ($n_8=1$) field, one $SU(2)_L$ triplet ($n_3=1$) field, and a pair of $H$- or $L$-type (Higgs- or Lepton-like, respectively) messenger fields ($n_H+n_L=1$). There could also be a pair of $e^c$ or $\nu^c$ type fields ($n_{e^c}+n_{\nu^c}=0,1$). There are thus a total of six choices for the messenger multiplicities. However, each of these messenger sectors produces a very similar mass spectrum for the MSSM, making our scheme extremely predictive. We have listed the spectrum obtained from these models in Table~I (for $\Lambda_{SUSY}=100~TeV, \Lambda_M=100~\Lambda_{SUSY}$) and Table~II (for $\Lambda_{SUSY}=50~TeV,~\Lambda_M=10~\Lambda_{SUSY}$). In both Tables the value of the bilinear scalar coupling ${\tilde B}_{\mu}$ is kept fixed and $\tan \beta$ is allowed to vary. Note that in the second case only one model survives and the stau mass is barely positive. In the first Table the solutions for 5 out of 6 scenarios are obtained. As can be seen from these Tables, both of the representative cases require a large $\tan \beta \approx 35-40$. In every case studied the $SU(2)_L$ and the $SU(3)_C$ gauge couplings meet at $M_G=2.0 \times 10^{16}$GeV (which we take to be our GUT scale) for $10^5{\text{GeV}}<M_R<10^7{\text{GeV}}$. For the choice $n_{e^c}+n_{\nu^c}=0$ the coupling $\alpha_1$ corresponding to the $U(1)_{I_{3R}} \times U(1)_V$ gauge groups and $\alpha_3$ corresponding to the color gauge group meet at $M_G'=1.9 \times 10^{16}$GeV, or within 6\% of the GUT scale. For the choice $n_{e^c}+n_{\nu^c}=1$ the mismatch is much worse: $\alpha_1$ meets the colour gauge coupling at $M_G' \simeq 4-5 \times 10^{15}$GeV. However, this mismatch can be accounted for and explained through the threshold effects and so we will consider this model as well. Finally, in left-right models the ratio: \begin{equation} \tan^2 \theta_R=\frac{\alpha_{B-L}(M_R)}{\alpha_R(M_R)}, \end{equation} is a very important phenomenological parameter which is completely determined by the unification condition. Here the value of $\tan^2 \theta_R$ lies in the interval $1.3 \le \tan^2 \theta_R \le 1.6$. The solutions presented above do not include the fields $\Delta+\overline{\Delta} =(1,3,0,-2) +conj.$ and $\Delta^c+\overline {\Delta^c} =(1,1,-1, 2)+conj.$ at the MSSM scale. These fields are essential for the see-saw mechanism and conservation of R-parity. The contribution of these fields is included in the renormalization-group equations between $M_G$ and $M_R$. Including them in the interval between $M_{MSSM}=10^3$ GeV and $M_R=10^5$ GeV would affect only slightly the $\beta_1$ function, translating in a change of the messenger scale by a factor of 1.5, from $10^5-10^7$ GeV to $1.5 \times 10^5-10^7$ GeV, which is insignificant compared to the accuracy we are working with \cite{chacko}. \section{The sparticle spectrum} Given a messenger sector, and once the messenger scale and $\tan^2 \theta_R$ are fixed, one can calculate the sparticle masses in the corresponding left-right model (see appendix A). The left-right model is then matched to the MSSM (appendix B). Knowing the MSSM parameters at the messenger scale, we can then calculate the full MSSM particle spectrum as a function of $\tan \beta$~\cite{hamidian}. One could reduce the parameters further and solve the supersymmetric $CP$-problem, as was done in~\cite{kaiandotherfriends}, by requiring that the bilinear scalar coupling ${\tilde B}_\mu$ vanish at the messenger scale. This would fix $\tan \beta$. However, in this work we wish to take a broader course by treating $\tan \beta$ (or equivalently the bilinear scalar coupling) as a free parameter. We list representative results for the sparticle spectrum obtained for the six models by choosing $\tan \beta = 15$ in Table~III and $\tan \beta = 2$ in Table~IV. >From Table~\ref{tab:beta1}, where we list the spectrum in terms of $\Lambda_{\text SUSY}$, $\Lambda_M=\lambda S$ (the messenger scale: see appendix A), and ${\tilde B}_\mu$, one can see that the squarks are quite heavy in this model. The heavy top squark drives the radiative symmetry breaking and we find that for all values of the parameters the mass-squared term of the up-type Higgs boson indeed acquires a negative value, thus always leading to the desired (radiatively-induced) symmetry breaking. As the gauge symmetry is always radiatively broken, one must check that the resulting vacuum is a physical one, i.e. that all mass-squared eigenvalues of the charged scalars remain positive and above the current experimental limits. Here, as in other GMSB models, the lightest scalar fermion turns out to be the lighter stau mass eigenstate. For low values of $\Lambda_{\text{SUSY}}$ the lighter stau has a mass below the experimental limit of about $72$~GeV \cite{LEP2}. In Figure~\ref{fig:mass6} we have plotted the masses of the lighter sparticles as a function of scale $\Lambda_{\text{SUSY}}$. One can see that the stau is always light and the lower bound on the stau mass sets a lower limit on the scale of supersymmetry breaking and the sparticle masses. We have plotted this limit in Figure~\ref{fig:limit3} as a function of $\tan \beta$. It is seen that the squark mass scale must always be heavy, of the order of $\sim$~1~TeV. At high values of $\tan \beta$ ($\tan \beta \agt 30$) one would require multi-TeV squarks in order to satisfy the LEP2 exclusion limit on stau mass. Despite the fact that our restrictions allow six different solutions, we find that all of them are somewhat remarkably similar in predictions for the supersymmetric spectrum. The characteristic features of the obtained mass spectrum for the supersymmetric partners (and $H^{\pm}$) are: \begin{itemize} \item Depending on the exact messenger content, the next-to-lightest supersymmetric particle (NLSP) can be either the lighter stau or the lightest neutralino, which is essentially a gaugino. The choice of $n_{e^c}=1$ favours the neutralino as the NLSP, while solutions with $n_{e^c}=0$ favour the stau as the NLSP. One can notice, from eq. (C4) and (C5), that in the latter case, the Bino is the NLSP. If $n_{e^c}=1$, the Bino mass is a triplet. One has approximately the folowing ratios for the gaugino masses ($M_k\propto \alpha_kN_k$ and the 1,2,3 stand for $U(1)_Y,~SU(2)_L$ and $SU(3)_C$ respectively): $$n_{e^c}=0~~~~M_1:M_2:M_3=0.029:0.29:1$$ $$n_{e^c}=1~~~~M_1:M_2:M_3=0.088:0.29:1$$ \item As expected, the lighter selectron is always heavier than the lighter stau. Also, the scalar leptons are always lighter than the squarks, which turn out to be very heavy in this model. From Figure~\ref{fig:limit3}, one can see that the theory always predicts a light stau and, equivalently, heavy squarks ($\agt 1$~TeV for small $\tan \beta$ and $\agt 2$~TeV for large $\tan \beta$). Note that the usual hierarchy $m_{{\tilde e}_{1,2}} \leq m_{{\tilde d}_{1,2}} \approx m_{{\tilde u}_{1,2}} \leq m_{{\tilde t}_{1,2}}$ holds. In this model, $m_{{\tilde l}_{1,2}}$ and $m_{{\tilde q}_{1,2}}$ are mixed states of left and right sleptons or squarks. \item The sneutrinos are always heavier than the lighter of the charged sleptons, unlike in supersymmetric models without GMSB. In fact $m_{{\tilde \nu}_{e,\tau}} \approx m_{{\tilde \tau}_2}$. The charged sleptons are lighter than the Higgs/Higgsinos. (The masses of the right-handed sleptons are around 100~GeV and the masses of the left-handed sleptons around 300 GeV for minimal squark masses.) \item The bilinear Higgs coupling, the so-called $\mu$ parameter in the superpotential, lies in the $400-500$~GeV region and can be {\em either} positive {\em or} negative, as shown in Table~\ref{tab:modellist}, as opposed to the values obtained in \cite{nandi}. As expected in GMSB theories, $\mu^2 > m_{{\tilde l}_{1,2}}^2$. Here $\mu \cong M_{\tilde q}/100$, meaning that $|{\mu}|$ is always at least 50~GeV. This makes the Higgsino at least twice as heavy as the wino, so that the light charginos are mostly gauginos. If we tune the value of the bilinear scalar coupling (the ${\tilde B}_{\mu}$ parameter) to zero at the messenger scale, we obtain a positive value for $\mu$. In general, the sign of $\mu$ does not seem to make much difference to the mass spectrum. However, we obtain sizable effects in, for example, the $b \rightarrow s\gamma$ decay width, since for positive $\mu$ the interference between the SM and chargino contributions is destructive, whereas for negative $\mu$ it is constructive. \item The heavy spartner masses are quite accurately directly proportional to the scale $\Lambda_{\text{SUSY}}= F/S$. We have listed the particle content for all our six models in Table~\ref{tab:modellist}. For the choice of $\tan \beta=15$ and $\Lambda_{\text{SUSY}}=50$ TeV the squark masses are around $1$~TeV, while the charged Higgs boson mass is about $0.5$~TeV in all cases. The heavy Higgs mass is of the order of the $\mu$-term. The heavy sleptons, neutralinos and charginos all have masses between $310$~GeV and $470$~GeV. \item Finally, if one includes a supergravity sector, one would expect a light (mass in the eV range) gravitino. As in other GMSB theories, we assume the gravitino to be the LSP. \end{itemize} \section{Summary and Conclusion} If the low-energy world of the standard model indeed descends from a supersymmetric theory, there could be a plethora of experimental signals at future facilities and the results of theoretical analyses will be needed to distinguish among a large variety of supersymmetric scenarios. Most of the parameters in SUSY are associated with the SUSY breaking sector. Within the framework of perturbative unification, one starts with a SUSY GUT scenario, specifies the scale and mechanism for SUSY breaking, and imposes phenomenological constraints to reduce the number of free parameters (as many as 124 in the MSSM only! \cite{h}). Here we have chosen to study supersymmetric $SO(10)$ GUTs with GMSB as viable SUSY GUT candidates, since these are free from the problems that plague $SU(5)$-based theories, with or without GMSB. By breaking $SO(10)$ to left-right symmetric gauge groups, and by taking advantage of the extreme predictive power of the GMSB mechanism, we calculate and discuss a number of important phenomenological results. As in any other SUSY GUT, with GMSB or otherwise, a number of conditions must be imposed before a detailed study is made. We have kept the number of such (boundary) conditions to a minimum to examine {\em all} such SUSY GUTs with GMSB in a large class of $SO(10)$ theories. The conditions imposed were that the gauge couplings remain perturbative ($\alpha_k<1$) up to the Planck scale, and that they unify at the $GUT$ scale. Except for the left-right symmetry breaking and the messenger scales, no other intermediate scales are assumed, and we take the SUSY and the left-right breaking scales, $\Lambda_{\text{SUSY}}$ and $M_R$ respectively, to be the same. From the condition that sparticle masses be of the order of $\sim 1$~TeV, the SUSY breaking scale must be $\Lambda_{\text{SUSY}} \sim 100$~TeV. Remarkably, these conditions rule out a large number of scenarios, and the ones that survive exhibit surprisingly similar features. The NLSP is either the neutralino (with a large gaugino component) or the lighter stau, with masses of about 50~GeV. This would be testable either at LEP~II through direct production of neutralinos \cite{lnz}, or at the Tevatron/LHC through the production of two lepton jets plus missing energy \cite{nandi}. However, one could see that the chargino mass is above 300 GeV, which is just below the experimental limit in $b \rightarrow s \gamma$ decay. This would make it more likely to observe a signal at the LHC, whereas we expect LEP to be able to observe at most the stau or the lightest Higgs boson. Future collider experiments can put all these predictions fully to test. Finally, let us note that although here we have studied $SO(10)$ GUTs equipped with the GMSB mechanism as the next logical choice beyond similar theories based on the $SU(5)$ gauge group, the supersymmetric spectrum obtained is not unlike the one obtained in gauge-mediated models with a simpler breaking chain. This raises the attractive possibility that gauge-mediated breaking imposes similar features on a variety of GUT scenarios. It would certainly be interesting to study other GUT scenarios with GMSB to test what appear to be general features of this mechanism. \section*{Aknowledgements} We thank Z. Chacko, M. Raidal, A. Riotto and T. Kobayashi for comments and for raising interesting questions. The work of M.F. was partially funded by NSERC of Canada (SAP0105354). The work of H.H. is supported by the U.S. Department of Energy under Grant Number DE-FG02-91ER40676 and that of K.P. partially by the Academy of Finland (No.~37599).
\section*{Abstract} A formal symplectic structure on $R \times M$ is constructed for the unsteady flow of an incompressible viscous fluid on a three dimensional domain $M$. The evolution equation for the helicity density is expressed via the divergence of the associated Liouville vector field that generates symplectic dilation. For an inviscid fluid this equation reduces to a conservation law. As an application the symplectic dilation is used to generate Hamiltonian automorphisms of the symplectic structure which are then related to the symmetries of the velocity field. \newpage The helicity which is first discovered in \cite{skr66} has been recognized to be an important ingradient of the problem of relationship between invariants of fluid motion and the topological structure of the vorticity field \cite{FRI}-\cite{ark92}. For three-dimensional flows its ergodic and topological interpretations were introduced and investigated in \cite{mof69}-\cite{gik92}. It has also been studied in the context of Noether theorems \cite{mor61}-\cite{pad98}. Kinematical aspects of helicity invariants in connection with the particle relabelling symmetries were discussed in \cite{hg99}. In this work, we shall show that there is also a dynamical content of the helicity density in the sense that the information contained in the Eulerian dynamical equations can be represented in the framework of symplectic geometry by a current vector field governing the dynamics of helicity. More precisely, starting from the Navier-Stokes equations of incompressible fluids we shall construct helicity four-vector whose divergence will define the time-evolution of helicity density. The dynamical properties of the fluid, such as viscosity, are implicit in this vector field. The evolution equation for the helicity density reduces to a conservation law for inviscid Euler flows. For fluid dynamical content of this work we shall refer to \cite{FRI} and the necessary mathematical background can be found in \cite{via}-\cite{LM}. The Navier-Stokes equation for a viscous incompressible fluid in a bounded domain $M \subset R^3$ is \begin{equation} {\partial {\bf v} \over \partial t} + {\bf v} \cdot \nabla {\bf v} = - \nabla p +\nu \nabla^2 {\bf v} \label{euler} \end{equation} where ${\bf v}$ is the divergence-free velocity field tangent to the boundary of $M$, $p$ is the pressure per unit density and $\nu$ is the kinematic viscosity. The identity ${\bf v} \cdot \nabla {\bf v} = {1 \over 2} \nabla | {\bf v}|^{2} - {\bf v} \times (\nabla \times {\bf v} )$ can be used to bring the equation (\ref{euler}) into the form \begin{equation} {\partial {\bf v} \over \partial t} - {\bf v} \times (\nabla \times {\bf v}) = \nu \nabla^2 {\bf v} - \nabla (p + {1 \over 2} v^{2} ) \label{reuler} \end{equation} and in terms of the vorticity field ${\bf w} \equiv \nabla \times {\bf v}$ this gives \begin{equation} {\partial {\bf w} \over \partial t} - \nabla \times ( {\bf v} \times {\bf w} ) = \nu \nabla^2 {\bf w} \;. \label{weq} \end{equation} For a fluid with a potential $\varphi$ and velocity field ${\bf v}$ the densities \begin{equation} {\cal H} = {1 \over 2} {\bf v} \cdot \nabla \times {\bf v} \;,\;\;\; {\cal H}_w = {1 \over 2} {\bf w} \cdot \nabla \times {\bf w} \;,\;\;\; q= {\bf w} \cdot \nabla \varphi =w(\varphi) \end{equation} will be called helicity, vortical helicity and potential vorticity, respectively. \begin{theorem} For a velocity field satisfying Eqs.(\ref{euler}) and for $q \neq 2 \nu {\cal H}_w$ the two-form \begin{equation} \Omega_{\nu} = - (\nabla \varphi + {\bf v} \times {\bf w} - \nu \nabla \times {\bf w} ) \cdot d{\bf x} \wedge dt + {\bf w} \cdot (d{\bf x} \wedge d{\bf x}) \label{sympd} \end{equation} is symplectic on $I \times M$ where $I$ is an open interval in $R$. Moreover, it is exact, $\Omega_{\nu} = d \theta$ with the Liouville (or canonical) one-form \begin{equation} \theta = - ( \varphi + p + {1 \over 2} v^{2} ) \, dt + {\bf v} \cdot d{\bf x} \label{cone2} \end{equation} which is independent of the viscosity $\nu$. \end{theorem} {\bf Proof:} $\Omega_{\nu}$ is closed by Eq. (\ref{weq}) and the divergence-free property of the vorticity field. The non-degeneracy follows from the recognition that the density in the symplectic volume $\Omega_{\nu} \wedge \Omega_{\nu} /2$ is the function $q-2\nu {\cal H}_w$ which is assumed to be non-zero. The exactness can be verified using Eq. (\ref{reuler}). $\bullet$ For an arbitrary smooth function $f$ of $(t,x)$ the unique Hamiltonian vector field $X_f$ defined by the symplectic two-form (\ref{sympd}) via $i(X_f)(\Omega_{\nu})= -df$ is given by \begin{equation} X_{f}={1 \over q-2 \nu {\cal H}_{w} } [ - w(f) ({\partial \over \partial t}+v) +{df \over dt} w + (( \nabla \varphi -\nu \nabla \times {\bf w}) \times \nabla f) \cdot \nabla ] \;. \end{equation} Here, $d/dt$ denotes the convective derivative $\partial_t+ {\bf v} \cdot \nabla$ which, viewed as a vector field on $I \times M$, is not Hamiltonian. In fact, with the notation $v \equiv {\bf v} \cdot \nabla$, one can check that the one-form \begin{equation} i(\partial_t+v)(\Omega_{\nu})=(\nabla \varphi - \nu \nabla \times {\bf w}) \cdot (d {\bf x} - {\bf v} dt) \end{equation} is not closed and hence $\partial_t+v$ is not even locally Hamiltonian. Next proposition describes invariantly the connection between the symplectic structure (\ref{sympd}) and the the helicity density. \begin{theorem} \label{ee} The identity \begin{equation} d (\theta \wedge \Omega_{\nu} ) - \Omega_{\nu} \wedge \Omega_{\nu} \equiv 0 \label{helc} \end{equation} gives the equation \begin{equation} { \partial {\cal H} \over \partial t} + \nabla \cdot ( {\cal H} {\bf v} + {1 \over 2} ( p - {1 \over 2} {\bf v}^2 ) {\bf w} ) = {\nu \over 2} {\bf v} \cdot \nabla^2 {\bf w} - \nu {\cal H}_w \label{hel} \end{equation} for the evolution of helicity density. \end{theorem} {\bf Proof:} We have $\Omega_{\nu} \wedge \Omega_{\nu} = -2(q-2\nu {\cal H}_w) dx \wedge dy \wedge dz \wedge dt$ and we compute \begin{equation} \nabla \cdot [ \varphi {\bf w} + {\bf v} \times ( \nabla \varphi - \nu \nabla \times {\bf w}) ] = 2q - 2 \nu {\cal H}_w - \nu {\bf v} \cdot \nabla^2 {\bf w} \end{equation} for the derivative of certain terms in the expression \begin{eqnarray} & & \theta_{\nu} \wedge \Omega_{\nu} = 2{\cal H} \, dx \wedge dy \wedge dz - \nonumber \\ & & \;\;\;\; [ ( \varphi +p - {1 \over 2} v^2 ) {\bf w} + 2{\cal H} {\bf v} + {\bf v} \times ( \nabla \varphi - \nu \nabla \times {\bf w}) ] \cdot d{\bf x} \wedge d{\bf x} \wedge dt \;\;\; \end{eqnarray} for the three-form. Putting them together in the identity (\ref{helc}) we obtain Eq. (\ref{hel}). Upon integration, the term $\nu {\bf v} \cdot \nabla^2 {\bf w} /2$ in Eq. (\ref{hel}) gives the integral of $- \nu {\cal H}_w$ and one obtains the usual expression for the time change of total helicity as given in, for example, Ref. \cite{FRI}. $\bullet$ Note that the helicity flux in Eq. (\ref{hel}) is independent of the function $\varphi$ which we have introduced by hand to make the symplectic form non-degenerate. Using the invariant description (\ref{helc}) of the evolution of helicity density, we shall introduce a current vector $J_{\nu}$ and show that it is an infinitesimal symplectic dilation of $\Omega_{\nu}$. $J_{\nu}$ will be defined as the one-dimensional kernel of the three-form $\theta \wedge \Omega_{\nu}$. Since the symplectic two-form is nondegenerate, it can be obtained as the unique solution of \begin{equation} i(J_{\nu})(\Omega_{\nu} \wedge \Omega_{\nu} /2)= \theta \wedge \Omega_{\nu} \;, \label{defj} \end{equation} that is, as the dual of the three-form $\theta \wedge \Omega_{\nu}$ with respect to the symplectic volume. We find \begin{equation} J_{\nu} = {1 \over q-2 \nu {\cal H}_{w} } [ 2{\cal H} (\partial_t + v) + ( \varphi +p - {1 \over 2} v^2 ) w + {\bf v} \times ( \nabla \varphi - \nu \nabla \times {\bf w}) \cdot \nabla ] \end{equation} as the expression for the helicity current. \begin{theorem} $J_{\nu}$ is a vector field of divergence $2$ with respect to the symplectic volume and it is an infinitesimal symplectic dilation for $\Omega_{\nu}$. The evolution of helicity density ${\cal H}$ can be described by the identity \begin{equation} div_{\Omega_{\nu}}(J_{\nu}) -2 \equiv 0 \;. \label{divj} \end{equation} \end{theorem} {\bf Proof:} The exterior derivative of Eq. (\ref{defj}) gives \begin{eqnarray} d i(J_{\nu})(\Omega_{\nu} \wedge \Omega_{\nu} /2)&=& {\cal L}_{J_{\nu}}( \Omega_{\nu}\wedge \Omega_{\nu}/2) \equiv div_{\Omega_{\nu}}(J_{\nu}) \, \Omega_{\nu}\wedge \Omega_{\nu}/2 \\ &=& d (\theta \wedge \Omega_{\nu} ) \; = \; \Omega_{\nu}\wedge \Omega_{\nu} \end{eqnarray} where we used the identity ${\cal L}_{J} = i(J) \circ d + d \circ i(J)$ in the first equality and the second equality is the definition of the divergence. We see that $J_{\nu}$ is a vector field whose divergence is $2$. From the last equality, we conclude that the equation (\ref{divj}) is equivalent to Eq. (\ref{hel}) describing the evolution of helicity density. $J_{\nu}$ is the unique vector field satisfying \begin{equation} i(J_{\nu})(\Omega_{\nu}) = \theta \label{teta} \end{equation} and it follows from this that $J_{\nu}$ fulfills the condition \begin{equation} {\cal L}_{J_{\nu}} (\Omega_{\nu}) = d i(J_{\nu}) (\Omega_{\nu}) = d \theta = \Omega_{\nu} \label{ome} \end{equation} of being an infinitesimal symplectic dilation for $\Omega_{\nu}$ \cite{alan}. $J_{\nu}$ is also called to be the Liouville vector field of $\Omega_{\nu}$ \cite{LM}. $\bullet$ We observed that the local existence of a Hamiltonian function for $\partial_t+v$ is being prevented by the viscosity term \cite{hg97}. Moreover, the viscosity term causes the helicity not to be conserved. We shall now show that, for the case of inviscid incompressible fluids described by the Euler equation, namely Eq. (\ref{euler}) with $\nu=0$, $\partial_t+v$ is Hamiltonian and that the helicity density ${\cal H}$ is conserved. To this end, we assume that the scalar field $\varphi$ is advected by the fluid motion \begin{equation} {\partial \varphi \over \partial t} + {\bf v} \cdot \nabla \varphi =0 \label{he} \end{equation} and that the potential vorticity $ q \neq 0$. \begin{theorem} \cite{hg97} Let $v$ and $\varphi$ satisfy Eq. (\ref{euler}) with $\nu=0$ and Eq.(\ref{he}), respectively. Then, the suspended velocity field $\partial_{t}+v$ on $I \times M$ and $q^{-1}w$ are Hamiltonian vector fields for the exact symplectic two-form \begin{equation} \Omega_{0} = - (\nabla \varphi + {\bf v} \times {\bf w}) \cdot d{\bf x} \wedge dt + {\bf w} \cdot (d{\bf x} \wedge d{\bf x}) = d \theta \label{sympo} \end{equation} with the Hamiltonian functions $\varphi$ and $t$, respectively. The evolution equation (\ref{hel}) reduces to the conservation law in divergence form for the helicity density. \end{theorem} {\bf Proof:} Using Eq. (\ref{he}) $\partial_{t}+v$ can be written in Hamiltonian form $i(\partial_t+v)(\Omega_0)= - d\varphi$. More generally, the Hamiltonian vector field with the symplectic two-form (\ref{sympo}) for an arbitrary function $f$ on $I \times M$ is given by \begin{equation} X_{f}={1 \over q } [ - w(f) ({\partial \over \partial t}+v) +{df \over dt} w + (\nabla \varphi \times \nabla f) \cdot \nabla ] \end{equation} which clearly reduces to $\partial_t+v$ for $f=\varphi$ and to $q^{-1}w$ for $f=t$. The conservation of helicity density is obvious. $\bullet$ For the inviscid flow of the Euler equation the helicity current takes the form \begin{equation} J_{0} ={1 \over q} [ 2{\cal H} (\partial_t + v) + ( \varphi +p - {1 \over 2} v^2 ) w + {\bf v} \times \nabla \varphi \cdot \nabla ] \end{equation} while the canonical one-form remains to be the same. That means, the difference between the dynamics of fluid motion with $\nu =0$ and $\nu \neq 0$ is contained in the helicity current. Thus, the dynamical content of the helicity is encoded in its current and this, in turn, is connected with the symplectic structure on $I \times M$ which was constructed as a consequence of the Eulerian dynamical equations. The realization of dynamics of fluid motion in the symplectic framework is useful in the study of the geometry of the motion on $M$ and of the hypersurfaces in $I \times M$ defined by the time-dependent Lagrangian invariants, that is, the invariants of the velocity field. The present framework also provides geometric tools for the investigation of scaling properties of the fluid motion because the action by the Lie derivative of helicity current on tensorial objects corresponds to infinitesimal scaling transformations \cite{LM}. Leaving the discussions of these issues elsewhere, we shall conclude this work with an application to the symmetry structure of the velocity field which is also related to the results presented in \cite{hg98}. \begin{theorem} Let $X_f$ be a Hamiltonian vector field for $\Omega_{\nu}$. Then, the vector fields $({\cal L}_{J_{\nu}})^k(X_f), \; k=0,1,2,...$ are infinitesimal Hamiltonian automorphisms of $\Omega_{\nu}$. \end{theorem} {\bf Proof:} The symplectic two-form is invariant under the flows of Hamiltonian vector fields because ${\cal L}_{X_f}(\Omega_{\nu}) = d i(X_f)(\Omega_{\nu}) = d^2f \equiv 0$ where we used the identity ${\cal L}_{X}=i(X) \circ d+d \circ i(X)$ for the Lie derivative, $d \Omega_{\nu} =0$ and the Hamilton's equations $i(X_f)(\Omega_{\nu}) = -df$. It then follows from the identity \begin{equation} {\cal L}_{[ J_{\nu},X_f ]}= {\cal L}_{J_{\nu}} \circ {\cal L}_{X_f} - {\cal L}_{X_f} \circ {\cal L}_{J_{\nu}} \label{iden} \end{equation} evaluated on $\Omega_{\nu}$ that $[ J_{\nu},X_f ]$ also leaves $\Omega_{\nu}$ invariant. Replacing $X_f$ with $[ J_{\nu},X_f ]$ in Eq. (\ref{iden}) we see that one can generate an infinite hierarchy of invariants of the symplectic two-form $\Omega_{\nu}$. To see that these are Hamiltonian vector fields we compute \begin{eqnarray} i([ J_{\nu},X_f ])(\Omega_{\nu}) &=& {\cal L}_{J_{\nu}} ( i(X_f) (\Omega_{\nu})) - i(X_f) ( {\cal L}_{J_{\nu}} (\Omega_{\nu})) \label{jx} \\ &=& - d ( J_{\nu}(f) -f) \end{eqnarray} where we used Eq. (\ref{ome}). Thus, $[ J_{\nu},X_f ]$ is Hamiltonian with the function $J_{\nu}(f) -f$. By induction one can find similarly that $({\cal L}_{J_{\nu}})^2(X_f)$ is Hamiltonian with $(J_{\nu})^2(f)-2J_{\nu}(f)+f$ and so on. Interchanging $J_{\nu}$ and $X_f$ in the identity (\ref{jx}) we also obtain $i(X_f)(\theta) = J_{\nu}(f) $. $\bullet$ In particular, we let $\nu=0$, $f=t$ so that $X_t=q^{-1}w$ and consider the infinitesimal Hamiltonian automorphisms $({\cal L}_{J_{0}})^k(q^{-1}w), \; k=0,1,2,...$ of $\Omega_0$. The identity (\ref{iden}) evaluated on the vector field $\partial_t+v$ gives \begin{equation} {\cal L}_{[ J_{0},q^{-1}w ]}(\partial_t+v)= - {\cal L}_{q^{-1}w} ([ J_{0},\partial_t+v ]) \label{sym} \end{equation} where the vector field $[ J_{0},\partial_t+v ]$ is, by proposition (5), Hamiltonian with the function $J_{0}(\varphi) -\varphi = p-v^2/2$. By the Lie algebra isomorphism $[X_f,X_g]= X_{ \{ f,g \} }$ defined by the symplectic structure $\Omega_0$, the right hand side of Eq. (\ref{sym}) is a Hamiltonian vector field with the function \begin{equation} \{ t,p-{1 \over 2}v^2 \} = {1 \over q} w(p-{1 \over 2}v^2) \;. \label{ff} \end{equation} On the level surfaces defined by the constant values of the function (\ref{ff}) we have $[[ J_{0},q^{-1}w ], \partial_t+v]= 0$ In fact, if we restrict to the constant values of the function $p-v^2/2$ the hierarchy of Hamiltonian automorphisms of $\Omega_0$ can be identified as the infinitesimal symmetries of the velocity field. This can be seen by replacing $q^{-1}w $ with $[ J_{0},q^{-1}w ]$ in Eq. (\ref{sym}). We thus proved that \begin{theorem} For the Euler flow, the hierarchy of infinitesimal Hamiltonian automorphisms $({\cal L}_{J_{0}})^k(q^{-1}w), \; k=0,1,2,...$ of $\Omega_0$ generate infinitesimal time-dependent symmetries of the velocity field on the level surfaces $p-v^2/2 = constant$. \end{theorem} As a matter of fact, the function $p-v^2/2$ is related, in Ref. \cite{pam}, to the invariance under particle relabelling symmetries of the Lagrangian density of the variational formulation of the Euler equation.
\section{Introduction} Studying Quantum Chromodynamics (QCD) at finite baryon density does not require a special motivation: this is, after all, the traditional subject of nuclear physics. Nevertheless, the investigation of cold quark matter lay dormant for some time and was only revived recently when it was realized that a number of exciting new phenomena can be predicted with some certainty. The first is that we expect the high density phase of QCD to be a color superconductor, with sizeable gaps on the order of 100~MeV around the phase transition~\cite{ARW_98,RSSV_98}. The structure of this phase depends sensitively on the number of active flavors. For three or more flavors, color and flavor quantum numbers are locked and chiral symmetry is broken even at large chemical potential $\mu$~\cite{ARW_98b,SW_98b}. In addition to that, it was argued that the second order chiral phase transition at finite temperature $T$ and zero density is likely to turn first order at some critical density. This entails the existence of a tricritical point in the $\mu-T$ phase diagram, which would persist even if the light quarks are not massless~\cite{BR_98,HJS*_98,SRS_98}. The idea that asymptotic freedom and the presence of a sharp Fermi surface imply that high density QCD should be a color superconductor goes back to the work of Frautschi, Barrois, Bailin and Love~\cite{Frau_78,Bar_77,BL_84}. Color superconductivity has many features of the standard model, such as dynamical gauge symmetry breaking and the Higgs phenomenon. It is different from electroweak symmetry breaking in the standard model in the sense that the Higgs is composite. And it is different from models with compositeness (such as technicolor) in that it does not require strong interactions. Color superconductivity takes place even in weak coupling. This, of course, is a consequence of the BCS instability. Detailed numerical calculations of color superconducting gaps were carried out by Bailin and Love, who concluded that one-gluon exchange (OGE) induces gaps on the order of 1~MeV at several times nuclear matter density (at asymptotically large chemical potentials, however, magnetic gluon exchanges generate increasingly large gaps~\cite{Son_98}). The main new feature pointed out in \cite{ARW_98,RSSV_98} is that instanton-induced interactions can lead to substantially larger gaps, on the order of 100~MeV. Furthermore, it was realized that the phase structure of QCD at finite baryon density is very rich. Besides the dominant order parameter for the superconducting phase transition, which is a scalar-isoscalar color antitriplet diquark operator, many other forms are possible. Previous work mostly concentrated on two or three massless flavors and was based on the mean-field approximation (MFA). In the present work, we go beyond these approximations in several important respects. In the case of two flavors we replace schematic zero range interactions with the full momentum dependent instanton-induced interaction~\cite{CD_98,CD_99,Vel_98}. We study the three-flavor case and show that the ground state exhibits color-flavor locking. We show that chiral symmetry is broken, calculate the chiral condensate, and also assess the effects of a finite strange quark mass. In order to go beyond the mean field approximation we study the role of instanton-antiinstanton clusters. In particular, we consider the competition between random instantons and clusters employing a statistical mechanics treatment of the partition function for the instanton liquid. Finally, we consider more speculative possibilities such as phases with diquark Bose condensation, and give a general discussion of the phase structure of QCD with different quark masses. Throughout the article we assume that instanton-induced effects are the predominant source of strong non-perturbative interactions in cold quark matter at small and moderate densities. This assumption is based on the success of the instanton model at zero temperature and zero chemical potential, as well as at $\mu=0$ and $T\neq 0$, see \cite{SS_98} for a review. The other major type of interaction that has been widely used is the (perturbative) OGE. While the latter should be prevailing at high densities, it encounters conceptual difficulties at low and moderate densities since the involved momentum transfers at the Fermi surface $q^2\le p_F^2$, and the running coupling constant of QCD might not be sufficiently small as required for a perturbative treatment. On the other hand, the Debye-screening of electric fields suppresses instanton effects at large densities; however, this suppression seems not to be effective below the chiral phase transition, as has been explicitely demonstrated in finite-$T$ lattice studies~\cite{CS_95,Alles} Further evidence for the importance of instantons in this context is related to the existence of fermion zero modes in the spectrum of the Dirac operator, which arise as a consequence of the axial anomaly: In a topologically non-trivial background field with topological charge +1 there is -- for each flavor -- a left-handed state that emerges out of the Dirac sea, and a right-handed one that moves from positive to negative energy. As a result, the axial charge is violated by 2$N_f$ units. For $N_f=1$ this immediately implies chiral condensation. For more than one flavor, chiral condensation is a collective effect. The quark condensate is determined by the number density of (almost) zero modes of the Dirac operator. These anomalously small eigenmodes can originate from the interaction of exact zero modes associated with isolated instantons and antiinstantons. The wave function of the condensate is the collective state built from instanton and antiinstanton zero modes. To check this mechanism for chiral symmetry breaking on the lattice has lately attracted appreciable attention: the results indeed support the suggested picture~\cite{IN_97,TFM_97}. At finite baryon chemical potential the axial anomaly is connected with fermion zero modes in exactly the same way as in vacuum. The only difference is that the zero modes now correspond to extra states appearing at the Fermi surface, rather than the surface of the Dirac sea. As in the vacuum the effect of the zero modes can be represented as an effective $(2N_f)$-quark interaction that operates near the Fermi surface. For two flavors, this interaction directly leads to the BCS instability, the formation of Cooper pairs and the appearance of a gap. For three and more flavors the instanton vertex does not directly support a Cooper pair; some of the chiral condensates have to be non-zero to close off external quark legs, reducing the $N_f$-body instanton interaction to a two-body one. Instanton-antiinstanton molecules, on the other hand, lead to an effective four quark operator for any number of flavors, which, if attractive, will trigger the formation of a gap. Nevertheless, direct instantons play an important role even for $N_f\ge 3$. In particular, instantons provide a novel mechanism for chiral condensation: A diquark-driven $\bar q q$ (chiral) condensate. The investigation of chiral symmetry restoration and color superconductivity at finite density should also be placed in a broader context, {\it e.g.}, including finite temperatures. Moreover, we would like to understand the phase structure as a function of parameters that we cannot control in the real world, such as the number of flavors and their masses. After all, the underlying mechanisms for the various transitions and the role of non-perturbative effects (such as instantons) in the different phases have to be clarified. At high temperature we expect to find a quark-gluon plasma phase in which chiral symmetry is restored, {\it i.e.}, the density of (almost) zero modes has to vanish. This can be realized if the instanton liquid changes from a random ensemble of instantons and antiinstantons to a correlated system with finite clusters, {\it e.g.}, instanton-antiinstanton ($I$-$A$) molecules. The formation of molecules and other correlated clusters was observed in numerical simulations of the instanton liquid~\cite{SS_96}, where a number of consequences of this scenario were explored. On the lattice the disappearance of the quasi-zero modes in the vicinity of $T_c$ is well established, and the formation of clusters has been observed~\cite{DeForcrand}. Nevertheless, many details of the transition remain to be understood. In the case of many flavors the instanton calculations \cite{SV_97} suggest a chirally restored vacuum state already for a fairly small number of flavors, around 5. Again, the transition is associated with the formation of correlated clusters. In this article we would like to understand the interplay of the three major phases that have been considered: (i) the hadronic (H) phase, with (strongly) broken chiral symmetry (ii) the color superconductor (CSC) phase, with broken color symmetry, and (iii) the quark-gluon plasma (QGP) phase. All three phases are associated with three specific instanton-induced interactions. Chiral symmetry breaking is caused by the strong $\bar qq$ attraction. The binding energy of the lightest baryon, the nucleon, is mostly associated with the $qq$ interaction. The same interaction is responsible for superconductivity at large baryon density. Finally, quark exchanges between instantons and antiinstantons drive their pairing, which is expected to become the predominant feature in the instanton liquid as temperature increases (at any chemical potential). The structure of our paper is as follows. The first part, comprised of sects. II-VII contains a mean-field analysis of color superconductivity in finite-density QCD with 2 and 3 flavors. We begin with a brief introduction to the structure of the effective instanton-induced interaction in sect.~\ref{sec_L_eff}. In sect.~\ref{sec_diq}, we study the physical effects of this interaction in different diquark channels at $\mu=0$. In sect.~\ref{sec_int} we discuss the modifications of the instanton-induced interactions at non-zero chemical potential. The interplay of chiral symmetry breaking and quark superconductivity in two-flavor QCD is studied in sect.~\ref{sec_nf2}. This section employs the mean-field approximation but is based on the exact form of the interaction. Using a simplified version of the form factors we then consider different $\langle\bar qq\rangle$ and $\langle qq\rangle$ condensates of increasing complexity: the three-flavor problem in the chiral limit (sect.~\ref{sec_nf3}) and the effect of flavor symmetry breaking due to a finite strange quark mass (sect.~\ref{sec_fsb}). The second part of the article (sect.~\ref{sec_cocktail} and \ref{sec_phdia}) addresses effects due to clustering, which are beyond the mean-field approximation. In sect.~\ref{sec_cocktail} we quantitatively discuss only one type of cluster, the instanton-antiinstanton molecule, which we believe is the most important cluster in the chirally restored phase. In sect.~\ref{sec_phdia}, we also discuss the role of correlations between quarks -- in particular non-condensed diquarks and (the most obvious cluster of all!) nucleons in nuclear matter --, and comment on possible experimental consequences for heavy-ion reactions and neutron stars. We summarize and conclude in sect.~\ref{sec_concl}. \section{Effective instanton-induced interactions in vacuum} \label{sec_L_eff} \subsection{Single-Instanton Interactions} \label{oneinst} Our starting point is the euclidean QCD partition function \begin{equation} {\cal Z}=\int{\cal D}\psi{\cal D} \psi^\dagger {\cal D}A \exp(-{S_{QCD}})=\int {\cal D}A \det(\not\!\!{D}) \exp(-{S_{gauge}}) \ . \end{equation} The main assumption of the instanton model is that the gauge field is saturated by classical (anti-)instanton solutions. If the instanton ensemble is sufficiently dilute, the gauge field can be approximated by a sum of individual instanton gauge potentials \begin{equation} A=\sum_{k \in I,\bar I} A_k \ . \label{sumansatz} \end{equation} Collective effects related to chiral symmetry breaking are generated through the low-momentum part of the fermion determinant. In particular, we will concentrate on the fermion determinant in a basis spanned by the zero modes of the individual instantons. Matrix elements of the Dirac operator in this basis are given by the overlap integrals \begin{eqnarray} T_{IA}(z,u)=\int d^4x \ \phi_{I}^\dagger(x-z_I) \ \not\!\!{D} \ \phi_{A}(x-z_A) \ . \label{Tiavac} \end{eqnarray} Here, $z_I$ and $z_A$ denote the positions of the instanton and antiinstanton, and $\phi_{I,A}$ the corresponding zero mode wave functions, which are solutions of the Dirac equation \begin{eqnarray} \not\!\!{D}_{I,A} \ \phi_{I,A}(x)=0 \ , \label{dirac} \end{eqnarray} where the covariant derivative $\not\!\!\!{D}_{I,A}$ includes the gauge potential of the (anti-) instanton $I$ ($A$). Using the Dirac equation (\ref{dirac}) and the sum ansatz (\ref{sumansatz}), Eq.~(\ref{Tiavac}) can be simplified by replacing the covariant derivative by an ordinary one. The overlap matrix element can also be viewed as the quark ``hopping'' amplitude from an instanton to an antiinstanton. To extract effective $2N_f$-quark interaction vertices, we follow the approach of Diakonov and Petrov~\cite{DP}, who suggested to reintroduce free fermion fields according to \begin{eqnarray} {\cal Z}=\int d\psi d \psi ^\dagger {\exp \{\int d^4 x \psi ^\dagger i \not\!\partial \psi\} \over N_+!N_-!}\prod_{I=1}^{N_+} \theta_+ \prod_{\bar I=1}^{N_-} \theta_- \ , \end{eqnarray} where in the two-flavor case \begin{eqnarray} \theta_+=\int d\Omega_I \prod_{f=1}^{2}\big[\int d^4x \psi^ \dagger_f(x) i\not\!\partial \phi_I (x-z_I) \int d^4y\phi_I^\dagger(y-z_I)i\not\!\partial \psi_f (y)\big] \ , \end{eqnarray} and the integrals are over the collective coordinates $\Omega_I=\{z_I,\rho_I,u_I\}$ (position, size and color orientation) of the instantons. The original zero mode determinant can be recovered by calculating a Green's function with $N_f (N_++N_-)$ external legs. In order to perform the integration over the centers of the instantons it is convenient to proceed to momentum space. This automatically induces a four-momentum conserving $\delta$-functions at each vertex. An effective interaction is most easily derived by exponentiating the fermion terms. This is accomplished by applying an inverse Laplace transformation which gives the following partition function: \begin{eqnarray} {\cal Z} &=& const \int d\psi d \psi ^\dagger d\beta_+d\beta_- \exp\left\{ -(N_++1)\log\left(\frac{\beta_+}{c_{\rho}}\right) - (N_-+1)\log\left(\frac{\beta_-}{c_{\rho}}\right) \right. \nonumber \\ & & \hspace{1cm}\left. + \int d^4 x\, (\psi^\dagger i \not\!\partial \psi\ + \beta_+\theta_++\beta_-\theta_- )\right\}. \end{eqnarray} The integrations over $\beta_\pm$ can be performed by the saddle point method, which becomes exact in the thermodynamic limit as the coefficients in the exponent are extensive quantities ($N_\pm=n_\pm V_4$). For an equal number of instantons and antiinstantons, one may consider $g= \beta_+=\beta_-$ as an effective fermion coupling. $\beta_\pm$ are then eliminated through the final minimization of the free energy, leaving the total instanton density $N/V=n_++n_-$ as the physical parameter. In the remainder of this section we restrict ourselves to two flavors. In this case, four quarks participate at each vertex, and the pertinent vertex operator $\theta_\pm$ takes the form \begin{eqnarray} {\cal O}_{\theta_+}=\prod_{f=1}^2 d\Omega_I (\Omega_I \chi_L)\otimes (\chi_R^\dagger \Omega_I^\dagger) \ . \end{eqnarray} Its non-locality can be expressed through a momentum-dependent formfactor ${\cal F}(k)$, which is also a matrix in the Dirac space, attached to each fermion field. We will analyze the formfactors, including their dependence on density, in sect.~\ref{sec_formfactors}. After color-averaging, one obtains the effective interaction lagrangian \begin{eqnarray} \label{l_nf2} {\cal L} &=& g\frac{1}{4(N_c^2-1)} \Big\{ \frac{2N_c-1}{2N_c}\left[ (\bar\psi {\cal F}^\dagger\tau_\alpha^{-}{\cal F}\psi)^2 + (\bar\psi {\cal F}^\dagger\gamma_5\tau_\alpha^{-}{\cal F}\psi)^2 \right]\cr & & \hspace{0.5cm} + \frac{1}{4N_c}(\bar\psi{\cal F}^\dagger \sigma_{\mu\nu} \tau_\alpha^{-}{\cal F}\psi)^2 \Big\} \ , \end{eqnarray} where $N_c$ is the number of colors and $\tau^-=(\vec\tau,i)$ is an isospin matrix. In the pseudoscalar channel the interaction combines attraction for the isospin-1 (pion) channel with repulsion (due to the extra $i$) for isospin-0 ($\eta'$). Similarly, one finds attraction in the scalar isospin-0 ($\sigma$) channel (responsible for spontaneous chiral symmetry breaking) together with repulsion in the scalar isospin-1 channel ($a_0$). In practice we will calculate correlation functions and the mean field effective potential in the Hartree-Fock approximation. For this purpose, it is convenient to construct an effective $s$-channel kernel including the exchange term. This is made possible by the simple (separable) form of the momentum dependence. Using this kernel, one can reproduce the result of a Hartee-Fock calculation by evaluating the Hartree term only. In short-hand notation we will refer to the kernel as the effective meson or diquark lagrangian. From the Fierz identities given in appendix \ref{fierz}, we obtain the following kernel for color singlet and octet $\bar qq$ states \begin{eqnarray} \label{l_mes} {\cal L}_{mes} &=& {g \over 8 N_c^2} \Big\{ \left[(\bar\psi{\cal F}^\dagger \tau^- {\cal F}\psi)^2+ (\bar\psi{\cal F}^\dagger \tau^- \gamma_5 {\cal F}\psi)^2 \right] \nonumber \\ & & \mbox{} + {N_c-2\over 2(N_c^2-1)} \left[(\bar\psi{\cal F}^\dagger \tau^- \lambda^a {\cal F}\psi)^2 +(\bar\psi{\cal F}^\dagger \tau^- \lambda^a \gamma_5 {\cal F}\psi)^2 \right]\cr & & \mbox{} -{N_c \over 4 (N_c^2-1)}(\bar\psi{\cal F}^\dagger \tau^- \sigma_{\mu \nu} \lambda^a {\cal F}\psi)^2 \Big\} \ , \end{eqnarray} again being attractive in the $\sigma$ and $\pi$ channel, repulsive in the $\eta'$ and $a_0$ channel. Analogously, we can construct the effective interaction for color-antisymmetric ${\bf \bar 3}$ and -symmetric ${\bf 6}$ diquarks. The result is \begin{eqnarray} \label{l_diq} {\cal L}_{diq} &=& {g\over 8 N_c^2} \left\{ -{1\over N_c-1} \left[ (\psi^T{\cal F}^T C \tau_2 \lambda_A^a {\cal F}\psi) (\bar\psi{\cal F}^\dagger\tau_2 \lambda_A^a C {\cal F}^*\bar\psi^T \right. \right.\nonumber\\ & & \left.\left. \hspace{1.5cm}\mbox{} +(\psi^T{\cal F}^T C \tau_2 \lambda_A^a \gamma_5 {\cal F}\psi) (\bar\psi{\cal F}^\dagger \tau_2 \lambda_A^a \gamma_5 C {\cal F}^* \bar\psi^T) \right]\right. \nonumber\\ & & \left. \mbox{} +{1\over 2(N_c+1)} (\psi^T{\cal F}^T C \tau_2 \lambda_S^a \sigma_{\mu \nu} {\cal F}\psi) (\bar\psi{\cal F}^\dagger \tau_2 \lambda_S^a \sigma_{\mu \nu} C {\cal F}^*\bar\psi^T) \right\} \ , \end{eqnarray} where $\tau_2$ is the antisymmetric Pauli matrix, and $\lambda_{A,S}$ are the antisymmetric (color ${\bf\bar3}$) and symmetric (color {\bf 6}) color generators (normalized in an unconventional way, ${\rm tr}(\lambda^a \lambda^b)=N_c\delta^{ab}$, in order to facilitate the comparison between mesons and diquarks). In the color ${\bf\bar3}$ channel, the interaction is attractive for scalar $(\psi^TC\gamma_5\psi)$ diquarks, and repulsive for pseudoscalar $(\psi^TC\psi)$ diquarks. As we have already discussed in our previous paper~\cite{RSSV_98}, in the case of two-color ($N_c$=2) QCD there exists an additional Pauli-G{\"u}rsey symmetry (PGSY)~\cite{PG,DFL_96} which mixes quarks with antiquarks. It also manifests itself in the lagrangians given above, as in this case the coupling constants in $\bar qq$ and $qq$ channels are identical, {\it i.e.}, diquarks (the baryons of the $N_c$=2-theory) are degenerate with the corresponding mesons. Chiral symmetry breaking then implies that scalar diquarks are also Goldstone bosons, with their mass vanishing in the chiral limit ({\it i.e.}, for current quark masses $m=0$). \subsection{$I$-$A$-Molecule Induced Interactions} \label{molint} Using the 't Hooft interaction introduced in the last section one can calculate correlation functions in a systematic expansion in multi-instanton interactions, starting from direct instantons graphs and proceeding to two-instanton or instanton-antiinstanton graphs, as well as more complicated clusters. In simple cases, like the set of RPA diagrams discussed in sect.~\ref{sec_diq}, one can sum a whole series of terms involving infinitely many instantons. But if the instanton ensemble is strongly correlated, this method may become very inefficient. In that case it is more useful to determine the effective vertex for a given cluster, and fix the strength of the vertex by calculating the concentration of clusters from the partition function. The simplest kind of cluster that can arise in the instanton ensemble are instanton- antiinstanton molecules. We have observed the formation of these clusters at high temperature and at large $N_f$ in both analytic~\cite{IS_94,VS_97} and numerical simulations of the instanton ensemble~\cite{SS_96}. In these cases, molecules are intimately connected with chiral symmetry restoration. An ensemble of molecules does not have delocalized zero modes or collective eigenstates, and the chiral condensate is zero. In the high density problem, the role of molecules is twofold. First, the concentration of instanton-antiinstanton molecules in the ensemble may be dynamically enhanced for similar reasons as in the case of high temperature or large number of flavors. We will discuss this problem in detail in sect.~\ref{sec_cocktail}. Second, the BCS instability is due to quark-quark scattering, or four-fermion operators. The 't Hooft vertex is a $(2N_f)$-fermion operator and does not automatically lead to an instability for $N_f>2$. However, an instanton-antiinstanton molecule can always generate an effective four-fermion interaction, with the additional $(2N_f-4)$ fermion lines being internal. The effective four-fermion vertex induced by instanton-antiinstanton molecules was evaluated in \cite{SSV_95}. The result is particularly simple if the relative color orientation is fixed such that the instanton-antiinstanton interaction is most attractive. In that case one has \begin{eqnarray} {\cal L}_{IA} &=& G_{mol} \left\{ \frac{1}{N_c^2}\left[ (\bar\psi\gamma_\mu\psi)^2+ (\bar\psi\gamma_\mu\gamma_5\psi)^2 \right] -\frac{1}{2N_c(N_c-1)} \left[ (\bar\psi\gamma_\mu\lambda^a\psi)^2+ (\bar\psi\gamma_\mu\gamma_5\lambda^a\psi)^2 \right] \right. \nonumber \\ & & \left.\hspace{0.4cm} -\frac{1}{N_c^2} \left[ (\bar\psi\gamma_\mu\psi)^2- (\bar\psi\gamma_\mu\gamma_5\psi)^2\right] -\frac{2N_c-1}{2N_c(N_c^2-1)} \left[ (\bar\psi\gamma_\mu\lambda^a\psi)^2- (\bar\psi\gamma_\mu\gamma_5\lambda^a\psi)^2\right] \right\} \ . \label{l_mesmol} \end{eqnarray} Similar to the procedure leading to Eq.~(\ref{l_diq}), this interaction can be rearranged into an effective $qq$ vertex. In the color antitriplet channel the result is \begin{eqnarray} {\cal L}_{IA}^{ \underline 3} &=& G_{mol} \left\{ \frac{1}{N_c(N_c-1)}\left[ (\psi^T C\gamma_5 \tau_2 \lambda_A^a \psi) (\bar\psi\gamma_5\tau_2 \lambda_A^a C \bar\psi^T) -(\psi^T C \tau_2 \lambda_A^a \psi) (\bar\psi \tau_2 \lambda_A^a C \bar\psi^T) \right] \right. \nonumber\\ & & \left. \mbox{} + \frac{1}{4N_c(N_c-1)}\left[ (\psi^T C\gamma_\mu\gamma_5 \tau_2\lambda_A^a \psi) (\bar\psi\gamma_\mu\gamma_5\tau_2 \lambda_A^a C \bar\psi^T) -(\psi^T C \gamma_\mu\tau_2\vec\tau \lambda_A^a \psi) (\bar\psi \gamma_\mu\tau_2\vec\tau \lambda_A^a C \bar\psi^T) \right] \right\} \ . \end{eqnarray} Even in the case $N_f=2$, there are two important differences as compared to the single-instanton vertex. First, since molecules are topologically neutral, the interaction is $U(1)_A$ invariant. This implies that it does not distinguish between scalar and pseudoscalar diquarks. Second, whereas the 't Hooft vertex only operates in scalar (and tensor) channels, molecules also provide an interaction in vector-meson and diquark channels. The coupling constant is related to the density of molecules and has to be determined from the partition function of the instanton liquid. We will study this problem in sect.~\ref{sec_coupl}. \section{Diquarks in the Random Phase Approximation} \label{sec_diq} Color superconductivity implies that the high density phase is composed of diquark Cooper pairs. In a weakly coupled BCS system, the expression 'Cooper pair' should not be taken too literally: it is not tightly bound and the range of its wave function is large compared to average inter-particle separations ({\it i.e.}, the cube root of the inverse particle density, $d=n^{-3/2}$). In QCD this is not necessarily the case. The gap can be quite large, and the existence of an intermediate phase of diquarks with or without Bose condensation is not a priori excluded. For this reason we first study the possibility of diquark bound states in vacuum~\cite{BL_89,SSV_94,DFL_96}. In QCD with $N_c>2$ there are, of course, no gauge invariant diquark states. Instead, one can study correlation functions of heavy-light $Qqq$ states, where the heavy quark $Q$ serves to neutralize color. Effectively, this corresponds to a diquark correlator in the presence of a Wilson line. The ``gauge invariant'' diquark masses extracted from these correlators then measure the mass of the heavy $Qqq$ state minus the mass of the heavy quark. In a dense medium, the Wilson line is not necessary, and color is neutralized by light quarks of the third color. In the effective fermionic theory derived in the previous section, diquarks can appear as physical bound states. These states should be interpreted as building blocks in the formation of baryons and dense matter. In practice, we study diquark correlation functions and look for poles in the diquark propagators at momenta $|p|< 2M$, where $M$ is the constituent quark mass. For simplicity, instead of the exact instanton formfactors ${\cal F}(p)$, we employ in this section an euclidean $O(4)$ symmetric cutoff $\Lambda$, being adjusted to yield a realistic constituent quark mass $M$. \begin{figure}[tbp] \begin{center} \psfig{file=diq01.eps,angle=0,height=6. cm} \end{center} \vspace{0.2cm} \caption{The quantity $KJ$ entering the denominators of the $T$ matrix for the scalar (S), pseudoscalar (P) and tensor (T) diquark channels.} \label{poles} \end{figure} The Bethe-Salpeter equation for the two-body ${\cal T}$ matrix in a given channel can be written as \begin{eqnarray} {\cal T}(q)&=&\sum_i K_i(C^{-1}{\cal O}^i) \times \left\{ ({\cal O}^i C) +i \rm tr \int {d^4p\over (2 \pi)^4}S_F(p+q/2)({\cal O}^i C)S_F^T(-p+q/2){\cal T}(q) \right\} \ . \label{BS} \end{eqnarray} Introducing the notation \begin{eqnarray} {\cal T}(q)=\sum_{k,k'}(C^{-1}{\cal O}^k)T_{k,k'}(q) ({\cal O}^{k'}C) \ , \end{eqnarray} Eq.~(\ref{BS}) can be expressed schematically as \begin{eqnarray} T=K(1+JT) \ . \end{eqnarray} This equation has the solution \begin{eqnarray} T=(1-KJ)^{-1}K. \end{eqnarray} In both diquark and mesonic channels $J$ takes the form \begin{eqnarray} J_{k,k'}(p)=i \rm tr \int {d^4p\over (2 \pi)^4}S_F(p+q/2){\cal O}^k S_F(p-q/2){\cal O}^{k'} \ , \end{eqnarray} since $C S_F^T(p)C^{-1}=S_F(-p)$. From Eq.~(\ref{l_diq}) one has the three different operator structures, ${\cal O}= \tau_2 \lambda_A^a i \gamma_5$ (scalar channel), ${\cal O}= \tau_2 \lambda_A^a$ (pseudoscalar channel) and ${\cal O}= \tau_2 \lambda_S^a \sigma_{\mu \nu}$ (tensor channel), which lead to \begin{eqnarray} J_{SS}&=& -2 I_1(M)+2 q^2 I_2(q^2,M) \nonumber\\ J_{PP}&=& 2I_1(M)+2(4 M^2-q^2)I_2(q^2,M) \\ J_{TT}&=&-16 M^2 I_2(q^2,M) \nonumber \end{eqnarray} with the two standard integrals \begin{eqnarray} I_1(M)&=&8N_c\int_0^\Lambda {d^4p\over (2 \pi)^4}{1\over p^2+M^2} \\ I_2(-q^2,M)&=&4N_c\int_0^\Lambda {d^4p\over (2 \pi)^4}{1\over (p+1/2q)^2+M^2}{1\over (p-1/2q)^2+M^2} \ . \end{eqnarray} They are readily evaluated for Euclidean momenta and analytically continued to Minkowski space to yield \begin{eqnarray} I_1(M)&=&{N_c\over 2 \pi^2}\left[ \Lambda^2 +M^2 \ln {M^2\over \Lambda^2 +M^2}\right], \\ I_2(q^2,M)&=&{N_c\over 4 \pi^2}\left[\ln {M^2\over \Lambda^2 +M^2} +2\sqrt{{4 M^2 -q^2\over q^2}}\arctan \sqrt{{q^2\over 4 M^2 -q^2}}\right.\cr & &\left. -2 \left( 1-{2\Lambda^2 \over4(\Lambda^2+M^2)-q^2}\right) \sqrt{{4(\Lambda^2+M^2)-q^2\over q^2}} \arctan \sqrt{{q^2\over 4(\Lambda^2+M^2) -q^2}}\right]. \end{eqnarray} The chiral gap equation in the scalar $\bar q q$ channel then becomes \begin{eqnarray} {g\over 8 N_c^2}I_1(M)=1 \ , \end{eqnarray} which provides a relation between the coupling $g$ and the constituent quark mass $M$. The conditions for the existence of poles in the corresponding channels of the diquark ${\cal T}$ matrix are \begin{eqnarray} -{g\over 8 N_c^2(N_c-1)}(-2 I_1(M)+2 q^2 I_2(q^2,M))=1 \end{eqnarray} for the scalar channel, \begin{eqnarray} -{g\over 8 N_c^2(N_c-1)}(2I_1(M)+2(4 M^2-q^2)I_2(q^2,M))=1, \end{eqnarray} for the pseudoscalar one and \begin{eqnarray} {g\over 16 N_c(N_c+1)}(-16 M^2 I_2(q^2,M))=1 \end{eqnarray} for the tensor one. If the {\it l.h.s.}~crosses 1, there is a bound state. This is illustrated in Fig.~\ref{poles} where we have taken $M=350$~MeV, $\Lambda=900$~MeV for definiteness. One finds that only the scalar diquark is bound, with a binding energy of about 200~MeV. This result agrees with numerical simulations of the instanton liquid~\cite{SSV_94}, which included all diagrams in this interaction. However, it is at variance with the conclusion drawn in \cite{DFL_96}, where no bound scalar diquark was found in the same model. We believe that the discrepancy is due to the fact that the authors of \cite{DFL_96} only used part of the interaction in the diquark channel. In this work, we have performed a Hartree-Fock calculation with the full one-instanton interaction. A lattice calculation of diquark masses was performed in \cite{HKLW_98}. These authors find a significant scalar-vector diquark mass splitting, but no scalar diquark bound state. On the other hand, they also obtained a too small $N$-$\Delta$ mass splitting, and a mass ratio $m_N/m_\rho$ that is too large. One can also argue that there should be some continuity when going from the theory with $N_c=2$ to $N_c=3$. In the former case the scalar diquark is the partner of the pion and the vector diquark is the partner of $\rho$. This implies that the scalar diquark binding is large, $M_{dq,V}-m_{dq,S}= m_\rho-m_\pi\simeq 600 MeV$. It is then natural that some remnant of the binding is left at $N_c=3$, since the coupling constant in the scalar diquark channel for $N_c=3$ is only reduced by a factor of 2 as compared to the $N_c=2$ case. This corresponds to a ``$N_c=2+\epsilon$'' picture of the baryon octet: a tightly bound scalar diquark loosely coupled to the third quark. A number of phenomenological observations (reviewed, {\it e.g.}, in~\cite{diquarks}) actually supports the validity of this picture for real QCD. The decuplet baryons, on the other hand, do not contain scalar diquarks, and therefore should be generic 3-body objects. This picture is quite contrary to another (and much better known) view of baryonic structure, the large $N_c$ limit. Here both $N,\Delta$ as well as other members of the octet and decuplet are basically the same heavy object, slowly rotating with slightly different angular momenta. Indeed, as one reads off from the lagrangians given in the previous section, in this limit the diquark coupling tends to zero, and the scalar diquark binding disappears. \section{Instanton-induced interaction in dense matter } \label{sec_int} \subsection{Quark Zero Modes} \label{sec_zeromodes} In the previous section the instanton-induced interactions between quarks were approximated by effective {\em local} 4-fermion vertices. In the microscopic treatment of sect.~\ref{sec_L_eff}, the external quarks couple to the quark zero modes in the instanton field, leading to a {\em nonlocal} profile function for the interaction vertex with a size characterized by the typical instanton radius of about $\rho$=0.33~fm. In dense matter at zero temperature, the single-instanton solution itself is not affected by the surrounding quarks. The zero-mode wave functions, however, {\em are} density dependent leading to important modifications of the instanton-induced interactions in the medium. They can be constructed from the Dirac equation at finite (quark-) chemical potential, \begin{eqnarray} (i\not\!\!D_I-i\mu\gamma_4)\phi_I=0 \ . \end{eqnarray} The correct solution was obtained in \cite{Car_80,Abr_83}: \begin{eqnarray} \phi_I=i\frac{\rho}{2\pi}{e^{\mu\,t}\over x}\sqrt {\rho^2+{x}^2}\not\!\partial \frac{[\cos(\mu r) + {t\over r} \sin(\mu r)] e^{-\mu\,t}}{\rho^2+{x}^2 } \ \chi_L \ , \label{zm} \end{eqnarray} where the spinor $\chi_L$ arises from an antisymmetric (singlet) coupling of spin and color wave functions, as before. Note that the solution of the adjoint Dirac equation, \begin{eqnarray} \phi_I^\dagger(x;-\mu) \ (i\not\!\!D_I-i\mu\gamma_4)=0 \ , \end{eqnarray} carries the chemical potential argument with opposite sign. This is necessary for a consistent definition of expectation values at finite $\mu$, and in particular renders a finite norm, \begin{eqnarray} \int d^4x \ \phi_I^\dagger(x;-\mu) \ \phi_I(x;\mu) =1 \ , \end{eqnarray} whereas without the extra sign one has \begin{eqnarray} \int d^4x \ \phi_I^\dagger(x;\mu) \ \phi_I(x;\mu) =\infty \ . \end{eqnarray} This singularity, corresponding to the particle-particle channel, is in fact directly related to well-known BCS singularity one encounters when resumming an effective (attractive) particle-particle interaction around the Fermi surface (see, {\it e.g.} , ref.~\cite{AGD}). \subsection{Instanton Form Factors} \label{sec_formfactors} Using the explicit form of the zero mode wave function, we calculate the form factor from the Fourier transform \begin{eqnarray} \tilde \phi&=&\int d^4x \phi(x) e^{-i k\cdot x} \nonumber \\ &=& 2i \rho \int_0^{\infty}dR\, R^3 \int_0^{\pi} d\eta \sin^2 \eta \int_0^{\pi} d \theta \sin \theta {e^{\mu\,t}\over x} \sqrt{\rho^2+x^2}(\gamma_0 \partial_t+\vec \gamma \cdot \hat k \cos \theta \partial_r)\nonumber \\ & & \hspace{1cm}{(\cos(\mu r) + {t\over r} \sin(\mu r)) e^{-\mu\,t} \over \rho^2 + x^2} e^{-i(\omega t+ k r \cos \theta)}\chi_L, \end{eqnarray} where $x^2=r^2+t^2$ and $\hat k=\vec k / k$. Introducing hyper-spherical coordinates for the integration, $r=R\sin\eta$, $t=R\cos\eta$, the result can be expressed through two scalar functions $A(\omega,k,\mu)$ and $B(\omega,k,\mu)$, given in appendix~\ref{app_ff}, as \begin{eqnarray} \tilde \phi=[\gamma_0 B(\omega,k,\mu)+\vec\gamma\cdot \hat k A(\omega,k,\mu)] \chi_L\equiv \tilde \psi \chi_L \ , \end{eqnarray} The finite density zero mode wave functions $\tilde \psi$ have the symmetry properties \begin{eqnarray} \tilde \psi(\omega,\vec k,\mu)=\tilde \psi^*(-\omega,-\vec k,\mu)i =\tilde \psi^* (\omega,\vec k,-\mu) \ . \label{sym} \end{eqnarray} The combination ${\cal F}(\omega,\vec k,\mu)= \psi^*(\omega,\vec k,\mu)G_0^{-1}(\omega, \vec k,\mu)$ appears in the effective quark interaction on each propagator entering or exiting the instanton-induced vertex. In the mean field approach, when two of the propagators participating in the vertex have the same momentum, it is useful to combine them into two new formfactors. For a propagator entering the instanton vertex and another exiting with the same momentum one obtains \begin{eqnarray} \alpha={\cal F}(-\omega,-\vec k,\mu)^\dagger {\cal O} {\cal F}(\omega,\vec k,\mu) \ , \end{eqnarray} where $ {\cal O}$ is a matrix with Dirac, color and flavor indices. For an overall unit matrix, one has \begin{eqnarray} \label{alpha} \alpha&=&(\gamma_0 i(\omega-i\mu)+i\vec\gamma\cdot \vec k) (\gamma_0 B^*(\omega,k,\mu)+\vec\gamma\cdot \hat k A^*(\omega,k,\mu))^2\cr & &(\gamma_0 i(\omega-i\mu)+i\vec\gamma\cdot \vec k)\cr &=&({A^*}^2+{B^*}^2)(k^2+(\omega-i\mu)^2)\cr &\equiv & \alpha_r+i \alpha_i \ . \end{eqnarray} For a propagator entering the instanton vertex and a transposed one exiting with the same momentum one finds \begin{eqnarray} \beta={\cal O}{\cal F}(-\omega,-\vec k,\mu)^T{\cal O} {\cal F}(\omega,\vec k,\mu) \ . \end{eqnarray} When the Dirac part of $ {\cal O}$ is $ C \gamma_5$, where $C$ is the charge conjugating matrix, \begin{eqnarray} \label{beta} \beta&=& C \gamma_5 (\gamma_0^T i(-\omega-i\mu)-i\vec\gamma^T\cdot \vec k) (\gamma_0^T B^*(-\omega,k,\mu)-\vec\gamma^T\cdot \hat k A^*(-\omega,k,\mu))\cr & & C \gamma_5(\gamma_0 B^*(\omega,k,\mu)+\vec\gamma\cdot \hat k A^*(\omega,k,\mu)) (\gamma_0 i(\omega-i\mu)+i\vec\gamma\cdot \vec k)\cr &=& (\gamma_0(\omega+i\mu)+\vec\gamma\cdot \vec k) (\gamma_0 B(\omega,k,\mu)+\vec\gamma\cdot \hat k A(\omega,k,\mu))\cr & &(\gamma_0 B^*(\omega,k,\mu)+\vec\gamma\cdot \hat k A^*(\omega,k,\mu)) (\gamma_0 (\omega-i\mu)+\vec\gamma\cdot \vec k)\cr &=& (\omega^2+k^2+\mu^2)(|A|^2+|B|^2)+2\mu k i(A^*B-AB^*)\cr & &+i\gamma_0\vec\gamma\cdot \hat k[2\mu k(|A|^2+|B|^2)+(\omega^2+k^2+\mu^2) i(A^*B-AB^*)]\cr &\equiv & \beta_r+i\gamma_0\vec\gamma\cdot\hat k\beta_i \ . \end{eqnarray} We note that $\alpha$ and $\beta$ have the same symmetry as in Eq.~(\ref{sym}). \begin{figure} \begin{center} \psfig{file=al0.eps,angle=0,height=5 cm} \hspace{0.5cm} \psfig{file=al05.eps,angle=0,height=5 cm} \psfig{file=bet0.eps,angle=0,height=5 cm} \hspace{0.5cm} \psfig{file=bet05.eps,angle=0,height=5 cm} \end{center} \vspace{0.2cm} \caption{The moduli squared of the form factors, $|\alpha|^2=\alpha_r^2+\alpha_i^2$ (top panels) and $|\beta|^2=\beta_r^2-\beta_i^2$ (bottom panels), for two different values of the chemical potential: $\mu=0$ (left column) and $\mu=300$~MeV (right column) .} \end{figure} \subsection{Quark Overlap Matrix Elements} \label{sec_tia} The instanton formfactors discussed above are designed for momentum space calculations, in particular for extracting effective interactions between quarks in the mean-field framework. However, in the statistical mechanics treatment of the instanton liquid partition function presented in sect.~\ref{sec_cocktail}, the coordinate space description is the more suitable one. For that we will need the explicit form of the fermionic overlap matrix element $T_{IA}$ which at finite density takes the form \begin{eqnarray} T_{IA}(z,u;\mu) & = & \ \int d^4x \ \phi_I^\dagger(x-z_I;-\mu) \ (i\not\!\!{D}-i\mu\gamma_4) \ \phi_A(x-z_{A};\mu) \nonumber\\ & = & - \int d^4x \ \phi_I^\dagger(x-z_I;-\mu) \ (i\not\!\partial-i\mu\gamma_4) \ \phi_A(x-z_{A};\mu) \ . \label{tiamu} \end{eqnarray} The second line is again obtained by virtue of the Dirac equation when choosing the sum ansatz for the gauge-field configurations, $A_\mu=A_\mu^I+ A_\mu^A$. $T_{IA}$ plays a crucial role in the fermionic determinant of the instanton partition function, where it generates the fermionic interaction ('quark hopping amplitude') between $I$'s and $A$'s and is therefore responsible for correlations in the instanton liquid. In particular, $T_{IA}$ controls the probability of forming molecules. The definite chirality of the zero modes (in the limit of vanishing current quark masses) entails that $I$-$I$ and $A$-$A$ matrix elements are zero. In the vacuum Lorentz invariance implies that the overlap matrix element can be characterized by a single scalar function~\cite{SV91}, {\it e.g.}, $T_{IA}\equiv i \ u\cdot \hat{z} \ f(z)$. In the medium this is no longer true and $T_{IA}$ must be calculated in terms of two independent scalar functions $f_1,f_2$ according to \begin{eqnarray} T_{IA}(z,u;\mu) \equiv i \ u_4 \ f_1(\tau,r;\mu) + i \ \frac{(\vec u \cdot \vec r)}{r} \ f_2(\tau,r;\mu) \ . \label{tiamu2} \end{eqnarray} They are shown in Fig.~\ref{fig_tia}, see also \cite{Rapp_98,Sch_98}. Similar to the finite temperature case, we observe a strong enhancement with increasing $\mu$ in the temporal direction. Moreover, the fermionic interaction becomes very long range, $\sim \mu^2/x_4$ (at finite temperature it was limited to the Matsubara box of size $1/T$ enforced by periodic boundary conditions). In the spatial direction, the exponential damping $\exp[-\pi T r]$ in the finite-$T$ case is replaced by oscillations $\sim \sin(\mu r)$. The latter also effectively suppress the hopping amplitude once the $r$-integration in the partition function is performed. \begin{figure}[tb] \begin{center} \epsfig{file=f1tau.eps,width=6.5cm,angle=-90} \epsfig{file=f2r.eps,width=6.5cm,angle=-90} \end{center} \caption{Quark-induced $I$-$A$-interaction at finite density for the most attractive color orientation $u_4$=1, $\vec u$=0 as well as $r$=0 (left panel) and for $u_4$=0, $|\vec u|$=1 and $\tau$=0 (right panel). } \label{fig_tia} \end{figure} From the strong enhancement of $T_{IA}$ with increasing $\mu$ one may already anticipate the relevance of molecule configurations at finite densities~\cite{Rapp_98}. This issue will be quantitatively investigated in the 'cocktail' model in sect.~\ref{sec_cocktail}. \section{The two-flavor problem} \label{sec_nf2} This section will be devoted to study finite density two-flavor QCD in the mean-field approximation using the exact momentum dependent instanton profile functions as discussed in sect.~\ref{sec_formfactors}. As density increases the basic competition will be between the chiral condensate $\langle \bar qq\rangle$ and the scalar $ud$ diquark condensate in the $\langle qq\rangle$ channel, representing the color superconducting state as discussed in Refs.~\cite{RSSV_98,ARW_98}. We will first discuss the corresponding coupled gap equations formalism (sect.~\ref{sec_nf2A}) and then proceed to the numerical results and their interpretation in sect.~\ref{sec_nf2B}. \subsection{Mean-Field Grand Canonical Potential at Finite $\mu$} \label{sec_nf2A} For the evaluation of the grand canonical potential we employ the Cornwall-Jackiw-Tomboulis (CJT)~\cite{CJT} effective action, which is elucidated in more detail in Appendix~\ref{app_CJT}. It involves 3 type of propagators (including their antiparticle pendants) corresponding to single (anti-) quarks carrying color charge that participates in the diquark condensate ($G_1$, $\bar G_1$), single (anti-) quarks carrying color charge that is not part of the diquark condensate ($G_2$, $\bar G_2$), and (anti-) diquarks ($F$ $\bar F$). The minimization of the action with respect to ({\it w.r.t.}) these propagators generates the following six gap equations: \begin{eqnarray} -(G_1-F\bar G_1^{-1}\bar F)^{-1}+G_0^{-1}-M_1 \alpha&=&0 \nonumber\\ -(\bar G_1-\bar F G_1^{-1}F)^{-1}+\bar G_0^{-1}+M_1 \alpha^*&=&0 \nonumber\\ -G_2^{-1}+G_0^{-1}-M_2 \alpha^*&=&0 \nonumber\\ -\bar G_2^{-1}+\bar G_0^{-1}+M_2 \alpha^*&=&0 \nonumber\\ \bar G_1^{-1}\bar F(G_1-F\bar G_1^{-1}\bar F)^{-1}+i\Delta \beta&=&0 \nonumber\\ G_1^{-1}F(\bar G_1-\bar F G_1^{-1}F)^{-1}+i\Delta \beta^*&=&0 \ , \label{pre-gap} \end{eqnarray} where \begin{eqnarray} M_1&=&2g\left({1\over 8 N_c^2}(\rm tr(G_1+G_2)\alpha)+{1\over \sqrt{2}}{N_c-2\over 16 N_c^2 (N_c^2-1)}(\rm tr \lambda_8(G_1+G_2)\alpha)\right),\nonumber \\ M_2&=&2g\left({1\over 8 N_c^2}(\rm tr(G_1+G_2)\alpha)-{2\over \sqrt{2}}{N_c-2\over 16 N_c^2 (N_c^2-1)}(\rm tr \lambda_8(G_1+G_2)\alpha)\right),\nonumber \\ \Delta&=&2g{1\over 8 N_c^2 (N_c-1)}\rm tr(FC \gamma_5 \lambda_2 \tau_2 \beta), \label{mass_gap} \end{eqnarray} are the two chiral masses and the diquark gap, and $G_0$ ($\bar G_0$) is the bare (anti-) quark propagator defined through Eq.~(\ref{S_source}). As before, the traces involve momentum integrations. The chiral masses are linear combinations of the $\bar q q$ condensates, $ \rm tr (G_1 \alpha)$ and $ \rm tr (G_2 \alpha)$, while the gap $\Delta$ is proportional to the $qq$ condensate, $\rm tr (FC \gamma_5 \lambda_2 \tau_2\beta)$. Eqs.~(\ref{mass_gap}) represent a coupled system of gap equations in the chiral and diquark masses, $M_1, M_2$ and $\Delta$. To determine their solutions, one needs to know the explicit form of the propagators $G_1, G_2$ and $F$. They are constructed from the coupled set of Eqs.~(\ref{pre-gap}). Using the relation $\bar G(p)=-G^T(-p)$ and the transposition property $C\gamma_5 \gamma_\mu^T= \gamma_\mu C\gamma_5$, one can rearrange Eqs.~(\ref{pre-gap}) into Gorkov-type equations (note that $G,F,\bar F$ do not commute) as \begin{eqnarray} G_2(p)&=&G_0(p)+G(p)M_2\alpha(p) G_0(p)\nonumber \\ G_1(p)&=&G_0(p)+G(p)M_1\alpha(p) G_0(p)+F(p)(i\Delta C \gamma_5 \lambda_2 \tau_2 \beta(p)G_0(p)\nonumber \\ F(p)&=&F(p)M_1\alpha^*(p)G_0^T(-p)+G(p)i\Delta\beta^*(p) C \gamma_5 \lambda_2 \tau_2 G_0^T(-p) \ . \label{Gorkov} \end{eqnarray} A graphic representation of these equations is displayed in Fig.~\ref{fig_Gorkov}. The Gorkov equations can be solved in algebraic form yielding \begin{eqnarray} G_2(p)&=&(G_0^{-1}(p)-M_2\alpha(p))^{-1}\nonumber \\ G_1(p)&=&\left(G_0^{-1}(p)-M_1\alpha(p)+\Delta^2\beta^*(p)(G_0^{-1}(-p)- M_1\alpha^*(p))^{-1}\beta(p) \right)^{-1}\nonumber \\ F(p)&=&i\Delta G_1(p) \beta^*(p)(G_0^{-1}(-p)-M_1\alpha^*(p))^{-1} C \gamma_5 \lambda_2 \tau_2\nonumber \\ &=& i\Delta (G_0^{-1}(p)-M_1\alpha(p))^{-1}\beta^*(p)G_1(-p) C \gamma_5 \lambda_2 \tau_2\nonumber \\ &\equiv & \tilde F(p)C \gamma_5 \lambda_2 \tau_2\nonumber \\ \bar F(p)&=&i\Delta C \gamma_5 \lambda_2 \tau_2 (G_0^{-1}(-p)-M_1\alpha^*(p))^{-1} \beta(p) G_1(p)\nonumber \\ &=&i\Delta C \gamma_5 \lambda_2 \tau_2 G_1(-p) \beta(p)(G_0^{-1}(p)- M_1\alpha(p))^{-1}\nonumber \\ &\equiv & C \gamma_5 \lambda_2 \tau_2 \tilde{\bar F}(p) \ . \label{prop} \end{eqnarray} \begin{figure}[htb] \begin{center} \psfig{file=prop.eps,angle=0,height=6 cm} \end{center} \caption{Diagrammatic representation of the Gorkov Eqs.~(\protect\ref{Gorkov}).} \label{fig_Gorkov} \end{figure} To obtain the thermodynamic state variables such as pressure and energy density we need to know the explicit dependence of the thermodynamic potential on the mass parameters. This can be achieved by reinserting the explicit solutions of the propagators into the effective action, Eq.~(\ref{Gamma}), constituing the grand canonical potential times the 4-volume, $-V_4\Omega(M_1,M_2,\Delta)$. For that purpose we evaluate the propagators more explicitly. After some algebra, one can rewrite $G_1$ from Eq.~(\ref{prop}) as \begin{eqnarray} \label{G_1} G_1(p)&=&1\!\!1_2^{color}\otimes 1\!\!1_2^{flavor}\otimes \left\{ -i\gamma_0 \left[(\omega-i\mu)((\omega+i\mu)^2+k^2+ \tilde M_1^{* 2})\right.\right. \nonumber \\ & & \left. \left. +(\omega+i\mu) \Delta^2 (\beta_r^2+\beta_i^2)-2ik\Delta^2 \beta_r \beta_i \right]-i \vec \gamma \cdot \hat k \right. \nonumber \\ & & \left. \times \left[k((\omega+i\mu)^2+k^2+\tilde M_1^{* 2})+k \Delta^2 (\beta_r^2+\beta_i^2)+2i(\omega+i\mu)\Delta^2 \right] \right. \nonumber \\ & & \left. -\tilde M_1\left[ ((\omega+i\mu)^2+k^2+\tilde M_1^{* 2})+\tilde M_1^*\Delta^2(\beta_r^2- \beta_i^2)\right] \right\} {\cal D}^{-1}, \end{eqnarray} where $\tilde M_1(p)=M_1 \alpha(p)$, and \begin{eqnarray} \label{Denom} {\cal D}&=&|(\omega-i\mu)^2+k^2+\tilde M_1^{2})|^2 + \Delta^4(\beta_r^2- \beta_i^2)^2-8 k\mu \Delta^2 \beta_r \beta_i \nonumber \\ & & +2\Delta^2 (\omega^2+k^2+\mu^2)(\beta_r^2+ \beta_i^2)+2|\tilde M_1|^2\Delta^2(\beta_r^2- \beta_i^2) . \end{eqnarray} Using the relation (\ref{det_trick}) for the Dirac part ($\rm tr_D$) of the trace-log in the kinetic part of Eq.~(\ref{Gamma}) (here the trace does not include the momentum integration), one finds \begin{eqnarray} {1 \over 2}\rm tr_D \ln (-\bar G_1 G_1+\bar G_1 F \bar G_1^{-1}\bar F) =-4\ln {\cal D}, \end{eqnarray} and from Eq.~(\ref{G_1}), \begin{eqnarray} {\rm Re} \left[\rm tr_D(G_0^{-1}G_1-1)\right] &=&-4\left[{\rm Re}[((\omega+i\mu)^2+k^2)\tilde M_1^2]+|\tilde M_1|^4+ \Delta^4(\beta_r^2- \beta_i^2)^2 \right. \nonumber \\ & &\left. -4 k\mu \Delta^2 \beta_r \beta_i +\Delta^2 (\omega^2+k^2+\mu^2)(\beta_r^2+ \beta_i^2)\right. \nonumber \\ & &\left.+2|\tilde M_1|^2\Delta^2(\beta_r^2- \beta_i^2)\right]{\cal D}^{-1}, \\ {\rm Re} \left[\rm tr_D(G_1 \alpha)\right] &=& -{4\over M_1}\left[{\rm Re}[((\omega+i\mu)^2+k^2)\tilde M_1^2]+|\tilde M_1|^4\right. \nonumber \\ & &\left.+|\tilde M_1|^2\Delta^2(\beta_r^2- \beta_i^2)\right]{\cal D}^{-1}, \\ {\rm Re} \left[\rm tr_D(\tilde F \beta)\right] &=& {4\over \Delta}\left[\Delta^4(\beta_r^2- \beta_i^2)^2-4 k\mu \Delta^2 \beta_r \beta_i +\Delta^2 (\omega^2+k^2+\mu^2) \right. \\ & & \left. \times (\beta_r^2+ \beta_i^2) +|\tilde M_1|^2\Delta^2(\beta_r^2- \beta_i^2)\right]{\cal D}^{-1} \ . \end{eqnarray} The analogous quantities involving $G_2$ and $M_2$ are obtained from the above by substituting $\Delta\to 0$, $M_1 \to M_2$. Using the above relations our final expression for $\Omega$ becomes \begin{eqnarray} \label{Omega} \Omega(M_1,M_2,\Delta)&=& \int_0^{\infty} d\omega {k^2 dk \over 2 \pi^3}\biggl\{ -4 \ln {\cal D}-2 \ln|(\omega-i\mu)^2+k^2+\tilde M_2^{2})|^2 \nonumber\\ & & -4{\rm Re}\left[\rm tr_D(G_0^{-1}G_1-1)\right] +8{{\rm Re}[((\omega+i\mu)^2+k^2)\tilde M_2^2]+|\tilde M_2|^4 \over|(\omega-i\mu)^2+k^2+\tilde M_2^{2})|^2} \biggr\} \nonumber\\ & & -{g \over 18}\Big\{\int_0^{\infty} d\omega {k^2 dk \over 2 \pi^3} {\rm Re}\left[ \rm tr_D (2 G_1 \alpha +G_2 \alpha)\right] \Big\}^2 -{g \over 144}\Big\{\int_0^{\infty} d\omega {k^2 dk \over 2 \pi^3} {\rm Re}\left[ \rm tr_D ( G_1 \alpha -G_2 \alpha)\right] \Big\}^2 \nonumber\\ & & -{g \over 6}\Big\{\int_0^{\infty} d\omega {k^2 dk \over 2 \pi^3} {\rm Re}\left[ \rm tr_D (\tilde F \beta)\right] \Big\}^2 \ . \end{eqnarray} The global minimum of $\Omega$ {\it w.r.t.}~$M_1,M_2,\Delta$ at each $\mu$ determines the thermodynamically stable phase and the values of $M_1,M_2,\Delta$ are the chiral masses and color superconducting gap in that phase. The extrema of $\Omega$ at each $\mu$ can be found by equating the derivatives {\it w.r.t.}~$M_1,M_2,\Delta$ to zero. Equivalently, one can verify that after differentiating the expression for $\Omega$ {\it w.r.t.}~$M_1,M_2$ and $\Delta$ one recovers Eqs.~(\ref{mass_gap}). Their solutions correspond to the local extrema of $\Omega$ and represent different possible phases. We shall discuss them in the next subsection. Phase transitions correspond to two distinct minima of $\Omega$ having an equal value (first order), or merging together (second order). One should recall that in this fomulation the coupling constant $g$ is an integration variable: an inverse Laplace transformation was used to exponentiate the instanton vertex. However, in the thermodynamic limit the saddle-point approximation for the $g$-integration becomes exact (since it is multiplied by the 4-volume $V_4$). Identifying the potential energy of $\Omega$ in Eq.~(\ref{Omega}) as $-gU$, the integral in question is \begin{eqnarray} {\cal Z} \propto \int dg \ \exp[-gU+\frac{N}{V} \ \ln(g)] \ , \end{eqnarray} where $N/V$ is the total instanton density. The saddle point is then found to be at \begin{eqnarray} \label{g} g_{max}= \frac{N}{V} \ \frac{1}{U} \ . \end{eqnarray} Thus, at the saddle point the new potential energy is $- N/V \ln(U)$, up to a constant. The real question is how to determine the $\mu$ dependence of $N/V$. This will be addressed in sect.~\ref{sec_cocktail} within a statistical mechanics treatment taking into account correlations in the instanton ensemble. It will be shown there that the simplifying assumption of a constant total instanton density is indeed reasonably justified. Another approximation concerns the density-dependence of the second key parameter, the average instanton radius $\rho$, which defines the scale in all instanton calculations. Lacking better knowledge, we also assume that it does not vary significantly at the chemical potentials under consideration. \subsection{$N_f=2$ Phase Diagram} \label{sec_nf2B} Solutions of the gap equations are extrema of $\Omega$, but only minima represent a thermodynamic phase. There are 4 types of solutions to (\ref{mass_gap}): \begin{enumerate} \item There is a chiral condensate, but no diquark one, {\it i.e.}, $M_1=M_2=M, \Delta=0$, \item There is a diquark condensate, but no chiral one, {\it i.e.}, $M_1=M_2=0, \Delta\neq 0$, \item Both condensates are present, {\it i.e.}, all $M_1,M_2,\Delta\neq 0$, \item No condensates at all, {\it i.e.}, $M_1=M_2=\Delta=0$. \end{enumerate} The last case corresponds to free quarks and it is easy to see that, at $T=0$, having at least one condensate is always more favorable. The rather complicated expressions for the formfactors require the integrals in Eq.~(\ref{mass_gap}), together with Eq.~(\ref{g}), to be evaluated numerically for each value of $\mu$ varying from 0 to 500~MeV. We shall call the three types of solutions described above phase 1, 2 and 3, and fix our two input numbers as $\rho=1/3$~fm and $(N/V)\rho^4=0.0116$, corresponding to $N/V=0.94$~fm$^{-4}$. With these values we find that at $\mu=0$ the system is in phase 1 with the familiar value of about 330~MeV for the chiral mass M at $\mu=0$. We normalize the grand-canonical potential such that the minimum at $\mu=0$ has zero pressure, $\Omega=0$. We find minima for phase 1 in the range of $\mu$ from 0 to 360~MeV, for phase 2 for all values of $\mu$ and for phase 3 only for values of $\mu$ between 250 and 290~MeV. The values of $\Omega$ for all three phases are shown in Fig.~\ref{omega_pict}. \begin{figure}[htb] \epsfig{file=omega.eps,angle=0,width=8cm} \epsfig{file=omega1.eps,angle=0,width=8cm} \caption{$\Omega$ for phases 1, 2 and 3. The right panel is a magnification of the region between 250 and 290~MeV, where phase 3 exists.} \label{omega_pict} \end{figure} Phase 1 dominates until 250~MeV where the system makes a transition to phase 3, and then, at 288~MeV, there is a transition to phase 2. The coupling constant $g$ for all three phases changes little (not shown). The mass $M$ for phase 1, the gap $\Delta$ for phase 2 and the two masses $M_1,M_2$ and the gap $\Delta$ for phase 3 are shown in Fig.~\ref{mgap_pict}. \begin{figure}[htbt] \psfig{file=gap.eps,angle=0,width=8cm} \psfig{file=gap1.eps,angle=0,height=8cm} \caption{The left panel shows $M$ for phase 1 and $\Delta$ for phase 2. The right panel is a magnification of the region between 250 and 290 MeV, with the masses and gaps of all three phases.} \label{mgap_pict} \end{figure} To understand the nature of the phase transitions we study the dependence of $\Omega$ on the three parameters $M_1,M_2$ and $\Delta$. As seen in the right panel of Fig.~\ref{mgap_pict}, phase 3 starts at the values of the masses and the gap of phase 1 at $\mu=250$ MeV. This is an indication of a second order phase transition, as the corresponding chiral and diquark condensates (which are proportional to $M_1,M_2$ and $\Delta$) are the first derivatives of $\Omega$ {\it w.r.t.}~these masses and the gap. The emergence of the phase transition at $\mu=250 MeV$ s exhibited even more clearly in the left panel of Fig.~\ref{hom13}. \begin{figure}[htb] \psfig{file=hom13.eps,angle=0,width=8cm} \psfig{file=hom23.eps,angle=0,width=8cm} \caption{The left panel shows profiles of $\Omega$ at $\mu=270$ and $246$~MeV along the direction that at $\mu=270$~MeV connects the solution for phase 1 ($\xi=0$) and phase 3 ($\xi=1$). The right panel shows $\Omega$ at $\mu=284, 288$ and $292$~MeV along the direction that at $\mu=288$~MeV connects the solution for phase 1 ($\xi=0$) and phase 3 ($\xi=1$).} \label{hom13} \end{figure} At $\mu=270$~MeV, we have plotted the values of $\Omega$ on a straight line in the parameter space $(M_1=M_{1,ph1}+\xi \delta M_1,M_2=M_{2,ph1}+\xi \delta M_2 ,\Delta=\Delta_{ph1}+\xi \delta \Delta)$, connecting phase 1 ($\xi=0$) with phase 3 ($\xi=1$). We see that the solution for phase 1 is not a minimum but a local maximum of $\Omega$. The second curve shows $\Omega$ just before the transition, where phase 3 has not yet emerged, but the minimum is quite flat indicating that the second derivative of $\Omega$ is close to 0. When moving along the same direction $(\delta M_1,\delta M_2 ,\delta \Delta)$ in the parameter space we see a classic example of a second order phase transition, when at certain value of the parameter $\mu$ the symmetry changes (the SU(3) color symmetry is broken to SU(2)), a diquark condensate appears, the second derivative of $\Omega$ goes through 0 and the old solution (with the higher symmetry) turns from a minimum into a maximum. The new absolute minimum is phase 3. Of course any new extrema can only turn up in pairs, but all solutions are symmetric {\it w.r.t.}~the sign of $\Delta$ so that the two new minima must be at $\pm \Delta$. The second phase transition at $\mu=286$~MeV is analyzed in the right panel of Fig.~\ref{hom13}. The middle curve shows the values of $\Omega$ at $\mu=270 MeV$ on a straight line in the parameter space $(M_1=M_{1,ph2}+\xi \delta M_1,M_2=M_{2,ph2}+\xi \delta M_2 ,\Delta=\Delta_{ph2}+\xi \delta \Delta)$, connecting phase 2 ($\xi=0$) with phase 3 ($\xi=1$). At this value of $\mu$ phase 2 already dominates over phase 3. The opposite situation is observed from the upper curve which is for $\mu=284$~MeV with the cross-section of $\Omega$ in the same direction in the parameter space. Obviously a first order phase transition occurs for some $\mu$ between these two values. However, the barrier between the phases is quite low and the values of the parameters $(M_1,M_2,\Delta)$ (and the condensates) of the two phases are quite close, so it is a rather weak first order transition. This is further supported by the fact that at $\mu=290$~MeV the minimum for phase 3 disappears quite rapidly (by going through an inflection point), as seen from the right panel of Fig.~\ref{hom13}. For the lowest curve only phase 2 exists, but the remnant of the inflection point is visible. There are some shortcomings in the mean-field analysis as presented here. Below some critical $\mu_{c}$, which marks the onset transition, no physical quantity should depend on $\mu$. Due to the fact that the instanton zero modes explicitly depend on $\mu$ -- however small -- this is not respected in our calculation. Nevertheless, the variation of $M$ and $\Omega$ below the phase transition is quite small. The result might be further improved by taking into account the dependence of the instanton density and size on $\mu$. Another problem is that the onset transition happens quite early, at $\mu\simeq 250$~MeV, whereas the expected onset corresponds to a third of the nucleon mass minus the binding energy of nuclear matter $(939-16)/3 {\rm MeV} \simeq 308$~MeV. Again, this might be related to various approximations employed. It is interesting to note that we do find all three phases to exist, not just the chirally broken and superconducting phases, but also a phase with chiral symmetry breaking and diquark condensation. The existence of the latter phase is not a very firm prediction as the difference in energy density of the three phases in the transition regions is rather small. In fact, phase 3 was not observed in the NJL calculation of \cite{BR_98} or the instanton calculation of \cite{Rapp_98} or \cite{CD_99}. The latter work uses slightly different techniques to evaluate the grand canonical potential. We will return to the (speculative) phase with simultaneous diquark condensation and chiral symmetry breaking in the discussion in sect.~\ref{sec_more}. \section{Three flavor QCD in the chiral limit} \label{sec_nf3} The situation becomes more involved if one includes the strange quark. Since the critical chemical potential $\mu_c\sim 300-350$ MeV is larger than the strange quark mass $m_s\simeq 140$ MeV, strange quarks have to be included whenever there is time for strangeness to equilibrate. There are several qualitatively new features in going from $N_f=2$ to $N_f=3$. First, since $N_f=N_c$, there are new order parameters in which the color and flavor orientation of the condensate is locked~\cite{ARW_98b}. Second, the instanton-induced interaction is a six-fermion vertex, so it does not directly induce the BCS instability. In this section we will consider $N_f=3$ flavor QCD in the chiral limit. In that case, we expect that in the low density phase chiral symmetry is broken by a quark condensate $\langle \bar uu\rangle = \langle \bar dd\rangle = \langle \bar ss\rangle$. In the high density phase, quark pairs are condensed. One possible form of ordering is the analog of the $N_f=2$ diquark condensate, $\langle q_i^aC\gamma_5 q_j^b\rangle = \Delta^k_c \epsilon_{ijk}\epsilon_{abc}$. This order parameter breaks color $SU(3)_C\to SU(2)_C$ and the chiral $SU(3)_L\times SU(3)_R\to SU(2)_L\times SU(2)_R$. A more attractive possibility is provided by the following order parameter~\cite{ARW_98b} \begin{eqnarray} \label{qq_MFA} \langle q_{i,R}^{a\alpha} q_{j,R}^{b\beta} \rangle &=& \frac{1}{2}(C^\dagger\gamma_5P_R)^{\alpha\beta} \left(\Delta_1 \delta_{ia}\delta_{jb} + \Delta_2 \delta_{ib}\delta_{ja} \right) . \end{eqnarray} Here, $P_R$ is the projector on right-handed quark fields, and there is an analogous expression for left-handed fields also. This order parameters breaks color and chiral symmetry down to the diagonal subgroup $SU(3)_{C+L+R}$. Since the color symmetry is completely broken, there is a gap in the spectrum for all 9 quarks and 8 gluons. This already suggests that the phase characterized by Eq.~(\ref{qq_MFA}) should be preferred over the $N_f=2$ like phase, in which only 4 quark states are gaped. We will see this more explicitly in the next section. The order parameter (\ref{qq_MFA}) breaks chiral symmetry since the residual symmetry couples flavor rotations of right and left handed quarks. The diagonal symmetry acts on the quark fields as \begin{eqnarray} q_i^a \to (U^*)_{ij} U^{ab} q_j^b , \end{eqnarray} where $U$ is an element of $SU(3)$. The most general form of the quark condensate that is consistent with this symmetry is \begin{eqnarray} \label{qbarq_MFA} \langle \bar q_{L,i}^{a\alpha} q_{R,j}^{b\beta} \rangle &=& \frac{1}{2}(P_R)^{\alpha\beta} \left( \left( \Sigma_0- \frac{2}{3} \Sigma_8\right) \delta^{ab}\delta_{ij} + 2\Sigma_8 \delta^a_{\;i} \delta^b_{\;j} \right). \end{eqnarray} At zero density we expect $\Sigma_8$ to be zero, but in the high density phase both $\Sigma_0$ and $\Sigma_8$ will in general be non-zero. It is important to note that even though the above argument establishes that chiral symmetry is broken, it does not show how a chiral condensate is actually formed. From the superfluid order parameter (\ref{qq_MFA}), we can directly form the chiral order parameter $\langle(\bar q_Lq_R)^2\rangle \sim \langle \bar q_L\bar q_L\rangle \langle q_R q_R\rangle$, but not the chiral condensate $\langle\bar q_L q_R\rangle$. This is because (\ref{qq_MFA}) violates right (and left) handed quark number by two units, whereas $\langle\bar q_L q_R\rangle$ violates right and left handed quark number by one unit. In other words, the order parameter leaves a discrete chiral symmetry unbroken, and this symmetry prevents the quark condensate from acquiring an expectation value. But this discrete symmetry is explicitly broken by instantons. In the color-flavor-locked phase, we can saturate four of the external legs of the instanton vertex $(\bar q_L q_R)^3$ with the condensate and obtain an effective interaction $\langle \bar q_L\bar q_L\rangle\langle q_R q_R \rangle (\bar q_L q_R)$ which leads to the formation of a quark condensate. In the case of three massless flavors, the 't Hooft interaction is a flavor antisymmetric six-fermion interaction \cite{SVZ_80b,NVZ_89c,Dia_95} \begin{eqnarray} \label{l_nf3} {\cal L} &=& G_6 (2\pi\rho)^6 \frac{1}{6N_c(N_c^2-1)} \epsilon_{f_1f_2f_3}\epsilon_{g_1g_2g_3} \left( \frac{2N_c+1}{2N_c+4} (\bar\psi_{L,f_1} \psi_{R,g_1}) (\bar\psi_{L,f_2} \psi_{R,g_2}) (\bar\psi_{L,f_3} \psi_{R,g_3}) \right. \\ & & \left. + \frac{3}{8(N_c+2)} (\bar\psi_{L,f_1} \psi_{g_1}) (\bar\psi_{L,f_2} \sigma_{\mu\nu} \psi_{R,g_2}) (\bar\psi_{L,f_3} \sigma_{\mu\nu}\psi_{R,g_3}) + ( L \leftrightarrow R ) \right) \nonumber . \end{eqnarray} In the following, we will consider the somewhat more general case of a $U(1)_A$ violating six-fermion interaction characterized by two independent coupling constants $G_{6,1}$ and $G_{6,2}$, corresponding to the scalar and tensor terms in Eq.~(\ref{l_nf3}). In the vicinity of the Fermi surface, six-fermion terms are suppressed {\it w.r.t.} four-fermion interactions. This is the Cooper phenomenon: Near the Fermi surface, the only interaction that is not suppressed is $2\to 2$ scattering, where the two particles are back-to-back. In the more modern language of the renormalization group one finds that the strength of the six-fermion interaction is reduced as one integrates out states away from the Fermi surface~\cite{Sha_95,EHS,SW_98}. In the context of the mean-field approximation employed here, we will see that the gap equation has a logarithmic enhancement in the case of four-fermion interactions, but not for six- (or even higher) fermion vertices. For this reason we will have to consider the effect of four-fermion interactions. We already stressed that instanton-antiinstanton pairs provide a four-fermion interaction for any number of flavors. In terms of left- and right-handed fermions, the interaction is \begin{eqnarray} \label{l_mol} {\cal L}_{4}&=& G_4 \left\{ \frac{2}{N_c^2}\left[ (\bar\psi_L\gamma_\mu\psi_L)^2+ (\bar\psi_R\gamma_\mu\psi_R)^2 \right] -\frac{1}{N_c(N_c-1)} \left[ (\bar\psi_L\gamma_\mu\lambda^a\psi_L)^2+ (\bar\psi_R\gamma_\mu\lambda^a\psi_R)^2 \right] \right. \nonumber \\ & & \left.\hspace{0.4cm} -\frac{4}{N_c^2} (\bar\psi_L\gamma_\mu\psi_L) (\bar\psi_R\gamma_\mu\psi_R) -\frac{2(2N_c-1)}{N_c(N_c^2-1)} (\bar\psi_L\gamma_\mu\lambda^a\psi_L) (\bar\psi_R\gamma_\mu\lambda^a\psi_R) \right\} \ . \end{eqnarray} In the following, we shall study the condensates (\ref{qq_MFA}) and (\ref{qbarq_MFA}) for an interaction given by the sum of the four-fermion vertex (\ref{l_mol}) and the six-fermion vertex (\ref{l_nf3}). In this section, we will consider the coupling constants $G_4$ and $G_6$ to be arbitrary parameters, constrained mainly by the known value of the quark condensate at zero density. In sect.~\ref{sec_coupl} we shall try to determine these couplings from the partition function of the instanton liquid. The system of gap equations for the three-flavor case can be derived along the same lines as the two-flavor case discussed in the previous section. However, the resulting equations are algebraically much more involved. In order to keep the presentation reasonably simple, we will ignore the instanton form-factors and take the interaction to be point-like. As we saw in the last section in the case of $N_f=2$, this approximation does not qualitatively affect the results. \begin{figure}[htt] \vspace{-2cm} \centerline{\epsfig{file=gap_nf3.ps,width=120mm}} \vspace{-2cm} \caption{Chiral and superconducting gaps $\sigma_{1,2}$ and $\delta_{1,2}$ as a function of the chemical potential for the three-flavor model in the chiral limit.} \label{fig_nf3} \end{figure} We shall calculate the thermodynamic potential in the mean-field approximation. This calculation is somewhat complicated by the fact that the color-flavor structure of the propagators is quite involved. The first step is to determine the quadratic part of the action in the mean-field approximation. For this purpose, we close off all except two legs of the interaction. The result is \begin{eqnarray} \label{mass_nf3} {\cal M} &=& \left( \bar q_{L,i}^a q_{R,j}^b \right) \left\{ \left( \delta^{ab}\delta_{ij} \right) \left[ G_4 \left( \frac{16}{3} \Sigma_0 - \frac{2}{9} \Sigma_8 \right) + G_{6,1} \left( 84 \Sigma_0^2 + 8\Sigma_0\Sigma_8 -\frac{74}{3} \Sigma_8^2 + \frac{9}{2}\Delta_A^2\right) \right. \right. \nonumber \\ & & \hspace{6cm}\mbox{} + G_{6,2} \left.\bigg( 144 \Sigma_0^2 + 48\Sigma_0\Sigma_8 - 168 \Sigma_8^2 - 30 \Delta_A^2\bigg) \right] \nonumber \\ & & \hspace{1.5cm}\mbox{} + \left( \delta^a_{\;i} \delta^b_{\;j} \right) \left[ G_4 \frac{2}{3}\Sigma_8 + G_{6,1}\left( -24 \Sigma_0\Sigma_8 + 10 \Sigma_8^2 - \frac{3}{2} \Delta_A^2 \right) \right. \nonumber \\ & & \hspace{6cm}\mbox{} + G_{6,2}\left.\left. \bigg( -144 \Sigma_0\Sigma_8 + 120 \Sigma_8^2 + 18 \Delta_A^2 \bigg)\right] \right\} \nonumber \\ & & +\left( q_{R,i}^a C\gamma_5 q_{R,j}^b \right) \left\{ \left( \delta^a_{\;i} \delta^b_{\;j} - \delta^a_{\;j} \delta^b_{\;i} \right) \Delta_A \left[ \frac{2}{3}G_4 + G_{6,1} \left( 3 \Sigma_0 - 2 \Sigma_8 \right) + G_{6,2} 12\left( 3\Sigma_0 - 4 \Sigma_8 \right) \right] \right\} \ . \end{eqnarray} We note that the interaction is only sensitive to the antisymmetric part $\Delta_A=\Delta_1-\Delta_2$ of the $\langle qq\rangle$ order parameter. This is different from the OGE interaction considered in \cite{ARW_98b}, but does not make much of a difference in practice, since even in that case the solution of the gap equation has $\Delta_S/\Delta_A\ll 1$, where $\Delta_S=\Delta_1+\Delta_2$. We also note that there is a chiral symmetry breaking $\bar qq$ interaction proportional to $\Delta_A^2$. This is as expected: Color-flavor-locking combined with instantons leads to chiral symmetry breaking. We also stress that both ingredients, instantons and color-flavor-locking, are essential. The $\langle qq\rangle$ and $\langle \bar qq\rangle$ mass terms can be diagonalized simultaneously. In general, the mass term can be decomposed as \begin{eqnarray} \label{mass_def} \left( \bar q_{L,i}^a q_{R,j}^b \right) \left\{ g_0 M_0 + g_1 M_1 \right\} +\left( q_{R,i}^a C\gamma_5 q_{R,j}^b \right) \left\{ f_1 M_1 + f_2 M_2 \right\} \ , \end{eqnarray} where we introduced the color-flavor matrices \begin{eqnarray} \label{mat_def} M_0= \delta^{ab}\delta_{ij}, \hspace{1cm} M_1= \delta^a_{\;i}\delta^b_{\;j}, \hspace{1cm} M_2= \delta^a_{\;j}\delta^b_{\;i} \ . \end{eqnarray} The matrices $M_{1},M_2,M_3$ commute. This means that there is a color-flavor basis in which (\ref{mass_def}) becomes diagonal. We denote the quark fields in this basis by $\phi_\rho$, with $\rho=1,\ldots,9$. The mass term becomes \begin{eqnarray} \label{nf3_diag} \left(\sum_{\rho=1}^{8} \left\{ \sigma_1\left(\bar\phi_{\rho,L}\phi_{\rho,R}\right) + \delta_1\left(\phi_{\rho,R}C\gamma_5\phi_{\rho,R}\right) \right\} \right)+ \left\{ \sigma_2\left(\bar\phi_{9,L}\phi_{9,R}\right) + \delta_2\left(\phi_{9,R}C\gamma_5\phi_{9,R}\right) \right\} \ , \end{eqnarray} where $\sigma_1=g_0,\, \delta_1=-f_2$ is eightfold degenerate and $\sigma_2=g_0+3g_1, \,\delta_2=3f_1+f_2$. In our case \begin{eqnarray} \label{nf3_evals} \sigma_1 &=& G_4\left(\frac{16}{3}\Sigma_0-\frac{2}{9}\Sigma_8\right) +G_{6,1}\left(84\Sigma_0^2+8\Sigma_0\Sigma_8 -\frac{74}{3}\Sigma_8^2+\frac{9}{2}\Delta_A^2\right) \nonumber \\ & & \hspace{5cm}\mbox{} +G_{6,2}\bigg( 144\Sigma_0^2+48\Sigma_0\Sigma_8 - 168 \Sigma_8^2 - 30 \Delta_A^2\bigg) \\ \sigma_2 &=& G_4\left(\frac{16}{3}\Sigma_0+\frac{16}{9}\Sigma_8\right) +G_{6,1}\left(84\Sigma_0^2-64\Sigma_0\Sigma_8 +\frac{16}{3}\Sigma_8^2\right) \nonumber \\ & & \hspace{5cm} \mbox{} +G_{6,2}\left(144\Sigma_0^2- 384\Sigma_0\Sigma_8 + 192\Sigma_8^2 + 24 \Delta_A^2 \right) \\ \delta_1 &=& \frac{1}{2}\delta_2 \;=\; \Delta_A \left( \frac{2}{3}G_4 + G_{6,1} \left(3\Sigma_0-\Sigma_8\right) + G_{6,2} \left(-3\Sigma_0+4\Sigma_8 \right)\right) \ . \end{eqnarray} The potential for the mean field is given by closing of all external legs of the interaction. This way we get two-loop graphs proportional to $G_4$ and three-loop graphs proportional to $G_6$. In the mean-field approximation, we have \begin{eqnarray} \label{V_nf3} V &=& G_4 \left( 2\Delta_A^2 + 12 \Sigma_0^2 + \frac{8}{3} \Sigma_8^2 \right) + G_{6,1} \left( 504 \Sigma_0^3 - 384 \Sigma_0\Sigma_8^2 + \frac{320}{3} \Sigma_8^3 + 24\left(3\Sigma_0 - 2\Sigma_8 \right) \Delta_A^2 \right) \nonumber \\ & & \hspace{2cm}\mbox{} + G_{6,2}\Big( 864 \Sigma_0^3 - 2304 \Sigma_0\Sigma_8^2 + 1280 \Sigma_8^3 - 144\left(3\Sigma_0 - 4\Sigma_8 \right) \Delta_A^2 \Big) \ . \end{eqnarray} In the quadratic part of the interaction we can now integrate over the fermion fields. Since the color-flavor structure is already diagonal, we get a sum of nine terms, each (in principle) with different gap parameters. We finally obtain the following result for the free energy \begin{eqnarray} F &=& -8\epsilon(\sigma_1,\delta_1) -\epsilon(\sigma_2,\delta_2)+V \ . \end{eqnarray} Here, the single particle energy is given by \begin{eqnarray} \epsilon(\sigma,\delta) = \int \frac{d^3p}{(2\pi)^3} \left\{ \sqrt{(E_p-\mu)^2+\delta^2} +\sqrt{(E_p+\mu)^2+\delta^2} \right\} \ , \end{eqnarray} and $E_p^2=p^2+\sigma^2$. The mean-field parameters $\Sigma_0,\Sigma_8, \Delta_A$ are determined by making the free energy stationary $(\partial F)/(\partial \Sigma_i)=(\partial F)/(\partial \Delta_i) = 0$. This gives three coupled gap equations that have to be solved numerically. Before we present the results we have to discuss how to fix the coupling constants $G_4$ and $G_6$. We take $G_{1,2}$ to have the relative size implied by the instanton interaction (\ref{l_nf3}). If we were to ignore random instantons, and only had instanton-antiinstanton pairs, the four-fermion interaction would break chiral symmetry for $G_4>7.5\, {\rm \Lambda}^{-2}$. We consider this to be the upper limit on this interaction. In order to see how large the gaps in the three-flavor case can possibly be, we take $G_4$ just below this limit $G_4=7.4 {\rm \Lambda}^{-2}$. $G_6$ is then fixed by the requirement that for $m_s=150$ MeV (see next section) we get a reasonable constituent $u,d$ mass of 400 MeV. This gives $G_6=12.0 {\rm \Lambda}^{-5}$. Results for the various gaps are shown in Fig.~\ref{fig_nf3}. Note that the superconducting gap is smaller as compared to the two-flavor case. This is because diquark condensation is now due to pairs, not individual instantons, and we restricted the size of the corresponding coupling such that it does not lead to chiral condensation at $\mu=0$. This is similar to the scenario of Alford et al.~\cite{ARW_98b}, where superconductivity for $N_f=3$ is driven by one-gluon exchange. Again, reason dictates that the corresponding coupling is below the critical coupling for chiral condensation. Instantons lead to chiral condensation in the diquark condensed phase. This is immediately clear because if we take the six-fermion vertex and close off four legs by two diquark insertions, the remaining interaction violates chiral symmetry. Color-flavor locking is nevertheless essential. For the two-flavor superfluid order parameter, instantons only lead to a non-zero $\langle \bar ss\rangle$. Alford et al. realized that chiral symmetry would be broken, but could not calculate the size of the effect in their model. Here we find it to be very small. The maximum constituent mass generated in the diquark condensed phase is less than 10 MeV. Qualitatively, this is not hard to understand: the constituent mass arises from terms in (\ref{mass_nf3}) that are proportional to $\Delta_A^2$. These terms arise as exchange terms from the original interaction (\ref{l_nf3}), so they are suppressed by degeneracy factors $2N_f N_c$. In addition to that, the constituent mass is driven by the superconducting gap squared, which is already about an order of magnitude smaller than the zero density chiral gap. There is one more important direct instanton effect in the high density phase. For $N_f=3$ all chirally invariant four-fermion interactions (molecules, OGE, etc.) are $U(1)_A$ invariant, and do not distinguish between scalar and pseudoscalar diquarks. This means that a parity broken vacuum characterized by the order parameter \begin{eqnarray} \label{qq_odd} \langle q_i^a C q_j^b\rangle = \bar\Delta_1 \delta_{ia}\delta_{bj} + \bar\Delta_2 \delta_{ib}\delta_{ja} \end{eqnarray} is degenerate with the parity conserving vacuum considered here. The same is true for an arbitrary linear combination of positive and negative parity condensates. The degeneracy is lifted by the six-fermion interaction in conjunction with finite quark masses or non-vanishing chiral condensates. This implies that the difference in energy density between the parity broken and parity conserving vacua is small. This is different from the $N_f=2$ case, where the four-fermion interaction distinguishes between (\ref{qq_odd}) and (\ref{qq_MFA}), and the energy difference is big. This effect is also different from the scenario considered by Pisarski and Rischke \cite{PR_98}, who argued that the parity broken vacuum is degenerate with the parity conserving one if instanton effects are small. In three-flavor QCD in the chiral limit parity broken and parity conserving vacua are almost degenerate, even if instanton effects are not small. \section{Flavor symmetry breaking} \label{sec_fsb} The situation is even more complicated if we take flavor symmetry breaking into account. For simplicity, we will restrict ourselves to $m_u=m_d=0$ and $m_s\neq 0$. It is clear that as $m_s\to\infty$, we have to recover the two-flavor scenario, with the order parameter given by \begin{eqnarray} \label{del_ud} \langle q_i^aC\gamma_5q_j^b\rangle = \Delta_{ud} \epsilon_{ij3}\epsilon^{ab3} \ . \end{eqnarray} Note that in the two-flavor case the color orientation of the condensate is arbitrary, but for three flavors the choice (\ref{del_ud}) is preferred because it preserves an $SU(2)$ subgroup of the diagonal $SU(3)_{C+L+R}$. We might also consider additional gap parameters that have a different color orientation, but the corresponding gap equation simply decouples and the solution (except in the limit $m_s\to\infty$) is not energetically favored. Since flavor symmetry is broken, the structure of the quark condensate is also more complicated. The following ansatz generalizes Eq.~(\ref{qbarq_MFA}): \begin{eqnarray} \label{qbarq_fsb} \langle \bar q_{L,i}^{a\alpha} q_{R,j}^{b\beta} \rangle = \frac{1}{2}(P_R)^{\alpha\beta} \left( \left( \Sigma_0- \frac{2}{3} \Sigma_8\right) \delta^{ab}\delta_{ij} + \Sigma_s \delta^{ab}\delta_{i3}\delta_{j3} + 2\Sigma_8 \delta^a_{\;i} \delta^b_{\;j} + \Sigma_{8,1} P_1 + \Sigma_{8,2}P_2 \right) \ , \end{eqnarray} where $P_1=\delta^{a3}\delta_{i3}\delta^{b3}\delta_{j3}$ and $P_2=\delta^{a3}\delta_{i3}\delta^b_{\;j}+\delta^a_{\;i} \delta^{b3} \delta_{j3}$. There are a number of complications that occur once flavor symmetry is broken, and it is hard to take into account all of these effects at the same time. In the following we will concentrate on the dynamical interplay between a flavor symmetric four-fermion interaction generated by one-gluon exchange or instanton pairs and the flavor symmetry breaking four-fermion vertex that comes from the six-fermion 't Hooft interaction and a strange mass insertion. In addition to that, we have to take into account that there is no pairing between strange and non-strange quarks if the mismatch between the Fermi momenta is too big. The BCS instability arises for pairs with total momentum zero where both of the individual momenta are on the Fermi surface. This is not possible if the masses are different and the Fermi surfaces are shifted. In the presence of pairing the Fermi surface is not sharp, but smeared out over an energy range given by the gap. This means that pairing between strange and non-strange quarks is suppressed if the mismatch between the Fermi momenta exceeds the gap, \begin{eqnarray} \label{mismatch} m_s^2/(4p_F) >\Delta \ . \end{eqnarray} Moreover, there is the problem that the color-flavor matrix characterizing the most general diquark mass term does not commute with the color-flavor structure of the $\langle \bar qq\rangle$ mass term. This means that we cannot simultaneously diagonalize the two mass terms, and write the free energy in the simple form (\ref{Omega}). Instead it seems unavoidable to deal with the full (spin, color, flavor, and $\langle qq\rangle$ versus $\langle \bar qq\rangle$) matrix structure of the quark propagator. On the other hand, we found that, except possibly in a small regime, quark and diquark condensates do not coexist below the critical chemical potential, and that the quark condensate in the high density phase is small. In the following we will therefore treat the quark condensate in the high density phase as a small perturbation. The important new ingredient if $m_s\neq 0$ is the presence of a four-fermion interaction which operates exclusively in the $u,d$ quark sector. This interaction arises from the 't Hooft interaction, Eq.~(\ref{l_nf3}), by closing off two external legs by a strange mass insertion. The result is \begin{eqnarray} \label{l_nf2_ms} {\cal L} &=& G_6 (m_s\rho)(2\pi\rho)^4 \frac{1}{2N_c(N_c^2-1)} \epsilon_{f_1f_2}\epsilon_{g_1g_2} \left\{ \frac{2N_c-1}{2N_c} (\psi_{L,f_1}^\dagger \psi_{R,g_1}) (\psi_{L,f_2}^\dagger \psi_{R,g_2}) \right. \nonumber\\ & & \hspace{1cm}\left. + \frac{1}{8N_c} (\psi_{L,f_1}^\dagger \sigma_{\mu\nu} \psi_{R,g_1}) (\psi_{L,f_2}^\dagger \sigma_{\mu\nu}\psi_{R,g_2}) + ( L \leftrightarrow R ) \right\} \ , \end{eqnarray} which (of course) has the form of the $N_f=2$ 't Hooft interaction, but with a coupling constant controlled by the parameter $3m_s/(4\pi^2 \rho^2)$. So, unlike in the OGE-based works, the value of the strange quark mass has not just kinematical but also dynamical significance. Our model then consists of a flavor symmetric four-fermion interaction, the flavor symmetry breaking four-fermion interaction (\ref{l_nf2_ms}), and the flavor symmetric six-fermion interaction (\ref{l_nf3}). In order to compare with work of ARW~\cite{ARW_98}, we take the flavor symmetric four-fermion interaction to be one-gluon exchange. We could equally well have used the instanton-antiinstanton induced interaction -- qualitatively this makes very little difference. In this model, the mass term becomes \begin{eqnarray} \label{mass_fsb} {\cal M} &=& \left( \bar q_{L,i}^a q_{R,j}^b \right) \left\{ \left( \delta^{ab}\delta_{ij} \right) \left[ \frac{16}{3}K\Sigma_0 + \left(7 G_{4,1} + 6 G_{4,2}\right)\Sigma_0 + \left(84 G_{6,1} + 144 G_{6,2}\right) \left( \Sigma_0^2+\Sigma_0\Sigma_s \right) \right] \right. \nonumber \\ & & \hspace{1.2cm}\mbox{} + \left. \left( \delta^{ab}\delta_{i3}\delta_{j3} \right) \left[ m_s + \frac{16}{3}K\Sigma_s - \left(7 G_{4,1} + 6 G_{4,2}\right)\Sigma_0 - \left(84 G_{6,1} + 144 G_{6,2}\right) \Sigma_0\Sigma_s \right] \right\} \nonumber \\ & & +\left( q_{R,i}^a C\gamma_5 q_{R,j}^b \right) \left\{ \left( \delta^a_{\;i} \delta^b_{\;j} \right) \left[ \frac{K}{3} \left( 2\Delta_A -\Delta_S\right)\right] + \left(\delta^a_{\;j} \delta^b_{\;i} \right) \left[-\frac{K}{3} \left( 2\Delta_A +\Delta_S\right)\right] \right. \nonumber \\ & & \hspace{2.5cm}\mbox{} + \left. \left( \epsilon^{3ab} \delta_{3ij} \right) \left[ \frac{4}{3}K\Delta_{ud} + \frac{1}{2} \left( G_{4,1} -12 G_{4,2}\right) \left(\Delta_A+2\Delta_{ud}\right) \right] \right\} \ . \end{eqnarray} The quark mass term is already diagonal, while the diquark mass term is of the form \begin{eqnarray} \label{dmass_def} \left( q_{R,i}^a C\gamma_5 q_{R,j}^b \right) \Big\{ f_1 M_1 + f_2 M_2 + f_3 M_3 \Big\} \ \end{eqnarray} with $M_{0,1,2}$ as before and $M_3=\epsilon^{3ab}\epsilon_{3ij}$. This mass matrix has four eigenvalues, \begin{eqnarray} \delta_1 &=& \pm f_2 \\ \delta_2 &=& \pm (f_2-f_3) \\ \delta_{3,4} &=& \frac{1}{2} \left( 3f_1+2f_2+f_3 \pm \sqrt{9f_1^2 + 2f_1f_3 + f_3^2} \right) \ , \end{eqnarray} with degeneracies $d_i=4,3,1,1$. The free energy is given by $F=-\sum_i d_i\epsilon(\sigma_i,\delta_i)+V$ as before, where the potential is \begin{eqnarray} V &=& 16K\Big( 3\Sigma_0^2 + 2 \Sigma_0\Sigma_s + \Sigma_s^2\Big) + 6\Big(7G_{4,1}+6G_{4,2}\Big)\Sigma_0^2 + \Big(504 G_{6,1} + 864 G_{6,2}\Big) \Sigma_0^2 \Big( \Sigma_0+\Sigma_s \Big) \nonumber \\ & & \mbox{} + 4K\Big(3\Delta_A^2-3\Delta_S^2 + 4\Delta_{ud}^2 + 4\Delta_A\Delta_{ud} \Big) + \Big(G_{4,1}-12G_{4,2}\Big) \Big(\Delta_A+2\Delta_{ud}\Big)^2 \ . \end{eqnarray} An important difference as compared to the flavor symmetric case is that we have to take into account the kinematic restriction discussed above. In channels that involve pairing between strange and non-strange quarks we restrict the integration to the regime $(E_p-\mu)^2>(m_s^2/(4p_F)-\delta^2)$. For a more detailed discussion, we refer the reader to \cite{SW_99} and \cite{ABR_99}. Again, the gap equation follows from the requirement that $F$ is stationary {\it w.r.t.}~the gap parameters $\Sigma_i$ and $\Delta_i$. \begin{figure}[t] \epsfig{file=gap_fsb_mu350.ps,width=8.5cm,angle=0} \epsfig{file=gap_fsb_mu500.ps,width=8.5cm,angle=0} \caption{Superfluid order parameters and gaps as a function of $m_s$ for $\mu=350$~MeV (left panel) and $\mu=500$~MeV (right panel). The upper panels show the up-down and up-strange components $\Delta_{ud}$ and $\Delta_{us}$ of the color-flavor locked state. The lower panels show the maximum, minimum, and average superconducting gap.} \label{fig_fsb} \end{figure} Numerical results are shown in Fig.~\ref{fig_fsb}. The parameters were fixed as described in the last section. In the figure we plot $\Delta_{ud}$ and $\Delta_{us}$ as a function of $m_s$ for two different chemical potentials $\mu=0.35$ and 0.5 GeV. In particular we show the value of the largest, the smallest, as well as the average gap. In the two-flavor case, the smallest gap is zero and the average gap is $4/9$ of the maximum gap. In the three-flavor case, the smallest gap is $1/2$ and the average gap $5/9$ of the maximum gap. We observe that there is a sharp transition between the two-flavor scenario ($\Delta_{us}=0, \Delta_{ud}\neq 0$) and the three-flavor scenario ($\Delta_{ud}=\Delta_{us}\neq 0$) that takes place around the physical value of the strange quark mass, $m_s^{crit}=65$ MeV for $\mu=0.35$ GeV, and $m_s^{crit}=160$ MeV for $\mu=0.5$ GeV. The ratio $\Delta_{ud}/\Delta_{us}$ grows roughly linearly already for small $m_s$. This is different from the results of \cite{SW_99,ABR_99}, and an instanton effect. As discussed above, instantons induce a four-fermion interaction among light quarks which is proportional to $m_s$. We should note that we have not included the possibility of a dynamically generated contribution to the strange quark mass in the superfluid phase. In terms of the current mass, this effect will shift the critical mass to smaller values. \section{Statistical Mechanics of the Instanton Liquid} \label{sec_cocktail} \subsection{The Cocktail Model at non-zero Density} In this section we take a step beyond the mean-field approximation in that we allow for possible clustering effects in the instanton ensemble. This is achieved by employing a somewhat different formalism. In the previous sections we started from an effective quark interaction obtained by integrating out the gluonic (instanton-) fields in the underlying partition function. In this section we reverse the strategy and integrate over the fermion (quark) fields first. This leads to the following partition function for the instanton ensemble at finite density: \begin{eqnarray} {\cal Z}_{inst}(\mu)=\sum\limits_{N_+,N_-} \frac{1}{N_+! N_-!} \prod\limits_{I=1}^{N_+,N_-} \int d\Omega_I \ n(\rho_I) \ e^{-S_{int}} \ \rho_I^{N_f} \prod\limits_{f=1}^{N_f} \det (i \not\!\!D+im_f-i\mu\gamma_4) \ . \label{zinst} \end{eqnarray} The original QCD path-integral over all possible gluonic field configurations has been converted into an integration over the collective coordinates $\Omega_I=\{z_I,\rho_I,u_I \}$ (position, size and color orientation) of $N_+$ instantons and $N_-$ antiinstantons. The single-instanton amplitude $n(\rho_I)$ contains the semi-classical tunneling rate (including one-loop quantum corrections), as well as the Jacobian arising from the introduction of collective coordinates. The instanton interactions can be divided into a gluonic part $S_{int}$ and a fermionic part represented by the determinant of the Dirac operator. It is usually approximated in the subspace of zero modes, {\it i.e.}, \begin{eqnarray} \det(i\not\!\!D-i\mu\gamma_4) \simeq \det \left( \begin{array}{cc} 0 & T_{IA}(\mu) \\ T_{AI}(\mu) & 0 \end{array} \right) \ , \end{eqnarray} where $T_{IA}$ is the fermionic overlap matrix element discussed in sect.~\ref{sec_tia}, see Eq.~(\ref{tiamu}). When restricted to the zero mode basis, the fermionic determinant is in fact equivalent to the sum of all closed loop diagrams to all orders in the 't Hooft effective interaction. The non-hermiticity of the finite-$\mu$ Dirac operator is reflected by the fact that \begin{eqnarray} T_{AI}(\mu) &=& T_{IA}^\dagger(-\mu) \nonumber\\ &\ne& T_{IA}^\dagger(\mu) \ , \label{sign} \end{eqnarray} {\it i.e.}, the fermionic determinant is complex, entailing the well-known 'sign'-problem in the partition function, which will be addressed below. In the statistical mechanics treatment chosen here, spontaneous chiral symmetry breaking in the vacuum is generated by randomly distributed uncorrelated anti-/instantons which allow for a delocalization of the associated quark quasi-zero modes corresponding to the formation of a nonzero $\langle q\bar q\rangle$ condensate state. In other words, quarks can travel arbitrarily long distances by randomly jumping from one instanton to another (antiinstanton), and thus may carry their chiral charge to spatial infinity where it effectively becomes `lost'. In this picture chiral restoration can in principle proceed in two ways: either instantons disappear altogether, or they rearrange into some {\em finite} clusters which do no longer support any finite $\langle q\bar q\rangle$ condensate. In the limit of very large temperature or density instantons will eventually disappear, because Debye screening of the large gluon fields inside the instantons leads to a strong suppression of the tunneling rate. Nevertheless, in the case of finite temperature QCD it was argued~\cite{IS_94} that this cannot be the relevant mechanism for chiral restoration, since lattice simulations have observed the transition at rather low temperatures of $T_c^\chi\simeq 150$~MeV. This, in turn, led to the suggestion that chiral restoration proceeds through the formation of $I$-$A$ molecules. Further support for this idea is provided by lattice measurements of the instanton density at finite temperatures, showing no depletion below $T_c^\chi$ and a smooth onset of the expected Debye-screening above~\cite{CS_95,Alles}. More recent studies~\cite{DeForcrand,Ilg_new} seem to find more direct indications for $I$-$A$ molecule formation at $T\approx T_c$, but their quantitative role in the transition remains to be clarified. At finite density additional possibilities for clustering of the (chirally asymmetric) random instanton liquid into chirally symmetric configurations are available. In particular, random instantons can be supported by diquark condensates for $N_f=2$ or by a combination of diquark and quark-antiquark ones for larger $N_f$. However, as we will show in this section, effects of $I$-$A$ molecule formation may still be an important element in the finite-$\mu$ chiral restoration transition, as was shown in Ref.~\cite{Rapp_98}. In some sense this is not really surprising: when both T and $\mu$ are sufficiently large so that the all condensates are absent ({\it i.e.}, in the true QGP phase), $I$-$A$ molecules should constitute the preferred configuration for any remaining instanton component in the system. To investigate the interplay between the various components in the finite density partition function more quantitatively we resort to the 'cocktail-model' introduced in Refs.~\cite{IS_89,IS_94}. Here, the instanton ensemble is decomposed into a mixture of random (``atomic'') and ``molecular'' configurations, which yields a grand canonical partition function of the form \begin{eqnarray} {\cal Z}_{inst}^{a+m}=\sum_{N_a,N_m} {(z_a V_4)^{N_a} \over N_a!} \ {(z_m V_4)^{N_m} \over N_m!} \ . \end{eqnarray} In the thermodynamic limit $V_4\to \infty$, and using the Stirling formula, the free energy (thermodynamic potential) becomes \begin{eqnarray} \Omega_{inst}^{a+m}(n_a,n_m;\mu)= -\frac{\ln[{\cal Z}^{a+m}_{inst}]}{V_4} =-n_a \ \ln\left[ \frac{{\rm e} z_a}{n_a} \right] - n_m \ \ln\left[ \frac{{\rm e} z_m}{n_m} \right] \ . \label{Ominst} \end{eqnarray} The atomic and molecular 'activities' are~\cite{IS_89,IS_94} \begin{eqnarray} z_a & = & 2 \ C \ \rho^{b-4} \ {\rm e}^{-S_{int}} \ \langle T_{IA}(\mu) T_{AI}(\mu)\rangle^{N_f/2} \nonumber\\ & = & 2 \ C \ \rho^{b-4} \ {\rm e}^{-S_{int}} \ \left(\frac{n_a}{2} \int d^4z \ du \ [T_{IA}(\mu) T_{AI}(\mu)] \ \rho^2 \right)^{N_f/2} \nonumber\\ z_m & = & C^2 \ \rho^{2(b-4)} \ {\rm e}^{-2S_{int}} \ \langle [T_{IA}(\mu) T_{AI}(\mu)]^{N_f}\rangle \ \nonumber\\ & = & C^2 \ \rho^{2(b-4)} \ {\rm e}^{-2S_{int}} \ \int d^4z \ du \ \left[ T_{IA}(\mu) T_{AI}(\mu)\right]^{N_f} \ \rho^{2N_f} \ . \label{zam} \end{eqnarray} The underlying approximation in this approach is that the values of the hopping amplitudes $T_{IA}$ in each individual configuration are replaced by a product of their mean square values in an uncorrelated ensemble. To establish a connection to the (chiral) quark condensate and constituent mass, one can use the mean-field approximation to express them via the ``atomic'' density $n_a$ as \begin{eqnarray} \langle \bar qq\rangle &=& -\frac{1}{\pi\rho} \left(\frac{3}{2}n_a\right)^{1/2} \\ M &=& - C_{M} \ \frac{2}{3} \ (\pi \rho)^2 \langle \bar qq\rangle = - C_{M} \ (\pi \rho) \ \sqrt{\frac{2}{3}n_a} \ , \label{Mq} \end{eqnarray} respectively. In the original work of Ref.~\cite{Shu_82}, where these relations have been first derived, the coefficient $C_M=1$, leading to an effective quark mass of $M^*\simeq 200$~MeV. The latter is to be understood as the average mass of a constituent quark at finite (euclidean) momentum participating in tunneling processes associated with instantons. With increasing four-momentum $M^*$ is appreciably reduced and therefore does not correspond to the usual constituent quark mass $M$ (defined at zero momentum). We account for this by using $C_M=2$, yielding $M=400$~MeV. The minimization of the total $\Omega$ over $n_a$ and $n_m$ determines their equilibrium values in the ensemble for given $T,\mu$. The corresponding gap equation for the constituent quark mass is the direct analog of the mean-field equation conjugated to the $\langle\bar q q\rangle $ condensate, but expressed in terms of different variables. For a refined treatment at finite densities we supplement the cocktail model by two additional components. Following the arguments given above, we have to account for the possibility that the random instanton component can become engaged in diquark chains, first observed in numerical simulations of the instanton ensemble in the high-density limit~\cite{Sch_98}. The pertinent term in the free energy reads \begin{eqnarray} \Omega_{d}(n_d;\mu)=-n_d \ln\left[ \frac{{\rm e} z_d}{n_d}\right] \label{Omegad} \end{eqnarray} with the associated activity \begin{eqnarray} z_d = 2 \ C \ \rho^{b-4} \ {\rm e}^{-S_{int}} \ \left(\frac{n_a}{2} \int d^4z \ du \ [T_{IA}(\mu) T_{IA}(\mu)]_{\bar 3} \ \rho^2 \right)^{N_f/2} \ . \label{zd} \end{eqnarray} The subscript ``$\bar 3$'' indicates the color projection in some predefined direction characterizing the color vector of the diquark. In the same way that $n_a$ determines the constituent quark mass $M$, the superconducting gap is related to the density $n_d$ as \begin{equation} \Delta = C_\Delta \frac{\pi\rho}{(N_c-1)} \sqrt{\frac{2}{3} n_d} \label{deltaMF} \end{equation} In analogy to Eq.~(\ref{Mq}) we use $C_\Delta=2$, but also perform calculations with $C_\Delta=1.5$ to assess the inherent uncertainty of this mean-field estimate. Notice that in contrary to Eqs.~(\ref{zam}), the fermionic overlap matrix element enters Eq.~(\ref{zd}) as $(T_{IA})^2$, which, in fact, causes the $z$-integration to diverge. This is precisely the BCS-singularity of an attractive interaction in the particle-particle channel, here encountered in coordinate space. The standard procedure to treat this singularity is to start from a new ground state which a priori has the gap built into the fermion propagators, thereby regulating the integrals. The net effect of the gap on the overlap integrals is a damping factor for the intermediate quark propagators, which is delineated in appendix~\ref{sec_damp}. The second refinement consists of including a Fermi sphere of quark-'quasi-particles' in the free energy, representing the contribution of quark non-zero modes. The final expression for the thermodynamic potential then becomes \begin{eqnarray} \Omega(n_a,n_m,n_d;\mu)=\Omega_{inst}^{a+m+d}(n_a,n_m,n_d;\mu) + \Omega_{quark}^{QP}(M,\Delta;\mu) \ , \label{Omega_cktl} \end{eqnarray} where the quark contribution is simply given by \begin{eqnarray} \Omega_{quark}^{QP}(M,\Delta;\mu) = \epsilon_q(M,\Delta;\mu) -\mu \ n_q(M,\Delta;\mu) \end{eqnarray} with \begin{eqnarray} n_q(M;\mu) &=& g_q \int d^3k \ n_F(\mu-\omega_k) \nonumber\\ \epsilon_q(M;\mu) &=& g_q \int d^3k \ \omega_k \ n_F(\mu-\omega_k) \ , \end{eqnarray} $g_q=2 N_c N_f$ and $\omega_k^2=M^2+k^2$ (in the superconducting phase the dispersion relations of paired quarks picks up an additional dependence on the gap $\Delta$, see below). At this point it is instructive to compare the 'cocktail model' with other approaches. The (simpler) philosophy employed in the first part of this paper introduces the multi-fermion interactions with a {\em constant} coupling $g$, independent of condensates and temperature/density. However, since $g$ is itself generated through instantons, density variations of the latter will affect the coupling. A step towards including this feature was made by Carter and Diakonov~\cite{CD_99}: their coupling parameter, called $\lambda$ (which essentially corresponds to $z_a$ without the fermionic determinant part), was subjected to minimization. The resulting gap equation relates the mean instanton density $n_a$ to $\lambda$ and the condensates. However, instead of calculating $\lambda$ and then finding $n_a$ (as done here), they stayed in the mean-field approximation in which the instanton density remains unchanged. Our results to be discussed below support the (approximate) validity of this assumption. To investigate possible mechanisms for chiral symmetry restoration at finite density in some detail, in particular the competition between diquark and molecule formation, we will in the following separately discuss various versions of the cocktail with increasing complexity, {\it i.e.}, the two-flavor model including (anti-) instantons, molecules and a Fermi sphere of constituent quarks (sect.~\ref{CKTnf2}), additionally including the simplest $ud$-pairing as discussed in Refs.~\cite{ARW_98,RSSV_98} (sect.~\ref{CKTnf2+}), and the three-flavor case (sect.~\ref{CKTnf3}). In sect.~\ref{sec_coupl} we also give estimates for the density-dependence of effective coupling constants for molecule-induced (anti-)quark-quark interactions, which naturally emerge from the formalism employed in this section. \subsection{Two Flavor Cocktail Model without Diquark Condensates} \label{CKTnf2} In this subsection we basically follow the approach of Ref.~\cite{IS_94}, generalizing it to finite density~\cite{Rapp_98}. The issue here is to assess the potential role of $I$-$A$ molecule formation in chiral symmetry restoration at finite density, without the additional complication of superconducting gaps. For the actual calculations we now have to face the problem of the complex fermionic determinant appearing in the various activities, Eqs.~(\ref{zam}) and (\ref{zd}). As has been suggested in Ref.~\cite{Rapp_98}, it can be solved under the assumption that the gluonic interaction does not exhibit a pronounced dependence on the color angles, approximating it by an average value (see below). As a result, the color dependence in the activities only enters through the combinations of $T_{IA}$'s, which then can be integrated analytically, rendering the fermionic determinant real. For two flavors one obtains \begin{eqnarray} z_a(z_4,r) & \propto & \int du T_{IA}(\mu) T_{IA}^\dagger(-\mu) = \frac{1}{2N_c} [f_1^+f_1^-+f_2^+f_2^-] \nonumber\\ z_m (z_4,r) & \propto & \int du [T_{IA}(\mu) T_{IA}^\dagger(-\mu)]^{N_f} =\frac{(2N_c-1) \{f_1^+f_1^-+f_2^+f_2^-\}^2 +\{f_1^+f_2^--f_1^-f_2^+\}^2}{4N_c(N_c^2-1)} \ \label{colav2} \end{eqnarray} where $f_i^{\pm}\equiv f_i(\pm\mu)$ are defined through Eqs.~(\ref{tiamu}) and (\ref{tiamu2}). Note that the color averaging of the former complex expressions is sufficient to yield real-valued activities. Also note that, whereas $z_m(z_4,r)$ is a positive definite quantity, this is not the case for $z_a(z_4,r)$ due to the oscillations in $f_{1,2}(r)$. In fact, when further integrating $z_a(z_4,r)$ over space-time (the remaining four collective coordinates), delicate cancellations occur, which require accurate numerical values for the $f_{1,2}^{\pm}$; otherwise one easily encounters negative/incorrect results for both activities at finite chemical potential. The gluonic interaction entering Eqs.~(\ref{zam}) has been approximated by an average repulsion $S_{int}=\kappa \rho^4(n_a+2n_m)$, where $\kappa=\beta/2\bar\rho^4 (N/V)$, $\beta=b/2+3N_f/4-2$, $b=\frac{11}{3}N_c-\frac{2}{3}N_f$. The free parameter $\kappa$ characterizes the diluteness of the ensemble. In the case $N_c=3$ and $N_f=2$ we have chosen $\kappa\simeq 130$ in order to reproduce the phenomenological value for the diluteness of the instanton vacuum. The normalization constant $C\propto (\Lambda_{QCD})^b$ can also be fixed in the vacuum by requiring that the absolute minimum of the thermodynamic potential $\Omega_{inst}(\mu=0;n_a,n_m)$ appears at a total instanton density of $N/V=n_a+2n_m=1.4$~fm$^{-4}$, being realized for $n_a$=1.34~fm$^{-4}$ and $n_m$=0.03~fm$^{-4}$, which gives $\Lambda_{QCD}\simeq 260$~MeV. The smallness of the molecular component in the vacuum is a consequence of the large entropy associated with quantum fluctuations of the color angles, randomizing the system. Available lattice data agree that such correlations are indeed small at $T=\mu=0$. \begin{figure}[tp] {\makebox{\epsfig{file=zam32.eps,width=90mm,angle=0}}} \hspace{-0.7cm} {\makebox{\epsfig{file=omtotmq30.eps,width=75mm,angle=0}}} \caption{Results for the two-flavor cocktail model {\it without} diquark pairing; the left part shows the densities of random instantons ('atomic' component) and $I$-$A$ molecules (upper left panel), the pressure $p=-\Omega$ (middle left panel) and the constituent quark mass (lower left panel) after minimizing the free energy, Eq.~(\ref{Omega_cktl}), {\it w.r.t.}~$n_a$ and $n_m$. In the right panel the free energy is displayed as a function of the constituent quark mass $M \propto n_a^{1/2}$, indicating a first order transition from the minimum at finite $M$ to the one at $M=0$. } \label{fig_ct2} \end{figure} Our numerical findings for the $N_c=3, N_f=2$ atomic+molecular instanton cocktail model at finite density are summarized in Fig.~\ref{fig_ct2}. At small $\mu$ essentially nothing happens until, at a critical value $\mu_c\simeq 310$~MeV, the system jumps into the chirally restored phase, the latter being characterized by $n_a=0$. The transition is of first order, as can be seen by inspection of the $M$-dependence of the free energy (right panel of Fig.~\ref{fig_ct2}). Below $\mu_c$, the pressure actually decreases slightly with increasing $\mu$ indicating a mixed phase-type instability, similar to what has been discussed in Refs.~\cite{ARW_98,Bub_96}. The total instanton density at the transition (residing in $I$-$A$-molecules) is appreciable, $N/V=2n_m\simeq 1.1$~fm$^{-4}$, providing the major part of the pressure at this point. In other words: a substantial part of the non-perturbative vacuum pressure persists in the chirally restored phase. The vacuum pressure of $p(\mu=0)=0.6$~GeV~fm$^{-3}$ (indicated by the dotted line in the middle left panel of Fig.~\ref{fig_ct2}) is not recovered until a chemical potential of $\mu_0=350$~MeV, corresponding to a free quark number density of $n_q^0=1.15$~fm$^{-3}$ (naively, this translates into a nucleon density of $n_N=2.4n_0$ with the normal nuclear matter density $n_0=0.16~$fm$^{-3}$). For all chemical potentials in between, $0<\mu<\mu_0$, the system is mechanically instable, possibly indicating droplet formation in the region $M<\mu<\mu_0$ (as suggested in Ref.~\cite{ARW_98}), where, with a finite quark density, the pressure $p(\mu)$ is below its vacuum value. A further comment concerning the numerical value of the critical chemical potential, which is very close to a third of the nucleon mass, is in order. Given the various approximations applied it should be regarded as a coincidence. At the same time it most likely provides a {\it lower} bound for the true value, as has been the case for the finite temperature calculations of Ref.~\cite{SS_96} (resulting in a critical temperature which is about 20\% lower than observed on the lattice and might well be related to the fact that the instanton model lacks explicit confinement). On the other hand, the inclusion of superconducting gaps may further reduce the critical chemical potential and in fact dominate the transition, as will be discussed in the following section. \subsection{Two Flavor Cocktail Model Including Pairing} \label{CKTnf2+} We now address the effects of diquark condensation. For clarity, let us first ignore the $I$-$A$ molecule component in the cocktail to study the competition between chiral and diquark condensates only. The additional inclusion of a pairing gap resulting from $ud$ diquark formation at the Fermi surface as discussed in Ref.~\cite{RSSV_98} modifies the quark quasiparticle contribution according to \begin{eqnarray} \Omega_q^{\Delta}(\mu)=\rm tr \log\left[D(M,\Delta)\right] \ . \label{Omqdel} \end{eqnarray} Here, $D$ denotes the quasiparticle quark-propagator now including the BCS gap $\Delta$. All interaction contributions are effectively accounted for through the instanton part $\Omega_{inst}^d$, cf.~Eqs.~(\ref{Omegad}), (\ref{zd}). The resulting expression for the free energy then becomes \begin{eqnarray} \Omega(n_a,n_d;\mu)= \Omega_{inst}^{a+d}(n_a,n_d;\mu) +\frac{1}{N_c} \left[2\Omega_q^\Delta(M,\Delta;\mu)+(N_c-2) \Omega_q^{QP}(M;\mu)\right] \ \end{eqnarray} (the last term accounting for unpaired quarks), which now has to be minimized {\it w.r.t.}~$n_a$ and $n_d$. The results for two different values of the vacuum instanton density and the (not precisely determined) coefficient $C_\Delta$ for calculating the pairing gap (cf.~Eq.~(\ref{deltaMF})) are summarized in Tab.~\ref{tab_ad}. We find that color superconductivity appears at critical chemical potentials around $\mu_c\simeq$~300~MeV, very similar to the values of the previous section where only $I$-$A$ molecule formation was considered. The associated gaps range between 120-180~MeV. These results are consistent within 20\% with the findings of sect.~\ref{sec_nf2} and those of Ref.~\cite{CD_99}. We should also note that the calculated gaps in our original work~\cite{RSSV_98} are significantly smaller (below $\sim$~100~MeV) as those were effectively obtained for 2+1 flavors, {\it i.e.}, in Ref.~\cite{RSSV_98} we included the effect of a reduced (constituent) strange quark mass $M_s$ in the closed-off strange quark loop of the six-fermion instanton vertex, which decreases the effective instanton-induced coupling constant for the four-quark interaction by about 60\% in the chirally restored phase. Another feature that emerges here is that the total instanton density changes little across the transition, which a posteriori justifies to assume it as constant in the mean-field calculations of sect.~\ref{sec_nf2}. As in the previous section, there is an intermediate constituent quark-diquark phase, similar to what was found in sect.~\ref{sec_nf2}, but here it again has small negative pressure which makes it mechanically unstable against the formation of a mixed phase. Let us now turn to the full two-flavor cocktail model with simultaneous account for the chiral and diquark condensates as well as $I$-$A$ molecules. The total free energy \begin{eqnarray} \Omega(n_a,n_d,n_m;\mu)= \Omega_{inst}^{a+d+m}(n_a,n_d,n_m;\mu) +\frac{1}{N_c} \left[2\Omega_q^\Delta(M,\Delta;\mu)+(N_c-2) \Omega_q^{QP}(M;\mu)\right] , \ \end{eqnarray} is to be minimized {\it w.r.t.}~$n_a$, $n_m$ and $n_d$. The results are shown in Fig.~\ref{fig_ct2+}, using the coefficient of $C_\Delta=1.5$ in Eq.~(\ref{deltaMF}) (which most closely resembles the results of the calculations in sect.~\ref{sec_nf2}). \begin{table}[htb] \begin{tabular}{ccccc} $n_a(\mu=0)$ [fm$^{-4}$] & $C_\Delta$ & $\mu_c$ [MeV] & $\mu_0$ [MeV] & $\Delta(\mu_c)$ [MeV] \\ \hline 1 & 2 & 300 & 355 & 158 \\ 1 & 1.5 & 265 & 290 & 120 \\ 1.4 & 2 & 310 & 375 & 188 \\ 1.4 & 1.5 & 270 & 305 & 142 \\ \end{tabular} \caption{Parameter dependence of the critical chemical potential for chiral restoration in a cocktail model with chiral and diquark condensates (no $I$-$A$ molecules included). } \label{tab_ad} \end{table} \begin{figure}[htb] \begin{center} {\makebox{\epsfig{file=zabm3215.eps,width=120mm,angle=0}}} \end{center} \caption{The full two-flavor cocktail model including $I$-$A$ molecules and diquark pairing; upper panel: densities of random instantons ('atomic' component, full line), $I$-$A$ molecules (dashed line) and instantons engaged in 'diquark chains' (dashed-dotted line); middle panel: pressure $p(\mu)=-\Omega(\mu)$, maximized over $n_a, n_m$ and $n_d$ at each value of $\mu$; lower panel: constituent quark mass (full line, using $C_M=2$ in Eq.~(\protect\ref{Mq}) to give $M(\mu=0)=400$~MeV) and diquark gap (dashed-dotted line, using $C_\Delta=1.5$ in Eq.~(\protect\ref{deltaMF}), as discussed in the text).} \label{fig_ct2+} \end{figure} As to be expected from the previous analysis, there is a delicate competition for chiral restoration between random instantons engaged in diquarks and $I$-$A$ molecule formation. The former do, in fact, induce the chiral transition for all parameter ranges considered. Towards higher densities it may happen that a second transition occurs within the chirally restored phase characterized by a substantial jump in the molecule density and an accompanied drop of the diquark gap, (its location is somewhat sensitive to parameter choices). The singularity in $z_d$, on the other hand, guarantees that there is always a finite $\langle qq\rangle$ condensate present, albeit possibly strongly reduced in molecule-dominated phases. In Fig.~\ref{fig_ct2+}, the transition occurs at $\mu_c=270$~MeV into a chirally broken diquark phase, which, again, is mechanically unstable. The vacuum pressure is only recovered at $\mu_0=295$~MeV, {\it i.e.}, the combined effect of diquarks and molecules further lowers the $\mu_0$-values found when including only molecules (Fig.~\ref{fig_ct2}, where $\mu_0=350$~MeV) or only diquarks (Tab.~\ref{tab_ad}, where $\mu_0=305$~MeV). Another important point to note is that the total instanton density, $N/V=n_a+n_d+2n_m$, is indeed essentially independent of the chemical potential within $\sim$15\% or so. {\it E.g.}, in Fig.~\ref{fig_ct2}, the upward jump in the density of molecules is close to half the drop in density of the random instanton component. A similar continuity is seen in Fig.~\ref{fig_ct2+} for the more realistic case of the transition between the two kinds of random instanton liquids, associated with the two condensates $\langle\bar qq\rangle$ and $\langle qq\rangle$. After all this is not too surprising since, at least for chemical potentials $\mu\le 0.4$~GeV, the by far dominant fraction of the total energy in the system is carried by the gluonic component residing in instantons, and the free energy itself must of course be continuous across the transition. Finally, a comment concerning the finite temperature behavior is in order. On the zero-density and finite-$T$ axis it has been shown that molecules drive the chiral restoration. This implies that at finite density, starting from low $T$, the role of molecules should become more and more relevant; at the same time, color-superconducting gaps are well-known to be suppressed. Thus we expect that for {\em any} value of $\mu$ where there is diquark condensed phase for $T=0$, a rearrangement into a molecule-dominated phase should occur when raising the temperature. As indicated by our zero temperature results, for chemical potentials $\mu\ge 300$~MeV or so, the corresponding $T_c(\mu)$ line in the phase diagram might actually reach down to fairly low temperatures. \subsection{Chiral Restoration for more Flavors} \label{CKTnf3} As was already mentioned in the introduction, in vacuum the tendency towards chiral restoration in the instanton model strongly increases with the number of flavors, leading to a chirally symmetric vacuum state for $N_f$ as low as $\sim$~5. The reason for that has been discussed in sect.~\ref{molint}: the increased number of quark lines enhances the interactions between instantons and antiinstantons, making the random liquid less favorable. More specifically, the integral over color orientations within the molecule, being proportional to $\langle (\cos \theta)^{2N_f} \rangle$, increases more strongly than that for the averaged 'random'-instanton configurations raised to the $N_f$-th power, $\langle (\cos \theta)^2\rangle^{N_f}$. The former integral is strongly peaked at $\theta=0$, creating a ``locking'' of the color orientation within a molecule. From continuity one may thus expect that the critical chemical potential for chiral restoration should be further reduced when moving from two to more massless flavors. Here we would like to pursue the question in how far a $third$ massless flavor impacts the results of the two-flavor case. For simplicity, we now ignore color superconductivity and consider an interplay between random and molecular components only. As discovered in \cite{ARW_98b} and further elaborated in the first part our article, starting from $N_f=3$ a color-flavor locking phenomenon sets in, leading to a complicated set of $\langle qq\rangle$ and $\langle \bar q q\rangle$ condensates. However, they are relatively small in magnitude, and can therefore be neglected for the present purpose. Compared to the $N_f$=2 case, the color integration for $N_f$=3 is substantially more involved. Using the appropriate relations for the integration over a string of six SU(3)-color matrices (see, {\it e.g.}, Ref.~\cite{Now_91}), one obtains \begin{eqnarray} z_m(z_4,r) & \propto & \int du [T_{IA}(\mu) T_{IA}^\dagger(-\mu)]^{N_f} \nonumber\\ & = & \frac{3 \left( f_1^+ f_1^- + f_2^+ f_2^- \right)} {16N_c(N_c+2_)(N_c^2-1)} \left( 3N_c \left[ (f_1^+)^2 (f_1^-)^2 + (f_2^+)^2 (f_2^-)^2 \right] + \left[ 4-N_c \right] \left[ f_1^+f_2^- - f_1^- f_2^ +\right]^2 \right) \ . \label{colav3} \end{eqnarray} Note that $z_m(z_4,r)$ in the three-flavor case has no definite sign before integration over space-time (similar to the $N_f$=1 case, {\it i.e.}, $z_a(z_4,r)$ from Eq.~(\ref{colav2})), which inevitably entails partial cancellations. Apparently, only for an even number of flavors the space-time integrand of the finite density activities is positive definite. Performing again the minimization procedure in $n_a$ and $n_m$ for the thermodynamic potential we find that the critical chemical potential is indeed further reduced, by about 10\% to $\mu_c\simeq$~270~MeV (as compared to 310~MeV for $N_f$=2). This indeed complies with the above mentioned expectation that at a sufficiently large number of flavors a purely ``molecular'' vacuum is more preferable than a ``random'' one, and thus chiral symmetry would be unbroken even in vacuum. As this phenomenon reflects itself also along the $\mu$-axis, we expect $\mu_c$ to be further reduced at $N_f$=4, possibly crossing zero at $N_f$=5. \subsection{Molecule-Induced Effective Couplings} \label{sec_coupl} In this section we apply the cocktail model to a microscopic estimate of the density-dependence in the effective coupling constants for (anti-)quark-quark interactions. In the mean-field framework employed in sects.~\ref{sec_nf2},~\ref{sec_nf3},~\ref{sec_fsb} such density-dependencies were not accounted for. To begin with let us recall the expression for the effective coupling constant for single-instanton induced interactions as, {\it e.g.}, used in Ref.~\cite{RSSV_98}: \begin{eqnarray} G_{inst} = \int d\rho \ n(\rho,\mu) \ \rho^{N_f} \ (2\pi\rho)^4, \label{Ginst} \end{eqnarray} with \begin{eqnarray} n(\rho,\mu) = C_{N_c} \ \left(8\pi^2/g^2\right)^6 \ \exp\left[-8\pi^2/g(\rho)^2\right]\ \rho^{-5} \label{nrho} \end{eqnarray} denoting the single-instanton distribution, {\it i.e.}, the semi-classical tunneling amplitude. Replacing the $\rho$-integration by an average value for $\bar \rho$, one has \begin{eqnarray} G_{inst} = n^{\pm} \ (\bar \rho)^{N_f} \ (2\pi\bar\rho)^4 \end{eqnarray} with the anti-/instanton density $n^{\pm}$. Notice that (for $N_f=2$) six additional powers of $\rho$ appear in the integral of Eq.~(\ref{Ginst}) as compared to the usual expression for the density, $n^\pm=\int d\rho \ n(\rho)$. This entails a significantly larger average value $\bar\rho$ than the typical instanton size of about $\rho\simeq 1/3$~fm. In order to be consistent the standard value of the zero-density constituent quark mass $M \simeq 400$~MeV (which requires $G_{inst}\simeq 19.1$~fm$^2$), one should have the effective size saturating this integral to be $\bar\rho=0.51$~fm. The effective coupling constant for molecule-induced interactions has been first derived in Ref.~\cite{SSV_95}, where it is written as \begin{eqnarray} G_{mol}= \int n(\rho_1,\rho_2) \ d\rho_1 \ d\rho_2 \ \frac{1}{T_{IA}^2} \ (2\pi\rho_1)^2 \ (2\pi\rho_2)^2 \label{Gmol} \end{eqnarray} with the total molecule amplitude \begin{eqnarray} n(\rho_1,\rho_2)=\int du \ d^4z \ n(\rho_1) \ n(\rho_2) \ T_{IA}(u,z)^{2N_f} \ \rho_1^{N_f} \ \rho_2^{N_f} \ , \label{nmol} \end{eqnarray} which is nothing but the molecular activity given in Eq.~(\ref{zam}). The graphical interpretation of Eq.~(\ref{Gmol}) is quite transparent: starting from the molecule amplitude, where {\em all} 2$N_f$ quark legs are closed within the molecule, one 'cuts' open two of them (corresponding to the division by $T_{IA}^2$) which provides an effective interaction between four external (anti-) quarks. Inserting (\ref{nmol}) into (\ref{Gmol}), and replacing again the size integrals by using average values for $\bar\rho$ we obtain \begin{eqnarray} G_{mol}=\frac{(2\pi\bar\rho)^4 {\bar\rho}^{2N_f} n^+ n^-}{8} \int du \ d^4z \ T_{IA}(u,z)^{2N_f-2} \ . \label{Gmol2} \end{eqnarray} We see that for $N_f=2$ the individual $\rho$ integrations carry additional powers of $\rho^4$, while it is $\rho^5$ for $N_f=3$. Those powers should be compared to $\rho^6$ in Eq.~(\ref{Ginst}) and $\rho^0$ for the usual instanton density. As these last two integrals lead to $\bar\rho\simeq$ 0.51 and 1/3 fm, respectively, we interpolate between them (lacking a more accurate determination of the size distribution), {\it i.e.}, estimate the appropriate average size values entering Eq.~(\ref{Gmol2}) to be $\bar\rho\simeq 0.43$~fm for $N_f=2$ and $\bar\rho\simeq 0.47$~fm for $N_f=3$. The corresponding zero-density values for the coupling constants are (including a factor of 16 accounting for the difference in color coefficients between Eqs.~(\ref{l_mes}) and (\ref{l_mesmol})): \begin{eqnarray} G_{inst}(\mu=0,\bar\rho=0.51{\rm fm}) & = & 19.1~{\rm fm}^2 \nonumber\\ 16~G_{mol}^{N_f=2}(\mu=0,\bar\rho=0.47{\rm fm}) & = & 0.86~{\rm fm}^2 \nonumber\\ 16~G_{mol}^{N_f=3}(\mu=0,\bar\rho=0.43{\rm fm}) & = & 0.044~{\rm fm}^2 \ . \end{eqnarray} Obviously, $G_{inst}\simeq 20\times(16~G_{mol}^{N_f=2})$, and $(16~G_{mol}^{N_f=2})\simeq20\times(16~G_{mol}^{N_f=3})$. Recalling that the fermionic matrix element, averaged over positions and sizes of an $I$-$A$-pair, is given by~\cite{SS_98} \begin{eqnarray} \rho^2\langle |T_{IA}|^2\rangle &=& \frac{2\pi^2}{3N_c} \frac{N\rho^4}{V} \simeq {1\over 40} \end{eqnarray} (for $\rho=1/3$~fm), one readily understands the decreasing magnitudes of the coupling constants in terms of the additional powers of dimensionless diluteness of the ensemble entering Eq.~(\ref{Gmol2}). Note that the effect of color integrations favors molecules, and the corresponding factors compensate, to some degree, powers of the small diluteness parameter. We are now in position to evaluate the density-dependence of the effective coupling constants. Using the fermionic matrix elements at finite $\mu$, Eq.~(\ref{tiamu}), and performing the color integrations as discussed in sects.~\ref{CKTnf2},~\ref{CKTnf3} we obtain the results shown in Fig.~\ref{fig_Gmu}; we recall that throughout this paper possible effects from the Debye-screening of instantons at high densities have been ignored. Therefore $G_{inst}$ is in fact a constant as it does not depend on $T_{IA}(\mu)$. On the other hand, the behavior of the molecule-induced couplings depends on the number of flavors: whereas in the $N_f=2$ case it decreases with $\mu$, the opposite is found for $N_f=3$. This is directly related to the $\mu$-dependence of the activities calculated in sect.~\ref{CKTnf2}, since $G_{mol}^{N_f=2}(\mu)\propto T_{IA}^2$ (corresponding to $z_a$) and $G_{mol}^{N_f=3}(\mu)\propto T_{IA}^4$ (corresponding to $z_m^{N_f=2}$). It is instructive to compare the values of the coupling constants with the perturbative OGE interaction. For large-angle scattering\footnote{For small-angle scattering there appears an additional logarithmic enhancement, which becomes relevant for color superconductivity at asymptotically high $\mu$~\cite{Son_98}.} the coupling is \begin{eqnarray} G_{OGE}={4\pi\alpha_s \over p^2_{eff}} \ , \end{eqnarray} where $p_{eff}$ is some effective momentum transfer averaged over the Fermi sphere. With $p_{eff}=0.5-1$~GeV and $\alpha_s(p_{eff}) \approx 0.3$ we have $G_{OGE}=0.15-0.7$~fm$^2$. This is comparable to the effect of molecules for $N_f=2$, and significantly larger than that for $N_f=3$ at low $\mu$. One should note, however, that the structure of the interactions is different. For example, OGE is flavor independent whereas instanton-induced interactions are flavor antisymmetric. \begin{figure}[htb] \begin{center} \epsfig{file=Gmolmu.eps,width=75mm,angle=-90} \end{center} \caption{Instanton-/molecule-induced effective 4-quark coupling constants as a function of chemical potential.} \label{fig_Gmu} \end{figure} \section{More Phases, Outlook and Experimental Consequences} \label{sec_phdia} \subsection{The phase diagram} \label{sec_phd} In this section we would like to put the main results obtained in this work into perspective and discuss the emerging picture of the QCD phase diagram. The main feature of the phase diagram for two-flavor QCD in the chiral limit and at $T=0$ is that at a critical chemical potential $\mu_c$ the system undergoes a transition from the chirally broken phase to the superconducting phase~\cite{ARW_98,BR_98,CD_99}. With instanton-induced formfactors we also found a small window in which a chirally broken `constituent quark' plus diquark condensate phase may exist. Unfortunately, this is not a robust prediction of the model, since the free energy differences between the phases are small (in fact, in the cocktail model analysis of sect.~\ref{sec_cocktail} this phase is mechanically unstable). We will discuss this issue from a slightly different perspective in the next section. In any case, the mean-field approach predicts a strong first order phase transition, either from the vacuum phase to superfluid quark matter phase, or from the chirally broken diquark phase to the superfluid phase. This implies the existence of an inhomogenous phase at intermediate density, with dense quark matter bubbles immersed in the chirally broken phase. Of course, a more refined treatment should reproduce the fact that matter clusters into nucleons and nuclei. In the high density phase chiral symmetry is restored, but color-$SU(3)$ is broken to $SU(2)$. We have not explored the possibility of further breaking $SU(2)$ via color-6 condensates. These condensates seem to generate very small gaps for the up and down quarks of the third color~\cite{ARW_98}. For three massless flavors we also find a first order phase transition from the chirally broken vacuum phase to the superconducting phase. In spite of the fact that instanton-induced dynamics are very different from OGE considered in Ref.~\cite{ARW_98b}, we also find that the preferred order parameter in the superconducting phase exhibits color-flavor-locking. Both color-$SU(3)$ and chiral $SU(3)_L\times SU(3)_R$ are broken, while the diagonal $SU(3)_{C+L+R}$ is preserved. But even though color-flavor locking in general implies that chiral symmetry is broken, instantons are crucial for generating a non-zero value of $\langle \bar qq\rangle$. In practice, the value of the chiral condensate turns out to be small. Furthermore, we have considered the more general case of QCD with 2+1 flavors allowing for a massive strange quark. Of course, the two cases $N_f=2,3$ discussed above emerge as limiting cases for $m_s\to 0$ and $m_s\to\infty$. For $m_s\simeq 2\sqrt{\Delta\mu}$ pairing between light and strange quarks becomes impossible, and there is a phase transition between the color-flavor locked phase and the two-flavor superconductor. Again, a significant difference between the instanton model and schematic interactions abstracted from one-gluon exchange appears. In the case of OGE, the strange quark mass has a purely kinematical effect. Instantons induce four-fermion interactions of the type $m_s(ud)(\bar u\bar d)$, which can generate large asymmetries between the the $\langle ud\rangle$ and $\langle us\rangle =\langle ds\rangle$ components of the color-flavor locked state even for $m_s<m_s^{crit}$. Similar effects are well known from hadronic spectroscopy. \begin{figure}[htb] \begin{center} \epsfig{file=phases_CSC2.eps,width=6cm,angle=270} \end{center} \vspace{1cm} \caption{Schematic QCD phase diagram in the chemical potential $\mu$ - temperature $T$ plane. The small $T$-/$\mu$-region corresponds to ordinary hadronic matter, with broken chiral symmetry. The point M (from ``multifragmentation'') is the endpoint of the nuclear liquid-gas phase transition. The point E indicates where the first order line either terminates in a second order endpoint (for $m_u,m_d=0$) or disappears (for finite light quark masses). CSC2 and CSC3 label the $N_f=2$ and $N_f=3$-type superconducting phases. The hypothetical intermediate quark-diquark phase is indicated by QDQ.} \label{fig_phases} \end{figure} In the present work we have not addressed the effects of finite temperature. One would expect that the first order chiral phase transition at $\mu\neq 0$ will persist for some range in temperature, until the transition becomes second order at a tricritical point~\cite{HJS*_98,BR_98}. In BCS theory one can also estimate that superfluidity disappears at a critical temperature $T_c(\mu)\simeq 0.6\Delta(\mu,T=0)$ (however, as discussed in sect.~\ref{sec_cocktail} the impact of instanton-antiinstanton molecules presumably reduces the BCS-coefficient appreciably). As explained above, the boundary between the two types of superconductors should be approximately determined by the condition (\ref{mismatch}). Since, asymptotically, the gap is expected to slowly grow as a function of chemical potential, the critical temperature will also grow. As a function of chemical potential the system will eventually reach the color-flavor locked state for any value of the strange quark mass. On the other hand, as a function of temperature (at given $\mu$) one might expect that, since the gap in the two-flavor superfluid is somewhat larger than in the color-flavor locked phase, the system first makes a transition to the two-flavor superconductor, and then to the quark-gluon plasma phase. Combinig these conjectures leads to the schematic phase diagram dsiplayed in Fig.~\ref{fig_phases} (where we have also indicated the nuclear liquid-gas transition line). \subsection{More Phases?} \label{sec_more} Finally, we would like to make a few comments on the range of applicability and the limitations of our approach. First, we have restricted ourselves to an instanton model. Even though many of our conclusions apply to any phenomenologically successful effective interaction, many of our numbers are indeed model-dependent. Furthermore, we have made extensive use of the mean-field approximation (MFA). The MFA is valid provided the condensates are sufficiently smooth, and their fluctuations can be neglected. This assumption becomes better at large density, because in this limit the coupling is weak and Cooper pairs are large compared to the inter-quark distances (as is the case in ordinary superconductors). In the opposite limit of small density the simplest phase one can generate in the mean-field approach is a chirally asymmetric Fermi gas of constituent quarks with masses on the order of 400~MeV. Remarkably enough, our approach predicts that this phase is unstable against separation into a mixed phase, with high density quark droplets separated by pure vacuum~\cite{Bub_96,ARW_98}. While this result appears to be very suggestive, the MFA cannot predict the actual composition of the clusters. Of course, we know that the ``correct'' clusters are nucleons. Below nuclear matter saturation density, nucleons themselves will form a mixed phase of clusters (nuclei), but for larger density one has homogeneous nuclear matter. In the vicinity of nuclear saturation density ($n_N=n_0=0.16$~fm$^{-3}$), the equation of state is known experimentally. The behavior of the energy density as a function of density is commonly parameterized as \begin{eqnarray} \epsilon= \left(m_N-\delta m_N\right)n_N + \frac{K n_0}{18} \left(1-{n_N \over n_0}\right)^2, \end{eqnarray} where $\delta m_N\approx 16$~MeV is the binding energy (per nucleon) of nuclear matter and the compression modulus is on the order of $K^{-1}=200-300$~MeV. This means that nuclear matter is rather stiff, that is, the pressure grows very fast as a function of density. The physical reason for the steep rise in pressure is not just Fermi motion, but, more importantly, a strong repulsive core in the nucleon-nucleon interaction. The standard lore is that the steep rise in the pressure with density above saturation density will eventually stop, due to the transition to some other phase of matter, either previously considered scenarios such as pion and kaon condensation~\cite{mesonic_cond}, or superconducting quark matter. In neutron stars, $K^-$ condensation is additionally favored because of a large electron chemical potential. Although lacking explicit confinement, the instanton model does provide the interaction to bind quarks into nucleons~\cite{SSV_94}. The question whether the model provides the repulsive core necessary to prevent nucleons from collapsing into 6 and more quark clusters remains open. Nevertheless, it seems plausible that a sufficiently sophisticated treatment could lead to nuclear matter as the correct ground state at small density. Can there be other phases, in addition to nuclear matter and superfluid quark matter? In the remainder of this section we will discuss a number of possibilities connected with diquark fluids or Bose condensates. These diquarks would not be Cooper pairs, but tightly bound states. As discussed in sect.~\ref{sec_diq}, the instanton model seems to predict such states as bound scalar $ud$ diquarks. The corresponding energy per baryon $E/B= 3M_{dq} /2 \sim 800-900$ MeV is lower than the nucleon mass, so it seems natural to look for a diquark phase. Naively, one would expect diquarks at $T=0$ to be Bose-condensed in the zero momentum state, because in addition to the gain in binding energy, there is a gain over nuclear or quark matter because no Fermi motion is required. However, since diquarks are colored, this phase could not be color neutral. This means that we have to consider either (i) significant motion of diquarks or (ii) add color-neutralizing quarks. Let us start with the first idea, with only diquarks at $T=0$. A good starting point is the question why -- if the scalar diquark is bound -- a two-baryon state, such as the deuteron, does not decay into a more tightly bound three-diquark state. The simplest color singlet combination of three $(ud)_i=\phi_i$ scalar diquark fields is $\phi_{i1} \phi_{i2}\phi_{i3}\epsilon_{i1,i2,i3}$, but this wave function is antisymmetric, violating Bose statistics. This means that we have to consider p-wave diquark states $\phi_i^m$ ({\it e.g.}, in a bag), where $m$ is the third component of angular momentum, in a symmetric combination: \begin{eqnarray} \psi=\phi_{i1}^{m1} \phi_{i2}^{m2}\phi_{i3}^{m3}\epsilon^{i1,i2,i3}\epsilon_{m1,m2,m3}. \end{eqnarray} Simple estimates show that such a state is no longer more economical. Similarly, one can construct a wave function for infinite diquark matter starting from a set of plane wave states. The ground state would then be a quite peculiar ``color crystal'', where the diquark momentum is determined from the balance of kinetic and potential energy, the latter resulting from the color-electric fields. We have not attempted to calculate the energy in this phase, but it seems much more natural to consider a quark-diquark phase. The quark-diquark (QDQ) phase could occur as an intermediate phase between nuclear matter and the color superconducting phase. Let us focus on a flavor composition corresponding to neutron matter (relevant for dense stars): $ud$ diquarks plus an equal amount of $d$ quarks. In this case, the total color and electric charge are zero. However, if the density is very low, color has to be neutralized over large distances, and confinement prohibits a phase like this. It is amusing to note that even without confinement, there is no low density QDQ phase. One reason is that the nucleon is bound {\it w.r.t.}~a diquark and a quark. Another reason is that at very low density Fermi motion in the QDQ phase is more costly than in the nuclear phase. For example, let us set the threshold for the QDQ phase, $M_{dqq}=M_{dq}+M$ equal to the nucleon mass $m_N$. Then the density of color-compensating $d$ quarks is equal to the density of neutrons in the nuclear phase at the same baryon density. Furthermore, since both have the same degeneracy factor ($g_s$=2 due to spin, the color being fixed by the color of the diquark condensate), both the $d$ quark and the neutron have the same Fermi momentum $p_f$. The kinetic energy $p_F^2/(2 M)$ is then smaller for the neutron because of its larger mass. Nevertheless, it is not obvious which phase is preferred at densities a few times nuclear matter density. The QDQ phase is very different from both nuclear matter and color superconducting ($N_f=2$) quark matter, in particular both chiral symmetry and color are broken. This suggests that if such a phase exits, it is probably an isolated minimum, separated by first order transitions on both sides. Let us try to estimate if such a window may exist. We assume that $\langle\bar qq\rangle$ is large in this phase, still providing an effective quark mass $\cal O$(400 MeV), and therefore this phase should have no strange component, in contrast to the CSC3. The interactions in nuclear matter crucially combine long range attraction with short range repulsion. A similar description may be approximately valid for diquarks as well. Let us use a simple model, with only repulsive interactions represented by the scattering length\footnote{The radius of the nucleon repulsive core is about 0.4 fm, but diquarks (and constituent quarks) are smaller objects. If they are instanton-generated, their core should be of the order of the typical instanton radius $\rho \approx 1/3$~fm~\cite{Shu_82}, which we took as a representative value.} $a$. The energy per baryon of the diquark Bose gas is \begin{eqnarray} {\epsilon_{dq} \over n_B}={12 \pi a n_{dq} \over m} \left( 1+ {128 \over 15\sqrt{\pi}}(a^3n_{dq})^{1/2}\right) \ , \end{eqnarray} where $n_{dq}$ denotes the diquark density. The first term is just the mean-field interaction of the condensed diquarks, the second term stems from non-condensed bosons, as follows from the classic Lee-Yang paper~\cite{LY}. In this approximation the unphysical behavior of the ideal Bose gas is overcome, the chemical potential grows with the density, and at some density it becomes favorable to split some diquarks into quarks. We are then led to a mixture of a Bose gas of diquarks and a Fermi gas of quarks (now of the same color), with chemical potentials related by the equilibrium condition $\mu_{dq} = 2 \mu$. Such a description leads to a more natural transition to quark matter with Cooper pairs at high density. \begin{figure}[t] \epsfxsize=4.0in \centerline{\epsffile{p_mu3.eps}} \caption{ Pressure versus baryonic chemical potential (3 times $\mu$ for quarks as used above) for 3 phases: nuclear matter, the quark-diquark phase made of a Bose gas of interacting diquarks and a Fermi gas of constituent quarks (no bag constant), and a Fermi gas of $u,d,s$ quarks with current masses (for two different bag constants). } \label{fig1_fb} \end{figure} So far we have ignored the role of confinement. One possibility is that the transition to the quark-diquark phase leads to deconfinement. If not, one has to include the energy associated with separated color charges. It is well known that the confining potential in vacuum is linear $V(r)=Kr$, with a string tension $K\simeq 1$~GeV/fm. However, for small $r$ this relation only holds for heavy, point-like quarks, not for light ones. Various non-relativistic models of hadronic structure use some effective $K$, reduced by a significant factor, in order to obtain a good description of hadronic masses. Indeed, it was was found on the lattice -- by measuring the string tension after smoothing the gauge fields\footnote{For example, the results of \cite{FIMT_98} can be approximated as $V=K |R-R_0| \theta(R-R_0)$ with the standard string tension $K\simeq 1$~GeV/fm but a rather large $R_0\simeq 0.7$~fm. } -- that the effective potential between constituent quarks is very small at small $r$, but approaches $V(r)=Kr$ at large distances. The motivation for smoothening is the due to the extended nature of constituent quarks, as opposed to essentially point-like heavy flavors ($c,b$). To put it differently: color-strings only form if constituent quarks do not overlap. For such a potential the average energy of colored strings seems to be negligible for the relevant densities considered above. \subsection{Observable signatures} In this section we would like to discuss potential experimental signatures of the quark-diquark and quark superconducting phases, in particular with regard to heavy-ion experiments and neutron stars. Let us start with the quark-diquark phase. It is easy to estimate the critical temperature for the quark-diquark mixture by applying Einstein's ideal gas expression for Bose condensation \begin{eqnarray} T_c=3.31 n_{dq}^{2/3}/M_{dq} \ . \end{eqnarray} Assuming $n_{dq}=3*n_0$ and $M_{dq}=0.6$~GeV we find $T_c=120$~MeV, which is expected to be further reduced by the short range repulsive core. Heavy-ion collisions at SIS/BEVALAC energies (1-2~AGeV), where comparable compression is reached, lead to a heating of the system of up to $T\approx 100$~MeV, and so we conclude that even if this phase exists and our estimates are valid, it can be involved only peripherally. The same is even more true for the color superconducting phase. Because of the high density and low critical temperature of the color superconductor, it is not likely to be produced in heavy-ion collisions. Even if the gap is still larger than expected so that matter in the color superconducting phase would be produced in heavy-ion collisions, its presence will be hard to establish. The two most spectacular manifestations of superconductivity, perfect conductivity and the Meissner effect, are very difficult to detect for a short-lived sample. The transition to the superconducting state has an effect on the equation of state, but the condensation energy $\epsilon\sim\mu \Delta^2 p_F/(2\pi^2)$ is small compared to the energy density of a Fermi gas. In this context, the effect of quark-diquark phase is likely to be larger. Finally, the superconducting phase, in particular color-flavor-locking, may have an effect on the flavor composition. But this effect is not very specific since strangeness enhancement is a general consequence of quark matter formation. For these reasons, compact stars and stellar explosions are probably a more appropriate place to search for observational consequences of quark superconductivity. For a recent review on the structure of neutron stars we refer the reader to Refs.~\cite{HHJ_99,HPS_94}. Owing to theoretical uncertainties in the nuclear equation of state, the central density of neutron stars is not very well known. The main experimental constraint comes from the fact that neutron stars with masses $M_{NS}=1.45M_{Sun}$ have definitely been observed. This still allows for central densities as low as $3n_0$ or as high as $10n_0$. Ultimately, a better handle on the central density will come from measurements of neutron star radii. In neutron star structure calculations quark matter is almost always treated as a simple Fermi gas, confined by a bag constant $B$. Let us only note here that there are two distinct scenarios: (i) For sufficiently small values of $B$ strange quark matter is absolutely stable, and the entire star is in this phase, while (ii) for large $B$ the outer part of the star consists of nuclear matter, while the interior contains various mixed phases (quark matter sheets, rods or clusters). These funny arrangements owe their existence to the possibility to move charge from the quark phase to the hadronic phase. The shape is then determined by the interplay of the equation of state and long range Coulomb forces. Quark superconductivity will again have an influence on the equation of state, but as before we expect this to be a correction on the order of ${\cal O}((\Delta/\mu)^2)$, small compared to the uncertainties in the bag constant. In neutron stars, a more direct measure of the gap is provided by the cooling history of the star. Without a gap, neutron stars can efficiently cool by $\beta$-decay of thermally excited $u,d$ quarks. This process seems to lead to unacceptably large cooling rates~\cite{Iwa_82}. Such a problem emerged already for neutrino emission from nuclear matter. In that case it is solved due to nuclear superfluidity. Since both neutrons and protons are gapped, there are no single-particle states in the vicinity of the Fermi surface. For two-flavor quark matter, a similar problem arises since the quarks of the third color have no or only very small gaps. It is absent in the color-flavor-locked phase, because all quarks acquire a gap. For realistic values of the strange quark mass, the system is likely to be in the two-flavor phase at moderate densities, so that the ``cooling-problem'' for neutron stars might still persist. Another feature of the CSC-phases is that Cooper pairs are electrically charged, so one might expect the color superconductor to be electrically superconducting as well. This is not quite true. For both the $N_f=2$ and color-flavor locked phase, there is a modified charge operator that is not broken, so there is a linear combination of the photon and the diagonal gluons that remains massless. Nevertheless, since the photon inside the superconducting phase is different from the photon outside, magnetic flux will be $partially$ expelled. Magnetic fields in pulsars are very large, up to $10^{12}$ Gauss. Fields on the order of $10^{15}$ Gauss were recently suggested to drive the so called ``magnetars'', and $10^{18}$ gauss is the absolute upper limit allowed by star stability (virial theorem). However, even such fields are not yet large enough to significantly influence the CSC phase. On the other hand, simple scaling considerations show that in order to have the field completely expelled from the quark matter core would cost energy of order ${\cal O}(R^3)$ (where $R$ is the star radius), while transferring the field lines through some channels into the superconductor only costs ${\cal O}(R)$. Inside these channels the field should be at the critical value $B_c\sim 10^{19}$ Gauss. They can be either macroscopic (if the superconductor is of the first kind) or microscopic Abrikosov vortices (if it is of the second kind). \section{Summary and Conclusions} \label{sec_concl} We have studied the interplay of instantons, superfluidity/-conductivity and chiral symmetry breaking in QCD at finite density. Unlike many schematic models based on short-range interactions abstracted from one gluon exchange, the instanton model has the virtue of providing a realistic phenomenology of the zero-density ground state, including such features as spontaneous chiral symmetry breaking and a correct description of (light) hadron spectroscopy. This gives us some confidence for a semi-quantitative investigation of cold quark matter at moderate densities as might be encountered, {\it e.g.}, in the core of neutron stars. We started out by reviewing some properties of the instanton-induced quark-quark interaction at $\mu=0$: it originates from the same effective interaction that leads to chiral condensation and a light pion in the QCD vacuum, predominantly acting in the scalar-isoscalar, color antitriplet $qq$ channel. We found these correlations to be sufficiently strong to generate a bound-state pole in the corresponding diquark propagator. In a second step we studied the density dependence of the instanton-induced (`t Hooft) interaction. It arises through the modification of the quark zero modes as we introduce a chemical potential. Both the density-dependence of the instanton form factors, governing the effective quark-(anti-)quark interactions, and of the instanton-antiinstanton overlap matrix elements, relevant for a statistical treatment of the instanton liquid, have been assessed. The main difference to earlier calculations (using local approximations or schematic formfactors) manifests itself in quantitatively somewhat larger superconducting gaps ($\sim$~150-200~MeV). In the first main part of this article we have performed a systematic study of the finite-$\mu$, $N_c=3$-QCD phase structure for 2, 3 and 2+1 flavors within the mean-field approximation: \begin{itemize} \item[(i)] For two massless flavors we confirmed the usual transition to a $ud$ superconductor at $\mu_c\simeq 300$~MeV with gaps $\Delta\simeq 200$ MeV. Apart from some quantitative deviations the use of microscopic instanton formfactors entailed the possible existence of a novel intermediate phase, namely chirally broken quark-diquark matter (characterized by simultaneous chiral symmetry breaking and diquark condensation). \item[(ii)] For three massless flavors we returned to a simplified treatment based on a idealized sharp formfactor (amenable to the more involved calculations with somewhat less accuracy). We found that, like schematic one-gluon exchange interactions, instantons lead to color-flavor locking, despite their very different color-flavor vertex structure. In addition, there emerged an interesting new feature: in contrast to OGE, instantons generate a non-zero (albeit small) chiral condensate in the superfluid phase. \item[(iii)] The consequences of finite (current) strange quark masses $m_s$ have been investigated. With increasing $m_s$ a sharp phase transition from the color-flavor locked phase to the two-flavor superconductor occurs. At small chemical potential, the critical strange quark mass is smaller than the physical mass. As the chemical potential grows, the critical mass is also expected to grow. We find that the color-flavor locked phase will appear only at $\mu>450-500$~MeV. This implies that chiral symmetry would be restored initially and then broken again at higher density, but with much smaller condensates. \end{itemize} In the second main part of our article a statistical mechanics treatment of the instanton liquid has been employed. This enabled us to incorporate effects that go beyond the standard mean-field approximation, in particular those associated with instanton-antiinstanton molecules. We demonstrated how to handle a complex fermion determinant in this context (using a gluonic interaction that has been averaged over the relative color orientation). Without molecules, the results were shown to be in reasonable agreement with the mean-field analysis in the first part. Including correlations, molecule formation constitutes a $\sim$10\% effect in the free energy/critical chemical potential of the $T$=0-, $N_f=2$-transition. However, with rising temperature $I$-$A$ molecules are expected to play an increasingly important role in the transition between the superconducting and the plasma phases, corresponding to the critical temperature for superconductivity. Furthermore, the statistical mechanics approach allows to assess the $\mu$-dependence of the total instanton density. It turns out that the latter is indeed approximately constant for all $\mu$-values under consideration, which is not really surprising as the gluonic energy within the instanton component carries the dominant fraction of the free energy of the system. This supports the respective assumption made in earlier works as well as in the first part of this article. Based on our results, we conjectured a qualitative picture of the QCD phase diagram in the $T$-$\mu$ plane. We argued that for realistic values of the strange quark mass, there are both finite-$\mu$ and finite-$T$ transitions between the color-flavor locked (CSC3) and two-flavor superconducting (CSC2) states. With increasing density and at small $T$, chiral symmetry is first restored and then (weakly) broken again. We also elucidated on the possibility of a new quark-diquark phase, in which nucleons are dissolved into a Bose gas of $ud$-diquarks and a Fermi gas of unpaired quarks, characterized by simultaneously broken color and chiral symmetry. It represents a natural possibility for an intermediate phase between the hadronic and superconducting phase, although numerical estimates for its existence are rather uncertain. Concerning experimental consequences, we conclude that heavy-ion reactions are unlikely to reach into the rich high-density/low-temperature phase structure discussed here. In this respect neutron stars are much more promising. The pairing gaps should leave their traces in cooling rates and even in the equation of state provided the pairing gaps are large enough. Decomposition of the ``external'' into ``internal'' magnetic fields presumably imply complicated configurations inside the stars. As an outlook, we expect that it should be rather straightforward to generalize our approach to a simultaneous account of finite $\mu$ and $T$. Another major challenge is to go beyond the mean-field approximation in the quark sector in order to address clustering of quarks into nucleons, and, even more difficult, the nucleon-nucleon interaction. The instanton liquid model does provide the correct properties on the one-nucleon level. Turning the instanton liquid into a realistic description of nuclear matter, however, requires a long way to go. These and related issues certainly provide exciting opportunities for future research in the field of finite-density QCD. \vspace{1cm} {\bf ACKNOWLEDGEMENTS} \\ R. R. acknowledges support from the A.-v.-Humboldt foundation as a Feodor-Lynen fellow. T. S. is supported by NSF-PHY-513835. This work is supported in part by US-DOE grants DE-FG02-88ER40388, DE-FG06-90ER40561 and DE-FG02-91-ER40682. \newpage \begin{appendix} \section{Fierz Transformations} \label{fierz} Let us denote a general four-fermion interaction of the type (\ref{l_nf2}) by $(\bar \psi_a {\cal O} \psi_b)(\bar \psi_c {\cal O} \psi_d)$. An effective ``mesonic'' interaction corresponds to the sum of the direct and exchange ($s$+$u$) channels. The corresponding kernel is the sum of the original interaction and its Fierz transform: \begin{eqnarray} "s+u"=(\bar \psi_a {\cal O} \psi_b)(\bar \psi_c {\cal O} \psi_d)+ (\bar \psi_a {\cal O}_1 \psi_d)(\bar \psi_c {\cal O}_1 \psi_d) \ . \end{eqnarray} An effective diquark ($t$-channel) interaction is obtained by first transposing and then performing the Fierz transformation \begin{eqnarray} T=(\bar \psi_a {\cal O} \psi_b)(\psi_d^T {\cal O}^T \bar \psi_c^T)= \epsilon(\bar \psi_a {\cal O} \psi_b)(\psi_d^T C {\cal O} C \bar \psi_c^T)= \epsilon(\bar \psi_a {\cal O}_2 C \bar \psi_c^T)(\psi_d^T C {\cal O}_2 \psi_b) \ . \end{eqnarray} We use Euclidean gamma matrices in the standard or chiral representations, $C=-i \gamma_2\gamma_0, C^2=C^\dagger C=1\!\!1$. If the isospin and color parts of ${\cal O}$ are $1\!\!1$, $\epsilon=-1$ if the Dirac part of ${\cal O}$ is $\gamma_\mu, \sigma_{\mu \nu}$, and $\epsilon=1$ when it is $1\!\!1_D, \gamma_5,\gamma_\mu \gamma_5$. For the flavor part, $\tau_2^T=- \tau_2$ and $\tau_{1,3}^T= \tau_{1,3}$. For the Fierz transformation the following completeness relations are used: \begin{eqnarray} \delta_{ii'}\delta_{jj'}={1\over N_c}\delta_{ij'}\delta_{ji'}+2t^n_{ij'} t^n_{ji'} \end{eqnarray} for color, \begin{eqnarray} \delta_{AA'}\delta_{BB'}&=&{1\over 2}\delta_{AB'}\delta_{BA'}+{1\over 2} \tau^a_{AB'}\tau^a_{BA'} \\ \tau^a_{AA'}\tau^a_{BB'}&=&{3\over 2}\delta_{AB'}\delta_{BA'}-{1\over 2} \tau^a_{AB'}\tau^a_{BA'} \end{eqnarray} for isospin, and, defining $S=1\!\!1\otimes 1\!\!1$,$V=\gamma_\mu \otimes \gamma_\mu$, $P=\gamma_5 \otimes \gamma_5$, $A=\gamma_5\gamma_\mu \otimes\gamma_\mu$,$T=\sigma_{\mu \nu} \otimes\sigma_{\mu \nu}$, with $\sigma_{\mu \nu}=1/2[\gamma_\mu, \gamma_\nu]$, \begin{eqnarray} \pmatrix{S\cr V\cr T \cr A\cr P\cr}'=\pmatrix{1/4&1/4&-1/8&-1/4&1/4\cr 1&-1/2&0&-1/2&-1\cr -3&0&-1/2&0&-3\cr-1&-1/2&0&-1/2&1\cr 1/4&-1/4&-1/8&1/4&1/4\cr}\pmatrix{S\cr V\cr T \cr A\cr P\cr} \ \end{eqnarray} for the Dirac structures. Due to the tensor product structure of the matrices ${\cal O}$, one has to Fierz transform separately the Dirac, color and isospin parts by using the above relations and multiply them accounting for the fact that fermion fields anticommute. The latter generates an additional (-1) in the $u$-channel, and for the $t$-channel it excludes any symmetric parts of the ${\cal O}'$ matrices. The final results are given in the main text, Eqs.~(\ref{l_mes}) and (\ref{l_diq}). \section{Instanton Form Factors at Finite $\mu$} \label{app_ff} The two structures in the form factor are \begin{eqnarray} B&=&4\rho i \int_0^\infty dR R^3 \int_0^\pi d\eta\, \sin^2 \eta{1\over R\sqrt{R^2+\rho^2}} \left[\left(1-2{t^2\over R^2+\rho^2}-\mu t\right) {\sin(\mu r)\over r}\right.\cr & &\hspace{0.5cm}\left. -\left(2{t\over R^2+\rho^2}+\mu\right) \cos(\mu r)\right]{\sin(kr) \over kr}e^{-i\omega t} \end{eqnarray} \begin{eqnarray} A&=&-4\rho\int_0^\infty dR R^3 \int_0^\pi d\eta\, \sin^2 \eta{1\over R\sqrt{R^2+\rho^2}} \left[-\left(\mu+{t\over r^2}+{2t\over R^2+\rho^2}\right) \sin(\mu r)\right.\cr & & \hspace{0.5cm}\left. +\left({\mu t \over r}-{2 r \over R^2+\rho^2}\right) \cos(\mu r)\right] \left({\cos(kr)\over k r}-{\sin(k r) \over k^2 r^2}\right) e^{-i\omega t}. \end{eqnarray} Let us separate the real and imaginary parts of $A$ and $B$: $A=A_c+i A_s, B=B_s+i B_c,$ the subscript referring to $\cos(\omega t)$ or $\sin(\omega t)$. We also introduce $k^\pm=k\pm\mu$. Then \begin{eqnarray} A_c&=&{2 \rho\over k}\int_0^\infty dR \int_0^\pi d\eta {R \sin \eta \cos(\omega R \cos \eta) \over \sqrt{ R^2+\rho^2}}\cr & & \left[ \left( {2 R \sin \eta \over R^2+\rho^2}+{\mu \over k R \sin \eta}\right) \cos(k^+ R \sin \eta)\right.\cr & &\left.+\left(\mu -{2\over k( R^2+\rho^2)}\right)\sin (k^+ R \sin \eta)\right] + (\mu \to -\mu) \end{eqnarray} \begin{eqnarray} A_s&=&{-i 2 \rho\over k}\int_0^\infty dR \int_0^\pi d\eta {R\cos \eta \sin(\omega R \cos \eta) \over \sqrt{ R^2+\rho^2}}\cr & & \left[ \left( {1\over k R^2 \sin^2 \eta}+{2 \over k (R^2+\rho^2)}-\mu\right)\cos(k^+ R \sin \eta) \right.\cr & &\left . +\left({k^+\over kR \sin \eta}+{2 R \sin \eta \over R^2+\rho^2}\right)\sin (k^+ R \sin \eta)\right] - (\mu \to -\mu) \end{eqnarray} \begin{eqnarray} B_c&=&{-i 2 \rho\over k}\int_0^\infty dR \int_0^\pi d\eta {\cos(\omega R \cos \eta)\over \sqrt{ R^2+\rho^2}} \left[ \left( 1-{2 R^2 \cos^2 \eta \over R^2+\rho^2}\right)\right.\cr & &\left . \cos (k^+ R \sin \eta)+\mu R \sin \eta\sin (k^+ R \sin \eta)\right] - (\mu \to -\mu) \end{eqnarray} \begin{eqnarray} B_s&=&{2 \rho\over k}\int_0^\infty dR \int_0^\pi d\eta {R \cos \eta \sin(\omega R \cos \eta)\over \sqrt{R^2+\rho^2}}\cr & & \left[\mu\cos (k^+ R \sin \eta)-{2 R \sin \eta \over R^2+\rho^2}\sin (k^+ R \sin \eta)\right] + (\mu \to -\mu) \ . \end{eqnarray} Using the following basic integrals, \begin{eqnarray} \int_0^\pi d\eta \cos( \alpha \cos \eta)\cos(\beta \sin \eta)= \pi J_0(\sqrt{ \alpha^2+\beta^2}) \end{eqnarray} and, denoting $\omega^\pm=\sqrt{\omega^2+(k^\pm)^2}$, \begin{eqnarray} \int_0^\infty dR{1\over \sqrt{R^2+\rho^2}} J_0(R \omega^\pm)=I_0({\rho \omega^\pm \over 2})K_0({\rho \omega^\pm \over 2}) \equiv I_0^\pm K_0^\pm , \end{eqnarray} one can rewrite the above expressions as derivatives of $I_0^\pm K_0^\pm $ according to \begin{eqnarray} A_c&=&{2 \pi \rho \over k}\left[ {2 \over \rho}{d\over d \rho} \left({d^2 \over d k^2}-{1 \over k}{d \over d k} \right) + \mu\left({1 \over k}- {d \over d k} \right)\right]I_0^+ K_0^+ +(\mu \to -\mu), \\ A_s&=&{i 2 \pi \rho \over k}\left[ {d\over d \omega} \left({2 \over \rho}{d\over d \rho}\left({d \over d k}-{1 \over k}\right)-\mu\right)I_0^+ K_0^+ + \int_0^\mu d \mu' {k^+\over k}{d\over d \omega}I_0^+ K_0^+\right]\cr & & -(\mu \to -\mu) \ . \end{eqnarray} Using ${d \over d \omega}={\omega \over \omega^+}{d \over d \omega^+}$, ${d \over d \mu}={k^+ \over \omega^+}{d \over d \omega^+}$, the integral over $\mu'$ is just ${\omega \over k}I_0^+ K_0^+$. Furthermore, \begin{eqnarray} B_c&=&{i 2 \pi \rho \over k}\left(1-{2 \over \rho}{d\over d \rho}{d^2 \over d \omega^2}-\mu {d \over dk}\right)I_0^+ K_0^+ - (\mu \to -\mu), \\ B_s&=&{- 2 \pi \rho \over k}\left(\mu-{2 \over \rho}{d\over d \rho}{d \over dk} \right) {d\over d \omega}I_0^+ K_0^+ +(\mu \to -\mu). \end{eqnarray} Denoting ${d\over dz}I_0^+ K_0^+=I_1^+ K_0^+-I_0^+ K_1^+\equiv \Delta^+$, we get \begin{eqnarray} A_c&=&{2 \pi \rho \over k}\left[ \left({\mu \over k}+{2 (k^+)^2 \over (\omega^+)^2}\right)I_1^+ K_1+ {\rho k^+ \over 2 \omega^+}(2k+\mu) \Delta^+\right]+(\mu \to -\mu), \\ A_s&=&{i2 \pi \rho \over k}\left[ \left({\omega \over k}+{2\omega k^+\over (\omega^+)^2}\right)I_1^+ K_1+ {\rho \omega \over 2 \omega^+}(2k+\mu) \Delta^+\right]-(\mu \to -\mu), \\ B_c&=&{i 2 \pi \rho \over k}\left[{\omega^2- (k^+)^2 \over (\omega^+)^2}I_1^+ K_1+{\rho \over 2 \omega^+}(2\omega^2+\mu k^+) \Delta^+\right]-(\mu \to -\mu), \\ B_s&=&{2 \pi \rho \over k}\left[{2\omega k^+\over (\omega^+)^2}I_1^+ K_1+ {\rho \omega \over 2 \omega^+}(2k+\mu) \Delta^+\right]+(\mu \to -\mu). \end{eqnarray} These Fourier transforms of the fermion zero modes at finite $\mu$ agree with the results obtained in \cite{CD_99}. \section{Grand Canonical Potential in the Cornwall-Jackiw-Tomboulis (CJT) Formalism} \label{app_CJT} To derive the grand canonical potential in MFA we start from a generating functional with bilocal meson and diquark sources, \begin{eqnarray} \exp(W[J,\bar J,K,\bar K])&=&\int {\cal D}\psi{\cal D}\bar \psi \exp\left\{(S_0+U+\int\bar\psi J\psi +\int\psi^T \bar J \bar\psi^T+\int\bar\psi \bar K\bar\psi^T + \int\psi^T K\psi)\right\} \ , \label{Z_source} \end{eqnarray} where $S_0$ is the free fermion action and $U=\int({\cal L}_{mes}+{\cal L}_{diq})$ with ${\cal L}_{mes}, {\cal L}_{diq}$ given in Eqs.~(\ref{l_mes},\ref{l_diq}). Using a matrix representation, we write the free part of the action in momentum space as \begin{eqnarray} S_0&=&\int {d^4p\over (2 \pi)^4}\left[ \bar\psi(p)(J(p)+{1 \over 2}G_0^{-1}(p))\psi(p) +\psi^T(-p)(\bar J(p) \right. \nonumber\\ && \qquad \qquad \left. +{1 \over 2}G_0^{-1T}(-p))\bar\psi^T(-p)+ \bar\psi(p)\bar K(p)\bar\psi^T(-p) +\psi^T(-p) K\psi(p)\right] \nonumber\\ &=& \int {d^4p\over (2 \pi)^4} \pmatrix{\bar\psi(p),&\psi^T(-p)\cr}\pmatrix{ \bar K(p) &J(p)+{1 \over 2}G_0^{-1}(p) \cr \bar J(p)-{1 \over 2}G_0^{-1T}(-p) & K(p)\cr}\pmatrix{\bar\psi^T(-p) \cr \psi(p)\cr}, \label{S_source} \end{eqnarray} where the sources $J,\bar J, K, \bar K $ have the following properties: \begin{eqnarray} \label{sour_sym} -\bar J^T(-p)=J(p)\nonumber \\ -\bar K^T(-p)=\bar K(p)\nonumber \\ - K^T(-p)= K(p) \ . \end{eqnarray} The path integral over the fermion fields is \begin{eqnarray} e^{W_0}&=&\int {\cal D}\psi{\cal D}\bar \psi e^S={\det} ^{1\over2}\pmatrix{ \bar K &J+{1\over 2}G_0^{-1}\cr \bar J- {1\over 2}G_0^{-1T}& K\cr}\equiv \det {\cal M} \nonumber \\ W_0&=&{1 \over 2} \rm tr \ln {\cal M} \ , \end{eqnarray} where the trace includes the momentum integration, and the products between the fields (sources) include convolutions in momentum as well as all other indices. Using the identities \begin{eqnarray} \pmatrix{ \bar A & B \cr \bar B & A\cr}&=&\pmatrix{ 0 & B \cr \bar B & 0\cr} \pmatrix{ 1\!\!1 &\bar B^{-1} A \cr B^{-1}\bar A & 1\!\!1\cr}, \nonumber\\ \ln \det \pmatrix{ \bar A & B \cr \bar B & A\cr}&=&\ln \det(-B\bar B)+\rm tr \ln(1+\pmatrix{ 0 &\bar B^{-1} A \cr B^{-1}\bar A & 0\cr} \nonumber\\ &=&\ln \det(-B\bar B)+\rm tr \sum_{n=1}^{\infty}{(-1)^{n-1}\over n }\pmatrix{ 0 &\bar B^{-1} A \cr B^{-1}\bar A & 0\cr}^n \nonumber\\ &=&\ln \det(-B\bar B)+\rm tr \sum_{n=1}^{\infty}-{1\over 2n}(\bar B^{-1} AB^{-1}\bar A +B^{-1}\bar A\bar B^{-1} A) \nonumber\\ &=& \ln \det(-B\bar B)+\rm tr \ln(1-\bar B^{-1} AB^{-1}\bar A) \nonumber\\ &=&\rm tr \ln(-B\bar B+BAB^{-1}\bar A) \ , \label{det_trick} \end{eqnarray} we obtain \begin{eqnarray} W_0={1\over 2}\rm tr \ln[-(J+{1\over 2}G_0^{-1})(\bar J- {1\over 2}G_0^{-1T})+ (J+{1\over 2}G_0^{-1})K(J+{1\over 2}G_0^{-1})^{-1}\bar K] \ . \end{eqnarray} Next, one introduces the classical two-point fields $\bar F= {\delta W \over \delta \bar K}$, $F= {\delta W \over \delta K}$, $G= {\delta W \over \delta J}$, $\bar G= {\delta W \over \delta \bar J}$ and performs a Legendre transformation \begin{eqnarray} \Gamma[G, \bar G, F,\bar F] =W[J,\bar J,K,\bar K] -\rm tr(GJ+\bar G \bar J +KF+\bar K\bar F) \end{eqnarray} with $J,\bar J,K,\bar K$ expressed as functionals of $G, \bar G, F,\bar F$. From the non-interacting part $W_0$ we have \begin{eqnarray} G&=&{1\over2}( J+{1\over 2}G_0^{-1} -\bar K(\bar J-{1\over 2}G_0^{-1T})^{-1} K)^{-1 } \nonumber\\ \bar G&=&{1\over2}( \bar J-{1\over 2}G_0^{-1T} - K(J+{1\over 2}G_0^{-1})^{-1} \bar K)^{-1 } \nonumber\\ F&=&- (J+{1\over 2}G_0^{-1})^{-1}\bar K \bar G \nonumber\\ \bar F&=&-(\bar J-{1\over 2}G_0^{-1T})^{-1} K)^{-1 }KG \ , \end{eqnarray} and \begin{eqnarray} \label{Gamma_0} \Gamma_0=-{1 \over 2}\rm tr \ln (-\bar G G+\bar G F \bar G^{-1}\bar F)+ {1 \over 2}\rm tr(G_0^{-1} G - G_0^{-1T} \bar G-2) \ . \end{eqnarray} Following the CJT approach~\cite{CJT} for the four-fermion interaction, one can show that \begin{eqnarray} \label{gamma_expansion} \Gamma=\Gamma_0+{1\over 4!}\hbar^2 \left(\rm tr(G{\delta^2 \over \delta \psi \delta \bar \psi}) S_{int} \rm tr({\stackrel{\gets}{\delta^2}\over \delta \psi \delta \bar \psi}\bar G) + \rm tr(F{\delta^2 \over \delta \bar\psi^T \delta \bar \psi}) S_{int}\rm tr({\stackrel{\gets}{\delta^2} \over \delta \psi \delta \psi^T}\bar F)\right) + \Gamma_4 \ , \end{eqnarray} where $S_{int}$ is the original four-fermion interaction and we have explicitly indicated the dependence on $\hbar$. Now it becomes clear that, in order to perform the functional derivatives, it is convenient to Fierz-rearrange $S_{int}$ into the 3 channels $s$, $t$ and $u$ as outlined above. The sum of the four derivatives in Eq.~(\ref{gamma_expansion}) naturally suggests the use of $U$ in Eq.~(\ref{Z_source}) with two-by-two derivatives {\it w.r.t.}~the fermion fields that are contracted with the same matrix ${\cal O}^i$. The Hartree-Fock scheme is equivalent to (i) neglecting $\Gamma_4$, which is the sum of all four-particle irreducible diagrams and is of order $O(\hbar^4)$, and (ii) considering only translationally invariant solutions ({\it e.g.}, $G(p,p')=G(p)\delta^4(p-p')$, and the same for $F, \bar F$). The mesonic and the diquark terms in the above lowest-order interaction term have the form $V_4({\bf M}_i)({\bf\bar M}_i)$, $V_4 ({\bf D}_i)({\bf\bar D}_i)$, where \begin{eqnarray} {\bf M}_i&=&\int {d^4p\over (2 \pi)^4} \rm tr (G {g^{mes}_i}{\cal F}^\dagger {\cal O}^i {\cal F})\nonumber \\ {\bf D}_i&=&\int {d^4p\over (2 \pi)^4}\rm tr( F^\dagger {g^{diq}_i}{\cal F}^T {\cal O}^i C {\cal F}^*) \ , \end{eqnarray} {\it i.e.}, they are products of mesonic and diquark condensates. A priori one might consider condensation in any channel, and then study if it is energetically favored. It is reasonable to start with the channels that are most attractive ({\it e.g.}, those which generate bound states in vacuum). As we have seen in sects.~\ref{sec_L_eff} and \ref{sec_diq}, these are the scalar color singlet mesonic channel and the diquark color-${\bf \bar 3}$ one. For the $N_f=2$ case there is only one pattern of diquark condensation, which necessarily breaks the color symmetry down from $SU(3)_C\to SU(2)_C$. The diquark condensate is proportional to a unit vector in $SU(3)_C$. Moreover, because the flavor part of the vertex is antisymmetric (due to $\tau_2$), the Pauli principle requires that the above unit vector belongs to the antisymmetric part of $SU(3)_C$. Without loss of generality we can choose this vector to be the Gell-Man matrix $\lambda_2$ (recall that all Gell-Mann matrices in this paper are normalized to 3, {\it i.e.}, $\rm tr \lambda_i^2 =3$). We can identify the unbroken $SU(2)_C$ with the upper $2\times 2$ corner of the $SU(3)_C$ group. If both $qq$ and $\bar qq$ condensates are present, the latter should consist of two parts -- one that involves the two colors from the unbroken $SU(2)_C$ and the second one which involves quarks and antiquarks from the third color. One should note that there are two terms in ${\cal L}_{mes}$ contributing to the $\bar qq$ condensates: the isoscalar color singlet one and the isoscalar $SU(3)_C$ octet one proportional to $\lambda_8$. Both have projections onto the $SU(2)_C$ singlet as well as onto the subgroup represented by third color of (unpaired) quarks. Owing to the two different chiral condensates, it is convenient to split the propagators $G, \bar G$ into a $SU(2)_C$ part $G_1, \bar G_1$ and a part involving the third color only, $G_2, \bar G_2$. The $F$ and $\bar F$ propagators are proportional to the diquark condensate and hence to $\lambda_2$. Putting everything together, Eq.~(\ref{gamma_expansion}) becomes \begin{eqnarray} \label{Gamma} \Gamma&=&-{1 \over 2}\rm tr \ln (-\bar G_1 G_1+\bar G_1 F \bar G_1^{-1}\bar F)+ {1 \over 2}\rm tr(G_0^{-1} G_1 - G_0^{-1 T} \bar G_1-2)\nonumber \\ & & -{1 \over 2}\rm tr \ln (-\bar G_2 G)+ {1 \over 2}\rm tr(G_0^{-1} G_2 - G_0^{-1 T} \bar G_2-2)\nonumber \\ & &+g{1\over 8 N_c^2}|\rm tr(G_1+G_2)\alpha|^2+g {N_c-2\over 16 N_c^2 (N_c^2-1)}|\rm tr \lambda_8(G_1+G_2)\alpha|^2\nonumber \\ & &+g {1\over 8 N_c^2 (N_c-1)}\rm tr(FC \gamma_5 \lambda_2 \tau_2 \beta)\rm tr(\beta^*C \gamma_5 \lambda_2 \tau_2 \bar F), \end{eqnarray} where $\alpha$ and $\beta$ are the formfactors defined in Eqs.~(\ref{alpha}),(\ref{beta}). Since $G,\bar G, F, \bar F$ are classical fields, $\Gamma$ should vanish under the respective variations. \section{Gap-Induced Damping of Single-Quark Propagators} \label{sec_damp} In the superconducting phases, single-quark propagation in the vicinity of the Fermi surface is damped in the temporal direction due to the presence of the finite energy gap. This effect has to be included in the evaluation of the $I$-$A$ overlap matrix elements, Eq.~(\ref{tiamu}), at finite density (this is particularly crucial for the activity $z_d$ of Eq.~(\ref{zd}), which otherwise would diverge). Rewriting the overlap matrix elements as \begin{eqnarray} T_{IA} = - \int d^4x \ \bigl[\phi_I^\dagger(x-z_I;-\mu)(i\not\!\partial-i\mu\gamma_4)\bigr] \ (i\not\!\partial-i\mu\gamma_4)^{-1} \ \bigl[(i\not\!\partial-i\mu\gamma_4) \Psi_{0,A}(x-z_{A};\mu)\bigr] \ , \end{eqnarray} they are readily interpreted as a quark hopping amplitude, represented by two amputated (anti-) instanton vertices and an intermediate (free) quark propagator $(i\not\!\partial-i\mu\gamma_4)^{-1}$. In the following we will evaluate an approximate damping factor for this propagator. Starting from the expression of the (massless) momentum space propagator at finite $\mu$ in the superconducting phase (ignoring the Dirac structure) \begin{eqnarray} G(p,\mu,\Delta) & = & \frac{p_0+\xi}{p_0^2-\xi^2-\Delta^2+i\eta} \nonumber\\ & = & \frac{u_p^2}{p_0-\epsilon(p)+i\eta} + \frac{v_p^2}{p_0+\epsilon(p)-i\eta} \end{eqnarray} ($\xi=\omega_p-\mu, \epsilon(p)^2=\xi^2+\Delta^2, u_p^2=\frac{1}{2}[1+\xi/\epsilon(p)], v_p^2=\frac{1}{2}[1-\xi/\epsilon(p)]$), we compute its Fourier transform as \begin{eqnarray} G(z;\mu,\Delta) & = & \int \frac{d^4p}{(2\pi)^4} e^{-ipz} G(p,\mu,\Delta)\nonumber\\ & = & \frac{1}{2\pi^2r} \int\limits_{p_F}^{\infty} p \ dp \ \sin(pr) \ e^{-\epsilon(p) z_4} \ u_p^2 \ . \end{eqnarray} For $\Delta\to 0$ one recovers the result~\cite{Sch_98} \begin{eqnarray} G(z;\mu)=\frac{1}{2\pi^2 z^4} \left[ (2z_4+\mu z^2) \cos(\mu r) +(z_4^2-r^2+\mu z_4 z^2) \frac{\sin(\mu r)}{r} \right] \ . \end{eqnarray} An approximate correction to $T_{IA}(z;\mu)$ is thus obtained by supplying it with the ratio \begin{eqnarray} R(z_4;\mu,\Delta)\equiv \frac{G(z_4;\mu,\Delta)}{G(z_4;\mu)} \ , \label{Rdelta} \end{eqnarray} where we have restricted the {\it l.h.s}~of Eq.~(\ref{Rdelta}) to $r$=0 to avoid artificial singularities caused by oscillations in $G(z_4,r;\mu)$. However, since (for $N_f=2$) only two out of three quarks at the Fermi surface can participate in the diquark condensate, the correction (\ref{Rdelta}) enters on average with a smaller power, {\it i.e.}, \begin{eqnarray} T_{IA}(z;\mu,\Delta)\simeq R(z_4;\mu,\Delta)^{2/3} \ T_{IA}(z;\mu) \ . \end{eqnarray} The net effect of $R(z_4;\mu,\Delta)$ is a damping of the zero mode propagation, which, in fact, is much less pronounced than the naive expectation, $\propto e^{-\Delta z_4}$, would suggest. On the other hand, $\Omega_{inst}$ {\it disfavors} finite values for $\Delta$. \end{appendix}
\section{Introduction} Brill-Noether theory is concerned with the study of the subvarieties of the moduli space of stable bundles, determined by bundles having at least a specified number of sections. More precisely, if $\mo$ is the moduli space of stable vector bundles of rank $n$ and degree $d$ over a non-singular algebraic curve $X$ of genus $g\geq 2$ over {\bf{C}}, and $k\geq 1$, the corresponding {\it Brill-Noether locus} is $$\bn := \{ E\in \mo | h^0(E)\geq k \}. $$ The main questions in Brill-Noether theory regard the nonemptiness, dimension, connectedness, irreducibility, cohomology classes, etc., of these varieties. (Similar statements can be made for semistable bundles.) For line bundles, Brill-Noether theory has been studied since the last century, and for a generic curve the basic questions have been answered (see \cite{ACGH}). However, the corresponding theory for vector bundles of higher rank is far from being complete even for the generic case. (In section 2 we will recall the known results for this case.) In this paper we will be concerned with the nonemptiness question. We will prove (see Corollary 3.2 and Theorem 3.9) that {\it if $d=nd'+d''$, $0<d''<2n$, and $d' \ge0 $ then, {\it for any} $X$, $$\bn \neq \emptyset \ \ \ \ \ \text{if} \ \ \ \ n\leq d''+(n-k)g\text{ and } (d'',k)\ne(n,n).\eqno(A)$$} More generally, {\it if $1\le s\le g$ and there exists a line bundle $L$ on $X$ of degree $d'$ with $h^0(L)\ge s$, then $${\cal W}^{sk-1}_{n,d} \neq \emptyset \ \ \ \text{if} \ \ \ 0<d''<2n,\ n\leq d''+(n-k)g \text{ and } (d'',k)\ne(n,n).\eqno(B)$$ } Observe that such a line bundle $L$ always exists if $d'\geq (s-1)(s+g)/s$. If $X$ is hyperelliptic, we see by considering powers of the hyperelliptic line bundle that $L$ exists if $d' \geq 2s-2.$ To get a clear picture of the triples $(n,d,k)$ for which $\bn$ is not empty, as Alastair King has pointed out, it is easier to represent this kind of result if we write $\mu=d/n$, $\lambda=k/n$ and plot points in the $\pa$-plane; we will refer to this representation as the {\it Brill-Noether map} (or BN map). It will be convenient to call a point $\pa\in{\bold Q}^2$ a {\it $n$-Brill-Noether} point (or $n$-BN point), $n\in\Bbb N$, if $d=n\mu$ and $k=n\lambda$ are both integers and $\bn\ne\emptyset$. If $\pa$ is $n$-BN for all $n$ such that $d=n\mu$ and $k=n\lambda$ are both integers, we just say it is a BN point. The results $(A)$ and $(B)$ define a region in the $\pa$-plane, which we denote by $BMNO$, where the points are $n$-BN for ``many values'' of $n$, and thus the corresponding Brill-Noether loci are non-empty (for an explanation of what we mean by ``many values'' see Remarks 4.3 and 4.4). Actually, in the hyperelliptic case we give a precise description of which points are $n$-BN. The results, in this case, come close to a complete solution of the nonemptiness problem. In particular, the boundary of the region in which stable bundles of rank $>1$ can exist is completely determined, and, as might be expected, it is close to the Clifford line. On the other hand, the results in \cite{Te1}, later refined in \cite{M2}, show that one can define a polygonal region $T$, the so-called ``Teixidor's parallelograms," such that all the points $\pa$ inside $T$ are BN, except perhaps at certain vertices. The region $BMNO$ covers a large part of $T$, but more importantly it extends beyond $T$. Our methods, and especially Theorem 3.9, give stronger results for special curves. Furthermore, since we do not use the results of \cite{Te1} and \cite{M2}, our results give another proof of nonemptiness for those parts of $T$ which are included in $BMNO$. \vskip 6pt The paper is organized as follows: In section 2 we give a brief survey of what is known about $\bn$. In section 3, we prove assertions $(A)$ and $(B)$. In section 4 we describe the region $BMNO$ in the $\pa$-plane for the general case. In section 5 we compare the regions $BMNO$ and $T$. In section 6 we study the case where $X$ is hyperelliptic. \section{A survey of the known theory} In this section we recall the known results of Brill-Noether theory for vector bundles of higher rank (see also \cite{BGN} and \cite{M1}), and thereby also fix notations. We will use the Brill-Noether map to indicate the regions where not only the nonemptiness is known but also some of the topology of $\bn$. \bigskip Let $X$ be a non-singular algebraic curve of genus $g\geq 2$ over {\bf{C}} and $\mo$ the moduli space of stable vector bundles over $X$ of rank $n$ and degree $d$. We define the Brill-Noether loci for $(n,d,k)$, $$\bn := \{ E\in \mo | h^0(E)\geq k \} $$ as in the introduction. Denoting by $\wmo$ the moduli space of equivalence classes of semistable bundles over $X$ of rank $n$ and degree $d$, we can define similarly Brill-Noether loci $\wbn$ in $\wmo$. In what follows we will concentrate on stable bundles, but we will also explicitly indicate where semistable bundles are allowed. Since for $k>0,d<0$, $\bn =\emptyset$ and for $ k\leq 0, \ \bn $ is the whole moduli space $\mo $, we will assume that $d \geq 0$ and $k \geq 1.$ It follows from the theory of determinantal varieties that every non-empty component of $\bn$ has dimension greater than or equal to the Brill-Noether number $$\bnn := n^2(g-1)+1-k(k-d+n(g-1)),$$ and for generic $X$ this number is the expected dimension of $\bn$. However, there is no similar formula for the expected dimension of $\wbn$. For $n=1$, it is a classical result (see \cite{ACGH}) that $\bnl\ne\emptyset$ if $ \bnnl\geq 0$; moreover, for a generic curve the converse is also true. However, for $n\geq 2$, it is known (see \cite{BGN}) that $\bnn \geq 0$ does not imply that $\bn\ne\emptyset$. A point $\pa\in{\bold Q}^2$ will be called a {\it $n$-Brill-Noether} point (or $n$-BN point), $n\in\Bbb N$, if $d=n\mu$ and $k=n\lambda$ are both integers and $\bn\ne\emptyset$. If it is $n$-BN for all $n$ such that $d=n\mu$ and $k=n\lambda$ are both integers we just say it is a BN point. In the Brill-Noether map, using the Riemann-Roch Theorem and Clifford's Theorem, one can define a region such that outside this region the problem becomes trivial, in the sense that $\bn$ is either empty or the whole moduli space $\mo$. More precisely, consider the following lines: \vskip 4pt {\it i\/}) $\mu =\lambda + g-1$ \ \ \ (Riemann-Roch line) \vskip 4pt {\it ii\/}) $\mu = 2\lambda -2 $ \ \ \ \ (Clifford line ) \vskip 4pt These lines, together with the positive axes and the line $\mu=2g-2$, define a bounded pentagonal region, which we denote by $P$. That is, we define $P$ to be the region defined by the inequalities $$\mu<\lambda+g-1,\ \mu\geq 2\lambda-2,\ 0\leq\mu\leq2g-2,\ \lambda>0$$ (see Figure 1). Below and to the right of $P$, Riemann-Roch implies that $\bn$ is the whole space. Above and to the left of $P$, Riemann-Roch and Clifford's Theorem, together with the definition of stability, imply that $\bn$ is empty. Thus, we are interested in studying only the points inside $P$. \begin{rem}\begin{em} {\it i)} For $\mu=0$, the only stable bundle in $P$ is the trivial line bundle $\cal O$ at the point $(0,1)$, while for $\mu=2g-2$, the only such bundle is the canonical line bundle $K$ at the point $(2g-2,g)$. However semistable bundles exist at all points of these two edges of $P$ \cite{BGN}. Note that, according to our definition, the points $(0,1)$ and $(2g-2,g)$ are only $1$-BN. \noindent {\it ii)} The inequalities defining $P$ are all sharp except for the Clifford bound. The exact r\^ole of this bound is not clear but it has, for example, been improved by Re (see \cite{R}) for non-hyperelliptic curves. In this case, if we restrict to the range $1\leq\mu\leq2g-3$, we can replace the Clifford line by the line $\mu =2\lambda -1$. For further improvements, see for example \cite{BN} and \cite{M4}. \end{em}\end{rem} By Serre duality we know that, if $\pa$ is BN, then so is $$\sigma\pa:=(2g-2-\mu,\lambda +g-1-\mu).$$ Though it is not readily apparent from Figure 1, this gives a symmetry in $P$ through the line $\mu = g-1$. For later purposes, it will be convenient to write $$R=\{\pa\in P: \mu\le g-1\},$$ so that in particular $P=R\cup\sigma(R)$ (see Figure 1). An important feature of the BN map is the curve defined by the equation $$\widetilde{\rho}=\frac{1}{n^2}(\bnn -1) =0,$$ called the {\it Brill-Noether curve} (or BN curve). From what was said earlier, this represents the boundary of the region where one would expect the Brill-Noether loci to have positive dimension, though it is known that this analogy to the case of line bundles is not valid in general (see \cite{BGN}, \cite{M1}, etc.). The BN curve is a portion of a hyperbola, with equation $$\widetilde{\rho}\pa =(g-1)-\lambda (\lambda -\mu +g-1)=0.$$ The results of \cite{Te1} and \cite{M2} allow us to define a polygonal region which we denote by $T$, contained in the interesting region $P$ in the $\pa$-plane, such that any point $\pa$ in $T$ is BN except possibly for certain vertices. This region was described in detail in \cite{BGN} and \cite{M1} in its original form, and we will recall its construction in section 4 incorporating the results of \cite{M2}. For the time being, we just point out that $T$ has sides parallel to the lines $\lambda =0,\mu =\lambda$, and vertices at points with integer coordinates, on or below the BN curve (see Figure 2). The most significant results for our purposes are the following, which hold for slopes restricted to $0\leq\mu < 2$: \begin{enumerate} \item{} For $0< \mu \leq 1$, Brambila-Paz, Grzegorczyk and Newstead proved in \cite{BGN} that $\pa$ is BN if and only if $1\leq\mu+ (1-\lambda )g $ and $(\mu,\lambda)\ne(1,1)$. \item{} For $1< \mu <2$, Mercat in \cite{M1} proved that $\pa$ is BN if and only if $1\leq \mu+ (1-\lambda )g $. \end{enumerate} In the BN map these results define two trapezoidal regions inside $R$, which we will denote by $BGN$ and $M$ respectively (see Figure 3). Ballico, Mercat and Newstead have recently proved the existence of stable bundles at some points outside the regions defined above; in particular, these bundles can have negative Brill-Noether number \cite{BMN}. For $X$ generic, Teixidor also proved that $\bn$ has an irreducible component of dimension $\bnn$. For any curve, Brambila-Paz, Grzegorczyk and Newstead in \cite{BGN} also proved that for $ 0<\mu \leq 1$, if $\bn$ is nonempty, then it is irreducible of dimension $\bnn$, and $\text{Sing\,}\bn ={\cal W}^k_{n,d}$ (\cite{BGN}, Theorem A). For $1<\mu <2$, Mercat in \cite{M1} also proved that, if $n=d+(n-k)g $ or $n<d<n+g$, $\bn $ is irreducible (\cite{M1}, 2-B-1 and 3-A-1); in any case all components have the expected dimension and $\text{Sing\,}\bn ={\cal W}^k_{n,d}$ (\cite{M1}, 2-C-1). So, for slopes $0\leq \mu < 2$, the results are very complete. For $k=1$, Sundaram \cite{Su} proved that ${\cal W}^0_{n,d}$ is irreducible of dimension $\rho^0_{n,d}$, and Laumon \cite{L} showed $\text{Sing\,}{\cal W}^0_{n,d}$= ${\cal W}^1_{n,d}$. For rank $2$ there are also results of nonemptiness and irreducibility in \cite{T}, \cite{Te2} and \cite{Su} and for rank $3$ in \cite{NB}. \begin{rem}\begin{em} If $X$ is not hyperelliptic, the results of \cite{M1} can be extended to cover the case $\mu=2$ \cite{M3}. For further details, see Remark 4.8. \end{em}\end{rem} >From the results of \cite{BGN} and the symmetry of the region $P$, for $g=2$ one has a complete description of Brill-Noether loci: nonemptiness, irreducibility and singularities. \section{ Nonemptiness of Brill-Noether loci} In this section we will prove assertions $(A)$ and $(B)$. \smallskip Note first that, for any line bundle $L$ of degree $d'$, the formula $E\mapsto E\otimes L $ defines an isomorphism $$ \Phi _L :{\cal M}(n,d) \rightarrow {\cal M}(n,nd'+d). \eqno(1)$$ We shall make repeated use of this idea of tensoring stable bundles by line bundles. If $V$ is a subvariety of $\mo$, then $\Phi _L(V)$ is a subvariety of ${\cal M}(n,nd'+d)$; in particular, if $V$ is a non-empty Brill-Noether locus and $h^0(L)>0$, then $\Phi _L(V)$ will meet certain Brill-Noether loci in ${\cal M}(n,nd'+d)$, which will therefore be non-empty. Actually, if $d'\ge0 $ then there always exists a line bundle $L$ of degree $d'$ that has at least one section, so we have the following theorem. \begin{thm} If $\bn \neq \emptyset$ (respectively $\wbn\neq\emptyset$), then ${\cal W}^{k-1}_{n,d+nd'}\neq \emptyset$ (respectively $\widetilde{\cal W}^{k-1}_{n,d+nd'}\neq\emptyset$) for any $d'\ge0$. \end{thm} \smallskip {\it Proof:} Choose a line bundle $L$ of degree $d'$ with $h^0(L) >0$. Then, for any $E$ with $h^0(E)\ge k$ we have $h^0(E\otimes L)\ge k$. \bigskip \begin{cor}{\bf (Assertion (A))} Suppose $d=nd'+d''$ with $0<d''<2n$, $d'\ge0$, $n\leq d''+(n-k)g$ and $(d'',k)\ne(n,n)$. Then $\bn\ne\emptyset$. \end{cor} \smallskip {\it Proof:} This follows from the results of \cite{BGN} and \cite{M1} and Theorem 3.1. \bigskip \begin{cor} If $k<n$, then ${\cal W}^{k-1}_{n,rn} \neq \emptyset $ for $r\geq 1$. \end{cor} \smallskip {\it Proof:} Take $d''=n$ and $d'=r-1$ in Corollary 3.2. \bigskip In general, the multiplication map $$ \mu _{E,F} : H^0(E)\otimes H^0(F) \rightarrow H^0(E\otimes F) $$ is not injective and $h^0(E)^.h^0(F)$ does not give a lower bound for $h^0(E\otimes F)$. However, if $E$ is a point in $\bn$ with $d/n <2$ and $L$ a line bundle of degree $d'\geq0 $ with at least $s$ independent sections, then we have the following lemmas. \bigskip \begin{lemma} If $d<n+g$, then $h^0(E\otimes L) \geq ks.$ \end{lemma} \smallskip {\it Proof:} From \cite{BGN} (for $d\le n$) and \cite{M1}, 3-A-1 (for $n<d<n+g$) we know that any such bundle has ${\cal O}^k $ as a subsheaf. Hence, $\oplus ^kL $ is a subsheaf of $E\otimes L$. Therefore $h^0(E\otimes L) \geq h^0(\oplus ^kL) \geq ks.$ \bigskip \begin{rem}\begin{em} Note that the existence of $E$ implies that $n\le d+(n-k)g$, so the hypothesis $d<n+g$ implies that $k\le n$. \end{em}\end{rem} \bigskip \begin{lemma} If $k>n$, $d=n +(k-n)g$ and $d'\le2g$, then $h^0(E\otimes L) \geq ks.$ \end{lemma} \smallskip {\it Proof:} From \cite{M1}, 2-B-1 we know that any such bundle fits in an exact sequence $$0\rightarrow F^* \rightarrow {\cal O} \otimes H^0(X,F)^* \rightarrow E\rightarrow 0\eqno(2)$$ where $F$ is a stable bundle of slope $>2g$ and $h^0(F)=h^0(E)=k$. Tensor $(2)$ by $L$ and take the cohomology sequence. Since $\deg(F^*\otimes L)<0$, we have $H^0(X,F^*\otimes L) =0$ and hence $$h^0(E\otimes L) \geq h^0(L)^.h^0(F) \geq ks.$$ \bigskip \begin{lemma} If $n+g\le d<2n$ and $d'\le2g$, then there exists $E\in\bn$ with $h^0(E\otimes L) \geq ks.$ \end{lemma} \smallskip {\it Proof:} In this case we have two exact sequences $$0\rightarrow {\cal O}^{l'} \rightarrow E'\rightarrow E\rightarrow 0\eqno(3)$$ $$0\rightarrow D(E')^* \rightarrow {\cal O}^{n+l+l'} \rightarrow E' \rightarrow 0\eqno(4)$$ where $k\le n+l$, $E\in{\cal W}_{n,d}^{n+l-1}$, $E'$ and $D(E')$ are stable and $\mu(D(E'))>2g$ (see \cite{M1}, 3-B-1 and its proof). Tensor both sequences by $L$ and take the cohomology sequences. Since $H^0(X,L\otimes D(E')^*)=0$, $$h^0(E'\otimes L) \geq h^0(\oplus ^{n+l+l'}L)= (n+l+l')h^0(L)$$ Thus $$h^0(E\otimes L)\ge(n+l)h^0(L) \geq ks.$$ \bigskip \begin{rem}\begin{em}{\it i)} From the Brill-Noether theory for line bundles we know that, if $d'\geq\parti:= (s-1)(s+g)/s$, then there exists a line bundle $L$ of degree $d'$ with at least $s$ independent sections. Actually, ${\cal W}^{s-1}_{1,d'}$ is the variety of such bundles and, for $X$ generic, it has dimension $g-s(s-d'+g-1)=s(d'-\eta(s))$. \noindent {\it ii)} If $X$ is hyperelliptic, we see by considering powers of the hyperelliptic line bundle that $L$ exists if $d' \geq 2s-2.$ \end{em}\end{rem} We deduce the following theorem that proves assertion ($B$): \bigskip \begin{thm} Suppose $d=nd'+d''$ with $d'\ge0$, $0<d''<2n$ and that $1\le s\le g$. If $n\leq d'' +(n-k)g$ and $(d'',k)\ne(n,n)$ and there exists a line bundle $L$ on $X$ of degree $d'$ with $h^0(L)\ge s$, then $\bns \neq \emptyset$. \end{thm} \smallskip {\it Proof:} Note first that, for fixed $n$, $k$, $d''$, $s$, if the theorem is true for one value of $d'$, it is true for all larger values. Since $\eta$ is an increasing function of $s$ and $\eta(g)=2g-2$, it is therefore sufficient by Remark 3.8 to prove the theorem with the additional hypothesis that $d'\le2g-2$. Now, from \cite{BGN} and \cite{M1} we know that under the given hypotheses $\bnp$ is non-empty. It follows from the lemmas that $\bns $ is non-empty. \bigskip \begin{rem}\begin{em} In the semistable case, the theorem can be extended to the cases $d''=0$, $k\le n$ and $(d'',k)=(n,n)$. \end{em}\end{rem} \bigskip \begin{cor} If $k<n$ and $r\geq s+g-g/s $, then ${\cal W}^{sk-1}_{n,rn}\neq\emptyset.$ \end{cor} \smallskip {\it Proof:} Take $d''=n$ and $d'=r-1$ in Theorem 3.9. \bigskip We finish this section by describing the above results for $g=3$. \bigskip \begin{exa}\begin{em} Suppose $X$ has genus $3$. It follows from \cite{BGN}, \cite{M1} and Corollary 3.3 that a point $\pa\in R$ is BN if $\mu>0$, $1\le \mu+3(1-\lambda)$, $\pa\ne(1,1)$, except possibly when $\mu=2$ and $1\le \lambda\le 4/3$. In fact $(2,1)$ is also BN by \cite{Te1} and \cite{M2}. Moreover, any $\bn$ corresponding to such a point has pure dimension $2n^2+1-k(k-d+2n)$ and $\text{Sing\,}\bn ={\cal W}^k_{n,d}$; it is irreducible if $0<d<n+3$ or if $d=n+3(k-n)$. If $X$ is not hyperelliptic, then by \cite{M3} we can remove the exceptional case and say that $\pa\in R$ is BN if and only if $\mu>0$, $1\le \mu+3(1-\lambda)$, $\pa\ne(1,1)$. The statements about dimension and singularities still apply. Apart from the trivial bundle $\cal O$, there is also precisely one further stable bundle on $X$ in $R$, namely the bundle $E_K$ with $n=2, d=4, k=3$ (see \cite{M1}, 2-A-4). \end{em}\end{exa} \section{ The region $BMNO$ in the $(\mu,\lambda$)-plane} In this section, using the results of the previous section, combined with Serre duality, we describe the region $BMNO$. Throughout the section $X$ is an arbitrary non-singular algebraic curve of genus $g\geq 3$. \bigskip In order to translate the results of section 2 into geometric form, we introduce for each $d'$ and $s$ such that $d'\ge\parti=(s-1)(s+g)/s $, the affine map $T_{d',s}$ given by $$T_{d',s}(\mu,\lambda)=(\mu+d',s\lambda).$$ Notice that it shifts points to the right, and, if $s>1$, also expands in the $\lambda$ direction. (We shall refer to these maps as {\it translations} although strictly speaking only the $T_{d',1}$ are translations.) The idea is to use the regions $BGN$ and $M$ as ``tiles'' to cover a larger region, the tiling being obtained by translating $BGN$ and $M$ by the maps $T_{d',s}$. Then we will apply Serre duality to obtain the $BMNO$ region. More precisely, recall that $\parti=(s-1)(s+g)/s$, and set $\hparti=\lceil\eta(s)\rceil$ (we will use the notation $\lceil\cdot\rceil$, and $\lfloor\cdot\rfloor$, respectively, for the least integer not smaller, and the largest integer not greater than a given number, the so-called ``ceiling'' and ``floor'' functions). We will consider $d'$ and $s$ such that $d' \geq \hparti$ for the affine maps $T_{d',s}$. For the description of the regions $BGN$ and $M$, we consider the trapezia $$ BGN'=\{\pa: 0<\mu\leq1,\quad 0<\lambda\leq\frac{1}{g}(\mu+g-1)\}$$ $$M'=\{\pa: 1<\mu<2,\quad 0<\lambda\leq\frac{1}{g}(\mu+g-1)\}.$$ Note however that the point $(1,1)\in BGN'$ is only $1$-BN, so we define $$BGN=BGN'-\{(1,1)\}.$$ Moreover, a geometrical interpretation of Corollary 3.3 shows that, if we translate $BGN$ by $T_{1,1}$, we obtain the points in the boundary of $M'$ with $\mu=2, 0<\lambda<1$, and so these also give BN points. We therefore define $$M=M'\cup\{(2,\lambda): 0<\lambda<1\}.$$ \begin{rem}\begin{em} Notice that there are still some points in the boundary line $\mu=2$ of $M$ that are not covered this way, namely those for which $1\leq\lambda\leq 1+1/g$, and therefore we do not know whether they are BN points or not. For the non-hyperelliptic case see Remark 4.3. \end{em}\end{rem} >From Theorem 3.9 and Remark 3.8 we have \begin{thm} If $\pa\in BGN\cup M$, $1\le s\le g$ and $d'\ge\hparti$, then $T_{d',s}\pa$ is $n$-BN for all $n$ such that $(\mu ,\lambda )$ is $n$-BN. \end{thm} \begin{rem}\begin{em} One can in fact obtain many points in $T_{d',s}(BGN\cup M)$ which are BN points. Let $(\mu,\lambda)\in T_{d',s}(BGN\cup M)$ and write $\mu=\frac{a}{b}$, $\lambda=\frac{c}{e}$ in their lowest terms. To get bundles of rank $n$, the conditions we need are that $b|n$ and $es|cn$. If $s|c$, these conditions reduce to $b|n$ and $e|n$, so $\pa$ is BN. Points of this form are dense in $T_{d',s}(BGN\cup M)$. \end{em}\end{rem} \begin{rem}\begin{em} {\it i)} If we take a point $\pa$ in $T_{d',s}(BGN\cup M)$ lying strictly below the top boundary, we can obtain an improvement to Theorem 4.2. Suppose $n$ is a positive integer such that $n\mu$ and $n\lambda$ are both integers and define $\lambda'=\frac{s}{n}\lceil\frac{n\lambda}{s}\rceil$. If $(\mu,\lambda')\in T_{d',s}(BGN\cup M)$, then ${\cal W}^{n\lambda'-1}_{n,n\mu}\ne\emptyset$; hence ${\cal W}^{n\lambda-1}_{n,n\mu}\ne\emptyset$. Now $$\lambda'\le\frac{s}{n}\left(\frac{n\lambda+s-1}{s} \right)=\lambda+\frac{s-1}{n};$$ so this holds for all sufficiently large $n$ (even if $n\lambda/s$ is not an integer). \noindent {\it ii)} By a similar method, taking $\lambda ' =s$, we can show also that the region $$\mu>\hparti+1,\ \lambda\le s,\ \mu\not\in{\bold N}$$ consists entirely of BN points. \end{em}\end{rem} \bigskip The translates of the regions $BGN$ and $M$ can be described explicitly as follows: $$ T_{d',s}(BGN)=\{\pa: d'<\mu\leq d'+1,\ 0<\lambda\leq\frac{s}{g}(\mu-d'-1)+s,$$ $$\pa\ne(d'+1,s)\}$$ $$ T_{d',s}(M)=\{\pa: d'+1<\mu<d'+2,\ 0<\lambda\leq\frac{s}{g}(\mu-d'-1)+s\}$$ $$\quad\quad\cup\{(d'+2,\lambda): 0<\lambda<s\}$$ Furthermore, it follows from these formulae that $T_{d'+1,s}(BGN)$ is strictly included in $T_{d',s}(M)$, while $T_{d'+1,s+1}(BGN)$ strictly includes $T_{d',s}(M)$, i.e. $$T_{d'+1,s}(BGN) \subset T_{d',s}(M) \subset T_{d'+1,s+1}(BGN).\eqno(5)$$ Therefore, if $d'\geq\hparti$, the translate $T_{d',s}(BGN)$ covers a larger region than does $T_{d'-1,s-1}(M)$. \bigskip We will use these relations to translate, in a convenient way, the known regions. From Example 3.12, we can assume $g>3$. \bigskip We can {\it obtain a new} region of BN points, {\it either by translating} $BGN$ by $T_{2,1}$, or $M$ by $T_{1,1}$; by $(5)$, the latter covers a larger area than the former, so we use $T_{1,1}(M)$ to enlarge the region. We now {\it continue the process}, translating $M$ by $T_{d',1}$ with increasing $d'$ (but always keeping $d'<g-2$ to remain in $R$) (as illustrated in Figure 4; there, $T_{2,1}(BGN)$ is represented by the lighter part in the first diagram, so we can compare it to $T_{1,1}(M)$). Now, for exactly the same reasons as before, this is the best we can do {\it as long as} $d'<\hat\eta(2)-1$. However, when $d'=\hat\eta(2)-1$, $T_{\hat\eta(2)-1,1}(M)$ covers a smaller region than $T_{\hat\eta(2),2}(BGN)$, so we now use the latter (see Figure 5). Of course we can now only guarantee to get $n$-BN points, for some values of $n$; however, see Remarks 4.3 and 4.4. If we have not yet arrived at $d'=g-2$, at the {\it next step} we cover a larger region using now $T_{\hat\eta(2),2}(M)$. We then {\it continue the process} until we arrive at $\hat\eta(3)-1$ or $g-2$. In the latter case {\it we stop}, in the former {\it we use now} $T_{\hat\eta(3),3}(BGN)$, {\it and repeat the process}. The union of trapezia obtained in this way is therefore a polygonal region which consists entirely of $n$-BN points for some values of $n$. This is the best we can do purely by translating, but there is a possibility of {\it obtaining further} $n$-BN points by first translating beyond $\mu=g-1$ and then {\it applying the Serre duality} map $\sigma$. Thus we consider the affine maps $U_{d',s}=\sigma\circ T_{d',s}$, given explicitly by $$U_{d',s}\pa=(2g-2-\mu-d',\ s\lambda+g-1-\mu-d').$$ We now have $d'\ge g-2$ and $T_{d',s}$ maps part of $BGN\cup M$ below the Riemann-Roch line and hence outside $P$. So $U_{d',s}\pa$ will not lie entirely in $R$; in fact the second coordinate in the above formula can be $\le0$. However it is easy to see that $$ U_{d',s}(BGN)\cap R\ne\emptyset\iff d'\ge\hparti\text{ and }g-1\le d'\le\min\{s+g-2,2g-3\},$$ $$ U_{d',s}(M)\cap R\ne\emptyset\iff d'\ge\hparti\text{ and }g-2\le d'\le s+g-3.$$ If $s\ge g$, then $\parti\ge s+g-2$, so there are no $d'$ satisfying the above conditions; we shall therefore assume that $s<g$. We write for convenience $U_{d',s}(1,1)=(d_1,s_1)$, so that$$d_1=2g-3-d',\ s_1=s+g-2-d'.$$ Then $U_{d',s}(BGN)\cap R$ is given by $$d_1\le\mu<d_1+1,\quad0<\lambda\le(1-\frac sg)(\mu-d_1)+s_1$$with the point $(d_1,s_1)$ omitted, while $U_{d',s}(M)\cap R$ is given by $$d_1-1<\mu<d_1,\quad0<\lambda\le(1-\frac sg)(\mu-d_1)+s_1$$together with the line segment $$\ell =\{(d_1-1,\lambda): 0<\lambda<s_1-1\}.$$ \bigskip The following lemmas will show that we can gain an extra triangle by replacing $T_{\hat\eta(s_1+1)-2,s_1}(M)$ by the appropriate $(U_{d',s}(BGN)\cap R)\cup \ell ''$ where $\ell ''$ is a line segment. \begin{lemma}Suppose $d'\ge\hparti$ and $g-2\le d'\le s+g-3$. Then $s_1\ge1$, $d_1-1\ge\hat\eta(s_1)$ and $$U_{d',s}(M)\cap R\subset T_{d_1-1,s_1}(BGN)\cup \ell .$$ \end{lemma} {\it Proof:} Note first that $s_1=s+g-2-d'\ge1$. The inequality $d_1-1\ge\hat\eta(s_1)$ is equivalent to $$s_1(d_1-1)\ge(s_1-1)(s_1+g),$$ or, substituting for $s_1$, $d$, $$(s+g-2-d')(2g-4-d')\ge(s+g-3-d')(s+2g-2-d').$$ This simplifies to $$(s+1)d'\ge s^2+(g-1)s-2=(s-1)(s+g)+g-2.$$ But by hypothesis $sd'\ge(s-1)(s+g)$ and $d'\ge g-2$. This proves the inequality. Comparing the formulae for $U_{d',s}(M)$ and $T_{d_1-1,s_1}(BGN)$, we see that it is now sufficient to prove that $\frac{s_1}g\le1-\frac sg$, i.e.~$s_1+s\le g$. Now $s_1+s=2s+g-2-d'$, so we need to show that $d'\ge2s-2$. Since $s<g$, this follows from the hypothesis $d'\ge\hparti$.\vskip 10pt Note that, if $s_1=1$, then $\ell =\emptyset$, while, if $s_1>1$, $$\ell \subset T_{d_1-2,s_1-1}(BGN).$$ Combined with the lemma, this tells us that $U_{d',s}(M)$ gives nothing new. \begin{lemma} Suppose $d'\ge\hparti$ and $g-1\le d'\le s+g-2$. If $d_1\ge \hat\eta(s_1+1)$, then $$U_{d',s}(BGN)\cap R\subset T_{d_1,s_1+1}(BGN)\cup \ell '$$ where $$\ell '=\{(d_1,\lambda): 0<\lambda<s_1\}.$$ \end{lemma} {\it Proof:} This follows easily from the formulae for the two sets. \vskip 10pt If $s_1=0$, then $\ell '=\emptyset$, while, if $s_1\ge1$,$$\ell '\subset T_{d_1-1,s_1}(BGN).$$So again we get nothing new. It remains therefore to consider the case where $d_1<\eta(s_1+1)$. Now $(d_1+1,s_1+1)=\sigma(d',s)$. Since the Brill-Noether number $\rho$ is invariant under $\sigma$ and $d'\ge\parti$, it follows that $d_1+1\ge\eta(s_1+1)$. So the only case we need to consider is $$d_1+1=\hat\eta(s_1+1)\le g-1.$$ \begin{lemma}In the above circumstances, $$T_{d_1-1,s_1}(M)\subset (U_{d',s}(BGN)\cap R)\cup \ell ''$$ where $$\ell ''=\{(d_1+1,\lambda): 0<\lambda<s_1\}.$$ \end{lemma} {\it Proof:} From the formulae for the two sets, we see that it is sufficient to prove that $\frac{s_1}g\le1-\frac sg$. For this, see the proof of Lemma 4.5. \bigskip Thus in each chain $$T_{\hat\eta(s_1),s_1}(M), \ldots, T_{\hat\eta(s_1+1)-2,s_1}(M)$$ in the construction described earlier, we can gain an extra triangle by replacing\linebreak $T_{\hat\eta(s_1+1)-2,s_1}(M)$ by the appropriate $(U_{d',s}(BGN)\cap R)\cup \ell ''$. \bigskip Finally, then, we {\it define} $BMNO$ to be the union of the trapezia constructed above together with their Serre duals. The region $BMNO\cap R$ is bounded from below by $\lambda=0$, on the sides by $\mu=0$ and $\mu=g-1$, and from above by the graph of a seesaw-like function $f_g$ defined on the interval $(0,g-1]$ by $$f_g(\mu)=\left\{\begin{array}{lll} \frac{s}{g}(\mu-\lceil\mu\rceil)+s &\mu\in(\hat\eta(s),\hat\eta(s)+1]\\ \frac{s}{g}(\mu-\lceil\mu\rceil+1)+s &\mu\in(\hat\eta(s)+1,\hat\eta(s+1)-1]\\ \frac{\hat\eta(s+1)-s}g(\mu-\lceil\mu\rceil+1)+s &\mu\in(\hat\eta(s+1)-1,\hat\eta(s+1)]. \end{array}\right.$$ We extend $f_g$ to the whole interval $(0,2g-2)$ by insisting that its graph is invariant under $\sigma$; the graph of $f_g$ is then the top boundary of $BMNO$. Figure 6 shows a typical $BMNO$ region. We stress the fact that we have to exclude from $BMNO$ those points corresponding to translates of those parts of the boundaries of $BGN$ or $M$ which are not included in the original regions, and we can summarize as follows: {\it If $\pa$ lies in or on the polygon defined above, $\pa$ is $n$-BN, for many values of $n$, except for $\mu=0$ and $\pa=(1,1)$, and possibly for $\mu\in{\Bbb N}-\{0,1\}$, $\mu\in(\hparti,\hat\eta(s+1)]$, $\lambda\ge s$.} \begin{rem}\begin{em} When $X$ is not hyperelliptic, Mercat has proved recently \cite{M3} that the results of \cite{M1} extend to the case $\mu=2$. The constructions of \cite{M3} are the same as those of \cite{M1}, so the proofs of section 2 still work, except that in Lemmas 3.6 and 3.7, we should replace the condition $d'\le2g$ by $d'\le2g-1$. The effect of this is that those of the points excluded from $BMNO$ as above which arise as translates of the right-hand boundary of $M$ can be restored. However those points arising from the left-hand boundary of $BGN$ cannot be restored. Thus the only points of the boundary which must be excluded are those of the form $(\hparti,\lambda)$ with $\lambda>(s-1)(1+\frac 1g)$ and the points $(\hparti+1,s)$ which arise as translates of $(1,1)$. \end{em}\end{rem} \begin{rem}\begin{em} For semistable bundles, the results of \cite{BGN} and \cite{M3} allow us to include both left-hand and right-hand boundaries of $BGN\cup M$ (and indeed the point $(1,1)$). So in this case the whole boundary of $BMNO$ can be included. Moreover one can include the whole of the line segments $\{(\hparti,\lambda): 0<\lambda\le s\}$. \end{em}\end{rem} \begin{rem}\begin{em} The analysis in Remarks 4.3 and 4.4 works also for $U_{d',s}$ and hence whenever $\lambda<f_g(\mu)$ (with the usual exceptions for integral values of $\mu$). So there is certainly a dense subset of $BMNO$ consisting of BN points. \end{em}\end{rem} \bigskip Finally, we have the following proposition, showing that the region $BMNO$ always ``stays close'' to the BN curve: \begin{propn} Let $$\rho_g(\mu)=\frac{\sqrt{(\mu-g+1)^2+4(g-1)}+\mu-g+1}2$$ denote the function whose graph is the BN curve. Then, for $\mu\in(0,2g-2)$, $$0\le\rho_g(\mu)-f_g(\mu)<1.$$ \end{propn} {\it Proof:} Since the graphs of $\rho_g$ and $f_g$ are both invariant under $\sigma$, it is sufficient to prove this for $\mu\le g-1$. For the first inequality we need to prove that every point of $BMNO$ lies on or below the BN curve. Since this is certainly true for points of $BGN\cup M$ and $\tilde\rho$ is invariant under $\sigma$, it is sufficient to prove that, whenever $\pa\in BGN\cup M$ and $d'\ge\parti$, $$I=\frac1{\lambda}\left(\tilde\rho(T_{d',s}\pa)- \tilde\rho\pa\right)\ge0.$$ A simple calculation shows that $$ \begin{array}{lll} I&= sd'-(s-1)((s+1)\lambda-\mu+g-1)\\ &\ge(s-1)(s-(s+1)\lambda+\mu+1)\\ &\ge(s-1)\left(\mu+\dfrac{s+1}g(1-\mu)\right)\ge0. \end{array} $$ For the second inequality, note first that both $\rho_g$ and its derivative $\rho_g'$ are strictly increasing (this is easy to see either geometrically or by calculus). It follows from the formulae for $f_g(\mu)$ and the fact that, by definition of $\hparti$, $$\rho_g(\mu)<s+1\text{ for }\mu\le\hat\eta(s+1)-1,$$ that it is sufficient to prove the inequalities $$\rho_g(\hat\eta(s+1))-\left( \frac{\hat\eta(s+1)-s}g+s\right)<1$$ $$ \rho_g(\hat\eta(s+1))-\left(s+1- \frac{s+1}g\right)<1$$ for $s\ge1$ and $\hat\eta(s+1)\le g-1$. Since $\rho_g'(g-1)=\frac12$ and $\rho_g'$ is strictly increasing, $$\rho_g(\hat\eta(s+1))<\rho_g(\hat\eta(s+1)-1)+ \frac12<s+\frac32.$$ On the other hand $$ \begin{array}{lll} \dfrac{\hat\eta(s+1)-s}g+s&\ge& \dfrac{\eta(s+1)-s}g+s\\ &=&\dfrac s{s+1}+s\ge s+\dfrac12, \end{array} $$ proving the first of the required inequalities. Also $\eta(s+1)\le g-1$ implies that $(s+1)^2\le g$; hence $\frac{s+1}g\le\frac1{s+1}\le\frac12$. Thus $$s+1-\frac{s+1}g\ge s+\frac12$$and we are done. \begin{rem}\begin{em} A careful analysis of this proof shows that the worst cases for $\rho_g(\mu)-f_g(\mu)$ are as $\mu\to\hat\eta(s+1)-1$ from above. Thus in fact $$\rho_g(\mu)-f_g(\mu)<\max[\rho_g(\hat\eta(s+1)-1)-s]$$ taken over values of $s\ge1$ for which $\hat\eta(s+1)\le g-1$, and this inequality is best possible. The best possible inequality which is independent of $g$ is the one stated in the proposition. \end{em}\end{rem} Examples of stable bundles which are outside the range to which the constructions of this section apply are given in \cite {BF} and \cite{Mu}, and some different examples in \cite{BMN}. \section{ Comparison with Teixidor's region} We now compare $BMNO$ with the corresponding region $T$ constructed by the results of Teixidor \cite{Te1} and Mercat \cite{M2}, mainly by means of some examples. \smallskip In the stable case, Teixidor's original result excluded from $\bn$ the vertical segments of length 1, with upper end at a point {\it on\/} the BN curve $\tilde\rho=0$ with integer coordinates. However Mercat in \cite{M2} removed this restriction except for the topmost point of each segment, although he needs also to exclude all the points described in the last sentence of the following theorem, while Teixidor excluded only those segments whose topmost point lies on the BN curve. We will quote the results of both as follows: \begin{thm}{\bf (Teixidor/Mercat)} A point $\pa$ determines a non-empty locus $\wbn$ if any of the following three conditions holds: \item $\tilde\rho(\lceil\mu\rceil, \lceil\lambda\rceil)\geq0\text{ and } 0\neq\lambda-\lfloor\lambda\rfloor\leq\mu-\lfloor\mu\rfloor$ \item $ \tilde\rho(\lfloor\mu\rfloor, \lceil\lambda\rceil)\geq0 \text{ and } \lambda-\lfloor\lambda\rfloor>\mu-\lfloor\mu\rfloor$ \item $ \tilde\rho(\lfloor\mu\rfloor, \lfloor\lambda\rfloor)\geq0 \text{ and }\lambda=\lfloor\lambda\rfloor.$ Moreover, under the same conditions, $\bn$ is non-empty except possibly for points $\pa$ with $\mu$, $\lambda$ integers and $\tilde\rho(\mu-1,\lambda)<0$. \end{thm} \begin{rem}\begin{em} In the semistable case, this theorem is a mere translation of a result of Teixidor (\cite{Te1}, Theorem 1, p. 386) to the $\pa$ language; note that Teixidor's result is stated for $X$ generic, but for semistable bundles this automatically implies the result for any $X$. Observe that conditions (1) and (2) in fact define triangles in the $\pa$-plane, with all their vertices at points with integer coordinates, as illustrated in Figure 7, where the lighter area corresponds to the first condition and the darker to the second. Condition 3 describes a horizontal segment of length 1, starting at the point $(\lfloor\mu\rfloor, \lfloor\lambda\rfloor)$. \end{em}\end{rem} As shown in Figure 7, for any point with integer coordinates on or below the BN curve, the first two conditions together determine a parallelogram; hence, the region defined by Theorem 5.1 is sometimes referred to as ``Teixidor's parallelograms''. We denote this region by $T$. We can describe the region $T$ in a similar way to $BMNO$ by first defining, for any integer $s$, $$\hparti'=\lceil\parti+\tfrac1s\rceil-1.$$ Then $d'\ge\hparti'$ if and only if $\tilde\rho(d'+1,s) \ge0$. (Recall that $d'\ge\hparti$ if and only if $\tilde\rho(d',s)\ge-1$.) The region $T$ is then bounded below by $\lambda=0$, on the sides by $\mu=0$ and $\mu=2g-2$ and from above by the graph of a function $t_g$ defined by $$t_g(\mu)=\cases \mu-\lceil\mu\rceil+s &\mu\in (\hparti',\hparti'+1]\\ s&\mu\in(\hparti'+1,\hat\eta(s+1)'] \endcases$$ Unlike $f_g$, the function $t_g$ is in fact continuous and non-decreasing, so the shape of $T$ is simpler than that of $BMNO$. Note also that the region covered by Teixidor's parallelograms is invariant under $\sigma$, so we do not obtain anything new by using Serre duality. Finally it is easy to check that $0\le\rho_g(\mu)-t_g(\mu)<1$ (compare Proposition 4.11). Figure 8 shows a typical Teixidor polygon (here, $g=10 $ and the only vertex on the BN curve is $(3,9)$, since $3$ is the only divisor of $g-1=9$). \bigskip To compare the upper boundaries of $T$ and $BMNO$, we first note that $$\hparti'=\cases\hparti\text{ if }\hparti=\parti\\\hparti-1\text{ otherwise}.\endcases$$ For $\mu\le g-1$, it follows that $f_g(\mu)\ge t_g(\mu)$ except possibly in the intervals $(\hparti-1,\hparti+1)$. If $\hparti=\parti$ (or equivalently $\tilde\rho(\hparti,s)=-1$), then $f_g(\mu)\ge t_g(\mu)$ in this interval as well. On the other hand, if $\hparti\ne\parti$, then $t_g(\mu)>f_g(\mu)$ on $(\hparti-1,\hparti+1)$. Thus $BMNO$ always extends outside $T$ and, for almost all values of $g$, $T$ also extends outside $BMNO$. At any rate, for a given (small) genus, it is easy to compute both $\hparti$ and $\hparti'$ explicitly. The figures 9, 10 and 11, illustrate the cases $g=10$, $g=12$, and $g=13$, respectively, where different situations can be appreciated. There the shaded area is $BMNO$, and Teixidor's polygons are only outlined. \section{ The hyperelliptic case} Suppose now that $X$ is a non-singular hyperelliptic curve of genus $g\geq 3$. If we denote by $L$ the hyperelliptic line bundle on $X$ then $h^0(L^{\otimes(s-1)})=s$ for $1\le s\le g$, so we can take $d'=2s-2$ in Theorem 3.9. The analogue of Theorem 4.2 is \begin{thm} Let $X$ be a non-singular hyperelliptic curve of genus $g\geq 3$. If $\pa\in BGN\cup M$ and $1\le s\le g$, then $T_{2s-2,s}\pa$ is $n$-BN for all $n$ such that $(\mu ,\lambda )$ is $n$-BN. \end{thm} We now define $$BMNO_h=\bigcup_{1\le s\le g-1}\left(T_{2s-2,s}(BGN\cup M)\cap P\right).$$ It will be convenient to include the point $(2,1)$ in $M$ (see \cite{M2}). This region is already invariant under Serre duality, so we do not need to invoke the transformations $U_{d',s}$ in this case. The top boundary of $BMNO_h$ is given by the graph of the function $h_g$ defined on $(0, 2g-2)$ by $$h_g(\mu)=\frac{s}{g}(\mu-2s+1)+s\text{ for }\mu\in(2s-2,2s].$$ The analogues of Remarks 4.3 and 4.4 hold and indeed we can improve Remark 4.4 (ii). For $1\le s\le g-1$, the region $$2s-1<\mu\le2s,\ \ \ \lambda\le s$$ consists entirely of BN points. By Serre duality, so also does $$2g-2-2s\le\mu<2g-1-2s,\ \ \ \lambda\le s+\mu-g+1,$$ i.e. (replacing $s$ by $g-s$) $$2s-2\le\mu<2s-1,\ \ \ \lambda\le \mu-s+1.$$ Of course, all points of $BGN\cup M$ are BN, hence also all points of its Serre dual. These results are illustrated in Figure 12. In the semistable case, we can include the points $(2s-1,s)$ and also the line segments $\{(2s,\lambda):s<\lambda\le s+1\}$. \bigskip The next step is to show that all special stable bundles, except for certain line bundles, lie in $BMNO_h$. \begin{thm} Let $X$ be a hyperelliptic curve, $E$ a stable bundle on $X$ of rank $n$, degree $d$ and slope $\mu=\frac{d}{n}$, and $s$ an integer. \item {\it 1)} If $0\le s\le g$ and $2s-2<\mu<2s$, then $$h^0(E)\le sn+\frac{s}{g}(d-(2s-1)n).$$ \item {\it 2)}If $0\le s\le g-1$, $\mu=2s$ and $E\not\cong L^{\otimes s}$, then $h^0(E)\le sn$. \end{thm} {\it Proof:} (1) We begin by writing $$F_s(n,d)=sn+\frac{s}{g}(d-(2s-1)n).$$ We check easily that $$2F_s(n,d)=F_{s-1}(n,d-2n)+F_{s+1}(n,d+2n).$$ To prove the theorem, we argue by induction on $s$. For $s=0$, the result is obvious, since $E$ stable with $\mu<0$ implies $h^0(E)=0$. The result for $s=g$ follows from this by Serre duality and Riemann-Roch. Now suppose $0<s<g$. Suppose that there exists a stable bundle $E$ of slope $\mu$ with $2s-2<\mu<2s$ and such that $H^0(E)=F_s(n,d)+b_0$ with $b_0>0$. Tensoring the exact sequence $0\to L^*\to H^0(L)\otimes{\cal O}\to L\to 0$ by $E$, we get $$0\to L^*\otimes E\to H^0(L)\otimes E\to L\otimes E\to0.$$ Since $h^0(L)=2$, this gives $$2h^0(E)\le h^0(E\otimes L^*)+ h^0(E\otimes L).$$ By inductive hypothesis, we have $$h^0(E\otimes L^*)\le F_{s-1}(n,d-2n);$$ hence $$h^0(E\otimes L)\ge2F_s(n,d)+2b_0-F_{s-1}(n,d-2n)=F_{s+1}(n,d+2n)+2b_0.$$ Thus $h^0(E\otimes L)=F_{s+1}(n,d+2n)+b_1$, with $b_1\ge2b_0$. Continuing in this way, we construct a sequence $(b_i)$, defined by $$h^0(E\otimes L^{\otimes i})=F_{s+i}(n,d+2in)+b_i,$$ with $$b_{i+1}\ge 2b_i-b_{i-1}.$$ We deduce that this sequence is strictly increasing. On the other hand, by the result for $s=g$, we have $$h^0(E\otimes L^{\otimes (g-s)})\le F_g(n,d+2(g-s)n).$$ So $b_{g-s}=0$, which is a contradiction. The result follows. \bigskip (2) \ \ Again we proceed by induction. For $s=0$, the only stable bundle of slope $0$ with $h^0(E)>0$ is $\cal O$. Similarly, the only stable bundle of slope $2g-2$ with $h^0(E)>(g-1)n$ is $K$. For $0<s<g-1$, we proceed as in (1). If there exists a stable bundle $E$ of slope $2s$ such that $h^0(E)=sn+b_0$ with $b_0>0$, we define the sequence $(b_i)$ for $1\le i\le g-s-1$ by $h^0(E\otimes L^{\otimes i}) = (s+i)n+b_i$ and prove that $(b_i)$ is strictly increasing. On the other hand, since by hypothesis $E\otimes L^{\otimes (g-s-1)}\not\cong K$, it follows that $b_{g-s-1}=0$. Again we have a contradiction. \begin{rem}\begin{em} It follows from the proof of Theorem 6.2 that, if $1\le s\le g-1$ and $h^0(E)$ takes its maximum value $F_s(n,d)$ (or $sn$), then also $h^0(E\otimes L^*)=F_{s-1}(n,d-2n)$ (or $(s-1)n$) and $h^0(E\otimes L)=F_{s+1}(n,d+2n)$ (or $(s+1)n$). \end{em}\end{rem} \begin{cor}If $\pa \in BMNO_h$, then $\pa$ is $n$-BN for infinitely many values of $n$. The only special stable bundles which lie outside $BMNO_h$ are the line bundles $L^{\otimes(s-1)}$ for $1\leq s\leq g$ and $L^{\otimes(s-1)}(p)$ for $1\leq s\leq g-1$ and $p \in X.$ \end{cor} {\it Proof:} By Theorems 6.1 and 6.2, it is sufficient to prove that the points $(2s-1,s)$ are only $1$-BN. By \cite{BGN} Theorem B, $(1,1)$ is only $1$-BN; hence, by Remark 6.3, $(2s-1,s)$ is also only $1$-BN. \bigskip According to this Corollary, there do not exist stable bundles of rank $n>1$ and slope $2s-1$ with $1\leq s\leq g-1$ and $h^0(E)=sn.$ However \begin{propn} Let $X$ be a hyperelliptic curve. For any integers $n$, $s$ with $n>0$, $1\le s\le g-1$, there exist stable bundles $E$ of rank $n$ and slope $2s-1$ with $h^0(E)=sn-1$. \end{propn} {\it Proof:} For $s=1$, this is a special case of \cite{BGN}, Theorem B. If $1<s\le g-1$, a result of \cite{BMN} says that, if $\Delta$ is a torsion sheaf of length $n$ with support $n$ distinct points of $X$, and if $M$ is a line bundle of degree $2$ on $X$ such that $h^0(M)=1$ then a sufficiently general extension $$0\to L^{\otimes(s-1)}\oplus\cdots\oplus L^{\otimes(s-1)}\oplus L^{\otimes(s-2)}\otimes M\to E\to\Delta\to0$$ is stable, and clearly $h^0(E)=sn-1$. \bigskip We have now completely settled the nonemptiness problem for bundles of integral slope. For bundles of non-integral slope, however, we still have an indeterminate region of points which we know to be $n$-BN but which may fail to be BN. The next example shows that this can indeed happen. \begin{exa}\begin{em} Suppose that $X$ has genus $g\geq 4$. Suppose that $1\le s\le g-1$ and that $E$ is a stable bundle of rank $n$ and degree $d$ with $2s-1<\mu=\frac{d}{n}<2s$. Write $$d-(2s-1)n=gl+l'\text{ with }0\le l'<g.$$ By Theorem 6.2, we have $$h^0(E)\le sn+\frac{s}{g}(d-(2s-1)n)=sn+sl+\frac{sl'}{g},$$ in other words $$h^0(E)\le sn+sl+\lfloor\frac{sl'}{g}\rfloor.$$ If $\lfloor\frac{sl'}{g}\rfloor<1$, then Theorem 6.1 gives the existence of a bundle $E$ with the maximum possible number of sections. Suppose now that $s=2$ and $\frac{g}2\le l'<g-1$. We claim that, in this case, $$2h^0(E)<2n+2l+\lfloor\frac{2l'}{g}\rfloor=2n+2l+1.$$ \vskip 6pt{\it Proof of the claim:} Suppose that there exists a stable bundle $E$ as above with $h^0(E)=2n+2l+1$. We know that $$2h^0(E)\le h^0(E\otimes L)+h^0(E\otimes L^*).$$ Hence $$4n+4l+2\le3n+3l+\lfloor\frac{3l'}{g}\rfloor+n+l.$$ So $\lfloor\frac{3l'}{g}\rfloor=2$ and $h^0(E\otimes L)=3n+3l+2$. Beginning again with $E\otimes L$ and continuing in this way for a total of $g-3$ steps, we obtain $$\lfloor\frac{(g-1)l'}{g}\rfloor=g-2,$$ hence $l'=g-1$. This contradicts our assumption and proves that there are points which fail to be BN. \end{em}\end{exa} \begin{rem}\begin{em} In the exceptional case $l'=g-1$ of Example 6.6, we can prove that $E$ does exist. In fact, since $3<\mu<4$, by \cite{BGN} we can find a stable bundle $F$ of rank $n$ and slope $4-\mu$ with $h^0(F)=n-l-1$. Then $K\otimes F^*$ has slope $2g-6+\mu$ and $$h^0(K\otimes F^*)=(2g-3)n+gl+l'-(g-1)n+n-l-1=(g-1)n+(g-1)l+g-2.$$ Now take $E=K\otimes F^*\otimes L^{*\otimes(g-3)}=F^*\otimes L^{\otimes2}$ and use the argument of Example 6.6 in reverse. We obtain $h^0(E)=2n+2l+1$ as required. \end{em}\end{rem} Re's improvement of the Clifford bound for $X$ non-hyperelliptic \cite{R} is intriguingly close to the boundary of $BMNO_h$. The results of this section show the extent to which Re's bound fails for a hyperelliptic curve. We finally remark that, in the hyperelliptic case, the upper boundary of the region where $n$-BN points exist is not the graph of a continuous function; possibly this extends to other cases. \vspace{1 true cm} {Fig. 1} The BN map \vspace{1 true cm} {Fig. 2} Teixidor's parallelograms \vspace{1 true cm} {Fig. 3} The regions $BGN$ and $M$ \vspace{1 true cm} {Fig. 4} First steps in the construction of the region $BMNO$ \vspace{1 true cm} {Fig. 5} Gain by translating $BGN$ with bundle with 2 sections \vspace{1 true cm} {Fig. 6} Construction of a typical $BMNO$ region (genus 10) \vspace{1 true cm} {Fig. 7} Teixidor's triangles \vskip 10pt \vspace{1 true cm} {Fig. 8} Teixidor's region for genus 10 \vspace{1 true cm} {Fig. 9} $BMNO$ and $T$ regions for genus $10$ \vskip 10pt \vspace{1 true cm} {Fig. 10} $BMNO$ and $T$ regions for genus $12$ (restricted to $R$) \vspace{1 true cm} {Fig. 11} $BMNO$ and $T$ regions for genus $13$ (restricted to $R$) \vspace{1 true cm} {Fig. 12} The hyperelliptic case.
\section{Introduction} Let $\Sigma$ be a closed Riemann surface of genus $g \geq 2$ and let $G$ be a connected Lie group. Consider the space of reductive representations of the fundamental group of $\Sigma$ in $G$ modulo the action of $G$ by conjugation, \begin{displaymath} \mathcal{M}_{G} = \Hom(\pi_1(\Sigma),G)^{+}/G, \end{displaymath} the superscript ``$+$'' indicating reductive representations. As is well known, $\mathcal{M}_G$ can also be identified with the moduli space of reductive flat $G$-bundles over $\Sigma$ and it has an algebro-geometric interpretation as a moduli space of Higgs bundles (see Hitchin \cite{hitchin,hitchinduke}). In this paper we study the connected components of $\mathcal{M}_{G}$ in the cases $G = \Sp(4,\mathbb{R})$ and $G = \mathrm{SU}(2,2)$. Previous work on this type of problem includes the determination of the number of connected components of $\mathcal{M}_G$ for the groups $\mathrm{PSL(2,\mathbb{R})}$ and $\mathrm{PSL(2,\mathbb{C})}$ by Goldman \cite{goldman-top}, for the groups $\mathrm{PSL}(n,\mathbb{R})$ ($n \geq 3$) by Hitchin \cite{hitchinlie}, and for the groups $\mathrm{PGL(2,\mathbb{R})}$, $\mathrm{PU}(2,1)$, and $\mathrm{U}(p,1)$ by Xia \cite{xia,xia:pu21,xia:up1}. The first observation is that there is a characteristic number, $d$, which comes from a characteristic class of the bundle obtained from a reduction of structure group to the maximal compact subgroups $\mathrm{U}(2) \subseteq \Sp(4,\mathbb{R})$ and $\mathrm{S}(\mathrm{U}(2)\times\mathrm{U}(2)) \subseteq \mathrm{SU}(2,2)$, respectively. It is well known that this satisfies the Milnor-Wood type inequality $\abs{d} \leq 2g-2$ (cf.\ \secref{sec:milnorwood}). This allows one to write \begin{displaymath} \mathcal{M}_G = \mathcal{M}_{-(2g-2)} \cup \cdots \cup \mathcal{M}_{2g-2}, \end{displaymath} where each $\mathcal{M}_d$ is a union of connected components. We can then state our main result as follows. \begin{thm*} The subspaces $\mathcal{M}_0 \subseteq \mathcal{M}_{G}$ are connected for $G = \Sp(4,\mathbb{R})$ and $G = \mathrm{SU}(2,2)$ and the subspaces $\mathcal{M}_{\pm (2g-2)} \subseteq \mathcal{M}_{\mathrm{SU}(2,2)}$ are connected. The subspaces $\mathcal{M}_{\pm (2g-2)} \subseteq \mathcal{M}_{\Sp(4,\mathbb{R})}$ have $3\cdot 2^{2g} + 2g - 4$ connected components. \end{thm*} The most remarkable aspect of this result is that $\mathcal{M}_{\pm (2g-2)} \subseteq \Sp(4,\mathbb{R})$ breaks up into a number of different connected components, which are not detected by the first Chern class given by reduction of structure group to $\mathrm{U}(2) \subseteq \Sp(4,\mathbb{R})$. It seems likely that the remaining $\mathcal{M}_d$ are connected and we hope to come back to this question on a later occasion. The method we use for studying the connected components is via the algebro-geometric interpretation of $\mathcal{M}_G$ as a moduli space of Higgs bundles, due to Hitchin \cite{hitchin,hitchinlie}. (A Higgs bundle is a pair $(E,\Phi)$, where $E$ is a holomorphic rank $n$ degree $d$ vector bundle and $\Phi \in H^0(\Sigma; \End(E)_0 \otimes K)$, see \secref{sec:higgs} for more details.) From this point of view one can define a Hamiltonian circle action on the moduli space and one uses a moment map for this action as a Morse function, in the sense of Bott, to obtain topological information about the space (cf.\ Hitchin \cite{hitchin,hitchinlie} and Gothen \cite{higgs}). The central point is then to identify the critical submanifolds of the Morse function and to obtain topological information about them. In particular, to obtain information about connected components, one needs to consider the local minima of the Morse function. It should be remarked that the moduli spaces have singularities and so one cannot directly apply Morse theory, however, in the case of the determination of connected components this difficulty can be circumvented (see Sections \ref{sec:morse} and \ref{sec:morseindices}). In this paper we show that certain critical submanifolds, corresponding to local minima of the Morse function, can be identified with moduli spaces of \emph{stable triples}, or spaces closely related to them, as studied by Bradlow and Garc{\'\i}a-Prada \cite{bradlow-garcia-prada,garcia-prada}. In the cases $d = 0$ and $\abs{d} = 2g-2$ the structure of the moduli spaces of triples is particularly simple and this allows us to prove the theorem above. In the case of $G = \Sp(4,\mathbb{R})$ and $\abs{d} = 2g-2$ we further need to use a spectral curve (see Hitchin \cite{hitchinduke}) which is an unramified covering of $\Sigma$ and the mod 2 index theorem of Atiyah-Singer to identify the local minima of the Morse function as certain Prym varieties associated to the covering of $\Sigma$. This paper is organized as follows. In \secref{sec:basics} we recall the basics of the theory of Higgs bundles and their relation to representations of the fundamental group of a surface in a non-compact Lie group. We also recall the concept of a \emph{$Q$-bundle}, of which a holomorphic triple is a special case, and prove a theorem (\thmref{thm:quiver}) which is essential for the identification of the subspace of local minima with a moduli space of holomorphic triples. Finally we describe the Morse theory on the moduli space and, in particular, we describe how to find the Morse indices. It was observed by Hausel \cite{hausel} that our results on the Morse indices, together with a theorem of his, imply a theorem of Laumon \cite{laumon} in this context: the nilpotent cone in the moduli space $\mathcal{M}$ of rank $n$ Higgs bundles is a Lagrangian subvariety with respect to the holomorphic symplectic form on $\mathcal{M}$. We end this section by briefly describing this. In \secref{sec:milnorwood} we reprove the known bound $\abs{d} \leq 2g-2$, using Higgs bundles; we include the proof because it gives some extra information which is important later on (cf.\ \propref{prop:iso}). In \secref{sec:minima} we analyze the local minima of the Morse function on the space $\mathcal{M}_G$ for $G = \Sp(4,\mathbb{R})$ and $G = \mathrm{SU}(2,2)$ in detail. Finally, in \secref{sec:count}, we finish the proof of our main theorem, using the previous results. \emph{Acknowledgments.} Part of this paper is based on my Ph.D.\ thesis and I would like to thank my supervisor Nigel Hitchin. I also benefited from workshops organized by the VBAC research group under the European algebraic geometry networks AGE and Europroj, supported by the EU. This work was partially supported by Statens Naturvidenskabelige Forskningsr{\aa}d (Denmark), and by the Funda{\c c}{\~a}o para a Ci{\^e}ncia e a Tecnologia (Portugal) through the Centro de Matem{\'a}tica da Universidade do Porto and the project Praxis 2/2.1/MAT/63/94. \section{Higgs bundles and the topology of moduli spaces} \label{sec:basics} \subsection{Higgs bundles} \label{sec:higgs} In this section we review some basic facts about Higgs bundles and set up notation. For details see Hitchin \cite{hitchin} and Simpson \cite{simpson:higgs}. Let $G_{\mathbb{C}}$ be a complex semi-simple Lie group with Lie algebra $\mathfrak{g}_{\mathbb{C}}$. Let $G \subset G_{\mathbb{C}}$ be a maximal compact subgroup with Lie algebra $\mathfrak{g}$. Thus there is a compact real structure $\tau \colon \mathfrak{g}_{\mathbb{C}} \to \mathfrak{g}_{\mathbb{C}}$ whose fixed point set is $\mathfrak{g}$. Denoting the $-1$-eigenspace of $\tau$ by $\mathfrak{g}^{\perp}$ we then have $\mathfrak{g}_{\mathbb{C}} = \mathfrak{g} \oplus \mathfrak{g}^{\perp}$. Non-abelian Hodge theory gives an equivalence between reductive representations of $\pi_1(\Sigma)$ in $G_{\mathbb{C}}$ and Higgs bundles over $\Sigma$, which we now describe. Let \begin{displaymath} \rho \colon \pi_1(\Sigma) \to G_{\mathbb{C}} \end{displaymath} be a reductive representation. This data is equivalent to having a principal bundle \begin{displaymath} P_{\mathbb{C}} = \tilde{\Sigma}\times_{\rho}G_{\mathbb{C}} \end{displaymath} with a reductive flat connection $B \in \Omega^1(P_{\mathbb{C}};\mathfrak{g}_{\mathbb{C}})$ (here $\tilde{\Sigma}$ is the universal cover of $\Sigma$). If we have a metric in $P_{\mathbb{C}}$, i.e.\ a reduction of structure group from $G_{\mathbb{C}}$ to $G$, we can write \begin{displaymath} i^*B = A + \theta \end{displaymath} where $i \colon P \hookrightarrow P_{\mathbb{C}}$ is the inclusion of the principal $G$-bundle $P$ given by the reduction of structure group, $A$ is a connection on $P$, and $\theta \in \Omega^1(P;\mathfrak{g}^{\perp})$ is a tensorial form, which can therefore be thought of as an element of $\Omega^1(\Sigma; P\times_{\Ad} \mathfrak{g}^{\perp})$. Given a complex representation of $G$ (e.g.\ the adjoint representation on $\mathfrak{g}_{\mathbb{C}}$), we have the usual decomposition of the covariant derivative $\d_A$ in its $(1,0)$- and $(0,1)$-parts: \begin{displaymath} \d_A = \partial_A + \bar{\partial}_A. \end{displaymath} Similarly, we can write \begin{displaymath} \theta = \Phi - \tau(\Phi), \end{displaymath} for a unique $\Phi \in \Omega^{1,0}(\Sigma;\Ad P_{\mathbb{C}})$ (by abuse of notation, we denote by $\tau$ the combination of the compact real structure $\tau$ on $\mathfrak{g}_{\mathbb{C}}$ and conjugation on the form component). Corlette \cite{corlette} and Donaldson \cite{Donaldson} proved that there exists a harmonic metric in $P$, that is, a metric such that $(A,\Phi)$ obtained via the above procedure satisfy Hitchin's equations \begin{align*} F(A) - [\Phi,\tau(\Phi)] &= 0, \\ \bar{\partial}_A\Phi &= 0. \end{align*} This, in turn, gives a principal \emph{Higgs bundle}, i.e.\ a pair $(P_{\mathbb{C}},\Phi)$ consisting of a holomorphic principal bundle $P$ (with holomorphic structure defined by $\bar{\partial}_A$) and a \emph{Higgs field} $\Phi \in H^0(\Sigma;\Ad P_{\mathbb{C}}\otimes K)$, where $K$ denotes the canonical bundle of $\Sigma$. Given a representation, $V$, of $G_{\mathbb{C}}$ one then obtains a Higgs vector bundle $(E,\Phi)$, where $E = P_{\mathbb{C}} \times_{G_{\mathbb{C}}} V$ and $\Phi \in H^0(\Sigma;\End(E)\otimes K)$. The two main examples we have in mind are the adjoint representation $V = \mathfrak{g}_{\mathbb{C}}$ and the fundamental representation of $G_{\mathbb{C}} = \mathrm{SL}(n,\mathbb{C})$. If the original representation $\rho$ of $\pi_1(\Sigma)$ is irreducible then $(E,\Phi)$ is \emph{stable}, i.e.\ \begin{displaymath} \mu(F) < \mu(E) \end{displaymath} for any proper non-trivial $\Phi$-invariant subbundle $F$ of $E$ (here $\mu(E) = \deg(E)/\rk(E)$ is the \emph{slope} of the holomorphic bundle $E$). Allowing equality in the above inequality gives the notion of a semi-stable Higgs bundle. Finally, if $\rho$ is reductive, then the corresponding Higgs bundle is poly-stable, i.e.\ it is a direct sum of lower rank Higgs bundles, all of the same slope. By a theorem of Hitchin \cite{hitchin} and Simpson \cite{simpson}, the above procedure can be reversed, by finding a harmonic metric in the Higgs bundle. This produces a reductive representation of $\pi_1(\Sigma)$ from a poly-stable Higgs bundle. This gives a homeomorphism \begin{displaymath} \mathcal{M}_{G_{\mathbb{C}}} \to \Hom(\pi_1(\Sigma),G_{\mathbb{C}})^{+}/G_{\mathbb{C}}. \end{displaymath} where $\mathcal{M}_{G_{\mathbb{C}}}$ is the moduli space of poly-stable principal $G_{\mathbb{C}}$ Higgs bundles. We finish by recalling the description of the Zariski tangent space to $\mathcal{M}_{G_{\mathbb{C}}}$ at $(P_{\mathbb{C}},\Phi)$ given by Biswas and Ramanan \cite{biswasramanan}. This is the first hyper-cohomology $\mathbb{H}^1(C_{\mathbb{C}}^{\bullet})$ of the complex of sheaves \begin{equation} \label{eq:cxtangent} C_{\mathbb{C}}^{\bullet} : \mathcal{O}(\Ad P_{\mathbb{C}}) \xrightarrow{\ad(\Phi)} \mathcal{O}(\Ad P \otimes K). \end{equation} From this they deduce the long exact sequence \begin{multline} \label{eq:cxlong} 0 \to \mathbb{H}^0(C_{\mathbb{C}}^{\bullet}) \to H^0(\Sigma;\Ad P_{\mathbb{C}}) \to H^0(\Sigma;\Ad P \otimes K) \to T_{(P_{\mathbb{C}},\Phi)}\mathcal{M}_{G_{\mathbb{C}}} \\ \to H^1(\Sigma;\Ad P_{\mathbb{C}}) \to H^1(\Sigma;\Ad P \otimes K) \to \mathbb{H}^2(C_{\mathbb{C}}^{\bullet}) \to 0. \end{multline} $(P_{\mathbb{C}},\Phi)$ is a smooth point of the moduli space if $\mathbb{H}^0(C_{\mathbb{C}}^{\bullet})$ and $\mathbb{H}^2(C_{\mathbb{C}}^{\bullet})$ vanish (by Serre duality, it is sufficient to check that $\mathbb{H}^0(C_{\mathbb{C}}^{\bullet}) = 0$). From this one sees that stable Higgs bundles represent smooth points of the moduli space. The dimension of $\mathcal{M}_{G_{\mathbb{C}}}$ can be calculated using Riemann-Roch to be \begin{displaymath} \dim_{\mathbb{C}}\mathcal{M}_{G_{\mathbb{C}}} = \chi(\mathcal{O}(\Ad P \otimes K)) - \chi(\mathcal{O}(\Ad P_{\mathbb{C}})) = 2\dim_{\mathbb{C}} \mathfrak{g}_{\mathbb{C}}(g-1) \end{displaymath} \subsection{Real groups} Hitchin \cite{hitchin,hitchinlie} showed how to use Higgs bundles to study representations of $\pi_1(\Sigma)$ in real (non-compact) Lie groups. Next we recall the relevant parts of this theory. Let $G_r \subset G_{\mathbb{C}}$ be a real form, given by a real structure $\sigma \colon \mathfrak{g}_{\mathbb{C}} \to \mathfrak{g}_{\mathbb{C}}$. Let $K \subset G_{r}$ be a maximal compact subgroup and let $\mathfrak{k} \subset \mathfrak{g}_{r}$ be the corresponding inclusion of Lie algebras. Let $\mathfrak{k}^{\perp}$ be the orthogonal complement to $\mathfrak{k}$ with respect to the Killing form, then we can write $\mathfrak{g}_{r} = \mathfrak{k} \oplus \mathfrak{k}^{\perp}$, where the Killing form is negative definite on $\mathfrak{k}$ and positive definite on $\mathfrak{k}^{\perp}$. Define a complex linear involution $\phi \colon \mathfrak{g}_{\mathbb{C}} \to \mathfrak{g}_{\mathbb{C}}$ by $\phi_{\vert \mathfrak{k}_{\mathbb{C}}} = 1$ and $\phi_{\vert \mathfrak{k}^{\perp}_{\mathbb{C}}} = -1$. Define another real structure on $\mathfrak{g}_{\mathbb{C}}$ by $\tau = \sigma \phi = \phi \sigma$. It is then easy to see the that the corresponding real subgroup $G \subset G_{\mathbb{C}}$ is a maximal compact subgroup and, clearly, $\mathfrak{k} = \mathfrak{g} \cap \mathfrak{g}_r$. Now suppose that we have a reductive representation of $\pi_1(\Sigma)$ in $G_r$ and let $B$ be the associated flat connection on the principal $G_r$-bundle $P_{G_r}$. The theorem of Donaldson and Corlette also applies in this case and gives a reduction of structure group to $K \subset G_r$: let $i \colon P_K \hookrightarrow P_{G_{\mathbb{C}}}$ be the inclusion of principal bundles given by combining the reduction of structure group with the inclusion $G_r \subset G_{\mathbb{C}}$. In the decomposition $i^*B = A + \theta$, $A$ and $\theta$ will be fixed by $\sigma$, while $\tau(A) = A$ and $\tau(\theta) = -\theta$. Thus $\bar{\partial}_A$ is fixed by $\phi$, and $\Phi$ is in the $-1$-eigenspace of $\phi$. This means that the corresponding Higgs bundle is of the form \begin{math} (P_{K_{\mathbb{C}}},\Phi), \end{math} where \begin{itemize} \item $P_{K_{\mathbb{C}}}$ is a holomorphic principal $K_{\mathbb{C}}$-bundle, \item $\Phi \in H^0(\Sigma; \Ad_{\mathfrak{k}_{\mathbb{C}}^{\perp}} P_{K}\otimes K)$ (where we use the notation $\Ad_{\mathfrak{k}_{\mathbb{C}}^{\perp}} P_{K} = P_K \times_{\Ad} \mathfrak{k}_{\mathbb{C}}^{\perp}$). \end{itemize} Conversely, such a Higgs bundle gives a representation of $\pi_1(\Sigma)$ in $G_r$. We then have a homeomorphism \begin{displaymath} \mathcal{M}_{G_{r}} \to \Hom(\pi_1(\Sigma),G_{r})^{+}/G_{r}, \end{displaymath} where $\mathcal{M}_{G_{r}}$ is the moduli space of poly-stable Higgs bundles of the above type. Alternatively $\mathcal{M}_{G_r}$ can be thought of as the moduli space of solutions $(A,\Phi)$ to Hitchin's equations modulo $K$-gauge equivalence: then $A$ is a connection on a principal $K$-bundle $P_K$ and $\Phi \in \Omega^{1,0}(\Sigma;\Ad_{\mathfrak{k}_{\mathbb{C}}^{\perp}} P_{K})$. The analogue to \eqref{eq:cxtangent} in this context is that the Zariski tangent space to $\mathcal{M}_{G_{r}}$ is the first hyper-cohomology of the complex of sheaves \begin{equation} \label{eq:realtangent} C_r^{\bullet} : \mathcal{O}(\Ad_{\mathfrak{k}_{\mathbb{C}}} P_{K}) \xrightarrow{\ad(\Phi)} \mathcal{O}(\Ad_{\mathfrak{k}_{\mathbb{C}}^{\perp}} P_{K}\otimes K), \end{equation} where we use the notation $\Ad_{\mathfrak{k}_{\mathbb{C}}} P_{K} = P_K \times_{\Ad} \mathfrak{k}_{\mathbb{C}} = \Ad P_{K_{\mathbb{C}}}$. The analogue to the long exact sequence \eqref{eq:cxlong} is \begin{multline} \label{eq:reallong} 0 \to \mathbb{H}^0(C_{r}^{\bullet}) \to H^0(\Sigma;\Ad_{\mathfrak{k}_{\mathbb{C}}} P_{K}) \to H^0(\Sigma;\Ad_{\mathfrak{k}_{\mathbb{C}}^{\perp}} P_{K} \otimes K) \to T_{(P_{K},\Phi)}\mathcal{M}_{G_{r}} \\ \to H^1(\Sigma;\Ad_{\mathfrak{k}_{\mathbb{C}}} P_{K}) \to H^1(\Sigma;\Ad_{\mathfrak{k}_{\mathbb{C}}^{\perp}} P_{K} \otimes K) \to \mathbb{H}^2(C_{r}^{\bullet}) \to 0. \end{multline} The smooth points of the moduli space are those for which $\mathbb{H}^0(C_r^{\bullet}) = \mathbb{H}^2(C_r^{\bullet}) = 0$ and again the stable Higgs bundles represent smooth points. The dimension of $\mathcal{M}_{G_r}$ can be calculated as before to be \begin{displaymath} \dim_{\mathbb{C}}\mathcal{M}_{G_r} = \dim_{\mathbb{C}} \mathfrak{g}_{\mathbb{C}}(g-1) = \tfrac{1}{2} \dim_{\mathbb{C}}\mathcal{M}_{G_{\mathbb{C}}}. \end{displaymath} We finish this section by giving two examples of this setup. First consider $G_r = \mathrm{SU}(n,n)$ which is a real form of $\mathrm{SL}(2n,\mathbb{C})$. The Higgs vector bundles $(E,\Phi)$ corresponding to representations of $\pi_1(\Sigma)$ in $\mathrm{SU}(n,n)$ are of the form \begin{equation}\label{eq:suhiggsbundles} E = V \oplus V' \quad \text{and} \quad \Phi = \begin{pmatrix} 0 & b \\ c & 0 \end{pmatrix}, \end{equation} where $V$ and $V'$ are rank $n$ vector bundles with $\Lambda^n V \otimes \Lambda^nV' \cong \mathcal{O}$, $b \in H^0(\Hom(V',V) \otimes K)$, and $c \in H^0(\Hom(V,V') \otimes K)$. Two $\mathrm{SU}(n,n)$ representations are conjugate if and only if the corresponding Higgs bundles of this form are isomorphic by an isomorphism which is of the form \begin{math} \left( \begin{smallmatrix} g & 0 \\ 0 & g' \end{smallmatrix} \right) \end{math} and of determinant one. The second example is $G_r = \Sp(2n,\mathbb{R})$; this is a split real form of $\Sp(2n,\mathbb{C})$. The Higgs vector bundles $(E,\Phi)$ obtained from the standard representation of $\Sp(2n,\mathbb{C})$ on $\mathbb{C}^{2n}$, and corresponding to representations of $\pi_1(\Sigma)$ in $\Sp(2n,\mathbb{R})$ are of the form \begin{equation}\label{eq:sphiggsbundles} E = V \oplus V^* \quad \text{and} \quad \Phi = \begin{pmatrix} 0 & b \\ c & 0 \end{pmatrix}, \end{equation} where $V$ is a rank $n$ vector bundle, $b \in H^0(S^2 V \otimes K)$, and $c \in H^0(S^2 V^* \otimes K)$. Two $\Sp(2n,\mathbb{R})$ representations are conjugate if and only if the corresponding Higgs bundles of this form are isomorphic by an isomorphism which is of the form \begin{math} \left( \begin{smallmatrix} g & 0 \\ 0 & g^{t} \end{smallmatrix} \right) \end{math}. \subsection{$Q$-bundles and triples} \label{sec:q-bundles-triples} The special forms \eqref{eq:suhiggsbundles} and \eqref{eq:sphiggsbundles} suggest a different point of view, that of $Q$-bundles. This notion, due to Alastair King, provides a general framework for considering a large number of the various kinds of vector bundles with extra structure, which have been studied in recent years. The vortex pairs of Bradlow \cite{bradlow}, the triples of Garc{\'\i}a-Prada, introduced in \cite{garcia-prada} and studied systematically by him and Bradlow in \cite{bradlow-garcia-prada} (and also Higgs bundles), are all examples of $Q$-bundles. Let $Q$ be a quiver, that is, $Q$ is a directed graph, specified by a set of vertices $Q_0$ and a set of arrows $Q_1$, together with head and tail maps $h,t \colon Q_1 \to Q_0$. \begin{defn} A \emph{$Q$-bundle\/} over a Riemann surface $\Sigma$ is a collection of holomorphic vector bundles $\{E_i\}_{i \in Q_0}$ over $\Sigma$ and a collection of holomorphic maps $\{\phi_a \colon E_{t(a)} \to E_{h(a)} \}_{a \in Q_1}$. A \emph{twisted $Q$-bundle} is given by in addition specifying a linebundle $L_a$ for each arrow $a$. The maps $\phi_a$ are then required to be holomorphic maps \begin{math} \phi_a \colon E_{t(a)} \to E_{h(a)} \otimes L_a \end{math}. \end{defn} We shall only consider $Q$-bundles of a particularly simple form: we let $Q$ be a quiver with $2$ vertices and exactly one arrow connecting the vertices in each direction (see fig.\ \ref{fig:1}). \begin{figure}[h] \centering \font\thinlinefont=cmr5 \begingroup\makeatletter\ifx\SetFigFont\undefined% \gdef\SetFigFont#1#2#3#4#5{% \reset@font\fontsize{#1}{#2pt}% \fontfamily{#3}\fontseries{#4}\fontshape{#5}% \selectfont}% \fi\endgroup% \mbox{\beginpicture \setcoordinatesystem units <1.00000cm,1.00000cm> \unitlength=1.00000cm \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) \setshadesymbol ({\thinlinefont .}) \setlinear \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 4.847 24.408 5.080 24.289 4.930 24.504 / \circulararc 82.224 degrees from 5.080 24.289 center at 3.810 22.834 \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 2.773 24.011 2.540 24.130 2.690 23.915 / \circulararc 82.224 degrees from 2.540 24.130 center at 3.810 25.585 \put{$\bullet$} [lB] at 2.381 24.130 \put{$\bullet$} [lB] at 5.080 24.130 \put{$a_{21}$} [lB] at 3.810 24.924 \put{$a_{12}$} [lB] at 3.810 23.178 \linethickness=0pt \putrectangle corners at 2.381 25.076 and 5.097 23.139 \endpicture} \caption{The quiver $Q$} \label{fig:1} \end{figure} We denote the the arrows by $a_{ij}$, where $a_{ij}$ is the arrow going from $j$ to $i$. Also, the maps will be twisted by the canonical bundle $K$. Thus, from now on, a $Q$-bundle is a pair \begin{displaymath} \mathbf{E}=(\underline{E},\underline{\Phi}), \end{displaymath} where $\underline{E}=\{E_1,E_2\}$ and $\underline{\Phi}=\{\phi_{ij}\}$. Here, each $E_i$ is a holomorphic vector bundle on $\Sigma$ and $\phi_{ij}$ is a holomorphic section of $\Hom(E_j,E_i \otimes K)$. A particularly interesting special case occurs when $\phi_{12} = 0$. The data of the above type of $Q$-bundle then comes down to a triple $(E_1,E_2,\phi)$, where $\phi \in H^0(\Sigma;\Hom(E_1,E_2)\otimes K)$. If we define $\tilde{E}_2 = E_2 \otimes K$ then this is equivalent to a holomorphic triple $(E_1,\tilde{E}_2,\phi)$ in the sense of Bradlow and Garc{\'\i}a-Prada \cite{bradlow-garcia-prada}. Given a $Q$-bundle $\mathbf{E}=(\underline{E},\underline{\Phi})$, we can define an associated Higgs bundle $(E,\Phi)$ by putting \begin{equation} E = E_1 \oplus E_2 \quad\text{and}\quad \Phi = (\phi_{ij}), \end{equation} where $(\phi_{ij})$ is the matrix of $\Phi$ with respect to the above direct sum decomposition of $E$. Note that the Higgs bundles of the form \eqref{eq:suhiggsbundles} or \eqref{eq:sphiggsbundles} arise in this way. Conversely, given a Higgs bundle of the special form \eqref{eq:suhiggsbundles} or \eqref{eq:sphiggsbundles} we get an associated $Q$-bundle. There are equations for preferred special metrics in a $Q$-bundle, the \emph{$Q$-vortex equations}. Choose a metric compatible with the complex structure on $\Sigma$ and, for convenience, normalize it so that $\mathrm{vol}(\Sigma) = 2\pi$. This of course also gives a Hermitian metric in the canonical bundle $K$. The $Q$-vortex equations are equations for Hermitian metrics in $E_1$ and $E_2$ and in our case they take the form \begin{equation} \label{eq:qvortex} \left\{ \begin{split} \mathit{i} \Lambda F(A_1) + \phi_{12}\phi_{12}^* - \phi_{21}^*\phi_{21} &= \tau_1 \mathrm{Id}_{E_1} \\ \mathit{i} \Lambda F(A_2) + \phi_{21}\phi_{21}^* - \phi_{12}^*\phi_{12} &= \tau_2 \mathrm{Id}_{E_2} \\ \end{split}\right. \end{equation} where $F(A_i)$ is the curvature of the metric connection in $E_i$, $\Lambda$ denotes contraction with the K{\"a}hler form of $\Sigma$, and $\phi_{ij}^*$ denotes the adjoint taken with respect to the metric obtained from the metrics on $E_i$ and $K$. The parameters $(\tau_1,\tau_2)$ are real, subject to the condition \begin{displaymath} \sum_{i=1}^2 \bigl(\deg(E_i) - \tau_i \rk(E_i) \bigr) = 0, \end{displaymath} obtained by taking traces in the equations \eqref{eq:qvortex}, summing and integrating over $\Sigma$ (thus there is really only one real parameter involved, which is usually taken to be $\tau=\tau_2$). There is a stability condition for $Q$-bundles, such that any $Q$-bundle which supports a solution to the $Q$-vortex equations is a direct sum of stable $Q$-bundles. In our case the condition is \begin{equation} \label{eq:taustability} \sum_{i=1}^2 \bigl(\deg(F_i) - \tau_i \rk(F_i) \bigr) < 0 \end{equation} for any proper $Q$-subbundle $\mathbf{F}$ of $\mathbf{E}$. Note that the condition depends on the parameters $(\tau_1,\tau_2)$. Bradlow and Garc{\'\i}a-Prada \cite{bradlow-garcia-prada} constructed moduli spaces of stable triples, varying with the parameter $\tau$. We shall only need to consider the case $\tau_1 = \tau_2 = \mu(E)$ so we shall assume this from now on. The stability condition \eqref{eq:taustability} can then be reformulated as \begin{equation} \label{eq:alphastability} \mu(F) < \mu(E) \end{equation} for any proper $Q$-subbundle $\mathbf{F}=(\{F_1,F_2\},\{\phi_{12},\phi_{21}\})$ of $\mathbf{E}$, and where we write $F = F_1 \oplus F_2$. But, obviously, $F = F_1 \oplus F_2 \subset E$ is a $\Phi$-invariant subbundle, thus stability of the Higgs bundle $(E,\Phi)$ implies stability of the $Q$-bundle $\mathbf{E}$. The following lemma will allow us to conclude that the converse also holds. \begin{lemma} \label{lemma:quiver} Let $(E,\Phi)$ be a Higgs bundle of the form $E = E_1 \oplus E_2$ and \begin{displaymath} \Phi = \begin{pmatrix} 0 & \phi_{12} \\ \phi_{21} & 0 \end{pmatrix}. \end{displaymath} Let $\mathbf{E} = (\{E_1,E_2\} , \{\phi_{12},\phi_{21}\})$ be the associated $Q$-bundle. Let $F' \subset E$ be a $\Phi$-invariant subbundle. Then there is a $Q$-subbundle $\mathbf{E}'= (\{E'_1,E'_2\} , \{\phi_{12},\phi_{21}\})$ of $\mathbf{E}$ such that \begin{displaymath} \mu(F') \leq \mu(E'), \end{displaymath} where $E'=E'_1 \oplus E'_2$. \end{lemma} \begin{proof} Let $\pi_{i}\colon E \to E_{i}$ be the projection on the $i$th factor. Let $F_{i} \subset E_i$ and $G_{i} \subset F'$ be the subbundles which are generated by the image and kernel of $\pi_i$, respectively. Then $F_1$ and $G_2$ are contained in $E_1$, $F_2$ and $G_1$ are contained in $E_2$, and we have sequences of vector bundles \begin{displaymath} 0 \to G_i \to F' \to F_i \to 0, \end{displaymath} which are generically short exact. Hence, \begin{math} \deg(F') \leq \deg(G_i) +\deg(F_i), \end{math} and putting $F = F_1 \oplus F_2$ and $G = G_2 \oplus G_{1}$, it follows that \begin{displaymath} 2 \deg(F') \leq \deg(F) + \deg(G). \end{displaymath} Clearly \begin{math} 2 \rk(F') = \rk(F) + \rk(G), \end{math} so that \begin{equation} \label{eq:convex} \mu(F') \leq \frac{\rk(F)}{\rk(F) + \rk(G)} \mu(F) + \frac{\rk(G)}{\rk(F) + \rk(G)} \mu(G), \end{equation} and therefore either $\mu(F) \geq \mu(F')$ or $\mu(G) \geq \mu(F')$. Provided that $F$ and $G$ give $Q$-subbundles of $\mathbf{E}$ we can then take $\mathbf{E}'$ to be the $Q$-bundle associated to either $F$ or $G$. It thus remains to see show that $F$ and $G$ are $\Phi$-invariant and, therefore, define $Q$-sub\-bundles of $\mathbf{E}$. First, let $x_1 \in F_1$. If we write $x_1 = \pi_{1}(x)$ for some $x = x_1 + x_2$ in $F'$, then \begin{displaymath} \Phi(x) = \Phi(x_1) + \Phi(x_2). \end{displaymath} By our assumption on the matrix for $\Phi$, it follows that $\Phi(x_1) \in E_2$ and $\Phi(x_2) \in E_1$. Then $\pi_{1} (\Phi (x)) = \Phi (x_2) \in E_{1}$ and $\pi_2 (\Phi(x)) = \Phi (x_1) \in E_2$. But $\Phi(x) \in F'$ because $F'$ is $\Phi$-invariant, and thus $\Phi(x_2) \in F_{1}$ and $\Phi(x_1) \in F_{2}$. Of course, we can repeat the argument with $x_2 \in F_2$ and hence, $F$ is $\Phi$-invariant. The proof that $G$ is $\Phi$-invariant is similar. Let $x_1 \in G_2$. By assumption, $\Phi(x_1) \in E_{2}$. But $G_2 \subset F'$, so $\Phi(x_1) \in F'$ as well. It follows that $\Phi(x_1) \in G_{1}$ and thus, $G$ is $\Phi$-invariant. We have thus seen that $F$ and $G$ define $Q$-subbundles of $\mathbf{E}$ and this finishes the proof. \end{proof} As an immediate consequence we have the following theorem. \begin{thm} \label{thm:quiver} Let $Q$ be a quiver with two vertices and one arrow connecting the vertices in each direction, and let $\mathbf{E} = (\{E_1,E_2\} , \{ \phi_{12},\phi_{21}\})$ be a $Q$-bundle. Let $(E,\Phi)$ be the associated Higgs bundle as above; thus $E = E_1 \oplus E_2$ and \begin{displaymath} \Phi = \begin{pmatrix} 0 & \phi_{12} \\ \phi_{21} & 0 \end{pmatrix}. \end{displaymath} Then $\mathbf{E}$ is stable if and only if $(E,\Phi)$ is. Furthermore, if $(E,\Phi)$ is poly-stable, i.e.\ the direct sum of lower rank stable Higgs bundles, these lower rank Higgs bundles are $Q$-subbundles of $\mathbf{E}$. \end{thm} \begin{proof} The only assertion that requires proof is the final one. Suppose that $(E,\Phi)$ is poly-stable and that $(F',\Phi')$ is a proper stable Higgs subbundle of $(E,\Phi)$ with $\mu(F') = \mu(E)$. By semi-stability of $(E,\Phi)$ the bundle called $G$ in the proof of \lemref{lemma:quiver} must then satisfy $\mu(G) = \mu(E) = \mu(F')$ and, since $G \subseteq F'$, it follows by stability of $F'$ that $G=F'$. But $G$ is a $Q$-subbundle so this finishes the proof. \end{proof} \subsection{Connected components and Morse theory} \label{sec:morse} We shall use Hitchin's method \cite{hitchin,hitchinlie}, which we shall now review, for finding the connected components of $\mathcal{M}_{G_{r}}$. The idea is to use discrete invariants of flat bundles for dividing $\mathcal{M}_{G_{r}}$ into subspaces which are unions of connected components. We then show that these subspaces are, in fact, connected. For this consider the moduli space $\mathcal{M}_{G_r}$ of Higgs bundles as the space of solutions $(A,\Phi)$ to Hitchin's equations modulo gauge equivalence. The function \begin{align*} f \colon \mathcal{M} &\to \mathbb{R} \\ (A,\Phi) &\mapsto \norm{\Phi}^2 = \int_{\Sigma}\abs{\Phi}^2\d\mathrm{vol} \end{align*} is proper. Thus, a subspace $N$ of $\mathcal{M}$ is connected if the subspace of local minima of $f$ on $N$ is connected. Restrict for a moment attention to irreducible solutions to Hitchin's equations (i.e.\ stable Higgs bundles); these are smooth points of $\mathcal{M}$. In order to identify the subspaces of local minima of $f$ one uses the fact that it is a moment map for the $S^1$ action on $\mathcal{M}$, given by $(A,\Phi) \mapsto (A,\mathrm{e}^{\mathit{i}\theta}\Phi)$: this implies that the critical points of $f$ are exactly the fixed points of the circle action. Now, $(A,\Phi)$ represents a fixed point if and only if there is an infinitesimal gauge transformation $\psi \in \Omega^0(\Sigma; P_K\times_{\Ad}\mathfrak{k})$ such that \begin{align} d_A\psi &=0, \\ [\psi,\Phi] &= \mathit{i}\Phi. \label{eq:psiPhi} \end{align} Let $(E,\Phi)$ be a Higgs vector bundle obtained from a complex representation of $K$, then this can be decomposed in eigenspaces for the covariantly constant gauge transformation $\psi$. Thus \begin{equation}\label{eq:decomp} E = \bigoplus_{m} U_m, \end{equation} where $\psi_{\vert U_m} = \mathit{i} m$. Then \eqref{eq:psiPhi} shows that $\Phi \colon U_m \to U_{m+1} \otimes K$. The case \begin{displaymath} E = \Ad P_{\mathbb{C}} = \Ad_{\mathfrak{k}_{\mathbb{C}}} P_{K} \oplus \Ad_{\mathfrak{k}_{\mathbb{C}}^{\perp}} P_{K}, \end{displaymath} is of particular interest. Since the adjoint action of $\psi \in \mathfrak{k}$ and the involution $\phi$ on $\mathfrak{g}_{\mathbb{C}}$ commute, this decomposition of $\Ad P_{\mathbb{C}}$ is compatible with the decomposition $\Ad P_{\mathbb{C}} = \bigoplus_{m} U_m$. From $\ad(\Phi) \colon U_m \to U_{m+1} \otimes K$ we conclude that $\Phi \in H^0(\Sigma; U_1 \otimes K)$ and since $\Phi \in H^0(\Sigma;\Ad_{\mathfrak{k}_{\mathbb{C}}^{\perp}} P_{K})$ it follows that $U_1 \subset \Ad_{\mathfrak{k}_{\mathbb{C}}^{\perp}} P_{K}$. Furthermore, $\ad(\Phi)$ interchanges $\Ad_{\mathfrak{k}_{\mathbb{C}}} P_{K}$ and $\Ad_{\mathfrak{k}_{\mathbb{C}}^{\perp}} P_{K}$ and we therefore have \begin{align*} \Ad_{\mathfrak{k}_{\mathbb{C}}} P_{K} &= \bigoplus_{k} U_{2k}, \\ \Ad_{\mathfrak{k}_{\mathbb{C}}^{\perp}} P_{K} &= \bigoplus_{k} U_{2k+1}, \end{align*} where $k$ is integer. Two additional pieces of information will be useful. The first is that there is an isomorphism $\Ad P_{\mathbb{C}} \xrightarrow{\cong} \Ad P_{\mathbb{C}}^*$ from the adjoint bundle to the co-adjoint bundle given by the Killing form on $\mathfrak{g}_{\mathbb{C}}$ and it is trivial to check that under this \begin{equation} \label{eq:UmtoU-m*} U_m \xrightarrow{\cong} U_{-m}^*. \end{equation} The second useful piece of information is that one can calculate the value of the Morse function $f$ at a Higgs bundle of the form \eqref{eq:decomp}: denoting the component of $\Phi$ mapping $U_i$ to $U_{i+1} \otimes K$ by $\phi_i$ one shows easily, using Hitchin's equations, that \begin{equation} \label{eq:vanishing} \norm{\phi_i}^2 = \norm{\phi_{i-1}}^2 + \bigl(\deg(U_i) - \mu(E)\rk(U_i) \bigr). \end{equation} In particular, if $\mu(U_i) = \mu(E)$ for all $i$ then $\Phi$ must vanish. Finally, consider the case of local minima of $f$ which are not represented by stable Higgs bundles. For simplicity we restrict attention to bundles of the form \eqref{eq:suhiggsbundles} or \eqref{eq:sphiggsbundles}. These are then direct sums of stable Higgs bundles of lower rank. But from \thmref{thm:quiver} we can conclude that these lower rank Higgs bundles decompose as direct sums of subbundles of the bundles appearing in the direct sum decomposition of the original Higgs bundle. \subsection{Morse indices} \label{sec:morseindices} Given a poly-stable Higgs bundle $(P_{K_{\mathbb{C}}},\Phi)$ which represents a critical point of $f$ it is necessary to decide whether it is a local minimum. This can be done using the observation of Hitchin \cite{hitchinlie} that if $\psi$ acts with weight $m$ on an element of $\Ad_{\mathfrak{k}_{\mathbb{C}}}P_K$ then the corresponding eigenvalue of the Hessian of $f$ is $-m$, while a weight $m$ on $(\Ad_{\mathfrak{k}_{\mathbb{C}}}P_K)\otimes K$ gives the eigenvalue $1-m$. It follows that the subspace of $T_{(P_{K_{\mathbb{C}}},\Phi)}\mathcal{M}_{G_r}$ on which the Hessian of $f$ is negative definite is $\mathbb{H}^1(C^{\bullet}_{-})$, where $C^{\bullet}_{-}$ is the complex of sheaves \begin{displaymath} C^{\bullet}_{-} : \mathcal{O}\Bigl(\bigoplus_{k \geq 1} U_{2k}\Bigr) \xrightarrow{\ad(\Phi)} \mathcal{O} \Bigl(\bigoplus_{k \geq 1} U_{2k+1} \otimes K\Bigr). \end{displaymath} In other words, the Morse index is $\dim_{\mathbb{R}}\mathbb{H}^1(C^{\bullet}_{-})$. At a smooth point of $\mathcal{M}_{G_r}$, where $\mathbb{H}^0(C^{\bullet}_{-})=\mathbb{H}^2(C^{\bullet}_{-})=0$, the Morse index can be calculated using the Riemann-Roch theorem: \begin{multline}\label{eq:morseindex} \dim_{\mathbb{C}} \mathbb{H}^1(C^{\bullet}_{-}) = \chi\biggl(\mathcal{O} \Bigl(\bigoplus_{k \geq 1} U_{2k+1} \otimes K\Bigr)\biggr) - \chi\biggl(\mathcal{O}\Bigl(\bigoplus_{k \geq 1} U_{2k}\Bigr)\biggr)\\ = (g-1) \sum_{k \geq 1} \bigl(\rk(U_{2k}) + \rk(U_{2k+1})\bigr) + \sum_{k \geq 1} (\deg(U_{2k+1}) - \deg(U_{2k})). \end{multline} The stable Higgs bundle $(P_{K_{\mathbb{C}}},\Phi)$ represents a local minimum of $f$ if and only if this number is zero. Finally consider the case of reducible Higgs bundles. It is shown in \cite{hitchinlie} that if $\mathbb{H}^1(C^{\bullet}_{-}) = 0$ then we continue to have a local minimum and, on the other hand, if there is an element of $\mathbb{H}^1(C^{\bullet}_{-})$ on which $\ddot{f}$ is negative and which is tangent to a smooth family of deformations of the Higgs bundle, then this does not represent a local minimum. \subsection{A theorem of Laumon} \label{sec:laumon} We conclude this section by a digression to the moduli space of flat $G_{\mathbb{C}}$ bundles. This space has a holomorphic symplectic form at its smooth points (it is hyper-K{\"a}hler). Of course the theory outlined above applies in this case. In particular, if the moduli space is smooth (for example the moduli space of stable Higgs vector bundles with rank and degree co-prime) one can consider the Morse flow on it. At a critical point, we again have the decomposition $\Ad P_{\mathbb{C}} = \bigoplus_{m}U_m$. The subspace $T_{\leq 0}$ of the tangent space to the moduli space on which the Hessian is less than or equal to zero is $\mathbb{H}^1$ of the following complex of sheaves \begin{displaymath} \mathcal{O}\Bigl(\bigoplus_{m \geq 0} U_{m}\Bigr) \xrightarrow{\ad(\Phi)} \mathcal{O} \Bigl(\bigoplus_{m \geq 1} U_{m} \otimes K\Bigr). \end{displaymath} The dimension of this is given by Riemann-Roch as \begin{align*} \dim_{\mathbb{C}} T_{\leq 0} &= \chi \biggl(\mathcal{O} \Bigl(\bigoplus_{m \geq 1} U_{m} \otimes K\Bigr)\biggr) - \chi \biggl(\mathcal{O}\Bigl(\bigoplus_{m \geq 0} U_{m}\Bigr)\biggr)\\ &= (g-1)\Bigl(\rk(U_0) + \sum_{m \geq 1} 2\rk(U_m)\Bigr) - \deg(U_0). \end{align*} But from \eqref{eq:UmtoU-m*} we have $\deg(U_0)=0$ and $\rk(U_0) + \sum_{m \geq 1} 2\rk(U_m) = \dim_{\mathbb{C}} \mathfrak{g}_{\mathbb{C}}$ so that \begin{displaymath} \dim_{\mathbb{C}} T_{\leq 0} = \dim_{\mathbb{C}}(\mathfrak{g}_{\mathbb{C}}) = \tfrac{1}{2}\dim_{\mathbb{C}}\mathcal{M}_{G_{\mathbb{C}}}. \end{displaymath} It was pointed out by Hausel \cite{hausel} that this fact, together with his theorem that the downwards Morse flow coincides with the nilpotent cone (the pre-image of $0$ under the Hitchin map), implies a theorem of Laumon \cite[Th.\ 3.1]{laumon} in this context: The nilpotent cone in $\mathcal{M}$ is a Lagrangian subvariety with respect to the holomorphic symplectic form on $\mathcal{M}$. \section{Milnor-Wood inequalities} \label{sec:milnorwood} For any $G$ there is a locally constant obstruction map \begin{displaymath} o_2 \colon \Hom(\pi_1(\Sigma);G) \to H^2(\Sigma;\pi_1(G)). \end{displaymath} Note that in both cases, $G=\mathrm{SU}(n,n)$ and $G=\Sp(2n,\mathbb{R})$, $\pi_1(G) = \mathbb{Z}$ so that we have an integer valued function. In the case of representations in $\Sp(2n,\mathbb{R})$, we have $o_2(\rho) = c_1(V)$, where $V$ is the vector bundle appearing in the decomposition \eqref{eq:sphiggsbundles} of the Higgs bundle associated to $\rho$. In the case of $\mathrm{SU}(n,n)$ representations, we have $o_2(\rho) = c_1(V)$, where $V$ is the vector bundle appearing in \eqref{eq:suhiggsbundles}. In both cases we thus have an integer valued function $d = \deg(V) = \langle c_1(V), [\Sigma] \rangle$ whose fibres are unions of connected components. There is an outer automorphism of $\mathcal{M}_G$ given by exchanging $V$ with $V^*$ (in the $\Sp(2n,\mathbb{R})$ case), or exchanging $V$ and $V'$ (in the $\mathrm{SU}(n,n)$ case). Thus, in both cases we have an isomorphism between $o_2^{-1}(d)$ and $o_2^{-1}(-d)$, and it therefore suffices to consider the case $d \geq 0$, whenever convenient. It is well known that there are bounds on the possible values of characteristic numbers of flat bundles, known as Milnor-Wood inequalities and, using Higgs bundles, we shall prove one for flat $\mathrm{SU}(n,n)$- and $\Sp(2n,\mathbb{R})$-bundles. The original inequality proved by J.\ Milnor \cite{milnor} concerns $\mathrm{SL}(2,\mathbb{R})$-bundles, while J.\ Wood \cite{wood} considered $\mathrm{SU}(1,1)$-bundles. J.\ Dupont \cite{dupont} found a bound for any semi-simple group with finite centre, however, the inequality of \propref{prop:milwood} below for $G = \Sp(2n,\mathbb{R})$ is sharper than his. Using ideas of Gromov, A.\ Domic and D.\ Toledo \cite{domic-toledo} proved a general result for mappings of a surface into manifolds covered by bounded symmetric domains, and their work implies \propref{prop:milwood} below. Hitchin obtained a proof in the case of flat reductive $\mathrm{SL}(2,\mathbb{R})$-bundles, using Higgs bundles, in \cite{hitchin}, and we obtain our inequality in a similar way. The reason why we include the proof here is, that it gives crucial extra information about the poly-stable Higgs bundles of the form \eqref{eq:suhiggsbundles} and \eqref{eq:sphiggsbundles} (see \propref{prop:iso}). \begin{prop} \label{prop:milwood} Let $\rho$ be a reductive representation of $\pi_1(\Sigma)$ in $\Sp(2n,\mathbb{R})$ or $\mathrm{SU}(n,n)$. Then the characteristic number $d = \langle o_2(\rho),[\Sigma] \rangle$ satisfies the inequality \begin{displaymath} \abs{d} \leq n(g-1). \end{displaymath} \end{prop} \begin{proof} We give the proof in case of $\Sp(2n,\mathbb{R})$-representations, the $\mathrm{SU}(n,n)$ case being completely analogous. Let $(E,\Phi)$ be the poly-stable Higgs bundle of the form \eqref{eq:sphiggsbundles} corresponding to $\rho$, as we already noticed $d = \deg(V)$. Without loss of generality we can assume that $d>0$. In this case $c\neq 0$, as otherwise $V$ would be $\Phi$-invariant, and therefore violate the stability condition. Let $U$ be the subbundle of $V^*$, such that $U \otimes K$ is the vector bundle generated by the image of $c$. Similarly, let $U' \subset V$ be the subbundle, which is generated by the kernel of $c$. Then the bundles $U'$ and $V \oplus U$ are both $\Phi$-invariant. We therefore get the following inequalities from semi-stability of $(E,\Phi)$: \begin{align} \deg(U') &\leq 0 \label{ineqone} \\ d+\deg(U) &\leq 0. \label{ineqtwo} \end{align} Note that these inequalities also hold in the case when $U' = 0$ and $U = V^*$. Next, we note that $c$ induces a non-trivial global section of the linebundle \begin{displaymath} \det(V/U')^{-1} \otimes \det(U \otimes K), \end{displaymath} which therefore has positive degree, i.e. \begin{equation} \label{degline} \deg(U')-d+\deg(U)+(2g-2)\rk(c) \geq 0, \end{equation} where $\rk(c) = \rk(U)$ is the generic rank of $c$. Combining this with the inequalities (\ref{ineqone}) and (\ref{ineqtwo}), we obtain \begin{equation} \label{eq:degreeVleq} d \leq (g-1) \rk(c), \end{equation} so $d\leq n(g-1)$ as claimed. \end{proof} The above proof gives some important additional information: from \eqref{eq:degreeVleq} it follows that $\rk(c) = n$ for $d > (n-1)(g-1)$. In particular, in the extremal case $d=n(g-1)$, we have $\rk(c)=n$, and furthermore equality holds in (\ref{degline}). Hence, $\det(c)$ is a non-zero section of a linebundle of degree $0$, and we conclude that $c$ is an isomorphism. We thus have the following proposition: \begin{prop}\label{prop:iso} Let $(E,\Phi)$ be the poly-stable Higgs bundle of the form \eqref{eq:sphiggsbundles} corresponding to a reductive representation of $\pi_1(\Sigma)$ in $\Sp(2n,\mathbb{R})$. If $\deg(V)=n(g-1)$ then $c \colon V \to V^* \otimes K$ is an isomorphism. Let $(E,\Phi)$ be the poly-stable Higgs bundle of the form \eqref{eq:suhiggsbundles} corresponding to a reductive representation of $\pi_1(\Sigma)$ in $\mathrm{SU}(n,n)$. If $\deg(V)=n(g-1)$ then $c \colon V \to V' \otimes K$ is an isomorphism. \qed \end{prop} This has as a consequence that there is another discrete invariant on $\mathcal{M}_{\Sp(2n,\mathbb{R})}$ and we shall come back to this in \secref{sec:spcomp}. \section{Minima of $f$} \label{sec:minima} In this section we determine the poly-stable Higgs bundles which represent local minima of the function $f$ on $\mathcal{M}_{G_r}$ in the cases $G_r = \mathrm{SU}(2,2)$ and $G_r = \Sp(4,\mathbb{R})$. We shall determine which stable Higgs bundles correspond to critical points of $f$ and then identify those which are local minima using \eqref{eq:morseindex}. It will be convenient to consider the decomposition \eqref{eq:decomp}, $E = \bigoplus_m F_m$ of the Higgs bundles $(E,\Phi)$ of the form \eqref{eq:suhiggsbundles} and \eqref{eq:sphiggsbundles}, which then gives rise to the decomposition of the adjoint bundle: note that we have $\Ad P_{\mathbb{C}} = \End(V \oplus V')_0$ (the subscript $0$ indicating traceless endomorphisms) when $G_r = \mathrm{SU}(n,n)$, while $\Ad P_{\mathbb{C}} = \End(V) \oplus S^2V \oplus S^2V^*$ for $G_r = \Sp(2n,\mathbb{R})$. We begin by finding the minima on $\mathcal{M}_{\mathrm{SU}(2,2)}$ which are stable Higgs bundles, leaving the reducible ones for later. As noted in \secref{sec:milnorwood}, we only need to consider Higgs bundles $E= V \oplus V'$ with $\deg(V) \geq 0$. \begin{prop}\label{prop:suminima} The stable Higgs bundles of the form \eqref{eq:suhiggsbundles} with $\deg(V) \geq 0$, which correspond to a local minimum of $f$ on $\mathcal{M}_{\mathrm{SU}(2,2)}$ are the ones which have $b = 0$, $c \neq 0$, and $\deg(V) > 0$. \end{prop} \begin{proof} Let $(E,\Phi)$ be a Higgs bundle of the form \eqref{eq:suhiggsbundles} which represents a critical point of $f$. $E$ comes from the standard representation of $\mathrm{S}(\mathrm{U}(2)\times\mathrm{U}(2))$ on $\mathbb{C}^2 \oplus \mathbb{C}^2$ and the infinitesimal gauge transformation $\psi$ which produces the decomposition $E=\bigoplus_m F_m$ is fibrewise in $\mathfrak{s}(\mathfrak{u}(2)\times\mathfrak{u}(2))$. Hence each of the bundles $F_m$ is of the form $F_m = F_{m1} \oplus F_{m2}$ where $F_{mi} = V_i \cap F_m \subseteq V_i$ for $i=1,2$. We claim that either $F_{m1}$ or $F_{m2}$ must be zero (unless $E=U_0$). To see this, let $m_0$ be the smallest $m$ such that $F_{m1}$ and $F_{m2}$ are both non-zero. Then $\Phi(U_{m_0-1})$ is contained in either $V$ or $V'$, since the same is true for $U_{m_0-1}$ and $\Phi$ interchanges $V$ and $V'$. Without loss of generality we may suppose that $\Phi(U_{m_0-1}) \subseteq V$. Then each of the bundles \begin{displaymath} \bigoplus_{m < m_0} F_m \oplus \Bigl(V \cap \bigoplus_{m \geq m_0} F_m\Bigr) \end{displaymath} and \begin{displaymath} V' \cap \bigoplus_{m \geq m_0} F_m \end{displaymath} is $\Phi$-invariant, and so we have a decomposition of $(E,\Phi)$ as a direct sum of lower rank Higgs bundles. This is impossible because $(E,\Phi)$ is stable. Let $r = (\rk(F_m))$ be the \emph{rank vector} whose entries are the ranks of the bundles $F_m$. We analyze the possibilities for $r$ case by case. \emph{1st case: $r=(1,1,1,1)$.} Note that $0= \tr(\psi) = \mathit{i}\sum m\rk(F_m)$. In this case we therefore have $E = F_{-3/2} \oplus F_{-1/2} \oplus F_{1/2} \oplus F_{3/2}$, where each $F$ is a linebundle. Hence the decomposition \eqref{eq:decomp} of $\Ad P_{\mathbb{C}}$ is of the form \begin{displaymath} \Ad P_{\mathbb{C}} = U_{-3} \oplus \cdots \oplus U_{3}, \end{displaymath} where \begin{align*} U_2 &= \Hom(F_{-3/2},F_{1/2}) \oplus \Hom(F_{-1/2},F_{3/2}), \\ U_3 &= \Hom(F_{-3/2},F_{3/2}). \end{align*} The formula \eqref{eq:morseindex} for the Morse index then takes the form \begin{displaymath} \begin{split} \dim_{\mathbb{C}} \mathbb{H}^1(C^{\bullet}_{-}) &= 3(g-1) + \deg(F_{3/2}) - \deg(F_{-3/2}) \\ & \quad - \bigl(\deg(F_{1/2}) - \deg(F_{-3/2}) + \deg(F_{3/2}) - \deg(F_{-1/2}) \bigr) \\ &= 3(g-1) + \deg(F_{-1/2}) - \deg(F_{1/2}). \end{split} \end{displaymath} Now we note that $F_{1/2} \oplus F_{3/2}$ is a $\Phi$-invariant subbundle of $E$ and thus, by stability, $\deg(F_{1/2}) + \deg(F_{3/2}) < 0$. Combining this with the above result we get \begin{displaymath} \dim_{\mathbb{C}} \mathbb{H}^1(C^{\bullet}_{-}) > 3(g-1) + \deg(F_{-1/2}) + \deg(F_{3/2}). \end{displaymath} But since $\Phi$ interchanges $V$ and $V'$ we must have $V = F_{-3/2} \oplus F_{1/2}$ and $V' = F_{3/2} \oplus F_{-1/2}$ or vice-versa. Therefore $\abs{d} = \abs{\deg(V)} = \abs{\deg(V')} = \abs{\deg(F_{-1/2}) + \deg(F_{3/2})}$. Combining this with the above inequality we get \begin{displaymath} \dim_{\mathbb{C}} \mathbb{H}^1(C^{\bullet}_{-}) > 3(g-1) - \abs{d} \geq 0, \end{displaymath} where the last inequality comes from the Milnor-Wood inequality $\abs{d} \leq 2(g-1)$ of \propref{prop:milwood}. We conclude that a critical point of this type always has strictly positive Morse index and hence it cannot be a local minimum of $f$. \emph{2nd case: $r=(1,2,1)$.} Again using $\sum m \rk(F_m) = 0$ we see that in this case $E = F_{-1} \oplus F_{0} \oplus F_{1}$. We then have $\Ad P_{\mathbb{C}} = U_{-2} \oplus \cdots \oplus U_2$ and \begin{displaymath} U_2 = \Hom(F_{-1},F_{1}), \end{displaymath} and so, from \eqref{eq:morseindex}, we get \begin{align*} \dim_{\mathbb{C}} \mathbb{H}^1(C^{\bullet}_{-}) &= g-1 - \bigl( \deg(F_1) - \deg(F_{-1})\bigr) \\ &= g-1 - \bigl(2\deg(F_1) + \deg(F_{0})\bigr), \end{align*} where the second equality is due to the fact that $\deg(E) = 0$. Since $F_0 \oplus F_1$ and $F_1$ are $\Phi$-invariant we get from stability that $\deg(F_0) + \deg(F_1) < 0$ and $\deg(F_1)<0$. Hence $2\deg(F_1) + \deg(F_{0}) <0 $, which shows that $\dim_{\mathbb{C}} \mathbb{H}^1(C^{\bullet}_{-}) > 0$. Therefore a critical point of this type cannot be a minimum of $f$ either. \emph{3rd case: $r=(1,1,2)$ (or $r=(2,1,1)$).} In this case $E = F_{m_1} \oplus F_{m_2} \oplus F_{m_3}$ where $V = F_{m_1} \oplus F_{m_2}$ and $V' = F_{m_3}$ (or vice-versa). Since $\Phi$ interchanges $V$ and $V'$, it follows that $\Phi_{\vert F_{m_1}} = 0$ and so, $(E,\Phi)$ is reducible. Thus this case cannot occur. The case $r=(2,1,1)$ is analogous. \emph{4th case: $r=(2,2)$.} In this case $E = F_{-1/2} \oplus F_{1/2}$. Then $U_m = 0$ for $m \geq 2$ and hence we see from \eqref{eq:morseindex} that these critical points are local minima of $f$. Clearly $F_{-1/2} = V$ and $F_{1/2} = V'$, or vice-versa. If $V = F_{1/2}$ it would be $\Phi$-invariant and so $\deg(V) < 0$ which is absurd. Thus, in fact, $V = F_{-1/2}$ and $V' = F_{1/2}$. In the notation of \eqref{eq:suhiggsbundles} this means that $c = \Phi_{\vert V}$ and $b = \Phi_{\vert V'} = 0$. This gives the minima with $\deg(V) > 0$. Finally, note that if $\deg(V) = \deg(V') = 0$ then either $V$ or $V'$ is a $\Phi$-invariant subbundle which violates stability. Thus there are no stable Higgs bundles with $\deg(V) = 0$ which are local minima of $f$. \end{proof} The fact that there is another discrete invariant for flat $\Sp(2n,\mathbb{R})$-bundles (cf.\ \secref{sec:spcomp}) is reflected in the difference between the previous and the following result. \begin{prop}\label{prop:spminima} The stable Higgs bundles of the form \eqref{eq:sphiggsbundles} with $\deg(V) \geq 0$, which correspond to a local minimum of $f$ on $\mathcal{M}_{\Sp(4,\mathbb{R})}$ are the ones which have \begin{enumerate} \item $b = 0$, $c \neq 0$, and $\deg(V) > 0$. \item $\deg(V) = 2g-2$, $V=L_1 \oplus L_2$, and $\Phi$ of the form \begin{displaymath} \begin{pmatrix} 0 & 0 & 0 & \tilde{c} \\ 0 & 0 & \tilde{c} & 0 \\ 0 & 0 & 0 & 0 \\ 0 & \tilde{b} & 0 & 0 \end{pmatrix} \end{displaymath} with respect to the decomposition $E = V \oplus V^* = L_1 \oplus L_2 \oplus L_1^{-1} \oplus L_2^{-1}$. \end{enumerate} \end{prop} \begin{proof} This is analogous to the proof of \propref{prop:suminima}. However, in this case the infinitesimal gauge transformation $\psi$ which produces the decomposition $E = \bigoplus F_m$ of the Higgs bundle of the form \eqref{eq:sphiggsbundles} belongs to $\Omega^0(\Sigma;\Ad P_K)$, that is, it is fibrewise in $\mathfrak{u}(2)$. Thus there are only two possibilities: either $V = F_{-1/2}$ and $V^* = F_{1/2}$ with $\Phi \colon V \to V^* \otimes K$, that is, $b=0$ (here we are using that $\deg(V) \geq 0$. These Higgs bundles are seen to be minima as before. The other possibility is that $V = F_{m_1} \oplus F_{m_2}$ and $V^* = F_{-m_1} \oplus F_{-m_2}$, where $F_{-m} = F_{m}^*$. Note that either $(m_1,m_2) = (-3/2,1/2)$ or $(m_1,m_2) = (-1/2,3/2)$. In this case the decomposition \eqref{eq:decomp} has the form \begin{displaymath} \Ad P_{\mathbb{C}} = U_{-3} \oplus \cdots \oplus U_{3}, \end{displaymath} where \begin{align*} U_2 &= \Hom(F_{-3/2},F_{1/2}) \cong \Hom(F_{-1/2},F_{3/2}), \\ U_3 &= \Hom(F_{-3/2},F_{3/2}). \end{align*} From \eqref{eq:morseindex} we therefore get the Morse index \begin{displaymath} \begin{split} \dim_{\mathbb{C}} \mathbb{H}^1(C^{\bullet}_{-}) &= 2(g-1) + \deg(F_{3/2}) - \deg(F_{-3/2}) \\ & \quad - \bigl(\deg(F_{1/2}) - \deg(F_{-3/2})\bigr) \\ &= 2(g-1) - \bigl(\deg(F_{-3/2}) + \deg(F_{1/2})\bigr) \\ &= 2(g-1) \pm \deg(V). \end{split} \end{displaymath} Thus we cannot have a minimum unless $\deg(V) = 2g-2$, and in this case, from \propref{prop:iso}, we have $V = F_{-3/2} \oplus F_{1/2}$ since otherwise $c$ would not be of rank $2$. This gives the second case of the proposition. \end{proof} It remains to identify the local minima of $f$ which are not stable Higgs bundles. \begin{prop} \label{prop:suminima2} The reducible Higgs bundles of the form \eqref{eq:suhiggsbundles} with $\deg(V) \geq 0$ which correspond to a local minimum of $f$ on $\mathcal{M}_{\mathrm{SU}(2,2)}$ either have $\Phi = 0$ and $\deg(V) = \deg(V') = 0$, or, if $\Phi \neq 0$, they are direct sums of rank $2$ Higgs bundles $(E_1,\Phi_1)$ and $(E_2,\Phi_2)$, where $E_i = L_i \oplus L_i'$, $L_i$ a line-bundle with $\deg(L_i) \geq 0$ and $\Phi_i \colon L_i \to L_i' \otimes K$. If $\Phi_i \neq 0$ then $\deg(L_i) > 0$. \end{prop} \begin{proof} Let $(E,\Phi)$ be a reducible Higgs bundle of the form \eqref{eq:suhiggsbundles} which is a local minimum of $f$. Consider $\mathcal{M}_{\mathrm{SU}(2,2)}$ as the space of solutions $(A,\Phi)$ to Hitchin's equations modulo $\mathrm{S}(\mathrm{U}(2)\times \mathrm{U}(2))$ gauge equivalence. First consider the case $\Phi=0$. Then, by poly-stability, $\deg(V) = \deg(V') = 0$, and $V$ and $V'$ are poly-stable vector bundles. On the other hand, it is clear that such Higgs bundles are, in fact, reducible (absolute) minima of $f$. This gives the first case of the proposition. Suppose now that $\Phi \neq 0$. The possible reductions of structure group are the following. \textit{Reduction to $\mathrm{S}\bigl(\bigl(\mathrm{U}(1) \times \mathrm{U}(1)\bigr) \times \bigl(\mathrm{U}(1) \times \mathrm{U}(1)\bigr)\bigr)$}. In this case we have $V = L_1 \oplus L_2$ for linebundles $L_1$ and $L_2$, while $V'= L_1'\oplus L_2'$. Thus $(E,\Phi)$ is the direct sum of two Higgs bundles $(E_1,\Phi_1)$ and $(E_2,\Phi_2)$, where $E_i = L_i \oplus L_i'$, $L_1L_1'L_2L_2'= \mathcal{O}$, and the Higgs field $\Phi_i$ has zeros along the diagonal. Note also that $\deg(E_i) = 0$ by poly-stability of $(E,\Phi)$. Each of the bundles $(E_i,\Phi_i)$ is a minimum on the moduli space of rank $2$ Higgs bundles of this form and hence of the form \eqref{eq:decomp}, in other words all components of $\Phi_i$ are zero, except one off-diagonal entry. (cf.\ Hitchin \cite{hitchin}, Section 10). There are now two cases to consider. The first case is when $\Phi$ is zero on one of the bundles $V$ and $V'$; since $\deg(V) \geq 0$ we must have $\Phi \colon V \to V' \otimes K$. In other words, $\Phi$ is of the form \begin{displaymath} \begin{pmatrix} 0 & 0 \\ c & 0 \end{pmatrix} \end{displaymath} with respect to the decomposition $E = V \oplus V'$. Thus $(E,\Phi)$ is of the form considered in \propref{prop:suminima}. As in the proof of that proposition one sees that there is no subspace of the Zariski tangent space with negative weights and, therefore, these Higgs bundles represent local minima of $f$. This case includes the case of one of the $\Phi_i$ being equal to zero. Note that, if $\deg(L_i) = 0$ then it follows from \eqref{eq:vanishing} and the remark following it that $\Phi_i = 0$. This case gives the remaining local minima of the statement of the proposition. The other case is when $\Phi$ is non-zero on both $V$ and $V'$, say that $\Phi_1 \colon L_1 \to L_1'\otimes K$ and $\Phi_2 \colon L_2' \to L_2\otimes K$. By stability, and since $\Phi \neq 0$, we then have $\deg(L_1') < 0$ and $\deg(L_2) < 0$, and so $\deg(L_1) > 0$ and $\deg(L_2') > 0$. We shall show that in this case $(E,\Phi)$ is not a local minimum of $f$. Let the infinitesimal gauge transformation $\psi$ which produces the decomposition $E_i = L_i \oplus L_i'$ of \eqref{eq:decomp} have weights $m_i$ on $L_i$, and weights $m_i'$ on $L_i'$. We then have the following equations relating these numbers: \begin{align*} m_1' &= m_1 + 1, \\ m_2 &= m_2' + 1, \\ \end{align*} From these equations it follows that $(m_2 - m_1) + (m_1' - m_2') = 2$ and hence, either $m_2 - m_1 \geq 1$ or $m_1'-m_2'\geq 1$. For definiteness suppose that $m_2 - m_1 \geq 1$ (the other case is entirely similar). This means that \begin{displaymath} \Hom(L_1,L_2) \subseteq \Ad_{\mathfrak{k}_{\mathbb{C}}} P_K = \bigl(\End(V) \oplus \End(V') \bigr)_0 \end{displaymath} has weight $\geq 1$ and that this is a subspace of the highest weight space of $\psi$. Note that $\ad(\Phi)$ is zero restricted to the highest weight space and so $H^1(\Sigma;\Hom(L_1,L_2))$ gives a subspace of $\mathbb{H}^1(C^{\bullet}_{-})$ on which $\ddot{f}$ is negative. But since $\deg(L_1) \geq 0$ and $\deg(L_2) < 0$ we have $H^0(\Sigma;\Hom(L_1,L_2)) = 0$ and therefore, from Riemann-Roch, $H^1(\Sigma;\Hom(L_1,L_2)) \neq 0$. It only remains to find a smooth family of Higgs bundles in $\mathcal{M}_{\mathrm{SU}(2,2)}$ to which an element in $H^1(\Sigma;\Hom(L_1,L_2))$ is tangent (cf.\ Section \ref{sec:morseindices}). By hypothesis $(E,\Phi)$ is the direct sum of the stable Higgs bundles $(E_1,\Phi_1)$ and $(E_2,\Phi_2)$. All extensions \begin{displaymath} 0 \to E_2 \to E \to E_1 \to 0 \end{displaymath} are parametrized by $H^1(\Sigma; \Hom(E_1,E_2))$ so, in particular, $t \in H^1(\Sigma; \Hom(L_1,L_2))$ defines an extension \begin{displaymath} 0 \to E_2 \to E_t \to E_1 \to 0 \end{displaymath} which is non-trivial if $t \neq 0$. Note that $E = V_t \oplus V'$, where $V_t$ is the non-trivial extension \begin{displaymath} 0 \to L_2 \to V_t \to L_1 \to 0 \end{displaymath} defined by $t$. We define a Higgs field \begin{math} \Phi = \left( \begin{smallmatrix} 0 & b_t \\ c_t & 0 \end{smallmatrix} \right) \end{math} on $E_t$ of the appropriate form in the following way. To define $b_t \colon V'\to V_t \otimes K$ we use the composition \begin{displaymath} V' \xrightarrow{b} L_2 \otimes K \to V_t \otimes K, \end{displaymath} while to define $c_t \colon V_t\to V' \otimes K$ we use the composition \begin{displaymath} V_t \to L_1 \xrightarrow{c} V' \otimes K. \end{displaymath} For $t \neq 0$ the Higgs bundle $(E_t,\Phi_t)$ is a non-trivial extension of stable Higgs bundles and therefore stable. For $\alpha \in \mathbb{C}$, the family $(E_{\alpha t},\Phi_{\alpha t})$ is thus a smooth family of Higgs bundles in $\mathcal{M}_{\mathrm{SU}(2,2)}$ to which $t \in H^1(\Sigma;\Hom(L_1,L_2))$ is tangent. \textit{Reduction to $\mathrm{S}\bigl(\bigl(\mathrm{U}(1) \times \mathrm{U}(1)\bigr) \times \mathrm{U}(2)\bigr)$}. In this case we have a decomposition of $(E,\Phi)$ as a direct sum of Higgs bundles $(E_1,\Phi_1)$ and $(E_2,\Phi_2$), where $E_1 = V \oplus L_1$ and $E_2 = L_2$ with $L_1$ and $L_2$ linebundles. Again $(E_1,\Phi_1)$ and $(E_2,\Phi_2$) represent local minima on lower rank moduli spaces and so, $\Phi_2 = 0$. If $(E_1,\Phi_1)$ is reducible we are back in one of the previous cases so we may assume that $(E_1,\Phi_1)$ is stable. Since we are at a minimum it must be of the form \eqref{eq:decomp} and again there are several possibilities. If $\Phi_1$ is zero on either $V$ or $L_1$ then it is zero on either $V$ or $V'$ and $(E,\Phi)$ is a minimum as above. Thus the only case that remains is when $V = F_{-1} \oplus F_{1}$ and $\Phi \colon F_{-1} \to L_1 \otimes K$, and $\Phi \colon L_1 \to F_1 \otimes K$. The weights of the infinitesimal gauge transformation producing this decomposition are $-1$, $0$, and $1$ on $F_{-1}$, $L_1$, and $F_1$, respectively. As above one sees that $H^1(\Sigma ; \Hom(F_{-1}, F_1))$ gives a subspace of $\mathbb{H}^1(C^{\bullet}_{-})$ on which $\ddot{f}$ is negative and that $t \in H^1(\Sigma ; \Hom(F_{-1}, F_1))$ is tangent to a smooth family of Higgs bundles, so that $(E,\Phi)$ does not represent a local minimum. We omit the details. \textit{Reduction to $\mathrm{S}\bigl(\mathrm{U}(2) \times \bigl(\mathrm{U}(1) \times \mathrm{U}(1)\bigr)\bigr)$}. This case is analogous to the previous one. \end{proof} In an analogous manner one can prove the following proposition. \begin{prop} \label{prop:spminima2} The reducible Higgs bundles of the form \eqref{eq:sphiggsbundles} with $\deg(V) \geq 0$ which correspond to a local minimum of $f$ on $\mathcal{M}_{\Sp(2,\mathbb{R})}$ either have $\Phi = 0$ and $\deg(V) = 0$, or, if $\Phi \neq 0$, they are direct sums of rank $2$ Higgs bundles $(E_1,\Phi_1)$ and $(E_2,\Phi_2)$, where $E_i = L_i \oplus L_i^{-1}$, $L_i$ a line-bundle with $\deg(L_i) \geq 0$ and $\Phi_i \colon L_i \to L_i^{-1} \otimes K$. If $\Phi_i \neq 0$ then $\deg(L_i) > 0$.\qed \end{prop} \section{Connected Components} \label{sec:count} \subsection{Components of $\mathcal{M}_{\mathrm{SU}(2,2)}$} \label{sec:sucomp} In this section we consider the connected components of $\mathcal{M}_{\mathrm{SU}(2,2)}$. Using \propref{prop:milwood} we can write \begin{displaymath} \mathcal{M}_{\mathrm{SU}(2,2)} = \mathcal{M}_{-(2g-2)} \cup \cdots \cup \mathcal{M}_{2g-2}, \end{displaymath} where $\mathcal{M}_d$, the subspace of Higgs bundles of the form \ref{eq:suhiggsbundles} with $\deg(V) = d$, is a union of connected components. Denote the subspace of local minima of $f$ on $\mathcal{M}_d$ by $\mathcal{N}_d$. Note that, since $f$ is proper, connectedness of $\mathcal{N}_d$ implies connectedness of $\mathcal{M}_d$. As noted in \secref{sec:milnorwood}, we can without loss of generality assume that $d \geq 0$. The results of the previous section then give the following identification of $\mathcal{N}_d$. \begin{prop} \label{prop:allsuminima} The subspace of local minima of $f$ on $\mathcal{M}_d$, $\mathcal{N}_d$, is the space of poly-stable Higgs bundles of the form \ref{eq:suhiggsbundles} with $b = 0$. \end{prop} \begin{proof} Immediate from Propositions \ref{prop:suminima} and \ref{prop:suminima2}. \end{proof} We can use this to identify $\mathcal{N}_d$ with a moduli space of triples, as studied by Bradlow and Garc{\'\i}a-Prada \cite{bradlow-garcia-prada, garcia-prada}. Denote the moduli space of stable triples $(V,\tilde{V'},\phi)$ (where $\phi \in H^0(\Sigma;\Hom(V,\tilde{V'}))$, $\deg(V) = d$, $\deg(\tilde{V'}) = 2g-2-d$, and $\rk(V) = \rk(\tilde{V'}) = 2$) by $\mathcal{M}_{d}^{\mathrm{triples}}$ (cf.\ \secref{sec:q-bundles-triples}). To each such triple we associate a Higgs bundle $ (V \oplus V', \left( \begin{smallmatrix} 0 & 0 \\ \phi & 0 \end{smallmatrix} \right)) $, where $V' = \tilde{V'} \otimes K^{-1}$. The following theorem is then an immediate consequence of \thmref{thm:quiver}. \begin{thm} $\mathcal{N}_d$ is isomorphic to the fibre, over the trivial bundle $\mathcal{O}$, of the map \begin{align*} \mathcal{M}_{d}^{\mathrm{triples}} &\to \Jac(\Sigma) \\ (V,\tilde{V'},\phi) &\mapsto \Lambda^2(V)\otimes \Lambda^2(\tilde{V'}) \otimes K^{-2}. \end{align*} \qed \end{thm} Thus information about connectedness of moduli spaces of stable triples would give information about connectedness of the $\mathcal{M}_d$ and we hope to come back to this on a later occasion. At present we can prove the following theorem. \begin{thm} The subspaces $\mathcal{M}_d$ of $\mathcal{M}_{\mathrm{SU}(2,2)}$ are connected for $d=0$ and $d=\pm (2g-2)$. \end{thm} \begin{proof} First consider the case $d=0$. To see that $\mathcal{N}_0$ is connected, consider the continuous map \begin{align*} N_0 \times N_0 \times \Jac(\Sigma) &\to \mathcal{N}_0 \\ (E,E',L) &\mapsto ((E \otimes L) \oplus (E' \otimes L^{-1}), 0), \end{align*} where $N_0$ denotes the moduli space of rank $2$ poly-stable vector bundles with fixed trivial determinant bundle. From \propref{prop:allsuminima} we see that this is surjective and, since $N_0$ and $\Jac(\Sigma)$ are connected, that $\mathcal{N}_0$ is connected. Next consider the case $d = 2g-2$ (as already noticed, this also takes care of the case $d= -(2g-2)$). From Propositions \ref{prop:iso} and \ref{prop:allsuminima} we see that $\mathcal{N}_{2g-2}$ is isomorphic to the moduli space of rank $2$, degree $2g-2$ vector bundles with fixed determinant, which is known to be connected. \end{proof} \subsection{Components of $\mathcal{M}_{\Sp(4,\mathbb{R})}$} \label{sec:spcomp} In this section we consider the connected components of $\mathcal{M}_{\Sp(4,\mathbb{R})}$. Again using \propref{prop:milwood} we can write \begin{displaymath} \mathcal{M}_{\Sp(4,\mathbb{R})} = \mathcal{M}_{-(2g-2)} \cup \cdots \cup \mathcal{M}_{2g-2}, \end{displaymath} where $\mathcal{M}_d$, the subspace of Higgs bundles of the form \ref{eq:sphiggsbundles} with $\deg(V) = d$, is a union of connected components. Again we denote the subspace of local minima of $f$ on $\mathcal{M}_d$ by $\mathcal{N}_d$, and connectedness of $\mathcal{N}_d$ implies connectedness of $\mathcal{M}_d$. We can also identify $\mathcal{N}_d$ with a moduli space of triples, using \thmref{thm:quiver}, as follows. \begin{thm} For $-(2g-2) \leq d \leq 2g-2$, $\mathcal{N}_d$ is the isomorphic to the fixed point set of the involution on the moduli space $\mathcal{M}_{d}^{\mathrm{triples}}$ of poly-stable triples (as defined in the previous section), defined by \begin{align*} \mathcal{M}_{d}^{\mathrm{triples}} &\to \mathcal{M}_{d}^{\mathrm{triples}} \\ (V,\tilde{V'},\phi) &\mapsto \Lambda^2(V)\otimes \Lambda^2(\tilde{V'}) \otimes K^{-2}. \end{align*} \qed \end{thm} With regard to connectedness, we consider the cases $\abs{d} < 2g-2$ and $\abs{d} = 2g-2$ separately. \paragraph{The case $\abs{d} < 2g-2$.} In this case everything is completely analogous to the case of $\mathrm{SU}(2,2)$-bundles. To begin with, we have the following result. \begin{prop} \label{prop:allspminima} For $\abs{d} < 2g-2$, the subspace of local minima of $f$ on $\mathcal{M}_d$, $\mathcal{N}_d$, is the space of poly-stable Higgs bundles of the form \ref{eq:sphiggsbundles} with $b = 0$. \end{prop} \begin{proof} Follows from Propositions \ref{prop:spminima} and \ref{prop:spminima2}. \end{proof} We have the following result about connectedness of $\mathcal{M}_0$. \begin{thm} The subspace $\mathcal{M}_0$ of $\mathcal{M}_{\Sp(4,\mathbb{R})}$ is connected. \end{thm} \begin{proof} From \propref{prop:allspminima} it follows in particular that $\mathcal{N}_0$ is isomorphic to the moduli space of rank $2$, degree $0$ poly-stable vector bundles. Since this space is connected, the result is proved. \end{proof} \paragraph{The case$\abs{d} = 2g-2$.} In this case the results are entirely different from those of $\mathrm{SU}(2,2)$-bundles, due to \propref{prop:spminima}. Let $(E,\Phi)$ be a Higgs bundle of the form \eqref{eq:sphiggsbundles} with $d = n(g-1)$. Choosing a square root $L_0$ of the canonical bundle on $\Sigma$, we can define a rank $n$ vector bundle $W$ by \begin{displaymath} W = V \otimes L_0^{-1}, \end{displaymath} and we can define $C \in H^0(\Sigma; S^2 W^*)$ and $\phi \in H^0(\Sigma; \End(W) \otimes K^2)$ by \begin{displaymath} C = c \otimes 1_{L_0^{-1}}, \end{displaymath} and \begin{displaymath} \phi = (b \otimes 1_{L_0}) \circ (c \otimes 1_{L_0^{-1}}). \end{displaymath} Note that $\phi$ is symmetric with respect to the quadratic form $C$. From \propref{prop:iso} we know that $c$ is an isomorphism when $(E,\Phi)$ is poly-stable, and thus we can recover $(E,\Phi)$ from this data. Therefore the set of isomorphism classes of Higgs bundles of the form \eqref{eq:sphiggsbundles} is equal to the set of isomorphism classes of Higgs bundles \begin{equation} \label{eq:symhiggs} (W,C,\phi), \end{equation} where $W$ has a non-degenerate quadratic form $C$, and the Higgs field $\Phi$ is twisted by $K^2$ and symmetric with respect to $C$. There is an obvious stability condition for $(W,C,\phi)$, namely that \begin{equation} \label{eq:Wstability} \mu(U) < \mu(W) \end{equation} for all $\phi$-invariant subbundles $U$ of $W$. Next, we shall prove that $(W,C,\phi)$ is stable, if and only if $(E,\Phi)$ is. \begin{thm} \label{thm:extremalcase} The subspace $\mathcal{M}_{n(g-1)} \subset \mathcal{M}_{\Sp(2n,\mathbb{R})}$ of Higgs bundles of the form \eqref{eq:sphiggsbundles}, with $d = n(g-1)$ is isomorphic to the moduli space of poly-stable Higgs bundles of the form \eqref{eq:symhiggs}. \end{thm} \begin{proof} We have to prove that $(E,\Phi)$ is stable if and only if $(W,C,\phi)$ is. From \thmref{thm:quiver} we know that stability of $(E,\Phi)$ is equivalent to stability of the $Q$-bundle $\mathbf{E} = (\underline{E},\underline{\Phi})$. Thus, all we need to prove is that $\mathbf{E}$ is stable if and only if $(W,C,\phi)$ is. Because stability is unaffected by tensoring with a line bundle, we can equally well prove that $(V,b \circ c)$ is stable. Note, that $\mu(V) = g-1$. Assume $\mathbf{E}$ is a stable $Q$-bundle. Let $U \subset V$ be a $\phi$-invariant subbundle. Let $U' \subset V^*$ be the subbundle such that $U' \otimes K$ is generically the image of $U$ under $c$. Then $b$ maps $U'$ to $U$, because of the $\phi$-invariance of $U$. Hence, $\mathbf{F} = (\{U,U'\},\{b,c\})$ defines a $Q$-subbundle of $\mathbf{E}$, and it follows that \begin{equation} \label{eq:stable} \mu(\mathbf{F}) < \mu(\mathbf{E}). \end{equation} But, as $c$ is an isomorphism \begin{align*} \mu(\mathbf{E}) &= \mu(E) \\ &= \mu(V \oplus V \otimes K^{-1}) \\ &= \mu(V) - (g-1), \end{align*} and similarly $\mu(\mathbf{F}) = \mu(U) - (g-1)$. Therefore $\mu(U) < \mu(V)$ and so, $(W,C,\phi)$ is stable. Conversely, assume that $(W,C,\phi)$ is stable. Let $\mathbf{F} = (\{U,U'\},\{b,c\})$ be a $Q$-subbundle of $\mathbf{E}$. Let $\tilde{U} \subset V^*$ be the subbundle which is generically the image of $U' \otimes K$ under $c^{-1}$. Both $U$ and $\tilde{U}$ are $\phi$-invariant subbundles of $V$, because $\mathbf{F}$ is a $Q$-subbundle. Hence, $\mu(U) < \mu(V)$ and $\mu(U') < \mu(V)$, by stability of $(W,C,\phi)$. Recalling that $\mu(V) = g-1$ and $\mu(\tilde{U}) = \mu(U') - (2g-2)$, we get \begin{align} \mu(U) &< g-1, \label{eq:muU} \\ \intertext{and} \mu(U') &< -(g-1). \label{eq:muUprime} \end{align} Note also that \begin{equation}\label{eq:rkUprime} \rk(U') \geq \rk(U), \end{equation} because $c$ is an isomorphism, and the image of $U$ under $c$ is contained in $U' \otimes K$, by the assumption that $\mathbf{F}$ is a $Q$-subbundle. Combining \eqref{eq:muU}, \eqref{eq:muUprime}, and \eqref{eq:rkUprime}, we get: \begin{align*} \mu(\mathbf{F}) &= \mu(U \oplus U') \\ &= \frac{\rk(U)}{\rk(U \oplus U')} \mu(U) + \frac{\rk(U')}{\rk(U \oplus U')} \mu(U') \\ &< \frac{\rk(U) - \rk(U')}{\rk(U \oplus U')} (g-1) \\ &\leq 0. \end{align*} Of course, $\mu(\mathbf{E}) = 0$ and hence the proof is finished. \end{proof} The existence of the quadratic form $C$ on $W$ means that the structure group is $\mathrm{O}(n,\mathbb{C})$. The maximal compact subgroup of $\mathrm{O}(n,\mathbb{C})$ is $\mathrm{O}(n)$ and, therefore, we have the Stiefel-Whitney classes $w_1$ and $w_2$ as topological invariants. We now specialize to the case $n=2$. The first Stiefel-Whitney class can then be seen in holomorphic terms as follows: the quadratic form $C$ gives an isomorphism $(\Lambda^2 W)^2 \cong \mathcal{O}$; hence, $\Lambda^2 W$ gives an element of $H^1(\Sigma; \mathbb{Z} / 2)$, and it is easy to see that this element is $w_1(W)$. It follows that $\Lambda^2 W = \mathcal{O}$ if and only if $w_1(W) = 0$. This, in turn, is equivalent to the existence of a reduction of structure group to $\mathrm{SO}(2,\mathbb{C}) \subset \mathrm{O}(2,\mathbb{C})$. Using the identification $\mathbb{C}^\times \cong \mathrm{SO}(2,\mathbb{C})$ via \begin{displaymath} \lambda \mapsto \begin{pmatrix} \lambda & 0 \\ 0 & \lambda^{-1} \end{pmatrix}, \end{displaymath} we see that this happens exactly when $W$ decomposes as a direct sum \begin{displaymath} W = L \oplus L^{-1}, \end{displaymath} and $C$ is of the form \begin{displaymath} \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix} \end{displaymath} with respect to this decomposition. Now it is clear that, in this case, $w_2(W)$ is given by \begin{displaymath} w_2 = c_1(L) \mod{2}. \end{displaymath} By interchanging $L$ with its dual if necessary, we may assume that $\deg(L) \geq 0$. Furthermore, when $\deg(L) > 0$, the Higgs field $\phi$ must induce a non-zero holomorphic map \begin{displaymath} L \to L^{-1} K^2, \end{displaymath} because otherwise $L \subset W$ would violate stability. Hence, we have \begin{displaymath} \deg(L) \leq 2g-2. \end{displaymath} We, therefore, have a decomposition of $\mathcal{M}_{2g-2}$ into subspaces, each of which is a union of connected components, as follows: \begin{displaymath} \mathcal{M}_{2g-2} = \Bigl( \bigcup_{u,v} \mathbf{M}^v_u \Bigr)\cup \Bigl( \bigcup_{l=0}^{2g-2} \mathbf{M}_0^l \Bigr), \end{displaymath} where $\mathbf{M}^v_u$ is the moduli space of poly-stable Higgs bundles $(W,C,\phi)$ with $w_1(W) = u \in H^1(\Sigma; \mathbb{Z}/2)-\{0\}$ and $ w_2(W) = v \in H^2(\Sigma; \mathbb{Z}/2)$, and where $\mathbf{M}_0^l$ is the moduli space of poly-stable Higgs bundles $(W,C,\phi)$ with $w_1(W) = 0$ and $\deg(L) = l$. We shall prove that each of these subspaces is connected, except $\mathbf{M}_0^{2g-2}$: in this case the Higgs field $\phi$ induces a non-zero section of the degree $0$ linebundle $L^{-2}K^2$ and thus $L^2=K^2$. We therefore have a further decomposition of of $\mathbf{M}_0^{2g-2}$ into subspaces $\mathbf{M}_{0,L}^{2g-2}$, indexed by the $2^{2g}$ square roots $L \in \Jac^{2g-2}(\Sigma)$ of $K^2$ (note the analogy with the breaking up of $\mathcal{M}_{2g-2}$ into several connected components). We can now state our main result, to be proved in the remaining part of this section. \begin{thm} \label{thm:components} \begin{itemize} \item[$i)$] The spaces $\mathbf{M}_{0,L}^{2g-2}$ are connected. \item[$ii)$] The spaces $\mathbf{M}^l_0$ are connected for $0 \leq l < 2g-2$. \item[$iii)$] The spaces $\mathbf{M}^v_u$ are connected. \end{itemize} \end{thm} \begin{rem} Hitchin showed in \cite{hitchinlie} that for any split real form $G_r$ of a complex simple Lie group the moduli space of reductive representations of $\pi_1(\Sigma)$ in $G_r$ contains a connected component which is homeomorphic to an Euclidean space of dimension $(2g-2)\dim G_r$. This component is called the Teichm{\"u}ller component. The group $\Sp(4,\mathbb{R})$ is a split real form of $\Sp(4,\mathbb{C})$ so there is a Teichm{\"u}ller component in this case. As a matter of fact, each of the subspaces $\mathbf{M}_{0,L}^{2g-2}$ is isomorphic to a vector space: note that $W = L \oplus L^{-1}$ is completely determined by $L$ and that any $(W,C,\phi)$ is stable. Hence $\mathbf{M}_{0,L}^{2g-2}$ is isomorphic to the space of Higgs fields $\phi \in H^0(\Sigma; \End(W) \otimes K^2)$ which are symmetric with respect to $C = \left( \begin{smallmatrix} 0 & 1 \\ 1 & 0 \end{smallmatrix} \right)$, that is, of the form $\phi = \left( \begin{smallmatrix} \phi_{11} & \phi_{12} \\ \phi_{12} & \phi_{22} \\ \end{smallmatrix} \right)$. It follows that $\mathbf{M}_{0,L}^{2g-2}$ is isomorphic to the vector space \begin{displaymath} H^0(\Sigma;K^2) \oplus H^0(\Sigma;K^2) \oplus H^0(\Sigma;K^4). \end{displaymath} Note that this proves $i)$ of the theorem. \end{rem} \begin{rem} One can see (see \cite{thesis} for details), that the subspaces $\mathcal{M}_d$, $\mathbf{M}^v_u$, $\mathbf{M}^l_0$, and $\mathbf{M}^{2g-2}_{0,L}$ are non-empty. Therefore, \thmref{thm:components} shows that $\mathcal{M}_{\Sp(4,\mathbb{R})}$ has at least $3\cdot 2^{2g} + 8g-13$ connected components. \end{rem} \subparagraph{Proof that the subspaces $\mathbf{M}^l_0 \subset \mathcal{M}_{2g-2}$ are connected.} Recall that any $(W,C,\phi)$ in $\mathbf{M}^l_0$ is of the form \begin{displaymath} W = L \oplus L^{-1}, \end{displaymath} with $l = \deg(L)$ and $C$ of the form $ \left( \begin{smallmatrix} 0 & 1 \\ 1 & 0 \end{smallmatrix} \right) $. First, we consider the case of $l > 0$. In this case, the Higgs field $\phi$ must be non-zero, as otherwise the subbundle $L \subset W$ would violate stability. But any critical point of the type described in $i)$ of \propref{prop:spminima} has $\phi = 0$ so, it follows that all the critical points in $\mathbf{M}^l_0$ for $l > 0$ are of the type described in $ii)$ of \propref{prop:spminima}. We therefore see that the critical points correspond to Higgs bundles $(W,C,\phi)$, which are of the form described above and where, furthermore, $\phi$ is of the form \begin{displaymath} \phi = \begin{pmatrix} 0 & 0 \\ \tilde{\phi} & 0 \end{pmatrix}, \end{displaymath} with $\tilde{\phi} \in H^0(\Sigma; L^{-2}K^2)$. Using this, it is now easy to give an explicit description of the subspace of local minima of $f$ on $\mathbf{M}^l_0$. \begin{prop} The subspace of local minima $N^l_0 \subset \mathbf{M}^l_0$ fits into a pull-back diagram \begin{displaymath} \begin{CD} N^l_0 @>>> \Jac^l(\Sigma) \\ @VV{\pi}V @VV{L \mapsto L^{-2}K^2}V \\ S^{4g-4-2l}\Sigma @>{D \mapsto [D]}>> \Jac^{4g-4-2l}(\Sigma), \end{CD} \end{displaymath} where $\pi(W,C,\phi) = (\phi)$. \end{prop} \begin{proof} The only thing there is to remark is that any $(W,C,\phi)$, of the form given above, is stable. But, $L^{-1} \subset W$ is the only $\phi$-invariant subbundle so, this is obvious. \end{proof} From this proposition, it is clear that $N^l_0$ is connected so, from the properness of $f$, it follows that $\mathbf{M}^l_0$ is connected for $l > 0$. In the case $l = 0$, we have the following result. \begin{prop} Any local minimum of $f$ on $\mathbf{M}^0_0$ has $\phi = 0$ and is, therefore, of the type described in $i)$ of \propref{prop:spminima}. \end{prop} \begin{proof} Suppose we have a critical point of the type described in $ii)$ of \propref{prop:spminima}, with $\phi \neq 0$. Then, $L^{-1} \subset W$ is $\phi$-invariant and therefore, $(W,C,\phi)$ is semi-stable, but not stable. Since we are considering the moduli space of poly-stable Higgs bundles, $(W,C,\phi)$ decomposes as a direct sum of rank $1$ Higgs bundles of degree $0$. The only subbundles of $W$ of rank $1$ and degree $0$ are $L$ and $L^{-1}$, and $L$ is not $\phi$-invariant so, we conclude that this situation cannot occur. \end{proof} Consequently, we have the following description of the subspace of local minima of $f$ on $\mathbf{M}^0_0$. \begin{prop} The subspace $N^0_0 \subset \mathbf{M}^0_0$ of local minima of $f$ is isomorphic to the moduli space of poly-stable $(W,C)$, where $W$ is of the form \begin{displaymath} W = L \oplus L^{-1}, \end{displaymath} for a linebundle $L$ of degree $0$, and $C$ is of the form $ \left( \begin{smallmatrix} 0 & 1 \\ 1 & 0 \end{smallmatrix} \right) $, with respect to this decomposition. \end{prop} Note that the pair $(W,C)$ decomposes into a direct sum of Higgs linebundles exactly when $L^2 = \mathcal{O}$, and it is then poly-stable, but not stable. All other $(W,C)$ are stable. It follows that there is a surjective continuous map \begin{displaymath} \Jac^0(\Sigma) \to N^0_0, \end{displaymath} given by taking $L$ to $(W,C)$ of the form given above. Therefore, $N^0_0$ is connected, finishing the proof that the subspaces $\mathbf{M}^l_0$ are connected. \subparagraph{Proof that the subspaces $\mathbf{M}^v_u \subset \mathcal{M}_{2g-2}$ are connected.} We begin by noting that the $(W,C,\phi)$ corresponding to a critical point of the type described in $ii)$ of \propref{prop:spminima} has $w_1(W) = 0$, thus we see that the subspaces of local minima $N^v_u \subset \mathbf{M}^v_u$ consist of critical points of the type described in $i)$ of \propref{prop:spminima}. Recall that for these $b = 0$; in terms of the Higgs bundle $(W,C,\phi)$, this means that $\phi = 0$. Thus, $N^v_u$ is the moduli space of stable pairs $(W,C)$ with the given characteristic classes. From $\Lambda^2 W \neq \mathcal{O}$, one sees easily that any such pair is stable. There is a connected double cover $\tilde{\Sigma} \overset{\pi}{\to} \Sigma$ given by \begin{displaymath} w_1(W) \in H^1(\Sigma; \mathbb{Z}/2) = \Hom(\pi_1 \Sigma,\mathbb{Z}/2). \end{displaymath} Clearly, the pull-back of $W$ to $\tilde{\Sigma}$ is of the form $\pi^*W = M \oplus M^{-1}$ with $ \pi^*C = \left( \begin{smallmatrix} 0 & 1 \\ 1 & 0 \end{smallmatrix} \right) $ and $M$ a linebundle. Let $\tau \colon \tilde{\Sigma} \to \tilde{\Sigma}$ be the involution interchanging the sheets of the covering, then, clearly, \begin{displaymath} \tau^*M = M^{-1}. \end{displaymath} Conversely, if $M$ is a linebundle on $\tilde{\Sigma}$ which satisfies this condition, then $W = \pi_*M$ is a rank $2$ vector bundle with a non-degenerate quadratic form $C$. In fact $\tilde{\Sigma}$ is the spectral curve associated to $(W,C)$ (see Hitchin \cite{hitchinduke} and Beauville, Narasimhan and Ramanan \cite{beauville}). Hence $N^0_u \cup N^1_u$ can be identified with the kernel of the map \begin{displaymath} 1 + \tau^* \colon \Jac(\tilde{\Sigma}) \to \Jac(\tilde{\Sigma}), \end{displaymath} where $\tilde{\Sigma}$ is the unramified double cover of $\Sigma$ given by $u \in H^1(\Sigma; \mathbb{Z}/2)$. It remains to distinguish between $w_2$ being equal to 0 or 1. When the cover is unramified, the kernel of $1 + \tau^*$ splits into two components, \begin{displaymath} \ker(1 + \tau^*) = P^+ \cup P^-, \end{displaymath} each of them a translate of the Prym variety of the covering. It is a classical theorem of Wirtinger, that the function $\delta \colon P^+ \cup P^- \to \mathbb{Z}/2$, defined by \begin{align*} \delta(M) &= \dim_{\mathbb{C}} H^0(\tilde{\Sigma}; M \otimes \pi^* L_0) \mod 2 \\ &= \dim_{\mathbb{C}} H^0(\tilde{\Sigma}; \pi_*M \otimes L_0) \mod 2, \end{align*} is constant on each of $P^+$ and $P^-$ and takes different values on them. For proofs of these facts, see Mumford \cite{mumford:theta} or \cite{mumford:prym}. Now, let $F \to \Sigma$ be a real vector bundle. Choosing a metric on $F$, the complexification $F^c = F \otimes_{\mathbb{R}} \mathbb{C}$ acquires a holomorphic structure and therefore, there is a $\bar{\partial}$-operator \begin{displaymath} \bar{\partial}_{L_0}(F) \colon \Omega^0(\Sigma; L_0 \otimes F^c) \to \Omega^{0,1}(\Sigma; L_0 \otimes F^c). \end{displaymath} Atiyah \cite{atiyah:spin} shows that the function \begin{displaymath} \delta_{L_0}(F) = \dim_{\mathbb{C}} \ker (\bar{\partial}_{L_0}(F)) \mod 2 \end{displaymath} is independent of the choice of the metric, and that it extends to give a group homomorphism \begin{displaymath} \delta_{L_0} \colon KO(\Sigma) \to \mathbb{Z}/2. \end{displaymath} Define $\gamma \in \widetilde{KO} (\Sigma)$ to be the pull-back of the generator of $\widetilde{KO} (S^2)$ under a map $\tilde{\Sigma} \to S^2$ of degree $1$. Atiyah \cite[Lemma (2.3)]{atiyah:spin} shows that \begin{displaymath} \delta_{L_0}(\gamma) = 1. \end{displaymath} Furthermore, the total Stiefel-Whitney class gives an isomorphism \begin{displaymath} w \colon \widetilde{KO}(\Sigma) \to \{1\} \oplus H^1(\Sigma; \mathbb{Z}/2) \oplus H^2(\Sigma; \mathbb{Z}/2) \end{displaymath} of the additive group $\widetilde{KO}(\Sigma)$ onto the multiplicative group of the cohomology ring $H^*(\Sigma; \mathbb{Z}/2)$ (see \cite[Remark, p.\ 54]{atiyah:spin}). Clearly, \begin{displaymath} w(\gamma) = (1,0,1), \end{displaymath} where we identify $H^2(\Sigma; \mathbb{Z}/2) = \mathbb{Z}/2$. We may, therefore, think of $\delta_{L_0}$ as a homomorphism of the multiplicative group of $H^*(\Sigma; \mathbb{Z}/2)$ to $\mathbb{Z}/2$, which takes the value $1$ on the element $(1,0,1)$. Let $u \in H^1(\Sigma; \mathbb{Z}/2)$; then, \begin{displaymath} (1,u,0) = (1,u,1) \cdot (1,0,1) \end{displaymath} in $H^*(\Sigma; \mathbb{Z}/2)$. Therefore, \begin{equation} \label{eq:w2} \delta_{L_0}(1,u,0) = \delta_{L_0}(1,u,1) + 1. \end{equation} Returning to $(W,C)$ with $W = \pi_* M$ for $M \in \ker(1 + \tau^*)$, we see that \begin{displaymath} \delta(M) = \delta_{L_0}(W^r), \end{displaymath} where $W^r$ is a real rank two bundle, whose complexification is $W$. It follows from \eqref{eq:w2}, that $\delta$ takes different values for different values of $w_2(W)$ and hence, that $w_2(W)$ determines whether $M$ lies in $P^+$ or $P^-$. From this discussion, we obtain the following explicit description of the subvariety $N^v_u \subset \mathbf{M}^v_u$ of local minima of $f$. \begin{prop} Let $u \in H^1(\Sigma; \mathbb{Z}/2) - \{0\}$, let $v \in H^2(\Sigma; \mathbb{Z}/2) = \mathbb{Z}/2$ and let $P^+$ and $P^-$ be the Abelian varieties associated to the double cover of $\Sigma$, given by $u$ as above. Then, the subvariety $N^v_u \subset \mathbf{M}^v_u$ of local minima of $f$ is equal to $P^+$ and $P^-$, respectively, for the two values of $v$. \qed \end{prop} Consequently, $N^v_u$ is connected and, from the properness of $f$, it follows that $\mathbf{M}^v_u$ is connected, finally finishing the proof of \thmref{thm:components}. \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
\section{Introduction} Stimulated in part by the upcoming launches of the next generation of astronomical instruments, the epoch of reionization has become the subject of increasingly intense theoretical effort in recent years. Reionization affects observations of the Cosmic Microwave Background, both as a source of new anisotropies, and in suppressing small-scale fluctuations from the surface of last scatter (Haiman \& Knox, 1999, and references therein). In addition, the Next Generation Space Telescope (NGST) offers the prospect of directly detecting the first luminous objects. Nonetheless, the large number of complex, non-linear physical processes involved means that it is extremely difficult to make firm theoretical predictions with any kind of confidence. While it appears quite likely that NGST will be able to image Population III objects in the infrared continuum (Haiman \& Loeb 1997, 1998), and detection of the Gunn-Peterson trough may yield the redshift at which the IGM is not yet fully reionized (Miralda-Escude \& Rees 1998), this information alone falls far short of being able to fully constrain models of reionization. Key among the uncertainties is the ionizing emissivity of collapsed objects, and the degree of gas clumping. In this paper, we suggest the observation of diffuse gas and Population III objects in thermal bremstrahhlung and recombination line emission as a direct probe of these quantities. Due to their $n_{e}^{2}$ dependence, free-free and recombination line fluctuations trace the clumpiness and ionization fraction of the gas. While the intensities of these signals are directly proportional to one another, their joint detection should prove invaluable, given the highly different nature of the foreground contaminants and instrumental systematics in each case. In addition, we examine the prospects for detecting ionized gas at high redshift via the kinetic Sunyaev-Zeldovich effect. Below, we briefly summarize the observational case for each signal. {\bf Balmer line emission} Discussion in the literature on detecting line emission from high redshift galaxies has thus far focussed predominantly on Ly$\alpha$ (Miralda-Escude \& Rees 1998, Kunth et al 1998). However, detection in Ly$\alpha$ faces problems due to attenuation by dust and resonant scattering by neutral hydrogen. Indeed, in the dark ages before the universe was completely reionized, it is unlikely that we will be able to detect sources in Ly$\alpha$ emission, due to the large column density of neutral hydrogen along the line of sight. By contrast, Balmer lines face none of these problems, and provide a much cleaner probe of ionizing emissivities. Thus, for instance, the intensity of H$\alpha$ emission is intimately linked to star formation rates. In this paper, we show that for equivalent integration times with NGST, moderate resolution spectroscopy (${\rm R \sim 1000}$) should yield H$\alpha$ line detection with the same signal to noise as continuum IR imaging for $z<6.6$, and 10\% signal to noise compared to IR imaging for $z>6.6$. Thus, if a high-redshift galaxy is detected in imaging mode, H$\alpha$ line detection should be eminently feasible. In addition, higher order Balmer lines such as H$\beta$ should also be detectable. The redshift information provided by Balmer line emission should provide an invaluable complement to free-free and IR imaging. {\bf Free-free emission} In addition to Compton scattering off hot gas via the Sunyaev-Zeldovich effect, free-free emission from the ionized IGM can also produce a spectral distortion of the CMB (Barlett \& Stebbins 1991). This distortion increases quadratically at low frequencies, and is poorly constrained by COBE observations near the peak of the CMB spectrum ($\nu \sim 56$ GHz). One of the drivers behind the development of the DIMES satellite (Kogut et al 1996; see http://ceylon.gsfc.nasa.gov/DIMES) is the hope of detecting this distortion produced by the ionized IGM at low frequencies ($\nu \sim 2$ GHz). The magnitude of the distortion is directly related to the optical depth of ionized gas in the IGM, and thus has the potential to constrain the epoch of reionization. However, the ionized IGM is not the only source of free-free emission. Lyman limit systems in ionization equilibrium with the external intergalactic ionizing background emit thermal bremsstrahlung radiation (Loeb 1996), as do ionized halos with an internal source of ionizing radiation (Haiman \& Loeb 1997). In this paper, we argue that if the escape fraction of ionizing photons from starburst galaxies is small (as is indicated by local observations), then the integrated contribution of ionized halos to the free-free background overwhelms that of the IGM, regardless of the details of the reionization history of the universe. Thus, the detection of a spectral distortion by DIMES is likely to probe the integrated ionizing emissivity and clumping of gas in halos, rather than the optical depth of the reionized IGM. A natural way to tell if an observed spectral distortion is produced by ionised halos rather than an ionised IGM is to observe fluctuations in the free-free background. If the spectral distortion of the CMB is produced by the IGM, it should exhibit a fairly smooth spatial signature across the sky, whereas if ionized halos predominate then the free-free background should exhibit a great deal of small scale power. Since Galactic foregrounds in free-free and synchrotron emission damp rapidly at small scales, it should be possible to directly image the predicted population of high redshift ionized halos with a high resolution instrument. This is an application well suited for the proposed Square Kilometer Array, or SKA (Braun et al 1998) which will operate in the frequency range $0.03-20$ GHz with an angular resolution comparable to that of HST ($\sim 0.1''$) and a sensitivity 100 times better than the VLA, down to the $\sim 20$ nJy level. Such a quantum leap in performance should reap immense scientific rewards. At present, the appearance of the radio sky at nano Jansky levels is not known. We make a first stab at characterising observable properties of a predicted population of free-free emitters. In our Press-Schechter based model, we find that $\sim 10^{4}$ sources at $z>5$ should be directly detectable as discrete sources in the $1^{\circ}$ SKA field of view. In addition, the integrated background from sources too faint to be directly identified creates a fluctuating signal, similar to the Extragalactic Background Light (EBL) in the optical and the Cosmic Infrared Background (CIB) in the IR, which may be characterised by its power spectrum and skewness. Finally, we investigate the clustering of sources at high redshift and find that while it is fairly substantial, the observed signal is still likely to be dominated by Poisson fluctuations. In all numerical estimates, we assume a background cosmology given by the 'concordance' values of Ostriker \& Steinhardt (1995): $(\Omega_{m},\Omega_{\Lambda},\Omega_{b},h,\sigma_{8 h^{-1}},n)=(0.35,0.65,0.04,0.65,0.87,0.96)$. We use expressions from Carroll, Press \& Turner (1992) and Eisenstein (1997) to compute cosmological expressions such as the growth factor and luminosity distance. \section{Modelling the Ionizing Sources} \subsection{Estimating Halo Luminosities} Since the free-free and H$\alpha$ emissivity are both proportional to $\langle n_{e}^{2} \rangle$, observational prospects hinge crucially on this quantity at high redshift. The signal is dominated by the small-scale clumping of the gas at high redshift. One approach to estimating $\langle n_{e}^{2} \rangle$ is to model the properties of halos at high redshift. It is not known to what extent the baryons can cool and condense in halos at high redshift. A minimal assumption would be that gas within a halo traces the distribution of the dark matter, generally overdense by $\delta \sim 180$ at collapse; this would indicate $\langle n_{e}^{2} \rangle \propto (1+z)^{6}$. Observations of the local universe indicate that the maximum luminosity density of dynamically hot stellar systems varies as $M^{-4/3}$(Burstein et al 1997), which implies that the gas is more centrally condensed in low mass systems (which are prevalent at high redshift), presumably because dissipational effects are more efficient. Thus, the above scaling with redshift is likely an underestimate. If the gas settles to form a disk it could be overdense by $\delta \sim 18 \pi^{2} \lambda^{-3} \sim 10^{6}$ (Dalcanton, Spergel \& Summers 1997), where the spin parameter $\lambda \sim 0.05$ typically (Barnes \& Efstathiou 1987). Note, however that $\langle n_{e}^{2} \rangle$ is likely to be dominated by the local clumping of the ISM in ionization bounded HII regions (in other words, the clumping factor $\langle n_{e}^{2} \rangle/ \langle n_{e} \rangle^{2}$ of the gas is significantly greater than unity), in which case these limits would have to be revised upwards. A simpler and likely more accurate approach is to assume that most of the ionizing radiation produced is absorbed locally in the halo, with only a small fraction escaping ($< 20\%$). This is consistent with our observations of the nearby universe. For instance, the observation of 4 starburst galaxies with the Hopkins Ultraviolet Telescope (HUT) by Leitherer et al (1995) report an escape fraction of only $3\%$, by comparing the observed Lyman continuum flux with that predicted from theoretical spectral energy distributions. In our own galaxy, the models of Dove, Shull \& Ferrara (1999) have an escape fraction for ionizing photons of $6\%$ and $3\%$ (for coeval and Gaussian star formation histories respectively) from OB associations in the Milky Way disk, while Bland-Hawthorn \& Maloney (1999) find an escape fraction of $6\%$ is necessary for consistency with the observed line emission from the Magellanic stream and high velocity clouds. The increasing density of the gas in halos with collapse redshift indicates that the escape fraction should decrease, strengthening our bound. The production rate of recombination line photons $\dot{N}_{\rm recomb}$ is directly related to the production rate of ionizing photons: \begin{equation} \dot{N}_{\rm recomb}= \alpha_{B} \langle n_{e}^{2} \rangle V \approx (1-f_{esc})\dot{N}_{\rm ion} \label{nesq} \end{equation} where $\dot{N}_{\rm ion}$ is the number of ionizing photons emitted per second, and $\alpha_{B}$ is the case B recombination coefficient. In this manner the luminosities of a source in H$\alpha$ and free-free emission scale directly with the production rate of ionizing photons. For H$\alpha$, Hummer \& Storey (1987) find that 0.45 H$\alpha$ photons are emitted per Lyman continuum photon over a wide range of nebulosity conditions. We thus obtain directly \begin{equation} {\rm L(H \alpha)} = 1.4 \times 10^{41} \left(\frac{\dot{N}_{\rm ion}}{10^{53} {\rm photons \, s^{-1}}} \right) {\rm ergs \, s^{-1}}. \label{L_Halpha} \end{equation} For free-free emission, the volume emissivity for a plasma is (Rybicki \& Lightman 1979): \begin{equation} \epsilon_{\nu} = 3.2 \times 10^{-39} n_{e}^{2} \left( \frac{T}{10^{4} K} \right)^{-0.35} \ {\rm erg \,s^{-1} \,cm^{-3} \, Hz^{-1}} \label{free_emissivity} \end{equation} where we adopt the velocity averaged Gaunt factor $\bar{g}_{ff}=4.7$ and approximate its mild temperature dependence with a power law. We can then use the estimate of $n_{e}^{2}V$ derived from equation (\ref{nesq}) to obtain: \begin{equation} L_{\nu}^{ff}=1.2 \times 10^{27} \left(\frac{\dot{N}_{\rm ion}}{10^{53} {\rm photons \, s^{-1}}} \right) {\rm ergs \, s^{-1} \, Hz^{-1}} \label{L_free} \end{equation} We thus wish to estimate $\dot{N}_{\rm ion}(M)$, the production rate of ionizing photons as a function of halo mass. We can either assume that this function is independent of collapse redshift, or that it evolves with redshift. In this paper, we adopt the models of Haiman \& Loeb (1998), which assume the former. In particular, we adopt their starburst model, which is normalised to the observed metallicity $ 10^{-3} Z_{\odot} \le Z \le 10^{-2} Z_{\odot}$ of the IGM at $z \sim 3$. In this model, a constant fraction of the gas mass turns into stars, $1.7 \% \le f_{star} \le 17\%$ in a starburst which fades after $\sim 10^{7}$ years. In this paper, we shall assume the upper value $ Z \sim 10^{-2} Z_{\odot}$, corresponding to $f_{star} \sim 17\%$, in our discussion and plots. We show the results of the calculations for the lower limit star formation efficiency $f_{star} \sim 1.7 \%$ in figure (\ref{lower_SF_efficiency}) (discussed in section (\ref{lower_SF_section})). During the starburst period, the rate of production of ionizing photons is: \begin{equation} \dot{N}_{\rm ion}(M)=2 \times 10^{53} \, (f_{star}/0.17) \, (M/10^{9} M_{\odot}) \, {\rm photons} \, {\rm s^{-1}} \label{Ndot_scaling} \end{equation} Their mini-quasar model yields similar estimates, $\dot{N}_{\rm ion}(M)=4 \times 10^{53} (M/10^{9} M_{\odot})$photons ${\rm s^{-1}}$, with the quasar fading after $\sim 10^{6}$ yrs. The quasar model is normalised to the observed quasar luminosity function at low redshift, assuming a constant ratio of black hole mass to halo mass. We consider the adopted model to be a reasonably conservative estimate for high redshift sources, as it seems quite possible that the efficiency of ionizing photon production increases with redshift. The star formation rate is tied to the dynamical timescale of the gas, SFR$\propto M_{g}/t_{dyn}$. Since high redshift sources are denser, the free-fall time is shorter and $\dot{N}_{ion} \propto {\rm SFR} \propto (1+z)^{1.5}$. Alternatively, one can use the empirical Schmidt law SFR$\propto \Sigma_{g}^{1.4}$ (Kennicutt 1998), where $\Sigma_{g}$ is the gas surface density, in conjunction with an isothermal sphere model to deduce $\dot{N}_{ion} \propto {\rm SFR} \propto (1+z)^{0.8}$ (all scalings are for an object of fixed collapse mass). Intriguingly, Weedman et al (1998) find from observations of galaxies in the HDF that high redshift starbursts have intrinsic UV surface brightnesses typically four times brighter than low redshift starbursts. Furthermore, the absence of metals to provide cooling in the early universe would tend to lead to a top-heavy IMF (Larson 1998), further increasing the efficiency of ionizing photon production. Note also that in our adopted model the escape fraction should fall with redshift, since $\dot{N}_{\rm ion}$ is independent of redshift, whereas $\dot{N}_{\rm recomb} \propto \langle n_{e}^{2} \rangle V \propto (1+z)^{3}$ (using $r_{vir} \propto (1+z)^{-1}$). We pause to note the possible caveat that supernovae could drive gas out of the galaxy (Dekel \& Silk 1986), thereby drastically increasing the escape fraction of ionizing photons and vitiating our assumptions. The expulsion of gas would also shut off further star formation. Although there is enough energy from supernovae to drive such a wind, recent numerical simulations (Mac Low \& Ferrara 1999) indicate that mass ejection is in fact only efficient in galaxies with $M_{g} < 10^{6} M_{\odot}$ (such galaxies are excluded from our analysis), where the small depth of the potential well means that gas is extremely easily unbound. In higher mass galaxies, the supernovae produce a hole in the interstellar medium through which the shock heated gas is expelled, while the cold gas remains bound. We thus ignore the effect of gas expulsion by supernovae. Note also that our emissivities are calibrated to a quiescent, steady-state mode of star formation, as is observed in the disk of many spiral galaxies today. There also exists the possibility of massive starbursts, with star formation rates of up to 100 $M_{\odot} \, yr^{-1}$, in which a large fraction of the gas mass is converted into stars in the short dynamical time of the dense galactic center. This is particularly likely to happen as halos merge, and would produce significantly brighter objects, easily detectable with the proposed observations. When computing surface brightnesses, it is necessary to assume a characteristic size for an ionized halo $r$. We can assume that the gas profile extends out to the virial radius ($r \sim r_{vir}$) or that the gas is confined to a disk with spin parameter $\lambda$ ($r \sim \lambda r_{vir}$), where the virial radius $r_{vir}= v_{c} /(10 \surd{2} H(z))= 2.1 {\rm Kpc} (M/10^{9} M_{\odot})^{1/3} (1+z/10)^{-1}$, for an isothermal sphere. In subsequent calculations we shall assume $r \sim \lambda r_{vir}$. The small corresponding angular size means that in general we do not resolve the galaxy, which may thus be treated as a point source. We quote angular size as $\theta \equiv {\rm max(\theta_{telescope}},r/d_{A}(z))$, where $\theta_{\rm telescope}$ is the angular resolution of the telescope. Note that this assumption only affects our surface brightness calculations, since $L_{\rm H\alpha, \, free-free} \propto \dot{N}_{\rm ion}$, independent of $r$. \subsection{Statistics of Halos} \label{halo_stat} We employ a Press-Schechter based model of the abundance of ionizing halos similar to that adopted by Haiman \& Loeb (1998). In our simplified physical model, halos collapse, form a starburst lasting $t_{o}=10^{7}$ years, and then recombine and no longer contribute to the free-free background (FFB). We ignore sources ionized by an external background radiation field, since the limiting column density before these sources become self shielding corresponds to a undetectably small flux at high redshift. We adopt a different expression for the halo collapse from Haiman \& Loeb (1998); their expression slightly underestimates the halo formation rate as it includes a negative contribution from merging halos. The more accurate expression we use is given by Sasaki (1994): \begin{equation} \frac{d\dot{N}^{form}}{dM}(M,z)=\frac{1}{D}\frac{dD}{dt}\frac{d n_{PS}}{dM}(M,z)\frac{\delta_{c}^{2}}{\sigma^{2}(M)D^{2}} \end{equation} where $D(z)$ is the growth factor, and $\delta_{c}=1.7$ is the threshold above which mass fluctuations collapse (note that at high redshift, the difference between this expression and the one used by Haiman \& Loeb (1998) is negligible, since mergers are unimportant). Here, $\frac{d\dot{N}^{form}}{dM}$ is the collapse rate of halos per mass interval per unit comoving volume, while $\frac{d n_{PS}}{dM}$ is the standard Press-Schechter number density of collapsed halos per mass interval (Press \& Schechter 1974, Bond et al 1991). Then, given an expression for the expected flux from a halo of mass M at redshift z, $S=S(M,z)$, we can calculate the expected comoving number density of ionized halos in a given flux interval as a function of redshift: \begin{equation} \frac{dN_{\rm halo}}{dS dV}(S,z)= \int_{t(z)}^{t(z)-t_{o}} dt \frac{d \dot{N}^{form}}{dM} \frac{dM}{dS} \end{equation} where $t_{o}$ is the duration of a starburst. Note, however, that star formation or quasar activity is only able to proceed if gas is able to cool and condense in the halo. Thus, we set a cut-off mass for a halo to be ionized of $M_{*}=10^{8} (1+z/10)^{-3/2} M_{\odot}$, which is the critical mass needed to attain a virial temperature of $10^{4}$ K to excite atomic hydrogen cooling. With this in mind, we can compute moments of the intensity distribution due to sources above a redshift $z_{\rm min}$ via: \begin{equation} \langle S^{n}(>z_{min},S_{c}) \rangle= \int_{z_{min}}^{\infty} dz \int_{S_{\rm min}(z)}^{S_{max}} dS \frac{dN_{\rm halo}}{dSdV} \frac{dV}{dz d\Omega} S^{n} \label{moments} \end{equation} where $S_{c}$ is a flux limit above which discrete source identification and removal is possible. We shall use this expression entensively in ensuing sections. The zeroth moment of equation (\ref{moments}) can be used to compute the number counts of sources above the flux limit $S_{c}$. In this case, $S_{max} \rightarrow \infty$ and $S_{min}(z)={\rm max}(S_{c},S_{*}(z))$, where we denote $S_{*}(z)$ as the flux from a halo of minimum mass $M_{*}$ at redshift z. One can also use equation (\ref{moments}) to compute the moments of the residual background signal due to sources below the flux limit $S_{c}$, when all discrete sources above the flux limit have been identified and removed. In this case, we set $S_{max}=S_{c}$ and $S_{min}(z)=S_{*}(z)$. \subsection{Comparing Halo Emissivity with the IGM} Throughout this paper we only compute the contribution of ionized halos to free-free and recombination line emission, ignoring the emission from filaments and the IGM. While it is obvious that small scale fluctuations will be dominated by ionized halos, it is not immediately obvious that the mean sky-averaged signal (as is probed by spectral distortion measurements of the CMB) from ionized halos is greater than that from other sources. This in fact is a direct consequence of our assumption that the escape fraction of ionising photons from halos is small. This implies that the degradation of ionizing photons into recombination line photons take place largely in halos. The fraction of ionizing photons produced which are degraded in halos as opposed to the IGM (including ionized filaments and Lyman limit systems) scales as $(1-\langle f_{esc}\rangle)/\langle f_{esc} \rangle $, where $f_{esc}$ is averaged over all halos. This ratio is large due to the small assumed $f_{\rm esc}$, which (as we've argued above) also is likely to decrease with redshift. We can compare the volume averaged emissivities of the IGM and ionized halos more quantitatively. The comoving free-free emissivity of the ionized halos is given by: \begin{equation} \epsilon_{\nu}^{halos}(z)=\int_{M_{*}}^{\infty} dM \frac{dN}{dMdV}(M,z) L_{\nu}(M,z) \end{equation} where $L_{\nu}(M,z)$ is given by equations (\ref{L_free}) and (\ref{Ndot_scaling}). The comoving free-free emissivity of the IGM may be estimated directly from equation (\ref{free_emissivity}), assuming the entire IGM is ionized. For the purpose of this calculation, we have assumed a uniform IGM and ignored the contribution of overdense structures without an internal source of ionizing radiation. This is a good approximation as the low ambient radiation field in the early universe means that these objects become self-shielding at low column densities and are the last objects to become reionized (Miralda-Escude, Haehnelt, \& Rees, 1998). In Figure (\ref{clump}) we display the relative halo-IGM emissivity $\tilde{\epsilon} \equiv \langle \epsilon \rangle_{\rm halo}/ \langle \epsilon \rangle_{\rm IGM} = \langle n_{e}^{2} \rangle_{\rm halo}/ \langle n_{e}^{2} \rangle_{\rm IGM}$. For $z<20$, the halo emissivity dominates over the IGM emissivity (at high redshift, we underestimate $\tilde{\epsilon}$ as the IGM is only partially reionized. As computed by Haiman \& Loeb (1998), full reionization in this model only occurs at $z \sim 13$). Note that this ratio also applies directly to recombination line emission. We conclude that the free-free and recombination line emission is dominated by discrete sources rather than a diffuse background. \section{Recombination line emission} \subsection{H$\alpha$ emission} The H$\alpha$ flux density from an ionized halo as detected with a R=1000 filter may be estimated as: \begin{equation} J_{\rm H \alpha}= \frac{L_{H \alpha}}{4 \pi d_{L}^{2}} \frac{1}{\Delta \nu} \approx 625 \left( \frac{1+z}{10} \right)^{-1} \left ( \frac{R}{1000} \right) \left( \frac{\rm M} {10^{9} {\rm M_{\odot}}} \right) {\rm nJy} \label{J_Halpha} \end{equation} where we have estimated $L_{H \alpha}$ from equations (\ref{L_Halpha}) and (\ref{Ndot_scaling}). Note that the line width of H$\alpha$ is simply given by Doppler broadening $\Delta \lambda / \lambda \sim 10^{-4} (b/30 {\rm kms^{-1}})$ (unlike the Ly$\alpha$ line, which will have broad damping wings), so it is still not resolved at R=1000. Using a narrow bandpass enables us to reduce the amount of background noise. Let us compare with the prospects for observability in the infra-red, which corresponds to rest-frame UV emission. The UV continuum emission in the rest frame 1500--2800 \AA (longward of the Lyman limit) for a Salpeter IMF is $L_{\nu} = 8 \times 10^{27} \left(\frac{{\rm SFR}}{\rm 1 M_{\odot} yr^{-1}}\right) = 8 \times 10^{27} \left(\frac{\rm \dot{N}_{ion}} {\rm 10^{53} photons \, s^{-1}} \right) {\rm erg \, s^{-1} Hz^{-1}}$ (Madau 1998). Using equation (\ref{Ndot_scaling}), this translates into an observed IR flux density: \begin{equation} J_{\rm IR}= \frac{L_{\nu}}{4 \pi d_{L}^{2}} (1+z) \approx 16 \left( \frac{1+z}{10} \right)^{-1} \left( \frac{\rm M} {10^{9} {\rm M_{\odot}}} \right) {\rm nJy} \label{J_UV} \end{equation} Hence, $J_{\rm H \alpha}/J_{\rm R}\approx 40 (R/1000)$. Note that the spectral energy distribution/IMF of a luminous source can be probed by its color, while the ratio of the H$\alpha$ luminosity to the rest frame UV luminosity will constrain the escape fraction of ionizing photons. Let us examine the competitiveness of H$\alpha$ with continuum IR imaging with NGST. The signal to noise ratio of an observation with exposure time $t$ is given by (Gillett \& Mountain 1998): \begin{equation} \frac{\rm S}{\rm N}=\frac{I_{s}t}{(I_{s}t +I_{bg}t+n\cdot I_{dc} \cdot t + n \cdot N_{r}^{2})^{1/2}} \label{signal_to_noise} \end{equation} where $I_{s}=1.5 \cdot 10^{7} A \cdot \epsilon \cdot 1/R \cdot j_{\nu} \cdot f$ electrons ${\rm s^{-1}}$ is the signal photocurrent ($A$ is the area of the primary mirror in $m^{2}$, $\epsilon$ is the detector quantum efficiency ($\epsilon > 80\%$ for $\lambda < 5 \mu$m, $\epsilon > 50\%$ for $\lambda > 5 \mu$m), $j_{\nu}$ is the source strength in Jy, and f is the fraction of source photons collected ($f \approx 0.7$ for $\lambda < 3.5 \mu$m and $f \approx 0.5$ for $\lambda > 3.5 \mu$m)), $I_{bg}=1.5 \cdot 10^{7} A \cdot \epsilon \cdot 1/R \cdot \beta_{\nu} \cdot \Omega$ electrons ${\rm s^{-1}}$ is the background photocurrent ($\beta_{\nu}$ is the background in Jy ${\rm sr}^{-1}$), $n\approx 4$ is the number of pixels required to cover $\Omega \approx 0.01 \, {\rm arcsec^{2}}$, $I_{dc}$ is the detector dark current, and $N_{r} < 15 \, e \, {\rm read^{-1}}$ (Stockman et al 1997) is the read noise per pixel. We assume a point source subtends $\Omega \approx {\rm max} (1,(\lambda/3.5 \mu m)) \, 0.01 \, {\rm arcsec^{2}}$. For $\lambda=0.5-5 \mu$m, the projected dark current is $I_{dc} <0.02 {\rm e \, s^{-1}}$ for a Si:As IBC array, while for $\lambda > 5 \mu$m with an InSb array, $I_{dc} < 1{\rm e \, s^{-1}}$ (Stockman et al 1997). To estimate the sky background, we consider twice the sky background observed by COBE (Hauser et al 1998). The sky background is measured at $1.25, 2.2, 3.5, 4.9, 12, 25, 60 \mu$m. We assume the full wavelength dependence of the sky background in our calculations, employing a spline interpolation between the observed points. Roughly speaking, the background is $< 10^{-5} \ {\rm Jy \ arcsec}^{-2}$ in the range $\lambda \sim 1-5 \mu$m, and rises to $\sim 5 \times 10^{-4} \ {\rm Jy \ arcsec}^{-2}$ for $\lambda > 10 \, \mu$m. An additional source of noise is thermal emission from the primary mirror, which we approximate as a 75K blackbody (Barsony, 1999). This exceeds the sky background at $\lambda \sim 10 \mu$m, and thus only affects sources in H$\alpha$ emission with $z>15$. For IR continuum imaging ($R = 3$) at $\lambda \sim 1-5 \mu$m, the sky background is the dominant source of noise. Thus, we have ${\rm S/N(IR)} \approx 0.5 (\Omega/{\rm 0.01 \, arcsec^{2}})^{-1/2} \left( \frac{1+z}{10} \right)^{-1} \left( \frac{\rm M} {10^{9} {\rm M_{\odot}}} \right) t^{1/2}$. So, for instance, a 10$\sigma$ detection of our canonical $10^{9} M_{\odot}$ source at z=9 would take 400s. For spectroscopic applications with $R>1000$, the detector noise is the dominant source of noise. The increase in dark current at $\lambda > 5 \mu$m corresponds to reduced a signal to noise ratio for sources with $z>6.6$, leading to ${\rm S/N(H \alpha; z<6.6, z>6.6)} \approx (0.7,0.06) \, (\Omega/{\rm 0.01 \, arcsec^{2}})^{-1/2} \left( \frac{1+z}{10} \right)^{-1} \left( \frac{\rm M} {10^{9} {\rm M_{\odot}}} \right) t^{1/2} $. A 10$\sigma$ detection of our canonical $10^{9} M_{\odot}$ source at z=9 in H$\alpha$ would take $3 \times 10^{4}$s. Note that since the flux density $j_{\nu}^{H \alpha} \propto R$, the signal to noise is independent of the spectral resolution $R$ in the dark current limited regime $R>1000$. Thus, even higher resolution spectroscopy, up to the point when the $H \alpha$ line is resolved ($R \sim 10^{4}$), should be possible. We derive the useful rule of thumb that: \begin{equation} \frac{\frac{\rm S}{\rm N}({\rm H} \alpha)}{\frac{\rm S}{\rm N}({\rm IR})}(z<6.6,z>6.6) = (1.4, 0.12) \left( \frac{t_{{\rm H} \alpha}}{t_{\rm IR}} \right)^{1/2} \label{halpha_vs_uv} \end{equation} Thus, for sources which sit at $z<$6.6, a source may be detected in IR imaging and H$\alpha$ spectroscopy in equal amounts of time. For sources at z$>$6.6, when the detector dark current for H$\alpha$ detection is signficantly larger, detection in H$\alpha$ requires an integration time 100 times longer than in IR imaging to achieve the same signal to noise ratio. This ratio will be reduced if advances in technology permit smaller dark currents at $\lambda > 5 \mu$m. Note that the detection rate depends on the angular extent of the sources, even in the dark current limited regime. An extended source covers more pixels, resulting in a higher dark current (in very extended sources, it may be best to focus on the central pixels, which contain most of the light). For this reason, we quote the relative signal to noise in equation (\ref{halpha_vs_uv}), which is independent of the angular size. In figure (\ref{Halpha_fig}), we show the number of sources detectable with a S/N of 10 with $R>1000$ spectroscopy in $10^{4}$s ($\sim 3$ hours) of integration time, all computed using the full expression (\ref{signal_to_noise}), for both disk-like ($r\sim \lambda r$) and extended ($r \sim r_{vir}$) objects. At least several thousand objects with $z>5$ may potentially be detected in the NGST $4' \times 4'$ field of view. The actual detection rate, of course, depends on the targeting strategy for spectroscopy (see discussion below). Note that if the star formation efficiency is lower, then the detection rate falls. In figure (\ref{lower_SF_efficiency}), we show the detection rate for $f_{star} \sim 1.7\%$, rather than $f_{star} \sim 17\%$. For a star formation efficiency lower by an order of magnitude, an integration time 100 times longer is required to obtain the same number counts. It should be possible to observe higher order Balmer lines as well. For case B recombination with $n_{e}=100 \, {\rm cm^{-3}}$ (the density dependence is fairly weak), $f_{\rm H \beta}/f_{\rm H \alpha}=0.35$ (Osterbrock 1989). Accounting also for the different $\delta \nu= \nu/R$ for H$\alpha$ and H$\beta$, we find that $S/N({\rm H \beta})/S/N({\rm H \alpha})=0.25$ for a given source, which implies an integration time 16 times longer is required to observe a source in H$\beta$ with the same signal to noise ratio as in H$\alpha$ (note that we assume that attenuation effects are still unimportant for H$\beta$, which is of shorter wavelength than H$\alpha$ and thus more suspectible to dust extinction or stellar absorption). However, in the redshift range $6.6<z<9.5$, the H$\beta$ line falls in the regime $\lambda < 5 \mu$m with low dark current, while the H$\alpha$ line falls in the $\lambda > 5 \mu$m regime with high dark current. In this case $S/N({\rm H \beta})/S/N({\rm H \alpha})=2$, i.e., the H$\beta$ line is potentially easier to observe than the H$\alpha$ line. Observing the H$\beta$ line should provide a firmer redshift identification, while the H$\alpha$/H$\beta$ ratio gives a useful measure of extinction. In figure (\ref{Halpha_fig}), we show the detection rate of the H$\beta$ line as a function of redshift. How can we pick out potential high redshift targets for spectroscopy? They may be quickly identified via continuum imaging using the Lyman-break technique successfully used at low redshifts (Steidel et al 1996). Beyond the redshift of reionization, an analogous technique using the Gunn-Peterson trough to detect dropouts can be used. Photometric redshifts may be supplemented by a narrow band filter search ($R \sim 100-400$) to isolate $H\alpha$ bright galaxies in a narrow redshift range. The filter width involves some trade-offs: the filter should be broad enough to cover a sufficiently large comoving volume, yet narrow enough that the flux density in H$\alpha$ greatly exceeds continuum flux densities. Note that current plans for NGST allow for multi-object spectroscopy in the NIR $1-5 \mu$m range, using micro-mirrors but only single slit spectroscopy in the MIR range, $5-20 \mu$m (Stockman et al 1997). However, even in the NIR a grating spectrograph will be able to perform spectroscopy for only a few hundred ($\sim$ several percent) of the objects in the field of view at a given time, thus requiring many pointings to complete a survey. Alternatively, a promising proposal is the Infrared Fourier Transform Spectrograph (IFTS) (Graham et al 1998) which will be able to perform panchromatic observations in the $1-15 \mu$m range (which is the desired $z<20$ range for H$\alpha$ observations). It will be able to acquire both broad band imaging data and higher spectral resolution data (R=1-10000) for all objects in the field of view simultaneously, with very high (near $100\%$) throughput. Thus, the need for object preselection is obviated, and such an instrument would be extremely well suited for our purposes. \subsection{Comparison with Ly$\alpha$} While ionizing objects are expected to be more luminous in Ly$\alpha$ than in H$\alpha$, detection in H$\alpha$ has a number of nice features. Firstly, since H$\alpha$ is emitted in rest-frame optical, it is less susceptible to attenuation by dust than Ly$\alpha$, or rest-frame UV continuum. It is thus a significantly cleaner tracer of the star formation rate. Furthermore, the resonant scattering of Ly$\alpha$ photons within the HII region and in surrounding HI regions means that they will diffuse over a large region, resulting in an unobservably low surface brightness. The large optical path of Ly$\alpha$ photons as they random walk out of these regions greatly increases their probability of absorption by dust (Charlot \& Fall 1993). Finally, the observability of Ly$\alpha$ is affected by complex velocity flows of the neutral gas, which may be outflowing from the ionizing source (thus, for instance, the observation of P-Cygni type profiles in local star-forming galaxies (see Kunth el al 1998 and references therein)). Note that 50\% of the objects selected by the Lyman-break technique show no emission in Ly$\alpha$, while the rest have weak rest-frame equivalent widths $< 20 \, {\rm \AA}$ (Steidel et al 1996). By comparison, the expected rest-frame equivalent widths of star-forming galaxies from modelling their spectral energy distribution is ${\rm W_{\alpha}} \sim 100-200 \, {\rm \AA}$ (Charlot \& Fall 1993). Also, note that quasars are expected to have equivalent widths in the quasar rest frame of ${\rm W_{\alpha}}= 0.68 \frac{c}{\nu_{\alpha}f_{c}} \int^{\infty}_{\nu_{L}} d\nu f_{\nu}/\nu \approx 827 \alpha^{-1} (3/4)^{\alpha} {\rm \AA} = 360 {\rm \AA} (\alpha = 1.5)$ (Charlot \& Fall 1993), where $\nu_{L}, \nu_{\alpha}$ are the Lyman limit and Ly$\alpha$ frequencies, $f_{c}$ is the continuum flux near Ly$\alpha$, and $f_{\nu} \propto \nu^{-\alpha}$ blueward of Ly$\alpha$. In fact, most bright quasars are observed to have rest-frame equivalent widths of $50 \leq {\rm W_{\alpha}} \le 150 {\rm \AA}$ (Schneider, Schmidt \& Gunn 1991, Baldwin, Wampler \& Gaskell 1989). This could be due to covering of the AGN source by HI clouds or dust extinction effects. We tentatively conclude that even in the post-reionization epoch, a Ly$\alpha$ survey is an inefficient means to search for ionizing sources. Until recently, H$\alpha$ surveys of high-z objects have not been carried out due to the difficulties of infrared spectroscopy from the ground. It is intriguing to note that recent such H$\alpha$ surveys (Glazebrook et al 1998) show a star formation rate approximately three times as high as that inferred from the 2800 ${\rm \AA}$ continuum luminosity by Madau et al (1996). This reinforces the expectation that surveys in UV continuum and Ly$\alpha$ tend to underestimate the star formation rates as they are subject to highly uncertain dust extinction effects. Most importantly for our purposes, a H$\alpha$ survey would be able to detect ionizing sources prior to the epoch when HII regions overlap, when the universe was optically thick to Ly$\alpha$ photons. In order for Ly$\alpha$ emission from a source to be unaffected by resonant scattering, the ionizing source must ionize a region large enough for the Ly$\alpha$ photon to redshift out of the Gunn-Peterson damping wing. This problem has been considered by Miralda-Escude (1998). Given that the optical depth of the Gunn-Peterson damping wing as a function of displacement $\Delta \lambda$ from line center, $\tau(\Delta \lambda)= 1.7 \times 10^{-3} (1+z/10)^{3/2} (\Delta \lambda / \lambda)^{-1}$, a region $r \sim 1$ Mpc in proper size must be ionized, independent of redshift, in order to have $\tau(\Delta \lambda) \sim 0.3$. Note that this is somewhat greater than the proper radius of influence of a single ionizing source at z=9, $r_{\rm influence} \sim 0.1 {\rm Mpc} (M/10^{9} M_{\odot})^{1/3} (1+z/10)^{-2}$ (assuming the source never reaches its equilibrium Stromgren radius). Nonetheless, the clustering of ionizing sources means that many should exist in large, overlapping ionized regions. Even if an ionized region $r \sim 1 \, $Mpc is created, the high recombination rate due to the higher IGM densities are high redshift results in a residual neutral fraction which is optically thick to Ly$\alpha$. If we take the column density to be $N_{HI} \sim x n_{b} r$, where $x$ is the ionization fraction in ionization equilibrium, then the optical depth to Ly$\alpha$ scattering in transversing this region is: \begin{equation} \tau_{o} \simeq 100 \left( \frac{1+z}{10} \right)^{6} (1+ \delta_{b})^{2} \left( \frac{J_{21}}{0.5} \right) \left ( \frac{T}{10^{4} K} \right)^{-0.7} \end{equation} where $J_{21}$ is the background ionizing intensity in units of $10^{-21} {\rm erg \, s^{-1} \, cm^{-2} \, Hz^{-1} \, sr^{-1}}$, and $\delta_{b}=\delta \rho_{b}/\rho_{b}$ is the gas overdensity. It therefore seems dubious that unscattered Ly$\alpha$ can be observed except in highly underdense regions, or unless ionizing sources are considerably more luminous than we have considered. Loeb \& Rybicki (1999) have considered the observational signature of Ly$\alpha$ emission of an ionizing halo, given that the Ly$\alpha$ photons are likely to resonantly scatter in the surrounding neutral IGM, before redshifting out of resonance. They find that the Ly$\alpha$ photons emitted by a typical source at $z \sim 10$ scatter over a characteristic angular radius of $\sim 15''$ around the source and compose a line which is broadened and redshifted by $\sim 10^{3} {\rm km \, s^{-1}}$. However, the large angular scale leads to a very low surface brightness. Adopting the same estimates as the previous section for instrumental noise and the IR background, we find that the signal is likely to be overwhelmed by instrumental noise. Assuming $\dot{N}_{\rm Ly \alpha} \approx \dot{N}_{ion}$, so that the flux $f_{\rm Ly \alpha} \approx 12 f_{\rm H \alpha}$ and the flux density $j_{\rm Ly \alpha} \approx 2.2 j_{\rm H \alpha}$ (Case B), the relative signal to noise ratio at the same spectral resolution is given by: \begin{equation} \frac{\frac{\rm S}{\rm N}({\rm Ly} \alpha)}{\frac{\rm S}{\rm N}({\rm H \alpha})}(z<6.6,z>6.6)= (8 \times 10^{-3} ,0.1) \left(\frac{\Omega_{\rm H_{\alpha}}}{0.01 {\rm arcsec^{2}}} \right)^{1/2} \left(\frac{\Omega_{\rm Ly_{\alpha}}}{700 {\rm arcsec^{2}}} \right)^{-1/2} \label{compare_ly_h} \end{equation} (in fact, line broadening for Ly$\alpha$ is likely to further lower this ratio by a factor of a few). Thus, with $ {\rm S/N(Ly \alpha)} = 6 \times 10^{-3} (\Omega/700 {\rm arcsec^{2}})^{-1/2} \left( \frac{1+z}{10} \right)^{-1} \left( \frac{\rm M} {10^{9} {\rm M_{\odot}}} \right) t^{1/2}$, our canonical source $10^{9} {\rm M_{\odot}}$ source at z=9 would require an unacceptably long integration time of 32 days for a 10$\sigma$ detection. If even trace amounts of dust are present, the large number of resonant scatterings means that the Ly$\alpha$ signal is likely to be severely attenuated. We conclude that this would be a very difficult observation, although extremely interesting if carried out successfully. In particular, it would be very interesting to view the same bright source in Ly$\alpha$ and H$\alpha$. Their relative intensities would establish the importance of dust extinction, while their relative spatial extent would constrain the topology of neutral gas surrounding the source. Note that Loeb \& Rybicki (1999) assume a uniform IGM surrounding the source. If the surrounding IGM is inhomogeneous, the Ly$\alpha$ photons are likely to diffuse along the underdense directions, leading to a signal of even lower surface brightness. In addition, Lyman limit systems will absorb a certain fraction of the Ly$\alpha$ photons (since photons scattering in these systems are no longer redshifting out of resonance). \section{Free-free emission} Free-free emission from galaxies forms a fluctuating background in radio frequencies similar to the EBL sought in the optical (Vogeley 1997) and the CIB sought after in the infra-red (Kashlinsky et al 1996, Dwek et al 1998). It is has the potential to yield a great deal of useful information in that it is a direct tracer of star formation and clumping, it is unaffected by dust, there are no unknown K-corrections to be applied (due to the flat spectrum nature of free-free emission), and sources at very high redshifts (z $>$ 10) make a significant contribution and are still potentially observable. We adopt two approaches in characterizing the Free-Free Background (FFB): detecting the mean signal (via spectral distortion of the CMB) and detecting the fluctuating signal (by direct imaging of ionized halos or detecting temperature fluctuations across the sky). \subsection{Spectral distortion} Consider first the mean sky averaged signal $\langle S \rangle$. The integrated emission from ionized halos creates a spectral distortion of the CMB which is potentially observable. Note that for a given mean surface brightness $\langle S \rangle$, the temperature perturbation rises quadratically at low frequencies, where (in the Rayleigh-Jeans limit) $\Delta T_{ff}=c^{2}\langle S \rangle /2k\nu^{2}$. This spectral distortion may be characterised by the optical depth to free-free emission (Bartlett \& Stebbins 1991) $Y_{ff}=\frac{\Delta T_{ff}}{T_{\gamma}} x^{2}$ where $T_{\gamma}$ is the undistorted CMB photon temperature, $x \equiv h\nu/kT_{\gamma}$ and $\Delta T_{ff}=T_{\rm eff}-T_{\gamma}$ is the mean temperature perturbation created by free-free emission (note that the $Y_{ff}$ has a different spectral form from Compton y-distortion (parameterized by $y$), arising from Compton scattering of CMB photons off hot gas). The quadratic nature of the distortion means that a bound on $Y_{ff}$ is best obtained by comparing temperature measurements at high frequencies (when the perturbation due to free-free emission is small) and low frequencies (when it becomes important). The CMB blackbody temperature near the peak of emission has been measured by the FIRAS instrument aboard COBE to be $T_{\gamma}= 2.768 \pm 0.004$K (95\% CL) (Fixsen et al 1996), while at low frequencies the effective temperature has been measured as $T_{\rm eff}=2.65^{+0.33}_{-0.30}$K at 1.4 GHz (Staggs et al 1996a) and $T_{\rm eff}=2.730 \pm 0.014$K at 10.7 GHz (Staggs et al 1996b). Combining the limits on extant data to compute $\Delta T_{ff}$, one obtains $Y_{ff} < 1.9 \times 10^{-5}$ (95\% CL) (Smoot \& Scott 1996). Significant amounts of free-free emission can still be present without violating this constraint. At low frequencies, it corresponds to a limit on the mean temperature perturbation: \begin{equation} \Delta T_{ff}\le 4.4 \times 10^{-2} \left(\frac{\nu}{2 GHz} \right)^{-2} \ K \label{ydistort} \end{equation} One of the chief goals of the DIMES satellite (Kogut 1996) is to measure spectral distortions from free-free emission at low frequencies. DIMES will have a frequency coverage 2--100 GHz and a sensitivity $0.1$mK (at 2 GHz). It is hoped that a measurement of $\Delta T_{ff}$ will indicate the optical depth of the ionized IGM and thus constrain the redshift of reionization $z_{r}$. However, we have already seen from quite general arguments in section 2.2 that the mean emission from ionized halos should dominate that of the IGM. One can examine this in more detail. The equation of cosmological radiative transfer (Peebles 1993) gives the observed surface brightness of the ionized IGM as: \begin{equation} J(\nu_{o}) = \frac{1}{4 \pi} \int_{0}^{z_{r}} dz \frac{dl}{dz} \frac{\epsilon(\nu(1+z))}{(1+z)^{3}} \label{pathint} \end{equation} where $\epsilon$ is given by equation (\ref{free_emissivity}). If $z_{r}=13$, as in our model, then this translates into $\Delta T_{ff}=c^{2}J/2k\nu^{2}= 6 \times 10^{-6}$K at 2GHz, which is beyond the 0.1mK sensitivity of DIMES, which will only be able to detect $z_{r} > 85$ for the $\Lambda$CDM cosmology we have adopted. On the other hand, if we compute $\langle S \rangle$ from equation (\ref{moments}) (setting $z_{\rm min},S_{c}=0$ since point source removal is not possible with the $10^{\circ}$ beam of DIMES), we obtain $\Delta T_{ff}= 3.4 \times 10^{-3}$K, well within the capability of DIMES, while still satisfying the y-distortion constraint by an order of magnitude. Note that we have assumed that free-free emitters are the only radio-bright sources in the sky. The existence of radio-loud galaxies and AGNs(see section 4.3.3 for discussion) is likely to increase the spectral distortion due to radio sources. Thus, it is likely that any spectral distortion detected by DIMES will not constrain the reionization of the IGM but rather the integrated free-free and other radio emission from galaxies, which provides a stronger signal by at least an order of magnitude. \subsection{Direct Detection of Ionized Halos} A natural way to determine if an observed spectral distortion is due to free-free emission from the first ionized halos or the IGM is to attempt to directly detect sources with a high angular resolution instrument. \subsubsection{Detection of point sources} The free-free flux from an ionized halo may be estimated as: \begin{equation} S_{\rm ff}= \frac{L_{\nu}^{\rm ff}}{4 \pi d_{L}^{2}} (1+z) \approx 2.5 \left( \frac{1+z}{10} \right)^{-1} \left( \frac{\rm M} {10^{9} {\rm M_{\odot}}} \right) \left( \frac{\rm T} {10^{4} \, {\rm K}} \right)^{-0.35} {\rm nJy} \label{J_free_free} \end{equation} where we have estimated $L_{\nu}^{\rm ff}$ from equations (\ref{L_free}) and (\ref{Ndot_scaling}). Assuming a point source, and the 0.1'' resolution of the Square Kilometer Array (SKA), this translates into a brightness temperature of: \begin{equation} T_{b}^{ff}= 0.09 {\rm K} \left( \frac{\nu}{2 \, {\rm GHz}} \right)^{-2.1} \left( \frac{T} {10^{4} {\rm K}} \right)^{-0.35} \left( \frac {M}{10^{9} M_{\odot}} \right) \left( \frac{1+z}{10} \right)^{-1} \left(\frac{\theta}{0.1''} \right)^{-2} \end{equation} On the other hand, the detector noise may be estimated as (Rohlfs \& Wilson 1996): \begin{equation} S_{\rm instrum}= \frac{2 kT_{\rm sys}}{A_{\rm eff} \sqrt{2 t \Delta \nu}}= 10 \left(\frac{\Delta \nu}{1 \, {\rm GHz}} \right)^{-1/2} \left( \frac{t} {10^{5} \, s} \right)^{-1/2} {\rm nJy} \label{noise} \end{equation} where we have used $A_{\rm eff}/T_{\rm sys}=2 \times 10^{8} {\rm cm^{2}/K}$ for the SKA (Braun et al 1998), and we assume a bandwidth $\Delta \nu \approx 0.5 \nu$. Thus, in 10 days, a 5$\sigma$ detection of a 16 nJy source is possible. This is a tremendous leap in sensitivity; the deepest observations with the VLA at 8.4 GHz (Partridge et al, 1997) with approximately a week's total integration time were able to identify point sources at the 7 $\mu$Jy level (the rms sensitivity was $1.5 \mu$Jy), with a resolution of 6''. For a given integration time, the sensitivity for detection of a given halo in free-free emission with the SKA is significantly less than that for H$\alpha$ detection (with $R> 1000$) with NGST. The relative signal to noise is: \begin{equation} \frac{\frac{\rm S}{\rm N}({\rm free-free)}}{\frac{\rm S}{\rm N}({\rm H \alpha})}(z<7,z>7) =(1.1 \times 10^{-3},1.3 \times 10^{-2}) \left( \frac{\Delta \nu}{1 \, {\rm GHz}} \right)^{1/2} \left( \frac{t_{\rm ff}}{t_{\rm H\alpha}} \right)^{1/2} \label{compare_free_h} \end{equation} where the above ratio holds in the regime where free-free detection is limited by instrumental noise (for a discussion of confusion noise, see section (\ref{foreground_section})). Thus, the SKA will only be able to detect bright sources. However, note that its field of view is considerably larger than that of NGST--$1^{\circ}$ rather than 4'. In figure (\ref{Num_sources_free}) we compute the number of sources detectable above a given flux threshold $S_{c}$. We find that $\sim 10^{4}$ sources with $z>5$ should be present in the 1 square degree field of view of the Square Kilometer Array above a source detection threshold of $70$nJy. \subsubsection{Power spectrum of unresolved sources} Identification of individual sources may not always be feasible. Many sources will too faint to be individually detected, or they may be crowded and blended together in a confused field. Such sources can still be detected statistically by observing the fluctuations in the FFB due to sources below the flux limit. The discreteness of sources create large amounts of small scale power. The power spectrum is white noise ($C_{l} \sim$ constant) until the angular scale of the beam (or the typical scale of a halo, if it is larger), beyond which it damps rapidly. We can compute compute the statistics of the background below the flux cut $S_{c}$, assuming the sources are unclustered. In particular, we compute the power spectrum of the fluctuations from equation (\ref{moments}): \begin{equation} C_{l}^{\rm Poisson}=\langle S^{2} \rangle= \int^{S_{c}}_{0} \frac{\partial N}{\partial S} S^{2} dS \end{equation} (independent of $l$), and thus compute the rms temperature fluctuations from $ T_{\rm rms}^{2}= \sum_{l} \frac{2l+1}{4\pi} C_{l}^{\rm Poisson} B_{l}^{2} \approx \frac{C_{l}^{\rm Poisson}}{4 \pi \theta^{2}}$ where we have taken the beam of the telescope to be Gaussian: $B_{l}^{2}=e^{-l(l+1)\theta^{2}}$. Note that the rms temperature fluctuations increase with increasing resolution. In Figure (\ref{background}) we display the computed rms temperature fluctuations at 2 GHz and with $4''$ resolution. Also shown is the mean signal $\bar{T}$, computed from the first moment of equation (\ref{moments}), which is independent of angular resolution. Our model is consistent with present observations. With $S_{c}=7 \mu$Jy and $\theta=6''$ we obtain $\Delta T/T= 9.5 \times 10^{-5}$ at 8.4GHz, as compared with the measurement $\Delta T/T \approx 7 \pm 8 \times 10^{-5}$ of Partridge et al (1997), and within their 95\% confidence limit $\Delta T/T < 1.3 \times 10^{-4}$. Asssuming that galactic foregrounds are negligible (see section (\ref{foreground_section})), the main source of noise is instrumental. From equation (\ref{noise}), the rms sensitivity to temperature fluctuations for a beam of angular size $\theta$ is: \begin{equation} T_{noise}=0.35 \, {\rm K}\left(\frac{\nu} {2 \,{\rm GHz}} \right)^{-2.1} \left(\frac{\theta}{0.1''} \right)^{-2} \left(\frac{\Delta \nu}{1 \, {\rm GHz}} \right)^{-1/2} \left( \frac{t} {10^{5} \, s} \right)^{-1/2} \end{equation} For the same parameters, and assuming a cut-off flux of $S_{c}=70$nJy, our expected signal is only $0.02$K. However, note that our expected signal $T_{rms} \propto \theta^{-1}$, whereas the instrumental noise $T_{\rm instrum} \propto \theta^{-2}$. By varying the size of the interferometer, we can vary the angular size of the beam, optimising it for this observation. Referring to figure (\ref{angular}), we see that on angular scales $\sim 1''-10''$, the temperature fluctuations due to the free-free background exceeds both the instrumental noise level and the galactic synchrotron foreground. Therefore, observing a radio field at coarser resolution increases the signal to noise ratio of temperature fluctuations sufficiently to enable the fluctuating free-free background to be detected. Note that if we simply smooth the map, the $T_{\rm instrum} \propto \theta^{-2}$ scaling is not strictly applicable. Smoothing the map corresponds to downweighting longer baseline information, which essentially means that some information is thrown away. This decreases the effective area and increases the detector noise, so $T_{\rm instrum}$ declines more gradually with $\theta$ (although still faster than $T_{rms} \propto \theta^{-1}$). The best solution would be to observe in a more compact configuration with resolution $\sim 1''-10''$ to begin with, so no data is thrown away. One wants to reach a compromise between the need for the high angular resolution required to beat down source confusion in discrete source identification and removal, and the somewhat lower angular resolution optimal for analysing the fluctuating residual signal, or confusion noise itself. A natural solution would be to work at an angular resolution $\theta_{\rm equal}$ where instrumental and confusion noise are comparable. Working at higher resolutions than this does not yield significant advantages for discrete source extraction, since the instrumental noise background in flux units is independent of angular resolution. By fiat, after source removal the instrumental noise and fluctuating signal are now about equal, and smoothing the fluctuating signal increases the signal to noise ratio sufficiently for an analysis to be undertaken. Note that source removal need not always be in the instrument noise dominated regime, $S_{c}=n_{\sigma} S_{\rm instrum}$. Source identification in a crowded field can be difficult, and not all sources can be removed down to the flux limit. If the angular size of typical objects is large, $\theta_{\rm typ} > \theta_{\rm equal}$, then confusion noise exceeds instrumental noise and $S_{c} > n_{\sigma} S_{\rm instrum}$. In this case, increasing the angular resolution beyond $\theta_{\rm typ}$ yields no advantage. Both because of the higher cut-off flux $S_{c}$, as well as the lower angular resolution, detecting the fluctuating signal will not be a problem. Furthermore, as we discuss in section (\ref{foreground_section}), our model probably underestimates the temperature fluctuations which will be observed, which is likely to have components from other radio souces in the sky such as low-redshift radio galaxies and radio-loud AGNs. The free-free component may be separated from other contaminants in multi-frequency observations by its spectral signature (flat spectrum, as opposed (for instance) to the $B_{\nu} \propto \nu^{-0.9}$ spectrum of synchrotron radiation). In addition, besides merely computing RMS temperature fluctuations, it would be desirable to compute the angular power spectrum of the observed signal and confirm its white noise character. In performing this analysis one has to take care to ensure that there is no aliasing of power on scales larger than the field of view (due to Galactic foregrounds and the CMB). This can be done by high-pass filtering and edge tapering the map. Besides the various sources of noise depicted in figure (\ref{angular}), one also has to contend with sample variance, given by: \begin{equation} (\Delta C_{l})^{2}= \frac{1}{f_{\rm sky}} \frac{2}{2 l+1} C_{l}^{2} \label{sample_variance} \end{equation} where $f_{\rm sky}$ is the fraction of sky covered (this expression assumes the $C_{l}$'s are Gaussian distributed, as they should be by the Central Limit Theorem, since a large number of modes contribute). In addition, one probably will bin the computed $C_{l}$'s in bins of size $\Delta l$ to increase the signal to noise. Using $l \, \theta \sim \pi$, an angular resolution of $\theta \sim 4''$ corresponds to $l \sim 10^{5}$. The signal to noise for the sample variance term is $\frac{C_{l}}{\Delta C_{l}}=20 \left( \frac{l}{10^{5}} \right)^{1/2} \left(\frac{\Omega}{1^{\circ} \, \Box} \right)^{1/2} \left(\frac{\Delta l}{100} \right)^{1/2}$, which may be increased by increasing the bin size $\Delta l$. Thus, sample variance should not pose any problems. \subsubsection{Clustering of ionized sources} Does clustering of the ionized halos make a significant contribution to fluctuations in the FFB? We find that on arcsecond scales and with a cutoff flux $S_{c}=70$nJy, Poisson fluctuations are more important than the contribution due to clustering. However, with a lower cutoff flux $S_{c}$, or on somewhat larger angular scales, the clustering term dominates. First, let us consider the evolution of the halo correlation function with redshift. The matter correlation function may be computed as: \begin{equation} \xi_{mm}(r,z) = \int_{0}^{\infty} \frac{dk}{k} 4 \pi k^{3} P(k,z) \frac {sin \ kr} {kr} \label{correl_func} \end{equation} However, halos which collapse early in the history of the universe consitute rare density peaks which are highly biased. In a linear bias model, the correlation function of halos is given by: \begin{equation} \xi_{hh}(M_{1},M_{2},r,z)=b(M_{1},z)b(M_{2},z) \xi_{mm}(r,z) \end{equation} where the bias factor is given by Mo \& White (1996): \begin{equation} b(M,z)=1+ \frac{\nu^{2}-1}{\delta_{c}} = 1+ \frac {\delta_{c}}{\sigma^{2}(M,z)} - \frac{1}{\delta_{c}} \label{bias} \end{equation} where $\delta_{c}=1.68$ and $\nu=\delta_{c}/\sigma(M,z)$. For simplicity we assume the correlation function of halos has the same shape at different redshifts, and differs only in amplitude. This assumption is shown to hold in numerical simulations (Kravtsov \& Klypin 1998, Ma 1999), due to a cancellation of non-linear effects in the evolution of the power spectrum and bias factor. We assume $\xi_{mm}(r)$ is a power law on scales below the non-linear lengthscale (since this is shown to hold in numerical simulations), and is given by equation (\ref{correl_func}) on scales above the non-linear lengthscale. In Figure (\ref{cluster}), we compute the halo correlation length $r_{o}^{halo}$ (defined as $\xi_{hh}(r_{o}^{halo},z)=1$) at each epoch $z$, assuming a number weighted bias factor \begin{equation} \tilde{b}(z)= \frac {\int_{M_{*}}^{\infty} b(M,z) n(M,z) dM} {\int _{M_{*}}^{\infty} n(M,z) dM} \end{equation} The cutoff mass $M_{*}=10^{8} (1+z/10)^{-3/2} M_{\odot}$ introduces strong bias at high redshift, with the net result that the correlation length of halos decreases very slowly with redshift, despite the fall in the matter correlation length. For a higher cutoff mass, bias becomes important at lower redshift, and the initial dip seen in Figure (\ref{cluster}) moves to lower redshift and is less pronounced. Thus, the observed clustering in the Lyman-break galaxy sample of Giavalisco et al (1998), with a comoving correlation length at $z \sim 3$ comparable to that in the local universe, is consistent with a detection of the most luminous and massive ($M \sim 8 \times 10^{11} h^{-1} M_{\odot}$) galaxies at that redshift. The clustering of ionising sources at high redshift is therefore non-negligible, and a potentially important effect in contributing to free-free background fluctuations. The angular correlation function of the free-free background below a flux cut-off $S_{c}$ is: \begin{eqnarray} C(\theta,S_{c})&=& \left( \frac{1}{4 \pi} \right)^{2} \int_{0}^{\infty} dz \frac{dl}{dz} \frac{j_{eff}^{2}(z,S_{c})}{(1+z)^{6}} \int_{-\infty}^{\infty} d\Delta \xi_{hh}(r,z) \\ \nonumber &=& \left( \frac{1}{4 \pi} \right)^{2} \int_{0}^{\infty} dz \frac{dl}{dz} \left[ \int_{S_{min}(z)}^{S_{c}} dS \frac{dN}{dVdS} \left( \frac{4 \pi d_{L}^{2} S}{(1+z)} \right) b(S,z) \right]^{2} \ \int_{-\infty}^{\infty} d\Delta \xi_{mm}(r,z) \end{eqnarray} where $j_{eff}(S_{c},z)$ is the proper free-free emissivity of halos at redshift $z$ below the flux limit $S_{c}$, and $b(S,z)=b(M(S,z),z)$. We use the small angle approximation $r^{2}=\Delta^{2} +d_{A}^{2}\theta^{2}$, an excellent approximation since the clustering length covers a small part of the sky, $r_{o}/d_{A} \ll 1$. The contribution to the angular power spectrum due to clustering $C_{l}^{\rm cluster}$ is then straightforwardly computed from the Legendre transform of the correlation function: \begin{equation} C_{l}^{\rm cluster}= 2 \pi \int d(cos \theta) P_{l}(cos \theta) C(\theta) \approx 2 \pi \int_{0}^{\infty} \theta d\theta J_{o}(l\theta) C(\theta) \end{equation} where the last step is valid for $l \gg 1$. In Figure (\ref{background}), we show the dependence of the clustering term on the flux cutoff at fixed angular scale $\theta \sim 1''$. As we lower the flux cutoff $S_{c}$, the Poisson term declines more sharply than the clustering term, until the clustering term dominates. We can understand this intuitively: the Poisson term is dominated by rare bright objects just below the detection threshold. As we lower the detection threshold, such objects are excluded from the residual FFB, leaving a more uniform distribution of faint objects. By contrast, the clustering term, like the mean brightness term, is dominated by faint, low mass objects. In particular, the amplitude of the clustering term is determined by the bias of objects at the mass cut-off, $M_{c}=10^{8} (1+z/10)^{-3/2} \, M_{\odot}$, rather than the bias of the rarer (though more highly biased) high mass objects. Thus, subtraction of bright sources decreases the Poisson term much more effectively than the clustering term. In Figure (\ref{angular}), we compute the amplitude of the clustering contribution to temperature fluctuations as a function of angular scale at fixed flux cutoff $S_{c}=70$nJy. We find that on the arcsecond scales on which the surface brightness of free-free fluctations exceeds instrumental noise and Galactic foregrounds (see below), the Poisson term is more important than the clustering term. However, at a lower flux cutoff the clustering term would dominate on these scales. Note the different angular dependence of the two terms: $C_{l}^{Poisson} \sim $const, whereas $C_{l}^{clustering} \propto l^{\gamma-3}$, where $\xi_{hh} \propto r^{-\gamma}$. On large scales, the matter correlation function falls below a power law, and eventually $C_{l}^{cluster} \propto l^{n}$ (where the linear power spectrum $P(k) \propto k^{n}$). This downturn is unimportant on the scales plotted, though it becomes appreciable on larger scales. At such low fluctuation amplitudes, both galactic foregrounds and intrinsic CMB anistropies dominate over the free-free signal. \subsubsection{Skewness} Another signature of residual discrete sources is the skewness they induce in maps (Refregier, Spergel \& Herbig 1998). We compute this as the third moment of equation (\ref{moments}), $I_{3}= \Omega_{pix} \int_{0}^{S_{c}} dS \frac{dn}{dS} S^{3} \, {\rm Jy^{3}}$. The instrumental noise is Gaussian distributed, and thus has vanishing skewness. The 1$\sigma$ measurement uncertainty on this vanishing skewness is $\sigma[\eta_{instrum}]=\sqrt{\frac{6}{N_{pix}}} \sigma_{instrum}^{3}$, where $\sigma_{instrum}$ is the rms noise per pixel. Thus, the signal to noise of a skewness measurement is $\frac{S}{N}= \left(\frac{N_{pix}}{6} \right)^{1/2} \left(\frac{I_{3}} {\sigma_{instrum}^{3}} \right)$. We assume $\sigma_{instrum} \sim 1 \mu$Jy for the VLA, and $\sigma_{instrum} \sim 10 $nJy for the SKA, and also that point sources may be detected and removed at $S_{c}=7 \mu$Jy for the VLA and $S_{c}=70$nJy for the SKA. The number of pixels is given by $N_{pix}=(FOV/\theta_{pix})^{2}$, where FOV=40' for the VLA and FOV=$1^{\circ}$ for the SKA. We assume pixel sizes of order $\sim 10''$ for the VLA and $\sim 1''$ for the SKA. From equation (\ref{moments}) we obtain:$I_{3}=(1 \mu{\rm Jy})^{3}$ for the VLA and $I_{3}= (12 {\rm nJy})^{3}$ per pixel for the SKA. This translates into $S/N\approx 5^{3}(\theta/10'') (\sigma_{\rm instrum}/1 \mu Jy)^{3}(S_{c}/7 \, {\rm \mu Jy})^{(4-\beta)}$ for the VLA and $S/N\approx 13^{3}(\theta/1'') (\sigma_{\rm instrum}/10 \, {\rm nJy})^{3} (S_{c}/70 {\rm nJy})^{(4-\beta)}$ for the SKA, where $\frac{dn}{dS} \propto S^{-\beta}$ in the neighbourhood of $S_{c}$. The skewness is thus a very efficient way of detecting the residual field, potentially much more powerful than attempting to use the second moment. It uses the fact that free-free sources can only contribute positive temperature fluctuations, whereas fluctuations due to instrumental noise are symmetric about zero. It is thus a fairly robust statistic. Depending on how accurately Gaussian the instrumental noise is, the same exercise can be carried out with kurtosis. Note that we have assumed that Galactic foregrounds have no skewness. Given that their second moment damps rapidly at small scales (see below), both the amplitude and variance of the third moment is likely to be small. It would be interesting to look at existing VLA maps for residual skewness. \subsubsection{Foreground contamination} \label{foreground_section} A possible concern is foreground source contamination. Note that due to the finite thickness of the surface of last scattering, CMB fluctuations damp rapidly beyond $l \sim 10^{3} (\theta \sim 0.1^{\circ})$. Thus, intrinsic CMB fluctations from the surface of last scattering contribute negligibly on the scales probed here, $C_{l}^{CMB} \approx 0$. One can filter out the large scale power with a filter function $F_{l}=1-e^{l(l+1)\sigma_{filter}^{2}}$, where $\sigma_{filter} \sim 0.1^{\circ}$. At small scales, the Galactic foregrounds $C_{l} \propto B(\nu)l^{-3}$ should damp rapidly(Tegmark \& Efstathiou 1996, Bersanelli et al 1996), although note that the empirical $l^{-3}$ power law fit has only been observed for $l<300$ (Wright 1998); we assume this behaviour may be extrapolated to smaller scales. Below, we briefly summarise the estimated contributions from various foregrounds. {\bf Galactic Free-free emission} This component is dangerous as its spectral signature is identical to our signal. However, the large scale component of the Galactic free-free emission can be removed by cross-correlation with H$\alpha$ maps. For instance, the WhaM survey (Reynolds et al 1995; see http://www.astro.wisc.edu/wham) will map the northern sky at $1^{\circ}$ resolution down to 0.01 Rayleighs (1 Rayleigh=$2.41 \times 10^{-7} {\rm erg \, s^{-1} cm^{-2} sr^{-1}}$), or $T_{ff}= 17 \mu{\rm K} T_{4}^{-0.35} \left(\frac{\nu}{2 {\rm GHz}} \right)^{-2.15}$. We assume $B(\nu) \propto \nu^{-0.15}$ and normalise to the DIRBE analysis of Kogut et al (1996), which obtained rms fluctuations of 7.1 $\mu K$ at 53 GHz (COBE's finite beam of course causes an underestimate of the rms fluctuations, but because $C_{l} \propto l^{-3}$ this is a negligible effect). This yields: \begin{equation} C_{l}^{free-free}=4.2 \times 10^{-4} \left( \frac{\nu}{2 {\rm GHz}} \right)^{-4.3} l^{-3} \ K^{2} \ sr \end{equation} which translates into $ \delta T^{free-free}= 2.3 \times 10^{-6} \left( \frac{\theta}{0.1''} \right)^{0.5} \left( \frac{\nu}{2 {\rm GHz}} \right)^{-2.15}$ K. {\bf Galactic Synchrotron} We assume $B(\nu) \propto \nu^{-0.9}$ and normalise to an rms fluctuation of 11 $\mu K$ at 31 GHz, the upper limit of Kogut et al (1996), as indicated by a cross-correlation between 408 MHz and 19 GHz emission (de Oliveira-Costa et al 1998). This yields: \begin{equation} C_{l}^{syn}=6.1 \times 10^{-3} \left( \frac{\nu}{2 {\rm GHz}} \right)^{-5.8} l^{-3} \ K^{2} \ sr \end{equation} which yields $\delta T^{syn}= 8.6 \times 10^{-6} \left( \frac{\theta}{0.1''} \right)^{0.5} \left( \frac{\nu}{2 {\rm GHz}} \right)^{-2.9}$ K. The synchrotron emission may be identified and eliminated by its distinct spectral signature. {\bf Extragalactic Point Sources} Extragalactic point sources are essentially the signal we are looking for. In this paper, we predict the number of sources in a field of view above a certain flux $S_{c}$ which may be directly detected, as well as attempt statistical characterizations of the fluctuating background signal due to faint undetected sources. Here we pause to note two facts. Firstly, there are many other sources in the radio sky (radio galaxies, AGNs, etc) than the free-free emitters we have attempted to characterize. Secondly, when detecting point sources, the fluctuating signal due to undetected sources is also a source of confusion noise. How do the figures in our model compare with existing observations? The deep radio survey by Partridge et al (1997) obtain a maximum likelihood estimate of the integral source count as: \begin{equation} N(\ge S) = (17 \pm 2)\left(\frac{S}{1 \mu{\rm Jy}} \right)^{-1.2 \pm 0.2} {\rm arcmin^{-2}} \label{source_counts} \end{equation} Assuming that most sources have flat spectra $I_{\nu} \propto \nu^{\alpha}$, where $\alpha \sim 0$ in this frequency range, we see from figure (\ref{Num_sources_free}) that our model yields fewer sources than the observed number counts at 7$\mu$Jy by a factor of a few. Our admittedly crude model is not meant to be accurate at low redshift and ignores many contributions to the radio sky. Our low source counts, together with the fact that we satisfy present y-distortion bounds by an order of magnitude, indicate that the model employed is very conservative and probably underestimates the observable signal. On the other hand, extrapolating the observed source counts to fainter flux levels clearly overestimates the signal, as the mean flux $\langle S \rangle$ diverges and violates the y-distortion constraint. Evidently the source counts must turn over at low flux levels (in our model, this occurs because of the low mass cutoff $M_{*}$). To be conservative, we underestimate the signal to noise ratio of projected observations. Thus, we employ our semi-analytic model to estimate the fluctuating background signal, whereas when estimating confusion noise in point source detectation, we use the extrapolated source counts. Note that we are not completely powerless to distinguish against contaminants, such as low redshift AGNs. If star formation occurs in an extended fashion across galactic halos, we should be able to spatially resolve our target sources, whereas AGNs would only show up as point sources. In addition, by observing at several frequencies, the observed spectral index $\alpha$ (where $S_{\nu} \propto \nu^{-\alpha}$) can be used to diagnose the origin of radio emission. In particular, it can be used to distinguish between thermal and non-thermal emission. Finally, as a zeroth order cut, our target high redshift galaxies should be amongst the faintest objects in a sample, both because of increased cosmological dimming, and the fact that collapsed objects at high redshift are intrinsically less massive and luminous. We now estimate the confusion noise. Using equation (\ref{source_counts}), we can place an upper bound on the noise from residual undetected sources from $\langle S^{2} \rangle= \int^{S_{c}}_{0} \frac{\partial N}{\partial S} S^{2} dS=25.5 \, S^{0.8} \mu{\rm Jy^{2}} \, {\rm arcmin^{-2}}$, which yields the white noise power spectrum: \begin{equation} C_{l}^{PS}=2.4 \times 10^{-15} \left( \frac{S_{c}}{70 {\rm nJy}} \right)^{0.8} \left( \frac{\nu}{2 \ {\rm GHz}} \right)^{-4 +2\alpha} \ {\rm K^{2} \, sr} \end{equation} This gives $\delta T^{PS}= 0.09 {\rm K} \left ( \frac{\theta}{0.1''} \right)^{-1} \left( \frac{S_{c}}{70 {\rm nJy}} \right)^{0.4} \left( \frac{\nu}{2 {\rm GHz}} \right)^{-2+\alpha}$, by far the leading contribution to foreground contamination at these frequencies and angular scales. We can compare the relative contribution of instrumental and confusion noise by demanding that the source flux cutoff be $n_{\sigma} \sim 7$ times the rms noise. In this case, using equation (\ref{noise}), we obtain $\frac{S_{confusion}}{S_{instrum}}=1. \left( \frac{\theta}{0.4''} \right) \left( \frac{n_{\sigma}}{7} \right)^{0.8} \left( \frac{\Delta \nu}{1 \, {\rm Ghz}} \right)^{0.5} \left( \frac{t}{10^{5} s} \right)^{0.5} \left( \frac{\nu}{2 \, {\rm GHz}} \right)^{\alpha}$. In the confusion noise dominated regime, the appropriate cut-off flux may be determined by directly comparing the flux received from a source at the cut-off flux $S_{c}$ and the background flux below this cut-off: \begin{equation} \frac{S}{N} \approx \frac{S_{c}}{(\langle S^{2} \rangle|_{S_{c}} \theta^{2})^{1/2}} = 6 \left(\frac{S_{c}}{70 \, {\rm nJy}} \right)^{0.6} \left( \frac{\theta}{0.4''} \right)^{-1} \end{equation} Confusion noise thus becomes the leading source of noise in discrete source identification and extraction at low angular resolution or if sources are extended rather than point-like, i.e., $\theta ={\rm max(\theta_{beam},\theta_{object})} > 0.4''$. Will source crowding be a problem? If we extrapolate the Partridge et al (1997) counts down to $70$ nJy, then there will be $\sim 400$ sources per square arcmin, or $\sim 9 {\rm arcsec^{2}}$ per source. With a maximum resolution of 0.1'', source crowding of bright point sources such as disks or mini-quasars should not be a problem. Even if the sources are extended, the angular extent of typical high redshift $M_{*}$ objects is small, $\theta_{vir} \sim r_{vir}/d_{A} < 1''$. However, it is important to note that nothing is known about the appearance of the radio sky below $1 \mu$Jy. It is possible that a host of extended low redshift, low surface brightness objects may suddenly surface at these flux levels. \subsubsection{Lower Star Formation Efficiency Case} \label{lower_SF_section} Let us consider the lower star formation efficiency case where only $1.7\%$, rather than $17\%$, of the gas mass in halos fragments to form stars. This corresponds to an IGM metallicity at z=3 of $Z \sim 10^{-3} \, Z_{\odot}$, rather than $Z \sim 10^{-2} \, Z_{\odot}$. The net effect of the lower value is that while the number of sources remain the same, each source is an order of magnitude fainter. Thus, to obtain the same number counts as previous plots, an integration time 100 times longer is required. Also, the brightness of the fluctuating background is reduced, and is equivalent to removing discrete sources an order of magnitude fainter. The net result is that point source detection is likely to be instrument noise limited rather than confusion noise limited. Thus, after discrete source removal, the fluctuating residual background will be significantly harder to detect. In figure (\ref{lower_SF_efficiency}), we show results for the lower star formation efficiency, assuming the same integration time and source removal threshold as before. Note from the plot of foregrounds that the fluctuating background signal is only marginally detectable. Besides the RMS signal, the skewness is also reduced, to $I_{3}= (0.5 \mu {\rm Jy})^{3}$ for the VLA (down from $I_{3}=(1 \mu{\rm Jy})^{3}$) and $I_{3}= (5.5 {\rm nJy})^{3}$ for the SKA (down from $I_{3}= (12 {\rm nJy})^{3}$). Skewness at this level is still detectable, with S/N=$2.5^{3}(\theta/10'') (\sigma_{\rm instrum}/1 \mu Jy)^{3}(S_{c}/7 \, {\rm \mu Jy})^{(4-\beta)}$ (VLA) and S/N=$5^{3}(\theta/1'') (\sigma_{\rm instrum}/10 \, {\rm nJy})^{3} (S_{c}/70 {\rm nJy})^{(4-\beta)} $ (SKA). The situation in reality is likely lie between the two scenarios we have considered, and thus the actual number counts is probably bracketed by our calculations. The temperature fluctuations in the residual field are detectable if discrete source removal is confusion limited; otherwise, it seems likely that only the non-gaussian signature of the residual field will yield useful information. \section{Kinetic Sunyaev-Zeldovich effect} Sunyaev-Zeldovich effects arise from the scattering of CMB photons off moving electrons, which result in brightness fluctuations of the CMB. They are a very attractive means of detecting high redshift objects, in part because of the well-known fact that the surface brightness of SZ effects is independent of redshift. This is because the increased energy density $(1+z)^{4}$ of the CMB at high redshift keeps pace with surface brightness dimming effects. Thus, since the angular size of an object of fixed physical size increases with redshift beyond $z \sim 2$, the SZ flux from a given object {\it increases} with redshift. The kinetic effect is given by $\frac{\Delta T}{T_{CMB}}=\frac{v}{c} \tau$ while the thermal effect is given by $(kT/m_{e} c^{2})\tau$, where the optical depth $\tau= \int n_{e} \sigma_{T}\, dr$. For the photoionized halos we are considering, typical temperatures are $T \sim 10^{4}$K. Their peculiar velocities are $v_{rms} \approx v_{o} D(z)^{1/2} \, {\rm km \, s^{-1}}$, where $D(z)$ is the growth factor (normalised to 1 today) and the characteristic one-dimensional velocity dispersion of galaxies in the local universe is $v_{o}\sim 600 \, {\rm km \,s^{-1}}$ (Strauss \& Willick 1995). Thus, $(v/c)/(kT/m_{e} c^{2}) \sim 400 (1+z/10)^{-1/2}$ and the kinetic effect dominates over the thermal effect. For an isothermal halo, we obtain $T^{SZ}_{b} \approx 3 \times 10^{-7} \left( \frac{1+z}{10} \right)^{2} \left( \frac{M} {10^{9} M_{\odot}} \right)^{1/3} $ K, similar to the temperature fluctuation produced by a cluster in kinetic SZ at low redshift (the high redshift halo is $\sim 1000$ times smaller, but $\sim 1000$ times denser). Unfortunately, this is likely to be difficult to detect. The flux from a halo is related to the temperature shift by: \begin{equation} S_{\nu}=\frac{2k^{3}T_{\gamma}^{2}}{h^{2}c^{2}}q(x) \int d\Omega |\Delta T_{\nu} (\theta)| = \frac{2k^{3}T_{\gamma}^{3}}{h^{2}c^{2}}q(x) \frac{v}{c} \sigma_{T} \frac{N_{e}}{d_{A}^{2}} \label{SZ_flux} \end{equation} where $q(x)= x^{4}/4{\rm sinh}^{2}(x/2)$ is the spectral function, ,$x \equiv h\nu/kT_{\gamma}$, $T_{\gamma}=2.7 {\rm K}$ is the CMB temperature, and $N_{e}=\int n_{e} d^{3}r \approx M/m_{p}(\Omega_{b}/\Omega_{m})$ is the total number of free electrons in the system. From the form of the spectral function, the kinetic SZ effect peaks at $\nu=217$ GHz. This is well beyond the frequency coverage of the SKA, which has an upper limit of 20 GHz. At this frequency, and taking $v \sim v_{rms}$, we obtain a flux density $S_{\nu}(20 GHz)=0.24 (\frac{M}{10^{9} M_{\odot}}) (1+z/10)^{1.5}$nJy. Comparing with equation (\ref{J_free_free}), the SZ flux density at 20 GHz is $\sim 10$ times less than the free-free flux density. We can maximise the SZ flux by going to higher frequencies. At the peak frequency, $S_{\nu}(217 GHz)= 9.6 (\frac{M}{10^{9} M_{\odot}}) (1+z/10)^{1.5}$nJy, and the kinetic SZ signal dominates over free-free emission. At these frequencies, the MMA (see http://www.mma.nrao.edu) would be the premier instrument, with a resolution of 0.1'' at 230 GHz. However, its rms sensitivity at this frequency is 0.21 mJky/${\rm min}^{0.5}$, which means that even in a week's integration time it will only be able to go down to $\mu$Jy levels. We conclude that SZ detection of high-redshift ionized halos is not possible in the near future. However, it is worth noting that the SKA and MMA {\it will} be able to detect the kinetic SZ effect from the HII regions in the IGM blown by the first luminous sources. If the escape fraction is $f_{esc} \sim 20\%$ for our $10^{9} M_{\odot}$ source, it will be able to create an ionized region with $N_{e} \sim \dot{N}_{ion} t_{o} \sim 2 \times 10^{67}$ electrons (using equation (\ref{Ndot_scaling}) and assuming the source lasts for $t_{o} \sim 10^{7}$yrs). This corresponds to a region $r_{comoving}=(3 N_{e}/4 \pi n_{comoving})^{1/3} \sim 1.2$ Mpc in size (note that $\dot{N}_{ion}/V_{\rm proper} \alpha_{B} n_{\rm proper}^{2} \sim 10^{-7} \ll 1$, so the effect of recombinations is negligible). Since this is much smaller than the coherence length of the velocity field, we assume all parts of the bubble are moving with a uniform peculiar velocity. From (\ref{SZ_flux}), we find that the flux from such an ionized bubble is $S_{\nu}(20 {\rm GHz})= 35$nJy (detectable by the SKA, which, using equation (\ref{noise}) and $A_{eff}/T_{sys}=1 \times 10^{8} {\rm cm^{2}/K}$ at 20 GHz (Braun et al 1998) has a $S_{instrum}=6$nJy rms sensitivity) and $S_{\nu}(217 GHz)=1.4 \, \mu$Jy (detectable by the MMA). However, along any given line of sight one intersects not one but a multitude of ionized bubbles. Thus, the signal has to be detected statistically. Calculating the CMB fluctuations due to inhomogeneous reionization has been the subject of much recent work (Agahanim et al 1996, Grusinov \& Hu 1998, Knox et al 1998; see also Haiman \& Knox 1999, Natarajan \& Sigurdsson 1999), and is beyond the scope of this paper. However, we make two observations. Firstly, predictions have focussed on the possibility of detection by Planck (detection by MAP seems unlikely). The temperature fluctuation power spectrum is generally white noise with a peak on the scale of a typical bubble; the induced $\Delta T/T \sim 10^{-6}-10^{-7}$. For the anistropies to be detectable by Planck, the bubbles need to have an unrealistically large size $r_{\rm comoving} \sim 10$Mpc. For bubbles with $r_{\rm comoving} \sim 1$Mpc, the anisotropies peak on sub-arcminute scales, beyond the angular resolution of Planck ($\theta_{\rm FWHM} \sim 5.5'$ at 217 GHz; see http://astro.estec.esa.nl/Planck/). High resolution ground based instruments such as the SKA or MMA, which can easily resolve the bubbles, might be better suited to the task. Furthermore, the significantly larger collecting area of these instruments means that they will be more sensitive than Planck by about 3 orders of magnitude: with a year of observing time, the rms sensitivity of Planck at 217 GHz is at best $\sim 1$mJy. If we require that the sample variance be sufficiently low that $\delta \equiv C_{l}/ \Delta C_{l} =10$, then using equation (\ref{sample_variance}), we find that a map of size $\Omega= 2 (\delta/10)^{2} (l/10^{4})^{-1} (\Delta l/10^{2})^{-1}$ square degrees is sufficient. This is easily achievable with the SKA. Secondly, the signal due to inhomogeneous reionization (IR) is quite possibly smaller than that due to the Ostriker-Vishniac effect, which arises from the coupling between density and velocity fluctuations in a fully reionized medium (Haiman \& Knox 1999, Jaffe \& Kamionkowski 1998). To isolate the IR signal, it would be interesting to cross-correlate observed CMB fluctuations with the observed distribution of ionizing halos. They could either correlate positively (if halos blow ionized bubbles around themselves) or negatively (if underdense voids are reionized first). \section{Conclusions} In this paper, we have used a simple model based on Press-Schechter theory to make observational predictions for the detection of high redshift galaxies in free-free and H$\alpha$ emission. These signals are valuable as they directly trace the distribution of dense ionised gas at high redshift. Furthermore, they are relatively unaffected by dust or resonant scattering, and their intensities are directly related to one another. They can thus serve as a very powerful probe of the dark ages, and constitute promising applications of the new generation of instruments designed to probe this epoch, in particular the Next Generation Space Telescope (NGST) and the Square Kilometer Array (SKA). A fairly robust prediction of our model is that the integrated free-free and other radio emission from ionizing sources will swamp the free-free emission from the ionized IGM by an least an order of magnitude. Thus, any spectral distortion of the CMB observed by DIMES at low frequencies will constrain the former quantity. This claim depends only upon the small escape fraction of ionizing photons from sources, as is observed in galaxies in the local universe. To evade it, the ionizing sources must themselves contain very little ionized gas, either because gas has been driven out by supernovae, or ionizing photons are able to escape through a hole in the ISM, leaving most of the gas in the host galaxy neutral. The escape fraction of ionizing photons $f_{esc}$ from high redshift halos may be constrained by comparing the rest frame UV luminosities longward of the Lyman limit of the sources as observed by NGST, with the expected luminosities in H$\alpha$ and free-free emission. Given a model for the relative spectral energy distribution, the relative luminosities then directly constrain the escape fraction in halos at high redshift, an important quantity in theories on reionization. If this fraction is low, as we argue, then in a given amount of integration time, the H$\alpha$ emission from an ionized halo should be detected (with R=1000) with the same signal to noise as IR continuum imaging with NGST for $z<6.6$. For $z>6.6$, due to the increase in dark current noise in the $\lambda > 5 \mu$m regime, it will be detected with $10 \%$ of signal to noise ratio of IR imaging. Higher order lines such as H$\beta$ should also be detectable. In addition, the SKA should be able to directly detect the same ionizing sources in free-free emission. While the NGST will be able to detect fainter sources than the SKA, the SKA has a larger field of view: $\sim$200 times larger in solid angle. We find that the SKA should be able to detect $\sim 10^{4}$ individual free-free emission sources with $z>5$ in its field of view. However, a large population of other radio bright sources (AGNs at low redshift, etc) will also be detected. NGST (which can obtain redshift information from the H$\alpha$ line) must therefore be used in conjunction with the SKA to distinguish between these two populations. Free-free sources below the flux limit of the SKA constitute a fluctuating background which can be detected statistically. At arcsecond scales, Poisson fluctuations in the number of free-free sources create temperature fluctuations detectable by SKA. At such small scales, galactic foregrounds are expected to damp rapidly. We find that the expected clustering of high-redshift ionizing sources makes a negligible impact on the observed power spectrum of temperature fluctuations. These temperature fluctuations have a highly non-Gaussian signature. In particular, they should result in a detectable skewness. Finally, ionized gas at high redshift may also be detected by the kinetic Sunyaev-Zeldovich effect. While individual ionized halos are probably too faint to be detectable by the SKA or MMA, the ionized IGM around these halos create CMB anisotropies on sub-arcminute scales which should be detectable by these instruments. I am grateful to my advisor, David Spergel, for his encouragement and guidance. I also thank Michael Strauss for many detailed and helpful comments on an earlier manuscript. Finally, I thank Nabila Aghanim, Gillian Knapp, Alexandre Refregier and Stephen Thorsett for helpful conversations. This work is supported by the NASA Theory program, grant number 120-6207.
\section{INTERACTION BETWEEN A PREDICTIVE MANAGEMENT SYSTEM AND A PREDICTIVE MOBILE NETWORK} There is an interesting interaction between the predictive management system and the predictive mobile network. A predictive mobile network such as the Rapidly Deployable Radio Network \cite{BushICC96, BushThesis} will have cached results in advance of use for many configuration parameters. These results should be part of the Management Information Base for the mobile network and includes the predicted time of the event requiring the result, the value of the result, and the probability that the result will be within tolerance. Thus there will be a triple associated with each predicted event: {\em (time, value, probability)}. Network management protocols, e.g. SNMP \cite{SNMP} and CMIP \cite{CMIP}, include the time as part of the Protocol Data Unit, however this time indicates the real time the poll occurred. A predictive management system could simply use Logical Processes to represent the predictive mobile processes as previously described, however, this is redundant since the mobile network itself has predicted events in advance as part of its own management and control system. Therefore, managing a predictive mobile network with a predictive network management system provides an interesting problem in trying to get the maximum benefit from both of these predictive systems. Combining the two predictive systems in a low level manner, e.g. allowing the Logical Processes to exchange messages with each other, raises questions about synchronization between the mobile network and the management station. However, the predicted mobile network results can be used as additional information to refine the management system results. The management system will have computed {\em (time, value, probability)} triples for each predicted result as well. The final result by the management system would then be an average of the {\em times} and {\em values} weighted by their respective probabilities. An additional weight may be added given the quality of either system. For example the network management system might be weighted higher because it has more knowledge about the entire network. Alternatively, the mobile network system may weighted higher because the mobile system may have better predictive capability for the detailed events concerning handoff. Thus the two systems do not directly interact with each other, but the final result is a combination of the results from both predictive systems. A more complex method of combining results from these two systems would involve a causal network such as the one described in \cite{Lehmann}. \section{INTRODUCTION} Recently proposed mobile networking architectures and protocols involve predictive mobility management schemes. For example, an optimization to a Mobile IP-like protocol using IP-Multicast is described in \cite{Seshan}. Hand-offs are anticipated and data is multicast to nodes within the neighborhood of the predicted handoff. These nodes intelligently buffer the data so that no matter where the mobile host (MH) re-associates after a handoff no data will be lost. Another example \cite{Liu, Liuphd} proposes deploying mobile floating agents that decouple services and resources from the underlying network. These agents are pre-assigned and pre-connected to predicted user locations. This paper focuses on the Rapidly Deployable Radio Networks Project \cite{BushICC96, BushRDRN} as an example of a specific predictive mobile network. The Virtual Network Configuration algorithm developed as part of RDRN uses a predictive mechanism for every phase of configuration, including location and handoff. Progress is being made in research involving predictive system and network management \cite{Schiaffino}. This paper develops a variation of the Virtual Network Configuration Algorithm as proposed for the RDRN \cite{BushRDRN} for a predictive network management system. The predictive capability of such a system can be used to help optimize its own operation by controlling the management of the polling rate. Finally, a discussion of how predictive mobile networks and predictive network management systems should interact is presented. \section{A PREDICTIVE NETWORK MANAGEMENT SYSTEM} Systems management means the management of heterogeneous subsystems of network devices, processing platforms, distributed applications, and other components found in communications and computing environments. Current system management relies on presenting a model to the user of the managed system that should accurately reflect the current state of the system and should ideally be capable of predicting the future health of the system. System management relies on a combination of asynchronously generated alerts and polling to determine the health of a system \cite{Steinber91}. The management application presents state information such as link state, buffer fill, and packet loss to the user in the form of a model \cite{Sylor}. The model can be as simple as a passive display of nodes on a screen, or a more active model that allows displayed nodes to change color based on state changes, or react to user input by allowing the user to manipulate the nodes, causing values to be set on the managed entity. This model can be made even more active by enhancing it with predictive capability. This enables the management system to manage itself, for example, to optimize its polling rate. The two major management protocols, Simple Network Management Protocol (SNMP) \cite{SNMP} and Common Management Information Protocol (CMIP) \cite{CMIP}, allow the management station to poll a managed entity to determine its state. To accomplish real-time and predictive network management in an efficient manner, the model should be updated with real-time state information when it becomes available, while other parts of the model work ahead in time. Those objects working ahead of real-time can predict future operation so that system management parameters such as polling times and thresholds can be dynamically adjusted and problems can be anticipated. The model will not deviate too far from reality because those processes found to deviate beyond a certain threshold are rolled back, as explained in detail later. The processes' messages must obey the following rules for consistency in \cite{Lamport78}: \newtheorem{lamportr}{Rule} \begin{lamportr} If two events are scheduled for the same process, then the event with the smaller timestamp must be executed before the one with the larger timestamp. \end{lamportr} \begin{lamportr} If an event executed at a process results in the scheduling of another event at a different process, then the former must be executed before the latter. \end{lamportr} To determine the characteristics and performance of this predictive network management algorithm, we will review the research on performance and modeling of other lookahead algorithms and Time Warp in particular. A comparison of the conservative Chandy-Misra approach and the optimistic Time Warp is presented in \cite{Felderman90}. This is done using a typical queuing theory approach which assumes exponential service times. There have been several other detailed comparisons between conservative and optimistic methods of simulation. These studies also make simplifying assumptions. In \cite{Lin90a}, it is shown that in a feed-forward network, the time of execution of a message will occur earlier in Virtual Time than its corresponding message in the synchronous parallel algorithm described in \cite{Lamport78}. In \cite{Lipton90}, it is shown that Time Warp can out-perform the conservative technique known as Chandy-Misra by a factor of $P$, $P$ being the number of processors, but that no such model\index{simulation!model} in which Chandy-Misra\index{Chandy-Misra} out-performs Time Warp\index{Time Warp} by a factor the number of processors used exists. Past work has examined the performance of Time Warp by comparing it to conservative mechanisms \cite{Lin90a} or simulating the Time Warp mechanism itself \cite{Turnbull92}. In this paper the focus is not only on analyzing and optimizing speed of execution but also using the algorithm to maintain network management prediction accuracy. One goal of this research is to minimize polling overhead in the management of large systems \cite{Takagi86}. Instead of basing the polling rate on the characteristics of the data itself, the entity is emulated some time into the future to determine the characteristics of the data to be polled. Polling is still required with this predictive network management system to verify the accuracy of the emulation. \section{RELATIONSHIP BETWEEN NETWORK MANAGEMENT AND PARALLEL DISCRETE EVENT SIMULATION} Management information from standards-based managed entities must be mapped into this predictive network management system. Network management systems rely upon standard mechanisms to obtain the state of their managed entities in near real-time. These mechanisms, SNMP \cite{SNMP} and CMIP \cite{CMIP} for example, use both solicited and unsolicited methods. The unsolicited method uses messages sent from a managed entity to the manager. These unsolicited messages are called traps or notifications; the former are not acknowledged while the latter are acknowledged. These messages are very similar to messages used in distributed simulation algorithms; they contain a timestamp and a value, they are sent to a particular destination, i.e. a management entity, and they are the result of an event which has occurred. Information requested by the management system from a particular managed entity is solicited information. It also corresponds to messages in a distributed simulation. It provides a time and a value; however, not all such messages are equivalent to messages in distributed simulation. These messages provide the management station with the current state of the managed entity, even though no change of state may have occurred or multiple state changes may have occurred. The design of a management system that requests information concerning the state of its managed entities at the optimum time has always been a problem in network management. If management information is requested too frequently, bandwidth is wasted, if not requested frequently enough, critical state change information will be missed. We will assume for simplicity that each managed entity is represented in the predictive management system by a Logical Process. It would greatly facilitate system management if vendors provide not only the standards based SNMP MIBs as they do now, but also standard simulation code that models entity or application behavior and can be plugged into the management system just as Management Information Bases are used today. Vendors have such simulation models of their managed devices readily available from product development \section{THE PREDICTIVE NETWORK MANAGEMENT SYSTEM ALGORITHM} Terminology borrowed from previous distributed simulation algorithms has a slightly different meaning in this predictive network management system\index{predictive network management system}. In addition, new terminology must be introduced. Thus it is important that the terminology be precisely defined. The predictive network management system management system algorithm encapsulates Physical Process simulating managed network devices within an Logical Process. A Physical Process is nothing more than the executing process defined by the program code. An Logical Process consists of the Physical Process and additional data structures and instructions to maintain message order and correct operation as the system executes ahead of real time. An Logical Process contains a Receive Queue, Send Queue, and State Queue. The Receive Queue maintains newly arriving messages in order by their Receive Time. The Send Queue maintains copies of previously sent messages in order of their send times. The state of the Logical Process is periodically saved in the State Queue. The Logical Process also contains its notion of time known as Local Virtual Time and a Tolerance ($\Theta$) that is the allowable deviation between actual and predicted values of incoming messages. Also, the Current State of a Logical Process will be the current state of the Logical Process and its encapsulated Physical Process. The predictive network management system\index{predictive network management system} contains a notion of the complete system time known as Global Virtual Time and a sliding window known as the {\bf Lookahead\index{Lookahead}} time ($\Lambda$\index{$\Lambda$|see{Look Ahead}}). Messages contain the Send Time, Receive Time, Anti-toggle\index{Anti-toggle}, and the message contents. The Receive Time is the time this message should be received by the destination Logical Process\index{Logical Process}. The Send Time is the time this message was sent by the originating Logical Process\index{Logical Process}. The Anti-toggle field is the anti-toggle and is used for creating an anti-message\index{anti-message} to remove the effect of false messages\index{false message} as described later. A message will also contain a field for the current Real Time. This is used to differentiate a real message\index{real message} from a virtual message\index{virtual message}. A {\bf driving process\index{predictive network management system!driving process}} is required to predict future events and inject them into the system. For example, in a mobile system such as the Rapidly Deployable Radio Network \cite{BushRDRN}, the Global Positioning System is used to provide each node with its current position. The Global Positioning System receiver process runs in real-time\index{real-time} and inject future predicted location messages. In the predictive network management system, the driving process may be the number of expected users and their estimated bandwidth usage. The driving process(es)\index{predictive network management system!driving process} originate virtual messages via internal prediction. The remaining Physical Processes\index{PP} react to these messages as though they are real messages\index{real message}. A message which is generated and time-stamped with the current time will be called a {\bf real message\index{real message}}. Messages which contain future event information and are time-stamped with a time greater than current time are called {\bf virtual messages\index{virtual message}}. If a message arrives at a Logical Process\index{LP} out-of-order or with invalid information, it is called a {\bf false message\index{false message}}. A false message\index{false message} causes an Logical Process\index{LP} to rollback\index{rollback}. Rollback\index{predictive network management system!rollback} is a mechanism by which a Logical Process\index{LP} returns to a known correct state. The rollback\index{rollback} occurs in three phases. In the first phase, the Logical Process\index{LP} state is restored to a time strictly earlier than the time stamp of the false message\index{false message}. In the second phase, anti-messages are sent to cancel the effects of invalid messages that had been generated before the arrival of the false message\index{false message}. An {\bf anti-message\index{anti-message}} contains exactly the same contents as the original message with the exception of an anti-toggle bit\index{Anti-toggle!anti-toggle bit} that is now set. When the anti-message\index{anti-message} and original message meet, they are both annihilated\index{annihilation}. The final phase consists of executing the Logical Process\index{Logical Process} forward in time from its rollback\index{rollback} state to the time the false message\index{false message} arrived. No messages are canceled or sent between the time to which the Logical Process\index{Logical Process} rolled back and the time of the false message\index{false message}. Because these messages are correct there is no need to cancel or re-send them. This increases performance, and it reduces the number of messages causing roll-back. Note that another false message\index{false message} or anti-message\index{anti-message} may arrive before this final phase has completed without causing any problems. \section{CHARACTERISTICS OF THE PREDICTIVE NETWORK MANAGEMENT SYSTEM} There are two types of false messages generated in this predictive network management system; those produced by messages arriving in the past Local Virtual Time of an Logical Process and those produced because the Logical Process is generating results which do not match reality. If rollbacks occur for both reasons the question arises as to whether system will be stable. A stable predictive network management system is one in which rollbacks do not have a significant impact on system performance. A stable system is able to make reasonably accurate predictions far enough into the future to be useful. An unstable system will have its performance degraded by rollbacks to the point where it is not able to predict ahead of real-time. Initial results, shown later, indicate that predictive network management systems using the algorithm described in this paper can be stable. There are several parameters in this predictive network management system which must be determined. The first is how often the predictive network management system should check the Logical Process to verify that past results match reality. There are two conditions which cause Logical Processes in the system to have states which differ from the system being managed and to produce inaccurate predictions. The first is that the predictive model which comprises an Logical Process is most likely a simplification of an actual managed entity and thus cannot model the entity with perfect fidelity. The second condition occurs when events outside the scope of the model may occur which lead to inaccurate results. However, a benefit of this system is that it self-adjusts to both of these conditions. The optimum choice of verification query time, $T_{query}$, is important because querying entities is something the predictive management system should minimize while still guaranteeing that the accuracy is maintained within some predefined tolerance, $\Theta$. For example, the network management station may predict user location as explained later. If the physical layer attempts spatial reuse via antenna beamforming techniques as in the Rapidly Deployable Radio Network project, then there is an acceptable amount of error in the steering angle for the beam and thus the node location is allowed to be within a tolerance. The tolerances are set for each state variable sent from a Logical Process. State verification can be done in one of at least two ways. The Logical Process state can be compared with previously saved states as real time catches up to the saved state times or output message values can be compared with previously saved output messages in the send queue. In the prototype implemented for this predictive network management system state verification is done based on states saved in the state queue. This implies that all Logical Process states must be saved from the Logical Process LVT back to the current time. The amount of time into the future that the emulation will attempt to venture is another parameter which must be determined. This lookahead sliding window width, $\Lambda$, should be preconfigured based on the accuracy required; the farther ahead this predictive network management system attempts to predict past real time, the more risk that is assumed. \subsection{Tolerance and Accumulated Simulation Error} In order to consider the impact that out-of-tolerance rollback will have on the predictive system, consider how simulation error occurs. A predictive management system Logical Process may deviate from the real object because either the Logical Process does not accurately represent the actual entity or because events outside the scope of the predictive network management system may effect the entities being managed. Ignore events outside the scope of the simulation for now and consider error from inaccurate simulation modeling only. Because of the possibility for prediction error, a method of determining the amount of error in a predicted result needs to be developed. A function of total accumulated error in a predicted result, $AC(\cdot)$, is described by the following Equations: \begin{equation} AC_n(n) = \sum_{i = 1}^N CE_{lp_i}(ME_{lp_{i-1}}, t_{lp_i}) \label{acn} \end{equation} \begin{equation} AC_t(\tau) = \liminf_{\sum t_{lp_i} \rightarrow \tau} \sum_{i = 1}^N CE_{lp_i}(ME_{lp_{i-1}}, t_{lp_i}) \label{actau} \end{equation} $ME_{dp}$ is the error introduced by the virtual message injected into the predictive system by the driving process. The error introduced by the output message produced by the computation of each Logical Process is represented by the computation error function, $CE(\cdot)$. The actual time taken by the $n^{th}$ Logical Process to calculate and output the next virtual message is $t_{lp_n}$. Note that the Logical Process topology may not necessarily be a feed-forward network as described by Equations \ref{acn} and \ref{actau}; it may include a cycle. Note also that the right side of Equation \ref{actau} is the greatest lower bound of all sub-sequential limits of $\sum_{i = 1}^N CE_{lp_i}(ME_{lp_{i-1}}, t_{lp_i})$ as $\sum t_{lp_i} \rightarrow \tau$. The driving process is indicated by $lp_0$. $AC_n(n)$ is the total accumulated error in the virtual message output by the $n^{th}$ Logical Process from the driving process. $AC_t(\tau)$ is the accumulated error in $\tau$ actual time units from generation of the virtual message from the driving process. For example, if a prediction result is generated in the third Logical Process from the driving process, then the total accumulated error in the result is $AC_n(3)$. If 10 represents the number of time units after the initial message was generated from the driving process then $AC_t(10)$ would be the amount of total accumulated error in the result. \subsection{Optimum Choice of Verification Query Times} As previously stated, the prototype system performs the verification based on the states in the state queue. One method of choosing the verification query time is to query the managed entity based on the frequency of the data we are trying to monitor. Assuming the simulated data is correct, query or sample in such a way as to perfectly reconstruct the data, e.g. based on the maximum frequency component of the monitored data. A possible drawback is that the actual data may be changing at a multiple of the predicted rate. The samples may appear to be accurate when they are invalid. \subsection{Verification Tolerance} The verification tolerance, $\Theta$, is the amount of difference allowed between the Logical Process state and the actual entity state. A large tolerance decreases the number of false messages and rollbacks, thus increasing performance and requires fewer management queries. Allowing a larger probability of error between predicted and the actual values will cause rollbacks in each Logical Process at real times of $t_{vfail}$ from the start of execution of each Logical Process. The error throughout the simulated system may be randomized in such a way that errors among Logical Processes cancel. However, if the simulation is composed of many of the same class of Logical Process, the errors may compound rather than cancel each other. The tolerance of a particular Logical Process, $\Theta_{lp_n}$, will be reached in time $t_{vfail_n} = \{ \mbox{lub\ } \tau \mbox{ s.t. } AC_t(\tau) > \Theta_{lp_n} \}$. The verification query period ($\Upsilon$) should be periodic with period less than or equal to $t_{vfail_n}$ in order to maintain accuracy within the tolerance. The accuracy of any predicted event must be quantified. This could be quantified as the probability of occurrence of a predicted event. The probability of occurrence will be a function of the verification tolerance, the time of last rollback due to verification error, the error between the simulation and actual entity, and the sliding lookahead window. Every Logical Process will be in exact alignment with its Physical Process as a result of a state verification query. This occurs every $T_{query} = t_{vfail}$ time units. \subsection{Length of Lookahead Window} The length of the lookahead window, $\Lambda$, should be as large as possible while maintaining the required accuracy. The total error is a function of the chain of messages which lead to the state in question. Thus, the farther ahead of real-time the predictive network management system advances, $t_{ahead} = GVT - t_{current-time}$, the greater the number of messages before a verification query can be made and the greater the error. The maximum error is $AC_t(\Lambda)$. \subsection{Simulation Time} Since the verification query time is less than or equal to the current time, $t_{current-time}$, rollbacks due to the verification query will take the Logical Process back to the current time. Thus, Global Virtual Time as defined in \cite{Jefferson82} is no longer a lower bound on the simulation rollback time. The lower bound is now $t_{current-time}$. Global Virtual Time is still required in order to determine how far into the future the predictive network management system has gone. \subsection{Calibration Mode of Operation} It may be helpful to run the predictive network management system in a mode such that error between the actual entities and the predictive network management system are measured. This error information can be used during the normal predictive mode in order to help set the above parameters. This has an effect similar to back-propagation in a neural network, i.e. the predictive network management system automatically adjusts parameters in response to output in order to become more accurate. This calibration mode could be part of normal operation. The error can be tracked simply by keeping track of the difference between the simulated messages and the result of verification queries. \section{MODEL AND SIMULATION} \begin{comment} Predictive Network Management Maisie Simulation is in /home/bushsf/oldvnc. It can be run with: tw.run. \end{comment} The algorithm described in this paper has been implemented and analyzed in \cite{BushThesis} that describes a predictive mobile network. This paper extends the algorithm to network management. An initial test of this algorithm in a network management environment has been performed in a simulation\index{Sequential Simulation} of a predictive management system implemented with Maisie \cite{bagrodia}. Its suitability has been demonstrated in the RDRN\index{RDRN} network management and control design and development and in \cite{Short} to develop a mobile wireless network parallel simulation\index{Parallel Simulation} environment. The parallel simulation\index{Parallel Simulation} environment shows a speedup over the currently used commercial sequential simulation packages. The environment and a set of modules which have been developed for mobile network simulation are described in \cite{Short}. Maisie uses a language which has been influenced by a classic work describing the characteristics of a parallel programming\index{Parallel Programming} language structure \cite{Hoare81}. Since every Maisie entity has a built-in input queue, each Logical Process is comprised of three additional Maisie entities: \begin{itemize} \item An entity which represents the Physical Process\index{Physical Process} \item An entity for the Logical Process state queue\index{State Queue} \item An entity for the Logical Process output message queue\index{Send Queue} \end{itemize} There is also a gvt entity for the calculation of Global Virtual Time\index{VNC!Global Virtual Time}. All three of the above entities work together to implement Virtual Time as described in \cite{Jefferson82}. The first entity above, representing the Physical Process, contains a delay mechanism in order to implement the sliding lookahead window\index{Sliding Lookahead Window|see{LA}}. The gvt process should notify all processes to cease forward simulation when Global Virtual Time reaches the end of the window. However, in this version of the predictive management system, each Logical Process simply compares its Local Virtual Time to the current time and holds processing until current time is back within the lookahead sliding window. Determination of Global Virtual Time should be done as defined by \cite{Lazowaska90}. This algorithm allows Global Virtual Time to be determined in a message-passing environment as opposed to the easier case of a shared memory environment. It also allows normal processing to continue during the Global Virtual Time determination phase. However, in this implementation each output message is sent to the gvt entity as well as to its proper destination. In addition, the gvt entity checks all Logical Processes for their current Local Virtual Time and chooses the minimum message send time and Local Virtual Time as the current Global Virtual Time. The gvt entity is allowed to execute in parallel with other entities in this simulation, thus it may not always be perfectly accurate. This is because messages may be in transit when the Local Virtual Time request poll takes place, and because Logical Processes are changing while the Global Virtual Time computation is taking place. However, the results are close enough for the purpose of these experiments. \subsection{Verification Query Rollback Versus Causality Rollback} Verification query\index{State Adjustment|see{VNC}} rollbacks are the most critical part of the predictive management system. They are handled in a slightly different fashion from causality\index{VNC!causality} failure rollbacks. A state verification failure causes the Logical Process state to be corrected at the time of the state verification that failed. The state, $S_{v}$, has been obtained from the actual device from the verification query at time $t_{v}$. The Logical Process rolls back to exactly $t_{v}$ with state, $S_{v}$. States greater than $t_{v}$ are removed from the state queue\index{State Queue}. Anti-messages\index{Anti-Message} are sent from the output message queue for all messages greater than $t_{v}$. The Logical Process continues forward execution from this point. Note that this implies that the message and state queues cannot be purged of elements that are older than the Global Virtual Time. Only elements which are older than real time can be purged. \subsection{The Prototype System Simulation} A simple network management system was simulated to test the concept of the predictive management protocol just described. Note that none of the previous assumptions are made in the simulation. The purpose of this simulation is to determine if the concept is feasible\index{VNC!feasibility simulation}. A key question this simulation attempts to answer is whether the overhead/performance ratio results in a useful system. A small closed queuing network with First Come First Serve servers represents the actual system. Figure \ref{qrsim} shows the real system to be managed and the predictive management model. In this initial feasibility study, the managed system and the predictive management model are both modeled with Maisie. The verification query between the real system and the management model are explicitly illustrated in Figure \ref{qrsim}. \begin{figure*}[htbp] \centerline{\psfig{file=figures/qrmodrel.eps,width=5.0in}} \caption{Initial Feasibility Network Model} \label{qrsim} \end{figure*} The system consists of three switch-like entities, each switch contains a single queue and switches consisting of 10 exponentially distributed servers that sequentially service each packet. A mean service time of 10 time units is assumed. The servers represent the link rate. The packet is then forwarded with equal probability to another switch, including the originating switch. Each switch is a driving process; the switches forward real and virtual messages. The cumulative number of packets which have entered each switch's queue is the state. This is similar to Simple Network Management Protocol \cite{SNMP} statistics monitored by Simple Network Management Protocol Counters, for example, the {\bf ifInOctets} counter in MIB-II interfaces \cite{RFC1156}. Both real and virtual messages contain the time service ends and a count of the number of times a packet has entered a switch. An initial message enters each queue upon startup to associate a queue with its switch. This is the purpose of the {\bf idmsg} that enters the queues in Figure \ref{qrsim}. The predictive system parameters are more compactly identified as a triple consisting of Lookahead Window Size (seconds), Tolerance (counter value), and Verification Query Period (seconds) in the form $(\Lambda, \Theta, \Upsilon)$. The effect of these parameters are examined on the system of switches previously described. The simulation was run with the following triples: $(5,10,5)$, $(5,10,1)$, $(5,3,5)$, $(400,5,5)$. The graphs that follow show the results for each triple. The first run parameters were $(5, 10, 5)$. There were no state verification rollbacks although there were some causality induced rollbacks as shown in Figure \ref{rb5105}. Global Virtual Time increased almost instantaneously versus real time; at times the next event far exceeded the look-ahead window. This is the reason for the nearly vertical jumps in the Global Virtual Time as a function of real-time as shown in Figure \ref{rb5105}. The state graph for this run is shown in Figure \ref{s5105}. \begin{figure*}[htbp] \centerline{\psfig{file=figures/tg510510.ps,width=5.0in}} \caption{Rollbacks Due to State Verification Failure (5, 10, 5)} \label{rb5105} \end{figure*} \begin{figure*}[htbp] \centerline{\psfig{file=figures/ts510510.ps,width=5.0in}} \caption{State (5, 10, 5)} \label{s5105} \end{figure*} In the initial implementation, state verification was performed in the Logical Process immediately after each new message was received. However, the probability that an Logical Process had saved a future state, while processing at its Local Virtual Time, with the same state save time as the time at which a real message arrived was low. Thus, there was frequently nothing to compare the current state with in order to perform the state verification. However, it was observed that the predictive system was simulating up to the lookahead window very quickly and spending most of its time holding, during which time it was doing nothing. The implementation was modified so that each entity would perform state verification during its hold time\index{VNC!hold time}. This design change better utilized the processors and resulted in more accurate alignment between actual and logical processes. The results for the $(5, 10, 1)$ run were similar, except that the predictive and actual system comparisons were more frequent because the state verification period had been changed from once every 5 seconds to once every second. Error was measured as the difference in the predicted Logical Process state versus the actual system state. This run showed errors that were greater than those in the first run, great enough to cause state verification rollbacks. The error levels for both runs are shown in Figures \ref{e5105} and \ref{e5101}. The state graph for this run is shown in Figure \ref{s5101}. \begin{figure*}[htbp] \centerline{\psfig{file=figures/te510510.ps,width=5.0in}} \caption{Amount of Error (5, 10, 5)} \label{e5105} \end{figure*} \begin{figure*}[htbp] \centerline{\psfig{file=figures/te510110.ps,width=5.0in}} \caption{Amount of Error (5, 10, 1)} \label{e5101} \end{figure*} \begin{figure*}[htbp] \centerline{\psfig{file=figures/ts510110.ps,width=5.0in}} \caption{State (5, 10, 1)} \label{s5101} \end{figure*} The next run used $(5, 3, 5)$ parameters. Here we see many more state verification failure rollbacks as shown in Figure \ref{rb535}. This is expected since the tolerance has been reduced from 10 to 3. The cluster of causality rollbacks near the state verification rollbacks was expected. These clusters of causality rollbacks do not appear to significantly reduce the feasibility of the system. The real-time versus Global Virtual Time plot as shown in Figure \ref{rb535} shows much larger jumps as the Logical Processes were held back due to rollbacks. The entities had a larger variance in their hold times than the $(5, 10, 5)$ run. The state graph for this run is shown in Figure \ref{s535}. \begin{figure*}[htbp] \centerline{\psfig{file=figures/tg53510.ps,width=5.0in}} \caption{Rollbacks Due to State Verification Failure (5, 3, 5)} \label{rb535} \end{figure*} \begin{figure*}[htbp] \centerline{\psfig{file=figures/ts53510.ps,width=5.0in}} \caption{State (5, 3, 5)} \label{s535} \end{figure*} A $(400, 5, 5)$ run showed the Global Virtual Time jump quickly to 400 and then gradually increase as the sliding lookahead window maintained a 400 time unit lead as shown in Figure \ref{rb40055}. The Logical Process hold times were shorter than an any previous run. The state graph for this run is shown in Figure \ref{s40055}. \begin{figure*}[htbp] \centerline{\psfig{file=figures/tg4h5510.ps,width=5.0in}} \caption{Rollbacks Due to State Verification Failure (400, 5, 5)} \label{rb40055} \end{figure*} \begin{figure*}[htbp] \centerline{\psfig{file=figures/ts4h5510.ps,width=5.0in}} \caption{State (400, 5, 5)} \label{s40055} \end{figure*} This set of results is interesting because it shows the system to be stable with the introduction of state verification rollbacks. The overhead introduced by these rollbacks did not greatly impact the performance, because as previously shown in the Global Virtual Time versus time graphs in Figures \ref{rb5105}, \ref{rb535}, and \ref{rb40055}, the system was always able to predict up to its lookahead time very quickly. \section{OPTIMIZING MANAGEMENT POLLING WITH THE PREDICTIVE MANAGER} Since the predictive network management system provides a good approximation of the future behavior of the data to be managed as shown in the Global Virtual Time versus real time values of state in Figures \ref{s5105}, \ref{s5101}, \ref{s535}, and \ref{s40055}, the verification query period can be automatically determined as a function of the look-ahead window and tolerance, with the goal of minimizing the frequency of verification queries thus solving the polling problem in network management. In most standards based approaches, network management stations are sampling counters in managed entities that simply increment in value until they roll over. A management station which is simply plotting data will have some fixed polling interval and record the absolute value of the difference in value of the counter. Such a graph is not a perfectly accurate representation of the data, it is merely a statement that sometime within a polling interval the counter has monotonically increased by some amount. Spikes in this data, which may be very important to the current state of the system, may not be noticed if the polling interval is long enough such that a spike followed by low data values averages out to a normal or low value. One of the goals of a predictive management system is to determine the minimum polling interval required to accurately represent the data. \begin{comment} \newcommand{T}{T} It is clear that the counter, $c(t)$, represents the total area under the graph of the actual data, $d(t)$. Thus $c(t) = \int_{t_1}^{t_2} d(t)$ and \newcommand{d(t) = {dc(t) \over dt}}{d(t) = {dc(t) \over dt}} $d(t) = {dc(t) \over dt}$ where $t_2 - t_1$ is the polling interval and define $T$ as ${t_2 - t_1} \over 2$. Since the polling interval, $t_2 - t_1$, cannot be instantaneous in practice, $c(t)$ is a moving average ${1 \over 2 T} \int_{t_1}^{t_2} d(t)$. The relation of c(t) to the actual data d(t) will depend on the maximum frequency component of d(t) and the polling interval, $T$. The impulse response is a rectangle centered at the origin, and its Fourier Transform is shown in Equation \ref{ft}. \begin{equation} H(\tau) = {1 \over {2 T}} \int^{- T}_{+ T} e^{- j \tau t} dt = {{sin T \tau} \over {T \tau}} \label{ft} \end{equation} \begin{equation} S_{yy} (\omega) = S_{xx} (\omega) {sin^2 {T \omega} \over {T^2 \omega^2}} \label{spectral} \end{equation} Working from the spectral density in Equation \ref{spectral}, the variance can be determined. The autocorrelation, $R ( t_1 , t_2 ) = E [ x ( t_1 ) x ( t_2 ) ]$ of the output (assuming WSS) is a function of the input autocorrelation, $R_{yy} (\theta) = R_{xx} (\theta) * h(\theta) * h^* (- \theta)$. Thus, $R_{yy} (\theta) = {1 \over {2 T}} \int_{- 2 T}^{+ 2 T} ( 1 - {|\alpha| \over {2 T}}) {R_{xx} (\theta - \alpha) d\alpha}$. The variance, $\sigma^2 = \int_{ - \infty}^{ + \infty} (x - \eta)^2 f(x) dx = E [ x^2 ] - E^2 [ x ]$ can be found from the auto-covariance, $C ( t_1, t_2 ) = R ( t_1, t_2 ) - \eta( t_1 ) \eta ( t_2 )$ it is also know that $E [ x^2 (t) ] = R(t,t)$. Thus with $t_1 = t_2$ the variance is \begin{equation} {\sigma_y}^2 = C_{yy} (0) = {1 \over {2 T}} \int_{- 2 T}^{+ 2 T} (1 - {|\alpha| \over {2 T}} ) {C_{xx} (\alpha) d\alpha} \end{equation} also \begin{equation} {\sigma_y}^2 = {1 \over {2 \pi}} \int_{ - \infty}^{+ \infty} S_{xx} (\tau) |H(\tau)|^2 d\tau \end{equation} We would like to see how $\sigma^2$ varies as a function of $\omega$ and T. In order to get rid of the integral on the right consider the Fourier Transform of the auto-covariance, $S^c(\tau)$. Thus \begin{equation} S^c ( \tau ) = (1 - {|\alpha| \over {2 T}} ) C_{xx} ( \alpha ) \end{equation} \end{comment} From the information provided by the predictive management system, a polling interval which provides the desired degree of accuracy can be determined and dynamically adjusted; however, the cost must be determined. An upper limit on the number of systems that can be polled is $N \le {T \over \Delta}$ where $N$ is the number of devices capable of being polled, $T$ is the polling interval, and $\Delta$ is the time required for a single poll. Thus although the data accuracy will be constrained by this upper limit, taking advantage of characteristics of the data to be monitored can help distribute the polling intervals efficiently within this constraint. Assume that $\Delta$ is a calculated and fixed value, as is $N$. Thus this is a lower bound on the value of $T \ge \Delta N$. The overhead\index{Virtual Network Configuration!overhead} bandwidth required for use by the management system to perform polling is shown in Equation \ref{bw}. The packet size will vary depending upon whether it is an SNMP or CMIP packet and the MIB object(s) being polled. The number of packets varies with the amount of management data requested. Let $P$ be the number of packets, $S$ be the bits/packet, $N$ be the number of devices polled, and $T$ be the polling period. $Bw$ is the total available bandwidth and $Bw_{oh}$ is the overhead\index{Virtual Network Configuration!overhead} bandwidth of the management traffic. \begin{figure*} \begin{equation} Bw_{oh}\% = {100.0 P N S Bw \over T} \label{bw} \end{equation} \end{figure*} It may be desirable to limit the bandwidth used for polling system management data to be no more than a certain percentage of total bandwidth. Thus the optimum polling interval will use the least amount of bandwidth while also maintaining the least amount of variance due to error in the data signal. All the required information to maintain the cost versus accuracy at a desired level is provided by the predictive\index{Predictive Network Management!predictive} network management system. \section{CONCLUSION} Network management systems capable not only of passive monitoring but also of active prediction capability are undergoing research and development. Work on prediction mechanisms for mobile communication networks is also underway. A method used by standards-based network management systems to cope with these two developments has been proposed in this paper. A predictive network management algorithm and the characteristics of a predictive network management system have been presented. The predictive capability of the network management system is used to solve the polling rate problem for network management. The Rapidly Deployable Radio Network \cite{BushThesis} is presented as an example of a predictive mobile communications network. Finally, interaction between the predictive capabilities of the network management and mobile network systems has been discussed.
\section{Introduction} Ashtekar's formalism of general relativity[1] has led to a considerable progress in loop quantum gravity[2]. A special feature of this framework is that degenerate triads, and hence degenerate metrics, are admitted, and the degenerate metrics play an important role in the quantum description of gravity[3,4]. The significance of understanding degenerate metrics was emphasized in Refs.5 and 6. Various kinds of degenerate solutions to classical Ashtekar's equations have been studied[5-11], and the local causal structure of degenerate Ashtekar theory has also been established[12]. Using a ``covariant approach'', Bengtsson and Jacobson[6] investigated the structure of the ``phase boundaries'' between degenerate and nondegenerate space-time regions, and conjectured that the phase boundaries should always be null provided that the metric is a ``regular'' solution to Ashtekar's equations, that is, solutions in which the canonical variables $(A_i^I,% \tilde{e}_I^i)$ , the shift vector $N^i$, and the lapse density, $\underline{% N}$ (weight $-1$), all take finite value which, except for $\underline{N}$, are allowed to vanish. In a recent paper[13], however, a degenerate phase boundary is distinguished from its image, and moreover, it is shown that the definition of the nullness of the image of the phase boundary used in Ref.6 could not be generalized to the phase boundary itself. The main focus of the present paper, on the other hand, is first to give a reasonable definition of the nullness of the boundary, and then to prove the conjecture under certain circumstances. The suggestion of Ref.6 to create a space-time with a degenerate region by the covariant approach is as follows. Start off with a nondegenerate metric which solves Einstein's equations and reparametrize one of the coordinates. This reparametrization is chosen so that it is not a diffeomorphism at some particular value of the coordinate. Adopting the new coordinate, the solution can be smoothly matched to a solution to the Ashtekar equations with a degenerate metric at the surface where the transformation misbehaves. To make things clearer we reformulate this procedure as follows. Let $M$ be a 4-dimensional manifold and $M_1$ a 4-dimensional submanifold with a 3-dimensional boundary $\partial M_1$. Suppose $\hat{M}$ is a 4-dimensional manifold with a nondegenerate metric $\hat{g}_{\mu \nu }$ which solves the Einstein's equations, and $\phi $ is a diffeomorphism from $M_1$ to some open set $\hat{M}_1\subset \hat{M}$. Extend the domain of $\phi $ smoothly to the whole of $M$ so that $M-M_1$ is mapped onto $\phi [\partial M_1]$, and the pushforward $\phi _{*}$ restricted to the tangent bundle of $% \partial $$M_1$ to that of $\phi [\partial M_1]$ is nondegenerate. Then the pullback $g_{\mu \nu }\equiv \phi ^{*}\hat{g}_{\mu \nu }$ is nondegenerate on $M_1$ and degenerate on $M-M_1$. One therefore has a space-time $% (M,g_{\mu \nu })$ with a ``phase boundary'', $\partial M_1$, separating a nondegenerate region from a degenerate one. It is clear that the ``reparametrization procedure'' mentioned above is a special case of this treatment. Inspired by Ref.12, we try to define timelike and null hypersurfaces in a degenerate space-time in Sec.2. Armed with this new definition for a null hypersurface, we then give a proof of the conjecture that the phase boundary is null in Sec.3 under the circumstances where the degenerate space-time with a phase boundary is obtained by the reparametrization procedure mentioned above. \section{ Defining null hypersurfaces in degenerate space-times} Suppose the boundary $\phi [\partial M_1]$ is given by $f=0$, where $f$ is a smooth function on $\hat{M}$ with $\nabla _\mu f|_{\phi [\partial M_1]}\neq 0 $ , then $\phi [\partial M_1]$ is defined in Ref.[6] to be null if $g^{\mu \nu }\nabla _\mu f\nabla _\nu f\rightarrow 0$ as $\phi [\partial M_1]$ ``is approached from the nondegenerate side''. However, as pointed out in Ref.[13], this definition is inappropriate to $\partial M_1$ since it depends upon the choice of the function $f$ on $M$, and concrete examples show that there exist functions $f$ and $\bar{f}$ with $\lim g^{\mu \nu }\nabla _\mu f\nabla _\nu f=0$ while $\lim g^{\mu \nu }\nabla _\mu \bar{f}% \nabla _\nu \bar{f}\neq 0$. This obstacle could be overcome if we use $\sqrt{% -g}g^{\mu \nu }\nabla _\mu f\nabla _\nu f$ instead of $g^{\mu \nu }\nabla _\mu f\nabla _\nu f$. In Ashtekar theory there is a well-defined densitized inverse metric, $\tilde{g}^{\mu \nu }$, with components in any coordinate system of a $3+1$ decomposition[12]: \begin{equation} \tilde{g}^{\mu \nu }=\left( \begin{array}{cc} \tilde{g}^{tt} & \tilde{g}^{ti} \\ \tilde{g}^{jt} & \tilde{g}^{ji} \end{array} \right) =\left( \begin{array}{cc} -\underline{N}^{-1} & \underline{N}^{-1}N^i \\ \underline{N}^{-1}N^j & \ \underline{N}\tilde{h}^{ji}-\underline{N}% ^{-1}N^jN^i \end{array} \right) , \end{equation} where $\underline{N}$ and $N^i$ are respectively the lapse density and the shift vector, and $\tilde{h}^{ji}$ is the densitized inverse 3-metric of weight $+2$, and in the nondegenerate case one has $\tilde{g}^{\mu \nu }=% \sqrt{-g}g^{\mu \nu }$. Eq.(1) implies that $\tilde{g}^{\mu \nu }$ remains finite in the Ashtekar theory of degenerate space-times, and therefore $\lim \sqrt{-g}g^{\mu \nu }\nabla _\mu f\nabla _\nu f=\tilde{g}^{\mu \nu }\nabla _\mu f\nabla _\nu f|_{f=0}.$ Let $f$ and $\bar{f}$ be two distinct functions on $M$ with $f|_{\partial M_1}=\bar{f}|_{\partial M_1}=0$ and $\nabla _\mu f|_{f=0}\neq 0$ and $\nabla _\mu \bar{f}|_{\bar{f}=0}\neq 0$, then there exists a function $\lambda $ on $M$ such that $\nabla _\mu f|_{f=0}=\lambda \nabla _\mu \bar{f}|_{\bar{f}=0}$, and hence \[ \tilde{g}^{\mu \nu }\nabla _\mu f\nabla _\nu f|_{f=0}=\lambda ^2\tilde{g}% ^{\mu \nu }\nabla _\mu \bar{f}\nabla _\nu \bar{f}|_{\bar{f}=0}. \] Hence, $\tilde{g}^{\mu \nu }\nabla _\mu f\nabla _\nu f|_{f=0}=0$ if and only if $\tilde{g}^{\mu \nu }\nabla _\mu \bar{f}\nabla _\nu \bar{f}|_{\bar{f}% =0}=0 $. We therefore obtain a self-consistent definition of null hypersurfaces in a degenerate space-time in Ashtekar's theory as follows: {\bf Definition }${\bf 1}${\bf :} A hypersurface described by $f=0$ with $% \nabla _\mu f|_{f=0}\neq 0$ in space-time $(M,g_{\mu \nu })$ is said to be null if $\tilde{g}^{\mu \nu }\nabla _\mu f\nabla _\nu f|_{f=0}=0$. In the following we will use the symbol, $\tilde{E}_A^\mu $, to denote the vierbein of vector densities weighted $+1/2$, i.e., the square roots of $% \tilde{g}^{\mu \nu }$, namely, \[ \tilde{g}^{\mu \nu }=\eta ^{AB}\tilde{E}_A^\mu \tilde{E}_B^\nu , \] where $\eta ^{AB}$ is the Minkowski metric to raise (and $\eta _{AB}$ to lower) the interior indices ``$A$'' and ``$B$''. Note that there is $SO(3,1)$ gauge freedom for $\tilde{E}_A^\mu $, and the components of certain choice of $\tilde{E}_A^\mu $ in any coordinate system associated with a $3+1$ decomposition are \begin{equation} \tilde{E}_A^\mu =\left( \begin{array}{cc} \tilde{E}_0^t & \tilde{E}_0^i \\ \tilde{E}_I^t & \tilde{E}_I^i \end{array} \right) =\left( \begin{array}{cc} \sqrt{\underline{N}^{-1}} & \ -\sqrt{\underline{N}^{-1}}N^i \\ 0 & -\sqrt{\underline{N}}\tilde{e}_I^i \end{array} \right) , \end{equation} where $\tilde{e}_I^i$ is the densitized triad of weight $+1$ in Ashtekar theory (with ``$i$'' and ``$I$'' the spatial and interior indices respectively), and the columns of $(\tilde{E}_A^\mu )$ are labelled by space-time indices. Given a vierbein, one may consider $\tilde{E}_A^\mu (x):% {\rm M}^4\rightarrow T_xM$ as a map from the 4-dimensional Minkowski space into the tangent bundle of the manifold $M$. The ``future'', ${\cal F}(x)$, of a point, $x\in M$, can therefore be defined[12] as the set of all tangent vectors at $x$ which are images of some vectors, $\varsigma ^A$, in ${\rm M}% ^4$ satisfying $\varsigma ^A\varsigma _A\leq 0$ and $\varsigma ^0>0$ , i.e., \[ {\cal F}(x)\equiv \{v^\mu (x)\in T_xM\ |\ \exists \ \varsigma ^A\in {\rm M}% ^4,\varsigma ^A\varsigma _A\leq 0,\varsigma ^0>0,\ such\ that\ v^\mu (x)=\varsigma ^A\tilde{E}_A^\mu (x)\}. \] Thus, depending on the rank of the vierbein, the future, ${\cal F}(x)$, is either a (4-dimensional) hypercone ($rank\ \tilde{E}_A^\mu =4$), a (3-dimensional) cone ($rank\ \tilde{E}_A^\mu =3$), an angle ($rank\ \tilde{E}% _A^\mu =2$), or a half-line ($rank\ \tilde{E}_A^\mu =1$) in Ashtekar theory. This local causal structure can be used to define the timelike and null hypersurfaces as follows. {\bf Definition }${\bf 2}${\bf :} A hypersurface $\Sigma $ is said to be timelike if for any point $x\in \Sigma $ the tangent space, $T_x\Sigma $ (tangent to $\Sigma $), of $x$ contains a nonzero vector, $v^\mu $, which is the image under the mapping $\tilde{E}_A^\mu $ of a timelike vector, $% \varsigma ^A$, in the Minkowski space, i.e., $\exists \ v^\mu =\varsigma ^A% \tilde{E}_A^\mu \in T_x\Sigma $ such that $\eta _{AB}\varsigma ^A\varsigma ^B<0$. {\bf Definition }${\bf 3}${\bf : }A hypersurface $\Sigma $ is said to be null if for any point $x\in \Sigma $ the tangent space, $T_x\Sigma ,$ of $x$ contains a nonzero vector that is the image of a null vector in the Minkowski space, and there exists no timelike vector, $\varsigma ^A$, in the Minkowski space such that $\varsigma ^A\tilde{E}_A^\mu \in T_x\Sigma $. Definitions $2$ and $3$ are consistent with the causal structure and can be re-formulated in terms of ${\cal F}(x)$ as follows: Let $i({\cal F}(x))$ and $\partial {\cal F}(x)$ represent respectively the interior and the boundary of ${\cal F}(x)$, then a hypersurface $\Sigma $ is timelike if and only if $% T_x\Sigma \cap i({\cal F}(x))\neq \emptyset $, while $\Sigma $ is null if and only if $T_x\Sigma \cap {\cal F}(x)\neq \emptyset $ and $T_x\Sigma \cap {\cal F}(x)\subset \partial {\cal F}(x)$. Note that the definition of a spacelike hypersurface $\Sigma $ given by Ref.12 is equivalent to requiring $% T_x\Sigma \cap {\cal F}(x)=\emptyset $. Note also that both Definitions $2$ and $3$ are applicable to the cases where the ranks of $\tilde{E}_A^\mu $ are two, three, and four. In the case where $\tilde{E}_A^\mu $ is of rank one, the timelike and null hypersurfaces become indistinguishable. Now it is natural to ask whether Definition $3$ is equivalent to Definition $% 1$. Suppose a hypersurface defined by $f=0$ with $\nabla _\mu f|_{f=0}\neq 0$ is null according to Definition $3$, then any vector field, $v^\mu $, tangent to the hypersurface satisfies \begin{equation} 0=v^\mu \nabla _\mu f=\varsigma ^A\tilde{E}_A^\mu \nabla _\mu f=\varsigma ^A\omega _A, \end{equation} where $\varsigma ^A$ is any inverse image of $v^\mu $ under $\tilde{E}_A^\mu $, and $\omega _A\equiv \tilde{E}_A^\mu \nabla _\mu f$. Since $\varsigma ^A$ can be null but not timelike, it follows from Eq.(3) that $\omega ^B\equiv \eta ^{BA}\omega _A$ must be a null vector. Consequently on $f=0$ we have \[ \tilde{g}^{\mu \nu }\nabla _\mu f\nabla _\nu f=\eta ^{AB}\tilde{E}_A^\mu \tilde{E}_B^\nu \nabla _\mu f\nabla _\nu f=\eta ^{AB}\omega _A\omega _B=0, \] i.e., the hypersurface $f=0$ is also null according to Definition $1$. However, the degeneracy of $\tilde{E}_A^\mu $ implies the possibility of $% \omega _A\equiv \tilde{E}_A^\mu \nabla _\mu f=0$ , in this case the hypersurface is null according to Definition $1$ while might well be nonnull according to Definition $3$. The above arguments lead to the following equivalent definition of Definition $3$: {\bf Definition }${\bf 3}^{\prime }${\bf :} A hypersurface $f=0$ (with $% \nabla _\mu f|_{f=0}\neq 0$) is null if $\omega _A\equiv \tilde{E}_A^\mu \nabla _\mu f|_{f=0}$ is a nonzero null covector in the Minkowski space. Since Definition $3^{\prime }$ is consistent with the local causal structure and convenient to use, we will use it to judge whether the phase boundary $% \partial M_1$ is null in the next section. It should be noted that the choice of the gauge as well as the coordinate system for $\tilde{E}_A^\mu $ is irrelevant. The interior gauge transformation preserves the Minkowski metric and hence does no harm to the previous discussions. Since $\tilde{E}_A^\mu $ are vector densities, for a vector $\varsigma ^A$ in Minkowski space, the image $v^\mu (x)\equiv \tilde{E% }_A^\mu \varsigma ^A$ , viewed as a vector at $x\in M$, will change under a coordinate transformation to $v^{\prime \mu }(x)\equiv \tilde{E}_A^{\prime \mu }\varsigma ^A$. However, the transformation law for the components of a vector density guarantees that $v^{\prime \mu }(x)$ and $v^\mu (x)$ have the same direction, therefore coordinate transformations do no harm to the previous discussions either. \section{ Nullness of the degenerate phase boundary $\partial M_1$} We assume in this section that the degenerate phase boundary is obtained through the covariant approach mentioned in Sec.1. As shown in Ref.13, the hypersurface $\phi [\partial M_1]$ in $\hat{M}$ must be null if the pullback metric $g_{\mu \nu }$ on $M$ is to be a regular solution to Ashtekar's equations. The argument is as follows in short. In any $3+1$ decomposition of the space-time $(M,g_{\mu \nu })$ one has $h_{ij}N^j=g_{0i}\ (i=1,2,3)$, where $h_{ij}$ and $N^j$ are respectively the 3-metric and the shift vector. Since $h\equiv \det (h_{ij})=0$ in $M-M_1$, the last three columns of $% (g_{\mu \nu })$ must be linearly dependent to ensure the finiteness of $N^i$% . Hence there exists a 4-vector $T^\nu =(0,\lambda ^i)$ at each point of $% M-M_1$ with $\lambda ^i$ a non-vanishing 3-vector such that $g_{\mu \nu }T^\nu =0$. Furthermore, the lapse scalar $N$ must vanish to keep the lapse density $\underline{N}$ finite in $M-M_1$, it then follows from $% g_{00}=-N^2+g_{0i}N^i$ that there exists another 4-vector $S^\nu =(1,-N^i)$ at each point of $M-M_1$ such that $g_{\mu \nu }S^\nu =0$. Therefore $T^\nu $ and $S^\nu $ represent two independent degenerate directions of $g_{\mu \nu } $, and hence there must be some degenerate vector field, $W^\nu $, that is tangent to $\partial M_1$. It follows from the nondegeneracy of the pushforward $\phi _{*}$ (restricted to $\partial M_1$ ) that there is a vector field, $\phi _{*}W^\nu $, on $\phi [\partial M_1]$ that is the desired null normal to $\phi [\partial M_1]$. It is also clear from this argument that the rank of $g_{\mu \nu }$ is two on $\partial M_1$, from which one can argue that $rank(\tilde{g}^{\mu \nu })=2$, and hence $rank(% \tilde{E}_A^\mu )$ must be two or three on $\partial M_1$. We could therefore use Definition $3^{\prime }$ in Sec.2 to judge whether $\partial M_1$ is null. Without loss of generality, we choose a ``time orthogonal'' $3+1$ decomposition of the space-time $(\hat{M},\hat{g}_{\mu \nu })$, and the line element reads \begin{equation} d\hat{s}^2=-\hat{N}^2dT^2+\hat{h}_{ij}dX^idX^j. \end{equation} Let $U$ be a smooth function on $M$ with $\nabla _\mu U\neq 0$ and $U=0$ represents $\phi [\partial M_1]$, then \begin{equation} dU=\hat{\beta}(T,X^j)dT+\hat{\alpha}_i(T,X^j)dX^i, \end{equation} where $\hat{\beta}\equiv \partial U/\partial T$ and $\hat{\alpha}_i\equiv \partial U/\partial X^i,i=1,2,3$. It follows from Eqs.(4), (5) and the nullness of $\phi [\partial M_1]$, i.e., $\hat{g}^{\mu \nu }\nabla _\mu U\nabla _\nu U|_{U=0}=0$ that \begin{equation} \lbrack \hat{h}^{ij}\hat{\alpha}_i\hat{\alpha}_j-(\hat{\beta}/\hat{N}% )^2]|_{U=0}=0, \end{equation} where $\hat{h}^{ij}$ is the inverse of the 3-metric $\hat{h}_{ij}$, and $% \hat{N}$ is the lapse scalar. The line element (4) in the domain of $U$ can be re-expressed as \begin{equation} d\hat{s}^2=(\hat{N}/\hat{\beta})^2[-dU^2+2\hat{\alpha}_idX^idU-(\hat{\alpha}% _idX^i)^2]+\hat{h}_{ij}dX^idX^j. \end{equation} The mapping $\phi :M\rightarrow \hat{M}$ induces four functions $\phi ^{*}U,\phi ^{*}X^i(i=1,2,3)$ on $M$ with $\phi ^{*}U|_{M-M_1}=0$. Without essential loss of generality, let $(u,x^i)$ be a local coordinate system on $% M$ covering a neighborhood of $\partial M_1$ with $u|_{\partial M_1}=0,u|_{M_1}>0$ [hereafter $M_1$ (or $M$) is short for ``the interaction of $M_1$ (or $M$) and the coordinate patch''] and $x^i=\phi ^{*}X^i$, then one has a function $U(u)$ [short for $(\phi ^{*}U)(u)$ ] with $U^{\prime }(u)|_{M-M_1}\equiv [dU/du]_{M-M_1}=0$. It then follows from Eq.(7) that the line element of $g_{\mu \nu }\equiv \phi ^{*}\hat{g}_{\mu \nu }$ in this coordinate system reads \begin{equation} ds^2=(\hat{N}/\hat{\beta})^2[-(U^{\prime })^2du^2+2U^{\prime }\hat{\alpha}% _idx^idu-(\hat{\alpha}_idx^i)^2]+\hat{h}_{ij}dx^idx^j. \end{equation} Let $u=u(t,x^i)$, where $t$ is the time coordinate of certain $3+1$ decomposition of $(M,g_{\mu \nu })$, and \[ du=\beta (t,x^i)dt+\alpha _i(t,x^i)dx^i, \] where $\beta \equiv \partial u/\partial t$ and $\alpha _i\equiv \partial u/\partial x^i,i=1,2,3$, then the $3+1$ decomposition of the metric (8) reads \begin{equation} ds^2=-(\hat{N}/\hat{\beta})^2[(U^{\prime }\beta )^2dt^2+2U^{\prime }\beta \gamma _idx^idt+(\gamma _idx^i)^2]+\hat{h}_{ij}dx^idx^j, \end{equation} where $\gamma _i\equiv U^{\prime }\alpha _i-\hat{\alpha}_i$. The determinant, $g$, of the line element (9), the spatial 3-metric, $h_{ij}$, induced by metric (9), and the determinant, $h$ , of $h_{ij}$ can be obtained through straightforward calculations as \begin{equation} g\equiv \det \left( g_{\mu \nu }\right) =-(U^{\prime }\beta \hat{N}/\hat{% \beta})^2\hat{h},\quad \hat{h}\equiv \det \left( \hat{h}_{ij}\right) , \end{equation} \begin{equation} h_{ij}=\hat{h}_{ij}-(\hat{N}/\hat{\beta})^2\gamma _i\gamma _j, \end{equation} \begin{equation} h\equiv \det \left( h_{ij}\right) =\hat{h}[1-(\hat{N}/\hat{\beta})^2\gamma _i\gamma _j\hat{h}^{ij}]. \end{equation} Since $g=-N^2h$, where $N$ is the lapse scalar, it follows from Eqs.(10) and (12) that the Ashtekar's lapse density on $M_1$ is \begin{equation} \underline{N}=\frac N{\sqrt{h}}=\frac{\sqrt{-g}}h=\frac{\left| \beta \hat{% \beta}U^{\prime }\right| }{\hat{N}\sqrt{\hat{h}}[(\hat{\beta}/\hat{N}% )^2-\gamma _i\gamma _j\hat{h}^{ij}]}. \end{equation} Since the shift vector, $N^i$, relates to the metric components of Eq.(9) via \[ h_{ij}N^j=g_{oi}=-(\hat{N}/\hat{\beta})^2\beta U^{\prime }\gamma _i, \] a straightforward calculation shows that the shift vector on $M_1$ reads \begin{equation} N^i=g_{0i}h^{ij}=\frac{U^{\prime }\beta \gamma _j\hat{h}^{ij}}{\hat{h}% ^{lm}\gamma _l\gamma _m-(\hat{\beta}/\hat{N})^2}. \end{equation} Using Eq.(6), it is not difficult to show that Eqs.(13) and (14) imply \begin{equation} \underline{N}|_{\partial M_1}=\lim_{u\rightarrow 0^{+}}\frac{\left| \beta \hat{\beta}U^{\prime }\right| }{\hat{N}\sqrt{\hat{h}}[(\hat{\beta}/\hat{N}% )^2-\gamma _i\gamma _j\hat{h}^{ij}]}=\frac{\left| \beta \hat{\beta}\right| }{% \hat{N}\sqrt{\hat{h}}\left| 2\hat{\alpha}_i\alpha _j\hat{h}^{ij}+B\right| }, \end{equation} and \begin{equation} N^i|_{\partial M_1}=\lim_{u\rightarrow 0^{+}}\frac{U^{\prime }\beta \gamma _j% \hat{h}^{ij}}{\hat{h}^{lm}\gamma _l\gamma _m-(\hat{\beta}/\hat{N})^2}=\frac{% \beta \hat{\alpha}_j\hat{h}^{ij}}{2\hat{h}^{lm}\hat{\alpha}_l\alpha _m+B}, \end{equation} where \begin{equation} B\equiv \lim_{u\rightarrow 0^{+}}\frac 1{U^{\prime }}[(\hat{\beta}/\hat{N}% )^2-\hat{h}^{ij}\hat{\alpha}_i\hat{\alpha}_j]. \end{equation} If we choose the vierbein, $\tilde{E}_A^\mu $, as Eq.(2), then, according to Definition $3^{\prime }$, the key quantity for judging whether the hypersurface $u=const.$ is null is \begin{equation} u_A\equiv \tilde{E}_A^\mu (du)_\mu =\left( \begin{array}{c} \sqrt{\underline{N}^{-1}}(\beta -\alpha _iN^i) \\ -\sqrt{\underline{N}}\alpha _i\tilde{e}_I^i \end{array} \right) \equiv \left( \begin{array}{c} u_0 \\ u_I \end{array} \right) ,\quad I=1,2,3, \end{equation} where \[ (du)_\mu =\left( \begin{array}{c} \beta \\ \alpha _i \end{array} \right) ,\quad i=1,2,3. \] It follows from Eqs.(15), (16), and (18) that \begin{equation} (u_0)^2|_{\partial M_1}=\left| \frac \beta {\hat{\beta}}\right| \frac{\hat{N}% \sqrt{\hat{h}}(\hat{\alpha}_i\alpha _j\hat{h}^{ij}+B)^2}{\left| 2\hat{h}^{lm}% \hat{\alpha}_l\alpha _m+B\right| }, \end{equation} and \begin{equation} (u_1)^2+(u_2)^2+(u_3)^2=\underline{N}\alpha _i\alpha _j\tilde{e}_I^i\tilde{e}% ^{Ii}=\underline{N}\alpha _i\alpha _j\tilde{h}^{ij}. \end{equation} Through a non-trivial calculation, which will be given in the Appendix, we get \begin{equation} \alpha _i\alpha _j\tilde{h}^{ij}|_{\partial M_1}=\hat{h}(\hat{N}/\hat{\beta}% )^2(\hat{\alpha}_i\alpha _j\hat{h}^{ij})^2. \end{equation} Hence Eq.(20) evaluated on $\partial M_1$ gives \begin{equation} \lbrack (u_1)^2+(u_2)^2+(u_3)^2]|_{\partial M_1}=\left| \frac \beta {\hat{% \beta}}\right| \frac{\hat{N}\sqrt{\hat{h}}(\hat{\alpha}_i\alpha _j\hat{h}% ^{ij})^2}{\left| 2\hat{h}^{lm}\hat{\alpha}_l\alpha _m+B\right| }, \end{equation} which equals Eq.(19) if and only if $B=0$. To show that this condition is indeed satisfied we first obtain from Eq.(17) the following expression of $B$% : \begin{equation} B=\lim_{u\rightarrow 0^{+}}[\frac{\partial (\hat{\beta}^2/\hat{N}^2-\hat{h}% ^{ij}\hat{\alpha}_i\hat{\alpha}_j)}{\partial U}\frac{U^{\prime }}{U^{\prime \prime }}], \end{equation} where $U^{\prime \prime }\equiv dU^{\prime }/du$. If \begin{equation} b\equiv \lim_{u\rightarrow 0^{+}}\frac{U^{\prime }}{U^{\prime \prime }}\neq 0, \end{equation} there exists a smooth function $L(u)$ on $M$ such that $U^{\prime \prime }/U^{\prime }=L(u)$ on $M_1$, and $\lim_{u\rightarrow 0^{+}}(U^{\prime \prime }/U^{\prime })=L(0)=1/b$. Therefore one has \begin{equation} U^{\prime }=C\exp [\int_0^uL(\tau )d\tau ]\quad on\ M_1 \end{equation} where $C=const.$ . Since $\lim_{u\rightarrow 0^{+}}U^{\prime }=0$, Eq.(25) implies that $C=0$ , and hence $U^{\prime }|_{M_1}=0$, contradicting the above mentioned requirement of the mapping $\phi $. Therefore the assumption (24) is false, and we have \[ \lim_{u\rightarrow 0^{+}}\frac{U^{\prime }}{U^{\prime \prime }}=0, \] and hence it follows from Eq.(23) that $B=0$ since $\partial (\hat{\beta}^2/% \hat{N}^2-\hat{h}^{ij}\hat{\alpha}_i\hat{\alpha}_j)/\partial U$ is regular on $M$. Consequently the vector $u^A|_{\partial M_1}$ is null in the Minkowski space. Furthermore, since $\underline{N}$ should be finite for a regular solution to Ashtekar's equations, Eq.(15) implies that Eq.(22) does not vanish, ensuring that $u^A|_{\partial M_1}\neq 0$. We therefore conclude that the phase boundary, $\partial M_1$, represented by $u=0$, is null. It is also worth noting that, in contrast with what the authors of Ref.6 argued, (This is originally, in Ref.6, referred to the image rather than the phase boundary itself.) the Ashtekar's evolution equations are not at all a necessary condition for the degenerate phase boundary $\partial M_1$ to be null. \acknowledgments The authors would like to thank Prof. Zhiquan Kuang for his enlightening suggestions and valuable comments. One of the authors, Y. Ma, also would like to thank Prof. Ted Jacobson for helpfull discussions. This work has been supported by the National Science Foundation of China.
\section{Introduction} The understanding of the most general coupling of vector multiplets in $N=2$ supersymmetry or supergravity is important in very different contexts. It is used in the context of the low-energy effective actions of supersymmetric gauge theories and their coupling to hypermultiplets \cite{SW}. It is also an essential element in the compactification of string theory on Calabi--Yau manifolds \cite{CY}. This general coupling has been studied for the supergravity case in \cite{dWLPQVP,dWVP} and has been given the name special geometry \cite{strom}. The similar coupling in rigid supersymmetry was obtained in \cite{STG} and is referred to as `rigid special geometry'. \par Historically, the coupling of several $N=2$ matter multiplets to $N=2$ supergravity in four dimensions was found using superconformal tensor calculus \cite{dWVHVP, dRvHdWVP, dWLPQVP, dWVP, dWLVP}. Superconformal actions are built with representations of a larger algebra, the $N=2$ superconformal one. The residual symmetries are broken by introducing two compensating multiplets (one vector multiplet to give rise to the graviphoton in the Poincar\'e gravity multiplet and a hyper-, a linear or a nonlinear multiplet) to end up with an $N=2$ Poincar\'e supergravity theory coupled to $N=2$ matter multiplets. Reviews on superconformal tensor calculus can be found in \cite{sctc}. The superconformal construction clarifies the origin of many terms in the action. Here we will confine ourselves to the coupling of vector multiplets to supergravity, where after breaking the superconformal symmetry, the complex scalars of the vector multiplets form a special K\"ahler\ manifold. \par A prepotential, a holomorphic function of second order, was an essential ingredient to construct the theory. In \cite{CDFF,FL,ABCDFF}, other approaches were used to describe the coupling of vector multiplets to supergravity. \par Electric--magnetic duality transformations in four dimensions manifest themselves by symplectic transformations \cite{dual}. Symplectic transformations in a special K\"ahler\ manifold have been studied in \cite{dWVP,CecFerGir}. The duality symmetry is not a symmetry of the complete action, but only of the equations of motion. In particular, the prepotential is not an invariant of the symplectic transformations. On the other hand, for a coordinate-free formulation of special geometry \cite{CDFF}, the symplectic symmetry is an essential ingredient. The symplectic set-up also clarified the link to Calabi--Yau manifolds \cite{CY}. In \cite{CDFVP}, vector multiplet couplings to supergravity were constructed for which no prepotential existed, by performing a symplectic transformation of an action based on a prepotential. The resulting action was thus not based on a prepotential. The first and main purpose of this paper is to obtain a symplectic covariant formulation of the coupling of vector multiplets to $N=2$ supergravity which at the same time uses superconformal tensor calculus. In particular, it should thus contain the coupling of \cite{CDFVP}. For obtaining an action in superconformal tensor calculus, one needs a prepotential, and hence to give up the symplectic covariance. The combination of superconformal and symplectic covariance will, however, be possible if we only construct equations of motions without an action. \par The various possible actions and geometric formulations were compared in \cite{CRTVP}, and one has arrived at a new definition of special geometry (also the definition for the case of rigid supersymmetry is given there). Remarkably, it was also noticed that one part of the definition, expressed by differential constraints, can be formulated in two different ways. These two forms are equivalent when more than one vector multiplet is coupled to supergravity, but inequivalent if only one vector multiplet is coupled. The presentation in \cite{CRTVP} contained the constraints such that for one vector multiplet the coupling known previously (e.g.\ from superconformal tensor calculus) is obtained. However, it was noted that another form of the constraints is possible, which is also symplectic covariant. Obvious physical arguments could not exclude the existence of hitherto unknown couplings of 1 vector multiplet to supergravity which obey the weaker constraint, and not the stronger one. The second aim of this paper is to show that such new couplings are indeed possible. \par To be more explicit, let us repeat here one of the formulations of the 3 equivalent definitions of a special K\"ahler\ manifold of \cite{CRTVP}\footnote{The first one is not explicitly symplectic covariant, but we could as well have discussed here definition~2, where the constraint relevant for the discussion below was formulated as $\langle v , \partial_{\alpha} v \rangle =0$. The alternative form is then $\langle \partial_\alpha v , \partial_\beta v \rangle =0$.}. The most suitable definition for our discussion is formulated in terms of symplectic products. It is the one which was denoted as definition~3. \vspace{2mm} \par \noindent{\it Definition of a special K\"ahler\ manifold}\vspace{2mm} \par Take a complex manifold ${\cal M}$. Suppose we have in every chart a $2(n+1)$-component vector $V(z^{\alpha},\bar z^{\alpha})$ such that on overlap regions there are transition functions of the form \begin{equation} e^{\frac{1}{2}(f(z^{\alpha}) -\bar f(\bar z^{\alpha}))}\,S\,, \end{equation} with $f$ a holomorphic function and $S$ a constant $Sp(2(n+1),\Rbar)$ matrix. (These transition functions have to satisfy the cocycle condition.) Take a $U(1)$ connection of the form $\kappa_{\alpha}\,dz^{\alpha}+ \kappa_{\bar \alpha}\,d\bar z^{\alpha}$ with \begin{equation} \kappa_{\bar\alpha}= -\overline{\kappa_{\alpha}}\,, \label{defkappabar} \end{equation} under which $\bar V$ has opposite weight as $V$. Denote the covariant derivative by ${\cal D}$: \begin{equation} \begin{array}{ll} U_\alpha\equiv {\cal D}_{\alpha} V\equiv \partial_{\alpha} V+\kappa_{\alpha} V \,, \qquad& {\cal D}_{\bar \alpha}V\equiv \partial_{\bar\alpha} V+\kappa_{\bar\alpha} V\,,\\ \bar U_{\bar \alpha}\equiv {\cal D}_{\bar\alpha}\bar V\equiv \partial_{\bar\alpha}\bar V-\kappa_{\bar\alpha}\bar V \,,\qquad& {\cal D}_{\alpha}\bar V\equiv \partial_{\alpha}\bar V-\kappa_{\alpha}\bar V \,. \end{array} \label{defD} \end{equation} We impose the following conditions: \begin{eqnarray} &1.& \langle V,\bar V\rangle ={\rm i}\,, \label{VVbar} \\ &2.& {\cal D}_{\bar \alpha}V= 0\,, \label{DbarV} \\ &3.& {\cal D}_{[\alpha}U_{\beta]}=0\,, \label{DU=0}\\ &4.& \langle V,U_{\alpha}\rangle =0 \,, \label{VUal} \end{eqnarray} where $\langle \cdot ,\cdot \rangle$ denotes the symplectic inner product, e.g.\ $\langle V,\bar V\rangle= V^T\Omega\bar V$, with an antisymmetric matrix $\Omega$, which has as standard form \begin{equation} \Omega_{st}=\pmatrix{0&\unity \cr -\unity &0}\,.\label{standardOmega} \end{equation} Define \begin{equation} \label{defg} g_{\alpha\bar\beta}\equiv {\rm i}\langle U_{\alpha},\bar U_{\bar\beta}\rangle \,, \end{equation} where $\bar U_{\bar\beta}$ denotes the complex conjugate of $U_{\alpha}$. If this is a positive-definite metric, ${\cal M}$ is called a special K\"ahler\ manifold. \par It can then be shown that locally a function $K'$ exists such that \begin{equation} \label{kappa} \kappa_{\alpha}=\ft{1}{2}\partial_\alpha K' \,, \qquad \kappa_{\bar\alpha}=-\overline{\kappa_{\alpha}}= -\ft{1}{2}\partial_{\bar\alpha}\bar K'\,. \end{equation} The real part of $K'$ is the K\"ahler\ potential $K$. If there is an imaginary part ${\rm Im ~} K'$, then \begin{equation} V'=e^{{\rm i}{\rm Im ~} K'/2}V \,,\label{V'V} \end{equation} satisfies the same constraints with $K'$ replaced by the real $K$. \par As discussed at the end of section~4.2.2 in \cite{CRTVP}, the constraints have a clear physical interpretation, related to the positivity of kinetic terms in the action. However, as suggested there already, the fourth constraint \eqn{VUal} could be replaced by \begin{equation} 4'.\ \ \langle U_\alpha, U_\beta\rangle =0 \,, \label{UU} \end{equation} without violating the physical arguments. The constraint $4$ implies $4'$, by taking a covariant derivative and antisymmetrizing, and with $4'$ it was shown that $4$ follows when $n>1$. However, for $n=1$, equation $4'$ is empty. Taking $4'$ as constraint thus allows $\langle V, U_z\rangle \neq 0$. Such $N=2$ models would be new, and this possibility will be investigated in this paper. It will be called `the special case'. The case with $n>1$ or $\langle V, U_\alpha \rangle =0$ will be called `the generic case'. \par In appendix~C of \cite{CRTVP}, two $n=1$ examples are given where the condition \eqn{VUal} is not fulfilled. In these examples it was shown that the relaxation of that constraint leads to models not allowed by other definitions of special geometry. Here we will first give further evidence of the non-triviality of this condition. The main result will be that indeed models which violate \eqn{VUal}, still allow an $N=2$ supersymmetric formulation. \par The scalars of the special K\"ahler\ manifolds are contained in the $\theta=0$ sector of chiral superfields of $N=2$ supersymmetry. A chiral multiplet is a reducible representation of $N=2$ supersymmetry. After imposing suitable constraints, it gives a vector multiplet. In rigid supersymmetry these constraints can be written in superspace for a symplectic section of superfields or, in components, as a linear multiplet of constraints of symplectic sections \cite{CRTVP,LW}. With a standard symplectic metric \eqn{standardOmega}, the rigid special K\"ahler\ constraints can be used to write the lower part of the symplectic sections $V$ in terms of the upper one. The reducibility constraints for the lower parts of the sections then give rise to the field equations of the fields in the upper parts. We want to use this approach to construct the field equations of vector multiplets, coupled to $N=2$ supergravity. \par Chiral and vector multiplets can also be defined as representations of the local superconformal algebra. Then the fields of the gauge multiplet of the superconformal gauge invariances, which is the Weyl multiplet, enter in the transformation rules of the multiplets \cite{dWVHVP, dRvHdWVP}. To describe the coupling of $n$ on-shell vector multiplets to supergravity, we will start from $2n+2$ chiral multiplets. The linear multiplet of constraints which reduce these chiral multiplets to vector multiplets in supergravity will contain additional terms with fields of the Weyl multiplet \cite{dWVHVP}. \par The equations that follow by supersymmetry from this weak definition of special K\"ahler\ geometry are derived for the complete set of $2n+2$ chiral multiplets. The constraints defining special K\"ahler\ geometry involve a breaking of dilatations and the $U(1)$ transformations in the superconformal group. We also choose a symplectic fermionic constraint as the gauge choice for $S$-supersymmetry. Special conformal symmetry is broken by a choice for the dilatation gauge field as in previous approaches. So finally, this leads to the breaking of superconformal to super-Poincar\'e spacetime symmetry with a residual internal $SU(2)$ in a consistent way, without relying on a prepotential or an action. Combining the reducibility constraints with the constraints of special K\"ahler\ geometry we find $n$ on-shell vector multiplets, coupled to $24+24$ supergravity components, remnants of the Weyl multiplet. \par These $24+24$ components reside in a `current multiplet', which we identify as a reduced chiral self-dual superfield. The full supergravity equations, however, would rely on a second compensating multiplet, which is independent of the symplectic formulation. For these aspects we refer to the 3 known constructions of auxiliary field formulations \cite{auxf,dWVHVP}. \par In section~\ref{s:bb}, the building blocks of the construction, the Weyl multiplet and the chiral multiplet, are given. Their supersymmetry transformation rules and the constraint to make a vector multiplet out of a chiral are recapitulated. In section~\ref{s:kahler} the special K\"ahler constraints and the supersymmetric relatives are treated for the most general case. In section~\ref{s:gBI} we combine the constraints imposed on the chiral multiplets and those found in section~\ref{s:kahler} to find on-shell vector multiplets. Finally, we comment on the remaining off-shell components of supergravity and their field equations. We recapitulate our results in section~\ref{ss:concl}. \section{The building blocks of the construction} \label{s:bb} In this section we review the Weyl multiplet, i.e.\ the gauge multiplet of the $N=2$ superconformal symmetry, and the superconformal chiral multiplet, coupled to the Weyl multiplet. Most of the material presented here is well known (see e.g., \cite{dRvHdWVP, vecten})\footnote{However, here we use different normalizations, more suited for a manifestly symplectic formulation of the theory. We use the notations of \cite{trsummer}. So the old supersymmetry parameters are $1/\sqrt{2}$ the new ones and the old fermionic fields are $\sqrt{2}$ the new ones. Also keep in mind that $\varepsilon^{0123} = {\rm i}$.}. \subsection{The Weyl multiplet}\label{s:Weyl} The Weyl multiplet is the gravitational multiplet of $N=2$ superconformal gravity. It contains the gauge fields $e^a_\mu, \omega^{ab}_\mu, b_\mu, f_\mu^a, {\cal V}_\mu^i{}_j, A_\mu, \psi^i_\mu$ and $\phi^i_\mu$. They are, respectively, gauge fields of general coordinate transformations, Lorentz rotations, dilatations, special conformal boosts, chiral $SU(2)$ and $U(1)$, supersymmetry and special supersymmetry. The representation is completed by the Lorentz tensor $T^{ij}_{ab}$, antisymmetric in $[ij]$, the spinor $\chi^i$ and the scalar $D$. Note that $T^{ij}_{ab}$ is the antiself-dual tensor, and its complex conjugate $T_{abij}$ is self-dual. The spin connection and the gauge fields for the special conformal transformations and special supersymmetry are composite gauge fields, given by \begin{eqnarray} \omega_\mu^{ab} &=& -2e^{\nu[a}\partial_{[\mu}e_{\nu]}{}^{b]} -e^{\nu[a}e^{b]\sigma}e_{\mu c}\partial_\sigma e_\nu{}^c -2e_\mu{}^{[a}e^{b]\nu}b_\nu \nonumber\\ & & -\ft{1}{2}(2\bar{\psi}_\mu^i\gamma^{[a}\psi_i^{b]} +\bar{\psi}^{ai}\gamma_\mu\psi^b_i+{\rm h.c.})\,, \nonumber\\ \phi_\mu^i &=& (\sigma^{\rho\sigma}\gamma_\mu-\ft{1}{3}\gamma_\mu\sigma^{\rho\sigma}) \left({\cal D}_\rho\psi_\sigma^i-\ft{1}{8}\sigma\cdot T^{ij}\gamma_\rho\psi_{\sigma j}\right) +\ft{1}{2}\gamma _\mu \chi^i\,, \nonumber\\ f_\mu{}^a &=& \ft12 {\cal R}_\mu^a -\ft12e_\mu{}^a f_\nu{}^\nu -{\rm i}\ft14 e^{a\nu}\varepsilon_{\mu\nu}{}^{\rho\sigma}{\hat R}_{\rho\sigma}(U(1)) +\ft1{16}T_{ij}^{ab}T_{\mu b}^{ij} -\ft34 e_\mu{}^a D\label{dependent} \\ &&+ \Big({\bar \psi}^i_{[\mu}\sigma^{ab}\phi_{\nu]i} + \ft12{\bar\psi}^i_{[\mu}T^{ab}_{ij}\psi^j_{\nu]} -\ft32 {\bar \psi}^i_{[\mu}\gamma_{\nu]}\sigma^{ab}\chi_i - {\bar\psi}^i_{[\mu}\gamma_{\nu]}{\hat R}^{ab}({\rm Q})_i + {\rm h.c.}\Big)e_b{}^\nu \,.\nonumber \end{eqnarray} The following expressions are used in $f_\mu{}^a$: \begin{eqnarray} f_\mu{}^\mu &=& \ft{1}{6}{\cal R}-D -\Big(\ft{1}{6}e^{-1}\varepsilon^{\mu\nu\rho\sigma}\bar{\psi}_\mu^i\gamma_\nu {\cal D}_\rho\psi_{\sigma i} -\ft{1}{6}\bar{\psi}_\mu^i\psi_\nu^j\,T_{ij}^{\mu\nu} -\ft{1}{2}\bar{\psi}_\mu^i\gamma^\mu\chi_i+{\rm h.c.}\Big)\,,\nonumber \\ {\hat R}_{\rho\sigma}(U(1)) & = & 2\partial_{[\mu}A_{\nu]} -{\rm i}\left(2{\bar \psi}^i_{[\mu}\phi_{\nu]i} +\ft32 {\bar \psi}^i_{[\mu}\gamma_{\nu]}\chi_i +{\rm h.c.}\right)\,,\nonumber \\ {\hat R}^{ab}( Q)^i & = & 2{\cal D}_{[\mu} \psi_{\nu]}^i -\gamma_{[\mu}\phi_{\nu]}^i-\ft14 \sigma\cdot T^{ij}\gamma_{[\mu}\psi_{\nu] j} \,. \label{dependent2} \end{eqnarray} Also, ${\cal D}_\mu$ is covariant with respect to the linearly realized symmetries: Lorentz transformations, dilatations, $U(1)$ and $SU(2)$, i.e. \begin{equation} {\cal D}_\mu \psi_\nu^i= \left( \partial_\mu -\ft12\omega_{\mu}^{ab}\sigma_{ab} +\ft12 b_\mu +\ft 12{\rm i} A_\mu \right) \psi_\nu^i +\ft12 {\cal V}_\mu{}^i{}_j\psi_\nu^j\ . \end{equation} Furthermore, ${\cal R} = e^\mu_a e^\nu_b{\cal R}_{\mu\nu}^{ab}$ is the Ricci scalar derived from the Riemann tensor \begin{equation} {\cal R}_{\mu\nu}{}^{ab} = 2\partial_{[\mu}\omega_{\nu ]}^{ab} - 2\omega_{[\mu}^{ac}\omega_{\nu ]c}{}^b \end{equation} and \begin{equation} {\cal R}^a_\mu = e^\nu_b{\cal R}_{\mu\nu}^{ab}\ . \end{equation} The transformation rules of the independent fields of the Weyl multiplet under supersymmetry, special supersymmetry and special conformal transformations (with parameters $\epsilon^i$, $\eta^i$ and $\Lambda^a_K$) are \begin{eqnarray} \delta e_\mu{}^a &=& \bar{\epsilon}^i\gamma^a\psi_{\mu i}+{\rm h.c.}\,, \nonumber\\ \delta\psi_\mu^i &=& {\cal D}_\mu\epsilon^i -\ft18 \sigma\cdot T^{ij}\gamma_\mu\epsilon_j -\ft12 \gamma_\mu\eta^i \,,\nonumber\\ \delta b_\mu &=& \ft12\bar{\epsilon}^i\phi_{\mu i} -\ft34\bar{\epsilon}^i\gamma_\mu\chi_i -\ft12\bar{\eta}^i\psi_{\mu i}+{\rm h.c.} +\Lambda_K^a\,e_{\mu\,a} \,,\nonumber\\ \delta A_\mu &=& \ft{1}{2}{\rm i} \bar{\epsilon}^i\phi_{\mu i} +\ft{3}{4}{\rm i}\bar{\epsilon}^i\gamma_\mu\chi_i +\ft{1}{2}{\rm i} \bar{\eta}^i\psi_{\mu i}+{\rm h.c.}\,, \nonumber\\ \delta {\cal V}_\mu^i{}_j &=& 2\bar{\epsilon}_j\phi_\mu^i-3\bar{\epsilon}_j\gamma_\mu\chi^i+2\bar{\eta}_j\psi_\mu^i -({\rm h.c.} \, ; \, {\rm traceless})\,, \nonumber\\ \delta T_{ab}^{ ij} &=& 8 \bar{\epsilon}^{[ i} \hat{R}_{ab}(Q)^{j]}\,, \nonumber\\ \delta\chi^i &=& -\ft1{12}\sigma\cdot\hbox{\ooalign{$\displaystyle D$\cr$\hspace{.03in}/$}} T^{ij}\epsilon_j +\ft{1}{6}\hat{R}(SU(2))^i{}_{j}\cdot\sigma\epsilon^j -\ft{1}{3}{\rm i} \hat{R}(U(1))\cdot\sigma\epsilon^i \nonumber\\ && +\ft12D\,\epsilon^i +\ft1{12}\sigma\cdot T^{ij}\eta_j \,,\nonumber\\ \delta D &=& \bar\epsilon^i\hbox{\ooalign{$\displaystyle D$\cr$\hspace{.03in}/$}}\chi_i+{\rm h.c.}\,. \label{transfo4}\end{eqnarray} The expression for the superconformal covariant curvatures $\hat R$ and the other transformation rules (in terms of the old conventions) were given in \cite{dWVHVP}. \subsection{The chiral multiplet} A chiral multiplet is a reducible representation of the superconformal algebra \cite{dRvHdWVP}. By imposing a linear multiplet of constraints it becomes a vector multiplet. This is an irreducible representation of the superconformal algebra. The constraints are called the generalized Bianchi identities, because they contain a Bianchi identity for the tensor in the chiral multiplet. \par Later we want to couple vector multiplets to conformal supergravity. The scalars of these vector multiplets form a symplectic section. They are the lowest components of a multiplet. Therefore, all the components of the multiplets have to form such a symplectic section. This is the reason to start from a $(2n+2)$-dimensional section of chiral multiplets: \begin{eqnarray} \tilde \Phi &=& V + \bar \theta^i \tilde \Omega_i + \ft14 \bar \theta^i \theta^j \tilde Y_{ij} + \ft14 \varepsilon_{ij} \bar \theta^i \sigma\cdot \tilde {\cal F}^- \theta^j\nonumber\\ & &+ \ft16 \varepsilon_{ij} (\bar \theta^i \sigma_{ab} \theta^j) \bar \theta^k \sigma^{ab} \tilde \Lambda_k + \ft1{48} (\varepsilon_{ij} \bar \theta^i \sigma_{ab} \theta^j)^2 \tilde C\,. \label{chiralsup} \end{eqnarray} The components of the section are denoted by \begin{equation} \tilde \Phi =\left(\begin{array}{c} \phi^I\\\phi_{F,I} \end{array}\right)\,, \end{equation} with $I=0,\dots,n$, which gives for the components of the chiral superfield \begin{eqnarray} V&=&\left(\begin{array}{c}X^I\{\cal F}_I\end{array}\right)\,, \hspace{.5in} \tilde \Omega_i = \left(\begin{array}{c}\Omega_i^I \\ \Omega_{F,Ii}\end{array}\right)\,, \hspace{.5in} \tilde Y_{ij} = \left(\begin{array}{c}Y_{ij}^I\\Y_{F,I ij}\end{array}\right)\,, \nonumber\\ \tilde {\cal F}^-_{ab} &=& \left(\begin{array}{c} {\cal F}_{ab}^{I-} \\ {\cal G}_{F,I ab}^-\end{array}\right)\,, \hspace{.34in} \tilde \Lambda_i =\left(\begin{array}{c} \Lambda^I_i \\ \Lambda_{F,Ii}\end{array}\right)\,, \hspace{.57in} \tilde C = \left(\begin{array}{c} C^I\\C_{F,I}\end{array}\right)\,. \end{eqnarray} This section of multiplets is independent of the existence of a prepotential. The multiplets starting with $F_I$ are on an equal footing with the ones starting with $X^I$. As long as we do not impose the constraints to obtain a vector superfield or the special K\"ahler constraints these are $2n+2$ independent chiral multiplets. The full superconformal transformation rules are given by \begin{eqnarray} \delta V &=& \bar \epsilon^i \tilde \Omega_i + (\Lambda_D - {\rm i} \Lambda_A)V\,,\nonumber\\ \delta \tilde \Omega_i &=& \hbox{\ooalign{$\displaystyle D$\cr$\hspace{.03in}/$}} V \epsilon_i + \ft12 \tilde Y_{ij} \epsilon^j + \ft12 \sigma\cdot \tilde {\cal F}^- \varepsilon_{ij} \epsilon^j + V\eta_i + (\ft32 \Lambda_D - \ft 12{\rm i} \Lambda_A) \tilde \Omega_i + \Lambda_{SU(2)\, i}{}^j \tilde \Omega_j\,,\nonumber\\ \delta \tilde Y_{ij} &=& 2 \bar{\epsilon}_{(i} \hbox{\ooalign{$\displaystyle D$\cr$\hspace{.03in}/$}} \tilde \Omega_{j)} - 2\bar{\epsilon}^k \tilde\Lambda_{(i} \varepsilon_{j)k} + 2 \Lambda_D \tilde Y_{ij} +2 \Lambda_{SU(2)\, (i}{}^k \tilde Y_{j)k}\,,\nonumber\\ \delta \tilde {\cal F}_{ab}^-&=&\varepsilon^{ij} \bar{\epsilon}_i \hbox{\ooalign{$\displaystyle D$\cr$\hspace{.03in}/$}} \sigma_{ab} \tilde \Omega_j + \bar{\epsilon}^i \sigma_{ab} \tilde \Lambda_i -2 \varepsilon^{ij} \bar \eta_i \sigma_{ab} \tilde \Omega_j + 2 \Lambda_D \tilde {\cal F}^-_{ab} \,,\nonumber\\ \delta \tilde \Lambda_i &=&-\ft12 \sigma \cdot \tilde {\cal F}^- \stackrel{\leftarrow} {\hbox{\ooalign{$\displaystyle D$\cr$\hspace{.03in}/$}}} \epsilon_i - \ft12 \hbox{\ooalign{$\displaystyle D$\cr$\hspace{.03in}/$}} \tilde Y_{ij} \epsilon_k \varepsilon^{jk} + \ft12 \tilde C \epsilon^j \varepsilon_{ij}\nonumber\\ & & -\ft18 \hbox{\ooalign{$\displaystyle D$\cr$\hspace{.03in}/$}} (V \varepsilon^{jk}T_{jk} \cdot \sigma)\epsilon_i - \ft32(\bar \chi_{[i}\gamma_a \tilde \Omega_{j]}) \gamma^a \epsilon_k \varepsilon^{jk} \nonumber\\ & & - \tilde Y_{ij} \varepsilon^{jk} \eta_k + \ft52 \Lambda_D \tilde \Lambda_i + \ft 12{\rm i}\Lambda_A \tilde \Lambda_i + \Lambda_{SU(2)\, i}{}^{j} \tilde \Lambda_j \,,\nonumber\\ \delta \tilde C &=& - 2 \varepsilon^{ij} \bar{\epsilon}_i \hbox{\ooalign{$\displaystyle D$\cr$\hspace{.03in}/$}} \tilde \Lambda_j - 6 \bar{\epsilon}_i \chi_j \tilde Y_{kl} \varepsilon^{ik}\varepsilon^{jl} + \ft12 \bar{\epsilon}_i \sigma\cdot T_{jk} \hbox{\ooalign{$\displaystyle D$\cr$\hspace{.03in}/$}} \tilde \Omega_l \varepsilon^{ij}\varepsilon^{kl} + 2 \varepsilon^{ij} \bar\eta_i \tilde \Lambda_j\nonumber\\ & & + 3 \Lambda_D \tilde C + i \Lambda_A \tilde C\,. \label{transfo} \end{eqnarray} This superconformal chiral superfield can be reduced to a vector superfield with the constraints \begin{eqnarray} 0&=&\tilde Y_{ij} - \varepsilon_{ik}\varepsilon_{jl} \tilde Y^{kl}\,,\label{bian1}\\ 0&=&\hbox{\ooalign{$\displaystyle D$\cr$\hspace{.03in}/$}} \tilde \Omega^i - \varepsilon^{ij} \tilde \Lambda_j\,,\label{bian2}\\ 0&=& D^a(\tilde {{\cal F}}_{ab}^+ - \tilde {{\cal F}}_{ab}^- + \ft14 V T_{ab\, ij}\varepsilon^{ij} - \ft14 \bar V T^{ij}_{ab}\varepsilon_{ij}) - \ft32 ( \varepsilon^{ij} \bar \chi_i\gamma_b \tilde \Omega_j - \mbox{h.c.})\,,\label{bian3}\\ 0&=&-2 \Box \bar V - \ft 14 \tilde {\cal F}^+_{\mu\nu} T^{\mu\nu}_{ij}\varepsilon^{ij} - 6 \bar \chi_i \tilde \Omega^i - \tilde C\,. \label{bian4} \end{eqnarray} The symplectic vector of chiral multiplets with these constraints define $2n+2$ vector multiplets in superconformal gravity. The special K\"ahler constraints will relate them such that one ends up with $n+1$ vectors and $n$ complex scalars and spinors obeying field equations. \setcounter{equation}{0} \section{Gauge choices and special K\"ahler constraints\label{s:kahler}} To obtain a Poincar\'e supergravity theory of $n$ vector multiplets, we start from the assumption that the components in the symplectic sections $V$ are the lowest components of reduced chiral multiplets, as is the case in previous constructions of matter couplings in $N=2$ supergravity. To achieve that, we have to impose the reducibility constraints \eqn{bian1}--(\ref{bian4}) on the chiral multiplets and suitable constraints that impose restrictions on the sections such that the resulting theory contains $n$ physical vector multiplets and the gravity multiplet. The superfluous symmetries of the superconformal construction need to be broken by suitable gauge choices. The symplectic section $V$ can be seen as a function of $n$ scalars $z^\alpha$ and their complex conjugates $\bar z^{\bar\alpha}$ ($\alpha =1,..., n$). These scalars can be interpreted as the coordinates of a special K\"ahler manifold. \par Having introduced $K'$ in \eqn{kappa}, we have exhausted constraint \eqn{DU=0}. The remaining relevant constraints are then \eqn{VVbar}, \eqn{DbarV}, and we will take the formulation with \eqn{UU}. Condition (\ref{VVbar}) gauge fixes the dilatations, choosing the canonical kinetic term for the graviton. Equation~(\ref{DbarV}) imposes the holomorphicity of the scalar fields. For the symmetry of the kinetic matrix of the vectors, one needs another constraint, which is \eqn{UU}. In all previous papers on special geometry, one imposed instead \eqn{VUal}, which is equivalent for $n>1$, but not for $n=1$ as mentioned in the introduction. There is no physical argument known to demand \eqn{VUal}, but up to now, no physical applications have been found that do not fulfil it. \par We have thus seen that we can look upon equations \eqn{VVbar} and \eqn{DbarV} in two ways. They are the defining equations of special geometry, as well, they can be considered as gauge choices for the dilatations and chiral $U(1)$ transformations present in the superconformal algebra. As we will see below a supersymmetric extension of these constraints will include the gauge choice of $S$-supersymmetry. \par {}From (\ref{defD}) follows \begin{equation} g_{\alpha \bar \beta}\equiv \partial_\alpha \partial_{\bar \beta} K = {\rm i} \< U_\alpha, \bar U_{\bar \beta}\>\,. \end{equation} Furthermore, we impose the \lq physical' condition (positivity of the kinetic energy terms of the vectors \cite{CRTVP}) that\footnote{ Keep in mind that for $n>1$ one always has $Z_z = 0$.} \begin{eqnarray} \det g_{\alpha\bar\beta} > 0 \ &\qquad\mbox{if }& \< V,U_\alpha\> = 0\,, \nonumber\\ g'_{z\bar z} \equiv g_{z{\bar z}}- Z_z\bar Z_{\bar z}\ > 0 &\qquad\mbox{for }& Z_z \equiv \< V, U_z\> \,. \label{physicalcond} \end{eqnarray} Using the constraints it can be shown that \begin{equation} {\cal W}=(V,U_\alpha, \bar V, \bar U_{\bar \alpha}) \label{basiscW} \end{equation} forms for every $z,\bar z$ a basis for symplectic vectors. More information about the expansion coefficients can be found in appendix~\ref{s:basis}. This expansion will be used in the derivation of the supersymmetric extension of the special K\"ahler constraints and of the field equations. \subsection{The constraint on the curvature} Covariant derivatives involve the K\"ahler\ connection as in \eqn{defD}, and after choosing a real K\"ahler\ potential one may define a K\"ahler\ weight\footnote{$V$ and $U_\alpha$ have weight 1, while $Z_z$ has weight 2, and for their complex conjugates respectively $-1$ and $-2$.} $p$ for a symplectic section $W$, such that \begin{equation} {\cal D}_\alpha W= \left( \partial_\alpha +\ft p2 (\partial_\alpha K)\right) W\,,\qquad {\cal D}_{\bar \alpha} W= \left( \partial_{\bar \alpha} -\ft p2 (\partial_{\bar \alpha} K)\right) W \,. \label{covderp} \end{equation} If $W$ carries indices $\alpha$ or $\bar \alpha$ there is a further metric connection, defined such that ${\cal D}_\alpha g_{\beta\bar \gamma}=0$. The curvature of the special K\"ahler\ manifold is then defined by \begin{equation} [{\cal D}_\alpha, {\cal D}_{\bar\beta}] X_\gamma = - p g_{\alpha\bar\beta} X_\gamma - R_{\alpha\bar\beta\gamma}{}^\delta X_\delta\,, \end{equation} where $X_\alpha$ is a generic vector with K\"ahler weight $p$. Applying this for $X_\alpha$ replaced by $U_\alpha$ and taking the symplectic inner product with $\bar U_{\bar \delta}$ one finds \begin{equation} R_{\alpha\bar\beta\gamma\bar\delta}\equiv g_{\delta\bar\delta} R_{\alpha\bar\beta\gamma}{}^\delta = - 2 g_{(\alpha|\bar\beta|} g_{\gamma)\bar\delta} - {\rm i} \< {\cal D}_\alpha U_\gamma , {\cal D}_{\bar\beta} \bar U_{\bar\delta}\>\,. \label{curvatureSK} \end{equation} If we introduce a symmetric tensor \begin{equation} C_{\alpha\beta\gamma} = \< U_\alpha, {\cal D}_\beta U_\gamma\>\,, \label{defC} \end{equation} and expand the last term of \eqn{curvatureSK} in the basis ${\cal W}$ according to appendix~\ref{s:basis} we obtain the following two cases:\\ \noindent $1. \underline{{\rm The\, generic\, case:}}$ \begin{equation} {\cal D}_\alpha U_\beta={\rm i} C_{\alpha\beta\gamma}g^{\gamma\bar \gamma}\bar U_{\bar \gamma}\ , \end{equation} and the curvature is constrained to \begin{equation} R_{\alpha\bar\beta\gamma\bar\delta} \equiv g_{\beta\bar\beta}R^\beta_{\alpha\gamma\bar\delta} = - 2 g_{(\alpha|\bar\beta|} g_{\gamma)\bar\delta} + C_{\alpha\gamma\epsilon} g^{\epsilon\bar\epsilon} \bar C_{\bar\beta\bar\delta\bar\epsilon}. \end{equation} \noindent $2. \underline{{\rm The\, special\, case:}}$\newline In a similar way one finds that in this case \begin{equation} \mathcal{D}_z U_z = {\rm i} g'^{z\bar z}(C_{zzz}\bar U_{\bar z} - g_{z\bar z}{\cal D}_z Z_z \bar V')\,. \end{equation} The curvature becomes \begin{equation} R_{z\bar zz\bar z}= -2 g_{z\bar z}^2 - g_{z\bar z}({\cal D}_z Z_z)g^{\prime z\bar z}({\cal D}_{{\bar z}}\bar Z_{\bar z} ) + C_{zzz} g^{\prime z\bar z}\bar C_{\bar z\bar z\bar z} \,. \end{equation} \subsection{An adapted basis and metric for the special case} When $Z_z\neq 0$, one may diagonalize the matrix of symplectic products between $V,\bar V,U_z$ and $\bar U_{\bar z}$ by defining \begin{equation} U'_z=U_z +{\rm i} Z_z\bar V \ ; \qquad \bar U'_{\bar z}=\bar U_{\bar z} -{\rm i} \bar Z_{\bar z}V \,. \end{equation} We then have symplectic products \begin{eqnarray} &&\< V,\bar V\>={\rm i}\,, \qquad \< V,U'_z\>= \< V,\bar U'_{\bar z}\>= \< \bar V,U'_z\> = \< \bar V,\bar U'_{\bar z}\>=0\,,\nonumber\\ && \<\bar U'_{\bar z}, U'_z\>={\rm i} (g_{z\bar z}-Z_z\bar Z_{\bar z})= {\rm i} g'_{z\bar z} \,. \end{eqnarray} In this way we find the Hermitian metric $g'_{z\bar z}$ which is invertible because of \eqn{physicalcond}, but is not the second derivative of the K\"ahler\ potential $K$, used to define the covariant derivatives in \eqn{covderp}. With this definition, covariant derivatives on the above equations lead to \begin{equation} \< {\cal D}_zU'_z,\bar V \>= \< {\cal D}_zU'_z,V \>= 0 \,.\label{DU12} \end{equation} The defining expressions for $U_z$ and $Z_z$ imply \begin{equation} {\cal D}_z \bar U_{\bar z}=g_{z{\bar z}}\bar V\,,\qquad {\cal D}_{\bar z} U_z=g_{z{\bar z}} V\,,\qquad {\cal D}_z \bar Z_{\bar z}= {\cal D}_{\bar z} Z_z=0\,, \end{equation} which in the new basis give \begin{equation} {\cal D}_z \bar U'_{\bar z}=g'_{z{\bar z}}\bar V-{\rm i}\bar Z_{\bar z} U'_z\,,\qquad {\cal D}_{\bar z} U'_z=g'_{z{\bar z}} V+{\rm i} Z_z\bar U'_{\bar z}\,. \label{DbUmod} \end{equation} When we define ${\cal D}'$ with metric connection such that ${\cal D}'_z g'_{z\bar z}=0$, all the above relations remain valid for ${\cal D}'$, as the non-zero connections are just $\Gamma_{zz}^z$ and $\Gamma_{\bar z \bar z}^{\bar z}$. The new definition now implies \begin{equation} \< \bar U'_{\bar z},{\cal D}'_z U'_z\> =0\,. \end{equation} The analogue of \eqn{defC} is then the definition \begin{equation} C'_{zzz}\equiv \< U'_z,{\cal D}'_z U'_z\>=C_{zzz}\,. \end{equation} This leads again to \begin{equation} {\cal D}'_z U'_z ={\rm i} C_{zzz} g'^{z\bar z} \bar U'_{\bar z}\,. \end{equation} We define then the curvature based on the metric $g'$ by \begin{equation} [{\cal D}'_z, {\cal D}'_{\bar z}] X_z = - p g_{z\bar z} X_z - R'_{z\bar zz}{}^z X_z\,. \end{equation} Observe that the first term has $g$ and not $g'$ as this is the K\"ahler\ curvature. Calculating as before the curvature $R'$ by replacing $X_z$ with $U'$, and an inner product with $\bar U'$, the last terms in \eqn{DbUmod} lead to extra terms such that we find \begin{equation} R'_{z\bar zz\bar z}\equiv R'_{z\bar zz}{}^z g'_{z\bar z}= -2 g_{z\bar z} g'_{z\bar z} +C_{zzz}g'^{z\bar z}\bar C_{\bar z\bar z\bar z}\,. \end{equation} Rephrasing as much as possible in terms of the metric $g'_{z\bar z}$, we thus recover another geometry as for other special K\"ahler\ models. There is an essential difference in the product of metrics in \eqn{curvatureSK} and here. We tried to extend our analysis in the basis ${\cal W'}=(V, U'_z, \bar V , \bar U'_{\bar z})$, but ran into problems with the transformation rules because we want $V$ to be the lowest component of a chiral multiplet. So, it is not possible to get rid of $Z_z\neq 0$ by choosing another basis while keeping a section of chiral multiplets. The model with $Z_z\neq 0$ is really another model compared to those studied in the past. \par For some calculations below, it is also useful to introduce another basis with symplectic vectors orthogonal to $U$. That is, we introduce \begin{eqnarray} V'=V +{\rm i} Z_zg^{z\bar z}\bar U_{\bar z} &\,,\qquad& \bar V'=\bar V -{\rm i}\bar Z_{\bar z}g^{z\bar z} U_z\,, \nonumber\\ \< V',\bar V'\> = {\rm i}(1-g^{z{\bar z}}Z_z\bar Z_{\bar z}) &\,, \qquad&\< V',U_z\>= \< V',\bar U_{\bar z}\> =\< \bar V',U_z\> = \< \bar V',\bar U_{\bar z}\>=0\,,\nonumber\\ \<\bar U_{\bar z}, U_z\>={\rm i} g_{z\bar z}\,. \end{eqnarray} \subsection{Supersymmetric extension of special K\"ahler constraints} \label{ss:susyKaconstr} It is clear that the constraint (\ref{VVbar}) breaks the superconformal symmetry. The constraints and their supersymmetric partners therefore play the role of gauge conditions for some of the superconformal symmetries. The residual symmetry should then still contain the symmetries of Poincar\'e supergravity. In this subsection we will derive the supersymmetric partners of the constraints \eqn{VVbar}, \eqn{DbarV} and \eqn{UU}, and compute the decomposition rule for the resulting supergravity, i.e.\ the rule which gives the remaining symmetry as a linear combination of the original, superconformal, symmetry. \subsubsection{Gauge choices and decomposition rule} Before we go to these constraints, we {\it break the special conformal symmetry} by imposing a constraint on $b_\mu$: \begin{equation}\label{EC} K\mbox{-gauge:}\quad b_\mu =0\, . \end{equation} This does not alter the number of degrees of freedom as $b_\mu$ is pure gauge in the Weyl multiplet (cf table~\ref{dofbeforeConstr}). \begin{table}[ht] \begin{center} \begin{tabular}{|c|c|l|}\hline fields&d.o.f.&comments\\ \hline\hline \multicolumn{3}{|c|}{The Weyl multiplet ($24+24$)}\\ \hline $e_\mu^{\ a}$&5&16 - 4(translation.) - 6(Lorentz) - 1(dilatation)\\ \hline $b_\mu$&0&4 - 4(special conformal.)\\ \hline $A_\mu$&3&4 - 1($U(1)$)\\ \hline ${\cal V}_{\mu}{}^i{}_j$&9&12 - 3($SU(2)$)\\ \hline $\psi_\mu^i$&16&32 - 8($Q$-supersymmetry) - 8($S$-supersymmetry)\\ \hline $T^{ij}_{ab}$&6&complex antiself-dual \\ \hline $\chi_i$&8&\\ \hline $D$&1&real scalar\\ \hline\hline \multicolumn{3}{|c|}{Symplectic section of chiral multiplets ($16(2n+2)+16(2n+2)$)}\\ \hline $V$&2(2n+2)&\\ \hline $\tilde \Omega_i$&8(2n+2)&\\ \hline $\tilde Y_{ij}$&6(2n+2)&\\ \hline $\tilde {\cal F}^-_{ab}$&6(2n+2)&\\ \hline $\tilde \Lambda_i$ & 8(2n+2)&\\ \hline $\tilde C$ & 2(2n+2)&\\ \hline \end{tabular} \caption{Degrees of freedom in the model before the constraints. \label{dofbeforeConstr}} \end{center} \end{table} \par The decomposition rule for the special conformal symmetry is \begin{equation} \Lambda_K^a = -e^{\mu a} \left( \ft12 \bar{\epsilon}^i \phi_{\mu i} - \ft34 \bar{\epsilon}^i \gamma_\mu \chi_i - \ft12 \bar \eta^i \psi_{\mu i} + \mbox{h.c.}\right)\,. \label{LK} \end{equation} Constraint (\ref{VVbar}) {\it breaks the dilatations}. Indeed, the superconformal transformation of (\ref{VVbar}) gives \begin{equation} \< \bar V, \bar{\epsilon}^i \tilde \Omega_i \> - \< V,\bar{\epsilon}_i\tilde \Omega^i\> + 2{\rm i} \Lambda_D =0\,, \label{DecompD} \end{equation} and the dilatations are now a combination of other symmetries. We choose as $S$-gauge \begin{equation} S\mbox{-gauge:}\quad\< \bar V, \tilde \Omega_i\> =0 \quad {\rm and} \quad \< V ,\tilde \Omega^i\> =0\,.\label{Sgauge} \end{equation} Remark that after this gauge choice the decomposition rule \eqn{DecompD} simplifies to \begin{equation} \Lambda_D = 0\label{LD}\,, \end{equation} such that we can forget about the original dilatations completely. Demanding that the sections $V$ depend on $z^\alpha$ and $\bar z^{\bar \alpha}$ in the way described in \eqn{defD}--\eqn{DbarV}, is a {\it gauge choice for the chiral $U(1)$-transformations}. In fact, consider the transformation of the first line of \eqn{transfo} using these equations: \begin{equation} \delta V = U_\alpha\delta z^\alpha -\ft 12 (\partial_\alpha K' \delta z^\alpha - \partial_{\bar \alpha} K' \delta \bar z^{\bar \alpha})V\,. \end{equation} An inner product with $\bar V$ gives (using (\ref{VVbar}) and its covariant derivative) a decomposition rule for the $U(1)$-transformations, i.e.\ \begin{equation} \Lambda_A= {\rm Im ~}(\partial_\alpha K' \delta z^\alpha)\,, \label{LA} \end{equation} where we have already used \eqn{kappa}. \par The decomposition rule for $\delta_S(\eta_i)$ follows from the variation of the $S$-gauge: \begin{eqnarray} \eta_i &=& -{\rm i} \< \bar V, \hbox{\ooalign{$\displaystyle D$\cr$\hspace{.03in}/$}} V\> \epsilon_i - \ft 12{\rm i} \< \bar V, \tilde Y_{ij}\> \epsilon^j \nonumber\\ & &- \ft 12{\rm i} \< \bar V, \tilde {\cal F}_{ab}^-\> \varepsilon_{ij} \sigma^{ab} \epsilon^j - {\rm i} \< \bar{\epsilon}_j \tilde \Omega^j, \tilde \Omega_i\> \,. \label{SS} \end{eqnarray} {}From now on, we only request that the constraints are invariant under the resulting Poincar\'e supersymmetry \begin{equation} \delta(\epsilon_i) = \delta_Q(\epsilon_i) + \delta_S (\eta_i) + \delta_A (\Lambda_A) + \delta_K(\Lambda_K)\,, \end{equation} with $\Lambda_K$, $\Lambda_A$ and $\eta_i$ defined in \eqn{LK}, (\ref{LA}) and (\ref{SS}). \par Having the symplectic sections as functions of $z$ and $\bar z$, we can consider the transformations of the bosonic constraints \eqn{VVbar}--\eqn{DU=0} and \eqn{UU}. The variation of the first one determined the breaking of dilatations. The constraints \eqn{DbarV} and \eqn{DU=0} are used to determine the $z,{\bar z}$ dependences of $V$, $U$ and $K$ and their supersymmetry transformations are thus trivial if we compute them in terms of $\delta z$ and $\delta\bar z$. The constraint (\ref{UU}) is only non-trivial for $n>1$. Its variation is \begin{equation} \delta \< U_\alpha, U_\beta \> = 2\< {\cal D}_\gamma U_{[\alpha},U_{\beta]}\> \delta z^\gamma\,, \end{equation} which is $0$ due to the symmetry of \eqn{defC}. This finishes the supersymmetry variations of the bosonic special K\"ahler\ constraints. \subsubsection{Physical fermions and fermionic constraints} The first line of \eqn{transfo}, using \eqn{LD} and \eqn{LA}, is in terms of $\delta z$: \begin{equation} \bar \epsilon^i \Omega_i=U_\alpha \delta z^\alpha\,. \end{equation} Therefore, the supersymmetry transformation of $z$ is chiral, and we define $\lambda_i^\alpha$ as \begin{equation} \bar{\epsilon}^i \lambda_i^\alpha \equiv\delta z^\alpha \,, \end{equation} leading to \begin{equation} \tilde \Omega_i = U_\alpha \lambda_i^\alpha \,, \label{Oml} \end{equation} compatible with the $S$-gauge. The relation (\ref{Oml}) can be inverted to \begin{equation} \lambda_i^\alpha = -{\rm i} g^{\alpha\bar \alpha}\< \bar U_{\bar \alpha} , \tilde \Omega_i\> \,. \label{lambdainOmega} \end{equation} That $\tilde \Omega_i$ has only components in the $U$ direction implies the constraints (the primes here and below are irrelevant for $n>1$ or $Z_z=0$) \begin{equation} \< V ,\tilde \Omega_i\> =Z_z\lambda_i^\alpha \, \mbox{ or }\, \< V' ,\tilde \Omega_i\> = 0\,,\qquad \< U_\alpha ,\tilde \Omega_i\> =0\,. \label{UO} \end{equation} The transformation rules for $z^\alpha$ and $\lambda_i^\alpha$ are\footnote{In the transformation laws below, there is still the $SU(2)$ transformation which is not gauge fixed and thus independent of the other transformations. We will not indicate these transformations explicitly, as they follow from the position of the $i$ indices.} \begin{eqnarray} \delta z^\alpha &=& \bar{\epsilon}^i \lambda_i^\alpha\,,\nonumber\\ \delta \lambda_i^\alpha &=& - \Gamma^\alpha_{\beta\gamma} \lambda_i^\beta \delta z^\gamma + \ft14 (\partial_\beta K \delta z^\beta - \mbox{h.c.})\lambda^\alpha_i\nonumber\\ &&+/\!\!\!\!\nabla z^\alpha \epsilon_i - \ft 12{\rm i} g^{\alpha\bar \alpha} \< \bar U_{\bar \alpha}, \tilde Y_{ij}\> \epsilon^j - \ft 12{\rm i} g^{\alpha \bar \alpha} \< \bar U_{\bar \alpha} ,\tilde {\cal F}_{ab}^-\> \varepsilon_{ij} \sigma^{ab} \epsilon^j\,, \label{Psusy} \end{eqnarray} where \begin{equation} \nabla_\mu z^\alpha = \partial_\mu z^\alpha - \bar \psi^i_\mu \lambda_i^\alpha\,. \end{equation} \subsubsection{Further variations of constraints in the generic case} At the first fermionic level we have imposed the gauge choice \eqn{Sgauge}, and found furthermore the constraints \eqn{UO}, leaving $n$ physical fermions as shown in \eqn{Oml} and \eqn{lambdainOmega}. The variation of the $S$-gauge leads to the decomposition rule. Here we will determine the further constraints on the $2(n+1)$ chiral multiplets in the symplectic vector. We first perform this analysis for the generic case where $\< V , U_\alpha\> = 0$, and treat the case $n=1$ separately afterwards. \par The Poincar\'e transformations on (\ref{UO}) give \begin{eqnarray} \< V, \tilde Y_{ij}\> &=&0\,, \nonumber\\ \< U_\alpha , \tilde Y_{ij}\> & =& -C_{\alpha\beta\gamma} \bar \lambda_i^\beta \lambda_j^\gamma\,, \nonumber\\ \< V, \tilde {\cal F}_{ab}^-\>&=&0 \,,\nonumber\\ \< U_\alpha ,\tilde {\cal F}^-_{ab}\> &=& -\ft12 C_{\alpha\beta\gamma} \varepsilon^{ij} (\bar\lambda_i^\beta \sigma_{ab}\lambda_j^\gamma)\,. \label{dUOF} \end{eqnarray} To analyse the content of these equations, we make use of lemma~B.1 of \cite{CRTVP}. This says that the $2(n+1)\times (n+1)$ matrix $(V,U_\alpha )$ has rank $(n+1)$. Thus we can solve (\ref{dUOF}) for half of the components of $\tilde Y_{ij}$ and $\tilde {\cal F}_{ab}^-$. \par Straightforward variation of these two equations under Poincar\'e supersymmetry yields a set of new constraints: \begin{eqnarray} \< V,\tilde\Lambda_i\>&=& - \ft16 C_{\alpha\beta\gamma} \varepsilon^{kl} (\bar \lambda_k^\alpha \sigma^{ab} \lambda_l^\beta) \sigma_{ab} \lambda_i^\gamma\,,\label{condL3}\\ \< U_\alpha, \tilde \Lambda_i \> &=& \ft 12{\rm i} C_{\alpha\beta\gamma} g^{\beta\bar\beta} \left( \< \bar U_{\bar \beta} , \tilde Y_{ij} \> \varepsilon^{jk} \lambda_k^\gamma + \< \bar U_{\bar\beta} , \sigma\cdot \tilde {\cal F}^-\> \lambda_i^\gamma\right)\nonumber\\ & & + \ft16 {\cal D}_\alpha C_{\beta\gamma\delta} \cdot \varepsilon^{kl} (\bar \lambda_k^\beta \sigma_{ab} \lambda^\gamma_l) \sigma^{ab} \lambda_i^\delta\,. \label{condL1} \end{eqnarray} Varying constraint (\ref{condL3}) yields \begin{eqnarray} \< V,\tilde C\>&=& \ft 12{\rm i} \varepsilon^{ik} \varepsilon^{jl} C_{\alpha\beta\gamma} g^{\alpha\bar\alpha} \< \bar U_{\bar\alpha} , \tilde Y_{ij}\> \bar \lambda_k^\beta \lambda_l^\gamma\nonumber\\ &&- \ft 12{\rm i} C_{\alpha\beta\gamma} g^{\alpha\bar\alpha} \< \bar U_{\bar\alpha} , \tilde {\cal F}_{ab}^-\> \varepsilon^{kl} (\bar \lambda_k^\beta \sigma^{ab} \lambda_l^\gamma)\nonumber\\ &&- \ft16 {\cal D}_\alpha C_{\beta\gamma\delta}\cdot \varepsilon^{ij} (\bar \lambda_i^\alpha \sigma_{ab} \lambda_j^\beta) \varepsilon^{kl} (\bar \lambda_k^\gamma \sigma^{ab} \lambda^\delta_l)\,.\label{VC} \end{eqnarray} The variation of (\ref{condL1}) gives \begin{eqnarray}\label{UC} \< U_\alpha, \tilde C\>&=& \ft14 C_{\alpha\beta\gamma} g^{\beta\bar\beta} g^{\gamma\bar\gamma}\varepsilon^{ik}\varepsilon^{jl} \< \bar U_{\bar\beta}, \tilde Y_{ij}\> \< \bar U_{\bar \gamma}, \tilde Y_{kl}\> \nonumber\\ & & - \ft12 C_{\alpha\beta\gamma} g^{\beta\bar\beta} g^{\gamma\bar\gamma} \< \bar U_{\bar\beta}, \tilde {\cal F}^-_{ab}\> \< \bar U_{\bar \gamma}, \tilde {\cal F}^{-ab}\> \nonumber\\ & & - \ft 12{\rm i} \varepsilon^{ik}\varepsilon^{jl} [C_{\alpha\beta\gamma} \< \bar V, \tilde Y_{ij}\> + {\cal D}_\alpha C_{\beta\gamma\delta}\cdot g^{\delta\bar\delta} \< \bar U_{\bar \delta}, \tilde Y_{ij}\> ]\bar \lambda_k^\beta \lambda_l^\gamma\nonumber\\ & & + \ft 12{\rm i} [ C_{\alpha\beta\gamma}\< \bar V, \tilde {\cal F}_{ab}^-\> + {\cal D}_\alpha C_{\beta\gamma\delta}\cdot g^{\delta\bar\delta} \< \bar U_{\bar \delta}, \tilde {\cal F}_{ab}^-\> ] \varepsilon^{ij}(\bar \lambda_i^\beta \sigma^{ab} \lambda_j^\gamma)\nonumber\\ & &-2{\rm i} C_{\alpha\beta\gamma} g^{\beta\bar\beta} \varepsilon^{ij} \bar \lambda_i^\gamma \< \bar U_{\bar \beta}, \tilde \Lambda_j\>\nonumber\\ & & +\ft1{12} {\cal D}_\alpha {\cal D}_\beta C_{\gamma\delta\epsilon} \cdot\varepsilon^{ij} (\bar \lambda_i^\beta \sigma_{ab} \lambda_j^\gamma )\,\varepsilon^{kl} (\bar \lambda_k^\delta \sigma^{ab} \lambda_l^\epsilon)\,. \end{eqnarray} \par These are all the possible \lq K\"ahler' constraints on the sections. Let us review the degrees of freedom. Before imposing the constraints, we have the degrees of freedom as in table~\ref{dofbeforeConstr}. First of all there is the Weyl multiplet with $24+24$ degrees of freedom. The gauge invariances have been used to determine the counting. Indeed, the dilation invariance can be seen as removing the trace of the vierbein $e_\mu^\mu$ and $\gamma^\mu \psi_\mu^i$ is pure gauge under special supersymmetry. Similarly the vectors $A_\mu $ and ${\cal V}_\mu{}^j{}_ i$ lose a degree of freedom because of their gauge transformations. Secondly, we have the symplectic vectors of $2n+2$ chiral multiplets, which altogether consist of $(2n+2) 16 + (2n+2) 16$ degrees of freedom. \par Then we have imposed the constraints \eqn{VVbar}--\eqn{DU=0}, \eqn{UU} and their supersymmetry partners. The new counting is in table~\ref{dofAfterConstr}. All the symplectic sections are first reduced to $(n+1)$ rather than $(2n+2)$ degrees of freedom, as inner products with $V$ and with $U_\alpha $ are removed by the constraints. The symplectic vector $V$ is further reduced to $n$ complex variables $z^\alpha $, by constraints which we have interpreted as gauge choices of dilatations and chiral $U(1)$. These invariances have thus disappeared, and in the upper part of the table we should thus no longer subtract from degrees of freedom of the vierbein and of $A_\mu $. Similarly, the constraint $\< \bar V, \tilde \Omega_i\>=0$ removed a spinor doublet from the degrees of freedom of $\tilde\Omega_i$, but this breaks the $S$--symmetry, and thus the gravitino still has 24 degrees of freedom. As a result, the superconformal invariance is reduced to super-Poincar\'e. The super-Poincar\'e\ multiplet contains the graviphoton, which resides in $\< \bar V,\tilde {\cal F}^-_{ab}\> $. Similarly the other internal products with $\bar V$ can be seen as auxiliary fields of the $40+40$ off-shell super-Poincar\'e\ multiplet. In other formulations \cite{auxf,dWVHVP} they are expressed in terms of another compensating multiplet. This compensating multiplet is then also used to gauge fix the $SU(2)$ invariance which we have not broken here. \begin{table}[ht] \begin{center} \begin{tabular}{|c|c|l|}\hline fields&d.o.f.&comments\\ \hline\hline \multicolumn{3}{|c|}{The Gravity Multiplet (40+40)}\\ \hline $e_\mu^{\ a}$&6&16 - 4(translation) - 6(Lorentz)\\ \hline $A_\mu$&4&gauge vector $\rightarrow$ vector\\ \hline ${\cal V}_{\mu}{}^i{}_j$&9&12 - 3($SU(2)$)\\ \hline $\psi_\mu^i$&24&32 - 8($Q$-supersymmetry)\\ \hline $T^{ij}_{ab}$&6&complex antiself-dual \\ \hline $\chi_i$&8&\\ \hline $D$&1&real scalar\\ \hline $\< \bar V,\tilde Y_{ij}\> $&6&\\ \hline $\< \bar V,\tilde {\cal F}^-_{ab}\> $&6&\\ \hline $\< \bar V,\tilde \Lambda_i\> $ &8&\\ \hline $\< \bar V,\tilde C\> $ &2&\\ \hline \hline \multicolumn{3}{|c|}{Symplectic section of constrained chiral multiplets ($16\,n+16\,n)$}\\ \hline $z^\alpha$&2n&(d.o.f. of $V$)/2 - 2 (=trace vierbein + extra comp. of $A_\mu$)\\ \hline $\lambda^\alpha_i$&8n&(d.o.f. of $\tilde \Omega_i$)/2 - 8 (=$\gamma^\mu \psi_\mu^i$)\\ \hline $\< \bar U_{\bar\alpha},\tilde Y_{ij}\> $&6n&\\ \hline $\< \bar U_{\bar\alpha},\tilde {\cal F}^-_{ab}\> $&6n&\\ \hline $\< \bar U_{\bar\alpha} ,\tilde \Lambda_i\> $ &8n&\\ \hline $\< \bar U_{\bar\alpha} ,\tilde C\> $ &2n&\\ \hline \end{tabular} \caption{Degrees of freedom in the model after the special K\"ahler\ constraints \label{dofAfterConstr}} \end{center} \end{table} \subsubsection{Further variations of constraints in the special case} Now we continue the analysis of the supersymmetry transformations on special K\"ahler constraints for supergravity theories with $Z_z = \< V, U_z \> \neq 0$. This can only happen for $n=1$, because that is the only case where this condition is not equivalent with (\ref{UU}). Because $n=1$, $U_\alpha$ and ${\cal D}_\alpha$ can be replaced by $U_z$ and ${\cal D}_z$. \par The computation of the special K\"ahler \ constraint goes along the same track as for the generic case, but extra terms appear because of the weaker constraint. The new contributions appear for the first time after the supersymmetry variation of (\ref{UO}): \begin{eqnarray} \< V', \tilde Y_{ij}\> &=& - {\cal D}_z Z_z\cdot \bar \lambda_i^z \lambda^z_j\,,\nonumber \\ \< U'_z , \tilde Y_{ij}\> &=& -C_{zzz} \bar \lambda_i^z \lambda_j^z\,, \nonumber \\ \< V', \tilde {\cal F}^-_{ab}\> &=& - \ft12 {\cal D}_z Z_z\cdot \varepsilon^{ij} (\bar \lambda_i^z \sigma_{ab}\lambda_j^z )\,,\nonumber\\ \<U'_z , \tilde {\cal F}^-_{ab}\> &=& -\ft12 C_{zzz} \varepsilon^{ij} (\bar \lambda_i^z \sigma_{ab}\lambda_j^z )\,.\label{dUOFn} \end{eqnarray} Define a new vector $V''$ \footnote{Note that $g_{z\bar z} {\cal D}_z Z_z = {\cal D}_{\bar z} C_{zzz} $ is not necessarily $0$ in this case.} \begin{equation} V'' \equiv V'- \frac{g^{z\bar z} {\cal D}_{\bar z}C_{zzz}\cdot U'_z}{C_{zzz}} = V' - \frac{{\cal D}_z Z_z\cdot U'_z}{C_{zzz}} \,. \end{equation} In terms of $V''$ the constraints will have the same form as before: \begin{equation} \< V'' , \tilde Y_{ij}\> = \< V'' , \tilde {\cal F}_{ab}^- \> = 0 \, . \end{equation} Then, one finds \begin{eqnarray} \< V',\tilde\Lambda_i\> &=& \ft 12{\rm i} ({\cal D}_z Z_z) g^{z\bar z}(\< \bar U_{\bar z}, \tilde Y_{ij}\> \varepsilon^{jk} \lambda_k^z + \< \bar U_{\bar z}, \sigma \cdot \tilde {\cal F}^-\> \lambda_i^z)\nonumber\\ && - \ft16 (C_{zzz} - {\cal D}_z{\cal D}_z Z_z)\, \varepsilon^{kl} \bar \lambda_k^z \sigma^{ab} \lambda_l^z \sigma_{ab} \lambda_i^z\,,\label{condL3n} \\ \< U'_z, \tilde \Lambda_i \>&=& \ft 12{\rm i} C_{zzz} g^{z\bar z} (\< \bar U_{\bar z} , \tilde Y_{ij} \> \varepsilon^{jk} \lambda_k^z + \< \bar U_{\bar z} , \sigma\cdot \tilde {\cal F}^-\> \lambda_i^z)\nonumber\\ & & + \ft16 {\cal D}_z C_{zzz} \cdot \varepsilon^{kl} (\bar \lambda_k^z \sigma_{ab} \lambda^z_l) \sigma^{ab} \lambda_i^z\,, \label{condL1n} \end{eqnarray} where we have used that \begin{equation} \<V,{\cal D}_z{\cal D}_zU_z\> = -C_{zzz} + {\cal D}_z{\cal D}_z Z_z \,. \end{equation} Note that these are the analogues of \eqn{condL1} and \eqn{condL3}.\\ Equation~\eqn{condL3n} can be replaced by \begin{equation} \< V'', \tilde \Lambda_i\> = \ft16 \,\varepsilon^{kl} \bar \lambda_k^z \sigma_{ab} \lambda_l^z \sigma^{ab} \lambda_i^z \frac{1}{C_{zzz}}( (-C_{zzz} + {\cal D}_z{\cal D}_z Z_z)\, C_{zzz} - {\cal D}_z Z_z \cdot {\cal D}_z C_{zzz} )\,. \end{equation} Using the notation \begin{eqnarray} {\cal O}_{zzzzz} &\equiv & g^{\prime z\bar z} \bar Z_{\bar z} \left[(-C_{zzz} + {\cal D}_z{\cal D}_z Z_z)\, C_{zzz} - {\cal D}_z Z_z \cdot {\cal D}_z C_{zzz} \right] \, , \end{eqnarray} the variation of (\ref{condL1n}) now gives \begin{eqnarray} \< U'_z, \tilde C\> &=& \ft14 C_{zzz} g^{z\bar z} g^{z\bar z}\varepsilon^{ik}\varepsilon^{jl} \< \bar U_{\bar z}, \tilde Y_{ij}\> \< \bar U_{\bar z}, \tilde Y_{kl}\> \nonumber\\ & & - \ft12 C_{zzz} g^{z\bar z} g^{z\bar z} \< \bar U_{\bar z}, \tilde {\cal F}^-_{ab}\> \< \bar U_{\bar z}, \tilde {\cal F}^{-ab}\>\nonumber\\ & & - \ft 12{\rm i} \varepsilon^{ik}\varepsilon^{jl} ( C_{zzz} \< \bar V, \tilde Y_{ij}\> + {\cal D}_z C_{zzz} \cdot g^{z\bar z} \< \bar U_{\bar z}, \tilde Y_{ij}\> ) \bar \lambda_k^z \lambda_l^z\nonumber\\ & & + \ft 12{\rm i} ( C_{zzz} \< \bar V, \tilde {\cal F}_{ab}^-\> + {\cal D}_z C_{zzz} \cdot g^{z\bar z} \< \bar U_{\bar z}, \tilde {\cal F}_{ab}^-\> )\, \varepsilon^{kl}(\bar \lambda_k^z \sigma^{ab} \lambda_l^z)\nonumber\\ & & -2{\rm i} C_{zzz} g^{z\bar z} \varepsilon^{ij} \bar \lambda_i^z \< \bar U_{\bar z}, \tilde \Lambda_j\> \nonumber\\ & & +\ft1{12} ( {\cal D}_z {\cal D}_z C_{zzz} + \ft12 {\cal O}_{zzzzz})\, \varepsilon^{ij} (\bar \lambda_i^z \sigma_{ab} \lambda_j^z) \varepsilon^{kl} (\bar \lambda_k^z \sigma^{ab} \lambda_l^z) \,. \end{eqnarray} Straightforward variation of constraint (\ref{condL3n}) gives \begin{eqnarray} \< V', \tilde C\> &=& \ft14 {\cal D}_z Z_z \cdot g^{z\bar z} g^{z\bar z}\varepsilon^{ik}\varepsilon^{jl} \< \bar U_{\bar z}, \tilde Y_{ij}\> \< \bar U_{\bar z}, \tilde Y_{kl}\> \nonumber\\ & & - \ft12 {\cal D}_z Z_z \cdot g^{z\bar z} g^{z\bar z} \< \bar U_{\bar z},\tilde {\cal F}^-_{ab}\> \< \bar U_{\bar z}, \tilde {\cal F}^{-ab}\>\nonumber\\ & & + \ft 12{\rm i} \varepsilon^{ik}\varepsilon^{jl} \left((C_{zzz} - {\cal D}_z{\cal D}_z Z_z) g^{z\bar z} \< \bar U_{\bar z}, \tilde Y_{ij}\> - {\cal D}_z Z_z \cdot \< \bar V, \tilde Y_{ij}\> \right) \bar \lambda_k^z \lambda_l^z\nonumber\\ & & - \ft 12{\rm i} \left((C_{zzz} - {\cal D}_z{\cal D}_z Z_z)\, g^{z\bar z} \< \bar U_{\bar z}, \tilde {\cal F}_{ab}^-\> - {\cal D}_z Z_z \cdot \< \bar V, \tilde {\cal F}_{ab}^-\> \right) \varepsilon^{kl}(\bar \lambda_k^z \sigma^{ab} \lambda_l^z)\nonumber\\ & & - 2{\rm i} {\cal D}_z Z_z \cdot g^{z\bar z} \varepsilon^{ij} \bar \lambda_i^z \< \bar U_{\bar z}, \tilde \Lambda_j\> \nonumber\\ & & -\ft1{12} (2 {\cal D}_z C_{zzz} - {\cal D}_z {\cal D}_z {\cal D}_z Z_z) \varepsilon^{ij} (\bar \lambda_i^z \sigma_{ab} \lambda_j^z) \varepsilon^{kl} (\bar \lambda_k^z \sigma^{ab} \lambda_l^z)\,. \end{eqnarray} This can be rewritten in \begin{eqnarray} \< V'', \tilde C\> & = & \frac{{\rm i} g^{z\bar z}}{2C_{zzz}} \left[ (C_{zzz} - {\cal D}_z{\cal D}_z Z_z)\, C_{zzz} + {\cal D}_z C_{zzz}\cdot{\cal D}_z Z_z\right] \nonumber \\ & &\times \left[\varepsilon^{ik}\varepsilon^{jl}\<\bar U_{\bar z}, \tilde Y_{ij}\>\bar \lambda^z_k\lambda^z_l -\<\bar U_{\bar z}, \tilde {\cal F}_{ab}\>\varepsilon^{kl} (\bar \lambda^z_k\sigma^{ab}\lambda^z_l)\right] \nonumber \\ & & -\ft{1}{12}\left[\left(2{\cal D}_z C_{zzz} - {\cal D}_z{\cal D}_z{\cal D}_z Z_z \right) + \frac{{\cal D}_z Z_z}{C_{zzz}}\left({\cal D}_z {\cal D}_z C_{zzz} + \ft12 {\cal O}_{zzzzz}\right)\right] \nonumber \\ & & \times \varepsilon^{ij} (\bar \lambda_i^z \sigma_{ab} \lambda_j^z) \varepsilon^{kl} (\bar \lambda_k^z \sigma^{ab} \lambda_l^z)\,. \end{eqnarray} \section{The generalized Bianchi identities combined with the special K\"ahler\ constraints}\label{s:gBI} In this section we start by imposing the reduction constraints on the chiral multiplets. Because the constraints on the field strengths are Bianchi identities, this linear multiplet of constraints is called the generalized Bianchi identities. We combine these constraints with the special K\"ahler constraints of section \ref{s:kahler}. Together they give the field equations of $n$ vector multiplets and expressions for the auxiliary fields $\chi_i$ and $D$. We derive this for the generic case $\< V, U_\alpha \> =0$. We comment on the supergravity equations of motion in this generic case. Finally, we give the equations for the special case where $\< V, U_\alpha \> \neq 0$. \subsection{The field equations for the generic case}\label{s:genBI} To see what follows from equations~(\ref{bian1})--(\ref{bian4}), we take the symplectic inner product of these equations with the basis ${\cal W}$. The $4$ components of equation~(\ref{expn>1}) give $4$ equations for each constraint. \par {}From the first identity we learn that the section $\tilde Y_{ij}$ is totally constrained, as it should be because it is auxiliary. We have \begin{equation} \tilde Y_{ij} = -{\rm i} g^{\alpha\bar\alpha} \left[ C_{\alpha\beta\gamma} \bar \lambda_i^\beta \lambda_j^\gamma \bar U_{\bar\alpha} - \bar C_{\bar\alpha\bar\beta\bar\gamma} \varepsilon_{ik} \varepsilon_{jl} \bar \lambda^{\bar\beta k} \lambda^{\bar\gamma l} U_\alpha \right ]\,. \label{coY} \end{equation} \par It is more interesting to take a look at (\ref{bian2}). Taking the symplectic inner product of this equation with all components of the basis ${\cal W}$ (\ref{basiscW}), and using the special K\"ahler constraints of section~\ref{ss:susyKaconstr} and \eqn{coY} gives: \begin{eqnarray} \label{coL1} \< \bar V, \tilde \Lambda_i\> &=& 0\,,\nonumber\\ \< \bar U_{\bar \alpha} , \tilde \Lambda_i\> &=& - \varepsilon_{ij} \bar C_{\bar\alpha\bar\beta\bar\gamma} /\!\!\!\!\nabla \bar z^{\bar \beta}\cdot \lambda^{\bar\gamma j}\,,\nonumber\\ 0 &=& \chi^i - \ft23 \sigma^{\mu\nu} ({\cal D}_\mu \psi^i_\nu - \ft18 \sigma\cdot T^{ij} \gamma_\mu \psi_{\nu j}) - \ft12 g_{\alpha\bar\beta} \hbox{\ooalign{$\displaystyle\partial$\cr$/$}} z^\alpha\cdot \lambda^{i\bar\beta} \nonumber\\ && + \ft 12{\rm i} \gamma^\mu (/\!\!\!\! {\cal Q} - /\!\!\!\! A) \psi^i_\mu - \ft 14{\rm i} \varepsilon^{ij} \gamma^\mu \< V,\sigma\cdot \tilde {\cal F}^+\> \psi_{\mu\,j} + \ft 1{12}{\rm i} C_{\alpha\beta\gamma} \varepsilon^{ij} \varepsilon^{kl} (\bar \lambda^\alpha_k \sigma^{ab} \lambda_l^\beta) \sigma_{ab} \lambda_j^\gamma\,,\nonumber\\ 0 &=& {\rm i} g_{\alpha\bar\beta} \left(/\!\!\!\!\nabla \lambda^{i\bar\beta} + \ft 12{\rm i} (/\!\!\!\! {\cal Q} - /\!\!\!\! A)\lambda^{i\bar\beta}\right) + \ft 12{\rm i} \varepsilon^{ij} C_{\alpha\beta\gamma} g^{\beta\bar\beta}\< \bar U_{\bar\beta}, \sigma \cdot \tilde {\cal F}^-\> \lambda^\gamma_j \nonumber\\ &&+\ft 12{\rm i} C_{\alpha\beta\gamma} \bar C_{\bar\alpha\bar\beta\bar\gamma}g^{\gamma\bar\gamma} (\bar\lambda^{i\bar\alpha} \lambda^{j\bar\beta}) \lambda_j^\beta + \ft16 {\cal D}_\alpha C_{\beta\gamma\delta}\cdot \varepsilon^{ij}\varepsilon^{kl} (\bar \lambda_k^\beta \sigma^{ab} \lambda_l^\gamma )\sigma_{ab} \lambda_j^\delta\,,\label{bianchi2} \end{eqnarray} where \begin{eqnarray} \nabla_\mu \lambda_i^\alpha &=& \partial_\mu \lambda_i^\alpha -\ft12 \omega_\mu^{ab} \sigma_{ab} \lambda_i^\alpha +\ft12 {\cal V}_{\mu\, i}^{\ \ \ j}\lambda_j^\alpha + \Gamma_{\beta\gamma}^\alpha \partial_\mu z^\beta \cdot\lambda_i^\gamma - \ft 12{\rm i} {\cal Q}_\mu \lambda_i^\alpha\nonumber\\ && - /\!\!\!\!\nabla z^\alpha \cdot\psi_{\mu i} + \ft 12{\rm i} g^{\alpha\bar\alpha} \left[ \< \bar U_{\bar\alpha}, \tilde Y_{ij}\> + \varepsilon_{ij} \< \bar U_{\bar \alpha}, \sigma \cdot \tilde {\cal F}^-\> \right] \psi_\mu^j\,, \end{eqnarray} and \begin{equation} {\cal Q}_\mu = - \ft 12{\rm i} (\partial_\alpha K \cdot\partial_\mu z^\alpha - \mbox{h.c.}) \end{equation} is the K\"ahler 1-form. \par The first two equations in (\ref{bianchi2}) imply with (\ref{condL1}) and (\ref{condL3}) that all components of $\tilde \Lambda_i$ are expressed in terms of other fields, and thus they contain no independent degrees of freedom. The third expresses $\chi^i$ in terms of other fields. In a superconformal calculation using a Lagrangian, this expression for $\chi^i$ can be found from the equation of motion of the fermion of the compensating vector multiplet. The fourth equation is the field equation for $n$ fermion doublets $\lambda^\beta_i$.\vspace{5mm} \par We now proceed with the analysis of (\ref{bian3}). We first repeat that (\ref{dUOF}) implies that there are $(n+1)$ independent antisymmetric tensors in the symplectic vector ${\tilde {\cal F}}_{ab}$. Apart from these there is another antiself-dual tensor in the Weyl multiplet $T^{ij}_{ab}$. A few definitions make (\ref{bian3}) more transparent. First define the combination in brackets as \begin{equation} \label{FTsec1} \tilde {\hat F}_{ab} = \tilde {\cal F}_{ab} + \ft14 V T_{ab\, ij}\varepsilon^{ij} + \ft14 \bar V T^{ij}_{ab} \varepsilon_{ij}\,. \end{equation} Then we take out covariantization terms: \begin{equation} \label{FHsec1} \tilde F_{ab} =\tilde {\hat F}_{ab} - 2 (\bar {\tilde \Omega^i} \gamma_{[a} \psi_{b]}^j \varepsilon_{ij} + \bar V \bar \psi_a^i \psi^j_b \varepsilon_{ij} + \mbox{h.c.})\,. \end{equation} This is chosen such that covariant derivatives in (\ref{bian3}) can be rewritten as ordinary derivatives, and the equation reduces to \begin{equation} \partial_\mu \varepsilon^{\mu\nu\rho\sigma} {\tilde F}_{\rho\sigma} = 0\,. \label{realBianchi} \end{equation} Applying this on the $n+1$ independent components of $\tilde F_{\mu \nu }$, implies that they can be expressed in terms of $n+1$ vectors. The other $(n+1)$ equations of (\ref{realBianchi}) are the equivalent of field equations for these vectors. Here, it is clear how our formulation keeps the symplectic covariance. Only in the interpretation do we distinguish one half of the equations as Bianchi identities and the others as field equations. These could have been interchanged giving the `magnetic dual formulation'. Also the fact of whether or not a prepotential exists is hidden here. The difference is seen only when breaking the symplectic formulation in finding an explicit solution of equations~(\ref{dUOF}). If the $(n+1) \times (n+1)$-matrix, formed by the upper part of $(V\,,U_\alpha )$ is invertible, then (\ref{dUOF}) expresses the $(n+1)$ lower components of ${\cal F}^-_{ab}$ in terms of the upper ones. This is the case where there is a prepotential. When this matrix is not invertible\footnote{ As proven in \cite{CRTVP}, it is only the matrix $\left(f_\alpha^I \bar X^I\right)$ that is always invertible.}, then one can still solve (\ref{dUOF}) for other $(n+1)$ components of $\tilde {\cal F}^-_{ab}$. \par We thus conclude that we have $n+1$ on-shell vectors and their field equations also depend on the $6$ degrees of freedom of the tensor $T^{ij}_{ab}$ of the Weyl multiplet.\vspace{5mm} \par Now let us have a look at \eqn{bian4}. It involves the covariant Laplacian, \begin{eqnarray} \Box \bar V &\equiv& \eta^{mn} D_m D_n \bar V\nonumber\\ &=& e^{-1} \partial_\mu (e D^\mu \bar V) + (b^\mu - {\rm i} A^\mu) D_\mu \bar V + f_\mu^\mu \bar V + 2\bar \psi^{i[\mu} \gamma_\mu \psi_i^{\nu]}D_\nu\bar V\nonumber\\ & & - \bar \psi^\mu_i D_\mu \tilde \Omega^i + \ft12 \bar \phi_i^\mu \gamma_\mu \tilde \Omega^i + \ft18 \bar \psi^i_\mu \gamma^\mu \sigma\cdot T_{ij} \tilde \Omega^j - \ft32 \bar \psi^i_\mu \gamma^\mu \chi_i \bar V\,. \label{boxVbar} \end{eqnarray} To derive this expression, we used a theorem on covariant derivatives in the second reference of \cite{sctc}. We can again take the symplectic inner product of (\ref{bian4}) with ${\cal W}$: \begin{eqnarray}\label{eomsc} \< \bar V , \tilde C\> &=& 0\,,\nonumber\\ \< \bar U_{\bar\alpha}, \tilde C\> &=& - 2\bar C_{\bar\alpha\bar\beta\bar\gamma} \nabla_\mu \bar z^{\bar\beta} \cdot\nabla^\mu \bar z^{\bar \gamma} +\ft 14 \bar C_{\bar\alpha\bar\beta\bar\gamma}(\bar\lambda^{k\bar\beta}\sigma\cdot T_{kl} \lambda^{l\bar\gamma}) \,,\nonumber\\ 0 &=& -2e^{-1}\partial_\mu \left(e({\cal Q}^\mu - A^\mu) \right) + 2{\rm i} g_{\alpha\bar\alpha}\partial_\mu z^\alpha\cdot \nabla^\mu{\bar z}^{\bar\alpha} -2{\rm i} g_{\alpha\bar\alpha}\partial_\mu z^\alpha \cdot(\bar \psi_i^\mu \lambda^{i\bar\alpha}) \nonumber\\ &&-2{\rm i}({\cal Q}^\mu - A^\mu)({\cal Q}_\mu - A_\mu) +2{\rm i} f_\mu^\mu -4{\bar \psi}^{i[\mu}\gamma_\mu\psi_i^{\nu]}({\cal Q}_\nu - A_\nu) \nonumber\\ &&+{\bar \psi}_i^\mu \< V,\sigma\cdot \tilde {\cal F}^+ \> \varepsilon^{ij}\psi_{\mu j} +2{\rm i}{\bar \psi}_i^\mu \phi^i_\mu -2 \bar \psi_i^\mu(/\!\!\!\! {\cal Q} - /\!\!\!\! A)\psi^i_\mu \nonumber \\ &&-3{\rm i}{\bar\psi}^i_\mu \gamma^\mu \chi_i +\ft 14 \< V, \tilde{\cal F}^+_{\mu\nu}\> T^{\mu\nu}_{ij}\varepsilon^{ij} -\ft 12{\rm i} C_{\alpha\beta\gamma}g^{\alpha\bar\alpha} \bar C_{\bar\alpha\bar\beta\bar\gamma}(\bar\lambda^{\bar\beta k}\lambda^{\bar\gamma l})(\bar\lambda^\beta_k\lambda^\gamma_l)\nonumber \\ &&-\ft 12{\rm i} C_{\alpha\beta\gamma}g^{\alpha\bar\alpha} \< {\bar U}_{\bar\alpha}, \tilde {\cal F}^-_{ab}\> \varepsilon^{kl}(\bar\lambda^\beta_k \sigma^{ab} \lambda_l^\gamma) -\ft 16 {\cal D}_\alpha C_{\beta\gamma\delta}\cdot\varepsilon^{ij}(\bar\lambda^\alpha_i\sigma_{ab}\lambda_j^\beta)\, \varepsilon^{kl} (\bar\lambda^\gamma_k\sigma^{ab}\lambda_l^\delta)\,,\nonumber\\ 0 &=& -2{\rm i} e^{-1}g_{\alpha\bar\alpha}\partial_\mu (e \nabla^\mu \bar z^{\bar\alpha}) +4({\cal Q}_\mu - A_\mu)g_{\alpha\bar\alpha}\nabla^\mu\bar z^{\bar\alpha} +g_{\alpha\bar\alpha}({\cal Q}_\mu - A_\mu)(\bar\psi_i^\mu\lambda^{i\bar\alpha}) \nonumber \\ &&-2{\rm i} g_{\alpha \bar \alpha }\Gamma_{\bar\beta\bar\gamma }^{\bar\alpha } \partial_\mu \bar z^{\bar\beta} \cdot\nabla^\mu{\bar z}^{\bar\gamma } -4{\rm i} g_{\alpha\bar\alpha}{\bar \psi}^{i[\mu} \gamma_\mu \psi_i^{\nu]}\nabla_\nu\bar z^{\bar\alpha} +2{\rm i} g_{\alpha\bar\alpha}\bar \psi_i^\mu\nabla_\mu \lambda^{i\bar\alpha}\nonumber \\ &&-{\rm i} g_{\alpha\bar\alpha}\bar\phi_i^\mu \gamma_\mu \lambda^{i\bar\alpha} +2C_{\alpha\beta\gamma}\varepsilon^{ik}\varepsilon^{jl}(\bar\lambda^\beta_k\lambda^\gamma_l)(\bar \psi^\mu_i\psi_{j\mu}) +2\bar\psi^\mu_i \< U_\alpha, \sigma \cdot \tilde{\cal F}^+\varepsilon^{ij}\> \psi_{\mu j}\nonumber\\ &&-\ft 14{\rm i} g_{\alpha\bar\alpha}\bar\psi^i_\mu\gamma^\mu\sigma\cdot T_{ij}\lambda^{j\bar\alpha} -6{\rm i} g_{\alpha\bar\alpha}\bar \chi_i\lambda^{i\bar\alpha}\nonumber \\ &&+\ft 14 \< U_\alpha, \tilde {\cal F}^+_{\mu\nu}\> T^{\mu\nu}_{ij}\varepsilon^{ij} +\ft 14 C_{\alpha\beta\gamma}g^{\beta\bar\beta}g^{\gamma\bar\gamma}\varepsilon_{ik} \varepsilon_{jl}\bar C_{\bar\beta\bar\delta\bar\varepsilon} (\bar\lambda^{i\bar\delta}\lambda^{j\bar\varepsilon})\bar C_{\bar\gamma\bar\alpha\bar\zeta} (\bar\lambda^{k\bar\alpha}\lambda^{l\bar\zeta}) \nonumber \\ &&-\ft 12 C_{\alpha\beta\gamma}g^{\beta\bar\beta}g^{\gamma\bar\gamma}\< \bar U_{\bar\beta}, \tilde {\cal F}^-_{ab}\> \< \bar U_{\bar\gamma}, \tilde {\cal F}^{-ab}\> -\ft 12{\rm i} {\cal D}_\alpha C_{\beta\gamma\delta}\cdot g^{\delta\bar\delta}\bar C_{\bar\delta\bar\beta\bar\gamma} (\bar\lambda^{k\bar\beta}\lambda^{l\bar\gamma})(\bar\lambda_k^\beta\lambda^\gamma_l) \nonumber \\ &&+\ft 12{\rm i} \left[ C_{\alpha\beta\gamma}\< \bar V, \tilde {\cal F}^-_{ab}\> +{\cal D}_\alpha C_{\beta\gamma\delta}\cdot g^{\delta\bar\delta} \< \bar U_{\bar\delta}, \tilde {\cal F}^-_{ab}\> \right] \varepsilon^{ij}(\bar\lambda^\beta_i\sigma^{ab}\lambda_j^\gamma) \\ &&+\ft 1{12} {\cal D}_\alpha{\cal D}_\beta C_{\gamma\delta\varepsilon} \cdot\varepsilon^{ij}(\bar\lambda^\beta_i\sigma_{ab}\lambda_j^\gamma) \varepsilon^{kl} (\bar\lambda^\delta_k\sigma^{ab}\lambda_l^\varepsilon) -2{\rm i} C_{\alpha\beta\gamma}g^{\beta\bar\beta}\bar C_{\bar\beta\bar\gamma\bar\delta}\bar\lambda^\gamma_i\,/\!\!\!\!\nabla\bar z^{\bar\gamma}\cdot\lambda^{i\bar\delta}\,.\nonumber \end{eqnarray} The first two equations, together with (\ref{UC}) and (\ref{VC}) (which can be simplified using the second equations in (\ref{bianchi2})) determine that $\tilde C$ is completely determined in terms of other fields. The real and imaginary part of the third equation have both to be $0$. The real part constrains the divergence of $\mathcal{Q}_\mu - A_\mu$, and the imaginary part gives an expression for the $D$-field of the Weyl multiplet. ($D$ is hidden in the $f^\mu_\mu$-term.) The fourth equation of the expansion in terms of ${\cal W}$ gives the field equations for $n$ complex scalars. So we find the same structure in the equations as for the fermions: $n+1$ equations express $\tilde C$ in terms of other fields, while the $n+1$ other equations give the field equations for $n$ complex scalars $z^\alpha$, an expression for $D$ and a constraint for $({\cal Q}_\mu - A_\mu)$. The degrees of freedom are described in table~\ref{dofAfterConstrBianchi}. \begin{table}[ht] \begin{center} \begin{tabular}{|c|c|l|}\hline fields&d.o.f.&comments\\ \hline\hline \multicolumn{3}{|c|}{The Gravity Multiplet}\\ \hline $e_\mu^{\ a}$&6&16 - 4(translation) - 6(Lorentz)\\ \hline $A_\mu$&3&$\partial ^\mu A_\mu$ constrained\\ \hline ${\cal V}_\mu{}^j{}_i$&9&12 - 3($SU(2)$)\\ \hline $\psi_\mu^i$&24&32 - 8($Q$-supersymmetry)\\ \hline $T^{ij}_{ab}$&6&$n$ on-shell and $2$ off-shell vectors \\ \hline $\chi_i$&0& expressed in terms of other fields\\ \hline $D$&0&expressed in terms of other fields\\ \hline\hline \end{tabular} \caption{Off-shell degrees of freedom after the special K\"ahler\ constraints and the generalized Bianchi identities: $24 + 24$ d.o.f.. All other variables are expressed in terms of these fields or have a field equation (for $n$ complex scalars, $n$ doublet spinors and $n+1$ vectors). \label{dofAfterConstrBianchi}} \end{center}\end{table} \subsection{Comments on the supergravity equations} Using the results of the previous section, we comment on the appearance of equations of motion for the remaining $24+24$ components of table~\ref{dofAfterConstrBianchi} from one symplectic invariant constraint \begin{equation} \< V , \tilde {\cal F}^+_{ab} \> \approx 0\label{FET}\,, \end{equation} which gives rise to a $24+24$ `current' multiplet. The $\approx$-sign is used to denote that we only expose the linear terms. This already shows the essential features of this symplectic covariant formulation. With the linearized approximation we mean that we keep terms with an arbitrary power of undifferentiated scalar fields or metric, but only linear in other fields. In a full treatment of $N=2$ supergravity couplings the right-hand side of (\ref{FET}) would, for example, contain an additional coupling to hypermultiplets. \par To discuss the supersymmetry partners of (\ref{FET}), we derive a new $N=2$ multiplet with 24 + 24 components. The multiplet starting with the symplectic expression $\<V , \tilde {\cal F}^+_{ab} \> $ is a supergravity realization of this multiplet. As shown below the supermultiplet of constraints derived from (\ref{FET}) is only equivalent to the supergravity equations of motion, up to integration `constants'. These $8+8$ remaining unknowns can be determined when one of the three possibilities of a second compensating multiplet is introduced as in \cite{auxf,dWVHVP, dWLVP}. In our approach this is the place where the second compensating multiplet, which is also needed for consistency in the Lagrangian formulation, comes into play. \subsubsection{A restricted chiral self-dual tensor multiplet} The supermultiplet structure of the `current' multiplet from (\ref{FET}) is that of a chiral self-dual tensor multiplet, \begin{eqnarray} W_{ab}^+ &=& A_{ab}^+ + \bar \theta^i \psi_{abi} + \ft14 \bar \theta^i \theta^j B_{abij} + \ft14 \varepsilon_{ij} \bar \theta^i \sigma_{cd} \theta^j F_{ab}{}^{cd} \nonumber\\ && + \ft16 \varepsilon_{ij} (\bar \theta^i \sigma_{cd} \theta^j) \bar \theta^k \sigma^{cd} \chi_{abk} + \ft1{48} (\varepsilon_{ij} \bar \theta^i \sigma_{cd} \theta^j)^2 C^+_{ab}\,. \label{chasdsup} \end{eqnarray} It has the following field content. $A_{ab}^+$ is a self-dual complex tensor with 6 degrees of freedom. $\psi_{abi}$ has 24 left-handed fermionic components. The tensor $B_{abij}$ has 18 components. The tensor $F_{ab}{}^{cd}$ is self-dual in its first and antiself-dual in its second pair of indices, leading to 9 complex components. It also satisfies the following properties: \begin{eqnarray} F_{ab,cd}+F_{cd,ab} &=& \ft12 \varepsilon _{ab}{}^{ef}\left( F_{ef,cd} - F_{cd,ef}\right)\,, \nonumber\\ F_{ab,cd}-F_{cd,ab} & = & \varepsilon _{abe[d}\left( F^e{}_{c]} +F_{c]}{}^e\right)\,, \nonumber\\ F_{ab,cd} &=& \delta_{a[c}F_{d]b} -\delta_{b[c} F_{d]a} - \varepsilon_{abe[c}F^e{}_{d]}\,,\nonumber\\ F_{[ab]} & = & 0 \nonumber\,,\\ F^a{}_{a} & = & 0\label{Fprop} \,, \end{eqnarray} where \begin{equation} F_a{}^c = F_{ab}{}^{cd} \delta^b_d\,. \end{equation} A general component of this self-dual--antiself-dual tensor $F_{ab}{}^{cd}$ can thus be written in terms of the traceless symmetric part $F_{(ab)}$ with 9 components. The fermion $\chi_{abi}$ has again 24 left-handed components and $C_{ab}^+$ has 6. So, this is a chiral multiplet with $48 + 48$ components. \par The transformation rules of this multiplet are the same as for a chiral multiplet with a complex scalar as lowest component, but with the components replaced straightforwardly: \begin{eqnarray} \delta A^+_{ab} & = & \bar{\epsilon}^i \psi_{abi}\,, \nonumber \\ \delta \psi_{abi} &=& \hbox{\ooalign{$\displaystyle\partial$\cr$/$}} A_{ab}^+ \epsilon_i + \ft12 B_{abij}\epsilon^j + \ft12 \sigma_{cd} F_{ab}{}^{cd} \varepsilon_{ij} \epsilon ^j\,, \nonumber\\ \delta B_{abij} &=& 2 \bar{\epsilon}_{(i} \hbox{\ooalign{$\displaystyle\partial$\cr$/$}} \psi_{abj)} - 2\bar{\epsilon}^k \chi_{ab(i} \varepsilon_{j)k}\,,\nonumber\\ \delta F_{ab}{}^{cd}&=&\varepsilon^{ij} \bar{\epsilon}_i \hbox{\ooalign{$\displaystyle\partial$\cr$/$}} \sigma^{cd} \psi_{abj} + \bar{\epsilon}^i \sigma^{cd} \chi_{abi}\,, \nonumber\\ \delta \chi_{abi} &=&-\ft12 \sigma_{cd} F_{ab}^{cd} \stackrel{\leftarrow} {\hbox{\ooalign{$\displaystyle\partial$\cr$/$}}} \epsilon_i - \ft 12 \hbox{\ooalign{$\displaystyle\partial$\cr$/$}} B_{abij} \varepsilon^{jk} \epsilon_k + \ft12 C_{ab}^+ \varepsilon_{ij}\epsilon^j\,, \nonumber\\ \delta C_{ab}^+ &=& - 2 \varepsilon^{ij} \bar{\epsilon}_i \hbox{\ooalign{$\displaystyle\partial$\cr$/$}} \chi_{abj}\,.\label{susy} \end{eqnarray} Since we have broken superconformal symmetry to super-Poincar\' e and $SU(2)$, we only need a super-Poincar\' e version of this multiplet. Note that it cannot be extended to a superconformal one. The commutator of a supersymmetry and a special supersymmetry has to give a Lorentz transformation that can never be realized because of the duality and chirality properties of the spinors. For this reason, it is only possible to construct an antiself-dual chiral tensor multiplet, realizing the superconformal algebra, as given in \cite{BDRDW}. \par To study the field equations of the fields of table~\ref{dofAfterConstrBianchi}, we need a multiplet with $24+24$ components. A suitable multiplet of constraints is: \begin{eqnarray} 0 & = & \partial^a(B_{abij} + \varepsilon_{ik}\varepsilon_{jl}{\bar B}_{ab}^{kl})\,, \label{conB}\\ 0 & = & \partial^a( \chi_{ab}^i -\varepsilon^{ij}\hbox{\ooalign{$\displaystyle\partial$\cr$/$}} \psi_{abj})\,, \label{conchi}\\ 0 & = & \partial^a( C^-_{ab} - \Box A^+_{ab})\,,\label{conC}\\ 0 & = & \partial^a\partial_c(F_{ab}{}^{cd} + {\bar F}_{ab}{}^{cd})\,. \label{conF} \end{eqnarray} These are the analogues of the constraints (5.4) in \cite{BDRDW}. This set contains $(9+6+9) + 24$ equations. The constraint for $F_{ab}{}^{cd}$ splits up in a part symmetric in $(bd)$ (6 independent equations) and an antisymmetric part in $[bd]$ (3 independent equations), which correspond to the real and imaginary part of $F_{ac}$: \begin{eqnarray}\label{conF'} 0 & = & -\partial^c\left(\partial_{(b}(F_{d)c} +{\bar F}_{d)c})\right)+\ft 12 \delta_{bd}\partial^a\partial^c(F_{ac}+{\bar F}_{ac})+\ft12 \Box(F_{bd}+{\bar F}_{bd})\nonumber \\ & & +\ft 12\varepsilon_{bdae}\partial^a\partial^c(F^e{}_c - {\bar F}^e{}_c)\,. \end{eqnarray} As far as we know, this reduced multiplet is a new representation of the rigid $N=2$ algebra. An explicit supergravity realization of this reduced multiplet is given by \begin{eqnarray} A^+_{ab} &=& \langle V, {\tilde {\cal F}}_{ab}^+\rangle\,,\nonumber\\ \psi_{abi} &\approx& -{\rm i}\varepsilon_{ij}\gamma^\rho\sigma_{ab}\phi_\rho^j\,,\nonumber\\ B_{abik} & \approx & 2{\rm i}\varepsilon_{ij}R_{SU(2)}{}_{ab}^+{}^j{}_k\,,\nonumber \\ F_{ab}{}^{cd} & \approx & 2 \delta^{[c}_{[a} \left(\partial_{b]}({\cal Q}^{d]}-A^{d]}) + (\partial^{d]}({\cal Q}_{b]}-A_{b]})) -2{\rm i}{\cal R}_{b]}{}^{d]} +\ft 12{\rm i}\delta_{b]}^{d]}{\cal R}\right)\nonumber \\ & & -\varepsilon^{cd}{}_{ef}\delta^{[e}_{[a} \left(\partial_{b]}({\cal Q}^{f]}-A^{f]}) + (\partial^{f]}({\cal Q}_{b]}-A_{b]})) -2{\rm i} {\cal R}_{b]}{}^{f]} +\ft 12{\rm i}\delta_{b]}^{f]}{\cal R}\right)\,. \label{currmult} \end{eqnarray} In deriving this multiplet we used the constraints of sections~\ref{s:kahler} and \ref{s:gBI}. The expression for $B_{abij}$ satisfies constraint (\ref{conB}), which is a Bianchi identity that expresses the existence of $SU(2)$-vectors. The expression for $F_{ab}{}^{cd}$ fulfils (\ref{Fprop}). It also satisfies (\ref{conF}) when the third equation of \eqn{eomsc} for $({\cal Q}_\mu - A_\mu)$ is used. Therefore, the multiplet derived from $\langle V,{\tilde{\cal F}}_{ab}^+\rangle$ has 24 + 24 components. \subsubsection{Some comments on the multiplet of equations from $\langle V,{\tilde{\cal F}}_{ab}^+\rangle \approx 0$} Putting the `current' multiplet (\ref{currmult}) to zero, will give rise to some supergravity field equations. These are 24 + 24 equations for the 24 + 24 remaining degrees of freedom of table~\ref{dofAfterConstrBianchi}. The counting in this table subtracts the gauge degrees of freedom. The multiplet here is a multiplet of curvatures and the counting is equivalent if we take into account the Bianchi identities. \par However, our equations are not equivalent to the complete supergravity equations of motion. They differ modulo `integration constants'. These can be determined when a second compensating multiplet is coupled \cite{dWVHVP,auxf}. Since this step is independent of the symplectic formulation of the coupling of vector multiplets to supergravity, we do not treat it here. \par Let us give a brief discussion of the content of the equations following from \eqn{FET}. Equation \eqn{FET} reduces 6 degrees of freedom. It expresses the `graviphoton' field strength $T_{abij}$ as a combination of the $n+1$ on-shell vectors obtained above. It is the symplectic expression for the algebraic equation of motion that one finds in the Lagrangian approach, (4.11) in \cite{dWLVP}. Using \eqn{bianchi2} in (\ref{dependent}) with \begin{equation} R^{\mu i} \equiv e^{-1}\varepsilon^{\mu\nu\rho\sigma}\gamma_5\gamma_\nu \left({\cal D}_\rho\psi_\sigma^i -\ft 18 \sigma\cdot T^{ij}\gamma_\rho\psi_{\sigma j}\right) \end{equation} in the second component of the current multiplet gives that \begin{equation}\label{phi0} \phi_\rho^i = R_\rho^i -\ft 14 \gamma_\rho\gamma\cdot R_\rho^i \approx 0\,, \label{FEpsi} \end{equation} the traceless part of the field equation of the gravitini. Therefore, this equation cannot determine the trace-part $\gamma\cdot R^i$. However, combining \eqn{FEpsi} with the Bianchi identity for the gravitino field strength $\partial^\mu R_\mu^i \approx 0$, yields \begin{equation} \hbox{\ooalign{$\displaystyle\partial$\cr$/$}} \gamma\cdot R^i\approx 0\,, \end{equation} which determines $\gamma \cdot R^i$ in terms of 8 `integration constants'. The $B_{abij}$ component yields \begin{equation} R_{SU(2)}{}_{ab}{}^i{}_j \approx 0\,. \end{equation} Together with the Bianchi identity for the $SU(2)$ curvature it states that the gauge fields ${\cal V}_\mu{}^i{}_j$ are pure gauge, i.e.\ \begin{equation} {\cal V}_\mu{}^i{}_j = (\varphi^{-1} \partial_\mu \varphi)^i{}_j\,, \end{equation} where $\varphi$ is a group element of $SU(2)$. The three local parameters defining $\varphi$ are left undetermined. $F_{ab}{}^{cd}$ has its components in the traceless part of $F_{(ac)}$. {}From $F_{ab}{}^{cd} \approx 0$ follows \begin{equation} F_{ac} = 2\partial_{(a}({\cal Q}_{c)}-A_{c)}) - 2{\rm i}{\cal R}_{ac} +\ft 12{\rm i} g_{ac}{\cal R}\approx 0\,. \end{equation} The imaginary part is the traceless part of the Einstein equation. Again we cannot determine the scalar curvature ${\cal{R}}$ from this equation. However, combining this equation with the Bianchi identity for the Einstein tensor \begin{equation} \partial^a({\cal R}_{ab} -\ft 12 g_{ab}{\cal R}) = 0\,, \end{equation} gives \begin{equation} \partial^a {\cal{R}} \approx 0 \end{equation} and again $\cal{R}$ is determined up to a constant. The real part of the $F$-component gives that $A_\mu \approx {\cal Q}_\mu $ up to a constant vector. Also in the Lagrangian approach \cite{dWLVP}, one finds \begin{equation} A_\mu \approx {\cal Q}_\mu \,. \end{equation} The additional $8+8$ remaining unknowns can be determined through the field equations of a second compensating multiplet. This concludes the short discussion of the supergravity equations of motion. \subsection{The field equations for the special case} In this subsection, the expressions of section \ref{s:genBI} are generalized for the case $n=1$ where further $Z_z = \< V, U_z \> \neq 0$. This is the case that was excluded by the former definitions and where our less restrictive definitions becomes relevant. The equations are found by expanding the constraints in terms of the basis of symplectic vectors using the methods mentioned at the end of the appendix.\vspace{5mm} \par The section $\tilde Y_{ij}$ remains totally constrained: \begin{eqnarray} \tilde Y_{ij} & = & g'^{z\bar z}\Big( -{\rm i}\varepsilon_{ik}\varepsilon_{jl} g_{z\bar z}{\cal D}_{\bar z} \bar Z_{\bar z}\cdot \bar\lambda^{k\bar z} \lambda^{l\bar z}V +{\rm i} g_{z\bar z}{\cal D}_z Z_z\cdot\bar\lambda^z_i\lambda^z_j\bar V \nonumber\\ && \hspace{1cm}+{\rm i}\varepsilon_{ik}\varepsilon_{jl} \bar C_{\bar z\bar z\bar z} \bar\lambda^{k{\bar z}}\lambda^{l{\bar z}} U_z -{\rm i} C_{zzz}\bar\lambda^z_i\lambda^z_j\bar U_{\bar z}\Big)\,. \end{eqnarray} This equation reduces to the former equation~(\ref{coY}) when $\< V, U_z \> =0$. \vspace{5mm} \par The equations that can be derived from the constraints for the fermions are the following ones: \begin{eqnarray} \<\bar V',\tilde \Lambda _i\>&=& - {\cal D}_{\bar z}\bar Z_{\bar z}\cdot \varepsilon_{ij} /\!\!\!\!\nabla \bar z\cdot \lambda^{j\bar z}\,, \nonumber \\ \<\bar U'_{\bar z},\Lambda_i\>&=& - \varepsilon_{ij} \bar C_{\bar z\bar z\bar z} /\!\!\!\!\nabla {\bar z} \cdot\lambda^{j{\bar z}}+ \ft12\gamma ^\mu\varepsilon_{ij} \left( \<\bar U'_{\bar z}, \tilde Y^{jk}+\sigma \cdot\tilde {\cal F}^+\varepsilon^{jk}\> \right) \psi_{\mu k} \,, \nonumber \\ 0 & = & \chi^i - \ft23 \sigma^{\mu\nu} ({\cal D}_\mu \psi^i_\nu - \ft18 \sigma\cdot T^{ij} \gamma_\mu \psi_{\nu j}) - \ft12 \left(g_{z\bar z} \hbox{\ooalign{$\displaystyle\partial$\cr$/$}} z \cdot \lambda^{i\bar z} -i \gamma^\mu (/\!\!\!\! {\cal Q} - /\!\!\!\! A)\right) \psi^i_\mu \nonumber\\ & &- g'^{z\bar z}\left[ \ft12 Z_z \bar C_{\bar z\bar z\bar z}\hbox{\ooalign{$\displaystyle\partial$\cr$/$}}\bar z\cdot\lambda^{i\bar z} + \ft 14{\rm i} \gamma^\mu (g_{z\bar z}\< V',\tilde Y^{ij} + \sigma\cdot \tilde {\cal F}^+\varepsilon^{ij}\> \psi_{\mu\,j}\right. \nonumber \\ &&- \ft 14 \varepsilon^{ij} {\cal D}_z Z_z\cdot (\<\bar U_{\bar z},\tilde Y_{jk}\>\varepsilon^{kl}\lambda^z_l +\<\bar U_{\bar z},\sigma \cdot\tilde {\cal F}^-\>\lambda^z_j) \nonumber \\ &&\left. - \ft 1{12}{\rm i} \varepsilon^{ij} g_{z\bar z} (C_{zzz} - {\cal D}_z{\cal D}_z Z_z)\, \varepsilon^{kl}(\bar \lambda^z_k\sigma_{ab}\lambda^z_l)\sigma^{ab}\lambda_j^z \right] \,,\nonumber\\ 0 & = & {\rm i} g'_{z\bar z}\left(/\!\!\!\!\nabla \lambda^{i\bar z} + \ft{1}{2}{\rm i} \left(/\!\!\!\! {\cal Q} - /\!\!\!\! A\right)\lambda^{i\bar z}\right)\nonumber\\ && + \ft{1}{2}{\rm i}\varepsilon^{ij} C_{zzz} g^{z\bar z}\left(\< \bar U_{\bar z},\tilde Y_{jk}\>\varepsilon^{kl}\lambda^z_l +\<\bar U_{\bar z}, \sigma \cdot \tilde {\cal F}^-\> \lambda^z_j\right) \nonumber\\ && + \ft16 {\cal D}_z C_{zzz}\cdot \varepsilon^{ij}\varepsilon^{kl} (\bar \lambda_k^z \sigma^{ab} \lambda_l^z) \sigma_{ab} \lambda_j^z -{\rm i} Z_z {\cal D}_{\bar z}\bar Z_{\bar z}\cdot/\!\!\!\!\nabla \bar z\cdot \lambda^{i\bar z}\,. \end{eqnarray} Also here, the equations reduce to those that we have found for the generic case where $\<V,U_z\> = 0$. The same comments as in section \ref{s:genBI} are valid here.\vspace{5mm} \par Repeating the analysis for the equations of the vectors, it appears that there is no information used about $Z_z$. This means that the analysis of the equations for the vectors of section \ref{s:genBI} remains valid. This is no surprise because the equations for the vectors are a symplectic section of equations. All the other equations are singlets for the symplectic group and can therefore be written as symplectic invariant equations.\vspace{5mm} \par Also the last constraint can be decomposed with respect to the symplectic basis. Then the equations become: \begin{eqnarray} \< \bar V' , \tilde C\> &=&-2 \bar\psi_i^\mu\<\bar V,\tilde Y^{ij} +\sigma\cdot\tilde{\cal F}^+\varepsilon^{ij}\>\psi_{\mu j} -2 \partial_\mu\bar z \cdot{\cal D}_{\bar z} \bar Z_{\bar z}\cdot \left(\nabla^\mu\bar z-\bar\psi^\mu_i\lambda^{i\bar z}\right) \nonumber \\ && +\ft 14 {\cal D}_{\bar z}\bar Z_{\bar z}\bar\lambda^{i\bar z}\sigma \cdot T_{ij}\lambda^{j\bar z}\,,\nonumber\\ \< \bar U'_{\bar z}, \tilde C\>&=& - 2\bar C_{\bar z\bar z\bar z} \nabla_\mu \bar z \cdot \nabla^\mu \bar z +\ft 14\bar C_{\bar z\bar z\bar z} (\bar\lambda^{k\bar z}\sigma\cdot T_{kl} \lambda^{l\bar z}) \,,\nonumber\\ 0 &=& g^{z\bar z}g'_{z\bar z} \Big( 2e^{-1}\partial_\mu \left(e({\cal Q}^\mu - A^\mu)\right) +2{\rm i}({\cal Q}^\mu - A^\mu)({\cal Q}_\mu - A_\mu) -2{\rm i} f_\mu^\mu +3{\rm i}{\bar\psi}^i_\mu \gamma^\mu \chi_i \nonumber \\ &&- 2{\rm i} g_{z\bar z}\partial_\mu z \cdot(\nabla^\mu{\bar z} -\bar \psi_i^\mu \lambda^{i\bar z})\nonumber\\ && +4{\bar \psi}^{i[\mu}\gamma_\mu\psi_i^{\nu]}({\cal Q}_\nu - A_\nu) +2 \bar \psi_i^\mu(/\!\!\!\! {\cal Q} - /\!\!\!\! A)\psi^i_\mu -2{\rm i}{\bar \psi}_i^\mu \phi^i_\mu \Big) \nonumber\\ && +\bar\psi^\mu_i\psi_{\mu j} \varepsilon^{ik}\varepsilon^{jl}{\cal D}_z Z_z\cdot\bar\lambda^z_k\lambda^z_l - {\bar \psi}_i^\mu \< V',\sigma\cdot \tilde {\cal F}^+ \> \varepsilon^{ij} \psi_{\mu j} -\ft 14 g^{z\bar z} g_{z\bar z}\< V', \tilde{\cal F}^+_{\mu\nu}\> T^{\mu\nu}_{ij}\varepsilon^{ij} \nonumber \\ && +2{\rm i} g^{z\bar z}Z_z \left({\rm i}{\cal D}_{\bar z}\bar Z_{\bar z}\cdot\partial_\mu\bar z\cdot \left({\cal Q}_\mu - A_\mu\right) +\bar\psi^\mu_i\bar C_{\bar z\bar z\bar z} \partial_\mu \bar z \cdot \lambda^{i\bar z}\right)\nonumber \\ && +\ft 12 g^{z\bar z}{\cal D}_z Z_z\cdot g^{z\bar z}\left(\ft 12 \varepsilon^{ik}\varepsilon^{jl} \<\bar U_{\bar z},\tilde Y_{ij}\>\<\bar U_{\bar z}, \tilde Y_{kl}\> - \<\bar U_{\bar z},\tilde {\cal F}^-_{ab}\>\<\bar U_{\bar z}, \tilde {\cal F}^{-ab}\>\right) \nonumber \\ &&-\ft {1}{2}{\rm i}\varepsilon^{ik}\varepsilon^{jl}\left({\cal D}_z Z_z\cdot\<\bar V,\tilde Y^{ij}\> +(-C_{zzz} + {\cal D}_z{\cal D}_z Z_z)\, g^{z\bar z}\<\bar U_{\bar z},\tilde Y_{ij}\>\right) \bar\lambda^z_k\lambda^z_l \nonumber \\ && -2{\rm i} g^{z\bar z}{\cal D}_z Z_z\cdot\varepsilon^{ij}\bar\lambda^z_i\<\bar U_{\bar z},\tilde\Lambda_j\> \nonumber\\ && -\ft {1}{12}\left( {\cal D}_z{\cal D}_z{\cal D}_z Z_z -2{\cal D}_z C_{zzz}\right) \varepsilon^{ij}(\bar\lambda^z_i\sigma^{ab}\lambda^z_j)\,\varepsilon^{kl}(\bar\lambda^z_k\sigma_{ab}\lambda^z_l) \nonumber \\ &&-\ft 12{\rm i}\left({\cal D}_z Z_z\cdot\<\bar V,\tilde {\cal F}^-_{ab}\> +(-C_{zzz} + {\cal D}_z{\cal D}_z Z_z)\, g^{z\bar z}\<\bar U_{\bar z},\tilde {\cal F}^-_{ab}\>\right) \varepsilon^{kl}(\bar\lambda^z_k\sigma^{ab}\lambda^z_l)\,, \nonumber \\ 0 &=& -2{\rm i} g'_{z\bar z}\Big[ e^{-1}\partial_\mu (e \nabla^\mu \bar z) +2{\rm i}({\cal Q}_\mu - A_\mu)\nabla^\mu\bar z\nonumber\\ && +\ft 12{\rm i}({\cal Q}_\mu - A_\mu)(\bar\psi_i^\mu\lambda^{i\bar z}) +2{\bar \psi}^{i[\mu}\gamma_\mu\psi_i^{\nu]}\nabla_\nu {\bar z}\nonumber \\ && +3\bar \chi_i\lambda^{i\bar z} -\ft 12 \bar\lambda^{i\bar z} \gamma_\mu \phi_i^\mu + \Gamma_{\bar z\zb}^{\bar z} \partial_\mu \bar z\cdot\nabla^\mu{\bar z} +\ft 18 \bar\psi^i_\mu\gamma^\mu\sigma\cdot T_{ij}\lambda^{j\bar z}\nonumber \\ &&-\bar \psi_i^\mu\left(\nabla_\mu \lambda^{i\bar z} +\ft 12{\rm i}({\cal Q}_\mu-A_\mu)\lambda^{i\bar z} +\ft 12{\rm i} g^{z\bar z}\<U_z,\tilde Y^{ij} +\sigma\cdot\tilde{\cal F}^+\varepsilon^{ij}\>\psi_{\mu j}\right)\Big] \nonumber \\ &&+2{\rm i} Z_z\left(\partial_\mu\bar z\cdot{\cal D}_{\bar z} \bar Z_{\bar z} (\nabla^\mu\bar z -\bar\psi^\mu_i\lambda^{i\bar z}) +\bar\psi^\mu_i\<\bar V,\tilde Y^{ij} +\sigma\cdot\tilde {\cal F}^+\varepsilon^{ij}\>\psi_{\mu j}\right)\nonumber \\ &&+\ft 14 C_{zzz}g^{z\bar z}g^{z\bar z}\varepsilon^{ik}\varepsilon^{jl}\<\bar U_{\bar z},\tilde Y_{ij}\> \<\bar U_{\bar z},\tilde Y_{kl}\> -\ft 12 C_{zzz}g^{z\bar z}g^{z\bar z}\<\bar U_{\bar z},\tilde {\cal F}^-_{ab}\>\<\bar U_{\bar z}, \tilde {\cal F}^{-ab}\> \nonumber \\ &&-\ft 12{\rm i} \varepsilon^{ik}\varepsilon^{jl}\left(C_{zzz}\<\bar V,\tilde Y_{ij}\> +{\cal D}_z C_{zzz} \cdot g^{z\bar z}\<\bar U_{\bar z}, \tilde Y_{ij}\>\right)\bar\lambda^z_k\lambda^z_l +\ft 14 \<U'_z, \tilde {\cal F}^+_{\mu\nu}\> T^{\mu\nu}_{ij}\varepsilon^{ij}\nonumber \\ &&+\ft 12{\rm i} \left(C_{zzz}\<\bar V,\tilde {\cal F}^-_{ab}\> +{\cal D}_z C_{zzz}\cdot g^{z\bar z}\<\bar U_{\bar z}, \tilde {\cal F}^-_{ab}\>\right)\, \varepsilon^{ij}(\bar\lambda^z_i\sigma^{ab}\lambda^z_j) \nonumber \\ &&+\ft 1{12} \left({\cal D}_z {\cal D}_z C_{zzz} +\ft 12 {\cal O}_{zzzzz}\right) \varepsilon^{ij}(\bar\lambda^z_i\sigma_{ab}\lambda^z_j) \varepsilon^{kl}(\bar\lambda^z_k\sigma_{ab}\lambda^z_l) \nonumber \\ && -2{\rm i} C_{zzz}g^{z\bar z}\varepsilon^{ij}\bar\lambda^z_i\<\bar U_{\bar z},\tilde\Lambda_j\>\,. \label{eomz} \end{eqnarray} The metric in front of the kinetic term of the scalar in the fourth equation is positive because of the physical condition \eqn{physicalcond}. Again, all these equations reduce to the equations of section \ref{s:genBI} if $Z_z=0$ and the same conclusions can be drawn as in section \ref{s:genBI}. Therefore, we conclude at this point that the `special case' is a valid alternative for a theory with $N=2$ supergravity and one vector multiplet. \section{Conclusions}\label{ss:concl} We have presented a fully symplectic invariant formulation of the coupling of an arbitrary number $n$ of vector multiplets to $N=2$ supergravity in 4 dimensions by using superconformal tensor calculus. This approach does not start from a prepotential, but rather from a $2(n+1)$ symplectic vector of chiral superconformal multiplets. We imposed the reducibility constraints \eqn{bian1}--\eqn{bian4} of chiral multiplets in supergravity to end up with vector multiplets in a superconformal background. Furthermore, we imposed the symplectic covariant defining equations of special K\"ahler\ geometry, and supersymmetric partners thereof. The bosonic defining equations include a breaking of dilatations and $U(1)$-transformations, the special conformal symmetry being broken as usual by imposing a constraint \eqn{EC} on the dilatational gauge field. In the fermionic sector, one of the constraints, \eqn{Sgauge}, breaks special supersymmetry. This results in unbroken Poincar\' e supersymmetry and $SU(2)$ gauge symmetry. The other constraints are determined by demanding the preservation of the Poincar\'e\ supersymmetry. \par The combination of all the special K\"ahler\ constraints \eqn{VVbar}--\eqn{VUal} and their supersymmetric partners with the generalized Bianchi identities of the chiral multiplets gave rise to a full set of field equations for the vector multiplet fields. \par Furthermore, we also discussed part of the equations of motion for the gravity sector. This could be done by imposing a new symplectic constraint \eqn{FET} and its supersymmetric partners. The full analysis would need a second compensating multiplet. \par Finally, we did the same analysis for a weaker definition of the special K\"ahler\ constraints where the constraint \eqn{VUal} is replaced by \eqn{UU}. This is a weaker constraint for the case of one physical vector multiplet. In appendix~C of \cite{CRTVP} two examples were given where (\ref{VUal}) is not satisfied. They are not suitable for illustrating non-trivial aspects of our construction. The first example does not fulfil the positivity condition (\ref{physicalcond}). The second example does agree with our definition, but is trivial in the sense that there ${\cal D}_z U_z= 0$. Therefore, the extra terms that appear in this paper are absent for this model. A nontrivial realization of our new models can e.g.\ be obtained by taking \begin{equation} V= \left(\begin{array}{c} 1\\ az\\ -z^3 \\ 3z^2 \end{array}\right)e^{K/2}\,. \end{equation} For $a=1$ it is the well-known $SU(1,1)/U(1)$ symmetric space with positive metric in the complex upper half plane. Deviating from this value gives a non-zero value to \begin{equation} \langle V, U_z \rangle = \frac{3i(a-1)z^2}{z^3 - 3az^2{\bar z} +3az{\bar z}^2 -{\bar z}^3}\,. \end{equation} Then the new metric \begin{equation} g'_{z{\bar z}}=\frac{-3a(z-\bar z)^2}{(z^2 + (1-3a)z\bar z + \bar z^2)^2} \end{equation} has a well-defined positivity domain, and \begin{equation} C_{zzz} = \frac{6ia(z -\bar z)}{(z^2 + (1 - 3a)z\bar z + \bar z^2)^2} \label{Czzznew} \end{equation} is not covariantly constant. \medskip \section*{Acknowledgments.} \noindent This work was supported by the European Commission TMR programme ERBFMRX-CT96-0045. \newpage
\section{Introduction} Since the recent experimental observation of Bose-Einstein condensation in dilute atomic gases \cite{BEC}, much interest has been raised about the coherence properties of the condensates. Considerable attention has been devoted to the matter of the relative phase between two condensates: how this phase manifests itself in an interference experiment \cite{Javanainen,Ketterle}, how it can be established by measurement \cite{Zoller,Jean}, and how it evolves in presence of atomic interactions \cite{Jean,Walls,Maciek} and in presence of particle losses \cite{resur}. As it was proved in recent experiments performed at JILA, binary mixtures of condensates represent an ideal system to study the phase coherence properties of Bose-Einstein condensates \cite{JILA_phase}. In these experiments two condensates in two different internal atomic states are created with a well defined relative phase. After a time $\tau$ during which the condensates evolve in the trapping potentials, one mixes coherently the two internal atomic states which makes the two condensates interfere; from the spatial interference pattern one gets the relative phase of the two condensates. By repeating the whole experimental process, one has access to the distribution of the relative phase after an evolution time $\tau$, so that one can investigate phase decoherence as function of time. The interaction between the two condensates in the JILA experiment gives rise to a rich spatial separation dynamics between the two condensates \cite{JILA_demix}, which complicates the theoretical study of the relative phase dynamics. As a consequence previous theoretical treatments of the phase decoherence processes, dealing essentially with steady state condensates, as in \cite{Bigelow}, cannot {\it a priori} be applied to the experimental situation. A treatment of the phase coherence of two interacting, non stationary, condensates can be found in \cite{Villain}, with two important differences as compared to the present situation of interest: (1) In \cite{Villain} the condensates are subject to a continuous coherent coupling of amplitude $\Lambda$; results are obtained from a perturbative expansion in powers of $1/\Lambda$ and cannot be simply extended to the present $\Lambda=0$ case. (2) In \cite{Villain} all the coupling constants $g_{aa},g_{ab},g_{bb}$ between the two internal atomic states $a$ and $b$ are assumed to be equal. In this paper we propose a formalism to study the relative phase dynamics of interacting and dynamically evolving Bose-Einstein condensates initially at zero temperature. We present the general method in section \ref{sec:gen-method}. It consists in expanding the initial state on Fock states, and in evolving each Fock state in the Hartree-Fock approximation. We determine the time dependence of the phase collapse for a binary mixture of condensates, due to (1) fluctuations in the relative number of particles between the condensates, intrinsic to the initial state with well defined relative phase, and (2) fluctuations in the total number of particles. In the next two sections we apply this general formalism to two limiting cases that can be treated analytically. The first case, in section \ref{sec:breath-together}, considers a particular solution of the non-linear Schr\"odinger equations for the condensates wavefunctions; in this solution the two condensates remain spatially superimposed as they breathe in phase, provided that dynamical stability conditions (that we determine) are satisfied. We find that phase decoherence can be highly reduced with respect to non mutually interacting condensates when the three coupling constants $g_{aa}, g_{ab},g_{bb}$ between atoms in the two internal states $a,b$ are close to each other. In section \ref{sec:closeg's}, we therefore study in a more general case (not restricted to the breathe-together solution) the dynamics of the relative phase for a mixture of condensates for close coupling constants. Our treatment requires also in this case the absence of demixing instability, a point that we discuss in detail. Finally we discuss the case of the JILA experiment in section \ref{sec:JILA}. This case, that corresponds to close coupling constants in a regime of demixing instability, is more difficult to analyze. The predicted phase collapse time depends on the fluctuations of the total number of particles ; it is on the order of 0.4 second for Gaussian fluctuations of 8\%. \section{General method} \label{sec:gen-method} In this section, we introduce a gedanken experiment to characterize phase coherence between two condensates: the relevant quantity is the interference term $\langle\hat{\psi_b}^\dagger(\vec{r},t)\hat{\psi_a}(\vec{r},t)\rangle$ between the atomic fields of the two condensates $a$ and $b$. Subsequently we express this interference term in the Hartree-Fock approximation, assuming an initially well defined relative phase between the condensates. After a further approximation on the modulus and the phase of the condensate wavefunctions, we determine the decay with time of the interference term due to atomic interactions; we arrive at the simple results Eq.(\ref{eq:inter_court}) for a fixed total number of particles and Eq.(\ref{eq:gaus_court}) for Gaussian fluctuations in the total number of particles. \subsection{Considered gedanken experiment} The experimental procedure we consider to measure the phase coherence is inspired by recent experiments at JILA \cite{JILA_phase}. A condensate is first created in a trap in some internal atomic state $a$; the corresponding condensate wavefunction in the zero temperature mean-field approximation is $\phi_0$, a stationary solution of the Gross-Pitaevskii equation: \begin{equation} \mu\phi_0 = -{\hbar^2\over 2m} \Delta\phi_0 + [U_a(\vec{r})+Ng_{aa}|\phi_0|^2]\phi_0. \label{eq:phi0} \end{equation} In this equation $N$ is the number of particles, $g_{aa}$ is the coupling constant between the atoms in the internal state $a$, related to the scattering length $a_{aa}$ by $g_{aa}=4\pi\hbar^2 a_{aa} /m$; $U_a$ is the trapping potential seen by the atoms in $a$ and $\mu$ is the chemical potential. Note that we have normalized $\phi_0$ to unity. At time $t=0$ a resonant electromagnetic pulse transfers in a coherent way part of the atoms to a second internal state $b$. The state of the system is then given in the Hartree-Fock approximation by \begin{equation} |\Psi(0)\rangle = [c_a |a,\phi_0\rangle + c_b |b,\phi_0\rangle]^N \label{eq:init} \end{equation} with $|c_a|^2 + |c_b|^2 =1$. As we assume a Rabi coupling between $a$ and $b$ much more intense than $\mu/\hbar$ the atomic interactions have a negligible effect during the transfer so that the amplitudes $c_{a,b}$ depend only on the pulse parameters, not on the number $N$ of particles. In the $N$-particle state Eq.(\ref{eq:init}) the condensate in state $a$ and the condensate in state $b$ have a well defined relative phase; we therefore call this state a {\sl phase state}, in analogy with \cite{Leggett}. The two condensates evolve freely in their trapping potentials during the time $\tau$. During this evolution we assume that there is no coherent coupling between $a$ and $b$ to lock the relative phase of the condensates; in particular the only considered interactions between the particles are elastic, of the type $a+a\rightarrow a+a$ (coupling constant $g_{aa}>0$), $a+b\rightarrow a+b$ (coupling constant $g_{ab}>0$), $b+b\rightarrow b+b$ (coupling constant $g_{bb}>0$). We therefore expect a collapse of the relative phase for sufficiently long times, due to atomic interactions. To test the phase coherence at time $\tau$, a second electromagnetic pulse is applied to mix the internal states $a$ and $b$. We assume that this second pulse is a $\pi/2$ pulse, so that the atomic field operators in the Heisenberg picture are transformed according to \begin{eqnarray} \hat{\psi}_a(\tau^+) & = & {e^{-i\delta}\over\sqrt{2}} \hat{\psi}_a(\tau^-) + {e^{i\delta}\over\sqrt{2}} \hat{\psi}_b(\tau^-) \\ \hat{\psi}_b(\tau^+) & = & -{e^{-i\delta}\over\sqrt{2}} \hat{\psi}_a(\tau^-) + {e^{i\delta}\over\sqrt{2}} \hat{\psi}_b(\tau^-), \end{eqnarray} $\delta$ being an adjustable phase. One then measures the mean spatial density $\rho_a$ in the internal state $a$, averaging over many realizations of the whole experiment: \begin{equation} \rho_a = \langle \hat{\psi}^\dagger_a(\tau^+)\hat{\psi}_a(\tau^+)\rangle. \end{equation} The signature of a phase coherence between the two condensates at time $\tau$ is the dependence of the mean density $\rho_a$ on the adjustable phase $\delta$. More precisely we define the contrast \begin{equation} C ={ \mbox{max}_\delta \rho_a -\mbox{min}_\delta\rho_a \over \mbox{max}_\delta\rho_a +\mbox{min}_\delta\rho_a} = {2|\langle\hat{\psi}_b^\dagger(\tau^-)\hat{\psi}_a(\tau^-)\rangle| \over \sum_{\varepsilon=a,b} \langle \hat{\psi}^\dagger_\varepsilon(\tau^-) \hat{\psi}_\varepsilon(\tau^-)\rangle} \end{equation} The contrast involves the interference term $\langle\hat{\psi}_b^\dagger(\tau^-)\hat{\psi}_a(\tau^-)\rangle$ which carries the information about the relative phase between the two condensates. \subsection{Approximate evolution of an initial phase state} The time evolution in the phase state representation is not simple, as an initial phase state is mapped onto a superposition of phase states. It is more convenient to introduce Fock states, that is states with a well defined number of particles in $a$ and in $b$, these numbers being preserved by the time evolution. We therefore expand the initial phase state over the Fock states: \begin{equation} |\Psi(0)\rangle = \sum_{N_a=0}^{N} \left({N!\over N_a! N_b!}\right)^{1/2} c_a^{N_a}c_b^{N_b}|N_a:\phi_0,N_b:\phi_0\rangle \end{equation} where we set $N_b=N-N_a$. By calculating the evolution of each Fock state in the simplest Hartree-Fock approximation, we get the following mapping: \begin{equation} |N_a:\phi_0,N_b:\phi_0\rangle \rightarrow e^{-iA(N_a,N_b;t)/\hbar} |N_a:\phi_a(N_a,N_b;t),N_b:\phi_b(N_a,N_b;t)\rangle \label{eq:evol_Fock} \end{equation} where the condensates wavefunctions evolve according to the coupled Gross-Pitaevskii equations: \begin{equation} i\hbar\partial_t\phi_\varepsilon = \left[-{\hbar^2\over 2m}\Delta + U_\varepsilon(\vec{r}) +N_\varepsilon g_{\varepsilon\varepsilon}|\phi_\varepsilon|^2 +N_{\varepsilon'} g_{\varepsilon\varepsilon'}|\phi_{\varepsilon'}|^2\right] \phi_\varepsilon \label{eq:gpe} \end{equation} (where $\varepsilon'\neq\varepsilon$) with the initial conditions \begin{equation} \phi_a(0)=\phi_b(0)=\phi_0 \label{eq:init_cond} \end{equation} and where the time dependent phase factor $A$ solves: \begin{equation} {d\over dt} A(N_a,N_b;t) = -{1\over 2} N_a^2 g_{aa}\int d\vec{r}\;|\phi_a|^4 -{1\over 2} N_b^2 g_{bb}\int d\vec{r}\;|\phi_b|^4 -N_aN_b g_{ab} \int d\vec{r}\;|\phi_a|^2|\phi_b|^2. \label{eq:apoint} \end{equation} Equation (\ref{eq:apoint}) is derived in Appendix \ref{app:A}. Physically $dA/dt$ is simply the opposite of the mean interaction energy between the particles in the Fock state. In the case where the Fock state is a steady state, the need for the phase factor $A$ additional to the Gross-Pitaevskii equation is obvious; the exact phase factor is indeed $e^{-iEt/\hbar}$, where $E$ is the energy of the Fock state, whereas the phase factor obtained from the Gross-Pitaevskii evolution is $e^{-i(N_a\mu_a+N_b\mu_b)t/\hbar}$, where $\mu_{a,b}$ is the chemical potential in $a,b$. Using the evolution of the Fock states, and other approximations valid in the limit of large numbers of particles (as detailed in the Appendix \ref{app:interf}) we obtain for the interference term between the condensates with a well-defined total number $N$ of particles: \begin{equation} \langle \hat{\psi}_b^\dagger \hat{\psi_a}\rangle _{N} = c_a c_b^* \sum_{N_a=1}^{N} {N!\over (N_a-1)!N_b!} |c_a|^{2(N_a-1)}|c_b|^{2N_b} \phi_a(N_a,N_b)\phi_b^*(N_a-1,N_b+1) \label{eq:complique} \end{equation} where $N_b=N-N_a$. The exact computation of this sum remains a formidable task, since it involves in principle the solution of $N$ different sets of two coupled Gross-Pitaevskii equations. We introduce some simplifying approximations in the next subsection. \subsection{Phase collapse for a mixture} In the present experiments the total number of particles fluctuates from one realization to the other, so that Eq.(\ref{eq:complique}) has to be averaged over $N$. We assume that the fluctuations of the total number of particles have a standard deviation $\Delta N$ much smaller than the mean total particle number $\bar{N}$. As the distributions of the number of particles in $a$ and $b$ in a phase state have also a width much smaller than $\bar{N}$ (typically on the order of $\bar{N}^{1/2}$) we can assume than the number of particles in $a$ and in $b$ are very close to their average values $\bar{N}_\varepsilon = |c_\varepsilon|^2\bar{N}$. We now take advantage of this property to simplify Eq.(\ref{eq:complique}). We split the condensate wavefunctions in a modulus and a phase $\theta_\varepsilon$; we assume that the variation of the modulus can be neglected over the distribution of $N_{a,b}$, and that the variation of the phase can be approximated by a linear expansion around $\bar{N_\varepsilon}$. We thus get the approximate form for the condensate wavefunctions: \begin{equation} \phi_\varepsilon(N_a,N_b) \simeq \bar{\phi_\varepsilon} \exp\left[i\sum_{\varepsilon'=a,b}(N_{\varepsilon'}-\bar{N_{\varepsilon'}}) (\partial_{N_{\varepsilon'}}\theta_{\varepsilon})(\bar{N_a},\bar{N_b})\right] \end{equation} where $\bar{\phi_\varepsilon} = \phi_\varepsilon(N_a=\bar{N_a}, N_b=\bar{N_b})$. To this level of approximation the mean densities in the internal states $a,b$ are simply given by \begin{equation} \langle\hat{\psi}_\varepsilon^\dagger\hat{\psi}_\varepsilon\rangle_N \simeq \bar{N_\varepsilon}|\bar{\phi_\varepsilon}|^2 \end{equation} whereas the interference term between the condensates is: \begin{eqnarray} \langle \hat{\psi}_b^\dagger \hat{\psi_a}\rangle _{N} &\simeq& \bar{N}c_ac_b^*\bar{\phi_a}\bar{\phi_b}^* \exp\left\{i[(N-\bar{N})\chi_s-\bar{N}(|c_a|^2-|c_b|^2)\chi_d] \right\}e^{i\chi_0} \nonumber\\ &&\left[|c_a|^2 e^{i\chi_d} + |c_b|^2 e^{-i\chi_d}\right]^{N-1}. \label{eq:inter_N} \end{eqnarray} In this last expression we have introduced the time and position dependent quantities \begin{eqnarray} \chi_s &=& {1\over 2} \left[\left(\partial_{N_a}+\partial_{N_b}\right) (\theta_a-\theta_b)\right](\bar{N_a},\bar{N_b}) \label{eq:chi_s} \\ \chi_d &=& {1\over 2} \left[\left(\partial_{N_a}-\partial_{N_b}\right) (\theta_a-\theta_b)\right](\bar{N_a},\bar{N_b}). \label{eq:chi_d} \end{eqnarray} The phase $\chi_0 = {1\over 2} (\partial_{N_a}-\partial_{N_b})(\theta_a +\theta_b)(\bar{N_a},\bar{N_b})$ in Eq.(\ref{eq:inter_N}) is less important as contrarily to $\chi_{s,d}$ it is not multiplied by $N$. At time $t=0$ all the $\chi$'s vanish. In the large $N$ limit, the $\chi$'s are expected to be on the order of $\bar{\mu} t/\hbar\bar{N}$. The factor responsible for the collapse of the contrast at a fixed value of $N$ is the second line of Eq.(\ref{eq:inter_N}), the factors in the first line being of modulus one. As $N$ is large a small variation of $\chi_d$ from its initial value $\chi_d(t=0)=0$ is sufficient to destroy the interference term. Over the range of the collapse we can therefore expand the exponential of $\pm i\chi_d$ to second order in $\chi_d$, obtaining: \begin{equation} \langle \hat{\psi}_b^\dagger \hat{\psi_a}\rangle _{N} \simeq \bar{N}c_ac_b^*\bar{\phi_a}\bar{\phi_b}^* \exp\left\{i[(N-\bar{N})[\chi_s+(|c_a|^2-|c_b|^2)\chi_d]\right\}\ \exp\left[-2N\chi_d^2|c_a|^2|c_b|^2\right]. \label{eq:inter_court} \end{equation} The second exponential factor in this expression allows to determine the collapse time $t_c^{\mbox{\scriptsize fix}}$ for a fixed number of particles, through the condition \begin{equation} 4N|c_a|^2|c_b|^2\chi_d^2(t_c^{\mbox{\scriptsize fix}}) \simeq 1 \label{eq:deftc} \end{equation} such that the modulus of the interference term is reduced by a factor $e^{-1/2}$ from its initial value. The first exponential factor in Eq.(\ref{eq:inter_court}) accounts for the phase difference of the interference term for $N$ particles and $\bar{N}$ particles, as shown by the identity: \begin{equation} \chi_s+(|c_a|^2-|c_b|^2)\chi_d = {d\over dN} \left[(\theta_a-\theta_b)(N|c_a|^2,N|c_b|^2)\right]_{N=\bar{N}}. \label{eq:magique} \end{equation} This phase factor can also be understood as a consequence of a drift of the relative phase between two condensates at a velocity $v(N)$ depending on the total number of particles: \begin{equation} v(N) = {\bar{\mu_b}-\bar{\mu_a}\over \hbar} + (N-\bar{N}) \left[\chi_s+(|c_a|^2-|c_b|^2)\chi_d\right]. \label{eq:vdrift} \end{equation} As we shall see in the next subsection fluctuations in the total number of particles $N$ result in fluctuations of this phase factor, providing an additional source of smearing of the phase, as already emphasized in \cite{resur}. \subsection{Effect of fluctuations in the total number of particles} The effect on the phase collapse of fluctuations in the total number of particles is obtained by averaging Eq.(\ref{eq:inter_court}) over the probability distribution of $N$. To be specific we assume a Gaussian distribution for $N$. The average can be calculated by replacing the discrete sum over $N$ by an integral; we neglect a term proportional to $(\Delta N\chi_d^2)^2$ scaling as $(\Delta N/\bar{N})^2$ at the collapse time $t_c^{\mbox{\scriptsize fix}}$; the resulting modulus of the interference term reads: \begin{equation} |\langle \hat{\psi}_b^\dagger \hat{\psi_a}\rangle^{\mbox{\scriptsize Gauss}}| \simeq \bar{N}|c_ac_b^*\bar{\phi_a}\bar{\phi_b}^*| \exp\left\{ -{1\over 2}(\Delta N)^2 \left[{d\over dN}(\theta_a-\theta_b)\right]^2_{N=\bar{N}} \right\} \ \exp\left[-2\bar{N}\chi_d^2|c_a|^2|c_b|^2\right]. \label{eq:gaus_court} \end{equation} The first exponential factor in this expression represents the damping of the interference term due to fluctuations in the total number of particles; the second exponential factor, already present in Eq.(\ref{eq:inter_court}), gives the damping due to fluctuations in the relative number of particles between $a$ and $b$, as can be seen in Eq.(\ref{eq:chi_d}). \subsection{The steady state case and comparison with previous treatments} \label{subsec:sta} Our treatment can be easily adapted to the case of two initially different condensate wavefunctions $\phi_a(t=0)$ and $\phi_b(t=0)$. In the particular case of condensates in stationary states, the formulas for the interference term $\langle\hat{\psi_b}^\dagger\hat{\psi_a}\rangle$ remain the same, and one has $\theta_\varepsilon= -\mu_\varepsilon(N_a,N_b)t/\hbar$. We can give in this case the explicit expression for the collapse time $t_c^{\mbox{\scriptsize fix}}$ defined in Eq.(\ref{eq:deftc}), assuming a fixed total number of atoms $N=\bar{N}$: \begin{equation} t_c^{\mbox{\scriptsize fix}} = \hbar \left[\bar{N}^{1/2}|c_a c_b| |(\partial_{N_a}-\partial_{N_b})(\mu_a-\mu_b)| \right]^{-1}. \label{eq:geniale} \end{equation} For the particular case of non mutually interacting steady state condensates $\mu_\varepsilon$ depends on $N_\varepsilon$ only, so that the partial derivatives in the denominator of Eq.(\ref{eq:geniale}) reduce to $d\mu_a/dN_a + d\mu_b/dN_b$, and we recover the results of \cite{Jean,resur}. From Eq.(\ref{eq:geniale}) we see that what matters physically is the change in the {\it difference} between the chemical potentials of the two condensates when one transfers one particle from one condensate to the other. For this reason the case of mutually interacting condensates with close coupling constants can lead to much larger $t_c$'s as compared to the case of non-mutually interacting condensates. For example, in the case of the JILA experiment \cite{JILA_phase}, assuming that the condensates are in steady state, one finds $t_c^{\mbox{\scriptsize fix}} \simeq 3.1$ s; by ignoring the interaction between the condensates (setting by hand $g_{ab}=0$) one obtains the much shorter time $ \simeq 0.25$ s. The JILA case is analyzed in more detail in our section \ref{sec:JILA}. A similar prediction on the reduction of decoherence due to mutual interactions between the two condensates, in trapping potentials with different curvatures, was obtained numerically in \cite{Bigelow}. The treatment in \cite{Maciek} considers the absolute phase dynamics of a single condensate (in our formalism $c_b=0$) in a coherent state. When the condensate wavefunction is stationary one has simply $\theta_a = -\mu_a t/\hbar$. From Eq.(\ref{eq:gaus_court}) with $\Delta N=\bar{N}^{1/2}$ (as the coherent state has a Poisson distribution for $N$) we then find that the phase of the condensate order parameter is damped as $\exp[-\bar{N}(d\mu_a/dN)^2t^2/2\hbar^2]$ as in \cite{Maciek}. \section{Application to the breathe-together solution} \label{sec:breath-together} In this section we consider a particular solution of the coupled Gross-Pitaevskii equations for which an approximate scaling solution is available when the chemical potential is much larger than the energy spacing between trap levels, the so-called Thomas-Fermi regime. We first give the set of parameters for which this solution, that we call the breathe-together solution, exists. We then linearize the Gross-Pitaevskii equations around this solution to determine its stability with respect to demixing and to obtain the phase coherence dynamics. \subsection{Description of the breathe-together solution} We now determine the set of parameters such that the coupled Gross-Pitaevskii equations Eq.(\ref{eq:gpe}) for \begin{equation} N_\varepsilon = \bar{N_\varepsilon}\equiv \bar{N} |c_\varepsilon|^2 \end{equation} have a solution with $\bar{\phi_a}(\vec{r},t)=\bar{\phi_b}(\vec{r},t) \equiv\bar{\phi}(\vec{r},t)$. The general condition is that the effective potential, that is the trapping potential plus the mean field potential, seen by the atoms in $a$ and in $b$ should be the same. This condition is satisfied when: \begin{eqnarray} U_a(\vec{r}) = U_b(\vec{r}) &\equiv& U(\vec{r}) \\ \bar{N_a} g_{aa} + \bar{N_b} g_{ab} = \bar{N_b} g_{bb} + {\bar N_a} g_{ab} &\equiv& \bar{N} g. \label{eq:angle} \end{eqnarray} The resulting Gross-Pitaevskii equation for the condensate wavefunction $\bar{\phi}$ common to $a$ and $b$ is then: \begin{equation} i\hbar\partial_t\bar{\phi} =\left[-{\hbar^2\over 2m}\Delta + U(\vec{r}) +\bar{N}g |\bar{\phi}|^2\right]\bar{\phi} \end{equation} with the initial condition $\bar{\phi}(\vec{r},0)=\phi_0[N=\bar{N}](\vec{r}) \equiv\bar{\phi_0}$, where $\phi_0$ is defined in Eq.(\ref{eq:phi0}). By rewriting Eq.(\ref{eq:angle}) as $\bar{N_a}/\bar{N_b}=(g_{bb}-g_{ab})/(g_{aa}-g_{ab})$ we see that this equality can be satisfied by choosing properly the mixing angle between $a$ and $b$ provided that \begin{equation} g_{ab} < g_{aa},g_{bb} \ \ \ \ \mbox{or} \ \ \ \ \ \ g_{ab} > g_{aa},g_{bb}. \end{equation} As we shall see below, only the first case is relevant here, since the second case corresponds to an unstable solution with respect to demixing between $a$ and $b$. \subsection{Linearization around the breathe-together solution} The strategy to obtain the quantities $\chi_{s,d}$ relevant for the phase dynamics is to calculate in the linear approximation the deviations $\delta\phi_\varepsilon$ between the breathe-together solution $\bar{\phi}$ and neighboring solutions $\phi_{\varepsilon}$ for $N_\varepsilon$ slightly different from $\bar{N_\varepsilon}$: \begin{equation} \delta\phi_\varepsilon \equiv \phi_\varepsilon(\bar{N_a}+\delta N_a,\bar{N_b} +\delta N_b)-\phi_\varepsilon(\bar{N_a},\bar{N_b}). \label{eq:def_delta_phi} \end{equation} From the definitions Eq.(\ref{eq:chi_s}) and Eq.(\ref{eq:chi_d}) one indeed realizes that in the limit of small $\delta N_a$: \begin{eqnarray} \chi_s &=&\left[{\delta\theta_a-\delta\theta_b\over 2\delta N_a} \right]_{\delta N_b=\delta N_a} \label{eq:chi1}\\ \chi_d &=&\left[{\delta\theta_a-\delta\theta_b\over 2\delta N_a} \right]_{\delta N_b=-\delta N_a} \label{eq:chi2} \end{eqnarray} where $\delta\theta_{a,b}$ are the deviations of the phase of the neighboring solutions $\phi_{\varepsilon}$ from the phase of the breathe-together solution: \begin{equation} \delta\theta_a-\delta\theta_b = \mbox{Im} \left[{\delta\phi_a\over\bar{\phi}} -{\delta\phi_b\over\bar{\phi}}\right]. \label{eq:phase_rel_def} \end{equation} It turns out that homogeneous rather than inhomogeneous linear equations can be obtained for the deviations $\delta\phi_\epsilon$ by introducing the quantities: \begin{equation} \delta\varphi_\varepsilon \equiv{\delta[\sqrt{N_\varepsilon}\phi_\varepsilon] \over \sqrt{\bar{N_\varepsilon}}}= {\delta N_\varepsilon\over 2 \bar{N_\varepsilon}} \bar{\phi} + \delta\phi_\varepsilon. \label{eq:def_varphi} \end{equation} Furthermore a partial decoupling occurs for the linear combinations \begin{eqnarray} \delta\varphi_s &\equiv &\delta\varphi_a + \delta\varphi_b \\ \delta\varphi_d &\equiv &\delta\varphi_a - \delta\varphi_b. \end{eqnarray} The sum $\delta\varphi_s$ obeys a linear equation involving $\delta\varphi_d$ as a source term: \begin{eqnarray} i\hbar\partial_t \delta\varphi_s& =&\left[-{\hbar^2\over 2m}\Delta +U + 2\bar{N}g |\bar{\phi}|^2\right]\delta\varphi_s\nonumber\\ & +& \bar{N}g\bar{\phi}^2\delta\varphi_s^* +(\bar{N_a}g_{aa}-\bar{N_b}g_{bb}) (|\bar{\phi}|^2\delta\varphi_d + \bar{\phi}^2\delta\varphi_d^*). \end{eqnarray} The part of this equation involving $\delta\varphi_s$ is identical to the one obtained for a single condensate with $\bar{N}$ particles and a coupling constant $g$. The corresponding modes have minimal frequencies on the order of the trap frequency $\omega$ for an isotropic harmonic trap \cite{Stringari}. The difference $\delta\varphi_d$ obeys the closed equation: \begin{equation} i\hbar\partial_t \delta\varphi_d = \left[-{\hbar^2\over 2m}\Delta +U + \bar{N}g |\bar{\phi}|^2\right]\delta\varphi_d + {\bar{N_a}\bar{N_b} \over \bar{N}} (g_{aa}+g_{bb}-2g_{ab}) (|\bar{\phi}|^2\delta\varphi_d + \bar{\phi}^2\delta\varphi_d^*) \end{equation} where we have used the identity: \begin{equation} \bar{N_b}(g_{bb}-g_{ab}) = \bar{N_a}(g_{aa}-g_{ab}) = {\bar{N_a}\bar{N_b} \over \bar{N}} (g_{aa}+g_{bb}-2g_{ab}). \end{equation} As shown in \cite{Nous} minimal eigenfrequencies of this equation can be much smaller than $\omega$; e.g.\ when all the coupling constants are equal, the minimal eigenfrequencies in a harmonic isotropic trap of frequency $\omega$ scale as $\hbar\omega^2/\mu\ll\omega$ in the Thomas-Fermi limit. For the derivation of the $\chi$'s it is sufficient to calculate $\delta\varphi_d$. The relative phase between the two condensates for the considered neighboring solution with $N_\varepsilon=\bar{N_\varepsilon}+\delta N_\varepsilon$ particles in the state $\varepsilon$ is in fact given by: \begin{equation} \delta\theta_a-\delta\theta_b = {1\over 2i}\left[ {\delta\varphi_d\over \bar{\phi}} -{\delta\varphi_d^*\over \bar{\phi}^*}\right]. \label{eq:rel_phase} \end{equation} as can be checked from the definition Eq.(\ref{eq:phase_rel_def}). \subsection{Approximate equations of evolution in the Thomas-Fermi limit} In the remaining part of this section we assume an isotropic harmonic trapping potential $U(\vec{r})=m\omega^2 r^2/2$ and we restrict to the Thomas-Fermi limit $\mu\gg\hbar\omega$. In the Thomas-Fermi limit it is known \cite{Kagan,Yvan} that most of the time dependence of the wavefunction $\bar{\phi}$ can be absorbed by a time dependent gauge and scaling transform; here we apply this transform to both $\bar{\phi}$ and $\delta\varphi_d$: \begin{eqnarray} \bar{\phi}(\vec{r},t) & \equiv & {e^{-i\eta(t)}\over\lambda^{3/2}(t)} e^{imr^2\dot{\lambda}(t)/2\hbar\lambda(t)} \tilde{\bar{\phi}}(\vec{r}/\lambda(t),t) \label{eq:gau1}\\ \delta\varphi_d(\vec{r},t) & = & {e^{-i\eta(t)}\over\lambda^{3/2}(t)} e^{imr^2\dot{\lambda}(t)/2\hbar\lambda(t)} \tilde{\delta\varphi_d}(\vec{r}/\lambda(t),t). \label{eq:gau2} \end{eqnarray} The scaling factor $\lambda(t)$ solves the second order Newton-type differential equation \begin{equation} \ddot{\lambda}={g\over g_{aa}}{\omega^2\over\lambda^4} -\omega^2\lambda \label{eq:ev_lambda} \end{equation} with the initial condition $\lambda(0)=1,\dot{\lambda}(0)=0$. The \lq\lq force" seen by $\lambda$ in Eq.(\ref{eq:ev_lambda}) derives from the sum of an expelling $1/\lambda^3$ potential due to repulsive interactions between atoms and an attractive $\lambda^2$ potential due to the harmonic confinement of the atoms. It leads to periodic oscillations of $\lambda$, that is to a periodic breathing of the condensates. We have also introduced a phase factor involving the time dependent function $\eta$ such that $\dot{\eta} = \bar{\mu} g/(g_{aa}\lambda^3\hbar)$. In the Appendix \ref{app:TF} we derive approximate evolution equations for $\tilde{\bar{\phi}}$ and $\tilde{\delta\varphi_d}$; we give here only the result. To lowest order in the Thomas-Fermi approximation $\tilde{\bar{\phi}}$ does not evolve and can be approximated by the Thomas-Fermi approximation for $\bar{\phi_0}$: \begin{equation} \tilde{\bar{\phi}}(\vec{r},t) \simeq \bar{\phi_0}(\vec{r}) \simeq \left({15\over 8\pi R_0^3}\right)^{1/2} \left[1-{r^2\over R_0^2}\right]^{1/2} \label{eq:phi_TF} \end{equation} with a Thomas-Fermi radius $R_0=\sqrt{2\bar{\mu}/m\omega^2}$. The approximate evolution for $\tilde{\delta\varphi_d}$ is conveniently expressed in terms of the real function $\alpha$ and the purely imaginary function $\beta$: \begin{eqnarray} \alpha &=& \tilde{\bar{\phi}}^*\tilde{\delta\varphi_d}+ \tilde{\bar{\phi}}\tilde{\delta\varphi_d}^* \label{eq:def_alpha}\\ \beta &=& {1\over 2}\left[ {\tilde{\delta\varphi_d}\over\tilde{\bar{\phi}}} -{\tilde{\delta\varphi_d}^*\over\tilde{\bar{\phi}}^*}\right]. \label{eq:def_beta} \end{eqnarray} These functions have a clear physical meaning. The first one $\alpha$ corresponds to the deviation $\delta\rho_a/\bar{N_a}-\delta\rho_b/\bar{N_b}$ written in the rescaled frame, $\delta\rho_\varepsilon$ being the deviation of spatial density in the condensate $\varepsilon$ from the breathe-together solution. Apart from a factor $i$ the second function $\beta$ is the deviation of the relative phase Eq.(\ref{eq:rel_phase}) written in the rescaled frame: \begin{equation} (\delta\theta_a-\delta\theta_b)(\vec{r},t)= -i\beta(\vec{r}/\lambda,t). \label{eq:sens_beta} \end{equation} The equations of evolution for $\alpha,\beta$ are: \begin{equation} i\hbar\partial_t \left( \begin{array}{c} \alpha \\ \beta \end{array} \right) = L(t) \left( \begin{array}{c} \alpha \\ \beta \end{array} \right) \label{eq:evol_ab} \end{equation} where the operator $L(t)$ in the Thomas-Fermi approximation reads: \begin{equation} L(t)= \left(\begin{array}{cc} 0 & -{\hbar^2\over m\lambda^2} \; \mbox{div}\left[\bar{\phi_0}^2\;\vec{\mbox{grad}}(\cdot)\right] \\ {1\over\lambda^3} {\bar{N_a}\bar{N_b} \over \bar{N}} (g_{aa}+g_{bb}-2g_{ab}) & 0 \end{array}\right). \label{eq:op_ab} \end{equation} The initial conditions for $\alpha,\beta$ at time $t=0$ obtained from Eq.(\ref{eq:def_varphi}) and Eq.(\ref{eq:init_cond}) are: \begin{eqnarray} \label{eq:init_alpha} \alpha(0) &=& \left({\delta N_a\over \bar{N_a}} -{\delta N_b\over \bar{N_b}}\right) \bar{\phi_0}^2 \\ \beta(0) &=& 0. \label{eq:init_beta} \end{eqnarray} \subsection{Solution of the Thomas-Fermi evolution equations: stability against demixing} The strategy to determine the time evolution of $\alpha,\beta$ is (1) to expand the vector $(\alpha(0),\beta(0))$ on eigenmodes of the operator $L(0)$, and (2) to calculate the time evolution of each eigenmode. \subsubsection{Expansion on modes of $L(0)$} Consider an eigenvector $(\alpha,\beta)$ of the operator $L(0)$ with the eigenvalue $\hbar\Omega$. For $\Omega\neq 0$ one can express the component $\beta$ as function of $\alpha$: \begin{equation} \beta = {\alpha\over\hbar\Omega} {\bar{N_a}\bar{N_b} \over \bar{N}} (g_{aa}+g_{bb}-2g_{ab}) \label{eq:elimine_beta} \end{equation} and obtain the eigenvalue problem for $\alpha$: \begin{equation} \Omega^2\alpha = \left({\bar{N_a}\bar{N_b} \over \bar{N}^2} {(g_{aa}+g_{bb}-2g_{ab})\over g_{aa}}\right)S[\alpha] \label{eq:eigen} \end{equation} where we have introduced the Stringari operator: \begin{equation} S[\alpha] \equiv -{\bar{N}g_{aa}\over m}\;\mbox{div}[\bar{\phi_0}^2\; \vec{\mbox{grad}}\;\alpha]. \label{eq:stringari} \end{equation} This operator has been studied in \cite{Stringari}. It is an Hermitian and positive operator, with a spectrum $q\omega^2$, $q$ non negative integer; $q$ is given by \begin{equation} q=2n^2+2nl+3n+l \end{equation} as function of the radial quantum number $n$ and the angular momentum $l$. This allows the determination of the eigenfrequencies $\Omega$: \begin{equation} \Omega_q = \pm\left({\bar{N_a}\bar{N_b} \over \bar{N}^2} {(g_{aa}+g_{bb}-2g_{ab})\over g_{aa}}\right)^{1/2}q^{1/2}\omega, \end{equation} with $q>0$ as we have assumed $\Omega\neq 0$. The case of a vanishing $\Omega$ corresponds to the zero energy mode $\alpha_0=0,\beta_0=1$ of the operator $L(0)$, as it can be checked from a direct calculation. All the eigenmodes of $L(0)$ have been identified. They do not form a complete family of vectors however. The vector $(\alpha=1,\beta=0)$ cannot be expanded on the eigenmodes of $L(0)$. Its first component $\alpha$ is indeed in the kernel of the operator $S$ (as $S[\alpha]=0$) whereas none of the $\alpha_q$ is in the kernel of $S$ ($S[\alpha_q]=q \omega^2 \alpha_q$ is not identically zero) except when $\alpha_q$ is identically zero (for $q=0$). The family of eigenvectors of $L(0)$ completed by the additional vector $(\alpha=1,\beta=0)$ forms a basis. The additional vector is called an anomalous mode, and we set $\alpha_{\mbox{\scriptsize anom}} = 1$, $\beta_{\mbox{\scriptsize anom}} = 0$; the action of $L(0)$ on the anomalous mode gives the zero energy mode times the constant factor $\bar{N_a}(g_{aa}-g_{ab})$ \cite{anomal}. The mode functions $\alpha_q$ of the operator $S$ are given in \cite{Stringari}. It turns out that in the expansion of the initial conditions for $\alpha,\beta$ Eq.(\ref{eq:init_alpha},\ref{eq:init_beta}), only the isotropic eigenmodes of $L(0)$ with $q=5$ and the anomalous mode are involved: \begin{equation} \left(\begin{array}{c}\alpha(0) \\ \beta(0)\end{array}\right)= C_5\left[ \left(\begin{array}{c}\alpha_{q=5}\\ \beta_{q=5}\end{array}\right)+ \left(\begin{array}{c}\alpha_{q=5}\\ -\beta_{q=5}\end{array}\right) \right] +C_{\mbox{\scriptsize anom}} \left(\begin{array}{c} 1 \\ 0 \end{array}\right). \label{eq:devt_t=0} \end{equation} The isotropic eigenmode of $S$ with $q=5$, the so-called breathing mode, reads \begin{equation} \alpha_{q=5}(\vec{r}) = \left[{r^2\over R_0^2} -{3\over 5}\right]. \label{eq:mode5} \end{equation} By Eq.(\ref{eq:elimine_beta}) we have $\beta_{q=5} = \alpha_{q=5} \bar{N_a}\bar{N_b} (g_{aa}+g_{bb}-2g_{ab})/\bar{N}\hbar \Omega_{q=5}$. For the coefficients of the modal expansion of $(\alpha(0),\beta(0))$, we obtain \begin{eqnarray} C_{\mbox{\scriptsize anom}} &=& {3\over 4\pi R_0^3} \left({\delta N_a\over \bar{N_a}} -{\delta N_b\over \bar{N_b}}\right) \label{eq:coef} \\ C_{5} &=& -{5\over 4} C_{\mbox{\scriptsize anom}}. \end{eqnarray} \subsubsection{Evolution of the modes and stability against demixing} As a second step we determine the time evolution of the modes of the operator $L(0)$. If we consider an eigenstate $(\alpha_q(0),\beta_q(0))$ of $L(0)$ with the eigenenergy $\hbar\Omega_q$ and evolve it according to Eq.(\ref{eq:evol_ab}), we find that the evolution reduces to multiplication by purely time dependent factors $A_q(t),B_q(t)$: \begin{eqnarray} \alpha_q(\vec{r},t) &=& A_q(t) \alpha(\vec{r},0) \\ \beta_q(\vec{r},t) &=& B_q(t) \beta(\vec{r},0) \end{eqnarray} where the factors satisfy the differential equations: \begin{eqnarray} i\dot{A_q} &=& {\Omega_q\over\lambda^2} B_q \\ i\dot{B_q} &=& {\Omega_q\over\lambda^3} A_q \label{eq:syst} \end{eqnarray} with the initial conditions $A_q(0) = B_q(0) =1$. Note that the zero energy eigenmode does not evolve, as $\Omega_q=0$. The anomalous mode has to be integrated separately, leading to \begin{eqnarray} \alpha_{\mbox{\scriptsize anom}} (\vec{r},t) &=& 1 \\ \beta_{\mbox{\scriptsize anom}} (\vec{r},t) &=& {\bar{N_a}\bar{N_b} \over \bar{N}} {(g_{aa}+g_{bb}-2g_{ab}) \over i\hbar} \int_0^t {dt'\over\lambda^3(t')}. \label{eq:ev_anom} \end{eqnarray} We are now able to address the problem of dynamical stability of the breathe-together solution. Dynamical stability requires that any small deviation of the $\phi_\varepsilon$'s from the breathe-together solution $\bar{\phi}$ should not grow exponentially with time. Here an exponential growth of $\alpha$ may correspond to a demixing of the two condensates $a$ and $b$. A first case of instability occurs when $g_{ab}>g_{aa},g_{bb}$. In this case the eigenfrequencies $\Omega_q$ are purely imaginary and $A_q,B_q$ diverge exponentially with time \cite{demons}. We have checked by a numerical integration of the Gross-Pitaevskii equations with spherical symmetry that the spatial distribution then acquires a structure of alternating shells of $a$ atoms and $b$ atoms (see Fig.\ref{fig:stab}). We suppose from now on that $g_{ab}<g_{aa},g_{bb}$. Instability may still occur in this case due to the periodic time dependence of the coefficients in the system Eq.(\ref{eq:syst}), as shown in \cite{Ralph}. We have studied in more detail the stability of the mode $q=5$, which is the one populated initially (see Eq.(\ref{eq:devt_t=0})); we have found non-zero instability exponents $\sigma$ ($C_5(t)\sim e^{\sigma t}$) in a very limited region of the plane $(g_{ab}/g_{aa},g_{bb}/g_{aa})$, with very small exponents ($\sigma<3\times 10^{-2}\omega$). A direct numerical integration of the Gross-Pitaevskii equations did not show any demixing of $a$ and $b$ even at times $\gg \sigma^{-1}$ \cite{Alice_est_contre}. This suggests that the finite instability exponent is an artifact of the Thomas-Fermi approximation. We assume in what follows the dynamical stability of the breathe-together solution. \begin{figure} \caption{Modulus squared of the condensate wavefunctions $|\phi^2_{a,b}|(N_a,N_b)$ as function of the distance $r$ to the trap center at a time $\omega t\simeq 29.5$, from a numerical solution of the coupled Gross-Pitaevskii equations in the case of a dynamically unstable breathe-together solution. We have taken $g_{bb}/g_{aa}=1.2$ and $g_{ab}/g_{aa}=1.5$. We have applied a deviation $\delta N_a = -\delta N_b=-0.05 \bar{N_a}$ from the exact breathe-together condition. The chemical potential is $\bar{\mu}=28.9\hbar\omega$. The curve in solid line corresponds to $\phi_a$, the dotted curve corresponds to $\phi_b$. \label{fig:stab}} \end{figure} \subsection{Phase dynamics} In order to calculate the functions $\chi_d,\chi_s$ relevant for the relative phase dynamics, we calculate the evolution of the deviation $\delta\varphi_d$ due to a small change in $N_a,N_b$ with respect to $\bar{N_a}, \bar{N_b}$, that is we evolve the initial state Eq.(\ref{eq:devt_t=0}) according to the results of the previous subsection. As we assume dynamical stability of the breathe-together solution, the modes with $q=5$ perform only oscillations in time \cite{proviso}. The relevant contribution for the phase dynamics therefore comes from the anomalous mode, which from Eq.(\ref{eq:ev_anom}) has a $\beta$ diverging linearly with time. Assuming $\beta(\vec{r},t) \sim C_{\mbox{\scriptsize anom}} \beta_{\mbox{\scriptsize anom}} (t)$ and using Eqs.(\ref{eq:coef}), (\ref{eq:sens_beta}) we obtain: \begin{equation} (\delta\theta_a -\delta\theta_b)(\vec{r},t) \sim -{2\bar{\mu}\over 5}{\bar{N_a}\bar{N_b} \over \bar{N}^2} {(g_{aa}+g_{bb}-2g_{ab})\over g_{aa}} \left({\delta N_a\over \bar{N_a}} -{\delta N_b\over \bar{N_b}}\right) \int_0^t {dt' \over\lambda^3(t')}. \end{equation} We specialize this formula with $\delta N_b=\pm\delta N_a$ and we get from Eq.(\ref{eq:chi1}), Eq.(\ref{eq:chi2}): \begin{eqnarray} \chi_d &\sim & -{1\over 2\hbar} \left({d\mu\over dN}\right)_{N=\bar{N}} {g_{aa}+g_{bb} -2 g_{ab}\over g_{aa}} \int_0^t {dt' \over\lambda^3(t')} \\ \chi_s &\sim & (|c_b|^2-|c_a|^2) \chi_d. \end{eqnarray} We have introduced the derivative of the chemical potential with respect to the total number of particles ($(d\mu/dN)(N=\bar{N})\simeq 2\bar{\mu}/5\bar{N}$ in the Thomas-Fermi limit) in order to recover the characteristic time scale for the phase collapse of steady state non mutually interacting condensates. Our formula reveals the interest of close coupling constants, such that $g_{aa}+g_{bb}-2g_{ab}\ll g_{aa}$. In this case $\chi_d$ is strongly reduced with respect to non mutually interacting condensates; $\lambda$ performs small oscillations around the value $\lambda=1$ so that the integral over $t'$ can be replaced by $t$. The more general case of close $g$'s not necessarily satisfying the breathe together condition is analyzed in the next section. We note that the value of $\chi_s$ as function of $\chi_d$ could be expected {\sl a priori} from Eq.(\ref{eq:magique}): when Eq.(\ref{eq:angle}) is satisfied, the condensate wavefunctions form a breathe-together solution and have therefore a vanishing relative phase for $N_a=N|c_a|^2,N_b=N|c_b|^2$, whatever the value of $N$ is. An important consequence is that there is no extra damping of the phase coherence due to the fluctuations of the total number of particles (see Eq.(\ref{eq:gaus_court})). \section{Case of close coupling constants} \label{sec:closeg's} We consider in this section the case of close coupling constants which leads to a dramatic reduction of the relative phase decoherence with respect to the case of non mutually interacting condensates. The strategy is to solve approximately the Gross-Pitaevskii equations Eq.(\ref{eq:gpe}) for $\phi_a(N_a,N_b)$ and $\phi_b(N_a,N_b)$ and apply the formulas Eq.(\ref{eq:chi_s},\ref{eq:chi_d}) directly. For all equal $g$'s the initial state is indeed a steady state for the Eq.(\ref{eq:gpe}) and $\chi_s=\chi_d=0$. For close $g$'s we linearize the Gross-Pitaevskii equations around the initial value in the hydrodynamic point of view. \subsection{Linearization in the classical hydrodynamics approximation} We first rewrite the Gross-Pitaevskii equations Eq.(\ref{eq:gpe}) in terms of the hydrodynamic variables: \begin{eqnarray} \rho_\varepsilon &\equiv& N_\varepsilon |\phi_\varepsilon(N_a,N_b)|^2 \\ \vec{v_\varepsilon} &\equiv& {\hbar\over m} \; \vec{\mbox{grad}}\; \theta_\varepsilon (N_a,N_b) \end{eqnarray} that is densities and velocity fields of the two condensates. We further assume the Thomas-Fermi limit $\mu \gg \hbar \omega$ and neglect the quantum pressure terms as in \cite{Stringari} in the time evolution of the velocity fields: \begin{eqnarray} \partial_t \rho_\varepsilon+\mbox{div}(\rho_\varepsilon\vec{v_\varepsilon} ) & = & 0 \label{eq:cont_ex}\\ \partial_t\vec{v_\varepsilon} +{1\over 2}\;\vec{\mbox{grad}}\;v_\varepsilon^2 & = & -{1\over m}\;\vec{\mbox{grad}}\;[U(\vec{r}) +\rho_\varepsilon g_{\varepsilon\varepsilon} +\rho_{\varepsilon'}g_{\varepsilon\varepsilon'}] . \label{eq:hydro} \end{eqnarray} At this point we introduce the deviations of the densities and velocity fields from their initial values: \begin{eqnarray} \rho_\varepsilon (t) & =& \rho_\varepsilon(0) +\delta\rho_\varepsilon(t)\\ \vec{v_\varepsilon} (t) & =& \vec{v_\varepsilon}(0) +\delta\vec{v_\varepsilon}(t) \end{eqnarray} where the initial values are given by: \begin{eqnarray} \rho_\varepsilon(t=0) &=& N_\varepsilon |\phi_0|^2(N) \\ \vec{v_\varepsilon}(t=0) &=& \vec{0}. \end{eqnarray} By expanding Eqs.(\ref{eq:cont_ex}),(\ref{eq:hydro}) to first order in the small quantities $\delta\rho_\varepsilon,\delta\vec{v_\varepsilon}$, we obtain: \begin{eqnarray} \partial_t \delta\rho_\varepsilon + \mbox{div}[N_\varepsilon |\phi_0|^2 \delta \vec{v_\varepsilon}] & = & 0 \label{eq:cont}\\ \partial_t \delta \vec{v_\varepsilon} +{1\over m}\;\vec{\mbox{grad}} \; [\delta\rho_\varepsilon g_{\varepsilon\varepsilon} +\delta\rho_{\varepsilon'} g_{\varepsilon\varepsilon'}] &=& -{1\over m}\;\vec{\mbox{grad}}\; [|\phi_0|^2] \times \nonumber\\ && (N_\varepsilon g_{\varepsilon\varepsilon}+ N_{\varepsilon'}g_{\varepsilon\varepsilon'}-Ng_{aa}). \label{eq:navier} \end{eqnarray} By taking the first time derivative of Eq.(\ref{eq:cont}) we eliminate the velocity field and we get: \begin{equation} \partial_t^2\delta\rho_\varepsilon +\sum_{\varepsilon'} M_{\varepsilon\varepsilon'} S[\delta\rho_{\varepsilon'}] +\sigma_\varepsilon =0 . \label{eq:rhoseul} \end{equation} The source terms of these inhomogeneous equations are: \begin{equation} \sigma_\varepsilon = -{N_\varepsilon\over m} \mbox{div}[|\phi_0|^2\; \vec{\mbox{grad}}\;|\phi_0|^2] (N_\varepsilon g_{\varepsilon\varepsilon}+N_{\varepsilon'} g_{\varepsilon\varepsilon'} - N g_{aa}). \end{equation} The homogeneous part of Eq.(\ref{eq:rhoseul}) involves the $2\times 2$ matrix $M$: \begin{equation} M = {1\over N g_{aa}} \left(\begin{array}{cc} N_a g_{aa} & N_a g_{ab} \\ N_b g_{ab} & N_b g_{bb} \end{array} \right) \end{equation} and the Stringari operator defined in Eq.(\ref{eq:stringari}). In order to solve Eq.(\ref{eq:rhoseul}) we introduce the eigenvectors $\vec{e}_{\pm}$ of the matrix $M$ with corresponding eigenvalues $g_{\pm}$. Consistently with our previous approximations, we calculate, to leading order in the differences between the coupling constants, the eigenvalues: \begin{eqnarray} g_+ &\simeq& g_{aa} \\ g_- &\simeq& {N_a N_b\over N^2}(g_{aa}+g_{bb}-2g_{ab}) \end{eqnarray} and the components of $(\delta \rho_a, \delta \rho_b)$ on the eigenvectors of $M$: \begin{eqnarray} \delta\rho_+ &\simeq& \delta\rho_a + \delta\rho_b \\ \delta\rho_- &\simeq& {N_b\over N}\delta\rho_a - {N_a\over N}\delta\rho_b . \end{eqnarray} For those linear combinations we get the decoupled equations: \begin{equation} \partial_t^2\delta\rho_{\pm} +{g_\pm\over g_{aa}} S[\delta\rho_\pm] +\sigma_\pm = 0. \label{eq:dec} \end{equation} To study the dynamics of the system we expand $\rho_{\pm}$ and the source terms $\sigma_\pm$ on the eigenmodes of the Stringari operator. It turns out that the source terms are simply proportional to the breathing mode $\alpha_{q=5}$ already introduced in Eq.(\ref{eq:mode5}). The solution of Eq.(\ref{eq:dec}) with the initial conditions $\delta \rho_{\pm}=\partial_t \delta \rho_{\pm}=0$ is then: \begin{equation} \delta\rho_\pm(\vec{r},t) = N|\phi_0(\vec{0})|^2 {\cal A}_\pm{g_{aa}\over g_\pm} [1-\cos\Omega_\pm t] \alpha_{q=5}(r) \label{eq:res} \end{equation} with eigenfrequencies and amplitudes given by: \begin{eqnarray} \Omega_\pm &=& \left({5g_\pm\over g_{aa}}\right)^{1/2}\omega\\ {\cal A}_+ &=& {N_a^2 g_{aa} + N_b^2 g_{bb} +2 N_a N_b g_{ab}\over N^2 g_{aa}} -1\\ {\cal A}_- &=& {N_a N_b\over N^2} \left[{N_a g_{aa}+N_b g_{ab}-N_b g_{bb}-N_a g_{ab}\over N g_{aa}}\right]. \end{eqnarray} We note that when the numbers of atoms $N_{a,b}$ satisfy the breathe-together condition Eq.(\ref{eq:angle}) the amplitude ${\cal A}_-$ vanishes as expected, since $\delta\rho_- \equiv 0$ in this case. \subsection{Validity of the linear approximation} In order for our linearized treatment to be valid the deviations $\delta \rho_{\pm}$ should remain small as compared to the initial densities. A first necessary condition to be satisfied is that the eigenfrequencies $\Omega_\pm$ should be real. This imposes the positivity of the matrix $M$, ensured by the positivity of its determinant: \begin{equation} g_{ab}^2 \leq g_{aa} g_{bb} . \label{eq:stab} \end{equation} This condition is known in the case of homogeneous mixtures of condensates as a stability condition against demixing \cite{homogene}. To the leading order in the difference between the coupling constants, the condition Eq.(\ref{eq:stab}) is equivalent to $g_{aa}+g_{bb}-2g_{ab}>0$. We note at this point that the amplitude ${\cal A}_-/g_-$ in the expression for $\delta \rho_-$ is a ratio of two small numbers. When this ratio is large the system can evolve far from its initial state even in the stable case $g_->0$: numerical solutions of the Gross-Pitaevskii equations confirm this expectation, showing the formation of a crater at the center of one of the condensates. We therefore have to impose a second condition: \begin{equation} |{\cal A}_\pm{g_{aa}\over g_{\pm}}| \ll 1 . \label{eq:seconde_cond} \end{equation} Finally the present treatment is based on the classical hydrodynamic approximation; by including the quantum pressure terms in the hydrodynamic equation for the velocity field one can show that this imposes on the eigenfrequencies $\Omega_-$: \begin{equation} \frac{\hbar \omega^2}{\mu} \ll \Omega_{-} \end{equation} (see also Appendix \ref{app:TF}). This condition can be violated even in the Thomas-Fermi limit, when the $g_-$ eigenvalue almost vanishes. In this case one has to include the quantum pressure terms; the decoupling property of $\delta \rho_\pm$ is unaffected; for the evolution of $\delta \rho_-$ similar results as in Eq.(\ref{eq:res}) are obtained; we find e.g.\ $\Omega_-\simeq 63\hbar\omega^2/8\mu$. \subsection{Phase dynamics} We assume that all the conditions for the validity of the linearized treatment are satisfied so that we can proceed to the analysis of the relative phase dynamics. To this aim we write the equation of evolution for the phases $\theta_\varepsilon$ of the condensate wavefunctions $\phi_\varepsilon$ in the classical hydrodynamic approximation: \begin{equation} \partial_t \theta_\varepsilon +{\hbar\over 2m} \left(\;\vec{\mbox{grad}}\;\theta_\varepsilon\right)^2= - [ U + g_{\varepsilon \varepsilon} \rho_\varepsilon + g_{\varepsilon \varepsilon'} \rho_{\varepsilon' }]/\hbar . \label{eq:theta_hydro} \end{equation} The equations for the velocity fields previously given are simply the gradient of Eq.(\ref{eq:theta_hydro}). By linearizing Eq.(\ref{eq:theta_hydro}) around the initial state $\theta_\varepsilon =0$ we obtain for the relative phase: \begin{eqnarray} \hbar \partial_t(\theta_a- \theta_b) &\simeq& - |\phi_0|^2 ( N_a g_{aa} + N_b g_{ab} - N_b g_{bb} - N_a g_{ab})\\ & +& (g_{ab} - g_{aa}) \delta \rho_a + (g_{bb} - g_{ab}) \delta \rho_b . \end{eqnarray} The right hand side of this equation is a sum of terms constant in time and of oscillatory functions of time. The function $\theta_a-\theta_b$ then has two components: an oscillating component and a component diverging linearly with time which will dominate for long times. By using the result Eq.(\ref{eq:res}) and the Thomas-Fermi approximation for $|\phi(0)|^2$ Eq.(\ref{eq:phi_TF}) we can calculate the time diverging component and we obtain to leading order in the $g$'s difference: \begin{equation} \theta_a - \theta_b \sim - \frac{2 \mu}{5 N g_{aa}} [N_a g_{aa} - N_b g_{bb} + (N_b -N_a) g_{ab} ] \; t/\hbar . \label{eq:evol_theta_hydro} \end{equation} We now use Eq.(\ref{eq:chi_d}) and Eq.(\ref{eq:magique}) to obtain: \begin{eqnarray} \chi_d &\sim & -{1\over 2} \left({d\mu\over dN}\right)_{N=\bar{N}} {g_{aa}+g_{bb} -2 g_{ab}\over g_{aa}}\;t/\hbar \\ \chi_s+(|c_a|^2-|c_b|^2)\chi_d & \sim& - \frac{2 }{5 g_{aa}} \left(\frac{d \mu}{dN}\right)_{N=\bar{N}} \times\nonumber \\ && (|c_a|^2 g_{aa} + |c_b|^2 g_{ab} -|c_b|^2 g_{bb} - |c_a|^2 g_{ab}) \;t/\hbar \end{eqnarray} where we introduced the derivative of the chemical potential with respect to the total number of particle $(d\mu/dN)(N=\bar{N}) \simeq (2/5) \bar{\mu}/\bar{N}$ in the Thomas-Fermi limit. As we already found in the particular case of the breathe-together solution the constants $\chi_d$ and $\chi_s$ governing the relative phase collapse are highly reduced for close $g$'s with respect to the case of non mutually interacting condensates. \subsection{Physical interpretation of the results} We now show that all the previous results of this section can be interpreted in terms of small oscillations of the condensates around the steady state. Let us introduce the steady state densities $\rho_{\varepsilon}^{\mbox{\scriptsize st}}$ for the condensates with $N_a$ particles in $a$ and $N_b$ particles in $b$. As we are in the case of quasi complete spatial overlap between the two condensates we can use the Thomas-Fermi approximation to determine these densities: \begin{eqnarray} \mu_a -U &=& \rho_a^{\mbox{\scriptsize st}} g_{aa} + \rho_b^{\mbox{\scriptsize st}} g_{ab} \\ \mu_b -U &=& \rho_a^{\mbox{\scriptsize st}} g_{ab} + \rho_b^{\mbox{\scriptsize st}} g_{bb} \end{eqnarray} where $\mu_\varepsilon$ are the chemical potentials in steady state. We rewrite these equations in terms of the deviations $\delta\rho_{\varepsilon}^{\mbox{\scriptsize st}}$ of the steady-state densities from the initial state densities $N_\varepsilon|\phi_0|^2$ and in terms of the deviations $\delta\mu_\varepsilon$ of the chemical potentials from $\mu$ defined in Eq.(\ref{eq:phi0}): \begin{eqnarray} \label{eq:syst_sta1} \delta\mu_a &= & (N_a g_{aa}+N_b g_{ab}-Ng_{aa}) |\phi_0|^2 +\delta\rho_a^{\mbox{\scriptsize st}} g_{aa} +\delta\rho_b^{\mbox{\scriptsize st}} g_{ab} \\ \delta\mu_b &= & (N_b g_{bb}+N_a g_{ab}-Ng_{aa}) |\phi_0|^2 +\delta\rho_a^{\mbox{\scriptsize st}} g_{ab} +\delta\rho_b^{\mbox{\scriptsize st}} g_{bb}. \label{eq:syst_sta2} \end{eqnarray} Using the fact that the spatial integral of $\delta\rho_\varepsilon$ vanishes, we get from integration of Eq.(\ref{eq:syst_sta1},\ref{eq:syst_sta2}) over the volume of $|\phi_0|^2$ the approximate relations: \begin{eqnarray} \delta\mu_a &=& {2\mu\over 5Ng_{aa}}(N_a g_{aa} + N_b g_{ab} -N g_{aa}) \\ \delta\mu_b &=& {2\mu\over 5Ng_{aa}}(N_b g_{bb} + N_a g_{ab} -N g_{aa}). \end{eqnarray} We can therefore check that the relative phase of the condensates in steady state, given by $\theta_a^{\mbox{\scriptsize st}}- \theta_b^{\mbox{\scriptsize st}}= -i(\delta\mu_a-\delta\mu_b)t/\hbar$, evolves as in Eq.(\ref{eq:evol_theta_hydro}). The phase decoherence properties of the evolving mixture are then essentially the same as in steady state. Moreover we now show that the average $\langle\delta\rho_\varepsilon \rangle$ of $\delta\rho_\varepsilon$ over the oscillations at frequencies $\Omega_\pm$ coincide with $\delta\rho_\varepsilon^{\mbox{\scriptsize st}}$. First, by averaging Eq.(\ref{eq:cont}) over time we find that the velocity fields have a vanishing time average \cite{details}. Second, we average Eq.(\ref{eq:navier}) over time; we find equations for the spatial gradient of $\langle\delta\rho_\varepsilon \rangle$, which coincide with the spatial gradient of Eq.(\ref{eq:syst_sta1}, \ref{eq:syst_sta2}), so that $\langle\delta\rho_\varepsilon\rangle =\delta\rho_\varepsilon^{\mbox{\scriptsize st}}$ \cite{proviso2}. \section{Discussion of the JILA case} \label{sec:JILA} In the JILA experiment the values of the three coupling constants between the atoms are known with good precision; they are in the ratio \cite{JILA_demix}: \begin{equation} g_{aa} : g_{ab} : g_{bb} = 1.03\ :\ 1\ :\ 0.97 . \end{equation} No breathe-together solution exists in this case, as $g_{ab}$ lies within $g_{aa}$ and $g_{bb}$. Experimentally half of the particles are in the state $a$ so that $|c_a|^2=|c_b|^2=1/2$, and the mean total number of particles is $\bar{N}=5\times 10^5$. Although the coupling constants are close, the linearized treatment presented in section \ref{sec:closeg's} does not apply either, because condition Eq.(\ref{eq:seconde_cond}) is violated. It is actually found experimentally that the two condensates evolve far from the initial state, with formation of a crater in the $a$ condensate while the $b$ condensate becomes more confined at the center of the trap; eventually the condensates separate in some random direction \cite{JILA_demix}. To avoid the crater formation and trigger the spatial separation of the two condensates in a reproducible direction a small spatial shift is applied to the trapping potential of one of the two states. The two condensates separate, with a relative motion exhibiting strongly damped oscillations \cite{JILA_demix}. The system then reaches a steady state that still exhibits phase coherence, up to times on the order of $150$ ms after the phase state preparation \cite{JILA_phase}. \subsection{Time dependent calculations} We have already studied in \cite{Nous} the damping of the relative motion between the condensates, by numerical integration of the coupled Gross-Pitaevskii equations Eq.(\ref{eq:gpe}). The agreement with the experimental results of \cite{JILA_demix} is qualitatively good, although the damping in the theory is weaker and incomplete, small oscillations of the condensate wavefunctions remaining undamped even at long times. We have applied the formalism of section \ref{sec:gen-method} by numerically integrating the Gross-Pitaevskii equations for the parameters of the JILA experiment. The coefficients $\chi_s,\chi_d$, now complicated functions of time and space, are obtained by evolving wavefunctions with slightly different numbers of atoms in $a$ and $b$. In order to facilitate the comparison with the experiments, in which the $x$-integrated atomic density $\bar{\rho}_a(y,z)$ in the internal state $a$ is measured after the $\pi/2$ pulse applied at time $\tau$, we calculated the following contrast: \begin{equation} C_{\mbox{\scriptsize JILA}}(y,z) = { \mbox{max}_\delta \bar{\rho}_a -\mbox{min}_\delta\bar{\rho}_a \over \mbox{max}_\delta\bar{\rho}_a +\mbox{min}_\delta\bar{\rho}_a} = {2|\int dx\;\langle\hat{\psi}_a^\dagger(\tau^-)\hat{\psi}_b(\tau^-)\rangle ^{\mbox{\scriptsize Gauss}}| \over \sum_{\varepsilon=a,b} \int dx\; \bar{N_\varepsilon}|\bar{\phi_\varepsilon}|^2 (\tau^-)} \end{equation} where the interference term Eq.(\ref{eq:inter_N}) is averaged over a Gaussian distribution of the total number of particles with a standard deviation $\Delta N$. A direct comparison with the experiment would require the inclusion of the 22 ms ballistic expansion, not included in the present numerical calculations. Our numerical result for $C_{\mbox{\scriptsize JILA}}$ at the center of the trap for the species $a$, $y=z=0$, is presented in Fig.\ref{fig:JILAa}, for Gaussian fluctuations in the total number of particles $\Delta N/\bar{N} =8 \%$ corresponding to the JILA experiment \cite{private}, together with the pure Gross-Pitaevskii prediction $C_{\mbox{\scriptsize GPE}}$ obtained by setting all the $\chi$'s to 0. The Gross-Pitaevskii prediction oscillates around $\langle C_{\mbox{\scriptsize GPE}}\rangle =0.63$. On the contrary the result of the more complete calculation including fluctuations in the relative and total number of particles exhibits a damping of the contrast, that we have fitted by convenience with the formula $C_{\mbox{\scriptsize JILA}} =C_0 e^{-\gamma t}$; we obtain $C_0\simeq \langle C_{\mbox{\scriptsize GPE}}\rangle$ and $\gamma^{-1}=0.42$s. Note the oscillatory aspect of the curves in Fig.\ref{fig:JILAa}. More understanding of the structure of the condensate wavefunctions given by Eq.(\ref{eq:gpe}) is required as this point: as detailed in \cite{Nous} $\bar{\phi_\varepsilon}$ is a sum of a smooth part, performing oscillations with frequencies expected to be close to eigenfrequencies of the steady state condensates \cite{verif}, and of a noisy quasi-stochastic part. The slow oscillatory structure evident on $C_{\mbox{\scriptsize GPE}}$ comes from this smooth oscillating part of the wavefunctions. We have also considered the ideal case of a well defined total number of particles. The numerical prediction for the contrast $C_{\mbox{\scriptsize JILA}}$ in this case corresponds to a very long lived phase coherence: after a time of 1 second, the contrast is still very close to the pure Gross-Pitaevskii prediction. \begin{figure} \caption{For the parameters of the JILA experiment (not including the 22 ms ballistic expansion), phase contrasts $C_{\protect\mbox{\protect\scriptsize JILA}}$ (lower curve) and $C_{\protect\mbox{\protect\scriptsize GPE}}$ (upper curve) defined in the text, at $y=z=0$, as function of time in seconds, for the evolving binary mixture, with $\Delta N=0.08\bar{N}$. \label{fig:JILAa}} \end{figure} \subsection{Steady state calculations and effect of particle losses} As the wavefunctions at long times perform mainly oscillations around the steady state we have also tried a much simpler steady state calculation (see subsection \ref{subsec:sta}). During the collapse time the contrast $C_{\mbox{\scriptsize JILA}}$ is a Gaussian in time Eq.(\ref{eq:gaus_court}), with an initial value $0.958$ and with a half-width $t_c$ at the relative height $e^{-1/2}$. We plot in Fig.\ref{fig:JILAb} the variation of $t_c$ as function of the standard deviation $\Delta N$. As we find $\chi_s/t=-7.7\times 10^{-5} $ s${}^{-1} $ and $\chi_d/t= -4.5\times 10^{-4}$ s${}^{-1} $, one has $|\chi_s|\simeq |\chi_d|/6$, so that relatively high values of $\Delta N$ are required to observe a significant effect of the fluctuations of the total number of particles on phase decoherence. For $\Delta N =0.08 \bar{N}$ the phase decoherence time is $t_c=0.32$ second, close to the result of the time-dependent calculation of Fig.\ref{fig:JILAa}. Note that for such a high value of $\Delta N/\bar{N}$ the decay of the phase contrast in Eq.(\ref{eq:gaus_court}) is essentially due to the first exponential factor accounting for the smearing of the phase by fluctuations of the total number of particles, the spreading of the phase for a fixed number of particles being very small ($\bar{N}\chi_d^2(t_c)/2\simeq 0.005$). We now briefly consider the issue of losses of particles. An intrinsic source of losses in the JILA experiment are the inelastic collisions between $a$ atoms and $b$ atoms, resulting in the simultaneous loss of two particles. We estimate the mean number $\langle\delta N\rangle$ of lost particles from the rate constant $K_2$ for binary inelastic collisions between the states $|F=1,m=-1\rangle$ and $|F=2,m=2\rangle$ \cite{overlap} and from a numerical calculation of the overlap integral $\int d^3\vec{r}\,|\bar{\phi_a}|^2|\bar{\phi_b}|^2$. For the JILA parameters we find $\langle\delta N\rangle/\bar{N}=0.04$ at time $t_c=0.32$ second. One could then naively expect the effect of losses on phase coherence to be comparable with the effect of fluctuations of $N$. To test this naive expectation we use the following simple model, inspired by the two-mode model developed in \cite{resur}, and focusing on the effect of the losses on the drift velocity $v(N)$ of the relative phase of the two condensates given in Eq.(\ref{eq:vdrift}). Imagine that the system has initially $\bar{N}$ condensate atoms and that $k$ binary inelastic collisions have taken place at times $t_1 < \ldots <t_k$ between time $0$ and time $t$. The shift of the relative phase during $t$ is then given by: \begin{equation} \Theta = \int_0^t d\tau\; v(N(\tau)) = v(\bar{N}) t_1 + v(\bar{N}-2) (t_2-t_1) +\ldots + v(\bar{N}-2k)(t-t_k). \end{equation} As we do in \cite{resur} we assume a constant mean number of collisions $\lambda$ per unit of time and we average the phase factor $e^{i\Theta}$ multiplying the interference term $\langle\hat{\psi_b}^\dagger\hat{\psi_a}\rangle$ over the probability distribution of the times $t_1,\ldots, t_k$ and of the number of loss events $k$, \begin{equation} P_t(t_1,\ldots,t_k;k) = \lambda^k e^{-\lambda t} \end{equation} to obtain: \begin{eqnarray} |\langle e^{i\Theta}\rangle| &=& \exp \left\{-\langle k\rangle[1-\sin(2\chi_s)/ (2\chi_s)]\right\} \\ &\simeq & \exp \left[-{2\over 3}\langle k\rangle\chi_s^2\right] \ \ \ \ \mbox{for} \ |\chi_s| \ll 1 \label{eq:perte_court} \end{eqnarray} where $2\langle k\rangle=2\lambda t=\langle \delta N\rangle$ is the mean number of lost particles during $t$. At time $t=t_c=0.32$ second the corresponding modulus of the averaged phase factor is on the order of $[1-4\times 10^{-6}]$, very close to one: particle losses have a negligible effect on the phase coherence at the considered time $t_c$, even if $\langle\delta N\rangle$ and $\Delta N$ have the same order of magnitude. Actually an inspection of the $\chi_s$ dependent factor in Eq.(\ref{eq:gaus_court}) and of Eq.(\ref{eq:perte_court}) reveals that these equations have the same structure; replacing in Eq.(\ref{eq:gaus_court}) the variance $\Delta N^2$ of the total number of particles by the variance $\Delta k^2$ of the number of loss events ($\Delta k^2=\langle k\rangle$ as $k$ obeys a Poisson law) one recovers Eq.(\ref{eq:perte_court}) up to a numerical factor inside the exponential. For equally large values of $\Delta N$ and $\langle k\rangle$ the effect of losses on phase coherence is less important than that of fluctuations of $N$ because $\Delta k^2 = \langle k\rangle \ll \Delta N^2$. We have also investigated another source of losses, the collisions of condensate atoms with the background gas of the cell. Assuming a lifetime of the particles in the cell of 250 seconds as in \cite{decay} we find as well that this loss mechanism has a negligible effect on the phase coherence for a time $t_c=0.32$ second. \begin{figure} \caption{For the parameters of the JILA experiment (except the 22 ms ballistic expansion), collapse time $t_c$ for $C_{\protect\mbox{\protect\scriptsize JILA}}$ at $y=z=0$ as a function of $\Delta N/\bar{N}$ for zero temperature steady state condensates in the shifted traps. \label{fig:JILAb}} \end{figure} \section{Conclusion and perspectives} We have extended previous treatments of the phase dynamics of Bose-Einstein condensates at zero temperature to the case of mutually interacting and dynamically evolving binary mixtures of condensates, for a measurement scheme of the phase coherence inspired by the JILA experiment. We have first applied this extended formalism to the interesting breathe-together solution of the Gross-Pitaevskii equations, in which the two condensates oscillate in phase, remaining always exactly spatially superimposed. The analytical results for the phase show that a dramatic increase of the phase coherence time can be obtained for close coupling constants $g_{aa},g_{ab},g_{bb}$ describing the elastic interactions between $a$ atoms and $b$ atoms. We have also treated analytically the case of close $g$'s, in the absence of demixing instability. Basically the phase collapse is identical to the steady state case for the two mutually interacting condensates. Finally, we have investigated numerically the more difficult case of JILA. We find a collapse time of the phase on the order of 0.4 second, both by a dynamical and a steady state calculation, in the case of Gaussian fluctuations of the total number of particles, corresponding to $\Delta N/\bar{N} = 8\%$. This result for the collapse time is significantly larger than the experimental results (no phase coherence measured after 150 ms). We have also estimated in a simple way the effect of collisional losses on phase coherence in the JILA experiment. A possible extension of this work could include the effect of the presence of a thermal component in the experiment. \section*{Acknowledgements} Part of this work (the breathe-together solution) would have not been possible without the contribution of G.\ Shlyapnikov, J.\ Dalibard and P.\ Fedichev. We thank A.\ Leggett, Y.\ Kagan for very useful discussions on the role of fluctuations in the total number of particles. We thank Ralph Dum for help in the numerical calculations. A.\ S.\ acknowledges financial support from the European Community (TMR individual research grant).
\section{Introduction} The exclusive process of photodisintegration of the deuteron addresses an interesting interplay of nuclear and particle physics. At low energies (say, below $E_\gamma$ = 0.5 GeV) conventional nuclear models based upon meson exchange which fit the NN phase shifts give a satisfactory description of both the energy dependence and the angular distribution of the experimental cross section \cite{Lee,Aren,Laget}. However, at higher energies nuclear potential models fail to explain the data \cite{Belz,Bochna}. This is not unexpected since at $E_\gamma >1$ GeV small distances of the order of 0.2 fm play a role. \\ Alternatively it has been attempted to describe the cross section in terms of quark-gluon degrees of freedom \cite{BF,BH,Kon}. A possible signature for the emergence of quark-gluon degrees of freedom would be the observation of the onset of scaling of the cross section \cite{BF}. Several examples where this happens have been found. For example, the $p(\gamma ,\pi) n$ reaction above 3 GeV appears to be consistent with counting rules at all angles {\cite{And}. \\ In case of the exclusive process $d(\gamma,p)n$ the dimensional counting analysis of \cite{BF} leads to a differential cross section of the form \begin{equation} d\sigma/dt = s^{-(n-2)} f(\theta), \end{equation} where $s (\theta)$ is the cm energy (angle), and $n$ denotes the number of elementary fields in initial and final state (i.e. $n=13$ in case of the deuteron). \\ Previous data from SLAC \cite{Belz} up to $E_\gamma$ =3 GeV at $\theta =90^\circ$ indicated a scaling behavior consistent with $n=13.1 \pm 0.3.$ More recent data from Jefferson Lab up to 4GeV confirmed \cite{Bochna} the scaling behavior with $n \approx 13$ at $\theta=90^\circ$ and $69^\circ; $ whereas at smaller angles, $\theta= 36^\circ , 52^\circ, $ the best fit yielded $n \approx$ 11.5 and 11.6, respectively. This constitutes a deviation from the scaling behavior predicted by simple counting rules. \\ A more refined approach is the reduced nuclear amplitude (RNA) approach \cite{BH} which is also based upon parton exchange between the two nucleons, but takes into account some finite mass and higher twist effects. However, if normalized at $E_\gamma =1.0 $GeV, the prediction falls below the data at $\theta=90^\circ $ for $E_\gamma > 3$GeV. On the other hand the quark-gluon string (QGS) model proposed in \cite{Kon} and based upon Regge phenomenology appears to describe the data \cite{Bochna} only at small $t$ values, corresponding to small angles. \\ This indicates that in practice the situation is more complex. \\ The aim of this paper is to study the question of the possible origin of the apparent scaling and scaling violation in more detail; in particular we address whether the occurrence of scaling is an exclusive pQCD phenomenom or whether it can arise from different mechanisms. \\ The approach in the present paper, in which the basic degrees of freedom are taken to be hadronic, is an extension of the one in ref. \cite{Nagorny} which predicted the cross section at $\theta=90^\circ$ for $E_\gamma >1$ GeV in fair agreement with experiment, but failed to describe the angle dependence. Using a simple covariant parametrization of the deuteron vertex in terms of a hard component (suggested by an effective counting rule) and imposing gauge invariance the cross section for large but not infinite $s$ shows a ``preasymptotic" scaling behavior at $\theta=90^\circ$. This resembles the counting rule prediction, however, with an angle dependent value of $n$ (which decreases away from $90^\circ$). \section{Formalism} The general covariant half off-shell d-pn vertex ($p^2_2=m^2$) can be expressed as \cite{SB} \begin{equation} A^\mu= \Gamma_1 \gamma^\mu +\frac{(p_1-p_2)^\mu}{2m} \Gamma_2 +\frac{\slash\hspace{-2.2mm}p_1-m} {2m}[\Gamma_3\gamma^\mu + \frac{(p_1-p_2)^\mu }{2m}\Gamma_4], \label{Avertex} \end{equation} where the $\Gamma_i$ are scalar functions: invariant form factors. For large virtualities the contribution of the first term in eq.(\ref{Avertex}), proportional to $\Gamma_1, $ dominates over the last three. \\ Assuming that at large $s$ only tree-type diagrams survive, the $ \gamma +d \rightarrow n+p$ (Born) amplitude can be written as the sum of the pole diagrams in the $s,t$ and $u$ channels ($\Gamma_1 \equiv G$): $$ T^{\mu}_{pole}= \xi_{\nu}(d) \bar{u}(p') [F^p_\mu \frac{ \slash\hspace{-2.2mm}p +m}{ p^2-m^2} \gamma^\nu G(-k_t^2) + F^n_\mu \gamma^\nu \frac{ \slash\hspace{-2.2mm}n-m}{n^2-m^2} G(-k_u^2) $$ \begin{equation} + G(-k_s^2)\gamma^\alpha \frac{ -g_{\beta \alpha} + d'_\beta d'_\alpha /d'^2}{ s-m_d^2} F^d_{\mu\nu\beta}] C \bar{u}^T(n'). \label{Tpole} \end{equation} Here $s$-,$t$-, $u$- variables for the d-pn vertex are: $k_t= (k_s-q)/2, \ k_u=(k_s+q)/2 $, \ $k_s=(p'-n')/2, \ p'=p+q, \ d'=d+q; $ $u(p)$ is the nucleon spinor, $\xi_{\nu}(d)$ the polarization vector of the deuteron ($\xi^{\nu}(d) d_{\nu} = 0$), $C $ is the charge conjugation operator, $F_\mu^i \ (i=n,p) $ denotes the electromagnetic coupling to the nucleon, $ F^i_\mu= (e_i+ \slash\hspace{-2.2mm}q \frac{ K_i}{2m}) \gamma^\mu $, and $ F^d_{\mu\nu\beta}$ the corresponding one for the deuteron \cite{Nagorny} ($f_d = 2\mu_d-1+Q_d$): $$-F^d_{\mu\alpha\beta}= 2d_\mu(g_{\alpha\beta} - q_\alpha q_\beta {f_d \over 2m_d^2}) +2 \mu_d(g_{\mu\alpha} q_\beta- g_{\mu \beta} q_\alpha) + (s-m_d^2)( g_{\mu\alpha} q_\beta+g_{\mu \beta} q_\alpha) {f_d \over 4m_d^2}.$$ Both $\gamma NN$ and $\gamma dd$ vertices satisfy the identities: $q^{\mu} F^i_{\mu} = e_i \slash\hspace{-2.2mm}q$, $q^{\mu} F^d_{\mu \alpha \beta} = -(s - m_d^2) g_{\alpha \beta}$. Note, $s$-channel accounts for the pole part of the $T$-matrix of the final state n-p interaction. \\ In the presence of a momentum dependent vertex function the pole diagrams themselves are not gauge invariant. Indeed, using the Ward-Takahashi identity for the 3-point electromagnetic vertices, eq.(\ref{Tpole}), and the Dirac equation one finds that the contraction of $q_\mu$ with the sum of the $s$-, $t$-, $u$-pole amplitudes does not vanish (i.e. the corresponding Born current is not conserved), if the strong d-pn vertices contain momentum dependent form factors $$q_{\mu} T^{\mu}_{pole} = - \xi_{\nu}(d) \bar{u}(p') \gamma^{\nu} [ G(-k^2_t) - G(-k^2_s) ] C \bar{u}^T(n') \neq 0.$$ Therefore the Born current in eq.(\ref{Tpole}) is not complete and a {\it contact} contribution (which should not contain any pole-type singularities!) must be added to provide current conservation on the tree-level. To this end we use {\it minimal insertion} of the gauge field directly into the d-pn vertex \cite {Nag89}, which gives rise to a {\it contact amplitude}, $T_c^{\mu}$: \begin{equation} T^{\mu}_c= \xi_{\nu}(d) \bar{u}(p') \gamma^\nu \int_0^1 \frac{d\lambda}{\lambda} \frac{ \partial} {\partial q_\mu} \{e_p G(-(k_s-\lambda q/2)^2) + e_n G(-(k_s+\lambda q/2)^2)\} C \bar{u}^T(n') . \label{Tc} \end{equation} Calculating the contraction $q_{\mu} T_c^{\mu}$ with $e_p=1$ and $e_n=0$, one can easily check that the total current is conserved \cite{Nagorny}: $ q_\mu (T^{\mu}_{pole}+ T^{\mu}_c) =0, $ irrespective of the explicit form of the strong form factor $G(-k^2)$, and hence the total amplitude is gauge invariant. \\ To proceed we need make a choice for the vertex function $G(-k_i^2).$ Previous work \cite{Lee} showed that in the energy region above $E_\gamma >$ 1GeV a conventional potential model wave function cannot describe the data. \\ In the present approach it is hypothesized that the d-np vertex can be separated into two parts: a {\it soft} part corresponding to conventional meson exchange theory, and a small {\it hard} component caused by short-range phenomena. It is assumed that the soft part describes all low-energy (static) properties of the NN system and provides the dominant contribution to the normalization of the bound state wave function, while the hard part dominates the cross section at large virtualities. Since the microscopic structure of the short-range dynamics is poorly known we will use an effective counting rule prescription \cite{Gross} to describe the hard part of the d-pn vertex: \begin{equation} G(p^2)= \frac{C}{(\Lambda^2/2+m^2-p^2)^g }, \label{vertex} \end{equation} where $p$ is the momentum of the off-shell nucleon, $C$ is a normalization parameter and $\Lambda$ is related to the inverse of the range. For the special case $g=3$ \cite{Gross} the three-pole vertex represents one meson propagator and two (monopole) nucleon-meson form factors. \\ At the large virtualities involved obviously relativistic effects play an important role. In practice there exist various relativistic formulations, such as the instant form (if}) formalism and the light-front (lf) approach. Whereas in an exact calculation these are expected to yield the same result, this is not true in a truncated Fock space scheme. Indeed it has been noted \cite{Coes,FS} that in lowest order (IA) the lf and if approaches lead to different results. In particular in the lowest order Fock states in the lf approach negative energy states do not enter. To illustrate this model dependence in the following we distinguish the covariant (instant form) and the light-front approach. \\ {\bf Covariant Approach: Instant form kinematics} \\ In terms of the variables $k_i$ the vertex in eq.(\ref{vertex}) takes on the form ($i=s,t,u)$ \begin{equation} G^{\rm{if}}(-k_i^2)= \frac{C}{ (\delta^2-k_i^2)^g}, \label{Gif} \end{equation} where $k^2_t=-\alpha \vec{k}^2$, $k_u^2=-(2-\alpha) \vec{k}^2, $ $ k_s^2=-\vec{k}^2$, with $\vec{k}^2 =s/4-m^2$ and $\alpha=2pq/(dq)= 1-\frac{k}{E_k} \cos \theta. $ Furthermore $\delta^2= \Lambda^2/4+ \alpha_0^2 $ with $\alpha_0^2 = m^2- m_d^2/4.$ \\ Substituting eq.(\ref{Gif}) in eq.(\ref{Tpole}) the cross section can be obtained straightforwardly. The absolute value of the cross section cannot be predicted and therefore the (only) parameter $C$ is fixed by fitting the data at $E_\gamma$=1 GeV. We have chosen this energy for the normalization since it is in this region that the microscopic d-np vertex gives a reasonable description of the absolute value of the empirical cross section. \\ The resulting cross section $s^{11} \frac {d\sigma}{dt} $ at $\theta=90^\circ$ is compared with data in fig. 1 for $g=3$ and $g=4$. For comparison the results for the reduced amplitude approach of ref \cite{BH} which is very close to that $g=4$, are also shown. It is seen that the observed overall energy dependence of the cross section in the energy region $1 < E_\gamma < $ 4GeV is described well with $g=3$. Only at the highest energies a larger value of $g$ would fit better. \\ Turning to the angular distribution one sees from fig. 2 (dot-dashed curve) that in the if approach the predicted angular dependence on $\theta$ increases rapidly away from $ 90^\circ.$ Although the data are available only for a few angles this is clearly not observed experimentally. \\ {\bf Light-front approach} \\ Here the light-cone variables $\alpha$ and $k_\perp$ have a direct physical interpretation as the longitudinal and transverse momentum fraction carried by the nucleon in the deuteron, with $ \vec{k}^2=( s-4m^2)/4= \frac{m^2(1-\alpha)^2+ k_{\perp}^2}{ \alpha(2-\alpha)}. $ We will oriente the normal vector of the light-cone hypersurface along $\vec{q}$ to suppress $Z$-graphs \cite{Nagorny}. In general strong lf form factors are functions of two variables, for which we can use any convenient pair from the set $(\alpha,k_\perp,k_3, \vec{k}^2)$. For simplicity we will assume a factorized form of the lf form factor: \begin{equation} G^{\rm lc}(\alpha, \vec{k}^2)= \frac{C} {( \beta^2-\vec{k}^2)^g} \phi_\kappa(\alpha), {\rm \ with \ } \phi_\kappa (\alpha)= \alpha^{-\kappa}. \label{Glc} \end{equation} Here the functional dependence on $\vec{k}^2$ is the same as in the instant form, eq.(\ref{Gif}). The function $\phi(\alpha)$ in eq.(\ref{Glc}) goes to unity in the nonrelativistic limit. The simplest choice is $\kappa=0,$ (as in ref \cite{FS}), but a more realistic choice is $\kappa=\frac{1}{2} $ \cite{Kar} (basically corresponding to the Wick Cutkovsky model). At $\theta= 90^\circ $ (where $\alpha=1$) both vertices are identical and the lf and if formalisms lead to the same cross cross section (apart from a slight difference coming from the contact term, eq.(\ref{Tc})). However, at $\theta \neq 90^\circ$ the form factors (6) and (7) reproduce a completely different dynamics. Note, that the angular dependence of the cross section arises mainly from the dependence of the arguments of $G$ in eq.(\ref{Tpole}) on $\alpha= (1 - \frac{k}{E_k} \cos \theta). $ \\ In $G^{\rm if}$ in eq.(\ref{Gif}) these arguments are $k_t^2= -\alpha \vec{k}^2$ and $k_u^2= -(2-\alpha) \vec{k}^2 $. This gives rise to a strong increase of the cross section at forward and backward angles, which has a dynamic origine. Namely because of the requirement of covariance in the instant form the vertex dependence on $\alpha$ and $\vec{k}^2$ effectively reduces to the dependence on one covariant variable only, say $k_t^2$ (or $k_u^2$); the latter is a function of both $s$ and $\theta$, and hence in this case the angle dependence is essentially dictated by the value of the vertex parameter $g$ in eq.(\ref{vertex}), which also determines the $s$- dependence. \\ On the other hand in the lf approach the angular distribution is flatter since in this case the vertex depends on two variables, $\alpha$ and $\vec{k}^2,$ which are in principle independent. Thus, a steep $s$-dependence of the cross section may be consistent with a flat $\theta$-dependence. One sees from the curves in fig.2 that with $\kappa =\frac{1}{2}$ the observed angular dependence at $E=3.2$ GeV is described well. \\ In fig.3 the calculated energy dependence of the cross sections at 4 different angles is compared with the data. While the overall agreement is reasonable the data at the forward angles suggest a steeper increase of the cross section with energy than predicted, and moreover there is a discrepancy for the highest energy at $69^\circ.$ Also shown are the results of the RNA \cite{BH} and (except at $69^\circ$) the QGS approach \cite{Kon}; the latter is expected to be applicable only at small angles. \section{Discussion} It is of interest to discuss the underlying mechanism for scaling in the present approach. Using eqs.(\ref{Avertex}-\ref{Tc}) we can express the cross section for large (but finite) $s$ \begin{equation} \frac{d\sigma}{dt}= \frac{m_dC} {\sqrt{s}(s-m_d^2)^{2g-1}(s-m_d^2)^{3/2}} f(\theta,s) \label{as}. \end{equation} For $\theta =90^\circ$ one has $\alpha=1$ and $f(\theta,s)=1$. In the relevant region of $s$ the contribution of $m_d^2$ in eq.(\ref{as}) is still non-negligible; for $g=3$ the calculated ``prescaling" behavior in the relevant energy region can be approximately written as $\frac{d\sigma}{dt} \approx s^{-n+2},$ with $n-2=10. $ \\ For other angles $f(\theta,s)$ depends upon $s$ through $\alpha(\theta,s)$ and therefore is a different function of $s$ at different angles, effectively giving an additional power of $s$ in eq.(8) when $\theta \neq 90^\circ$. Therefore only for $s \rightarrow \infty $ the longitudinal fractions $\alpha$ and $\alpha' = 2-\alpha$ do not depend upon $s$: $\alpha(s \rightarrow \infty) \rightarrow 1- \cos(\theta)$, and one expects an (angular independent) scaling. We find that in the present model in the energy region 1-4 GeV $n$ decreases slowly from $n-2 \approx 10 $ at $90^\circ$ to $n-2 \approx 8 $ at $10^\circ$ (or $170^\circ$). In practice from eqs.(\ref{Avertex}-\ref{Tc}) this behavior can be expressed as an angular dependence of the cross section on the lf variable $\alpha$, namely $\frac{d\sigma}{dt} \propto \alpha^{-2} \ (\alpha >1)$ or $\propto \ (1-\alpha)^{-2} \ (\alpha < 1). $ \\ In the past Brodsky \cite{BH} has discussed the concept of a ``hadron helicity conservation law". In case of the $\gamma+d \rightarrow p+n$ reaction it states that only helicity amplitudes satisfying the condition $\lambda_p + \lambda_n = \lambda_d$ contribute at $s \gg m^2$. Taking into account that only the Dirac coupling ($ \approx \gamma_{\mu}$) in the $\gamma NN$-vertex conserves hadron helicity, while Pauli coupling ($\approx \sigma_{\mu \nu}$) does not, we see that in the present approach asymptotic ``hadron helicity conservation" will occur only in case the Pauli couplings are negligible in the limit $s \gg m^2$, i.e. if the latter fall off at least as $1/s^2.$ Note that the gauge constraint for the 3-point EM Green function, which does not allow any form factors in the Dirac coupling (in case of a reducible vertex with real photon), does not lead to any restrictions for the Pauli one. \\ In practice helicity conservation predicts that for $s \gg m^2$ the cross section in the backward hemisphere receives no contribution from the neutron pole (located near 180$^\circ)$ but only from the deuteron pole which does not depend on the angle. Hence at backward angles the cross section would scale with $s$ independent of angle. On the other hand at forward angles, where we have a competition of two different pole contributions, namely the proton one (which depends on $\theta$ and $s$) and the deuteron one, one does not expect a unified $s$-dependence but rather an angle dependent scaling. For this reason it would be of interest to extend the experiment to backward angles.\\ As to the differences in the results in the lf and if formalisms we note that we used an effective counting rule for the hard part of the d-pn vertex; this is based on the assumption of a {\it fixed number of constituents} which is only well defined in the lf approach, but not in the if. Therefore we consider the results of the latter (in this particular model) less reliable at extreme angles, which involve large $t$-, $u$-virtualities. \\ We note that the deuteron vertex at large virtualities can in principle be addressed in more detail in semi-exclusive reactions, such as $ d(e,e'p)X$ at large $Q^2$ in which the spectator proton is observed in the backward hemisphere to avoid contamination from hadronization products. In the past this reaction has been proposed \cite{Melnit} to discriminate between various models for the nuclear EMC effect. In the IA the deuteron hadronic tensor factorizes in terms of a neutron structure function and a nuclear spectral function determining the variables $\alpha, k_\perp$ of the observed recoiling proton. Hence the cross section is directly proportional to the square of the deuteron vertex. In this respect we note that the relation between the spectral function and the deuteron wave function is different for the lf and if formalism \cite{Melnit,FS}. (In fact in the covariant formulation it is not well possible to define a proper normalization.) This difference tends to increase with increasing $\alpha$, and depend also on $k^2_\perp$ \cite{Lex}. \\ Finally we note that in this work we have assumed (as an extreme model) that the hard part of the NN interaction resides only in the (initial) deuteron vertex, and not in the final np state; in future work we will explore the contribution from hard contributions in the final np rescattering. \\ {\bf Acknowledgement \\} The authors are indebted to R.J. Holt, A. Radushkin and B.P.Terburg for stimulating discussions. This work is supported in part by the Stichting voor Fundamenteel Onderzoek der Materie (FOM) with financial support from the Nederlandse Organisatie voor Wetenschappelijk Onderzoek (NWO). \\ {\bf Figure Captions} \\ Fig. 1. Calculated energy dependence of $\frac{d\sigma}{dt}s^{11}$ at $\theta=90^ \circ $ for $g=3$ (dashed), 4 (dotted curve); the solid curve is the result from Brodsky and Hiller \cite{BH}; the data are from \cite{Belz,Bochna}. \\ Fig. 2. Calculated angular dependence of $s^{11} \frac{d\sigma}{dt}$ at $E_\gamma= 3.2$ GeV , for $g=3,$ in the light-front approach with $\kappa $=0 (solid),$\frac{1}{2}$ (dotted), 1 (dashed), and instant form (dashed-dotted curve). \\ Fig. 3. The scaled cross section $s^{11}\frac{d\sigma}{dt}$ as a function of photon energy for (a) $\theta =89^\circ $, (b) $69^\circ, $ (c) $52^\circ$, (d) $ 36^\circ$. The data are from ref. \cite{Belz,Bochna}, the solid curve shows the present results, the dashed curve those from \cite{Kon} (not available for $69^\circ),$ and the dotted curve those from \cite{BH}. \newpage \epsfxsize=20cm \epsfbox{DEUTFIG1A.PS} \newpage \epsfxsize=20cm \epsfbox{DEUTFIG2A.PS} \newpage \epsfxsize=18cm \epsfbox{DEUTFIG3A.PS}
\section{Introduction} \indent The pure leptonic decays of the heavy meson $B_c$ are very interesting\cite{cch,du,lcd,aliev,aliev1,lih,gch}. In principle, the pure-leptonic decay $B_{c}{\longrightarrow} {\ell}{{\nu}_{\ell}}$(see, Fig.1) can be used to determine the decay constant $f_{B_{c}}$ if the fundamental Cabibbo-Kobayashi-Maskawa matrix element $V_{bc}$ of Standard Model (SM) is known. Conversely if we know the value of decay constant $f_{B_{c}}$ from other method, these process also can be used to extract the matrix element $V_{bc}$. The $B_{c}$ meson has recently been observed at Fermilab\cite{cdf}, that opens a `new page' for the experimental study of $B_c$ meson. Based on the estimates in Refs.\cite{cch1,k}, numerous $B_{c}$ mesons will be produced in Tevatron and also in LHC. Considering the schedule for the new runs at Tevatron and LHC constructing, we may expect that experimental studies on the $B_{c}$ meson with more care and much higher statistics will be accessible in the foreseeable future. The pure leptonic weak decays of the pseudoscalar meson corresponding to Fig.1 are helicity suppressed by factor of $m_{\ell}^{2}/m_{B_{c}}^{2}$: \begin{equation} \Gamma(B_{c}{\longrightarrow}{\ell}{\overline{\nu}_{\ell}})= {\frac{G_{F}^{2}}{8\pi}}|V_{bc}|^{2}f_{B_{c}}^{2}m_{B_{c}}^{3}{\frac {{m_{\ell}}^{2}}{{m_{B_{c}}}^{2}}} \left(1-{\frac{{m_{\ell}}^{2}}{{m_{B_{c}}}^{2}}}\right)^{2}, \end{equation} whereas to study them with their radiative decays $B_{c}{\longrightarrow} {l}{\nu}_l\gamma$ simultaneously is attractive\cite{lcd,aliev,aliev1}. Of them only the process $B_{c}{\longrightarrow} {{\tau}}{\nu}_{\tau}$ is special i.e. it does not suffer so much from the helicity suppression thus its branching ratio may reach to $1.5\%$\cite{cch} in SM. However the produced $\tau$ will decay promptly and one more neutrino is generated in the cascade decay at least, thus it makes the decay channel difficult to be observed, especially, in such a strong background of hadronic collisions. Fortunately, having an extra real photon emitted in the leptonic decays, the radiative pure leptonic decays can escape from the suppression, furthermore, as pointed out in \cite{lcd}, with the extra photon to identify the produced meson $B_c$ experimentally in hadronic collisions from the backgrounds has advantages, namely in Tevatron and LHC to observe the radiative decays certainly has advantages to compare with the pure leptonic ones. Although the radiative corrections are suppressed by an extra electromagnetic coupling constant $\alpha$, it will not be suppressed by the helicity suppression. Therefore, the radiative decay may be comparable, even larger than the corresponding pure leptonic decays. Considering the possibility to measure the decay constant $f_{B_c}$ and the CKM matrix element $V_{cb}$ in LHC in the foreseeable future, the problem to increase the accuracy of the theoretical calculation, at least up to the first order radiative corrections, emerges. The radiative pure leptonic decays, theoretically, have inferred divergences and will be canceled with those from loop corrections of the pure leptonic decays, thus we are also interested in considering the radiative decays and the pure leptonic decays with one-loop radiative corrections together. Moreover, all of them can be divided into two components: 1). The so-called short distance contributions, for instance, those of the radiative pure leptonic decays correspond to the four diagrams in Fig.2, i.e. a real photon is attached to any of the charged lines of the pure-leptonic decay Feynman diagrams Fig.1. 2). The so-called long distance contributions which are relevant to a virtual heavy hadronic state as an intermediate state. To increase the accuracy of the theoretical calculation, both components should consider precisely. Recently, there are a few papers\cite{lcd,aliev,aliev1} on the radiative processes with different methods in literature, but inconsistent results have been obtained no matter all of them just the lowest order calculations. To clarify the inconsistency, i.e. further to re-examine the process is certainly needed. Furthermore, since in Ref.\cite{aliev} an approach of QCD sum rule is adopted and in Ref.\cite{lcd} a kind of quark model is adopted, QCD sum rule is supposed some long distance effects may be taken into account, so it is reasonable to doubt that the disagreement between the results of Refs.\cite{lcd,aliev}could be due to the long distance contributions. We essentially will adopt the same approach as Ref.\cite{lcd}, thus in the paper we especially make some estimates on the long distance contributions precisely. Namely in the present paper, baring the problems and the situation pointed out above in mind, we will investigate the $B_{c}$ leptonic decays up-to one loop radiative corrections carefully. The paper is organized as follow: in Section II, we present the calculations of all the short distance contributions: {\it i)} besides the leptonic decays shown as Fig.1, the radiative decays into three family leptons with a real photon appearing in the final state as shown in Fig.2; {\it ii)} the virtual photon corrections to the pure leptonic decays i.e. a photon in loop as shown in Figs.3.1, 3.2, 3.3, 3.4. In Section III, we estimate a typical contributions from the long distance contributions. In Sec.IV, we evaluate the values of the decays. As uncertainties the dependence of the results on the parameters appearing in the considered model is discussed. Comparison with the others' results are made. Finally, preliminary conclusion is obtained. \section{The Short Distance Contributions} Let us start with the radiative decay. The short distance contributions are corresponding to the four diagrams in Fig.2. According to the constituent quark model which is formulated by Bethe-Salpeter (B.-S.) equation, the amplitude turns out to be the four terms $M_i (i=1,2,3,4)$: $$M_1=Tr \left[\int\frac{d^{4}q}{(2\pi)^{4}}\chi(p, q) i \left(\frac{G_{F}m_{w}^{2}}{\sqrt{2} }\right)^{\frac{1}{2}}{\gamma}_{\mu}(1-\gamma_{5})V_{bc}\right]\times $$ $${\frac{ i \left(-g^{\mu\nu}+ \frac{p^{\mu}p^{\nu}}{m_{w}^{2}}\right)}{p^{2}-m_{w}^{2}}} i e [(p'+p)_{\lambda}g_{\nu\rho}+ (k-p')_{\nu}g_{\rho\lambda}+(-p-k)_{\rho}g_{\nu\lambda} ] {\epsilon}^{\lambda}\times$$ \begin{equation} \frac{ i \left(-g^{{\rho}{\sigma}}+ \frac{(p-k)^{\rho}(p-k)^{\sigma}}{m_{w}^{2}}\right)}{(p-k)^{2}- m_{w}^{2}}\overline { \ell }\frac{ig}{2\sqrt{2}}\gamma_{\sigma} (1-{\gamma}_5)\nu_\ell, \end{equation} $$M_2=Tr \left [ \int\frac{d^{4}q}{(2\pi)^{4}}\chi(p, q) i \left (\frac{G_{F}m_{w}^{2}}{\sqrt{2} }\right )^{\frac{1}{2}}{\gamma}_{\mu}(1-\gamma_{5})V_{bc} \right ]{\frac{ i \left(-g^{\mu\nu}+ \frac{p^{\mu}p^{\nu}}{m_{w}^{2}}\right)}{p^{2}-m_{w}^{2}}}$$ \begin{equation} \times\overline { \ell }(-ie)\not\! { \epsilon } \frac{i}{\not\! {k}_{\ell}-m_{\ell}}\frac{ig}{2\sqrt{2}}\gamma_{\nu}(1-{\gamma}_5)\nu_\ell, \end{equation} $$M_3=Tr \left [ \int\frac{d^{4}q}{(2\pi)^{4}}\chi(p, q) i\left (\frac{G_{F}m_{w}^{2}}{\sqrt{2} }\right )^{\frac{1}{2}}{\gamma}_{\mu}(1-\gamma_{5})V_{bc} \frac{i}{\frac{m_{b}}{m_{b}+m_{c}}\not\! {p} +\not\! {q}-\not\! {k}-m_{b}}\left(-i \frac{e}{3}\not\! {\epsilon}\right)\right ] $$ \begin{equation} \times\frac{ i \left(-g^{{\mu}{\sigma}}+ \frac{(p-k)^{\mu}(p-k)^{\sigma}}{m_{w}^{2}}\right)}{(p-k)^{2}- m_{w}^{2}}\overline { \ell }\frac{ig}{2\sqrt{2}} \gamma_{\sigma}(1-{\gamma}_5)\nu_\ell, \end{equation} $$M_4=Tr \left [ \int\frac{d^{4}q}{(2\pi)^{4}}\chi(p, q) \left( i \frac{2e}{3}\not\! {\epsilon}\right) \frac{i}{-(\frac{m_{c}}{m_{b}+m_{c}}\not\! {p} +\not\! {q}-\not\! {k})-m_{c}} i \left (\frac{G_{F}m_{w}^{2}}{\sqrt{2} }\right )^{\frac{1}{2}}{\gamma}_{\mu}(1-\gamma_{5})V_{bc}\right] $$ \begin{equation} \times\frac{ i \left(-g^{{\mu}{\sigma}}+ \frac{(p-k)^{\mu}(p-k)^{\sigma}}{m_{w}^{2}}\right)}{(p-k)^{2}- m_{w}^{2}}\overline { \ell } \frac{ig}{2\sqrt{2}}\gamma_{\sigma} (1-{\gamma}_5)\nu_\ell, \end{equation} where $\chi(p, q)$ is Bethe-Salpeter wave function of the meson $B_{c}$; $p$ is the momentum of $B_{c}$; ${\epsilon}, k$ are the polarization vector and momentum of the emitted photon. In the quark model, the momenta of $\bar {b}, c$-quarks inside the bound state i.e. the $B_{c}$ meson, are: $$\displaystyle p_{b}={\frac{m_{b}} {m_{b}+m_{c}}}p+q; \;\;\; p_{c}={\frac{m_{c}} {m_{b}+m_{c}}}p-q ,$$ where $q$ is the relative momentum of the two quarks inside the $B_{c}$ meson. As the $B_c$ meson is a nonrelativistic bound state in nature, so the higher order relativistic corrections may be computed precisely, but being an approximation for a $S$-wave state, and focusing the light on the radiative decay corrections only at this moment now, we ignore $q$ dependence i.e. we may still have the non-relativistic spin structure for the wave function of the meson $B_c$ (a $^1S_0$ state) correspondingly: $$\int\frac{d^{4}q}{(2\pi)^{4}}\chi(p, q)= \frac{\gamma_{5}({/}\!\!\! {p}{+m})}{2\sqrt{m}}\psi(0).$$ Here $\psi(0)$ is the wave function at origin, and by definitions it connects to the decay constant $f_{B_{c}}$: $$f_{B_{c}}=\frac{\psi(0)}{2\sqrt {m}},$$ where m is the mass of $B_{c}$ meson. Moreover we note that for convenience we take unitary gauge for weak bosons to do the calculations throughout the paper. With a straightforward computation, the amplitude can be simplified as: $$M_{1}={\frac{-4Ai}{(m^{2}-m_{w}^{2})(m^{2}-2p\cdot k-m_{w}^{2})}}\times$$ \begin{equation} \overline {\ell}\left (-1+\frac{m^{2}}{m_{w}^{2}}\right ) \left[ p\cdot\epsilon(\not\! {k}-\not\! {p})-(2p\cdot{k}-m^{2}) \not\! {\epsilon} \right](1-\gamma_{5})\nu_{\ell}, \end{equation} \begin{equation} M_{2}={\frac{4Ai}{\left(m^{2}-m_{w}^{2}\right)(2k_{1}\cdot{k})}} \overline {\ell}\left(-1+\frac{m^{2}}{m_{w}^{2}}\right) \not\! {\epsilon}(\not\! {k}_{1}+\not\! {k}+m_{e})\not\! {p}(1-\gamma_{5})\nu_{\ell}, \end{equation} \begin{equation} \begin{array}{lcr} M_{3}+M_{4}&=&\frac{-4Ai}{(-p\cdot{k})(m^{2}-2p\cdot k-m_{w}^{2})} \overline {\ell} \left\{\left[-(p\cdot{\epsilon})\not\! {p}+ (p\cdot{\epsilon}\not\! {k}-p\cdot{k}\not\! {\epsilon})s_{2} \right.\right.\\ [2mm] &+& \left. \left. is_{1}\varepsilon^{\alpha\mu\beta\nu}p_{\alpha} \epsilon_{\mu}k_{\beta}\gamma_{\nu}\right]+ \frac{(\not\! {p}-\not\! {k})}{m_{w}^{2}}p\cdot{\epsilon}(m^{2}-p\cdot{k}) \right\}(1-\gamma_{5})\nu_{\ell}, \end{array} \end{equation} where $k_{1}$ is the momentum of the charged lepton, and $$A=\frac{\psi(0)\left(\frac{G_{F}m_{w}^{2}}{\sqrt{2}}\right)^{\frac{1}{2}} V_{bc}eg}{2\sqrt{m}2\sqrt{2}}= \frac{\psi(0)\left(\frac{G_{F}m_{w}^{2}}{\sqrt{2}}\right)V_{bc}e}{2\sqrt{m}};$$ $$s1=-\frac{m_{b}+m_{c}}{6m_{b}}+\frac{m_{b}+m_{c}}{3m_{c}}; \;\; s2=\frac{m_{b}+m_{c}}{6m_{b}}+\frac{m_{b}+m_{c}}{3m_{c}}.$$ \par As a matter of fact, there is infrared infinity when performing phase space integral about the square of matrix element at the soft photon limit. It is known that the infrared infinity can be cancelled completely by that of the loop corrections to the corresponding pure leptonic decay $B_c\to l\nu$. In Eqs.(7) and (8), the infrared terms can be read out: \begin{equation} M^{i}=M_{2}^{i}+M_{3}^{i}+M_{4}^{i} =\frac{4Ai}{m_{w}^{2}}\overline{\ell}\left[\frac{k_{1}\cdot{\epsilon}\not\! {p}}{k_{1}\cdot{k}}- \frac{p\cdot{\epsilon}\not\! {p}}{p\cdot{k}}\right](1-\gamma_{5})\nu_{\ell}. \end{equation} As the diagrams (g), (h), (i), (j) in Figs.3.3, 3.4 always have a further suppression factor $m^{2}/{m_{w}}^{2}$ to compare with the other loop diagrams, we may ignore the contributions from these four diagrams savely. Furthermore we should note that in our calculations throughout the paper, the dimensional regularization to regularize both infrared and ultraviolet divergences is adopted, while the on-mass-shell renormalization for the ultraviolet divergence is used. If Feynman gauge for photon is taken (we always do so in the paper), the amplitude corresponding to the diagrams (a), (b) of Fig.3.1 can be written as: $$M_{(2)}(a)={\frac{2}{3}}eA\int\frac{d^{4}l}{(2\pi)^{4}} \left[\frac{-4i\varepsilon^{\alpha\mu\beta\nu}p_{\alpha}l_{\beta}-4(p_{\mu}l_{\nu} -{p\cdot l}g_{\mu\nu}+p_{\nu}l_{\mu})+\frac{8m_{c}}{m_{b}+m_{c}}p_{\mu}p_{\nu}} {l^{2}(l^{2}-2p\cdot l-m_{w}^{2})(l^{2}-{\frac{2m_{c}}{m_{b}+m_{c}}}p\cdot {l} )[l^{2}-2l\cdot {(p-k_{2})}]}\right]$$ \begin{equation} \times\overline{\ell}[2(p-k_{2})_{\mu}-\gamma_{\mu}{\not\! {l}}] (-\gamma_{\nu})(1-\gamma_{5})\nu_{\ell}, \end{equation} $$M_{(2)}(b)={-\frac{1}{3}}eA\int\frac{d^{4}l}{(2\pi)^{4}} \left[\frac{-4i\varepsilon^{\alpha\mu\beta\nu}p_{\alpha}l_{\beta}+4(p_{\mu}l_{\nu} -{p\cdot l}g_{\mu\nu}+p_{\nu}l_{\mu})-\frac{8m_{b}}{m_{b}+m_{c}}p_{\mu}p_{\nu}} {l^{2}(l^{2}-2p\cdot l-m_{w}^{2})(l^{2}-{\frac{2m_{b}}{m_{b}+m_{c}}}p\cdot {l} )[l^{2}-2l\cdot {(p-k_{2})}]}\right]$$ \begin{equation} \times\overline{\ell}[2(p-k_{2})_{\mu}-\gamma_{\mu}{\not\! {l}}] (-\gamma_{\nu})(1-\gamma_{5})\nu_{\ell}, \end{equation} where the $l$, $k_{2}$ denote the momenta of the loop and the neutrino respectively. These two terms also have infrared infinity when integrating out the loop momentum $l$. After doing the on-mass-shell subtraction, the terms corresponding to vertex and self-energy diagrams (c), (d), (e), (f) can be written as: $$M_{(2)}(c+d+e+f)=\frac{ieA}{4\pi^{2}}\overline{\ell}{\not\! p} (1-\gamma_{5})\nu_{\ell}\times\left[ ln(4)-\frac{8}{9}+\frac{2}{9}{\frac{m_{b}-m_{c}}{m_{b}+m_{c}}} ln\left(\frac{m_{b}}{m_{c}}\right)\right.$ $$+\left(\frac{2}{9}+\frac{8}{9}{\frac{m_{c}}{m_{b}+m_{c}}}\right) ln\left(\frac{m_{b}+m_{c}}{m_{b}}\right) +\left(\frac{8}{9}+\frac{8}{9}{\frac{m_{c}}{m_{b}+m_{c}}} \right)ln\left(\frac{m_{b}+m_{c}}{m_{c}}\right)$ \begin{equation} +\left.\frac{2}{\varepsilon_{I}}-2\gamma+ ln\left(\frac{4\pi\mu^{2}}{m^{2}}\right)+ ln\left(\frac{4\pi\mu^{2}}{m_{e}^{2}}\right)\right]. \end{equation} Now let us see the cancellation of the infrared divergencies precisely. The infrared parts of the decay widths which are from the interference of the self-energy and vertex correction diagrams with the tree diagrams: $$\delta\Gamma_{s,v}^{infrared}= \displaystyle\left(\frac{\alpha V_{bc}^{2}f^2_{B_{c}}G^{2}_{F}Mm^{2}_{\ell}}{16{\pi}^2}\right)\left[ -\frac{34}{9}-\frac{4}{9}{\frac{m_{b}-m_{c}}{m_{b}+m_{c}}}ln\left(\frac{m_{b}} {m_{c}}\right)\right.$$ $$-\left(\frac{2}{9}+\frac{4}{9}{\frac{m_{c}}{m_{b}+m_{c}}}\right)ln\left(\frac{ m_{b}+m_{c}}{m_{b}}\right) -\left(-\frac{4}{9}+\frac{4}{9}{\frac{m_{b}}{m_{b}+m_{c}}}\right)ln\left(\frac{m _{b}+m_{c}}{m_{c}}\right)$$ \begin{equation} -\left.\frac{2}{\varepsilon_{I}}+2\gamma-ln\left(\frac{4\pi\mu^{2}}{m^{2}}\right )- ln\left(\frac{4\pi\mu^{2}}{m_{e}^{2}}\right)\right] \; ; \end{equation} the infrared part of the decay widths from the interference of the "box" correction diagrams with the tree diagrams: \begin{equation} \delta\Gamma_{box}^{infrared}= \left(\frac{\alpha V_{bc}^{2}f^2_{B_{c}}G^{2}_{F}Mm^{2}_{\ell}} {16{\pi}^2}\right)\left[ -\left(\frac{1}{\varepsilon_{I}}-\gamma\right)ln\left(\frac{m^2_{\ell}}{M^2} \right)- ln\left(\frac{m^2_{\ell}}{M^2}\right)ln\left(\frac{4\pi\mu^2}{M^2}\right)\right]\; ; \end{equation} the infrared part of the decay width from the real photon emission: $$ \delta\Gamma_{real}^{infrared}= \left(\frac{\alpha V_{bc}^{2}f^2_{B_{c}}G^{2}_{F}Mm^{2}_{\ell}} {16{\pi}^2}\right)$$ \begin{equation} \cdot \left[\frac{2}{\varepsilon_I}-2\gamma+\left(\frac{1}{\varepsilon_{I}}- \gamma\right)ln\left(\frac{m^2_{\ell}}{M^2}\right) +\left( 2+ ln\left( \frac{m^2_{\ell}}{M^2} \right) \right) ln \left(\frac{4\pi\mu^2}{4(\Delta E)^2} \right) \right] \;. \end{equation} Here $\Delta{E}$ is a small energy, which corresponds to the experimental resolution of a soft photon so that the phase space of the emitting photon in fact is divided into a soft and a hard part by $\Delta{E}$. Where $\mu$ is the dimensional parameter appearing in the dimensional regularization. It is easy to check that when adding up all the parts: the real photon emission $\delta\Gamma_{real}^{infrared}$ and the virtual photon corrections $\delta\Gamma_{s,v}^{infrared}, \delta\Gamma_{box}^{infrared}$, the infrared divergences $\frac{2}{\varepsilon_{I}}-\gamma$ are canceled totally and the $\mu$ dependence is also cancelled. Hence we may be sure that we finally obtain the pure leptonic widths for the $B_c$ meson decays to the three families of leptons which are accurate up-to the `next'-order corrections and `infrared finite' but depend on the experimental resolution $\Delta{E}$. \section{Estimate of the Long Distance Contributions} We have done the calculations of the radiative `short distance' corrections corresponding to the energy scale around $m_{B_c}$, whereas, there are corrections which correspond to much softer nature than what we have considered. Typically, some of them in the radiative decay $B_c \rightarrow l\nu\gamma$ may be described by Fig.4. People, generally, call the soft corrections as the `long distance contributions' correspondingly. To have a rough estimate on the contributions of the softer part and to be typical, let us only consider those corresponding to Fig.4. The amplitude for the long distance radiative corrections corresponding to Fig.4 may be written as: $$M={\frac{-ie(2\pi)^{4}\delta^{4}(P_{\small{B}_{c}}-k-k_{1}-k_{\nu}) G_{F}\epsilon_{\mu}(k)}{\sqrt 2}}\overline{\ell}(k_{1})\gamma^{\lambda} (1-\gamma_{5})\nu_{e}(k_{\nu})$$ \begin{equation} \cdot \left[\sum\limits_{\stackrel{\rightarrow}{p_{n}}=\stackrel{\rightarrow}{k}} \frac{\langle{0}|j^{\mu em}(0)|p_{n}\rangle\langle{p_{n}}|j_{q\lambda}^{+}(0)|P_ {B_{c}}\rangle}{2p_{n0}(k_{0}-p_{n0}+i\epsilon)}+ \sum\limits_{\stackrel{\rightarrow}{p_{n}}=\stackrel{\rightarrow} {P_{\small{B}_{c}}}-\stackrel{\rightarrow}{k}} \frac{\langle{0}|j_{q\lambda}^{+}(0)|p_{n}\rangle\langle{p_{n}}|j^{\mu em}(0)|P_ {\small{B}_{c}}\rangle}{2p_{n0}(P_{\small{B}_{c0}}-k_{0}-p_{n0}+i\epsilon)} \right],\\[2mm] \end{equation} where $j^{\mu em}$, $j_{q\lambda}^{+}$ are electromagnetic current and weak current respectively, and $\epsilon(k)$ is polarization vector of the real photon. The intermediate states $|p_{n}\rangle$ are all the possible physical states, and in the summation $\stackrel{\rightarrow}{p_{n}}=\stackrel{\rightarrow}{k}$ for the first term and $\stackrel{\rightarrow}{p_{n}}= \stackrel{\rightarrow}{P_{B_{c}}}-\stackrel{\rightarrow}{k}$ for the second term are kept. Note that for the time-component, generally, $k_{0}\neq p_{n0}$ for the first term and $p_{n0}\neq P_{B_{c0}}-k_{0}$ for the second term. To compute the whole amplitude, let us compute the current matrix elements appearing in Eq.(16) with the intermediate meson states being on mass-shell. However to have a rough estimate instead of a precise one, here only the intermediate meson states $J/\psi$ and $B_{c}^{*}$, being of typical long distance contributions, are computed by means of the three Feynman diagrams shown in Fig.4. According to Eq.(16), we have: \begin{equation} \langle{0}|j^{\mu em}(0)|p_{J/\psi}\rangle=-{\frac{2}{3}}{\frac{1} {\sqrt{2\pi p_{J/\psi 0}}}}\phi_{J/\psi}(0)M_{J/\psi}\epsilon_{\mu}(J/\psi), \end{equation} \begin{equation} \langle{0}|j_{q\lambda}^{+}(0)|p_{B^{*}_{c}}\rangle=-{\frac{1} {\sqrt{2\pi p_{B^{*}_{c} 0}}}}\phi_{B^{*}_{c}}(0) M_{B^{*}_{c}}\epsilon_{\mu}(B^{*}_{c}). \end{equation} The matrix element $\langle p_{J/\psi}|j_{q\lambda}^{+}(0)|P_{B_{c}}\rangle$ can be decomposed into two parts: \begin{equation} \langle p_{J/\psi}|j_{q\lambda}^{+}(0)|P_{B_{c}}\rangle= \langle p_{J/\psi}|V_{\lambda}|P_{B_{c}}\rangle -\langle p_{J/\psi}|A_{\lambda}|P_{B_{c}}\rangle, \end{equation} namely those of the vector current $V_{\lambda}={\frac{1}{2}}\overline{c} \gamma _{\lambda}b$ and the axial current $A_{\lambda}={\frac{1}{2}}\overline{c} \gamma _{\lambda}\gamma_{5}b$. They are related to the form factors\cite{cch,s11,s12} as follows: \begin{equation} \langle p_{J/\psi}|V_{\mu}|P_{B_{c}}\rangle= ig{\frac{1}{2}}\varepsilon_{\mu\nu\rho\sigma} \epsilon^{*\nu}(P_{B_{c}}+p_{J/\psi})^{\rho}(P_{B_{c}}-p_{J/\psi})^{\sigma}, \end{equation} \begin{equation} \langle p_{J/\psi}|A_{\mu}|P_{B_{c}}\rangle={\frac{1}{2}}[f\epsilon^{*}_{\mu}+a_{+} (\epsilon^{*}\cdot P_{B_{c}})(P_{B_{c}}+p_{J/\psi})_{\mu}+a_{-} (\epsilon^{*}\cdot P_{B_{c}})(P_{B_{c}}-p_{J/\psi})_{\mu}], \end{equation} where $g, f, a_+, a_{-}$ are the possible form factors. For convenience, here we actually calculate the matrix element $\langle p_{B_{c}}|j^{\mu em}(0)|P_{B_{c}^{*}}\rangle$ instead of $\langle p_{B_{c}^{*}}|j^{\mu em}(0)|P_{B_{c}}\rangle$: \begin{equation} \langle p_{B_{c}}|j^{\mu em}(0)|P_{B^{*}_{c}}\rangle=i(c_{b}g_{b}+c_{c}g_{c}) \varepsilon_{\mu\nu\rho\sigma}\epsilon^{*\nu}({B_{c}^{*}})(P_{B_{c}^{*}}+ p_{B_{c}})^{\rho}(P_{B_{c}^{*}}-p_{B_{c}})^{\sigma} \end{equation} Now we just present the final results of the form factors $g, f, a_{+}, a_{-}, g_{b}, g_{c}$ here, as the detail calculations on them can be found in Ref.[1]. Let us introduce further definitions, which is convenient for presenting the results. If $p_{1}, p_{2}$ are the momenta of the constituent particles 1 and 2 respectively, the total and the relative momenta $p$ and $q$ are defined as: $$p_{1}={\alpha}_{1}p+q, \;\; {\alpha}_{1}=\frac{m_{1}}{m_{1}+m_{2}};$$ $$p_{2}={\alpha}_{2}p+q, \;\; {\alpha}_{2}=\frac{m_{2}}{m_{1}+m_{2}}.$$ The momenta $p$, $p'$ are replaced to denote those for the initial and final mesons $P_{B_{c}}$, $p_{B_{c}^{*}}$ (or $p_{J/\psi}$), and $M, M'$ are denoted the masses of the initial and final mesons. Furthermore $q_{p}$, $q_{pT}$ are denoted the two Lorentz covariant variables as follows: $$q_{p}={\frac{p\cdot q}{M_{p}}}, \;\;\; q_{pT}=\sqrt{q^{2}_{p}-q^{2}}$$ and: $$\omega_{ip}=\sqrt{m_{i}^{2}+q^{2}_{pT}}$$ $${{\omega}'}_{ip'}=\sqrt {{m'}_{i}^{2}+{q'}^{2}_{p'T}}.$$ Now we can give the formulas of form factor using the above covariant variables: $$ g=\xi{\frac{\omega_{1}^{'}+\omega_{2}^{'}}{MM^{'}m_{1}^{'}}} $$ $$ f=\xi \left[{\frac{(p\cdot p_{1}^{'})}{Mm_{1}^{'}}}+1\right] $$ \begin{equation} a_{\pm}=\xi\left\{{\frac{2m_{2}}{M^{2}m'_{1}}}+ \delta\mp\left[{\frac{{\omega}'_{1}+{\omega}'_{2}} {MM'm'_{1}}}+{\frac{(p\cdot{p'})}{M'^{2}}}\delta\right]\right\} \end{equation} where $$\delta=-{\frac{C[1+(p\cdot{p'_{1}})/Mm'_{1}]}{p^{2}-(p\cdot{p'})^{2}/M'^{2}}}$$ $$C={\frac{1}{L_{+}\left[{\frac{({p'}\cdot p_{1})({p'} \cdot {p'}_{1})}{{M'}^{2}}}-{\frac{1}{L_{-}^{2}}}\right]^{1/2}-1}}$$ where $$L_{\pm}=\left[{\frac{1}{2}}(p_{1}\cdot {p'}_{1}\pm m_{1}{m'}_{1} ) \right]^{-1/2}.$$ The common factor: \begin{equation} \xi=\left[{\frac{2{\omega}'_{2}m^{2}_{1}m'^{2}_{1}}{[(p_{1} \cdot{p'_{1}})+m_{1}m'_{1}] {\omega}_{1}{\omega}'_{1}{\omega}_{2}}}\right]^{1/2} \times\int\frac{d^{3}\stackrel\rightarrow {q}}{(2\pi)^{3}}{\phi}'^{*}_{p'}(q'_{p'_{T}}) \cdot{\phi_{p}(|\stackrel {\rightarrow}{q}|)} \end{equation} if it is written in c.m.s. of the initial meson ($\stackrel{\rightarrow}{p}=0$). Here ${\phi}'^{*}_{p'}(q'_{p'_{T}})$ and ${\phi_{p}(|\stackrel {\rightarrow}{q}|)}$\cite{s14} correspond to the radius wave functions of the mesons in the initial and final states respectively, but both are presented in c.m.s. of the initial meson. The equations about $g_{a}$, $g_{b}$ are similar as $g$, we will not repeat them here. \section{Numerical Results and Discussion} First of all, let us show the `whole' leptonic decay widths i.e. the sum of the corresponding radiative decay widths and the corresponding pure leptonic decay widths with radiative corrections, and put them in Table (1). As firstly we would like to see the facts of `short distance' contributions, so here only the `short distance' contributions are taken into account. Why we put the radiative decay and the pure leptonic decay with radiative corrections together here is to make the width not to depend on the experimental resolution for a soft photon at all. To compare with the earlier computations\cite{lcd,aliev,aliev1} in the numerical evaluation, the values for the parameters $\alpha=1/132$ and $|V_{bc}|=0.04$\cite{data} are taken, and two possible values for the rests are selected as bellow:\\ (1) $m_{B_{c}}=6.258$ GeV, $m_{b}=4.758$ GeV, $m_{c}=1.500$ GeV, $f_{B_{c}}=0.360$ GeV\cite{lcd};\\ (2) $m_{B_{c}}=6.258$ GeV, $m_{b}=4.700$ GeV, $m_{c}=1.400$ GeV, $f_{B_{c}}=0.350$ GeV\cite{aliev,aliev1}. \begin{center} Table (1) The `Whole' Leptonic Decay Widths (in unit GeV)\\ \noindent {\small (short distance contributions only)}\\ \vspace{2mm} \begin{tabular}{|c|c|c|} \hline &(1) &(2) \\ \hline $\Gamma_{e}(10^{-17})$ & 6.444 & 6.902 \\ \hline $\Gamma_{\mu}(10^{-16})$ & 1.383 & 1.389 \\ \hline $\Gamma_{\tau}(10^{-14})$ & 1.871 & 1.782 \\\hline \end{tabular} \end{center} For comparison, the width of each pure leptonic decay at tree level with the same parameters as those in Table (1) is put in Table (2). \begin{center} Table (2) The Pure Leptonic Decay Widths (in unit GeV) of Tree Level\\ \vspace{2mm} \begin{tabular}{|c|c|c|} \hline &(1)&(2) \\ \hline $\Gamma_{e}(10^{-21}$)& 1.827& 1.727 \\ \hline $\Gamma_{\mu}(10^{-16}$) & 0.7841& 0.7412\\ \hline $\Gamma_{\tau}(10^{-14}$)& 1.862& 1.773 \\ \hline \end{tabular} \end{center} If the lifetime of $B_{c}$ meson is (a). $\tau(B_{c})=0.46\times10^{-12}s$ as indicated by the first observation\cite{cdf}; (b). $\tau(B_{c})=0.52\times 10^{-12}s$ as adopted in Ref.\cite{lcd}, the corresponding branching ratios are showed in Tables (3), (4). \begin{center} Table (3) Branching Ratios of the `Whole' Leptonic Decays \\ \noindent {\small (short distance contributions)}\\ \vspace{2mm} \begin{tabular}{|c|c|c|c|c|} \hline &(1-a)&(2-a)&(1-b)&(2-b) \\ \hline $B_{e}(10^{-5})$ & 5.09 & 5.45&4.5&4.82\\ \hline $B_{\mu}(10^{-5})$ & 10.93& 10.98 & 9.69 & 9.76 \\ \hline $B_{\tau}(10^{-2})$& 1.477 & 1.407 & 1.306 &1.246 \\ \hline \end{tabular} \end{center} \begin{center} Table (4) Tree Level Branching Ratios of The Pure Leptonic Decays\\ \vspace{2mm} \begin{tabular}{|c|c|c|c|c|} \hline &(1-a)&(2-a)&(1-b)&(2-b) \\ \hline $B_{e}(10^{-9})$&1.44&1.36&1.28&1.21\\ \hline $B_{\mu}(10^{-4})$ &0.62&0.586&0.55&0.52\\ \hline $B_{\tau}(10^{-2})$&1.47&1.40&1.30&1.24\\ \hline \end{tabular} \end{center} To see the contributions of the radiative decays precisely we present the radiative decay widths with a cut of the photon energy i.e. the widths of the radiative decays $B_c\rightarrow l\nu\gamma$ with the photon energy $E_\gamma \geq k_{min}$ as the follows: $k_{min}=0.1$ GeV, $k_{min}=0.2$ GeV, $k_{min}=0.5 $ GeV and $k_{min}=1.0 $ GeV respectively in Table (5). \begin{center} Table (5): The Radiative Decay Widths (in unit $10^{-17}$GeV)\\ \noindent {\small (with cuts of the photon momentum and the angle between photon and lepton)}\\ \vspace{2mm} \begin{tabular}[t]{|c|c|c|c|} \hline & & (1) & (2) \\ \hline $k_{min(GeV)}$ & & $\quad \;5^0\quad \quad \;\;\;15^0\quad \quad \;\;30^0\,\quad $ & ${}{}5^0\;\;\qquad \;15^0\;\;\qquad 30^0$ \\ \hline $0.1$ & $\Gamma _e$ & 6.384\qquad 6.370\qquad 6.297 & \ 6.832\qquad 6.819\qquad 6.752\ \\ \hline $0.2$ & $\Gamma _e$ & \thinspace 6.317\qquad 6.303\qquad 6.242 & 6.762\qquad 6.750\qquad 6.693 \\ \hline $0.5$ & $\Gamma _e$ & 5.931\qquad 5.918\qquad 5.883 & 6.351\qquad 6.340\qquad 6.307 \\ \hline $1.0$ & $\Gamma _e$ & 4.807\qquad 4.800\qquad 4.790 & 5.151\qquad 5.143\qquad 5.136 \\ \hline $0.1$ & $\Gamma _\mu $ & 6.613\qquad 6.518\qquad 6.385 & 7.049\qquad 6.958\qquad 6.834 \\ \hline $0.2$ & $\Gamma _\mu $ & 6.484\qquad 6.412\qquad 6.306 & 6.920\qquad 6.850\qquad 6.753 \\ \hline $0.5$ & $\Gamma _\mu $ & 6.018\qquad 5.977\qquad 5.917 & 6.433\qquad 6.394\qquad 6.340 \\ \hline $1.0$ & $\Gamma _\mu $ & 4.843\qquad 4.824\qquad 4.802 & 5.184\qquad 5.165\qquad 5.146 \\ \hline $0.1$ & $\Gamma _\tau $ & 13.75\qquad 13.66\qquad 12.88 & 13.60\qquad 13.52\qquad 12.78 \\ \hline $0.2$ & $\Gamma _\tau $ & 10.87\qquad 10.82\qquad 10.34 & 10.86\qquad 10.81\qquad 10.36 \\ \hline $0.5$ & $\Gamma _\tau $ & 7.139\qquad 7.121\qquad 6.970 & 7.282\qquad 7.265\qquad 7.122 \\ \hline $1.0$ & $\Gamma _\tau $ & 4.169\qquad 4.165\qquad 4.146 & 4.340\qquad 4.335\qquad 4.318 \\ \hline \end{tabular} \end{center} \medskip For the convenience to compare with experiments, we present the photon spectrum of the radiative decays in Fig.6 and give the decay widths of the so-called physical pure leptonic decays: $\Gamma _{phys}=\Gamma _{whole}-\Gamma _{cut}$ where $\Gamma _{cut}$ are showed in Table (5); $\Gamma _{whole}$ are showed in Table (1). We put the values of $\Gamma _{phys}$ in Table (6). \begin{center} Table (6): Decay widths of $\Gamma _{phys}$ ($10^{-17}$GeV) \\ \vspace{2mm} \begin{tabular}{|c|c|c|c|} \hline & & (1) & (2) \\ \hline $k_{min}$(GeV) & & $5^0\;\qquad \;\;\;15^0\qquad \;\;30^0$ & $5^0\;\qquad \;\;15^0\;\qquad \;30^0$ \\ \hline $0.1$ & $\Gamma _e$ & 0.060\qquad 0.074\qquad 0.147 & 0.070\qquad 0.083\qquad 0.150 \\ \hline $0.2$ & $\Gamma _e$ & 0.127\qquad 0.141\qquad 0.202 & 0.140\qquad 0.152\qquad 0.209 \\ \hline $0.5$ & $\Gamma _e$ & 0.513\qquad 0.526\qquad 0.561 & 0.551\qquad 0.562\qquad 0.595 \\ \hline $1.0$ & $\Gamma _e$ & 1.637\qquad 1.644\qquad 1.654 & 1.751\qquad 1.759\qquad 1.766 \\ \hline $0.1$ & $\Gamma _\mu $ & 7.217\qquad 7.312\qquad 7.445 & 6.837\qquad 6.928\qquad 7.052 \\ \hline $0.2$ & $\Gamma _\mu $ & 7.346\qquad 7.418\qquad 7.524 & 6.966\qquad 7.036\qquad 7.133 \\ \hline $0.5$ & $\Gamma _\mu $ & 7.812\qquad 7.853\qquad 7.913 & 7.453\qquad 7.492\qquad 7.546 \\ \hline $1.0$ & $\Gamma _\mu $ & 8.987\qquad 9.006\qquad 9.028 & 8.702\qquad 8.721\qquad 8.740 \\ \hline $0.1$ & $\Gamma _\tau (10^3)$ & 1.857\qquad 1.857\qquad 1.858 & 1.768\qquad 1.768\qquad 1.769 \\ \hline $0.2$ & $\Gamma _\tau (10^3)$ & 1.860\qquad 1.860\qquad 1.860 & 1.771\qquad 1.771\qquad 1.772 \\ \hline $0.5$ & $\Gamma _\tau (10^3)$ & 1.864\qquad 1.864\qquad 1.864 & 1.775\qquad 1.775\qquad 1.775 \\ \hline $1.0$ & $\Gamma _\tau (10^3)$ & 1.866\qquad 1.866\qquad 1.867 & 1.777\qquad 1.777\qquad 1.778 \\ \hline \end{tabular} \end{center} Let us select the parameters as in Ref.\cite{lcd}: $m_{B_c}=6.258$ GeV, $m_b=4.758$ GeV, $m_c=1.500$ GeV, $f_{B_c}=0.360$ GeV, so as to compare with the short distance contributions and to show the contributions from the typical long distance component concerned in the paper. For each of the family, the contribution to the width: $$\delta\Gamma _e=1.828\times10^{-18} GeV,$$ $$\delta\Gamma _\mu =1.827\times10^{-18} GeV,$$ $$\delta\Gamma _\tau =1.190\times10^{-18} GeV.$$ From Table (5), we my see that the present results confirm those in Ref.\cite{lcd}, and the slight differences are maily due to the radiative corrections and the fact that instead of ignoring in Ref.\cite{lcd}, we keep the lepton mass precisely in the numerical evaluation. In comparison with the short contributions, for the radiative leptonic decays of the $B_{c}$ meson the long distance contributions $\delta\Gamma _l, (l=e,\mu,\tau)$ are quite small. It may be the reason that the $b$ and $c$ quark in $B_{c}$ meson are heavy. Their heavy mass causes the long distance contributions small. Therefore, it seems that we can conclude that the disagreement between Ref.\cite{lcd} and Refs.\cite{aliev,aliev1} is not due to the long distance effects. In addition, we should note that the widthes are quite sensitive to the decay constant $f_{B_{c}}$, and the values of the quark masses $m_b$ and $m_c$. Note that when we almost completed this paper, one similar paper\cite{s17} appeared and its results supported those in Ref.\cite{lcd} that is in agreement with this paper. \vspace{2cm} {\Large\bf Acknowledgement} This work was supported in part by the National Natural Science Foundation of China and the Grant No. LWLZ-1298 of the Chinese Academy of Sciences. One of the author (C.-D. L\"u) would also like to thank JSPS for supporting him in research.
\section*{Figure Captions} \noindent FIG. 1. Strand/Loop decomposition for a 36-mer \\ \\ \noindent FIG. 2. Symmetric potential spectrum of a 36-mer \\ \\ \noindent FIG. 3. Principal contribution to 36-mer spectrum with u = -0.3 \\ \\ \noindent FIG. 4. Change in expected designability due to 1st-order perturbation \pagebreak \pagestyle{empty} \begin{center} \par \leavevmode\def\epsfsize#1#2{2.0 #1}\epsffile{Fig1.eps} \end{center} \vspace{10pt} \begin{flushright} Kassel and Shakhnovich Fig.1 \end{flushright} \pagebreak \begin{center} \par \leavevmode\def\epsfsize#1#2{2.0 #1}\epsffile{Fig2.eps} \end{center} \vspace{10pt} \begin{flushright} Kassel and Shakhnovich Fig.2 \end{flushright} \pagebreak \begin{center} \par \leavevmode\def\epsfsize#1#2{2.0 #1}\epsffile{Fig3.eps} \end{center} \vspace{10pt} \begin{flushright} Kassel and Shakhnovich Fig.3 \end{flushright} \pagebreak \begin{center} \par \leavevmode\def\epsfsize#1#2{2.0 #1}\epsffile{Fig4.eps} \end{center} \vspace{10pt} \begin{flushright} Kassel and Shakhnovich Fig.4 \end{flushright} \end{document} \end{document}
\section{Introduction} People have long considered that the almost isotropy of the cosmic microwave background radiation (CMB) indicates that the universe on very large scales is isotropic and homogeneous, or nearly so. The first rigorous result supporting this conjecture was the theorem proved by Ehlers, Geren and Sachs (1968) (EGS): ``If a family of freely falling observers measure self-gravitating background radiation to be everywhere exactly isotropic, then the universe is exactly Friedmann-Lema\^{i}tre-Robertson-Walker (FLRW).'' Later, Grishchuk and Zel'dovich (1978), analyzing the FLRW perturbations, argued that, if CBR anisotropies are very small, all anisotropies and inhomogeneities on scales larger than the horizon should also be very small. Much more recently Stoeger et al. (1995) (SME) proved a significant generalization of this theorem: ``If the Einstein-Liouville equations are satisfied in an expanding universe, where there is present pressure-free matter with 4-velocity vector field $u^a$ ($u_au^a = -1$) such that (freely propagating) background radiation is everywhere almost-isotropic relative to $u^a$, then spacetime is almost FLRW.'' These results obviously provide a fundamental link between observational and theoretical cosmology -- one which promises to reveal key aspects of the very large scale structure of the universe without relying on the often uncertain observational measurements of local and intermediate scale structures. It should be noted that both the EGS and the SME results depend on assuming that the cosmological principle holds -- here expressed in terms of the almost-isotropy of the background radiation relative to {\it every} fundamental observer in the space-time. Goodman (1995) and Maartens et al. (1995a), however, have recently pointed out the important fact that the cosmological principle itself is partially testable via the Sunyaev-Zel'dovich effect. \\ Employing this connection between CBR anisotropies and the full range of cosmological isotropies and inhomogeneities, Maartens et al. (1995a,b) (MES) have developed a detailed scheme demonstrating how limits on the temperature anisotropies of the CMB imply rigorous limits on the anisotropy and inhomogeneity of the universe. In this paper we show how CMB anisotropy data is inserted into the theoretically derived limits of this scheme to constrain strongly the vorticity, shear, spatial gradients, Weyl-tensor components, and other measures of deviation from FLRW for the observable universe. We then introduce the limits COBE places on the dipole, quadrupole and octupole of the CBR to determine those limits quantitatively. \\ In the next section we shall briefly give the key MES equations which constitute the CMB limits on the anisotropy and inhomogeneity of the universe. Then, in the third section, we shall discuss the relationship of the usual CMB multipole results with the temperature anisotropy multipoles in the MES equation. Finally, in section 4, we shall present the COBE quadrupole and octupole observational results, and use these to arrive at limits on all the cosmological anisotropy and inhomogeneity measures mentioned above. \\ Maartens, Ellis and Stoeger (1996) recently provided such limits by introducing order-of-magnitude COBE results of $\epsilon_2 \approx \epsilon_3 \approx 10^{-5}$, but they did not discuss the relationship between temperature anisotropy multipoles used in their scheme and those in terms of which the COBE results are given. Here we fill that gap, and then use the up-to-date COBE results for the rms quadrupole and octupole to obtain improved limits. \section{The MES Equations} MES have written the covariant temperature isotropy multipoles as $\tau_{a_1 ...a_L}$, where $L$ give the multipole number. Thus, for instance, $\tau_a$, $\tau_{ab}$, $\tau_{abc}$ are, respectively, the components of the dipole, quadrupole, and octopole of the CMB temperature anisotropy. This form of the harmonic decomposition, which is given in detail in Maartens et al. (1995a), is equivalent to that formulated in terms of spherical harmonics (Ellis, Matravers and Treciokas 1983; Ellis Treciokas and Matravers 1983). \\ MES have assumed that there are observed bounds on these temperature anisotropy multipoles, so that there exist O[1] constants $\epsilon _L$ such that $$\mid \tau_{a_1...a_L} \mid < \epsilon_L. \eqno(1) $$ The absolute-value brackets have been {\it defined} to be the squareroot of the sum of the squares of the components of a given vector or tensor (Maartens et al. 1995a, p. 1526). Thus, $\epsilon_1$ gives the limits on the dipole components, $\epsilon_2$ on the quadrupole components, $\epsilon_3$ on the octupole moments, and so on. \\ Then, using the strong observational assumptions on the spatial gradients and time-derivatives of the temperature harmonics they adopted in Maartens et al. (1995b), they have the observational limit equations on various kinematic, dynamic and geometric indicators of anisotropy and inhomogeneity (Maartens et al. 1995a,b): $$\frac{\mid \hat{\bigtriangledown}_a \mu \mid}{\mu} = 4\frac{\mid \hat{ \bigtriangledown}_a T \mid}{T} < H (12 \epsilon_1 + \frac{24}{5} \epsilon_2 ), \eqno (2) $$ $$\frac{\mid \sigma_{ab} \mid}{\Theta} < \frac{5}{3}\epsilon_1 + 3 \epsilon_2 + \frac{3}{7} \epsilon_3, \eqno(3) $$ $$\frac{\mid \omega_{ab} \mid}{\Theta} < \frac{10}{3}\epsilon_1 + \frac{2}{15} \epsilon_2, \eqno(4) $$ $$\frac{\mid \hat{\bigtriangledown}_a \rho \mid}{\Theta} < \frac{9}{2}H \epsilon_2 + (\frac{H}{\Omega_M})[60 \epsilon_1 + 134 \epsilon_2 + 6\epsilon_3] + (\frac{\Omega_R}{\Omega_M})H[16 \epsilon_1 + \frac{61}{3} \epsilon_2], \eqno(5) $$ $$\frac{\mid \hat{\bigtriangledown}_a \Theta \mid}{\Theta} < H(\frac{205}{3} \epsilon_1 + 8 \epsilon_2) + 4(2 \Omega_R + \Omega_M) H \epsilon_1, \eqno(6)$$ $$\frac{\mid E_{ab} \mid}{\Theta} < H(\frac{55}{3} \epsilon_1 + \frac{103}{3} \epsilon_2 + \frac{23}{7} \epsilon_3) + \frac{4}{45}(11 \Omega_R + 15 \Omega_M)H\epsilon_2, \eqno(7) $$ $$\frac{\mid H_{ab} \mid}{\Theta} < H(\frac{16}{3} \epsilon_1 + \frac{52}{15} \epsilon_2 + \frac{1}{21} \epsilon_3). \eqno(8) $$ \noindent Here $\mu$ and $\rho$ are the radiation and matter densities, respectively, and $\Omega_R$ and $\Omega_M$ are the ratios of the radiation energy density and the matter densities, respectively, to the critical density of the universe. $H$, of course, the Hubble parameter, and $T$ is the CMB temperature. $\sigma_{ab}$ is the shear of the congruence of timelike geodesics in the universe, $\omega_{ab}$ its vorticity, $\Theta = u^a_{;a}$ its expansion scalar, and $E_{ab}$ and $H_{ab}$ the electric and magnetic components, respectively, of the Weyl tensor. These quantities, along with the spatial gradients of $\rho$, $\mu$ and $\Theta$ describe the anisotropy and inhomogeneity of the space-time. The $\hat {\bigtriangledown}_a$ operator expresses the covariant spatial gradient in the spacetime. It is defined by $$\hat{\bigtriangledown}_c Q_{a...b} \equiv h_c^d h_a^e h_b^f \bigtriangledown _d Q_{e...f},$$ where $h_{ab} = g_{ab} + u_au_b$ is the projection tensor in the rest spaces of the geodesically moving observers and $\bigtriangledown_a$ is the covariant derivative define by the metric $g_{ab}$. Finally, it should be noted that equations (2) - (8) hold independently of the statistics of the underlying matter-density fluctuations -- they do not assume that the fluctuations are Gaussian. \\ If we can determine $\epsilon_1$, $\epsilon_2$ and $\epsilon_3$ from CMB measurements, and have observational estimates of $H$, $\Omega_R$ and $\Omega_M$ from other observations, then we can determine limits on all these anisotropy and inhomogeneity indicators. We shall proceed to do this in Section 4. But first, in the next section, we need to discuss the relationship between the $\mid\tau_{a_1...a_L} \mid$ given in the equations and the multipole moment results determined by the COBE Differential Microwave Radiometers (DRM) and other CMB anisotropy detectors. \section{CMB Multipole Anisotropy Data} We have mentioned above that the harmonic decomposition represented by the $\mid\tau_{a_1...a_L}\mid$ is equivalent to that in terms of spherical harmonics. But the question then is whether or not the multipole results recovered from, say, COBE DRM data can be simply substituted into our limit equations. Are the multipole results presented in the COBE papers measurements of the $\mid\tau_{a_1...a_L}\mid$ in our limit equations? The answer to this question is `Yes,' as long as we use the real rms dipole, quadrupole, octupole moments they obtain, and not those associated with obtaining the power spectrum, which are often the focus of their reported results, determine the numerical factors relating these moments, defined in terms of Legendre polynomials, to the $\mid \tau_{a_1 . . . a_L} \mid$ (see section 5 below), and as long as we realize that they are usually given as the squareroot of the sum of the squares of the $(2L + 1)$ components of the L-pole, the rms L-pole -- not as values of each separate component of the multipole in question. Furthermore, in the COBE data the multipoles are quoted for $\delta T$, instead of $\frac{\delta T}{T}$, for which our $\mid\tau_{a_1...a_L}\mid$ are the multipoles. The COBE rms values published all have units of $\mu K$, therefore. To translate these values into what we need we must thus divide them by $T$, the average CMB background temperature over the sky. \\ As a relevant example, Bennett el al. (1994) give results for the square of the rms quadrupole amplitude $Q^2_{rms}$, $$Q^2_{rms} = \frac{4}{15}(\frac{3}{4}Q_1^2 + Q_2^2 + Q_3^2 + Q_4^2 + Q_5^2), \eqno(9)$$ \noindent where the five quadrupole components $Q_i$ are given in terms of Galactic coordinates $l$ and $b$ (Galactic longitude and latitude respectively) by the expansion \setcounter{equation}{9} \begin{eqnarray} Q(l, b) & = & Q_1(3 \sin^2 b - 1)/2 + Q_2 \sin 2b \cos l + Q_3 \sin 2b \sin l + \nonumber \\ & & Q_4 \cos^2 b \cos 2l + Q_5 \cos^2 b \sin 2l. \end{eqnarray} \noindent The `strange' coefficients in equation (9) are due to the fact that the basis vectors are orthogonal but not orthonormal (see Partridge 1995); $Q^2_{rms}$ does {\it not} contain a factor of $(2L + 1)^{-1} = 1/5$. >From four years of data the COBE workers give us a best fit value of (Smoot, private communication; Kogut et al., 1996) $$Q_{rms} = 10.7 \pm 7 \mu K , \eqno(11)$$ with a 95\% confidence limit. We can modify this result (see below) for use in equations (2) to (8). This quantity is to be carefully distinguished in the COBE results from $Q_{rms-PS}$ which is often referred to and which is {\it not} the true best fit value of the quadrupole, but rather the value of the quadrupole derived from a power-spectrum fit (when a power law is assumed) of the other higher-order multipole moments [see Bennett et al. 1994; Smoot et al. 1992] \\ The dipole moment can be neglected, according to the fairly well justified assumption that it is all due to our peculiar motion with respect the rest frame of the microwave background. In fact this is what is done in the COBE anisotropy analysis (Bennett et al. 1994). However, we should be aware that there could in principle be a small non- Doppler contribution to the CMB dipole (see Maartens et al. 1995a and Maartens et al. 1996 concerning this). Thus, in our calculations below we shall set $$\epsilon_1 = 0. \eqno(12)$$ \\ The COBE workers have not yet published the rms octupole from their data, but G. Smoot (private communication) has kindly informed us that the rms octupole results from COBE are: $$ O_{rms} = 16 \pm 8 \mu K. \eqno(13)$$ \\ This is consistent with the results given by Wright et al. (1994) in their Figure 1. \\ There are several important observational and data-reduction issues, and one theoretical issue, which we should briefly mention here, to provide the background against which we can understand and appreciate these COBE multipole results. \\ The first is that the COBE DMR experiment does not directly measure the dipole, quadrupole, octupole moments of the temperature anisotropy, but rather the two-point correlation function of the temperature anisotropy (Smoot et al. 1992, Padmanabham 1993, Partridge 1995) $$C(\alpha) = <S({\bf n})S({\bf m})> = \sum_{\ell = 1}^\infty \bigtriangleup T_{\ell}^2 W(\ell)^2 P_{\ell}[\cos \alpha], \eqno(14)$$ where $\alpha$ is the angle between the two points, ${\bf n}$ and ${\bf m}$ are unit vectors denoting the two different directions, so that $\cos \alpha = {\bf n} \cdot {\bf m}$, and the averaging brackets signify the average over all pairs of points on the sky with separation angle $\alpha$. Furthermore, $S({\bf n}) \equiv \delta T ({\bf n})$, the temperature anisotropy in a given direction ${\bf n}$ on the plane of the sky. The $\bigtriangleup T_{\ell}^2$ (with $\ell \equiv L$) are the squares of the rotationally invariant rms multipole moments (thus, $\bigtriangleup T_2^2 = Q^2_{rms}$ -- see Bennett et al. 1994, for instance) $$\bigtriangleup T_{\ell}^2 = \frac{1}{4 \pi} \sum_m \mid a_{\ell m} \mid^2, \eqno(15)$$ where the $a_{\ell m}$ are the coefficients of the expansion of the temperature anisotropy in spherical harmonics, i. e. $$S({\bf n}) \equiv S(\theta, \phi) = \sum_{\ell,m} a_{\ell m} Y_{\ell m}(\theta, \phi), \eqno(16) $$ $\theta$ and $\phi$ being, of course, the angular coordinates of the direction at which the temperature anisotropy is being measured. In equation (14), finally, $P_{\ell}[\cos \alpha]$ is just the Legendre polynomial of degree $\ell$ given as a function of $\cos \alpha$, and $W(\ell)$ is the window function, which describes the smoothing properties of the instrument's beam. For detectors measuring large-angle CMB anisotropies it essentially weights the multipole moments in such a way that the higher multipoles are smoothed over -- that is, the instument is insensitive to anisotropies on angular scales less than a certain $\ell$-pole, and $W(\ell)$ describes that insensitivity and resulting transfer of power in the measurement from higher multipoles to lower ones. Thus, when the actual quadrupole or octupole is determined from the data, $\bigtriangleup T_\ell$ must be deconvolved from $W(\ell)$. When the rms dipole, quadrupole, octupole results are given by the COBE researchers, this deconvolution has already been performed. This is one of the procedures which must be effected to give us the real rms multipole moments. \\ The second major observational and data-reduction problem is that in the COBE measurements there is a great deal of contamination by experimental systematic errors, including Galactic emission (see Bennett et al. 1994 and Smoot et al. 1992, and references therein) And the lower multipole moments are the most susceptible to these distortions. Furthermore, when the data is processed, the entire region containing the Galaxy is removed from the data set. This `Galactic cut' destroys the orthogonality of the spherical harmonics and leads to further aliasing of higher order multipole power onto the lower multipoles (dipole, quadrupole, octupole, etc.). Corrections for this are estimated on the basis of Monte Carlo simulations (Bennett et al. 1994) and included in the published values for the rms multipoles. There are a number of other complex issues which it has been necessary to resolve in arriving at these values ( see Bennett et al. 1994, Smoot et al. 1992, and Wright et al. 1994 for further discussion). \\ Finally, there is the theoretical-observational issue of `cosmic variance.' Actually, cosmic variance does not affect what we are concerned with here, as we shall see. But it is important to realize why it does not. It may explain why the multipoles we measure have the values they have relative to theoretical models of the perturbation spectrum, but it is does not lead to observational errors, which would have to be corrected for. It {\it does} affect the comparison of the measured multipole power spectrum with the theoretical power spectrum predicted from, say, inflationary models (Abbott \& Wise 1984). If the primordial perturbation spectrum originated due to fluctuations in the inflaton field, as we think it did, then the values of the temperature anisotropy multipoles they induce will be random variables with a certain distribution, probably Gaussian with zero mean (Liddle \& Lyth 1993), and thus with a certain variance. Our observable universe is only one realization of that ensemble of universes represented by the probability distribution. Therefore, the value for each multipole we obtain from our observations will give us just one point of the distribution, which, in general, will not reflect the ensemble averaged value. It will deviate from it by a certain amount, which can be theoretically estimated by the variance (Liddle \& Lyth 1993). This variance goes as $2/(2 \ell + 1)$ and so will be more significant for the lower multipoles (Abbott \& Wise 1984; Smoot et al. 1992). Our concern here, however, is not to compare the observed power spectrum of CMB anisotropies with the theoretical spectrum of density perturbations generated by an inflationary scenario. It is merely to use the best values of the CMB multipole moments we have available -- however they are generated and whatever their spectrum -- to set definite limits on the large scale anisotropy and inhomogeneity of the observable universe itself. Thus, cosmic variance falls outside those issues which we need to consider in arriving at those limits. \\ \section{Calculating the Anisotropy and Inhomogeneity Limits} We are now ready to use the values for the rms dipole, quadrupole and octupole the COBE team have so far obtained to determine the anisotropy and inhomogeneity of the universe on very large scales. As indicated above, we set the dipole equal to zero -- equation (12). Thus, we shall set $\epsilon_1 = 0$ in our equations (2) to (8). In order to transform equations (11) and (13) into values of $\mid \tau_{ab} \mid$ and $\mid \tau_{abc} \mid$, respectively, we need to divide the results in equations (11) and (13) by $T = 2.73 K$, since our multipoles are for $\delta T/T$. We do not need to divide them by $(2 \ell + 1)^{1/2}$, since our multipole quantities are given as the absolute values, which we {\it defined} as the squareroot of the sum of the squares of the components (see above). Finally, we also need to relate the $\bigtriangleup T_{\ell}^2$, the coefficients in the usual Legendre polynomial expansion, equation (14) to the $\mid \tau_{a_1. . .a_{\ell}}\mid$. The numerical relationship is (Gebbie \& Ellis 1998) $$\left < {\tau_{A_\ell} \tau^{A_\ell}} \right > = (2 \ell + 1){{(2 \ell )!\over 2^{\ell} ({\ell}!)^2}} \Delta T_{\ell}^2. \eqno (17)$$ This gives $$ \mid \tau_{ab} \mid^2 = 7.5 Q_{rms}^2, \eqno (18) $$ and $$ \mid \tau_{abc}\mid^2 = 17.5 O_{rms}^2, \eqno (19)$$ We obtain, therefore, $$ \langle \epsilon_2 \rangle = 1.1 (\pm 0.8) \times 10^{-5}, \eqno(20)$$ and $$ \langle \epsilon_3 \rangle = 2.5 (\pm 1.3) \times 10^{-5}. \eqno(21)$$ \\ In examining equations (2) to (8), we notice that they are not expressed in terms of average values of the components. Let us now consider them to be equations for the rms averages of the indicators they represent, which we can do, since our $\epsilon$'s are all positive. Now writing the Hubble parameter as $H = 100 h \; km/sec/Mpc $, $0.4 < h < 1.0$, and neglecting terms containing the factor $\Omega_R$, since this is presently so small -- $\Omega_R = 4.11 h^{-2} \times 10^{-5}$ (Roos, 1994) -- we obtain: $$\langle \frac{\vert \hat{\bigtriangledown}_a \mu \vert}{\mu}\rangle < 1.8h \times 10^{-8} \; Mpc^{-1}, \eqno(22)$$ $$\langle \frac{\vert \sigma_{ab} \vert}{\Theta}\rangle < 4.4 \times 10^{-5}, \eqno(23)$$ $$\langle \frac{\vert \omega_{ab} \vert}{\Theta}\rangle < 1.5 \times 10^{-6}, \eqno(24) $$ $$\langle \frac{\vert \hat{\bigtriangledown}_a \rho \vert}{\Theta}\rangle < (0.02 + 0.54 \Omega_M^{-1} ) h \times 10^{-6} \; Mpc^{-1}, \eqno(25)$$ $$\langle \frac{\vert \hat{\bigtriangledown}_a \Theta \vert}{\Theta}\rangle < 3.0 h \times 10^{-8} \; Mpc^{-1}, \eqno(26)$$ $$ \langle \frac{\vert E_{ab} \vert}{\Theta}\rangle < (1.5 + 48 \Omega_M) h \times 10^{-7} \; Mpc^{-1}, \eqno(27)$$ $$\langle \frac{\vert H_{ab} \vert}{\Theta}\rangle < 1.6 h \times 10^{-8} \; Mpc^{-1}, \eqno(28)$$ \noindent In equations with $H$ we have put a factor of $c^{-1}$ back into the equations to give the correct order of magnitude and the correct dimensions. This factor is hidden in the normalization of the 4-velocity $u^a$, which figures implicitly in the equations. \\ These are the limits we desire on the anisotropies and the inhomogeneities of the universe. In the past other groups have put limits on the shear and the vorticity using limits on CBR isotropies (Bajtlik et al. 1986; Martinez- Gonzalez \& Sanz 1995; Barrow et al. 1985). But in doing so they assumed exact spatial homogeneity, which yields a limit on shear which is too strong (Maartens et al. 1996). Our limiting scheme, as we have mentioned, does not assume exact homogeneity, or even that the inhomogeneities and anisotropies are small. The result that they are small is due to the fact that the CBR anisotropies are small. The primary assumption we have made is a weak form of the Copernican principle -- that all fundamental observers in the relevant space-time domain measure at most the same level of CBR anisotropy. This, as we have pointed out, is at least partially testable -- and fully falsifiable. A single observation demonstrating that CBR anisotropies relative to some other cluster of galaxies is large compared to those we observe would banish that assumption. Finally, our approach provides limits on the full range of possible isotropies and inhomogeneities, including spatial gradients of the radiation and matter densities, and of the expansion parameter, and the electric and magnetic components of the Weyl tensor. It is worth pointing out that the Weyl tensor components measure those parts of the gravitational curvature which are not determined locally by the mass-energy distribution(that is, via the Einstein field equations). Instead, they are determined by the Bianchi identities, and so, in a sense, are due to the mass-energy distribution at other, more distant points (Hawking \& Ellis 1973; Maartens et al. 1996). $H_{ab}$, the magnetic Weyl tensor, measures the amount of gravitational radiation in the space-time. It vanishes in the case in which exact spatial homogeneity and isotropy are assumed. $E_{ab}$, the electric Weyl tensor, which also vanishes in that case, measures the tidal, shear-inducing force of the global gravitational field. It thus manifests its presence by the shear it introduces in timelike and null congruences of geodesics -- worldlines of particles. In fact, the measurement of null shear in bundles of light rays would be the clearest signature of the presence of non-zero $E_{ab}$. (see Hawking \& Ellis 1973) \\ We see clearly from our results, equations (19) to (25), that in every case the anisotropy and inhomogeneity measures for the universe on the largest scales are very small, despite the significant inhomogeneities which have been detected on intermediate scales. This provides very strong justification for considering the observable universe to be almost-FLRW on large scales -- that is, the deviations from FLRW are small enough to be considered perturbations. \\ Finally, we might wonder what the relevance of these limits is to ``the averaging problem'' (see Zotov and Stoeger 1995 and references therein) in cosmology. Do these results demonstrate that the universe is almost-FLRW whatever the resolution of those issues happen to be? Unfortunately, that is not the case. The analysis we have implemented here {\it assumes} that the effective theory of gravity on cosmological scales is general relativity. The averaging problem calls that assumption into question. \\ We are very grateful to George Smoot for providing recently derived COBE values for the rms quadrupole and octupole, as well as for clarifying the principal source of error and the conventions in the COBE results. Special thanks also to Roy Maartens and George Ellis for encouragement and helpful comments. Marcelo Araujo thanks the Vatican Observatory Research Group for support and hospitality during the course of this work.\\ \newpage \noindent {\bf REFERENCES} \\ \noindent Abbott, L. F. \& Wise, M. B. 1984, ApJ, 282, L47. \\ \noindent Bajtlik, S., Juszkiewicz, R., Pr\'{o}szy\'{n}dski, M., \& Amsterdamski, P. 1986, ApJ, 300, 463. \\ \noindent Barrow, John D., Juszkiewicz, R., \& Sonoda, D. H. 1985, MNRAS, 213, 917. \\ \noindent Bennett, C. L. et al. 1994, ApJ, 436, 423. \\ \noindent Ehlers, J., Geren, P. \& Sachs, R. K. 1968, J. Math. Phys. 9, 1344. \\ \noindent Ellis, G. F. R., Matravers, D. R. \& Treciokas, R. 1983, Ann. Phys. (N. Y.), 150, 455. \\ \noindent Ellis, G. F. R., Treciokas, R., \& Matravers, D. R. 1983, Ann. Phys. (N. Y.), 150, 487. \\ \noindent Gebbie, T., Ellis, G. F. R.,1998 astro-ph/9804316. \\ \noindent Goodman, J. 1995, Phys. Rev. D, 52, 1821. \\ \noindent Grishchuk, L. P., \& Zel'dovich, Ya. B. 1978, Sov. Astron. AJ, 22, 125. \\ \noindent Hawking, S. W., \& Ellis, G. F. R. 1973, {\it The Large Scale Structure of Space-Time}, Cambridge University Press, pp. 85-88. \\ \noindent Kogut, A. et al. 1996, COBE preprint. \noindent Liddle, A. R., \& Lyth, D. H. 1993, Phys. Rep., 231, No. 1-2, 1 (pp. 57-59).\\ \noindent Maartens, R., Ellis, G. F. R., \& Stoeger, W. R. 1995a, Phys. Rev. D, 51, 1525. \\ \noindent Maartens, R., Ellis, G. F. R., \& Stoeger, W. R. 1995b, Phys. Rev. D, 51, 5942. \\ \noindent Maartens, R., Ellis, G. F. R., \& Stoeger, W. R. 1996, Submitted to A\&A. \\ \noindent Martinez-Gonzalez, E., \& Sanz, J. L. 1995, A\&A, 300, 346. \\ \noindent Padmanabhan, T. 1993, {\it Structure Formation in the Universe}, Cambridge University Press, pp 220ff. \\ \noindent Partridge, R. B. 1995, {\it 3 K: The Cosmic Microwave Background Radiation}, Cambridge University Press, p. 189. \\ \noindent Roos, M. 1994, {\it Introduction to Cosmology}, John Wiley and Sons, p. 99 \\ \noindent Smoot, G. F. et al. 1992, ApJ, 396, L1. \\ \noindent Stoeger, W. R., Maartens, R., \& Ellis, G. F. R. 1995, ApJ, 443, 1. \\ \noindent Wright, E. L. et al. 1994, ApJ, 436, 443. \\ \noindent Zotov, N. V. and Stoeger, W. R. 1995, ApJ, 453, 574. \\ \end{document} In the paper ``The Limits on Cosmological Anisotropies and Inhomogeneities from {\it COBE} Data,'' by William R. Stoeger, Marcelo E. Araujo, and Tim Gebbie [Ap. J. 476, 435 (1997)], the numerical relationship in equation (17) should be corrected to (Gebbie and Ellis 1998) to be publish in Apj Sept 1999.
\section{#1}} \newcommand{\vs}[1]{\rule[- #1 mm]{0mm}{#1 mm}} \newcommand{\hs}[1]{\hspace{#1mm}} \newcommand{\mb}[1]{\hs{5}\mbox{#1}\hs{5}} \newcommand{\begin{eqnarray}}{\begin{eqnarray}} \newcommand{\end{eqnarray}}{\end{eqnarray}} \newcommand{\wt}[1]{\widetilde{#1}} \newcommand{\underline}[1]{\underline{#1}} \newcommand{\ov}[1]{\overline{#1}} \newcommand{\sm}[2]{\frac{\mbox{\footnotesize #1}\vs{-2}} {\vs{-2}\mbox{\footnotesize #2}}} \newcommand{\partial}{\partial} \newcommand{\epsilon}\newcommand{\p}[1]{(\ref{#1})}{\epsilon}\newcommand{\p}[1]{(\ref{#1})} \newcommand{\mbox{\rule{0.2mm}{2.8mm}\hspace{-1.5mm} R}}{\mbox{\rule{0.2mm}{2.8mm}\hspace{-1.5mm} R}} \newcommand{Z\hspace{-2mm}Z}{Z\hspace{-2mm}Z} \newcommand{{\cal D}}{{\cal D}} \newcommand{{\cal G}}{{\cal G}} \newcommand{{\cal K}}{{\cal K}} \newcommand{{\cal W}}{{\cal W}} \newcommand{\vec{J}}{\vec{J}} \newcommand{\vec{\lambda}}{\vec{\lambda}} \newcommand{\vec{\sigma}}{\vec{\sigma}} \newcommand{\vec{\tau}}{\vec{\tau}} \newcommand{\stackrel{\otimes}{,}}{\stackrel{\otimes}{,}} \newcommand{\theta_{12}}{\theta_{12}} \newcommand{\overline{\theta}_{12}}{\overline{\theta}_{12}} \newcommand{{1\over z_{12}}}{{1\over z_{12}}} \newcommand{{(1 + i\sqrt{3})\over 2}}{{(1 + i\sqrt{3})\over 2}} \newcommand{{(1 - i\sqrt{3})\over 2}}{{(1 - i\sqrt{3})\over 2}} \newcommand{\NP}[1]{Nucl.\ Phys.\ {\bf #1}} \newcommand{\PLB}[1]{Phys.\ Lett.\ {B \bf #1}} \newcommand{\PLA}[1]{Phys.\ Lett.\ {A \bf #1}} \newcommand{\NC}[1]{Nuovo Cimento {\bf #1}} \newcommand{\CMP}[1]{Commun.\ Math.\ Phys.\ {\bf #1}} \newcommand{\PR}[1]{Phys.\ Rev.\ {\bf #1}} \newcommand{\PRL}[1]{Phys.\ Rev.\ Lett.\ {\bf #1}} \newcommand{\MPL}[1]{Mod.\ Phys.\ Lett.\ {\bf #1}} \newcommand{\BLMS}[1]{Bull.\ London Math.\ Soc.\ {\bf #1}} \newcommand{\IJMP}[1]{Int.\ J.\ Mod.\ Phys.\ {\bf #1}} \newcommand{\JMP}[1]{Jour.\ Math.\ Phys.\ {\bf #1}} \newcommand{\LMP}[1]{Lett.\ Math.\ Phys.\ {\bf #1}} \renewcommand{\thefootnote}{\fnsymbol{footnote}} \newpage \setcounter{page}{0} \pagestyle{empty} \vs{12} \begin{center} {\LARGE {\bf Real Structures in Clifford Algebras and}}\\ {\LARGE {\bf Majorana Conditions in Any Space-time}}\\ [0.8cm] \vs{10} {\large M.A. De Andrade$^{(a,\ast )}$ and F. Toppan$^{(b,\ast )}$} ~\\ \quad \\ {\em {~$~^{(a)}$} UCP, FT, Rua Bar{\~{a}}o do Amazonas, 124, cep 25685-070, Petr\'{o}polis (RJ), Brazil}\\ ~\quad\\ {\em ~$~^{(b)}$ UFES, CCE Depto de F{\'{\i}}sica, Goiabeiras cep 29060-900, Vit\'oria (ES), Brazil}\\ ~\quad \\ {\em {~$^{(\ast ) }$} CBPF, DCP, Rua Dr. Xavier Sigaud 150, cep 22290-180 Rio de Janeiro (RJ), Brazil} \end{center} \vs{6} \centerline{ {\bf Abstract}} \vs{6} Clifford algebras and Majorana conditions are analyzed in any spacetime. An index labeling inequivalent $\Gamma$-structures up to orthogonal conjugations is introduced.\par Inequivalent charge-operators in even-dimensions, invariant under Wick rotations, are considered. The hermiticity condition on free-spinors lagrangians is presented. The constraints put by the Majorana condition on the free-spinors dynamics are analyzed. Tables specifying which spacetimes admit lagrangians with non-vanishing kinetic, massive or pseudomassive terms (for both charge-operators in even dimensions) are given. The admissible free lagrangians for free Majorana-Weyl spinors are fully classified. \vs{6} \vfill \rightline{April 1999} \rightline{CBPF-NF-013/99} \rightline{hep-th/xxx} {\em E-Mail:\\ 1) [email protected]\\ 2) <EMAIL>} \newpage \pagestyle{plain} \renewcommand{\thefootnote}{\arabic{footnote}} \setcounter{footnote}{0} \vs{8} \section{Introduction.} The theory of Clifford algebras is an old subject which has been extensively investigated both in the mathematical and in the physicists' literature. For obvious reasons physicists mainly dealt with the theory of spinors in Minkowskian or Euclidean spacetimes \cite{{Glio},{Sche}}. Nevertheless in some papers [3--6] spinors in pseudoeuclidean spacetimes with arbitrary signature $(t,s)$, $D=t+s$ being the dimensionality of the spacetime, have been analyzed. In particular \cite{Kugo} can be regarded as the reference work on the subject since it presents a rather complete list of results in this topic.\par The development of supergravities and superstring theories which emphasize the Kaluza-Klein aspect of compactification to lower dimensional space led investigating properties of spinors (and supersymmetries) in arbitrary dimensional spaces. However, apart some special papers as the ones previously recalled, the great majority of works were still devoted to standard-signature spacetimes. The question of providing a physical interpretation for the extra-times somehow masked the fact that from a strictly mathematical point of view consistent superstring theories can be formulated in exotic signatures (like e.g. $5+5$). This negative attitude towards exotic signatures seems at present time changing and their possible physical implications find increasing attentions (see e.g. \cite{Bars}). Various reasons are at the basis of this shift of attitude. Some recent works \cite{Hull} for instance pointed out the existence of dualities relating theories formulated in different signatures. On the other hand the still-mysterious $M$-theory suggests that we need investigating along all possible directions. A rather formal argument can also be invoked, a reasonable demand for any possible theory which could claim to be a genuine ``theory of everything" is that the signature of the spacetime should be determined by the properties of the spacetime itself rather than imposed a priori. The Minkowskian signature should therefore be selected after confrontation with the other signatures. \par Motivated by the above considerations in this paper we analyze the real Clifford structures and Majorana conditions in any signature spacetime with arbitrary dimensions. Without loss of generality, the analysis here presented is based on a specific Weyl realization of Clifford algebras. The technique employed allows to recover the results of \cite{Kugo} in a considerably simplified manner. Besides that, extra-informations, not presented in \cite{Kugo}, are gained. As an example discussed in the text we mention the correct choice of the charge operator which preserves the Majorana condition under a Wick rotation to the Euclidean. \par A list of further topics here discussed, some of them we are not aware to be found elsewhere, is the following. An index is introduced to label and discriminate among classes of inequivalent Clifford algebras up to orthogonal conjugations. Such index could in principle be relevant to physical applications whenever some kind of reality conditions are imposed on the fields. \par Moreover the compatibility of the Majorana condition with the free massive equations of motion is thoroughly investigated. The conditions upon free lagrangians for Majorana spinors in order to be non-vanishing, hermitian and charge-conjugated are presented. Explicit and easy-to-consult tables of spacetimes supporting massive (or pseudomassive in the even-dimensional case) Majorana spinors are provided. The list of results here presented is more complete than the one given in reference \cite{DeAn}.\par \par The scheme the present work is as follows: the next section is devoted to notations and preliminary results. In section $3$ the specific Weyl representation employed is introduced. Even-dimensional inequivalent charge operators, invariant under Wick rotations, are constructed. The index labeling $\Gamma$-structures up to orthogonal conjugation is discussed in section $4$. The hermiticity condition on free-spinors lagrangians is discussed in section $5$. Section $6$ presents an exhaustive list of the constraints put by the Majorana condition on the free-spinors dynamics, both at the level of the equations of motion and of the action. Tables specifying which spacetimes admit lagrangians with non-vanishing kinetic, massive or pseudomassive terms (for both charge-operators in even dimensions) are given. Finally in section $7$ the problem of determining the admissible free lagrangians (kind of terms, non-vanishing conditions, type of coefficients) for Majorana-Weyl spinors in even-dimensional spacetimes is fully solved. \section{Notations and preliminary results.} Let $\eta^{\mu\nu}$ be the (pseudo)-euclidean metric associated to an $M^{t,s}$ generalized Minkowski space-time with $t$ time-directions and $s$ space-directions. The space-time dimension being $D=t+s$. In the following we will denote as time (space) directions those which are related to the $+$ (and respectively $-$) sign in $\eta^{\mu\nu}$.\par A $\Gamma$-structure associated to the $M^{t,s}$ spacetime is a matrix-representation of the Clifford algebra generators $\Gamma^\mu$ ($\mu =1,...,D$) satisfying the anticommutation relations \begin{eqnarray} \Gamma^\mu\Gamma^\nu +\Gamma^\nu\Gamma^\mu &=& 2\eta^{\mu\nu} \cdot {\bf 1}_{\Gamma} \label{1} \end{eqnarray} The representation is realized by $2^{[{D\over 2}]}\times 2^{[{D\over 2}]}$ matrices ($[{D\over 2}]$ denoting the integral part of ${D\over 2}$) which can be further assumed to satisfy the unitarity requirement \begin{eqnarray} {\Gamma^{\mu}}^\dag &=& {\Gamma^{\mu}}^{-1} \label{2} \end{eqnarray} A tracelessness condition holds \begin{eqnarray} tr \Gamma^\mu &=& 0 \label{3} \end{eqnarray} for any $\mu$.\par According to the fundamental Pauli theorem \cite{Saku} the above matrix-representation is uniquely realized up to unitary conjugation. \par Notice that choosing the ($+$) sign in the r.h.s. of (\ref{1}) is a matter of convention. The opposite choice (i.e. r.h.s. $\equiv -2\eta^{\mu\nu}$) is admissible. However it is just sufficient to look at results and tables obtained for an $(s,t)$-signature within the $+$ convention since the results for the $-$ convention-case are immediately recovered by interchanging $t$ and $s$: $t\leftrightarrow s$. It should be therefore clear that the complete set of solutions for $(s,t)$ spacetimes are recovered from the tables produced below only after such ($t,s$)-``dualization" has been taken into account. \par The introduction of lagrangians and charge conjugations for the spinor fields require the presence of three (only two of them mutually independent) unitary matrices, denoted in the literature as $A,B,C$, associated to each one of the three conjugations (hermitian, complex-conjugation and transposition respectively) acting on the $\Gamma^\mu$-matrices, according to \begin{eqnarray} A\Gamma^\mu A^\dag &=& (-1)^{t+1} {\Gamma^\mu}^\dag\nonumber\\ B\Gamma^\mu B^\dag &=&\eta{\Gamma^\mu}^\ast\nonumber\\ C\Gamma^{\mu}C^\dag &=& \eta (-1)^{t+1}{\Gamma^\mu}^T \label{4} \end{eqnarray} As discussed later $\eta$, as well as $\varepsilon$ introduced below, is a sign ($\pm 1$) specifying the assignment of a $\{\Gamma^\mu, A, B, C\}$ structure up to unitary transformations. $\eta$ and $\varepsilon$ will be explicitly computed in the next section. The introduction of $\eta$ as defined in (\ref{4}) corresponds to the standard convention in the literature.\par The equation relating $A,B$ and $C$ can be expressed through \begin{eqnarray} C&=& B^T A \label{5} \end{eqnarray} with the transposed matrix $B^T$ satisfying \begin{eqnarray} B^T &=& \varepsilon B \label{6} \end{eqnarray} An useful form of restating the above equation is \begin{eqnarray} B^\ast B &=& \varepsilon \cdot {\bf 1} \label{7} \end{eqnarray} The $A$-matrix can be expressed through the position \begin{eqnarray} A &=& \prod_{i=1,...,t}\Gamma^i \label{8} \end{eqnarray} where the product (the order is irrelevant since $A,B,C$ can always be determined up to an arbitrary phase) is restricted to time-like $\Gamma$-matrices, i.e. those satisfying the ${\Gamma^i}^2 =+{\bf 1}$ equation (conversely the spacelike $\Gamma$-matrices are those belonging to the complementary set satisfying ${\Gamma^j}^2=-{\bf 1}$).\par The $A$-matrix allows constructing in generic flat spacetimes $M^{t,s}$ the conjugated $\overline{\psi}$ spinor as $\overline{\psi}=\psi^{\dag}A$, and generalizes the $\Gamma^{0}$-matrix of the standard Minkowskian spacetime.\par The matrix $C$ corresponds to the charge-conjugation matrix, while $B$ is employed in introducing the charge-conjugated spinors $\psi^{c}$ according to \begin{eqnarray} \psi^{c}&=& B^{\dag}\psi^{\ast} \label{9} \end{eqnarray} Quantum mechanical states are rays in a Hilbert space. A physical spinorial state can be equally well described by a spinor transformed via an unitary matrix $U$, $\psi \mapsto U\psi$.\par It is easily proven that under such a unitary transformation $\Gamma^{\mu}, A,B,C$ are mapped as follows \begin{eqnarray} \Gamma^{\mu} &\mapsto& U\Gamma^\mu U^\dag\nonumber\\ A&\mapsto& UAU^{\dag}\nonumber\\ B&\mapsto& U^\ast B U^{\dag}\nonumber\\ C&\mapsto & U^\ast C U^\dag \label{10} \end{eqnarray} Notice that the unitary transformations acting upon $B,C$ {\em do not} coincide with their unitary conjugations.\par If we introduce the notion of a $\{\Gamma^{\mu}, A, B, C\}$-structure assignment associated to a given spacetime $M^{t,s}$ and we look for inequivalent classes of such assignments under the (\ref{10}) transformations, we easily realize that the $\eta, \varepsilon$ signs introduced above label inequivalent classes of assignments. Indeed $\eta,\varepsilon$ can be equivalently introduced in a unitary-invariant trace form as \begin{eqnarray} tr(B^\ast B) &=& \varepsilon \cdot tr {\bf 1}\nonumber\\ tr(B\Gamma^{\mu}B^\dag{\Gamma_\mu}^{\ast} )&=& \eta D\cdot tr{\bf 1} \label{11} \end{eqnarray} where the convention on the repeated indices is understood.\par With a slight abuse of language we can say that $\eta,\varepsilon$ label inequivalent choices of charge-conjugations.\par In an even-dimensional spacetime ($D=2n$) we can introduce a timelike generalized $\Gamma^5$ matrix (i.e. the matrix generalizing the one associated to the ordinary Minkowskian spacetime), through the position \begin{eqnarray} \Gamma^5 &=& (-1)^{{s-t\over 4}}\prod_{\mu = 1, ..., D} \Gamma^\mu \label{12}\end{eqnarray} The sign is chosen in order to guarantee ${\Gamma^5}^2={\bf 1}$.\par Let us conclude this section by presenting some further useful identities \begin{eqnarray} A^\dag&=& (-1)^{{t\over 2}(t-1)} A\nonumber\\ A^\ast &=&\eta^t B A B^\dag\nonumber\\ A^T &=& \eta^t (-1)^{{t\over 2}(t-1)} CAC^\dag \label{13} \end{eqnarray} and \begin{eqnarray} C^T &=& \varepsilon \eta^t (-1)^{{t\over 2}(t-1)} C \label{14} \end{eqnarray} \section{Clifford algebras and the Majorana condition.} The allowed values for the signs $\eta,\varepsilon$ labelling inequivalent $\{\Gamma^\mu, A,B,C\}$-structures associated to any given spacetime have been computed in \cite{Kugo}. A very efficient and much simpler method of computing $\eta,\varepsilon$ is at disposal by explicitly using a $\Gamma$-structure in a given Weyl representation. The choice of working within such Weyl representation can always be done and, due to the fundamental property that $\Gamma$-structures are all equivalent up to unitary conjugation \cite{Saku}, by no means affects the generality of the results so obtained. More than just reproducing previous results the computation with the choice we made encodes further information. Indeed we will show that our $C$ charge-conjugation operators are left unchanged under a Wick rotation. When inequivalent charge-conjugation operators are present the tables provided below inform which charge-conjugations should be correctly chosen in performing analytical continuation to let's say the Euclidean space. \par It is always possible to find (for a constructive approach see \cite{Cola}) for any even-dimensional ($D=2n$) $\Gamma$-structure a Weyl representation with the further property that the $\Gamma^\mu$ matrices are all symmetric or antisymmetric under transposition ($\Gamma^\mu = \pm {\Gamma^\mu}^T$) and moreover the number of symmetric equal the number of antisymmetric $\Gamma^\mu$ matrices ($=n)$.\par In odd dimensional spacetimes an extra symmetric matrix, the $\Gamma^5$ introduced in (\ref{12}) is presents.\par A Wick rotation of a timelike $\overline{\mu}$ direction into a spacelike direction is represented on $\Gamma$-matrices by the rescaling $\Gamma^{\overline{\mu}}\mapsto i\Gamma^{\overline{\mu}}$, while the remaining $\Gamma$-matrices are left unchanged. Clearly the symmetric or antisymmetric character of the $\Gamma^{\overline{\mu}}$ matrix is not affected by a Wick rotation.\par In a Weyl-represented even-dimensional $\Gamma$-structure we can introduce two inequivalent charge operators (i.e. realizing inequivalent $\{\Gamma^\mu, A, B,C\}$ assignments, see the discussion in the previous section) $C_S$ and $C_A$ defined as follows \begin{eqnarray} C_S &=&\prod_{i_S= 1,...,n}\Gamma^{i_S} \nonumber \\ C_A &=&\prod_{i_A=1,...,n}\Gamma^{i_A} \label{15} \end{eqnarray} the products being restricted to symmetric (and respectively antisymmetric) $\Gamma$-matrices. As in the definition of the matrix $A$ (\ref{8}), the ordering of the products is irrelevant. Please notice that the index ($S$ or $A$) labeling $C$ reflects the construction, via symmetric or antisymmetric matrices, of the corresponding charge-conjugation operator and not its (anti)-symmetry property which is expressed by formula (\ref{14}). {}From (\ref{8}) and (\ref{15}) we obtain the relation $C_A\equiv C_S\Gamma^5$. \par \par Clearly $C_S$ and $C_A$ are left invariant by Wick rotations up to an arbitrary phase, implying the convenience of the Weyl basis in discussing such an issue. \par In odd-dimensional spacetimes a charge operator $C$ can be introduced by using both formulas in (\ref{15}). Due to the presence in this case of the extra $\Gamma^5$ among the symmetric matrices, the two definitions indeed collapse into a single one (modulo an arbitrary phase), recovering the well-known result that there exists a unique $\{\Gamma^\mu, A, B, C \}$-assignment, up to unitary transformations, in odd dimensions.\par We recall that the $A$ matrix is defined in (\ref{8}), while $B_{S,A}$ are introduced from (\ref{5}) as \begin{eqnarray} B_{S,A} &=& A\cdot{C_{S,A}}^T \label{16} \end{eqnarray} If we take into account the fact that timelike $\Gamma$-matrices are hermitian, it is just a matter of tedious but straightforward computations to check for both $(S,A)$-cases, which ($\pm 1 $)-signs correspond to $\eta_S$, $\eta_A$, as well as $\varepsilon_S$, $\varepsilon_A$, introduced in the formulas (\ref{4}) and (\ref{6}). \par In an ($s,t$) even-dimensional spacetime we obtain the following table \begin{center} \begin{tabular}{|c|c|c|c|c|} \hline $\spadesuit$& $0$ & $2$ &$ 4$ &$ 6$ \\ \hline $\eta_S$ & $+$ &$ -$ & $+$ & $-$ \\ \hline $\eta_A $ & $-$ &$ +$ &$ -$ & $+$ \\ \hline $\varepsilon_S $ & $+$ &$ + $& $- $&$ -$ \\ \hline $ \varepsilon_A $ &$ +$ &$ -$ & $- $& $+$ \\ \hline \end{tabular} \end{center} \begin{eqnarray} &&\label{17} \end{eqnarray} where the even values characterizing the columns correspond to \begin{eqnarray} X&=& s-t\quad mod\quad 8 \label{18} \end{eqnarray} A similar table can be produced for odd-dimensional spacetimes. In this case no splitting between the $S,A$-cases is produced \begin{center} \begin{tabular}{|c|c|c|c|c|} \hline $\spadesuit $ & $1$ & $3$ & $5$ & $7$ \\ \hline $\eta$ & $-$ & $+$ &$ -$ & $+$ \\ \hline $\varepsilon $& $+ $&$ -$ & $-$ & $+$ \\ \hline \end{tabular} \end{center} \begin{eqnarray} &&\label{19} \end{eqnarray} As above the columns are marked by $X$ given by (\ref{18}).\par Another sign, denoted by $\xi $ and important for later considerations, is introduced through the position \begin{eqnarray} B\Gamma^5 B^\dag &=& \xi \Gamma^5 \label{20} \end{eqnarray} where $\Gamma^5$ is the timelike extra-$\Gamma$ matrix given in (\ref{12}). We obtain \begin{center} \begin{tabular}{|c|c|c|c|c|} \hline $\spadesuit$ & $0$ & $2$ & $4$ & $6$ \\ \hline $\xi$ & $+$ & $-$ & $+$ & $-$ \\ \hline \end{tabular} \end{center} \begin{eqnarray} &&\label{21}\end{eqnarray} Here as well columns correspond to $X$ given in (\ref{18}).\par The Majorana reality condition on spinors is a constraint on the charge-conjugated spinor $\psi^c$ introduced in (\ref{9}), imposed to satisfy \begin{eqnarray} \psi^c &=& \psi \label{22} \end{eqnarray} Such a constraint can be consistently set only when $\varepsilon = +1$ (due to the combined result of applying the complex conjugation on (\ref{9}) and the formula (\ref{7})).\par One of the consequences read from the table (\ref{17}) is the well-known result that Majorana spinors do not exist in the euclidean $4$-dimensional space.\footnote{Enlarged reality conditions which are applicable when $\varepsilon=-1$, like the $SU(2)$-reality condition proposed be Wetterich \cite{Wett}, will not be discussed in the present paper.}\par The table (\ref{17}) is useful for another purpose. It allows reading which choice of the $C_S$, $C_A$ charge-conjugation operators should be adopted to mantain a Majorana reality condition if a Wick analytical continuation is performed. Indeed in the $0$-column both the $C_S$ and $C_A$ charge-conjugation operators are consistent with the Majorana condition. Therefore e.g. the $(2,2)$ spacetime supports inequivalent Majorana spinors, based either on $C_S$ or on $C_A$; conversely for the standard $(3,1)$-Minkowskian spacetime the Majorana condition is only defined w.r.t. $C_S$. Recalling the property that $C_S$, $C_A$ are left unchanged by Wick rotation, it turns out that only the $(2,2)$ $C_S$-Majorana spinors are Wick related to the $(3,1)$ Majorana spinors.\par An euclidean space which supports Majorana spinors is the $10$-dimensional one. We obtain the two following chains of Wick-related Majorana spinors: \begin{eqnarray} C_S&:& (10,0)\rightarrow (9,1) \rightarrow (6,4)\rightarrow (5,5) \rightarrow (2,8) \rightarrow (1,9)\nonumber\\ C_A&:& (0,10) \rightarrow (1,9) \rightarrow (4,6)\rightarrow (5,5) \rightarrow (8,2) \rightarrow (9,1) \label{23} \end{eqnarray} The three potentially ambiguous cases are $(9,1)$, $(5,5)$ and $(1,9)$ which present both kinds of Majorana spinors. \section{Inequivalent real Clifford-Weyl structures.} Let the Clifford $\Gamma$-structure in the specific Weyl basis given above be denoted a Clifford-Weyl structure. In this section we provide an answer to the question: how many inequivalent real Clifford-Weyl structures do exist? To provide a solution we introduce an appropriate index labeling inequivalent structures. \par The mathematical formulation of the problem is better phrased as finding the classes of equivalence of $\Gamma$-matrices up to orthogonal conjugation \begin{eqnarray} \forall \mu , \quad \Gamma^\mu &\mapsto & O\Gamma^\mu O^T\label{24} \end{eqnarray} with $O$ $2^{[{D\over 2}]}\times 2^{[{D\over 2}]}$ real-valued, orthogonal ($OO^T=O^TO={\bf 1}$) matrices.\par We already mentioned that the fundamental Pauli theorem guarantees that $\Gamma$-matrices are uniquely represented up to unitary conjugation; they however fit into different classes when just orthogonality is concerned. \par One could ask whether this well-posed mathematical problem has sensible physical consequences. Indeed, as far as quantum mechanics is concerned, equivalent descriptions are provided by unitary-transformed states in a given Hilbert space. However, if some reality condition has to be imposed, it may well restrict the class of allowed transformations to be the orthogonal ones. Indeed this happens when e.g. the Majorana reality condition is imposed on spinors. Later we comment more on that. \par The above mathematical problem finds the following solution. \par Let an index $I$ be defined for a $D$-dimensional $(s,t)$-spacetime ($D=s+t$) through the position \begin{eqnarray} I&=& {1\over 2^{([{D\over 2}]+1)}}\cdot tr ( \Gamma^\mu {\Gamma_\mu}^\ast ) \label{25} \end{eqnarray} The sum over repeated indices is understood. The normalization is chosen for a matter of convenience and as before $[{D\over 2}]$ denotes the integral part of ${D\over 2}$.\par $I$ is clearly left invariant by orthogonal transformations (\ref{24}) while it is affected by unitary conjugations of $\Gamma$-matrices. It can be therefore used to label inequivalent classes of $\Gamma$-matrices up to orthogonal conjugation.\par In a Weyl basis $I$ can be easily computed. Indeed, as previously recalled, in such a basis $\Gamma$-matrices are either symmetric or antisymmetric. ${\Gamma^\mu}^\ast$ coincides with $\Gamma^\mu$ up to a sign which is determined by both the time-like or space-like character of the $\mu$ direction, as well as the (anti)-symmetry nature of $\Gamma^\mu$. It is a matter of straightforward computations to check the following results.\par $i)$ Let us consider at first an even ($D=2n$) dimensional spacetime. We denote as $t_A$ ($s_A$) the number of time-like (space-like) directions associated to antisymmetric $\Gamma$-matrices. The number of symmetric timelike (spacelike) matrices is therefore $t-t_A$ ($s-s_A$). In a Weyl basis the equality $t_A + s_A = n$ holds. For a Weyl assignment with $t_A$ antisymmetric timelike matrices the index $I$ takes the value \begin{eqnarray} I &=& t-2t_A \label{26} \end{eqnarray} Let us introduce $m$ given by $m=\min (s,t)$. We are free to choose among $m+1$ different Weyl assignments, $t_A=0,..., m$, each one leading to a different value for $I$ and therefore inequivalent under orthogonal conjugations. Indeed we obtain $m+1$ possible values for $I$, \begin{eqnarray} -m + 2j &,&\quad\quad\quad j=0,...,m\label{27} \end{eqnarray} in the even-dimensional case.\par $ii)$ Let $D=2n+1$ ($s+t=2n+1$) be an odd-dimensional spacetime ($s+t=2n+1$). An extra (the generalized-$\Gamma^5$ matrix here denoted $\Gamma^{2n+1}$) symmetric matrix is present w.r.t. the previous case. It could be associated either to a time-like or to a space-like direction according to the sign \begin{eqnarray} {\Gamma^{2n+1}}^2 &=& \kappa = \pm 1\label{28} \end{eqnarray} Let as before $t_A$ be the number of antisymmetric timelike $\Gamma$-matrices. The computation of the index $I$ follows the same steps. We obtain \begin{eqnarray} I &=& t - 2t_A -{\textstyle{1\over 2}} (1-\kappa ) \label{29} \end{eqnarray} Notice that for a fixed $(s,t)$-spacetime the parity of $I$ is determined by the sign of $\kappa$: \begin{eqnarray} (-1)^I &=& \kappa (-1)^t\label{30} \end{eqnarray} or conversely \begin{eqnarray} \kappa &=& (-1)^{t+I} \label{31} \end{eqnarray} which implies that the timelike or spacelike character of $\Gamma^{2n+1}$ cannot be reverted by orthogonal conjugations. As before let us put $m=\min (s,t)$.\par $\kappa$ can be arbitrary chosen unless $m=0$. $I$ can assume $2m+1$ different values labeling corresponding inequivalent classes. For $m\neq 0$ they are given by \begin{eqnarray} &&m-2j -{\textstyle{1\over 2}} (1-\kappa )\nonumber\\ &&\kappa =\pm 1, \quad\quad j = 0,1,..., m-{\textstyle{1\over 2}} (1+\kappa ) \label{32} \end{eqnarray} For $m=0$ either we have $I= +1$ in the $(2n+1,0)$ case or $I=-1$ in the $(0,2n+1)$ case.\par Let us discuss now a possible application of the above construction to the Majorana reality condition. A standard result (see \cite{Kugo}) states that for $\varepsilon = 1$, i.e. the consistency requirement for the Majorana condition, the $B$ matrix introduced in (\ref{5}) can be unitary-transformed (\ref{10}) to the identity matrix \begin{eqnarray} && \exists U \quad s.t. \quad U^\ast B U^\dag ={\bf 1} \label{33} \end{eqnarray} The choice $B\equiv {\bf 1}$ corresponds to the so-called Majorana representation ($\psi^c =\psi^\ast$). The orthogonal transformations are the unitary transformations acting on $B$ and preserving the Majorana representation \begin{eqnarray} U^\ast {\bf 1} U^\dag ={\bf 1} &\Rightarrow & U U^T = U^T U ={\bf 1} \label{34} \end{eqnarray} From (\ref{4}) in the Majorana representation we have ${\Gamma^{\mu}}^\ast=\eta \Gamma^\mu $, so that the index $I$ takes the value \begin{eqnarray} I &=& \eta D \label{35} \end{eqnarray} In this particular case the information furnished by the index $I$ is reduced to the same information provided by $\eta, \varepsilon$. The logics behind is however different. $\eta,\varepsilon$ label inequivalent classes under unitary transformations of a richer $\{\Gamma^\mu, A, B, C\}$-structure, while $I$ corresponds to inequivalent classes of orthogonal transformations of just a $\Gamma$-structure (in the Majorana realization). It deserves a careful investigation to determine whether for other choices of reality conditions which can select physical fields (such as the $SU(2)$-Majorana condition on spinors) the index $I$ can refine the standard classification and be physically meaningful. In a different but related context we already found \cite{Cola} a physical application where the index plays a non-trivial role. \section{Free Hermitian actions.} The most general lagrangian involving free spinorial fields is given by \begin{eqnarray} &&\alpha\cdot {\overline \psi}\Gamma^\mu\partial_\mu\psi +\beta \cdot{\overline\psi}\psi +\gamma\cdot{\overline \psi}\Gamma^\mu\Gamma^5\partial_\mu\psi +\delta\cdot {\overline\psi}\Gamma^5\psi \label{36} \end{eqnarray} The third (pseudokinetic) and the fourth (pseudomassive) term involve the $\Gamma^5$ matrix defined in (\ref{12}) and are present in even $D$-dimensional spacetimes only.\par The transposition acting on anticommuting fields $\zeta,\psi$ satisfies \begin{eqnarray} (\zeta\cdot\psi)^T&=& -\psi^T\cdot\zeta^T \label{37} \end{eqnarray} while the hermitian conjugation can be conventionally defined, without losing generality, according to \begin{eqnarray} (\zeta\cdot\psi)^\dag &=& \psi^\dag\cdot\zeta^\dag \label{38} \end{eqnarray} (as for the complex conjugation, it follows from (\ref{37}) and (\ref{38})).\par Demanding the hermiticity of the action, i.e. of the (\ref{36}) lagrangian, fixes unambiguously the nature, real or imaginary, of the coefficients in (\ref{36}). Straightforward computations lead to the table \begin{center} \begin{tabular}{|c|c|c|c|c|} \hline $\spadesuit $ & $0$ & $1$& $2$ & $3$ \\ \hline $\alpha$ & ${\bf R }$& ${\bf I}$ & ${\bf I}$ & ${\bf R}$ \\ \hline $\beta$ & ${\bf R}$ & ${\bf R}$ & ${\bf I}$ & ${\bf I}$ \\ \hline $\gamma$& ${\bf I}$ & ${\bf I}$ & ${\bf R}$ & ${\bf R}$ \\ \hline $\delta $& ${\bf R}$ & ${\bf I}$ & ${\bf I}$ & ${\bf R}$ \\ \hline \end{tabular} \end{center} \begin{eqnarray} &&\label{39} \end{eqnarray} the columns are labeled by $t\quad mod\quad 4,\quad$ $t$ being the number of time-like dimensions.\par The table above is useful e.g. in finding mass-shell properties. Indeed in the case of a theory involving, let's say, massive and/or pseudomassive terms, the mass-shell condition reads as follows \begin{eqnarray} p^2 &=& {{1\over \alpha^2}} (\delta^2-\beta^2) \end{eqnarray} where $\alpha,\beta,\delta$ enter (\ref{36}). The hermiticity requirement (\ref{39}) allows to set e.g. in the ($3,1$) Minkowski spacetime $\alpha = i$, $\beta = m$, $\delta = i m_5$, with $m,m_5$ real, so that $p^2=m^2 + {m_5}^2$ is necessarily positive. In a $(2,2)$-spacetime we only need to change the definition of $\beta$, which must be imaginary, by setting $\beta = i m $. The mass-shell condition reads $p^2 = {m_5}^2 - m^2$. A vanishing value can be found even for $m, m_5 \neq 0$ provided that $m=m_5$. \section{Majorana constraints on the dynamics.} In this section we analyze the constraints put by the (\ref{9}) Majorana condition on the dynamics of free spinors.\par {} From the (\ref{36}) lagrangian we derive the equation of motion \begin{eqnarray} \alpha\Gamma^\mu\partial_\mu\psi +\beta\psi +\gamma\Gamma^\mu\Gamma^5\partial_\mu\psi +\delta\Gamma^5\psi &=&0 \label{41} \end{eqnarray} The above equation of motion is compatible with the (\ref{9}) Majorana condition provided the coefficients are constrained to satisfy \begin{eqnarray} {\alpha^\ast} &=& \chi\cdot (\eta \alpha ) \nonumber\\ {\beta^\ast}&=& \chi\cdot\beta\nonumber\\ {\gamma^\ast} &=& \chi\cdot (\eta\xi\gamma )\nonumber\\ \delta^\ast &=& \chi \cdot (\xi \delta ) \end{eqnarray} where $\eta$, $\xi$ have been introduced in (\ref{4}) and (\ref{6}) respectively. The common factor $\chi$, as far as the equation of motion alone is concerned, is an arbitrary phase \begin{eqnarray} \mid \chi \mid^2 &=& 1 \label{43} \end{eqnarray} The derivation of the (\ref{41}) equation of motion from a lagrangian puts further constraints. Each Majorana-constrained ${\cal L}_i$ ($i=1,...,4$) term appearing in (\ref{36}), in order to be non-vanishing, must be symmetric, i.e. \begin{eqnarray} {{\cal L}_i}^T &=& {\cal L}_i \end{eqnarray} For an $(s,t)$-spacetime ($s+t=D$) this so happens when the following signs assume the $+1$ value: \par $i)$ for the kinetic term the sign is $\lambda$ given by \begin{eqnarray} \lambda &=& -\varepsilon\eta^{t+1}(-1)^{\textstyle{t\over 2}(t+1)}\label{45} \end{eqnarray} $ii)$ for the massive term, $\mu$ \begin{eqnarray} \mu &=& -\varepsilon \eta^t (-1)^{\textstyle{t\over 2}(t-1)}\label{46} \end{eqnarray} $iii)$ for the pseudokinetic term, $\lambda_5$ \begin{eqnarray} \lambda_5 &=& \lambda (-1)^{\textstyle{D\over 2}} \end{eqnarray} $iv)$ for the pseudomassive term, $\mu_5$ \begin{eqnarray} \mu_5 &=& \mu (-1)^{\textstyle{D\over 2}} \end{eqnarray} The following tables, specifying which spacetimes support the existence of non-vanishing kinetic and massive terms, can be produced. For even-dimensional spacetimes we have\\ \begin{center} \begin{tabular}{|c|c|c|c|c|}\hline $ \spadesuit$ & $0$ & $2$ & $4$ & $6$ \\ \hline $\lambda_S$ & $1,2$ & $0,1$ & & \\ \hline $\lambda_A$ & $0,1$ & & & $1,2$ \\ \hline $\mu_S$ & $2,3$ & $1,2$ & & \\ \hline $\mu_A$ & $1,2$ & & & $2,3$ \\ \hline \end{tabular} \end{center} \begin{eqnarray} &&\label{49} \end{eqnarray} Some comments are in order. The columns are labeled by $X=s-t\quad mod\quad 8$. The index $S$ or $A$ is referred to the corresponding charge-conjugation (either $C_S$ or $C_A$). The entries are evaluated only when $\varepsilon =1$ (Majorana consistency requirement); the $\sharp$'s in the entries specify for which number of $t$ time-like directions \begin{eqnarray} t&=&\sharp \quad mod \quad 4 \end{eqnarray} the sign in the associated row assumes the $+1$ value.\par Similarly, for odd-dimensional spacetimes, we have\\ \begin{center} \begin{tabular}{|c|c|c|c|c|} \hline $\spadesuit$ & $1$ & $3$ & $5$ & $7$ \\ \hline $\lambda$ & 0,1 & & & $1,2$ \\ \hline $\mu$ & $1,2$ & & & $2,3$ \\ \hline \end{tabular} \begin{eqnarray} &&\label{51} \end{eqnarray} \end{center} (same meaning for the symbols).\par The next question to be answered is whether the action associated to the (\ref{36}) lagrangian admits a charge-conjugation which allows to consistently introduce the Majorana condition. This point has been raised in \cite{DeAn}. The existence of a charge conjugation requires \begin{eqnarray} {\cal L}^\ast &=& {\cal L} \end{eqnarray} and is automatically guaranteed from the non-vanishing condition ${\cal L}^T={\cal L}$ once assumed the hermiticity of the action (${\cal L}^\dag = {\cal L}$), i.e. when the coefficients are chosen to satisfy the table (\ref{39}).\par It turns out that the phase $\chi$ appearing in (\ref{43}) is no longer arbitrary now but fixed to be \begin{eqnarray} \chi &=& -\eta^t \end{eqnarray} {} From the (\ref{49}) and (\ref{51}) tables above we can extract some particular results, e.g. that massive lagrangians for Majorana spinors exists in\\ $i)$ $t= 1 \quad mod\quad 4$ spacetimes (for $\eta=-1$) when\par $ia)$ $s-t = 0\quad mod \quad 8$ (for the $C_A$ charge-operator),\par $ib)$ $s-t = 2\quad mod\quad 8$ (for the $C_S$ charge-operator),\par $ic)$ $s-t = 1 \quad mod\quad 8 $, \\ as well as in\\ $ii)$ $t=2\quad mod\quad 4$ (for $\eta=+1$) when\par $iia)$ $s-t=0\quad mod\quad 8$ (for the $C_S$ charge-operator),\par $iib)$ $s-t = 6 \quad mod \quad 8 $ (for the $C_A$ charge-operator), \par $iic)$ $s-t=7 \quad mod\quad 8$.\par The role of $s,t$ can be interchanged as recalled in section $2$.\par In the case of odd-dimensional spacetimes the table (\ref{51}) provides further information. Kinetic ($K$) or massive ($M$) terms are only allowed in $D$-dimensional spacetimes according to \begin{eqnarray} && D= 1\quad mod\quad 8 \quad \{K\}\nonumber\\ && D=3 \quad mod\quad 8\quad \{K,M\}\nonumber\\ && D=5 \quad mod \quad 8\quad \{M\}\nonumber\\ && D=7 \quad mod \quad 8 \quad \{...\} \end{eqnarray} Up to $D=11$ dimensions the list of odd-dimensional spacetimes supporting Majorana spinors is given by \begin{eqnarray} \{K,M\}&:& (2,1), \quad (10,1), \quad (9,2), \quad(6,5)\nonumber\\ \{K\} &:& (1,0),\quad (9,0),\quad (8,1),\quad (5,4)\nonumber\\ \{M\} &:& (2,3)\nonumber\\ \{...\} &:& (7,0),\quad (4,3) \end{eqnarray} For even-dimensional spacetimes (up to $D=10$) an useful table can be written\\ \begin{center}\begin{tabular}{|c|c|c|c|c|c|} \hline $\spadesuit$ & $2$ & $4$ & $6$ & $8$ & $10$ \\ \hline $0$ & $S- K P$ & $$ & $A+$ & $\begin{array}{l} S+ \\ A-K \end{array}$ & $S - K P$ \\ \hline $1$ & $\begin{array}{l} S + K P \\ A - K M \end{array}$& $S-KMP$ & $$ & $A+K$ & $\begin{array}{l} S+KP \\ A-KM \end{array}$ \\ \hline $2$ & $A+KM$ & $\begin{array}{l} S+KMP \\ A- \end{array}$ & $S-$ & $$ & $A+KM$ \\ \hline $3$ & $\bullet$ & $A+$ & $\begin{array}{l} S+ \\ A- \end{array}$ & $S-$ & $$ \\ \hline $4$ & $\bullet$ & $$ & $A+$ & $\begin{array}{l} S + \\ A-K \end{array}$ & $S-KP$ \\ \hline $5$ & $\bullet$ & $\bullet$ & $$ & $A+K$ & $\begin{array}{l} S+KP \\ A-KM \end{array}$\\ \hline $6$ & $\bullet$ & $\bullet$ & $S-$ & $$ & $A+KM$ \\ \hline $7$ & $\bullet$ & $\bullet$ & $\bullet$ & $A+$ & $$ \\ \hline $8$ & $\bullet$ & $\bullet$ & $\bullet$ & $\begin{array}{l} S + \\ A-K \end{array}$ & $S-KP$ \\ \hline $9$ & $\bullet$ & $\bullet$ & $\bullet$ & $\bullet$ & $\begin{array}{l} S +KP \\ A-KM \end{array}$ \\ \hline $10$& $\bullet$ & $\bullet$ & $\bullet$ & $\bullet$ & $A+KM$ \\ \hline \end{tabular} \end{center} \begin{eqnarray} && \end{eqnarray} It contains the following informations. Columns are labeled by $D$, rows by $t$. Each entry is evaluated for $\varepsilon =1$. The presence of $S$ or $A$ denotes if the corresponding charge-operator defines a Majorana spinor. The sign ($\pm$) represents the corresponding value of $\eta$. The presence of $K,M,P$ denotes if the kinetic ($K$), massive ($M$) or pseudomassive ($P$) term in the (\ref{36}) lagrangian can be non-vanishing. These last two terms have been evaluated only when the corresponding kinetic term is nonzero. The pseudokinetic term has not been inserted here since its physical interpretation is problematic (due to the presence of negative-normed states, which however can be eliminated if projected out, as in the Majorana-Weyl case discussed in the next section). \par Since the same results are repeated $mod\quad 8$, both in $s$ and in $t$, the following compact information can be extracted. Majorana spacetimes with a non-vanishing kinetic term can be found in \begin{eqnarray} D&=& 0\quad mod \quad 8: \quad\{ A, K \}\nonumber\\ D&=&2\quad mod \quad 8: \quad either\quad \{S, KP\}\quad or \quad \{A, KM\}\nonumber\\ D&=& 4 \quad mod\quad 8: \quad \{S, KMP\} \end{eqnarray} In $D=2, 10,...$ either a massive or a pseudomassive term could be present, according to the choice of the charge-conjugation operator. Simultaneous presence of massive and pseudomassive terms is allowed in $D=4,12,...$ dimensions only. \section{The Majorana-Weyl conditions.} In order to make this paper self-consistent we review in this section the status of Majorana-Weyl spinors and present a complete list of results.\par In $D=2n$ even-dimensional spacetimes the projectors \begin{eqnarray} P_{R,L} &\equiv & (\frac{{\bf 1} \pm \Gamma^5}{2}) \end{eqnarray} (where $\Gamma^5$ has been introduced in (\ref{12})) allow defining chiral (Weyl) spinors $\psi_{R,L}$ as \begin{eqnarray} \psi_{R,L} &=& P_{R,L}\psi \label{59} \end{eqnarray} Majorana-Weyl spinors, satisfying both the condition (\ref{9}) and the projection (\ref{59}), can be consistently defined (see \cite{Kugo}) in spacetimes such that \begin{eqnarray} s-t &=& 0\quad mod\quad 8 \label{60} \end{eqnarray} therefore in particular in all $(n,n)$ spacetimes. Up to $10$ dimensions the remaining spacetimes supporting Majorana-Weyl spinors are the euclidean $(8,0)$ space and the minkowskian $(9,1)$ spacetime.\par In any spacetime satisfying (\ref{60}) Majorana-Weyl spinors can be introduced for both $C_S$ and $C_A$ charge-conjugation operators. The list of results presented below holds in both cases.\par Let us first recall that ${\overline\psi} =\psi^T C$ under the condition (\ref{9}) and that moreover the $C_{S,A}$ charge-operator is respectively block-diagonal or block-antidiagonal according if $n$ is even or odd. As a consequence kinetic ($K$) and massive ($M)$ terms can either mix (denoted in such case as $K_{xy}, M_{xy}$) chiralities or not ($K_{xx}, M_{xx})$. We can write \begin{eqnarray} K_{xx} &\equiv & {\psi_{R,L}}^T C \Gamma^\mu \partial_\mu \psi_{R,L}\nonumber\\ M_{xx} &\equiv & {\psi_{R,L}}^T C\psi_{R,L} \label{61} \end{eqnarray} and \begin{eqnarray} K_{xy} &\equiv & {\psi_R}^T C\Gamma^\mu \partial_\mu \psi_L + \lambda {\psi_L}^T C\Gamma^\mu\partial_\mu \psi_R\nonumber\\ M_{xy} &\equiv & {\psi_R}^T C\psi_L +\mu {\psi_L}^T C\psi_R \label{62} \end{eqnarray} The ``mixed" terms $K_{xy}, M_{xy}$ can always be chosen to be non-vanishing. It is sufficient for this purpose to conveniently fix the relative sign between the two terms in the r.h.s. of (\ref{62}). This is done in (\ref{62}), the two signs $\lambda$ and $\mu$ coincide with their values given in (\ref{45}) and (\ref{46}) respectively.\\ Conversely the $K_{xx}$ and $M_{xx}$ terms could be identically zero according to the (anti-)symmetry properties of $C$. \par Let us now introduce $\upsilon= (-1)^n$. From the previous remarks on the block-character of $C$ we have that \begin{eqnarray} K_{\upsilon = +1} \equiv K_{xy} \quad &,& \quad M_{\upsilon =+1} \equiv M_{xx}\nonumber\\ K_{\upsilon=-1} \equiv K_{xx} \quad &,& \quad M_{\upsilon = -1} \equiv M_{xy} \label{63} \end{eqnarray} The most general free lagrangian for Majorana-Weyl spinors in $D=2n$ dimensions can be expressed as \begin{eqnarray} {\cal L} &=& \alpha K_{\upsilon} +\beta M_{\upsilon} \end{eqnarray} The formula (\ref{63}) specifies which kind of kinetic and which kind of massive term could appear in $D=2n$. The coefficients $\alpha,\beta$ are either real or imaginary according to the table (\ref{39}).\par The last feature to be computed is in which dimensions the $K_{xx}$, $M_{xx}$ terms are not identically vanishing. The final results can be summarized in the following table, which presents the types of allowed kinetic and massive terms in accordance with the dimensionality $D$ of the spacetime \begin{eqnarray} && D = 0\quad mod\quad 8,\quad \{ K_{xy}\}\nonumber\\ && D=2\quad mod\quad 8, \quad \{ K_{xx}, M_{xy} \}\nonumber\\ && D=4\quad mod \quad 8, \quad \{ K_{xy}, M_{xx}\}\nonumber\\ && D=6 \quad mod\quad 8, \quad \{ M_{xy} \} \end{eqnarray} The list of results presented in this section removes any possible ambiguities and completely determines all features of the free actions for Majorana-Weyl spinors in any space-time. \section{Conclusions.} This paper has been devoted to discuss real structures in Clifford algebras and Majorana conditions in any space-time. Without loss of generality a specific Weyl representation, useful for dealing with Wick rotations, of Clifford algebras has been employed to analyze $\Gamma$-structures and Majorana spinors. An index, which to our knowledge has not been discussed before at least in the physicists' literature, has been introduced. It classifies $\Gamma$-structures up to orthogonal conjugations. \par For what concerns Majorana spinors, some of the issues here discussed have not been considered in previous papers. We can mention e.g. the interplay between the hermiticity condition, the charge-conjugation and the non-vanishing condition for the (\ref{36}) lagrangian. The different role played by the even-dimensional charge operators $C_S$, $C_A$ (\ref{15}), invariant under Wick rotations, is another example.\par We have furnished a series of tables presenting an exhaustive list of results concerning Majorana and Majorana-Weyl spinors. They include in particular the non-vanishing conditions in any given space-time for kinetic, massive and pseudomassive terms, in association with each charge operator $C_S$, $C_A$ (in even dimensions), as well as the type of coefficients (\ref{39}) and the kind of terms (in the Majorana-Weyl case) entering the free lagrangian (\ref{36}).\par One of our main motivations for presenting here such a systematic list of results concerns their relevance in analyzing supersymmetries in generic pseudoeuclidean spacetimes. Their connection with supergravities, strings, brane dynamics, etc., could be explored (for a recent review of this topic in standard Minkowskian spacetimes see e.g. \cite{West}). In the introduction we mentioned why this issue could be important. Problems like Kaluza-Klein compactifications, dimensional reductions, analytical continuations to the euclidean spaces, are among those which have to be addressed. \vskip1cm \noindent{\Large{\bf Acknowledgments}} \\ {\quad}\\ We are pleased to acknowledge J. A. Helay\"{e}l-Neto and L.P. Colatto for both encouragement and helpful discussions. We are grateful to C. Preitschopf for having pointed out an ambiguous statement made in the preliminary version, which is now fully clarified. We are grateful to DCP-CBPF for the kind hospitality.
\section{Introduction}\label{intro} Quantum stochastic calculus (QSC) \cite{1HP84,GarC85,Gar,Partha92}, a noncommutative analog of the classical Ito's stochastic calculus, revealed to be a powerful tool to construct mathematical models of quantum optical systems \cite{Gar,Gar86,Bar87,KW88,AMW88,LRW88,MRW88,CW88,WM94} and to develop a theory of photon detection \cite{BarL85,Bar86,Bar90,BPag}. Just at the beginning of QSC, Hudson and Parthasarathy proposed a quantum stochastic Schr{\"o}dinger equation for quantum open systems \cite{1HP84,Partha92}. Such an equation has found applications in quantum optics, but not in its full generality \cite{Bar87,Bar90,Gar}. It has been used to give, at least approximately, the dynamics of photoemissive sources such as an atom absorbing and emitting light, or matter in an optical cavity, which exchanges light with the surrounding free space. But in these cases the possibility of introducing the so called gauge (or number) process in the dynamical equation has not been considered; roughly speaking, the gauge process is a quadratic expression in the field operators which preserves the number of quanta, but changes their wave functions. In this paper we want to show, in the case of the simplest photoemissive source, namely a two-level atom stimulated by a laser, how the full Hudson-Parthasarathy equation allows to describe in a consistent way the scattering of the light by the atom not only through the absorption/emission channel, but also through another process which can be called ``direct scattering'', which can be included in the interaction via a term containing the gauge process. When the atom is approximated by a two-level system, the introduction of an interaction term which preserves the number of photons allows to simulate also the scattering processes involving virtual transitions to states different from the two ones responsible of the real absorption/emission process. So, a first aim is to show how the full Hudson-Parthasarathy equation is able to give a reasonable and rich model for the dynamics of an atom interacting with the electromagnetic field. A second one will be the study of the elastic, inelastic and total cross sections for the scattering of monochromatic coherent light by the atom. The resulting line-shapes are very interesting. For instance, the dependence of the total cross section on the frequency of the stimulating laser can present not only a Lorentzian shape, but the full variety of Fano profiles $(\!\!{}$\nobreak\cite{Fano}, \cite{CT92} pp.\ 61--63). Moreover, the dependence of the line shape on the intensity of the stimulating laser is computed and power broadening and intensity dependent shifts are found. The study of the inelastic cross section, instead, shows possible modifications to the known triplet structure of the fluorescence spectrum \cite{Moll}. Some preliminary results on the total cross sections where reported in \cite{BarL98}. \subsubsection*{Fock space and QSC} Let us recall some notions of QSC and the Hudson-Parthasarathy equation; this is just to fix our notations, while for the proper mathematical definitions and the rules of QSC we refer to the book by Parthasarathy \cite{Partha92}. We denote by ${\cal F} = {\cal F}({\cal X})$ the Boson Fock space over the ``one-particle space'' ${\cal X} = {\cal Z} \otimes L^2({\mathbb R}_+) \simeq L^2({\mathbb R}_+; {\cal Z})$, where ${\cal Z}$ is another separable complex Hilbert space. A vector $f$ in $\cal X$ is a function from ${\mathbb R}_+$ into $\cal Z$; we fix a c.o.n.s.\ $\{e_i,\ i\geq 1\}$ in ${\cal Z}$ and we denote by \begin{equation}\label{1.3} f_j(t) = \langle e_j| f(t) \rangle \end{equation} the components of a vector $f(t)$ in $\mathcal Z$. The Fock space $\mathcal F$ is spanned by the exponential vectors $E(f)$, whose components in the $0,1,\ldots,k,\ldots$ particle spaces are \begin{equation}\label{1.5} E(f) = \left(1,f,(2!)^{-1/2}f\otimes f,\ldots,(k!)^{-1/2} f^{ \otimes k}, \ldots \right), \end{equation} $f \in \mathcal X$; the inner product between two exponential vectors is given by \begin{eqnarray}\nonumber \langle E(g) | E(f) \rangle &=& \exp\, \langle g | f \rangle \equiv \exp \biggl[ \int_{-\infty}^{+\infty} \langle g(t)|f(t) \rangle \,{\mathrm d} t \biggr] \\ \label{1.6} &\equiv& \exp \biggl[ \sum_j \int_{-\infty}^{+\infty} \overline {g_j(t)}\,f_j(t) \,{\mathrm d} t \biggr], \end{eqnarray} where an overline means complex conjugation, and we get normalized vectors by defining \begin{equation}\label{1.7} e(f)= \exp\left(-\textstyle\frac{1}{2}\|f\|^2\right) E(f)\,. \end{equation} The annihilation, creation and gauge (or number) processes are defined by \begin{eqnarray}\nonumber A_j(t) E(f) &=& \int_0^t f_j(s)\,{\mathrm d} s\, E(f)\,, \\ \label{1.8} \langle E(g)| A^\dagger_j(t)E(f)\rangle &=& \int_0^t \overline{g_j(s)}\, {\mathrm d} s \, \langle E(g)|E(f)\rangle\,, \\ \nonumber \langle E(g)|\Lambda_{ij}(t)E(f)\rangle &=& \int_0^t \overline{g_i(s)}f_j(s)\, {\mathrm d} s \, \langle E(g)|E(f)\rangle\,. \end{eqnarray} Eqs.~(\ref{1.8}) allow to write formally \begin{eqnarray}\nonumber A_j(t) &&{}= \int_0^t a_j(s) \,{\mathrm d} s\,, \\ \label{1.4} A_j^\dagger(t) &&{}= \int_0^t a_j^\dagger(s) \,{\mathrm d} s\,, \\ \nonumber \Lambda_{ij}(t) &&{}= \int_0^t a_i^\dagger(s)a_j(s) \,{\mathrm d} s\,, \end{eqnarray} where $a_j(t)$, $a_j^\dagger(t)$ are usual Bose fields, satisfying the canonical commutation rules \begin{equation}\label{1.9} [a_j(t),a_i(s)] = 0\, , \qquad [a_j(t),a_i^{\dagger} (s)] = \delta_{ji}\, \delta(t - s)\,, \end{equation} and whose coherent vectors are the normalized exponential vectors: \begin{equation}\label{1.10} a_j(t)\,e(f) = f_j(t)\, e(f)\,. \end{equation} In particular the vector $e(0)\equiv E(0)$ is the Fock vacuum. The Bose fields introduced here represent a good approximation of the electromagnetic field in the so called \emph{quasi-monochromatic paraxial approximation} \cite{Yuen,Bar90}. Now, ${\cal F}$ is interpreted as the Hilbert space of the electromagnetic field; $A^\dagger_j(t)$ creates a photon with state $e_j$ in the time interval $[0,t]$, $A_j(t)$ annihilates it, $\Lambda_{jj}(t)$ is the self\-ad\-joint operator representing the number of photons with state $e_j$ in the time interval $[0,t]$ and \begin{equation}\label{1.1} N(t)= \sum_j \Lambda_{jj}(t) \end{equation} is the observable ``total number of photons entering the system up to time $t$''. Moreover, in the approximation we are considering, the fields behave as monodimensional waves, so that a change of position is equivalent to a change of time and viceversa. If we forget polarization, the one-particle space ${\cal Z}$ has to contain only the degrees of freedom linked to the direction of propagation \cite{Bar-direct}, so that we can take \begin{eqnarray}\nonumber {\cal Z} &=& L^2 \big(\Upsilon,\,\sin \theta\, {\mathrm d} \theta\,{\mathrm d} \phi \big)\,, \\ \Upsilon &=& \{0\leq \theta \leq \pi, \ 0 \leq \phi < 2\pi\} \,;\label{3.22} \end{eqnarray} the angular coordinates $(\theta,\phi)$ represent the direction of propagation. Now, a vector $f$ in the one-particle space $\cal X$ can be identified with a function $f(\theta, \phi, t)$ such that $\int_0^{+\infty} {\mathrm d} t \int_0^\pi {\mathrm d} \theta \, \sin \theta \int_0^{2\pi} {\mathrm d} \phi \left| f(\theta, \phi, t)\right|^2 < +\infty$. In QSC integrals of ``Ito type'' with respect to ${\mathrm d} A_j(t)$, ${\mathrm d} A_j^\dagger(t)$, ${\mathrm d} \Lambda_{ij}(t)$ are defined. The main practical rules to manipulate ``Ito differentials'' are the facts that ${\mathrm d} A_j(t)$, ${\mathrm d} A_j^\dagger(t)$, ${\mathrm d} \Lambda_{ij}(t)$ commute with anything contain the fields only up to time $t$ and that the products of the fundamental differentials satisfy \begin{eqnarray}\nonumber &&{\mathrm d} A_j(t)\, {\mathrm d} A_i^{\dagger}(t) = \delta_{ji}\, {\mathrm d} t\,, \\ \nonumber &&{\mathrm d} A_j(t)\, {\mathrm d} \Lambda_{ki}(t) = \delta_{jk}\, {\mathrm d} A_i(t)\,, \\ \label{1.11} &&{\mathrm d}\Lambda_{ji}(t)\, {\mathrm d} A_k^\dagger(t) = \delta_{ik}\, {\mathrm d} A_j^\dagger(t)\,, \\ \nonumber &&{\mathrm d} \Lambda_{ji}(t)\, {\mathrm d} \Lambda_{lk}(t) = \delta_{il}\, {\mathrm d} \Lambda_{jk}(t)\,, \\ \nonumber &&{\mathrm d} A_i^\dagger(t)\, {\mathrm d} A_j(t) = {\mathrm d}\Lambda_{ki}(t)\, {\mathrm d} A_j(t) = {\mathrm d} A_k^\dagger(t)\, {\mathrm d} \Lambda_{ji}(t) = 0\,; \end{eqnarray} all the products of ${\mathrm d} A_j(t)$, ${\mathrm d} A_j^\dagger(t)$ or ${\mathrm d} \Lambda_{ij}(t)$ with ${\mathrm d} t$ vanish. \subsubsection*{The evolution equation} Let ${\cal H}$ be a separable complex Hilbert space (the system space) and let $R_i$, $i\geq 1$, $S_{ij}$, $i,j\geq 1$, $H$ be bounded operators in ${\cal H}$ such that $H^\dagger =H$, $\sum_i R_i^{\dagger } R_i$ is strongly convergent to a bounded operator, and $\sum_{i,j} S_{ij} \otimes |e_i \rangle \langle e_j|= S\in {\cal U}({\cal H}\otimes {\cal Z})$ (unitary operators in ${\cal H} \otimes {\cal Z}$); we set also \begin{equation}\label{1.12} K=H - \frac{{\mathrm i}}{2} \sum_j R_j^{\dagger }R_j\,. \end{equation} Then $(\!\!{}$\cite{Partha92} Theor.\ 27.8 p.\ 228) there exists a unique unitary operator-valued adapted process $U(t)$ satisfying $U(0) =\openone$ and \begin{eqnarray} {\mathrm d} U(t) = {}&&\bigg\{ \sum_j R_j \,{\mathrm d} A_j^\dagger(t) + \sum_{i,j} \left(S_{ij}- \delta_{ij}\right) {\mathrm d} \Lambda_{ij}(t) \nonumber \\ \label{1.2} &&{}- \sum_{i,j} R_i^{\dagger } S_{ij }\,{\mathrm d} A_j(t) -{\mathrm i} K\,{\mathrm d} t \bigg\} \, U(t)\,. \end{eqnarray} The operator $U_t$ will be the evolution operator for the atom-field system, in the interaction picture with respect to the free dynamics of the field. In order to describe a two-level atom, we take ${\cal H}= {\mathbb C}^2$; then, to fix the model, we have to determine the atomic operators $H$, $R_i$, $S_{ij}$ on the basis of physical considerations. In the next section we shall require: (a) the existence of a ground state to which the atom decays by emitting at most one photon when it is not stimulated, (b) a balance equation [Eq.\ (\ref{2.15})] between the numbers of ingoing and outgoing photons when there is some coherent source. This suffices to determine the structure of the atomic operators [Eqs.\ (\ref{2.11}), (\ref{2.16})]. \subsubsection*{Contents} As said before, in Section \ref{model} we fix the model by physical considerations; here, a central role is played by a balance equation saying that the mean number of outgoing photons plus the mean number of photons stored in the atom is equal to the mean number of ingoing photons. In Section \ref{master} we consider the case of a spherically symmetric atom stimulated by monochromatic coherent light, we obtain the master equation which gives the time evolution of the reduced atomic density matrix and we study the large-time behavior of its solutions. In Section \ref{direct} we study the differential (with respect to the angle) and total cross sections for the scattering of laser light by the atom, as a function of the frequency of the stimulating laser; in this section such cross sections are obtained from the direct detection scheme. In Section \ref{heter}, starting from the balanced heterodyne detection scheme, we obtain the power spectrum of the fluorescence light and the elastic and inelastic cross sections. Section \ref{Dis} is devoted to a discussion of the main features of the integral cross sections and of the power spectrum. \section{The model and the balance equation for the number of photons} \label{model} First of all we want a model for an atom stimulated by a laser (coherent light, not necessarily monochromatic); this means to choose as initial state $\Psi(\xi, f)\in {\cal H}\otimes {\cal F}$ a generic state for the atom and a coherent vector for the field \cite{Bar90}, i.e. \begin{eqnarray}\label{2.1} \Psi(\xi,f) &=& \xi \otimes e(f)\,, \\ \nonumber \xi \in {\cal H}\,, \quad \|\xi\|&=&1\,, \quad f\in L^2({\mathbb R}_+;{\cal Z})\,. \end{eqnarray} Then, the atomic reduced statistical operator $\rho(t;\xi,f)$ is defined by the partial trace over the Fock space \begin{equation}\label{2.2} \rho(t;\xi,f) = {\mathrm Tr}_{{}_{\cal F}} \left\{ U(t) |\Psi(\xi,f) \rangle \langle \Psi(\xi,f)| U(t)^\dagger \right\}. \end{equation} Moreover, the quantity \begin{equation}\label{2.3} \langle N(t)\rangle_f = \langle U(t) \Psi(\xi,f) | N(t)\, U(t) \Psi(\xi,f) \rangle \end{equation} represents the mean number of photons up to time $t$, after the interaction with the atom, while \begin{equation}\label{2.4} \langle N(t)\rangle_f^0 =\langle \Psi(\xi,f) | N(t)\, \Psi(\xi,f)\rangle = \int_0^t \|f(s)\|^2 {\mathrm d} s \end{equation} is the same quantity before such an interaction \cite{Bar90}; we can also say that Eq.\ (\ref{2.4}) gives the mean number of ingoing photons entering the system in the time interval $[0,t]$ and that Eq.\ (\ref{2.3}) gives the mean number of outgoing photons leaving the system in the same time interval. \end{multicols} \vspace{-0.4cm} \noindent\rule{0.49\textwidth}{0.4pt}\rule{0.4pt}{\baselineskip} \widetext \begin{proposition}\label{prop1} The reduced statistical operator $\rho(t;\xi,f) $ satisfies the master equation \begin{equation}\label{2.5} \frac{{\mathrm d}\ }{ {\mathrm d} t} \, \rho(t;\xi,f) = {\cal L}\big(f(t)\big) [\rho(t;\xi,f)]\,, \end{equation} where \begin{mathletters}\label{2.6} \begin{equation}\label{2.6a} {\cal L}\big(f(t)\big)[\rho] = -{\mathrm i} \left[H\big(f(t)\big)\,,\, \rho\right] + \frac 1 2 \sum_j \left( \left[ R_j \big(f(t)\big)\rho\,,\, R_j\big(f(t)\big)^\dagger \right] +\left[ R_j\big(f(t)\big)\,,\, \rho R_j\big(f(t)\big)^\dagger \right]\right), \end{equation} \begin{equation}\label{2.6b} R_j\big(f(t)\big) = R_j +\sum_i S_{ji} f_i(t)\,, \end{equation} \begin{equation} \label{2.6c} H\big(f(t)\big) = H - \frac {\mathrm i} 2 \sum_{ij} \Bigl( R^\dagger_j S_{ji} f_i(t) - \overline{f_i(t)} \, S_{ji}^{\ \dagger} R_j \Bigr); \end{equation} \end{mathletters} moreover, we have \begin{equation} \label{2.7} \langle N(t) \rangle_f = \int_0^t {\mathrm Tr}_{{}_{\cal H}} \Bigl\{ \sum_j R_j\big(f(s)\big)^{\dagger } R_j\big(f(s)\big) \rho(s;\xi,f) \Bigr\} {\mathrm d} s\,. \end{equation} \end{proposition} \noindent{}\hfill{ \rule[-\baselineskip]{0.4pt}{\baselineskip}\rule{0.49\textwidth}{0.4pt}} \vspace{-0.4cm} \begin{multicols}{2} \narrowtext {\smallskip\bf \noindent Proof. } By using the rules of QSC, it is possible to differentiate $\langle N(t) \rangle_f$ and $\left\langle U(t) \Psi(\xi,f) | \,a \,U(t) \Psi(\xi,f) \right\rangle$, where $a$ is a generic system operator. Then, one gets the results by recalling that the increments of the field operators commute with $U(t)$ and that ${\mathrm d} A_j(t) \Psi(\xi ,f)= f_j(t){\mathrm d} t\, \Psi(\xi,f)$ and by using the definition of $\rho(t;\xi,f)$ given in Eq.\ (\ref{2.2}). {\hfill{$\square$} \vskip4pt} In order to formulate physical requirements, let us start by considering the case when no photon is injected into the system, i.e. $f=0$. In these conditions it is natural to ask that the atom can emit at most one photon; moreover, we require the existence of a unique equilibrium state which we denote by $\rho_g$ (it will be the ground state). We take as canonical basis $\{|+\rangle,\,|-\rangle\}$ in ${\cal H}$ an orthonormal basis which diagonalises $\rho_g$, so that we can write \begin{equation}\label{2.8} \rho_g= pP_+ +(1-p)P_- \end{equation} for some $p$ in $[0,1]$, where $P_\pm$ are the orthogonal projections over the vectors $|\pm\rangle$. We shall use also the Pauli matrices \begin{mathletters}\label{2.9} \begin{eqnarray}\label{2.9a} \sigma_z = \left(\begin{array}{cc} 1& 0 \\ 0 &-1 \end{array}\right)&&, \qquad \sigma_y = \left(\begin{array}{cc} 0& - {\mathrm i} \\ {\mathrm i} & 0 \end{array}\right), \\ \label{2.9b} \sigma_+ = \left(\begin{array}{cc} 0& 1 \\ 0 &0 \end{array}\right)&&, \qquad \sigma_- = \left(\begin{array}{cc} 0& 0 \\ 1 &0 \end{array}\right), \end{eqnarray} by which the two orthogonal projections $P_\pm$ can be written as \begin{equation}\label{2.9c} P_+ = \frac 1 2 (\openone + \sigma_z) = \sigma_+ \sigma_-\,, \quad P_- = \frac 1 2 (\openone - \sigma_z) = \sigma_- \sigma_+\,. \end{equation} \end{mathletters} \begin{proposition} We require \begin{mathletters}\label{2.10} \begin{eqnarray}\label{2.10a} && \langle N(t)\rangle_{f=0} \leq 1\,, \qquad \forall \xi,\ \forall t, \\ \label{2.10b} && \rho(t;\xi,0) \stackrel{t \to +\infty}{ \longrightarrow } \rho_g\,, \qquad \forall \xi. \end{eqnarray} \end{mathletters} \noindent Then, apart from an exchange of roles between the two states $|+\rangle$ and $|-\rangle$, we obtain \begin{equation}\label{2.11a} \rho_g= P_-\,, \end{equation} \begin{mathletters}\label{2.11} \begin{equation}\label{2.11b} H= \frac 1 2 \, \omega_0 \sigma_z\,, \qquad \omega_0 \in {\mathbb R}\,, \end{equation} \begin{equation}\label{2.11c} R_j = \langle e_j|\alpha \rangle\, \sigma_-\,, \qquad \alpha \in {\cal Z} \,, \ \alpha\neq 0\,. \end{equation} \end{mathletters} \noindent Viceversa, Eqs.\ (\ref{2.11b}) and (\ref{2.11c}) imply Eqs.\ (\ref{2.10}) and (\ref{2.11a}). \end{proposition} {\smallskip\bf \noindent Proof. } By Eqs.\ (\ref{2.7}) and (\ref{2.10a}), $\langle N(t)\rangle_f$ is a bounded and non decreasing function of $t$, so $\lim_{t\to +\infty}\langle N(t)\rangle_f$ exists; then, Eqs.\ (\ref{2.7}) and (\ref{2.10b}) give $\sum_j{\mathrm Tr}_{{}_{\cal H}} \left\{ R_j^{\dagger } R_j \rho_g \right\}=0$. By the cyclic property of the trace and the positivity of $\rho_g$ and of $R_j \rho_g R_j^{\dagger }$, we get that this condition is equivalent to $ R_j\, \rho_g= 0 $, $\forall j$. Now, let us set $R_j = x_j\openone +y_j \sigma_z + z_j \sigma_+ + \alpha_j \sigma_-$ (every operator on ${\mathbb C}^2$ can be written in this way). Then, Eq.\ (\ref{2.8}) and $R_j\rho_g=0$ give $p(x_j+y_j)=0$, $(1-p)(x_j-y_j)=0$, $(1-p)z_j=0$, $p\alpha_j=0$. For $p\in (0,1)$ this system of equations gives $R_j=0$, which is not acceptable because in this case the equilibrium state is not unique. For $p=0$ we get $x_j=y_j$ and $z_j=0$; we need also \begin{equation}\label{2.12} \sum_j|\alpha_j|^2\neq 0 \end{equation} to have decay to an equilibrium state. We do not consider the case $p=1$, because it is analogous to the previous one, apart from the exchange of $|+\rangle$ and $|-\rangle$. Therefore we have Eq.\ (\ref{2.11a}) and \begin{equation}\label{2.13} R_j = \alpha_j \sigma_- + \beta_j P_+\,, \end{equation} with $\beta_j = 2 x_j$; by the convergence of $\sum_j R_j^\dagger R_j$, the complex numbers $\alpha_j$ and $\beta_j$ can be seen as the components of two vectors $\alpha$ and $\beta$ in $\cal Z$. Eq.\ (\ref{2.5}) and (\ref{2.10b}) give ${\cal L}(0)[\rho_g]=0$; by Eqs.\ (\ref{2.6}), (\ref{2.11a}) and (\ref{2.13}) this condition reduces to $[H,\rho_g]=0$. Because $H$ is selfadjoint and defined up to a constant, we obtain Eq.\ (\ref{2.11b}). Finally, let us choose $\xi=|+\rangle$. By using the relation $\sum_j R_j^\dagger R_j = \left( \|\alpha\|^2 + \|\beta\|^2 \right) P_+$ and by differentiating $\langle N(t) \rangle_{f=0}$ two times, we obtain \[ \frac{{\mathrm d}^2\ }{{\mathrm d}t^2}\, \langle N(t)\rangle_{f=0} + \|\alpha\|^2 \, \frac{{\mathrm d}\ }{{\mathrm d}t}\, \langle N(t) \rangle_{f=0} =0\,, \] together with the initial conditions \[ \langle N(0)\rangle_{f=0}=0\,, \] \[ \frac{{\mathrm d}\ }{{\mathrm d}t}\,\langle N(0)\rangle_{f=0} =\|\alpha\|^2 + \|\beta\|^2\,. \] This gives \[ \langle N(t)\rangle_{f=0} = \left( 1+ \frac{\|\beta\|^2}{\|\alpha\|^2} \right) \left(1- {\mathrm e}^{- \|\alpha\|^2t}\right); \] then, condition (\ref{2.10a}) implies $\beta=0$ and Eqs.\ (\ref{2.12}) and (\ref{2.13}) give Eq.\ (\ref{2.11c}). The last statement of the proposition follows by direct computations. {\hfill{$\square$} \vskip4pt} Now we have to find some physical restrictions on the possible forms of the operator $S\in {\mathcal U}( {\mathcal H}\otimes {\mathcal Z})$. In \cite{BarL98}, the case $f(t)= \lambda(t) \equiv \exp(-{\mathrm i}\omega t)\theta(T-t)\, \lambda$ is considered, where $\theta(x)$ is the usual step function and $\lambda \in {\mathcal Z}$; for $T\to +\infty$, $\lambda(t)$ represents a monochromatic coherent wave. Then, in \cite{BarL98} we asked \begin{equation} \lim_{t\to +\infty} \lim_{T\to +\infty} \frac {\langle N(t) \rangle_\lambda} {\langle N(t) \rangle_\lambda^0} =1\,, \qquad \forall \lambda\in {\cal Z}\,, \quad \forall \omega\,, \label{2.14} \end{equation} which is a form of flux conservation in the mean: if the possible physical processes are absorption/emission of single photons and direct scattering without change of atomic state, for large times the mean number of injected photons $\langle N(t)\rangle_\lambda^0= \|\lambda\| t$ should be equal to the mean number of outgoing photons $\langle N(t)\rangle_\lambda$. The same restrictions on $S$ are obtained by requiring a balance equation on the number of photons: the mean number of outgoing photons up to time $t$ plus the mean number of photons stored in the atom must be equal to the mean number of ingoing photons. \begin{proposition} Under assumptions (\ref{2.11}), the balance equation \begin{eqnarray}\nonumber \langle N(t)\rangle_f &+&\frac 1 2\, {\mathrm Tr}_{{}_{\cal H}} \left\{ \sigma_z \bigl[ \rho(t;\xi,f) - \rho(0;\xi,f) \bigr] \right\} \\ &=& \langle N(t)\rangle_f^0 \label{2.15} \end{eqnarray} holds $\forall t$, $\forall \xi$, $\forall f$ if and only if one has \begin{equation}\label{2.16} S = P_+\otimes S^+ + P_-\otimes S^-\,, \qquad S^\pm \in {\cal U} ({\cal Z})\,. \end{equation} \end{proposition} {\smallskip\bf \noindent Proof. } Any bounded operator on $\mathcal H \otimes Z$, like $S$, can always be decomposed as \begin{equation}\label{2.17} S=P_+\otimes S^+ + P_-\otimes S^- + \sigma_+ \otimes F^+ +\sigma_- \otimes F^-\,, \end{equation} where $ S^\pm,\, F^\pm$ are bounded linear operators on $\cal Z$; the unitarity of $S$ implies some simple relations among $S^\pm$, $F^\pm$. By using Eqs.\ (\ref{2.5}), (\ref{2.6}), (\ref{2.11}), we compute the time derivative of ${\mathrm Tr}_{{}_{\cal H}} \{ \sigma_z \rho(t;\xi,f) \}$. Then, we insert Eq.\ (\ref{2.6b}) into Eq.\ (\ref{2.7}) and, by using also Eq.\ (\ref{2.17}), we get \begin{eqnarray*} \langle N(t)\rangle_f &-& \langle N(t)\rangle_f^0 \\ &&{}+\frac 1 2\, {\mathrm Tr}_{{}_{\cal H}} \left\{ \sigma_z \bigl[ \rho(t;\xi,f) - \rho(0;\xi,f) \bigr] \right\} \\ &=&\int_0^t {\mathrm d}s \, {\mathrm Tr}_{{}_{\cal H}} \Bigl\{ \Bigl[ \left\| F^+ f(s) \right\|^2 P_- - \left\| F^- f(s) \right\|^2 P_+ \\ &&{}+ \left\langle S^+ f(s) \big| F^+ f(s) \right \rangle \sigma_+ \\ && {}+ \left\langle F^+ f(s) \big| S^+ f(s) \right \rangle \sigma_- \Bigr] \rho(s;\xi,f) \Bigr\}. \end{eqnarray*} By the arbitrariness of $t$, $f$ and $\xi$, condition (\ref{2.15}) is equivalent to $F^\pm=0$ and Eq.\ (\ref{2.16}) is proved; the unitarity of $S^\pm$ follows from the unitarity of $S$. {\hfill{$\square$} \vskip4pt} From now on we assume Eqs.\ (\ref{2.1}), (\ref{2.11}), (\ref{2.16}) to hold and, always for physical reasons, we take \begin{equation}\label{2.20} \omega_0 >0\,. \end{equation} In order to have an atom stimulated by a monochromatic coherent wave we take \begin{equation}\label{3.1} f(t)= \lambda(t) \equiv {\mathrm e}^{-{\mathrm i}\omega t}\theta(T-t)\, \lambda\,, \quad \lambda \in {\mathcal Z}\,, \quad \omega>0\,. \end{equation} The step function $\theta$ is defined by $\theta(x)=1$ for $x\geq 0$ and $\theta(x)=0$ for $x<0$, so that $\lambda(t)$ represents a monochromatic wave for $T\to +\infty$. Quantities like $\omega_0$, $\alpha$, $S^\pm$ are phenomenological parameters, or, better, they have to be computed from some more fundamental theory, such as some approximation to quantum electrodynamics. The whole model is meaningful only for $\omega$ not too ``far'' from $\omega_0$ and $\omega_0$ must include the Lamb shifts. In the final results one can admit a slight $\omega$-dependence in the direct scattering matrices $S^\pm$. \section{The master equation}\label{master} In this section we study the master equation (\ref{2.5}) and the long time behavior of the atom; the relations (\ref{2.6}), (\ref{2.11}), (\ref{2.16}), (\ref{2.20}), (\ref{3.1}) hold. \subsubsection*{The reduced statistical operator} First of all, by setting \begin{eqnarray}\nonumber \rho_\lambda(t) &=& \lim_{T\to +\infty} \exp \left\{ {\textstyle \frac{\mathrm i}{2}}\, \sigma_z (\beta+\omega t)\right\} \rho(t; \xi,\lambda) \\ \label{3.2} &&{}\times \exp \left\{ - {\textstyle\frac{\mathrm i}{2}}\, \sigma_z (\beta+\omega t)\right\}, \\ \nonumber \beta &=& \arg \left\{ - \langle S^- \lambda |\alpha\rangle\right\}, \end{eqnarray} exactly as in Proposition \ref{prop1} we obtain the master equation \begin{equation}\label{3.3} \frac{{\mathrm d} \ }{{\mathrm d} t}\, \rho_\lambda(t) = {\cal L}_\lambda \left[\rho_\lambda(t)\right] \end{equation} with the time independent Liouvillian \begin{mathletters}\label{3.4} \begin{eqnarray}\nonumber {\cal L}_\lambda[\rho] &=& -{\mathrm i} [H_\lambda\,, \rho] +\frac 1 2 \sum_j \left(\left[ R^\lambda_j \rho\,,\, R^{\lambda \dagger }_j \right]\right. \\ \label{3.4a} &&{}+ \left.\left[ R^\lambda_j\,,\, \rho R^{\lambda \dagger }_j \right]\right), \end{eqnarray} \begin{eqnarray}\nonumber R_j^\lambda &=& {\mathrm e}^{- {\mathrm i}\beta}\langle e_j | \alpha \rangle \sigma_- + \langle e_j |S^+ \lambda \rangle P_+ \\ &&{}+\langle e_j |S^- \lambda \rangle P_-\,, \label{3.4b} \end{eqnarray} \begin{equation}\label{3.4c} H_\lambda = \frac 12\, (\omega_0 - \omega)\sigma_z - \frac {1}{ 2}\, |\langle \alpha | S^- \lambda \rangle |\sigma_y\,. \end{equation} \end{mathletters} The general master equation for a two-level system is studied in \cite{Lendi}; in the following we shall use similar techniques, apart from a different parametrization of the statistical operator, which turns out to be more convenient in our case. By setting \begin{mathletters}\label{3.5} \begin{equation}\label{3.5a} \rho_\lambda(t) = \left( \begin{array}{cc} u(t) & v(t) \\ & \\ \overline{v(t)} & 1-u(t) \end{array} \right), \end{equation} \begin{equation} \label{3.5b} \left\{\begin{array}{l} 0\leq u(t) \leq 1\,, \\ \\ u(t) \geq u^2(t) + | v(t)|^2 \,,\end{array}\right. \end{equation} \end{mathletters} \noindent where the conditions (\ref{3.5b}) express the fact that $\rho_\lambda (t)$ is a statistical operator, we obtain from the master equation \begin{equation}\label{3.6} \frac{{\mathrm d} \ }{{\mathrm d} t}\, \bbox{u}(t) = - \bbox{G\, u }(t) + \left( \begin{array}{c} 0\\ \Omega /2 \\ \Omega /2 \end{array}\right), \end{equation} where \begin{mathletters} \begin{equation}\label{3.7b} \bbox{u}(t) = \left( \begin{array}{c} u(t)\\ v(t)\\ \overline{v(t)} \end{array} \right), \end{equation} \begin{equation}\label{3.7c} \bbox{G}= \left( \begin{array}{ccc} \|\alpha\|^2 & -\Omega /2 & -\Omega /2 \\ & & \\ -{\mathrm e}^{{\mathrm i}\beta}\left \langle \alpha | \Delta S \lambda \right \rangle +\Omega & b & 0 \\ & & \\ -{\mathrm e}^{-{\mathrm i}\beta}\left \langle \Delta S \lambda | \alpha \right \rangle +\Omega & 0 & \overline b \end{array} \right), \end{equation} \begin{eqnarray}\label{3.7a} &&\Omega = 2\left|\left\langle \alpha | S^- \lambda \right \rangle \right| \\ \label{3.7d} && b= \frac { \kappa^2} 2\, \|\alpha\|^2 - {\mathrm i} \left( \Delta \omega - {\mathrm Im} \left \langle S^+ \lambda | P_\alpha S^- \lambda \right\rangle \right), \\ \label{3.7e} &&\kappa^2 = 1 + \|\Delta S \lambda\|^2\big/ \|\alpha\|^2 \\ \label{3.7g} && \Delta S = S^+ -S^-\,, \\ \label{3.7h} &&\Delta \omega = \omega - \left( \omega_0 + {\mathrm Im} \left\langle S^+ \lambda | P_\bot S^- \lambda\right\rangle \right), \\ \label{3.7i} && \widetilde \alpha = \frac \alpha{\|\alpha\|}\,, \qquad P_\alpha = \left| \widetilde \alpha \right\rangle \left \langle \widetilde \alpha \right|, \qquad P_\bot = \openone - P_\alpha\,. \end{eqnarray} \end{mathletters} \noindent The quantity $\Omega$ can be interpreted as the bare Rabi frequency. Moreover, we have \begin{equation}\label{3.8} {\rm det}\, \bbox{G} = \|\alpha\|^2 \left[ \left( \Delta \omega \right)^2 + \Gamma^2/4 \right], \end{equation} with \begin{eqnarray}\nonumber \Gamma^2 &=&\kappa^4 \|\alpha\|^4+4 \kappa^2 \|\alpha\|^2\, {\mathrm Re} \left\langle S^- \lambda | P_\alpha \left( S^+ + S^- \right) \lambda \right \rangle \\ \nonumber &&{}- 4 \left( {\mathrm Im} \left\langle S^+ \lambda | P_\alpha S^- \lambda \right \rangle \right)^2 \\ \nonumber & \equiv & \Bigl( \|\alpha\|^2 + \big\|P_\bot \Delta S \lambda\big\|^2+ 2 \left| \left\langle \widetilde \alpha | S^- \lambda \right \rangle \right |^2 \\ \nonumber &&{}- 2\, {\mathrm Re} \left \langle S^+ \lambda | P_\alpha S^- \lambda \right \rangle \Bigr)^2 + \left| \left \langle \widetilde \alpha \big| \left( S^+ + S^- \right) \lambda \right\rangle \right|^2 \\ &&{}\times \left[ \|\alpha\|^2\left(1+\kappa^2\right) + \big\|P_\bot \Delta S \lambda\big\|^2 \right]. \label{3.9} \end{eqnarray} Let us note that $\|\alpha\|>0$ implies ${\rm det}\, \bbox{G} > 0$ and $\Gamma^2 >0$. \subsubsection*{The equilibrium state and the general solution of the master equation} The equilibrium state is given by \begin{equation}\label{3.10} \lim_{t\to +\infty} \rho_\lambda(t) = \rho^\lambda_{\rm eq} = \left( \begin{array}{cc} u(\infty) & v(\infty) \\ {}&{}\\ \overline{v(\infty)} & 1-u(\infty) \end{array} \right), \end{equation} where $u(\infty)$ and $v(\infty)$ are computed by equating to zero the time derivative in Eq.\ (\ref{3.6}); then, we have $\bbox{u}(\infty)= \bbox{G}^{-1} \bbox{w}$, which gives \begin{mathletters}\label{3.11} \begin{equation}\label{3.11a} u(\infty) = \frac {\kappa^2 \Omega^2/4 }{ (\Delta \omega)^2 + \Gamma^2/4}\,, \end{equation} \begin{eqnarray}\label{3.11b} v(\infty) &=& \frac {\Omega/2} {(\Delta \omega)^2 + \Gamma^2/4} \\ \nonumber &\times& \left( \frac{\kappa^2} 2 \, \|\alpha\|^2+ {\mathrm i} \Delta \omega +{\mathrm i} \, {\mathrm Im}\, \langle S^+ \lambda | P_\alpha S^- \lambda \rangle \right). \end{eqnarray} \end{mathletters} For the computation of the fluorescence spectrum in Section \ref{heter}, we shall have to solve the master equation (\ref{3.3}) also when the initial condition is not a statistical operator. If \[ \sigma = \left( \begin{array}{cc} \sigma_{11} & \sigma_{12} \\ \sigma_{21} & \sigma_{22} \end{array} \right) \] is a generic $2\times 2$ matrix, we can always write \begin{equation}\label{3.27} {\mathrm e}^{{\mathcal L}_\lambda t}[\sigma] = (\sigma_{11} + \sigma_{22}) \rho_{\mathrm eq}^\lambda + \left( \begin{array}{cc} d_1(t) & d_2(t) \\ d_3(t) & - d_1(t) \end{array} \right), \end{equation} where \begin{equation}\label{3.28} \bbox{d}(t) = {\mathrm e}^{-\bbox{G}t} \bbox{d}(0)\,, \qquad \bbox{d}(t) \equiv \left( \begin{array}{c} d_1(t) \\ d_2(t) \\ d_3(t) \end{array} \right), \end{equation} \begin{eqnarray} d_1(0)&=& \sigma_{11} - (\sigma_{11}+ \sigma_{22}) u(\infty)\,, \nonumber \\ \label{3.29} d_2(0)&=& \sigma_{12} - (\sigma_{11}+ \sigma_{22}) v(\infty)\,, \\ \nonumber d_3(0)&=& \sigma_{21} - (\sigma_{11}+ \sigma_{22}) \overline{ v(\infty)}\,. \end{eqnarray} \subsubsection*{Spherically symmetric atom stimulated by a collimated laser} We end this section by particularizing our model to the case of a \emph{spherically symmetric atom} stimulated by a \emph{well collimated laser}. Let us recall that the Hilbert space $\mathcal Z$ contains the directions of propagation of the electromagnetic field [see Eq.~(\ref{3.22})]. So, in order to describe a laser beam propagating along the direction $\theta=0$, we have to take \begin{mathletters}\label{3.30} \begin{equation}\label{3.30a} \lambda =\eta \|\alpha\|\,{\mathrm e}^{{\mathrm i} \delta} \widetilde \lambda\,, \qquad \eta>0\,, \qquad \delta\in [0,2\pi) \,, \end{equation} \begin{equation}\label{3.30b} \widetilde \lambda (\theta,\phi) = \frac{ 1_{[0,\Delta \theta]}(\theta)} {\Delta \theta \sqrt{2\pi (1-\cos \Delta \theta)}} \,, \end{equation} \end{mathletters} \noindent where $1_{[0,\Delta \theta]}(\theta)=1$ for $0\leq \theta \leq \Delta \theta$, $1_{[0,\Delta \theta]}(\theta)=0$ elsewhere; in all the physical quantities the limit $\Delta \theta \downarrow 0$ will be taken. Note that the power of the laser $\hbar \omega\|\lambda\|^2=\hbar\omega \|\alpha\|^2 \eta^2/ {(\Delta\theta)^2}$ diverges for $\Delta \theta \downarrow 0$, because we need a not vanishing atom-field interaction in the limit. Let us denote by $Y_{lm}(\theta,\phi)$ the spherical harmonic functions; then, the spherical symmetry of the atom requires \begin{equation}\label{3.31} \widetilde \alpha(\theta,\phi) = Y_{00}(\theta,\phi) =1/\sqrt{4\pi}\,, \end{equation} \begin{equation}\label{3.32} S^\pm = \sum_{lm} {\mathrm e}^{2 {\mathrm i} \delta_l^\pm}\, |Y_{lm} \rangle \langle Y_{lm}|\,, \end{equation} where the quantities $\delta_l^+$ and $\delta_l^-$ are the phase shifts for the direct scattering in the up and down atomic states respectively. Let us note that we have \begin{mathletters}\label{3.33} \begin{equation}\label{3.33a} \lim_{\Delta \theta \downarrow 0} \left\langle Y_{lm}\Big| \widetilde \lambda \right \rangle = \delta_{m,0} \, \frac 1 2 \, \sqrt{2l+1}\,, \end{equation} \begin{equation} \label{3.33b} \lim_{\Delta \theta \downarrow 0} \left\langle Y_{lm}\Big| \left(S^\pm - \openone \right) \widetilde \lambda \right \rangle = \delta_{m,0}\, {\mathrm i} \sqrt{2l+1}\, {\mathrm e}^{{\mathrm i} \delta_l^\pm} \sin \delta_l^\pm\,. \end{equation} \end{mathletters} \noindent Let us recall that $Y_{l0}(\theta,\phi)=\sqrt{\frac{2l+1}{4\pi}}\, P_l(\cos \theta)$, where the functions $P_l(\xi)$ are the Legendre polynomials. Now, we set \begin{mathletters}\label{set} \begin{eqnarray}\nonumber g_\pm(\theta) &&{}= \lim_{\Delta\theta \downarrow 0}\left( \left( S^\pm - \openone \right) \widetilde \lambda \right) (\theta,\phi) \\ \label{3.34} &&{}= {\mathrm i} \sum_{l=0}^\infty \frac{2l+1}{\sqrt{4\pi}}\, {\mathrm e}^{{\mathrm i} \delta^\pm_{l}} \sin \delta_l^\pm \, P_l(\cos \theta)\,, \end{eqnarray} \begin{eqnarray}\label{3.35} &&\Delta g = g_+ - g_-\,,\qquad s = \delta_0^+ - \delta_0^-\,, \\ \label{5.28} &&z= \frac {2\Delta \omega} {\|\alpha\|^2} \,, \qquad y= z - \frac{\eta^2}{2}\, \sin 2 s\,, \\ \label{3.26} &&\varepsilon= -\frac{\|\alpha\|^2} 4 \sum_{l=1}^\infty (2l+1) \sin 2(\delta_l^+ - \delta_l^-)\,, \\ \nonumber &&\zeta^2 = \left( 1+ \eta^2 \left\| P_\bot \Delta g \right\|^2 \right)^2 \\ &&\quad \ {}+\eta^2 \left( 1+ \kappa^2 + \eta^2 \left\| P_\bot \Delta g \right\|^2 \right), \label{3.36m} \\ \label{set5} &&b^\prime= \kappa^2 - {\mathrm i} \left( z + \frac{\eta^2} 2\, \sin 2s \right), \\ \label{set6} &&\bbox{G}^\prime= \left( \begin{array}{ccc} 2 & - \eta & - \eta \\ & & \\ 2\eta {\mathrm e}^{{\mathrm i}s} \cos s & b^\prime & 0 \\ & & \\ 2\eta {\mathrm e}^{-{\mathrm i}s} \cos s & 0 & \overline{ b^\prime} \end{array} \right). \end{eqnarray} \end{mathletters} \noindent In order that all these quantities be finite, we require also \begin{mathletters}\label{3.37} \begin{eqnarray}\label{3.37a} &&\sum_{l=0}^\infty (2l+1) \sin^2 \delta_l^\pm < +\infty\,, \\ \label{3.37b} &&\sum_{l=0}^\infty (2l+1) \left| \sin 2 \left(\delta_l^+ -\delta_l^-\right) \right|< +\infty\,. \end{eqnarray} \end{mathletters} \noindent Then, we have \begin{mathletters}\label{3.36} \begin{eqnarray} \label{3.36k} &&\beta= \pi - \delta - 2\delta_0^-\,, \\ \label{3.36b} &&b= \frac {\|\alpha\|^2} 2 \, b^\prime\,, \\ \label{3.36a} &&\bbox{G}= \frac{ \|\alpha\|^2 } 2\, \bbox{G}^\prime\,, \\ \label{3.36g} &&\Delta \omega = \omega - \left(\omega_0 + \eta^2 \varepsilon\right), \\ \label{3.36y} && \Omega= \eta \|\alpha\|^2\,, \\ \label{3.36h} &&\Gamma^2 = \zeta^2\|\alpha\|^4\,, \\ \label{3.36c} &&\kappa^2 = 1 + \eta^2\|\Delta g\|^2\,, \\ \label{3.36i} &&u(\infty) = \frac{ \eta^2\kappa^2 }{z^2+\zeta^2}\,, \\ \label{3.36j} &&v(\infty) = \frac{ \eta} {z^2+\zeta^2} \left( \kappa^2 +{\mathrm i} y \right), \\ \label{3.36l} &&\Delta g(\theta) = {\mathrm i}\, \frac{{\mathrm e}^{{\mathrm i }\left(\delta_0^++ \delta_0^-\right)}} {\sqrt{4 \pi}} \, \sin s+\big( P_\bot\Delta g\big)(\theta)\,, \\ \nonumber &&\big( P_\bot\Delta g\big)(\theta)={\mathrm i} \sum_{l=1}^\infty \frac {2l+1}{ \sqrt{4\pi}}\, {\mathrm e}^{{\mathrm i}\left(\delta_l^+ + \delta_l^- \right)} \\ \label{3.36d} &&{}\qquad\qquad\qquad{}\times \sin \left(\delta_l^+ - \delta_l^- \right) P_l(\cos \theta)\,, \\ \label{3.36e} &&\|\Delta g\|^2 = \sin^2 s + \|P_\bot\Delta g\|^2 \\ \label{3.36f} &&\|P_\bot\Delta g\|^2 = \sum_{l=1}^\infty (2l+1) \sin^2\left(\delta_l^+ - \delta_l^- \right), \\ \label{3.36x} &&\|g_\pm\|^2 = \sum_{l=0}^\infty (2l+1) \sin^2 \delta_l^\pm\,, \end{eqnarray} \end{mathletters} \paragraph*{Low intensity laser.} For future use, it is useful to particularize the previous quantities to the case of a laser of vanishing intensity, i.e.\ $\eta=0$: \begin{mathletters}\label{3.23} \begin{eqnarray}\label{3.23a} &&b^\prime= 1 - {\mathrm i} z\,, \qquad \kappa^2=1\,, \qquad \zeta=1\,, \\ \label{3.23b} &&\Omega=0\,, \qquad \Gamma=\|\alpha\|^2\,, \qquad \Delta \omega = \omega - \omega_0\,. \end{eqnarray} \end{mathletters} \paragraph*{No direct scattering.} The usual model of a two-level atom, with only absorption/emission and no direct scattering, is characterized by $S^\pm = \openone$, so that the previous quantities reduce to \begin{mathletters}\label{3.38} \begin{eqnarray}\label{3.38a} && g_\pm = 0\,, \qquad s=0\,, \qquad y=z\,, \\ \label{3.38b} && \beta = \pi - \delta\,, \qquad \kappa^2=1\,, \qquad \zeta= \sqrt{1+2\eta^2} \\ \label{3.38c} && \Gamma^2=\|\alpha\|^4+ 2 \Omega^2\,, \qquad \Delta \omega = \omega - \omega_0\,, \\ \label{3.38d} && u(\infty) = \frac{ \eta^2}{ z^2+\zeta^2}\,, \\ \label{3.38e} && v(\infty) = \frac{ \eta} {z^2+\zeta^2} \left( 1 +{\mathrm i} z \right), \\ \label{3.38f} &&\bbox{G}^\prime= \left( \begin{array}{ccc} 2 & - \eta & - \eta \\ & & \\ 2\eta & 1 - {\mathrm i}z & 0 \\ & & \\ 2 \eta & 0 & 1 + {\mathrm i} z \end{array} \right). \end{eqnarray} \end{mathletters} \section{Direct detection and total cross section}\label{direct} By direct detection, it is possible to measure the intensity of the light (or to count the photons) propagating in a small solid angle $\Delta \Upsilon$ around some direction, which we take different from the direction $\theta=0$ of the incoming beem. The observable ``number of photons in $\Delta\Upsilon$ up to time $t$'' is represented by\begin{equation}\label{4.1} N(t;\Delta\Upsilon) = \sum_{i,j} \langle e_i| 1_{\Delta\Upsilon} \,e_j \rangle\, \Lambda_{ij}(t)\,, \end{equation} where $ 1_{\Delta\Upsilon}(\theta,\phi)=1$ for $(\theta,\phi)\in \Delta \Upsilon$ and $ 1_{\Delta\Upsilon}(\theta,\phi)=0$ elsewhere. The fact that the direction of detection is different from the beam direction is expressed by \begin{equation}\label{4.2} 1_{\Delta\Upsilon}\, \lambda=0\,. \end{equation} Then, the mean number of photons up to time $t$ per unit of solid angle around $(\theta, \phi)$ is given by \begin{equation}\label{4.3} \langle n(\theta,\phi;t)\rangle = \frac{1}{|\Delta\Upsilon|} \langle U(t) \Psi(\xi,\lambda) | N(t;\Delta\Upsilon) U(t) \Psi(\xi,\lambda)\rangle\,, \end{equation} where $\Psi$, $\lambda(t)$, $\lambda$ are given by Eqs.\ (\ref{2.1}), (\ref{3.1}), (\ref{3.30}) and $|\Delta\Upsilon|= \int\!\!\int_{\Delta \Upsilon} \sin \theta \,{\mathrm d}\theta {\mathrm d} \phi$; the limits $T\to +\infty $, $\Delta \theta \downarrow 0$, $\Delta \Upsilon \downarrow \{(\theta, \phi)\}$ are understood. The (angular) differential cross section is proportional to the outgoing flux per unit of solid angle $ \langle n(\theta,\phi;t)\rangle/t$ divided by the incoming flux $\langle \Psi(\xi,\lambda)|N(t) \Psi(\xi,\lambda)\rangle /t$; so we have \begin{eqnarray}\nonumber \sigma(\theta,\phi) &=& A_0 \lim_{t\to +\infty} \frac {\langle n(\theta,\phi;t) \rangle} {\langle \Psi(\xi,\lambda) | N(t) \Psi(\xi,\lambda) \rangle} \\ \label{4.4} &=& \frac {A_0}{ \|\lambda\|^2} \, \lim_{t\to +\infty} \frac 1 t \, \langle n(\theta,\phi;t) \rangle\,, \end{eqnarray} where $A_0$ is a kinematical factor to be determined and with dimensions of an area. To determine $A_0$ let us consider the cross section for direct photon scattering by the up or down atomic state, for which the Bohr-Peierls-Placzek formula (or optical theorem) gives $\sigma(\theta,\phi) = |q(\theta)|^2$, $\sigma_{{}_{\rm TOT}}= 2 \, \frac {2\pi c } \omega\, {\mathrm Im}\, q(0)$; the total cross section is the integral of the differential one on the whole solid angle. In our case we have to take $\alpha=0$ and from Eq.\ (\ref{4.4}) we get $\sigma(\theta,\phi)=A_0\left| \left( \left( S^\pm-\openone \right) \lambda \right) (\theta,\phi)\right|^2/ \|\lambda\|^2$ and, by the unitarity of $S^\pm$, \begin{eqnarray*} \sigma_{{}_{\rm TOT}}&=& \frac{A_0}{\|\lambda\|^2} \left\| \left( S^\pm -\openone \right)\lambda\right\|^2 \\ &=&-\frac{2A_0\sqrt\pi\, \Delta\theta}{\|\lambda\|} \,{\mathrm Im}\, {\mathrm i}{\mathrm e}^{-{\mathrm i}\delta} \left( \left( S^\pm -\openone \right)\lambda \right)(0,0)\,. \end{eqnarray*} Then, we must have $q(\theta)= -{\mathrm i} \sqrt{A_0}\, \Delta \theta \, g_\pm (\theta) $ and, by imposing the optical theorem, we get $A_0= \left(\frac{2\pi c}{\omega}\right)^2\frac1 {\pi(\Delta\theta)^2}$. Up to now we have not taken into account the polarization degrees of freedom. If they are taken into account and the cross section for not polarized light is considered, a $3/2$ extra-factor is obtained $(\!\!{}$\cite{CT92} pp.\ 532--533) and Eq.\ (\ref{4.4}) becomes \begin{equation}\label{4.5} \sigma(\theta,\phi) = \left(\frac{2\pi c}{\omega}\right)^2\frac 3 {2\pi\eta^2\|\alpha\|^2} \, \lim_{t\to +\infty} \frac 1 t \, \langle n(\theta,\phi;t) \rangle\,. \end{equation} To compute $\sigma(\theta,\phi)$ we differentiate Eq.\ (\ref{4.3}) by using the rules of QSC, we use Eq.\ (\ref{4.2}) and then we apply the transformation (\ref{3.2}); the final result is \begin{equation}\label{4.6} \frac{{\mathrm d} \ }{{\mathrm d} t} \left\langle n(\theta,\phi;t)\right\rangle = {\mathrm Tr} \left\{ R(\theta,\phi)^\dagger \,R(\theta,\phi)\, \rho_\lambda (t) \right\}, \end{equation} \begin{equation}\label{4.7} R(\theta,\phi)= {\mathrm e}^{{-\mathrm i}\beta} \, \frac{\|\alpha\|} {\sqrt{4\pi}} \,\sigma_- + {\mathrm e}^{{\mathrm i}\delta}\eta \|\alpha\|\bigl[ g_+(\theta) P_+ + g_-(\theta) P_-\bigr]. \end{equation} Then, Eq.\ (\ref{4.5}) gives \begin{equation}\label{4.8} \sigma(\theta,\phi) = \frac{6\pi c^2}{\eta^2\|\alpha\|^2\omega^2} \, {\mathrm Tr} \left\{ R(\theta,\phi)^\dagger \,R(\theta,\phi)\, \rho^\lambda_{\rm eq} \right\}. \end{equation} Finally, by computing the trace and by using the results of the previous section we obtain the differential cross section and, by integrating it, the total one: \begin{eqnarray}\nonumber \sigma(\theta,\phi) &=& \frac{6\pi c^2}{\omega^2} \biggl\{ |g_-(\theta)|^2 + \frac{\kappa^2}{z^2+\zeta^2} \\ \label{4.9}&\times& \left[ \frac{1}{ 4\pi} + \eta^2 \left( | g_+(\theta)|^2 - |g_-(\theta)|^2\right)\right] \\ \nonumber &-& \frac{2}{\sqrt{4\pi}\left (z^2+\zeta^2\right)}\, {\mathrm Re} \left[{\mathrm e}^{-2{\mathrm i} \delta_0^-} g_-(\theta) \left(\kappa^2 - {\mathrm i} y \right) \right] \biggr\}, \end{eqnarray} \begin{eqnarray}\nonumber \sigma_{{}_{\rm TOT}} &=& \frac{6\pi c^2}{\omega^2} \biggl\{ \| g_-\|^2 +\frac{\kappa^2}{ z^2+ \zeta^2} \\ \label{4.10} &\times& \left[1 + \eta^2\left(\|g_+\|^2 - \|g_-\|^2 \right)\right] \\ \nonumber &-&\frac{1}{ z^2+\zeta^2}\left(y \sin 2\delta_0^- + 2 \kappa^2 \sin^2 \delta_0^- \right) \biggr \}. \end{eqnarray} Let us note that the angular dependence in $\sigma(\theta,\phi)$ is entirely due to $g_\pm(\theta)$ (\ref{3.34}) and, so, to the presence of the $\Lambda$-term in Eq.\ (\ref{1.2}). By some algebraic manipulations $\sigma_{{}_{\rm TOT}} $ can be rewritten in a more perspicuous form: \begin{eqnarray}\nonumber \frac {\omega^2}{ 6\pi c^2}\, \sigma_{{}_{\rm TOT}} &=& \frac{\left(z\sin \delta_0^- - \cos \delta_0^- \right)^2 + \eta^2 A}{z^2 +\zeta^2} \\ &&{}+ \left\| P_\bot g_- \right\|^2 \, \frac{z^2+ B}{z^2 +\zeta^2}\,,\label{4.14} \end{eqnarray} \begin{eqnarray}\nonumber A&=& \sin^2 \delta_0^+ + \kappa^2 \|g_+\|^2 + \left\| P_\bot \Delta g \right\|^2 \\ \label{4.12} &&{}\times \left[ 1+ \eta^2 \left( 1+ \left\| P_\bot \Delta g \right\|^2 \right) \sin^2 \delta_0^- \right], \end{eqnarray} \begin{equation} B= \left( 1+ \eta^2 +\eta^2\left\| P_\bot \Delta g \right\|^2 \right) \left(1+\eta^2 \left\| P_\bot \Delta g \right\|^2 \right). \label{4.13} \end{equation} According to the values of the various coefficients different line shapes appear, which are known as Fano profiles $(\!\!{}$\nolinebreak\cite{CT92} pp.\ 61--63). These shapes are typical of the interference among various channels, when one of them has an amplitude with a pole near the real axis in the complex energy plane (see also Eq.\ (\ref{4.9bis}) below); in our case the channels are direct scattering in the up state, direct scattering in the down state and fluorescence. Some plots of $\frac{\omega^2}{6\pi c^2}\, \sigma_{{}_{\rm TOT}}$ are given in Fig.\ \ref{fig1}; the independent variable is the ``reduced" detuning $\widetilde z= (\omega -\omega_0)/\|\alpha\|^2$, the other parameters are given in the caption of Fig.\ \ref{fig1}; the same figure contains plots of elastic and inelastic cross sections, which will be discussed in Sections \ref{heter} and \ref{Dis}. Whichever the line shape be, there is a strong variation of the cross section for $\omega$ around $\omega_0 +\eta^2\varepsilon$ [see Eqs.\ (\ref{5.28}), (\ref{3.26}), (\ref{3.36g})]. The intensity dependent shift $\eta^2\varepsilon$ of the resonance frequency has received various names in the literature; a very suggestive one is \emph{lamp shift}, a name suggested by A.\ Kastler in \cite{Kast}. Note that in our two-level system the lamp shift is not vanishing only if the two states respond differently to direct scattering; moreover, only the contributions different from the $s$-wave ones do matter. Let us stress that also the width $\Gamma$ of the resonance and the whole line shape are intensity dependent. \paragraph*{No direct scattering.} Let us also note that when the direct scattering is negligible, i.e.\ when Eqs.\ (\ref{3.38}) hold, Eq.\ (\ref{4.10}) reduces to \begin{equation}\label{4.11} \sigma_{{}_{\rm TOT}} = \frac{6\pi c^2}{ \omega^2} \, \frac {\|\alpha\|^4/4}{(\Delta \omega)^2 + \Gamma^2/4}\,. \end{equation} For a laser with negligible intensity, i.e.\ when $\eta \downarrow 0$, Eq.\ (\ref{4.11}) reduces to the cross section for resonant scattering, given in \cite{CT92} pp.\ 530--533; for $\eta \neq 0$, we have a power broadening (see Eq.\ (\ref{3.38c})) of the resonance line, which maintains a Lorentzian shape \cite{Eze1}. By comparing the general case (\ref{4.14}) with the usual one (\ref{4.11}), we see that the main differences are that in the general case we have lamp shift, asymmetric line shape and bigger power broadening. \paragraph*{Low intensity laser.} For $\eta=0$ Eqs.\ (\ref{3.23}) hold and Eqs.\ (\ref{4.9}) and (\ref{4.14}) reduce to \begin{equation}\label{4.9bis} \frac {\omega^2} {6\pi c^2}\, \sigma(\theta,\phi) = \left| g_-(\theta) - \frac {{\mathrm i\, e}^{2{\mathrm i} \delta^-_0}} {\sqrt{4\pi} \left( z+ {\mathrm i} \right)} \right|^2, \end{equation} \begin{equation}\label{4.15} \frac {\omega^2} {6\pi c^2}\, \sigma_{{}_{\rm TOT}} = \|P_\bot g_-\|^2 + \frac{\left(z \sin \delta_0^- - \cos \delta_0^-\right)^2} {z^2 +1}\,. \end{equation} \section{Heterodyne detection}\label{heter} \subsection{Power spectrum}\label{spectrum} The best way to obtain the spectrum of our stimulated atom is by means of the \emph{balanced heterodyne detection} scheme; the output current of the detector is represented by the operator \cite{Bar90,Goslar} \begin{equation}\label{5.1} I(\nu,h;t) = \int_0^t F(t-s) j(\nu,h;{\mathrm d} s)\,, \end{equation} where $F(t)$ is the detector response function, say \begin{equation}\label{5.2} F(t) = k_1 \sqrt{\frac{\gamma}{4\pi}}\, \exp\left( -\frac{\gamma}{2}\, t \right), \qquad \gamma>0\,, \end{equation} $k_1\neq 0$ has the dimensions of a current, $j$ is essentially a field quadrature \begin{equation}\label{5.3} j(\nu,h;{\mathrm d} s) = \overline{ q}\, {\mathrm e}^{{\mathrm i} \nu s} \, {\mathrm d} A_h(s) + \text{h.c.}\,, \end{equation} \begin{equation}\label{5.9} {\mathrm d} A_h(t) = \sum_j \langle h|e_j\rangle\, {\mathrm d}A_j(t)\,, \end{equation} $q$ is a phase factor, $q\in \mathbb C$, $|q|=1$, $\nu$ is the frequency of the local oscillator and $h\in \mathcal Z$, $\|h\|=1$; $h$ contains information on the localization of the detector, say \begin{equation}\label{5.4} h(\theta^\prime, \phi^\prime) = \frac{1}{\sqrt{|\Delta \Upsilon|}}\, 1_{\Delta \Upsilon} (\theta^\prime,\phi^\prime)\,, \end{equation} where $\Delta \Upsilon$ is again the small solid angle around $(\theta,\phi)$ introduced in the previous section. From the canonical commutation relations for the fields one has \begin{eqnarray}\nonumber &&\big[I(\nu_1, h_1; t_1), I(\nu_2, h_2; t_2)] = \int_0^{\min\{t_1,t_2\}} {\mathrm d}s \, F(t_1-s) \\ && \qquad{}\times F(t_2-s) \left( {\mathrm e}^{{\mathrm i}(\nu_1 - \nu_2) s} \langle h_1| h_2 \rangle - \text{c.c.} \right) ; \label{5.6} \end{eqnarray} so, $I(\nu_1, h_1; t_1)$ and $I(\nu_2, h_2; t_2)$ are compatible observables for any choice of the times either if $\nu_1=\nu_2$ and $h_1=h_2$ either if $\langle h_1|h_2\rangle=0$. Under the same conditions also the $j$'s commute. In the following for the quantum expectation of any operator $B$ we shall use the notation \begin{equation}\label{5.11} \langle B\rangle_\lambda^T = \langle U(T) \Psi(\xi,\lambda)| B U(T) \Psi(\xi,\lambda) \rangle\,. \end{equation} In the long run the output mean power is given by \begin{equation}\label{5.5} P(\nu,h) = \lim_{T \to +\infty} \frac{k_2}{T} \int_0^T \big\langle \big( I(\nu,h;t)\big)^2 \big\rangle_\lambda^T \,{\mathrm d}t\,; \end{equation} $k_2>0$ has the dimensions of a resistance, it is independent of $\nu$, but it can depend on the other features of the detection apparatus. In this section $\lambda (t)$ is given by Eq.\ (\ref{3.1}); the limit case (\ref{3.30}) will be considered in the next one. As a function of $\nu$, $P(\nu,h)$ gives the \emph{power spectrum} observed in the ``channel $h$''; in the case of the choice (\ref{5.4}) it is the spectrum observed around the direction $(\theta,\phi)$. Proposition \ref{prop4} relates $P(\nu,h)$ to normal ordered quantum expectations of products of field operators and gives a sum rule which relates $P(\nu,h)$ to $\|\lambda\|^2$; let us note that $\hbar \omega \|\lambda\|^2$ is the total power of the input monochromatic state $\lambda(t)$ (\ref{3.1}). Proposition \ref{prop5} identifies an elastic and an inelastic contribution to the power and reduces the computation of $P(\nu,h)$ to the solution of the master equation (\ref{3.4}). For the use of QSC in the computation of the spectrum of a two-level atom see also Ref.\ \cite{Maass}. \begin{proposition}\label{prop4} The mean power $P(\nu,h)$ can be expressed as \begin{eqnarray}\label{5.8} P&&(\nu,h) = \frac k{4\pi} + \lim_{T\to +\infty} \frac{k}{ 2\pi T} \\ &&{}\times\left\{ \Big\langle \int_0^T {\mathrm d} A^\dagger_h(t) \int_0^t {\mathrm d} A_h(s) \, {\mathrm e}^{ - \left( \frac \gamma 2 +{\mathrm i} \nu \right) (t-s) } \Big\rangle_\lambda^T + \text{\rm c.c.} \right\},\nonumber \end{eqnarray} where $k= k_1^{\,2} k_2$; Eq.\ (\ref{5.8}) holds almost everywhere in $\nu$. We have also \begin{equation}\label{5.21} \int_{-\infty}^{+\infty} \left[ P(\nu,h) - \frac{k}{4\pi} \right]{\mathrm d}\nu = \lim_{T\to +\infty} \frac k T \, \langle \Lambda_{hh}(T) \rangle_\lambda^T\,, \end{equation} where $\Lambda_{hh}(T) = \sum_{ij} \langle e_i|h\rangle \Lambda_{ij} \langle h| e_j\rangle $; moreover, for any c.o.n.s.\ $\{h_j\}$ in $\mathcal Z$, the following sum rule holds: \begin{equation}\label{5.10} \sum_j \int_{-\infty}^{+\infty} \left[ P(\nu,h_j) - \frac{k}{4\pi} \right]{\mathrm d}\nu = k \|\lambda\|^2\,. \end{equation} \end{proposition} {\smallskip\bf \noindent Proof. } By inserting Eqs.\ (\ref{5.1}) and (\ref{5.2}) into the definition (\ref{5.5}) and by changing order of integration, one gets \begin{eqnarray*} P(\nu,h) &=& \lim_{T\to +\infty} \frac{k}{ 4\pi T} \int_0^T \int_0^T \left({\mathrm e}^{- \frac \gamma 2 \left| t-s\right|} - {\mathrm e}^{- \gamma \left( T- \frac{t+s} 2 \right)} \right) \\ &&{}\times \langle j(\nu,h;{\mathrm d}t) j(\nu,h;{\mathrm d}s) \rangle_\lambda^T\, . \end{eqnarray*} The term containing the factor $\exp \left[ - \gamma \left( T- \frac{t+s} 2 \right)\right]$ vanishes for $T\to +\infty$ and one obtains \begin{eqnarray}\nonumber P(\nu,h) &=& \lim_{T\to +\infty} \frac{k}{ 4\pi T} \int_0^T \int_0^T {\mathrm e}^{- \frac \gamma 2 \left| t-s\right|} \\ &&{}\times \langle j(\nu,h;{\mathrm d}t) j(\nu,h;{\mathrm d}s) \rangle_\lambda^T \,.\label{5.7} \end{eqnarray} By using the canonical commutation relations and normal ordering, we have \begin{eqnarray*} P(\nu,h) - \frac k {4\pi}&=& \lim_{T\to +\infty} \frac{k}{ 2\pi T} \int_{t\in (0,T)} \int_{s\in (0,t)} {\mathrm e}^{- \frac \gamma 2 \left( t-s\right)} \\ &&{}\times \langle\, :j(\nu,h;{\mathrm d}t) j(\nu,h;{\mathrm d}s):\,\rangle_\lambda^T \\ &=& \lim_{T\to +\infty} \frac{k}{ 2\pi T} \int_{t\in (0,T)} \int_{s\in (0,t)} {\mathrm e}^{- \frac \gamma 2 \left( t-s\right)} \\ &&{}\times \Big\langle \Bigl\{{\mathrm e}^{- {\mathrm i} \nu (t-s)}\, {\mathrm d} A_h^\dagger(t) {\mathrm d}A_h(s) \\ &&{}+ \overline{q}^2{\mathrm e}^{ {\mathrm i} \nu (t+s)}\, {\mathrm d} A_h(t) {\mathrm d}A_h(s)\Bigr\} \Big\rangle_\lambda^T + \text{c.c.} \end{eqnarray*} The factor $\exp[{\mathrm i} \nu(t+s)]$, when integrated over $\nu$ from $\nu_1$ to $\nu_2$, gives rise to $\big\{ \exp[{\mathrm i} \nu_2(t+s)] - \exp[{\mathrm i} \nu_1(t+s)]\big\}\big/ \{{\mathrm i} (t+s)\}$, which is not singular for $t>0$ and $s>0$; then, the integral containing this factor vanishes for $T\to +\infty$ and Eq.\ (\ref{5.8}) is proved. Now let us observe that \[ \int_0^T{\mathrm d} A^\dagger_h(t)\int_0^T {\mathrm d} A_h(s)\, \delta(t-s) = \Lambda_{hh}(T) \,. \] By integrating over $\nu$ the second term in the r.h.s.\ of Eq.\ (\ref{5.8}) a Dirac delta comes out and by adding the complex conjugated term a double integral for $s\in (0,T)$ and $t\in (0,T)$ is obtained; then, by the previous observation Eq.\ (\ref{5.21}) is obtained. By Eqs.\ (\ref{1.1}), (\ref{2.3}), (\ref{5.21}), we obtain \[ \sum_j \int_{-\infty}^{+\infty} \left[ P(\nu,h_j) - \frac{k}{4\pi} \right] {\mathrm d}\nu= \lim_{T\to +\infty} \frac{k}{T} \langle N(T)\rangle_\lambda^T \,. \] Finally, by Eqs.\ (\ref{2.15}), (\ref{2.4}), (\ref{3.1}), the sum rule (\ref{5.10}) is obtained. {\hfill{$\square$} \vskip4pt} \begin{proposition}\label{prop5} The mean power can be decomposed as the sum of three positive contributions \begin{equation}\label{5.13} P(\nu,h) = \frac k {4\pi} + P_{\mathrm el}(\nu,h) + P_{\mathrm inel}(\nu,h)\,, \end{equation} where \begin{equation}\label{5.17} P_{\mathrm el}(\nu,h) = k\left| r(h) \right|^2 \frac 1 \pi\, \frac {\gamma/2}{ (\nu - \omega)^2 + \gamma^2/4}\,, \end{equation} \begin{eqnarray}\nonumber P_{\mathrm inel}(\nu,h) &=& \frac k {2\pi} \int_0^{+\infty} {\mathrm d}t\, \exp \left[ - \left( \frac\gamma 2 + {\mathrm i} (\nu - \omega) \right) t \right] \\ &\times & {\mathrm Tr} \left\{ D(h)^\dagger \left( {\mathrm e}^{ {\mathcal L}_\lambda t } \left[ D(h) \rho_{\mathrm eq}^\lambda \right]\right) \right\} + \text{\rm c.c.}, \label{5.16} \end{eqnarray} \begin{mathletters}\label{5.18} \begin{eqnarray}\label{5.18a} D(h) &=& R(h) - r(h)\,, \\ \label{5.18b} r(h) &=& {\mathrm Tr} \left\{ R(h) \rho_{\mathrm eq}^\lambda \right\}, \\ \nonumber R(h) &=& {\mathrm e}^{- {\mathrm i}\beta}\langle h|\alpha \rangle \sigma_- + \langle h| S^+\lambda \rangle P_+ + \langle h| S^- \lambda\rangle P_- \\ &=& \sum_j \langle h|e_j \rangle R_j^\lambda\,. \label{5.18c} \end{eqnarray} \end{mathletters} \end{proposition} {\smallskip\bf \noindent Proof. } Let us start from Eq.\ (\ref{5.8}). We can write \begin{eqnarray*} \langle {\mathrm d} A_h^\dagger(t) {\mathrm d} A_h(s) \rangle_\lambda^T &=& \langle \Psi(\xi,\lambda)| U(T)^\dagger {\mathrm d} A_h^\dagger(t) U(T) \\ &\times& U(T)^\dagger {\mathrm d} A_h(s) U(T) \Psi(\xi,\lambda)\rangle \end{eqnarray*} with $T>t>s$. By the rules of QSC (see the ``output fields'' in \cite{Bar90}, Section 3), we obtain \begin{eqnarray*} U(T)^\dagger &&{\mathrm d} A_h(t) U(T) = U(t)^\dagger \bigl\{ \langle h|\alpha \rangle \sigma_- \, {\mathrm d}t \\ {}+&& \sum_j\bigl( \langle h|S^+e_j \rangle P_+ + \langle h|S^- e_j \rangle P_- \bigr) {\mathrm d}A_j(t) \bigr\} U(t)\,. \end{eqnarray*} By using this result we can write \begin{eqnarray}\label{bb} \Big\langle &&\int_0^T {\mathrm d} A_h^\dagger(t)\int_0^t {\mathrm d} A_h(s) \, {\mathrm e}^{-\left( \frac \gamma 2 + {\mathrm i} \nu \right)(t-s)} \Big \rangle_\lambda^T \\ \nonumber &&{}= \int_0^T {\mathrm d}t\int_0^t {\mathrm d}s \, {\mathrm e}^{-\left( \frac \gamma 2 + {\mathrm i}( \nu-\omega) \right)(t-s)} \big\langle {\mathrm e}^{\frac {\mathrm i} 2 \, \sigma_z \beta} \Psi(\xi,\lambda) \big| \\ \nonumber &&{}\times \widetilde U(t)^\dagger R(h)^\dagger \widetilde U(t) \widetilde U(s)^\dagger R(h) \widetilde U(s)\, {\mathrm e}^{\frac {\mathrm i} 2 \, \sigma_z \beta}\Psi(\xi,\lambda)\big\rangle, \end{eqnarray} where $R(h)$ is defined by Eq.\ (\ref{5.18c}) and \[ \widetilde U(t) = {\mathrm e}^{\frac {\mathrm i} 2 \, \sigma_z( \beta +\omega t)} U(t) {\mathrm e}^{-\frac {\mathrm i} 2 \, \sigma_z \beta}. \] By the quantum regression theorem, which holds for a dynamics like $\widetilde U(t)$ \cite{Frig}, we have \begin{eqnarray*} \big\langle &&{\mathrm e}^{\frac {\mathrm i} 2 \, \sigma_z \beta} \Psi(\xi,\lambda) \big| \widetilde U(t)^\dagger R(h)^\dagger \widetilde U(t) \widetilde U(s)^\dagger R(h) \widetilde U(s)\, {\mathrm e}^{\frac {\mathrm i} 2 \, \sigma_z \beta} \\ &&{}\times\Psi(\xi,\lambda)\big\rangle = {\mathrm Tr} \left\{ R(h)^\dagger \, {\mathrm e}^{{\mathcal L}_\lambda (t-s) } \left[ R(h) \, {\mathrm e}^{ {\mathcal L}_\lambda s} \left[ \rho_0\right] \right] \right\}, \end{eqnarray*} where $\rho_0 = \exp \left( \frac {\mathrm i} 2 \sigma_z \beta\right) |\xi\rangle \langle \xi | \exp \left(- \frac {\mathrm i} 2 \sigma_z \beta \right)$. By recalling that $\lim_{t\to +\infty} {\mathrm e}^{ {\mathcal L}_\lambda t} [ \rho]= \rho_{\mathrm eq}^\lambda$ for any state $\rho$, we obtain \begin{eqnarray}\nonumber \lim_{T\to +\infty} &&\frac{k}{ 2\pi T} \Big\langle \int_0^T {\mathrm d} A^\dagger_h(t) \int_0^t {\mathrm d} A_h(s) \, {\mathrm e}^{ - \left( \frac \gamma 2 +{\mathrm i} \nu \right) (t-s) } \Big\rangle_\lambda^T \\ \nonumber &&{}= \frac k {2\pi} \int_0^{+\infty } {\mathrm d}t\, {\mathrm e}^{ -\left( \frac \gamma 2 +{\mathrm i}(\nu -\omega) \right) t} \\ \label{aa} &&{}\times {\mathrm Tr} \left\{ R(h)^\dagger \, {\mathrm e}^{{\mathcal L}_\lambda t } [ R(h) \rho_{\mathrm eq}^\lambda] \right\}. \end{eqnarray} By inserting $R(h)=D(h)+r(h)$ into Eq.\ (\ref{aa}) and this equation into Eq.\ (\ref{5.8}), we obtain the decomposition (\ref{5.13})-(\ref{5.16}). The positivity of $k/(4\pi)$ and $P_{\mathrm el}(\nu,h)$ is apparent from their definitions, while to prove the positivity of $P_{\mathrm inel}(\nu,h)$ requires some transformations. By repeating in the reverse order the steps from Eq.\ (\ref{bb}) to Eq.\ (\ref{aa}), we obtain from Eq.\ (\ref{5.16}) \begin{eqnarray*} P_{\mathrm inel}(\nu,h) &=& \lim_{T\to +\infty} \frac 1 T \int_0^T {\mathrm d}t \int_0^t {\mathrm d}s \, {\mathrm e}^{-\frac \gamma 2 |t-s|} \langle \phi(t) | \phi(s) \rangle \\ &+& \lim_{T\to +\infty} \frac 1 T \int_0^T {\mathrm d}t \int_0^t {\mathrm d}s \, {\mathrm e}^{-\frac \gamma 2 |t-s|} \langle \phi(s) | \phi(t) \rangle\,, \end{eqnarray*} where \[ \phi(t) = \sqrt{ \frac k {2\pi}}\, {\mathrm e}^{{\mathrm i}(\nu - \omega)t}\, \widetilde U(t)^\dagger D(h)\widetilde U(t) {\mathrm e}^{\frac {\mathrm i}{2} \sigma_z \beta} \Psi(\xi,\lambda)\,. \] By exchanging the order of integration and the names of the variables $s$ and $t$ in the second term, we get \[ P_{\mathrm inel}(\nu,h) = \lim_{T\to +\infty} \frac 1 T \int_0^T {\mathrm d}t \int_0^T {\mathrm d}s \, {\mathrm e}^{-\frac \gamma 2 |t-s|} \langle \phi(t) | \phi(s) \rangle\,, \] which is positive because $\exp\left( -\frac \gamma 2 \left|t\right|\right)$ is a positive-definite function, i.e.\ the Fourier transform of a positive function. {\hfill{$\square$} \vskip4pt} Notice that in the decomposition (\ref{5.13}) the term $k/(4\pi)$, independent of $\nu$, is apparently a white noise contribution to the power; $P_{\mathrm el}(\nu,h)$ is the elastic contribution, as one sees from Eq.\ (\ref{5.17}) which gives $P_{\mathrm el}(\nu,h)\propto \delta(\nu-\omega)$ for $\gamma \downarrow 0$; finally, $P_{\mathrm inel}(\nu,h) $ is the inelastic contribution (from Eq.\ (\ref{5.16}) one can see that no delta term develops for $\gamma \downarrow 0$). By Eqs.\ (\ref{3.10}), (\ref{3.27})-(\ref{3.29}), (\ref{5.17})-(\ref{5.18}), we obtain \begin{eqnarray}\nonumber r(h)&=& \langle h| S^- \lambda \rangle + \langle h| \Delta S \lambda \rangle u(\infty) \\ &&{}+ {\mathrm e}^{-{\mathrm i}\beta} \langle h| \alpha \rangle v(\infty)\,, \label{5.12} \end{eqnarray} \begin{equation}\label{5.14} P_{\mathrm inel} (\nu,h) = \frac k {2\pi} \, \bbox{c}^{h\,\dagger} \, \frac 1 {\bbox{G} + \frac \gamma 2 + {\mathrm i} (\nu-\omega)}\, \bbox{d}^{h} + \text{c.c.}\,, \end{equation} \begin{equation}\label{5.15} \bbox{c}^h = \left( \begin{array}{c} \langle h| \Delta S \lambda \rangle \\ 0 \\ {\mathrm e}^{-{\mathrm i}\beta} \langle h| \alpha \rangle \end{array} \right), \end{equation} \begin{mathletters}\label{5.19} \begin{eqnarray}\nonumber d_1^h &=& \bigl[ \langle h| \Delta S \lambda \rangle \big(1-u(\infty)\big) \\ \label{5.19a} &&{}- {\mathrm e}^{-{\mathrm i}\beta} \langle h| \alpha \rangle v(\infty) \bigr] u(\infty)\,, \\ \nonumber d_2^h &=& \bigl[ \langle h| \Delta S \lambda \rangle \big(1-u(\infty)\big) \\ &&{}- {\mathrm e}^{-{\mathrm i}\beta} \langle h| \alpha \rangle v(\infty) \bigr] v(\infty)\,, \label{5.19b} \\ \nonumber d_3^h &=& {\mathrm e}^{-{\mathrm i}\beta} \langle h| \alpha \rangle \left(u(\infty) -|v(\infty)|^2 \right) \\ &&{}- \langle h| \Delta S \lambda \rangle u(\infty)\, \overline{v(\infty)}\,. \label{5.19c} \end{eqnarray} \end{mathletters} \subsection{Elastic and inelastic cross sections}\label{cross} Let us consider now the case of the spherically symmetric atom, stimulated by a well collimated laser beam, for which Eqs.\ (\ref{3.30})-(\ref{3.36}) hold. We also assume that the detector spans a small solid angle, so that $h$ is given by Eq.\ (\ref{5.4}) with $\Delta \Upsilon \downarrow \{(\theta,\phi)\}$, $| \Delta \Upsilon | \simeq \sin \theta \, {\mathrm d }\theta \, {\mathrm d} \phi$. Moreover, we assume that the transmitted wave does not reach the detector, i.e.\ $\theta>0$ and so \begin{equation}\label{5.22} \langle h | \lambda\rangle =0\,. \end{equation} From Eqs.\ (\ref{5.17}), (\ref{5.12})-(\ref{5.19}) we obtain the elastic and inelastic contributions to the power (per unit of solid angle) \begin{mathletters}\label{5.23} \begin{eqnarray}\label{5.23a} \frac 1 {|\Delta \Upsilon|} \, P_{\mathrm el}(\nu,h) &\simeq& P_{\mathrm el}(\nu;\theta,\phi)\,, \\ \label{5.23b} \frac 1 {|\Delta \Upsilon|} \, P_{\mathrm inel}(\nu,h) &\simeq& P_{\mathrm inel}(\nu;\theta,\phi)\,, \end{eqnarray} \end{mathletters} \noindent where \begin{mathletters}\label{5.23'} \begin{equation}\label{5.23'a} P_{\mathrm el}(\nu;\theta,\phi) = k \eta^2 \|\alpha\|^2 |a(\theta)|^2\, \frac {\gamma/(2\pi)} {(\nu - \omega)^2 + \gamma^2/4}\,, \end{equation} \begin{eqnarray}\nonumber P_{\mathrm inel}(\nu;\theta,\phi) &=& \frac{ k\eta^2 \|\alpha\|^2} {2\pi} \bbox{c}(\theta)^\dagger \, \frac 1 {\bbox{G} + \frac \gamma 2 + {\mathrm i} (\nu-\omega)}\, \bbox{d}(\theta) \\ &&{}+ \text{c.c.}\,,\label{5.23'b} \end{eqnarray} \end{mathletters} \begin{eqnarray}\nonumber a(\theta)&=& g_-(\theta) + \Delta g(\theta)\, \frac { \eta^2 \kappa^2} {z^2+\zeta^2} \\ \label{5.24} &&{}- {\mathrm e}^{2{\mathrm i}\delta_0^-} \, \frac {\kappa^2 + {\mathrm i} y } {\sqrt{4\pi} \left(z^2+\zeta^2\right)} \, , \end{eqnarray} \begin{equation}\label{5.25} \bbox{c}(\theta) = \left( \begin{array}{c} \eta \Delta g(\theta) \\ 0 \\ -{\mathrm e}^{2{\mathrm i}\delta_0^-} /\sqrt{4\pi} \end{array} \right), \end{equation} \begin{mathletters}\label{5.26} \begin{eqnarray}\label{5.26a} d_1(\theta)&=& \frac {\eta\kappa^2} {z^2+\zeta^2}\, m(\theta)\,, \\ \label{5.26b} d_2(\theta)&=& \frac {m(\theta)} {z^2+\zeta^2} \left( \kappa^2 + {\mathrm i} y\right) , \\ \nonumber d_3(\theta) &=& -\frac {\eta^2 } { \left( z^2 +\zeta^2\right)^2} \biggl\{ \frac{{\mathrm e}^{2{\mathrm i}\delta_0^-}}{\sqrt{4\pi}} \Bigl[ \|\Delta g\|^2 \left(y^2+ \kappa^4\right) \\ \nonumber &&{}+ \kappa^2 y\sin 2 s + 2\kappa^4 \cos^2 s \Bigr] \\ \label{5.26c} &&{}+ \Delta g(\theta) \kappa^2 \left( \kappa^2 - {\mathrm i}y \right) \biggr\}, \end{eqnarray} \end{mathletters} \begin{eqnarray}\nonumber m(\theta)&=& \Delta g(\theta) \left( 1 - \frac {\eta^2 \kappa^2} {z^2 +\zeta^2} \right) \\ \label{5.27} &&{}+{\mathrm e}^{2 {\mathrm i} \delta_0^-} \, \frac { \kappa^2+{\mathrm i}y} {\sqrt{4\pi} \left( z^2 +\zeta^2\right)}\,. \end{eqnarray} For the elastic and inelastic cross sections we shall have $\sigma_{\mathrm el}(\nu;\theta,\phi) \propto P_{\mathrm el}(\nu;\theta,\phi)$, $\sigma_{\mathrm inel}(\nu;\theta,\phi) \propto P_{\mathrm inel}(\nu;\theta,\phi)$. To find the constant of proportionality, let us observe that, from Eqs.\ (\ref{5.21}), (\ref{5.23}), (\ref{4.1}), (\ref{4.3}), (\ref{4.5}), we get \begin{eqnarray}\nonumber \int_{-\infty}^{+\infty}&& \left[P_{\mathrm el}(\nu;\theta,\phi) + P_{\mathrm inel} (\nu;\theta,\phi)\right] {\mathrm d} \nu \\ \label{5.29} {}=&& \lim_{t\to +\infty} \frac k t \, \langle n(\theta, \phi;t\rangle = \frac{\eta^2 k \|\alpha\|^2 \omega^2} {6\pi c^2}\, \sigma(\theta,\phi)\,. \end{eqnarray} Therefore, taking into account Eqs.\ (\ref{5.23'}), we obtain the expressions for the cross sections \begin{equation}\label{5.30} \sigma_{\mathrm el}(\nu;\theta,\phi)= \frac{3 c^2}{ \omega^2} \left| a(\theta)\right|^2\,\frac {\gamma} {(\nu - \omega)^2 + \gamma^2/4}\,, \end{equation} \begin{equation}\label{5.31} \sigma_{\mathrm inel}(\nu;\theta,\phi)= \frac {3 c^2} {\omega^2} \, \bbox{ c}(\theta)^\dagger \, \frac {1} {\bbox{G} + \frac \gamma 2 + {\mathrm i} (\nu-\omega)}\, \bbox{d}(\theta) + \text{c.c.}\,, \end{equation} and the relation \begin{equation}\label{5.32} \int_{-\infty}^{+\infty} \left[ \sigma_{\mathrm el}(\nu; \theta,\phi) + \sigma_{\mathrm inel}(\nu;\theta,\phi)\right] {\mathrm d}\nu = \sigma(\theta,\phi)\,, \end{equation} where $\sigma(\theta,\phi)$ is given by Eq.\ (\ref{4.9}). Finally, let us introduce the integral cross sections \begin{mathletters}\label{5.33} \begin{eqnarray}\label{5.33a} \sigma_{\mathrm el}(\nu) &=& \int_0^\pi {\mathrm d}\theta \, \sin \theta \int_0^{2\pi} {\mathrm d}\phi \, \sigma_{\mathrm el}(\nu;\theta,\phi)\,, \\ \label{5.33a'} \sigma_{\mathrm el} &=& \int_{-\infty}^{+\infty} \sigma_{\mathrm el}(\nu) \,{\mathrm d}\nu\,, \\ \label{5.33b} \sigma_{\mathrm inel}(\nu) &=& \int_0^\pi {\mathrm d}\theta \, \sin \theta \int_0^{2\pi} {\mathrm d}\phi \, \sigma_{\mathrm inel}(\nu;\theta,\phi)\,, \\ \label{5.33c} \sigma_{\mathrm inel} &=& \int_{-\infty}^{+\infty} \sigma_{\mathrm inel}(\nu) \,{\mathrm d}\nu\,; \end{eqnarray} \end{mathletters} \noindent the relation (\ref{5.32}) becomes \begin{equation}\label{tot} \sigma_{\mathrm el}+\sigma_{\mathrm inel} =\sigma_{{}_{\mathrm TOT}}\,, \end{equation} where $\sigma_{{}_{\mathrm TOT}}$ is given by Eqs.\ (\ref{4.12})-(\ref{4.14}). \section{Cross sections and fluorescence spectrum}\label{Dis} In this section we want to discuss the behavior of the integral cross sections and of the fluorescence spectrum. From Eqs.\ (\ref{5.24}), (\ref{5.30}), (\ref{5.33a}), (\ref{5.33a'}) we obtain \begin{equation}\label{6.1-} \sigma_{\mathrm el}(\nu)= \sigma_{\mathrm el}\, \frac{\gamma/(2\pi)} {(\nu -\omega)^2 + \gamma^2/4}\,, \end{equation} \begin{eqnarray}\nonumber \frac{\omega^2}{6\pi c^2}&&\, \sigma_{\mathrm el} = \frac 1 {\left(z^2+\zeta^2\right)^2} \left\| P_\bot \left[ \left( z^2 +B \right) g_- + \eta^2 \kappa^2 g_+ \right] \right\|^2 \\ \label{6.1} &&{}+ \left| {\mathrm e}^{-{\mathrm i}\delta_0^-} \sin \delta_0^- + \frac {\eta^2 \kappa^2 {\mathrm e}^{{\mathrm i}s} \sin s - y + {\mathrm i}\kappa^2} { z^2 + \zeta^2} \right|^2\,, \end{eqnarray} while from Eqs.\ (\ref{5.25})-(\ref{5.27}), (\ref{5.31}), (\ref{5.33b}), (\ref{5.33c}) we obtain \begin{equation}\label{6.2} \frac{\omega^2}{6\pi c^2}\, \sigma_{\mathrm inel} = \frac {\eta^2 \left( 1 + \kappa^2 \right) E(y)} { \left(z^2 +\zeta^2\right)^2}\,, \end{equation} \begin{equation}\label{6.3} E(y)= \left(y \sin s + \kappa^2 \cos s \right)^2 + \left\| P_\bot \Delta g \right\|^2 \left(y^2 + \kappa^4\right); \end{equation} one can check that the relation (\ref{tot}) holds true. Let us recall that the various quantities appearing in the previous formulas are given by Eqs.\ (\ref{3.7i}), (\ref{3.34})-(\ref{5.28}), (\ref{3.36m}), (\ref{3.36c}), (\ref{4.13}) and that $\|\alpha\|^2$ is the natural line width, $\omega_0$ is the atomic resonance frequency, $\Omega =\eta \|\alpha\|^2$ is the bare Rabi frequency, $\eta^2 \varepsilon$ is the intensity dependent shift, $\delta_0^\pm$, $\left\| P_\bot g_\pm \right\|^2$, $\left\| P_\bot \Delta g \right\|^2$ are parameters linked to the $S_\pm$ scattering matrices, satisfying \[ \Big| \left\| P_\bot g_+ \right\| - \left\| P_\bot g_- \right\| \Big| \leq \left\| P_\bot \Delta g \right\| \leq \left\| P_\bot g_+ \right\| + \left\| P_\bot g_- \right\| . \] We introduce also a ``reduced" detuning $\widetilde z$ \begin{equation}\label{6.red} \widetilde z = \frac{ \omega - \omega_0} {\|\alpha\|^2}\, ; \end{equation} we have also $z= 2 \widetilde z - 2 \eta^2 \varepsilon / \|\alpha\|^2$, $s= \delta_0^+ - \delta_0^-$, $y = z- \frac{\eta^2} 2 \sin 2s$. As an example, in Fig.\ \ref{fig1} we plot $\frac{\omega^2}{6\pi c^2}\, \sigma_{{}_{\rm TOT}}$, $\frac{\omega^2}{6\pi c^2}\, \sigma_{\rm inel}$, $\frac{\omega^2}{6\pi c^2}\, \sigma_{\rm el}$ as functions of the detuning $\widetilde z$ in the four cases $\eta^2 = 10,\, 18,\, 28,\, 40$; the other parameters are $\delta_0^+=-0.03$, $\delta_0^-=0.13$, $\left\| P_\bot g_\pm \right\|^2 = 0.005$, $\left\| P_\bot \Delta g \right\|^2= 0.02$, $\varepsilon/ \|\alpha\|^2=-0.001$. Let us note the strong asymmetry in $\widetilde z$ of the cross sections and the fact that $\displaystyle \lim_{\widetilde z \to \pm \infty} \frac{\omega^2}{6\pi c^2}\, \sigma_{{}_{\rm TOT}} = \lim_{\widetilde z \to \pm \infty} \frac{\omega^2}{6\pi c^2}\, \sigma_{\rm el} = \left\| P_\bot g_- \right\|^2 + \sin^2 \delta_0^-$, which is about 0.0218 with our parameters. Let us recall that the usual model with only the absorption/emission process corresponds to $\delta_0^\pm=0$, $\left\| P_\bot g_\pm \right\|^2=\left\| P_\bot \Delta g \right\|^2= 0$, $\varepsilon=0$, $z=2\widetilde z$; in this case, from Eqs.\ (\ref{4.11}), (\ref{6.1}), (\ref{6.2}), we have easily \begin{mathletters}\label{5.48} \begin{eqnarray}\label{5.48a} \frac{\omega^2}{6\pi c^2}\, \sigma_{{}_{\mathrm TOT}}&=& \frac{1}{4 \widetilde z^2+1+2\eta^2}\,, \\ \label{5.48b} \frac{\omega^2}{6\pi c^2}\,\sigma_{\mathrm el}&=& \frac{4 \widetilde z^2 +1}{\left(4 \widetilde z^2 +1+2\eta^2\right)^2}\,, \\ \label{5.48c} \frac{\omega^2}{6\pi c^2}\,\sigma_{\mathrm inel} &=& \frac{2\eta^2}{\left(4 \widetilde z^2+1+2\eta^2\right)^2}\,. \end{eqnarray} \end{mathletters} \noindent Now the cross sections are symmetric in $\widetilde z$ and $\displaystyle \lim_{\widetilde z \to \pm \infty} \frac{\omega^2}{6\pi c^2}\, \sigma_{{}_{\rm TOT}} = \lim_{\widetilde z \to \pm \infty} \frac{\omega^2}{6\pi c^2}\, \sigma_{\rm el} =0$. Then, we introduce the normalized inelastic spectrum \begin{equation}\label{6.4} \Sigma_{\mathrm inel}(x) = \frac{\omega^2 \|\alpha\|^2} {6\pi c^2} \, \sigma_{\mathrm inel}(\nu) \end{equation} and the total one \begin{equation} \label{6.21} \Sigma_{{}_{\mathrm TOT}}(x)= \frac {\omega^2 \sigma_{\mathrm el}} {6\pi c^2} \, \frac {\widetilde \gamma /(2\pi)} {x^2 + (\widetilde \gamma/2)^2} + \Sigma_{\mathrm inel}(x)\,, \end{equation} where we have introduced the ``reduced" frequency $x$ and the ``reduced" instrumental width $\widetilde \gamma$ \begin{equation}\label{6.5} x= \frac{ \nu-\omega} {\|\alpha\|^2}\,, \qquad \widetilde \gamma = \frac \gamma {\|\alpha\|^2}\,; \end{equation} the normalization we have chosen is \begin{mathletters}\label{6.15} \begin{eqnarray}\label{6.15a} \int_{-\infty}^{+\infty} \Sigma_{{}_{\mathrm TOT}}(x)\, {\mathrm d }x &=& \frac{\omega^2}{6\pi c^2}\, \sigma_{{}_{\rm TOT}}\,, \\ \label{6.15b} \int_{-\infty}^{+\infty} \Sigma_{\mathrm inel}(x)\, {\mathrm d }x &=& \frac{\omega^2}{6\pi c^2}\, \sigma_{\mathrm inel}\,. \end{eqnarray} \end{mathletters} \noindent The explicit expression of $\Sigma_{\mathrm inel}(x)$ is given by the following proposition. \begin{proposition}\label{prop6} The inelastic spectrum is given by \begin{eqnarray}\nonumber \Sigma_{\mathrm inel}(x) &=& \frac {\eta^2} {\pi \left(z^2+\zeta^2\right)^2} \biggr( \bbox{c}^{\prime \dagger} \, \frac 1 {\widetilde{\bbox{G}} + 2{\mathrm i} x}\, \bbox{d}^\prime \\ \label{6.9} &&{}+ \left\|P_\bot \Delta g \right\|^2 \bbox{c}^{\prime\prime \dagger} \, \frac 1 {\widetilde{\bbox{G}} + 2{\mathrm i} x}\, \bbox{d}^{\prime\prime} + \text{\rm c.c.} \biggr), \end{eqnarray} where \begin{equation}\label{6.6} \bbox{c}^\prime = \left( \begin{array}{c} {\mathrm i} {\mathrm e}^{{\mathrm i} s} \sin s \\ 0 \\ 1 \end{array} \right), \qquad \bbox{c}^{\prime\prime} = \left( \begin{array}{c} 1\\ 0 \\ 0 \end{array} \right), \end{equation} \begin{mathletters}\label{6.7} \begin{eqnarray}\label{6.7a} d_1^\prime&=& \kappa^2 m^\prime\,, \\ \label{6.7b} d_2^\prime&=& \left( \kappa^2 + {\mathrm i} y\right) m^\prime\,, \\ \nonumber d_3^\prime &=& \|\Delta g\|^2 \left(y^2+ \kappa^4\right) + \kappa^2 y\sin 2 s \\ \label{6.7c} &&{}+ 2\kappa^4 \cos^2 s + {\mathrm i} \kappa^2 \left( \kappa^2 - {\mathrm i}y \right) {\mathrm e}^{{\mathrm i}s}\sin s \,, \\ \label{6.7g} m^\prime &=& \kappa^2 + {\mathrm i} y + {\mathrm i}\left( z^2 + \zeta^2 - \eta^2\kappa^2 \right) {\mathrm e}^{{\mathrm i}s} \sin s \,, \\ \label{6.7d} d_1^{\prime\prime}&=& \kappa^2 \left(z^2+ \zeta^2 -\eta^2 \kappa^2 \right), \\ \label{6.7e} d_2^{\prime\prime} &=& \left( \kappa^2 + {\mathrm i} y\right) \left(z^2+ \zeta^2 -\eta^2 \kappa^2 \right), \\ \label{6.7f} d_3^{\prime\prime}&=& \kappa^2 \left( \kappa^2 - {\mathrm i}y \right) , \end{eqnarray} \end{mathletters} \begin{equation}\label{6.8} \widetilde{\bbox{G}}= \left( \begin{array}{ccc} 2 + \widetilde \gamma & - 1 & \eta^2 \\ & & \\ 2\eta^2 {\mathrm e}^{{\mathrm i}s} \cos s & b^\prime + \widetilde \gamma & 0 \\ & & \\ -2{\mathrm e}^{-{\mathrm i}s} \cos s & 0 & \overline{ b^\prime} + \widetilde \gamma \end{array} \right); \end{equation} $b^\prime$ is given by Eq.\ (\ref{set5}). \end{proposition} {\smallskip\bf \noindent Proof. } From Eqs.\ (\ref{3.36a}), (\ref{5.25}), (\ref{5.26}), (\ref{5.27}), (\ref{5.31}), (\ref{6.2}), (\ref{6.4}), (\ref{6.5}), we have \begin{eqnarray*} \Sigma_{\mathrm inel}(x) &=& \frac 1 {\pi \left(z^2+\zeta^2\right)^2} \biggr(\widetilde {\bbox{c}}^{\prime \dagger} \, \frac 1 {\bbox{G}^\prime + \widetilde \gamma+ 2{\mathrm i} x}\, \widetilde{\bbox{d}}^\prime \\ &&{}+ \eta \left\|P_\bot \Delta g \right\|^2 \bbox{c}^{\prime\prime \dagger} \, \frac 1 {\bbox{G}^\prime + \widetilde \gamma+ 2{\mathrm i} x}\, \widetilde{ \bbox{d}}^{\prime\prime} + \text{c.c.} \biggr), \end{eqnarray*} where, by using the definitions above, \[ \widetilde{\bbox{c}}^\prime = \left( \begin{array}{c} {\mathrm i} \eta{\mathrm e}^{{\mathrm i} s} \sin s \\ 0 \\ -1 \end{array} \right), \] \[ \widetilde{\bbox{d}}^\prime = \left( \begin{array}{c} \eta d_1^\prime \\ d_2^\prime \\ -\eta^2 d_3^\prime \end{array} \right), \qquad \widetilde {\bbox{ d}}^{\prime\prime} = \left( \begin{array}{c} \eta d_1^{\prime\prime} \\ d_2^{\prime\prime} \\ -\eta^2 d_3^{\prime\prime} \end{array} \right). \] Then, by using the transformation \[ \bbox{G}^\prime+ \widetilde \gamma= \bbox{M}\widetilde{ \bbox{ G}} \bbox{ M}^{-1}\,, \] where \[ \bbox{ M}= \left( \begin{array}{ccc} \eta & 0 & 0 \\ 0 & 1 &0 \\ 0 & 0 & -\eta^2 \end{array} \right), \] we get the Eq.\ (\ref{6.9}). {\hfill{$\square$} \vskip4pt} By direct computations we can get the inverse of the matrix $\widetilde {\bbox{ G}}+2{\mathrm i}x$. We can write \begin{equation}\label{6.16} \frac{1}{\widetilde{ \bbox{ G}} + 2{\mathrm i} x} = \frac{1}{\det \left(\widetilde{ \bbox{ G}} + 2{\mathrm i} x\right)}\, \bbox{ D}(x)\,, \end{equation} where \end{multicols} \vspace{-0.4cm} \noindent\rule{0.49\textwidth}{0.4pt}\rule{0.4pt}{\baselineskip} \widetext \begin{eqnarray}\nonumber \det \left( \widetilde{ \bbox{ G}} + 2{\mathrm i} x\right)&=& \left(2+ \widetilde \gamma+2{\mathrm i}x \right) \left[ \left( \kappa^2 + \widetilde \gamma+ 2{\mathrm i}x \right)^2 + \left( z + \frac {\eta^2} 2 \, \sin 2s \right)^2 \right] \\ &&{}+ 4 \eta^2 \cos s \left[ \left(\kappa^2 + \widetilde \gamma+ 2{\mathrm i}x \right)\cos s - \left( z + \frac {\eta^2} 2 \, \sin 2s \right)\sin s \right], \label{6.17} \end{eqnarray} \begin{mathletters}\label{6.18} \begin{eqnarray}\label{6.18a} D_{11}(x)&=& \left( \kappa^2 + \widetilde \gamma+ 2{\mathrm i}x \right)^2 + \left( z + \frac {\eta^2} 2 \, \sin 2s \right)^2, \\ \label{6.18b} D_{12}(x)&=& \kappa^2 + \widetilde \gamma+ {\mathrm i} \left(2x+ z + \frac {\eta^2} 2 \, \sin 2s \right), \\ \label{6.18c} D_{13}(x)&=& -\eta^2 \left[\kappa^2 + \widetilde \gamma+ {\mathrm i} \left(2x- z - \frac {\eta^2} 2 \, \sin 2s \right)\right], \\ \label{6.18g} D_{31}(x)&=& 2 {\mathrm e}^{-{\mathrm i}s}\, \cos s\left[\kappa^2 + \widetilde \gamma+ {\mathrm i} \left(2x- z - \frac {\eta^2} 2 \, \sin 2s \right)\right]\,, \\ \label{6.18h} D_{32}(x)&=& 2 {\mathrm e}^{-{\mathrm i}s}\, \cos s\,, \\ \label{6.18i} D_{33}(x)&=& \left(2+ \widetilde \gamma+2 {\mathrm i} x\right)\left[\kappa^2 + \widetilde \gamma+ {\mathrm i} \left(2x- z - \frac {\eta^2} 2 \, \sin 2s \right)\right]+2\eta^2 {\mathrm e}^{{\mathrm i}s}\, \cos s\,. \end{eqnarray} \end{mathletters} \begin{multicols}{2} \narrowtext \noindent The matrix elements $D_{2j}(x)$ are not needed in formula (\ref{6.9}). One can check that the inelastic spectrum is asymmetric, but it is invariant under the transformation: $x\to - x$, $s\to - s$, $z\to - z$. The formula for the inelastic spectrum given in Proposition \ref{prop6} becomes significantly simpler in the usual case ($g_\pm =0$) and when the intensity of the stimulating laser is low. \subsubsection*{The case $g_\pm =0$} Let us consider the usual model, when the direct scattering terms are negligible, i.e.\ $g_\pm=0$, which gives also $s=0$, $\zeta^2=1+2\eta^2$, $\kappa^2=1$, $y=z=2\widetilde z$; in this case the integral cross sections are given by Eqs.\ \ref{5.48} and, with some computations, the inelastic spectrum is obtained from Proposition \ref{prop6}: \begin{mathletters}\label{5.49} \begin{equation} \label{5.49a} \Sigma_{\mathrm inel}(x) = \frac {4\eta^2 p(x)} {\pi q(x) \left(z^2+1+2\eta^2\right)^2}\,, \end{equation} \begin{eqnarray}\nonumber p(x)&=& \left( 2 + \widetilde \gamma \right) \left[ \left( 1 + \widetilde \gamma \right)^2 +2 \eta^2 + z^2 \right] \\ \label{5.49b} {}&\times& \left[ \left( 2 + \widetilde \gamma \right)^2 + 2 \eta^2 +4x^2 \right] \\ \nonumber {}&+& 2\widetilde \gamma \left[ 2\left(2 x^2 - \eta^2\right)^2 + \left( 2 + \widetilde \gamma \right)^2\left(2 x^2 + \eta^2 \right) \right], \end{eqnarray} \begin{eqnarray}\nonumber q(x)&=& \biggl\{\left( 2 + \widetilde \gamma \right) \left[ \left( 1 + \widetilde \gamma \right)^2 + z^2 \right] \\ \label{5.49c} {}&+& 4 \left( 1 + \widetilde \gamma \right)\eta^2 - 4\left( 4 + 3 \widetilde \gamma \right) x^2 \biggr\}^2 \\ \nonumber {}&+& 4x^2 \left(3 \widetilde \gamma^2 +8 \widetilde \gamma +5 + z^2 + 4 \eta^2 - 4 x^2\right)^2 . \end{eqnarray} \end{mathletters} \noindent Now the inelastic spectrum is invariant either under the transformation $x\to - x$ either under the transformation $z\to - z$. If we put also $\widetilde \gamma=0$, which means that the instrumental width is negligible, then one can check that the fluorescence spectrum $\Sigma_{{}_{\mathrm TOT}}(x)$, given by Eqs.\ (\ref{6.21}), (\ref{5.48}), (\ref{5.49}), coincides exactly (apart from the different normalization) with the spectrum computed by Mollow $\big(\!$~\cite{Moll}, Eq.\ (4.15)$\big)$. Eq.\ (\ref{5.49}) is simply the convolution of the inelastic part of the Mollow spectrum with a Lorentzian of width $\widetilde \gamma$. If also $z=0$ (no detuning), the eigenvalues of $\widetilde{ \bbox{ G}}$ can be computed and, by using them, the denominator in Eq.\ (\ref{5.49}) can be factorized. In the case $\eta^2\leq 1/16$, $\widetilde{ \bbox{ G}}$ has real eigenvalues and $\Sigma_{\mathrm inel}(x)$ has a single peak in $x=0$, while for $\eta^2>1/16$ two complex eigenvalues appear; therefore, $\Sigma_{\mathrm inel}(x)$ has a three-peaked structure for $\eta^2$ sufficiently larger than $1/16$. For $\eta$ very large Eq.\ (\ref{5.49}) gives three peaks in $\nu\simeq \omega- \Omega$, $\nu=\omega=\omega_0$, $\nu\simeq \omega+ \Omega$ with height ratio $1 : \frac{3+2\widetilde \gamma} {1+ \widetilde \gamma} :1$ and widths $\frac 3 2 \, \|\alpha\|^2 + \gamma $, $\|\alpha\|^2+ \gamma $, $\frac 3 2 \, \|\alpha\|^2+ \gamma $ (see Ref.\ \cite{Moll} or Ref.\ \cite{CT92} pp.\ 387, 423-426, 437-441 for the case $\gamma =0$). \subsubsection*{Low intensity laser} From Eq.\ (\ref{6.2}) we see that the inelastic cross section vanishes in the limit of vanishing intensity of the laser; however, the first correction, proportional to $\eta^2$, presents some interesting aspects. We have immediately \begin{equation}\label{5.37} \frac {\omega^2}{6\pi c^2}\, \sigma_{\mathrm inel} \simeq \frac {2\eta^2 E_0\left( \widetilde z \right)} {\left(4 \widetilde z^2 + 1\right)^2}\,, \end{equation} \begin{eqnarray}\nonumber E_0\left( \widetilde z \right) = E(y)\big|_{\eta=0}&=& \left(2 \widetilde z\sin s + \cos s\right)^2 \\ &+& \left\|P_\bot \Delta g\right\|^2 \left(4 \widetilde z^2+1\right).\label{6.22} \end{eqnarray} The computation of the spectrum is straightforward, but long; the final result is: for small $\eta$ we have \end{multicols} \vspace{-0.4cm} \noindent\rule{0.49\textwidth}{0.4pt}\rule{0.4pt}{\baselineskip} \widetext \begin{eqnarray}\nonumber \Sigma_{\mathrm inel}(x) &\simeq& \frac{\eta^2} {2\pi} \left[ \frac{\left\|P_\bot \Delta g\right\|^2\left(1+ \widetilde \gamma\right)} {\widetilde z^2 +1/4} +\frac {\widetilde \gamma \left( 2 \widetilde z \sin s + \cos s \right)^2} {4 \left( \widetilde z^2+1/4 \right)^2} \right] \left[ \frac 1 {4\left(x +\widetilde z \right)^2 + \left(1+\widetilde \gamma \right)^2} + \frac 1 {4\left(x -\widetilde z \right)^2 + \left(1+\widetilde \gamma \right)^2} \right] \\ \label{5.35} &+&\frac{2\eta^2 \left(2 \widetilde z \sin s + \cos s \right)^2 \left[ \widetilde z^2 + \left(1+ \widetilde \gamma \right)^2/4 \right]} {\pi \left( \widetilde z^2 +1/4\right)^2 \left[4\left(x +\widetilde z\right)^2 + \left(1+ \widetilde \gamma \right)^2 \right] \left[4\left(x -\widetilde z\right)^2 + \left(1+ \widetilde \gamma \right)^2 \right] } \,. \end{eqnarray} \noindent{}\hfill{ \rule[-\baselineskip]{0.4pt}{\baselineskip}\rule{0.49\textwidth}{0.4pt}} \vspace{-0.4cm} \begin{multicols}{2} \narrowtext In this case the inelastic spectrum is invariant either under the transformation $x\to - x$ either under the transformation $s\to - s$ and $\widetilde z\to - \widetilde z$. The usual case $(g_\pm =0)$ was already discussed by Mollow $\big($\cite{Moll}, Eq.\ (4.30)$\big)$ for $\widetilde \gamma=0$ and can be obtained from Eqs.\ (\ref{5.49}) by letting $\eta^2$ vanish or from Eq.\ (\ref{5.35}) by taking $s=0$ and $\left\|P_\bot \Delta g\right\|=0$. In the Mollow case, for $|\Delta \omega|$ sufficiently large, the inelastic spectrum presents two peaks (see also Ref.\ \cite{CT92}, pp.\ 106-108, 386). The structure given by Eq.\ (\ref{5.35}) is similar also for $s\neq 0$, $\left\|P_\bot \Delta g\right\| \neq 0$: again two symmetric peaks appear for $|\Delta \omega|$ sufficiently large. \subsubsection*{Numerical computations} In the general case the total spectrum is given by Eqs.\ (\ref{6.21}), (\ref{6.9})--(\ref{6.18}); the analytic expression is involved, but plots can be easily obtained by numerical computations. According to the values of the various parameters, a well resolved triplet structure can appear, but also single-maximum structures can be shown. With the choice of parameters of Fig.\ \ref{fig1} and with an instrumental width $\widetilde \gamma = 0.6$, the on resonance spectrum for $\eta^2=10,\, 18,\, 28,\, 40$ is given in Fig.\ \ref{fig2} (solid lines); the dashed lines give the Mollow spectrum for the same values of $\eta^2$ and $\widetilde \gamma$. The parameters in Fig.\ \ref{fig2} have been chosen in such a way that a triplet structure appears, not too different from the usual one, but with a well visible asymmetry in the frequency $x$. Experiments \cite{Stroud,Walther1,Eze2,Eze1,Walther2} confirm essentially the triplet structure; some asymmetry has been found, whose origin has been attributed to various causes. In this connection it has also been observed that calculations for multilevel atoms indicate some asymmetry \cite{Eze2}. Indeed, the introduction in our model of the interaction term containing the gauge process simulates the presence of other levels and the virtual transitions to them. Finally, in Fig.\ \ref{fig3} we show some out of resonance spectra (detunings $\widetilde z = -4,\, -2,\, 3,\, 6$) for $\eta^2 =28$ and the other parameters as in Figs.\ \ref{fig1} and \ref{fig2} (solid lines); again, the dashed lines give the Mollow spectrum. Now, a strong difference from the usual case is shown, consistent with the strong asymmetry in $\widetilde z$ shown by the total and the elastic cross sections in Fig.\ \ref{fig1}.
\section{Introduction} The SN1987A in the Large Magellanic Cloud was a rare and unique event thanks to its nearness to us. It has been observed with all available modern instruments since its explosion (e.g., Chevalier 1992) and is expected to have another magnificent display in a few years when the expanding ejecta hits the circumstellar ring (e.g., Borkowski, Blondin, \& McCray 1997). Perhaps one of the greatest mysteries about SN1987A is the mysterious bright companion spot that was observed by optical speckle interferometry (Nisenson et al.\ 1987, N87 hereafter; Meikle, Matcher, \& Morgan 1987, M87 hereafter) about one month after the SN1987A explosion, with a projected displacement from SN1987A of about 17 light days. Its close proximity to SN1987A, the fact that it was seen for only a few weeks, and its high brightness (about one-tenth of the brightness of SN1987A itself) make it certain that the spot was related to SN1987A itself. Several models were proposed soon after its discovery (Burrow \& Subramanian 1987; Rees 1987; Piran \& Nakamura 1987; Goldman 1987; Felten, Dwek, \& Viegas-Aldrovandi 1989) but close examination showed that there are formidable difficulties with all these models (Phinney 1988). Recently, there was an interesting development in the observations of gamma-ray bursts: the supernova 1998bw was observed (Kulkarni et al.\ 1998b) to coincide spatially and temporally with the gamma-ray burst GRB980425. This has led to suggestions that gamma-ray bursts (GRBs) and supernovae (SNe) may be related (Wang \& Wheeler 1998; Cen 1998). Energetics dictate that if SNe are responsible for producing GRBs, GRBs have to be beamed, that is, GRBs are jets from SNe. Independently but consistently, it is also required that the jets have a beaming angle of a few degrees in order to reconcile the high rate of SN events with the low rate of GRB events. The pressing question that arises then is how to test this scenario, where the vast majority of SN jets would travel laterally and would not be seen as GRBs due to the small beaming angle. It is the goal of this {\it Letter} to examine the properties of such lateral jets, suggesting that the observed bright companion spot of SN1987A may be caused by such a jet from SN1987A. \section{A Possible GRB Jet from SN1987A} The bright SN1987A companion spot was observed independently by two groups (N87; M87). It was observed at H$_\alpha$ and several other optical wavelengths using speckle interferometry by the CfA group (N87) on days 30 and 38 after the SN1987A explosion at a separation of $0^{''}.059\pm 0^{''}.008$ from SN1987A. Adopting a fiducial value of $50$kpc for the distance to SN1987A (Panagia et al.\ 1991; Gould 1995; Sonneborn et al.\ 1997; Lundqvist 1999), one obtains a perpendicular separation of $r_\perp=17~$light-days. Assuming that the spot was due to an ultra-relativistic jet leaving SN1987A at the time of the explosion, it gives a travel time $\Delta t=34~$days and yields an apparent perpendicular velocity of $v_\perp=0.5c$ ($c$ is the speed of light). Because $v_\perp=c \sin\theta/(1+\cos\theta)$, where $\theta$ is the angle between the jet direction and the observer-SN1987A vector, one finds $\theta=53^\circ$. Thus, if the spot was due to the working surface of a relativistic jet, the jet was a receding one! The spot detected by M87 on day 50 at a separation $0^{''}.074\pm 0^{''}.008$ is fully consistent with the observations of N87 for a jet traveling at near the speed of light. Interestingly, new image reconstructions from the CfA speckle data show possible indications of a second, weaker jet, with a larger separation, on the opposite side of the SN1987A (Nisenson \& Papaliolios 1999). Although working surface models were disfavored earlier (Phinney 1988), in light of this new observation of a counter jet and possible association of supernovae with GRBs (see \S 1), it seems worthwhile to re-examine this type of models in the context of GRB jets. Let us now examine the spectral properties of an ultra-relativistic GRB jet (Cen 1998). The jet can be characterized by its initial equivalent isotropic energy $E_{iso}$, initial coasting Lorentz factor $\Gamma_i$ and opening solid angle $\Omega$. For the current analysis only an external shock model (Rees \& M\'esz\'aros 1992) is considered for the jet. The reverse shock is not considered). It is assumed that the external shocked electrons have a power-law distribution function: \begin{equation} N(\Gamma_e)d\Gamma_e=A(t)\Gamma_e^{-p}d\Gamma_e, \end{equation} \noindent where $\Gamma_e$ is the Lorentz factor of electrons in the jet comoving frame, and $A(t)$ is a coefficient (to be determined) that is assumed to be a function of time only. Time $t$ measured in the burster frame is used as the time variable to express various quantities in the derivations, but the final results are converted to be shown using observer's time. We will only consider synchrotron radiation from the shock heated electrons. For the analysis below we will assume that $p>1$ (Tavani 1996) so the integral of equation (1) is convergent at the high end. We set \begin{equation} \int_{\Gamma_{e}}^\infty N(\Gamma_e^\prime)d\Gamma_e^\prime=\Omega r^2 c n t_{cool}\Gamma(r), \end{equation} \noindent where $n$ the number density of the external medium into which the shock is propagating, $r$ is the distance of the shock from SN1987A ($r$ and $t$ are used interchangeably throughout the paper assuming $r=ct$) and $t_{cool}$ is the electron cooling time (see equation [4]). Equation (2) is equivalent to stating that the number of electrons with $\Gamma>\Gamma_{e}$ at time $t$ is the number of electrons that have been shocked within the last $t_{cool}$ time interval, and earlier shocked electrons have cooled to lower energies. The last factor $\Gamma(r)$ on the right hand side of equation (2) accounts for the time boost of a moving object. Integrating equation (2) yields \begin{equation} A(t) = (p-1)\Omega r^2 c n t_{cool} \Gamma_{e}^{p-1}(r) \Gamma(r)\ . \end{equation} \noindent The synchrotron cooling time measured in the comoving frame for an electron with $\Gamma_{e}$ is \begin{equation} t_{cool}={\Gamma_{e} m_e c^2\over P_{e}}, \end{equation} \noindent The majority of the freshly shocked electrons (as we will adopt $p\sim 6$) have a Lorentz factor \begin{equation} \Gamma_{e}(r)=\Gamma(r){m_p\over m_e}\xi_e, \end{equation} \noindent where $\Gamma(r)$ is the shock Lorentz factor, $m_p$ and $m_e$ are proton and electron mass and $\xi_e$ is an equipartition parameter (Waxman 1997). The synchrotron radiation power, $P_{e}$, for an average electron with $\Gamma_{e}$ in a randomly directed magnetic field $B$ is (Blumenthal \& Gould 1970): \begin{equation} P_{e}={4\over 3} \sigma_T c \Gamma_{e}^2 {B^2\over 8\pi}, \end{equation} \noindent where $\sigma_T=6.6\times 10^{-25}$cm$^2$ is the Thomson cross section. $B$ (Waxman 1997) is linked to the energy density of the postshock external nucleons, $4\Gamma(r)^2 n m_p c^2$, by \begin{equation} {B^2\over 8\pi} = 4\Gamma(r)^2 n m_p c^2 \xi_B, \end{equation} \noindent where $\xi_B$ is the equipartition parameter for the magnetic field. Now we may proceed to obtain the total emission. For the present purpose it is adequate to assume that the spectral emissivity of each electron is a delta function $P_\nu=P_e \delta (\nu-\nu_e)$, where $P_e$ can be expressed by equation (6), and the characteristic synchrotron radiation frequency $\nu_e$ for electrons with $\Gamma_e$ is (Rybicki \& Lightman 1979) \begin{equation} \nu_e = \Gamma_e^2 {e B\over 2\pi m_e c}\ . \end{equation} \noindent Multiplying equation (1) by $P_\nu$ and integrating over $\Gamma_e$, and using equations (3,4,6,7,8) give the total emission in the comoving frame \begin{equation} j(\nu,t) = {1\over 2} (p-1) \Omega r^2 n(r) m_e c^3 \Gamma_e(r)\Gamma(r)\nu_{e}^{-1}(r) \left({\nu\over\nu_e}\right)^{-{p-1\over 2}}\ . \end{equation} \noindent It is noted that the above expression for $j(\nu,t)$ is valid only above a lower cutoff frequency, $\nu_l$, since the total energy has to be finite. We observe the following simple ansatz to obtain $\nu_l$: the total radiation emitted during the time interval $t_{cool}$ (in the comoving frame) should not exceed the total energy input to the thermalized electrons during the same time interval, which translates to the following relation: \begin{equation} \int_{\Gamma_{e}}^\infty \Gamma_e^\prime m_e c^2 N(\Gamma_e^\prime)d\Gamma_e^\prime= t_{cool} \int_{\nu_{l}}^\infty j(\nu,t) d\nu. \end{equation} \noindent Integrating both sides of equation (10) and using equations (3,4,6,7,8) yield \begin{equation} \nu_l(t)=\left({p-2\over p-3}\right)^{2\over p-3} \nu_{e}(t), \end{equation} \noindent where $\nu_e$ is given by equation (8). Note that the derived $\nu_l(t)$ is slightly larger than $\nu_e(t)$. Below $\nu_{l}$, $j(\nu,t)$ scales as \begin{equation} j(\nu,t)=j(\nu_{l},t) ({\nu\over \nu_{l}})^{1/3}. \end{equation} \noindent Synchrotron self-absorption becomes important only at lower frequencies than those of interest here and is thus ignored in the present analysis. In order to compute $j(\nu,t)$ as a function of time, one needs to specify the circumstellar medium density distribution and the evolution of the bulk Lorentz factor of the shock. The standard steady wind model for the distribution of the circumstellar medium of a red supergiant is adopted: \begin{equation} \rho(r) = {\dot M\over 4\pi v_w r^2}, \end{equation} \noindent where $\dot M$ is the mass loss rate of the star and $v_w$ is the wind velocity. Using $\dot M=4\times 10^{-5}{\rm\,M_\odot}$yr$^{-1}$ and $v_w=10$km/s, as inferred from analysis of SN 1993J (Fransson, Lundqvist, \& Chevalier 1996) yields \begin{equation} n(r) = \left({r\over r_0}\right)^{-2} \hbox{atoms/cm$^3$} \end{equation} \noindent with $r_0=1.1\times 10^{19}\hbox{cm}$. This adopted density distribution is in fact quite consistent with the measured circumstellar density of SN1987A (e.g., Sonneborn et al.\ 1998). It is assumed that radiative losses are small, which is appropriate at the later times of the fireball evolution of interest here. Then, for our adopted $\rho(r)$, we find the following scaling solution for $\Gamma(t)$ (Blandford \& McKee 1976) \begin{equation} \Gamma(t)=\Gamma_i(t/t_{dec})^{-1/2} \end{equation} \noindent for $t>t_{dec}$. For $t\le t_{dec}$ we simply set $\Gamma(t)=\Gamma_i$. The transition time $t_{dec}$, measured in the burster frame, is set to be that when the mass of the swept-up circumstellar medium is equal to $1/\Gamma_i$ of the initial fireball rest mass, yielding \begin{equation} t_{dec}={E_{iso}\over 4\pi m_p \Gamma_i^2 c^3 r_0^2}. \end{equation} \noindent The flux density (in units of erg~cm$^{-2}$~s$^{-1}$~Hz$^{-1}$) of the jet at the observer at observed frequency $\nu_{obs}$ at observer's time $t_{obs}$ is (Blandford \& Konigl 1979) \begin{equation} S_\nu (\nu_{obs},t_{obs}) ={1\over 4\pi d_{SN}^2} j\left({\nu_{obs}\over D},{t_{obs}\over 1+\cos \theta}\right) D^3\left({t_{obs}\over 1+\cos\theta}\right), \end{equation} \noindent where $D(t)\equiv (1+\beta\cos\theta)^{-1}\Gamma^{-1}(t)$ is the Doppler factor of the moving surface. Flux density is then converted to magnitude to compare with observations. Figure 1 shows the magnitudes of the jet at $6560\AA$ (solid curve) and $4500\AA$ (dashed curve), as a function of time measured in the observer's frame, $t_{obs}$ [note $t_{obs}=t (1+\cos\theta)$]. Note that the open circle at day 98 is from a recent re-analysis of the observational data (Nisenson 1999). The observed points have been dereddened for extinction using the observed color excess $E(B-V)=0.19$ for SN1987A (Fitzpatrick \& Walborn 1990) and the extinction curve given by Seaton (1979). The following parameter values are used for the results shown in Figure 1: $\xi_e=1/3$, $\xi_B=1/4$, $p=6.0$, $E_{iso}=2\times 10^{54}$erg, $\Gamma_i=300$, $\Omega=1.5\times 10^{-3}~$sr, $\theta=53^\circ$ and $d_{SN}=50~$kpc. All the parameters used are characteristic of a supernova GRB jet proposed (Cen 1998) and are consistent with known GRB observations. Note that $E_{iso}=10^{54}$erg is capable of accounting for the most luminous GRBs observed (e.g., GRB971214, Kulkarni et al.\ 1998a). A detailed analysis of the jet in the context of a GRB and its afterglows will be given elsewhere. The GRB jet model fits the speckle observations of the spot at $6560\AA$ reasonably well over the entire period where observational data are available. However, the model appears to be too ``blue" in the sense, i.e., it appears to be too bright at shorter wavelengths. For example, the computed spot at $4500\AA$ appears to be too bright by about two magnitudes compared to the observed spot. While the model is consistent (not shown in the figure) with infared observations of SN 1987A (e.g., at $4.6\mu$m, Bouchet et al.\ 1987), it also appears to be too bright in the UV compared to the total flux of SN 1987A (e.g., $3100\AA$, Kirshner 1987) by about a factor of ten, consistent with Phinney (1988). Clearly, more work is needed to improve upon this simple model. One way to avoid excess flux at short wavelengths is to introduce a large, intrinsic color excess, say, $E(B-V)\sim 1.5$. The sharp turn near days 30-40 is due to the sharp turn in the spectrum at $\nu_l(t)$. The peak of the evolution of the jet brightness at a given wavelength corresponds to the epoch when $\nu(t) = \nu_l(t)$ and the sharp turn (to faint) of brightness of the jet at earlier times is primarily due to the fact that the $D^3$ term in equation (17) goes roughly as $t^{3/2}$ and $\nu_l$ increases rapidly with decreasing time (roughly $\propto t^{5/2}$) combined with the spectral form of $\nu^{1/3}$ below $\nu_l$. The evolution of the brightness of the jet past the peak is primarily determined by the combined effect of the evolution of $\nu_{l}$ and $p$. The quantity $p$ is well constrained by the observed evolution of the optical spot. We find that $p\sim 6$ is required in order to provide an acceptable fit to the observed optical spot. A larger $p$ ($>7$) would produce too steep a decline around $t_{obs}\sim 30~$days. A smaller $p$ ($<5$) would produce a flat to rising temporal evolution and is inconsistent with the observation, i.e., the spot should have been visible longer. This ``counterspot" on the opposite side (Nisenson \& Papaliolios 1999) has an apparent separation of $0^{''}.16$ at the same time when the first spot was seen, giving an apparent {\it superluminal} perpendicular velocity of $v_{\perp, second}=1.36c$. If one assumes that this weaker jet was in the exact opposite direction from the first (i.e., $\theta_{second}=180-\theta=127^\circ$), it is required that $v_{second}=0.84c$. However, due to the uncertainties in $d_{SN}$, it is possible that $v\sim c$ may be allowed for both jets. It is interesting and should be emphasized that the two jets have unequal strengths, a prediction of the model proposed by Cen (1998) to account for the asymmetrical natal kick of neutron stars (pulsars). The asymmetrical pair of jets would induce star-recoil with the induced bulk velocity of the star being $650 (m_{star}/10M_\odot)^{-1}~$km/s (Cen 1998), which moves about $0^".03(m_{star}/10M_\odot)^{-1}$ in ten years (using $d_{SN}=50$kpc). This effect might be observable by detecting a shift of the position of the neutron star/pulsar or the centroid of the debris (Garnavich 1999). Based on available debris data (Haas et al.\ 1990; Spyromilio, Meikle, \& Allen 1990; Jennings et al.\ 1993; Utrobin, Chugai, \& Andronova 1995; Wang et al.\ 1996), if seems that the debris does {\it not} share the recoil movement of the star but shares the movement of the jet. \section{Conclusion} It is shown here that the bright companion spot of SN1987A may be due to a receding ultra-relativistic jet traveling at $\sim 53^\circ$ to the observer-to-SN1987A vector, through a circumstellar medium with a stellar wind like density $\rho(r)\propto r^{-2}$. The model provides an adequate explanation for the evolution of the observed optical companion spot, at least energetically, although more modeling is required to produce a satisfactory color of the spot. The parameters for the jet are characteristic of or required by the observed GRBs (with $E_{iso}=2\times 10^{54}$erg, $\Gamma_i=300$) with an openning angle of a few degrees. If the jet traveled towards us along the line of sight, a very bright GRB would be seen with an inferred isotropic energy of $\sim 10^{54}~$erg. If this model is correct, it implies that at least some GRBs would be seen as going through a medium with density $\rho(r)\propto r^{-2}$, rather than a uniform density medium. It is urgent to systematically search for GRB-supernova associations or supernova-jet associations in order to test this hypothesis. \acknowledgments I thank Arlin Crotts, Dick McCray and Pete Nisenson for stimulating discussions, Jeremy Goodman for suggesting looking for evidence of lateral GRB jets and Micheal Strauss for carefully reading the manuscript. I want to thank the second referee, Jim Felten, for working tirelessly to help improve the paper. The work is supported in part by grants NAG5-2759 and AST93-18185, ASC97-40300.
\section{Introduction} \label{intro} Inflation is a compelling resolution to the cosmological puzzles, because it explains how a large class of initial states evolve into a unique final state that is consistent with our observed Universe. The background cosmology that defines inflation is a Friedmann-Robertson-Walker (FRW) cosmology in which the scale factor $R(t)$ has positive acceleration ${\ddot R}(t) > 0$. The importance of such a background cosmology for one of the cosmological puzzles, the horizon problem, has been appreciated for the longest time \cite{horizon,oldphaset} (for more details please see \cite{olive}). Later, Guth in his pivotal paper \cite{guth} pointed out the importance of this background cosmology, which he named inflation, to solving another cosmological puzzle, the flatness problem. Moreover, Guth's paper drew attention to the relevance of cosmological phase transitions in attaining a dynamical particle physics explanation for inflation. These ideas were foreshadowed by some earlier works \cite{oldphaset}. {}Furthermore, most of these ideas about cosmological phase transitions as well as suggestions about their importance in explaining the cosmological puzzles were initially developed by Kirzhnits and Linde \cite{kl}. The thermal field theory and its application to cosmological phase transitions was considered further by Weinberg \cite{weinberg} and Dolan and Jackiw \cite{dj}. These early ideas all converged into a single benchmark concept about inflationary dynamics, the slow-roll scalar field scenario \cite{ni,ci}. This picture postulates the existence of a scalar field $\phi({\bf x},t)$, named the inflaton, which governs inflationary dynamics. At some early time, it is assumed that the energy density $\rho$ and pressure density p were dominated by the homogeneous component of the inflaton $\varphi_0(t)$, where, in addition, $\varphi_0(t)$ is assumed classical, with \begin{equation} \rho = V(\varphi_0) + \frac{1}{2} {\dot \varphi}_0^2 + \rho_r \label{rho} \end{equation} and \begin{equation} p = -V(\varphi_0) + \frac{1}{2} {\dot \varphi}_0^2 + \frac{1}{3} \rho_r. \label{p} \end{equation} $V(\varphi_0)$ and ${\dot \varphi}_0^2/2$ are respectively the potential and kinetic energy of the inflaton and $\rho_r$ is a component of radiation energy density. The key observation for inflationary dynamics is when the potential energy dominates, $V(\varphi_0) > {\dot \varphi}_0^2/2, \rho_r$, the equation of state from Eq. (\ref{rho}) and (\ref{p}) becomes that of the vacuum $\rho \approx -p$, and this is the required condition in FRW cosmology for inflationary expansion. In the slow-roll inflationary scenario, considerations from observation generally require the stronger condition $V(\varphi_0) \gg {\dot \varphi}_0^2/2$. This condition along with the assumption that the potential is slowly varying implies the equation of motion for $\varphi_0$ is first order in time. The conditions on $\rho_r$ lead to two different types of thermodynamic regimes of inflation. {}For $\rho_r \approx 0$ expansion is isentropic during inflation, so that the Universe rapidly becomes supercooled. In this supercooled inflation regime, the termination of the inflationary period into a radiation dominated period occurs through a short intermediate reheating period, in which all the vacuum energy of the inflaton is converted into radiation energy. Alternatively, for $V(\varphi_0) \equiv \rho_v > \rho_r >0$, expansion is non-isentropic during inflation, so that the temperature of the Universe may still be sizable. In this warm inflation regime, conversion of vacuum energy into radiation energy occurs throughout the inflationary period and the inflationary regime smoothly terminates into a radiation dominated regime without an intermediate reheating period. Warm inflation cosmology, i.e., inflationary dynamics without reheating, was formulated in \cite{wi,ab1}. An earlier paper by Fang and the author \cite{bf2} developed some underlying ideas, although the statement of that paper is general to all of inflationary cosmology. The demonstration that non-isentropic expansion, the background cosmology of warm inflation, can be realized in FRW cosmology from a plausible field theoretic dynamics was given in \cite{ab2,rudnei,gmn}. Although the scale factor dynamics of the background cosmology in both supercooled and warm inflation regimes is similar, the microscopic scalar field dynamics in the two regimes is very different. In warm inflation dynamics, during the inflation period the inflaton interacts considerably with other fields. These interactions permit energy exchange and this is how the vacuum energy of the inflaton is converted into radiation energy. As the fields acquire the vacuum energy liberated by the inflaton and become excited, their reaction back on the inflaton damps its motion. To realize a viable inflationary regime, the inflaton must support the vacuum energy sufficiently long to solve the horizon/flatness problems, $N_e \stackrel{>}{\sim} 60$. In warm inflation this implies the reaction of all the fields on the inflaton must be sufficiently strong to overdamp its motion. Thus slow-roll motion in warm inflation is synonymous with overdamped motion. {}Furthermore, the effective dynamics of the inflaton is analogous to time dependent Ginzburg-Landau scalar order parameter kinetics. Such a kinetics has been derived from a near-thermal-equilibrium quantum field theory formulation \cite{bgr}. In this derivation, the classical inflaton is defined as the thermal expectation value of the inflaton field operator $\phi({\bf x},t)$, $\varphi({\bf x},t) \equiv <\phi({\bf x},t)>_{T}$. In supercooled inflation dynamics, the inflaton typically is modeled as noninteractive with other fields during the inflation period. Supercooling implies the Universe is in the ground state $|0>$ during inflation. It is assumed that the vacuum is translationally invariant. The classical inflaton that emerges in the slow-roll equation and energy-pressure densities, Eqs. (\ref{rho}) and (\ref{p}), is defined as the expectation value of the inflaton field operator with respect to $|0>$, $\varphi({\bf x},t) \equiv <0|\phi({\bf x},t)|0>$. Several authors \cite{guthpi2,lyth,halliwell} have argued on the basis of saturating the momentum-position uncertainty principle that the quantum equations for $\phi$ go over into classical equations for a given inflaton mode of comoving momentum ${\bf k}_c$, $\phi_{{\bf k}_c}(t)$, once the physical momentum associated with the mode ${\bf k}_p \equiv {\bf k}_c/R(t)$ crosses the Hubble radius $2\pi/{\bf k}_p > 1/H$. The slow-roll scalar field scenario has a second aspect to it. Small fluctuations of the inflaton about its homogeneous component provide the initial seeds of density perturbations \cite{guthpi1,amp}. These density perturbations produced during inflation evolve into the classical inhomogeneities observed in the Universe. {}For warm inflation dynamics, the fluctuations of the inflaton are thermally induced. As such, these initial seeds of density perturbations already are classical upon inception. In supercooled inflation dynamics, the fluctuations of the inflaton arise from zero-point quantum fluctuations and so are purely quantum mechanical. In this case, it must be explained how these initial quantum fluctuations evolve into the classical inhomogeneities observed in the Universe. This problem often is referred to as the quantum-to-classical transition problem of supercooled inflation. Early resolutions to this problem followed the same reasoning as for the homogeneous component. Based on uncertainty principle arguments for the field amplitude and its conjugate momenta, these arguments concluded that the fluctuations can be treated classically once the physical wavelength of a give mode is larger than the Hubble radius \cite{guthpi2,lyth,halliwell}. Sasaki noted \cite{sasaki} that this criteria for classical behavior is not invariant under canonical transformations. Thus unless a definite physical significance could be ascribed to the field amplitude and its conjugate momenta, such arguments are insufficient for explaining the classical realization of the quantum process. The complete resolution to the quantum-to-classical problem in supercooled inflationary dynamics has been understood to occur via a process which eliminates quantum interference between macroscopically distinguishable events. Such a process generally is termed decoherence \cite{decoher}. A common way to introduce decoherence is through an external environment with which the inflaton interacts. Various studies have examined decoherence along these lines in which external fields couple to the inflaton \cite{branddec} and even in which the short wavelength modes of the inflaton act to decohere the long wavelength modes \cite{nambu}. {}For warm inflation, since dissipation is a fundamental aspect of the dynamics, decoherence is an automatic consequence. So warm inflation contains a good example of decoherence within an inflationary dynamics. {}For supercooled inflation, decoherence is not a natural requirement of the dynamics, but rather it must be imposed an an additional condition. Thus to the extent of relevance of decoherence, warm inflation realizes the full consequences of it both to yield a classical dynamics and to exploit the fluctuation-dissipation effects associated with it. In the basic warm inflation picture, the only requirement is that radiation energy is produced during inflation and that the mechanism of production is via dissipative effects on the inflaton. Up to now, the only quantum field theory realization of this picture is when the radiation energy is near thermal equilibrium \cite{bgr,bgr2}. This is an interesting regime for developing a warm inflation dynamics, since it is the best understood regime of nonequilibrium quantum field theory. {}For the most part, there is limited understanding of nonequilibrium quantum field theory. However near thermal equilibrium, the advantage is that the state of the system can be studied as a perturbation about the thermal equilibrium state. {}Furthermore, if all macroscopic motion is slow relative to the relevant microscopic time scales, the macroscopic dynamics can be treated adiabatically. {}Finally, if the interactions are weak, perturbation theory is applicable. In such a thermalized, adiabatic, perturbative regime, a well formulated quantum field theory dissipative dynamics can be formulated. The foundations for this were developed by Kubo \cite{kubo} and Zubarev \cite{zubarev}, who basically examined the full dynamical consequences of the near-thermal-equilibrium fluctuation-dissipation theorem \cite{fdt}. Quantum mechanical models that implement the fluctuation-dissipation theorem within a dissipative dynamics have been known for a long time \cite{mag}, although in recent times such models generically have been termed Caldeira-Leggett models \cite{cl,kac}. They have been studied in relativistic models by various authors \cite{clstudies}. Within a realistic scalar quantum field theory model, Hosoya and Sagagami \cite{hs1} initially formulated dissipative dynamics. In their formulation, the near-thermal-equilibrium dynamics is expressed through an expansion involving equilibrium correlation functions. Although their formulation is physically transparent, formally it is cumbersome. Subsequently Morikawa \cite{morikawa} formulated the same problem in terms of an elegant real time finite temperature field theory formalism. These works were developed further by Lawrie \cite{lawrie} and Gleiser and Ramos \cite{gr}. {}For warm inflation, the overdamped regime of the inflaton is the one of interest. A realization of overdamped motion within a quantum field theory model based on this formulation was obtained in \cite{bgr}. An application of a near-thermal-equilibrium fluctuation-dissipation dynamics for warm inflation first was examined from a Caldeira-Leggett type model \cite{ab1}. {}For a realistic quantum field theory model, the results of \cite{bgr} were applied to warm inflation in \cite{bgr2} and a solution to the horizon/flatness problem was presented. The warm inflation solution in \cite{bgr2} was based on a specific quantum field theory model that generally has been termed distributed mass models (DM-models) \cite{bgr,abpas}. In this model, the inflaton interacts with several other fields through shifted couplings $g^2(\varphi_0-M_i)^2 \chi_i^2$ and $g(\varphi_0-M_i) {\bar \psi}_i \psi_i$ to bosons and fermions repectively. The mass sites $M_i$ are distributed over some range. As the inflaton relaxes toward its minimum energy configuration, it decays into all fields that are light and coupled to it. In turn this generates an effective viscosity. In order to satisfy the e-fold requirement of a successful inflation, $N_e > 60$, overdamping must be very efficient. The purpose of distributing the masses $M_i$ is to increase the interval for $\varphi_0$ in which light particles emerge through the shifted couplings. Within this simple near-thermal-equilibrium quantum field theory formulation, the distribution of mass sites appears necessary to sustain the inflaton's overdamped motion sufficiently long to satisfy the e-fold requirement. On the one hand, the DM-model may be regarded as an intermediate step towards realistic warm inflation models. On the other hand, the hierarchy of mass sites $M_i$ in this model is suggestive of mass levels of a fundamental string. Based on this hypothesis, it was shown in \cite{bk} that DM-models can be obtained from effective supersymmetric theories. {}Further development of the string interpretation of DM-models is given in \cite{bk} and \cite{bk2}. In this paper we study a wide range of warm inflation solutions for the DM-models, that extends the single case studied in \cite{bgr2}. The specifics of the extensions are clarified in the sections to follow. In addition, here the first estimates are given for thermally induced density perturbations \cite{bf2} in the DM-models. There is a second aspect to the present work, which perhaps is more important. It was mentioned above that the near-thermal-equilibrium quantum field theory formalism developed in \cite{hs1,morikawa,lawrie,gr} is unambiguously valid within a thermalized, adiabatic, perturbative regime. However in the basic formulation, the criteria for consistency are specified in terms of only a set of limiting inequalities (i.e., $\ll$, $\gg$). As such, these criteria are not specific about the extent to which the inequalities must be satisfied. In light of this, a convincing proof that a given dynamics is consistent with this formalism is if the solution space of interest exists for an arbitrary degree of validity for the consistency inequalities. {}For warm inflation dynamics, the solution space of interest is the regime of observational consistency with respect to expansion e-folds, $N_e \stackrel{>}{\sim} 60$, and density perturbations, $\delta \rho/\rho \stackrel{<}{\sim} 10^{-5}$ \cite{smoot} (for a review of the basic observational facts please see \cite{brandrev,olive,kolb}). In this paper, we show that the consistency inequalities of the underlying formalism are satisfied to an arbitrary degree and alongside this, an observationally consistent warm inflation regime always exists. In particular, as will be seen, the most restrictive consistency conditions involve the adiabaticity requirements. In the solutions, it will emerge that the degree of adiabaticity can be controlled by one parameter, $\alpha$, with adiabaticity improving as $\alpha \rightarrow 0$. What will be shown is in this limit, an observationally consistent warm inflation regime with respect to $N_e$ and $\delta \rho/\rho$ always exists. This result has a fundamental relevance, since it is an existence proof within quantum field theory for observationally consistent inflationary dynamics from start to finish: from an initial radiation dominated or inflationary regime, to an inflationary regime and finally into a radiation dominated regime. The paper is organized as follows. In Sect. \ref{sect2} the DM model Lagrangian is presented and useful results about its effective potential are reviewed. In Sect. \ref{sect3} the basic equations of warm inflation and the consistency conditions on the solutions are reviewed. Subsect. \ref{subsect3C} provides a convenient summary of all parameters, notation, and terminology used in this paper. In Sects. \ref{sect4} and \ref{sect5} the solutions for e-folds and density perturbations respectively are presented. Subsects. \ref{subsect4C} and \ref{subsect5C} discuss general features about the solutions in order to ease the effort to understand the calculations. This paper will not focus on phenomenological consequences of the model. In Sect. \ref{sect6} some example applications of the solutions are given. Subsect \ref{subsect6A} computes the dimensional scales of the relevant quantities in the cosmology. Subsect. \ref{subsect6B} analyzes this warm inflation model in the limit of arbitrary adiabaticity. Subsect. \ref{subsect6C} examines the solution's dependence on $\lambda$, the $\phi$ self coupling parameter. In the concluding section \ref{sect7}, first improvements to this calculation are discussed. Second, a particle/string physics interpretation of the model is discussed. {}Finally some concluding perspectives are given about the model and solutions. \section{Model} \label{sect2} Consider the following Lagrangian of a scalar field $\phi$ interacting with $N_M\times N_{\chi}$ scalar fields $\chi_{ik}$ and $N_M\times N_{\psi}$ fermion fields $\psi_{ik}$, \begin{eqnarray} {\cal L} [ \phi, \chi_{ik}, \bar{\psi}_{ik}, \psi_{ik}, \chi^r_{i}, \bar{\psi}^r_{i}, \psi^r_{i} ] & = & \frac{1}{2} (\partial_\mu \phi)^2 - \frac{m^2}{2}\phi^2 - \frac{\lambda}{4 !} \phi^4 - V_0 - V_1(\phi) \nonumber\\ & + & \sum_{i=i_{\rm min}}^{i_{\rm max}} \sum_{k=1}^{N_{\chi}} \left\{ \frac{1}{2} (\partial_\mu \chi_{ik})^2 - \frac{f_{ik}}{4!} \chi_{ik}^4 - \frac{g_{ik}^2}{2} \left(\phi-M_i\right)^2 \chi_{ik}^2 \right\} \nonumber \\ & + & \sum_{i=i_{\rm min}}^{i_{\rm max}} \sum_{k=1}^{N_{\psi}} \left\{ i \bar{\psi}_{ik} \not\!\partial \psi_{ik} - h_{ik} ( \phi - M_i ) \bar{\psi}_{ik} \psi_{ik} \right\} \nonumber\\ & + & \sum_{i=1}^{N_r} \left\{ \frac{1}{2} (\partial_\mu \chi^r_i)^2 - \frac{f^r_{i}}{4!} {\chi^r_{i}}^4 \right\} + \sum_{i=1}^{N_r/4} \left\{ i \bar{\psi}^r_{i} \not\!\partial \psi^r_{i} \right\} \: . \label{Nfields} \end{eqnarray} \noindent This model is described in \cite{bgr,abpas} and named the distributed mass model (DM model), since the interaction between $\phi$ with the $\chi_{ik}$ and $\psi_{ik}$ fields establishes a mass scale distribution for the $\chi_{ik}$ and $\psi_{ik}$ fields, which is determined by the mass parameters $\{M_i\}$. The finite temperature effective potential for this model was computed in \cite{bgr,bgr2} for a nonzero homogeneous field amplitude $\langle \phi ({\bf x},t) \rangle_T \equiv \varphi_0(t)$. Because of the shifted coupling arrangement, the self-coupling parameter $\lambda$ is not corrected by $\chi - \phi$ and $\psi-\phi$ interactions. However, these interactions will generate new terms $\propto (\phi-M_i)^n$ and they are discussed in the next section. The finite temperature, field dependent mass are \begin{equation} m_{\chi_{ik}}^2 (\varphi_0,T) = g^2(\varphi_0-M_i)^2 + \mu_{\chi_{ki}}^2(T), \label{rmchi} \end{equation} \begin{equation} m_{\psi_{ik}}^2 (\varphi_0,T) = [g(\varphi_0-M_i) + \mu_{\psi_{ki}}(T)]^2 \label{rmpsi} \end{equation} and \begin{equation} m_{\phi}^2(\varphi_0,T) = m^2 +\frac{\lambda \varphi_0^2}{2} +\mu_{\phi}^2(T). \label{rmphi} \end{equation} $\mu(T)$ are thermal mass corrections with $\mu_{\phi}(T) \sim \sqrt{\lambda}T$ and $\mu_{\chi,\psi}(T) \sim gT$. These thermal masses are nonzero only when the field dependent contribution to the respective particle's mass is below the temperature scale. Note that the $\chi_{ik}$ and $\psi_{ik}$ effective field-dependent masses, $m_{\chi_{ik},\psi_{ik}} (\varphi_0,T)$, can be constrained even when $\langle \phi \rangle = \varphi_0$ is large. The $\phi\chi$, $\phi\psi$ interactions can be made reflection symmetric, $\phi \rightarrow -\phi$, but for our purposes we consider all $M_i >0$ and $\varphi_0>0$. The parameters $M_i$ will be referred to as mass sites. The $\chi$ and $\psi$ fields will be referred to as dissipative heat bath fields. {}For our calculations we will consider all coupling constants to be positive: $\lambda$, $f_{ik},g_{ik}^2, h_{ik}, f^r_i$ $> 0$, and for simplicity choose them to be $f^r_i=f_{ik}=f$, $g_{ik}=h_{ik}=g$. Also, we will set $N_{\chi}=N$, $N_{\psi}=N/4$, which implies an equal number of bose and fermi degrees of freedom at each mass site $M_i$. This relation on the degrees of freedom along with our choice of coupling implies a cancellation of radiatively generated vacuum energy corrections in the effective potential \cite{bgr2}. {}For convenience we will choose the mass scales to be evenly spaced as $M_i = i M/g$, $i=i_{\rm min}, \ldots, i_{\rm max} \equiv i_{\rm min} + N_M$, where $M$ is the mass splitting scale between adjacent sites. Here $M_{i_{\rm min}}$ and $M_{i_{\rm max}}$ bound the interval for $\varphi_0$ in which dissipative dynamics, thus warm inflation, is possible. The Lagrangian Eq. (\ref{Nfields}) also contains $N_r$ bosonic $\chi^r$ and $N_r/4$ fermionic $\psi^r$ fields that do not interact with the inflaton $\phi$. These fields represent additional degrees of freedom in the heat bath that otherwise do not contribute to the dissipative dynamics of $\phi$. These fields will be referred to as non-dissipative heat bath fields. Our choice of equal numbers of bose and fermi degrees of freedom for the non-dissipative heat bath fields is not necessary, but done here for notational convenience. The non-dissipative heat bath fields as written in the above Lagrangian are completely noninteracting with either $\phi$ and the dissipative heat bath fields coupled to $\phi$. Thus, in the form written, none of the energy liberated by the inflaton $\phi$ is transferred into the non-dissipative heat bath fields. To properly realize such heat bath fields, decay channels are necessary between them and the system of fields coupled to $\phi$. We will not attempt to correct this shortcoming but assume that it is possible. Up to the level of approximation in \cite{gr,bgr,bgr2}, which is what is applied here, this problem can be solved. Here, simply we want to explore the consequences of this possibility. As will be seen, observationally consistent warm inflations can be achieved with no non-dissipative heat bath fields $N_r=0$, and in many cases this is the optimal situation. However, as also will be seen, interesting possibilities arise when there are a very large number of non-dissipative heat bath fields. The Lagrangian Eq. (\ref{Nfields}) has two extensions to the one in \cite{bgr2}. {}First the above Lagrangian allows for an overall constant shift in the vacuum energy $V_0$. Second, an additional term has been added to the interaction potential, $V_1(\phi)$. This additional term is defined to be zero within the interval of $\varphi_0$ in which we study warm inflation, $M_{i_{\rm min}} < \varphi_0 < M_{i_{\rm max}}$. However outside this region, $V_1(\phi)$ will be suitable to permit an absolute minimum of the $\varphi$-efffective potential at zero potential energy. In the calculations to follow, one case is treated where $V_0$ and $V_1(\phi)$ are nonexistent, $V_0=0$ and $V_1(\phi)=0$ for all $\phi$. A few comments are in order to motivate the extensions $V_0$ and $V_1(\phi)$. The focus of this study is to understand the quantum field theory dynamics that underlies warm inflation. In this respect, the extensions $V_0$ and $V_1(\phi)$, permit exploration of warm inflation dynamics in a variety of interesting regimes. In this paper, we will not attempt to motivate these extensions from particle physics. However, note that the effect of $V_0$ and $V_1(\phi)$ is similar to a plateau region that could arise from higher polynomial or non-polynomial potentials. Non-polynomial potentials can occur in particle physics from nonperturbative effects. {}For example, nonperturbative mechanisms for dynamical SUSY breaking involving instantons can yield non-polynomial potentials (for a review please see \cite{shifvan}). {}Finally note that the basic results in this paper do not rely on these extensions $V_0$ and $V_1(\phi)$. A warm inflation regime consistent with quantum field theory and observational requirements on e-folds and density perturbations can be obtained without $V_0$ and $V_1(\phi)$, i.e. $V_0=V_1(\phi) =0$. This case is treated in Subsects. \ref{subsubsect4B2} and \ref{subsubsect5B2}. \section{Statement of the Problem} \label{sect3} Let $t=t_{BI}=0$ define the time when warm inflation begins ($BI$) and $t=t_{EI}$ define the time when warm inflation ends ($EI$). The thermodynamic state of the Universe during warm inflation is given by the time dependent temperature $T(t)$. The results presented in the next sections require two temperature scales $T_{BI} \equiv T(t_{BI}=0)$ and $T_{EI} \equiv T(t_{EI})$, the temperatures of the universe at the beginning and end of warm inflation respectively. These two temperatures will be expressed through the ratios \begin{equation} \beta \equiv \frac{T_{EI}}{T_{BI}} \label{betadef} \end{equation} and \begin{equation} \kappa_M \equiv \frac{T_{BI}}{M}, \label{kapdef} \end{equation} where $M$, defined in Sect. \ref{sect2}, is the splitting scale between adjacent mass sites in the DM model. \subsection{Basic Equations} \label{subsect3A} The basic equation for warm inflation dynamics is an effective equation of motion for the scalar field $\varphi_0$, which in the warm inflation regime describes overdamped motion. This equation has been derived for DM-models in \cite{bgr2} with details about the bosonic disspative term in \cite{bgr} and the fermionic dissipative term in \cite{yl}. Intuitive explanations of the formalism are given in \cite{hs1,yl}. In our calculations, the contribution from fermionic dissipation is ignored. The only effect considered from fermions is their contributions to the temperature dependent $\varphi$-effective potential. The most important term they contribute is their $T=0$ radiatively generated vacuum energy term, since, due to our choice of bose-fermi degrees of freedom, it cancels a similar term from the bose sector. The high-T modifications from the fermion fields are not essential for the overall consistency of the calculation and they also do not cause any new problems. We will account for these contributions in the $\varphi$-effective potential. Had we accounted for the fermion effects to dissipation, the damping of $\varphi_0$ would increase, which in turn would increase the robustness of warm inflation solutions. However in our case, where there are one-forth the number of fermion versus bosonic fields, for the most part, based on the high-T expressions for the dissipative terms \cite{bgr,bgr2,yl}, their contribution would enhance disspation by only 20 percent. We believe one can configure situations in which the fermion versus bose contribution to dissipation dominates, which is interesting since the functional behavior of the two types of dissipation differ. This is another possibility to examine, especially since the analysis in \cite{yl} found greater success with fermionic dissipation. However here we will not pursue that direction. {}For all the cases studied below, the basic $\varphi_0$ effective equation of motion is from \cite{bgr,bgr2} \begin{equation} \sum_i^{t.e.}(\varphi_0-M_i)^2 \eta_{1i}^B(T) \frac{d \varphi_0}{dt} = - \frac{\partial V(\varphi_0,T)}{\partial \varphi_0}, \label{eom1} \end{equation} where $t.e.$ means sum over all thermally excited sites. Everything multiplying $d\varphi_0/dt$ will be referred to as the dissipative coefficient and $\eta_{1}^B(T)$ will be referred to as the dissipative function. Here $V(\varphi_0,T)$ is the effective potential for $\varphi_0$. Both the dissipative coefficient and $\varphi$-effective potential emerge upon integrating out the dissipative heat bath fields $\chi_{ik}$ and $\psi_{ik}$. For $\chi_{ik}$ - fields with $m_{\chi_i} < T$, the high-T expression for $\eta_{1i}^B$ is \begin{equation} \eta_{1i}^{B}(T) \stackrel{T \gg m_{\phi},m_{\chi_{ik}} }{\approx} \frac{\eta'Ng^4}{\pi T {\cal C}(g,f)}, \label{eta1B} \end{equation} where \begin{equation} {\cal C}(g,f) \equiv g^4\left[1-\frac{6}{\pi^2}{\rm Li}_2 \left(\frac{m_{\chi}(\varphi_0,T)}{m_{\phi}(\varphi_0,T)}\right) \right]+ \frac{f^2}{8} \approx g^4+\frac{f^2}{8} \label{calc} \end{equation} and $\eta'= 48 \ln(2T/\mu_{\chi})$. In an expanding background, the interaction of $\phi$ with the metric also yields a $3H{\dot \varphi}_0$ term in the $\varphi_0$-effective equation of motion. As discussed in \cite{bgr,bgr2}, this term is dropped in Eq. (\ref{eom1}) because the thermal dissipation term dominates. In order to be within an inflationary regime, $V(\varphi_0,T)$ must be vacuum energy dominated, so that $V(\varphi_0,T) \approx \rho_v(\varphi_0)$, where $\rho_v(\varphi_0)$ is the vacuum energy of the $\phi$-field, $\rho_v(\varphi_0) = V_0+\rho_{\phi}(\varphi_0)$, where $V_0$ is the vacuum energy density shift parameter and $\rho_{\phi}(\varphi_0) = V(\varphi_0,T=0)$. In this case, the rate of decay of vacuum energy is obtained as ${d \rho}_v/{dt} = V'(\varphi_0) {\dot \varphi_0}$. The energy balance equation describing the transfer of vacuum energy to radiation energy is \begin{equation} {\dot \rho}_r = -4H \rho_r - {\dot \rho}_v. \label{dotrex} \end{equation} In the warm inflation regime $\rho_r$ decreases slowly, so that to a good approximation Eq. (\ref{dotrex}) becomes \begin{equation} 4H \rho_r \approx - {\dot \rho}_v. \label{dotr} \end{equation} {}For the four cases considered in Sects. \ref{sect4} and \ref{sect5}, we examine the regime where the $\lambda \varphi_0^4$ term dominates the effective potential. \subsection{Consistency Conditions} \label{subsect3B} The validity of the basic equations of warm inflation given above requires four consistency checks, which we refer to as the thermalization, adiabatic, force and infra-red conditions. Below, these conditions are explained and exactly specified. The thermalization and adiabatic conditions both address the consistency of the macroscopic warm inflation equations with the underlying microscopic quantum field theory dynamics. Within the simple dissipative quantum field theory formulation in \cite{bgr}, the decay widths $\Gamma_{\phi}$, $\Gamma_{\chi}$, and $\Gamma_{\psi}$ characterize the time scale of microscopic dynamics. The basic self-consistency requirement is that all microscopic time scales are faster than all macroscopic time scales. Thus the slowest microscopic time scale is the important one for checking self-consistency. {}For dissipative dynamics, $\Gamma_{\phi}$ does not play a significant role. Since we are restricting to the bosonic contribution to dissipation \footnote{Since the fermion contribution to dissipation is being ignored, the behavior of $\Gamma_{\psi}$ is not very essential. If $\Gamma_{\psi} < \Gamma_{\chi}$, then the $\psi$-fields may not be thermally excited, which in turn would slightly modify the temperature dependent terms in the effective potential. However since the $\psi$-fields are free from the crucial self-consistency requirements for producing dissipative effects on $\varphi_0$, if $\Gamma_{\psi} < \Gamma_{\chi}$, is is easy to increase their number of decay channels, which thereby can increase $\Gamma_{\psi}$ to the level of $\Gamma_{\chi}$.}, $\Gamma_{\chi}$ is the relevant microscopic rate. {}From \cite{bgr}, the high temperature expression for it is \begin{equation} \Gamma_{\chi}(T) \stackrel{T \gg m_{\phi}, m_{\chi} }{\approx} \frac{\Gamma' T}{\pi} {\cal C}(g,f), \label{chidecay} \end{equation} where ${\cal C}$ is defined in Eq. (\ref{calc}) and $\Gamma' = 1/192$. {}For $\chi$-fields at site $i$, their decay rate is Eq. (\ref{chidecay}) when $m_{\chi_{ik}} \stackrel{<}{\sim} T$ and negligible when $m_{\chi_{ik}} \stackrel{>}{\sim} T$. Hereafter, rescaled time is used \begin{equation} \tau \equiv \Gamma_{\chi}(M) t. \label{ttau} \end{equation} The Hubble parameter at the beginning of warm inflation will be expressed as \begin{equation} H_{BI} = \alpha \beta \Gamma_{\chi}(T_{BI}) = \alpha \beta \kappa_M \Gamma_{\chi}(M), \label{hubdef} \end{equation} where the additional parameter $\alpha$ is introduced and $\beta$ and $\kappa_M$ are defined in Eqs. (\ref{betadef}) and (\ref{kapdef}) respectively. Warm inflation fundamentally has two macroscopic time scales, the Hubble expansion rate and the rate of change of the inflaton, which expressed in terms of rescaled time Eq. (\ref{ttau}) is ${\dot \varphi}_0 = \Gamma_{\chi}(d\varphi_0/d\tau)$. Self consistency requires the {\bf thermalization condition} \begin{equation} H(\tau) \ll \Gamma_{\chi}(T) \label{therm} \end{equation} and the $\varphi_0$-{\bf adiabatic condition} \begin{equation} \frac{1}{\varphi_0} \frac{d \varphi_0}{d\tau} \ll \frac{\Gamma_{\chi}(T)}{\Gamma_{\chi}(M)}. \label{padiab} \end{equation} {}For the cases in Sects. \ref{sect4} and \ref{sect5}, $H(\tau)$ changes little (less than a factor 5) from $T_{BI}$ to $T_{EI}$, so that $H_{BI}$ is its approximate scale throughout warm inflation. Therefore, the thermalization condition Eq. (\ref{therm}) requires $\alpha < 1$. A stronger adiabatic condition also requires that the $\varphi_0$ dependence in all Boltzmann factors varies slowly relative to the thermalization rate. The relevant thermal excitations in this model are the $\chi_{ki}$ and $\psi_{ki}$ fields when $m_{\chi_{ki},\psi_{ki}} <T$, for which the stringest form of the {\bf thermal-adiabatic condition} is \begin{equation} \frac{1}{T} \frac{d m_{\chi , \psi}}{d \tau} = \frac{g d\varphi_0 /d \tau}{T} \ll \frac{\Gamma_{\chi}(T)}{\Gamma_{\chi}(M)}. \label{thermadiab} \end{equation} The estimates to follow are based on this criteria, but it is much more restrictive than necessary for two reasons. {}First thermal excitation generally initiates once $m_{\chi,\psi} \stackrel{<}{\sim} 10T$. {}For the DM-model this implies mass sites have quite a long time to excite before their important role arises, which is when $m_{\chi,\psi} \stackrel{<}{\sim} T$ at which point they drive the viscosity. Second, within the thermally excited regime $m_{\chi,\psi} \stackrel{<}{\sim} T$, the Boltzmann factor is saturated at $\approx 1$, so is insensitive to mass variations. The next consistency check is the force condition. In our case this requires the $\lambda \varphi_0^3/6$ term to dominate the effective equation of motion. Due to vacuum energy cancellations between bosons and fermions, the one-loop $T=0$ radiative corrections \cite{dj,mgleiser,sher}, $\sim \sum_i^{all} g^4 N (\varphi_0-M_i)^4$, $\sum_{i}^{all}g^4 N (\varphi_0-M_i)^4 \ln[(\varphi_0-M_i)^2/\mu^2_{\rm renorm}]$, are eliminated \cite{bgr2}, where the sum here is over all mass sites.. The T-dependent force terms from the effective potential are 1. $(g^2 N/8) T^2 \sum_i^{t.e.}(\varphi_0-M_i)$, 2. $(g^3 N/(4 \pi)) T \sum_i^{t.e.}(\varphi_0-M_i)|\varphi_0-M_i|$, and 3. $(g^4 N/(64 \pi)) \sum_i^{t.e.}(\varphi_0-M_i)^3 \log(T/\mu_{\rm renorm})$. Our approximation makes a sharp division between thermally excited and unexcited sites. If the mass of a site $m_{\chi_{ik},\psi_{ik}}^2 \approx g^2(\varphi_0-M_i)^2 \leq T^2$, that site is thermally excited and for $m_{\chi_{ik},\psi_{ik}}^2>T^2$ it is thermally unexcited or cold. Here we assume $\varphi_0$ is always surrounded by sufficiently many mass sites on both sides so that there are no edge effects. There are two general features about all three types of force terms. {}First, the direction of the force flips for mass sites on opposite sides of $\varphi_0$ and this implies considerable cancellation in the summations. Second, all three forces are periodic in $\varphi_0$ with periodicity M. As $\varphi_0$ traverses any M-interval, it experiences an identical force profile in both directions. Thus the average force experienced by $\varphi_0$ from all three force terms is zero after traversing any M-interval. In our estimates, we use the most stringent consistency condition, which is to estimate the largest effect from these force terms and demand that the $\lambda \varphi_0^3/6$ term dominates it. We will find a large warm inflation solution space with this stringent condition. However estimates based on this force condition are overly conservative, since realistically $\varphi_0$ is not sharp, as we are treating it, but rather is smeared over some interval $\Delta \varphi_0$. Smearing would imply the average force experienced by the $\varphi_0$-packet diminishes due to the directionality dependence of the force terms. In fact if $\Delta \varphi_0 > M$, the force experienced by the wavepacket from these three force terms would be negligible. Nevertheless, for simplicity of the calculation, here this fact will not be treated and we will estimate the the upper bound of the three force terms for a sharp $\varphi_0$. Amongst the three force terms, note that 1 is the largest. Ignoring the premultiplying factors (which also is largest for 1), since for all thermally excited sites $g|\varphi_0-M_i| < T$, it implies $\sum_i^{te}(\varphi_0-M_i) T^2 > \sum_i^{t.e.} g (\varphi_0-M_i) |\varphi_0-M_i| T > \sum_i^{t.e.} g^2 (\varphi_0-M_i)^3$. Thus provided $\lambda \varphi_0^3/6$ dominates force term 1, it dominates them all\footnote{When all bosons and fermions are thermally excited, due to our choice of coupling and of ratio between bose and fermi fields, term 3, in fact, cancels amongst the bosons and fermions \cite{dj}.}. Approximating \begin{equation} g \sum_i^{t.e.} (\varphi_0-M_i) \approx T \end{equation} we obtain the {\bf force condition} \begin{equation} \frac{\lambda \varphi_0^3}{6} > \frac{g^2N}{8} \sum_i^{t.e.} (\varphi_0-M_i) T^2 \approx \frac{gN}{8} T^3 \label{forcec} \end{equation} The final consistency check is the infra-red condition. For the Minkowski space quantum field theory formalism used here to be valid, the Compton wavelength of all particle excitations should be much smaller than the Hubble radius $1/H(\tau)$ during inflation. {}For the $\chi_{ik}$,$\psi_{ik}$ fields, this condition is easily satisfied since their masses are generally $m_{\chi_{ik},\psi_{ik}} \approx T$ and $T \gg H$. This follows since for most of the time the $\varphi_0(\tau)$-induced portion of their masses is large. However, even when $\varphi_0(\tau) \approx M_i$, the thermal mass contribution $\sim gT$ generally is much larger than $H$. On the other hand, the $\phi$-mass is more concerning, since the self-coupling parameter $\lambda$ is usually tiny in warm inflation. Thus the essential constraint that must be checked to enforce the {\bf infra-red condition} is \begin{equation} m_{\phi}(\varphi_0,T) \gg H. \label{irc} \end{equation} In total, there are five consistency conditions. The thermal-adiabatic, $\varphi_0$-adiabatic and thermalization conditions all are adiabatic conditions. The latter, equivalently stated, requires the cosmological expansion rate to be adiabatic relative to the particle production rate. The force condition is not fundamentally required, but we have imposed it to simplify the $\varphi_0$ equation of motion. \subsection{Summary of the Parameters} \label{subsect3C} In Sects. \ref{sect4} and \ref{sect5} two cosmological parameters will be computed, the number of e-folds $N_e$ and amplitude of density perturbations $\delta \rho/\rho$, in terms of the microscopic and thermodynamic parameters of the model and one overall scale which will be the Planck scale $M_p$. Since the calculations are fairly detailed, for convenience here all definitions of parameters and related terminology are summarized. The following terminology is used in this paper: \begin{center} \begin{tabular}{p{1.5in} c p{4in}} dissipative heat bath fields & - & fields $\chi_{ik}$, $\psi_{ik}$ that are part of the heat reservoir and have significant effect on the dissipative dynamics of the inflaton. \\ non-dissipative heat bath fields & - & fields $\chi^r_i$, $\psi^r_i$ that only are part of the heat reservoir and have no significant effect on the dissipative dynamics of the inflaton. \\ mass site & - & ``location'' of dissipative heat bath fields in the DM model specified by the mass parameter $M_i \equiv M i/g$, with $i$ denoting the number of the site. \\ dissipative coefficient & - & the factor multiplying $d\varphi_0/dt$ in the $\varphi_0$ equation of motion, $\sum_i^{t.e.}(\varphi_0-M_i)^2 \eta_{1i}^B(T)$ \\ dissipative function & - & $\eta_1^B(T) = \eta'g^4/(\pi T {\cal C})$ with $\eta'$ and ${\cal C}$ defined below Eq. (\ref{eta1B}) \end{tabular} \end{center} The microscopic parameters have all been defined in detail in Sect. \ref{sect2}. Briefly, they are as follows: \begin{center} \begin{tabular}{p{1in} c p{4in}} N & - & number of $\chi_{ik}$ fields at every mass site $M_i$. Correspondingly the number of $\psi_{ik}$ fields at every mass site is $N/4$. \\ $N_r$ & - & number of nondissipative heat bath fields \\ $N_M$ & - & total number of mass sites crossed by the inflaton $\varphi_0$ during warm inflation. Ignoring small corrections at the two end points, this is equivalent to the total number of mass sites that were thermally excited at some point during warm inflation. \\ $\lambda$ & - & $\phi$ self coupling parameter \\ $g$ & - & $\phi$ - $\chi$ coupling $\sim g^2/2$ , $\phi$ - $\psi$ coupling $\sim g$ \\ $f$ & - & $\chi$ self coupling parameter \\ $V_0$ & - & vacuum energy density shift parameter \\ $M$ & - & splitting scale between adjacent mass sites $g|M_{i+1} - M_i| =M$ \\ $m_{\chi_{ik}, \psi_{ik}}$ & - & mass of $\chi_{ik}$ or $\psi_{ik}$ fields given in Eqs. (\ref{rmchi}), (\ref{rmpsi}) \\ $m_{\phi}$ & - & $\phi$ mass given in Eq. (\ref{rmphi}) \\ \end{tabular} \end{center} The thermodynamics of the warm inflation is expressed through the following quantities: \begin{center} \begin{tabular}{p{1.5in} c p{4in}} $T_{BI}$ & - & temperature at the beginning of warm inflation \\ $T_{EI}$ & - & temperature at the end of warm inflation \\ $\kappa_M$ & - & $T_{BI}/M$ \\ $\beta$ & - & $T_{EI}/T_{BI}$ \\ \end{tabular} \end{center} The cosmology is expressed through the following quantities: \begin{center} \begin{tabular}{p{1.5in} c p{4in}} $R$($\tau$) & - & scale factor, with $R(0) =1$ \\ $H_{BI}$ & - & Hubble parameter at the beginning of warm inflation, $H_{BI} \equiv H(\tau_{BI})$. In the four cases examined in this paper, the scale of $H(\tau)$ is of order $H_{BI}$ throughout the inflationary period. \\ $\alpha$ & - & $H_{BI} \equiv \alpha \beta \kappa_M \Gamma_{chi}(M)$. It will be seen that $\alpha$ is the adiabatic parameter with adiabaticity increasing as $\alpha \rightarrow 0$. \\ $\rho_v(\tau)$ & - & vacuum energy density $\equiv V_0 + \rho_{\phi}(\tau)$ where $\rho_{\phi}$ is the $\varphi_0$ dependent portion. \\ $\rho_r(\tau)$ & - & radiation energy density \\ $N_e$ & - & number of e-folds \\ $\delta \rho/\rho$ & - & amplitude of scalar density perturbation \\ \end{tabular} \end{center} Some additional notation used in this paper is as follows: \begin{center} \begin{tabular}{p{1.5in} c p{4in}} BI & - & begin warm inflation \\ EI & - & end warm inflation \\ $\Gamma_{\chi}(M)$ & - & decay width for $\chi$-fields when they are thermally excited, explicit expression in Eq. (\ref{chidecay}) \\ $\tau$ & - & rescaled time $\tau \equiv \Gamma_{\chi}(M) t$ \\ $n_{t.e.}$ & - & number of thermally excited ($t.e.$) mass sites \\ $\phi({\bf x}, \tau)$ & - & quantum inflaton field operator \\ $\varphi({\bf x}, \tau)$ & - & classical inflaton field $\langle \phi({\bf x}, \tau) \rangle_T \equiv \varphi({\bf x},\tau) =\varphi_0(\tau) + \delta \varphi({\bf x},\tau)$, where $\varphi_0(\tau)$ is the homogeneous background field \\ $y$ & - & slope parameter for $\varphi_0$. Since the evolution is overdamped, in all cases $d\varphi_0/d\tau \propto y \varphi_0$. \\ $k_F$ & - & freeze-out momentum for density perturbations, defined Eq. (\ref{kfcond}) \\ $\Delta \varphi^2_H(\tau)$ & - & amplitude of scalar field fluctuations, defined Eqs. (\ref{Dvp}) and (\ref{Dvplb}) \\ ${\bf k}_p, {\bf k}_c$ & - & physical ($p$) and comoving ($c$) wavenumber. By our convention they are the same at the end of warm inflation $\tau_{EI}$, ${\bf k}_p(\tau) \equiv {\bf k}_c/R(\tau -\tau_{EI})$. \\ $\delta \varphi_{{\bf k}_c}(\tau)$ & - & Fourier mode of the inflaton field $\delta \varphi({\bf x},\tau)$; equivalently this mode may be expressed as $\delta \varphi({\bf k}_p, \tau)$, where the relation between ${\bf k}_p (\tau)$ and ${\bf k}_c$ is specified. \\ \end{tabular} \end{center} {}Finally, for the following expressions, we simply state where they are defined: $\eta' = 48 \ln(2T/m_{\chi}) \approx 48$ - below Eq. (\ref{calc}); $\Gamma' \approx 1/192$ below Eq. (\ref{chidecay}); ${\cal C}$ - Eq. (\ref{calc}), also see first part of Sect. \ref{sect6}; $\Upsilon$ - Eq.(\ref{upsdef}). \section{Solutions for e-folds $N_e$} \label{sect4} In this section four warm inflation solutions are presented. The solutions we examine are for cases where $T_{BI} \gg T_{EI}$ in Subsect. \ref{subsect4A} and $T_{BI} \stackrel{>}{\sim} T_{EI}$ in Subsect. \ref{subsect4B}. We seek solutions in which $N_e$ e-folds of inflation occur with the temperature of the Universe $T_{BI}$ and $T_{EI}$ respectively at the beginning and end of warm inflation. The temperatures $T_{BI}$ and $T_{EI}$ are parameterized by $\beta$ and $\kappa_M$ in Eqs. (\ref{betadef}) and (\ref{kapdef}) respectively These conditions on the solutions imply two parameters of the model are determined in terms of the others. We choose these two quantities to be $\lambda$ and $\varphi_{BI}/M$. Under these conditions the solutions for $\varphi_0(\tau)$ and $T(\tau)$ are given as well as expressions for $\lambda$ and $\varphi_{BI}/M$. In addition, for ease of reference, expressions are written for $m_{\phi}^2/T^2$, $\rho_v(\tau)/\rho_r(\tau)$, $d\varphi_0 /d \tau$ and $\Gamma_{\chi}(\tau)/\Gamma_{\chi}(M)$. {}Finally the consistency conditions from Subsect. \ref{subsect3B} are evaluated. The calculation was designed so that the thermalization condition Eq. (\ref{therm}) is satisfied simply by requiring $\alpha < 1$. The adiabatic, force, and infra-red conditions result in two essential inequalities that restrict the self-consistent region of the parameter space. No overall scale is chosen in this section, since the results leave this choice completely arbitrary. Although all the calculations are very simple, due to the generality of the results, the final expressions may not appear transparent. It must be appreciated that despite the compactness of the final expressions, they contain the solutions under the very general situation in which the parameters can be varied over a wide range and in which four scales are adjustable, the initial temperature, the final temperature, the Hubble expansion rate and the duration of warm inflation. \subsection{$T_{BI} \gg T_{EI}$} \label{subsect4A} In this regime when the temperature from the beginning $T_{BI}$ to the end $T_{EI}$ of warm inflation changes significantly, the time dependence of the number of thermally excited sites is treated in the $\varphi_0$-equation of motion. One general parametric constraint for this regime is \begin{equation} \beta \ll 1. \end{equation} At temperature $T(\tau)$, any $\chi_{ik}$ fields with mass $m_{\chi_{ik}}^2 \approx g^2(\varphi_0-M_i)^2 < T^2$ is thermally excited. Due to our choice of spacings between mass sites $M_i$, this implies approximately $T(\tau)/M$ sites adjacent to the field amplitude $\varphi_0$ on either side are thermally excited. In our approximation for all sites with $m_{\chi_{ik}} > T$, the contribution to the dissipative dynamics is ignored. Eq. (\ref{eom1}) therefore becomes \begin{equation} \frac{\eta' \Gamma'}{\pi^2} N g^2 T^2 \frac{d \varphi_0}{d \tau} = - \frac{\lambda \varphi^3_0}{6}. \label{eomA} \end{equation} Also, we allow for an additional $N_r$ bosonic and $N_r/4$ fermionic non-dissipative heat bath fields that contribute to the radiation energy density but have no affect on the dissipative dynamics of the inflaton. Thus the radiation energy density at any instant during warm inflation is \begin{equation} \rho_r(\tau) = \frac{\pi^2}{16} [N_r + \frac{2NT(\tau)}{M}] T^4(\tau). \label{rhora} \end{equation} We examine the solutions for the cases $N_r=0$ and $N_r \gg 2T_{BI}/T_{EI}$ in Subsects. \ref{subsubsect4A1} and \ref{subsubsect4A2} respectively. {}For both cases, the vacuum energy density shift parameter will be $V_0 = \lambda \varphi^4_{BI}/24$, so that the vacuum energy is approximately constant throughout warm inflation. In this case the Hubble parameter, $H \equiv \sqrt{8\pi G \rho} \approx \sqrt{8 \pi G \rho_v}$, changes by an O(1)-factor during warm inflation and the number of e-folds is \begin{equation} N_e \approx H_{BI} t_{EI} = \alpha \beta \kappa_M \tau_{EI}. \label{ne} \end{equation} \subsubsection{\bf $N_r=0$} \label{subsubsect4A1} In this case there are no non-dissipative heat bath fields. \begin{center} {\bf Solutions} \end{center} Eq. (\ref{dotr}) and Eq. (\ref{eomA}) imply \begin{equation} T(\tau) = \left( \frac{\lambda^2}{18 \eta' \Gamma' \alpha \beta \kappa_M N^2 g^2 }\right)^{1/7} M^{1/7} \varphi_0^{6/7}(\tau). \label{tempa1t} \end{equation} Substituting into Eq. (\ref{eomA}), we find the solution \begin{equation} \varphi_0(\tau) = \frac{\varphi_{BI}}{(y \tau +1)^{7/2}}, \label{phia1} \end{equation} where $y=[\pi^2 2^{2/7}/(3^{3/7} 7 (\eta'\Gamma')^{5/7})] [(\alpha \beta \kappa_M)^2 \lambda^3/(N^3 g^{10})]^{1/7} (\varphi_{BI}/M)^{2/7}$. {}From Eq. (\ref{ne}) for $N_e$ and Eq. (\ref{betadef}) for $\beta$, \begin{equation} y=\frac{\alpha \beta^{2/3} \kappa_M}{N_e} (1-\beta^{1/3}). \end{equation} By equating this expression for $y$ with the one in terms of the parameters in the model, and applying Eq. (\ref{kapdef}) to eq. (\ref{tempa1t}), two parameters of the model are determined, of which we choose \begin{equation} \lambda= \left(\frac{1029 \eta^{'2} \Gamma^{'2}}{2 \pi^6}\right) \frac{\alpha^2 \beta \kappa_M (1-\beta^{1/3})^3 N g^4} {N_e^3} \label{lama1} \end{equation} and \begin{equation} \frac{\varphi_{BI}}{M}= (18\eta'\Gamma')^{1/6} \left(\frac{\alpha \beta N^2g^2}{\lambda^2}\right)^{1/6} \kappa_M^{4/3} =\left(\frac{\pi^2}{7}\sqrt{\frac{2}{\eta'\Gamma'}} \right) \frac{\kappa_M N_e} {\alpha^{1/2} \beta^{1/6} g (1-\beta^{1/3})}. \label{phima1} \end{equation} Using Eqs. (\ref{lama1}) and (\ref{phima1}), Eq. (\ref{tempa1t}) becomes \begin{equation} T(\tau) = \left( \frac{49 \eta' \Gamma'}{2 \pi^4} \right)^{3/7} \left(\frac{\alpha^{3/7} \beta^{1/7} \kappa_M^{1/7}(1-\beta^{1/3})^{6/7} g^{6/7}}{N_e^{6/7}} \right) M^{1/7} \varphi_0^{6/7}(\tau). \label{tempa1} \end{equation} Eqs. (\ref{phia1}) and (\ref{tempa1}) are the general solutions. Based on this solution, next some useful expressions are given. The number of sites that are thermally excited at a given instance is $n_{te} \equiv 2T(\tau)/M = 2\kappa_M/(y \tau+1)^3$. The number of mass sites $\varphi_0$ crosses during the warm inflation period is \begin{equation} N_M \equiv \frac{g|\varphi_{BI} -\varphi_{EI}|}{M} = \frac{\pi^2}{7}\sqrt{\frac{2}{\eta'\Gamma'}} \frac{\kappa_M N_e (1-\beta^{7/6})}{\alpha^{1/2} \beta^{1/6} (1-\beta^{1/3})} \end{equation} with \begin{equation} i_{\rm min} \equiv \frac{g \varphi_{EI}}{M} = \frac{\pi^2}{7}\sqrt{\frac{2}{\eta'\Gamma'}} \frac{\beta \kappa_M N_e}{\alpha^{1/2} (1-\beta^{1/3})}. \end{equation} The mass of $\phi$ is \begin{equation} {\bar m}_{\phi}^2 \equiv \frac{m_{\phi}^2}{T^2} \approx \frac{\lambda \varphi_0^2}{2T^2}= \left(\frac{21 \eta'\Gamma'}{2 \pi^2} \right) \frac{\alpha \beta^{2/3}(1-\beta^{1/3}) \kappa_M N g^2}{N_e (y \tau+1)}. \end{equation} The ratio of vacuum to radiation energy density is \begin{equation} \frac{\rho_{\phi}(\tau)}{\rho_r(\tau)} = \frac{2\beta^{1/3}N_e}{7(1-\beta^{1/3})}(y \tau+1) \label{rhorvA1} \end{equation} \begin{equation} \frac{V_0}{\rho_r(\tau)} = \frac{2\beta^{1/3}N_e}{7(1-\beta^{1/3})}(y \tau+1)^{15}. \end{equation} {}Finally, for examining the adiabatic condition \begin{equation} \frac{d \varphi_0}{d \tau} = \frac{7 \alpha \beta^{2/3} \kappa_M (1-\beta^{1/3})}{2N_e} \frac{\varphi_0(\tau)}{(y \tau+1)} = \left(\frac{\pi^2}{\sqrt{2\eta'\Gamma'}} \right) \frac{\alpha^{1/2} \beta^{1/2} \kappa_M}{g} \frac{T(\tau)}{(y \tau+1)^{3/2}} \label{dphia1} \end{equation} and \begin{equation} \frac{\Gamma_{\chi}(T)}{\Gamma_{\chi}(M)} = \frac{\kappa_M}{(y \tau+1)^3}. \label{therma1} \end{equation} \begin{center} {\bf Consistency Conditions} \end{center} The above solutions still are subject to consistency checks. {}For the $\varphi_0$-adiabatic condition Eq. (\ref{padiab}), since the thermalization rate Eq. (\ref{therma1}) decreases faster than $(d\varphi_0(\tau)/d\tau)/\varphi_0$ from Eqs. (\ref{phia1}) and (\ref{dphia1}), the most stringent test is at $\tau_{EI}$; if the $\varphi_0$-adiabatic condition holds at $\tau_{EI}$, then it is satisfied better at earlier times. We find the $\varphi_0$-adiabatic condition Eq. (\ref{padiab}) requires $7\alpha/(2N_e) < 1$. Since $\alpha<1$, for this to hold, it is sufficient that \begin{equation} N_e > 7/2. \end{equation} The thermal-adiabatic condition Eq. (\ref{thermadiab}) implies from Eqs. (\ref{tempa1}), (\ref{dphia1}) and (\ref{therma1}) that the most stringent time is $\tau_{EI}$ with the constraint \begin{equation} \frac{\pi^2}{\sqrt{2\eta'\Gamma'}} \alpha^{1/2} < 1. \label{thermaA1} \end{equation} By substituting for $\eta'$ and $\Gamma'$, it gives $\alpha < 1/199 $ (at $\tau=0$ the condition implies an additional factor of $\beta^{1/2}$ in Eq. (\ref{thermaA1}) ). The force condition Eq. (\ref{forcec}) implies \begin{equation} 4\sqrt{2\eta'\Gamma'} \alpha^{1/2} \beta \kappa_M > 1. \label{forcecA1} \end{equation} {}Finally the infra-red condition Eq. (\ref{irc}) implies \begin{equation} \frac{21 \eta'}{2 \Gamma'} \frac{\beta (1-\beta^{1/3}) \kappa_M N g^2} {\alpha N_e (g^4+f^2/8)^2} > 1. \end{equation} {}For the force and infra-red conditions, they also are evaluated at the most stringent instant during warm inflation, which turns out to be again $\tau=\tau_{EI}$. \subsubsection{\bf $N_r \gg 2NT/T_{EI}$} \label{subsubsect4A2} In this case the $N_r$ term dominates the radiation energy density in Eq. (\ref{rhora}). \begin{center} {\bf Solutions} \end{center} The procedure for solving this case is similar to the above case. The results are \begin{equation} \varphi_0(\tau) = \varphi_{BI} \exp(-y \tau) \end{equation} and \begin{equation} T(\tau) = \frac{1}{3^{1/3} (\eta' \Gamma')^{1/6}} \left(\frac{\lambda^2}{\alpha \beta \kappa_M N_rN g^2} \right)^{1/6} \varphi_0(\tau) = \kappa_M M \exp(-y\tau), \label{tempA2} \end{equation} where in terms of the parameters of the model $y=[\pi^2/(3^{1/3} 2 (\eta' \Gamma')^{2/3})][\lambda N_r \alpha \beta \kappa_M/(N^2 g^4)]^{1/3}$ and in terms of the expansion e-folds from Eq. (\ref{ne}) \begin{equation} y= \frac{\alpha \beta \kappa_M \ln(1/\beta)}{N_e}. \end{equation} {}From the specified conditions on temperature expressed through $\alpha$, $\beta$, and $\kappa_M$, two parameters in the model are determined \begin{equation} \lambda = \left(\frac{24 \eta^{'2}\Gamma^{'2}}{\pi^6}\right) \frac{\alpha^2 \beta^2 \kappa_M^2 \ln^3(1/\beta) N^2 g^4}{N_e^3 N_r} \end{equation} and \begin{equation} \frac{\varphi_{BI}}{M} = 3^{1/3}(\eta' \Gamma')^{1/6} \left(\frac{\alpha \beta \kappa_M N_r N g^2}{\lambda^2}\right)^{1/6} \kappa_M =\frac{\pi^2}{2(\eta'\Gamma')^{1/2}} \frac{\kappa_M^{1/2} N_eN_r^{1/2}}{\alpha^{1/2} \beta^{1/2} \ln(1/\beta) N^{1/2} g}. \end{equation} Based on these solutions, some useful expressions are as follows. The number of mass sites that are thermally excited at a given instance are $n_{t.e.} = 2 \kappa_M \exp(-y \tau)$. The number of mass sites that $\varphi_0$ crosses during the warm inflation period is \begin{equation} N_M=\frac{\pi^2}{2(\eta'\Gamma')^{1/2}} \frac{(1-\beta)\kappa_M^{1/2} N_e N_r^{1/2}}{\alpha^{1/2}\beta^{1/2} \ln(1/\beta) N^{1/2}}, \end{equation} with \begin{equation} i_{\rm min} = \frac{\pi^2}{2(\eta'\Gamma')^{1/2}} \frac{\beta^{1/2}\kappa_M^{1/2} N_e N_r^{1/2}}{\alpha^{1/2} \ln(1/\beta) N^{1/2}}. \end{equation} The mass of the $\phi$ particles is \begin{equation} {\bar m}_{\phi}^2 \approx \frac{\lambda \varphi_0^2}{2 T^2} =\frac{3\eta'\Gamma'}{\pi^2} \frac{\alpha \beta \kappa_M \ln(1/\beta) N g^2}{N_e}. \end{equation} The ratio of vacuum to radiation energy density is \begin{equation} \frac{\rho_{\phi}(\tau)}{\rho_r(\tau)}= \frac{1}{(\eta'\Gamma')^{1/2}} \frac{N_e}{\ln(1/\beta)} \label{rhorvA2} \end{equation} \begin{equation} \frac{V_0}{\rho_r(\tau)}= \frac{1}{(\eta'\Gamma')^{1/2}} \frac{N_e}{\ln(1/\beta)} \exp(4y\tau). \end{equation} {}For examining the adiabatic conditions \begin{equation} \frac{d \varphi_0}{d \tau} = -\frac{\alpha \beta \ln(1/\beta) \kappa_M}{N_e} \varphi_0(\tau) =\frac{\pi^2}{2(\eta'\Gamma')^{1/2}} \frac{\alpha^{1/2} \beta^{1/2} \kappa_M^{1/2} N_r^{1/2}} {N^{1/2} g} T(\tau) \end{equation} and \begin{equation} \frac{\Gamma_{\chi}(T)}{\Gamma_{\chi}(M)} = \kappa_M \exp(-y \tau). \end{equation} \begin{center} {\bf Consistency Conditions} \end{center} The parametric constraints from the consistency conditions are the $\varphi_0$-adiabatic condition Eq. (\ref{padiab}) with sufficiency condition \begin{equation} N_e > 1, \end{equation} the thermal-adiabatic condition Eq. (\ref{thermadiab}) \begin{equation} \frac{\pi^2}{2(\eta'\Gamma')^{1/2}} \left(\frac{\alpha N_r}{\beta \kappa_M N}\right)^{1/2} < 1, \label{thermaA2} \end{equation} the force condition Eq. (\ref{forcec}) \begin{equation} 4(\eta'\Gamma')^{1/2} \left(\frac{\alpha \beta \kappa_M N_r}{N}\right)^{1/2} > 1, \label{forcecA2} \end{equation} and the infrared condition Eq. (\ref{irc}) \begin{equation} \frac{3\eta'}{\Gamma'} \frac{\beta \ln(1/\beta) \kappa_M N g^2} {\alpha N_e (g^4+f^2/8)^2} > 1. \end{equation} It should also be recalled that the basic requirement for this regime, $N_r$-dominated, is \begin{equation} \frac{N_r}{N} \gg 2 \kappa_M. \end{equation} \subsection{$T_{BI} \stackrel{>}{\sim} T_{EI}$} \label{subsect4B} {}For this regime, we adopt the criteria \begin{equation} \beta \stackrel{>}{\sim} 0.5. \label{beta5} \end{equation} In this regime the number of thermally excited sites in the $\varphi_0$-equation of motion is approximated as constant, so that in Eq. (\ref{eom1}) $\sum_i^{t.e.}(\varphi_0-M_i)^2 \eta_{1i}^B(T) \approx \kappa_M^3 M^2 N \eta'/(g^2T)$. All other approximations are the same as in the previous section. Thus the $\varphi_0$-equation of motion is \begin{equation} \frac{\eta'\Gamma'\kappa_M^3 g^2 N M^3 }{\pi^2 T} \frac{d \varphi_0}{d \tau} = -\frac{\lambda \varphi_0^3}{6}. \label{eomB} \end{equation} Both cases studied below are for arbitrary number of dissipative and non-dissipative heat bath fields so that the radiation energy density is \begin{equation} \rho_r(\tau) = \frac{\pi^2}{16} (N_r + 2 N \kappa_M) T^4(\tau). \end{equation} The two cases examined in order are $V_0 >0$ and $V_0=0$. The latter case is the DM-model examined in \cite{bgr2}. However here the calculation is extended to permit an arbitrary mass splitting scale M (in \cite{bgr2} the mass splitting scale was restricted to $M \sim T$) and to treat the time dependence of the temperature in the $\varphi_0$-equation of motion. \subsubsection{$V_0> 0$} \label{subsubsect4B1} Similar to the previous subsection, we choose $V_0 \approx \lambda \varphi^4_{BI}/24$. The number of e-folds follows from the relation Eq. (\ref{ne}). \begin{center} {\bf Solutions} \end{center} The procedure for solving this case is similar to the above cases. The results are \begin{equation} \varphi_0(\tau) = \frac{\varphi_{BI}}{(y\tau+1)^{1/4}} \end{equation} and \begin{equation} T(\tau) = \frac{1}{(9 \eta' \Gamma')^{1/3}} \left(\frac{\lambda^2}{\alpha \beta \kappa_M^4 (N_r+2\kappa_M N) N g^2} \right)^{1/3} \frac{\varphi_0(\tau)^2}{M} = \frac{\kappa_M M}{(y \tau+1)^{1/2}}, \label{tempB1} \end{equation} where in terms of the parameters of the model $y=[2\pi^2/(3^{5/3}(\eta' \Gamma')^{4/3}] [\lambda^5/(\alpha \beta \kappa_M^{13} (N_r+2\kappa_M N) N^4 g^8]^{1/3} (\varphi_{BI}/M)^4$ and in terms of the expansion e-folds from Eq. (\ref{ne}) \begin{equation} y= \frac{\alpha \kappa_M (1-\beta^2)}{\beta N_e}. \end{equation} {}From the specified conditions of temperature expressed through $\alpha$, $\beta$, and $\kappa_M$ two parameters in the model are determined \begin{equation} \lambda = \left(\frac{3 \eta^{'2}\Gamma^{'2}}{8\pi^6}\right) \frac{\alpha^2 (1-\beta^2)^3 \kappa_M^2 N^2 g^4} {\beta^4 N_e^3 (N_r+2\kappa_M N) } \end{equation} and \begin{equation} \frac{\varphi_{BI}}{M} = (9 \eta' \Gamma')^{1/6} \left(\frac{\alpha^2 \beta^2 \kappa_M^7 (N_r+2\kappa_M N) N g^2} {\lambda^2}\right)^{1/6} = \frac{2 \pi^2}{(\eta'\Gamma')^{1/2}} \frac{\beta^{3/2} \kappa_M^{1/2} N_e (N_r+2\kappa_M N)^{1/2}} {\alpha^{1/2} (1-\beta^2) N^{1/2} g}. \end{equation} Based on these solutions, some useful expressions are as follows. The number of mass sites that are thermally excited at a given instance are $n_{t.e.} = 2 \kappa_M/(y \tau+1)^{1/2}$. The umber of mass sites that $\varphi_0$ crosses during the warm inflation period is \begin{equation} N_M=\frac{2 \pi^2}{(\eta'\Gamma')^{1/2}} \frac{\beta^{3/2}(1-\beta^{1/2})\kappa_M^{1/2} N_e (N_r+2\kappa_M N)^{1/2}}{\alpha^{1/2}(1-\beta^2) N^{1/2}}, \end{equation} with \begin{equation} i_{\rm min} = \frac{2 \pi^2}{(\eta'\Gamma')^{1/2}} \frac{\beta^2\kappa_M^{1/2} N_e (N_r+2\kappa_M N)^{1/2}}{\alpha^{1/2}(1-\beta^2) N^{1/2}}. \end{equation} The mass of the $\phi$ particles is \begin{equation} {\bar m}_{\phi}^2 \approx \frac{\lambda \varphi_0^2}{2 T^2} =\frac{3\eta'\Gamma'}{4 \pi^2} \frac{\alpha \kappa_M (1-\beta^2) N g^2}{\beta N_e}(y\tau+1)^{1/2}. \end{equation} The ratio of vacuum to radiation energy density is \begin{equation} \frac{\rho_{\phi}(\tau)}{\rho_r(\tau)}= \frac{4}{3^{4/3} (\eta'\Gamma')^{2/3}} \frac{\beta^2 N_e}{(1-\beta^2)} (y\tau+1) \label{rhorvB1} \end{equation} \begin{equation} \frac{V_0}{\rho_r(\tau)}= \frac{4}{3^{4/3} (\eta'\Gamma')^{2/3}} \frac{\beta^2 N_e}{(1-\beta^2)} (y\tau+1)^2. \end{equation} {}For examining the adiabatic conditions \begin{equation} \frac{d \varphi_0}{d \tau} = -\frac{\alpha (1-\beta^2) \kappa_M}{4 \beta N_e} \frac{\varphi_0(\tau)}{(y \tau+1)} =\frac{\pi^2}{2(\eta'\Gamma')^{1/2}} \frac{\alpha^{1/2} \beta^{1/2} \kappa_M^{1/2} (N_r+2\kappa_M N)^{1/2}} {N^{1/2}g(y\tau+1)^{3/4}} T(\tau) \end{equation} and \begin{equation} \frac{\Gamma_{\chi}(T)}{\Gamma_{\chi}(M)} = \frac{\kappa_M}{(y\tau+1)^{1/2}}. \end{equation} \begin{center} {\bf Consistency Conditions} \end{center} The parametric constraints from the consistency conditions are the $\varphi_0$-adiabatic condition Eq. (\ref{padiab}) with sufficiency condition \begin{equation} N_e > \frac{3}{8}, \end{equation} the thermal-adiabatic condition Eq. (\ref{thermadiab}) \begin{equation} \frac{\pi^2}{2(\eta'\Gamma')^{1/2}} \left(\frac{\alpha \beta (N_r+2\kappa_MN)}{\kappa_M N}\right)^{1/2} <1, \label{thermaB1} \end{equation} the force condition Eq. (\ref{forcec}) \begin{equation} 4(\eta'\Gamma')^{1/2} \left(\frac{\alpha \beta \kappa_M (N_r + 2\kappa_M N)}{N}\right)^{1/2} > 1, \label{forcecB1} \end{equation} and the infrared condition Eq. (\ref{irc}) \begin{equation} \frac{3\eta'}{4\Gamma'} \frac{(1-\beta^2) \kappa_M N g^2} {\alpha \beta^2 N_e (g^4+f^2/8)^2} > 1. \end{equation} \subsubsection{$V_0 = 0$} \label{subsubsect4B2} In this case, there is no shift parameter $V_0$ nor extra function $V_1(\phi)$ in the Lagrangian Eq. (\ref{Nfields}). This is the model examined in \cite{bgr2}, except here the calculation is extended to treat the time dependence of the temperature in the dissipative function. Also \cite{bgr2} only examined the regime where the mass level splitting $M\sim T$, whereas here the relation between $M$ and $T$ can be varied through the parameter $\kappa_M$. There are two differences in this calculation's procedures compared to the previous three cases. Both differences arise because here the Hubble parameter is $\varphi_0(\tau)$ dependent \begin{equation} H(\tau)=\alpha \beta \kappa_M \Gamma_{\chi}(M) \frac{\varphi_0^2(\tau)}{\varphi_{BI}^2}. \label{hubve0} \end{equation} {}First this dependence must be treated in the energy conservation equation (\ref{dotr}). Second the scale factor no longer grows exactly exponentially. However this calculation is also based on the assumption that the temperature during warm inflation does not change significantly, $\beta \stackrel{>}{\sim} 0.5$, and as will be seen this also implies $\varphi_0(\tau)$, thus $\rho_v(\tau)$, does not change significantly. In this case, the scale factor grows quasi-exponentially with e-folds \begin{equation} N_e \approx \int_0^{t_{EI}} H(t) dt = \alpha \beta \kappa_M \int_0^{\tau_{EI}} \frac{\varphi_0^2(\tau)} {\varphi_{BI}^2} d\tau. \label{nev00} \end{equation} \begin{center} {\bf Solutions} \end{center} The procedure for solving this case is similar to the above case. The results are \begin{equation} \varphi_0(\tau) = \frac{\varphi_{BI}}{(y\tau+1)^{3/10}} \end{equation} and \begin{equation} T(\tau) = \frac{1}{(9 \eta' \Gamma')^{1/3}} \left(\frac{\lambda^2}{\alpha \beta \kappa_M^4 (N_r+2\kappa_M N) N g^2} \right)^{1/3} \left(\frac{\varphi_{BI}}{M}\right)^{2/3} \frac{\varphi_0^{4/3}(\tau)}{M^{1/3}} = \frac{\kappa_M M}{(y \tau+1)^{2/5}}, \label{tempB2} \end{equation} where in terms of the parameters of the model $y=[5 \pi^2/(( 9\eta' \Gamma')^{4/3}] [\lambda^5/(\alpha \beta \kappa_M^{13} (N_r+2\kappa_M N) N^4 g^8)]^{1/3} (\varphi_{BI}/M)^4$ and in terms of the expansion e-folds from Eq. (\ref{nev00}) \begin{equation} y= \frac{5 \alpha \kappa_M (1-\beta)}{2 N_e}. \end{equation} {}From the specified conditions of temperature expressed through $\alpha$, $\beta$, and $\kappa_M$ two parameters in the model are determined \begin{equation} \lambda = \left(\frac{81 \eta^{'2}\Gamma^{'2}}{8\pi^6}\right) \frac{\alpha^2 (1-\beta)^3 \kappa_M^2 N^2 g^4} {\beta N_e^3 (N_r+2\kappa_M N) } \label{lamB2} \end{equation} and \begin{equation} \frac{\varphi_{BI}}{M} = 3 (\eta' \Gamma')^{1/2} \left(\frac{\alpha \beta \kappa_M^7 (N_r+2\kappa_M N) N g^2} {\lambda^2}\right)^{1/6} = \frac{2 \pi^2}{3(\eta'\Gamma')^{1/2}} \frac{\beta^{1/2} \kappa_M^{1/2} N_e (N_r+2\kappa_M N)^{1/2}} {\alpha^{1/2} (1-\beta) N^{1/2} g}. \end{equation} Based on these solutions, some useful expressions are as follows. The number of mass sites that are thermally excited at a given instance are $n_{t.e.} = 2 \kappa_M/(y \tau+1)^{2/5}$. The number of mass sites that $\varphi_0$ crosses during the warm inflation period is \begin{equation} N_M=\frac{2 \pi^2}{3(\eta'\Gamma')^{1/2}} \frac{\beta^{1/2}(1-\beta^{3/4})\kappa_M^{1/2} N_e (N_r+2\kappa_M N)^{1/2}}{\alpha^{1/2}(1-\beta) N^{1/2}}, \end{equation} with \begin{equation} i_{\rm min} = \frac{2 \pi^2}{3(\eta'\Gamma')^{1/2}} \frac{\beta^{5/4}\kappa_M^{1/2} N_e (N_r+2\kappa_M N)^{1/2}}{\alpha^{1/2}(1-\beta) N^{1/2}}, \end{equation} The mass of the $\phi$ particles is \begin{equation} {\bar m}_{\phi}^2 \approx \frac{\lambda \varphi_0^2}{2 T^2} =\frac{9 \eta'\Gamma'}{4 \pi^2} \frac{\alpha (1-\beta) \kappa_M N g^2}{N_e}(y\tau+1)^{1/5}. \end{equation} The ratio of vacuum to radiation energy density is \begin{equation} \frac{\rho_{\phi}(\tau)}{\rho_r(\tau)}= \frac{4}{3} \frac{\beta N_e}{(1-\beta)^4} (y\tau+1)^{2/5}. \label{rhorvB2} \end{equation} {}For examining the adiabatic conditions \begin{equation} \frac{d \varphi_0}{d \tau} = -\frac{3 \alpha (1-\beta) \kappa_M}{4 N_e} \frac{\varphi_0(\tau)}{(y \tau+1)} =\frac{\pi^2}{2(\eta'\Gamma')^{1/2}} \frac{\alpha^{1/2} \beta^{1/2} \kappa_M^{1/2} (N_r+2\kappa_M N)^{1/2}} {N^{1/2} g(y\tau+1)^{1/2}} T(\tau) \end{equation} and \begin{equation} \frac{\Gamma_{\chi}(T)}{\Gamma_{\chi}(M)} = \frac{\kappa_M}{(y\tau+1)^{2/5}}. \end{equation} \begin{center} {\bf Consistency Conditions} \end{center} The parametric constraints from the consistency conditions are the $\varphi_0$ adiabatic condition Eq. (\ref{padiab}) with sufficiency condition \begin{equation} N_e > \frac{3}{8} , \end{equation} the thermal-adiabatic condition Eq. (\ref{thermadiab}) \begin{equation} \frac{\pi^2}{2(\eta'\Gamma')^{1/2}} \left(\frac{\alpha \beta (N_r+2\kappa_MN)}{\kappa_M N}\right)^{1/2} <1, \label{thermaB2} \end{equation} the force condition Eq. (\ref{forcec}) \begin{equation} 4(\eta'\Gamma')^{1/2} \left(\frac{\alpha \beta \kappa_M (N_r + 2\kappa_M N)}{N}\right)^{1/2} > 1, \label{forcecB2} \end{equation} and the infrared condition Eq. (\ref{irc}) \begin{equation} \frac{9\eta'}{4\Gamma'} \frac{(1-\beta) \kappa_M N g^2} {\alpha N_e (g^4+f^2/8)^2} > 1. \end{equation} \subsection{Discussion} \label{subsect4C} In this subsection, some general comments are given about the four cases examined in the previous two subsections. The first noteworthy observation is that for arbitrary e-folds, $N_e$, the consistency conditions impose very mild restrictions on the parameter space in all four cases. Although, the precise restrictions vary amongst the four cases, a general set of restrictions that is valid for all four cases can be given. {}First recall that by construction of the solution, the thermalization condition Eq(\ref{therm}) always requires simply that $\alpha < 1$. The most stringent restrictions arise from the force Eq. (\ref{forcec}) and thermal-adiabatic Eq. (\ref{thermadiab}) conditions, which together require \begin{equation} \frac{1}{\beta \kappa_M^2} < \alpha \stackrel{<}{\sim} \frac{\beta \kappa_M N}{100(N_r+2\kappa_M N)}. \label{genalp} \end{equation} The other two consistency conditions impose very mild restrictions. The $\varphi_0$-adiabatic condition Eq. (\ref{padiab}) is always accommodated provided $N_e>1$ and the infra-red condition Eq. (\ref{irc}) is accommodated provided $\beta \kappa_M Ng^2/(\alpha N_e(g^4+f^2/8)^2 > 4 \times 10^{-5}$. With all the conditions combined, they are fairly unrestrictive to the parameter space. As such, it leaves considerable freedom for treating density perturbations. Another interesting point is to compare the solution in Subsect. \ref{subsubsect4B1} for $T_{EI} \stackrel{<}{\sim} T_{BI}$ with those in Subsect. \ref{subsect4A} for $T_{EI} \ll T_{BI}$ and see how well they match at some intermediate $T_{EI} < T_{BI}$. There clearly should be some overlapping region since the model is exactly the same and only the treatment of temperature dependence is different. By comparing the expressions for $\lambda$ and $\varphi_{BI}/M$, the functional dependence on the parameters is seen to be exactly the same except for with respect to $\beta$. In regards to $\beta$, both $\lambda$ and $\varphi_{BI}/M$ for the cases in Subsects. \ref{subsubsect4A1}, $N_r=0$, and \ref{subsubsect4A2}, $N_r \gg 2T/M$, equate to the expressions in \ref{subsubsect4B1} at $\beta=0.42$ and $0.53$ respectively. This is close to the approximate cut-off criteria we gave in 3B of $\beta \sim 0.5$ Eq. (\ref{beta5}). The final point is that cross comparison amongst the four cases indicates several similarities amongst the solutions. In the remainder of this subsection, the origin of these similarities are examined. Alongside this, a qualitative understanding of the solutions are developed. The two basic equations of warm inflation, the $\varphi_0$-equation of motion Eq. (\ref{eom1}) and the energy conservation equation Eq. (\ref{dotr}) have five properties, which are listed below. {}From the properties, all the similarities amongst the solutions then are explained. \noindent {\bf property 1}: The first time derivative of $\varphi_0$, $d\varphi_0(\tau)/d\tau$, is related to some product of $\varphi_0(\tau)^a T(\tau)^b$ in both basic equations of warm inflation, Eqs. (\ref{eom1}) and (\ref{dotr}). (In the $\varphi_0$-equation of motion, Eq. (\ref{eom1}), recall that we approximate the dissipative coefficient as $\sum_i^{t.e.} (\varphi_0-M_i)^2 \eta(T) \approx T^3 \eta(T)/(g^2 M)$, where for the $T^3(\tau)$ term, in Subsect. A the time dependence is treated and in Subsect. B it is treated as a constant $T^3(\tau)=T^3(0)$. In particular the dissipative coefficient in both subsections depends on some power of the temperature $T(\tau)$.) \noindent {\bf property 2}: In the $\varphi_0$-equation of motion, the direction of $d\varphi_0(\tau)/d\tau$ is always opposite to the sign of $\varphi_0(\tau)$. \noindent {\bf property 3}: Since the $\varphi_0$-equation of motion is first order, the solution requires one initial condition which then sets the overall scale for both $\varphi_0(\tau)$ and $T(\tau)$. In our approach, this scale is set by the condition $T(0)\equiv T_{BI}=\kappa_M M$. As such $\varphi_{BI}$ is then a derived quantity. \noindent {\bf property 4}: The force term in the $\varphi_0$ equation of motion in our approximation is always $\lambda \varphi_0^3/6$. \noindent {\bf property 5}: The number of e-folds is linearly related to the dimensionless time parameter at the end of warm inflation $\tau_{EI}$ in all four cases $N_e \propto \alpha \tau_{EI}$. Properties one and two imply that the solutions have the general behavior \begin{equation} \varphi_0(\tau) \equiv \varphi_{BI} D(\tau) \label{varphigs} \end{equation} where \begin{equation} D(\tau) = \frac{1}{(y\tau+1)^{\gamma_{\phi}}} ( \ \ {\rm or} \ \ \exp(-y\tau) ) \end{equation} and \begin{equation} T(\tau) = f(\lambda,g,f,N,\kappa_M,\alpha,\beta) \varphi^{\gamma_T}_0(\tau) M^{1-\gamma_T} \label{T1} \end{equation} with $\gamma_{\phi},\gamma_T >0$ and $f(\lambda,g,f,N,\kappa,\alpha,\beta)$ a function of the the model and thermodynamic parameters. Combining this deduction with property 3, we also can conclude that in general\footnote{Equating Eqs. (\ref{T1}) and (\ref{T2}) lead to one of the two parametric constraints that in the previous two subsections determined $\lambda$ and $\varphi_{BI}/M$.} \begin{equation} T(\tau) = \kappa_M D^{\gamma_T}(\tau) \label{T2} \end{equation} The noteworthy point for the present discussion is the solutions for $\varphi_0(\tau)$ and $T(\tau)$ are always a product of a mass-dimension one function which depends on the model and thermodynamic parameters and a time dependent decay function. The latter we represent through $D(\tau)$ taken to some positive power with $D(0)=1$ and $D(\tau>0) < 1$. {}For this discussion it is useful to think about the solutions this way, since the general features are contained in the mass-dimension one function. As such, the discussion to follow is not detailed about the decay function. The general behavior of the temperature permits two deductions. {}First from Eq.(\ref{T2}) and the definition of $\beta$ it follows that in general \begin{equation} y \sim \frac{\alpha h(\beta) \kappa_M}{N_e} \end{equation} where $h(\beta)$ depends on the specific time dependence of $T(\tau)$ \footnote{Equating this expression for $y$ with the one obtained with respect to the parameters of the model yields the second of two conditions that determine $\lambda$ and $\varphi_{BI}/M$ in the previous two subsections.}. Second, our approximation for the dissipative coefficient in all the cases studied has the general form \begin{equation} \sum_i^{t.e.}(\varphi_0-M_i)^2 \frac{\eta'}{\pi T} \approx \frac{T^3 \eta'}{\pi M T} \approx \frac{\kappa_M^2 M \eta'}{\pi} D(\tau)^{\gamma_{\eta}}, \label{dissgen} \end{equation} where $\gamma_{\eta}$ depends on the specific treatment of the temperature's time dependence. In Subsect. A $\gamma_{\eta} = 2\gamma_T$ and in Subsect. B $\gamma_{\eta} =-\gamma_{T}$. By this point it should begin to appear evident that the differences in the various cases emerge primarily in the exponent of the decay function. This becomes fully clear once the basic equations are examined below. In fact it can be recognized that the $\varphi_0$-equation of motion Eq.(\ref{eom1}) actually is a equation of motion for $D(\tau)$. Using Eq. (\ref{dissgen}) it has the general form \begin{equation} \frac{d\varphi_0(\tau)}{d\tau} =\varphi_{BI} \frac{dD(\tau)}{d\tau} =-\frac{\pi^2\lambda \varphi^3_{BI}}{6 \eta'\Gamma' \kappa_M^2 Ng^2M^2} D(\tau)^{3-\gamma_{\eta}} =-\frac{\pi^2\lambda \varphi_0^3(\tau)}{6 \eta'\Gamma' Ng^2 T^2(\tau)} D(\tau)^{2\gamma_{T}-\gamma_{\eta}}. \label{eomgen} \end{equation} On the other hand, from the general form of the $\varphi_0(\tau)$ solution Eq.(\ref{varphigs}), it also follows that \begin{equation} \frac{d\varphi_0(\tau)}{d\tau} \propto \frac{y \varphi_0(\tau)}{(y\tau+1)^{1-\delta_e}}, \label{phialt} \end{equation} where $\delta_e=1$ if $D(\tau) \sim exp(-y\tau)$ and zero otherwise. Based on this relation, we find that $d \rho_{\phi}(\varphi_0)/d\tau = \lambda \varphi_0^3(\tau) d\varphi_0(\tau)/d\tau/6 \propto y\rho_{\phi}(\varphi_0)/(y\tau+1)^{1-\delta_e}$ which applied to the energy conservation equation Eq.(\ref{dotr}) implies \begin{equation} \frac{\rho_{\phi}(\tau)}{\rho_r(\tau)} \equiv \frac{2\lambda \varphi_0^4(\tau)}{3\pi^2[N_r+2\kappa_M ND^{\gamma_{\rho'}} (\tau)] T^4(\tau)} \sim \frac{H(\tau)}{\Gamma_{\chi}(M)} \frac{y\tau+1}{y} \sim \frac{N_e \beta}{h(\beta)} (y\tau+1)D^{\gamma_{\rho''}}(\tau), \label{rhogen} \end{equation} where $\gamma_{\rho'}$ and $\gamma_{\rho''}$ are more exponents that depend on the specific solution regime. In Subsects. A and B $\gamma_{\rho'}=\gamma_T$ and $0$ respectively. $\gamma_{\rho''}$ represents the time dependence of $H$. Thus $\gamma_{\rho''}=0$ in all the cases except the last, IIIB2, where $\gamma_{\rho''}=2$. Now we can deduce all the general features of the solutions in the previous two subsections. In addition to examining the solutions in terms of the convenient parameters used in our analysis, it is interesting to see how the solutions depend on the parameters with direct physical interpretation, the Hubble parameter $H$ and the slope parameter for $\varphi_0(\tau)$, $y$ in exchange of $\alpha$ and $N_e$. {}For $\rho_{\phi}(\tau)/\rho_r(\tau)$, the general expressions already have been given in Eq. (\ref{rhogen}). {}From Eqs. (\ref{eomgen}) and (\ref{phialt}) we find \begin{equation} \frac{\lambda \varphi_0^2(\tau)}{2T^2(\tau)} \sim \eta' \Gamma' y N g^2 \frac{D^{\gamma_{\eta}-2\gamma_T}(\tau)}{(y\tau+1)^{1-\delta_e}} = \frac{\eta' \Gamma' \alpha h(\beta) \kappa_M N g^2}{N_e} \frac{D^{\gamma_{\eta}-2\gamma_T}(\tau)}{(y\tau+1)^{1-\delta_e}}. \label{mphigen} \end{equation} {}From Eqs. (\ref{rhogen}) and (\ref{mphigen}) we find \begin{equation} \lambda \sim (\eta' \Gamma')^2 \frac{y^3 N^2 g^4} {(H/\Gamma)(N_r+2\kappa_MN)} = (\eta' \Gamma')^2 \frac{\alpha^2 h^3(\beta) \kappa_M^2 N^2 g^4} {\beta N_e^3 (N_r+2\kappa_MN)} \label{lamgen} \end{equation} and \begin{equation} \frac{\varphi_{BI}}{M} \sim \frac{1}{(\eta'\Gamma')^{1/2}} \left(\frac{(H/\Gamma) \kappa_M^2 (N_r+2\kappa_MN)}{y^2 Ng^2} \right)^{1/2} = \frac{1}{(\eta'\Gamma')^{1/2}} \frac{\beta^{1/2} \kappa_M^{1/2} N_e (N_r+2\kappa_MN)^{1/2}} {\alpha^{1/2} h(\beta) N^{1/2} g}. \label{pomgen} \end{equation} Regarding the consistency conditions, the general forms are as follows: $\varphi_0$-adiabatic condition Eq. (\ref{padiab}) \begin{equation} \frac{y}{\kappa_M} \frac{D^{-\gamma_T}(\tau)} {(y\tau+1)^{1-\delta_e}} = \frac{\alpha h(\beta)}{N_e} \frac{D^{-\gamma_T}(\tau)} {(y\tau+1)^{1-\delta_e}} < O(1), \label{padaibgen} \end{equation} the thermal-adiabatic condition Eq. (\ref{thermadiab}) \begin{equation} \left(\frac{(H/\Gamma)(N_r+2\kappa_M N)}{\eta' \Gamma' \kappa_M^2 N} \right)^{1/2} \frac{D^{1-2\gamma_T}(\tau)} {(y\tau+1)^{1-\delta_e}} = \left(\frac{(\alpha \beta (N_r+2\kappa_M N)} {(\eta' \Gamma') \kappa_M N}\right)^{1/2} \frac{D^{1-2\gamma_T}(\tau)} {(y\tau+1)^{1-\delta_e}} < O(1), \label{tadaibgen} \end{equation} and the force condition Eq. (\ref{forcec}) \begin{eqnarray} (\eta'\Gamma')^{1/2} \left(\frac{(H/\Gamma) (N_r+2\kappa_MN)}{N}\right)^{1/2} D^{3-3\gamma_T}(\tau) & \sim & (\eta'\Gamma')^{1/2} \left(\frac{\alpha \beta \kappa_M (N_r+2\kappa_MN)}{N}\right)^{1/2} D^{3-3\gamma_T}(\tau) \nonumber\\ & > & O(1). \label{forcegen} \end{eqnarray} These expressions contain interesting insight into the solutions. The expression for $\rho_v/\rho_r \propto V(\varphi_0)/T^4$, Eq. (\ref{rhogen}), indicates that the relative content of radiation to vacuum energy is directly proportional to the Hubble expansion rate $H_{BI}$ and inversely proportional to the slope parameter $y$. These general trends can be explained. The former is expected since for a slower the Hubble expansion rate, the red-shifting of radiation is slower. The latter follows since as the slope parameter decreases, the decay of vacuum energy to radiation also becomes slower. The expression for $m_{\phi}^2/T^2$, Eq. (\ref{mphigen}), also has a simple explanation. It is independent of the Hubble parameter, which is indicative that this expression is entirely an outcome of the $\varphi_0$ equation of motion. Since the potential is a monomial in $\varphi_0$ and the $\varphi_0$ equation of motion is first order and overdamped, Eq. (\ref{phialt}) follows. {}Furthermore, in our case the $\varphi_0$ equation of motion always has two powers of the temperature in the form $d\varphi_0/d\tau \propto (dV(\varphi_0)/d\varphi_0)/T^2$. Comparing this with $d\varphi_0/d \tau \sim y \varphi_0$ from Eq. (\ref{phialt}) and noting that our potential is quartic, $V(\varphi_0) \sim \lambda \varphi_0^4$, the expression for $\lambda \varphi_0^2/T^2$ in Eq. (\ref{mphigen}) follows. Based on these two expressions, Eqs. (\ref{rhogen}) and (\ref{mphigen}), the remaining expressions follow. {}For example since $V(\varphi_0)/T^4 \propto H_{BI}/y$ and $(dV(\varphi_0)/d\varphi_0)/(\varphi_0 T^2) \propto y$, it must follow that $\varphi_0^2/T^2 \propto H_{BI}/y^2$. It then implies that for a quartic potential $\lambda \propto y^3/H_{BI}$. {}For a hypothetical monomial potential $V(\varphi_0) \sim \lambda M^{4-n} \phi^n$, it still follows that $\varphi_0^2/T^2 \propto H_{BI}/y^2$ but now $\lambda (\varphi_0/M)^{n-4} \propto y^3/H_{BI}$. The thermal adiabatic condition, $\propto (d\varphi_0/d\tau)/T$, also can be understood since $(d\varphi_0/d\tau)/T \propto y\varphi_0/T \propto H_{BI}^{1/2}$. The behavior of the other expressions as well as other parameter dependencies can be deduced by similar reasoning. \section{Estimates for amplitude of density Perturbations $\delta \rho/\rho$} \label{sect5} In this section estimates are made of scalar metric perturbations for flat spatial geometry, $\Omega =1$, for the four cases in the previous section. For the calculations to follow, the basic formulas can be deduced either from first principles arguments or from a perturbative quantum field theory calculation, and we will take the former route. This approach has limitations, within which it is indisputable. For our purposes these limitations are unimportant in obtaining the basic formulas and in making the estimates for density perturbations in the quantum field theory model studied in this paper. The limitation of this approach is in sacrificing some of the details that could be obtained by the perturbative calculation. These details, which will be described at an appropriate place below, are not useful for our present purposes. Nevertheless, at some stage, it will be important to understand these details and to do the perturbative calculation. For this purpose, the present approach also is useful as a cross-check for the perturbative calculations and in providing a necessary outline of the formal problems that must be addressed for that derivation. To estimate density perturbations, the classical background field now is treated with fluctuations, $\varphi({\bf x},t) = \varphi_0(t)+ \delta \varphi({\bf x},t)$, where $\varphi_0(t)$ is the zero mode and $\delta \varphi({\bf x},t)$ are small fluctuations about the zero mode. The calculations to follow are in momentum space. The Fourier transform of the fluctuations is defined as \begin{equation} \delta \varphi({\bf k},t) = \int_{V} d^3x \delta \varphi({\bf x},t) \exp(-i {\bf k} \cdot {\bf x}) \end{equation} where $({\bf k},{\bf x})$ can be either comoving coordinates or physical coordinates at a particular time. Hereafter we denote comoving coordinates as $({\bf k}_c,{\bf x}_c)$ and physical coordinates at time $\tau$ as $({\bf k}_p(\tau),{\bf x}_p(\tau))$, where the $\tau$- dependence sometimes is not shown explicitly. The comoving coordinates can be regarded as the intrinsic labels for the modes of the scalar field fluctuations. Thus the modes will be denoted as $\delta \varphi_{{\bf k}_c}(\tau)$. {}For definiteness, the comoving and physical coordinates will coincide at the end of warm inflation $\tau_{EI}$ so that for any component direction $k_p(\tau) = k_c/R(\tau-\tau_{EI})$, where by our convention $R(0)=1$. {}For the four cases treated in the last section $R(\tau) \approx \exp[H(\tau) \tau]$. Very often it will be necessary to consider modes in terms of their physical wavenumber at a given time $\tau$. Thus, the modes also may be denoted as $\delta \varphi({\bf k}_p,\tau)$, where the (redundant) ${\bf k}_c$ subscript has been dropped. Let us obtain the equation of motion for the fluctuations in flat nonexpanding spacetime. We argue below that this equation also suffices for our purposes in an expanding background. {}For the nonexpanding background case, no distinction is necessary between comoving and physical wavenumbers; the modes are denoted simply as $\delta \varphi({\bf k},\tau)$. As stated earlier, the equations of motion for the fluctuations can be obtained by two approaches, either by general first principles arguments or a direct perturbative calculation, and here the former approach is taken. The basic fact that permits the former approach is that the calculation of the zero mode and fluctuation equations of motion, which follow the methods of \cite{hs1,morikawa,lawrie,gr,bgr}, are done in a near-thermal-equilibrium approximation, thus respecting the fluctuation-dissipation theorem. This fact immediately determines the form of the fluctuation equation of motion, given that for the zero mode. In particular, assuming the fluctuations are small, their equation of motion is obtained through a linearization of the zero mode equation of motion, and with the inclusion of the gradient term, $\nabla^2 \delta \varphi$, and a noise term that represents the short distance dynamics of the heat bath. The fluctuation dissipation theorem uniquely determines the correlation statistics of the noise. The missing detail from this first principles procedure, which were eluded to earlier, is the relation of the noise function to the basic field variables of the Lagrangian. For our present purposes, this relation is not needed. Examples of such relations obtained from perturbative calculations are given in \cite{morikawa,gr} Turning to the equation of motion for the fluctuations, as stated above, it is obtained through the linearized deviation of the zero mode equation of motion, with account for the gradient term and noise. Although our interest in the zero mode dynamics is in the overdamped regime, where its equation of motion becomes first order in time, for obtaining the fluctuation equation of motion we must start with the initial second order zero mode equation. Compared to the zero mode equation of motion, the only additional term in the fluctuation equation of motion that could make it second order is the gradient term. This will happen if the gradient term creates a sufficiently large curvature ($\equiv -\nabla^2 \phi + m_{\phi}^2 \phi$) in the potential to overcome the large damping force term. However, this will not happen since our interest is in fluctuations that relatively are not very large. In particular, we will see below from explicit calculations that for the fluctuation wavenumbers of interest, the curvature term ($=\sqrt{{\bf k}^2+m_{\phi}^2}$) generally is smaller than the dissipative coefficient (recall this is the term multiplying $d\varphi_0/dt$ in Eq. (\ref{eom1})), and generically for a damped harmonic oscillator equation of motion this implies the overdamped (first order) regime. Thus, the equation of motion for the fluctuation $\delta \varphi({\bf k},\tau)$ is obtained from the equation of motion for $\varphi_0(\tau)$ that is given in Subsect. \ref{subsect3A} Eq. (\ref{eom1}) and derived in \cite{bgr}, \begin{equation} \Upsilon(\varphi_0,T) \frac{d \delta \varphi({\bf k},\tau)}{d \tau} = -({\bf k}^2+m^2_{\phi}(\varphi,T)) \delta \varphi({\bf k},\tau) +\xi({\bf k},\tau) \label{eomfluc} \end{equation} where \begin{equation} \Upsilon(\varphi_0,T) \equiv \sum_i^{t.e.}(\varphi_0-M_i)^2\eta_1(T) \Gamma_{\chi}(M) \label{upsdef} \end{equation} and $m_{\phi}(\varphi_0,T)$ is given in Eq. (\ref{rmphi}). This equation is obtained from Eq. ({\ref{eom1}) by substituting $\varphi({\bf x},t)$ and retaining terms linear in $\delta \varphi({\bf x},t)$ and Fourier transforming to k-space. In addition a noise function $\xi({\bf k}_p,t)$ is added which respects the fluctuation-dissipation theorem \cite{fdt} \begin{equation} <\xi({\bf k},t)>_{\xi} =0 \end{equation} \begin{equation} <\xi({\bf k},\tau) \xi(-{\bf k}^{\prime},\tau^{\prime}) >_{\xi} \stackrel{T\rightarrow \infty}{=} 2 \Upsilon(\varphi_0(\tau),T(\tau)) T(\tau) (2\pi)^3 \delta^{(3)}(\bf{k}-\bf{k}) \delta(\tau-\tau^{\prime}). \end{equation} The equation of motion (\ref{eomfluc}) represents the standard near-thermal-equilibrium dynamics in which $\xi({\bf k},\tau)$ drives the correlations of $\delta \varphi({\bf k},\tau)$ to thermal equilibrium with the relaxation rate of the initial conditions $({\bf k}^2+m^2_{\phi}(\varphi_0,T))/\Upsilon(\varphi,T)$. The relevance of Eq. (\ref{eomfluc}) to expanding background is that it approximates the equation of motion for a mode $\delta \varphi_{{\bf k}_c}(\tau)$ during a Hubble time interval say $\tau_i \stackrel{<}{\sim} \tau \stackrel{<}{\sim} \tau_{i+1}$ with $\tau_{i+1} - \tau_i \approx \Gamma_{\chi}(M)/H$. {}Furthermore, the momentum vector in the equation of motion Eq. (\ref{eomfluc}) is identified with the physical momentum of ${\bf k}_c$ which to a good approximation is fixed at one intermediate time during the respective time interval, such as $\tau_i$, ${\bf k} \rightarrow {\bf k}_p = {\bf k}_c \exp[-H(\tau)(\tau_i-\tau_{EI}]$. This approximation is valid for large ${\bf k}_p$, when the evolution of $H(\tau)$,${\bf k}_p(\tau)$ and $\varphi(\tau)$ is adiabatic relative to the evolution of $\delta \varphi_{{\bf k}_c}(\tau)$ during the respective time interval. Within the regime where this approximation is valid, the evolution of $\delta \varphi_{{\bf k}_c}(\tau)$ can be computed through piecewise construction of solutions for a sequence of Hubble time intervals, similar to the demonstration in \cite{ab1}. We will see below that the regime of ${\bf k}_p$ where the above approximation holds also is the appropriate regime for our purposes of estimating density perturbations. Consider what happens to a mode $\delta \varphi_{{\bf k}_c}(\tau)$ that is immersed in a heat bath and is in an expanding background. The larger ${\bf k}^2_p$ is, the faster is the relaxation rate. If ${\bf k}_p^2$ is sufficiently large for the mode to relax within a Hubble time, then that mode thermalizes. Thus at any instant during expansion, one can expect modes with physical momenta bigger than some lower bound $k_F$ to thermalize within a Hubble time interval. {}For these modes, within a single Hubble time interval, the flat-space equation of motion for the fluctuations Eq.(\ref{eomfluc}) is approximately valid. As soon as the physical wavenumber of a $\varphi({\bf x},\tau)$ field mode becomes less than $k_F$, it essentially feels no effect of the thermal noise $\xi({\bf k}_p,\tau)$ during a Hubble time. Thus for mode $\delta \varphi_{{\bf k}_c}(\tau)$, it essentially does not change once $|{\bf k}_p| \equiv |{\bf k}_c|\exp[-(H(\tau)/\Gamma_{\chi}(M))(\tau-\tau_{EI})] <k_F$, and at $|{\bf k}_p| =k_F$ the mode assumes its thermalized distribution. If $k_F > H(\tau)$, it implies the $\delta \varphi_{{\bf k}_c}(\tau)$ correlations that must be computed at time of Hubble radius crossing, $|{\bf k}_p(\tau)|=H(\tau)$, are the thermalized correlations that were fixed at $|{\bf k}_p(\tau)|=k_F$. This effect was clarified for warm inflation by Yokoyama and Linde \cite{yl} and they referred to it as ''freeze-out''. In order to determine $k_F$, consider the solution of Eq. (\ref{eomfluc}) for $\delta \varphi_{{\bf k}_c}(\tau)$ within the Hubble time interval $\tau_0 < \tau < \tau_0+\Gamma/H(\tau_0)$. We will ignore the time variation of ${\bf k}_p$, $\Upsilon(\varphi_0,T)$, and $m_{\phi}(\varphi_0,T)$ during this time interval. Their values at $\tau_0$ will be used ${\bf k}_p(\tau_0)$, $\Upsilon(\tau_0) \equiv \Upsilon(\varphi_0(\tau_0,T(\tau_o))$, and $m_{\phi}(\tau_0) \equiv m_{\phi}(\varphi_0(\tau_0),T(\tau_0))$. The approximate solution then is \begin{eqnarray} \delta \varphi_{{\bf k}_c}(\tau) & \approx & \frac{1}{\Upsilon(\tau_0)} \exp[-\frac{{\bf k}_p^2+m^2_{\phi}(\tau_0)}{\Upsilon( \tau_0)}(\tau-\tau_0)] \int_{\tau_0}^{\tau} \exp[\frac{{\bf k}_p^2+m^2_{\phi}(\tau_0)}{\Upsilon( \tau_0)}(\tau'-\tau_0)] \xi({\bf k}_p,\tau') d\tau' \nonumber\\ & + & \delta \varphi_{{\bf k}_c}(\tau_0) \exp[-\frac{{\bf k}_p^2+m^2_{\phi}(\tau_0)}{\Upsilon( \tau_0)}(\tau-\tau_0)]. \label{phiksol} \end{eqnarray} and for the corresponding correlation function \begin{eqnarray} \langle \delta \varphi_{{\bf k}_c}(\tau) \delta \varphi_{{\bf k}'_c}(\tau) \rangle_{\xi} & \approx & (2\pi)^3 \delta^{(3)}({\bf k}_p - {\bf k}_p') \frac{T}{[{\bf k}_p^2+m^2_{\phi}(\tau_0)]} \left[1.0-\exp\left(-\frac{2({\bf k}_p^2+m^2_{\phi}(\tau_0))} {\Upsilon(\tau_0)} (\tau-\tau_0)\right)\right] \nonumber\\ & + & \langle \delta \varphi_{{\bf k}_c}(\tau_0) \delta \varphi_{{\bf k}'_c}(\tau_0) \rangle_{\xi} \exp\left[-\frac{2({\bf k}_p^2+m^2_{\phi}(\tau_0))}{\Upsilon(\tau_0)} (\tau - \tau_0) \right]. \end{eqnarray} When the exponentially decaying terms are negligible, the above correlation is equivalent to the high-temperature correlation function $\langle \delta \varphi({\bf k}_p,\tau_0) \delta \varphi({\bf k}_p',\tau_0) \rangle_{T}$. In this solution Eq. (\ref{phiksol}), on the right hand side, the first term is the noise contribution that is driving the mode to thermal equilibrium and the second term contains the memory of the initial conditions at $\tau=\tau_0$, which are exponentially damping. By definition of freeze-out, for $|{\bf k}_p| \stackrel{>}{\sim} k_F$ the second term damps away within a Hubble time and for $|{\bf k}_p| \stackrel{<}{\sim} k_F$ it does not. To quantify the criteria, the freeze-out momentum $k_F$ is defined by the condition \begin{equation} \frac{{\bf k}_F^2+m^2_{\phi}(\varphi(\tau),T(\tau))}{\Upsilon( \varphi_0(\tau),T(\tau))} \frac{\Gamma_{\chi}(M)}{H(\tau)} = 1. \label{kfcond} \end{equation} This relation allows us to follow-up our discussion above Eq. (\ref{eomfluc}) about the fluctuation equation of motion being first order in time. From the above equation we see that the curvature term in the potential $\sqrt{{\bf k}^2 + m_{\phi}^2}$ is less than the dissipative coefficient (in this notation comparing to Eq. (\ref{eom1}) the dissipative coefficient is $\Upsilon/\Gamma_{\chi}$) since generically $H < \Upsilon/\Gamma_{\chi}$. As such, the overdamped equation of motion for the fluctuations, Eq. (\ref{eomfluc}), is justified for the wavenumbers of interest to us. Now we can write down the basic equations for calculating density perturbations during warm inflation. Properly this should be fully derived from the linearized General Relativity equations for perturbations \cite{bardeen,sasaki2,brandpert}. Nevertheless, an examination of the adiabatic density perturbations derivation for supercooled inflation \cite{guthpi1,amp} with the modification of a subdominant radiation component for the warm inflation case leaves unaltered the basic expression for $\delta \rho/\rho$. This conclusion was arrived at in our earlier papers \cite{bf2,wi} and independently by Yokoyama and Linde \cite{yl}. Therefore from this we find \begin{equation} \frac{\delta \rho}{\rho}(\tau) = \frac{V'(\varphi_0) \Delta \varphi_{H(\tau)}[k_F(\tau-{\tilde \tau}(\tau))]}{(\Gamma_{\chi}(M)d\varphi_0/d\tau)^2 +(4/3)\rho_r(\tau)} \approx \frac{6 H(\tau) \Delta \varphi_{H(\tau)}[k_F(\tau-{\tilde \tau}(\tau))]}{5 \Gamma_{\chi}(M)(d\varphi_0/d\tau)}. \label{delrho} \end{equation} The middle expression is the one used by Berera and Fang \cite{bf2} and the latter is the Guth and Pi expression \cite{guthpi1}. During warm inflation, since from Eq. (\ref{dotr}) $\rho_r \approx V'({\varphi_0}) \Gamma_{\chi}(M) (d\varphi_0/d\tau)/(4H(\tau))$ and $\rho_r \gg (\Gamma_{\chi}(M) d\varphi_0/d\tau)^2$, the middle and left expressions above are equivalent up to an O(1) constant. It would be interesting to investigate which of the two expressions is fundamentally more proper. In Eq. (\ref{delrho}), $\Delta \varphi_{H(\tau)}[k_F(\tau-{\tilde \tau}(\tau)]$ is the amplitude of the scalar field fluctuations and it is composed of all the modes whose physical wavelengths cross the Hubble radius within one Hubble time interval about time $\tau$. Recall that the comoving mode of physical wavenumber $|{\bf k}_p (\tau)| \sim H(\tau)$ had its amplitude frozen at an earlier time $\tau'$ when its physical wavenumber was $|{\bf k}_p (\tau')| \sim k_F(\tau')$. Since $|{\bf k}_p(\tau')|/|{\bf k}_p(\tau)| =\exp[(H(\tau) \tau - H(\tau') \tau')/\Gamma_{\chi}(M)]$, this implies $k_F(\tau') =H(\tau) \exp[(H(\tau)\tau - H(\tau') \tau')/\Gamma_{\chi}(M)]$. This defines ${\tilde \tau}(\tau) \equiv \tau - \tau'$. {}For the cases in the previous sections, $k_F(\tau)$ is slowly varying, so we will use the approximation $k_F(\tau-{\tilde \tau}(\tau)) \approx k_F(\tau)$. The expression for the scalar field amplitude is defined as the natural finite-temperature extension of the $T=0$ expression of supercooled scalar field inflation and with account for $k_F$. We use the definition \begin{equation} \Delta \varphi_H^2[k_F] \equiv \int_{k_F-shell}\frac{d^3{\bf k}_p}{(2\pi)^3} \int_{V} d^3{\bf x}_p \langle \delta \varphi({\bf x},\tau) \delta \varphi({\bf 0},\tau) \rangle_T \exp(-i{\bf k}_p \cdot {\bf x}_p) \stackrel{T \rightarrow \infty}{\approx} \frac{k_F T}{2 \pi^2}, \label{Dvp} \end{equation} where the $k_F$ - shell is defined as the spherical shell which is bounded between $k_F {\rm e}^{-1/2} < |{\bf k}_p| < k_F {\rm e}^{1/2}$ (we approximate the shell thickness simply to be $k_F$). The expression on the far right in the above equation is valid when $k_F < T$. The definition Eq. (\ref{Dvp}) is equivalent to the one given by Linde \cite{lindebook} in which one retains from $<\phi^2({\bf x}, \tau)>_T$ the contribution from wavenumbers within the $k_F$-shell. {}From Eq.(7.3.2) and Eq. (3.1.7) of Linde's book \cite{lindebook} this gives \begin{equation} \Delta \varphi_H^2[k_F] \equiv \frac{1}{2\pi^2} \int_{k_F-shell} \frac{k^2 dk}{\sqrt{k^2+m_{\phi}} \left[\exp(\sqrt{k^2+m_{\phi}}/T) -1 \right]}. \label{Dvplb} \end{equation} When $k_F > T$, Eq. (\ref{Dvp}) implies $\Delta \varphi_H^2(k_F)_{k_F > T} = k_F^2/(4\pi^2)$. {}For this region, it is a poor approximation to use the zero-mode dissipative coefficient in the k-mode equation of motion Eq. (\ref{eomfluc}). {}For this regime, a proper calculation of the k-mode dissipative function is important. Our calculations in the next two subsections consider only the high temperature expression on the right hand side of Eq. (\ref{Dvp}). Substituting the expression for $\Delta \varphi^2_H$ from Eq. (\ref{Dvp}) into Eq. (\ref{delrho}), the final expression for $\delta \rho/\rho$ for the mode that crosses the Hubble radius at time $\tau$ during warm inflation is\footnote{In our estimates of density perturbations, the equation of motion for $\delta \varphi_{{\bf k}_c}(\tau)$, Eq. (\ref{eomfluc}) uses the ${\bf k}=0$ dissipative function Eq. (\ref{eta1B}). The derivation in \cite{gr,bgr} suggests that the dissipative function $\eta_1(T)$ decreases as ${\bf k}$ increases, since it requires ``off-diagonal'' Green's functions. This would imply $k_F$ decreases since the relaxation rate is faster, thus any mode $\delta \varphi_{{\bf k}_c}(\tau)$ would thermalize faster at every physical wavenumber. Ultimately this means for a given ${\bf k}_c$ mode that is crossing the Hubble radius, $\delta \rho/\rho$ decreases relative to our estimates. In the future, a proper derivation of the dissipative function at non-zero k-vector would be useful.} \begin{equation} \frac{\delta \rho}{\rho}(\tau) \stackrel{k_F < T}{\approx} \frac{6[H(\tau)/\Gamma_{\chi}(M)] \sqrt{k_F(\tau) T(\tau)}} {5\sqrt{2} \pi d\varphi_0(\tau)/d\tau}. \label{delrhof} \end{equation} Recall, the comoving mode that crosses the Hubble radius at time $\tau$ is $|{\bf k}_c| \approx H(\tau) \exp[(H(\tau)/\Gamma_{\chi}(M)) (\tau-\tau_{EI})]$. {}From the above considerations, the final prescription for computing $\delta \rho/\rho$ is simple. {}First determine the freeze-out wavenumber from Eq. (\ref{kfcond}), and then substitute this in Eq.(\ref{delrhof}) along with expressions for $T(\tau)$, $H(\tau)$ and $d\varphi_0(\tau)/d\tau$ from the previous section. In the next two subsections this is done for the cases in the coinciding subsections of the last section. {}Finally Subsect. \ref{subsect5C} follows with a discussion of the results. \subsection{$T_{BI} \gg T_{EI}$} \label{subsect5A} This subsection considers the cases in Subsect. A of the last section. \subsubsection{\bf $N_r=0$} \label{subsubsect5A1} {}From Eq. (\ref{eomfluc}) and taking $\Upsilon$ in Eq. (\ref{upsdef}) the same as in Eq. (\ref{eomA}), the equation of motion for the fluctuation is \begin{equation} \frac{d \delta \varphi({\bf k}_p,\tau)}{d\tau} = -[{\bf k}^2_p + m_{\phi}^2(\varphi,T)] \frac{\pi^2}{\eta'\Gamma'ng^2 T^2(\tau)} \delta \varphi({\bf k}_p,\tau) =-21 y \left(\frac{\varphi_0(\tau)}{\varphi_{BI}}\right)^{2/7} \frac{{\bf k}^2_p}{\lambda \varphi_0^2(\tau)} \delta \varphi({\bf k}_p,\tau). \end{equation} The expression on the extreme right was obtained by substituting the expression for $y$ below Eq. (\ref{phia1}) and considering the regime where ${\bf k}_p^2 \gg m_{\phi}^2(\varphi_0,T) \approx \lambda \varphi_0^2(\tau)/2$. The freeze-out wavenumber from Eq. (\ref{kfcond}) is \begin{equation} k^2_F(\tau) = \frac{(H_{BI}/\Gamma_{\chi}(M)) \lambda \varphi_0^2(\tau)}{21y} \left(\frac{\varphi_0(\tau)}{\varphi_{BI}}\right)^{2/7}. \end{equation} Using this, from Eq.(\ref{delrhof}) we find \begin{eqnarray} \frac{\delta \rho}{\rho}(\tau) & \approx & \frac{6 (\eta'\Gamma')^{3/4}}{5 \pi^{7/2}} \left(\frac{H_{BI}}{\Gamma_{\chi}(M)}\right)^{3/4} \kappa_M^{-1/2} N^{1/4} g^{3/2} (y\tau+1)^{5/4} \nonumber\\ & = & \frac{6 (\eta'\Gamma')^{3/4}}{5 \pi^{7/2}} \alpha^{3/4} \beta^{3/4} \kappa_M^{1/4} N^{1/4} g^{3/2} (y\tau+1)^{5/4}. \end{eqnarray} The spectrum is very flat. Since the relation between time $\tau$ and the comoving wavenumber ${\bf k}_c$ crossing the Hubble radius at time $\tau$ is $\tau = -(\Gamma_{\chi}(M)/H(\tau))\ln(|{\bf k}_c|/H_{EI}) + \tau_{EI} \approx -(\Gamma_{\chi}(M)/H_{BI})\ln(|{\bf k}_c|/H_{BI}) + \tau_{EI}$, the variation of $\delta \rho/\rho \propto [y(-(\Gamma_{\chi}(M)/H_{BI})\ln(|{\bf k}_c|/H_{BI}) +\tau_{EI})+1]^{5/4}$ is only logarithmic. {}For example, for arbitrary e-folds $N_e$, $\delta \rho/\rho$ for the last scale that crosses the Hubble radius (smallest scale) is a factor $\beta^{-5/12}$ bigger than the first scale that crosses the Hubble radius (largest scale). Observe that this deviation from exact scale invariance is an outcome of the nontrivial thermodynamics and is not, in particular, initially inputed by hand by choosing a nonanalytic potential. \subsubsection{\bf $N_r \gg 2NT/T_{EI}$} \label{subsubsect5A2} {}From Eq. (\ref{eomfluc}) and taking $\Upsilon$ in Eq. (\ref{upsdef}) the same as in Eq. (\ref{eomA}), the equation of motion for the fluctuation is \begin{equation} \frac{d \delta \varphi({\bf k}_p,\tau)}{d\tau} = -[{\bf k}^2_p + m_{\phi}^2(\varphi,T)] \frac{\pi^2}{\eta'\Gamma'Ng^2 T^2(\tau)} \delta \varphi({\bf k}_p,\tau) =-6 y \frac{{\bf k}^2_p}{\lambda \varphi_0^2(\tau)} \delta \varphi({\bf k}_p,\tau). \end{equation} The expression on the extreme right was obtained by substituting the expression for $y$ below Eq. (\ref{tempA2}) and considering the regime where ${\bf k}_p^2 \gg m_{\phi}^2(\varphi_0,T) \approx \lambda \varphi_0^2(\tau)/2$. The freeze-out wavenumber from Eq. (\ref{kfcond}) is \begin{equation} k^2_F(\tau) = \frac{(H_{BI}/\Gamma_{\chi}(M)) \lambda \varphi_0^2(\tau)}{6y}. \end{equation} Using this and from Eq.(\ref{delrhof}) we find \begin{equation} \frac{\delta \rho}{\rho}(\tau) \approx \frac{6 \sqrt{2} (\eta'\Gamma')^{3/4}}{\pi^{7/2}} \frac{(H_{BI}/\Gamma_{\chi}(M))^{3/4} N^{3/4} g^{3/2}}{N_r^{1/2}} =\frac{6 \sqrt{2} (\eta'\Gamma')^{3/4}}{\pi^{7/2}} \frac{\alpha^{3/4} \beta^{3/4} \kappa_M^{3/4} N^{3/4} g^{3/2}} {N_r^{1/2}}. \end{equation} {}For this case, the spectrum is exactly flat, i.e. $\delta \rho/\rho$ is the same for all e-folds. \subsection{$T_{BI} \stackrel{>}{\sim} T_{EI}$} \label{subsect5B} This subsection considers the cases in Subsect. B of the last section. \subsubsection{$V_0> 0$} \label{subsubsect5B1} {}From Eq. (\ref{eomfluc}) and taking $\Upsilon$ in Eq. (\ref{upsdef}), the same as in Eq. (\ref{eomB}), the equation of motion for the fluctuation is \begin{equation} \frac{d \delta \varphi({\bf k}_p,\tau)}{d\tau} = -[{\bf k}^2_p + m_{\phi}^2(\varphi,T)] \frac{\pi^2 T(\tau)}{\eta'\Gamma'Ng^2 \kappa_M^3 M^3} \delta \varphi({\bf k}_p,\tau) =-\frac{3}{2}y \left(\frac{\varphi_0(\tau)}{\varphi_{BI}}\right)^4 \frac{{\bf k}^2_p}{\lambda \varphi_0^2(\tau)} \delta \varphi({\bf k}_p,\tau). \label{eomflucvog0} \end{equation} The expression on the extreme right was obtained by substituting the expression for $y$ below Eq. (\ref{tempB1}) and considering the regime where ${\bf k}_p^2 \gg m_{\phi}^2(\varphi_0,T) \approx \lambda \varphi^2_0(\tau)/2$. The freeze-out wavenumber from Eq. (\ref{kfcond}) is \begin{equation} k^2_F(\tau) = \frac{2(H(\tau)/\Gamma_{\chi}(M)) \lambda \varphi_0^2(\tau)}{3y} \left(\frac{\varphi_{BI}}{\varphi_0(\tau)}\right)^4. \end{equation} Using this and from Eq. (\ref{delrhof}) we find \begin{eqnarray} \frac{\delta \rho}{\rho}(\tau) & \approx & \frac{6 \sqrt{2} (\eta'\Gamma')^{3/4}}{5\pi^{7/2}} \frac{(H_{BI}/\Gamma_{\chi}(M))^{3/4} N^{3/4} g^{3/2}}{(N_r+2\kappa_MN)^{1/2}} (y\tau+1)^{9/8} \nonumber\\ & = &\frac{6 \sqrt{2} (\eta'\Gamma')^{3/4}}{5 \pi^{7/2}} \frac{\alpha^{3/4} \beta^{3/4} \kappa_M^{3/4} N^{3/4} g^{3/2}} {(N_r+2\kappa_M N)^{1/2}}(y\tau+1)^{9/8}. \end{eqnarray} {}The spectrum is nearly flat with deviations that are logarithmic. {}For arbitrary e-folds $N_e$, $\delta \rho/\rho$ at the last e-fold (smallest scale) is only a factor $\beta^{-9/4}$ bigger than at the first e-fold (largest scale). Recall that the regime for this approximation requires $1 > \beta \stackrel{>}{\sim} 0.5$. \subsubsection{$V_0 = 0$} \label{subsubsect5B2} {}From Eq. (\ref{eomfluc}) and taking $\Upsilon$ in Eq. (\ref{upsdef}) the same as in Eq. (\ref{eomB}), the equation of motion for the fluctuation is \begin{equation} \frac{d \delta \varphi({\bf k}_p,\tau)}{d\tau} = -[{\bf k}^2_p + m_{\phi}^2(\varphi,T)] \frac{\pi^2 T(\tau)}{\eta'\Gamma'Ng^2 \kappa_M^3 M^3} \delta \varphi({\bf k}_p,\tau) =-\frac{9}{5}y \left(\frac{\varphi_0(\tau)}{\varphi_{BI}}\right)^{10/3} \frac{{\bf k}^2_p}{\lambda \varphi_0^2(\tau)} \delta \varphi({\bf k}_p,\tau). \end{equation} The expression on the extreme right was obtained by substituting the expression for $y$ below Eq. (\ref{tempB2}) and considering the regime where ${\bf k}_p^2 \gg m_{\phi}^2(\varphi_0,T) \approx \lambda \varphi_0^2(\tau)/2$. The freeze-out wavenumber from Eq. (\ref{kfcond}) is \begin{equation} k^2_F(\tau) = \frac{5(H(\tau)/\Gamma_{\chi}(M)) \lambda \varphi_0^2(\tau)}{9y} \left(\frac{\varphi_{BI}}{\varphi_0(\tau)}\right)^{10/3}. \label{kfB2} \end{equation} In this case the Hubble parameter from Eq. (\ref{hubve0}) is $\tau$-dependent. Using this expression and the above condition for $k_F$, from Eq. (\ref{delrhof}) we find \begin{eqnarray} \frac{\delta \rho}{\rho}(\tau) & \approx & \frac{6 \sqrt{2} (\eta'\Gamma')^{3/4}}{5\pi^{7/2}} \frac{(H_{BI}/\Gamma_{\chi}(M))^{3/4} N^{3/4} g^{3/2}}{(N_r+2\kappa_MN)^{1/2}} (y\tau+1)^{9/20} \nonumber\\ & = &\frac{6 \sqrt{2} (\eta'\Gamma')^{3/4}}{5 \pi^{7/2}} \frac{\alpha^{3/4} \beta^{3/4} \kappa_M^{3/4} N^{3/4} g^{3/2}} {(N_r+2\kappa_M N)^{1/2}}(y\tau+1)^{9/20}. \label{delrB2} \end{eqnarray} {}The spectrum is nearly flat with deviations that are logarithmic. {}For arbitrary e-folds $N_e$, $\delta \rho/\rho$ at the last e-fold (smallest scale) is only a factor $\beta^{-9/8}$ bigger than at the first e-fold (largest scale). Again recall that the regime for this approximation requires $1 > \beta \stackrel{>}{\sim} 0.5$. \subsection{Discussion} \label{subsect5C} Some aspects of the results for density perturbations are discussed here. Recall from observation that $\delta \rho/\rho \sim 10^{-5}$ \cite{smoot}. {}From our results, for all four cases we find $\delta \rho/\rho \stackrel{<}{\sim} 10^{-2} \alpha^{3/4} \kappa_M^{3/4} N^{1/4} g^{3/2}$. Thus the parameters $\alpha$, $\kappa_M$, $N$ and $g$ must decrease $\delta \rho/\rho$ by at least three orders of magnitude. {}For observational consistency it is permissible if $\delta \rho/\rho < 10^{-5}$ from inflation, since post-inflationary mechanisms such as cosmic strings also can produce density perturbations. Overall consistency with respect to quantum field theory from Subsect. \ref{subsect3B} and observation, $\delta \rho/\rho \stackrel{<}{\sim} 10^{-5}$ and $N_e \stackrel{>}{\sim} 60$, can be achieved in a variety of ways. {}From the discussion Subsect. \ref{subsect4C}, Eq. (\ref{genalp}) is a general consistency regime for quantum field theory for arbitrary $N_e$. Combining this with the general form for $\delta \rho/\rho$ stated above, overall consistency can be achieved for all four solution regimes in general for \begin{equation} \frac{1}{\beta \kappa_M^2} \stackrel{<}{\sim} \alpha \stackrel{<}{\sim} {\rm min} \left(\frac{1}{\kappa_M^{1/3}}, \frac{\beta \kappa_M N}{100(N_r+2\kappa_M N)}\right). \label{genalpd} \end{equation} Alternatively, for arbitrary $\alpha$,$\beta$, and $\kappa_M$, $\delta \rho/\rho$ can be made arbitrarily small by requiring $g \rightarrow 0$, $Ng^a = {\rm const.}$ with $a<6$. Thus the model provides sufficient freedom to achieve consistency independently with respect to quantum field theory and observation. {}Furthermore, both alternatives for decreasing $\delta \rho/\rho$, $\alpha \rightarrow 0$ or $g \rightarrow 0$ improve the validity of the underlying approximations. In the former case the adiabatic regime is deepened and in the latter case perturbation theory is further justified. It is important to note the magnitude of the Hubble parameter within the consistency regime. {}From Eq.(\ref{hubdef}) recall that $H_{BI} = \alpha \beta \kappa_M \Gamma_{\chi}(M) \approx [\alpha \beta \kappa_M/(192 \pi)] (g^4 +f^2/8)$, with thermalization condition Eq. (\ref{therm}) requiring $\alpha < 1$. None of the consistency conditions so far have imposed any conditions on the self-coupling parameter $f$ except perturbation theory which requires $f \stackrel{<}{\sim} 1$. Thus irrespective of $g$, the bound can be given $H_{BI} \leq (2 \times 10^{-4}) \alpha \beta \kappa_M M$. There is some freedom to this bound. Observe that the entire calculation is independent of the number of bosonic decay channels for the respective $\chi_{ik}$-field. This reflects itself in the coupling constant dependent factor ${\cal C}$ in $\Gamma_{\chi}(T)$. Note that in the basic equations the product $\eta_1(T) \Gamma_{\chi}(M) $ always arises together, and because of this the coupling constant dependent factor, in our model ${\cal C} \approx (g^4+f^2/8)$, cancel. This is not a coincidence, since fundamentally $\eta_1(T) \sim 1/{\Gamma}$ as evident from the formalism of \cite{gr,bgr} and from heuristic arguments in \cite{hs1,yl}. {}For the model in this paper, each $\chi_{ik}$ field has only two decay channels, one to itself and the other to the $\phi$ field. The $\chi_{ik}$ field also may interact with other fields. {}For example consider the additional interaction $(g^2_{\sigma}/2) \sum_{j=1}^{N_{\sigma}} \chi_{ik}^2 \sigma_j^2$, where $\sigma_j$ are bosonic fields that only interact with $\chi_{ik}$. In our notation, these $\sigma_j$ fields are nondissipative heat bath fields\footnote{Such secondary interactions with $\phi$ also induce dissipative effects, though we will not treat that here}. Such interactions modify the $\chi_{ik}$ decay width as \begin{equation} \Gamma_{\chi_{ik}}(T) \stackrel{{\rm add} \ \sigma \ {\rm interactions}} {\rightarrow} \frac{T}{192 \pi} (g^4 +\frac{f^2}{8} + N_{\sigma} g^4_{\sigma}). \end{equation} If all the couplings $g^2_{\sigma} > 0$, then one must restrict the number of $\sigma$ fields, $N_{\sigma}$, since their interaction with $\chi_{ik}$ also adds a thermal mass contribution $\approx g^2_{\sigma} N_{\sigma} T^2/12$. In order to keep $m_{\chi_{ik}} < T$, it requires $g^2_{\sigma} N_{\sigma} \stackrel{<}{\sim} 12$. In the strong coupling limit $g^2_{\sigma} \sim 1$, this modification with its stated limits increases $\Gamma_{\chi}(M)$ by a factor $10^{2}$. In this case the bound on the Hubble parameter is $H_{BI} \leq (2 \times 10^{-2}) \alpha \beta \kappa_M M$. To go one step further, the thermodynamics of warm inflation in the previous sections would be unmodified if in total as many as $N_{\sigma} \stackrel{<}{\sim} \beta \kappa_M N$ $\sigma$-fields coupled to all the $\chi_{ik}$ fields that are thermally excited at a given instant. In this case $H_{BI} \leq (2 \times 10^{-3}) \alpha \beta^2 \kappa_M^2 N M$. {}For this case, precaution is necessary to keep $m_{\chi_{ik}} < T$. One way to achieve this is if the $\chi_i-\sigma_j$ couplings are not sign definite. Since the one-loop thermal mass correction is sensitive to the sign $ \propto \pm g^2_{\sigma}$ whereas $\Gamma_{\chi} \propto g^4_{\sigma}$ is not, in principle $\Gamma_{\chi}$ can be made arbitrarily large while the growth of the thermal mass correction is controlled. These modifications to $\Gamma_{\chi}$ may be useful for phenomenological applications. The final point addressed in this discussion is the general behavior in all four cases of the density perturbation formulas. The equation of motion for mode $\delta \varphi_{{\bf k}_c}({\bf k}_p,\tau)$ can be obtained from a linearized approximation to Eq. (\ref{eomgen}) and by replacing $\lambda \varphi_0^2 \rightarrow {\bf k}^2_p$ which gives \begin{equation} \frac{d \delta \varphi({\bf k}_p,\tau)}{d \tau} \sim \frac{{\bf k}_p^2}{\eta'\Gamma' N g^2 \kappa_M^2} D^{\gamma_{\eta}}(\tau) \delta \varphi({\bf k}_p,\tau) \ \ + \ \ {\rm noise}. \label{dphiapp} \end{equation} The decay function $D(\tau) \equiv (\varphi_0(\tau)/\varphi_{BI})$ and all exponents $\gamma$ are the same as defined in Subsect. \ref{subsect4C}. Once again, our focus is not on the (slowly varying) time dependence, but we have included the correct $\tau$-dependence for completeness. The quantity multiplying $\delta \varphi({\bf k}_p,\tau)$ on the right-hand-side of Eq. (\ref{dphiapp}) is the decay rate for the respective mode. Thus the freeze-out momentum for the modes is \begin{eqnarray} k_F & \sim & \left( \frac{H(\tau)}{\Gamma_{\chi}(M)}\right)^{1/2} (\eta'\Gamma')^{1/2} N^{1/2} g \kappa_M D^{(\gamma_{\eta}+\gamma_{\rho''})/2}(\tau) M \nonumber\\ & \sim & (\eta' \Gamma')^{1/2} \alpha^{1/2} \beta^{1/2} \kappa_M^{3/2} N^{1/2} g \kappa_M D^{(\gamma_{\eta}+\gamma_{\rho''})/2}(\tau) M. \label{kfgen} \end{eqnarray} Using this expression and the general expressions for $d\varphi_0(\tau)/d\tau$, Eq. (\ref{phialt}) and $\varphi_{BI}/M$, Eq.(\ref{pomgen}), implies \begin{eqnarray} \frac{\delta \rho}{\rho} & \sim & (\eta'\Gamma')^{3/4} \frac{(H_{BI}/\Gamma_{\chi}(M))^{3/4}N^{3/4} g^{3/2}}{(N_r+2\kappa_MN)^{1/2}} D^{(\gamma_{\eta}+\gamma_{\rho''}+2\gamma_T-4)/4}(\tau) (y\tau+1) \nonumber\\ & \sim & (\eta'\Gamma')^{3/4} \frac{\alpha^{3/4} \beta^{3/4} \kappa_M^{3/4} N^{3/4} g^{3/2}} {(N_r+2\kappa_MN)^{1/2}} D^{(\gamma_{\eta}+\gamma_{\rho''}+2\gamma_T-4)/4}(\tau) (y\tau+1). \label{delrgen} \end{eqnarray} Some features are worth mentioning. In Eq. (\ref{kfgen}), $k^2_F$ should be proportional to the Hubble expansion rate $H_{BI}$, since a slower expansion rate allows longer relaxation time, thus modes of lower $|{\bf k}_p|$ can equilibrate. The dependence of $\delta \rho/\rho$ on $H_{BI}$ in Eq. (\ref{delrgen}) is less than the naive linear behavior given by the defining formula Eq. (\ref{delrhof}). This arises because of $H_{BI}$ dependence induced by the dynamics on the other factors, $k_F^{1/2} \propto H_{BI}^{1/4}$ and $(T/\varphi_{BI})^{1/2} \propto H_{BI}^{-1/2}$. \section{Examples} \label{sect6} In this section the solutions from the last two sections are studied in a few examples. It is not the purpose of this paper to detail the phenomenological consequences of these solutions. However, here we would like to obtain some idea about the absolute scales for the various dimensional quantities and how they chance in various parametric regimes as well as in the limit of increasing adiabaticity. In much of this section, the numerical value of premultiplying constants are evaluated in various expressions and we set $\eta'=48$, $\Gamma'=1/192$. Before turning to the examples, note that in our construction, since the Hubble parameter, Eq.(\ref{hubdef}), is proportional to $\Gamma_{\chi}(M)$, the scales of all warm inflation quantities are controlled by this decay rate. Recalling our comments from the discussion Subsect. \ref{subsect5C} about modifying ${\cal C}$ in $\Gamma_{\chi}(M)$, in ``our model'' ${\cal C} = g^4 +f^2/8$ and in the ``extended model'' considered in Subsect. \ref{subsect5C} where $\sigma$-fields were introduced ${\cal C}=g^4+f^2/8 + g^4_{\sigma}N_{\sigma}$. In the three subsections that follow, we quote estimates for both ``our model'' and the ``extended model''. All the estimates to follow always are within the observationally consistent regime with respect to expansion e-folds $N_e \stackrel{>}{\sim} 60$ and density perturbation $\delta \rho / \rho \stackrel{<}{\sim} 10^{-5}$. In Subsect. \ref{subsect6A} a general estimate of the scales $M,T_{BI}, H_{BI}, m_{\phi_{BI}}$, and $i_{min}M$ are given that encompasses the four cases studied in the last two sections. Subsects. \ref{subsect6B} and \ref{subsect6C} focuses on the case studied in Subsects. \ref{subsubsect4B2} and \ref{subsubsect5B2} for $V_0 =0$ and $T_{EI} \stackrel{<}{\sim} T_{BI}$. In Subsect. \ref{subsect6B}, the dependence of the scales and parameters are examined in the limit of arbitrary adiabaticity $\alpha \rightarrow 0$. {}Finally in Subsect \ref{subsect6C} the warm inflation solutions are examined as the inflaton self-coupling parameter $\lambda$ varies over a wide range including $\lambda \sim 1$. Within the limits of the present analysis, we will find that $\lambda$ is restricted to be tiny. However, a possibility is examined that could increase $\lambda$ up to $10^{-4}$, within a regime that is consistent with observational requirements on e-folds and density perturbations. \subsection{Estimate of Scale} \label{subsect6A} The absolute scale of all dimensional warm inflation quantities in our solution are determined once the Planck mass is introduced $M_p \approx 10^{19} {\rm GeV}$. Our solutions are constructed such that all the dimensional quantities have been expressed in terms of M. To determine $M$, note that the Hubble parameter $H_{BI}$ can be expressed in two way, by its definition $H_{BI} \equiv \sqrt{8\pi \rho(0)/(3m_p^2)} \approx \sqrt{8\pi \rho_v(0)/(3m_p^2)}$ and the expression Eq. (\ref{hubdef}) $H_{BI} = \alpha \beta \kappa_M \Gamma_{\chi}(M)$. {}For $\rho_v(0)$, from the expressions for the ratio $\rho_r/\rho_v$ Eqs. (\ref{rhorvA1}),(\ref{rhorvA2}), (\ref{rhorvB1}), and (\ref{rhorvB2}), note that $\rho_v(0) \approx N_e \rho_r(0) =N_e \pi^2 (N_r+2\kappa_MN)T^4_{BI}/16 = N_e \pi^2 (N_r+2\kappa_MN)\kappa_M^4 M^4/16$. Using this in the first expression for $H_{BI}$ and Eq. ({\ref{chidecay}) in the second and equating the two, we find \begin{equation} M= \frac{\sqrt{3} \Gamma'}{\pi^{5/2}} \frac{\alpha \beta {\cal C}} {\sqrt{N_e N (1+N_r/(2\kappa_MN))} \kappa_M^{3/2}} M_p. \label{meqn} \end{equation} An approximate upper bound can be obtained using Eqs. (\ref{thermaA1}), (\ref{thermaA2}), (\ref{thermaB1}), and (\ref{thermaB2}) which imply $\alpha \stackrel{<}{\sim} \beta/[200(1+N_r/(2\kappa_MN))]$ and Eqs. (\ref{forcecA1}), (\ref{forcecA2}), (\ref{forcecB1}), and (\ref{forcecB2}) which imply $\kappa_M \stackrel{>}{\sim} 10\sqrt{1+N_r/(2\kappa_MN)}/\beta$. Substituting these bounds into the above equation (\ref{meqn}) gives \begin{equation} M \stackrel{<}{\sim} \frac{(8.2 \times 10^{-8}) \beta^{7/2} {\cal C}} {\sqrt{N_e N} (1+N_r/(2\kappa_MN))^{9/4}} M_p. \label{minequal} \end{equation} {}From this expression and the bound on $\kappa_M$ it follows that \begin{equation} T_{BI} \stackrel{<}{\sim} \frac{(8.2 \times 10^{-7}) \beta^{5/2} {\cal C}} {\sqrt{N_e N} (1+N_r/(2\kappa_MN))^{7/4}} M_p, \label{tinequal} \end{equation} \begin{equation} H_{BI} \stackrel{<}{\sim} \frac{(6.8 \times 10^{-12}) \beta^{9/2} {\cal C}^2} {\sqrt{N_e N} (1+N_r/(2\kappa_MN))^{11/4}} M_p, \label{hubinequal} \end{equation} \begin{equation} m_{\phi_{BI}} \stackrel{<}{\sim} \frac{(4.4 \times 10^{-8}) \beta^{5/2} (1-\beta)^{1/2} g {\cal C}} {N_e (1+N_r/(2\kappa_MN))^{4}} M_p, \label{mpinequal} \end{equation} and $g \varphi_{EI}$ or equivalently the scale of the mass levels $\sim i_{\rm min} M$ in the DM-model is \begin{equation} i_{\rm min} M \stackrel{<}{\sim} \frac{(1.0 \times 10^{-4}) \beta^{3 \pm {\rm few}} N_e^{1/2} {\cal C}} {(1-\beta^{O(1)}) N^{1/2} (1+N_r/(2\kappa_MN))^{3/4}} M_p. \label{imininequal} \end{equation} To quote some numbers, consider a typical case with $\beta =0.5$, $N_e=60$, $N_r=0$, $N=5$ and ``our model'' with $f \approx 1$ so that ${\cal C} \approx 1/8$ (``extended model'' with $g_{\sigma} \approx 1$, $N_{\sigma} \approx 12$, so that ${\cal C} \approx 12$). {}For this case we find $M \stackrel{<}{\sim} 5.2 \times 10^8 {\rm GeV}$ ($\stackrel{<}{\sim} 5.0 \times 10^{10} {\rm GeV}$), $H_{BI} \stackrel{<}{\sim} 2.7 \times 10^3 {\rm GeV}$ ($\stackrel{<}{\sim} 2.5 \times 10^7 {\rm GeV}$), $T_{BI} \stackrel{<}{\sim} 1.0 \times 10^{10} {\rm GeV}$ ($\stackrel{<}{\sim} 1.0 \times 10^{12} {\rm GeV}$), $i_{\rm min} M \stackrel{<}{\sim} 1.2 \times 10^{14} {\rm GeV}$ ($\stackrel{<}{\sim} 1.0 \times 10^{16} {\rm GeV}$), and $T_{EI} =T_{BI}/2$. We also find $N_M \stackrel{>}{\sim} 2.6 \times 10^5$ mass sites are crossed. If we set $\kappa_M$ at its lower bound, $\kappa_M = 1/\sqrt{2\alpha \beta}$, and require $\delta \rho/\rho \leq 10^{-5}$, then in all four cases within the high temperature regime, $k_F < T$, it requires $g \leq 0.2$. {}From this it follows that $\lambda \leq 10^{-16}$ and $m_{\phi_{BI}} \stackrel{<}{\sim} (2.1 \times 10^{-3}) T_{BI}$. The high temperature validity regime for the density perturbation results in Sect. \ref{sect5} require $k_F < T$. {}From the expressions for $k_F$ in the four cases, we find in general $k_F \sim \sqrt{N_e} m_{\phi}$. Thus the estimates given here involving density perturbations are valid for $N_e < 2 \times 10^5$. {}Finally, for both our and the extended models, the thermalization rate $\Gamma_{\chi}(T)$ is about 400 times faster than the Hubble expansion rate $H(\tau) \stackrel{<}{\sim} H_{BI}$, so that the thermalization approximation is well satisfied. Although $\lambda$ is tiny, the inflaton mass, $m_{\phi}$ is large relative to the Hubble parameter. In the above example $m_{\phi}$ is three orders of magnitude below the temperature scale but four orders of magnitude larger than the Hubble parameter. The smallness of $\lambda$ preempts questions about fine tuned potentials, similar to the situation in supercooled dynamics. This point briefly is addressed in Subsect. \ref{subsect6C}. However, it should be noted that for these tiny values of $\lambda$, when the thermal damping is removed after the mass site $M_{i_{\rm min}}$, the potential does not support inflation. Once the thermal damping is removed, the only damping term that remains is due to the coupling of the inflaton to the background cosmology. This yields a $3H {\dot \varphi}_0$ term that is familiar from supercooled inflationary dynamics. The inflaton equation of motion then becomes $3Hd\varphi_0/dt = -\lambda \varphi_0^3/6= -m_{\phi}^2 \varphi_0/3$. Thus in a Hubble time $\Delta t = 1/H$, $|\Delta \varphi_0/\varphi_0| \approx m_{\phi}^2/(9H^2) \gg 1$, so that $\varphi_0$ rapidly falls down the potential. In words, the curvature of the potential is huge relative to the scale of the Hubble expansion rate. As such, to terminate warm inflation in this model and go into a radiation dominated regime, it suffices simply to stop coupling the inflaton to mass sites. \subsection{The Limit of Arbitrary Adiabaticity} \label{subsect6B} The self consistency of the near-thermal-equilibrium quantum field theory formalism applied in this paper requires satisfying the conditions in Subsect. \ref{subsect3B}. Based on the solutions in the previous two sections, we find that the most constraining consistency conditions are the thermalization and thermal-adiabatic conditions Eqs. (\ref{therm}) and Eq. (\ref{thermadiab}) respectively. As it turns out, both these conditions are controlled by a single parameter in our solutions, $\alpha$, with the validity for both conditions improving as $\alpha \rightarrow 0$. In this subsection the observationally consistent regime with respect to $N_e$ and $\delta \rho/\rho$ is studied as a function of $\alpha$, in particular, in the limit of arbitrary adiabaticity $\alpha \rightarrow 0$. We focus on the case in Subsects. \ref{subsubsect4B2} and \ref{subsubsect5B2} for $V_0=0$, $T_{EI} \stackrel{<}{\sim} T_{BI}$. {}For the case of interest, from Subsect. \ref{subsubsect4B2} the consistency conditions are \begin{equation} \alpha < \frac{1}{2\beta \pi^4} \end{equation} and \begin{equation} \kappa_M > \frac{1}{2\sqrt{2}\alpha^{1/2} \beta^{1/2}}, \label{kaplimB2} \end{equation} where throughout this subsection we set $N_r=0$. The mass splitting scale parameter M is determined by the same procedure as in the previous subsection. We find \begin{equation} M=\frac{3\Gamma'}{2\pi^{5/2}} \frac{\alpha(1-\beta)^2 \beta^{1/2} {\cal C}} {\sqrt{NN_e}\kappa_M^{3/2}} M_p. \label{m6b} \end{equation} The other dimensional quantities can be determined easily from this. In Fig. 1, the $\alpha$ dependence of all dimensional scales are shown for two cases. $\kappa_M$ is set to its lower bound, $\kappa_M = 1/(2\sqrt{2}\alpha^{1/2} \beta^{1/2})$, with always the restriction $\kappa_M>1$. The limit of arbitrarily increasing adiabaticity is $\alpha \rightarrow 0$ ($-\log_{10}(\alpha) \rightarrow \infty$). All the scales, $M,T_{BI},H_{BI},i_{\rm min}M$, and $m_{\phi_{BI}}$, are in GeV with their $\log_{10}$ plotted. The solid lines are for the case $N=5$, $N_e=65$, $\beta=0.5$ and ${\cal C}=1/8$ (``our model''). The dashed lines are for the case $Ng^4=1/8$, $N_e=65$, $\beta=0.5$ and ${\cal C}=1/8$ (``our model''). {}For the ``extended model'', ${\cal C}=12$, all scales in Fig.1 are shifted up by a factor $\sim 10^2$ except for $H_{BI}$ which is shifted up by a factor $\sim 10^4$. For the region to the left of the dotted vertical line at 0.6 ($\alpha > 0.25$), $\kappa_M=1$. $M,T_{BI},H_{BI}$, and $i_{\rm min}M$ are independent of $\delta \rho/\rho$, whereas $m_{\phi} \propto g$, so it depends on $\delta \rho/\rho$ since from Eq. (\ref{delrB2}) \begin{equation} g \approx 30.45 \frac{(\delta \rho/\rho)^{2/3}} {\alpha^{5/12} \beta^{5/12} N^{1/6}}. \label{g6b} \end{equation} For $\alpha \rightarrow 0$ with everything else fixed, $g$ increases. Since $g<1$ is required by perturbation theory, the model requires $\delta \rho/\rho \rightarrow 0$ as $\alpha \rightarrow 0$. In Fig. 1, we set $\delta \rho/\rho= 10^{-5}$ down to the smallest $\alpha$ possible, which is given by the vertical solid and dashed lines for the two respective cases. To the right of these lines (smaller $\alpha$), $\delta \rho/\rho$ is less than $10^{-5}$. {}For the other parameters in the model, as $\alpha$ ranges from 0 to 1 ($-\log_{10}(\alpha)$ ranges from $\infty$ to 0), for the solid case the ranges are $g$ from 1 to 0.013, $\lambda$ from 0 ($\propto \alpha$) to $4 \times 10^{-17}$, $\kappa_M$ from $\infty$ ($\propto 1/\alpha^{1/2}$) to 1, and $N_M$ from $\infty$ ($\propto 1/\alpha$) to 350; for the dashed case the ranges are $g$ from 0.59 to $1 \times 10^{-5}$, $N$ from 1 to $3 \times 10^{18}$, $\lambda$ from 0 ($\propto \alpha$) to $4 \times 10^{-11}$, $\kappa_M$ from $\infty$ ($\propto 1/\alpha^{1/2}$) to 1, and $N_M$ from $\infty$ ($\propto 1/\alpha$) to 350. {}For $\alpha < 1/\pi^4$, which is in the region to the right of the dot-dashed vertical line at $-\log_{10}(\alpha)=1.9$, all adiabaticity conditions are valid. To the left of this vertical line, the thermal-adiabatic condition in its stringent form is not valid. However, as discussed in Subsect. \ref{subsect3C}, the thermal-adiabatic condition may still hold in some part of this region. To determine the extent to which the thermal-adiabatic condition can be relaxed requires details about thermalization that go beyond the simple high-temperature approximations applied in this paper. The DM-model warm inflation calculation in \cite{bgr2} is similar to the case in Subsect. \ref{subsubsect4B2} for $V_0=0$, $T_{EI} \stackrel{<}{\sim} T_{BI}$. The difference is \cite{bgr2} ignores the minor modifications that arise due to time dependence of the temperature, whereas this was treated in Subsect. \ref{subsubsect4B2}. The region studied in \cite{bgr2} is for $0.5 < \alpha \leq 1$. This was chosen for its simplicity in illustrating the basic features of the results. In this region, all the consistency conditions are satisfied except the stringent form of the thermal-adiabatic condition. Based on earlier discussions in this paper, this region is still within the plausible validity region. We have verified that in the region of overlap the results in \cite{bgr2} agree with thoses in Subsect. \ref{subsubsect4B2}. \subsection{$\lambda$ Dependence of the Solution} \label{subsect6C} In this subsection, we examine the $\lambda$ dependence of our solution for the case from Subsects. \ref{subsubsect4B2} and \ref{subsubsect5B2}. {}For this, we treat $\lambda$ as an independent variable in exchange for $\kappa_M$, which from Eq. (\ref{lamB2}) gives \begin{equation} \kappa_M = (1.2 \times 10^8) \frac{\beta^3 N_e^3 \lambda}{(1-\beta)^3 N g^4} \label{kappalam} \end{equation} where $\alpha = (2\beta \pi^4)^{-1}$ has been set to its upper bound and we only consider the regime with $N_r=0$. Here and throughout this subsection, the numerical values of all constants are quoted and we have used $\Gamma'=1/192$, $\eta'=48$. The parameters on the right-hand-side of Eq. (\ref{kappalam}) can be varied freely up to the mild constraints on $\kappa_M$, Eq. (\ref{kaplimB2}). Substituting Eq. (\ref{kappalam}) into our solutions in Subsects. (\ref{subsubsect4B2}) and (\ref{subsubsect5B2}) we find \begin{equation} M=(1.8 \times 10^{-18}) \frac{(1-\beta)^{13/2} Ng^6 {\cal C}} {\beta^5 N_e^5 \lambda^{3/2}} M_p, \label{mlam} \end{equation} \begin{equation} T_{BI}=(2.1 \times 10^{-10}) \frac{(1-\beta)^{7/2} g^2 {\cal C}} {\beta^2 N_e^2 \lambda^{1/2}} M_p, \label{tlam} \end{equation} \begin{equation} H_{BI}=(1.8 \times 10^{-15}) \frac{(1-\beta)^{7/2} g^2 {\cal C}^2} {\beta^2 N_e^2 \lambda^{1/2}} M_p, \label{hlam} \end{equation} \begin{equation} m_{\phi_{BI}}=(3.9 \times 10^{-8}) \frac{(1-\beta)^{5/2} g {\cal C}} {\beta N_e } M_p, \label{mplam} \end{equation} and $N_M=(3 \times 10^{10}) \beta^4 (1-\beta^{3/4}) \lambda N_e^4/[ (1-\beta)^4 N g^4]$ so that \begin{equation} i_{\rm min} M=(5.5 \times 10^{-8}) \frac{(1-\beta)^{5/2} (1-\beta^{3/4}) g^2 {\cal C}} {\beta^{17/4} N_e \lambda^{1/2}} M_p. \label{imlam} \end{equation} Also we express $g$ in terms of $\lambda$ and $\delta \rho/\rho$ to obtain \begin{equation} g= (4.4 \times 10^3) \frac{(1-\beta)^{3/2} (\delta \rho/\rho)^2} {\beta^{3/2} N_e^{3/2} \lambda^{1/2} } \label{gdelr} \end{equation} This can be substituted in the above expressions for the scales, but the perturbative restriction must be respected $g<1$. In the observationally consistent regime, $\delta \rho/\rho \stackrel{<}{\sim} 10^{-5}$, this perturbative restriction is comfortably satisfied for a wide range of the parameters. It is interesting to examine the maximum size of $\lambda$ within the observationally consistent regime of warm inflation. {}For satisfying just the horizon/flatness problems, the $\lambda \sim 1$ regime has solutions if also $g \sim 1$. However the primary restrictions arise from the density perturbation constraints. If we restrict to only the high temperature regime for the freeze-out momentum, the inequality $k_F < T$ must be imposed to Eq. (\ref{kfB2}). Reordering this inequality to isolate $\lambda$ and using the solutions from Subsect. \ref{subsubsect4B2}, we find the constraint \begin{equation} \lambda < (6.6 \times 10^{-5}) \frac{(1-\beta)^3 g^2}{\beta^3 N_e^3}. \label{lamlim} \end{equation} Setting $\lambda$ at its upper limit and substituting into Eq. (\ref{gdelr}), for $N_e=65$, $\delta \rho/\rho = 10^{-5}$, $\beta=0.5$, we find $g \approx 7.2 \times 10^{-3}$, which from Eq. (\ref{lamlim}) implies $\lambda \approx 1 \times 10^{-14}$. This value of $\lambda$ is two orders of magnitude larger than the limit in Subsect. \ref{subsect6A}, because here $\kappa_M \approx (1.4 \times 10^8)/N$ is not at its lower bound. With these values for $\lambda$,$g$, $N_e$, $\beta$ , and leaving $N$ unspecified, from Eqs. (\ref{mlam}) - (\ref{imlam}) for ''our model'' with ${\cal C}=1/8$ (''extended model'' with ${\cal C}=10$) we find the scales in GeV $M \approx 0.7N (6N)$, $T_{BI} \approx 1 \times 10^7$ ($8 \times 10^8$), $H_{BI} \approx 11$ ($7 \times 10^4$), $m_{\phi_{BI}} \approx 2 \times 10^6$ ($2 \times 10^8$), and $N_M= 10^{12}/N$ so that $i_{\rm min} M \approx 2 \times 10^{11}$ ($5 \times 10^{13}$). {}For $k_F > T$, as stated earlier, the high-temperature calculations in the previous section are not valid. Suppose that thermal dissipation is inactive for wavenumbers larger than the temperature $k_F > T$. In this case, if the freeze-out wavenumber from Eq. (\ref{kfB2}) is larger than $T$, the $\varphi$-amplitude should be set to its limiting value $\Delta \varphi^2_H \approx T^2/(2\pi^2)$. In this case, the density perturbation expression in Subsect. \ref{subsubsect5B2} is replaced by $\delta \rho/\rho \approx 3 \alpha^{1/2} \beta^{1/2} g (y\tau+1)^{3/10}/ (5\pi^3)$. {}From the previous paragraph, $k_F > T$ corresponds to the region of $\lambda$ in Eq. (\ref{lamlim}) with $<$ replaced by $>$. {}For this hypothetical case, if we set $N_e=65$ and $\beta=0.5$, we find $\kappa_M \approx (1 \times 10^{22}) \lambda /N$, $M \approx (8 \times 10^{-41}) N {\cal C} M_p/\lambda^{3/2}$, $T_{BI} \approx (9 \times 10^{-19}) {\cal C} M_p/\lambda^{1/2}$, $H_{BI} \approx (8 \times 10^{-24}) {\cal C}^2 M_p/\lambda^{1/2}$, $m_{\phi_{BI}} \approx (2 \times 10^{-12}) {\cal C} M_p$, and $N_M \approx (8 \times 10^{25}) \lambda/N$ so that $i_{\rm min}M \approx (6 \times 10^{-14}) {\cal C} M_p/\lambda^{1/2}$. If the temperature scale for inflation is assumed to be above 1 TeV, it requires ${\cal C}/\lambda^{1/2} > 100$, which for ${\cal C}=10$ implies the self-coupling can be as large as $\lambda \sim 0.0001$, although it also requires a very large number of fields. In any case, it is interesting to examine the difficulties that must be overcome in this model to avoid an ultra-flat inflationary potential. The fact that such a possibility in remotely realizable is interesting and motivates further investigation. In summary, the magnitude of $\lambda$ is dependent mutually on the overdamped dynamics of warm inflation and the density perturbation requirement. {}Further development of the simple dynamical framework used here may increase $\lambda$ by several orders. \section{Conclusion} \label{sect7} In this paper, warm inflation solutions have been obtained for the DM model that solve the cosmological horizon, flatness and scalar density perturbation problems. {}Furthermore such solution regimes exist for an arbitrarily slow evolution of all macroscopic variables in the inflaton field system and the background cosmology. {}For convenience, the DM models considered in this paper had all adjacent mass sites equally spaced, $g|M_{i+1}-M_i| = M$. This condition is not required to obtain warm inflation solutions. Spacings between adjacent sites can widely vary. The inflaton motion remains overdamped for a succession of sites in which adjacent spacings are all less than the temperature $T$. Spacings larger than $T$ are not inconceivable, but the inflaton motion is more complicated. {}For any DM model spacings, the ultimate test is the the usefulness of warm inflation solutions that it yields. We have examined only the single case of equal spacings. There are a few improvements to these calculations that can be made. Our calculations have adhered to the high-temperature approximation, which means fields with mass $m \leq T$ are thermally active and those with $m> T$ are thermally dormant. This approximation is over restrictive. Generally, fields participate in dissipative and thermalization dynamics once $m \leq 10T$. An example that treated dissipative heat bath fields with $m \leq 2.5T$ is in \cite{bgr2}. It is worthwhile to extend the present calculation beyond the high-temperature approximation. Another improvement to this calculation is to compute the ladder resummed dissipative function similar to the shear viscosity case computed by Jeon in \cite{jeon}. This point was noted in \cite{bgr} and recently has been verified by Jeon \cite{jeonpc}. The model studied in this paper has been developed into a string theory warm inflation scenario in \cite{bk,bk2}. The DM model has an essential feature for this interpretation. Its hierarchy of mass sites are reminiscent of the tower of mass levels of a string. The first step towards this interpretation was to obtain the DM model from a SUSY superpotential \cite{bk}. This has a relevance independent of the string interpretation. It establishes that the DM model is natural in the sense of nonrenormalization theorems, which means once the parameters are chosen, they stay fixed until SUSY is broken. The string picture developed in \cite{bk,bk2} interprets the DM model as an effective SUSY model in which the inflaton is a string zero mode and it interacts with higher string mass levels, which are the dissipative heat bath fields. Since the multiplicity of degenerate string states increases exponentially with excitation level, strings can provide an adequate supply of dissipative heat bath fields. Interestingly, the dispartity in scales that generally arises in DM model warm inflation realizations, $i_{\rm min} M \gg T_{BI} > M$, is readily explained in the string interpretation. $i_{\rm min} M$ of the DM model corresponds in the string interpretation to the mass scale of a string level for the unperturbed string, i.e., when the coupling to $\phi$ is switched off. The mass splitting scale $M$ corresponds in the string interpretation to a fine structure splitting of a initially degenerate string mass level. This fine structure splitting of the level arises from symmetry breaking. As temperature drops below the scale of the string mass levels, generally degeneracies in the mass levels will be lifted and thereby create fine structure splittings. Since for the warm inflation solutions in this paper $T_{BI} \ll i_{\rm min} M$, the conditions are adequate for various perturbations to break the degeneracy of the mass level. {}For example, if a symmetry breaking occurs at scale $i_{\rm min} M > v_1 > T_{BI}$, the states of any mass level split by characteristic scale $v_1$. Thus after symmetry breaking, the states of an initially degenerate mass level shift within a width of order $v_1$. A generic symmetry breaking typically will not lift all the degeneracy, so that a mass level with ${\cal N}$ degenerate states before symmetry breaking splits into ${\cal D} < {\cal N}$ finely split levels. In this case, the fine structure splitting scale is $M \sim v_1/{\cal D}$. Examples of symmetry breaking scenarios and estimates of ${\cal N}$, ${\cal D}$ and $v_i$ are given in \cite{bk2}. {}For the standard type I, II, heterotic and bosonic strings, the string scale is $M_S \sim 10^{17} {\rm GeV}$. Therefore the string interpretation requires $i_{\rm min} M \sim 10^{17} {\rm GeV}$. {}For the first mass level, $n=1$, of the heterotic string, for example, ${\cal N}\approx 2 \times 10^{7}$ states. ${\cal D}$ depends on the specific symmetry breakings that occurs. {}For a GUT motivated example in \cite{bk2} we found ${\cal D} \stackrel{<}{\sim} 10^5$. Comparing these estimates to the bounds in Subsect. \ref{subsect6A}, $i_{\rm min}M$ is at least one order of magnitude below $M_S$. The example in \cite{bk2} found $v_1/M \approx 10^5$. Since $T_{BI} < v_1$, to satisfy the example in \cite{bk2} requires $T_{BI}/M < 10^5$. In Subsect. \ref{subsect6A} we found $T_{BI}/M \approx 10^2$ with $T_{BI} \approx 10^{10-12} {\rm GeV}$ so that $v_1 \approx 10^{13-15} {\rm GeV}$. Thus the estimates from the model in this paper are somewhat compatible with the string scenario in \cite{bk2}. $i_{\rm min}M$ in our model still is below the string scale $M_S$. $v_1$ is within range of the GUT scales. The improvements to this calculation discussed above should elevate $i_{\rm min}M$ to $M_S$ and allow greater flexibility in the range of $T_{BI}$ and $M$. This good news comes at the expense of a finely tuned self-coupling parameter $\lambda$. The nature of this fine tuning problem is different from a similar problem in supercooled inflation scenarios \cite{ni,ci}. In our model, the dynamical parameter $m_{\phi}^2 \propto \lambda \varphi_0^2$ is quite large relative to the scale of the Hubble parameter. {}From Subsect. \ref{subsect6A} we find in our model $m_{\phi}/H_{BI} \approx 10^{2-4}$. In contrast, in supercooled scenarios $m_{\phi} \stackrel{<}{\sim} H$. {}Furthermore Subsect. \ref{subsect6C} demonstrated that $\lambda$ need not be tiny for observationally interesting warm inflation. However this required all the scales to be much smaller and it required an extremely large number of fields. Both these requirements conflict with the properties of conventional strings. These are disappointing features of this model. Nevertheless, there are indications that in this warm inflation dynamics, the density perturbation requirements do not mandate a tiny $\lambda$. Dissipative dynamics provides other means for bounding the density perturbation amplitude. The present model, despite some of the shortcomings, demonstrates the importance of the full decohereing dynamics in determining the density perturbations. In this paper, particle production has been treated to the minimal extent of its implications for strong dissipation and solving the cosmological puzzles. This comes short of attempting a detailed modeling of the produced particle spectrum. This direction has some interesting possibilities, which will be discussed here. An important fact to note is that the heat bath particles produced in this scenario are generally very massive, $ \sim 10^{16} - 10^5 {\rm GeV}$. Recall in this model, as the inflaton amplitude $\varphi_0$ relaxes to its minimum, the heat bath fields at a given site $i$ become thermally excited when their mass Eqs. (\ref{rmchi}), (\ref{rmpsi}) $m_{\chi_{ik},\psi_{ik}} \propto |\varphi_0-M_i|$ falls below the temperature and otherwise they are thermally unexcited. In order to realize this picture, an elementary requirement is that the microscopic time scale of thermalization must be faster than the characteristic time scale for the mass of the particle to change, as reflected through the thermal-adiabatic condition Eq. (\ref{thermadiab}). In general this condition becomes increasingly less valid for a given heat bath field, as its mass becomes increasingly larger than the temperature, because its decay width $\Gamma_{\chi,\psi}$ decreases exponentially $\propto \exp(-m_{\chi,\psi}/T)$. Thus, above some $m_{\chi,\psi} > T$ the thermal-adiabatic condition no longer will be satisfied. At this point, the existing abundance of the given heat bath particles then will freeze-out. After freeze-out, the mass of these heat bath particles continues to grow, due to their $\varphi_0$ dependence, but their total number is fixed. Since prior to freeze-out the thermal distribution of the heat bath fields also falls exponentially $\propto \exp(-m_{\chi,\psi}/T)$, at time of freeze-out, their abundance could be large if $m_{\chi,\psi} \sim T$ or small if $m_{\chi,\psi} \stackrel{>}{\sim} 10T$. As such, the magnitude of the freeze-out abundance depends on how quickly the thermal adiabatic condition becomes invalid, and our solutions cover a wide range of possibilities. The main point to observe here is that this dynamics generically produces very massive particles, in the range $ \sim 10^{16} - 10^5 {\rm GeV}$ once $\varphi_0$ equilibriates. Such heavy particle production at the early stages of warm inflation is not important in the post-inflation universe, since they will completely dilute during the inflationary period. However, this heavy particle productions from the last few mass sites, just before warm inflation ends, will leave significant abundances in the post-inflation universe. Thus, without further modifications to this model, particle production yields weakly interacting massive particles (WIMPs), which are a category of particles that play an important role in present-day notions about dark matter \cite{wimps}. In fact, exceptionally high mass particles $\sim 10^{12} - 10^{16} {\rm GeV}$, which corresponds to the upper mass range for the heat bath fields, sometimes are further distinguished in the literature as ``Wimpzillas'' \cite{wimpzilla}. In general, thermal production of very massive WIMPs is understood to lead to overabundance problems, which therefore restrict the upper bound on their mass to be $\stackrel{<}{\sim} 10^6 {\rm GeV}$ \cite{griest}. This implies thermal Wimpzilla production is prohibited for any generic situation, including the model in this paper. Thus most of the mass range for the heat bath fields in our model is in danger of an overabundance problem. In fact, as mentioned above, for the model in this paper the problem is worse since upon production, the mass of the massive heat bath particles can grow by as much as several orders of magnitude due to the $\varphi_0$-dependence in Eqs. (\ref{rmchi}), (\ref{rmpsi}). There are simple remedies to this overabundance problem, which provide a rich number of phenomenological possibilities in which the abundance of WIMPs and Wimpzillas can be controlled. The simplest solution is to couple the heat bath fields to secondary fields, such as the example of $\sigma$ fields in Subsect. VC. This provides decay channels for the $\chi,\psi$ heat bath fields into less massive or even light fields, thus providing a means to control the relative fractions of WIMPs, Wimpzillas, light particles etc... Furthermore, as discussed in Subsect. VC, this mechanism can be implemented with minimal effect on the dissipative dynamics. Another possibility, which can be applied separately or in conjunction with this, is to have a second and short stage of inflation after the initial warm inflation stage, which dilutes the WIMP and Wimpzilla abundances. Finally, the lower end of the mass range of the heat bath fields found from our calculations falls below the upper acceptability limit for thermal WIMP production. The present model has a narrow mass window that falls within this acceptability limit. However, with the improvements discussed earlier in this section, to the dissipative dynamics, this mass range could be substantially enlarged. The calculations in this paper were guided by observation and consistency with quantum field theory. ``Nice'' particle physics was not an {\it a priori} requirement. On the one hand, this is a constructive approach that attempts to find a toy model to study a very complicated dynamics. On the other hand, this is a predictive approach. Quantum field theory is postulated to hold at the scales of inflationary dynamics. Combining this theoretical tool with observation, a mutually consistent model has been deduced. We can conclude that this model is a good constructive tool for studying warm inflation dynamics. No complete conclusion is possible as yet about the predictive content of this model. The interesting connection between this model and strings found in \cite{bk,bk2} implies a revised meaning is required of ``nice'' particle physics. At the inflationary scale, nice particle physics may not be synonymous with simple particle physics. This model is evidence that inflationary dynamics can be a multi- or even ultramulti- field problem. \section*{Acknowledgments} I thank R. Brandenberger, M. Gleiser, A. Guth, S. Jeon, T. Kephart, A. Linde, R. Ramos, M. Sher and J. Yokoyama for helpful discussions. This work was supported by the U. S. Department of Energy under grant DE-FG05-85ER40226.
\section{Introduction} Interest in direct photon production at RHIC\cite{phenix,star} is motivated by the long mean free path of photons. Hard photon production can be calculated in the pQCD-improved parton model. Experiments need to separate the direct photons from the $\pi^0$ (and $\omega^0$) two-photon background. It is estimated that the direct photon yield has to be at least 10\% of the $\pi^0$ yield to be measurable. We have embarked on the following project: (i) fix the pQCD description of $\gamma$ and $\pi^0$ production in $pp$ collisions; (ii) address nuclear effects in $pA$ collisions; (iii) calculate direct $\gamma$ and $\pi^0$ production in $AA$ collisions at $\sqrt{s}=200$ GeV. Our first results, which pertain to (i) and (ii), are summarized in Ref. 3. Parton cross sections are calculable in pQCD at high energy to leading order (LO) or next-to-leading-order (NLO).\cite{Field,NLOAur} The parton distribution functions (PDFs) and fragmentation functions (FFs), however, need to be fitted to data. In recent NLO calculations the various scales are optimized.~\cite{OptimQres} Alternatively, an additional non-perturbative parameter, the width of the {\it intrinsic transverse momentum} ($k_T$) distribution of the partons is introduced.~\cite{Huston95} We choose the latter method and use $k_T$ phenomenologically in $pp$ collisions, expecting that its importance will decrease in higher orders of pQCD. \nopagebreak \section{Proton-proton collisions}\label{sec:pp} In the lowest-order pQCD-improved parton model, direct pion production can be described in $pp$ collisions by \be \label{fullpi} E_{\pi}\f{d\sigma_\pi^{pp}}{d^3p} = \sum_{abcd}\! \int\!dx_{1,2}\ f_{1}(x_1,Q^2)\ f_{2}(x_2,Q^2)\ K \f{d\sigma}{d\hat{t}}(ab\to c d)\, \frac{D(z_c,{\widehat Q}^2)}{\pi z_c} \ \ \ , \ee where $f_{1}(x,Q^2)$ and $f_{2}(x,Q^2)$ are the PDFs of partons $a$ and $b$, and $\sigma$ is the LO cross section of the appropriate partonic subprocess. The K-factor accounts for higher order corrections.~\cite{Owens87} Comparing LO and NLO calculations a nearly constant value, $K\approx 2$, is obtained as a good approximation of the higher order contributions in the $p_T$ region of interest.~\cite{Wong98} In eq.(\ref{fullpi}) $D(z_c,{\widehat Q}^2)$ is the FF of the pion, with ${\widehat Q} = p_T /z_c$, where $z_c$ is the momentum fraction of the final hadron. We use NLO parameterizations of the PDFs\cite{MRST98} and FFs~\cite{BKK} with fixed scales. Direct $\gamma$ production is described similarly.~\cite{Field} \begin{figure}[b] \centerline{\epsfysize=2in\epsfbox{kt.eps}} \caption{Best fits for the widths of the intrinsic transverse momentum distributions in $\gamma$ and $\pi^0$ production data in $pp$ collisions as a function of cm energy. See text for the lines. \label{fig:kTav}} \end{figure} To incorporate the intrinsic $k_T$, each integral is extended to $k_T$-space, $dx \ f(x,Q^2) \rightarrow dx \ d^2\!k_T\ g({\vec k}_T) \ f(x,Q^2)$.~\cite{Wong98,Wang9798} We approximate $g({\vec k}_T)$ as $g({\vec k}_T) = \exp(-k_T^2/\langle k_T^2 \rangle)/ {\pi \langle k_T^2 \rangle}$. Here $\langle k_T^2 \rangle$ is the 2-dimensional width of the $k_T$ distribution, related to the average $k_T$ of one parton as $\langle k_T^2 \rangle = 4 \langle k_T \rangle^2 /\pi$. We applied this model to data from $pp\rightarrow \pi^0 X$ and $pp\rightarrow \gamma X$.\cite{Cronin,pi,gamma} The calculations were corrected for the finite rapidity windows of the data. The Monte-Carlo integrals were carried out by the standard VEGAS routine.~\cite{VEGAS} We fitted the data minimizing $\Delta^2=\sum (Data-Theory)^2/Theory^2$ in the midpoints of the data. Fig. 1. shows the obtained fit values for $\la k_T^2 \ra$ and a calculated $\la k_T \ra$. The error bars display a $\Delta^2= \Delta^2_{min}\pm 0.1$ uncertainty in the fit procedure. The lines guide the eye and indicate a linear increase to a common value of $\la k_T \ra \approx 3.5$ GeV at $\sqrt{s}=1800$ GeV.\cite{Ziel98} These values of $\la k_T \ra$ provide a satisfactory description of the data up to energies of $\sqrt{s} \approx 65$ GeV. The data/theory ratios cluster around unity, with no systematic trend in the deviations (different experiments show different slopes).\cite{plf99,plf99;a} The maximum deviations are on the order of $\approx 50\%$. To estimate the $p_T$ value where the direct photon production cross section surpasses 10\% of the pion cross section in a $pp$ collision at $\sqrt{s}=200$ GeV, we use the $\sqrt{s}$ dependence of $\langle k_T \rangle$ from Fig.~1, and obtain the results displayed in Fig.~2. The threshold is reached in this approximation at $p_T \approx 6.5$ GeV. \begin{figure}[h] \centerline{\epsfxsize=2.5in\epsfbox{rhic.eps}} \caption{Extrapolated $\gamma$ (full line) and $\pi^0 (\times 0.1)$ (dashed) production cross sections as functions of transverse momentum in $pp$ collisions at $\sqrt{s}=200$ GeV. \label{fig:pp200}} \end{figure} \vspace*{-4mm} \section{Proton-nucleus collisions}\label{sec:pA} Having isolated the $\la k_T^2 \ra$ already present in $\pi^0$ and $\gamma$ production in $pp$ collisions at the present level of calculation, we turn to the nuclear enhancement observed\cite{Cronin} in $pA$ collisions. In minimum-biased $pA$ collisions the pQCD description of the inclusive pion cross section is based on \begin{eqnarray} \label{fullpipA} && E_{\pi}\f{d\sigma_\pi^{pA}}{d^3p} =\! \sum_{abcd}\! \int\!\! d^2\!b\, t_A(b)\! \int\!\!dx_{1,2} d^2k_{T_{1,2}}\ g_1(\vec{k}_{T_1},b) \nonumber \\ && \ \ g_2(\vec{k}_{T_2}) f_{1}(x_1,Q^2) f_{2}(x_2,Q^2)\ K \f{d\sigma}{d\hat{t}} \frac{D(z_c,{\widehat Q}^2)}{\pi z_c} \,. \end{eqnarray} Here $b$ is the impact parameter and $t_A(b)$ is the nuclear thickness function normalized as $\int d^2 b \ t_A(b) = A$. For simplicity, we use a sharp sphere nucleus with $t_A(b) = 2 \rho_0 \sqrt{R_A^2-b^2} $, where $R_A=1.14 A^{1/3}$ and $\rho_0=0.16$ fm$^{-3}$. \begin{figure}[b] \centerline{\epsfxsize=2.3in\epsffile{cronf3u.eps} \hfill \epsfxsize=2.3in\epsffile{cronf4u.eps}} \caption[]{{\it Left}: Cross section per nucleon in $pA \rightarrow \pi^0 X$ reactions ($A=Be,Ti,W$). We show $C^{sat}=1.2$ GeV$^2$ (full lines) and $C^{sat}=0$ (dashed lines). Lower panel: data/theory on a linear scale for the $pW$ collision. {\it Right}: Cross section per nucleon in the $pBe \rightarrow \gamma X$ reaction at two energies.\cite{E706} Solid lines indicate $C^{sat}=1.2$ GeV$^2$, dashed lines mean $C^{sat}=0$. Lower panel shows data/theory for $\sqrt{s}=31.6$ GeV.% } \label{fig:pApi} \end{figure} The standard physical explanation of the nuclear enhancement (Cronin effect) is that the proton traveling through the nucleus gains extra transverse momentum due to random soft collisions and the partons enter the final hard process with this extra $k_T$.\cite{Wang9798} We write the width of the transverse momentum distribution of the partons in the incoming proton as \begin{equation} \label{ktbroadpA} \la k_T^2 \ra_{pA} = \la k_T^2 \ra_{pp} + C \cdot h_{pA}(b) \ . \end{equation} Here $h_{pA}(b)$ is the number of {\it effective} nucleon-nucleon collisions at impact parameter $b$ imparting an average transverse momentum squared $C$. Naively all possible soft interactions are included, but such a picture leads to a target-dependent $C$.\cite{plf99} A more satisfactory description is obtained with a ``saturated'' $h_{pA}$, where it takes at most one semi-hard ($Q^2 \sim 1$ GeV$^2$) collision for the incoming proton to loose coherence, resulting in an increase of the width of its $k_T$ distribution. This is approximated by a smoothed step function with a maximum value of unity. The saturated Cronin factor is denoted by $ C^{sat}$. As displayed in the left of Fig. 3, $C^{sat}=1.2$ GeV$^2$ gives a good fit of $p A \rightarrow \pi^0 X$ data with $A=Be$, $Ti$ and $W$.\cite{Cronin} The right panel of Fig. 3 illustrates how $C^{sat}=1.2$ GeV$^2$ describes $\gamma$ production at $\sqrt{s} \approx 30-40$ GeV. We interpret $C^{sat}$ as the square of the typical transverse momentum imparted in {\it one} semi-hard collision prior to the hard scattering. This picture and the energy dependence of $C^{sat}$ need to be tested as functions of $A$ at different energies, and in particular at $\sqrt{s}=200$ GeV, for RHIC $AA$ predictions. \vspace*{-3mm} \section*{Acknowledgments} \vspace*{-1mm} Supported by DOE, DE-FG02-86ER40251, and by OTKA, F019689. \vspace*{-2mm}
\section{Introduction} \label{S:intro} This paper extend the arguments in the author's previous work on link homotopy \cite{me1} to link concordance. Many of the arguments are nearly identical; the main differences lie in the base cases to the main theorem, in Section~\ref{S:size}. We will begin with a brief overview of finite type invariants. In 1990, V.A. Vassiliev introduced the idea of {\it Vassiliev} or {\it finite type} knot invariants, by looking at certain groups associated with the cohomology of the space of knots. Shortly thereafter, Birman and Lin~\cite{bl} gave a combinatorial description of finite type invariants. We will give a summary of this combinatorial theory. For more details, see Bar-Natan~\cite{bn1}. \subsection{Singular Knots and Chord Diagrams} \label{SS:chord} We first note that we can extend any knot invariant to an invariant of {\it singular} knots, where a singular knot is an immersion of $S^1$ in 3-space which is an embedding except for a finite number of isolated double points. Given a knot invariant $v$, we extend it via the relation: $$\psfig{file=extend.ps}$$ An invariant $v$ of singular knots is then said to be of {\it finite type}, if there is an integer $d$ such that $v$ is zero on any knot with more than $d$ double points. $v$ is then said to be of {\it type} $d$. We denote by $V_d$ the space generated by finite type invariants of type $d$. We can completely understand the space of finite type invariants by understanding all of the vector spaces $V_d/V_{d-1}$. An element of this vector space is completely determined by its behavior on knots with exactly $d$ singular points. Since such an element is zero on knots with more than $d$ singular points, any other (non-singular) crossing of the knot can be changed without affecting the value of the invariant. This means that elements of $V_d/V_{d-1}$ can be viewed as functionals on the space of {\it chord diagrams}: \begin{defn} A {\bf chord diagram of degree d} is an oriented circle, together with $d$ chords of the circles, such that all of the $2d$ endpoints of the chords are distinct. The circle represents a knot, the endpoints of a chord represent 2 points identified by the immersion of this knot into 3-space. \end{defn} Functionals on the space of chord diagrams which are derived from knot invariants will satisfy certain relations. This leads us to the definition of a {\it weight system}: \begin{defn} A {\bf weight system of degree d} is a function $W$ on the space of chord diagrams of degree $d$ (with values in an associative commutative ring ${\bf K}$ with unity) which satisfies 2 relations: \begin{itemize} \item (1-term relation) $$\psfig{file=1-term.ps}$$ \item (4-term relation) $$\psfig{file=4-term.ps}$$ Outside of the solid arcs on the circle, the diagrams can be anything, as long as it is the same for all four diagrams. \end{itemize} We let $W_d$ denote the space of weight systems of degree d. \end{defn} Bar-Natan~\cite{bn1} defines maps $w_d: V_d \rightarrow W_d$ and $v_d: W_d \rightarrow V_d$. $w_d$ is defined by {\it embedding} a chord diagram $D$ in ${\bf R}^3$ as a singular knot $K_D$, with the chords corresponding to singularities of the embedding (so there are $d$ singularities). Any two such embeddings will differ by crossing changes, but these changes will not effect the value of a type $d$ Vassiliev invariant on the singular knot. Then, for any type $d$ invariant $\gamma$, we define $w_d(\gamma)(D) = \gamma(K_D)$. Bar-Natan shows that this is, in fact, a weight system. The 1-term relation is satisfied because of the first Reidemeister move, and the 4-term relation is essentially the result of rotating a third strand a full turn around a double point. $v_d$ is much more complicated to define, using the Kontsevich integral. For a full treatment of the Kontsevich integral, see Bar-Natan~\cite{bn1} and Le and Murakami~\cite{lm}. Using a Morse function, any knot (or link or string link) can be decomposed into elementary {\it tangles}: $$\psfig{file=tangles.ps}$$ Le and Murakami define a map $Z$ from an elementary tangle with $k$ strands to the space of chord diagrams on $k$ strands. This map respects composition of tangles: if $T_1\cdot T_2$ is the tangle obtained by placing $T_1$ on top of $T_2$, then $Z(T_1\cdot T_2) = Z(T_1)Z(T_2)$. Le and Murakami prove that this map gives an isotopy invariant of knots and links. Given a degree $d$ weight system $W$, and a knot $K$, we now define $v_d(W)(K) = W(Z(K))$. Bar Natan shows that $w_d$ and $v_d$ are ``almost'' inverses. More precisely, $w_d(v_d(W)) = W$ and $v_d(w_d(\gamma))-\gamma$ is a knot invariant of type $d-1$. As a result, (see \cite{bl,bn1,st}) the space $W_d$ of weight systems of degree $d$ is isomorphic to $V_d/V_{d-1}$. For convenience, we will usually take the dual approach, and simply study the space of chord diagrams of degree $d$ modulo the 1-term and 4-term relations. The dimensions of these spaces have been computed for $d \leq 12$ (see Bar-Natan~\cite{bn1} and Kneissler~\cite{kn}). It is useful to combine all of these spaces into a graded module via direct sum. We can give this module a Hopf algebra structure by defining an appropriate product and co-product: \begin{itemize} \item We define the product $D_1 \cdot D_2$ of two chord diagrams $D_1$ and $D_2$ as their connect sum. This is well-defined modulo the 4-term relation (see \cite{bn1}). $$\psfig{file=product.ps}$$ \item We define the co-product $\Delta(D)$ of a chord diagram $D$ as follows: $$ \Delta(D) = {\sum_J D_J' \otimes D_J''} $$ where $J$ is a subset of the set of chords of $D$, $D_J'$ is $D$ with all the chords in $J$ removed, and $D_J''$ is $D$ with all the chords not in J removed. $$\psfig{file=coproduct.ps}$$ \end{itemize} It is easy to check the compatibility condition $\Delta(D_1\cdot D_2) = \Delta (D_1)\cdot\Delta(D_2)$. \subsection{Unitrivalent Diagrams} \label{SS:ud} It is often useful to consider the Hopf algebra of bounded unitrivalent diagrams, rather than chord diagrams. These diagrams, introduced by Bar-Natan~\cite{bn1} (Bar-Natan calls them {\it Chinese Character Diagrams}), can be thought of as a shorthand for writing certain linear combinations of chord diagrams. We define a {\it bounded unitrivalent graph} to be a unitrivalent graph, with oriented vertices, together with a bounding circle to which all the univalent vertices are attached. We also require that each component of the graph have at least one univalent vertex (so every component is connected to the boundary circle). We define the space $A$ of bounded unitrivalent diagrams as the quotient of the space of all bounded unitrivalent graphs by the $STU$ relation, shown in Figure~\ref{F:stu}. \begin{figure} $$\psfig{file=STU.ps}$$ \caption{STU relation} \label{F:stu} \end{figure} As consequences of $STU$ relation, the anti-symmetry ($AS$) and $IHX$ relations, see Figure~\ref{F:ihx}, also hold in $A$. \begin{figure} $$\psfig{file=IHX.ps}$$ \caption{AS and IHX relations} \label{F:ihx} \end{figure} Bar-Natan shows that $A$ is isomorphic to the algebra of chord diagrams. We can get an algebra $B$ of {\it unitrivalent diagrams} by simply removing the bounding circle from the diagrams in $A$, leaving graphs with trivalent and univalent vertices, modulo the $AS$ and $IHX$ relations. Bar-Natan shows that the spaces $A$ and $B$ are isomorphic. The map $\chi$ from $B$ to $A$ takes a diagram to the linear combination of all ways of attaching the univalent vertices to a bounding circle, divided by total number of such ways (T. Le noticed that this factor, missing in \cite{bn1}, is necessary to preserve the comultiplicative structure of the algebras). The inverse map $\sigma$ turns a diagram into a linear combination of diagrams by performing sequences of ``basic operations,'' and then removes the bounding circle. The two basic operations are: $$\psfig{file=basic_op.ps}$$ \section{String Links, Links and Concordance} \label{S:string} \subsection{String Links} \label{SS:string_links} Bar-Natan~\cite{bn2} extends the theory of finite type invariants to string links. \begin{defn} (see \cite{hl}) Let D be the unit disk in the plane and let I = [0,1] be the unit interval. Choose k points $p_1,..., p_k$ in the interior of D, aligned in order along the the x-axis. A {\bf string link} $\sigma$ of k components is a smooth proper imbedding of k disjoint copies of I into $D \times I$: $$\sigma:\ \bigsqcup_{i=1}^k{I_i} \rightarrow D \times I$$ such that $\sigma|_{I_i}(0) = p_i \times 0$ and $\sigma|_{I_i}(1) = p_i \times 1$. The image of $I_i$ is called the ith string of the string link $\sigma$. \end{defn} Essentially, everything works the same way for string links as for knots. The bounding circle of the bounded unitrivalent diagrams now becomes a set of bounding line segments, each labeled with a color, to give an algebra $A^{sl}$ (the multiplication is given by placing one diagram on top of another). The univalent diagrams are unchanged, except that each univalent vertex is also labeled with a color to give the space $B^{sl}$. The isomorphisms $\chi$ and $\sigma$ between $A$ and $B$ easily extend to isomorphisms $\chi^{sl}$ and $\sigma^{sl}$ between $A^{sl}$ and $B^{sl}$, just working with each color separately. In addition, there are obvious maps $w_d^{sl}$ and $v_d^{sl}$ analogous to $w_d$ and $v_d$ (we just need to keep track of colors). \subsection{Links} \label{SS:links} The obvious definition of chord diagrams for links is simply to replace the bounding line segments with bounding circles. However, these diagrams are difficult to work with, and it is in particular unclear how to define the unitrivalent diagrams. Unlike for a knot, closing up the components of a string link of several components is not a trivial operation, so we need to place some relations on the space of unitrivalent diagrams. Since we understand the spaces of chord diagrams and unitrivalent diagrams for string links, it would be useful to be able to express these spaces for links as quotients of the spaces for string links. The question is then, what relations do we need? One relation is fairly obvious. When we construct the space $A^l$ of bounded unitrivalent diagrams for links, we replace the bounding line segments of $A^{sl}$ with directed circles. Bar-Natan et. al. observed (see Theorem 3, \cite{bgrt}) that this is exactly equivalent to saying that the ``top'' edge incident to one of the line segments can be brought around the circle to be on the ``bottom.'' So we can write $A^l$ as the quotient of $A^{sl}$ by relation (1), shown in Figure~\ref{F:link_rel} (where the figure shows {\it all} the chords with endpoints on the red component). \begin{figure} $$\psfig{file=link_rel.ps}$$ \caption{The link relation for chord diagrams} \label{F:link_rel} \end{figure} Then the Kontsevich integral for links, $Z^l$, is defined by cutting the link to make a string link, applying the Kontsevich integral for string links, and then taking the quotient by relation (1). Now $w_d^l$ and $v_d^l$ are defined similarly to $w_d$ and $v_d$. Given a link invariant $\gamma$ and a diagram $D$ in $A^l$, $w_d^l(\gamma)(D) = \gamma(L_{\hat{D}})$, where $\hat{D}$ is the closure of the diagram $D$ (i.e. the bounding line segments are closed to circles). $L_{\hat{D}}$ is well-defined by Theorem 3 of \cite{bgrt}. Defining $v_d^l$ is even easier, now that we have $Z^l$. Given a weight system (element of the graded dual of $A^l$) $W$ and a link $L$, we define $v_d^l(W)(L) = W(Z^l(L))$. One advantage of this formulation of $A^l$ is that it enables us to define the space $B^l$ of unitrivalent diagrams as a quotient of the already known (see \cite{bn2}) space $B^{sl}$. This was done by Bar-Natan et. al.~\cite{bgrt}. Using the $STU$ relation, we can rewrite relation (1) as in Figure~\ref{F:link_rel2}. \begin{figure} $$\psfig{file=link_rel2.ps}$$ \caption{The link relation for Bounded Unitrivalent Diagrams} \label{F:link_rel2} \end{figure} This suggests how we should define the space $B^l$. We will take the quotient of $B^{sl}$ by the relations (*) shown in Figure~\ref{F:link_rel3}, \begin{figure} $$\psfig{file=link_rel3.ps}$$ \caption{The link relation for unitrivalent diagrams} \label{F:link_rel3} \end{figure} where the univalent vertices shown are {\it all} the univalent vertices of a given color. With these definitions, Bar-Natan et. al. proved that $A^l$ and $B^l$ are isomorphic: \begin{thm} \label{T:isotopy-descends} (Theorem 3, \cite{bgrt}) The isomorphism between $A^{sl}$ and $B^{sl}$ descends to an isomorphism between $A^l$ and $B^l$. \end{thm} \subsection{Link Concordance} \label{SS:concordance} \begin{defn} \label{D:concordance} Consider two k-component links $L_0$ and $L_1$, and two k-component string links $SL_0$ $SL_1$. These can be thought of as embeddings (proper embeddings, in the case of the string links): $$L_i: \bigsqcup_{i=1}^k S^1 \hookrightarrow {\bf R}^3$$ $$SL_i: \bigsqcup_{i=1}^k I \hookrightarrow {\bf R}^2 \times I$$ A {\bf (link) concordance} between $L_0$ and $L_1$ is an embedding: $$H: \left({\bigsqcup_{i=1}^k S^1}\right) \times I \hookrightarrow {\bf R}^3 \times I$$ such that $H(x,0) = (L_0(x),0)$ and $H(x,1) = (L_1(x),1)$. Similarly, a {\bf (string link) concordance} between $SL_0$ and $SL_1$ is an embedding: $$H: \left({\bigsqcup_{i=1}^k I}\right) \times I \hookrightarrow ({\bf R}^2 \times I) \times I$$ such that $H(x,0) = (SL_0(x),0),\ H(x,1) = (SL_1(x),1)$, and $H(i,t) = (i,t)$ for $i=0,1$. A concordance is an isotopy if and only if H is level preserving; i.e. if the image of $H_t$ is a (string) link at level $t$ for each $t \in I$. \end{defn} We want to extend the results of the last section to string links and links considered up to concordance. For string links, this has already been done by Habegger and Masbaum~\cite{hm}. They describe the algebras $A^{csl}$ and $B^{csl}$ of bounded and unbounded unitrivalent diagrams for string links up to concordance (which they denote $A^t$ and $B^t$), and observe that they are isomorphic. In brief, we take the quotient of $A^{sl}$ (resp. $B^{sl}$) by the space of diagrams with non-trivial first homology. In other words, we are left with tree diagrams. It is then straightforward to define $w_d^{csl}$ and $v_d^{csl}$ in the usual way, and show that they are ``almost'' inverses in the same sense that $w_d$ and $v_d$ are. All of this extends to links just as it did for isotopy. We define $A^{cl}$ as the quotient of $A^{csl}$ by relation (1), and $B^{cl}$ as the quotient of $B^{csl}$ by relation (*). We then define $Z^{cl}$, $w_d^{cl}$, and $v_d^{cl}$ just as we did for links up to homotopy. Finally, the arguments of Bar-Natan et. al. carry through to show: \begin{thm} \label{T:concordance-descends} (Theorem 3, \cite{bgrt}) The isomorphism between $A^{csl}$ and $B^{csl}$ descends to an isomorphism between $A^{cl}$ and $B^{cl}$. \end{thm} {\bf Remark:} By results of Habegger and Masbaum (see Theorem 5.5 of \cite{hm}), $Z^{cl}$ is the {\it universal} finite type invariant of link concordance. By this we mean that it dominates all other such invariants. \section{The Size of $B^{cl}$} \label{S:size} Now that we have properly defined the space $B^{cl}$ of unitrivalent diagrams for link homotopy, we want to analyze it more closely. We will consider the case when $B^{cl}$ is a vector space over the reals (or, more generally, a module over a ring of characteristic 0). In particular, we would like to know exactly which diagrams of $B^{csl}$ are in the kernel of the relation (*) (i.e. are 0 modulo (*)). We will find that the answer is ``almost everything'' - to be precise, any unitrivalent diagram with a component of degree 2 or more. We will start by proving a couple of base cases, and then prove the rest of the theorem by induction. Let $B^{csl}(k)$ denote the space of unitrivalent diagrams for string link concordance with $k$ possible colors for the univalent vertices (i.e. we are looking at links with $k$ components). Consider a diagram $D \in B^{csl}(k)$. Recall from the previous sections that each component of $D$ is a tree diagram. {\bf Notation:} Before we continue, we will introduce two bits of notation which will be useful in this section. \begin{itemize} \item Given a unitrivalent diagrams $D$, we define $m(D;i,j)$ to be the number of components of $D$ which are simply line segments with ends colored $i$ and $j$, as shown below: $$i-----j$$ \item Components of a diagram with degree greater than one will be called {\it large} components. Components of degree one will be called {\it small} components. \end{itemize} \subsection{Knots and Two Component Links} \label{SS:1&2comp} We begin by considering the case of knot concordance, when $k = 1$. Ng~\cite{ng} has already shown that the only finite type invariant of knot concordance is the ${\bf Z}_2$-valued Arf invariant, so there are no real-valued finite type invariants of knot concordance. We will begin, as a warm-up, by showing this result using unitrivalent diagrams. Bar-Natan has shown that the spaces of unitrivalent diagrams for knots and string links of one component are isomorphic, so the relation (*) has no effect. \begin{lem} \label{L:knots} $B^{csl}(1) = B^{cl}(1) = 0.$ \end{lem} {\sc Proof:} First, we consider the case when $D \in B^{csl}(1)$ has a large component $C$. Since $C$ is a tree, we can use the $IHX$ relation as in Figure~\ref{F:treepf} to rewrite $C$ as a sum of diagrams: \begin{figure} $$\psfig{file=treepf.ps}$$ \caption{Using IHX to get ``standard'' tree diagram} \label{F:treepf} \end{figure} $$\psfig{file=tree.ps}$$ Clearly, twisting the branch at the right or left end of $T$ gives us the same diagram, but by the $AS$ relation, this flips the sign. Therefore, $T = -T = 0$, so $D = 0$. The case when all the components of D have degree one is somewhat more subtle, though it is essentially an application of the 1-term relation. If D has only small components, then $\chi(D)$ (see Section~\ref{SS:ud}) is a linear combination of bounded unitrivalent diagrams with no interior vertices (i.e. chord diagrams). By repeated applications of the $STU$ relation, we can isolate a chord in each of these diagrams, at the expense of adding a linear combination of diagrams which {\it do} have internal vertices. The diagrams with isolated chords disappear by the 1-term relation, so we are left with a linear combination of bounded unitrivalent diagrams with at least one internal vertex. Now we apply $\sigma$ to this linear combination to get a linear combination of (unbounded) unitrivalent diagrams. Since the basic operations $U$ and $S$ of $\sigma$ can never decrease the number of internal vertices, every unitrivalent diagram in the image of $\sigma$ will have at least one internal vertex; i.e. at least one large component. Hence, by the first case, all of these diagrams are 0. So we have shown that $\sigma(\chi(D)) = 0$. But $\sigma$ and $\chi$ are inverse isomorphisms, so this means $D = 0$, as desired. $\Box$ Next we consider links of two components, i.e. $k=2$. In this case we need to consider the effect of the relation (*). \begin{lem} \label{L:k=2} Let $D \in B^{csl}(2)$. If $D$ has a component $C$ of degree $d \geq 2$, then $D$ is trivial modulo (*). \end{lem} {\sc Proof:} $D$ is a diagram with all endpoints colored 1 or 2. Note that $C$ must have endpoints of both colors, or $D$ will be trivial by the same argument as in Lemma~\ref{L:knots}. In fact, any terminal branch of $C$ must have the form (where $\bar{C}$ denotes the remainder of $C$): $$C:\ \ \begin{matrix} \bar{C} \\ | \\ | \\ 1-----2 \end{matrix}$$ Otherwise, if the two endpoints have the same color, $C$ (and hence $D$) is trivial by the $AS$ relation. The proof is by induction on the number of large components of $D$. In the base case, there is only one such component, $C$. So all the other components $C_i$ are simply line segments labeled $a$ and $b$, where $a,b \in \{1,2\}$, as shown: $$C_i:\ \ a-----b$$ Now we apply the relation (*) to the branch of $C$ shown above using the color 1, as in Figure~\ref{F:arise} (where $\bar{C_i}$ represents the remainder of the component $C_i$). \begin{figure} $$\psfig{file=relation.eps}$$ \caption{Diagrams arising from relation (*)} \label{F:arise} \end{figure} We denote the merger of $C_i$ and $\bar{C}$ as $C'_i$. We do not need to consider other vertices of $C$ colored 1, since these terms result in a diagram with a loop, which are trivial in concordance. Since we are expanding using the color 1, we only need to consider $C_i$ where $a$ or $b$ is 1. If they are both 1, then by the $AS$ relation: $$C'_i:\ \ \begin{matrix} \bar{C} \\ | \\ | \\ 1-----1 \end{matrix} = 0$$ Therefore, we need only consider $C_i$ with one endpoint colored 1 and the other colored 2. But in this case, $C'_i = C$, so $D_i = D$, and we get an equation $D + \sum{D_i} = (1+m)D = 0$ for some $m \geq 0$. Hence $D = 0$ modulo (*). This concludes the base case. For the inductive step, we assume that the theorem is true for diagrams with $n$ large components, and consider a diagram $D$ with $n+1$ large components. Let $C$ be one of these components. As before, C has a branch: $$C:\ \ \begin{matrix} \bar{C} \\ | \\ | \\ 1-----2 \end{matrix}$$ Also, every small component $C_i$ looks like (where $a,b \in \{1,2\}$): $$C_i:\ \ a-----b$$ Once again, we apply the relation (*) to the color 1, using this branch of $C$, and find that $D + \sum{D_i} = 0$. Whenever $C_i$ is small, $D_i = D$ (as in the base case, we need only consider the case when $a = 1$ and $b = 2$). If $C_i$ is large, then $D_i$ has one fewer large component than $D$ does, since the bulk of $C$ has joined with $C_i$, leaving behind a line segment of degree one. So $D_i$ is trivial modulo (*) by the inductive hypothesis. Hence we are left with a sum of copies of $D$, and conclude that $(1+m)D = 0$ for some $m \geq 0$, so $D = 0$ modulo (*). This concludes the induction and the proof. $\Box$ \subsection{Three Component Links} \label{SS:3comp} The case when $k=3$ is our final special case before the proof of the general theorem, and it is significantly more complicated than the previous two lemmas. The main step is to show that no component of $D$ can have two endpoints of the same color (essentially, this reduces the problem to the case of link homotopy, treated in \cite{me1}). Once again, we will be inducting on the number of large components of the diagram. The base case contains most of the work of the proof, so we present it as a separate lemma. \begin{lem} \label{L:k=3_base} If k=3, D has exactly one large component C, and C has two endpoints of the same color, then D is trivial modulo (*). \end{lem} {\sc Proof:} Without loss of generality, we will say that $C$ has two endpoints colored 1. If these endpoints are on the same final branch, as shown below, then $D$ will be trivial by the $AS$ relation (since we will have $D = -D$). $\bar{C}$ denotes the remainder of C: $$C:\ \ \begin{matrix} \bar{C} \\ | \\ | \\ 1-----1 \end{matrix}$$ Otherwise, we can use the $IHX$ relation to move one of the endpoints colored 1 out to the ends of the component $C$, as shown in Figure~\ref{F:expand} (where we move the endpoint colored $k$). \begin{figure} $$\psfig{file=IHXexpand.ps}$$ \caption{Using the IHX relation to decompose a diagram} \label{F:expand} \end{figure} So it suffices, without loss of generality, to consider $C$ with a branch as shown below: $$C:\ \ \begin{matrix} \bar{C} \\ | \\ | \\ 1-----2 \end{matrix}$$ Our proof will be by induction on $m(D;2,3)$, inducting among diagrams with exactly one large component, which have a branch with colors 1 and 2 as shown above. For the base case, $m(D;2,3) = 0$. This means that there are no small components of $D$ colored 2 and 3. So the only degree 1 components with an endpoint colored 2 have their other endpoint colored 1 or 2: $$C_i:\ \ a-----2,\ a\in \{1,2\}$$ Now we apply the relation (*) to $D$ as in Figure~\ref{F:arise}, only now we are applying it using the color 2. So we only need to consider the components of $D$ with endpoints colored 2. Because of the loop relation for concordance, we can ignore other endpoints of $C$ which might be colored 2, and just consider the remaining components. These are all small components with endpoints colored $a$ and 2 as described above. As in Lemma~\ref{L:k=2}, the case when $a=2$ can be ignored by the $AS$ relation, so we are reduced to the case of small components with endpoints colored 1 and 2. Therefore, we obtain the equation $D + \sum{D_i} = 0$, where every $D_i = D$. So $D + m(D;1,2)D = (1+m(D;1,2))D = 0$. Since $m(D;1,2) \geq 0$, we conclude that $D = 0$ modulo (*). This concludes the base case of the induction. For the inductive step, we will have small components with endpoints colored 2 and 3. Our goal is to reduce the number of such components. Now, when we apply the relation (*), we can again ignore small components with both endpoints colored 2, by the $AS$ relation. We get an equation $D + m(D;1,2)D + m(D;2,3)D' = 0$, where $D'$ is identical to $D$ except that a small component with endpoints colored 2 and 3 has been replaced by one with endpoints colored 1 and 2, and the endpoint of $C$ colored 1 above has been colored 3. We denote this analogue of $C$ in $D'$ by $C'$. Diagramatically, we can represent $D'$ by showing the changes that have been made to $D$: $$D':\ \ \begin{matrix} \bar{C} \\ | \\ | \\ 3-----2 \end{matrix} \ (2,3) \rightarrow (1,2)$$ Note that $m(D';2,3) = m(D;2,3) - 1$. As we did in Lemma~\ref{L:knots}, we can apply the $IHX$ relation to $C'$, fixing the branch shown above. As a result, it suffices to consider the case when $C'$ has the form shown below (where $n$ is the degree of $C'$): $$\psfig{file=Cprime.ps}$$ We can assume that one of the endpoints $\alpha_3,...,\alpha_n$ is colored 1. If not, then $\alpha_{n+1} = 1$ (since we know there is a second endpoint of $C$ colored 1), and we can switch $\alpha_n$ and $\alpha_{n+1}$ using the $AS$ relation (at the cost of reversing the sign of $D'$). We can apply the $IHX$ relation as in Figure~\ref{F:treepf} to move the endpoint colored 2 along the ``spine'' of $C'$ (i.e. the path from the endpoint colored 3 to the endpoint colored $\alpha_{n+1}$) and obtain the decomposition $D' = \sum{D^i}$, where we transform $C$ into $C^i$ as shown below: $$\psfig{file=Ci.ps}$$ Notice that the spine of $C^i$ is shorter than the spine of $C$ by one edge; and that one of the branches along the spine has grown correspondingly. We observe that if $\alpha_i = 1$, then $D^i = 0$ by the inductive hypothesis, since it now has a branch colored 1 and 2, and $m(D^i;2,3) = m(D';2,3) = m(D;2,3) - 1$. So the endpoint colored 1 will not be incorporated into the larger branch, and has moved one position closer to the endpoint colored 3 at the far left. In the case when $\alpha_3 = 1$, $D^3 = 0$, and $C^i$ has the form below for $i > 3$: $$\psfig{file=Ci2.ps}$$ Now if we use the relation (*) to expand $D^i$ along the color 1, we find that $D^i + m(D^i;1,3)D^i + m(D^i;1,2)D^i_2 = 0$, where $D^i_2$ is the result of replacing a small component colored 1 and 2 with a small component colored 1 and 3, and changing the endpoint of $C^i$ colored 3 to one colored 2. Then $D^i_2$ has a large component with a branch colored 1 and 2, and $m(D^i_2;2,3) = m(D^i;2,3) = m(D;2,3)-1$, so $D^i_2$ is trivial modulo (*) by the inductive hypothesis. Therefore, we find that $(1+m(D^i;1,3))D^i = 0$, and hence $D^i = 0$ modulo (*). If $\alpha_3 \neq 1$, the first branch on the spine of $C^i$ (adjacent to the endpoint colored 3) will only have endpoints colored 2 and 3 (not 1). Now we repeat the process by expanding the (non-trivial) $C^i$'s, using the first branch on the spine of $C^i$. In order to continue the process, we need to show the following fact: {\sc Claim:} If $K$ is a unitrivalent diagram with all endpoints colored 2 or 3, then the following diagram is trivial modulo (*): $$\psfig{file=K.ps}$$ Here $K$ is assumed to be a subdiagram of the only component of degree greater than 1 in a diagram $E$ such that $m(E;2,3) = m(D;2,3)-1$. {\sc Proof of Claim:} Using the $IHX$ relation as in Figure~\ref{F:expand}, we can decompose this diagram into a linear combination of diagrams where the endpoint colored 1 has migrated out to one of the ends of $K$, leaving a diagram with a branch with endpoints colored 1 and 2 or 1 and 3. In the first case, the diagram is trivial by the inductive hypothesis. In the second case, the diagram is trivial by the argument used above for $\alpha_3 = 1$. So we conclude that $E$ is trivial modulo (*). $\Box$ (Claim) Using the claim, we can continue the process, moving the endpoint colored 1 to the left at each stage, until we are left with diagrams where the first branch, adjacent to the endpoint colored 3, consists of a single endpoint colored 1. These diagrams are trivial by the argument above (when $\alpha_3 = 1$). We can conclude that all of the $D^i$'s are trivial modulo (*). Therefore, $D'$ is trivial modulo (*), and we obtain the equation $D + m(D;1,2)D = (1+m(D;1,2))D = 0$. We conclude that $D = 0$ modulo (*), which finishes the induction and the proof. $\Box$ \begin{lem} \label{L:k=3_induct} If k=3 and D has a component C with two endpoints of the same color, then D is trivial modulo (*). \end{lem} {\sc Proof:} First, we assume that $C$ is large. We will induct on the number of large components of $D$. The base case of the induction has already been proved, in Lemma~\ref{L:k=3_base}. The general case follows exactly the same argument. The only modification is to notice that, whenever (*) is applied, the diagrams $D_i$ (as in Figure~\ref{F:arise}) which arise from large components $C_i$ will have fewer large components than D (since $C$ and $C_i$ have been joined, leaving behind a component of degree one). So by the inductive hypothesis, these diagrams can be ignored at every stage. Therefore, exactly the same proof shows that $D$ is trivial modulo (*). If $C$ is small, with both endpoints the same color, we can use the previous case together with the argument from Lemma~\ref{L:knots} to show $D$ is trivial. $\Box$ \begin{lem} \label{L:k=3} If k=3 and D has a large component C, then D is trivial modulo (*). \end{lem} {\sc Proof:} By Lemma~\ref{L:k=3_induct}, there is (up to sign) only one possible diagram for $C$: $$C = \ \begin{matrix} 3 \\ | \\ | \\ 1-----2 \end{matrix}$$ Now we apply the relation (*) to $D$ using C and the color 1. By Lemma~\ref{L:k=3_induct}, we need only consider components with endpoints colored 1 or 2, and no component can have two endpoints of the same color. Therefore, we need only consider components $C_i$ as shown: $$C_i = \ 1-----2$$ This gives us the equation $D + m(D;1,2)D = (1+m(D;1,2))D = 0$. We conclude that $D = 0$ modulo (*), which completes the proof for the case $k = 3$. $\Box$ \subsection{The General Case} \label{SS:general} We are now ready to begin our proof of the general case. As we did for the case when $k = 3$, we will induct on the number of large components of $D$. Once again, for clarity, we will prove the base case (which contains most of the work) as a separate lemma. \begin{lem} \label{L:base} If D has exactly one large component C, then D is trivial modulo (*). \end{lem} {\sc Proof:} The method of proof for this lemma is very similar to the earlier lemmas. We will successively apply (*) (and do a single expansion via IHX) until we obtain a set of diagrams which are all either trivial or repetitions of earlier diagrams. We can then backtrack to show that everything disappears. However, we will need to apply (*) four times. This unfortunately makes keeping track of the diagrams somewhat confusing - we have done our best. Without loss of generality, as before, we can assume that $C$ has a branch as shown: $$C:\ \ \begin{matrix} \bar{C} \\ | \\ | \\ 1-----2 \end{matrix}$$ We apply (*) using the color 1 and find that $D + m(D;1,2)D + \sum_{a\neq 1,2}{m(D;1,a)D_a} = 0$, where $D_a$ is the same as $D$ except that: \begin{itemize} \item $C$ has been replaced by a component $C_a$ identical to it except that the endpoint colored 2 is now colored $a$ (so $\bar{C_a} = \bar{C}$) \item A line segment with endpoints colored 1 and $a$ has been replaced by a line segment with endpoints colored 1 and 2. In other words, $m(D_a;1,a) = m(D;1,a)-1$ and $m(D_a;1,2) = m(D;1,2)+1$. \end{itemize} We will denote this as shown below: $$D_a:\ \ \begin{matrix} \bar{C} \\ | \\ | \\ 1-----a \end{matrix} \ (1,a) \rightarrow (1,2)$$ As in Lemma~\ref{L:k=3_base}, we use the IHX relation to decompose $D_a = \sum_{\alpha_i \neq a}{\pm D_a^i}$ (summing over the endpoints of $C_a$, with colors $\alpha_i$), where the analogue $C_a^i$ of $C_a$ in $D_a^i$ has a branch as shown, and the other components of the diagram are the same as $D_a$: $$D_a^i:\ \ \begin{matrix} \bar{C_a^i} \\ | \\ | \\ i-----a \end{matrix} \ (1,a) \rightarrow (1,2)$$ By abuse of notation, we write the color $\alpha_i$ as simply $i$. Since we will never have to compare different $D_a^ii$'s, this will not cause any confusion. Note that, aside from having endpoints of the same colors, $C_a^i$ looks nothing like $C_a$. Now we apply (*) to $D_a^i$, using the color $i$. In the pictures we use to describe the various diagrams that we produce in what follows, we will just be showing how the diagrams differ from $D_a^i$. This will involve showing how $C_a^i$ has been altered, and which line segments have been added or removed. At each stage, we will eliminate loop diagrams without comment. We obtain the relation: $$D_a^i + m(D_a^i;i,a)D_a^i + m(D_a^i;2,i)D_{a2}^i + \sum_{b\neq i,2,a}{m(D_a^i;i,b)D_{ab}^i} = 0$$ where: $$D_{a2}^i:\ \ \begin{matrix} \bar{C_a^i} \\ | \\ | \\ i-----2 \end{matrix} \ (2,i) \rightarrow (i,a)$$ $$D_{ab}^i:\ \ \begin{matrix} \bar{C_a^i} \\ | \\ | \\ i-----b \end{matrix} \ (i,b) \rightarrow (i,a)$$ Next we apply (*) to $D_{ab}^i$, using the color $b$, and find that: $$D_{ab}^i + m(D_{ab}^i;i,b)D_{ab}^i + m(D_{ab}^i;2,b)D_{ab2}^i + \sum_{c\neq b,i,2}{m(D_{ab}^i;b,c)D_{abc}^i} = 0$$ where: $$D_{ab2}^i:\ \ \begin{matrix} \bar{C_a^i} \\ | \\ | \\ 2-----b \end{matrix} \ \begin{matrix} (i,b) \rightarrow (i,a) \\ (2,b) \rightarrow (i,b) \end{matrix} \Rightarrow (2,b) \rightarrow (i,a)$$ $$D_{abc}^i:\ \ \begin{matrix} \bar{C_a^i} \\ | \\ | \\ c-----b \end{matrix} \ \begin{matrix} (i,b) \rightarrow (i,a) \\ (c,b) \rightarrow (i,b) \end{matrix} \Rightarrow (c,b) \rightarrow (i,a)$$ Now we apply (*) to $D_{ab2}^i$ using the color 2, and to $D_{abc}^i$, using the color $c$. We get two relations: $$D_{ab2}^i + m(D_{ab2}^i;2,b)D_{ab2}^i + m(D_{ab2}^i;2,i)D_{ab2i}^i + \sum_{c\neq b,i,2}{m(D_{ab2}^i;2,c)D_{ab2c}^i} = 0$$ $$D_{abc}^i + m(D_{abc}^i;b,c)D_{abc}^i + m(D_{abc}^i;2,c)D_{abc2}^i + m(D_{abc}^i;i,c)D_{abci}^i +$$ $$\sum_{d\neq b,c,i,2}{m(D_{abc}^i;c,d)D_{abcd}^i} = 0$$ where: $$D_{ab2i}^i:\ \ \begin{matrix} \bar{C_a^i} \\ | \\ | \\ 2-----i \end{matrix} \ \begin{matrix} (2,b) \rightarrow (i,a) \\ (2,i) \rightarrow (2,b) \end{matrix} \Rightarrow (2,i) \rightarrow (i,a)$$ $$D_{ab2c}^i:\ \ \begin{matrix} \bar{C_a^i} \\ | \\ | \\ 2-----c \end{matrix} \ \begin{matrix} (2,b) \rightarrow (i,a) \\ (2,c) \rightarrow (2,b) \end{matrix} \Rightarrow (2,c) \rightarrow (i,a)$$ $$D_{abc2}^i:\ \ \begin{matrix} \bar{C_a^i} \\ | \\ | \\ c-----2 \end{matrix} \ \begin{matrix} (c,b) \rightarrow (i,a) \\ (2,c) \rightarrow (c,b) \end{matrix} \Rightarrow (2,c) \rightarrow (i,a)$$ $$D_{abci}^i:\ \ \begin{matrix} \bar{C_a^i} \\ | \\ | \\ c-----i \end{matrix} \ \begin{matrix} (c,b) \rightarrow (i,a) \\ (i,c) \rightarrow (c,b) \end{matrix} \Rightarrow (i,c) \rightarrow (i,a)$$ $$D_{abcd}^i:\ \ \begin{matrix} \bar{C_a^i} \\ | \\ | \\ c-----d \end{matrix} \ \begin{matrix} (c,b) \rightarrow (i,a) \\ (c,d) \rightarrow (c,b) \end{matrix} \Rightarrow (c,d) \rightarrow (i,a)$$ We make several observations: \begin{itemize} \item $D_{ab2i}^i = -D_{a2}^i$. \item $D_{ab2c}^i = -D_{abc2}^i$. \item $D_{abci}^i = -D_{ac}^i$. \item $D_{abcd}^i = D_{adc}^i = -D_{acd}^i$. \end{itemize} Now that we have these recursive relations, we can plug them into our various equations. We will use the following equalities: \begin{eqnarray*} m(D_{ab}^i;i,b) + 1 & = & m(D_a^i;i,b) \\ m(D_{ab2}^i;2,b) + 1 & = & m(D_{ab}^i;2,b) = m(D_a^i;2,b) \\ m(D_{abc}^i;b,c) + 1 & = & m(D_a^i;b,c) \end{eqnarray*} And for all the other coefficients we have: $$m(D_*^i;x,y) = m(D_a^i;x,y)$$ For convenience, we will write $m(x,y) = m(D_a^i;x,y)$ in what follows: \begin{eqnarray*} (m(i,a)+1)D_a^i + m(2,i)D_{a2}^i & = & \sum_{b\neq a,i,2}{-m(i,b)D_{ab}^i} \\ & = & \sum_{b\neq a,i,2}{-(m(D_{ab}^i;i,b)+1)D_{ab}^i} \\ & = & \sum_{b\neq a,i,2}{\left({m(2,b)D_{ab2}^i + \sum_{c\neq b,i,2}{m(b,c)D_{abc}^i}}\right)} \end{eqnarray*} Note that: \begin{eqnarray*} m(2,b)D_{ab2}^i & = & (m(D_{ab2}^i;2,b)+1)D_{ab2}^i \\ & = & -m(2,i)D_{ab2i}^i + \sum_{c\neq b,i,2}{-m(2,c)D_{ab2c}^i} \\ m(b,c)D_{abc}^i & = & (m(D_{abc}^i;b,c)+1)D_{abc}^i \\ & = & -m(2,c)D_{abc2}^i - m(i,c)D_{abci}^i - \sum_{d\neq c,b,i,2}{m(c,d)D_{abcd}^i} \end{eqnarray*} Therefore: $$m(2,b)D_{ab2}^i + \sum_{c\neq b,i,2}{m(b,c)D_{abc}^i} = $$ $$-m(2,i)D_{ab2i}^i + \sum_{c\neq b,i,2}{\left({-m(2,c)(D_{ab2c}^i + D_{abc2}^i) - m(i,c)D_{abci}^i - \sum_{d\neq c,b,i,2}{m(c,d)D_{abcd}^i}}\right)} = $$ $$m(2,i)D_{a2}^i + \sum_{c\neq b,i,2}{\left({m(i,c)D_{ac}^i + \sum_{d\neq c,b,i,2}{m(c,d)D_{acd}^i}}\right)}$$ We plug this back in above to find: $$(m(i,a)+1)D_a^i + m(2,i)D_{a2}^i = \sum_{b\neq a,i,2}{\left({m(2,i)D_{a2}^i + \sum_{c\neq b,i,2}{\left({m(i,c)D_{ac}^i + \sum_{d\neq c,b,i,2}{m(c,d)D_{acd}^i}}\right)}}\right)}$$ We notice that: \begin{eqnarray*} \sum_{c\neq b,i,2\ }{\sum_{d\neq c,b,i,2}{m(c,d)D_{acd}^i}} & = & \frac{1}{2}\sum_{\substack{c\neq d \\ c,d\neq b,i,2}}{m(c,d)(D_{acd}^i + D_{adc}^i)} \\ & = & \frac{1}{2}\sum_{\substack{c\neq d \\ c,d\neq b,i,2}}{m(c,d)(D_{acd}^i - D_{acd}^i)} \\ & = & 0 \end{eqnarray*} Therefore: $$(m(i,a)+1)D_a^i + m(2,i)D_{a2}^i = \sum_{b\neq a,i,2}{\left({m(2,i)D_{a2}^i + \sum_{c\neq b,i,2}{m(i,c)D_{ac}^i}}\right)}$$ Returning to our first equation, we obtain (simply replacing $b$ by $c$ in the second equality): \begin{eqnarray*} (m(i,a)+1)D_a^i + m(2,i)D_{a2}^i & = & \sum_{b\neq a,i,2}{-m(i,b)D_{ab}^i} \\ & = & \sum_{c\neq a,i,2}{-m(i,c)D_{ac}^i} \\ & = & \left({\sum_{c\neq b,i,2}{-m(i,c)D_{ac}^i}}\right) + m(i,a)D_{aa}^i - m(i,b)D_{ab}^i \end{eqnarray*} Since $D_{aa}^i = D_a^i$, we can cancel and rearrange terms to write: $$\sum_{c\neq b,i,2}{m(i,c)D_{ac}^i} = -D_a^i - m(i,b)D_{ab}^i - m(2,i)D_{a2}^i$$ Therefore: \begin{eqnarray*} (m(i,a)+1)D_a^i + m(2,i)D_{a2}^i & = & \sum_{b\neq a,i,2}{\left({m(2,i)D_{a2}^i - D_a^i - m(i,b)D_{ab}^i - m(2,i)D_{a2}^i}\right)} \\ & = & \sum_{b\neq a,i,2}{\left({-D_a^i - m(i,b)D_{ab}^i}\right)} \\ & = & -\left({\sum_{b\neq a,i,2}{D_a^i}}\right) + (m(i,a)+1)D_a^i + m(2,i)D_{a2}^i \end{eqnarray*} Finally, cancelling gives us: $$\sum_{b\neq a,i,2}{D_a^i} = 0$$ We now need to consider the possibilities for the color $i$. If $i = 2$, then: $$\sum_{b\neq a,i,2}{D_a^i} = \sum_{b\neq a,2}{D_a^i} = (k-2)D_a^i = 0$$ However, if $i \neq 2$, then: $$\sum_{b\neq a,i,2}{D_a^i} = (k-3)D_a^i = 0$$ Since we have already dealt with the cases when $k \leq 3$ in Lemma~\ref{L:knots}, Lemma~\ref{L:k=2} and Lemma~\ref{L:k=3}, we can conclude that $D_a^i$ is trivial modulo (*) for every $i$. Hence, $D_a$ is trivial for every $a$. Finally, this means that $(1+m(D;1,2))D = 0$, so $D$ will also be trivial modulo (*). This completes the proof. $\Box$ \begin{thm} \label{T:induction} If D has any large components C, then D is trivial modulo (*). \end{thm} {\sc Proof:} As in Lemma~\ref{L:k=3_induct}, we induct on the number of large components of $D$, and the argument for the general case is almost identical to the argument for the base case. We just have to observe that whenever we apply the relation (*), we can ignore large components of $D$ (other than the one we're working with), since the diagrams they give rise to have fewer large components than $D$ does. This completes the induction and the proof. $\Box$ This theorem tells us that the only elements of $B^{csl}$ which are {\it not} in the kernel of the relation (*) are unitrivalent diagrams all of whose components are of degree 1 (i.e. line segments). By the arguments of Lemma~\ref{L:knots}, we also know that the degree 1 components with both endpoints of the same color are also trivial, so we only need to consider line segments with different colors on the two endpoints. Restricted to the space generated by these elements, (*) is clearly trivial, so $B^{cl}$ is in fact simply the polynomial algebra over the reals generated by these unitrivalent diagrams (since (*) is trivial on this space, $B^{cl}$ inherits a multiplication from $B^{csl}$). We formalize this as a corollary: \begin{cor} \label{C:algebra} $B^{cl}(k)$ (and hence $A^{cl}(k)$) is isomorphic to the algebra ${\bf R}[x_{ij}]$, where each $x_{ij}$ is of degree 1, and $1\leq i < j\leq k$. \end{cor} It is well-known that these diagrams correspond to the pairwise linking numbers of the components, so we conclude: \begin{cor} \label{C:reduction} The pairwise linking numbers of the components of a link are the only finite type link concordance invariants of the link. \end{cor} \section{Acknowledgements} I wish to thank my advisor, Robion Kirby, for his advice and support, and all the members of the Informal Topology Seminar at Berkeley.
\section{Introduction} The family of doped manganites, R$_{1-x}$X$_{x}$MnO$_{3}$ (where R = La, Pr, Nd; X=Sr, Ca, Ba, Pb), has renewed both experimental and theoretical interests due to the colossal magnetoresistance and its potential technological application to magnetic storage devices. Apart from their unusual magnetic transport properties, experimental observations of a series of charge, magnetic and orbital ordering states in a wide range of dopant also stimulate extensive theoretical curiosities. Early theoretical studies of manganites concentrated their effort on the existence of metallic ferromagnetism. From the so-called ``Double Exchange'' (DE) model, \cite {Zener51} in which the mobility of itinerant electrons forces the localized spins to align ferromagnetically, one can understand qualitatively the relation of transport and magnetism. However the rich experimental phase diagrams are far beyond the DE model. For example, according to the DE model, itinerant electrons have the lowest kinetic energy in a tight binding model, and should be driven to form a more stable ferromagnetic phase when the system is half doped, {\it i.e.}, x=0.5. On the contrary, it is insulating rather than metallic ferromagnetic at a low temperature as expected theoretically. Furthermore, a charge ordered state was observed, which is characterized by an alternating Mn$^{3+}$ and Mn$^{4+}$ ions arrangement in the real space. \cite{Jirak85} Usually when the repulsive interaction between charge carriers dominates over the kinetic energy the charge carriers are driven to form a Wigner lattice. It has been shown experimentally that the charge ordering is sensitive to an applied magnetic field at low temperatures: resistance of a sample may decrease in several order of magnitude and the charge ordering disappears at a low temperature, \cite{Tomioka95} which implies that the repulsive interaction should have a close relation to the spin background. Although there have been extensive theoretical efforts on anomalous magnetic properties, \cite{Millis95} a comprehensive understanding on the physical origin of ordered states and their relations to the transport properties are still awaited. To explore electronic origin of these phenomena, we try to establish a more unified picture to understand the physics starting from an electronic model, which has been used to investigate the magnetic properties of the system extensively. We derive an effective Hamiltonian in the case of the strong on-site Coulomb interaction and Hund coupling by means of a projective perturbation approach. It is found that the virtual process of electron hopping produces an antiferromagnetic superexchange coupling between localized spins and a repulsive interaction between itinerant electrons. The antiferromagnetic correlation will enhance the repulsive interaction and suppress the mobility of electrons. In the half-doped case, i.e., $x=0.5 $, relatively strong repulsion will drive electrons to form a Wigner lattice. In the case of the Wigner lattice, we prove that the electrons are fully saturated while the localized spins form an antiferromagnetic background. Strictly speaking, the ground state possesses both anti- and ferromagnetic, i.e., ferrimagnetic long-range orders. \section{Effective Hamiltonian} The electronic model for doped manganites studied in this paper is defined as \cite{Inoue95} \begin{eqnarray} H &=&-t\sum_{\langle ij\rangle ,\sigma }c_{i,\sigma }^{\dagger }c_{j,\sigma }+U\sum_{i}n_{i,\uparrow }n_{i,\downarrow } \nonumber \\ &-&J_{H}\sum_{i}{\bf S}_{i}\cdot {\bf S}_{ic}+J_{AF}\sum_{\langle ij\rangle }% {\bf S}_{i}\cdot {\bf S}_{j}. \label{ele-ham} \end{eqnarray} where $c_{i,\sigma }^{\dagger }$ and $c_{i,\sigma }$ are the creation and annihilation operators for $e_{g}$ electron at site $i$ with spin ${\sigma }$ $(=\uparrow ,\downarrow )$, respectively. $\langle ij\rangle $ runs over all nearest neighbor pairs of lattice sites. ${\bf S}_{ic}=\sum_{\sigma ,\sigma ^{\prime }}{\bf \sigma }_{\sigma \sigma ^{\prime }}c_{i,\sigma }^{\dagger }c_{i,\sigma ^{\prime }}/2$ and ${\bf \sigma }$ are the Pauli matrices. $% {\bf S}_{i}$ is the spin operator of three $t_{2g}$ electrons with the maximal value $3/2$. $J_{H}>0$ is the Hund coupling between the $e_{g}$ and $% t_{2g}$ electrons. The antiferromagnetic coupling originates from the virtual process of superexchange of $t_{2g}$ electrons. In reality, the $% e_{g}$ orbital is doubly degenerated. For the sake of simplicity, we only consider one orbital per site, which amounts to assuming a static Jahn-Teller distortion and strong on-site interactions (relative to kinetic energy). \begin{figure}[tbp] \epsfysize=6cm \epsfxsize=8.5cm \epsfbox{fig1.eps} \caption{ Tow virtual processes of electron hopping in the restricted Hilbert space which favors antiferromagnetic correlation between neighboring sites. On the left side are the initial states, and on the right side are the mediate states. The process (a) leads to an effective attraction between electron and hole, and the process (b) leads to an effective attraction between electrons.} \end{figure} Usually the Hund coupling in the doped manganites is very strong, i.e., $% J_{H}S\gg t$. Large $J_{H}S$ suggests that most electrons form spin $S+1/2$ states with the localized spins on the same sites, which makes it appropriate to utilize the projective perturbation technique to investigate the low-energy physics of the Hamiltonian (\ref{ele-ham}). The effect of finite and large $J_{H}S$ can be regarded as the perturbation correction to the case of infinite $J_{H}$, which is described by a quantum double exchange model. \cite{Kubo72} Up to the second-order perturbation correction, there are two types of the virtual processes which contribute to the low energy physics (See in Fig. 1): (a). An electron hops from one site to one of the nearest neighbor empty site to form a spin $S-1/2$ state and then hops backward. The intermediate state has a higher energy $\Delta E_{a}=J_{H}(S+1/2)$ than the initial state. (b). One electron hops from one site to one of the singly occupied sites and then backward. The intermediate state has a higher energy $\Delta E_{b}=J_{H}S+U$ than that of the initial state. Hence, by using a projective perturbation approach, \cite {Shen97,Shen98} the effective Hamiltonian is written as \cite{Note} \begin{eqnarray} H_{eff} &=&-t\sum_{ij,\sigma }\bar{c}_{i,\sigma }^{\dagger }\bar{c}% _{j,\sigma }+J_{AF}\sum_{\langle ij\rangle }\bar{{\bf S}}_{i}\cdot \bar{{\bf % S}}_{j} \nonumber \\ &+&\frac{2St^{2}}{J_{H}(2S+1)^{2}}\sum_{ij}\left( \frac{{\bf S}_{i}}{S}\cdot \frac{\tilde{{\bf S}}_{j}}{S+\frac{1}{2}}-1\right) P_{ih}P_{js}^{+} \nonumber \\ &+&\frac{t^{2}}{J_{H}S+U}\sum_{ij}\left( \frac{\tilde{{\bf S}}_{i}}{S+\frac{1% }{2}}\cdot \frac{\tilde{{\bf S}}_{j}}{S+\frac{1}{2}}-1\right) P_{is}^{+}P_{js}^{+} \label{eff-ham} \end{eqnarray} where $\bar{{\bf S}}_{i}={\bf S}_{i}P_{ih}+2S\tilde{{\bf S}}% _{i}P_{is}^{+}/(2S+1)$ and \[ \bar{c}_{i,\sigma }=\sum_{\sigma ^{\prime }}\frac{{\bf S}_{i}\cdot {\bf % \sigma }_{\sigma \sigma ^{\prime }}+(S+1)\delta _{\sigma \sigma ^{\prime }}}{% 2S+1}(1-n_{i,-\sigma ^{\prime }})c_{i,\sigma ^{\prime }}. \] $\tilde{{\bf S}}_{i}$ is a spin operator with spin $S+1/2$, and a combination of spin of electron and localized spin on the same site. $P_{ih}$ and $P_{is}^{+}$ are the projection operators for empty site and single occupancy of spin $S+1/2$. The first term in Eq.(\ref{eff-ham}) is the quantum double exchange model. \cite{Kubo72,Shen97} It enhances ferromagnetic correlation, and may be suppressed if the antiferromagnetic exchange coupling of localized spin is very strong. The second, third and fourth terms prefer antiferromagnetism to ferromagnetism. The third term describes an attractive particle-hole interaction since the value of the operator before $P_{ih}P_{js}^{+}$ is always non-positive. In another words, an repulsive interaction between electrons in the restricted space arises when the spin background deviates from a saturated ferromagnetic case. To simplify the problem, we take the large spin approximation, and keep $% J_{H}S=j_{h}$ and $J_{AF}S^{2}=j_{af}$. The spin operator is parameterized in polar angle $\theta $ and $\phi $. In the approximation, the Hamiltonian is further reduced to \begin{eqnarray} H_{cl} &=&-t\sum_{ij}c_{ij}\alpha _{i}^{\dagger }\alpha _{j}-2j_{af}\sum_{ij}\sin ^{2}\frac{\Theta _{ij}}{2} \nonumber \\ &+&\sum_{ij}2\sin ^{2}\frac{\Theta _{ij}}{2}\left( \frac{t^{2}}{2j_{h}}-% \frac{t^{2}}{j_{h}+U}\right) \alpha _{i}^{\dagger }\alpha _{i}\alpha _{j}^{\dagger }\alpha _{j} \nonumber \\ &-&\sum_{ij}\frac{t^{2}}{2j_{h}}\sin ^{2}\frac{\Theta _{ij}}{2}(\alpha _{i}^{\dagger }\alpha _{i}+\alpha _{j}^{\dagger }\alpha _{j}) \label{cl-ham} \end{eqnarray} where \[ c_{ij}=\cos \frac{\theta _{i}}{2}\cos \frac{\theta _{j}}{2}+\sin \frac{% \theta _{i}}{2}\sin \frac{\theta _{j}}{2}e^{-i(\phi _{i}-\phi _{j})} \] ; \[ \cos \Theta _{ij}=\cos \theta _{i}\cos \theta _{j}+\sin \theta _{i}\sin \theta _{j}\cos (\phi _{i}-\phi _{j}); \] \[ \alpha _{i}=\cos \frac{\theta _{i}}{2}(1-n_{i,\downarrow })c_{i,\uparrow }+\sin \frac{\theta _{j}}{2}(1-n_{i,\uparrow })c_{i,\downarrow }. \] Physically, $\alpha $ is an electronic operator which is fully polarized along the localized spin on the same site. $|c_{ij}|=\cos (\Theta _{ij}/2)$ and approaches to zero when $\Theta _{ij}\rightarrow \pi $. If we neglect the Berry phase in $c_{ij}$, the first term gets back to the classical DE model. Now it is clear that the ferromagnetism is always predominant in the ground state if other terms in the effective Hamiltonian (Eq. (\ref{cl-ham}% )) are neglected. The sign of the interaction \begin{mathletters} \begin{equation} V_{ij}=2\sin ^{2}\frac{\Theta _{ij}}{2}\frac{t^{2}}{2j_{h}}\frac{U-j_{h}}{% U+j_{h}} \label{interaction} \end{equation} is determined by the ratio $j_{h}/U$. If $U$ is less than $j_{h}$, the interaction is attractive, but if $U$ is greater than $j_{h}$, the interaction is repulsive. The attractive or repulsive interaction will lead to different physics. Hence $U=j_{h}$ is a quantum critical point. The influence of the on-site Coulomb interaction will change qualitatively (not just quantitatively) the physics of the doped manganites, which is usually ignored. In the case of small $U$, the attractive interaction will drive electrons to accumulate together to form an electron-rich regime. i.e., the phase separation may occur when the spin background becomes antiferromagnetism. \cite{Shen98} Monte Carlo simulation by Dagotto {\it et al.} \cite{Dagotto97} shows that the phase separation occurs in the case of $% U=0$. However, the phenomenon was not observed in the case of large $U$. The phase diagram of $U=0$ is also seen in Ref.\cite{Riera97}. From our analysis, the attractive interaction originates from the virtual process (b). Due to the double occupancy in the intermediate state, an extra energy $% U$ costs in the process. When $U$ is sufficiently large, the process (b) will be suppressed and the process (a) becomes predominant. The net interaction between electrons is repulsive. Therefore the phase separation may occur only if $U<j_{h}$. \section{Origin of Wigner lattice} We are now\ in the position to discuss the instability to the Wigner lattice. In the doped manganites, the on-site Coulomb interaction is much stronger than the Hund's rule coupling, {\it i.e.}, $U\gg J_{H}S.$ \cite {Satpathy96} In the case, the process (b) in Fig.1 needs a much higher energy to be excited than the process (a) does. The process (a) dominates over the process (b). The effective interaction is repulsive. Hence we shall focus on the case of strong correlation (i.e., $U\gg J_{H}S$). To simplify the problem, we take $U\rightarrow +\infty $ and neglect the term containing $U$ in Eq.(\ref{cl-ham}). A finite and large $U$ will produce minor quantitative (not qualitative) changes of the physics we shall discuss. The ratio of the repulsion to the hopping term $r=(t/j_{h})\sin ^{2}\frac{\Theta _{ij}}{2}/\cos \frac{\Theta _{ij}}{2}$ depends on not only $t/j_{h}$, which is usually very small, but also the angle of two spins. $r=0$ if $\Theta _{ij}=0$, and $+\infty $ if $\Theta _{ij}=\pi $. In other words, the ratio could become divergent in the antiferromagnetic spin background ($\Theta _{ij}=\pi $) even though $t/j_{h}$ is very small. Relatively large ratio will make a state with a uniform density of electrons unstable. To understand the physical origin for the Wigner lattice at $x=0.5$, we first see what happens in the antiferromagnetic background. When all $\Theta _{ij}\rightarrow \pi $, the average energy per bond is $-2j_{h}$ if the two sites are empty or occupied, and $-(2j_{af}+t^{2}/j_{h})$ if one site is empty and another one is occupied. The later has a lower energy. At $x=1/2$, $\langle \alpha _{i}^{\dagger }\alpha _{i}\rangle =1/2$. The average energy per bond is $-(2j_{af}+t^{2}/2j_{h})$ for a state with a uniform density of electrons. If the electrons form a Wigner lattice, i.e., $\langle (\alpha _{i}^{\dagger }\alpha _{i}-1/2)(\alpha _{j}^{\dagger }\alpha _{j}-1/2)\rangle =-1/4$, the average energy per bond is $% -(2j_{af}+t^{2}/j_{h})$, which is lower than that of the state with a uniform density. Therefore in the antiferromagnetic background a uniform density state is not stable against the Wigner lattice even for a small $% t/j_{h}$. The same conclusion can be reached by means of the random phase approximation. On the other hand, the formation of the Wigner lattice will also enhance the antiferromagnetic exchange coupling from $-j_{af}$ to $% -(j_{af}+t^{2}/2j_{h})$. The phase diagram of the ground state is determined by the mean field approach. Several of the features are determined in several limits: for example, the ground state is ferromagnetic at $t/j_{h}=0$ and $j_{af}=0$. Due to the instability to the Wigner lattice or charge density wave for finite $t/j_{h}$ and $j_{af}$ we take $\langle \alpha _{i}^{\dagger }\alpha _{i}-1/2\rangle =\Delta e^{i{\bf Q}\cdot {\bf r}_{i}}$ where ${\bf Q}=(\pi ,\pi ,\cdots )$ and $\langle \cdots \rangle $ is the ground state average. We also take $\langle c_{ij}\rangle =\cos (\Theta /2)$ and $\langle \sin ^{2}(\Theta _{ij}/2)\rangle =\sin ^{2}(\Theta /2)$.\cite{Calderon98} The free energy per bond is \end{mathletters} \begin{eqnarray} {\cal E}(\Delta ,\Theta ) &=&-\int \frac{dk}{(2\pi )^{d}}\sqrt{\epsilon ^{2}(k)\cos ^{2}\frac{\Theta }{2}+4\frac{t^{4}}{j_{h}^{2}}\sin ^{4}\frac{% \Theta }{2}\Delta ^{2}} \nonumber \\ &-&(j_{af}+\frac{1}{4}\frac{t^{2}}{j_{h}})\sin ^{2}\frac{\Theta }{2}+\frac{% t^{2}}{j_{h}}\sin ^{2}\frac{\Theta }{2}\Delta ^{2} \end{eqnarray} where $\epsilon (k)=-t(\sum_{\alpha =1}^{d}\cos k_{\alpha })/d$ and d is the number of dimension. The integration runs over the reduced Brillouin zone. The phase diagram (Fig. 2) is obtained by minimizing the energy ${\cal E}% (\Delta ,\Theta )$. $\Delta $ and $\Theta $ are the order parameters for charge and magnetic orderings, respectively. $\Delta =0$ and $\Theta =0$ represents a full ferromagnetic (FM) phase, $\Delta =0$ and $\Theta \neq 0$ represents a canted ferromagnetic (CF) phase, $\Delta =1/2$ and $\Theta =\pi $ represents the Wigner lattice (WL), and $\Delta <1/2$ and $\Theta \neq 0$ represents a mixture of charge and spin density waves. A full ferromagnetic phase diagram appears at smaller $t/j_{h}$ and $j_{af}$, which indicates that the double exchange ferromagnetism is predominant. The Wigner lattice appears at a larger $t/j_{h}$ and $j_{af}$. The antiferromagnetic coupling originating from the virtual precess (a) and superexchange coupling of the localized spins can suppress the double exchange ferromagnetism completely. A canted ferromagnetic phase is between the two phases. At $j_{af}=0$, the transition from ferromagnetism to the Wigner lattice occurs at $% t^{2}/j_{h}=2\int dk\epsilon (k)/{(2\pi )^{d}}$ which equals to $0.63662t$ for $d=1$, $0.405282t$ for $d=2$, and $0.336126t$ for $d=3$. When the effective potential energy $t/j_{h}$ becomes to dominate over the kinetic energy, the ferromagnetic phase is unstable against the Wigner lattice. For a finite $j_{af}$, a smaller $t/j_{h}$ is required to form a Wigner lattice. However $t/j_{h}$ must be nonzero, even for a large $j_{af}$. In the double exchange model, i.e. $j_{h}\rightarrow +\infty $, we do not expect that the Wigner lattice could appear at low temperatures at $x=1/2$ unless a strong long-range Coulomb interaction is introduced. \begin{figure}[tbp] \epsfxsize=8.5cm \epsfbox{fig2.eps} \caption{ The phase diagram of the ferromagnetic Kondo lattice model on a cubic lattice ($d=3$) at $x=1/2$.} \end{figure} \section{Ferrimagnetism and Wigner lattice} We go back to Eq. (\ref{eff-ham}) to discuss the magnetic properties of the ground state (or at zero temperature) in the case that the Wigner lattice is formed at $x=1/2$ ($1-x$ is the density of electrons). The charge ordering in the manganite is an alternating Mn$^{3+}$ and Mn$^{4+}$ arrangement rather than a charge density modulation, which means $n_{i}=1$ or 0. A d-dimensional hyper-cubic lattice can be decomposed onto two sublattice $% {\cal A}$ and ${\cal B}$. In the charge ordering state, suppose that all electrons occupy the sublattice ${\cal A}$, then \begin{equation} P_{ih}P_{js}^{+}=\left\{ \begin{array}{ll} 1, & \makebox{if $i\in {\cal B}$ and $j\in {\cal A}$}; \\ 0, & \makebox{otherwise.} \end{array} \right. \end{equation} The first term in Eq.(\ref{eff-ham}) must be suppressed completely as the Wigner lattice is a static real space pattern, {\it i.e.}, the hopping processes are forbidden. In the case, the Hamiltonian is reduced to \begin{equation} H_{AF}=J_{AF}^{\prime }\sum_{i\in {\cal B},j\in {\cal A}}\left( \frac{{\bf S}% _{i}}{S}\cdot \frac{\tilde{{\bf S}}_{j}}{S+\frac{1}{2}}-1\right) \label{anti-ham} \end{equation} where $J_{AF}^{\prime }=J_{AF}S^{2}+2St^{2}/J_{H}(2S+1)^{2}$ and the summation runs over the nearest neighbor pairs. This is an antiferromagnetic Heisenberg model. The spin on the sublattice ${\cal A}$ is $S+1/2$ as the electrons on the sites form spin $S+1/2$ state with the localized spins, and the spin on the sublattice ${\cal B}$ is $S$. According to Lieb-Mattis theorem, \cite{Lieb62} the ground state of Eq.(\ref{anti-ham}) is unique apart from spin SU(2) ($2S_{tot}+1$)-fold degeneracy. The total spin of the ground state $S_{tot}$ is equal to the difference of the maximal total spins of two sublattices. In the case, \begin{equation} S_{tot}=\frac{N_{e}}{2} \end{equation} which is also the maximal total spin of electrons ($N_{e}$ is the number of electrons). It seems to be that all electrons are saturated fully while the localized spins form a spin singlet state. Furthermore, it is shown rigorously that the ground state possesses antiferromagnetic long-range order as well as ferromagnetic one for any dimension. \cite{Shen94} \section{Discussion and summary} We wish to point out that, in the case that the Wigner lattice is formed, the magnetic structure established here is unlikely in full agreement with all experimental observations. \cite{Wollan55,Tokura96} The model discuss here is a simplified theoretical model which has neglected some effects, such as the orbital degeneracy of $e_{g}$ electrons, strong John-Teller effect and lattice distortion. Methodologically, we apply the projective perturbation approach to deal with the model. The strong electron-electron correlations has been successfully taken into account by the projection process. The perturbation process tells us that the effective Hamiltonian should be valid at small $t/j_{h}$, which requires a strong Hund coupling comparing with the hopping integral $t$. In practice, the parameters of the model for doped manganites are roughly estimated as $U\approx 5.5eV$, $% J_{H}\approx 0.76eV$, $t\approx 0.41eV$, $J_{AF}\approx 2.1meV.$\cite {Satpathy96} Thus,\ $U/J_{H}S\approx 4.82$ and $\ t/J_{H}S\approx 0.359$. For these parameters, the Wigner lattice at low temperatures is stable in the phase diagram in Fig.2. Therefore, the superexchange process in Fig.1(a) should play an important role in driving electrons to form the Wigner lattice no matter whether the direct nearest neighbor Coulomb interaction is strong. It is worth mentioning that the direct Coulomb interaction will always favor to form the Wigner lattice. \cite{Lee97} If the direct Coulomb interaction is also included in the electronic model, which is not much screened, the stability of Wigner lattice will be greatly enhanced. Note that the Coulomb interaction is independent of the magnetic structure, and should not be very sensitive to an external magnetic field. The effect of field-induced melting of the Wigner lattice suggests that the physical origin of the state may be closely related to the magnetic structure, which is an essential ingredient of the present theory. In the actual compounds, both the mechanisms should have important impact on the electronic behaviors. It is unlikely that only one of them is predominant. As for the mean field approximation, when it is sure that the instability of Wigner lattice occurs at low temperatures, it is an efficient and powerful tool to determine the phase diagram, although some other physical quantities, such as critical exponents, cannot be obtained accurately. Due to the strongly correlations of electrons, it is still lack of numerical results to verify the present theoretical prediction as this is the first time to discuss instability of the Wigner lattice in a model without nearest neighbor or long-range interactions. When the system is deviated from $x=0.5$, the superexchange interaction is still very important to determine the behaviors of electrons. Recently, it was observed that the charge stripes in (La,Ca)MnO% $_{3}$ pair. \cite{Mori98} However, the two pairing stripes of Mn$^{3+}$ ions are separated by a stripe of Mn$^{4+}$ ions. This fact suggests the nearest neighbor interaction should be very strong. Of course, for a comprehensive understanding of the phase diagram, including anisotropic properties of charge and magnetic orderings, we need to take other effects into account. The role of Hund's rule coupling in the doped manganites has been emphasized since the double exchange mechanism was proposed. However, the rich phase diagrams in the doped manganites go beyond the picture. Our theory shows that the on-site Coulomb interaction also has an important impact on the physical properties of the system. In the model we investigate, the sign of the effective interaction in Eq.(\ref{interaction}) depends on the ratio of j% $_{h}$/U. Repulsive or attractive interaction will lead to quite different physics. In one of our recent papers \cite{Shen98}, we proposed a mechanism of phase separation based on the attractive interaction caused by the virtual process (a) in Fig. 1, and neglect the on-site interaction U. The phase separation can occur in the high and low doping regions. As the mechanism of the phase separation is completely opposite to the mechanism of the Wigner lattice we discuss in this paper, we have to address the issue which one \ occurs for the doped manganites. From the estimation of the model parameters for the actual compound, $U/j_{h}\approx 4.82$. Thus, the effective interaction should be repulsive, not attractive. From this sense, the phase separation we predicted in Ref. \cite{Shen98} could not occur in doped manganites. In fact, both the phase separation and the Wigner lattice were observed in the family of samples with different dopings. For example the phase separation was observed in La$_{1-x}$Cu$_{x}$MnO$_{3}$ with x=0.05 and 0.08.\cite{Allodi97} It is worth pointing out that the electronic model is a simplified model for doped manganites since the degeneracy of e$_{g}$ electrons and the Jahn-Teller effect have been neglected. The importance of the orbital degeneracy of e$_{g}$ electron has been extensively discussed, especially for the ferromagnetism near $x=0$. If we take into account the orbital degeneracy, there may exist an superexchange virtual process in the ferromagnetic or A-type antiferromagnetic background, in which the superexchange coupling between different orbits instead of the spin indices in Fig. 1 could produce an attractive interaction as we predicted in Ref. \cite{Shen98}. The mechanism for phase separation may still be responsible for the experimental observation. The investigation along this direction is in progress. Before ending this paper, we would like to address the stability of the Wigner lattice with respect to the transfer $t$. Some experimental analysis suggested that a {\it relatively small} $t$ would favor to form the Wigner lattice,\cite{Tokura96} which seems to be unlikely in contradiction with the phase diagram in Fig. 2. In the present theory, the Wigner lattice occurs in a moderate value of $t$. On one hand, a large $t$ ($\gg j_{h}$), of course, will lead to the instability of the Wigner lattice and destroy the double exchange ferromagnetism. In the case, a paramagnetic phase should be favored at low temperatures. The perturbation technique used in this paper is also not valid. So the region of Wigner lattice in Fig.2 cannot be naively extended to the large $t$ case. On the other hand, when $t$ becomes very small comparing with $j_{h}$, the Wigner lattice should also be unstable since a small t means to enhance the ratio $j_{h}/t$ and a larger ratio is favorable to double exchange ferromagnetism. If the antiferromagnetism from $% t_{2g}$ electrons could compete over the double exchange ferromagnetism at $% x=0.5$, it would suppress ferromagnetism at all the range of $x$.\cite {Shen97} The effective transfer $t$cos$(\Theta /2)$ is determined by either t or $\Theta $ the angle of the two spins. The Wigner lattice is also accompanied by the strong antiferromagnetic correlation. The field-induced melting effect indicates that the Wigner lattice is unstable in the ferromagnetic background, which also indicates the important role of the antiferromagnetic correlation to stabilize the Wigner lattice. A smaller $% j_{af}$ will reduce the angle $\Theta $ and should also lead to the instability of the Wigner lattice. Thus, a small $t$ does not always favor to form the Wigner lattice. In short, we derived an effective Hamiltonian for an extended Kondo lattice model, based on which a physical mechanism for charge ordering in half-doped manganites is naturally put forward. \acknowledgments This work was supported by a CRCG research grant at the University of Hong Kong.
\section{ Introduction} \vskip 0.2 cm In this paper we compute Stokes' matrices and monodromy group for the Frobenius manifold given by the quantum cohomology of the projective space ${\bf CP}^d$. Our main motivation is to study the links between quantum cohomology and the theory of coherent sheaves. Stokes matrices first appeared in the theory of WDVV equations of associativity in the paper \cite{Dub4} by B. Dubrovin. WDVV equations were formulated in a geometrical setting: the theory of Frobenius manifolds. From then on, the notion of Frobenius manifold has been largely studied in many papers, of which we cite \cite{Dub1}, \cite{Dub2}, \cite{Dub3}. The Stokes' matrix is a part of the {\it monodromy data} for a semisimple Frobenius manifold. Monodromy data can serve as natural moduli of semisimple Frobenius manifolds. More precisely, any local chart of the atlas covering the manifold is reconstructed from the monodromy data. The glueing of the local charts is described by the action of the braid group on the data, particularly, on the central connection matrix and on the Stokes' matrix \cite{Dub1} \cite{Dub2}. One well-known example of Frobenius manifolds is the quantum cohomology of smooth projective varieties \cite{Manin} \cite{KM} \cite{RT} \cite{MDS} . It was conjectured \cite{Dub3} that the Stokes matrix for the quantum cohomology of a good Fano variety $X$ is equal to the Gram matrix of the bilinear form $\chi(E,F):= \sum_k ~(-1)^k~$dim$~Ext^k(E,F)$ computed on a full collection of exceptional objects in the derived category $Der^b(Coh(X))$ of coherent sheaves on $X$. More precisely, let $Der^b(Coh(X))$ be the derived category of coherent sheaves on a smooth projective variety $X$ of dimension $d$. An object $E$ of $Der^b(Coh(X))$ is called {\it exceptional} if $Ext^i (E,E)=0$ for $0<i<d$, $Ext^0(E,E)= {\bf C}$ and $Ext^d(E,E)$ is of the smallest dimension (if $X$ is a projective space, then $Ext^d(E,E)=0$). A collection $\{ E_1,...,E_s \}$ of exceptional objects is an {\it exceptional collection} if for any $1\leq m<n\leq s$ we have $Ext^{i}(E_n,E_m)=0$ for any $i\geq 0$, $Ext^i(E_m,E_n)=0$ for any $i\geq 0$ except possibly for one value of $i$. A {\it full exceptional collection} is an exceptional collection which generates $Der^b(Coh(X))$ as a triangulated category. This theory is developed in \cite{Rud1} \cite{Rud2} \cite{BP}. We say that a Fano variety is {\it good} if it has a full exceptional collection. It is known that $X={\bf CP}^d$ is good, the collection of sheaves on ${\bf CP}^d$ $\{ {\cal O}(n) \}_{n \in {\bf Z}}$ is exceptional, and $\{E_1,E_2,...,E_{d+1}\} :=\{ {\cal O}, {\cal O}(1), ~..., {\cal O}(d) \}$ is a full exceptional collection \cite{Bl}, \cite{GR}. In this case, $s_{ij}=\chi\left({\cal O}(i-1),{\cal O}(j-1)\right)$, $i,j=1,2,...,d+1$ has the ``canonical form'': $$ s_{ij}= \bin{d+j-i}{j-i}~~~~i<j$$ $$ s_{ii}=1,~~~~~s_{ij}=0~~~~i>j$$ The inverse to this matrix has entries $a_{ij}$ $$ a_{ij}=(-1)^{j-i} \bin{d+1}{j-i}~~~~i<j $$ $$ a_{ii}=1,~~~~~a_{ij}=0~~~~i>j$$ This matrix is equivalent to the one above with respect to the action of the braid group. We will also call it ``canonical''. The mentioned conjecture claims that the Stokes matrix of the quantum cohomology of ${\bf CP}^d$ is equal to the above Gram matrix (modulo the action of the braid group: remarkably, this action on the Stokes matrix for the Frobenius manifold coincides with the natural action of the braid group on the collections of exceptional objects \cite{Zas} \cite{Rud1}). \vskip 0.2 cm This conjecture has its origin in the paper by Cecotti and Vafa \cite{CV}, where another Stokes matrix introduced in \cite{Dub5} for the $tt^*$ equations was found in the case of the ${\bf CP}^2$ topological $\sigma$ model. It was suggested, on physical arguments, that the entries of the Stokes' matrix $S=\pmatrix{ 1&x&y\cr 0&1&z\cr 0&0&1 \cr } $ are integers. They must satisfy a Diophantine equation $x^2+y^2+z^2 -xyz=0$ whose integer solutions $(x,y,z)$ are all equivalent to (3,3,3) modulo the action of the braid group. The authors of \cite{CV} also suggested that their matrix must coincide with the Stokes matrix defined in the theory of WDVV equations of associativity, that is, in the geometrical theory of Frobenius manifolds for 2D topological field theories \cite{Dub4}, \cite{Dub1}. Later, in \cite{Zas}, the links between $N=2$ supersymmetric field theories and the theory of derived categories were further investigated and the coincidence of $\chi(E_i,E_j)$ with the Stokes matrix of $tt^*$ for ${\bf CP}^d$ was conjectured. The conjecture may probably be derived from more general conjectures by Kontsevich in the framework of categorical mirror symmetry. To my knowledge, the subject was discussed in \cite{Konpis} (I thank B. Dubrovin for this reference). \vskip 0.2 cm The main result of this paper is the proof (Theorem 2, $2^{\prime}$) that the conjecture about coincidence of the Stokes matrix for quantum cohomology of ${\bf CP}^d$ and the Gram matrix $\chi(E_i,E_j)$ of a full exceptional collection in $Der^b(Coh({\bf CP}^d))$ is true. In this way, we generalize to any $d$ the result obtained in \cite{Dub2} for $d=2$. We remark that it has not yet been proved that the Stokes' matrix for $tt^*$ equations and the Stokes' matrix for the corresponding Frobenius manifold coincide. This point deserves further investigation. \vskip 0.2 cm We also study the structure of the monodromy group of the quantum cohomology of ${\bf CP}^d$. The notion of {\it monodromy group} of a Frobenius manifold was introduced in \cite{Dub1}. We prove (Theorem 3) that for $d=3$ the group is isomorphic to the subgroup of orientation preserving transformations in the hyperbolic triangular group $[2,4,\infty]$. In \cite{Dub2} it was proved that for $d=2$ the monodromy group is isomorphic to the direct product of the subgroup of orientation preserving transformations in $[2,3,\infty]$ and the cyclic group of order 2, $C_2=\{\pm \}$. Our numerical calculations also suggest that for any $d$ even the monodromy group may be isomorphic to the orientation preserving transformations in $[2,d+1,\infty]$, and for any $d$ odd to the direct product of the orientation preserving transformations in $[2,d+1,\infty]$ by $C_2$. \vskip 0.2 cm \noindent {\bf Acknowledgments} \vskip 0.2 cm I am indebted to Prof. B Dubrovin who introduced me to the problem and constantly gave me suggestions and advice. I also thank M. Bertola and M. Mazzocco for useful discussions. \vskip 0.3 cm \section{ The system corresponding to ${\bf CP}^{k-1}$} \vskip 0.2 cm We introduce here the linear system of differential equations whose Stokes matrices are the Stokes matrices for the quantum cohomology of ${\bf CP}^{k-1}$ (we use the more convenient choice $k=d+1$). In the quantum cohomology of ${\bf CP}^{k-1}$ we choose flat coordinate $t^1,..,t^k$ for the symmetric non degenerate {\it bilinear form} $<~,~>$ : $$ \eta_{\alpha \beta}=<\partial_{\alpha},\partial_{\beta}>= \delta_{ \alpha+\beta,k+1} ~~~\hbox{ where }\partial_{\alpha}={\partial \over \partial t^{\alpha}} $$ Let $\eta$ be the matrix $(\eta_{\alpha \beta})$. In flat coordinates the {\it Euler vector field} is $$ E= \sum_{\alpha \neq 2} ~(1-q_{\alpha})t^{\alpha}{\partial \over \partial t^{\alpha} } +~k {\partial \over \partial t^2} $$ $$ q_1=0,~q_2=1,~q_3=2,~...,~q_{k}=k-1 $$ the {\it multiplication} is $$\partial_{\alpha}\cdot \partial_{\beta} = c^{\gamma}_{\alpha \beta}(t) \partial_{\gamma} ~~~~~\hbox{ where }~~ c_{\alpha \beta \gamma}(t):=\eta_{\delta \gamma} c^{\delta}_{\alpha \beta}(t) = \partial_{\alpha}\partial_{\beta} \partial_{\gamma} F(t)$$ $$ F(t)= {1 \over 6} \sum_{\alpha+\beta+\gamma=k+2}~t^{\alpha} t^{\beta} t^{\gamma} + \sum_{l=1}^{\infty} \Phi_{l}(t^3,...,t^k) e^{lt^2}$$ $$ \Phi_l(t^3,...,t^k)= \sum_{n=2}^{\infty} \left[ \sum_{\alpha_1+...+\alpha_n= (l+1)(k-1)+l-1+n } I(l;\alpha_1,...,\alpha_n) {t^{\alpha_1}...t^{\alpha_n} \over n!}\right] , ~~~~~~ I(1,k,k)=1 $$ and the {\it unity vector field } is $e = {\partial \over \partial t^1}$. Finally, let $${\mu}= \hbox{diag}(\mu_1,...,\mu_k)=\hbox{diag}(-{k-1\over 2},-{k-3 \over 2},...,{k-3 \over 2},{k-1 \over 2}), ~~~~~\mu_{\alpha}=q_{\alpha}-{d\over 2}$$ Consider the system of differential equations determining deformed flat coordinates (see \cite{Dub1} \cite{Dub2}): \begin{equation} \partial_z \xi= ({\cal U}(t)+{1\over z}{\mu}) \xi \label{zzzz} \end{equation} $$ \partial_{\alpha} \xi = z C_{\alpha}(t) \xi $$ where $z\in{\bf C}$, $\partial_z:={\partial \over \partial z}$ and $\xi$ is a column vector of components $$ \xi^{\alpha}= \eta^{\alpha\beta}\partial_{\beta}\tilde{t}(t,z) $$ Here $(\eta^{\alpha\beta})=(\eta_{\mu \nu})^{-1}$ and $\tilde{t}(t,z)$ is one of the $k$ (deformed) flat coordinates. ${\cal U}(t)$ is the matrix of multiplication by the Euler vector field $E(t)$, and $C_{\alpha}(t)$ is the matrix of entries $(C_{\alpha})^{\beta}_{~\gamma}:= c^{\beta}_{\alpha \gamma}$. The monodromy data of the system (\ref{zzzz}) are, by definition, the {\it monodromy data} of the quantum cohomology of ${\bf CP}^{k-1}$ in the local chart containing $t$. Let us compute $C_2(t)$ and ${\cal U}(t)$ at the semisimple point $(0,t^2,0,...,0)$: $$ E \cdot ~\partial_{\beta}=E^{\gamma} c^{\alpha}_{\gamma \beta} ~\partial_{\alpha} = E^2 c^{\alpha}_{2 \beta} ~\partial_{\alpha}= $$ $$ =k~c^{\alpha}_{2 \beta} \partial_{\alpha} \equiv {\cal U}^{\alpha}_{~~\beta} \partial_{\alpha} $$ Moreover $c_{2 \alpha \beta} (0,t^2,0,...,0)= \partial_2 \partial_{\alpha} \partial_{\beta} F (0,t^2,...,0)$. This immediately yields $$ C_2(0,t^2,0,...,0) = \pmatrix{ 0 & & & & e^{t^2} \cr 1 & 0 \cr & 1 & 0 \cr & & \ddots& \ddots \cr & & & 1 & 0 \cr }, ~~~~ {\cal U}(0,t^2,0,...,0) = \pmatrix{ 0 & & & & k e^{t^2} \cr k & 0 \cr & k & 0 \cr & & \ddots& \ddots \cr & & & k & 0 \cr } $$ Let $y_{\alpha} := \eta_{\alpha \beta} \xi^{\beta} \equiv \partial_{\alpha} \tilde{t}$. It satisfies \begin{equation} \partial_z y=\left(\hat{\cal U}(t^2) - {1 \over z} \mu \right) y \label{10p} \end{equation} \begin{equation} \partial_2 y= z \hat{C}_2(t^2) y \label{zzz1} \end{equation} where $$ \hat{C}_2(t^2):=\eta~C_2(0,t^2,...,0)~\eta^{-1}= \pmatrix{ 0 &1 \cr &0&1 \cr & & 0&1 \cr & & &\ddots& \ddots \cr & & & & 0 &1 \cr e^{t^2} & & & & & 0 \cr } $$ $$ \hat{\cal U}(t^2):=\eta~{\cal U}(0,t^2,...,0)~\eta^{-1} =\pmatrix{ 0 &k \cr &0&k \cr & & 0&k \cr & & &\ddots& \ddots \cr & & & & 0 &k \cr ke^{t^2} & & & & & 0 \cr } $$ \vskip 0.2 cm \noindent {\bf Lemma 1}: {\it Let ${y}(t^2,z)=(y_1(t^2,z),...,y_k(t^2,z))^T$ be a column vector solution of the above system (\ref{10p}) (\ref{zzz1}). With the following substitution \begin{equation} y_{\alpha}(t^2,z)= {1 \over k^{\alpha-1}} ~z^{{k-1\over 2}-\alpha+1}~ (z \partial_z)^{(\alpha-1)}~\varphi(t^2,z) ~~~ \label{40p} \end{equation} \begin{equation} \equiv z^{{k-1\over 2}-\alpha+1}~ \partial_2^{\alpha-1}\varphi(t^2,z),~~~ \alpha=1,2,...,k \label{zzz2} \end{equation} the above system is equivalent to the equations \begin{equation} (z \partial_z)^{k} \varphi = (k z)^k e^{t_2} ~\varphi \label{30p} \end{equation} \begin{equation} \partial_2^k\varphi =z^k e^{t_2} ~\varphi \label{zzz3} \end{equation} } \vskip 0.2 cm \noindent The proof is a simple calculation we leave to the reader. \vskip 0.2 cm The substitution of the lemma implies $\partial_2 \varphi= {1\over k} z \partial_z \varphi$. Then $$ e^{t^2\over k} {\partial \over \partial e^{t^2 \over k}}\varphi(t^2,z) = z {\partial \over \partial z } \varphi(t^2,z) $$ which implies (with abuse of notation) $$\varphi(t^2,z) \equiv \varphi(ze^{t^2\over k})$$ Namely, $\varphi$ (at $(0,t^2,...,0)$) depends on one argument $w=ze^{t^2\over k}$ and satisfies the {\it generalized hypergeometric equation} \begin{equation} \left(w {d \over dw}\right)^k \varphi(w) = (kw)^k~\varphi(w) \label{30pir} \end{equation} The equation is equivalent to the system \begin{equation} {d Y \over dw}= \left[ \hat{\cal U}+{\hat{\mu}\over w} \right] Y \label{10pir} \end{equation} where \begin{equation} Y_n(w)= {1 \over k^{n-1}} ~w^{{k-1\over 2}-n+1}~ (w \partial_w)^{(n-1)}~\varphi(w) ~~~n=1,2,...,k \label{40pir} \end{equation} $$ \hat{\cal U}:=\hat{\cal U}(0)= \left[ \matrix{ 0 & k & & & &\cr & 0 & k & & & \cr & & 0 & k & & \cr & & & \ddots & \ddots & \cr & & & & \ddots & k \cr k & & & & & 0 \cr } \right] $$ $$ \hat{\mu}:=-\mu=\hbox{ diag}\left({k-1\over 2}, {k-3\over 2}, {k-5 \over 2}, ..., -{k-3 \over 2}, -{k-1 \over 2} \right) $$ The system (\ref{10pir}) may also be interpreted as the system (\ref{10p}) with $t^2=0$. We will return later (section 4) on the connection between its monodromy data and the monodromy data of the system (\ref{10p}). \vskip 0.3 cm Let us study system (\ref{10pir}). We change notation and choose the more familiar letter $z$ instead of $w$. So, the system (\ref{10pir}) is re-written as \begin{equation} {d Y \over dz}= \left[ \hat{\cal U}+{\hat{\mu}\over z} \right] Y \label{10} \end{equation} and (\ref{40pir}) (\ref{30pir}) become \begin{equation} Y_n(z)= {1 \over k^{n-1}} ~z^{{k-1\over 2}-n+1}~ (z \partial_z)^{(n-1)}~\varphi(z) ~~~n=1,2,...,k \label{40} \end{equation} \begin{equation} \left(z {d \over dz}\right)^k \varphi(z) = (kz)^k~\varphi(z) \label{30} \end{equation} The point $z=0$ is e fuchsian singularity, and $z=\infty$ is a singularity of the second kind. (\ref{10}) has a fundamental matrix solution $Y_0(z)$ whose behaviour at $z=0$ is $$ Y_0(z)= (I+O(z)) z^{\hat{\mu}}z^{R}~~~~~~ R=\left[ \matrix{ 0 & k & & & &\cr & 0 & k & & & \cr & & 0 & k & & \cr & & & \ddots & \ddots & \cr & & & & \ddots & k \cr & & & & & 0 \cr } \right] $$ and the monodromy for a counter-clockwise loop around the origin is $e^{2 \pi i (\hat{\mu} +R)}$. The characteristic polynomial of the matrix $\hat{\cal U}$ is $0=\det( \hat{\cal U} - u)= (-u)^k+(-1)^{k+1} k^k$. It has $k$ distinct eigenvalues $u_{n}=k~e^{2 \pi i (n-1) \over k}$, $n=1,...,k$. The equations for the eigenvector ${\bf x}_n$ corresonding to $u_n$, namely $\hat{\cal U}{\bf x}_n=u_n~{\bf x}_n$, written for the components ${x^1}_n$, ...,${x^k}_n$ of the column vector ${\bf x}_n$ are $$ {x^{l+1}}_n=e^{2 \pi i (n-1) \over k}~{x^l}_n ~~~~ l=1,2,...,k-1 $$ $$ {x^1}_n= e^{2 \pi i (n-1) \over k} {x^k}_n $$ With the choice ${x^1}_n= e^{i \pi (n-1) \over k}$ we get ${\bf x}_n= (e^{i \pi (n-1) \over k},e^{i 3 \pi (n-1) \over k},e^{i 5 \pi (n-1) \over k},..., e^{-{i \pi (n-1) \over k}})^T$ ($T$ stends for transpose). The matrix $$ X= {1 \over \sqrt{k}}~[ {\bf x}_1 | {\bf x}_2 |...|{\bf x}_k]= {1 \over \sqrt{k}}~({x^j}_n)~~~~~{x^j}_n=e^{(2j-1)i \pi {n-1 \over k}}~~j,n=1,2,...,k $$ puts ${\cal U}$ in diagonal form: $$ U= X^{-1} \hat{\cal U} X = \hbox{diag}(u_1,u_2,...,u_n,...,u_k)~~~~u_n=k~e^{2 \pi i (n-1) \over k} $$ We stress that $u_i \neq u_j$ for $i \neq j$. The system (\ref{10}) is transformed by the gauge $X$ in an equivalent form \begin{equation} {d \tilde{Y} \over dz}= \left[ U+{V\over z} \right] \tilde{Y} \label{20} \end{equation} $$ \tilde{Y}= X^{-1}Y,~~~U=X^{-1}\hat{\cal U}X,~~~V= X^{-1} \hat{\mu} X $$ Observe that $$ \eta \hat{\mu} + \hat{\mu} \eta=0 ~~~~~~~ X X^T = \eta^{-1} $$ This implies that $V$ is skew-symmetric $$ V+V^T=0 $$ With the gauge $X$, $Y_0(z)$ transforms into $$ \tilde{Y}_0(z)= (X^{-1} +O(z)) ~z^{\hat{\mu}}z^R, ~~~~z\to 0 $$ \vskip 0.3 cm \section{Asymptotic Behaviour and Stokes' Phenomenon} \vskip 0.2 cm Our aim is to explicitely compute a Stokes' matrix for the above system (\ref{20}), or for the system (\ref{10}). The system (\ref{20}) has formal solution $$ \tilde{Y}_F = \left[ I + {F_1 \over z}+{F_2 \over z^2}+ ... \right]~e^{z~U} $$ where $F_j$'s are $k \times k$ matrices. It is a well known result that fundamental matrix solutions exist which have $\tilde{Y}_F$ as asymptotic expansion for $z \to \infty$ in some ``admissible'' sectors of the complex plane of angular width greater then $\pi$. In order to find such sectors we need the so called {\it Stokes' rays}, defined by $$ \Re ~((u_r-u_s)z)=0 ,~~r \neq s~~~ \Rightarrow ~~~\arg z= -\arg (u_r-u_s)+{\pi \over 2} + m \pi ,~~~m=0,1 $$ There exists a {\it unique} solution of the system asymptotic to $\tilde{Y}_F$ in a sector greater then $\pi$ and bounded by the first two Stokes' rays we meet extending over $\pi$ the angular width of the sector. The general theory of Stokes' phenomenon is found in the classical paper by W. Balser, W.B. Jurkat, D.A. Lutz \cite{BJL1}. Stokes matrices are also a natural monodromy datum in the theory of isomonodromy deformation \cite{ITS} \cite{JMU} \cite{JM1} \cite{JM2} \cite{Dub1} \cite{DM}. \vskip 0.2 cm A possible choice for the labelling of the rays is the following: we call $R_{rs}$ the Stokes' ray $$ R_{rs}= \{ z=- i \rho (\bar{u_r} - \bar{u_s}), ~~~\rho>0 \}~~~~~r \neq s $$ \vskip 0.3 cm \noindent {\bf Lemma 2}: {\it For $r<s$ the Stokes' rays of the system (\ref{20}) are $$ R_{rs}= \left\{z= \rho ~\exp \left( i\left[ {2 \pi \over k} - {\pi \over k} (r+s)\right] \right),~~~\rho>0 \right\} $$ $$ R_{sr}= -R_{rs} $$ } \vskip 0.2 cm \noindent {\bf Proof}: Just compute $$ -i (\bar{u_r} -\bar{u_s})= -i ( e^{-i {2 \pi \over k}(r-1) }-e^{-i{2 \pi \over k}(s-1)})= $$ $$ = 2 \sin \left({\pi \over k}(s-r)\right) ~e^{\left(i\left[ {2 \pi \over k} - {\pi \over k} (r+s)\right] \right)} $$ Then we note that $ \sin \left({\pi \over k}(s-r)\right)$ is positive because $0<s-r\leq k-1$. \rightline{$\Box$} \vskip 0.3 cm \noindent {\bf Remark 1}: $R_{rs}=R_{pq}$ for $r+s=p+q$. $R_{12}$ is at $\arg z = - {\pi \over k}$, $R_{13}$ is at $\arg z = - {2 \pi \over k}$, and so on. For $r+s = k+2$ the corresponding $R_{rs}$'s are at $\arg z = -\pi$ and the $R_{sr}$'s are at $\arg z =0$. $R_{k-1,k}$ is at the angle $-2 \pi +{3 \pi \over k}$ or, equivalently, at ${3 \pi \over k}$. See figure 1. \begin{figure} \epsfxsize=15cm \centerline{\epsffile{figr1.eps}} \caption{ Stokes' rays} \end{figure} \vskip 0.2 cm We choose two ``admissible'' overlapping sectors in a canonical way. Let $l$ be an ``admissible'' line through the origin, namely a line not containing Stokes' rays. For our purposes we take $$ l= \{ z ~|~ z= \rho e^{i \epsilon}, ~~~\rho \in{\bf R},~~~ 0<\epsilon< {\pi \over k} \} $$ $l$ has a natural orientation inherited from {\bf R}. We call $\Pi_R$ and $\Pi_L$ the half planes to the right/left of $l$ w.r.t its orientation. $$ \Pi_R=\{ - \pi +\epsilon < \arg z < \epsilon\}~~~\Pi_L=\{\epsilon<\arg z < \pi + \epsilon \}$$ We then define two different ``admissible'' sectors ${\cal S}_L$, ${\cal S}_R$ which contain $l$ $$ {\cal S}_L=\{ z\in {\bf C}~|~ 0<arg z < \pi + {\pi \over k} \}\supset \Pi_L $$ $$ {\cal S}_R=\{ z\in {\bf C}~|~ - \pi <arg z < {\pi \over k} \} \supset \Pi_R $$ We call che corresponding solutions $\tilde{Y}_L(z)$ and $\tilde{Y}_R(z)$. \vskip 0.2 cm \noindent {\bf Definition: } The {\it Stokes' matrix} of the system (\ref{20}) with respect to the admissible line $l$ is the connection matrix $S$ such that $$ \tilde{Y}_L(z)=\tilde{Y}_R(z)S~~~~~~0<\arg z<{\pi \over k} $$ On the opposite overlapping region one can prove (as a consequence of the skew-symmetry of $V$, see \cite{Dub1}) that $$ \tilde{Y}_L(z)=\tilde{Y}_R(z~e^{-2\pi i})S^T~~~~~~\pi<\arg z<\pi+{\pi \over k} $$ In \cite{BJL1} $S$ is called a `` Stokes' multiplier''. The terminology in this field changes from one author to the other... \vskip 0.2 cm \noindent {\bf Definition: } We call {\it central connection matrix} the connection matrix $C$ such that $$ \tilde{Y}_0(z)=\tilde{Y}_{R}(z) C ~~~z\in \Pi_R $$ \vskip 0.2 cm It is clear that the system (\ref{10}) has solutions $Y_0(z)=X \tilde{Y}_0(z)$, and $Y_L(z)= X \tilde{Y}_L(z)$, $Y_R(z)= X \tilde{Y}_R(z)$ asymptotic to $X \tilde{Y}_F(z)$ as $z \to \infty$ in ${\cal S}_L$ and ${\cal S}_R$ respectively, which are connected by the same $S$ and $C$. \vskip 0.2 cm In order to compute the entries of $S$ explicitely, we use the reduction of (\ref{10}) to the generalized hypergeometric equation (\ref{30}). If $\varphi^{(1)}(z)$, ..., $\varphi^{(k)}(z)$ is a basis of $k$ linearly independent solutions of (\ref{30}), then the matrix $Y(z)$ of entries $(n,j)$ defined by \begin{equation} Y^{(j)}_n(z):= {1 \over k^{n-1}} ~z^{{k-1\over 2}-n+1}~ (z \partial_z)^{(n-1)}~\varphi^{(j)}(z) \label{entries} \end{equation} is a fundamental matrix for (\ref{10}). \vskip 0.3 cm \noindent {\bf Lemma 3:} {\it The generalized hypergeometric equation (\ref{30}) has two bases of linerly independent solutions $\varphi_{L/R}^{(1)}(z)$, ..., $\varphi_{L/R}^{(k)}(z)$ having asymptotic behaviours $$ \varphi_{L/R}^{(n)}={1 \over \sqrt{k}} {e^{i{\pi \over k}(n-1)} \over z^{k-1 \over 2}} ~ \exp\left[k~e^{i{2 \pi \over k}(n-1)} ~z \right]~\left(1 + O\left({1 \over z}\right) \right),~~~~~z \to \infty $$ in ${\cal S}_L$ and ${\cal S}_R$ respectively. Let $\Phi(z)$ denote the row vector $ [ \varphi^{(1)}(z),...,\varphi^{(k)}(z)]$. The fundamental matrices $Y_L(z)$, $Y_R(z)$ of (\ref{10}) are expressed through formula (\ref{entries}) in terms of $\Phi_L(z)$ and $\Phi_R(z)$ and $$ Y_L(z)=Y_R(z)~S~~~~~~~0<\arg z < {\pi \over k} $$ if and only if $$ \Phi_L(z)= \Phi_R(z) S ~~~~~~~0<\arg z < {\pi \over k} $$ } \vskip 0.2 cm \noindent {\bf Proof:} Simply observe that for a fundamental solution in ${\cal S}_L$ or ${\cal S}_R$ (we omit subscripts $L$, $R$) $$ Y(z)= \left[ \matrix{ z^{k-1\over 2} \varphi^{(1)}& ... & z^{k-1\over 2} \varphi^{(k)}\cr \vdots & ... & \vdots \cr } \right] = X \tilde{Y} $$ which is asymptotic, for $z \to \infty$, to $$ \sim \left[ \matrix{ 1 & e^{i {\pi \over k}}&e^{i {2\pi \over k}} & e^{i {3\pi \over k}}& ...& e^{i {k-1 \over k}\pi } \cr \vdots &\vdots &\vdots & \vdots & ... & \vdots \cr } \right] ~ \left[ \matrix{ \exp(kz)& & & \cr &\exp(k e^{2 \pi i \over k}z) & & \cr & & \ddots & \cr & & & \exp( k e^{2 \pi i (k-1)\over k}) \cr } \right] $$ Now , the first row of $Y(z)$ is $z^{k-1 \over 2} \Phi(z)$ \rightline{$\Box$} \vskip 0.2 cm The Stokes' matrix $S$ has entries $$ s_{ii}=1 $$ $$ s_{ij}=0~~~\hbox{ if } R_{ij}\subset \Pi_R $$ This follows from the fact that on the overlapping reagion $0<\arg z <{\pi \over k}$ there are no Stokes' rays and $$ e^{zU}~S~e^{-zU} \sim I, ~~~~z \to \infty,~~~~ \hbox{ then }~~ e^{z(u_i-u_j)}~s_{ij}\to \delta_{ij} $$ Moreover, $\Re \left(z(u_i-u_j)\right)>0$ to the left of the ray $R_{ij}$, while $\Re \left(z(u_i-u_j)\right)< 0$ to the right (the natural orientation on $R_{ij}$, from $z=0$ to $\infty$ is understood). This implies $$ \left|e^{zu_i} \right| > \left|e^{zu_j} \right|~~~~~ \hbox{ and } % e^{z(u_i-u_j)}\to \infty ~~\hbox{ as } z \to \infty $$ on the left, while on the right $$ \left|e^{zu_i} \right| < \left|e^{zu_j} \right|~~~~~ \hbox{ and } % e^{z(u_i-u_j)}\to 0 ~~\hbox{ as } z \to \infty $$ With this observations in mind, we prove the following \vskip 0.3 cm \noindent {\bf Lemma 4 :} {\it $S$ has a column whose entries are all zero but one. More precisely: \par \noindent For $k$ even $$ s_{i,{k \over 2}+1}=0~~~~\forall i \neq {k \over 2}+1 , ~~~~~~~ s_{{k \over 2}+1,{k \over 2}+1}=1 $$ For $k$ odd $$ s_{i,{k+1 \over 2}}=0~~~~\forall i \neq {k+1 \over 2} , ~~~~~~~ s_{{k +1\over 2},{k+1 \over 2}}=1 $$ } \vskip 0.2 cm \noindent {\bf Proof: } Let us determine $n$ such that $s_{in}=0$ for any $i \neq n$ and $s_{nn}=1$. We need to find all rays in $\Pi_R$. We start with $R_{rs}$ with $r<s$. We know that for $r+s=k+2$ the ray is the negative real half-line (at angle $-\pi$). Then $R_{rs} \subset \Pi_R$ for $r+s\leq k+1$ ($r<s$). Then, in $\Pi_R$ we have $$ \left. \matrix{ R_{12} & R_{13} & ... & R_{1k} \cr R_{23} & R_{24} & ... & R_{2, k-1} \cr R_{34} & R_{35} & ... & R_{3,k-2} \cr \vdots & \vdots & & \cr R_{ab} & & & \cr } \right. $$ where $ R_{ab} = R_{{k \over 2},{k \over 2} +1}$ for $k$ even, and % $ R_{{k-1 \over 2},{k+3\over 2}}$ for $k$ odd. In $\Pi_R$ we have also $R_{rs}$ with $r+s\geq k+2$ and $r>s$. For fixed $n$ we require $R_{in} \subset \Pi_R$ for any $i$. Namely, $$ \forall i<n ~~~i+n \leq k+1 ~,~~~~~~\forall i>n ~~~i+n\geq k+2 $$ This yields $n= {k \over 2}+1 $ for $k$ even, $n={k+1 \over 2}$ for $k$ odd. \rightline{$\Box$} \vskip 0.3 cm Let $n(k)$ be $ {k \over 2}+1 $, or ${k+1 \over 2}$. Lemma 4 implies that the $n(k)^{th}$ columns of ${Y}_L$ and ${Y}_R$ coincide. In particular, their asymptotic representation holds for $-\pi < \arg z < \pi +{\pi \over k}$. Actually, this domain can be further enlarged, up to $$ -{\pi \over k} -\pi < \arg z < \pi +{\pi \over k} ~~~~k \hbox{ even} $$ $$ -\pi < \arg z < \pi + {2\pi \over k} ~~~~k \hbox{ odd } $$ To see this recall that $\left|e^{zu_i} \right|< \left|e^{zu_j} \right|$ on the right of $R_{ij}$, and conversely on the left. Then it is easy to see that for $k$ even $\left|\exp ({z~u_{{k \over 2}+1}}) \right|$ dominates all exponentials in the sector $ -{\pi \over k} - \pi < \arg z < {\pi \over k} - \pi$, while for $k$ odd $\left|\exp ({z~u_{{k +1\over 2}}})\right|$ dominates all exponentials in the sector $ \pi < \arg z< \pi + {2 \pi \over k}$. The first entry of the $n(k)^{th}$ column is $\varphi^{(n(k))}_L(z)\equiv \varphi^{(n(k))}_R(z)$ times $z^{k-1 \over 2}$. Then $\varphi^{(n(k))}$ has the estabilished asymtotic behaviour on the enlarged domains above. \vskip 0.3 cm We now introduce an integral representation for a solution $\varphi(z)$ of the generalized hypergeometric equation which will allow us to compute the entries of $S$. \vskip 0.2 cm \noindent {\bf Lemma 5:} { \it The function $$ g^{(n)}(z)= {1 \over (2 \pi )^{k+1 \over 2} ~e^{i \pi ( {k \over 2}-n-1)}} ~ \int_{-c -i \infty}^{-c+i \infty}ds~\Gamma^k(-s)~e^{-i\pi k s} ~e^{i2(n-1)\pi s}~z^{ks} $$ defined for ${\pi \over 2} -2(n-1){\pi \over k}< \arg z < {3 \pi \over 2} - 2 (n-1) {\pi \over k} $, $z \neq 0$ and for any positive number $c>0$, is a solution of the generalized hypergeometric equation (\ref{30}) ( the path of integration is a vertical line through $-c$). It has asymptotic behaviour $$ g^{(n)}(z)\sim {1 \over \sqrt{k}}~{e^{i{\pi \over k}(n-1)} \over z^{k-1\over 2}}~ \exp \left(k e^{i{2 \pi \over k}(n-1)}z \right)~~~~~~z\to \infty $$ In particular, for $n(k)= {k\over 2}+1$ ($k$ even), or $n(k)={k+1 \over 2}$ ($k$ odd), the analytic continuation of $g^{(n(k))}(z)$ has the above asymptotic behaviour in the domains $$ - \pi - {\pi \over k} < \arg z< \pi + {\pi \over k}~~~~~k \hbox{ even } $$ $$ - \pi < \arg z < \pi + { 2 \pi \over k} ~~~~~k \hbox{ odd } $$ and coincides with the solution $\varphi^{(n(k))}_L\equiv \varphi^{(n(k))}_R$ appearing in the first rows of the fundamental matrices $Y_L$ and $Y_R$ of the system (\ref{10}). The following identity holds \begin{equation} \sum_{m=0}^k ~(-1)^{m-k}~\pmatrix{ k \cr m \cr} ~g^{(n)}(z~e^{i {2 \pi \over k}m}) =0 \label{60} \end{equation} } We'll sketch the proof in Appendix 1. \vskip 0.2 cm \noindent {\bf Remark 2:} Observe that for basic solutions of the hypergeometric equation $\Phi_{L/R}= [ \varphi^{(1)}_{L/R},..., \varphi^{(k)}_{L/R}]$, $$ \varphi^{(n)}(z )\sim ~{1 \over \sqrt{k}} {e^{i {\pi \over k} (n-1)}\over z^{k-1\over 2} }\exp( k e^{i {2\pi \over k} (n-1)} z) $$ on some sector, and $$ \varphi^{(n)}(z e ^{2 \pi i \over k})\sim (-1)~{1 \over \sqrt{k}} {e^{i {\pi \over k} ([n+1]-1)}\over z^{k-1\over 2} }\exp( k e^{i {2\pi \over k} ([n+1]-1)} z) $$ like $- \varphi^{(n+1)}$, on the sector rotated by ${- 2 \pi \over k}$. Note however that $\varphi^{(k)} (ze^{2 \pi i \over k}) \sim (\sqrt{k} z^{k-1\over 2} )^{-1}~e^{kz} $, like $\varphi^{(1)}(z)$. Also, note that $g^{(n)}(z e^{2\pi i \over k})= - g^{(n+1)}(z)$ (we mean analytic continuations), with asymptotic behaviour on rotated domain. \section{ Monodromy Data of the Quantum Cohomology of ${\bf CP}^{k-1}$} \vskip 0.2 cm Let us return to the system (\ref{10p}). \begin{equation} \partial_z y=\left(\hat{\cal U}(t^2) + {1 \over z} \hat{\mu} \right) y \label{tuono} \end{equation} In this section we use the original notation $w=ze^{t^2\over k}$. The system has a fundamental matrix $$ y_0(t^2,z)=(I+A_1(t^2)z+A_2(t^2)z^2+...)~z^{\hat{\mu}}~z^R,~~~~z\to 0 $$ where $R$ is the same of system (\ref{10}). The series appearing in the solutions converges near $z=0$. The matrix $\hat{\cal U}(t^2)$ has eigenvalues and eigenvectors $$ u_n(t^2)=e^{i{2\pi\over k}(n-1)}~e^{t^2 \over k} ~\equiv~u_n e^{t^2 \over k}~~~~~n=1,...,k$$ $$ {\bf x}_n(t^2) \hbox{ of entries } {x^j}_n(t^2)= e^{i(2j-1)(n-1){\pi \over k}}~e^{{2j-1-k \over 2k}t^2} ~\equiv {x^j}_n e^{{2j-1-k \over 2k}t^2}$$ Let $X(t^2)=({x^j}_n(t^2))$. With the gauge $y=X(t^2)~\tilde{y}(t^2,z)$ we obtain the equivalent system \begin{equation} \partial_z \tilde{y} = \left[ U(t^2)+{V(t^2)\over z}\right] \tilde{y} \label{tuono1} \end{equation} $$ U(t^2)=X^{-1}(t^2)~\hat{\cal U}(t^2)~ X(t^2)=\hbox{diag}(u_1(t^2),...,u_k(t^2)) $$ $$ V(t^2) =X^{-1}(t^2)~\hat{\mu} ~X(t^2)) $$ $$ V(t^2)^T+V(t^2)=0 $$ The skew symmetry of $V$ follows from $\eta\hat{\mu} +\hat{\mu}\eta=0$ and from the choice of the normalization of the eigenvectors, such that $X(t^2)^T~X(t^2)=\eta^{-1}$. Let us fix an initial point $t_0=(0,t_0^2,0,...,0)$. The system (\ref{tuono1}) has fundamental matrices $y_R(t_0^2,z)$, $y_L(t_0^2,z)$, which are asymptotic to the formal solution $$ \tilde{y}_F(t_0^2,z)=\left[ I+{F_1(t_0^2)\over z}+{F_2(t_0^2)\over z}+ ... \right]~e^{z~U(t_0^2)} $$ in the sectors $$ {\cal S}_L(t_0)=\{ z\in {\bf C}~|~ 0<\arg \left[z \exp\left({t_0^2\over k}\right)\right] < \pi + {\pi \over k} \} $$ $$ {\cal S}_R(t_0)=\{ z\in {\bf C}~|~ - \pi < \arg \left[z \exp\left( {t_0^2\over k}\right)\right] < {\pi \over k} \} $$ and $$ \tilde{y}_L(t_0^2,z)=y_R(t_0^2,z)~ S ~~~~~~~~~~~0<\arg \left[z \exp\left({t_0^2\over k}\right)\right]<{\pi \over k} $$ with respect to the admissible line $$ l_{t_0}:= \{z~|~z=\rho \exp \left({i \epsilon} -{\Im m~ t_0^2\over k} \right),~~\rho>0\} $$ The Stokes' matrix is precisely the matrix $S$ of system (\ref{10}) with respect to the admissible line $l_{t_0}$. Also the central connection matrix defined by $$ y_0(t_0^2,z)=y_R(t_0^2,z)~C ~~~~~~~~~~ - \pi <\arg \left[ z e^{t_0^2 \over k} \right] < {\pi \over k} $$ is the the same of the system (\ref{10}). \vskip 0.2 cm \noindent {\bf Definition: } $C$ and $S$, together with $\hat{\mu}$, $R$, and $e={\partial \over \partial t^1}$ are the {\it monodromy data} of the quantum cohomology of ${\bf CP}^{k-1}$ in a local chart containing $t_0$. \vskip 0.2 cm Recall that we fixed a point $t_0=(0,t_0^2,...,0)$. When we consider a point $t$ away from $t_0$, the system (\ref{tuono}) acquires the general form \begin{equation} \partial_z y = \left[ \hat{\cal U}(t) +{\hat{\mu}\over z} \right]y \label{zzzz1} \end{equation} where $\hat{\cal U}(t^1,..,t^k)= \eta~ {\cal U}(t)~ \eta^{-1}$ and $y_{\alpha}^{(j)}(t^1,..,t^k;z)=\partial_{\alpha} \tilde{t}^{j}(t,z)$. The admissible line $l_{t_0}$ must be considered {\it fixed once and for all}. Instead, the Stokes' rays change. This is because they are functions of the eigenvalues $u_1(t)$, ..., $u_k(t)$ of the matrix $\hat{\cal U}(t^1,..,t^k)$. For example, if just $t^2$ varies, while $t^1=t^3=...=t^k=0$, the system (\ref{tuono}) has Stokes' rays $$ R_{rs}(t^2)=\{z~|~=\rho~ \exp\left(i{2\pi \over k} -i{\pi \over k}(r+s) -i {\Im m~t^2 \over k} \right),~~\rho>0 \} $$ The dependence of the coefficients of the system (\ref{zzzz1}) on $t$ is isomonodromic \cite{Dub1} \cite{Dub2}. Then $\mu$ and $R$ are the same for any $t$. $S$ and $C$ do not change if we move in a sufficently small neighbourhood of $t_0$. Problems arise when some Stokes' rays cross $l_{t_0}$. $S$ and $C$ must be modified by an action of the braid group. We will return to this point later. \vskip 0.2 cm \section{ Computation of S } \vskip 0.2 cm To compute $S$, we factorize it in ``Stokes' factors''. Our fundamental matrix $Y_L$ has the required asymptotic form on the sector between $R_{k2} ~~(\arg z=0)$ and $ R_{1k}~~(\arg z= \pi +{\pi \over k})$. $Y_R$ has the same behaviour between $R_{2k}~~(\arg z= -\pi)$ and $R_{k1} ~~( \arg z = {\pi \over k})$. \begin{figure} \epsfxsize=9cm \centerline{ \epsffile{figr2.eps}} \caption{ Sector where $Y_{k2}$ has the asymptotic behaviour $X~e^{zU}$ for $z \to \infty$} \end{figure} Of course, we can consider fundamental matrices with the same asymptotic behaviour on other sectors of angular width less then $\pi +{\pi \over k}$ and bounded by two Stokes' rays. We introduce the following notation: consider a fundamental matrix of (\ref{10}) having the required asymptotic behaviour on such a sector. If we go all over the sector clockwise we meet Stokes rays belonging to the sector at each displacement of ${\pi \over k}$. Let $R_{ij}$ be the last ray we meet before reaching the boundary (the boundary is still a Stokes ray {\it not } belonging to the sector). Then we will call the fundamental matrix $Y_{ij}$. For example, $Y_L=Y_{k1}$ and $Y_R=Y_{1k}$. See figure 2. Sometimes, a different labelling is used in the literature. The rays must be enumerated as in figure 3. The numeration refers to the line $l$: the rays are labelled in counter-clockwise order starting from the first one in $\Pi_R$ (which will be $R_0$; then $R_0$, $R_1$, ..., $R_{k-1}$ are in $\Pi_R$, and $R_k$, ..., $R_{2k-1}$ are in $\Pi_L$). For our particular choiche of $l$ , $R_0\equiv R_{1k}$ (at $\arg z = -\pi + {\pi \over k}$); $R_1\equiv R_{1,k-1}$ follows counter-clockwise... Then we proceed untill we reach $R_{k-1}\equiv R_{k2}$ before crossing $l$, and so on. The fundamental matrices are labelled as we prescribed above, namely $Y_j$ if its sector contains $R_j$ as the last ray met going all over the sector clockwise before the boundary. The sector itself is denoted by ${\cal S}_j$. See figure 3. \begin{figure} \epsfxsize=9cm \centerline{ \epsffile{figr3.eps}} \caption{ Sector ${\cal S}_k$ and labelling of Stokes' rays} \end{figure} We define {\it Stokes' factors} to be the connection matrices $K_j$ such that $$ Y_{j+1} (z) = Y_j(z) K_j $$ on the overlapping region of width $\pi$. We warn the reader that also the Stokes' factors will be labelled with both conventions above, according to our convenience (for example $K_0 \equiv K_{1k}$). As a consequence of the above definitions, we can factorize $S$ as follows $$ Y_L= Y_{k1}=Y_{k2} K_{k2}= Y_{k3}K_{k3}K_{k2}=...$$ $$ ...= Y_{1k} K_{1k}K_{1,k-1}K_{k,k-1} K_{k,k-2}...K_{k3} K_{k2} \equiv Y_R S$$ Then \begin{equation} S= K_{1k}K_{1,k-1}K_{k,k-1} K_{k,k-2}...K_{k3} K_{k2} \label{facto} \end{equation} We observe that, being the first row of $Y(z)$ equat to $z^{k-1 \over 2} \Phi(z)$, the following holds: $$ \Phi_{j+1}(z) = \Phi_j K_j $$ \vskip 0.3 cm {\bf Remark 3:} The Stokes' factors of the system (\ref{10}) and of the gauge-equivalent system (\ref{20}) are the same. From the skew-symmetry of $V$ it follows that $K_{ji}=(K_{ij}^{-1})^T$. \vskip 0.3 cm Before computing the Stokes factors explicitely, we show that just two of them are enough to compute all the others. Let $$ F(z):= \left( {1 \over \sqrt{k}} {1 \over z^{k-1 \over 2} } \exp(k z),~ {1 \over \sqrt{k}} {e^{{i \pi \over k}} \over z^{k-1 \over 2} } \exp(k e^{2\pi i \over k} z),~ ...,~ {1 \over \sqrt{k}} {e^{{i \pi \over k}(k-1)} \over z^{k-1 \over 2} } \exp(k e^{{2\pi i \over k}(k-1)} z) \right) $$ $$ = (F^{(1)}(z),~F^{(2)}(z),~...,~F^{(k)}(z))$$ be the row vector whose entries are the first terms of the asymptotic expansions of an actual solution $\Phi(z)$ of the generalized hypergeometric equation. By a straightforward computation we see that $$ F(z e^{2\pi i \over k })= F(z) T_F~,~~~~ T_F=\pmatrix{ 0 & & & &...& 1 \cr -1& 0 & & && \cr & -1 & 0 & && \cr & & -1 & && \cr & & & \ddots&\ddots& \vdots\cr & & & & -1 &0\cr } $$ We use now the convention of enumeration of Stokes' rays $R_0$, $R_1$, ... starting from $l$ (see above). Let $\Phi_m(z)$ be an actual solution of the hypergeometric equation having asymptotic behaviour $F(z)$ on ${\cal S}_m$. $$ \Phi_m(z) \sim F(z) ~~~z \to \infty ~~~ z \in {\cal S}_m $$ Then $$ \Phi_{m+2}(ze^{2\pi i \over k})\sim F(z e^{2 \pi i \over k} )= F(z) T_F ~,~~~~ z \in {\cal S}_m $$ Namely, $$ \Phi_{m+2}(ze^{2\pi i \over k})T_F^{-1}\sim F(z) ~,~~~~~ z \in {\cal S}_m$$ then, by unicity of actual solutions having asymptotic behaviour $F(z)$ in a sector wider then $\pi$, we have $$ \Phi_{m+2}(ze^{2\pi i \over k})= \Phi_m(z) T_F~,~~~~~z \in {\cal S}_m $$ \vskip 0.2 cm \noindent {\bf Lemma 6:} {\it For any $p \in {\bf Z}$ $$ K_{m+2p}= T_F^{-p}K_m T_F^p$$ } {\bf Proof: } For $z \in {\cal S}_m \cap {\cal S}_{m+1}$ $$ \Phi_{m+1}(z) = \Phi_m(z)~ K_m = \Phi_{m+2}(ze^{2\pi i \over k})~T_F^{-1}K_m =$$ $$ = \Phi_{m+3}(ze^{2\pi i \over k})~K_{m+2}^{-1}T_F^{-1}K_m = \Phi_{m+1}(z) ~~T_F~ K_{m+2}^{-1}~T_F^{-1}~K_m $$ Then $K_{m+2}=~T_F^{-1}~K_m~T_F$. By induction we prove the lemma. \rightline{$\Box$} \vskip 0.3 cm >From the lemma it follows that just two Stokes' factors are enought to compute all the others. We are ready to give a concise formula for $S$: \vskip 0.3 cm \noindent {\bf Theorem 1: } {\it Let $l$ be an admissible line (i.e. not containing Stokes' rays), and let us enumerate the rays in counter-clockwise order starting from the first one in $\Pi_R$ (which will be $R_0$, then $R_0$, $R_1$, ..., $R_{k-1}$ are in $\Pi_R$, and $R_k$, ..., $R_{2k-1}$ are in $\Pi_L$). Then the Stokes' matrix for (\ref{10}), (\ref{20}), (\ref{30}) and $k>3$ is \vskip 0.2 cm \begin{equation} S = \left\{ \matrix{ \left(K_0~K_1~T_F^{-1} \right)^{k \over 2} ~T_F^{k\over 2} ~\equiv~ T_F^{k \over 2}~\left( T_F^{-1}~K_{k-2}~K_{k-1} \right)^{k \over 2} , ~~~k \hbox{ even } \cr \cr \cr \left(K_0~K_1~T_F^{-1} \right)^{k-1 \over 2} ~K_0~T_F^{k-1\over 2} ~\equiv~ T_F^{k-1 \over 2}~K_{k-1}\left( T_F^{-1}~K_{k-2}~K_{k-1} \right)^{k-1 \over 2} , ~~~k \hbox{ odd } \cr } \right. \label{S100} \end{equation} } \vskip 0.3 cm \noindent {\bf Proof: } $S=K_0~K_1~K_2~...~K_{k-1}$, $K_{2p}= T_F^{-p}~K_0~ T_F^p$ and $K_{2p+1}= T_F^{-p}~K_1~T_F^p$. Then $$ S=K_0 K_1 ~( T_F^{-1}K_0T_F)~( T_F^{-1}K_1T_F)~K_4 K_5 ~...~K_{k-1}$$ $$ = (K_0 K_1 T_F^{-1})~K_0K_1~T_F~K_4 K_5 ~...~ K_{k-1}$$ $$ = (K_0 K_1 T_F^{-1})~K_0K_1~T_F~( T_F^{-2} K_0 T_F^2)~(T_F^{-2} K_1 T_F^2)~K_6~...~K_{k-1} $$ $$ = (K_0 K_1 T_F^{-1})(K_0 K_1 T_F^{-1})~K_0 K_1 T_F^2~K_6~...~ K_{k-1} $$ Now observe that $K_{k-1}= T_F^{-({k \over 2}-1)}~K_1~T_F^{{k\over 2} -1}$ for $k$ even, while $K_{k-1}= T_F^{-({k-1 \over 2})}~K_0~T_F^{k-1\over 2}$ for $k$ odd. Then, for $k$ even $$ S = (K_0 K_1 T_F^{-1})^{{k \over 2} -2}~K_0K_1T_F^{{k \over 2}-2}~K_{k-2}K_{k-1}$$ $$ = (K_0 K_1 T_F^{-1})^{{k \over 2} -2}~K_0K_1T_F^{{k \over 2}-2} ~ T_F^{-({k \over 2}-1)}K_0T_F^{{k\over 2}-1}~ T_F^{-({k \over 2}-1)}K_1T_F^{{k\over 2}-1}= \left(K_0~K_1~T_F^{-1} \right)^{k \over 2} ~T_F^{k\over 2} $$ For $k$ odd: $$ S= (K_0 K_1 T_F^{-1})^{{k -3\over 2} }~ K_0 K_1 T_F^{k-3 \over 2} ~K_{k-1} $$ $$ = (K_0 K_1 T_F^{-1})^{{k -3\over 2} }~ K_0 K_1 T_F^{k-3 \over 2}~ T_F^{-({k-1 \over 2})}K_0T_F^{k-1\over 2}= \left(K_0~K_1~T_F^{-1} \right)^{k-1 \over 2} ~K_0~T_F^{k-1\over 2} $$ If instead we write the Stokes' factors in term of $K_{k-2}$ and $K_{k-1}$ we obtain the other two formulas in the same way. \rightline{$\Box$} \vskip 0.3 cm \noindent {\bf Remark 4: } For our particular choiche of $l$, $K_0\equiv K_{1k}$, $K_1\equiv K_{1,k-1}$, $K_{k-2}\equiv K_{k3}$ and $K_{k-1}\equiv K_{k2}$. For $k=3$ $$ S=T_FK_{32}~(T_F^{-1}K_{12}K_{32})=K_{13}K_{12}K_{32}$$ \vskip 0.3 cm It is now worth deriving some properties of the monodromy of $Y(z)$ (for (\ref{10})) and $\Phi(z)$ (for (\ref{30})), which will be usefull later. Consider $\Phi_m(z)$ with asymptotic behaviour $F(z)$ on ${\cal S}_m$. Then $$ \Phi_m(z)= \Phi_{m-2}(z)~K_{m-2}~K_{m-1} \equiv \Phi_m(ze^{2 \pi i \over k})~T_F^{-1} ~K_{m-2}~K_{m-1} $$ On the other hand $$ \Phi_m(z)= \Phi_{m+2}(z)~K_{m+1}^{-1}~K_{m}^{-1} \equiv \Phi_m(ze^{-{2 \pi i \over k}})~T_F ~K_{m+1}^{-1}~K_{m}^{-1} $$ This proves the following \vskip 0.3 cm \noindent {\bf Lemma 7: } {\it The basic solution $\Phi_m(z)$ of the generalized hypergeometric equation (\ref{30}) with asymptotic behaviour $F(z)$ on ${\cal S}_m$, satisfies the identity $$ \Phi_m(z e^{2 \pi i \over k})= \Phi_m(z) ~T_m $$ where $$ T_m:= ~K_{m-1}^{-1}~K_{m-2}^{-1}~T_F~=~T_F~K_{m+1}^{-1}~K_m^{-1} $$ } \vskip 0.3 cm \noindent {\bf Corollary 1: } { \it The monodromy (at $z=0$) of $\Phi_m(z)$ is $$ \Phi_m(z e^{2\pi i})=\Phi_m(z)~(T_m)^k $$ } \vskip 0.2 cm Now, for our particular choiche of the line $l$ and for $m= k$, $\Phi_m(z)=\Phi_L(z)$. For the solution $Y_L(z)$ of (\ref{10}), the relations $Y_R(z)=Y_L(z)~S^{-1}$ ($0<\arg z<{\pi \over k}$), $Y_L(z)=Y_R(z e^{-2\pi i})~S^{T}$ ($\pi <\arg z <\pi+ {\pi \over k}$) immediately imply $$ Y_L(ze^{2 \pi i })= Y_L(z)~S^{-1} ~S^T $$ Recall that the $(n,j)$-th entry of $Y(z)$ is $ Y_{n,j}(z)\equiv Y^{(j)}_n(z)= {1 \over k^{n-1}} ~z^{{k-1\over 2}-n+1}~ (z \partial_z)^{(n-1)}~\varphi^{(j)}(z)$, and observe that $(z e^{2\pi i})^{k-1 \over 2}= (-1)^{k-1} ~z^{k-1\over2}$ Then, from Corollary 1 we get the following: \vskip 0.2 cm \noindent {\bf Corollary 2: } { \it Let $T$ be the $k$-monodromy matrix of $\Phi_L$ (namely, $T\equiv T_k$ for our choiche of $l$). Then $$ T^k = (-1)^{k-1}~S^{-1}~S^T $$ } Our formula (\ref{S100}) allows us to easily compute $S$. The recipe is simply to take $K_{k3}$, $K_{k2}$ (which we are going to compute explicitely) and substitute them into \begin{equation} S= T_F^{k \over 2}~(T_F^{-1}~K_{k3}~K_{k2})^{k \over 2}=T_F^{k \over 2} ~T^{-{k \over 2}} \label{S150} \end{equation} for $k$ even, or into \begin{equation} S= T_F^{k -1\over 2}~K_{k2}~(T_F^{-1}~K_{k3}~K_{k2})^{k-1 \over 2}= T_F^{k -1\over 2}~ K_{k2}~T^{-{k-1 \over 2}} \label{S200} \end{equation} for $k$ odd. \vskip 0.3 cm {\bf Computation of Stokes' factors: } We need to distinguish between $k$ odd and even. In the following $g(z)$ will mean $g^{(n(k))}(z)$ ($n(k)={k\over 2}+1$ or ${k+1 \over 2}$ for $k$ even or odd respectively). \vskip 0.2 cm \noindent {\bf $k$ odd:} $$ g(z)= \varphi^{({k+1 \over 2})}_{L}(z)\equiv \varphi^{({k+1 \over 2})}_{R}(z)$$ with asymptotic behaviour on $$ - \pi < \arg z < \pi +{2 \pi \over k} $$ If we iterate the map $ z \mapsto z e^{ {2\pi i \over k}} $ for $m=1, 2, ..., {k+1 \over 2}$ times, the domain of $g(z e^{{2\pi i\over k} m})$ for each $m$ covers ${\cal S}_R$. When we reach $m= {k+1 \over 2}$ a new iteration (i.e. a new rotation of the domain of $-{2 \pi \over k}$) will live the sector $ - {\pi \over k} < \arg z < {\pi \over k} $ of ${\cal S}_R$ uncovered. The same, if we do $ z \mapsto z e^{ -{2\pi i \over k}} $ the sector $-\pi< \arg z< -\pi +{2 \pi \over k}$ of ${\cal S}_R$ remains uncovered. Then, by remark 2: $$ \Phi_{1k}(z)\equiv \Phi_R(z)= \Big((-1)^{k-1 \over 2} g(z e^{i{2 \pi \over k}\left( {k+1 \over 2}\right)}), ~...\hbox{ ${k-3 \over 2}$ unknown terms }...~, $$ $$ g(z),~-g(z e^{2 \pi i \over k}), ~g(z e^{4 \pi i\over k}),~...~,~(-1)^{k-1 \over 2} g(z e^{i{2 \pi \over k}\left( {k-1 \over 2}\right)}) \Big) $$ In the same way we see that $$ \Phi_{k1}(z)\equiv \Phi_L(z)=\Big((-1)^{k-1 \over 2} g(z e^{-i{2 \pi \over k} \left( {k-1 \over 2}\right)}),~-(-1)^{k-1 \over 2} g(z e^{-i{2 \pi \over k}\left( {k-3 \over 2}\right)}),~...~,-g(ze^{-{2 \pi i \over k }}), $$ $$ g(z),~...\hbox{ ${k-1 \over 2}$ unknown terms }... \Big) $$ and similar expressions for $\Phi_{1,k-1}$, $\Phi_{k,k-1}$, $\Phi_{k,k-2}$, ..., $\Phi_{k3}$,$\Phi_{k2}$. The unknown terms are computed using the identity (\ref{60}) and simple considerations on the dominant exponentials $|e^{zu_i}|$ on the sectors which remain uncovered in the iterations of $ z \mapsto z e^{\pm i{2 \pi \over k}}$. \vskip 0.2 cm \noindent {\bf Example:} A simple example will clarify this procedure. Let $k=7$; then $g=\varphi^{(4)}$, $$ \Phi_{17}=\Phi_R= ( -g(ze^{8 \pi i \over 7}),~?,~?,~g(z),~ -g(ze^{2\pi i \over 7}),~ g(ze^{4 \pi i \over 7}),~ -g(ze^{6 \pi i \over 7})) $$ We look for $\varphi^{(3)}_R$. If we take $(-1)g(z e^{-2 \pi i \over k})$ we miss to cover $-\pi <\arg z< - \pi + {2 \pi \over 7}$ in ${\cal S}_R$. On $-\pi <\arg z< - \pi + { \pi \over 7}$ (between two nearby Stokes' rays) we have the relations $|e^{zu_4}|>|e^{zu_5}|>|e^{z u_3}|$. On $-\pi+{ \pi \over 7} <\arg z< - \pi + {2 \pi \over 7}$, $|e^{zu_4}|>|e^{zu_3}|$ (later on we will simply write $4>3$). Then $$ \varphi^{(3)}_R(z)= (-1)g(z e^{-2 \pi i \over 7}) +c_4 \varphi^{(4)}_R(z) +c_5\varphi^{(5)}_R(z) $$ To find $c_4$, $c_5$ we need another representation for $\varphi^{(3)}_R$. We consider $ (-1)g(z e^{12 \pi i \over 7})$, which has the correct asymptotic behaviour, but on a domain which leaves uncovered $ -{3 \pi \over 7}< \arg z < {\pi \over 7}$. The relations are: on $ 0< \arg z < {\pi \over 7}$, $1>7>2>6>3$; on $ -{ \pi \over 7}< \arg z < 0$, $1>2>7>3$; on $ -{2 \pi \over 7}< \arg z <- {\pi \over 7}$, $2>1>3$; on $ -{3 \pi \over 7}< \arg z < -{2\pi \over 7}$, $2>3$. Then $$ \varphi^{(3)}_R(z)= (-1)g(z e^{12 \pi i \over 7})+ d_1 \varphi^{(1)}_R(z)+ d_2 \varphi^{(2)}_R(z)+ d_6 \varphi^{(6)}_R(z)+ d_7 \varphi^{(7)}_R(z) $$ In the same way one finds $$ \varphi^{(2)}_R(z)= \left\{ \matrix{ g(z e^{-4 \pi i \over 7}) + a_3 \varphi^{(3)}_R(z)+ a_4 \varphi^{(4)}_R(z) + a_5 \varphi^{(5)}_R(z)+ a_6 \varphi^{(6)}_R(z) \cr g(z e^{10 \pi i \over 7})+ b_1\varphi^{(1)}_R(z)+b_2\varphi^{(4)}_R(z) \cr } \right. $$ $ \varphi^{(i)}$'s are known for $i=1,4,5,6,7$. Using the identity (\ref{60}) we compute $a$, $b$, $c$, $d$. We get $$ \varphi^{(2)}_R(z)= g(z e^{-4 \pi i \over 7}) -\bin{7}{1}g(z e^{-2 \pi i \over 7}) +\bin{7}{2}g(z)-\bin{7}{3}g(z e^{2 \pi i \over 7}) +\bin{7}{4} g(z e^{4 \pi i \over 7}) $$ $$ \varphi^{(3)}_R(z)=-g(z e^{-2 \pi i \over7}) + \bin{7}{1} g(z )- \bin{7}{2} g(z e^{2 \pi i \over7}) $$ A similar computation gives $\Phi_{71}=\Phi_L$. {\small $$ \Phi_{71}^T= \left[ \matrix{ -g(ze^{-{6\pi i \over 7}}) \cr g(ze^{-{4\pi i \over 7}}) \cr -g(ze^{{-2\pi i \over 7}}) \cr g(z) \cr -g(ze^{2\pi i \over 7})+\bin{7}{6} g(z) \cr g(ze^{4\pi i \over 7}) -\bin{7}{6} g(ze^{2\pi i \over 7})+\bin{7}{5}g(z) -\bin{7}{4}g(ze^{-{2 \pi i \over 7}}) \cr - g(ze^{{6 \pi i \over 7}}) +\bin{7}{6}g(ze^{{4 \pi i \over 7}}) -\bin{7}{5} g(ze^{{2 \pi i \over 7}})+\bin{7}{4}g(z) -\bin{7}{3} g(ze^{-{2 \pi i \over 7}})+ \bin{7}{2} g(ze^{-{4 \pi i \over 7}})\cr }\right] $$ } and {\small $$\Phi_{72}^T= \left[ \matrix{ -g(ze^{-{6\pi i \over 7}}) \cr g(ze^{-{4\pi i \over 7}}) \cr -g(ze^{{-2\pi i \over 7}}) \cr g(z) \cr -g(ze^{2\pi i \over 7})\cr g(ze^{4\pi i \over 7}) -\bin{7}{6} g(ze^{2\pi i \over 7})+\bin{7}{5}g(z) \cr - g(ze^{{6 \pi i \over 7}}) +\bin{7}{6}g(ze^{{4 \pi i \over 7}}) -\bin{7}{5} g(ze^{{2 \pi i \over 7}})+\bin{7}{4}g(z) -\bin{7}{3} g(ze^{-{2 \pi i \over 7}})\cr } \right] $$} Notice that in each of the last three entries of $\Phi_{72}$ there is a term missing w.r.t the corresponding entries of $\Phi_{71}$. This immediately implies {\tiny $$ \hbox{\small $K_{72}$} = \pmatrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 1 & 0 & 0 & 0 & 0 & \bin{7}{2} \cr 0 & 0 & 1 & 0 & 0 & \bin{7}{4}& 0 \cr 0 & 0 & 0 & 1 & \bin{7}{6} & 0 & 0 \cr 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } $$} The next step is the computation of $\Phi_{73}$ and $K_{73}$, through $\Phi_{72}=\Phi_{73}K_{73}$. It is done in the same way... \vskip 0.2 cm The above procedure is extended to the general case. In Appendix 2 we give, for example, the general expressions of $\Phi_R$ and $\Phi_L$. The factors of interest are: \vskip 0.3 cm {\bf $k$ odd:} \vskip 0.2 cm \noindent $(K_{k2})_{j,~k-j+2}=\bin{k}{2(j-1)}$ for $j=2,...,{k+1\over 2}$. $(K_{k2})_{j,j}=1$ for $j=1,...,k$. All the other entries are zero. \vskip 0.2 cm \noindent $(K_{k3})_{2,1}=-\bin{k}{1}$; $(K_{k3})_{j,~k-j+3}=\bin{k}{2j-3}$ for $j=3,...,{k+1 \over 2}$. $(K_{k3})_{j,j}=1$ for $j=1,...,k$. All the other entries are zero. Namely: \vskip 0.3 cm {\tiny $$ \hbox{\small $K_{k2}$}=\pmatrix{ 1& &&&&&&&& &0 \cr & 1&&&&&&&& &\bin{k}{2}\cr &&1&&&&&&&\bin{k}{4}&\cr &&&1&&&&&\bin{k}{6}&&\cr &&&&\ddots&&&\cdot^{\displaystyle{~\cdot}^{\displaystyle{~\cdot}} }&&&\cr &&&&&1&\bin{k}{k-1}&&&& \cr \cr &&&&&&1&&&\cr \cr & &&&&&&\ddots&&\cr & &&&&&&&1&\cr & &&&&&&&&1\cr & &&&&&&&&&1\cr} $$ $$ \hbox{\small $K_{k3}$}=\pmatrix{ 1&&&&&&&&&&&&0 \cr \cr -\bin{k}{1}&1&&&&&&&&&& &0\cr &&1&&&&&&& &&&\bin{k}{3}\cr &&&1&&&&&&&&\bin{k}{5}&\cr &&&&1&&&&&&\bin{k}{7}&&\cr &&&&&\ddots&&&&\cdot^{\displaystyle{~\cdot}^{\displaystyle{~\cdot}} }&&&\cr &&&&&&1&0&\bin{k}{k-2}&&&& \cr \cr &&&&&&&1&0 \cr \cr &&&&&&&&1\cr \cr & && &&&&&&\ddots&&\cr & && &&&&&&&1&\cr & &&&&&&&&&&1\cr & &&&&&&&&&&&1\cr} $$ } \vskip 0.3 cm {\bf $k$ even: } \vskip 0.2 cm \noindent $(K_{k2})_{j,~k-j+2}=\bin{k}{2(j-1)}$ for $j=2,...,{k\over 2}$. $(K_{k2})_{j,j}=1$ for $j=1,...,k$. All the other entries are zero. \vskip 0.2 cm \noindent $(K_{k3})_{2,1}=-\bin{k}{1}$; $(K_{k3})_{j,~k-j+3}=\bin{k}{2j-3}$, for $j=3,...,{k \over 2}+1$. $(K_{k3})_{j,j}=1$ for $j=1,...,k$. All the other entries are zero. Namely \vskip 0.3 cm {\tiny $$ \hbox{\small $K_{k2}$}=\pmatrix{ 1&&&&&&&&&&&0 \cr \cr &1&&&&&&& &&&\bin{k}{2}\cr &&1&&&&&&&&\bin{k}{4}&\cr &&&1&&&&&&\bin{k}{6}&&\cr &&&&\ddots&&&&\cdot^{\displaystyle{~\cdot}^{\displaystyle{~\cdot}} }&&&\cr &&&&&1&0&\bin{k}{k-2}&&&& \cr \cr &&&&&&1&0 \cr \cr &&&&&&&1\cr \cr && &&&&&&\ddots&&\cr && &&&&&&&1&\cr &&&&&&&&&&1\cr &&&&&&&&&&&1\cr} $$ $$ \hbox{\small $K_{k3}$}=\pmatrix{ 1& &&&&&&&& &&0 \cr \cr -\bin{k}{1}&1& &&&&&&&& &0\cr && 1&&&&&&&& &\bin{k}{3}\cr &&&1&&&&&&&\bin{k}{5}&\cr &&&&1&&&&&\bin{k}{7}&&\cr &&&&&\ddots&&&\cdot^{\displaystyle{~\cdot}^{\displaystyle{~\cdot}} }&&&\cr &&&&&&1&\bin{k}{k-1}&&&& \cr \cr &&&&&&&1&&&\cr \cr & &&&&&&&\ddots&&\cr & & &&&&&&&1&\cr & & &&&&&&&&1\cr & &&&&&&&&&&1\cr} $$ } \vskip 0.3 cm \section{ Reduction of $S$ to ``Canonical'' Form } Some examples of computations of $S$ are in Appendix 2. The reader may observe that $S$ is not in a nice upper triangular form (see also Lemma 4) and quite strange numbers (complicated combinations of sum and products of binomial coefficients) appear. Some natural operations are allowed on the Stokes matrices of a Frobenius manifold: \vskip 0.2 cm a) Permutations. Let us consider the system $$ { d \tilde{Y} \over dz}= \left[ U + {1 \over z} V \right]~\tilde{Y} $$ where $U = $ diag$( u_1,~u_2,~...,~u_k)$. Let $\sigma:~(1,2,..,k)$ $\mapsto$ $(\sigma(1),\sigma(2),...,\sigma(k))$ be a permutation. It is represented by an invertible matrix $P$ such that $$ P~U~P^{-1}~=~\hbox{ diag}(u_{\sigma(1)},u_{\sigma(2)},...,u_{\sigma(k)})$$ The system $$ {d {\cal Y} \over dz}= \left[PUP^{-1}+{1 \over z} PVP^{-1} \right] ~{\cal Y} $$ has solutions $${\cal Y}_{L/R}(z) := P~\tilde{Y}_{L/R}(z)~P^{-1} \sim ( I+O({1\over z}))~e^{z PUP^{-1}}~~~~{\cal Y}_{L}(z)={\cal Y}_{R}(z) ~PSP^{-1}$$ $$ {\cal Y}_0(z):= P \tilde{Y}_0(z)=(PX^{-1}+O(z))z^{\mu}z^R,~~~~{\cal Y}_0(z)= {\cal Y}_R(z) ~PC $$ $S$ and $C$ are then transformed in $PSP^{-1}$ and $PC$. For a suitable $P$, $PSP^{-1}$ is upper triangular. As a general result \cite{BJL1}, the good permutation is the one which put $u_{\sigma(1)}$, ..., $u_{\sigma(k)}$ in lexicographical ordering w.r.t. the oriented line $l$. The effect of the permutation $P$ corresponds to a change of coordinates in the given chart, consisting in the permutation $\sigma$ of the coordinates. \vskip 0.2 cm b) Sign changes of the entries. The construction at point a) is repeated, but now a diagonal matrix ${\cal I}$ with $1$'s or $-1$'s on the diagonal takes the place of $P$. In ${\cal I} S {\cal I}^{-1}$ some entries change sign. Note that ${\cal I}U{\cal I}^{-1}\equiv U$. Moreover, the formulae \cite{Dub1} \cite{Dub2} which define a local chart of the manifold in terms of monodromy data are not affected by $S\mapsto {\cal I}S{\cal I}$, $C \mapsto {\cal I} C$. \vskip 0.2 cm c) Action of the braid group. \noindent We first recall that the braid group is generated by $k-1$ elementary braids $\beta_{12}$, $\beta_{23}$, ..., $\beta_{k-1,k}$, with relations: $$ \beta_{i,i+1}\beta_{j,j+1}=\beta_{j,j+1}\beta_{i,i+1}~\hbox{ for } i+1 \neq j,~j+1\neq i$$ $$ \beta_{i,i+1}\beta_{i+1,i+2}\beta_{i,i+1}= \beta_{i+1,i+2}\beta_{i,i+1} \beta_{i+1,i+2} $$ This abstract group is realized as the fundamental group of (${\bf C}^k \backslash$diagonals)$/{\cal S}_k$ $:= \{(u_1,...,u_k)~|~u_i\neq u_j \hbox{ for } i\neq j \}/{\cal S}_k$, where ${\cal S}_k$ is the symmetric group of order $k$. In section 4 we proved that the Stokes' rays of the systems (\ref{tuono}) (\ref{tuono1}), and more generally the rays of the system (\ref{zzzz1}), depend on the point $t$ of the manifold. This is equivalent to the fact that the eigenvalues $u_1(t)$, ..., $u_k(t)$ of $\hat{\cal U}(t)$ depend on $t$. When $t$ changes, $u_1(t)$, ..., $u_k(t)$ change their position in the complex plane. Consequently, Stokes' rays move. Let us start from the point $t_0=(0,t_0^2,...,0)$. If we move in a sufficently small neighbourhood of $t_0$, the rays slightly change their positions. But if we move sufficently far away from $t_0$, some Stokes' rays cross the {\it fixed} admissible line $l_{t_0}$. Then, we must change ``Left'' and ``Right'' solutions of (\ref{zzzz1}). Then also $S$ and $C$ change. This is the reason why the monodromy data $S$ and $C$ change when we go from one local chart to another. The motions of the points $u_1(t)$, ..., $u_k(t)$ as $t$ changes represent transformations of the braid group. Actually, a braid $\beta_{i,i+1}$ can be represented as an ``elementary'' deformation consisting of a permutation of $u_i$, $u_{i+1}$ moving counter-clockwise (clockwise or counter-clockwise is a matter of convention). Suppose $u_1$, ..., $u_k$ are already in lexicographical order w.r.t. $l$, so that $S$ is upper triangular (recall that this configuration can be reached by a suitable permutation $P$). The effect on $S$ of the deformation of $u_i$, $u_{i+1}$ representing $\beta_{i,i+1}$ is the following : $$ S \mapsto~S^{\beta_{i,i+1}}:=A^{\beta_{i,i+1}}(S)~S~A^{\beta_{i,i+1}}(S) $$ where $$ \left(A^{\beta_{i,i+1}}(S) \right)_{nn}= 1~~~~~~n=1,...,k~~~n\neq~i,~i+1 $$ $$ \left(A^{\beta_{i,i+1}}(S) \right)_{i+1,i+1}=-s_{i,i+1} $$ $$ \left(A^{\beta_{i,i+1}}(S) \right)_{i,i+1}= \left(A^{\beta_{i,i+1}}(S) \right)_{i+1,i}=1 $$ and all the other entries are zero. For the inverse braid $\beta_{i,i+1}^{-1}$ ($u_i$ and $u_{i+1}$ move clockwise) the representation is $$ \left(A^{\beta_{i,i+1}^{-1}}(S) \right)_{nn}= 1~~~~~~n=1,...,k~~~n\neq~i,~i+1 $$ $$ \left(A^{\beta_{i,i+1}^{-1}}(S) \right)_{i,i}=-s_{i,i+1} $$ $$ \left(A^{\beta_{i,i+1}^{-1}}(S) \right)_{i,i+1}= \left(A^{\beta_{i,i+1}^{-1}}(S) \right)_{i+1,i}=1 $$ and all the other entries are zero. For a generic braid $\beta$, which is a product of $N$ elementary braids (for some $N$) $\beta=\beta_{j_1,j_1+1}...\beta_{j_N,j_N+1}$, the action is $$ S\mapsto S^{\beta}= A^{\beta}(S)~S~\left[A^{\beta}(S)\right]^T $$ where $$ A^{\beta}(S)= A^{\beta_{i_N,i_N+1}}(S^{ \beta_{j_1,j_1+1}...\beta_{j_{N-1},j_{N-1}+1} }) ~...~ A^{\beta_{i_2,i_2+1}}(S^{\beta_{i_1,i_1+1}}) A^{\beta_{i_1,i_1+1}}(S) $$ We remark that $S^{\beta}$ is still upper triangular. \vskip 0.3 cm In figure 4 we have drawn some lines $L_j=\{ \lambda= u_j + \rho e^{i({\pi \over 2} -\epsilon)},~~\rho>0\}$ ($0<\epsilon<{\pi \over k}$ is the angle of $l$), which help us to visualize the topological effect of the braids action ( they are the branch cuts for the fuchsian system which will be introduced in section 8). We are going to prove that the braid whose effect is to set the deformed points in cyclic order and the cuts in the configuration of figure 4 (namely, the last two cuts remain unchanged, the others are alternatively ``inverted''), brings $S$ in a {\it canonical form}: $s_{i,i+1}=\bin{k}{1}$ $\forall i=1,...,k-2$; $ s_{i,i+2}=\bin{k}{2}$ $\forall i=1,...,k-3$; ...; $ s_{i,i+n}=\bin{k}{n}$ $\forall i=1,...,k-n+1$; $ s_{i,k}=-\bin{k}{k-i}$. Namely {\small \begin{equation} S^{\beta}= \pmatrix{ 1&\bin{k}{1}&\bin{k}{2}&\bin{k}{3}&\bin{k}{4}& ...& -\bin{k}{k-1}\cr & 1 & \bin{k}{1} & \bin{k}{2} & \bin{k}{3}& ... & -\bin{k}{k-2}\cr & & 1 &\bin{k}{1} & \bin{k}{2} & ... & - * \cr & & & 1 & \bin{k}{1}&... & - * \cr &&&&&& \cr & & & & 1 & & - * \cr &&&&&\ddots & \cr &&&&&&1\cr } \label{SSCAN} \end{equation} } Note that the last column is negative. Its sign is inverted by $S \mapsto {\cal I} S {\cal I}$, where ${\cal I}:=$ diag$(1,1,...,-1)$. \begin{figure} \epsfxsize=14cm \centerline{ \epsffile{figr4.eps}} \caption{ Effect of the braid which brings $S$ to the canonical form on the lines $L_j$} \end{figure} \begin{figure} \epsfxsize=15cm \epsffile{figr5.eps} \caption{ Effect of the sequences of braids which bring $S$ to the canonical form (the figure refers to $k$ even)} \end{figure} \vskip 0.2 cm \noindent {\bf Lemma 8:} {\it Let the points $u_j$ ($j=1,..,k$) be in lexicographical order w.r.t the admissible line $l$. Then the braid $$ \beta:= (\beta_{k-5,k-4}~\beta_{k-6,k-5}~...~\beta_{12})~(\beta_{k-6,k-5}~\beta_{k-7,k-6} ~... ~\beta_{23})~(\beta_{k-7,k-6}~...~ \beta_{34})~...$$ $$...~\beta_{{k \over 2}-2,{k \over 2}-1}~ (\beta_{k-3,k-2}~\beta_{k-4,k-3}~...~\beta_{12})$$ for $k$ even, or $$ \beta:= (\beta_{k-5,k-4}~\beta_{k-6,k-5}~...~\beta_{12})~(\beta_{k-6,k-5}~\beta_{k-7,k-6} ~...~ \beta_{23})~(\beta_{k-7,k-6}~...~ \beta_{34})~...$$ $$...~(\beta_{{k-3 \over 2},{k-1 \over 2}}~\beta_{{k-5 \over 2},{k-3 \over 2}})~ (\beta_{k-3,k-2}~\beta_{k-4,k-3}~...~\beta_{12})$$ for $k$ odd, brings the points in cyclic counter-clockwise order, $u_1$ being the first point in $\Pi_L$ (figure 4, right side, or figure 5). } \vskip 0.2 cm Note that we have collected the braids in ${k\over 2}-1$ ($k$ even), or ${k-3 \over 2}$ ($k$ odd) sequences $(...)$. \vskip 0.2 cm \noindent {\bf Proof: } Let $k$ be even. The first braid $\beta_{k-5,k-4}$ iterchanges $u_{k-4}$ and $u_{k-5}$. The second braid interchanges $u_{k-5}$ and $u_{k-6}$. One easily sees that the effect of the first sequence of braids $(\beta_{k-5,k-4}~\beta_{k-6,k-5}~...~\beta_{12})$ is to bring $u_1$ in the (old) position of $u_{k-4}$, $u_{k-4}$ in the position of $u_{k-5}$, $u_{k-5}$ in the position of $u_{k-6}$, ..., $u_4$ in the position of $u_3$ and $u_2$ in the position of $u_1$ (figure 5). $u_{k}$, $u_{k-1}$, $u_{k-2}$, $u_{k-3}$ are not moved. The second sequence of braids $(\beta_{k-6,k-5}~\beta_{k-7,k-6} ~... ~\beta_{23})$ acts in a similar way, bringing $u_2$ in $u_{k-5}$, $u_{k-5}$ in $u_{k-6}$, ..., $u_3$ in $u_2$. $u_{k}$, $u_{k-1}$, $u_{k-2}$, $u_{k-3}$, $u_{k-4}$ are not moved. We go on in this way. After the action of $$(\beta_{k-5,k-4}~\beta_{k-6,k-5}~...~\beta_{12})~(\beta_{k-6,k-5}~\beta_{k-7,k-6} ~... ~\beta_{23})~(\beta_{k-7,k-6}~...~ \beta_{34})~...~\beta_{{k \over 2}-2,{k \over 2}-1}~$$ the points are as in figure 5: $u_k$ is on the positive real axis, $u_{k-2}$ is the first point met in counter-clockwise order, $u_1$ is the second, $u_2$ is the third; the points are in cyclic order up to $u_{k-3}$; finally, $u_{k-1}$ is the last point before reaching again the positive real axis from below. Then, $ (\beta_{k-3,k-2}~\beta_{k-4,k-3}~...~\beta_{12})$ brings $u_1$ in $u_{k-2}$, $u_{k-2}$ in $u_{k-3}$, $u_{k-3}$ in $u_{k-4}$, and so on. The cyclic order is reached. For $k$ odd the proof is similar. \rightline{$\Box$} \vskip 0.2 cm A careful consideration of the effect of the braid $\beta$ on the lines $L_j$ (which we leave as an exercise for the reader) shows that they are alternatively inverted as in figure 4. To reconstruct uniquely this configuration we just need to know the oriented line $l$, namely, its angle $\epsilon$ w.r.t the positive real axis. The points $u_{k-1}$, $u_k$ and the lines $L_{k-1}$ and $L_k$ are unchanged (angle ${\pi \over 2}-\epsilon$). The line at $u_1$ starts in the opposite direction, it goes around $u_2,...,u_{k-2}$ without intesecting other cuts, and then goes to $\infty$ with the original asymptotic direction ${\pi \over 2}-\epsilon$. Moving in the direction opposite to that of $l$ we meet $u_{k-2}$. Its line has the original direction ${\pi \over 2} -\epsilon$. Then we meet $u_2$, and the corresponding line starts with opposite direction, goes arouns $u_3,...,u_{k-3}$ and then goes to $\infty$ with asymptotic direction ${\pi \over 2}-\epsilon$. And so on. \vskip 0.2 cm Now we find the matrix representation for the braid $\beta$. \vskip 0.2 cm \noindent {\bf Proposition 1:} { \it The braid $\beta $ of Lemma 8 has the following matrix representation: {\tiny $$ \hbox{ \small $A^{\beta}(S)$} = \pmatrix{0&0&0&0&0&&&&0&0&0&0& 1&0 &0 \cr &&&&&&&&0&0& 1&0& \bin{k}{1}& 0&0\cr &&&&&&&&1&0&\bin{k}{1}&0&\bin{k}{2}&0& 0\cr &&&&&&&&\bin{k}{1}&0&\bin{k}{2} & 0&\bin{k}{3} &0& 0\cr &&&&&&&&*&.&*&.&*&.&. \cr &&&&&&&... &*&.&*&.&*&.&. \cr &&&&&&&&*&.&*&.&*&.&. \cr &&&&&&&&*&.&*&.&*&.&. \cr 0 &0&0&1&0&&&&*&.&*&.&*&.&. \cr 0 &1&0&\bin{k}{1}&0&&&...&*&.&*&.&*&.&. \cr 1&\bin{k}{1}&0&\bin{k}{2}&0&&&...&*&.&*&.&*&.&. \cr 0 &0&1&\bin{k}{3}&0&&&&*&.&*&.&*&.&. \cr &&&&&&&&&&&& \cr 0 &0&0&0&1&&&&*&.&*&.&*&.&. \cr &&&&&&&&*&.&*&.&*&.&. \cr .&.&.&.&.&&&... &*&.&*&.&*&.&. \cr .&.&.&.&.&&&&*&.&*&.&*&.&. \cr .&.&.&.&.&&&&*&.&*&.&*&.&. \cr &&&&&&&&\bin{k}{k-7}&0&\bin{k}{k-6}&0&\bin{k}{k-5}&0&0 \cr &&&&&&&&0&1&\bin{k}{k-5}&0&\bin{k}{k-4}&0&0 \cr &&&&&&&&0&0&0&1&\bin{k}{k-3}&0&0 \cr &&&&&&&&&&&& \cr &&&&&&&&0&0&0&0&0&1&0\cr 0&0&0&0&0&&&&0&0&0&0&0&0&1\cr } $$ } for $k$ even. {\tiny $$ \hbox{\small $A^{\beta}(S)$}= \pmatrix{0&0&0&0&0&&&&0&0&0&0& 1&0 &0 \cr &&&&&&&&0&0& 1&0& \bin{k}{1}& 0&0\cr &&&&&&&&1&0&\bin{k}{1}&0&\bin{k}{2}&0& 0\cr &&&&&&&&\bin{k}{1}&0&\bin{k}{2} & 0&\bin{k}{3} &0& 0\cr &&&&&&&&*&.&*&.&*&.&. \cr &&&&&&&... &*&.&*&.&*&.&. \cr &&&&&&&&*&.&*&.&*&.&. \cr &&&&&&&&*&.&*&.&*&.&. \cr 0 &0&0&0&1&0&&&*&.&*&.&*&.&. \cr 0&0 &1&0&\bin{k}{1}&0&&...&*&.&*&.&*&.&. \cr 1&0&\bin{k}{1}&0&\bin{k}{2}&0&&...&*&.&*&.&*&.&. \cr 0 &1&\bin{k}{2}&0&\bin{k}{3}&0&&&*&.&*&.&*&.&. \cr 0 &0&0&1&\bin{k}{4}&0&&&*&.&*&.&*&.&. \cr &&&&&&&&&&&&&\cr 0 &0&0&0&0&1&&&*&.&*&.&*&.&. \cr .&.&.&.&.&&&... &*&.&*&.&*&.&. \cr .&.&.&.&.&&&&*&.&*&.&*&.&. \cr .&.&.&.&.&&&&*&.&*&.&*&.&. \cr &&&&&&&&\bin{k}{k-7}&0&\bin{k}{k-6}&0&\bin{k}{k-5}&0&0 \cr &&&&&&&&0&1&\bin{k}{k-5}&0&\bin{k}{k-4}&0&0 \cr &&&&&&&&0&0&0&1&\bin{k}{k-3}&0&0 \cr &&&&&&&&&&&&&\cr &&&&&&&&0&0&0&0&0&1&0\cr 0&0&0&0&0&&&&0&0&0&0&0&0&1\cr } $$ } for $k$ odd. } The $``*''$ means $\bin{k}{j}$, and $j$ increases by one when we move downwards row by row. \vskip 0.3 cm \noindent {\bf Proof:} We need some steps. \vskip 0.2 cm 1) For an upper triangular Stokes' matrix $S$ (with entries $s_{ij}$) the entries of the matrix $A^{\beta}$ which are different from zero are the following: \vskip 0.2 cm \noindent {\bf $k$ odd:} \vskip 0.2 cm $$ A^{\beta}_{k-1,k-1}=A^{\beta}_{k,k}=1 $$ For $j$ even, $2+2i\leq j \leq k-2$, and $i=1,2,...,{k\over 2}-1$ $$ A^{\beta}_{{k\over 2}-i,2i}=1$$ $$ A^{\beta}_{{k\over 2}-i,j}= -s_{2i,j} + \sum_{n=1}^{{j\over 2}-i-1}~(-1)^{n+1} F_{n,i,j} $$ $$ F_{n,i,j}= \sum_{\alpha_1=1}^{{j\over 2}-i-n} \left( \sum_{\alpha_2=1}^{{j\over 2}-i-\alpha_1-n+1} \left( \sum_{\alpha_3=1}^{{j\over 2}-i-\alpha_1-\alpha_2-n+2} \left( ... \sum_{\alpha_p=1}^{{j\over 2}-i-\sum_{r=1}^{p-1}\alpha_r -n+p-1} \left(... \sum_{\alpha_n=1}^{{j\over 2}-i-\sum_{r=1}^{n-1}\alpha_r -1} f(i,j,\alpha) \right) \right) \right) \right) $$ $$ f(i,j,\alpha)=s_{2i,2i+2\alpha_1}~s_{2i+2\alpha_1,2i+2\alpha_1+2\alpha_2}~s_{2i+2\alpha_1+ 2\alpha_2,2i+2\alpha_1+2\alpha_2+2\alpha_3}~...~ s_{2i+2\alpha_1+...+2\alpha_n,j} $$ For $j$ even, $2i+2\leq j \leq k-2$ and $i=0,1,...,{k\over 2}-2$ $$ A^{\beta}_{{k\over 2}+i,2i+1}=1$$ $$ A_{{k\over 2}+i,j}=-s_{2i+1,j}+\sum_{n=1}^{{j\over 2}-i-1}(-1)^{n+1} G_{n,i,j} $$ $$ G_{n,i,j}= \sum_{\alpha_1=1}^{{j\over 2}-i-n} \left( \sum_{\alpha_2=1}^{{j\over 2}-i-\alpha_1-n+1} \left( \sum_{\alpha_3=1}^{{j\over 2}-i-\alpha_1-\alpha_2-n+2} \left( ... \sum_{\alpha_p=1}^{{j\over 2}-i-\sum_{r=1}^{p-1}\alpha_r -n+p-1} \left(... \sum_{\alpha_n=1}^{{j\over 2}-i-\sum_{r=1}^{n-1}\alpha_r -1} g(i,j,\alpha) \right) \right) \right) \right) $$ $$ g(i,j,\alpha)=s_{2i+1,2i+2\alpha_1}~s_{2i+2\alpha_1,2i+2\alpha_1+2\alpha_2}~s_{2i+2\alpha_1+ 2\alpha_2,2i+2\alpha_1+2\alpha_2+2\alpha_3}~...~s_{2i+2\alpha_1+...+2\alpha_n,j} $$ More explicitely, {\small \begin{equation} A^{\beta}(S)= \pmatrix{ .&.&.&.&.&.&... \cr .&.&.&.&.&.&... \cr .&.&.&.&0&0&... \cr .&.&0&0&0&1&...\cr .&0&0&1&0&-s_{46}& ...\cr 0&1&0&-s_{24}&0&-s_{26}+s_{24}s_{46}&... \cr 1&-s_{12}&0&-s_{14}+s_{12}s_{24}&0&-s_{16}+(s_{12}s_{26}+s_{14}s_{46}) -s_{12}s_{24}s_{46}&... \cr 0&0&1&-s_{34}&0&-s_{36}+s_{34}s_{46}&... \cr . &.&0&0&1&-s_{56}&... \cr .&.&.&.&0&0&... \cr .&.&.&.&.&.&... \cr .&.&.&.&.&.&... \cr } \label{As1} \end{equation} } \vskip 0.2 cm \noindent {\bf $k$ odd:} \vskip 0.2 cm $$ A_{k-1,k-1}=A_{k,k}=1$$ For $j$ odd, $2i+1 \leq j \leq k-2$, $i=1,...,{k-3 \over 2}$: $$ A_{{k-1\over 2},2i}=1 $$ $$ A_{{k-1\over 2}+i,j}=-s_{2i,j}+\sum_{n=1}^{{j-1\over 2}-i} (-1)^{n+1} H(n,i,j) $$ $$ H(n,i,j)=\sum_{\alpha_1=0}^{{j-1\over 2}-i-n}\left(\sum_{\alpha_2=1}^{ {j-1\over 2}-i-\alpha_1-n+1} \left( ...\sum_{\alpha_p=1}^{{j-1\over 2}-i -\sum_{\alpha_r=1}^{p-1}\alpha_r-n+p-1} \left( ... \sum_{\alpha_n=1}^{{j-1\over 2}-i -\sum_{\alpha_r=1}^{n-1}\alpha_r-1} h(i,j,\alpha) \right)\right) \right) $$ $$ h(i,j,\alpha)= s_{2i,2i+1+2\alpha_1}~s_{2i+1+2\alpha_1, 2i+1+2\alpha_1+2\alpha_2 }~....~s_{2i+1+2\alpha_1+...+2\alpha_n,j} $$ For $j$ odd, $2i+3\leq j \leq k-2$, $i=0,1,...,{k-3\over 2}$: $$ A_{{k-1\over 2} -i , 2i+1}=1$$ $$ A_{{k-1\over 2}-i,j}=-s_{2i+1,j}+\sum_{n=0}^{{j-1\over 2}-i}(-1)^{n+1} V(n,i,j) $$ $$ V(n,i,j)=\sum_{\alpha_1=1}^{{j-1\over 2}-i-n} \left( \sum_{\alpha_2=1}^{{j-1\over 2}-i-\alpha_1-n+1}\left(...\sum_{\alpha_p=1}^{{j-1\over 2}-i-\sum_{r=1}^{p-1}\alpha_r -n+p-1} \left(... \sum_{\alpha_n=1}^{{j-1\over 2}-i-\sum_{r=1}^{n-1}\alpha_r -1} v(i,j,\alpha) \right)\right) \right) $$ $$ v(i,j,\alpha)= s_{2i+1,2i+1+2\alpha_1}~s_{2i+1+2\alpha_1,2i+1+2\alpha_1+2\alpha_2} ~...~ s_{2i+1+2\alpha_1+...+2\alpha_n,j} $$ More explicitely: {\small \begin{equation} A^{\beta}(S)= \pmatrix{ .&.&.&.&.&.&. &... \cr .&.&.&.&.&.&0& ... \cr .&.&.&.&0&0&1 &...\cr .&.&0&0&1&0&-s_{57}& ...\cr 0&0&1&0&-s_{35}&0&-s_{37}+s_{35}s_{57}&... \cr 1&0&-s_{13}&0&-s_{15}+s_{13}s_{35}&0&-s_{17}+(s_{13}s_{37}+s_{15}s_{57}) -s_{13}s_{35}s_{57}&... \cr 0&1&-s_{23}&0&-s_{25}+s_{23}s_{35}&0&-s_{27}+(s_{23}s_{37}+s_{25}s_{57}) -s_{23}s_{35}s_{57}&... \cr .&0&0&1&-s_{45}&0&-s_{47}+s_{45}s_{57}&... \cr .&.&.&0&0&1&-s_{67}&...\cr .&.&.&.&.&0&0&... \cr .&.&.&.&.&.&.&... \cr .&.&.&.&.&.&.&... \cr } \label{As2} \end{equation} } In order to prove the above expressions, we have to find each matrix $A^{\beta_{i,i+1}}$ corresponding to the elementary braid $\beta_{i,i+1}$ appearing in $\beta$. This means computing its entry $(i+1,i+1)$. This is a rather complicated problem, since the entry $(i+1,i+1)$ of a given $A^{\beta_{i,i+1}}$ is minus the entry $(i,i+1)$ of the Stokes' matrix resulting from the action of the elementary braids acting before $\beta_{i,i+1}$, which in general is a sum of products of the elements $s_{ij}$ of the initial Stokes matrix $S$. First we recall that $S\mapsto A^{\beta_{i,i+1}}~S~A^{\beta_{i,i+1}}$ has the following effect on the entries of $S$: $$ s_{n,i}\mapsto s_{n,i+1} ,~~~~s_{n,i+1}\mapsto s_{n,i}-s_{i,i+1}s_{n,i+1},~~~~ n=1,2,...,i-1 $$ $$s_{i,i+1}\mapsto -s_{i,i+1}$$ $$ s_{i,n}\mapsto s_{i+1,n},~~~~s_{i+1,n}\mapsto s_{i,n}-s_{i,i+1}s_{i+1,n}, ~~~~ n=i+2,...,k $$ while all the other entries of $S$ remain unchanged. We start from $A^{\beta_{k-5,k-4}}$, whose non trivial entry is simply $-s_{k-5,k-4}$. Its action on $S$ brings $s_{k-6,k-5}$ to $s_{k-6,k-4}$ (the reader may compute all the elements of $S^{\beta_{k-5,k-4}}$). Then, the entry of $A^{\beta_{k-6,k-5}}$ is $-s_{k-6,k-4}$. Proceeding in this way, the reader may check that for the first sequence of braids $(\beta_{k-5,k-4}\beta_{k-6,k-5}~...~\beta_{12})$ the entries $(i,i+1)$ of the matrices are: $$ -s_{i,k-4}~\hbox{ for } A^{\beta_{i,i+1}} $$ Now, observe that {\small $$ \pmatrix{ 1& \cr &\ddots & \cr & &1 \cr & & & 0&1\cr & & & 1 & x_1 \cr & & & & & & 1& \cr & & & & & & & 1 \cr & & & & & & && \ddots \cr & & & & & & && & 1 \cr } \pmatrix{ 1& \cr &\ddots & \cr & &1 \cr & & &1 \cr && & & 0&1\cr && & & 1 & x_2 \cr && & & & & & 1& \cr && & & & & & & \ddots \cr && & & & & & & & 1 \cr } =\pmatrix{ 1& \cr &\ddots & \cr & &1 \cr & & &0 & 0 & 1 \cr & & &1 & 0 & x_1\cr & & &0 & 1 & x_2 \cr & & & & & & 1& \cr & & & & & & & \ddots \cr & & & & & & & & 1 \cr } $$ } and recall that $A^{\beta_1 \beta_2}= A^{\beta_2} A^{\beta_1}$. This implies for the first sequence of braids: {\small $$ m_1:= A^{\beta_{k-5,k-4}~\beta_{k-6,k-5}~...~\beta_{12}} \pmatrix{ 0& & && & 1 & & \cr 1&0 & && & * & & \cr & 1 &0&& & * & & \cr & & \ddots&\ddots &&\vdots & & \cr & & & \ddots &0& * & & \cr & & & & 1 &*& & \cr &&&&& &1& \cr &&&&& & &1 \cr &&&&& & & &1 \cr &&&&& & & & &1 \cr } $$} the entries $*$ are exactly those of the $A_{i,i+1}$'s, namely $-s_{1,k-4}$, $-s_{2,k-4}$, ..., $-s_{k-5,k-4}$ from the top to the bottom of the $(k-4)^{th}$ column. For the second sequence of braids $(\beta_{k-7,k-6}~...~\beta_{34})$, the entries are $$ -s_{i-1,k-6}~\hbox{ for } A^{\beta_{i,i+1}} $$ and, as above: {\small $$ m_2:= A^{\beta_{k-6,k-5}~\beta_{k-7,k-6} ~... ~\beta_{23}}= \pmatrix{1& & & && & 0 & & \cr & 0& & && & 1 & & \cr & 1&0 & && & * & & \cr & & 1 &0&& & * & & \cr & & & \ddots&\ddots &&\vdots & & \cr & & & & \ddots &0& * & & \cr & & & & & 1 &*& & \cr & &&&&& &1& \cr & &&&&& &&1 \cr & &&&&& && &1 \cr & &&&&& && & &1 \cr & &&&&& && & & &1 \cr }$$} and the entries $*$ are those of the $A_{i,i+1}$'s (namely, $-s_{1,k-6}$, ..., $-s_{k-7,k-6}$ from the top to the bottom). For the third sequence $(\beta_{k-6,k-5}~...~\beta_{23})$, they are $$ -s_{i-2,k-8}~\hbox{ for } A^{\beta_{i,i+1}} $$ and: {\small $$ m_3:= A^{\beta_{k-7,k-6}~...~ \beta_{34}}= \pmatrix{1& & & & && & 0 \cr & 1& & & && & 0 \cr && 0& & && & 1 & & \cr && 1&0 & && & * & & \cr && & 1 &0&& & * & & \cr && & & \ddots&\ddots &&\vdots & & \cr && & & & \ddots &0& * & & \cr && & & & & 1 &*& & \cr && &&&&& &1& \cr && &&&&& &&1 \cr && &&&&& && &1 \cr && &&&&& && & &1 \cr && &&&&& && & & &1 \cr && &&&&& && & & & &1 \cr } $$ } And so on. We reach the last but one ``sequence'', namely $\beta_{{k\over 2}-2,{k\over 2}-1}$ for $k$ even, or $(\beta_{{k-3\over 2},{k-1\over 2}}\beta_{{k-5 \over 2},{k-3 \over 2}})$ for $k$ odd. The entries are $-s_{12}$, or $-s_{23}$,$-s_{13}$ respectively. Then {\small $$ m_{{k \over 2} -2}:= A^{ \beta_{ {k \over 2}-2,{k\over 2}-1 } }= \pmatrix{ 1& \cr &\ddots & \cr & &1 \cr & & & 0&1\cr & & & 1 & -s_{12} \cr & & & & & & 1& \cr & & & & & & & \ddots \cr & & & & & & & & 1 \cr } $$} or {\small $$ m_{k-5 \over 2}:= A^{\beta_{{k-3\over 2},{k-1\over 2}}\beta_{{k-5 \over 2},{k-3 \over 2}}}= \pmatrix{ 1& \cr &\ddots & \cr & &1 \cr & & & 0&0&1\cr & & & 1 &0 & -s_{13} \cr & & & &1 & -s_{23} \cr & & & & & &1 \cr & & & & & & & \ddots \cr & & & & & & & & 1 \cr } $$ } The entries for the last braid are more complicated, because the entries on the first upper sub-diagonal of the Stokes' matrix have been shuffled by the preceeding braids. We give the result ($A_{i,i+1}$ stends for $A^{\beta_{i,i+1}}$ {\small $$ A_{k-3,k-2} :~~~ -s_{k-3,k-2}$$ $$ A_{k-4,k-3} :~~~ -s_{k-5,k-2}+s_{k-5,k-4}s_{k-4,k-2}$$ $$ A_{k-5,k-4} :~~~ -s_{k-7,k-2}+(s_{k-7,k-4}s_{k-4,k-2}+ s_{k-7,k-6}s_{k-6,k-2})-s_{k-7,k-6}s_{k-6,k-4}s_{k-4,k-2}$$ $$ \vdots$$ $$ A_{{k \over 2}-1,{k \over 2}} :~~~ -s_{1,k-2}+(s_{12}s_{2,k-2}+s_{14}s_{4,k-2} +...+ s_{1,k-4} s_{k-4,k-2})~+~...~+~(-1)^{{k\over2}-1} s_{12}s_{24}s_{48}...s_{k-4,k-2}$$ $$ A_{{k \over 2}-2,{k \over 2}-1} :~~~ -s_{2,k-2}+(s_{24}s_{4,k-2}+s_{26}s_{4,k-2} +...+ s_{2,k-4}s_{k-4,k-2})~+~...~+~(-1)^{{k\over2}-2} s_{24}s_{48}...s_{k-4,k-2}$$ $$ \vdots $$ $$ A_{34} :~~~ -s_{k-8,k-2}+(s_{k-8,k-4}s_{k-4,k-2}+s_{k-8,k-6} s_{k-6,k-2}) - s_{k-8,k-6}s_{k-6,k-4} s_{k-4,k-2} $$ $$ A_{23} :~~~ -s_{k-6,k-2}+s_{k-6,k-4}s_{k-4,k-2} $$ $$ A_{12} :~~~ -s_{k-4,k-2} $$ } For $k$ odd {\small $$ A_{k-3,k-2} :~~~ -s_{k-3,k-2}$$ $$ A_{k-4,k-3} :~~~ -s_{k-5,k-2}+s_{k-5,k-4}s_{k-4,k-2}$$ $$ A_{k-5,k-4} :~~~ -s_{k-7,k-2}+(s_{k-7,k-4}s_{k-4,k-2}+ s_{k-7,k-6}s_{k-6,k-2})-s_{k-7,k-6}s_{k-6,k-4}s_{k-4,k-2}$$ $$ \vdots$$ $$ A_{{k -1\over 2},{k+1 \over 2}} :~~~ -s_{2,k-2}+(s_{23}s_{3,k-2}+s_{25}s_{5,k-2} +...+ s_{2,k-4} s_{k-4,k-2})~+~...~+~(-1)^{{k-3\over2}} s_{23}s_{35}s_{57}...s_{k-4,k-2}$$ $$ A_{{k-3 \over 2},{k-1 \over 2}} :~~~ -s_{1,k-2}+(s_{13}s_{3,k-2}+s_{15}s_{5,k-2} +...+ s_{1,k-4} s_{k-4,k-2})~+~...~+~(-1)^{{k-3\over2}} s_{13}s_{35}s_{57}...s_{k-4,k-2}$$ $$ \vdots $$ $$ A_{34} :~~~ -s_{k-8,k-2}+(s_{k-8,k-4}s_{k-4,k-2}+s_{k-8,k-6} s_{k-6,k-2}) - s_{k-8,k-6}s_{k-6,k-4} s_{k-4,k-2} $$ $$ A_{23} :~~~ -s_{k-6,k-2}+s_{k-6,k-4}s_{k-4,k-2} $$ $$ A_{12} :~~~ -s_{k-4,k-2} $$ } and: $$ A^{\beta_{k-3,k-2}~\beta_{k-4,k-3}~...~\beta_{12}} = \pmatrix{ 0& & && & 1 & & \cr 1&0 & && & * & & \cr & 1 &0&& & * & & \cr & & \ddots&\ddots & &\vdots \cr & & & \ddots & 0 &* & \cr & & & & 1 &* & & \cr &&&&& &1&0 \cr &&&&& &0&1 \cr }$$ which we call $m_{{k\over 2}-1}$ for $k$ even, $m_{k-3\over 2}$ for $k$ odd. Then, for $k$ even $$A^{\beta}=m_{{k\over 2}-1}m_{{k\over 2}-2}~...~m_3m_2m_1$$ and for $k$ odd $$A^{\beta}=m_{{k-3\over 2}}m_{{k-5\over 2}-2}~...~m_3m_2m_1$$ Doing a careful computation, we obtain (\ref{As1}) and (\ref{As2}). \vskip 0.3 cm 2) The second step consists of expressing $A^{\beta}$ in terms of the entries of the Stokes' factors of $S$, which are simply binomial coefficients. First we prove the following \vskip 0.2 cm \noindent {\bf Lemma 9:} {\it Given an upper triangular $k \times k$ matrix $S$, with entries $s_{ii}=1$, we can uniquely determine numbers $a_{ij}$ such that, for $k$, $i$, $j$ all even or all odd: {\small $$ s_{ij}= a_{ij}~+~(a_{i,i+2}a_{i+2,j}+a_{i,i+4}a_{i+4,j}+~...~+a_{i,j-2}a_{j-2,j})~+$$ $$ +~(a_{i,i+2}a_{i+2,,i+4}a_{i+4,j}+a_{i,i+2}a_{i+2,i+6}a_{i+6,j}+~...~+a_{i,j-4}a_{j-4,j-2} a_{j-2,j})~+.... ~+~ a_{i,i+2}a_{i+2,i+4}a_{i+4,i+6}~...~a_{j-2,j} $$ } If $k$ is even, but $i$ is odd, just replace in the formula $i+2$ with $i+1$, $i+4$ with $i+3$, ect. If $k$ is even, but $j$ is odd, just replace $j-2$ with $j-1$, $j-4$ with $j-3$, ect. If $k$ is odd, but $i$ is even, or $j$ is even, just do the same replacements as above. More explicitely: {\tiny $$ \hbox{\small $S$}:=\pmatrix{ \cr 1& a_{12} & a_{13}+a_{12}a_{23} & a_{14}+a_{12}a_{24} & \matrix{a_{15}+(a_{12}a_{25}+a_{14}a_{45})\cr \cr + a_{12}a_{24}a_{45}\cr} & \matrix{a_{16}+(a_{12}a_{26}+a_{14}a_{46})\cr\cr + a_{12}a_{24}a_{46}\cr} & \matrix{ a_{17}+ (a_{12}a_{27}+a_{14}a_{47} +a_{16}a_{67}) \cr\cr +(a_{12}a_{24}a_{47}+a_{12}a_{26}a_{67}+a_{14}a_{46}a_{67}) \cr\cr + a_{12}a_{24}a_{46}a_{67} \cr} & ....\cr \cr \cr \cr & 1 & a_{23} & a_{24} & a_{25} + a_{24}a_{45} & a_{26}+ a_{24}a_{46} & a_{27}+(a_{24}a_{47}+ a_{26}a_{67})+a_{24}a_{46}a_{67}& .... \cr \cr \cr &&1&a_{34}& a_{35} + a_{34}a_{45} & a_{36}+ a_{34}a_{46} & a_{37}+(a_{34}a_{47}+ a_{36}a_{67})+a_{34}a_{46}a_{67}& .... \cr \cr \cr &&&1& a_{45} & a_{46} & a_{47} a_{46}a_{67} & .... \cr \cr \cr &&&&1&a_{56} & a_{57}+a_{56}a_{57}& .... \cr \cr\cr &&&&&1& a_{67}&....\cr\cr\cr &&&&&&1&.... \cr \cr\cr &&&&&&&.... \cr }$$ } for $k$ even. {\tiny $$ \hbox{\small $S$}:=\pmatrix{ \cr 1& a_{12} & a_{13} & a_{14}+a_{13}a_{34} & a_{15}+a_{13}a_{35} & a_{16}+(a_{13}a_{36}+a_{15}a_{56}) + a_{13}a_{35}a_{56} & a_{17}+ (a_{13}a_{37}+a_{15}a_{57}) +a_{13}a_{35}a_{57}& ....\cr \cr \cr & 1 & a_{23} & a_{24}+a_{23}a_{34} & a_{25} + a_{23}a_{35} & a_{26}+ (a_{23}a_{36}+a_{25}a_{56})+a_{23}a_{35}a_{56} & a_{27} + (a_{23}a_{37}+a_{25}a_{57})+a_{23}a_{35}a_{57}& .... \cr \cr \cr &&1&a_{34}& a_{35} & a_{36}+ a_{35}a_{56} & a_{37}+a_{35}a_{57} & .... \cr \cr \cr &&&1& a_{45} & a_{46}+a_{45}a_{56} & a_{47}+ a_{45}a_{57} & .... \cr \cr \cr &&&&1&a_{56} & a_{57}& .... \cr \cr\cr &&&&&1& a_{67}&....\cr\cr\cr &&&&&&1&.... \cr \cr\cr &&&&&&&.... \cr }$$ } for $k$ odd. } \vskip 0.2 cm \noindent {\bf Proof: } We have to solve a non linear system $F_{ij}(a)=s_{ij}$. The sum of the differences between the indices of the factors $a$ in $s_{ij}$ is equal to the difference of the indices of $s_{ij}$, namely $j-i$. >From this it follows that the trems non-linear in the $a$'s occur on sub-diagonals which lie above all the sub-diagonals containing the factors of the non linear terms. Then, the system $F_{ij}(a)=s_{ij}$ is uniquely solvable, starting from the first sub-diagonal and successively determining all the $a_{rs}$ going up diagonal by diagonal. \rightline{$\Box$} \vskip 0.2 cm \noindent {\bf Corollary 3:} {\it With the above factorization, the matrix $A^{\beta}$ becomes: {\small \begin{equation} A^{\beta}(S)= \pmatrix{ .&.&.&.&.&.&. &.&... \cr .&.&.&.&.&.&. &0&... \cr .&.&.&.&.&0&0& 1&... \cr .&.&.&0&0&1&0&-a_{68}&...\cr .&0&0&1&0&-a_{46}&0&-a_{48} & ...\cr 0&1&0&-a_{24}&0&-a_{26}&0&-a_{28} &... \cr 1&-a_{12}&0&-a_{14}&0&-a_{16}&0&-a_{18}&... \cr 0&0&1&-a_{34}&0&-a_{36}&0&-a_{38}&... \cr . &.&0&0&1&-a_{56}&0&-a_{58}&... \cr .&.&.&.&0&0&1&-a_{78}&... \cr .&.&.&.&.&.&0&0&... \cr .&.&.&.&.&.&.&.&... \cr } \label{As3} \end{equation}} for $k$ even; and {\small \begin{equation} A^{\beta}(S)= \pmatrix{ .&.&.&.&.&.&. &... \cr .&.&.&.&.&.&0& ... \cr .&.&.&.&0&0&1 &...\cr .&.&0&0&1&0&-a_{57}& ...\cr 0&0&1&0&-a_{35}&0&-a_{37}&... \cr 1&0&-a_{13}&0&-a_{15}&0&-a_{17}&... \cr 0&1&-a_{23}&0&-a_{25}&0&-a_{27}&... \cr .&0&0&1&-a_{45}&0&-a_{47}&... \cr .&.&.&0&0&1&-a_{67}&...\cr .&.&.&.&.&0&0&... \cr .&.&.&.&.&.&.&... \cr .&.&.&.&.&.&.&... \cr } \label{As4} \end{equation}} for $k$ odd. In other words, all the non linear terms in formulae (\ref{As1}) and (\ref{As2}) drop. } \vskip 0.2 cm \noindent {\bf Proof:} Just substitute the factorization of $S$ in (\ref{As1}), (\ref{As2}). \rightline{$\Box$} \vskip 0.2 cm In section 5 we computed the Stokes' factors for $S$. If we sum all the factors appearing in formula (\ref{facto}) we get a matrix of the form: $$M:= K_{1k}+K_{1,k-1}+K_{k,k-1}+...+K_{k3}+K_{k2}= \pmatrix{ k & & &&&& 0\cr * & k & &&&0& * \cr * & * & k &&&*&* \cr \vdots&\vdots&\vdots&\ddots& \cdot^{\displaystyle{~\cdot}^{\displaystyle{~\cdot}} }&\vdots&\vdots \cr * & * & * &\cdot^{\displaystyle{~\cdot}^{\displaystyle{~\cdot}} } &\ddots&\vdots&\vdots \cr * & * & 0 & & & k & *\cr * & 0 & & & & & k \cr }$$ The $*$ are the binomial coefficient appearing in the factors. If we know $M$, we can determine all the entries of the single Stokes' factors, because if the entry $(i,j)$ is not zero for one factor, then it is zero for all the other factors. Now, we rename the entries of the factors according to the following rule: {\small $$ M:= \pmatrix{ k & & &&&&&&&&& 0\cr a_{k-2,k}& k & &&&&&&&&0& a_{k-2,k-1} \cr a_{k-4,k} & a_{k-4,k-2} & k &&&&&&&&a_{k-4,k-3}&a_{k-4,k-1} \cr \vdots&\vdots&\vdots& \ddots&&&&&& \cdot^{\displaystyle{~\cdot}^{\displaystyle{~\cdot}} } &\vdots&\vdots& \cr a_{4k} &a_{4,k-2} & ...& a_{46} &k& 0&0&0&a_{45}&a_{47}&...&a_{4,k-1} \cr a_{2,k}&a_{2,k-2}&...& a_{26}&a_{24}&k&0&a_{23}&a_{25}&a_{27}&...&a_{2,k-1} \cr a_{1k} & a_{1,k-2}& ...&a_{16}&a_{14}&a_{12}&k&a_{13}&a_{15}&a_{17}&...&a_{1.k-1} \cr a_{3k}&a_{3,k-2}&...&a_{36}&a_{34}&0&0&k&a_{35}&a_{37}&...&a_{3,k-1} \cr a_{5k}&a_{5,k-2}&...&a_{56}&0&0&0&0&k&a_{57}&...&a_{5,k-1} \cr \vdots&\vdots&\cdot^{\displaystyle{~\cdot}^{\displaystyle{~\cdot}} }& &&&&&&\ddots&...&...\cr a_{k-3,k}&a_{k-3,k-2}&&&&&&&&&k&a_{k-3,k-1}\cr a_{k-1,k}& &&&&&&&&&& 1\cr }$$ } for $k$ even. {\small $$ M= \pmatrix{ k&&&&& & & & & & & &0 \cr a_{k-2,k}&k&&&& & & & & & & &a_{k-2,k-1} \cr a_{k-4,k}&a_{k-4,k-2}&k&&& & & & & & & a_{k-4,k-3}&a_{k-4,k-1} \cr \vdots&\vdots&\vdots&\ddots&& & & & & & \cdot^{\displaystyle{~\cdot}^{\displaystyle{~\cdot}} }& \vdots&\vdots \cr a_{5k}&a_{5,k-2}&a_{5,k-4}&...&k&0&0&0&0&a_{56}& ...& a_{5,k-3}&a_{5,k-1} \cr a_{3k}&a_{3,k-2}&a_{3,k-4}&...&a_{35}&k&0&0&a_{34}&a_{36}& ...& a_{3,k-3}&a_{3,k-1} \cr a_{1k}&a_{1,k-2}&a_{1,k-4}&...&a_{15}&a_{13}&k&a_{12}&a_{14}&a_{16}& ...& a_{1,k-3}&a_{1,k-1} \cr a_{2k}&a_{2,k-2}&a_{2,k-4}&...&a_{25}&a_{23}&0&k&a_{24}&a_{26}&... & a_{2,k-3}&a_{2,k-1} \cr a_{4k}&a_{4,k-2}&a_{4,k-4}&...&a_{45}&0&0&0&k&a_{46}&... & a_{4,k-3}&a_{4,k-1} \cr \vdots&\vdots&&\cdot^{\displaystyle{~\cdot}^{\displaystyle{~\cdot}} }&& & & & & & \ddots & \vdots &\vdots \cr a_{k-3,k}&a_{k-3,k-2}&&&& & & & & & & k &a_{ k-3,k-1} \cr a_{k-1,k}& &&&& & & & & & & & k \cr } $$ } for k odd. The strange labelling is simply the one such that $$P M P^{-1} = \pmatrix{k&a_{12}&a_{13}&a_{14}& ...\cr & k& a_{23}&a_{24}& ... \cr & & k&a_{34}&...\cr & & & \ddots& \vdots\cr & & & & k \cr} $$ where the matrix of permutation is $$P=\pmatrix{ &&&& & &&1&0&&&\cr &&&& &&1&0&0&&&&\cr &&&& &0&0&0&1&0&&&\cr &&&&&1&0&0&0&0& &&&\cr &&&& 0&0&0&0&0&1&0& && \cr \vdots&\vdots& &\vdots&\vdots&\vdots&\vdots&\vdots&\vdots&\vdots&\vdots& \vdots&& \vdots &\vdots \cr & 1 &&&&&&&&&&&&0&\cr 0&0&&&&&&&&&&&&0&1\cr 1&0& &&&&&&&&&&&0&0\cr } $$ for $k$ even (the 1 on the first row is on the ${k\over 2}+1$-th column) ; and $$P=\pmatrix{ &&&& & &0&1&&&&\cr &&&& &&0&0&1&&&&\cr &&&& &0&1&0&0&0&&&\cr &&&&&0&0&0&0&1& &&&\cr &&&&0&1&0&0&0&0&0 && \cr \vdots&\vdots& &\vdots&\vdots&\vdots&\vdots&\vdots&\vdots&\vdots&\vdots& \vdots&& \vdots &\vdots \cr 0 & 1 &&&&&&&&&&&&0&\cr 0&0&&&&&&&&&&&&0&1\cr 1&0& &&&&&&&&&&&0&0\cr } $$ for $k$ odd (the 1 on the first row is on the ${k+1\over 2}$-th column). With this choice of the labelling, the product $S_{upper}:=P \left( K_{1k}K_{1,k-1}K_{k,k-1}~...~ K_{k3}K_{k2}~\right)P^{-1}$ is precisely factorized as in lemma 9. Then we can write the entries of $A^{\beta}$ (formulae (\ref{As3}) (\ref{As4})) from the entries of the Stokes factors (which are binomial coefficient). The final result is precisely the claim of the proposition. \rightline{$\Box$} \vskip 0.3 cm We are ready to prove the main result: \vskip 0.2 cm \noindent {\bf Theorem 2:} {\it Consider the Stokes matrix $S=T_F^{k \over 2}~T^{-{k \over 2}}$ ($k$ even) or $S=T_F^{k-1 \over 2} K_{k2} T^{-{k-1 \over 2}}$ ($k$ odd) for the quantum cohomology of ${\bf CP}^{k-1}$ and set it in the upper triangular form $S_{upper}=P~S~P^{-1}$ by the permutation $P$. Then, there exists a braid $\beta$ (Lemma 8), represented by a matrix $A^{\beta}$ (Proposition), which sets $S_{upper}$ in the form (\ref{SSCAN}). The last column is negative, but conjugation by } ${\cal I} =$ diag$(1,1,...,1,-1)$ {\it makes it positive. We reach the canonical form: $$ s_{ij}=\bin{k}{j-i},~~~~i<j$$ Another conjugation by } diag$(-1,1,-1,1,-1,...)$ {\it brings the matrix in the equivalent canonical form $$ s_{ij}=(-1)^{j-i}\bin{k}{j-i}, ~~~~ i<j$$ Finally, by the action of the braid group, the last matrix can be put in the form $$s_{ij}=\bin{k-1+j-i}{j-i},~~~~i<j$$ In all the above matrices $$ s_{ii}=1,~~~~s_{ij}=0 ~~~~i>j$$ } \vskip 0.2 cm \noindent {\bf Proof:} First, we want to explain which is the braid which brings the upper triangular matrix with entries $ s_{ij}=(-1)^{j-i}\bin{k}{j-i}$ in the matrix $s_{ij}=\bin{k-1+j-i}{j-i}$. We make use of the following known result \cite{Zas}: \vskip 0.2 cm \noindent {\it Consider the upper triangular Stokes' matrix S, the braid $\beta= \beta_{12}(\beta_{23}\beta_{12})(\beta_{34}\beta_{23}\beta_{12})$~...~ $(\beta_{n-1,n}\beta_{n-2,n-1}...\beta_{23}\beta_{12})$ and the permutation $$P=\pmatrix{&&& 1\cr &&1&\cr &\cdot^{\displaystyle{~\cdot}^{\displaystyle{~\cdot}} }&& \cr 1&&& \cr}$$ Then, the relation \begin{equation} \left[S^{-1}\right]^{\beta}=PS^TP \label{Smeno1} \end{equation} holds. } \vskip 0.2 cm Observe that for the matrix $S$, whose upper triangular part has entries $s_{ij}=\bin{k-1+j-i}{j-i}$, we have $PS^TP\equiv S$. Moreover, $S^{-1}$ is upper triangular with entries $ s_{ij}=(-1)^{j-i}\bin{k}{j-i}$. This proves that $S$ and $S^{-1}$ are equivalent w.r.t the action of the braid group. \vskip 0.2 cm Let us now prove the theorem staring from $k$ even. We have to prove that $A^{\beta}~P~T_F^{k \over 2}~T^{-{k\over 2}}~P^{-1} [A^{\beta}]^T$ is in ``canonical form'' (\ref{SSCAN}). The proof ``reduces'' to the computation of products of matrices explicitely given. We do the products in an shrewd way. First we rewrite $$ S^{\beta} = A^{\beta}~\left( P~T_F~P^{-1}\right)^{k\over 2}~\left(P~ T^{-1}~ P^{-1} \right)^{k\over 2}~[A^{\beta}]^T$$ and we compute: {\tiny $$\hbox{\small $P~T_F~P^{-1}$}= \pmatrix{ &&&& & &&1&0&&&\cr &&&& &&1&0&0&&&&\cr &&&& &0&0&0&1&0&&&\cr &&&&&1&0&0&0&0& &&&\cr &&&& 0&0&0&0&0&1&0& && \cr \vdots&\vdots& &\vdots&\vdots&\vdots&\vdots&\vdots&\vdots&\vdots&\vdots& \vdots&& \vdots &\vdots \cr & 1 &&&&&&&&&&&&0&\cr 0&0&&&&&&&&&&&&0&1\cr 1&0& &&&&&&&&&&&0&0\cr } \pmatrix{ 0 & & & &...& 1 \cr -1& 0 & & && \cr & -1 & 0 & && \cr & & -1 & && \cr & & & \ddots&\ddots& \vdots\cr & & & & -1 &0\cr } \pmatrix{ &&&&&&0&1\cr &&&&&1&0&0\cr &&&&\vdots&\cr &&0&1 \cr &1&0&0 \cr 1&0&0&0\cr 0 &0 &1&0\cr &&0&0\cr &&&&\vdots\cr &&&&&0&0&0\cr &&&&&&1&0 \cr } $$ } (in $P^{-1}$ the 1 on the first column is on the ${k\over 2}+1$-th row) {\small $$ = \pmatrix{ 0&-1\cr 0&0&0& -1 \cr -1& 0&0&0& & \cr &0&0&0&0&-1 \cr & &-1&0&0&0\cr &&&&&&\vdots&\vdots \cr &&&&&&&&0 &0&0&-1\cr &&&&&&&&-1&0&0&0\cr &&&&&&&&&&1&0 \cr } $$} Thus $$ \left(P~T_F~P^{-1}\right)^{k\over 2}= \pmatrix{ &&&&&&&-1 \cr &&&&&&1\cr &&&&&-1\cr &&&&\cdot^{\displaystyle{~\cdot}^{\displaystyle{~\cdot}} }\cr &&&-1 \cr &&1\cr &-1\cr 1\cr } $$ Then, using the expression of $A^{\beta}$ from the proposition: {\tiny $$ \hbox{\small $F:= A^{\beta}~\left(P~T_F~P^{-1}\right)^{k\over 2}$}= \pmatrix{ 0&0&1 \cr &0&\bin{k}{1}&0&1 \cr &&\bin{k}{2}&0&\bin{k}{1} \cr &&\bin{k}{3}&0&\bin{k}{2}&0&1\cr &&\vdots & &\vdots& & \vdots& \cr &&&&&&&&...&1\cr &&*&&*&&*&&...&\bin{k}{1}&0&1 \cr &&*&&*&&*&&...&\bin{k}{2}&0&\bin{k}{1}&-1\cr &&*&&*&&*&&...&\bin{k}{3}&-1&0\cr &&\vdots&&\vdots&&\vdots \cr &&\bin{k}{k-4}&0&\bin{k}{k-5}\cr &&\bin{k}{k-3}&-1&0\cr \cr 0&-1&0\cr \cr -1&0\cr } $$ } (the -1 on the last column is on the ${k\over 2}$-th row). Using the explicit expressions for $K_{k3}$ and $K_{k2}$ of section 5 we compute $$ T^{-1}= T_F^{-1}~K_{k3}~K_{k2}=\pmatrix{0&-1\cr &0&-1\cr & & \ddots&\ddots \cr &&&& 0&-1 \cr &&&&&0&-1 \cr 1&&&&&&0 \cr }~K_{k3}~K_{k2}=$$ {\tiny $$ = \pmatrix{ \bin{k}{1}& -1 && &&&&&&&&& -\bin{k}{2} \cr & 0& -1& &&&&&&&& -\bin{k}{4}&-\bin{k}{3} \cr & & 0&-1 &&&&&&&-\bin{k}{6}&-\bin{k}{5}&0 \cr \cr & & & \ddots &\ddots &&&&&\cdot^{\displaystyle{~\cdot}^{\displaystyle{~\cdot}} }&\cdot^{\displaystyle{~\cdot}^{\displaystyle{~\cdot}} } \cr \cr & & & & 0 &-1&0& -\bin{k}{k-2}&-\bin{k}{k-3}\cr & & & & & 0& -1&-\bin{k}{k-1} \cr\cr &&&&&& 0& -1 \cr\cr &&&&&&& 0&-1 \cr \cr & & & & && & & \ddots&\ddots &\cr\cr\cr\cr\cr & & & & && & & & &&0&-1\cr \cr & & & & &&& & & && &0\cr } $$ } Then {\tiny $$ \hbox{\small $P~T^{-1}~P^{-1}$}= \pmatrix{ 0&0&-1\cr -1&0&-\bin{k}{k-1} \cr 0&0&0&0&-1\cr 0&-1&-\bin{k}{k-2}&0&-\bin{k}{k-3}&0 \cr &0&0&0&0&0&-1\cr &&&-1&-\bin{k}{k-4}&0&-\bin{k}{k-5} \cr &&& & & & & \ddots \cr &&& & & & & & 0&0&0&-1& 0\cr &&& & & & & & -1&-\bin{k}{4}&0&-\bin{k}{3}& 0\cr &&& & & & & & &0&0&0& 1\cr &&& & & & & & &0&-1&-\bin{k}{2}& \bin{k}{1}\cr } $$ } After this we computed $F~\left[P~T^{-1}~P^{-1}\right]$, $F~\left[P~T^{-1}~P^{-1}\right]~ \left[P~T^{-1}~P^{-1}\right]$, $F~\left[P~T^{-1}~P^{-1}\right]^3 $, ..., $ F~\left[P~T^{-1}~P^{-1}\right]^{k\over 2}$. We omit the intermediate computations and we give the final result: {\tiny $$ \hbox{\small $F_1:= F~\left(P~T^{-1}~P^{-1}\right)^{k\over 2}$}= \pmatrix{ &&&&&&&& &&1&\bin{k}{2}&-\bin{k}{1} \cr &&&&&&&& &\bin{k}{4}&0&\bin{k}{3}&-\bin{k}{2} \cr &&&&&&& &&\bin{k}{5}&0&\bin{k}{4}&-\bin{k}{3} \cr &&&&&&&... &&\vdots&\vdots &\vdots & \vdots \cr &&&&&1&\bin{k}{k-6}& ...&...&*&0&*&*\cr &&&1&\bin{k}{k-4}&0&\bin{k}{k-5}&...&...&*&0 &*&* \cr &1&\bin{k}{k-2}&0&\bin{k}{k-3}&0&\bin{k}{k-4}&...&...&*&0 & *&*\cr 1&0&\bin{k}{k-1}&0&\bin{k}{k-2}&0&\bin{k}{k-3}&...&...&*&0&*&* \cr &&1&0&\bin{k}{k-1}&0&\bin{k}{k-2}&...&...&*&0&*&*\cr &&&&1&0&\bin{k}{k-1}&...&...&*&0&*&*\cr &&&&&&1&...&...&*&0&*&*\cr &&&&&&&...&...&\vdots&\vdots&\vdots&\vdots \cr &&&&&&& &&\bin{k}{k-1}&0&\bin{k}{k-2}&-\bin{k}{k-4} \cr &&&&&&& &&&1&\bin{k}{k-1}&-\bin{k}{k-2} \cr &&&&&&& &&& &&-\bin{k}{k-1} \cr &&&&&&& && && & 1 \cr } $$ } Now, multiplying $F_1~\left[A^{\beta}\right]^T$, we obtain precisely the ``canonical form'' (\ref{SSCAN}). \vskip 0.2 cm For $k$ odd, we did a similar computation. We omit the detail, but we indicate the order of multiplications which yielded the most simple expressions to multiply step by step. Our aim is to compute $A^{\beta}~P~T_F^{k-1\over 2}~K_{k2}~(T_F^{-1}~K_{k3}~K_{k2})^{k-1\over 2}~P^{-1}~[A^{\beta}]^T$. First, we computed $PT_FP^{-1}$, then $(PT_FP^{-1})^{k-1\over 2}$, then $F:=A^{\beta}~(PT_FP^{-1})^{k-1\over 2}$. After this, we computed $PK_{k2}P^{-1}$, and $F_1:=F~PK_{k2}P^{-1}$. Finally, we calculated $m:=P~T_F^{-1}~K_{k3}~K_{k2}~ P^{-1}$ and $F_1~m$, $F_1~m~m$, ..., $F_2:=F_1~m^{k-1\over 2}$. The matrix $F_2~[A^{\beta}]^T$ proved to be in ``canonical form''. \rightline{$\Box$} \section{ Canonical form of $S^{-1}$} The matrix $S^{-1}$ such that $Y_R(z)=Y_{L}(z)~S^{-1}$ can be put in the same canonical form of $S$, as a consequence of the relation (\ref{Smeno1}). The only remarks we want to add concern the braid which brings $S^{-1}$ to the canonical form, because it arranges the lines $L_j$ in a ``beautiful'' shape. \vskip 0.2 cm \noindent {\bf Lemma $8^{\prime}$:} {\it Let the points $u_j$ be in lexicographical order w.r.t the admissible line $l$. Let us denote $\sigma_{i,i+1}:=\beta_{i,i+1}^{-1}$. Then, the following braid arranges the poins in cyclic clockwise order, $u_1$ being the first point in $\Pi_L$ for $k$ even, or the last in $\Pi_R$ for $k$ odd (w.r.t the clockwise order) (see figure 6): $$ \beta^{\prime}:= \left[~( \sigma_{34}\sigma_{56}\sigma_{78} ~...~\sigma_{k-3,k-2} \sigma_{k-1,k})~( \sigma_{45}\sigma_{67}\sigma_{89} ~...~\sigma_{k-4,k-3} \sigma_{k-2,k-1})~...~(\sigma_{{k\over 2},{k\over 2}+1}\sigma_{{k\over 2} +2,{k\over 2}+3})~\sigma_{{k\over 2}+1,{k \over 2}+2} \right]$$ $$\left[~( \sigma_{{k\over 2}+2,{k\over 2}+3}\sigma_{{k\over 2}+3,{k\over 2}+4}~...~\sigma_{k-2,k-1}\sigma_{k-1,k})~( \sigma_{{k\over 2}+2,{k\over 2}+3}\sigma_{{k\over 2}+3,{k\over 2}+4}~...~\sigma_{k-2,k-1})~....~(\sigma_{{k\over 2}+ 2,{k\over 2}+3}\sigma_{{k\over 2}+3,{k\over 2}+4})~\sigma_{{k\over 2}+ 2,{k\over 2}+3} \right] $$ for $k$ even, and $$ \beta^{\prime}:=\beta_{12}~ \left[ ~(\sigma_{45}\sigma_{67}\sigma_{89}~...~\sigma_{k-3,k-2} \sigma_{k-1,k})~(\sigma_{56} \sigma_{78}~...~\sigma_{k-4,k-3}\sigma_{k-2,k-1})~....~(\sigma_{{k+1\over 2},{k+3 \over 2}}\sigma_{{k+5 \over 2},{k+7\over 2}})~\sigma_{{k+3\over 2},{k+5\over 2}}\right]$$ $$ \left[~(\sigma_{{k+5\over 2},{k+7 \over 2}} \sigma_{{k+7 \over 2},{k+9 \over 2}}~...~\sigma_{k-2,k-1}\sigma_{k-1,k})~(\sigma_{{k+5\over 2},{k+7 \over 2}} \sigma_{{k+7 \over 2},{k+9 \over 2}}~...~\sigma_{k-2,k-1})~....~(\sigma_{{k+5\over 2},{k+7 \over 2}} \sigma_{{k+7 \over 2},{k+9 \over 2}})~\sigma_{{k+5\over 2},{k+7 \over 2}} \right] $$ for $k$ odd. } \begin{figure} \epsfxsize=15cm \epsffile{figr6.eps} \caption{ Lines $L_j$ (Branch cuts) after the braid which brings $S^{-1}$ to canonoical form} \end{figure} \vskip 0.2 cm A carefull consideration of the topological effect of the braid on the lines $L_j$ shows that they are arranged as in figure 6. To reconstruct the configuration it is enough to know the admissible line $l$ (at angle $\epsilon$ w.r.t. the positive real axis). In fact, $u_1$ is the first point in $\Pi_L$ (in clockwise order) for $k$ even, or the last in $\Pi_R$ for $k$ odd. The lines come out of the points in centrifugal directions. They go to infinity, without intersections (so preserving their lexicographical order w.r.t $l$) with the original asymptotic direction ${\pi \over 2}-\epsilon$. \vskip 0.2 cm \noindent {\bf Proposition $1^{\prime}$} {\it The matrix representing $\beta^{\prime}$ is {\tiny $$ A^{\beta^{\prime}}=\pmatrix{ 1\cr &1\cr &&-\bin{k}{k-3}&1\cr &&-\bin{k}{k-4}&0&-\bin{k}{k-5}&1\cr && -\bin{k}{k-5}&0&-\bin{k}{k-6}& \cr && \vdots & & \vdots & &... \cr && -* & & -* & & ... &1 \cr && -* & & -* & & ... &&-\bin{k}{3}&1\cr && -\bin{k}{{k\over2}-1} & & -\bin{k}{{k\over2}-2} & & ... && -\bin{k}{2}&0&-\bin{k}{1}&1 \cr && ~~\bin{k}{{k\over 2}-2}& & ~~\bin{k}{{k\over 2}-3}&&...&&~~\bin{k}{1}&0&~~1 \cr && ~~*& & ~~* & & ...&&~~1 \cr && ~~\vdots&&~~\vdots&& ... \cr &&~~\bin{k}{2} & &~~\bin{k}{1} \cr &&~~\bin{k}{1} &0&~~1 \cr && ~~1 &0 \cr } $$} for $k$ even, and {\tiny $$ A^{\beta^{\prime}}=\pmatrix{ 0&1\cr 1&\bin{k}{1}\cr &&1\cr &&&-\bin{k}{k-4}&1\cr &&&-\bin{k}{k-5}&0&-\bin{k}{k-6}&1\cr &&& -\bin{k}{k-6}&0&-\bin{k}{k-7}& \cr &&& \vdots & & \vdots & &... \cr &&& -* & & -* & & ... &1 \cr &&& -* & & -* & & ... &&-\bin{k}{3}&1\cr &&& -\bin{k}{{k-3\over2}} & & -\bin{k}{{k-5\over2}} & & ... && -\bin{k}{2}&0&-\bin{k}{1}&1 \cr &&& ~~\bin{k}{{k-5\over 2}}& & ~~\bin{k}{{k-7\over 2}}&&...&&~~\bin{k}{1}&0&~~1 \cr &&& ~~*& & ~~* & & ...&&~~1 \cr &&& ~~\vdots&&~~\vdots&& ... \cr &&&~~\bin{k}{2} & &~~\bin{k}{1} \cr &&&~~\bin{k}{1} &0&~~1 \cr &&& ~~1 &0 \cr } $$} for $k$ odd. } The proposition is proved as in the previous section. Finally, the analogous of Theorem 2 holds: \vskip 0.2 cm \noindent {\bf Theorem $2^{\prime}$} {\it Consider the Stokes matrix $S^{-1}=T^{k \over 2}~T_F^{-{k \over 2}}$ ($k$ even) or $S=T^{k-1 \over 2} K_{k2}^{-1} T_F^{k-1 \over 2}$ ($k$ odd) for the quantum cohomology of ${\bf CP}^{k-1}$ and set it in the upper triangular form $S_{upper}^{-1}=P~S^{-1}~P^{-1}$ by the permutation $P$. Then, there exists a braid $\beta^{\prime}$ (Lemma $8^{\prime}$), represented by a matrix $A^{\beta}$ (Proposition $1^{\prime}$), which sets $S_{upper}^{-1}$ in the ``canonical form''. The entries $s_{ij}$ have minus sign for $j>{k\over 2}+1,~~i\leq {k\over 2}+1$ and $k$ even, or $j>{k+3\over 2},~~i\leq {k+3 \over 2}$ and $k$ odd. A suitable conjugation by ${\cal I}=$ diag$( 1,1,...1,-1,...,-1)$ sets all signs positive and the final matrix has entries $$ s_{ij}=\bin{k}{j-i}, ~~~s_{ji}=0,~~~\hbox{ for }i<j,~~~~s_{ii}=1 $$ } \vskip 0.2 cm \noindent {\bf Proof: } The proof is similar to the one of theorem 2. \rightline{$\Box$} \vskip 0.2 cm Examples are found in Appendix 2. \vskip 0.3 cm \section{ Relation between Irregular and Fuchsian systems} \vskip 0.2 cm Let us consider the fuchsian system \begin{equation} \left( U-\lambda \right) ~{d \phi \over d \lambda} = \left( {1 \over 2} +V \right) ~\phi \label{Fuc20} \end{equation} which can also be written $$ {d \phi \over d \lambda}= \sum_{j=1}^k~{A_j \over \lambda - u_j} ~\phi $$ $$ A_j = - E_j ~ \left( {1 \over 2} +V \right),~~~~~~~(E_j)_{jj}=1,~~\hbox{ otherwise } (E_j)_{nm}=0 $$ Around the point $u_j$ a fundamental matrix has the form $$ \left[B_0 + O\left(\lambda - u_j \right) \right]~(\lambda - u_j)^M $$ where $M=$diag$(-{1\over 2},0,...,0)$ and the columns of $B_0$ are the eigenvectors of $A_j$; in particular, the first column is $(0,..,0,1,0,...,0)^T$, and $1$ occurs at the $j^{th}$ position. Then, the system has $k$ independent vector solutions, of which $k-1$ are regular near $u_j$ and the last is $$ \phi^{(j)}(\lambda)={ 1 \over \sqrt{ \lambda - u_j}}~\pmatrix{0\cr \vdots \cr 1 \cr \vdots \cr 0 \cr } + O \left( \sqrt{ \lambda - u_j} \right)~~~\lambda\to u_j $$ where $1$ occurs at the $j^{th}$ row. For any $u_j$ we can construct such a basis of solutions. The branch of $ \sqrt{ \lambda - u_j}$ is chosen as follows: let us consider an angle $\eta$ with a range of $2 \pi$, for example $-{\pi \over 2} \leq \eta < {3 \pi \over 2}$, such that $ \eta \neq \arg(u_i-u_j)$, $\forall i \neq j$. Then consider the cuts $L_j=\{ \lambda= u_j + \rho e^{i \eta},~~\rho>0\}$. Actually, the cuts have two sides, $L_j^{+}= \{ \lambda= u_j + \rho e^{i \eta},~~\rho>0\}$ and $L_j^{-}=\{ \lambda= u_j + \rho e^{i( \eta- 2 \pi)},~~\rho>0\}$. The branch is determined by the choice $ \log (\lambda -u_j)= \log |\lambda - u_j| + i \eta $ on $L_j^{+}$ and $ \log (\lambda -u_j)= \log |\lambda - u_j| + i (\eta- 2\pi) $ on $L_j^{-}$. On ${\bf C} \backslash \bigcup_j L_j$, $\sqrt{ \lambda -u_1}$, ..., $\sqrt{ \lambda -u_k}$ are single valued. For any two (column) vector solutions $\phi(\lambda)$, $\psi(\lambda)$ we define the symmetric bilinear form: $$ \left( \phi,\psi \right) := \phi(\lambda)^T (\lambda -U) \psi(\lambda) $$ which is independent of $\lambda$ and $u_1$, ...,$u_k$. Let $G$ be the matrix whose entries are $G_{ij}=\left( \phi^{(i)},\phi^{(j)} \right)$. In particular, $G_{ii}=1$. Then, it can be proved (see \cite{BJL2} and also \cite{Dub2} ) that near $u_j$ $$ \phi^{(i)}(\lambda) = G_{ij}\phi^{(j)}(\lambda)+ r_{ij}(\lambda) $$ where $ r_{ij}(\lambda)$ is regular near $u_j$. For a counter-clockwise loop around $u_j$ the monodromy of $\phi^{(i)}$ is $$ \phi^{(i)} \mapsto R_j \phi^{(i)} := \phi^{(i)} - 2{ \left( \phi^{(i)},\phi^{(j)} \right) \over \left( \phi^{(j)},\phi^{(j)} \right)} ~\phi^{(j)} \equiv \phi^{(i)} - 2 G_{ij} ~\phi^{(j)} $$ Then, the monodromy group of (\ref{Fuc20}) acts on $\phi^{(1)}$, ..., $\phi^{(k)}$ as a reflection group whose Gram matrix is $2G$. In particular, $\phi^{(1)}$, ..., $\phi^{(k)}$ are linearly independent (and then a basis) if and only if $\det G \neq 0$. \vskip 0.2 cm Now consider an oriented line $l$ of argument $\theta = {\pi \over 2} - \eta$, and for any $j$ define the following vector \begin{equation} \tilde{Y}^{(j)} = -{\sqrt{z} \over 2 \sqrt{\pi}} \int_{\gamma_j} d \lambda ~\phi^{(j)}(\lambda)~e^{\lambda z} \label{Laplace} \end{equation} which is a Laplace transform of $\phi^{(j)}$ . The path $\gamma_j$ comes from infinity near $L_j^{+}$, encircles $u_j$ and returns to infinity along $L_j^{-}$. We can define $\Pi_L=\{\theta< \arg z <\theta +\pi\}$ and $\Pi_R=\{ \theta - \pi< \arg z < \theta \}$. $\lambda=\infty$ is a regular singularity for (\ref{Fuc20}), then the integrals exist for $z \in \Pi_L$, and the non-singular matrix $\tilde{Y}(z):= \left[\tilde{Y}^{(1)}| ...|\tilde{Y}^{(k)} \right]$ has the asymptotic behaviour $$ \tilde{Y}(z)\sim \left(I + O\left( {1 \over z}\right)\right)~e^{zU}~~~z \to \infty,~~z \in \Pi_L$$ and satisfies the system (\ref{20}). Then it is a fundamental matrix $\tilde{Y}_L$. Note that $l$ is admissible, since it does not contain Stokes` rays. \vskip 0.2 cm It is a fundamental result \cite{Dub2} that the Stokes' matrix of (\ref{20}) satisfies $$ S+S^T = 2G $$ \vskip 0.3 cm \section{ Monodromy Group of the Quantum Cohomology of ${\bf CP}^{k-1}$} \vskip 0.2 cm A system like (\ref{Fuc20}) comes about in the theory of Frobenius manifolds (replace $U\mapsto U(t)$, $V\mapsto V(t)$). It determines flat coordinates $x^1(t,\lambda)$, ..., $x^k(t,\lambda)$ for a linear pencil of metrics $(~,~)-\lambda<~,~>^*$ ($(~,~)$ is the {\it intersection form} \cite{Dub1} \cite{Dub2}). We write a gauge equivalent form (gauge $X(t^2)$) at the semisimple point $(0,t^2,0,...,0)$ \begin{equation} (\hat{\cal U}(t^2)-\lambda)~{d\psi \over d \lambda}= \left({1\over 2} + \hat{\mu}\right)\psi \label{sticaz} \end{equation} A fundamental matrix $\psi(t,\lambda)$ has entries $\psi_{\alpha}^{(j)}(t,\lambda)=\partial_{\alpha}x^{(j)}(t,\lambda)$. Moreover, by (\ref{Laplace}) \begin{equation} \partial_{\alpha} \tilde{t}^j(t,z)= -{\sqrt{z}\over 2 \sqrt{\pi}} \int d \lambda~\partial_{\alpha}x^j(t,\lambda)~e^{\lambda z} \label{Basta!!} \end{equation} The {\it Monodromy group} of the Frobenius manifold $M$ is the group of the transformations which $(x^{1}(t,\lambda),...,x^{k}(t,\lambda))$ undergo when $t$ moves in $M\backslash \Sigma_{\lambda}$, where $\Sigma_{\lambda}= \{ t \in M~|$ $\det\left[ (~,~)-\lambda <~,~>\right]=0$ $\}$ is the {\it discriminant} of the linear pencil. Due to formula (\ref{Basta!!}), for ${\bf CP}^{k-1}$ this group is generated by the monodromy of the solutions of (\ref{sticaz}) when $\lambda$ describes loops around $u_1(t),...,u_k(t)$ (see \cite{Dub1} \cite{Dub2}). To these loops, we must add the effect of the displacement $t^2\mapsto t^2+2\pi i$. In fact, in this case $$[\varphi^{(1)}(ze^{t^2\over k}),..., \varphi^{(k)}(ze^{t^2\over k})] \mapsto [\varphi^{(1)}(ze^{t^2\over k}),..., \varphi^{(k)}(ze^{t^2\over k})] ~T$$ and the same holds for $\tilde{t}(z,(0,t^2,...,0))$. Then, the monodromy group of the quantum cohomology of ${\bf CP}^{k-1}$ is generated by the transformations $R_1$, $R_2$, ..., $R_k$, $T$ introduced in the preceeding sections. We are going to study the structure of the monodromy group of ${\bf CP}^{k-1}$ for any $k\geq 3$. Recall that the matrix $S$ for (\ref{20}) is not upper triangular, because in $U$ the order of $u_1$, ..., $u_k$ is not lexicographical w.r.t. the line $l$. Then, Coxeter identity is $-S^{-1}S^T=$ product of the $R_j$'s in the order refered to $l$. For example, for $k=3$, $S^{-1}S^T= -R_2 R_3 R_1$, since the lexicographical ordering would be $u_2$, $u_3$, $u_1$. From the identity $S^{-1}S^T= (-1)^{k-1}T^k$ it follows a first general relation in the group $$ T^{k} = (-1)^k \hbox{ product of }R_j \hbox{'s in suitable order } $$ Two cases must now be distinguished. \vskip 0.2 cm \noindent {\bf $k$ odd:} As a general result \cite{BJL2}, $\det G =0$ if and only if $V + {1\over 2} $ has an integer eigenvalue. The eigenvalues of $V$ are ${k-1 \over 2}$, ${k-3 \over 2}$, ..., $-{k-1 \over 2}$. Then, for $k$ odd, $\det G \neq 0$, and $\phi^{(1)}$, ..., $\phi^{(k)}$ are a basis. The matrices $R_j$ are $$ R_j = \pmatrix{ 1 & & & & & \cr &\ddots & & & & \cr & & \ddots & & & \cr -2G_{j1} & -2 G_{j2} & ... & -1& ...& -2G_{jk} \cr & & & & \ddots &\cr &&&&&1\cr } $$ where $S^T+S=2G$. In concrete examples, we have ``empirically'' found other relations like $$ R_2= p_1(T,R_1)$$ $$ R_2= p_2(T,R_1)$$ $$ \vdots$$ $$ R_k= p_k(T,R_1)$$ where $p_j(T,R_1)$ means a product of the elements $T$ and $R_1$. We have also found the relation $$ \left(T R_1 \right)^k = -I $$ We investigated the following cases: \vskip 0.2 cm \noindent {\small ${\bf CP}^2$ ($k=3$) $$ R_2=TR_1T^{-1},~~~R_3=T~(R_1 R_2 R_1)~T^{-1} $$ $$ (TR_1)^3 =-I $$ $$ T^3=-R_2 R_3 R_1 $$ \noindent ${\bf CP}^4$ ($k=5$) $$\left\{ \matrix{ R_2=TR_1T^{-1},~~~R_3=TR_2T^{-1} \cr R_4=T~(R_2R_3R_2)~T^{-1},~~~ R_5=T^{-1}~(R_2R_1R_2)~T \cr } \right. $$ $$ (TR_1)^5=-I $$ $$ T^5= - R_3R_4R_2R_5R_1 $$ \noindent ${\bf CP}^6$ ($k=7$) $$ \left\{ \matrix{ R _2=TR_1T^{-1},~~~R_3=TR_2T^{-1},~~~R_4=TR_3T^{-1} \cr R_5=T~(R_3R_4R_3)T^{-1},~~~R_6=T(TR_1)^3~R_2~[T(TR_1)^3]^{-1}\cr R_7=T^{-2}(R_3R_2R_3)T^2 \cr } \right. $$ $$ (TR_1)^7=-I$$ $$T^7 = -R_4R_5R_3R_6R_2R_7R_1$$ } Note that one ralation, for example that for $R_k$, can be derived from the others, and that just $R_1$, $T$, $-I$ are enought to generate the monodromy group in each of the examples. They satisfy ({\it in the examples}) the relations: $$ R_1^2=(-I~T~R_1)^k =(-I)^2 =I $$ $$ R_1(-I)~\left((-I)R_1 \right)^{-1}=I, ~~~ T(-I)~\left((-I)T \right)^{-1}=I $$ The last two relations mean simply the commutativity of $-I$ with $R_1$ and $T$. The relations are not only satisfied, but also ``fulfilled'' (namely, $(-I~T~R_1)^n\neq I$ for $n<k$). Now call $$ X:=R_1,~~~Y:=-ITR_1,~~~Z=-I$$ These elements generate the monodromy group of ${\bf CP}^{k-1}$ with {\it at least} the relations $$ X^2=Y^k=Z^2=1$$ $$ (ZX)(XZ)^{-1}=1,~~~(ZY)(YZ)^{-1}=1$$ Note that $Z$ generates the cyclic group $C_2$ of order 2. If there were no other relations (which we did not find ``empirically''), we would conclude that the monodromy group of the quantum cohomology of ${\bf CP}^k$ ({\it in the examples}) is isomorphic to the direct product $$ < X,Y~|~X^2=Y^k=1>\times C_2 $$ where $ < X,Y~|~X^2=Y^k=1>$ means the group generated by $X$, $Y$ with relations $X^2=Y^k=1$. \vskip 0.2 cm \noindent {\bf $k$ even:} Now $\det G=0$, since $V +{1\over 2}$ has integer eigenvalues. $G$ has rank $k-1$ and the eigenspace of its eigenvalue 0 has dimension 1. Let $(z^1,...,z^k)^T$ be an eigenvector of eigenvalue 1. The vector $v:= \sum_{j=1}^k z^j \phi^{(j)}$ is zero, because $$ \left(v, \phi^{(i)} \right)= \sum_{j=1}^k z^j G_{ji}=0~~~\forall i $$ then $$ z^1 \phi^{(1)}+z^{(2)}\phi^{(2)}+...+z^k\phi^{(k)} =0 $$ and $k-1$ of the $\phi^{(j)}$'s are linearly independent. The fuchsian system (\ref{Fuc20}) has a regular (vector) solution $\phi_0(\lambda)= \sum_{n=0}^d ~\phi_n \lambda^n$, where $\phi_n$ are constant (column) vectors, and $\phi_d$ is the eigenvector of $V+{1 \over 2}$ relative to the largest integer eigenvalue less or equal to zero; this eigenvalue is precisely $-d$ (see \cite{BJL2}). In our case, $d=0$ and $\phi_o(\lambda)=\phi_0$, a constant vector. $\phi^{(1)}$, $\phi^{(2)}$, ..., $\phi^{(k-1)}$, $\phi_0$ is then a possible choice for a basis of solutions. Observe that in the gauge equivalent form $\psi=X\phi$, $\psi_0$ is the eigenvector of ${1\over 2}+\hat{\mu}$ with eigenvalue zero. Then $$ \psi_0=\pmatrix{0\cr \vdots \cr 1 \cr \vdots \cr 0 \cr }~\equiv ~\pmatrix{\partial_1 x\cr \partial_2 x \cr \vdots \cr \vdots \cr \partial_k x \cr} $$ where all the entries are zero but the one at position ${k \over 2}+1$. $x$ is the flat coordinate for $(~,~)-\lambda <~,~>$ corresponding to $\psi_0$. Then, we can chose the following flat coordinates: $$ x^{1}(\lambda,t),~x^{2}(\lambda,t),~...~,~x^{k-1}(\lambda,t), ~ t^{{k \over2}+1}$$ The monodromy group then acts on a $k-1$ dimensional space. Let us determine the reduction of $R_1$, $R_2$, ..., $R_k$, $T$ to the $k-1$ dimensional space. The entries of $T$ on the vectors $\phi^{(j)}$ are: $T\phi^{(i)}= \sum_{j=1}^k ~T_{ji} \phi^{(j)}$, $i=1,...,k$. On the new basis $\phi^{(1)}$, ..., $\phi^{(k-1)}$, $\phi_0$ the matrices are rewritten $$ R_j \phi^{(i)}= \phi^{(i)} - 2 G_{ij} \phi^{(j)} ~~~i=1,..,k-1~~ ~j\neq k $$ $$ R_j \phi_0 = \phi_0 ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~j\neq k$$ $$ R_k \phi^{(i)} = \phi^{(i)} - 2 G_{ik} \left( -{1 \over z^k}~\sum_{j=1}^{k-1} ~z^j \phi^{(j)} \right) ~~~ i \neq k$$ $$ T \phi^{(i)} = \sum_{j=1}^{k-1} ~T_{ji} \phi^{(j)} +T_{ki} \left( -{1 \over z^k}~\sum_{j=1}^{k-1} ~z^j \phi^{(j)} \right) $$ Then the matrices assume a reduced form $\pmatrix{ * & * & 0 \cr * & * & 0 \cr 0 & 0 & 1 \cr }$ We studied two examples; besides Coxeter identity $T^k=$product of $R_j$'s, we found relations similar to the case $k$ odd: $$ R_2= p_1(T,R_1)$$ $$ R_2= p_2(T,R_1)$$ $$ \vdots$$ $$ R_k= p_k(T,R_1)$$ and $$ \left(T R_1 \right)^k = I $$ Namely: \vskip 0.2 cm {\small \noindent ${\bf CP}^3$ ($k=4$) $$ \left\{ \matrix{ R_2=TR_1T^{-1},~~~ R_3=T R_2T^{-1} \cr R_4 = T^{-1}~(R_2R_1R_2)~T \cr } \right. $$ $$ (TR_1)^4 =I $$ $$ T^4 =R_3R_2R_4R_1$$ \noindent ${\bf CP}^5$ ($k=6$) $$ \left\{ \matrix{ R_2=TR_1T^{-1},~~~ R_3=T R_2T^{-1} \cr R_4 = TR_3T^{-1},~~~R_5=T~(R_2R_3R_4R_3R_2)~T^{-1} \cr R_6= T^{-1}~(R_2 R_1 R_2)~T } \right. $$ $$ (TR_1)^6=I $$ $$ T^6= R_4R_3R_5R_2R_6R_1 $$ } The same remarks of $k$ odd hold here. Call $$ X:=R_1,~~~~Y:=R_1T$$ then, if there were no other hiden relations, the monodromy group of the quantum cohomology of ${\bf CP}^k$ ({\it in the examples}) would be isomorphic to $$ < X,Y,~|~ X^2=Y^k=1> $$ \vskip 0.2 cm Note that $ < X,Y,~|~ X^2=Y^k=1>$ is (isomorphic to) the subgroup of orientation preserving transformations of the hyperbolic triangular group $[2,k,\infty]$. \vskip 0.2 cm \noindent {\bf Lemma 10:} {\it The subgroup of the orientation preserving transformations of the hyperbolic triangular group $[2,k,\infty]$ is isomorphic to the subgroup of $PSL(2,{\bf R})$ generated by $$ \tau \mapsto -{1 \over \tau} $$ $$ \tau \mapsto {1 \over 2 \cos {\pi \over k} - \tau} $$ $ \tau \in H:= \{ z \in {\bf C} ~|~\Im z >0\}$ } \vskip 0.2 cm \noindent {\bf Proof: }Consider three integers $m_1$, $m_2$, $m_3$ such that $$ {1 \over m_1} +{1 \over m_2} +{1 \over m_3}<1$$ In the Bolyai-Lobatchewsky plane $H$, the triangular group $[m_1,m_3,m_3]$ of hyperpolic reflections in the sides of hyperbolic triangles of angles ${\pi \over m_1}$, ${\pi \over m_2}$, ${\pi \over m_3}$ is generated by three reflections $r_1$, $r_2$, $r_3$ satisfying the relations $$ r_1^2=r_2^2=r_3^2=(r_2r_3)^{m_1}=(r_3r_1)^{m_2}=(r_2 r_1)^{m_3}=1 $$ and the subgroup of orientation preserving transformation is generated by $X=r_2r_3$, $Y=r_3 r_1$, Then $$ X^{m_1}=Y^{m_2}=(XY)^{m_3}=1$$ For $m_1=2$, $m_2=k$, $m_3 = \infty$, a fundamental triangular region is $\{ 0< \Re z < \cos {\pi \over k} \} \cap \{ |z|>1 \}$. Then $$ r_1(\tau)=- \bar{\tau},~~~r_2(\tau)= {1 \over \bar{\tau}},~~~r_3(\tau)=2 \cos {\pi \over k} - \bar {\tau} $$ The bar means complex conjugation. Then $$X(\tau)=-{1 \over \tau},~~~ Y(\tau)= { 1 \over 2 \cos {\pi \over k} -\tau}$$ \rightline{$\Box$} \vskip 0.2 cm \noindent {\bf Remark :} The orientation preserving transformations of $[2,3,\infty]$ are the {\it modular group} $PSL(2, {\bf Z})$. \vskip 0.3 cm \noindent {\bf Theorem 3:} {\it The monodromy group of the quantum cohomology of ${\bf CP}^2$ is isomorphic to \begin{equation} < X,Y,~|~ X^2=Y^3=1>\times C_2 \cong PSL(2,{\bf Z}) \times C_2 \label{Re2} \end{equation} The monodromy group of the quantum cohomology of ${\bf CP}^3$ is isomorphic to \begin{equation} < X,Y,~|~ X^2=Y^4=1>\cong \hbox{ orient. preserv. transf. of } [2,4,\infty ] \label{Re3} \end{equation} } \vskip 0.2 cm The theorem for the case of ${\bf CP}^2$ is already proved in \cite{Dub2}. \vskip 0.2 cm \noindent {\bf Proof :} a) ${\bf CP}^2$: $$ R_1=\pmatrix{ -1&3&3 \cr 0&1&0 \cr 0&0&1 \cr }~~~~ T=\pmatrix{ 0&0&1 \cr -1 & 3 & 3 \cr 0 & -1 & 0 \cr } $$ and $X=R_1$, $Y= -I T R_1$ and $Z= -I$ satisfy the relations of (\ref{Re2}). They act on the column vector ${\bf x} =\pmatrix{x \cr y \cr z \cr}$. The quadratic form $q(x,y,z)={\bf x}^T G {\bf x}$ is $R_1$ and $T$ - invariant. Then $T$, $R_1$ act on two dimensional invariant subspaces $q(x,y,z)=$ constant. On each of these subspaces we introduce new coordinates $\chi\in {\bf R}$ and $\varphi\in [0,2\pi)$. Let $\tau = e^{\chi} ~e^{i\varphi}$ and $$ x= {a \over 2} ( \tau \bar{\tau} -{3\over 2} ( \tau +\bar{\tau})+1 ) ~{i \over \tau -\bar{\tau}} $$ $$ y= {a \over 2} ( \tau \bar{\tau} -{1\over 2} ( \tau +\bar{\tau})-1 ) ~{i \over \tau -\bar{\tau}} $$ $$ z= {a \over 2} ( -\tau \bar{\tau} -{1\over 2} ( \tau +\bar{\tau})+1 ) ~{i \over \tau -\bar{\tau}} $$ $a \in {\bf R}$, $a \neq 0$. Note that $q(x,y,z)=a^2>0$. Then, it is easily verified that $$ {\bf x}(-{1 \over \tau})= -X~{\bf x}(\tau) $$ $$ {\bf x}({1 \over 1-\tau})= Y~{\bf x}(\tau) $$ $$ {\bf x}(\tau, -a)= Z~ {\bf x}(\tau,a) $$. This implies the 1 to 1 correspondence between the generators of the modular group and $X$ and $Y$. \vskip 0.2 cm b) Case of ${\bf CP}^3$. $$ R_1=\pmatrix{ -1 & 4 & -10 & 0 \cr 0 & 1 & 0 & 0 \cr 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 1 \cr},~~~ T=\pmatrix{0 & 0 & 1 & 0 \cr -1 & 0 & 3 & 0 \cr 0 & -1 & 3 & 0 \cr 0 & 0 & 0 & 1 \cr } $$ The matrices are already written on $\phi^{(1)}$, $\phi^{(2)}$, $\phi^{(3)}$, $\phi_0$. Recall that the monodromy acts only on $x^1,x^2,x^3$, because the last flat coordinate is $t^3$. This action is given by the following three dimensional matrices, acting on a three dimensional space of vectors ${\bf x}= \pmatrix{ x \cr y \cr z \cr}$ $$ r_1= \pmatrix{ -1 & 4 & 10 \cr 0 & 1 & 0 \cr 0 & 0 & 1 \cr},~~~t:= \pmatrix{ 0 & 0 & 1 \cr -1 & 0 & 3 \cr 0 & -1 & 3\cr } $$ We redefine $X= r_1$ and $Y=t r_1$, which satisfy the relations (\ref{Re3}). We proceed as above, defining $$ x= {a } ( \tau \bar{\tau} -{1\over \sqrt{2}} ( \tau +\bar{\tau})+{1\over 3} ) ~{i \over \tau -\bar{\tau}} $$ $$ y= {a } (-{2 \over 3} \tau \bar{\tau} -{2 \sqrt{2}\over 3} ( \tau +\bar{\tau})+{2 \over 3} ) ~{i \over \tau -\bar{\tau}} $$ $$ z= {a } ( -{1 \over 3}\tau \bar{\tau} -{\sqrt{2}\over 6} ( \tau +\bar{\tau})+{1\over 3} ) ~{i \over \tau -\bar{\tau}} $$ $a\neq 0$. Note that ${\bf x}^T g {\bf x}= (8/9) a^2$, where $g$ is the $3 \times 3 $ reduction of $G$. It is easily verified that $$ {\bf x}(-{1\over \tau})= -X~ {\bf x}(\tau) $$ $$ {\bf x}({1\over \sqrt{2}-\tau})= -Y~ {\bf x}(\tau) $$ which proves the theorem. \rightline{$\Box$} \vskip 1 cm \noindent {\bf APPENDIX 1: Proof of Lemma 5} \vskip 0.3 cm Let us consider the function $$ g(z)= C~ \int_{-c-i \infty}^{-c+i\infty} ds ~\Gamma^k (-s) ~e^{i \pi f s } ~z^{ks} $$ where $C$ and $f$ are constants to be determined later, $c>0$. The path of integration is a vertical line through $-c$. {\bf a)} Domain of definition. We use Stirling formula $$ \Gamma(-s)= e^s e^{-(s+{1 \over 2}) \log(-s)} \sqrt{ 2 \pi} ( 1 + O(1/s)) ~~~~~ s \to \infty, ~~~|\arg(-s)|< \pi $$ where $ \log(-s) = \log(s) + i \pi $. The integrand is $$ \Gamma^k(-s)e^{i\pi f s} z^{ks} \sim e^{ks(1 + \log z)}e^{-k(s+{1 \over 2})\log s} e^{-ik \pi (s+{1 \over 2}) + i \pi s f} ~~~s \to \infty $$ Now let $s=-c + i \eta$, $ds = i d \eta$. The integral in $\eta$ is on the real axis. The dominant part in the integrand is $$ e^{ \eta \left[k \arg(i\eta -c ) + k \pi - k \arg z - \pi f \right] } $$ The condition of (uniform) convergence of the integral is $$ -{\pi \over 2} -f {\pi \over k} < \arg z< {\pi \over 2} - f {\pi \over k} $$ \vskip 0.2 cm {\bf b)} $g$ solves (\ref{30}). In fact (for simplicity $C=1$ here): $$ (z \partial_z)^k g(z)= k^k \int_{-c -i\infty}^{-c+i \infty} ds ~ \Gamma^k(-s) e^{i \pi f s } s^k z^{ks}= (-1)^k k^k \int_{-c -i\infty}^{-c+i \infty} ds ~\Gamma^k(1-s)e^{i \pi f s} z^{ks} $$ where we have used the identity $ s \Gamma(-s)= - \Gamma(1-s)$. Now let $t= s-1$: $$ (z \partial_z)^k g(z)= (-1)^k e^{i \pi f} (zk)^k \int_{-c-1 -i\infty}^{-c-1+i \infty} dt ~ \Gamma^{k}(-t) e^{i \pi f t} z^{kt} $$ $$ = (-1)^k e^{i\pi f} (kz)^k g(z)$$ Now we impose $(-1)^k e^{i \pi f}=1$, namely $$ f = k + 2 m~~~~~~~m \in {\bf Z} $$ Point b) is proved. \vskip 0.2 cm {\bf c)} Asymptotic behaviour of $g(z)$. We use Laplace method for analytic functions (the so called ``steepest descent method''). Put $C=1$. By Stirling's formula $$ g(z) = (2 \pi)^{k/2} e^{-i {\pi \over 2} k} ~ \int_{-c-i \infty}^{-c+i\infty} ds~ e^{\phi(s)} $$ $$ ~~~~~\phi(s)=s~[ k(1 + \log z) + i \pi f - i \pi k] - k\left( s + {1 \over 2}\right) \log s + O\left( {1 \over s} \right) $$ The the stationary point of $\phi(s)$ is $$ s_0 = z e ^{-i { \pi \over k } (k-f)} -{1 \over 2} + O\left({1 \over z} \right) ~~~~~z \to \infty $$ In the hypothesis $z \to \infty$ a straightforward computation gives $$ \phi(s_0)\sim k z e^{-i{\pi \over k} (k-f)}-{k \over 2} \log z + i {\pi \over 2} (k-f)~~~~~{d^2 \phi(s_0)\over ds^2}\sim -{k \over z} e^{i {\pi \over k}(k-f)} $$ We are ready to apply the steepest descent method. We deform the path of integration in such a way that it passes through $s_0$. Let us call it $\gamma$. $$ g(z)= (2 \pi)^{k \over 2} e^{-i {\pi \over 2} k} e^{\phi(s_0)} ~ \int_{\gamma} ds ~ e^{\phi(s)-\phi(s_0)} $$ $$ \sim (2 \pi)^{k \over 2} e^{-i {\pi \over 2} k} e^{\phi(s_0)}~ \int_{\gamma} ds~ e^{{1 \over 2} {d^2 \phi(s_0) \over ds^2} (s-s_0)^2} $$ Let us divide $\gamma$ in two paths: $\gamma_1$ from $s_0$ to $+ i \infty$ and $\gamma_2$ from $- i \infty$ to $s_0$. The integrals becomes the sum of two integrals. In $\int_{\gamma_1}$ we change variable. Let $\tau>0$ and $$ - \tau^2 = {1 \over 2} {d^2 \phi(s_0) \over ds^2} (s-s_0)^2= -{ k \over 2 z} \left[ e^{i {\pi \over k} (k-f)} (s-s_0)^2 \right] $$ Then $$ \int_{\gamma_1} e^{ {d^2 \phi(s_0) \over ds^2} (s -s_0)^2}= {1 \over 2} \sqrt{ \pi \over k} ( 2 z )^{1\over 2} e^{ - i {\pi \over 2 k} (k-f) -i \pi} + O(e^{- \alpha |z|})~~~~~\alpha>0$$ Note that $ \int_{\gamma_1} =\int_{\gamma_2}$. Now, recalling that $f= k + 2 m$, we conclude that $$ g(z) \sim ( 2 \pi) ^{k+1 \over 2} e^{-i(m+1) \pi} e^{ -i { \pi \over 2}k} ~ ~ { 1 \over \sqrt{k}}{ e^{i{\pi \over k}m} \over z^{k-1 \over 2}}~ \exp( k e^{i{2\pi \over k}m} z)~~~~~z\to \infty $$ If we choose $$ m= n-1 +kl ~~~~l \in {\bf Z} $$ and $C= \left[ (2\pi)^{k+1 \over 2} e^{i \pi ( l -{k\over 2} -n-kl)} \right]^{-1}$ we have $$ g(z) \sim {1 \over \sqrt{k}} { e^{i { \pi \over k}(n-1)} \over z^{k-1\over 2}} \exp( k e^{i { 2\pi \over k}(n-1)} z) $$ Then $g$ is a solution $\varphi^{(n)}$ on the domain $$ -{3 \pi \over 2} -2(n-1) { \pi \over k} -2 \pi l< \arg z < -{ \pi \over 2} -2(n-1) { \pi \over k} -2 \pi l $$ \vskip 0.2 cm {\bf d)} We now determine the domain where the analytic extension of $g(z)$ still has the above asymptotic behaviour. From Lemma 4 we derived that the first entries of the $n(k)^{th}$ columns of $Y_L$ and $Y_R$ are equal and coincide with $\varphi^{(n(k))}_L(z)\equiv \varphi^{(n(k))}_R(z)$ times $z^{k-1 \over 2}$, and $\varphi^{(n(k))}$ has the estabilished asymptotic behaviour on the enlarged domains $$ -{\pi \over k} - \pi < \arg z < \pi + {\pi \over k} $$ $$ - \pi < \arg z < \pi + {2\pi \over k} $$ If in $g(z)$ we chose $n=n(k)$ and $l=-1$ the integral representation holds for $-{\pi \over 2} <\arg z < {\pi \over 2}$ for $k$ even, and for $-{\pi \over 2}+{\pi \over k} < \arg z < {\pi \over 2}+{\pi \over k}$ for $k$ odd. Then its analytic continuation is precisely $\varphi^{(n(k))}_L(z)\equiv \varphi^{(n(k))}_R(z)$. \vskip 0.2 cm {\bf e)} We prove the identity (\ref{60}). Observe that $(z \partial_z)^k \varphi = (kz)^k \varphi$ is invariant for $z \mapsto z~e^{2\pi i \over k}$. A generic solution can be reperesented near $z=0$ as $$ \varphi(z)= \sum_{m=0}^{\infty} {z^{km}\over (m!)^k}~\left[ a^{(1)}_m+a^{(2)}_m\log z + ... + a^{(k)}_m \log^{k-1}z \right] $$ It follows that the operator $(A \varphi)(z)=\varphi(z e^{2 \pi i \over k})$ has eigenvalues 1, because on the basis obtained with $(a^{(1)}_0=1,~ a^{(2)}_0=...=a^{(k)}_0=0)$, $( a^{(1)}_0=0, ~a^{(2)}_0=1, ~a^{(3)}_0=...= a^{(k)}_0=0)$, ..., $(a^{(1)}_0=...= a^{(k-1)}_0=0, ~a^{(k)}_0=1)$ it is represented by a lower triangular matrix having 1's on the diagonal. Then $(A-1)^k=0$ and $$ (A-1)^k g(z)=0 $$ is precisely our identity. Lemma 4 is proved. \rightline{$\Box$} \vskip 1 cm \noindent {\bf APPENDIX 2: } \vskip 0.2 cm First we give $\Phi_R$ and $\Phi_L$. {\bf $k$ odd: } {\tiny $$ \Phi_R(z)^T= \left[ \matrix{ (-1)^{k-1 \over 2}\left[ g(ze^{-{2 \pi i \over k} ({k-1 \over 2})}) -\bin{k}{1} g(ze^{-{2 \pi i \over k} ({k-3 \over 2})})+ ... +\bin{k}{k-1} g(ze^{{2 \pi i \over k} ({k-1 \over 2})}) \right] \cr \cr \vdots \cr \cr g(ze^{-{4\pi i \over k} })-\bin{k}{1} g(ze^{-{2\pi i \over k} }) + \bin{k}{2} g(z)- \bin{k}{3} g(ze^{{2\pi i \over k} }) + \bin{k}{4} g(ze^{{4\pi i \over k} }) \cr - g(ze^{-{2\pi i \over k} }) + \bin{k}{1} g(z)- \bin{k}{2} g(ze^{{2\pi i \over k} }) \cr g(z) \cr - g(ze^{{2\pi i \over k} }) \cr g(ze^{{4\pi i \over k} }) \cr \cr \vdots \cr \cr (-1)^{k-1 \over 2} g(ze^{{2\pi i \over k} ({k-1\over 2})}) \cr } \right] $$ \vskip 0.2 cm $$ \Phi_L(z)^T= \left[ \matrix{ (-1)^{k-1 \over 2} g(ze^{-{2\pi i \over k} ({k-1\over 2})}) \cr -(-1)^{k-1 \over 2} g(ze^{-{2\pi i \over k} ({k-3\over 2})}) \cr \cr \vdots \cr \cr -g(ze^{-{2\pi i \over k} }) \cr g(z) \cr - g(ze^{{2\pi i \over k}} )+\bin{k}{k-1}g(z) \cr g(ze^{{4\pi i \over k} })- \bin{k}{k-1} g(ze^{{2\pi i \over k} })+ \bin{k}{k-2} g(z)- \bin{k}{k-3} g(ze^{-{2\pi i \over k} }) \cr \cr \vdots \cr \cr (-1)^{k-1\over 2} \left[ g(ze^{{2\pi i \over k} ({k-1\over 2})})- \bin{k}{k-1} g(ze^{{2\pi i \over k} ({k-3\over 2})}) + ...- \bin{k}{2} g(ze^{-{2\pi i \over k} ({k-3\over 2})}) \right] \cr }\right] $$ } \vskip 0.3 cm \noindent {\bf $k$ even:} {\tiny $$ \Phi_R(z)^T= \left[ \matrix{ (-1)^{k \over 2}\left[ g(ze^{-{i \pi } }) -\bin{k}{1} g(ze^{-i \pi +i{2 \pi \over k} })+ ... +\bin{k}{k-1} g(ze^{i(\pi-{2 \pi \over k})}) \right] \cr \cr \vdots \cr \cr g(ze^{-{4\pi i \over k} })-\bin{k}{1} g(ze^{-{2\pi i \over k} }) + \bin{k}{2} g(z)- \bin{k}{3} g(ze^{{2\pi i \over k} }) \cr - g(ze^{-{2\pi i \over k} }) + \bin{k}{1} g(z) \cr g(z) \cr - g(ze^{{2\pi i \over k} }) \cr g(ze^{{4\pi i \over k} }) \cr \cr \vdots \cr \cr (-1)^{{k \over 2}-1} g(ze^{{2\pi i \over k} ({k\over 2}-1)}) \cr } \right] $$ \vskip 0.2 cm $$ \Phi_L(z)^T= \left[ \matrix{ (-1)^{k \over 2} g(ze^{-i\pi }) \cr \vdots \cr \cr g(ze^{-i{4\pi \over k} }) \cr -g(ze^{-{2\pi i \over k} }) \cr g(z) \cr - g(ze^{{2\pi i \over k}} )+\bin{k}{k-1}g(z)-\bin{k}{k-2} g(ze^{-i{2\pi\over k}}) \cr g(ze^{{4\pi i \over k} })- \bin{k}{k-1} g(ze^{{2\pi i \over k} })+ \bin{k}{k-2} g(z)- \bin{k}{k-3} g(ze^{-{2\pi i \over k} })+\bin{k}{k-4} g(ze^{-i{4\pi \over k}}) \cr \cr \vdots \cr \cr (-1)^{{k\over 2}-1} \left[ g(ze^{i\pi -i{2\pi \over k} })- \bin{k}{k-1} g(ze^{i\pi -i{4\pi \over k}}) + ...- \bin{k}{2} g(ze^{-i\pi +i{2\pi i \over k}} ) \right] \cr }\right] $$ } \vskip 0.3 cm We give all the matrices of interest up to $k=10$. $S_{upper}$ is $PSP^{-1}$. $A$ stends for $A^{\beta}$, $A^{\prime}$ for $A^{\beta^{\prime}}$. $S^{\beta}=AS_{upper}A^T$, $S^{ \beta^{\prime}}= A^{\prime} S_{upper}^{-1} [A^{\prime}]^T$. $\sigma_{i,i+1}:=\beta_{i,i+1}^{-1}$ {\small ${\bf CP}^2$ $$ K_{12} = \pmatrix{ 1 & 0 & 0 \cr -3 & 1 & 0 \cr 0 & 0 & 1\cr } ~~~ K_{13} := \pmatrix{ 1 & 0 & 0 \cr 0 & 1 & 0 \cr -3 & 0 & 1\cr } ~~~ K_{32} := \left[ \matrix{ 1 & 0 & 0 \cr 0 & 1 & 3 \cr 0 & 0 & 1 \cr } \right] $$ $$ T := \left[ \matrix{ 0 & 0 & 1 \cr -1 & 3 & 3 \cr 0 & -1 & 0 \cr } \right] $$ $$ S := \left[ \matrix{ 1 & 0 & 0 \cr -3 & 1 & 3 \cr -3 & 0 & 1 \cr } \right] ~~~PSP^{-1}=\left[ \matrix{ 1 & 3 & -3 \cr 0 & 1 & -3 \cr 0 & 0 & 1 \cr } \right] $$ \vskip 0.2 cm ${\bf CP}^3$ $$ K_{42} := \left[ \matrix{ 1 & 0 & 0 & 0 \cr 0 & 1 & 0 & 6 \cr 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 1\cr } \right] ~~~ K_{43} := \left[ \matrix{ 1 & 0 & 0 & 0 \cr -4 & 1 & 0 & 0 \cr 0 & 0 & 1 & 4 \cr 0 & 0 & 0 & 1\cr }\right] $$ $$ T := \left[ \matrix{ 0 & 0 & 0 & 1 \cr -1 & 0 & 6 & 4 \cr 0 & -1 & 4 & 0 \cr 0 & 0 & -1 & 0 \cr } \right] ~~~ T_1 := \left[ \matrix{ 0 & 0 & 1 & 0 \cr -1 & 0 & 3 & 0 \cr 0 & -1 & 3 & 0 \cr 0 & 0 & 0 & 1 \cr } \right] $$ $$ S := \left[ \matrix{ 1 & 0 & 0 & 0 \cr -4 & 1 & 0 & 6 \cr 10 & -4 & 1 & -20 \cr -4 & 0 & 0 & 1 \cr } \right] $$ $$ {\it P } := \left[ \matrix{ 0 & 0 & 1 & 0 \cr 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 1 \cr 1 & 0 & 0 & 0 \cr }\right] ~~~ S_{upper}= \left[ \matrix{ 1 & -4 & -20 & 10 \cr 0 & 1 & 6 & -4 \cr 0 & 0 & 1 & -4 \cr 0 & 0 & 0 & 1 \cr } \right] $$ $$ A := \left[ \matrix{ 0 & 1 & 0 & 0 \cr 1 & 4 & 0 & 0 \cr 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 1 \cr } \right] ~~~A^{\prime}=\left[\matrix{ 1&0&0&0\cr 0&1&0&0\cr 0&0&-4&1\cr 0&0&1&0 \cr } \right] ~~~ S^{\beta}=S^{\beta^{\prime}} := \left[ \matrix{ 1 & 4 & 6 & -4 \cr 0 & 1 & 4 & -6 \cr 0 & 0 & 1 & -4 \cr 0 & 0 & 0 & 1 \cr } \right] $$ where $\beta=\beta_{12}$ \vskip 0.2 cm ${\bf CP}^4$ $$ K_{52} := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 \cr -5 & 1 & 0 & 0 & 0 \cr 0 & 0 & 1 & 0 & 10 \cr 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 1 \cr } \right] ~~~ K_{53} := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 \cr 0 & 1 & 0 & 0 & 10 \cr 0 & 0 & 1 & 5 & 0 \cr 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ T := \left[ \matrix{ 0 & 0 & 0 & 0 & 1 \cr -1 & 0 & 0 & 10 & 5 \cr 0 & -1 & 5 & 10 & 0 \cr 0 & 0 & -1 & 0 & 0 \cr 0 & 0 & 0 & -1 & 0 \cr } \right] ~~~ S= \left[ \matrix{ 1 & 0 & 0 & 0 & 0 \cr -5 & 1 & 0 & 0 & 10 \cr 15 & -5 & 1 & 5 & -40 \cr 40 & -10 & 0 & 1 & -95 \cr -5 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ P := \left[ \matrix{ 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 1 & 0 \cr 0 & 1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 1 \cr 1 & 0 & 0 & 0 & 0 \cr } \right] ~~~ S_{upper} := \left[ \matrix{ 1 & 5 & -5 & -40 & 15 \cr 0 & 1 & -10 & -95 & 40 \cr 0 & 0 & 1 & 10 & -5 \cr 0 & 0 & 0 & 1 & -5 \cr 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ A := \left[ \matrix{ 0 & 0 & 1 & 0 & 0 \cr 1 & 0 & 5 & 0 & 0 \cr 0 & 1 & 10 & 0 & 0 \cr 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 1 \cr } \right] ~~~ A^{\prime}:=\left[ \matrix{ 0 & 1 & 0 & 0 & 0 \cr 1 & 5 & 0 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & -5 & 1 \cr 0 & 0 & 0 & 1 & 0 \cr } \right] $$ $$ S^{\beta} =S^{\beta^{\prime}}:= \left[ \matrix{ 1 & 5 & 10 & 10 & -5 \cr 0 & 1 & 5 & 10 & -10 \cr 0 & 0 & 1 & 5 & -10 \cr 0 & 0 & 0 & 1 & -5 \cr 0 & 0 & 0 & 0 & 1 \cr } \right] $$ where $\beta=\beta_{23}\beta_{12}$, $\beta^{\prime}=\beta_{12}\sigma_{45}$. \vskip 0.2 cm ${\bf CP}^5$ $$ K_{62} := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 \cr 0 & 1 & 0 & 0 & 0 & 15 \cr 0 & 0 & 1 & 0 & 15 & 0 \cr 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] ~~~ K_{63} := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 \cr -6 & 1 & 0 & 0 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 & 20 \cr 0 & 0 & 0 & 1 & 6 & 0 \cr 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ T := \left[ \matrix{ 0 & 0 & 0 & 0 & 0 & 1 \cr -1 & 0 & 0 & 0 & 15 & 6 \cr 0 & -1 & 0 & 15 & 20 & 0 \cr 0 & 0 & -1 & 6 & 0 & 0 \cr 0 & 0 & 0 & -1 & 0 & 0 \cr 0 & 0 & 0 & 0 & -1 & 0 \cr } \right] ~~~ S := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 \cr -6 & 1 & 0 & 0 & 0 & 15 \cr 21 & -6 & 1 & 0 & 15 & -70 \cr -56 & 21 & -6 & 1 & -84 & 210 \cr 105 & -20 & 0 & 0 & 1 & -294 \cr -6 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ T_1 := \left[ \matrix{ 0 & 0 & 0 & 0 & 1 & 0 \cr -1 & 0 & 0 & 0 & 5 & 0 \cr 0 & -1 & 0 & 15 & 25 & 0 \cr 0 & 0 & -1 & 6 & 1 & 0 \cr 0 & 0 & 0 & -1 & -1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ P := \left[ \matrix{ 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 1 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 \cr 1 & 0 & 0 & 0 & 0 & 0 \cr } \right] $$ $$ A := \left[ \matrix{ 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 1 & 0 & 6 & 0 & 0 \cr 1 & 6 & 0 & 15 & 0 & 0 \cr 0 & 0 & 1 & 20 & 0 & 0 \cr 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] ~~~ A^{\prime}:= \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 \cr 0 & 1 & 0 & 0 & 0 & 0 \cr 0 & 0 & -20 & 1 & 0 & 0 \cr 0 & 0 & -15 & 0 & -6 & 1 \cr 0 & 0 & 6 & 0 & 1 & 0 \cr 0 & 0 & 1 & 0 & 0 & 0 \cr } \right] $$ $$ S^{\beta} := \left[ \matrix{ 1 & 6 & 15 & 20 & 15 & -6 \cr 0 & 1 & 6 & 15 & 20 & -15 \cr 0 & 0 & 1 & 6 & 15 & -20 \cr 0 & 0 & 0 & 1 & 6 & -15 \cr 0 & 0 & 0 & 0 & 1 & -6 \cr 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] ~~~S^{\beta^{\prime}} := \left[ \matrix{ 1 & 6 & 15 & 20 & -15 & -6 \cr 0 & 1 & 6 & 15 & -20 & -15 \cr 0 & 0 & 1 & 6 & -15 & -20 \cr 0 & 0 & 0 & 1 & -6 & -15 \cr 0 & 0 & 0 & 0 & 1 & 6 \cr 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ where $\beta= \beta_{12}(\beta_{34} \beta_{23} \beta_{12})$, $\beta^{\prime}= [(\sigma_{34}\sigma_{56})\sigma_{45}]\sigma_{56}$. \vskip 0.2 cm ${\bf CP}^6$ $$ K_{72} := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 1 & 0 & 0 & 0 & 0 & 21 \cr 0 & 0 & 1 & 0 & 0 & 35 & 0 \cr 0 & 0 & 0 & 1 & 7 & 0 & 0 \cr 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] ~~~ K_{73} := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 \cr -7 & 1 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 & 0 & 35 \cr 0 & 0 & 0 & 1 & 0 & 21 & 0 \cr 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ T := \left[ \matrix{ 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr -1 & 0 & 0 & 0 & 0 & 21 & 7 \cr 0 & -1 & 0 & 0 & 35 & 35 & 0 \cr 0 & 0 & -1 & 7 & 21 & 0 & 0 \cr 0 & 0 & 0 & -1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & -1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & -1 & 0 \cr } \right] ~~~ S := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 \cr -7 & 1 & 0 & 0 & 0 & 0 & 21 \cr 28 & -7 & 1 & 0 & 0 & 35 & -112 \cr -84 & 28 & -7 & 1 & 7 & -224 & 378 \cr -378 & 112 & -21 & 0 & 1 & -728 & 1638 \cr 224 & -35 & 0 & 0 & 0 & 1 & -728 \cr -7 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ P := \left[ \matrix{ 0 & 0 & 0 & 1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 1 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr 1 & 0 & 0 & 0 & 0 & 0 & 0 \cr } \right] $$ $$ A := \left[ \matrix{ 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 1 & 0 & 7 & 0 & 0 \cr 1 & 0 & 7 & 0 & 21 & 0 & 0 \cr 0 & 1 & 21 & 0 & 35 & 0 & 0 \cr 0 & 0 & 0 & 1 & 35 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] ~~~A^{\prime} := \left[ \matrix{ 0 & 1 & 0 & 0 & 0 & 0 & 0 \cr 1 & 7 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & -35 & 1 & 0 & 0 \cr 0 & 0 & 0 & -21 & 0 & -7 & 1 \cr 0 & 0 & 0 & 7 & 0 & 1 & 0 \cr 0 & 0 & 0 & 1 & 0 & 0 & 0 \cr } \right] $$ $$ S^{\beta} := \left[ \matrix{ 1 & 7 & 21 & 35 & 35 & 21 & -7 \cr 0 & 1 & 7 & 21 & 35 & 35 & -21 \cr 0 & 0 & 1 & 7 & 21 & 35 & -35 \cr 0 & 0 & 0 & 1 & 7 & 21 & -35 \cr 0 & 0 & 0 & 0 & 1 & 7 & -21 \cr 0 & 0 & 0 & 0 & 0 & 1 & -7 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] ~~~S^{\beta^{\prime}}:= \left[ \matrix{ 1 & 7 & 21 & 35 & 35 & -21 & -7 \cr 0 & 1 & 7 & 21 & 35 & -35 & -21 \cr 0 & 0 & 1 & 7 & 21 & -35 & -35 \cr 0 & 0 & 0 & 1 & 7 & -21 & -35 \cr 0 & 0 & 0 & 0 & 1 & -7 & -21 \cr 0 & 0 & 0 & 0 & 0 & 1 & 7 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ where $\beta= (\beta_{23}\beta_{12})( \beta_{45} \beta_{34} \beta_{23} \beta_{12})$, $\beta^{\prime}=\beta_{12}[(\sigma_{45}\sigma_{67})\sigma_{56}] \sigma_{67}$. \vskip 0.2 cm ${\bf CP}^7$ $$ K_{82} := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 1 & 0 & 0 & 0 & 0 & 0 & 28 \cr 0 & 0 & 1 & 0 & 0 & 0 & 70 & 0 \cr 0 & 0 & 0 & 1 & 0 & 28 & 0 & 0 \cr 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] ~~~ K_{83} := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr -8 & 1 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 & 0 & 0 & 56 \cr 0 & 0 & 0 & 1 & 0 & 0 & 56 & 0 \cr 0 & 0 & 0 & 0 & 1 & 8 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ T := \left[ \matrix{ 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr -1 & 0 & 0 & 0 & 0 & 0 & 28 & 8 \cr 0 & -1 & 0 & 0 & 0 & 70 & 56 & 0 \cr 0 & 0 & -1 & 0 & 28 & 56 & 0 & 0 \cr 0 & 0 & 0 & -1 & 8 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & -1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & -1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & -1 & 0 \cr } \right] $$ $$ S := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr -8 & 1 & 0 & 0 & 0 & 0 & 0 & 28 \cr 36 & -8 & 1 & 0 & 0 & 0 & 70 & -168 \cr -120 & 36 & -8 & 1 & 0 & 28 & -504 & 630 \cr 330 & -120 & 36 & -8 & 1 & -216 & 2100 & -1848 \cr -1512 & 378 & -56 & 0 & 0 & 1 & -3912 & 7476 \cr 420 & -56 & 0 & 0 & 0 & 0 & 1 & -1560 \cr -8 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ P := \left[ \matrix{ 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 1 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr } \right] ~~~ S_{upper} := \left[ \matrix{ 1 & -8 & -216 & 36 & 2100 & -120 & -1848 & 330 \cr 0 & 1 & 28 & -8 & -504 & 36 & 630 & -120 \cr 0 & 0 & 1 & -56 & -3912 & 378 & 7476 & -1512 \cr 0 & 0 & 0 & 1 & 70 & -8 & -168 & 36 \cr 0 & 0 & 0 & 0 & 1 & -56 & -1560 & 420 \cr 0 & 0 & 0 & 0 & 0 & 1 & 28 & -8 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & -8 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ A := \left[ \matrix{ 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 1 & 0 & 8 & 0 & 0 \cr 0 & 1 & 0 & 8 & 0 & 28 & 0 & 0 \cr 1 & 8 & 0 & 28 & 0 & 56 & 0 & 0 \cr 0 & 0 & 1 & 56 & 0 & 70 & 0 & 0 \cr 0 & 0 & 0 & 0 & 1 & 56 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] ~~~ A^{\prime}:= \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & -56 & 1 & 0 & 0 & 0 & 0 \cr 0 & 0 & -70 & 0 & -56 & 1 & 0 & 0 \cr 0 & 0 & -56 & 0 & -28 & 0 & -8 & 1 \cr 0 & 0 & 28 & 0 & 8 & 0 & 1 & 0 \cr 0 & 0 & 8 & 0 & 1 & 0 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 \cr } \right] $$ $$ S^{\beta} := \left[ \matrix{ 1 & 8 & 28 & 56 & 70 & 56 & 28 & -8 \cr 0 & 1 & 8 & 28 & 56 & 70 & 56 & -28 \cr 0 & 0 & 1 & 8 & 28 & 56 & 70 & -56 \cr 0 & 0 & 0 & 1 & 8 & 28 & 56 & -70 \cr 0 & 0 & 0 & 0 & 1 & 8 & 28 & -56 \cr 0 & 0 & 0 & 0 & 0 & 1 & 8 & -28 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & -8 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] ~~~ S^{\beta^{\prime}}:= \left[ \matrix{ 1 & 8 & 28 & 56 & 70 & -56 & -28 & -8 \cr 0 & 1 & 8 & 28 & 56 & -70 & -56 & -28 \cr 0 & 0 & 1 & 8 & 28 & -56 & -70 & -56 \cr 0 & 0 & 0 & 1 & 8 & -28 & -56 & -70 \cr 0 & 0 & 0 & 0 & 1 & -8 & -28 & -56 \cr 0 & 0 & 0 & 0 & 0 & 1 & 8 & 28 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 8 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ where $ \beta= ( \beta_{34} \beta_{23} \beta_{12}) \beta_{23} (\beta_{56} \beta_{45} \beta_{34}\beta_{23} \beta_{12})$, $\beta^{\prime}= [(\sigma_{34}\sigma_{56}\sigma_{78}) (\sigma_{45}\sigma_{67})\sigma_{56}] [(\sigma_{67}\sigma_{78})\sigma_{67}]$. \vskip 0.2 cm ${\bf CP}^8$ $$ K_{92} := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 36 \cr 0 & 0 & 1 & 0 & 0 & 0 & 0 & 126 & 0 \cr 0 & 0 & 0 & 1 & 0 & 0 & 84 & 0 & 0 \cr 0 & 0 & 0 & 0 & 1 & 9 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] ~~~ K_{93} := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr -9 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 & 84 \cr 0 & 0 & 0 & 1 & 0 & 0 & 0 & 126 & 0 \cr 0 & 0 & 0 & 0 & 1 & 0 & 36 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ T := \left[ \matrix{ 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr -1 & 0 & 0 & 0 & 0 & 0 & 0 & 36 & 9 \cr 0 & -1 & 0 & 0 & 0 & 0 & 126 & 84 & 0 \cr 0 & 0 & -1 & 0 & 0 & 84 & 126 & 0 & 0 \cr 0 & 0 & 0 & -1 & 9 & 36 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & -1 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & -1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & -1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & -1 & 0 \cr } \right] $$ $$ S := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr -9 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 36 \cr 45 & -9 & 1 & 0 & 0 & 0 & 0 & 126 & -240 \cr -165 & 45 & -9 & 1 & 0 & 0 & 84 & -1008 & 990 \cr 495 & -165 & 45 & -9 & 1 & 9 & -720 & 4620 & -3168 \cr 3168 & -990 & 240 & -36 & 0 & 1 & -3015 & 25740 & -19932 \cr -4620 & 1008 & -126 & 0 & 0 & 0 & 1 & -15867 & 25740 \cr 720 & -84 & 0 & 0 & 0 & 0 & 0 & 1 & -3015 \cr -9 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ P := \left[ \matrix{ 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr } \right] $$ $$ S_{upper} := \left[ \matrix{ 1 & 9 & -9 & -720 & 45 & 4620 & -165 & -3168 & 495 \cr 0 & 1 & -36 & -3015 & 240 & 25740 & -990 & -19932 & 3168 \cr 0 & 0 & 1 & 84 & -9 & -1008 & 45 & 990 & -165 \cr 0 & 0 & 0 & 1 & -126 & -15867 & 1008 & 25740 & -4620 \cr 0 & 0 & 0 & 0 & 1 & 126 & -9 & -240 & 45 \cr 0 & 0 & 0 & 0 & 0 & 1 & -84 & -3015 & 720 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 36 & -9 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & -9 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ A := \left[ \matrix{ 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 1 & 0 & 9 & 0 & 0 \cr 0 & 0 & 1 & 0 & 9 & 0 & 36 & 0 & 0 \cr 1 & 0 & 9 & 0 & 36 & 0 & 84 & 0 & 0 \cr 0 & 1 & 36 & 0 & 84 & 0 & 126 & 0 & 0 \cr 0 & 0 & 0 & 1 & 126 & 0 & 126 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 84 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] ~~~ A^{\prime}:= \left[ \matrix{ 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr 1 & 9 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & -126 & 1 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & -126 & 0 & -84 & 1 & 0 & 0 \cr 0 & 0 & 0 & -84 & 0 & -36 & 0 & -9 & 1 \cr 0 & 0 & 0 & 36 & 0 & 9 & 0 & 1 & 0 \cr 0 & 0 & 0 & 9 & 0 & 1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 \cr } \right] $$ $$ S^{\beta}= \left[ \matrix{ 1 & 9 & 36 & 84 & 126 & 126 & 84 & 36 & -9 \cr 0 & 1 & 9 & 36 & 84 & 126 & 126 & 84 & -36 \cr 0 & 0 & 1 & 9 & 36 & 84 & 126 & 126 & -84 \cr 0 & 0 & 0 & 1 & 9 & 36 & 84 & 126 & -126 \cr 0 & 0 & 0 & 0 & 1 & 9 & 36 & 84 & -126 \cr 0 & 0 & 0 & 0 & 0 & 1 & 9 & 36 & -84 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 9 & -36 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & -9 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] ~~~ S^{\beta^{\prime}}:= \left[ \matrix{ 1 & 9 & 36 & 84 & 126 & 126 & -84 & -36 & -9 \cr 0 & 1 & 9 & 36 & 84 & 126 & -126 & -84 & -36 \cr 0 & 0 & 1 & 9 & 36 & 84 & -126 & -126 & -84 \cr 0 & 0 & 0 & 1 & 9 & 36 & -84 & -126 & -126 \cr 0 & 0 & 0 & 0 & 1 & 9 & -36 & -84 & -126 \cr 0 & 0 & 0 & 0 & 0 & 1 & -9 & -36 & -84 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 9 & 36 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 9 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ where $\beta= (\beta_{45}\beta_{34}\beta_{23}\beta_{12})(\beta_{34}\beta_{23})(\beta_{67}\beta_{56} \beta_{45} \beta_{34} \beta_{23} \beta_{12})$, and $\beta^{\prime}=\beta_{12} [(\sigma_{45} \sigma_{67}\sigma_{89})(\sigma_{56}\sigma_{78}) \sigma_{67} ] [ (\sigma_{78} \sigma_{89})\sigma_{78}]$. ${\bf CP}^9$ $$ K_{10,2} := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 45 \cr 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 & 210 & 0 \cr 0 & 0 & 0 & 1 & 0 & 0 & 0 & 210 & 0 & 0 \cr 0 & 0 & 0 & 0 & 1 & 0 & 45 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ K_{10,3} := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr -10 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 120 \cr 0 & 0 & 0 & 1 & 0 & 0 & 0 & 0 & 252 & 0 \cr 0 & 0 & 0 & 0 & 1 & 0 & 0 & 120 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 10 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ T := \left[ \matrix{ 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr -1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 45 & 10 \cr 0 & -1 & 0 & 0 & 0 & 0 & 0 & 210 & 120 & 0 \cr 0 & 0 & -1 & 0 & 0 & 0 & 210 & 252 & 0 & 0 \cr 0 & 0 & 0 & -1 & 0 & 45 & 120 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & -1 & 10 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & -1 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & -1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & -1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & -1 & 0 \cr } \right] $$ $$ S := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr -10 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 45 \cr 55 & -10 & 1 & 0 & 0 & 0 & 0 & 0 & 210 & -330 \cr -220 & 55 & -10 & 1 & 0 & 0 & 0 & 210 & -1848 & 1485 \cr 715 & -220 & 55 & -10 & 1 & 0 & 45 & -1980 & 9240 & -5148 \cr -2002 & 715 & -220 & 55 & -10 & 1 & -440 & 10395 & -34320 & 15015 \cr 17160 & -4752 & 990 & -120 & 0 & 0 & 1 & -25190 & 177705 & -120120 \cr -11880 & 2310 & -252 & 0 & 0 & 0 & 0 & 1 & -52910 & 73755 \cr 1155 & -120 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & -5390 \cr -10 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ P := \left[ \matrix{ 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 \cr 0 & 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr } \right] $$ $$ S_{upper} := \left[ \matrix{ 1 & -10 & -440 & 55 & 10395 & -220 & -34320 & 715 & 15015 & -2002 \cr 0 & 1 & 45 & -10 & -1980 & 55 & 9240 & -220 & -5148 & 715 \cr 0 & 0 & 1 & -120 & -25190 & 990 & 177705 & -4752 & -120120 & 17160 \cr 0 & 0 & 0 & 1 & 210 & -10 & -1848 & 55 & 1485 & -220 \cr 0 & 0 & 0 & 0 & 1 & -252 & -52910 & 2310 & 73755 & -11880 \cr 0 & 0 & 0 & 0 & 0 & 1 & 210 & -10 & -330 & 55 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & -120 & -5390 & 1155 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 45 & -10 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & -10 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ A=\left[ \matrix{ 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 1 & 0 & 10 & 0 & 0 \cr 0 & 0 & 0 & 1 & 0 & 10 & 0 & 45 & 0 & 0 \cr 0 & 1 & 0 & 10 & 0 & 45 & 0 & 120 & 0 & 0 \cr 1 & 10 & 0 & 45 & 0 & 120 & 0 & 210 & 0 & 0 \cr 0 & 0 & 1 & 120 & 0 & 210 & 0 & 252 & 0 & 0 \cr 0 & 0 & 0 & 0 & 1 & 252 & 0 & 210 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 120 & 0 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ A^{\prime} := \left[ \matrix{ 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & -120 & 1 & 0 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & -210 & 0 & -252 & 1 & 0 & 0 & 0 & 0 \cr 0 & 0 & -252 & 0 & -210 & 0 & -120 & 1 & 0 & 0 \cr 0 & 0 & -210 & 0 & -120 & 0 & -45 & 0 & -10 & 1 \cr 0 & 0 & 120 & 0 & 45 & 0 & 10 & 0 & 1 & 0 \cr 0 & 0 & 45 & 0 & 10 & 0 & 1 & 0 & 0 & 0 \cr 0 & 0 & 10 & 0 & 1 & 0 & 0 & 0 & 0 & 0 \cr 0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \cr } \right] $$ $$ S^{\beta} := \left[ \matrix{ 1 & 10 & 45 & 120 & 210 & 252 & 210 & 120 & 45 & -10 \cr 0 & 1 & 10 & 45 & 120 & 210 & 252 & 210 & 120 & -45 \cr 0 & 0 & 1 & 10 & 45 & 120 & 210 & 252 & 210 & -120 \cr 0 & 0 & 0 & 1 & 10 & 45 & 120 & 210 & 252 & -210 \cr 0 & 0 & 0 & 0 & 1 & 10 & 45 & 120 & 210 & -252 \cr 0 & 0 & 0 & 0 & 0 & 1 & 10 & 45 & 120 & -210 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 10 & 45 & -120 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 10 & -45 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & -10 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ $$ S^{\beta^{\prime}} := \left[ \matrix{ 1 & 10 & 45 & 120 & 210 & 252 & -210 & -120 & -45 & -10 \cr 0 & 1 & 10 & 45 & 120 & 210 & -252 & -210 & -120 & -45 \cr 0 & 0 & 1 & 10 & 45 & 120 & -210 & -252 & -210 & -120 \cr 0 & 0 & 0 & 1 & 10 & 45 & -120 & -210 & -252 & -210 \cr 0 & 0 & 0 & 0 & 1 & 10 & -45 & -120 & -210 & -252 \cr 0 & 0 & 0 & 0 & 0 & 1 & -10 & -45 & -120 & -210 \cr 0 & 0 & 0 & 0 & 0 & 0 & 1 & 10 & 45 & 120 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 10 & 45 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 10 \cr 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \cr } \right] $$ where $\beta= (\beta_{56}\beta_{45}\beta_{34}\beta_{23}\beta_{12})(\beta_{45}\beta_{34} \beta_{23})\beta_{34} (\beta_{78}\beta_{67}\beta_{56}\beta_{45}\beta_{34}\beta_{23} \beta_{12}) $, \noindent and $\beta^{\prime}=[(\sigma_{34}\sigma_{56}\sigma_{78}\sigma_{9,10}) (\sigma_{45}\sigma_{67}\sigma_{89})(\sigma_{56}\sigma_{78})\sigma_{67}] [(\sigma_{78}\sigma_{89} \sigma_{9,10})(\sigma_{78}\sigma_{89})\sigma_{78}]$. }
\section{Introduction} The Palo Verde (Arizona) reactor neutrino experiment is a long baseline disappearance neutrino oscillation search~\cite{proposal}. Its goal is to test the $\nu_{\mu} \leftrightarrow \nu_e$ solution of the atmospheric neutrino anomaly~\cite{kamio_at} utilizing the inverse beta decay $\overline{\nu}_e + p \rightarrow e^+ + n$ as detection reaction. The Palo Verde detector is a finely segmented device minimizing correlated backgrounds by exploiting fast triple coincidences between detector elements. The positron energy deposit in coincidence with the annihilation quanta firing at least two neighboring elements identify an e$^+$. A delayed neutron capture signal completes the anti-neutrino signature.\\ The 11.3 tons of hydrocarbon based liquid scintillator (H:C$\approx$2) serve simultaneously as target and detector. The measured kinetic energy of the $e^+$, being stopped in the liquid scintillator, allows the reconstruction of the $\overline{\nu}_e$ energy. As the e$^+$ energy is rather low (1-5 MeV), a high light yield scintillator is vital for obtaining good energy resolution.\\ The correlated positron-neutron signature of the detection reaction is a powerful tool for the reduction of random backgrounds. The Gd loaded scintillator was chosen for the Palo Verde project because it offers two important advantages over pure hydrocarbon based scintillator: \begin{itemize} \item A large thermal neutron capture cross section of the isotopes $^{155,157}$Gd (61400 and 255000 barns) shorten the neutron capture time. For 0.1$\%$ Gd loading (by weight) the neutron capture time is $\tau \approx 30\;\; \mu s$, compared to $\tau \approx 180\;\; \mu s$ for capture on the proton in unloaded scintillators. \item A release of a high energy (8 MeV) gamma cascade after thermal neutron capture on Gd results in a neutron capture signal well above the radioactivity backgrounds. \end{itemize} The Palo Verde detector consists of 66 acrylic tanks ($900\; cm \times 12.7\; cm \times 25.4\; cm$) containing the liquid scintillator. Acrylic was chosen to obtain efficient light collection into the 5'' photomultiplier attached to either side of the tank by exploiting total reflection. This choice requires long term compatibility of the cells with the liquid scintillator and excludes the use of pure solvents like xylene (dimethylbenzene) or pseudocumene (1-2-4 trimethylbenzene, PC). Goal of the project was to formulate a liquid scintillator containing not more than 40$\%$ PC diluted with mineral oil and dissolve the Gd in this mixture in a way compatible with the acrylic. Early developments of this cocktail were done in collaboration with NE Technology Ltd. The relatively large tank length places strict requirements on the transparency of the scintillator.\\ Below we will discuss the scintillator formulation chosen to satisfy these requirements. We will present our measurements of the scintillator properties and, finally, comment on the stability of the liquid.\\ The technical details given in this paper might be of interest also for other projects; for example large (100 tons) liquid scintillation solar neutrinos detectors, loaded with Gd or Yb, are being discussed~\cite{ragha}. \section{Scintillator Formulation} To assure compatibility of the scintillator with the acrylic tanks a mixture of pseudocumene and mineral oil was chosen. As Gd dissolves in neither of the above liquids it was necessary to prepare a soluble and acrylic-compatible Gd compound. The compound prepared was gadolinium 2-ethylhexanoate [Gd(CH$_3$(CH$_2$)$_3$CH(C$_2$H$_5$)CO$_2$)$_3 \cdot $XH$_2$O]. It can be synthesized from gadolinium chloride, oxide, or nitrate. Early preparations in the project were made from the nitrate, since a sufficiently radio-pure source was available. However, solutions from the nitrate were found to be unstable (solid-liquid phase separation) when handled, especially when exposed to air. The oxide, Gd$_2$O$_3$, was converted to the 2-ethylhexanoate in an aqueous reaction, collected by filtration, and dried. The resultant compound was purity tested and then dissolved in a scintillation solvent, along with a primary fluorescent additive, a spectrum shifter, an antioxidant, and small amounts of two additional solvents to help keep the gadolinium compound in solution. The emission peaks of the primary fluor and spectrum shifter were 365 nm and 425 nm, respectively.\\ Pseudocumene was the solvent of choice, for its high scintillation efficiency, lower solvency toward acrylic, and higher flash point than xylene. To increase the hydrogen content, and further decrease the solvent attack on the acrylic tanks, a high purity mineral oil was added as a final component of the finished scintillator solution. The solution is optimized at 60$\%$ oil by volume as this is the minimum necessary to dilute the pseudocumene enough that it doesn't harm the acrylic tank. A higher concentration of oil than this decreases the scintillation light yield, since this is a function of pseudocumene content (an aromatic liquid which inherently scintillates, whereas the oil is a saturated hydrocarbon and does not). Also, the higher the oil content, the lower the possible Gd content, as this also is a function of the pseudocumene solvent content.\\ We found that the fully blended scintillator, contained in 200 l steel drums did not remain stable (a single phase liquid) during the transport over ground via truck from Ohio to Arizona. To solve this problem, pseudocumene based ``concentrate'' was blended in Ohio and additional pseudocumene and the mineral oil was blended on site in Arizona. This procedure is described below. \section{Scintillator Blending} The Gd ``concentrate'' (BC-521C) as prepared by Bicron at their Ohio factory was shipped to Palo Verde in 200 l stainless steel drums. There it was blended with the mineral oil (Witco Scintillator Fluid), and 1,2,4-trimethylbenzene (Koch Chemicals obtained through Bicron Inc.). The proportions used were 1:1:3 by volume for the Gd concentrate, PC and mineral oil to arrive at a Gd loading of 0.1$\%$ by mass for the blended scintillator. Special care was taken to select hardware being compatible with the used solvents. As the concentrate and mineral oil do not mix well (rapid addition of the mineral oil may lead to irreversible precipitation of the Gd compound) care was taken to mix those components in a controlled way.\\ The blending was done in two 200 l stainless steel drums equipped with air driven stirrers having stainless steel shafts and impellers. With each drum processing about 170 kg of scintillator; the load of one acrylic cell. The liquids were moved by air driven diaphragm pumps (Nylon body and Teflon diaphragm). Separate pumps were used for the mineral oil, the two other components and the blended scintillator. All piping to come in contact with pure Gd concentrate or PC was made from Teflon while PVDF valves were used. For the mineral oil polypropylene tubing was used. The blending drums were placed on electronic scales for a precise determination of the scintillator weight.\\ For the blending each component was first pumped into an intermediate stainless steel vessel placed on an electronic scale to determine its weight and then drained by gravity into the mixing drum. First the concentrate and PC were mixed. Then the mineral oil was added at a rate not exceeding $\sim$3 l/min. During the blending process the liquid was regularly inspected for precipitation of the Gd compound. The stirrers were operated at a low speed to avoid excessive agitation of the liquid. The final mixture was stirred for at least 1 hour. The liquid was then pumped through a 200 nm Gelman ``HiFLO Sol-Vent DCF'' filter directly into the acrylic tank. The filter was changed after every batch of scintillator. A 2 l sample was retained for every batch blended.\\ The liquid scintillator could not be bubbled with nitrogen for removal of dissolved oxygen as this bears the risk to destabilize the mixture leading to precipitation of the Gd compound. The acrylic tanks were flushed with Ar before the filling in order to remove any residual vapors left over from the bonding and to displace the air. In addition, all tanks were pressure tested prior to filling to detect possible leaks.\\ A total of 11343 kg of liquid scintillator was prepared in this way. \section{Scintillator Quality} The following scintillator properties were routinely tested for at least one sample of blended scintillator for every batch of Gd concentrate: \begin{itemize} \item Light attenuation length at 440 nm \item Light yield \item Gd loading \end{itemize} Achieving an acceptable light attenuation length was the most challenging task during the scintillator development. Before working with production size samples, discussed below, we went through four generations of prototypes. In this prototype phase emphasis was given to the optimization of the solvent balance and quality of the raw ingredients. While the light yield and Gd loading were satisfactory from the beginning, early samples did exhibit problems in their transparency.\\ In the following we will discuss our data on the scintillator properties. \subsection{Attenuation Length} \begin{figure}[htbp] \centering \leavevmode \mbox{\psfig{file=al_scint.eps,height=10cm,clip=}} \caption{Light attenuation length at 440 nm as measured for cleaning fluid (CF, star), NE scintillator (NE, circle) and Bicron scintillator (BC, square). The scintillator batches are identified by successive numbering, as given in the figure. All batches to the right of the vertical line were blended in Palo Verde, while the rest was mixed at Caltech or Bicron. } \label{al} \end{figure} The light attenuation length was measured using a vertically oriented 1.5 m long, 3 cm diameter stainless steel tube filled with liquid. A blue LED equipped with a 440 nm wavelength filter and focussed to parallel by means of a 7-cm-focal-length lens, placed one focal length away from a pinhole aperture, illuminates a phototube at the other end of the tube. The LED was operated in pulsed mode to minimize background through scintillation light. The average pulse hight, seen by the phototube, was then measured for varying liquid hight and fitted to an exponential to obtain the attenuation length (scattering plus absorption).\\ Figure~\ref{al} shows the results obtained through those measurements. The first entry labeled CF denotes a measurement done with a batch of fluid composed of 36$\%$ PC and 64$\%$ mineral oil mixed in the hardware described before. The good attenuation length obtained demonstrated that all used vessels and plumbing had the necessary purity before starting to work with the scintillator. This liquid was also used to test one of the acrylic tanks for surface impurities.\\ The batches labeled NE in figure~\ref{al} were early prototypes made by NE Technology Ltd. The first batches blended by Bicron (labeled BC1-4 in figure~\ref{al}) before transportation showed a rather high light absorption. The fact that samples BC5 and BC6, blended at Caltech from Gd concentrate like the NE samples NE1-8, were of much better transparency shows that the blended scintillator is too fragile for a long transport. After switching to the on-site blending attenuation lengths consistently larger than 10 m were achieved for the bulk of the scintillator. For the first three drums of Gd concentrate, the light attenuation length of every scintillator batch was tested (BC7-20), for the later samples only one per concentrate batch. The average attenuation length for samples blended from BC5 on was 11.4 m. \begin{figure}[htbp] \centering \leavevmode \mbox{\psfig{file=ly_scint.eps,height=10cm,clip=}} \caption{Light yield of different batches of Gd scintillator. } \label{ly} \end{figure} \subsection{Light Yield} The light yield of the scintillator was measured using a $^{207}$Bi conversion electron source, irradiating a 2'' diameter Petri dish containing 14 ml of the sampled scintillator. The resulting liquid hight of 7 mm was chosen to stop the $\sim$1 MeV conversion electrons (range $\sim$5 mm) within the liquid while optimizing the solid angle towards the photomultiplier. The sample was placed on a 5'' photomultiplier. To monitor PMT stability a 1'' diameter NaI detector, separated by a thin Pb shield from the source, registering its $\gamma$-radiation was placed next to the sample onto the same PMT and both spectra were collected in parallel. The position of the $^{207}$Bi conversion electron peak in the measured spectrum is a measure of the average pulse hight and hence the light yield. A mixture of PC with 4 g/l PPO and 100 mg/l bisMSB, deoxiginated by bubbling with Ar before the measurement, served as a light yield standard. We assume a light yield of 80$\%$ antracene for this mixture. Figure~\ref{ly} shows the measured light yield of scintillator samples. The light yield was verified for at least on scintillator batch per concentrate lot. The light yield was consistently 56$\%$ antracene for the 22 tested samples. \subsection{Gd Loading} The Gd concentration of the liquid was measured using x-ray fluorescence after thermal neutron capture on Gd, as described in~\cite{novikov}. The measurements were done in comparison to standard solutions of Gd(NO$_3$)$_3$ dissolved in alcohol. The absolute Gd concentration of one scintillator batch was verified by mass spectroscopy in a commercial laboratory. In addition to this cross check a few early samples were also tested by neutron activation of $^{158}$Gd, using a $^{252}$Cf spontaneous fission source placed in a water moderator. The 363 keV $\gamma$-radiation following the $\beta$-decay of $^{159}$Gd into $^{159}$Tb (T$_{1/2}$=18.6 h) was detected with a Ge detector.\\ \begin{figure}[htbp] \centering \leavevmode \mbox{\psfig{file=gd_scint.eps,height=10cm,clip=}} \caption{$\%$ Gd loading (by mass) of scintillator batches. } \label{gd} \end{figure} As for the light yield at least one scintillator batch per concentrate lot was tested for its Gd loading to make sure that the detector is of reasonable uniformity. As can be seen in figure~\ref{gd} the Gd loading was consistently 0.1$\%$. The small scattering shows that the production process of the concentrate is well controlled. The low value of NE2 was confirmed by the observation of Gd precipitation in the transport drum. The solvent balance of all following batches was readjusted after this observation. For prototype samples kept at Caltech for several years no instability of the Gd loading was observed. The described loading scheme is quite stable if the liquid is not disturbed or agitated. \section{Scintillator Stability} \begin{figure}[htbp] \centering \leavevmode \mbox{\psfig{file=stab_scint.eps,height=10cm,clip=}} \caption{Time development of the light attenuation length for Gd scintillator batches. Samples have been taken from the acrylic tank. } \label{stab} \end{figure} As neither light yield nor Gd loading showed any observable degradation over time we shall concentrate in the following on the time development of the scintillator transparency which is subject to aging. To test the stability of the transparency, samples from four scintillator batches have been taken from the acrylic tanks over time to study the transparency loss of the scintillator in its experimental environment. Figure~\ref{stab} gives the results of these repeated measurements. In all cases an initial decline stabilized, resulting in a loss rate of about 2 cm/day. This rate of aging is acceptable for our experiment. The early samples BC1-4 (see fig.~\ref{al}) showed a much more rapid loss of transparency. It is interesting to note that the aging process starts only after the blending. Test batches blended from concentrate retains exhibit the same attenuation length as the original liquid sample. The aging process is sped up by subjecting the liquid to elevated temperatures. In situ measurements made for all scintillator batches by scanning the detector elements with a $\gamma$ source confirm that the light attenuation has acceptable stability in time. \section{Conclusion} The formulation of a pseudocumene based 0.1$\%$ Gd loaded liquid scintillator having good long term stability and compatibility with acrylic has been reported. The measurements done during development and deployment of the scintillator have been discussed. The details of the scintillator preparation, which in our view were important for obtaining scintillator of high transparency and stability, were reported. {\bf Acknowledgement} The authors want to thank Prof. F. Boehm for his continued support and numerous stimulating discussions. We also want to acknowledge the efforts of Dr. H. Hunter who participated in the early development work and continued to help the project even after NE Technology Ltd. was no longer involved. The support of the Palo Verde Collaboration and the active help of S. Beckman, Prof. J. Busenitz, Prof. G. Gratta and Dr. J. Wolf was very much appreciated. Early development efforts were mainly driven by Prof. M. Chen and Dr. R. Hertenberger. Finally we acknowledge the excellent work of the Bicron/SGIC laboratory technicians Michael Berkley and Tim Hill who were responsible for the preparation of the Gd concentrate.
\section{Introduction}\label{sec:fss-1} In recent years off-equilibrium dynamics, particularly aging dynamics, of spin glasses has attracted much attention theoretically,\cite{BCKM} numerically,\cite{MPR} and experimentally.\cite{EXP,Nordblad,Weissman} Many experiments have confirmed that the response to small dc/ac magnetic field after quenching the system below the spin-glass transition temperature $T_{\rm c}$ shows aging effects persistently at least up to the largest possible time scale available in laboratories. On the side of theories, one of the most notable advances is the analytical theory\cite{BCKM,CK,FM} based on solvable mean-field spin-glass models, which has provided non-trivial predictions on aging effects. Presumably one expects that the slow relaxation is caused by some underlying collective spin excitations in the system. In the case of ordering processes in non-random systems like the conventional Ising ferromagnet, the relaxational dynamics after quenching the system below the phase transition temperature is related to coarsening of domain walls which separate ordered domains of two different pure states.\cite{Bray-94} The typical size of the domains, $R(t)$, at time $t$ usually grows by a certain power law. This behavior can be directly studied, for instance by scattering experiments. The intriguing question is if such a picture of the coarsening of domain walls also holds for aging processes in spin glasses. Obviously, the above mentioned mean-field theory, which is exact only in infinitely large dimension, is not helpful in this respect. The domain picture based on the so-called droplet theory,\cite{BM-84,FH-88-EQ,FH-88-NE,Huse-91} has provided an interesting phenomenology. However the pure states of spin glasses cannot be found in practice which makes it difficult to rationalize the domain picture in a direct way. In the present study,~\cite{Komori-D} we mainly focus on the energy relaxation in the 3-dimensional EA spin-glass model during the aging process simulated by Monte Carlo dynamics. The latter starts from arbitrary initial configurations at temperature $T$ below $T_{\rm c}$ which simulates instantaneous quench to $T$ from $T=\infty$. We have found that the relaxation of energy can be explained consistently in terms of coarsening of domain walls as in the case of usual phase ordering processes in non-random systems. Let us briefly describe the idea here. Suppose a system is aged by elapsed time $t$ after the quench, and a typical size of domain walls has grown to $R(t)$. In the presence of domain walls, the energy per spin, $e_{T}(t)$, is larger than the equilibrium energy $e_{T}^{(\infty)}$ at temperature $T$ by an amount \begin{equation} \delta e_{T}(t) \equiv e_{T}(t)-e_{T}^{(\infty)} \propto \frac{\Upsilon{}(T)(R(t)/l_{0})^{\theta}}{(R(t)/l_{0})^{d}}. \label{eq:fss-1-1} \end{equation} with $l_0$ being a certain microscopic unit of length. Here the numerator on the right-hand side is the domain wall energy where $\Upsilon{}(T)$ is the stiffness constant and $\theta$ is the characteristic exponent. The denominator is the volume of a domain with $d$ being the dimension of the space. We have studied systematically the energy per spin on relatively small system sizes $L$ and at relatively higher temperatures in the spin-glass phase. This enables us to see the behaviors of two time regimes within our computational time window; the size-independent behavior of domain growth in the dynamic regime $R(t) < L$ and the size-dependent equilibrium behavior in the static regime $L < R(t)$. It has turned out that the simulated data including in the crossover region between the two regimes can be concisely described by a finite-size-scaling (FSS) function. By multiplying eq.~(\ref{eq:fss-1-1}) by a scaling function $\tilde{\delta e}(R(t)/L)$, the scaling ansatz is proposed as \begin{equation} \delta e_{T}(t) = \Upsilon(T) \left(\frac{R(t)}{l_0}\right)^{\theta -d} \tilde{\delta e}\left( \frac{R(t)}{L}\right), \label{eq:fss-tmp} \end{equation} with $R(t)=t^{1/z(T)}$. The FSS analysis allows us to obtain the characteristic exponent $\theta$ for the domain wall energy. The value of $\theta$ turns out to be nearly equal to the result by Bray and Moore at $T=0$.\cite{BM-84} These results strongly support that the aging process in the present spin-glass model is described well in terms of coarsening of domain walls, i.e., the droplet picture. As described just above the FSS analysis on the energy relaxation yields the power-law growth in $t$ of the typical domain size $R(t)$ with the exponent $z(T)$, which is consistent with that obtained by the previous numerical studies using replica-overlap.\cite{Huse-91,Kisker-96,Marinari-98-VFDT} We have also performed analysis on the replica-overlap with greater statistical accuracy, and have established the consistency between the two analyses. A fundamental assumption of the droplet theory on aging processes\cite{FH-88-NE} is that coarsening of domain walls, in case of spin glasses, is driven by successive nucleation of thermally activated droplets. In order to test this assumption, we have studied largest relaxation times $\tau_{L}(T)$, which are those needed for a global spin-flip in each sample with relatively small sizes $L$. We have found that the $L$-dependence of $\tau_{L}(T)$ is consistent with the growth law of domains, i.e., $R(t \sim \tau_{L}(T)) \sim L$. Furthermore, the temperature dependence follows in fact an Arrhenius law. In our simulated results, however, there remain some discrepancies with the original droplet theory. Firstly, it is assumed to grow as $L^\psi$ in the droplet theory, while the barrier free energy simulated grows logarithmically with $L$. This results agree with the previous conjectures extraxted from the $t$-dependence of $R(t)$.\cite{Kisker-96,Rieger93}. Secondly, the width of the distribution of the barrier free energy does not increase with $L$. The present paper is organized as follows. After describing the model and numerical method in the next section, we present and discuss the results of our simulation in \S~\ref{sec:fss-3}. Section~\ref{sec:fss-4} is devoted to the concluding remarks on the present work. \section{Model and Numerical Method} \label{sec:fss-2} We carry on simulations on the 3D EA Ising spin-glass model with nearest-neighbor interactions $\{J_{ij}\}$ defined by the Hamiltonian, \begin{equation} H = -\sum_{\langle ij \rangle}J_{ij}S_i S_j, \label{eq:fss-2-1} \end{equation} on an $N=L\times L\times L$ simple cubic lattice. The periodic boundary conditions are adopted. The quenched random interaction $\{ J_{ij} \}$ are drawn from a Gaussian distribution with zero mean and variance one. It has been reported that this model has a spin-glass ordered phase below the critical temperature $T_{\rm c}$ whose most recent value is reported as $T_{\rm c}=0.95\pm 0.04$.\cite{Marinari-cd98-PS} We use the single-spin-flip heat-bath Monte Carlo method. As a unit time, which we call one Monte Carlo Step (MCS) hereafter, we adopt the time in which $N$ spins are updated. The energy per spin $e_{T}(t)$ at $t$ MCS after the temperature quench is calculated as \begin{equation} e_{T}(t) = \frac{1}{N}\left[ \frac{1}{2t_{\rm ta}+1} \sum_{\tau=t-t_{\rm ta}}^{t+t_{\rm ta}}H(\tau) \right]_{\rm av}, \label{eq:fss-2-6} \end{equation} where $H(\tau)$ is a value of Hamiltonian at $\tau$ MCS. In order to reduce thermal fluctuations we take here an average over short time $2t_{\rm ta}+1$ around $t$ with $t_{\rm ta}=t/1000$. The bracket $[\cdots]_{\rm av}$ indicates the average over $N_{\rm s}$ Monte Carlo runs with independent realizations of quenched random variables $\{J_{ij}\}$, initial spin configurations, and random numbers. We have simulated systems with relatively small sizes ($L=4 \sim 10$) in order to study the crossover between the dynamic and static regimes as mentioned in \S 1, as well as large systems with $L=24$ and $32$ in order to obtain bulk behavior which is not affected by finite size effects within our time window ($\sim 10^{6}$ MCS). Following the previous works,\cite{Huse-91,Kisker-96,Marinari-98-VFDT} the length scale of ordered domains $R(t)$ is estimated by analyzing the replica-overlap function, $G(r,t)$, defined as \begin{equation} G(r,t) = \frac{1}{N} \sum_{i=1}^{N} \left[ S_i^{(\alpha)}(t)S_i^{(\beta)}(t) S_{i+r}^{(\alpha)}(t)S_{i+r}^{(\beta)}(t) \right]_{\rm av}. \label{eq:fss-2-2} \end{equation} Here $\alpha$ and $\beta$ are indices of the two replicas which have different random spin configurations at $t=0$ (an instant of the temperature quench), and are updated independently. Its correlation length is considered to be proportional to $R(t)$. In eq.~(\ref{eq:fss-2-2}) $r$ is a spatial distance which we take only along the directions of the lattice axes. We have simulated large systems with $L=24$ and $32$ in this analysis. \section{Results and Discussions}\label{sec:fss-3} \subsection{Finite-size scaling of $e_{T}(t)$ }\label{subsec:fss-3-2} Let us begin with analysis on the relaxation of the energy per spin $e_{T}(t)$ of large systems with $L=32$ and $24$ after the quench. The log-log plots of $\delta e_{T}(t)\equiv e_{T}(t)-e_{T}^{(\infty)}$ vs. $t$ are shown in Fig.~\ref{fig:fss-3}, where $e_{T}^{(\infty)}$ denotes lim$_{t \rightarrow\infty}$lim$_{L \rightarrow \infty}e_T{}(t)$. We fitted the data to the following formula,\cite{Marinari-98-VFDT} \begin{equation} \delta e_{T}(t) \equiv e_{T}(t) - e_{T}^{(\infty)} = c t^{-\lambda{}(T)}, \label{eq:fss-3} \end{equation} with $e_{T}^{(\infty{})}$, $c$ and $\lambda$ being the fitting parameters. From eq.~(\ref{eq:fss-1-1}) with $R(t)\propto t^{1/z(T)}$ we obtain \begin{equation} \lambda = (d-\theta)/z. \label{eq:fss-scaling-relation} \end{equation} As seen in Fig.~\ref{fig:fss-3}, the fitting works very well. The values of exponent $\lambda$ depend on temperature. It is proportional to $T$ at low temperatures in agreement with the previous work\cite{Marinari-98-VFDT}. Below we determine $z(T)$ and check eq.~(\ref{eq:fss-scaling-relation}) numerically. \begin{figure} \leavevmode\epsfxsize=85mm \epsfbox{fss-e.eps} \caption{The power law decay of excess energy per spin $\delta e_{T}(t)$ in the dynamic regime. The constants $e_{T}^{(\infty)}$ are obtained by fitting the simulated date to eq.~(\ref{eq:fss-3}). The data are at $T=0.4, 0.5, 0.6, 0.7, 0.8$ from top to bottom. The systems simulated are with $L=32$ and $N_{\rm s}=1600$ for $T=0.8$ and $0.7$, and with $L=24$ and $N_{\rm s}=3200$ for $T=0.6 \sim 0.4$. They are sufficiently large to obtain size-independent (bulk) energy decay.} \label{fig:fss-3} \end{figure} \begin{figure} \leavevmode\epsfxsize=85mm \epsfbox{raw-Le.eps} \caption{ The decay of excess energy per spin $\delta e_T (t)$ for systems with $L=4$ (circles), $5$ (triangles), $6$ (squares), $7$ (solid circles), and $32$ (solid triangles). The asymptotic energy per spin $e_{T}^{(\infty)}$ is determined as $-1.66204$ by fitting data of $L=32$ to eq.~(\ref{eq:fss-3}). The sample numbers $N_{\rm s}$ are $10^5$ for $L=4$, $5$, $6$ , $3.5\times 10^5$ for $L=7$, and $640$ for $L=32$. } \label{fig:fss-4} \end{figure} In Fig.~\ref{fig:fss-4} we show the log-log plot of $\delta e_{T}(t)$ vs. $t$ at $T=0.7$ for small systems with $L=4,5,6,7$. It is seen that there exist two time regions of the relaxation curves. In the shorter time region the curves have no size dependence (the dynamic regime), while in the longer time region the curves depend on $L$ and $e_{T}(t)$ approaches to an equilibrium value $e_{T}^{(L)}$ which also depends on $L$ (the static regime). The crossover between the two regimes is expected to occur when a typical size of ordered domain $R(t)$ reaches $L$. We therefore have tried some finite-size-scaling (FSS) analyses based on eq.~(\ref{eq:fss-tmp}). It has turned out that the following FSS ansatz works best: \begin{eqnarray} \delta e_{T}(t) & = & e_{T}(t)-e_{T}^{(\infty)}=L^{\theta-d}\tilde{f}(t/L^{z(T)}), \label{eq:fss-5}\\ \tilde{f}(x) & = & \left\{ \begin{array}{lll} c x^{-\lambda{}(T)} & \mbox{for} & x \ll 1,\\ \Upsilon'(T) &\mbox{for} & x\gg 1, \end{array}\right. \label{eq:fss-6} \end{eqnarray} where $\Upsilon{}'{}(T)$ is a constant proportional to $\Upsilon{}(T)$ in eq.~(\ref{eq:fss-1-1}). In other words, the above FSS function has the following asymptotic limits: \begin{equation} e_{T}(t) - e_{T}^{(\infty)} = \left\{ \begin{array}{ll} \Upsilon{}(T) R(t)^{\theta-d} & \mbox{for dynamic regime},\\ \Upsilon'{}(T) L^{\theta-d} & \mbox{for static regime}, \end{array}\right. \label{eq:fss-3-2} \end{equation} with $R(t)$ being proportional to $t^{1/z(T)}$. In practice, we have performed the above FSS fit in the following way. By making use of $e_{T}^{(\infty)}$ evaluated above using the bulk data, and expecting that $\delta e_{T}(t)$ varies as a function of $x\equiv\log t$ and $y\equiv\log L$, we fitted the data of $\delta e_{T}(t)$ to eq.~(\ref{eq:fss-5}) by means of the following trial function with ten parameters, \begin{eqnarray} \log(e_{T}(t) - e_{T}^{(\infty)}) & = & (\theta-d) \log L + \log\tilde{f}(t/L^{z(T)})\\ & \simeq & (\theta-d) y + \sum_{n=0}^{7}\frac{c_n}{n!} \left(x-z(T)y\right)^n. \label{eq:fss-2-7} \end{eqnarray} Here we have expanded logarithm of $\tilde{f}$ with respect to logarithm of its argument up to the seventh order. The above fitting has been done for 10 independent $e_{T}(t)$, each of which is obtained by averaging over $N_{\rm s}/10$ samples. We have then estimated errors of exponent $\theta$ and $z(T)$ from its variance over the 10 independent values. \begin{figure} \leavevmode\epsfxsize=90mm \epsfbox{fit.eps} \caption{ The scaling plot of $\delta e_T(t)$ versus $t/L^{z(T)}$.} \label{fig:fss-5} \end{figure} The results of the above FSS are shown in Fig.~\ref{fig:fss-5}. We see that the scaling works quite satisfactorily. The similar results are obtained also at $T=0.8$, for which we have examined systems with $L=4,6,8,10$ and $N_{\rm s}$ up to $130000$. The exponents obtained are listed on Table~\ref{tbl:theta} for $\theta$ and Table~\ref{tbl:z} for $z(T)$. The values of $z(T)$ are consistent with those extracted from $R(t)$ in the larger systems, as will be discussed in the next subsection. We therefore consider that the present FSS with eqs.~(\ref{eq:fss-5}) and~(\ref{eq:fss-6}) is quite reasonable. The exponent $\theta$ seems almost independent on temperature as expected. Actually, its values derived by the present FSS at $T=0.7$ and $0.8$ are almost equal to the droplet exponent $\theta=0.19\pm 0.01$ extracted from the defect energy analysis at $T=0$ by Bray and Moore.\cite{BM-84} This result in the dynamic regime confirms the domain coarsening picture described in \S 1. It is worth noting that static energy of finite size systems is higher than that of the infinite one by an amount $L^{\theta}$, that is, \begin{equation} \left({L \over l_0}\right)^{d}e_{T}(t) = \left({L \over l_0}\right)^{d}e_{T}^{(\infty)} + \Upsilon'(T) \left({L \over l_0}\right)^{\theta}. \end{equation} This is interpreted that the imposed periodic boundary conditions distort the spin correlations with respect to what would be realized in an infinite system.\cite{Newman-98} A similar and more transparent situation can be seen in a pure ferromagnet with the {\it anti}-periodic boundary condition. The latter induces competition or frustration in the loop of the interactions winding the whole system. Clearly it leaves a domain wall even in equilibrium. The FSS of eqs.~(\ref{eq:fss-5}) and~(\ref{eq:fss-6}) works well also for this case and the well known exponents, $z=1/2$ and $\theta=d-1$ are deduced. The situation in the present spin-glass model is more complicated since we cannot separate the effects of frustrations in bulk and of the constraint by the boundary conditions. However, the near agreement of the exponents $\theta$ at $T=0.7$ and $0.8$ with the one at $T=0$ suggests that the periodic boundary condition alone introduces a `defect energy' similar to the one at $T=0$ discussed by Bray and Moore. In this respect let us note that a similar algebraic size dependence of finite size correction to the ground state energy with the periodic boundary condition has been reported recently.\cite{Hartmann-97} \subsection{Domain growth after quench}\label{subsec:fss-3-1} Here we examine the replica-overlap function $G(r,t)$ of eq.~(\ref{eq:fss-2-2}) and the domain growth $R(t)$ after the quench. Our analysis follows the work of Kisker et al.\cite{Kisker-96} We use an improved method of evaluating $R(t)$ from $G(r,t)$ as explained below. Because of the periodic boundary conditions, relation $G(r,t)=G(L-r,t)$ holds, and so $R(t)$ is evaluated as\cite{Cooper-82} \begin{equation} R(t) = 2\cdot\frac{L}{2\pi} \sqrt{\frac{\tilde{G}(0,t)}{\tilde{G}(2\pi/L,t)}-1}, \label{eq:fss-2-3} \end{equation} where \begin{equation} \tilde{G}(k,t) = 2\int_{0}^{L/2}G(r,t)\cos{}kr \mbox{\rm d}r. \label{eq:fss-2-4} \end{equation} If $G(r,t)$ is simply proportional to $\cosh[2(r-L/2)/R(t)]$ (or $\exp(-2r/R(t))$ in the limit $L \rightarrow \infty$), the estimation of eq.~(\ref{eq:fss-2-3}) gives $R(t)$ exactly. We consider that our estimation of $R(t)$ is superior to the method based on the moments $\int_0^{L/2}r^nG(r){\rm d}r$ with $n=0,1$ used by Kisker et al. since our estimation significantly reduces the irrelevant contribution from small $r$ which deviate from an exponential law. \begin{figure} \leavevmode\epsfxsize=85mm \epsfbox{fss-xi.eps} \caption{ The length scale $R(t)$ of ordered domains estimated from eq.~(\ref{eq:fss-2-3}) for temperatures $T=0.4, 0.5, 0.6, 0.7,$ and $0.8$ from bottom to top. The solid lines are fitting lines $R(t) \propto t^{1/z(T)}$.} \label{fig:fss-yobun} \end{figure} \begin{figure} \leavevmode\epsfxsize=85mm \epsfbox{scale-gr.eps} \caption{ Logarithms of the replica-overlap function $G(r,t)$ versus $r/t^{1/z(T)}$ at $t=5^{n}$ ($n=2,\dots,8$) and $r>3$ for $T=0.7$ with $L=32$ and $N_{\rm s}=3500$. The solid line is only a guide to the eye.} \label{fig:fss-1} \end{figure} The averaged domain size $R(t)$ estimated from eqs.~(\ref{eq:fss-2-3})~and~(\ref{eq:fss-2-4}), and shown in Fig.~\ref{fig:fss-yobun}, are well fitted to $R(t)=b t^{1/z(T)}$. The obtained exponents $z(T)$ are $7.86(3)$, $8.71(3)$, $9.84(5)$, $11.76(6)$ and $14.80(7)$ at $T=0.8$, $0.7$, $0.6$, $0.5$ and $0.4$, respectively. These values at $T=0.8$, $0.7$ listed on Table~\ref{tbl:z} are in good agreement with those extracted from the FSS analysis on $\delta e_{T}(t)$. \begin{table} \caption{Exponent $\theta$ evaluated by our finite-size-scaling (FSS) analysis on the energy relaxation. } \label{tbl:theta} \hspace*{\fill} \begin{tabular}{ccc}\hline $T$ & FSS & Bray and Moore\\ \hline $0.8$ & 0.20(2) & --- \\ $0.7$ & 0.19(3) & --- \\ $0.0$ & --- & 0.19(1) \\ \hline \end{tabular}\hspace{\fill} \end{table} \begin{table} \caption{Exponent $z(T)$ estimated from our finite-size-scaling (FSS) analysis on the energy relaxation and directly from correlation length of $G(r,t)$. The values are consistent with each others. The values $z(T)$ of the previous works \cite{Marinari-98-VFDT,Kisker-96} are also listed, where the values of Marinari et al. are calculated from their estimation as $1/z(T)\sim 0.16T$. The present result is consistent with $z(T)$ by Marinari et~al. but not with $z(T)$ by Kisker et al.}\label{tbl:z} \begin{tabular}{cccc}\hline $T$ & FSS & $R(t)$ & previous works \\ \hline $0.8$ & $7.73(7)$ & $7.86(3)$ & $\sim 7.81$ (Marinari et al.)\\ $0.7$ & $8.83(8)$ & $8.71(3)$ & $\sim 8.93$ (Marinari et al.) \\ & & & $12.4$ (Kisker et al.)\\ \hline \end{tabular} \end{table} The temperature dependence of $z(T)$ is given by $1/z(T) \simeq 0.17T$ at $T<0.7$. This linear $T$-dependence of $1/z(T)$ is consistent with the results by Marinari et al.,\cite{Marinari-98-VFDT} and those by Kisker et al.\cite{Kisker-96}, though its coefficient $0.17$ is different from the one estimated from those of Kisker et al. The replica-overlap function $G(r,t)$ is well scaled when it is plotted against $r/t^{1/z(T)}$ as shown in Fig.~\ref{fig:fss-1}. Furthermore it is seen that $G(r,t)$ in a longer length scale decays in a simple exponential form. The exponents $z(T)$ used in this plot are those extracted by the fit of $R(t)$. \subsection{Distribution of barrier free-energy of isolated droplets} \label{subsec:fss-3-3} As mentioned in \S 1, another important purpose of the present work is to study directly relaxational dynamics of droplets. Since, however, it is hard to specify each droplet in bulk, we examine here characteristic time scales $\tau_L(T)$ for spin configurations of systems with small $L$ to turn over as a whole in equilibrium. We regard such systems as isolated droplets of size $L$. At first, $t_0$ MCS is elapsed starting from a random spin configuration. Then the spin configuration $\{ S_i^{(\alpha)}(t_0) \}$ is copied to another replica $\beta$ and the two replicas are updated independently with different sequence of random numbers. The latter is continued until the clone-correlation function \begin{equation} Q(t)=\frac{1}{N}\sum_{i=1}^{N}\ S_i^{(\alpha)}(t)S_i^{(\beta{})}(t). \label{eq:fss-2-10} \end{equation} first crosses zero at some time $t_1$. Then the spin configuration $\{ S_i^{(\alpha)}(t_1) \}$ is copied to replica $\beta$ and the above process is repeated until $t_2$ when $Q(t)$ crosses zero. The procedure is repeated once more, and the time when $Q(t)$ crosses zero is denoted by $t_3$. In Fig.~\ref{fig:fss-6} we show distribution of the logarithmic times $P_{L}(\ln \tau_i)$, where $\tau_i=t_i - t_{i-1}$ with $t_0=0$. One can see clearly that $P_{L}(\ln \tau_2)$ and $P_{L}(\ln \tau_3)$ coincide with each other, but it is not the case for $P_{L}(\ln \tau_1)$. The result is interpreted that $P_{L}(\ln \tau_1)$ is affected by off-equilibrium effects, while $P_{L}(\ln \tau_2)$ can be regarded as an equilibrium distribution which we analyze in detail below with $\tau_L(T) = \tau_2\ (=\tau)$. In this analysis we have examined both the periodic and free boundary conditions. \begin{figure} \leavevmode\epsfxsize=85mm \epsfbox{first-dist-FSS.eps} \caption{The distribution of global-flip times $\tau_1$ (solid circles), $\tau_2$ (solid triangles) and $\tau_3$ (solid squares), at $T=0.8$ in $L=4$ systems. } \label{fig:fss-6} \end{figure} \begin{figure} \leavevmode\epsfxsize=85mm \epsfbox{Tdep-raw.eps}\\ \leavevmode\epsfxsize=85mm \epsfbox{Tdep-scale-FSS.eps} \caption{(a) The distribution of $\tau\ (=\tau_L(T))$ for $L=4$ at $T=0.8$ (open circles), $0.7$ (solid circles), $0.6$ (open triangles) and $0.5$ (solid triangles). (b) The plot of $\Phi_{L}(B)$ extracted from eq.~(\ref{eq:fss-2-9}). } \label{fig:fss-7} \end{figure} In Fig.~\ref{fig:fss-7}-(a) we show the log-log plot of $P_{L}(\ln \tau )$ versus $\tau$ of the system with $L=4$ at $T=0.8$, $0.7$, $0.6$ and $0.5$. If the relaxation process is of the Arrhenius-type, i.e., $\tau = \tau_0 \exp (B/T)$, the distribution of the free-energy barriers $\Phi_{L}(B)$ is written as \begin{equation} \Phi_{L}(B) = \frac{1}{T}P_{L}(\ln\tau{}/\tau_{0}) \label{eq:fss-2-9} \end{equation} where $\tau_{0}^{-1}$ is the attempt frequency. With $\tau_0=1$ the data in Fig.~\ref{fig:fss-7}-(a) lie top on each other as shown in Fig.~\ref{fig:fss-7}-(b). This result indicates that the relaxational dynamics of present interest is in fact an Arrhenius type whose barrier free energy $B$ little depends on temperature. Making use of relation $B=T\ln \tau$ confirmed above, we next examine the $L$-dependence of $\Phi_{L}(B)$. In Fig.~\ref{fig:fss-8} we show the plot of $\Phi_{L}(B)$ against $B-B_L$ at $T=0.8$, where $B_L$ is the value, at which $\Phi_{L}(B)$ is maximum for each $L$. The resultant $B_L$ is well fitted to \begin{equation} B_{L} = \Delta' \ln L, \label{eq:fss-3-5} \end{equation} as shown in Fig.~\ref{fig:fss-9} below. Another interesting observation in Fig.~\ref{fig:fss-8} is that the shape of the distribution $\Phi_{L}(B)$ does not depend on $L$, while its peak position does. This is certainly in disagreement with the original droplet theory,\cite{FH-88-EQ} in which the width of $\Phi_{L}(B)$ is conjectured to be of the same order as that of its mean. \begin{figure} \leavevmode\epsfxsize=85mm \epsfbox{Ldep-scale-FSS.eps} \caption{ The plot of $\Phi_{L}(B)$ versus $B-B_{L}$, at $T=0.8$ for $L=4$ (open circles), $6$ (solid circles), $8$ (open triangles), $10$ (solid triangles) and $12$ (open squares). } \label{fig:fss-8} \end{figure} \begin{figure} \leavevmode\epsfxsize=85mm \epsfbox{barrierLdep.eps} \caption{The characteristic barrier free energy corresponding to characteristic length scale $l$ (at $T=0.8$). The open circles are estimated from the growth of $R(t)$. The solid circles and solid triangles are estimated from the distribution of global flip times $\tau_L(T)$ with the periodic and open boundary conditions, respectively. } \label{fig:fss-9} \end{figure} If the coarsening process of domains represented by $R(t)$ analyzed before is also assumed due to nucleation of domains of the corresponding size by thermally activated process, the associated barrier free energy $B_R$ is given by\cite{Kisker-96} \begin{equation} B_{R} = \Delta \ln R. \label{eq:fss-3-8} \end{equation} In Fig.~\ref{fig:fss-9} we plot $B_{R}$ thus evaluated and $B_{L}$ of eq.~(\ref{eq:fss-3-5}) against the properly scaled length $l$. For $B_{R}$ we set $l=2R(t)$, while for $B_{L}$ we use $l=L/(2^{1/d})$ since the clone-correlation function can cross zero when a half of spins in the system turns over. The logarithmic dependence on $l$ of both $B_{R}$ and $B_{L}$ is clearly seen in the figure. Also it is seen that the coefficient $\Delta$ for $B_{R}$ is a little larger than $\Delta'$ for $B_{L}$. This difference may indicate that dynamics of ordered domains in a large system is not perfectly independent from their neighboring ones as was assumed in the original droplet theory. A similar conclusion has been also deduced experimentally.\cite{Weissman} \subsection{Discussions} Here let us comment on the magnitude of energy, length and time scales in the aging process simulated so far. Our result is consistent with $F_{L}\sim L^{\theta}$, while the barrier free energy $B_L$ associated with them is proportional to $\ln L$. It is then naturally expected that the latter is overcome by the former at a certain value of $L$, and so that the present results cannot provide us a proper picture in the thermodynamic limit. But we note that $F_{L}$ evaluated from the present simulation is given by \begin{equation} F_{L} \propto N\delta e_{T} \simeq 2.0 \times (L/l_0)^{\theta}, \end{equation} with $\theta = 0.19$, where the factor $2.0$ is determined from the saturated value in the scaling plot of Fig.~\ref{fig:fss-5}. Because of this small value of $\theta$, combined with a relatively large value of $\Delta\ (\simeq \Delta')$ in eq.~(\ref{eq:fss-3-8}) which in turn is due to a relatively large value of $z(T)(\simeq 8 \sim 10)$, $F_{L}$ and $B_R$ (or $B_L$) become of the same order of magnitude only when $L$ is of the order of $10^8\sim 10^9$ with $l_0=1$. The corresponding time scale, for $R(t)$ to reach to this length scale, is much more than an astronomical one. In this context, it is pointed out that $B_R$ experimentally extracted by Joh et al.\cite{Joh-cd98} recently is consistent with the simulated results expressed by eq.~(\ref{eq:fss-3-8}) (with $\tau/\tau_0 \sim 10^{12}$ and $R/l_0 \sim \ {\rm sevral}\ 10$). We therefore consider that the results obtained in the present work are applicable to aging phenomena observed within a time window of experiments on realistic spin glasses. However, mechanism which gives rise to the ln$R$ dependence of $B_R$ is not clear yet. It might be related to fluctuations governed by the $T=T_{\rm c}$ criticality since temperatures investigated in the present work may not be far enough from $T_{\rm c}$. This problem is left for future work. \section{Concluding Remarks}\label{sec:fss-4} By means of Monte Carlo simulations we have studied aging dynamics after the temperature quench from $T=\infty$ to $T$ in the spin-glass phase in the 3D Ising EA model. The results of the energy evolution $e_{T}(t)$, in particular, the finite-size scaling on $\delta e_{T}(t)\ (= e_{T}(t) - e_{T}^{(\infty)})$, strongly suggest that the droplet picture describes aging dynamics appropriately. It is emphasized that these results are obtained based on eq.~(\ref{eq:fss-1-1}) which is a most fundamental ingredient of the droplet picture, and that they are extracted by analyzing time evolution of one spin configuration as compared with the previous works which were making use of the replica-overlap. Also by analyzing relaxation times $\tau_L$ required for a global flip of spin configurations in small systems (isolated droplets) of size $L$, we have shown that the $L$-dependence of $\tau_L$ is consistent with the $t$-dependence of the mean domain size $R(t)$, i.e., $R(t \sim \tau_L(T) ) \sim L$. This result further supports the droplet picture that coarsening of domain walls is in fact driven by successive nucleation of thermally activated droplets. The explicit form of this $t\ (\tau)$-dependence of $R(t)\ (L)$, however, disagrees with the original droplet theory due to Fisher and Huse:\cite{FH-88-NE,FH-88-EQ} the barrier free energy associated with ordered domains of length scale $L$ is proportional not algebraically as predicted by the their theory but logarithmically to $L$. \section*{Acknowledgments} We would like to thank P. Nordblad, R. Orbach and E. Vincent for discussions on their experimental works. Two of the present authors (T. K. and H. Y.) were supported by Fellowships of Japan Society for the Promotion of Science for Japanese Junior Scientists. This work is supported by a Grant-in-Aid for International Scientific Research Program, ``Statistical Physics of Fluctuations in Glassy Systems'' (\#10044064) and by a Grant-in-Aid for Scientific Research Program (\#10640362), from the Ministry of Education, Science and Culture. The present simulation has been performed on FACOM VPP-500/40 at the Supercomputer Center, Institute for Solid State Physics, the University of Tokyo and on HITACHI SR2201 at the Computer Centre of the University of Tokyo.
\section{Introduction} In 1974, Kruskal \cite{kruskalpoles} considered the interaction of solitons governed by the Korteweg-deVries equation (KdV), \begin{equation}\la{kdv} u_t=6 u u_x+u_{xxx}. \end{equation} \noindent Each KdV soliton is defined by a meromorphic function in the complex $x$-plane ($i.e., \mbox{sech}^2k(x-x_0))$, so Kruskal \cite{kruskalpoles} suggested that the interaction of two or more solitons could be understood in terms of the dynamics of the poles of these meromorphic functions in the complex x-plane, where the poles move according to a force law deduced from \rf{kdv}. This was followed by the work of Thickstun \cite{thickstun} who considered the case of two solitons in great detail. Following a different line of thought, Airault, McKean and Moser \cite{amm} studied rational and elliptic solutions of the KdV equation. An elliptic solution of KdV is by definition a solution of the KdV equation that is doubly periodic and meromorphic in the complex $x$-plane, for all time. Note that the soliton case is an intermediary case between the elliptic and the rational case. It was treated as such in \cite{amm}. Airault, McKean and Moser \cite{amm} approached these elliptic solutions and their degenerate limits through the motion of their poles $x_i(t)$ in the complex $x$-plane. In particular, they looked for elliptic KdV solutions of the form \begin{equation}\la{sol} u(x,t)=-2\sum_{i=1}^N \wp(x-x_i(t)). \end{equation} \noindent Here $\wp(z)$ denotes the Weierstra\ss~ elliptic function. It can be defined by its meromorphic expansion \begin{equation}\la{wp} \wp(z)=\frac{1}{z^2}+\sum_{(m,n)\neq (0,0)}\left(\frac{1}{(z+2 m \omega_1+2 n \omega_2)^2}-\frac{1}{(2 m \omega_1+2 n \omega_2)^2}\right), \end{equation} \noindent with $\omega_1/\omega_2$ not real\footnote{In this paper it is always assumed that $\omega_1$ is real and $\omega_2$ is imaginary. This is necessary to ensure reality of the KdV solution \rf{sol} when $x$ is restricted to the real $x$-axis. Other considerations for reality of the elliptic KdV solutions will be discussed in Section \ref{sec:dis}.}. More properties of the Weierstra\ss~ function will be given as they are needed. It is shown in \cite{amm} that the dynamics of the poles $x_i(t)$ is governed by the dynamical system \setcounter{saveeqn}{\value{equation} \begin{equation}\la{dynsys} \dot{x_i}=12 \sum_{j=1, j\neq i}^N\wp(x_i-x_j),~~~i=1,2,\ldots, N, \end{equation} \noindent (the dot denotes differentiation with respect to time) with the invariant constraint \begin{equation}\la{constraint} \sum_{j=1, j\neq i}^N \wp'(x_i-x_j)=0.~~~i=1,2,\ldots, N. \end{equation} \setcounter{equation}{\value{saveeqn} \noindent Here the prime denotes differentiation with respect to the argument. The solutions \rf{sol} generalize an elliptic solution given earlier by Dubrovin and Novikov \cite{dubnov}, corresponding to the case $N=3$. These authors also recall the Lam\'{e}-Ince potentials \cite{ince2} \begin{equation}\la{lame} u(x)=-g(g+1)\wp(x), \end{equation} \noindent which are the simplest $g$-gap potentials of the stationary Schr\"{o}dinger equation \begin{equation}\la{schrodinger} \ppn{2}{\psi}{x}+u(x) \psi=\lambda \psi. \end{equation} \noindent The remarkable connection between the KdV equation and the stationary Schr\"{o}dinger equation has been known since the work of Gardner, Greene, Kruskal and Miura \cite{ggkm}. Dubrovin and Novikov show \cite{dubnov} that the $(N=3)$-solution discussed in \cite{dubnov} is a 2-gap solution of the KdV equation with a 2-gap Lam\'e-Ince potential as initial condition. If one considers the rational limit of the solution \rf{sol} ($i.e.$, the limit in which the Weierstra\ss~ function $\wp(z)$ reduces to $1/z^2$), then the constraint \rf{constraint} is only solvable for a triangular number of poles, \begin{equation}\la{triangular} N=\frac{n(n+1)}{2}, \end{equation} \noindent for any positive integer $n$ \cite{amm}. Notice that the Lam\'{e}-Ince potentials are given by $g(g+1)/2$ times an $N=1$ potential. Based on these observations, it was conjectured in \cite{amm} that also in the elliptic case given by \rf{sol}, the constraint \rf{constraint} is only solvable for a triangular number $N$, ``or very nearly so''. From the moment it appeared this conjecture was known not to hold, because it already fails in the soliton case, where the Weierstra\ss~ function degenerates to hyperbolic functions. This failure of the conjecture easily follows from the work of Thickstun \cite{thickstun}. A further understanding of the elliptic case had to wait until 1988, when Verdier provided more explicit examples of elliptic potentials of the Schr\"{o}dinger equation \cite{verdier}. Subsequently, Treibich and Verdier demonstrated that \begin{equation}\la{tv} u(x)=-2 \sum_{i=1}^4 \frac{g_i(g_i+1)}{2} \wp (x-x_0-\omega_i) \end{equation} \noindent ($\omega_3=\omega_1+\omega_2, \omega_4=0$, the $g_i$ are positive integers) are finite-gap potentials of the stationary Schr\"{o}dinger equation \rf{schrodinger}, and hence result in elliptic solutions of the KdV equation \cite{tv1,tv2,tv3}. The potentials of Treibich and Verdier were generalized by Gesztesy and Weikard \cite{gw1,gw2}. They showed that any elliptic finite-gap potential of the stationary Schr\"{o}dinger equation \rf{schrodinger} can be represented in the form \begin{equation}\la{gw} u(x)=-2 \sum_{i=1}^M \frac{g_i(g_i+1)}{2} \wp(x-\alpha_i), \end{equation} \noindent for some $M$ and positive integers $g_i$. Notice that this formula coincides with \rf{sol} if all the $g_i$ are 1. The focus of this paper is the constrained dynamical system (\ref{dynsys}-b). We return to the ideas put forth by Kruskal \cite{kruskalpoles} and Thickstun \cite{thickstun}. This allows us to derive the system (\ref{dynsys}-b) in a context which is more general than \cite{amm}: there it was obtained as a system describing a class of special solutions of the KdV equation. Here, it is shown that any meromorphic solution of the KdV equation which is doubly periodic in $x$ is of the form \rf{sol}. Hence the consideration of solutions of the form \rf{sol} and the system of equations (\ref{dynsys}-\ref{constraint}) leads to {\em all} elliptic solutions of the KdV equation. Simultaneously, some of the results of Gesztesy and Weikard \cite{gw1, gw2} are recovered here. Because of the connection between the KdV equation and the Schr\"{o}dinger equation, any potential of Gesztesy and Weikard can be used as an initial condition for the KdV equation, which determines any time dependence of the parameters of the elliptic KdV solution with that initial condition. Our approach demonstrates which parameters in the solutions \rf{gw} are time independent and which are time dependent. The following conclusions are obtained in this paper: \begin{itemize} \item All finite-gap\footnote{See Section \ref{sec:dynsys}} elliptic solutions of the KdV equation are of the form \rf{sol}, for almost all times (see below). In other words, if u(x,t) is a finite-gap KdV solution that is doubly periodic in the complex x-plane, then u necessarily has the form \rf{sol} except at isolated instants of time. \item Any number $N \neq 2$ of $x_j$ is allowed in \rf{sol}, if the ratio $|\omega_1/\omega_2|\gg 1$. It follows that in this case, the constraint \rf{constraint} is solvable for any positive integer\footnote{The constraint \rf{constraint} is solvable for $N=2$ \cite{amm}. As noted there, the corresponding solution reduces to a solution for $N=1$, with smaller periods. This case is therefore trivial and is disregarded} $N\neq 2$. Our method for demonstrating this is to use a deformation from the soliton limit of the system (\ref{dynsys}-b). Furthermore, this method is useful because it provides good initial guesses for the numerical solution of \rf{constraint} in Section \ref{sec:ex}. This method does not recover all elliptic solutions of the KdV equations, because only those solutions are found which have nonsingular soliton limits. In particular, the solutions corresponding to the Treibich-Verdier potentials \rf{tv} are not found. \item If $|\omega_1/\omega_2|\gg 1$, then for a given $N>4$, nonequivalent configurations satisfying the constraint exist that do {\em not} flow into each other under the KdV flow and which cannot be translated into each other. To the best of our knowledge this is a new result. \item The $x_i$ are allowed to coincide, but only in triangular numbers: if some of the $x_i$ coincide at a certain time $t_c$, then $g_i(g_i+1)/2$ of them coincide at that time $t_c$. At this time $t_c$, the solution can be represented in the form \rf{gw} with not all $g_i=1$. Such times $t_c$ are referred to as collision times and the poles are said to collide at the collision time. Before and after each collision time all $x_i$ are distinct, hence pole collisions are isolated events. At the collision times, the dynamical system \rf{dynsys} is not valid. The dynamics of the poles at the collision times is easily determined directly from the KdV equation. Gesztesy and Weikard \cite{gw1, gw2} demonstrate that \rf{gw} are elliptic potentials of the Schr\"{o}dinger equation. These potentials generalize to solutions of the KdV equation, but this requires the $g_i$ to be nonsmooth functions of time. Only at the collision times $t_c$ are the $g_i$ not all one. Furthermore, at all times but the collision times, the number of parameters $\alpha_i$ (which are time dependent) is $N=\sum_{i=1}^M g_i(g_i+1)/2$. This shows that the generalization from \cite{gw1, gw2} to solutions of the KdV equation is nontrivial. Using the terminology of Chapter 7 of \cite{belokolos1}, the poles are in {\em general position} if all $g_i$ are equal to one. Otherwise, if not all $g_i=1$, the poles are said to be in {\em special position}. We conclude that at the collision times the poles are in special position. Otherwise they are in general position. \item The solutions discussed here are finite-gap potentials of the stationary Schr\"{o}dinger equation \rf{schrodinger}, with $t$ treated as a parameter. In other words, each solution specifies a one-parameter family of finite-gap potentials of the stationary Schr\"{o}dinger equation. It follows from our methods that to obtain a $g$-gap potential that corresponds to a nonsingular soliton potential, one needs at least $N=g(g+1)/2$ poles $x_j$. The Lam\'{e}-Ince potentials show that this lower bound is sharp. If we consider potentials that do not have nonsingular soliton limits (such as the Treibich-Verdier potentials \rf{tv}) then it may be possible to violate this lower bound. \item The relationship between elliptic solutions and soliton solutions of (\ref{dynsys}-b) and hence of the KdV equation is made explicit. In fact, this relationship is essential to the method used here, as mentioned above. \item In Section \rf{sec:ex}, we present an explicit solution of the form \rf{sol} with $N=4$. Notice that if in \rf{tv} all $g_i$ are one, the solution is reducible to a solution with $N=1$ and smaller periods. To see this it is convenient to draw the pole configuration corresponding to \rf{tv} in the complex plane. This is actually true, even if $g_i$ is not one, but all $g_i$ are equal. In that case, \rf{tv} reduces to a Lam\'{e}-Ince potential \rf{lame}. Unlike any of the Treibich-Verdier solutions, the $(N=4)$-solution, presented in Section \rf{sec:ex}, has a nonsingular soliton limit. \end{itemize} The first five conclusions are all discussed in Section \ref{sec:dis}. In Section \ref{sec:thick}, the results of Kruskal \cite{kruskalpoles} and Thickstun \cite{thickstun} for the soliton solutions of the KdV equation are reviewed, but they are obtained from a point of view which is closer to the approach we present in Section \ref{sec:dynsys} for the elliptic solutions of the KdV equation. Finally, in Section \ref{sec:ex}, some explicit examples are given, including illustrations of the motion of the poles in the complex $x$-plane. \section{The soliton case: hyperbolic functions}\la{sec:thick} In this section, the results of Kruskal \cite{kruskalpoles} and Thickstun \cite{thickstun} for the dynamics of poles of soliton solutions are discussed from a point of view that will allow us to generalize to the periodic case. Consider the one-soliton solution of the KdV equation \begin{equation}\la{1sol} u(x,t)=2 k^2 \mbox{sech}^2 k(x+4 k^2 t-\varphi). \end{equation} \noindent Here $k$ is a positive parameter (the wave number of the soliton) determining the speed and the amplitude of the one-soliton solution. Using the meromorphic expansion \cite{tables} \begin{equation}\la{merosol} \frac{1}{T^2}\mbox{cosech}^2 \frac{x}{T}=\sum_{n=-\infty}^{\infty}\frac{1}{(x+ i n \pi T)^2}, \end{equation} \noindent (uniformly convergent except at the points $x=i n \pi T$) one easily obtains the following meromorphic expansion for the one-soliton solution of the KdV equation: \begin{equation}\la{mero1sol} u(x,t)=-2 k^2 \sum_{n=-\infty}^{\infty} \frac{1}{(k (x+4 k^2 t-\varphi)+i \frac{\pi}{2}+i n \pi)^2}. \end{equation} \noindent From this expression, one easily finds that the locations of the poles of the one-soliton solution of the KdV equation for all time are given by \begin{equation}\la{poles1sol} x_n=\varphi-4 k^2 t-\frac{i \pi}{k}\left(n+\frac{1}{2}\right). \end{equation} \noindent This motion is illustrated in Fig. \ref{fig1}a. Notice that the locations of the poles are symmetric with respect to the real $x$-axis. This is a consequence of the reality of the solution \rf{1sol}. In order for a solution to be real it is necessary and sufficient that if $x_n(t)$ is a pole, then so is $x_n^*(t)$, where $~^*$ denotes the complex conjugate\footnote{The vertical line of poles can be rotated arbitrarily. The expression \rf{mero1sol} still results in a solution of the KdV equation, but it is no longer real. Again, we will only consider real, nonsingular solutions, when restricted to the real $x$-axis}. The closest distance between any two poles is $d=\pi/k$ and is constant both along the vertical line $\mbox{Re}(x)=\varphi-4 k^2 t$ and in time. Note that the poles are moving to the left. This is a consequence of the form of the KdV equation \rf{kdv}, which has time reversed, compared to the version Kruskal \cite{kruskalpoles} and Thickstun \cite{thickstun} used. \begin{figure}[htb] \begin{tabular}{ccc} \psfig{file=1sol.eps,width=2.9in} &~~~& \psfig{file=2sol.eps,width=2.9in}\\ (a)&&(b) \end{tabular} \caption{\la{fig1}{\bf (a) The motion of the poles of a one-soliton solution in the complex plane. (b) The asymptotic motion of the poles of a two-soliton solution.}} \end{figure} Since two solitons of the KdV equation cannot move with the same speed, a two-soliton solution of the KdV equation asymtotically appears as the sum of two one-soliton solutions which are well-separated: the higher-amplitude soliton, which is faster, is to the right of the smaller-amplitude soliton as $t\rightarrow -\infty$. As $t\rightarrow \infty$, the higher-amplitude soliton is to the left of the smaller-amplitude soliton. Hence as $t\rightarrow -\infty$, the pole configuration of a two-soliton solution with wave numbers $k_1$ and $k_2$ is as in Fig. \ref{fig1}b. In this limit, the two-soliton solution is a sum of two one-soliton solutions. Each results in a vertical line of equispaced poles, with interpolar distance respectively $d_1=\pi/k_1$ and $d_2=\pi/k_2$. As long as the solitons are well-separated, these poles move in approximately straight lines, parallel to the real axis, with respective velocities $v_1=-4 k_1^2$ and $v_2=-4 k_2^2$. Since $|v_1|>|v_2|$, the solitons interact eventually. This interacting results in non-straight line motion of the poles. After the interaction, the situation is as in Fig. \ref{fig1}b, but with the two lines of poles interchanged. Thickstun \cite{thickstun} considered the case where $k_1$ and $k_2$ are rationally related, so $k_1/k_2=p/q$, where $p$ and $q$ are positive integers. In this case, one can define $D=p d_1=q d_2$. The complex $x$-plane is now divided into an infinite number of equal strips, parallel to the real $x$-axis, each of height $D$. The real $x$-axis is usually taken to be the base of such a strip. It is easy to show \cite{thickstun} that the motion of the poles in one strip is repeated in every strip. Hence, one is left studying the motion of a finite number $N$ ($=p+q$) of poles in the fundamental strip, whose base is the real $x$-axis. Thickstun examined this motion by analyzing the exact expression for a two-soliton solution of the KdV equation. Any two-soliton solution is expressible as \cite{as} \begin{equation}\la{twosol} u(x,t)=2 \partial_x^2 \ln \tau(x,t). \end{equation} \noindent It follows from this formula that the poles of $u(x,t)$ are the zeros of $\tau(x,t)$ if $\tau(x,t)$ is entire in $x$. Then the Weierstra\ss~ Factorization Theorem \cite{conway} gives a factorization for $\tau(x,t)$: \begin{equation}\la{factau} \tau(x,t)=C \prod_{k=1}^\infty \left(1-\frac{x}{x_k}\right)e^{x/x_k}. \end{equation} \noindent Since only the second logarithmic derivative of this function is relevant, the constant $C$ is not important. If the solution is periodic in the imaginary $x$-direction, this is rewritten as \begin{equation}\la{factau2} \tau(x,t)=C \prod_{n=1}^{N}\prod_{l=-\infty}^{\infty}\left(1-\frac{x}{x_n+i l D}\right)e^{x/(x_n+i l D)}, \end{equation} \noindent where the first product runs over the poles in the fundamental strip. The second product runs over all strips. Using the uniform convergence of \rf{factau2}, \begin{equation}\la{usol} u(x,t)=-2 \sum_{n=1}^{N}\sum_{l=-\infty}^{\infty}\frac{1}{(x-x_n-i l D)^2}, \end{equation} \noindent which, using \rf{merosol}, is rewritten as \begin{equation}\la{usolhyp} u(x,t)=-2 \frac{\pi^2}{D^2}\sum_{n=1}^{N}\mbox{cosech}^2\frac{\pi(x-x_n)}{D} =-2 \frac{k_2^2}{q^2}\sum_{n=1}^{p+q}\mbox{cosech}^2 \frac{k_2 (x-x_n)}{q} , \end{equation} \noindent where the pole locations $x_n$ depend on time: $x_n=x_n(t)$. One recovers the one-soliton solution \rf{1sol} easily, by equating $k_1=0, p=0,q=1$. Equation \rf{usolhyp} essentially expresses a two-soliton solution as a linear superposition of $N$ one-soliton solutions with nonlinearly interacting phases. Note that the first equality in \rf{usolhyp} is valid for arbitrary soliton solutions that are periodic in $x$ with period $i D$. This is the case for a g-soliton solution if its wavenumbers $k_i$, $i=1,2,\ldots, g$ are all commensurable: $(k_1:k_2:\ldots:k_g)=(p_1:p_2:\ldots:p_g)$, for positive distinct integers $p_i$, $i=1,2,\ldots, g$ which have no overall common integer factor. The total number of poles in a strip is then $N=p_1+p_2+\ldots+p_g$. In obtaining \rf{usol} and \rf{usolhyp}, we have deviated from Thickstun's approach \cite{thickstun} to an approach that is generalized to the elliptic case of the next section in a straightforward way. Next, we derive the dynamics imposed on the poles $x_n(t)$ by the KdV equation. This is conveniently done by substituting \rf{usol} into \rf{kdv} and examining the behaviour near one of the poles: $x=x_n+\epsilon$. This results in several singular terms as $\epsilon\rightarrow 0$, corresponding to negative powers of $\epsilon$. The dynamics of the poles is then determined by the vanishing of the coefficients of these negative powers and the zeroth power. This results in only two nontrivial equations, obtained at order $\epsilon^{-3}$ and $\epsilon^{-2}$ respectively: \setcounter{saveeqn}{\value{equation} \begin{eqnarray} \dot{x}_n&=&12 {\sum_{k=1, k\neq n}^N}\sum_{l=-\infty}^\infty \frac{1}{(x_k-x_n-i l D)^2}+12 \sum_{l\neq 0, l=-\infty}^{\infty}\frac{1}{(-i l D)^2},\\ 0&=&\sum_{k=1, k\neq n}^N \sum_{l=-\infty}^\infty \frac{1}{(x_k-x_n-i l D)^3}, \end{eqnarray} \setcounter{equation}{\value{saveeqn} \noindent for $n=1,2,\ldots, N$. Using \rf{merosol} and its derivative, \setcounter{saveeqn}{\value{equation} \begin{eqnarray}\la{soldynsys} \dot{x}_n&=&-4 \frac{\pi^2}{D^2}+12\frac{\pi^2}{D^2} {\sum_{k=1, k\neq n}^N} \mbox{cosech}^2 \frac{\pi (x_k-x_n)}{D},\\\la{solcons} 0&=&\sum_{k=1, k\neq n}^N \mbox{cosech}^2 \frac{\pi (x_k-x_n)}{D} \,\mbox{coth}\frac{\pi (x_k-x_n)}{D}, \end{eqnarray} \setcounter{equation}{\value{saveeqn} \noindent for $n=1,2,\ldots, N$. Hence the dynamics of the poles $x_n(t)$ is determined by \rf{soldynsys}. This dynamics is constrained by the equations \rf{solcons}. These constraint equations \rf{solcons} are invariant under the flow of \rf{soldynsys}. This follows from a direct calculation. \vspace*{12pt} \noindent {\bf Remarks:} \begin{itemize} \item Since the KdV equation has two-soliton solutions for any ratio of the wavenumbers $k_1/k_2\neq 1$, the constraint \rf{solcons} is solvable for any value of $N$, excluding $N=2$, which can only be obtained by $p=q=1$, resulting in equal wavenumbers $k_1$ and $k_2$. \item In particular, it follows that the minimum number of poles in a fundamental strip required to obtain a $g-soliton$ solution is $N=1+2+3+\ldots+g=g(g+1)/2$, corresponding to a $g$-soliton solution with wavenumbers which are related as $(k_1:k_2:\ldots:k_{g-1}:k_g)=(g:g-1:\ldots:2:1)$. \item Equating $k_1=0$, $p=0$, $q=1$, one obtains from \rf{soldynsys} $\dot{x}_1=-4 k_2^2$, corresponding to the dynamics of the one-soliton solution. The asymptotic behavior of the poles of a two-soliton solution also follows from \rf{soldynsys}: from the separation of the poles into distinct vertical lines, it follows from \rf{dynsys} that the velocity of these vertical lines is given by the one-soliton velocity for each line, as expected. This result follows from easy algebraic manipulation and the identity \begin{equation}\nonumber \frac{p^2-1}{3}=\sum_{n=1}^{p-1}\mbox{cosec}^2\left(\frac{n \pi}{p}\right), \end{equation} \noindent valid for any integer $p>1$. \item A full analysis of the interaction of the poles for the case of any two-soliton solution with $k_1/k_2=p/q$ is given in \cite{thickstun}. \end{itemize} \section{The elliptic case}\la{sec:dynsys} Consider the quasiperiodic finite-gap solutions of the KdV equation with $g$ phases \cite{itsmatveev} \begin{equation}\la{thetasol} u(x,t)=2 \partial_x^2 \ln \theta_g(\mbf{k}x+\mbf{\omega} t+\mbf{\phi}|\mbf{B}), \end{equation} \noindent where \begin{equation} \theta_g(\mbf{z}|\mbf{B})=\sum_{\mbf{m}\in\mbf{Z^g}} \exp\left(\frac{1}{2}\mbf{m}\cdot\mbf{B}\cdot\mbf{m}+i \mbf{m}\cdot\mbf{z}\right), \end{equation} \noindent a hyperelliptic Riemann-theta function of genus $g$. The $g\times g$ real Riemann matrix ($i.e.$, symmetric and negative definite) $\mbf{B}$ originates from a hyperelliptic Riemann surface with only one point at infinity. Furthermore, $\mbf{k}$, $\mbf{\omega}$ and $\mbf{\phi}$ are $g$-dimensional vectors. The derivation of equations \rf{usolhyp}, \rf{soldynsys} and \rf{solcons} is easily generalized to the case where the solution is not only periodic in the imaginary $x$-direction, but also in the real $x$-direction: \begin{equation}\la{uper} u(x+L_1,t)=u(x,t)=u(x+i L_2,t). \end{equation} \noindent This divides the complex $x$-plane into an array of rectangular domains, each of size $L_1\times L_2$. One of these domains, called the fundamental domain $S$, is conveniently placed in the lower left corner of the first quadrant of the $x$-plane. The theta function has the property \cite{dub} \begin{equation}\la{thetaprop} \theta_g(\mbf{z}+i \mbf{B}\cdot \mbf{M}+2 \pi \mbf{N}|\mbf{B})= \theta_g(\mbf{z}|\mbf{B})\exp\left(-\frac{1}{2}\mbf{M}\cdot \mbf{B} \cdot \mbf{M} +i \mbf{M}\cdot \mbf{z}\right), \end{equation} \noindent for any pair of $g$-component integer vectors $\mbf{M}, \mbf{N}$. This expression is useful to determine conditions on the wavevector $\mbf{k}$ in order for $u(x,t)$, given by \rf{thetasol}, to satisfy \rf{uper}: \begin{equation}\la{conper} \exists~ \mbf{N_0}, \mbf{M_0} \in \mbf{Z}^g:\mbf{k} L_1=2 \pi \mbf{N_0}, ~ \mbf{k} L_2=\mbf{B}\cdot \mbf{M_0}. \end{equation} \noindent These results are now used to determine the number of poles $N$ of $u(x,t)$ in the fundamental domain. The poles of \rf{thetasol} are given by the zeros of $\vartheta(x,t)=\theta_g(\mbf{k}x+\mbf{\omega} t+\mbf{\phi}|\mbf{B})$, regarded as a function of $x$: \begin{eqnarray}\nonumber N&=&\frac{1}{2 \pi i} \oint_{\partial S} d \ln \vartheta(x,t)\\\nonumber &=&\frac{1}{2 \pi i}\left( \int_{0}^{L_1}d \ln \vartheta(x,t)+\int_{0}^{i L_2} d \ln \vartheta(x+L_1,t)-\right.\\\nonumber &&\left.\int_{0}^{L_1} d \ln \vartheta(x+i L_2,t)-\int_{0}^{i L_2} d \ln \vartheta(x,t) \right)\\\nonumber &\since{thetaprop}&\frac{1}{2 \pi i}\int_{0}^{L_1} d \ln \frac{\vartheta(x,t)}{\vartheta(x+i L_2,t)}\\\nonumber &\since{thetaprop}&\frac{1}{2 \pi i} \int_{0}^{L_1} d (-i x \mbf{M_0}\cdot \mbf{k})\\\la{numberofpoles} &=&-\mbf{M_0}\cdot\mbf{N_0} =-\frac{L_1}{2 \pi L_2}\mbf{M_0}\cdot\mbf{B}\cdot\mbf{M_0}. \end{eqnarray} \noindent The first equality of \rf{numberofpoles} confirms that $N$ is an integer. The second equality shows that $N$ is positive, by the negative-definiteness of $\mbf{B}$. We now proceed to determine the dynamical system satisfied by the motion of the $N$ poles of $u(x,t)$ in the fundamental domain $S$. Again, the poles of $u(x,t)$ are the zeros of $\vartheta(x,t)$. Furthermore, simple zeros of $\vartheta(x,t)$ result in double poles of $u(x,t)$, as in the hyperbolic case. The Weierstra\ss~ Factorization theorem \cite{conway} gives the following form for $\vartheta(x,t)$: \begin{equation}\la{factheta} \vartheta(x,t)=e^{c x^2/2}\prod_{k}\left(1-\frac{x}{x_k}\right)e^{\frac{x}{x_k}+\frac{x^2}{2 x_k^2}}, \end{equation} \noindent where the product runs over all poles $x_k$. The additional exponential factors, as compared to \rf{factau}, are required because the poles now appear in a bi-infinite sequence: both in the vertical and horizontal directions. These exponential factors ensure uniform convergence of the product. The parameter $c$ is allowed to depend on time. It determines the behavior of $\vartheta(x,t)$ as $x$ approaches infinity in the complex $x$-plane \cite{dub}. Using \rf{uper}, this is rewritten as \begin{eqnarray}\nonumber \vartheta(x,t)=\exp(cx^2/2) \prod_{n=1}^N \prod_{m=-\infty}^\infty \prod_{l=-\infty}^\infty &&\!\!\!\!\!\!\!\!\!\!\left(1-\frac{x}{x_n+m L_1+i l L_2}\right)\times\\\la{facthetaper} &&\!\!\!\!\!\!\!\!\!\! \exp\left(\frac{x}{x_n+m L_1+i l L_2}+ \frac{x^2}{2(x_n+m L_1+i l L_2)^2}\right). \end{eqnarray} \noindent The first product runs over the number of poles ($N$) in the fundamental domain, the second and third products result in all translations of the fundamental domain. From the uniform convergence of \rf{facthetaper}, \begin{equation}\la{u1} u(x,t)=2 c-2 \sum_{n=1}^N \sum_{m=-\infty}^{\infty}\sum_{l=-\infty}^{\infty} \left(\frac{1}{(x-x_n-m L_1-i l L_2)^2}-\frac{1}{(x_n+m L_1+i l L_2)^2}\right). \end{equation} \noindent Using the definition of the Weierstra\ss~ function \rf{wp}, this is rewritten as \begin{equation}\la{u2} u(x,t)=2 c-2 \sum_{n=1}^N \wp(x-x_n)+2 \sum_{j=1}^N\wp(x_n), \end{equation} \noindent where the periods of the Weierstra\ss~ function are given by $2 \omega_1=L_1, 2 \omega_2=i L_2$. Define \begin{equation}\la{constant} \tilde{c}=2 c+2 \sum_{n=1}^N\wp(x_n). \end{equation} \noindent The dynamics of the poles $x_n=x_n(t)$ is determined by substitution of \rf{u2} or \rf{u1} into the KdV equation and expanding in powers of $\epsilon$ for $x$ near a pole: $x=x_k+\epsilon$. Equating the coefficients of $\epsilon^{-3}$, $\epsilon^{-2}$ and $\epsilon^0$ to zero result in \setcounter{saveeqn}{\value{equation} \begin{eqnarray}\la{eldynsys} \dot{x}_n&=&12 \sum_{j=1, j\neq n}^N \wp(x_j-x_n),\\\la{elcons} 0&=&\sum_{j=1, j\neq n}^N \wp'(x_j-x_n),\\ \dot{\tilde{c}}&=&0 \iff \tilde{c}(t)=\alpha=0, \end{eqnarray} \setcounter{equation}{\value{saveeqn} \noindent for $n=1,2,\ldots, N$. (The constant $\alpha$ can always be removed by a Galilean shift, so it is equated to zero, without loss of generality.) The constraints \rf{elcons} are invariant under the flow, as can be checked by direct calculation. Notice that (\ref{eldynsys}-b) are identical to the equations obtained by Airault, McKean and Moser \cite{amm}. These equations are obtained here in greater generality: any solution \rf{thetasol} which is doubly periodic in the $x$-plane gives rise to a system (\ref{eldynsys}-b). This allows us to reach the conclusions stated in the next section. \vspace*{12pt} \noindent {\bf Remarks} \begin{itemize} \item In the limit $L_1\rightarrow \infty$, the equations (\ref{eldynsys}--\ref{elcons}) reduce to (\ref{soldynsys}--\ref{solcons}). This limit is most conveniently obtained from the Poisson representation of the Weierstra\ss~ function: \begin{equation}\la{weierpois} \wp(x)=\left(\frac{\pi}{L_2}\right)^2 \left(\frac{1}{3}+\mbox{cosech}^2\frac{\pi x}{L_2}+\sum_{n=-\infty, n\neq 0}^{\infty}\left\{\mbox{cosech}^2 \frac{\pi}{L_2}(x+n L_1)-\mbox{cosech}^2\frac{n \pi L_1}{L_2}\right\}\right). \end{equation} \noindent This representation is obtained from \rf{wp} by working out the summation in the vertical direction. It gives the Weierstra\ss~ function as a sum of exponentially localized terms, hence few terms have important contributions in the fundamental domain. A Poisson expansion for $\wp'(x)$ is obtained from differentiating \rf{weierpois} term by term with respect to $x$. \item Define the one-phase theta function $\theta_1(z,q)$ \cite{tables}: \begin{equation}\la{theta1} \theta_1(z,q)=2 \sum_{n=0}^{\infty}(-1)^n q^{(n+1/2)^2}\sin(2n+1)z, \end{equation} \noindent with $|q|<1$. If $L_2<L_1$, then the relationship $\wp(z)=a-\partial_x^2 \ln \theta_1(\pi z/L_1, i L_2/L_1)$, with $a$ a constant \cite{tables}, allows us to rewrite \rf{sol} as \begin{equation}\la{connysol} u(x,t)=\hat{a}+2 \partial_x^2 \ln \prod_{j=1}^N \theta_1\left(\frac{\pi}{L_1} (x-x_j(t)),i \frac{L_2}{L_1}\right), \end{equation} \noindent with $\hat{a}=-2 a N$. Hence, for the doubly-periodic solutions of the KdV equation of the form \rf{thetasol}, it is possible to rewrite the $g$-phase theta function as a product of $N$ $1$-phase theta functions, with nonlinearly interacting phases. Note that this does not imply that the $g$-phase theta function appearing in \rf{thetasol} is reducible. Reducible theta-functions do not give rise to solutions of the KdV equation \cite{dub}. \item By taking another time derivative of \rf{eldynsys} and using \rf{elcons}, one obtains \begin{equation}\la{elsecond} \ddot{x}_n=-(12)^2 \sum_{j=1, j\neq n}^N \wp'(x_j-x_n)\wp(x_j-x_n). \end{equation} \noindent It is known that this system of differential equations is Hamiltonian \cite{chud}, with Hamiltonian \begin{equation}\la{hamil1} H=\frac{1}{2}\sum_{k=1}^{N}p_k^2+\frac{(12)^2}{2}\sum_{k=1}^N \sum_{j=1, j\neq k}^N \wp^2(x_k-x_j), \end{equation} \noindent and canonical variables $\{x_k,p_k=\dot{x}_k\}$. A second Hamiltonian structure for the equations \rf{elsecond} is given in \cite{chud}. A Lax representation for the system (\ref{eldynsys}-c) is also given there. This Lax representation is a direct consequence of the law of addition of the Weierstra\ss~ function \cite{tables}. It is unknown to us whether a Hamiltonian structure exists for the constrained first-order dynamical system (\ref{eldynsys}-b). The Hamiltonian structure \rf{hamil1} shows that the system (\ref{eldynsys}-b) is a (constrained) member of the elliptic Calogero-Moser hierarchy \cite{belokolos1}. \end{itemize} \section{Discussion of the dynamics}\la{sec:dis} In this section, the constrained dynamical system (\ref{eldynsys}-c) is discussed. In particular, the assertions made in the introduction are validated here. For reality of the KdV solution \rf{thetasol} when $x$ is restricted to the real line, it is necessary and sufficient that if $x_j(t)$ appears, then so does $x_j^*(t)$. Because the Weierstra\ss~ function is a meromorphic functions of its argument, this reality constraint is invariant under the dynamics \rf{eldynsys}. As a consequence, the distribution of the poles in the fundamental domain $S$ is symmetric with respect to the horizontal centerline of $S$. Poles are allowed on the centerline. Most of what follows is valid for both real KdV solutions\footnote{``Real KdV solution'' refers to a solution of the KdV equation which is real when $x$ is restricted to the real $x$-axis} and KdV solutions that are not real, but we restrict our attention to real KdV solutions. \subsection{All finite-gap elliptic solutions of the KdV equation are of the form \rf{sol}, up to a constant} A straightforward singularity analysis of the KdV equation \cite{amm} shows that any algebraic singularity of a solution of the KdV equation is of the type $u(x,t)=-2/(x-\alpha(t))^2+{\cal O}(x-\alpha)$, for almost all times t. At isolated times $t_c$, the leading order coefficient is not necessarily $-2$. It can be of the form $-g(g+1)$ (see below), but the exponent of the leading term is always $-2$. Hence, an elliptic function ansatz for $u(x,t)$ can only have second order poles and with the substitution $u(x,t)=2 \partial_x^2 \ln \vartheta(x,t)$ gives rise to a Weierstra\ss~ expansion of the form \rf{facthetaper}, with an arbitrary prefactor $\exp(c(x,t))$, for an arbitrary function $c(x,t)$, entire in $x$. Substitution of this ansatz in the KdV equation then determines that $c_{xx}$ is doubly periodic and meromorphic in $x$. The only such $c$ is a constant. Hence, all finite-gap elliptic solutions of the KdV equation are of the form \rf{sol}. \subsection{If $L_1/L_2\gg 1$, any number of poles $N\neq 2$ in the fundamental domain is allowed} We have already argued that the equations (\ref{usolhyp}) and (\ref{soldynsys}-b) are obtained from \rf{sol} and (\ref{eldynsys}-b) in the limit $L_1\rightarrow \infty$. On the other hand, \rf{weierpois} can be regarded as a perturbative expansion of the soliton case, which corresponds to its first two terms. Because \rf{weierpois} converges uniformly away from $x=n_1L_1+i n_2 L_2$, for arbitrary integers $n_1, n_2$ we conclude that for large (but finite) values of $L_1/L_2$ any real, nonsingular $g$-soliton solution with rationally related wave numbers $(k_1:k_2:\ldots:k_g)=(p_1:p_2:\ldots:p_g)$ has an elliptic deformation with real period $L_1$ and imaginary period $i L_2$, which is real and nonsingular\footnote{That this deformation is nonsingular follows from the fact that the limit of such a deformation is the original soliton solution: the limit of a singular solution results in a singular soliton solution. This is impossible, because only nonsingular soliton solutions are considered, hence the elliptic deformations of nonsingular soliton solutions are nonsingular}. Since $N=p_1+p_2+\dots+p_g\neq 2$ is arbitrary in the soliton case, this is also true for these elliptic deformations of the solitons. Hence the constraint \rf{elcons} is solvable for arbitrary $N$, for $L_1\gg L_2$. At this point, it is appropriate to remark that if one is interested in elliptic solutions of the Kadomtsev-Petviashvili (KP) equation, \begin{equation}\la{kp} \partial_x\left(-u_t+6 u u_x+u_{xxx}\right)+3 \sigma^2 u_{yy}=0 \end{equation} \noindent (with parameter $\sigma$), then the equations \rf{sol}, (\ref{eldynsys}-b) are replaced by \cite{chud} \setcounter{saveeqn}{\value{equation} \begin{eqnarray}\la{kp1} u(x,y,t)&=&-2 \sum_{n=1}^N\wp(x-x_n(y,t)),\\ \pp{x_n}{t}&=&3 \sigma^2 \left(\pp{x_n}{y}\right)^2+12 \sum_{j=1, j\neq n}^N\wp(x_j-x_n),\\ \sigma^2\ppn{2}{x_n}{y}&=&- 16 \sum_{j=1, j\neq n}^N\wp'(x_j-x_n). \end{eqnarray} \setcounter{equation}{\value{saveeqn} \noindent This clarifies the appearance of the constraint \rf{elcons} on the motion of the poles of elliptic solutions of the KdV equation, where the poles are independent of $y$. The search for $y$-independent solutions of the KP equation reduces it to the KdV equation and it reduces the equations (\ref{kp1}-c) to \rf{sol}, (\ref{eldynsys}-b), forcing the poles to remain on the invariant manifold defined by \rf{elcons}. For the KP equation, no such constraint exists and hence the number of poles $N$ in the fundamental domain can be any integer, not equal to two. \subsection{For any $N>4$, nonequivalent configurations exist, for $L_1/L_2$ sufficiently large} Again, we only consider solutions with nonsingular soliton limits; $i.e.,$ the elliptic deformations mentioned above. Consider the asymptotic behavior for $t\rightarrow -\infty$ of $ \lim_{L_1 \rightarrow \infty} u(x,t)$. In this soliton limit, as $t \rightarrow -\infty$, the poles are collected in groups corresponding to one-soliton solutions. In this section, two configurations are called {\em nonequivalent} if the above asymptotic behavior results in a different grouping of the poles. For $N=3$, all configurations are equivalent to one configuration. In the limit $L_1\rightarrow \infty$, this configuration corresponds to the two-soliton case with $k_1:k_2=2:1$. This configuration is discussed in Section \ref{sec:dubnov}. For $N=4$, all configurations are again equivalent to one configuration. This configuration corresponds to the two-soliton case with $k_1:k_2=3:1$. Recall that $k_1$ and $k_2$ are not allowed to be equal, hence a configuration with two poles to the left and two poles to the right does not exist. Another way of expressing that only one configuration exists is that $N=4$ can only be decomposed as the sum of distinct positive integers without common factor as $N=3+1$. Again, all $N=4$ configurations are equivalent. This configuration is discussed in Section \ref{sec:4poles}. That section also discusses another example of an $N=4$ potential which does not have a nonsingular soliton limit. This potential is a special case of one of the Treibich-Verdier potentials \rf{tv}. Any integer $N>4$ can be written as a sum of distinct positive integers without overall common factor in more than one way\footnote{$5=1+4=2+3$, $N=1+(N-1)=1+2+(N-3)$, for $N>5$}. Let the number of terms in the $m$-th decomposition of $N$ be denoted as $N_m$, then $N=\sum_{i=1}^{N_m}n_i$, with the $n_i$ distinct and having no overall common factor. This configuration corresponds to the $N_m$-soliton case with wavenumber ratios $(k_1:k_2:\ldots :k_{N_m})=(n_1:n_2:\ldots :n_{N_m})$. A solution with these wave numbers has $N_m$ phases and is an $N_m$-soliton solution. Hence for any $N>4$ there exist at least as many different configurations as there are decompositions of $N$ into distinct positive integers, without overall common factor. These configurations need not have the same number of phases. Two nonequivalent configurations corresponding to $N=5$ are discussed in Section \ref{sec:n=5}. \subsection{The poles only collide in triangular numbers} A collision of poles is a local process in which only the colliding poles play a significant role. The analysis of the collisions is identical to the rational and the soliton cases because close to the collision point, the Weierstra\ss~ function reduces to $1/x^2$. Kruskal \cite{kruskalpoles} already noticed that the poles do not collide in pairs, but triple collisions do occur. In fact, any triangular number of poles can participate in a collision, in which case the solution of the KdV equation at the collision time $t_c$, nearby the collision point $x_c$ is given by $u(x,t_c)=-g(g+1)/(x-x_c)^2$ \cite{amm}. Asymtotically near the collision point $x_c$, $i.e.,$ $t<t_c$, the poles lie on the vertices of a regular polygon with $g(g+1)/2$ vertices. For $t>t_c$, the poles emanate from the collision points, again forming a regular polygon with $g(g+1)/2$ vertices. If $g(g+1)/2$ is even, this polygon is identical to the polygon before the collision. If $g(g+1)/2$ is odd, the polygon is rotated around the collision point by $2 \pi/g(g+1)$ radians. Of all these collision types, the one where three poles collide (corresponding to $g=1$) is generic. It is the one observed in the examples illustrated in Section \ref{sec:ex}. Since the poles only collide in triangular numbers, it is possible that at any given time $t_c$ the solution of the KdV equation has the form \rf{gw}, with not all $g_i=1$. At almost every other time $t$, such a solution has $N=\sum_{i=1}^M g_i(g_i+1)/2$ distinct poles. \subsection{The solutions \rf{sol} are finite-gap potentials of the stationary Schr\"{o}dinger equation \rf{schrodinger}} By construction the solutions \rf{sol} are periodic in $x$ because they are obtained as a Weierstra\ss~ factorization of the theta function appearing in \rf{thetasol}, upon which we have imposed the double periodicity. Hence the solutions \rf{sol} are finite-gap potentials of the Schr\"{o}dinger operator. In \cite{gw2}, another proof of this can be found. For solutions that are elliptic deformations of the nonsingular solitons of Section \ref{sec:thick}, more can be said: an elliptic deformation of a $g$-soliton solution is a $g$-gap potential of the Schr\"{o}dinger equation. The reasoning is as follows: we already know that any elliptic deformation results in a finite-gap potential of the Schr\"{o}dinger equation. On the other hand, any finite gap potential of the Schr\"{o}dinger equation is of the form \rf{thetasol}. The soliton limit of such a finite-gap solution with $g$-phases is a $g$-soliton solution \cite{belokolos1}. Hence the number of phases of an elliptic deformation of a $g$-soliton solution is equal to $g$. This limit is the soliton limit of the periodic solutions, in which the fundamental domain reduces to the fundamental strip. In order to have a $g$-soliton solution of the KdV equation, we remarked in Section \ref{sec:thick} that at least $N=g(g+1)/2=1+2+\ldots+g$ poles are required in the fundamental strip. Hence, this many poles are required in the fundamental domain to obtain a $g$-gap potential of the Schr\"{o}dinger equation that is an elliptic deformation of a nonsingular soliton solution. \section{Examples}\la{sec:ex} In this section, some explicit examples of elliptic solutions of the KdV equation are discussed. These are illustrated with figures displaying the motion of the poles in the fundamental domain. Other figures display the solution of the KdV equation $u(x,t)$ as a function of $x$ and $t$. All these figures were obtained from numerical solutions of the corresponding constrained dynamical system. In all cases, the constrained dynamical system was solved using a projection method: the dynamical system \rf{eldynsys} is used to evolve the system for some time. Subsequently, the new solution is projected onto the constraints \rf{elcons} to correct numerical errors, after which the process repeats. In all examples given, $L_1=4$ and $L_2=\pi$. This seems to indicate that one can wander far away from $L_1/L_2\gg 1$ and still obtain soliton-like elliptic solutions of the KdV equation. This is not surprising as \rf{weierpois} indicates that as perturbation parameter on the soliton case one should use $\epsilon=\exp(-2 \pi L_1/L_2)$. For the values given above, this gives $\epsilon=0.00034$. \subsection{The solution of Dubrovin and Novikov \cite{dubnov}: $\mbf{N=3}$}\la{sec:dubnov} Dubrovin and Novikov \cite{dubnov} integrated the KdV equation with the Lam\'{e}-Ince potential $u(x,0)=-6\wp(x-x_c)$ as initial condition. They found the solution to be elliptic for all time, with $N=3$. They gave explicit formulae for the solution, which they remarked was probably the simplest two-gap solution of the KdV equation. The dynamics of the poles in the fundamental domain is displayed in Fig. \ref{fig3}a. Fig. \ref{fig3}b displays the corresponding two-phase solution of the KdV equation. Animations of the behavior of the poles and of $u(x,t)$ as $t$ changes are also available at {\tt http://amath-www.colorado.edu/appm/other/kp/papers}. Notice the soliton-like interactions of the two phases in the solution. In terms of the classification of Lax \cite{lax2}, these are interactions of type (c) ($i.e.,$ $u(x,t)$ has only one maximum while the larger wave overtakes the smaller wave). From Fig. \ref{fig3}a, it appears that the Dubrovin-Novikov solution is periodic in time. This was indeed proven by \`{E}nol'skii \cite{enolskii}. \begin{figure} \begin{tabular}{cc} \psfig{file=21pole.eps,width=3in}& \psfig{file=21u.eps,width=3in}\\ (a)&(b) \end{tabular} \caption{\label{fig3} {\bf The solution of Dubrovin and Novikov, with $L_1=4$ and $L_2=\pi$. (a) The motion of the poles in the fundamental domain. The initial position of the poles is indicated by the black dots. The arrows denote the motion of the poles. (b) The KdV solution $u(x,t)$}} \end{figure} For this specific solution only one of the three constraint equations is independent: since the derivative of the Weierstra\ss~ function is odd, the sum of the constraints is zero. Furthermore, labelling the three poles by $x_1, x_2$ and $x_3$, for reality $x_2=x_1^*+i L_2$ and $x_3$ is on the centerline. Hence the second constraint is the complex conjugate of the first constraint. The constraints \rf{elcons} reduce to the single equation \begin{equation}\la{dubnovcons} \wp'(x_1-x_1^*)+\wp'(x_1-x_3)=0. \end{equation} \noindent This equation was solved numerically to provide the initial condition shown in Fig. \ref{fig3}a. The initial guess required for the application of Newton's method is based on the knowledge of the soliton limit. In that case two poles on the right represent a faster soliton, one pole on the left represents the slower soliton. The periodic case is not that different: the vertical line of poles with the smallest vertical distance between poles has poles closer to the real $x$-axis than the others and correspond to the wave crest with the highest amplitude, as seen in Fig. \ref{fig3}b. We refer to the Dubrovin-Novikov solution as a $(2:1)$-solution because of the natural separation of the poles in a group of 2 poles $(x_1,x_2)$ and a single pole $(x_3)$. Equating $x_1=x_3+\epsilon$ and only condering the singular terms of \rf{dubnovcons}, it is possible to examine the location of the poles close to a collision points $x_c$. With $\wp'(x)=-2/x^3$ in this limit and $\epsilon=\epsilon_r+i \epsilon_i$, one finds \begin{equation}\la{closetocol} \epsilon_r^3-3 \epsilon_i^2\epsilon_r=0,~~~~ 9 \epsilon_i^3-3 \epsilon_r^2 \epsilon_i=0. \end{equation} \noindent This set of equations has three solutions, corresponding to the three distances between the poles: $\epsilon_r/\epsilon_i\in\left\{0, \sqrt{3}, -\sqrt{3}\right\}$. This allows for two triangular configurations of the poles: an equilateral triangle pointing left of the collision point and one pointing right. Using the dynamical system \rf{eldynsys} in the same way and only retaining singular terms results in \begin{equation}\la{dubnovdynsys} \dot{\epsilon}_r=-\frac{3}{4 \epsilon_i^2}+3 \frac{\epsilon_r^2-\epsilon_i^2}{\left(\epsilon_r^2+\epsilon_i^2\right)^2},~~~~ \dot{\epsilon}_i=\frac{6 \epsilon_r \epsilon_i}{\left(\epsilon_r^2+\epsilon_i^2\right)^2}. \end{equation} \noindent Since the constraints \rf{elcons} are invariant under the flow \rf{eldynsys}, the solutions to \rf{closetocol} give invariant directions of the system \rf{dubnovdynsys}. Along these invariant directions, one obtains ordinary differential equations for the motion of the poles as they approach the collision point. It follows from these equations that the poles approach the collision point $x_c$ with infinite velocity. Integrating the equations with initial condition $\epsilon(t_c)=0$ gives \begin{equation}\la{scalinglaw} \epsilon=\frac{3^{5/6}}{2}(\sqrt{3}+i)(t_c-t)^{1/3}. \end{equation} \noindent Using the three branches of $(t_c-t)^{1/3}$ results in the dynamics of each edge of the equilateral triangle. If $t<t_c$ this triangle is pointing left, for $t>t_c$ it is pointing right. \subsection{$\mbf{N=4}$: an elliptic deformation and a Treibich-Verdier solution}\la{sec:4poles} The next solution we discuss has 4 poles in the fundamental domain and is an elliptic deformation of a soliton solution. In the limit $L_1\rightarrow \infty$, this solution corresponds to a two-soliton solution with wavenumber ratio $k_1/k_2=3/1$, so this solution is refered to as a $(3:1)$- solution. \begin{figure}[ht] \centerline{\psfig{file=31pole.eps,width=3in}} \vspace*{-0.5in} \caption{\label{fig5} {\bf $N=4$: the pole dynamics of a $(3:1)$-solution. The initial positions of the 4 poles are indicated. The arrows on the centerline indicate that the poles there move in both directions.}} \end{figure} The motion of the poles in the fundamental domain is displayed in Fig. \ref{fig5}. Corresponding to the given wavenumber ratio, the amplitute ratio of the two phases present in the solution is roughly $k_1^2/k_2^2=9/1$. As a consequence, the form of $u(x,t)$ is not very illuminating and it has been omitted. Animations with the time dependence of both the positions of the poles and of $u(x,t)$ are again available at {\tt http://amath-www.colorado.edu/appm/other/kp/papers}. Note that the poles of the $(3:1)$-solution do not collide. This is in agreement with the results of Thickstun \cite{thickstun} who outlined which configurations lead to collisions and which do not, in the hyperbolic case. As mentioned before, the examination of collision behavior is essentially local and no differences appear between the rational, hyperbolic and elliptic cases. Another configuration with $N=4$ exists. Consider the potential \begin{equation}\la{tv4} u(x,t=t_c)=-2 \wp(x-x_0)-6 \wp(x-x_0-\omega_1), \end{equation} \noindent with $x_0$ on the centerline. This is a Treibich-Verdier potential, obtained from \rf{gw} with $M=2$, $g_1=1$, $g_2=2$, $\alpha_1=x_0$ and $\alpha_2=x_0+\omega_1$. It is referred to as a Treibich-Verdier potential because the position of the poles is given in terms of the periods of the Weierstra\ss~ function, as in \rf{tv}. Also, it can be obtained from \rf{tv} as a degenerate case. As before $N=\sum_{i=1}^{2}g_i(g_i+1)/2=4$, hence for all times that are not collision times, this solution has 4 distinct poles in the fundamental domain. The time $t=t_c$ is a collision time. Immediately after the collision time $t=t_c$, the 3 poles located at $x_0+\omega_1$ separate, as in the Dubrovin-Novikov solution, along an equilateral triangle. The result appears to be a three-phase solution. However, it is known that the potential \rf{tv4} is a two-gap potential of the Schr\"{o}dinger equation and its hyperelliptic Riemann surface is given explicitly in \cite{belokolos1}. This solution is not an elliptic deformation of a nonsingular soliton solution and the separation into different phases does not make sense. This is also seen from the following argument: if, for a fixed time which is not a collision time, we attempt to take the limit as $L_1=2 \omega_1\rightarrow \infty$, the poles seem to separate in three distinct solitons with respective wave numbers $(k_1:k_2:k_3)=(2:1:1)$. Such a nonsingular soliton solution does not exist for the KdV equation and the separation into different phases does not make sense. The dynamics of the poles is illustrated in Fig. \ref{figtv4}a. The corresponding KdV solution is shown in Fig. \ref{figtv4}b. \begin{figure} \begin{tabular}{cc} \psfig{file=211pole.eps,width=3in}& \psfig{file=211u.eps,width=3in}\\ (a)&(b) \end{tabular} \caption{\label{figtv4} {\bf An $N=4$, $M=2$ Treibich-Verdier solution, with $L_1=4$ and $L_2=\pi$. (a) The motion of the poles in the fundamental domain. The initial position of the poles is indicated by the black dots. The initial time $t=0$ was chosen different from the collision times $t_c$. The arrows denote the motion of the poles. (b) The KdV solution $u(x,t)$}} \end{figure} The dynamics of the poles illustrated in Fig. \ref{figtv4}a exhibits behavior that appears qualitatively different from any other solution discussed here. The trajectories traced out by the motion of the poles in the fundamental domain appear to have singular points (cusps), away from the collision points. Upon closer investigation, these ``cusps'' are only a figment of the resolution of the plot and the poles trace out a regular curve as a function of time, away from the collision times. Exactly why the global pole dynamics of the Treibich-Verdier potential \rf{tv4} under the KdV flow appears so different from the pole dynamics of elliptic deformations of soliton solutions of the KdV equation is an open problem. Another question one may ask is whether similar behavior is observed for other solutions originating from Treibich-Verdier potentials. \subsection{$\mbf{N=5}$: two different possibilities}\la{sec:n=5} For $N=5$, two soliton configurations are possible, and corresponding to each of these is an elliptic solution. The first solution is a $(4:1)$-solution. The second solution is a $(3:2)$-solution. \begin{figure}[p] \begin{tabular}{cc} \multicolumn{2}{c}{\psfig{file=41pole.eps,width=3in}}\\ \multicolumn{2}{c}{(a)}\\ \psfig{file=32polenew.eps,width=3in}& \psfig{file=32unew.eps,width=3in}\\ (b)&(c) \end{tabular} \caption{\label{fig6} {\bf $N=5$: (a) The pole dynamics of a $(4:1)$-solution in the fundamental domain. (b) The pole dynamics of a $(3:2)$-solution in the fundamental domain. (c) The KdV solution $u(x,t)$ corresponding to the pole dynamics in (b). In (a) and (b), the initial positions of the poles are indicated. Both solutions are quasiperiodic in time. $u(x,t)$}} \end{figure} The $(4:1)$-solution offers no new pole-dynamics: initially 1 pole is located on the centerline, at the left in the fundamental domain. The other poles are located at the right of the fundamental domain, symmetric with respect to the centerline. The three poles closest to the centerline interact as the $(2:1)$-solution. The two outer poles behave as the two outer poles of the $(3:1)$-solution. The pole dynamics of the $(4:1)$-solution is displayed in Fig. \ref{fig6}a. The $(3:2)$-solution is more interesting. It is displayed in Fig. \ref{fig6}c, together with the motion of the poles in the fundamental domain \ref{fig6}b. Again, the two crests of $u(x,t)$ interact in a soliton-like manner. In Lax's classification \cite{lax2}, this is an interaction of type (a), where at every time two maxima are observed. Fig. \ref{fig6}b only displays the motion of the poles for a short time, in order not to clutter the picture. The motion of the poles is presumably quasiperiodic in time, as is the case for the $(4:1)$-solution. It appears that the two poles above (or below) the middle line of the fundamental domain share a common trajectory. It is an open problem to establish whether or not this is the case. \subsection{$\mbf{N=6}$: two different possibilities. A three-phase solution} For $N=6$, two distinct pole configurations are possible. The first one corresponds to a $(5:1)$-solution and results in a two-gap potential of the Schr\"{o}dinger equation. It essentially behaves as the $(3:1)$-solution with two more poles added, which also behave as the outer poles of the $(3:1)$-solution. The second configuration is a $(3:2:1)$-solution, which limits to a three-soliton solution with wavenumber ratio $(k_1:k_2:k_3)=(3:2:1)$. This elliptic solution is a three-phase solution of the KdV equation. \begin{figure}[p] \centerline{\psfig{file=321pole.eps,width=4in}} \vspace*{-0.7in} \caption{\label{fig8} {\bf The pole dynamics of a $(3:2:1)$-solution. The black dots mark the initial position of the poles; the grey dots mark the position of the poles at $t=0.4$.}} \centerline{\psfig{file=321u.eps,width=4in}} \caption{\label{fig9} {\bf A $(3:2:1)$-solution of the KdV equation, corresponding to the pole dynamics in Fig. \ref{fig8}}} \end{figure} The amplitude ratio of the $(3:2:1)$-solution is $(9:4:1)$, which explains why the third phase is hard to notice in Fig. \ref{fig9}. Animations of the pole dynamics and of the time dependence of the $(3:2:1)$-solution are available at {\tt http://amath-www.colorado.edu/appm/other/kp/papers}. \section*{Acknowledgements} The authors acknowledge useful discussions with B. A. Dubrovin, S. P. Novikov, C. Schober and A. P. Veselov. This work was carried out at the University of Colorado and the Mathematical Sciences Research Institute. It was supported in part by NSF grants DMS 9731097 and DMS-9701755. \bibliographystyle{plain}
\section{Introduction} The study of viscoelastic fluids is of great scientific interest and industrial relevance. Viscoelastic fluids are fluids that show not only a viscous flow response to an imposed stress, as do Newtonian fluids, but also an elastic response. Viscoelastic effects are almost universally observed in polymeric liquids\cite{bird}, where they often dominate the flow behavior. They can also be observed in simple fluids, especially in high frequency testing\cite{boon} or in under-cooled liquids\cite{goetze}. Because most research into viscoelastic liquids, especially that with an eye toward engineering applications, is directed toward polymeric liquids, the viscoelastic behavior of simple liquids is not as well known among researchers. The fact that the manifestation of viscoelasticity does not require the presence of polymer molecules is at the heart of our approach, as will become clear in the description of the viscoelastic model. Although in most practical problems involving polymeric materials the viscosities of the materials involved are so large that the creeping flow approximation is valid, the non-linearity introduced by the viscoelastic response of the liquid makes it difficult to treat any but the most simple cases analytically. In engineering applications the situation is often further complicated by the fact that the system is comprised of several immiscible or partially miscible components with different viscoelastic properties. Examples of this include polymer blending, where two immiscible polymers are melted and mixed in an extruder, and the recovery of an oil-and-water mixture from porous bed rock. Simulation of these systems is very important, but due to the complexities only few numerical approaches exist to date. Boundary element methods have been used to simulate such systems with varying degrees of success, but the allowable complexity of the interface morphology is very limited in such approaches. Lattice Boltzmann simulations have been shown to be very successful for Newtonian two-component systems with complex interfaces\cite{PRL}, but for viscoelastic fluids the lattice Boltzmann models, derived by Giraud {et al.}\cite{giraud_epl,giraud}, are limited to one-component systems. \begin{figure} \begin{center} \begin{minipage}{4cm} \centerline{\psfig{figure=BubbleE2.eps,width=3.8cm}} \begin{center} (a) \end{center} \end{minipage} \begin{minipage}{4cm} \centerline{\psfig{figure=sBubbleS2.eps,width=3.94cm}} \begin{center} (b) \end{center} \end{minipage} \end{center} \caption{ (a) An air bubble rising in $\mbox{Ivory}^{\bigcirc \!\!\!\!\!\mbox{\tiny \rm R}}$ soap and (b) the simulation results for a low viscosity drop rising in a Oldroyd B fluid on a 256x1024 lattice for a drop of radius $R_0=35$. The simulated bubble has a cusp that is rounded at the tip due to the finite thickness of the interface ($\sim 3$ lattice spacings).} \label{fig:Bubble} \end{figure} In this article we report the successful combination of both two-component and viscoelastic features into a two-dimensional lattice Boltzmann model. We used this model to simulate a bubble rising in a viscoelastic liquid (see Figure 1) and in this letter report the first successful simulation of the experimentally observed cusp. \section{Lattice Boltzmann} We use a two-dimensional lattice Boltzmann model on a square lattice with a velocity set of $\{{\bf v}_i\}=\{(0,0)$, $(0,1)$, $(1,0)$, $(0,-1)$, $(-1,0),$ $(1,1),$ $(-1,1)$, $(-1,-1)$, $(1,-1)\}$ and a corresponding set of densities $\{ f_i \}$, but following Giraud {\it et al.}\cite{giraud} we introduce two densities for each non-zero velocity. We use a BGK lattice Boltzmann equation that contains the full collision matrix $\Lambda_{ij}$ \begin{eqnarray} \label{eqn:lb}\label{feqn} f_i(x+v_i \Delta t, t+\Delta t) &=& f_i(x,t)\nonumber\\ &&+\Delta t\Lambda_{ij} (f_j^0(x,t)-f_j(x,t)) \end{eqnarray} where the summation rule for repeated indices is implied and the required properties of the equilibrium distributions $f_i^0$ are discussed below. The local density is given by $\rho=\sum_i f_i$ and the momentum by $\rho {\bf u}=\sum_i f_i {\bf v}_i$. In order to simulate a two-component mixture we define a second set of nine densities, $\{ g_i\}$, with an appropriate equilibrium distribution, $\{ g_i^0\} $. These densities represent the density difference of the two components A and B as $\phi=\sum g_i=\rho_A-\rho_B$, where the total density introduced earlier is $\rho = \rho_A +\rho_B$. For the $g_i$s we choose a single relaxation time lattice Boltzmann equation \begin{eqnarray} g_i(x+v_i \Delta t,t+\Delta t)&=&g_i(x,t)\nonumber\\ &&+\frac{\Delta t}{\tau} (g_i^0(x,t)-g_i(x,t)),\label{geqn} \end{eqnarray} where $\tau$ is the relaxation time and $g_i^0$ is the equilibrium distribution. To use the lattice Boltzmann method in order to simulate fluid flow, mass and momentum conservation have to be imposed. Mass and momentum conservation are equivalent to constraints on the equilibrium distributions: \begin{equation} \sum_i f_i^0=\rho,\;\;\; \sum_i g_i^0= \phi,\;\;\; \sum_i f_i^0 {\bf v}_i= \rho {\bf u}. \label{momentum} \end{equation} There will be further constraints on the permissible equilibrium distributions in order for the corresponding macroscopic equations to be isotropic and to simulate the systems in which we are interested. In the next two subsections we will summarize the physics that we want to incorporate and then we will discuss how it imposes constraints on the equilibrium distributions and eigenvalues. \subsection{Binary mixtures} To simulate a binary mixture we follow the approach of Orlandini {\it et al.} \cite{enzo} and begin with a free energy functional $\Psi$ that consists of the free energy for two ideal gases and an interaction term as well as a non-local interface term: \begin{eqnarray} \Psi[\rho_A,\rho_B]&=&\int_{\bf x} \left[T\rho_A \ln(\rho_A) + T\rho_B \ln(\rho_B)\nonumber\right.\\ &&\left. + \lambda \rho_A \rho_B+\kappa |\partial_{\bf x} (\rho_A-\rho_B)|^2 \right] d{\bf x}, \end{eqnarray} where the densities $\rho_A$ and $\rho_B$ are functions of ${\bf x}$. The repulsion of the two components is introduced in the $\lambda$ term and $\kappa$ is a measure of the energetic penalty for an interface. When we write this free energy functional in terms of the total density, $\rho$, and the density difference, $\phi$, we can derive the chemical potential, $\mu$, and the pressure tensor, $P_{\alpha\beta}$, as\cite{shear}: \begin{eqnarray} \mu&=&\frac{\delta \Psi}{\delta \phi}=\partial_\phi \Psi -\kappa \partial_\gamma \partial_\gamma \phi,\\ P_{\alpha\beta} &=& (\rho \partial_\rho \Psi+ \phi \partial_\phi \Psi)\delta_{\alpha\beta} \nonumber \\ && + \kappa (\partial_\alpha \phi \partial_\beta \phi -\frac{1}{2} \partial_\gamma \phi \partial_\gamma \phi \delta_{\alpha\beta} - \phi \partial_\gamma \partial_\gamma \phi \delta_{\alpha\beta}), \end{eqnarray} where $\delta$ indicates a functional derivative and $\delta_{\alpha\beta}$ is the Kronecker delta. For a two-component model we fix the further moments of the equilibrium distributions\cite{shear}: \begin{eqnarray} \sum_i g_i^0 {\bf v}_i &=& \phi {\bf u},\\ \sum_i f_i^0 v_{i\alpha}v_{i\beta}&=&P_{\alpha\beta} + \rho u_\alpha u_\beta,\\ \sum_i g_i^0 v_{i\alpha}v_{i\beta}&=&\mu \delta_{\alpha\beta} +\phi u_\alpha u_\beta. \label{pab} \end{eqnarray} Thus far, the model allows us to simulate a binary mixture that phase separates below a critical temperature of $T_c=\lambda/2$. The surface tension, $\sigma$, can be calculated analytically for a flat equilibrium interface $\phi(y)$ orthogonal to the y direction as $ \sigma=\kappa \int_{-\infty}^{\infty} (\partial_y \phi(y))^2 dy $ where the equilibrium density profile of $\phi$ also depends on $\kappa$. \section{Viscoelasticity and the Boltzmann equation} Viscoelasticity was first proposed by Maxwell in his dynamic theory of gases\cite{maxwell}. He used the simple argument that in the limit where there are no intermolecular collisions the fluid in a container should behave like a solid: ``\dots Then it can easily be shown that the pressures on the sides of the vessel due to the impacts of the molecules are perfectly independent of each other, so that the mass of moving molecules will behave, not like a fluid, but like a solid.'' He goes on to deduce that the observed viscous behavior of fluids is due to binary collisions that randomize the directions of stress in the fluid. Since the collisions are fast, but not instantaneous, the elastic properties of the fluid are not completely lost, leading to the Maxwell model of viscoelasticity. Subsequently, derivations of hydrodynamics from the dynamic theory of gases have made the approximation of a purely viscous behavior because of the difficulties of deriving a continuum approach at the length scales of a mean free path of a molecule. In gases where lengths less than the mean free path are important kinetic theory for rarefied gases is used. There has recently been much activity in the research of the experimentally observed viscoelastic behavior of simple liquids that are undercooled. In this case, however, viscoelasticity is not obtained because the relevant length scales were of the order of a mean free path, but rather because of the correlations of subsequent collisions as described in the mode coupling theory\cite{goetze}. The arguments of Maxwell, however, are still valid for describing the behavior of the Boltzmann equation, and viscoelastic properties can be derived from the Boltzmann equation if the decay of viscous stresses is slow. The approach by Giraud {\it et al.} aims not at deriving a convected Maxwell fluid, but a convected Jeffreys fluid which is a mixture of a Maxwell fluid with a Newtonian fluid. A double set of densities is introduced allowing two stresses, one of which is chosen to relax quickly and is, therefore, a viscous stress, and the other, which is chosen to decay very slowly, represents a viscoelastic stress. The resulting model is a convected Jeffreys model that is often used to describe a polymeric fluid in a solvent. Care has to be taken for the choice of the collision matrix and the equilibrium distribution to ensure an isotropic model. The details of this one-component model are described in the publication by Giraud {\it et al.}\cite{giraud} A Chapman-Enskog expansion of the lattice Boltzmann equations (\ref{feqn}) and (\ref{geqn}) gives the macroscopic equations that our system simulates. Mass conservation gives the continuity equation: \begin{equation} \partial_t \rho + \partial_\alpha (\rho u_\alpha) = 0. \end{equation} Momentum conservation gives a Navier Stokes equation: \begin{equation} \rho \partial_t u_\alpha + \rho u_\beta \partial_\beta u_\alpha = -\partial_\beta \left(P_{\alpha\beta}+\sigma^v_{\alpha\beta}+ \sigma_{\alpha\beta} \right) \end{equation} where the viscous stress is given by \begin{equation} \sigma^v_{\alpha\beta}=\nu_\infty \partial_\beta (\rho u_\alpha) +\xi_\infty \partial_\gamma (\rho u_\gamma) \delta_{\alpha\beta}. \end{equation} The viscoelastic stress has the constitutive relation \begin{equation}\label{constitutive} \sigma_{\alpha\beta} + \theta \sigma_{(1)\alpha\beta} = -(\nu_0-\nu_\infty) \left(\partial_\alpha (\rho u_\beta)+\partial_\beta (\rho u_\alpha)\right) \end{equation} where $\sigma_{(1)}$ represents the upper convected derivative of $\sigma$. These equations are equivalent to the Navier-Stokes and Jeffreys equations only in the incompressible limit where $\partial_\alpha (\rho u_\beta)=\rho \partial_\alpha u_\beta$. The fully compressible equations can only be simulated when a larger set of velocities is used\cite{thesis}. The conservation of the density difference leads to the convection diffusion equation \begin{equation} \partial_t \phi + \partial_\alpha (\phi u_\alpha) = D \partial_\alpha \partial_\alpha \mu+ \partial_\beta\left(\frac{\phi}{\rho} \partial_\alpha (P_{\alpha\beta}-\sigma_{\alpha\beta})\right) \end{equation} where $D$ is a diffusion constant given by $D=(\tau-1/2)$. \section{Simulation of a bubble in a viscoelastic liquid} We applied our method to a system similar to the experimentally well-studied system of an air bubble rising in a viscoelastic fluid. In our simulation we represent the bubble using a phase-separated Newtonian drop of low viscosity in matrix which is viscoelastic by letting the relaxation time $\theta$ in equation (\ref{constitutive}) depend smoothly on the density difference $\phi$ between $\theta=0.05$ in the drop and $\theta=66$ in the surrounding fluid. We choose $\xi_\infty=0.06$, $\nu_\infty=0.01$, and $\nu_0=0.01$ in the drop and $\nu_0=0.175$ outside. For the thermodynamic parameters we select $T=0.5$, $\lambda=1.1$, and $\kappa=0.007$, which corresponds to a surface tension of $\sigma=0.02$. All units are in terms of the lattice spacings and the time steps $\Delta t=\Delta x=1$. We introduce a forcing dependent on $\phi$ so that the bubble is forced upward while the surrounding fluid is forced downward. We choose the total change in the momentum due to the forcing to be zero so that no walls are required in the simulations. We start the simulations without forcing and then periodically increase the forcing after $10,000$ to $40,000$ iterations. We observe the change in the velocity $u$ and store the distribution of $\phi$ so that we have a way of judging the deformation of the bubble. \begin{figure} \begin{center} \begin{minipage}{4cm} \centerline{\psfig{figure=sbub1.eps,width=4cm}} \begin{center} (a) \end{center} \end{minipage} \begin{minipage}{4cm} \centerline{\psfig{figure=sbub2.eps,width=4cm}} \begin{center} (b) \end{center} \end{minipage} \vspace{0.3cm} \begin{minipage}{4cm} \centerline{\psfig{figure=sbub4.eps,width=4cm}} \begin{center} (c) \end{center} \end{minipage} \begin{minipage}{4cm} \centerline{\psfig{figure=sbub10.eps,width=4cm}} \begin{center} (d) \end{center} \end{minipage} \end{center} \caption{ Shape for the simulated bubbles for different forcings in (lattice spacings)/$\mbox{(time step)}^2$: (a) $1.6 \cdot 10^{-5}$, (b) $3.6 \cdot 10^{-5}$, (c) $9.6 \cdot 10^{-5}$, and (d) $1.96 \cdot 10^{-4}$. In (c) a fit to the predicted form of the cusp ($|x|^{2/3}$ is also shown. All simulations are after a 20,000 iterations at this forcing.} \label{fig:shapes} \end{figure} Figure \ref{fig:shapes} shows the form of the drop for different forcings. At low forcings the drop is elongated in the flow direction. This is in direct contrast to a bubble in a Newtonian fluid, which is flattened in the flow direction. At a larger forcing the bubble forms a cusp at the lower tip of the drop. For even larger forcings the drop starts to flatten in the flow direction. This sequence is in agreement with the experimental findings\cite{liu}. The elongation of a rising bubble has been simulated before\cite{noh}, but this is the first time that the formation of a cusp has been simulated. In Figure \ref{fig:shapes}(c) it is shown that the cusp can be fitted to the functional form $|x|{2/3}$ predicted by Joseph {\it et al.}\cite{joseph} for a two-dimensional cusp created by the flow induced by two couter-rotating cylinders. Experimentally the formation of a cusp has been observed to coincide with a jump of nearly an order of magnitude in the terminal velocity of the bubble\cite{bird,liu}, although the mechanism remains disputed. On the one hand Bird {\it et al.} argue that surface-active impurities tend to immobilize bubble surfaces and hence retard the motion of gas bubbles. This discontinuous change in bubble shape may be responsible for the removal of the impurities, and thus lead to a jump in the final velocity. Liu\cite{liu} {\it et al.} alternatively suggest that the change in the shape of the bubble will make it more streamlined, and therefore increase the terminal velocity. \begin{figure} \centerline{\psfig{figure=utwo.eps,width=8.5cm}} \caption{Velocity, u, of the bubble in two different sets of simulations for a bubble of radius 35 in a 256x512 lattice. On the x-axis the iterations were multiplied by a scale factor so that corresponding forcings appear at the same point in the graph. No jump in the final velocity is observed at $\sim 0.5\; 10^6$ iterations where the formation of a cusp is observed (indicated by arrow). The scale factors were 0.25 and 1. Velocity is measured in lattice units per time step.} \label{fig:uall} \end{figure} We examined the velocity for the rising drop as described above and found no jump of about half an order of magnitude as observed by Liu {\it et al.} in their experiment (see Figure \ref{fig:Bubble}). Our simulations suggest that the jump in velocity they observe is not connected to a more streamline form of the bubble due to the cusp, but more likely to the presence of surfactants that are absent in our simulations. \section{Conclusion} We introduced a lattice Boltzmann model that can simulate viscoelastic two-component flows. We gave an intuitive explanation of the origin of viscoelasticity in our model and the model by Giraud {\it et al.} in terms of the original theory of Maxwell\cite{maxwell}. Simulations using this method have succeeded in reproducing the cusp at the end of a bubble rising in a viscoelastic medium that have eluded earlier numerical attempts with a more traditional boundary integral approach. The model has been successful in the qualitative simulation of the bubble problem in two dimensions. We intend to extend the model to three dimensions in the future. This will also enable us to compare the results quantitatively with experiment. \section*{Acknowledgements} One of us (A.W.) would like to thank Brad Chamberlain for his help in implementing the algorithm in ZPL\cite{ZPL}. We would also like to thank the Scientific Computing and Visualization center at Boston University for a Mariner grant. L.G. is grateful for its support by the ARC 97/02-210 project, Communaut\'e Fran{\c c}aise de Belgique. \vspace{-0.7cm} \def\jour#1#2#3#4{{#1} {\bf #2}, #3 (#4)}. \def\tbp#1{{\em #1}, in preparation.} \def\tit#1#2#3#4#5{{#1} {\bf #2}, #3 (#4).} \defAdv. Phys.{Adv. Phys.} \defEuro. Phys. Lett.{Euro. Phys. Lett.} \defPhys. Fluids A{Phys. Fluids A} \defPhys. Rev. Lett.{Phys. Rev. Lett.} \defPhys. Rev.{Phys. Rev.} \defPhys. Rev. A{Phys. Rev. A} \defPhys. Rev. B{Phys. Rev. B} \defPhys. Rev. E{Phys. Rev. E} \defPhysica A{Physica A} \defPhysica Scripta{Physica Scripta} \defZ. Phys. B{Z. Phys. B} \defJ. Mod. Phys. C{J. Mod. Phys. C} \defJ. Phys. C{J. Phys. C} \defJ. Phys. Chem. Solids{J. Phys. Chem. Solids} \defJ. Phys. Cond. Mat{J. Phys. Cond. Mat} \defJ. Fluids{J. Fluids} \defJ. Fluid Mech.{J. Fluid Mech.} \defAnn. Rev. Fluid Mech.{Ann. Rev. Fluid Mech.} \defProc. Roy. Soc.{Proc. Roy. Soc.} \defRev. Mod. Phys.{Rev. Mod. Phys.} \defJ. Stat. Phys.{J. Stat. Phys.} \defPhys. Lett. A{Phys. Lett. A} \defInt. J. Mod. Phys. C{Int. J. Mod. Phys. C}
\section{Introduction} Stochastic Burgers equation has been studied extensively because of its close connection with the Navier-Stokes equation. Recently it has been found that Burgers equation with correlated noise shows intermittency \cite{Chek1,Poly,Hayo}, however, these calculations have been done for specific degrees of noise correlation. In this paper we vary the applied noise from uncorrelated regime to strongly correlated regime and study the energy spectrum of $h$ and $u$, as well as intermittency exponents and the probability densities of $u$ and $du/dx$, where $u(x,t)$ is the velocity function appearing in the Burgers equation, and $h(x,t)$ is the surface height of KPZ equation. The one-dimensional Burgers equation \cite{Burg} is \begin{equation} \label{Burgers} \frac{\partial u}{\partial t}+ \lambda u \frac{\partial u}{\partial x} = \nu \frac{\partial^2 u}{\partial{x^2}} +\zeta, \end{equation} where $u$ is the velocity field, $\lambda$ is the strength of the nonlinear term ($\lambda=1$ in the standard equation), $\nu$ is the viscosity, and $\zeta$ is the noise. We assume that the noise $\zeta(k,t)$ is gaussian ($\zeta(k)$ is the Fourier transform of $\zeta(x)$), and in Fourier space follows distribution \begin{equation} \label{Burgnoisek} <\zeta(k,t)\zeta(k^{\prime },t^{\prime })>=2Dk^{-2\sigma }\left( 2\pi \right) ^2\delta (k+k^{\prime })\delta (t-t^{\prime }) \end{equation} Tatsumi and Kida \cite{Kida1}, Kida \cite{Kida2}, Gotoh \cite{Goto}, and Bouchaud et al. \cite{Bouc} have solved the noiseless Burgers equation ($\zeta=0$) exactly in limit of large time $t$ and zero viscosity. The solution is a linear profile with sharp discontinuities, which are called shocks. This shock solution yields structure function $<|u(x+r,t)-u(x,t)|^q> \propto r$ indicating that the $u$ is intermittent. The solution also yields the energy spectrum $<|u(k)|^{2}/2> \propto k^{-2}$ ($u(k)$ is the Fourier transform of $u(x)$). The noisy Burgers equation as well as the noisy Kardar-Parisi-Zhang (KPZ) equation have been studied by many authors. The Kardar-Parisi-Zhang (KPZ) equation, which describes a generic set of surface growth phenomena, is closely related to the Burgers equation. The replacement of $u$ by $-\partial h/ \partial x$ in the Burgers equation yields KPZ equation: \begin{equation} \label{KPZ} \frac{\partial h}{\partial t}=\frac{\lambda}{2} \left( \frac{\partial h}{\partial x}\right)^2 + \nu \frac{\partial^2 h}{\partial x^2} +f({\bf x},t), \end{equation} where $h(x,t)$ is the height of the surface profile at position $x$ and at time $t$, $\lambda $ is the strength of the nonlinearity, $\nu $ is the diffusion coefficient, and $f$ is the forcing function. We again assume that the noise $f(k,t)$ is gaussian, and in Fourier space follows distribution \begin{equation} \label{KPZnoisek} <f(k,t)f(k^{\prime },t^{\prime })>=2Dk^{-2\rho }\left( 2\pi \right) ^2\delta (k+k^{\prime })\delta (t-t^{\prime }). \end{equation} It is easy to see that $|u(k)|^{2} = k^2 |h(k)|^2$ and $\sigma=\rho-1$. In the following several paragraphs we will describe the work done by various researchers regarding energy spectrum, intermittency exponents, and probability densities of $h$ and $u$ when noise is present in the system. Eq. (\ref{KPZ}) has been solved by Kardar-Parisi-Zhang \cite{KPZ} for $\sigma=-1$ and by Medina et al. \cite{Medi} for $-1< \sigma \leq 0$ using renormalization group analysis. They have calculated the roughening exponents $\chi_{KPZ}$ and $\beta_{KPZ}$, which characterize the dynamical properties of the equation, e.g., $<|h(k)|^2> \propto k^{-2 \chi_{KPZ} -1}$. The RG analysis is not applicable beyond $\sigma =0$. Using the connection of the Burgers equation with the KPZ equation, the dynamical exponents of the Burgers equation is also automatically solved for this set of $\sigma$. Recently, Chattopadhyay and Bhattacharjee \cite{ChatJKB} have applied mode coupling scheme and obtained the same exponents for KPZ equation for $-1 \le \sigma \le 0$. For $\sigma=-1$, Barabasi and Stanley \cite{Bara} have shown that in 1D KPZ equation, the function $dh/dx (=-u)$ has a gaussian probability density under steady-state. Since $h'$ is gaussian, we can immediately determine the structure function $T_q(r)=<|h(x+r)-h(x)|^q>$ using the following arguments. Since $\Delta h \approx r h'$ for small $r$, $\Delta h [=h(x+r)-h(x)]$ will also have a gaussian probability density, i.e., \begin{equation} P(\Delta h)=\frac{1}{\sigma_r \sqrt \pi} \exp\left( -\frac{(\Delta h)^2}{\sigma_r^2} \right) . \end{equation} Therefore, \begin{eqnarray} <|\Delta h|^q> & = & \int_{-\infty}^{\infty} P(\Delta h) |\Delta h|^q d\Delta h \nonumber \\ & \propto & \sigma_r^q. \end{eqnarray} In the region where $<|\Delta h|^q>$ is a powerlaw (region of our interest), $\sigma_r$ will also be a powerlaw, say, $\sigma_r \propto r^\chi$. Therefore, $<|\Delta h|^q> \propto r^{\chi q}$. The above discussion clearly shows that if the derivative of the function has a gaussian probability density, then the exponent of the $q-$th order structure function (denoted by $\zeta_q$) is proportional to $q$. These kind of functions are called {\em non-intermittent} functions. On the other hand, if the probability density of the derivative is nongaussian, or the exponent of $q-$th order structure function is not proportional to $q$, then the function is said to be {\em intermittent}. The intermittent functions have typically powerlaw tails instead of gaussian tails in their probability densities, hence, probability of occurrence of large value of the function is higher for the intermittent functions then the non-intermittent functions. In other words, large events, which are absent in non-intermittent systems, occur in intermittent systems in a bursty manner. In fluid turbulence, it has been found that $P(u)$ is gaussian, whereas $P(u'_L)$ and $P(u'_T)$ (where $L$ and $T$ denote longitudinal and transverse components) have powerlaw tails for small $r$. The exponents of the structure functions are not proportional to $q$ either. Hence, fluid turbulence is said to have intermittency. These observations are explained using localized vortices. For further discussions on intermittency in fluid turbulence, refer to Frisch \cite{Frisbook} and references therein. Chekhlov and Yakhot \cite{Chek1}, and Hayot and Jayaprakash \cite{Hayo} numerically solved the structure function of the Burgers equation with correlated noise for a range of $\sigma$. For $\sigma=1/2$, Chekhlov and Yakhot \cite{Chek1} obtained $<|\Delta u(r)|^q> \propto r$ and claimed that the Burgers equation is intermittent according to the above definitions. Hayot and Jayprakash \cite{Hayo} found similar results for $0 \le \sigma \le 1/2$. This behaviour was attributed to the shocks present in the system \cite{Chek1,Hayo}. Regarding the energy spectrum, Chekhlov and Yakhot \cite{Chek1} obtained Kolmogorov's energy spectrum, i.e., $|u(k)|^{2} \propto k^{-5/3}$ for $\sigma=1/2$. They argued the constancy of energy flux to be the reason for Kolmogorov's spectrum in the noisy Burgers equation with $\sigma=1/2$. In this paper we will show that energy flux is constant for all $\sigma \ge 1/2$, yet the spectral exponent varies from 5/3 to 2. Hence, Chekhlov and Yakhot \cite{Chek1} claim that the constancy of flux implies Kolmogorov's spectrum for Burgers equation is incorrect. Hayot and Jayaprakash \cite{Hayo} have numerically calculated the energy spectrum for $-1 \le \sigma \le 1/2$. For $-1 \le \sigma \le 0$, they observed that at large $k$ the behaviour is that of free field, and at low $k$ is in agreement with the result derived using RG treatment. In our simulation we do not find any such crossover. For $0 \le \sigma \le 1/2$, Hayot and Jayaprakash argue that the shocks determine the exponents; their numerical exponent for energy is quite close to the renormalization group formula even though RG is not expected to work in this regime. They also find that the exponent $\beta_{Burg}=1/2$ for the Burgers equation. Polyakov \cite{Poly} has applied methods of quantum field theory to calculate probability density $P(u')$ of the Burgers equation with $\sigma \ge 3/2$, and showed that \begin{equation} P(u') = \left\{ \begin{array}{ll} \exp[{-u'^{3}/(3 B_1)}] & \mbox{if $u' \rightarrow \infty$} \\ u'^{-5/2} & \mbox{if $u' \rightarrow -\infty$} \end{array} \right. \end{equation} Clearly, the probability density of $u'$ is nongaussian and has a power-law tail for negative $u'$. Boldyrev \cite{Bold1,Bold2} extended Polyakov's method to the range $1/2 \le \sigma \le 3/2$ and calculated $P(u')$. He derived that $\chi = 1+2 \sigma /3$ for $1/2 \le \sigma \le 3/4$, and $\chi=3/2$ for $\sigma > 3/4$. Boldyrev \cite{Bold1} suggests that the above formula may also be valid for $0 \le \sigma \le 3/2$. Gotoh and Kraichnan \cite{GotoKrai} found the exponent of $P(u')$ for the negative $u'$ to be $-3$. Recently, E and Eijnden \cite{E} and Kraichnan \cite{KraiBurg} argue the exponent to be $-7/2$. Gurarie and Migdal \cite{Gura}, and Balkovsky et al. \cite{Balk} have applied instanton solutions for solving probability density for positive $u'$ and obtained qualitative agreement with the numerical results reported by Yakhot and Cheklov \cite{Chek2}. In this paper we will compare our findings with the above mentioned results. The roughening exponents for KPZ equation $\chi_{KPZ} $ and $\beta_{KPZ} $ have also been calculated by Zhang by replica method \cite{Zhan}, and by Hentschel and Family using scaling arguments \cite{HentFami}. They find a close agreement with the results of Medina et al. \cite{Medi}. Numerically, Peng et al. \cite{Peng} have calculated the exponents $\chi_{KPZ}$ and $\beta_{KPZ}$ using finite-difference method, and Amar et al. \cite{Amar} and Meakin and Jullien \cite{Meak1,Meak2} have calculated using lattice simulations. Their results are in qualitative agreement with the RG predictions. The energy cascade rate is one of the important quantities of interest in the statistical theory of turbulence. In this paper we have computed the energy flux of the Burgers equation for $-1 \le \sigma <6$. We find a constant flux for all $\sigma$ beyond 1/2. In our simulation we also varied the values of the parameters ($\nu,\lambda,D$) and analyzed its effects. We find that there are interesting crossover from KPZ to Edward Wilkinson (EW) equation depending on the values of the parameter. These crossover results are reported in the Appendix A. The outline of the paper is as follows: We restate the structure function and $P(u')$ of the noiseless Burgers and KPZ equations in section 2, and the energy spectrum of these equations in section 3. From the nature of the noise-noise correlation, we divide the $\sigma$ regime in three parts. This result is discussed in section 4. Section 5 and 6 contain the simulation method and results respectively. Here we calculate the energy spectrum and structure functions of the Burgers and KPZ equations for various values of $\sigma$. The probability densities of $u'$ and $u$ is also discussed in this section. Section 7 contains a brief discussion on the cascade rate of the noisy Burgers equation. Section 8 contains discussion and conclusions. \section{Structure Function Calculation for noiseless case} We will show later that the Burgers equation with strongly correlated noise has behaviour similar to the noiseless Burgers equation. Therefore, we will discuss the noiseless equation briefly before embarking on the noisy Burgers equation. Burgers \cite{Burg} solved the Burgers equation exactly at large time $t$ for vanishing viscosity. The velocity profile at a given time is linear except for the sharp discontinuities at the shock positions (see Fig.~\ref{profilefig}). The structure function of the noiseless Burgers equation $S_q(r)$ has been calculated using this solution (see Gotoh \cite{Goto}, Bouchaud et al. \cite{Bouc} and others). Assuming the ensemble average to be the same as the spatial average, Gotoh obtained \begin{eqnarray} \label{struct_Burg} S_{q}(r) & = & \frac{1}{L} \int_{0}^{L} |u(x+r)-u(x)|^{q} dx \nonumber \\ & = & \frac{1}{L} \sum_{i} \int_{\xi_{i}}^{\xi_{i+1}-r} \left( \frac{r}{t} \right)^{q} dx +\frac{1}{L} \int_{\xi_{i+1}-r}^{\xi_{i+1}} \left( \mu_{i+1} - \frac{r}{t}\right)^{q} dx \nonumber \\ & \approx & \left( \frac{r}{t} \right)^{q} + \frac{r}{L} \sum_{i} \left( \mu_{i}- \frac{r}{t} \right)^{q} \end{eqnarray} where $L$ is the length of the box, $\mu_i$ is the shock strength (defined as the velocity difference across the shock) of the $i$th shock, and $\xi_{i}$ is the position of the $i$th shock. The second term of the equation is due to the discontinuities at the shocks. For small $r$ and $q>1$, the second term will dominate the first one. Hence, for small $r$, $S_q(r)\propto r^{\zeta^{Burg}_q} $ with \begin{equation} \label{zeta_Burg} \zeta^{Burg}_q = 1. \end{equation} Note that this relationship is valid for $ \delta \ll r \ll \Delta$, where $\delta$ is the shock width, and $\Delta$ is the average distance between two shocks. For $\nu \rightarrow 0$, $\delta$ is finite but small (for details, see Saffman \cite{Saff}). Since $u$ is continuous within the shock, $S_q(r)=r^q$ for $r \ll \delta$. For large $r$, $S_q(r)$ is not proportional to $r$ because of $(\mu_{i}-r/t)^{q}$ term of Eq.~(\ref{struct_Burg}). A point is in order here. Tatsumi and Kida \cite{Kida1} showed that the number of shock fronts decrease with $t$ as $t^{-\gamma }$, where $0\leq \gamma <1$. Hence in the asymptotic state, there will be only several shocks, and the distance between the shock fronts will be of the order of box size. Therefore, asymptotically $\Delta$ will of the order of the box size $L$, and $r$ can be comparable to $L$. As mentioned in the introduction, structure function is closely related to intermittency. Since the exponent $\zeta^{Burg}_q$ is not proportional to $q$, the noiseless Burgers equation is classified as intermittent system. However, a point to note is that the velocity of the noiseless Burgers equation is not random. From the velocity profile it is clear that $P(u)= const$ for $u$ between $u_{min}$ and $u_{max}$. The slope $u'$ is a constant ($c$) for all $u$ except within the shock region where $u'$ is large but negative. Therefore, $P(u')$ will be a sum of a delta function at $u'=c$ and a small spiky function at large but negative $u'$ (see Fig.~\ref{Pnlessfig}). Since $u$ is not random, it is somewhat confusing to call the signal as intermittent. However, it is common practice to classify the noiseless Burgers equation as an intermittent system. For fluid turbulence it has been shown that the intermittent velocity field has a multifractal distribution. This was demonstrated by Meneveau and Srinivasan \cite{Sree1,Sree2} using the cascade model. We will apply the same model to Burgers turbulence for a further understanding of the intermittent nature of the Burgers solution. In the cascade model of Meneveau and Sreenivasan \cite{Sree1,Sree2} for fluid turbulence, the nonlinear energy flux $E_r$ is distributed unequally between two smaller eddies. The flux $E_r$ at length scale $r$ is divided into fractions $p_1E_r$ and $p_2E_r$ to two smaller eddies of length $r/2$, and so on. The generalized-dimension of the resulting multifractal is given by \begin{equation} \label{Dq} D_q=\log _2\{p_1^q+p_2^q\}^{1/(1-q)} . \end{equation} The constants $D_q$ are related to the exponents of the structure function of the energy flux $E_r^q$ in eddy of size $r$ by \begin{equation} \sum E_r^q=E_L^q\left( \frac rL\right) ^{\left( q-1\right) D_q} , \end{equation} where the sum is taken over all the eddies at the $n$th stage. When $p_{1} \neq p_{2}$, the flux function is unevenly distributed after several bifurcations of the original eddy, and that yields bursty behavior for $E_r$. A closer inspection shows that the Burgers solution corresponds to $p_1\rightarrow 0$ and $p_2\rightarrow 1$, which implies that $D_q\approx 0$. Hence, for Burgers turbulence, only one among the $2^n$ eddies has all the flux after $n$ bifurcation of the original eddy. This corresponds to maximal intermittency or maximal multiscaling. Using Eq. (5.6) of Meneveau and Sreenivasan \cite{Sree2}, we get \begin{equation} \zeta^{Burg}_q=\left( q/3-1\right) D_{q/3}+1=1, \end{equation} a result consistent with Eq. (\ref{zeta_Burg}). Note that when $p_1=p_2=1/2$, $E_r$ is constant and $D_q=1$, and there is no intermittency. In Burgers equation, the dissipation occurs only at the shocks, which are narrow regions. So, as we traverse along a line, there are regions of no dissipation, then suddenly a short region (shock) with intense dissipation appears. According to this observation, the noiseless Burgers equation exhibits strong intermittency, in fact maximal intermittency, in the light of Meneveau and Sreenivasan's model \cite{Sree1,Sree2}. The structure function of the noiseless KPZ equation can be easily calculated using the relationship $h(x,t)=-\int^xu(x^{\prime },t)dx^{\prime }$. The surface profile $h(x)$ at a asymptotic time (shown in Fig.~\ref{profilefig}) is \begin{equation} h(x,t)=\left\{ \begin{array}{ll} -\frac{x^2}{2t}+\frac{\eta _i}t\left( x-\xi _i\right) & \mbox{ for $\frac{1}{2} (\xi_{i-1}+\xi_{i}) < x < \xi_{i}$} \\ -\frac{x^2}{2t}+\frac{\eta _{i+1}}t\left( x-\xi _i\right) & \mbox{ for $ \xi_{i} < x < \frac{1}{2} (\xi_{i}+\xi_{i+1})$ } \end{array} \right. \end{equation} After some algebra we can obtain the structure function $T_q(r)$ for the KPZ equation: \begin{eqnarray} \label{zeta_KPZ} T_q(r) & \approx & \frac{1}{\left( q+1\right) L}\left( \frac{r}{t}\right)^q \sum_i \left[ \left(\eta _{i+1}-\xi _{i+1}\right)^{q+1} -\left( \eta _{i+1}-\xi _i\right)^{q+1} \right] \nonumber \\ & & +\frac{1}{\left( q+1 \right) L}\left( \frac{r}{t}\right)^q r \sum_i \frac{\left[ \left( \eta _{i+2}-2\xi _{i+1}\right)^{q+1} -\left(\eta _{i+1}-2\xi _{i+1}\right) ^{q+1}\right] } {\left( \eta _{i+2}-\eta_{i+1}\right) } \end{eqnarray} The solution of the KPZ equation has cusps at positions where shocks appear in Burgers equation. These cusps yield contributions proportional to $r^{q+1}$ to the structure function. Hence, to a leading order, $T_q(r)\propto r^q$. Therefore, $\zeta^{KPZ}_q=q$. Note that smooth continuous curves also yield $\zeta_q=q$ because $h(x+r)-h(x)\approx h' r$ for small $r$. Since the exponent $\zeta^{KPZ}_q$ is proportional to $q$, the solution of the noiseless KPZ equation will be classified to be non-intermittent. However, note that the probability density of $h'=-u$ is flat (nongaussion), hence indicating intermittent behavior for the noiseless KPZ equation. Clearly there appears to be a contradiction because the structure function indicates non-intermittency, but the $P(h')$ indicates intermittency. The resolution of this apparent contradiction is given below. The systems under investigation for intermittency have typically a random output. In non-intermittent systems, the probability density of the output is gaussian, i.e., probability of finding a large signal decreases as $\exp(-x^2)$. However, in the intermittent systems, the probability density of the random output signal deviates from the gaussian behaviour. The key is that the system under investigation for intermittency must have a random output, which is not the case for the noiseless KPZ and Burgers equation. In that sense, it is somewhat meaningless to ask whether the noiseless Burgers and KPZ equations exhibit intermittency or not. This is the reason why the conclusions from the structure function and those from the probability density differ for the noiseless KPZ equation. In literature, however, the noiseless Burgers equation is classified as an intermittent system. In the following section we will derive the energy spectrum of the noiseless Burgers and KPZ equation using Eqs.~(\ref{struct_Burg}, \ref{zeta_KPZ}). \section{Energy spectrum of noiseless Burgers and KPZ equation} We briefly discuss the energy spectrum of the noiseless Burgers and KPZ equation because they will be compared with the results obtained for the equations with coloured noise. We restate the earlier results by Kida \cite{Kida2}, Gotoh \cite{Goto}, and Bouchaud et al.~\cite{Bouc}. We can obtain the energy spectrum $E^u(k)=<|u(k)|^{2}/2>$ using $S_2(r)$. Clearly, \begin{equation} <u(x+r)u(x)>= <u^2> - \frac{1}{2} S_{2}(r). \end{equation} Therefore, \begin{eqnarray} E(k) & = & \frac{1}{L} \int_{-\infty}^{\infty} 1/2 <u(x+r)u(x)> exp(-i k r) dr \nonumber \\ & = & \delta_{k,0} \frac{<u^2>}{2} + \frac{1}{2 L t^2} \frac{d^2 \delta_{k,0}}{dk^2} + (Lk)^{-2} \sum_{i} \mu_{i}^2. \end{eqnarray} Hence, the noiseless Burgers equation has energy spectrum $E^{u}(k) \propto k^{-2}$ for $k>0$. In the above equation, the Fourier transform of $r^2$ is $d^2 \delta_{k,0}/dk^2 $ \cite{Ligh}. A similar procedure using $T_2(r)$ will yield $E^h(k)= <|h(k)|^2 /2>$, \begin{equation} E^{h}(k)=\delta_{k,0} \frac{<h^2>}{2} + B \frac{d^2}{d k^2} \delta_{k,0} + \frac{C}{(Lt)^2} k^{-4}, \end{equation} where $B$ and $C$ are constants \cite{Ligh}. Hence for $k>0$, we obtain $E^{h}(k) \propto k^{-4}$ for the noiseless KPZ equation. Most papers in the past implicitly assume that if $T_{2}(r) \propto r^{2 \chi}$, then $E^{h}(k)$ should be proportional to $k^{-2 \chi -1}$. Clearly this does not hold for the noiseless KPZ equation (check: $\chi=1$, but $E^{h}(k) \propto k^{-4}$). This apparent contradictions, which also occurs for $\sigma>1$, can be resolved using the following arguments. Suppose $E(k)=A k^{-2 \chi -1}$. The second-order structure function $S_{2}(r)$ is given by \begin{eqnarray} S_{2}(r) & = & \frac{1}{L} \int_{0}^{L} dx <|u(x+r)-u(x)|^2> \nonumber \\ & = & 16 \int_{0}^{\infty} dk E(k) \sin^2 (kr/2) \nonumber \\ & = & 16 \int_{0}^{\infty} dk A k^{-2 \chi -1} \sin^2 (kr/2). \end{eqnarray} We are interested in the leading order behaviour of $S_{2}(r)$ for small $r$. When $\chi < 1$, the above integral converges, and \begin{eqnarray} S_{2}(r) & \propto & 16 A r^{2 \chi} \int_{0}^{\infty} ds s^{-2 \chi -1} \sin^2 (s) \nonumber \\ & \approx & A C_1 r^{2 \chi}, \end{eqnarray} where $C_1$ is a dimensionless constant. When $\chi > 1$, the integral diverges from below (small $k$). However, we can cure this divergence by choosing the lower limit of the integral to be $2 \pi/L$, which yields \begin{eqnarray} S_{2}(r) & \propto & 16 A \int_{2 \pi /L}^{\infty} dk k^{-2 \chi -1} (kr/2)^2 \nonumber \\ & \approx & A C_2 (r/L)^2 L^{2 \chi}, \end{eqnarray} where $C_2$ is a constant. Hence, when $\chi <1$, $E(k) \propto k^{-2 \chi -1}$ and $S_{2}(r) \propto r^{2 \chi}$ as expected, however when $\chi >1$, $E(k) \propto k^{-2 \chi -1}$ and $S_{2}(r) \propto r^{2}$ . This analytical results are seen in our numerical simulations to be described below. After the discussion on the noiseless equation, now we turn to the noisy Burgers and KPZ equations. \section{Noisy Burgers equation: Various ranges of parameters $\sigma$} In this section we will attempt to divide the parameter range of $\sigma$ according to the properties of the solution. From Eq.~(\ref{Burgers}) the noise spectrum $|\zeta(k)|^2$ is proportional to $k^{-2 \sigma}$. Using this we can derive the noise-noise correlation as \begin{equation} \label{Burgernoisex} \left\langle \zeta (x,t)\zeta (x+r,t)\right\rangle \sim \left\{ \begin{array}{ll} B_0\delta(r) & \mbox{for $\sigma=0$} \\ B_1-C_1 r^{2 \sigma -1} &\mbox{for $0 \le \sigma \le 1/2$} \\ B_2-C_2\log(r) & \mbox{for $\sigma=1/2$} \\ B_3 - C_3 r^{2\sigma -1} & \mbox{for $1/2 < \sigma \le 3/2$} \\ B_4-C_4 r^2 L^{2\sigma -3} & \mbox{for $\sigma >3/2$ } \end{array} \right. \end{equation} where $L$ is the length of the system, and $B_i$ and $C_i$ are constants. For $-1 \le \sigma \le 0$, $|\zeta(k)|^2 \rightarrow \infty$ for large $k$, hence, the inverse Fourier transform which yields $<\zeta (x,t)\zeta (x^{\prime},t)>$ is not defined in this regime. However, Kardar et al. \cite{KPZ} and Medina et al. \cite{Medi} have solved the KPZ equation for this regime using $|f(k)|^2$ which is well defined for large $k$. We call the region $-1 \le \sigma \le 0$ as (a). It can be deduced from the above discussion that the noise-noise correlation increases with the increase of $\sigma$ till $\sigma=3/2$. Beyond $\sigma=3/2$, the correlation is proportional to $1-C r^2$ for all $\sigma$. Therefore, it is expected that the $h-h$ and $u-u$ correlation would increase with the increase of $\sigma$ till 3/2, beyond which the behaviour is expected to be somewhat similar. Keeping this in mind we have divided the $\sigma$ range beyond 0 in two regions: (b) $0 \le \sigma \le 3/2$, and (c) $\sigma > 3/2$. Well defined shocks develop in the parameter range (c) due to the large noise-noise correlation, and these shocks determine the exponents. In the intermediate parameter range (b), the exponents change slowly from RG dominated values to shock dominated values. The details are given in the next section. Polyakov \cite{Poly} has analytically solved Burgers equation with $\sigma=3/2$ using the methods of quantum field theory. Boldyrev \cite{Bold1,Bold2} extended Polyakov's method to the range $1/2 \le \sigma \le 3/2$ and derived the spectral exponents and the probability densities. Boldyrev derived that $\chi = 1+2 \sigma /3$ for $1/2 \le \sigma \le 3/4$, and $\chi=3/2$ for $\sigma > 3/4$. Boldyrev \cite{Bold1} suggests that the above formula may also be valid for $0 \le \sigma \le 1/2$. In the following section we will compare our numerical results with the theoretical predictions of Medina et al. \cite{Medi}, Polyakov \cite{Poly}, and Boldyrev \cite{Bold1,Bold2}. \section{Simulation method} Our calculations in this paper have been done using direct numerical simulations based on pseudo-spectral method. This method, commonly used in turbulence simulations, is expected to perform better than finite difference scheme because the derivatives can be calculated exactly in the spectral method \cite{Canu}. The finite difference scheme was adopted by Moser et al. \cite{Mose} and Peng et al. \cite {Peng} in their simulation of KPZ equation for $-1 \le \sigma \le 0$. We solve the KPZ equation in one dimension. The details of the simulation are as follows. A box of size $2\pi $ is discretized into $N=1024$ divisions. The KPZ equation is solved in Fourier space. However, to compute the nonlinear term, we go to real space, perform multiplication, then again come back to Fourier space. We time advance the Fourier components $h(k,t)$ using Adam Bashforth time marching procedure with flat surface as an initial condition. Two-third rule is used to remove aliasing \cite{Canu}. For details of the simulation refer to Canuto et al. \cite{Canu} and Verma et al. \cite{MKVmhdsim}. In our simulation we also add hyperviscosity term $(\kappa \nabla ^4h)$ to the KPZ equation to damp the large wavenumber modes strongly. The hyperviscosity term does not affect the intermediate scales which is of our interest \cite{MKVmhdsim}; this is because the $\nabla ^4h$ and higher order derivative terms are irrelevant in the renormalization group sense (cf. Barabasi and Stanley \cite{Bara} and references therein). In our simulation we take $\nu$ and $\kappa$ to be very small. Note that large $\nu$ correspond to Edward-Wilkinson (EW) equation. The dimensionless parameters used in our simulations are \begin{eqnarray} \label{para} \lambda & = & 1.0, \nonumber \\ \nu & = & 10^{-5}, \nonumber \\ \kappa & = & 10^{-6}, \nonumber \\ (2 \pi)^{2}D & = & 10^{-3}, \nonumber \\ dt & = & 1/2000 \end{eqnarray} with one exception. For $\sigma=-1$, we choose $\lambda=0.1$ for the stability of the code. In the Appendix A we have varied values of the parameters $\nu$ and $D$ and shown crossover from KPZ behaviour to EW equation. The values of $\sigma $ used in our simulation are -1, -0.85, -0.75, -0.60, -0.50, -0.25, 0, 1.25, 1.50, 1.75, 1, 2, and 6. We have time evolved the equation till 15 nondimensional time units. We find that the system reaches saturation in approximately 8 to 10 time units. For reference, $2 \pi$ time unit corresponds to one eddy turnover time in fluid turbulence. For ensemble averaging, we have performed averages over 100 samples which start with different random seeds for the noise. We have used $ran1$ of numerical recipes \cite{Recipe} as our random number generator. Each computer run for 15 time units and 100 samples takes approximately 7 hours on a Pentium machine (150 MHz). We have calculated $\chi $ and $\beta $ using the simulation data. The ensemble average $<.>$ have been obtained by taking averages over 100 runs. The width $W(L,t)=[<\sum_x h^2 (x)/N>]^{1/2}$ grows as a power law in time, i.e., $W(L.t) \propto t^{\beta_{KPZ}}$, in the early stages of growth. We obtain $\beta_{KPZ}$ by fitting a straight line in $log-log$ plot of $W(L,t)$ vs. $t$ over a range of $t=0.2:2.5$. The other exponent $\chi_{KPZ}$ is obtained from the asymptotic $<|h(k,t)|^2>$ averaged over $100$ runs. We find that powerlaw holds ($|h(k)|^2\propto k^{-2\chi_{KPZ} -1}$) for $k=10:8$; this range corresponds to the inertial range of turbulence. We perform the averaging at steady-state ($t=15$). In Fig.~\ref{ekfig} we plot $<|h(k)|^2>$ vs. $k$ for $\sigma =-1,0,1/2,$ and $2$. The lines of best fit are also shown in Figure. The computed values of the exponents are listed in Table 1. Our estimate of the error in the exponent is roughly 0.05. We have also calculated the structure functions $S_q(r)$ and $T_q(r)$ using the simulation data. The variable $r$ ranges from 1 to $N/2$. Near the boundaries we calculate $\Delta u$ or $\Delta h$ using wrap around scheme, i.e., $x+r$ is taken as $mod(x+r,N)$. we have reported the structure function exponents at the steady-state ($t=15$). The probability densities $P(h), P(u)$, and $P(u')$ have been calculated by averaging the histograms over 100 runs at every 0.2 time interval from initial time of 5 units to the final time of 15 units. Even though the solution has not reached the steady state at $t=5$, for more sampling we have taken the time interval from $t=5$ to $t=15$. In the following section we will describe the results of our simulation for various degrees of noise correlations, i.e., for different $\sigma$s. \section{Simulation results} As discussed in section IV, we divide the range of parameter $\sigma$ in three regions: (a) $-1 \le \sigma \le 0$, (b) $0 \le \sigma \le 3/2$, and (c) $\sigma \ge 3/2$. Fig.~\ref{uxpapfig} shows numerical $u(x,t)$ for various $\sigma$s. We find that noise dominates when $\sigma$ is negative. However, shock-like structures are clearly visible for $\sigma=2,6$ (compare with Fig.~\ref{profilefig}). In the the intermediate range of $\sigma$ (region (b)), both noise and shock structure coexist. We show below that the structures present in the profile contribute significantly to the determination of the spectral and intermittency exponents. Regarding energy spectrum, our numerical results are in good agreement with the RG predictions for $-1 \le \sigma \le 0$ or $0 \le \rho \le 1$. There is a gradual transition from $\chi_{KPZ}=1$ to $\chi_{KPZ}=3/2$ as $\sigma$ increases from 0 to 3/2. For $\sigma \ge 3/2$ we find that $\chi_{KPZ}=3/2$. Regarding probability density, we find that the probability densities $P(h)$ and $P(u)$ are gaussian for all $\sigma$'s. Fig.~\ref{Pufig} shows the plot of $P(u)$ vs. $u$ along with the best fit gaussian curves. Clearly $P(u)$ is gaussian for all the $\sigma$s. Therefore, the exponents of the structure functions $T_q(r)$ are expected to be proportional to $q$. Fig.~\ref{structfig} shows a typical log-log plots of $S_q(r)$ and $T_q(r)$ vs. $r$ (here shown for $\sigma=6$). The shock region is $r<10$ where both $S_q(r)$ and $T_q(r)$ are proportional to $r^q$ because $u(x)$ and $h(x)$ are continuous here. The region of our interest is $r=10:80$ where both $S_q(r)$ and $T_q(r)$ are power laws. The exponents for this range of $r$, $\zeta^{Burg}_q$ and $\zeta^{KPZ}_q$, are listed in Tables 1 and 2 respectively. Fig.~\ref{kpzfig} shows $\zeta^{KPZ}_q$ vs. $q$ plot, and Fig.~\ref{burgfig} shows $\zeta^{Burg}_q$ vs. $q$ plots. Clearly, $\zeta^{KPZ}_q$ is approximately proportional to $q$ as predicted in the previous paragraph. Hence, our results regarding $P(u)$ and $T_q(r)$ are consistent. We find, however, that $P(u')$ deviates significantly from the gaussian behaviour as we increase $\sigma$ from 0, thus signalling an intermittent behaviour for $u$. The details of our results for various ranges are given below. \subsection{$-1 \le \sigma \le 0$} In Table 1 we have listed our numerical values of $\chi_{KPZ}$ and $\beta_{KPZ}$. For comparison we also list the predicted values of Medina et al. \cite{Medi} below \begin{equation} \chi =\left\{ \begin{array}{ll} \frac 12 & \mbox{for $-1 \leq \sigma \leq -\frac{3}{4}$ } \\ 1+\frac 23 \sigma & \mbox{for $-\frac{3}{4} \leq \sigma \leq 0$ } \end{array} \right. \end{equation} and \begin{equation} \beta =\left\{ \begin{array}{ll} \frac 13 & \mbox{for $-1 \leq \sigma \leq -\frac{3}{4}$} \\ \frac{3+2\sigma }{3-2\sigma } & \mbox{for $-\frac{3}{4} \leq \sigma \leq 0$ } \end{array} \right. \end{equation} Medina et al.'s results are based on RG scheme that breaks down beyond $\sigma=0$. We compute $\chi_{KPZ}$ by a straight line fit to the $\log-\log$ plot of $<|h(k)|^2>$ over $k=10:80$ at $t=15$ (see Fig.~\ref{ekfig}). Note that $|h(k)|^2\propto k^{-2\chi-1}$. The $\chi_{KPZ}$s listed in Table 1 show that that our numerical $\chi_{KPZ}$s are in close agreement with Medina et al.'s \cite{Medi} and Chattopadhyay and Bhattacharjee's \cite{ChatJKB} theoretical predictions. Our findings are also consistent with earlier simulation results by Amar et al. \cite{Amar}, Peng et al. \cite{Peng}, and Meakin and Jullien \cite{Meak1,Meak2}. For $\sigma=-1/4$, Hayot and Jayaprakash \cite{Hayo} have reported a crossover in the wave number space---from RG dominated region for small $k$ to free field behaviour for large $k$. We do not find any such crossover in our simulation. Regarding the Burgers equations, $\chi_{Burg}=\chi_{KPZ}-1$. Regarding $\beta_{KPZ}$ calculations, our results are agreement with Medina et al.'s predictions for $\sigma=-1:-3/4$. However, our exponents differ significantly with Medina et al.'s exponents for higher $\sigma$s. For example, for $\sigma=0$, we obtain $\beta_{KPZ}=0.45$ contrary to the predicted $\beta_{KPZ}=1$. The reason for this discrepancy is not clear to us at this point. Regarding $\beta_{Burg}$, we find that it is approximately 0.5, consistent with Hayot and Jayaprakash's findings \cite{Hayo}. The probability density $P(u')$ vs. $u'$ for $\sigma=-1,-1/2,0$ is shown in Fig.~\ref{P0to1fig} on a semilog plot. The figure shows that $P(u')$ is gaussian for $\sigma=-1$, but it deviates from the gaussian behaviour with the increase of $\sigma$; the deviations becomes more and more prominent for higher $\sigma$. At $\sigma=0$, $P(u') \approx 0.01*\exp(-0.14 |u'|)$ for negative $u'$ (the curve of the best fit shown in the Fig.~\ref{P0to1fig}). Beyond $\sigma=0$, a powerlaw tails start appearing for negative $u'$. Table 2 and Fig.~\ref{burgfig} show that $\zeta^{Burg}_2$ is close to zero for $\sigma < 0$. The energy spectrum $ <|u(k)|^2/2> \propto k^{- 2 \chi_{KPZ}+1}$. Therefore, using the arguments of section 3 we can easily show that $<| u(x+r)-u(x)| ^2> \propto r^{2(\chi_{KPZ}-1)}$ for $\sigma < 0$. For $\sigma=0$, it can be easily shown that $<| u(x+r)-u(x)| ^2> \propto \log (r)$. We find in our simulation that $\zeta_2^{KPZ}$ is positive contrary to the above prediction. This deviation may be because of {\em intermittency} (due to the presence of small shocks). It is also possible that at large $r$ ($r \approx L$), the other sub-critical terms may become comparable to the leading order term and may change the exponent. Quantitative calculation of the exponents for this range of $\sigma$ is beyond the scope of this paper. \subsection{$\sigma \ge 3/2$} One of the important aspect of this paper is the discussion of the energy spectrum, probability density and structure functions of the Burgers equation with large $\sigma$. As shown in the Fig.~\ref{uxpapfig}, shocks are prominent for this range of $\sigma$. Due to these shocks, the exponent $\chi_{KPZ} \approx 3/2$ for $\sigma > 3/2$ (see Table 1 and Fig.~\ref{ekfig}). Note that $\chi_{KPZ}=3/2$ and $\chi_{Burg}=1/2$ for the noiseless Burgers equation due to the presence of shocks. The existence of shocks for large $\sigma$ can be argued from the noise-noise correlation discussed in section IV. For large $\sigma$, there is long-ranged noise-noise correlation. Physically, large fluid parcels are moved around by this noise. As a consequence, there will be regions where the parcels forced by appositely directed noise collide with each other and create ``strong'' shocks. Hence, it is not surprising that strong shocks will be generated for large $\sigma$. These shocks determine the dynamics and the spectral indices of Burgers and KPZ equations. Therefore, the noisy KPZ equation with large $\sigma$ yields the same energy spectrum and multiscaling exponents as the noiseless KPZ equation. The noise-noise correlation is proportional to $1-C r^2$ for all $\sigma \ge 3/2$, hence the spectral and multiscaling exponents are expected to be somewhat similar for all the $\sigma$ beyond 3/2. This is borne out by our numerical simulation. This is consistent with the trivial observation that for large $\sigma$ only $k=1$ mode is effective. It is interesting to note that $k=1$ mode (i.e., $\zeta(x)=sin(x)$) also yields noise-noise correlation proportional to $1-C r^2$, and approximately the same exponents as those with $\sigma \ge 3/2$. We have also analyzed the stability of the shocks at a preliminary level. For large $\sigma$, the significant contribution to the dynamics comes only from the $k=1$ mode of the noise. Since the first mode ($k=1$) is noisy, the the formation of a single shock delayed for some time because of the movement of the zero of the noise signal. However, after the shock is formed, it shifts around by only a small amount because the impulse due to the random noise is not strong enough to move the shock by a large distance. We have also calculated $\beta$ for both Burgers and KPZ equations when $\sigma>3/2$ and found them to be approximately 1/2 . This result is consistent with findings of Hayot and Jayaprakash \cite{Hayo}. Regarding the intermittency exponents, as shown in Fig.~\ref{burgfig}, $\zeta_q^{Burg} \approx 1$ for all $q$ for $\sigma \ge 3/2$. This result is consistent with existence of shocks (see Fig.~\ref{uxpapfig} and section II). For large $r$, $S_q(r)$ is not proportional to $r$. This is because of the term similar to $(\mu_{i}-r/t)^{q}$ of Eq.~(\ref{struct_Burg}), which becomes important at large $r$. The effects of this subcritical term is evident in the intermittency exponents listed in Table 2. We had anticipated $\zeta_q^{Burg}$ to be 1, but we consistently find them to be less that 1. The same trends were observed for the noiseless case. Fig.~\ref{P1gtfig} shows the probability density $P(u')$ for various $\sigma$s. The curves of best fit are also shown in the Figure. The probability density is clearly nongaussian for all $\sigma$s shown in the Figure. In our simulations we find that for all $\sigma \ge 3/2$, $P(u') \propto u'^{-\alpha} (\alpha=2.2-2.5)$ fits quite well with the numerical $P(u')$ for the intermediate range of the negative $u'$. This result is in agreement with the Polyakov's theoretical prediction for $\sigma \ge 3/2$ \cite{Poly}. Curiously, $P(u')$ for $u'>0$ is also nongaussian, but it does not match with the Polyakov's predictions for positive $u'$. We have shown in this section that both noiseless Burgers equation and the Burgers equation with strongly correlated noise have $S_q(r) \propto r$. However, the probability densities $P(u)$ and $P(u')$ are very different for these two cases (see Figs.~\ref{Pnlessfig} and \ref{P1gtfig}). This difference is because of the different nature of the noise. The entries in Table 2 show that $\zeta^{KPZ}_q=q$. This can be argued as follows. Since $\chi_{KPZ} \approx 3/2$, the second order structure function $T_2(r)$ will be proportional to $r^2$, or $\zeta^{KPZ}_2=2$ (see Section III). This is seen in our simulation for $\sigma \ge 3/2$. The Fig.~\ref{Pufig} also shows that $P(u)$ is gaussian. Therefore, $T_q(r) \propto [T_2(r)]^{q/2} \propto r^q$, hence $\zeta^{KPZ}_q=q$, a result seen in our numerics. Thus, structure function and energy spectrum calculations are consistent. After discussing the shock dominated region of $\sigma \ge 3/2$, we turn to the region $0 \le \sigma \le 3/2$. \subsection{$0 < \sigma \le 3/2$} The $\chi_{KPZ}$ calculated by our numerical simulation for the range $0 \le \sigma \le 3/2$ is listed in Table 1. The exponent increases from 1 and saturate at 3/2. That is, there is a gradual shift from RG dominated exponents to shock dominated exponents as we vary $\sigma$. The profile $u(x,t)$ of Fig. 4 shows that both fluctuations and shocks coexist in this range, with shocks becoming more and more important as $\sigma$ increases. Boldyrev \cite{Bold1,Bold2} has calculated $\chi_{KPZ}$ using methods of quantum field theory and derived that $\chi_{KPZ} = 1+2 \sigma /3$ for $1/2 \le \sigma \le 3/4$, and $\chi_{KPZ}=3/2$ for $\sigma > 3/4$. Boldyrev \cite{Bold1} suggests that the above formula may be valid for $0 \le \sigma \le 3/2$. Comparison of Boldyrev's predictions with the entries of Table 1 shows that the predictions work quite well till $0 \le \sigma \le 1/2$. After that there is a significant deviation, and the exponents appear to be outside numerical error bars. We need to probe the region $1/2 \le \sigma \le 3/4$ carefully to reach a definite conclusion. We presume that both fluctuations and the embedded structures are important in the determination of the spectral index. One would need to combine contributions from the structures and fluctuations to obtain the energy spectrum \cite{Supr1}. In Appendix B we sketch an elementary framework when both structure and fluctuations are present in the system. The Burgers equation for $\sigma=1/2$ has been studied extensively by Chekhlov and Yakhot \cite{Chek1}. They find in their high resolution simulation that $|u(k)|^{2} \propto k^{-5/3}$. The spectrum in our low resolution simulation is $|u(k)|^{2} \propto k^{-1.54}$ ($\chi_{KPZ}=1.27$), which is close to Chekhlov and Yakhot's \cite{Chek1} result. Chekhlov and Yakhot \cite{Chek1} have argued for a constant cascade of energy in the wavenumber space in this case and claimed that the Kolmogorov-like energy spectra is due to the constancy of the energy flux. However, we find the cascade rate to be constant for all $\sigma> 1/2$, but the spectral index of the energy spectrum is in the range of 5/3 to 2 (to be discussed in Section 7). Hence, the argument that the constant flux yields Kolmogorov's spectrum is incorrect. Regarding the probability distribution, in Fig.~\ref{P1gtfig} we plot $P(u')$ vs. $u'$ along with the curves of best fit. It is found that $P(u')$ is a powerlaw for $u'<0$. For $\sigma=1.25, P(u') \propto u'^{-2.86}$, but for $\sigma = 1/2$, $P(u') \approx u'^{-2.2}$, with an error of approximately $10\%$ for both the cases. Curiously, Polyakov's predictions ($\sigma=3/2$) for positive $u'$ appears to be applicable only for $\sigma=1/2$; here $P(u') \propto \exp(-C x^3)$. The roughening exponents $\chi_{KPZ}$ lies between 1 and 3/2. Therefore, $\zeta^{KPZ}_q$ should be proportional to $q$ as argued in the previous subsection. Our numerical simulations yield approximately the same exponents (see Table 3). Hence our energy spectrum calculations and the structure function calculations are consistent. \subsection{Summary of simulation results} From the above discussion we see that Burgers and KPZ equations are well understood for $-1 \le \sigma \le 0$ and $\sigma > 3/2$. In the intermediate range $0 < \sigma < 3/2$, Boldyrev's predictions appear to explain the numerical data in part of the regime; further analytic and numerical work is required in this regime. For all $\sigma > -1/2$, $\chi_{KPZ}+z$ $(z=\chi_{KPZ}/\beta_{KPZ})$ deviates significantly from 2. In fact, for larger $\sigma$, $\chi_{KPZ}+z=3/2+(3/2)/(1/2)=9/2$, quite far from 2. Note that Medina et al. \cite{Medi} argue that the identity $\chi_{KPZ}+z=2$ is a consequence of Galilean invariance. However, Hayot and Jayprakash \cite{Hayo}, Polyakov \cite{Poly} and others have speculated violation of this identity due to the presence and motion of the shocks. Meakin and Jullien's \cite{Meak1,Meak2} results also violate $\chi_{KPZ} +z=2$ condition for a range of $\sigma $. They find that for $\sigma =-1/4,$ $\chi_{KPZ}+z=2.37$, quite different than 2. The violation of Galilean invariance even in the region $-1 \le \sigma \le 0$ signals importance of structures \cite{Poly}. We also show that $P(u)$ is gaussian for all $\sigma$. However, $P(u')$ is gaussian at $\sigma=-1$, but continuously changes to $\exp(-x)$ then to a powerlaw behaviour. Therefore, $u$ field of Burgers equation exhibits intermittency. The degree of intermittency depends on $\sigma$. The energy flux play an in important role in turbulence analysis. In the following section we briefly report energy flux studies for the Burgers equation. \section{Energy flux of noisy Burgers equation} We derive an equation which gives us the energy transfer from the region $|k| \le K$. The derivation of the energy equation from Eq. (\ref{Burgers}) and averaging yields \cite{McCobook} \begin{equation} \label{flux_eq} \frac{\partial}{\partial t} \int_{0}^{K} E(k)dk = - \int_{0}^{K} 2 \nu k^2 E(k) dk \\ - \int_{0}^{K} \Re <u^*(k) \left[ FT \left(\frac{\partial}{\partial x} u^2 /2 \right) \right]_k> + \int_0^K \Re <u^*(k) f(k)>. \end{equation} From this equation, clearly the energy dissipation in the wavenumber sphere of radius $K$ is \begin{equation} D_{K}= \int_{0}^{K} 2 \nu k^2 E(k) dk, \end{equation} and the energy flux coming out the wavenumber sphere of radius $K$ is \begin{equation} \Pi_{k}=\int_{0}^{K} \Re <u^*(k) \left[ FT \left(\frac{\partial}{\partial x} u^2 /2 \right) \right]_k >. \end{equation} It can easily seen that for the random noise, the energy supplied by the forcing to the wavenumber sphere of radius $K$ is \cite{McCobook} \begin{eqnarray} \label{FK} F_{K} & = & \int_0^K \Re <u^*(k) f(k)> \nonumber \\ & = & \int_0^K <|f(k)|^2>. \end{eqnarray} From the Eq. (\ref{flux_eq}) it is clear that at the steady state \begin{equation} F_{K}=\Pi_{K}+D_{K}. \end{equation} We have plotted $D_{K}, \Pi_{K}$, and $F_{K}$ for various $\sigma$. Fig.~\ref{flux0fig} shows the plots for $\sigma = 0$. Here, we find that $D_{K} \approx F_{K}$ till $k \approx 250$, but $\Pi_{K} \ll F_{K}$ except for small $k$. Similar results are obtained for all $-1 \le \sigma \le 0$. Fig.~\ref{flux0.5fig} shows the plots of $D_{K}, \Pi_{K}$, and $F_{K}$ for $\sigma=1/2$. The forcing rate $F_{K}$ is proportional to $\log(K)$. We also find a range of $K$ for which $\Pi_{K}$ is constant. These results are consistent with the findings of Chekhlov and Yakhot \cite{Chek1}. Fig.~\ref{flux6fig} shows the plot of the above quantities for $\sigma =6$. Here again $\Pi_{K}$ is constant for a range of $K$. An interesting point to note is that for $\sigma=6$, $F_{K}$ is constant beyond $K=10$ or so; this is because effectively only first few modes are forced when $\sigma$ is large. Similar results are obtained for all $\sigma \ge 2$. We find that the energy flux is constant for the noisy Burgers equation for all $\sigma \ge 1/2$. However, the energy spectrum varies from $k^{-5/3}$ to $k^{-2}$ as we increase $\sigma$ from 1/2 to 3/2 and beyond. Hence our results show that constancy of energy flux is not a sufficient condition for the Kolmogorov's energy spectrum in noisy Burger equation. This indicates that Chekhlov and Yakhot's \cite{Chek1} argument that the constant energy cascade rate for $\sigma =1/2$ implies Kolmogorov's energy spectrum is incomplete. We find that for $\sigma \ge 2$, the flux rate $\Pi_{K} \approx 2 \times 10^{-3}$. This numerical value is consistent with value obtained using the formula derived by Saffman \cite{Saff} \begin{equation} \Pi_{K} \approx \frac{\mu^3}{24 L}, \end{equation} where $\mu$ is the velocity jump across the shock. In our simulations, $\mu \approx 0.3$ for large $\sigma$. Also, the Eq. (\ref{FK}) yields the same numerical value ($2 D=2 \times 10^{-3}$). \section{Discussion and Conclusions} In this paper we have numerically calculated the energy spectrum, structure function, and probability densities ($P(h), P(u), P(u')$) for the KPZ and Burgers equations in the presence of correlated noise. The Burgers equation with a strong correlated noise $(\sigma \ge 3/2)$ has distinct shock structures. Given this we have trivially worked out the roughening exponents for this range: $\chi_{Burg}=1/2, \chi_{KPZ}=3/2$. The $\chi$ calculated from the numerical simulation matches quite well with this result. It has been theoretically shown by many researchers that $\zeta^{Burg}_q \approx 1$ when shocks are present. Numerically we find the same $\zeta^{Burg}_q$ for $\sigma \ge 3/2$. The probability density $P(u') \propto u'^{-\alpha}$ with $\alpha=2.5 \pm 0.3$ for negative $u'$; this result is in agreement with the Polyakov's theoretical predictions. However, for positive $u'$, our results do not match with Polyakov's predictions for this range of $\sigma$. Note that our results differ from those of E and Eijnden \cite{E} and Kraichnan \cite{KraiBurg} who obtained $\alpha=7/2$, and Gotoh and Kraichnan \cite{GotoKrai} who argue for $\alpha =3$ In the $\sigma$ regime $-1 \le \sigma \le 0$, our numerical $\chi_{KPZ}$ matches with the RG predictions. The exponent $\beta_{KPZ}$, however, saturates at 0.5 contrary to the RG predictions. For $0 \le \sigma \le 3/2$, the exponent $\chi_{KPZ}$ varies smoothly from 1 to 3/2. Boldyrev's predictions appear to match with our numerical results for $0 \le \sigma \le 1/2$. After that there is a significant deviation. Our claims, however, are not on a very strong ground because of large uncertainties, and hence, more exhaustive simulations are needed to reach definite conclusions. We find in our simulation that $P(u)$ is gaussian for all $\sigma$s and also $\zeta^{KPZ}_q \propto q$. Therefore, according to the convention discussed in the introduction, the signal $h$ will be termed as non-intermittent for all $\sigma$s. Note, however, that according to the same convention, $u$ is as intermittent for $\sigma \ge 1/2$. These two statements appear to be contradictory because $u$ and $h$ are related by $u=-\nabla h$. The apparent contradiction is quite simple to resolve. The velocity signal $u$, being the derivative of $h$, is much more singular than $h$ because of the presence of the shocks. That is why probability density $P(u')$ is nongaussian (intermittent $u$), even though $P(u)$ is nongaussian (non-intermittent $h$). This observation cautions us to choose a right variable while investigating the system for intermittency. Note that for fluid turbulence also, $P(u)$ is gaussian, but $P(u')$ is not, and the variables used for structure function is $u$ (there is no corresponding $h$ anyway). The nonexistence of intermittency in KPZ equation will have relevance to other models of surface growth. Sneppen and Jensen \cite{SnepJens} and Tang and Leschhorn \cite{Tang} calculated the structure function in surface growth equation in presence of quenched disorder. They find no spatial multiscaling in their system; this result could be related to the conclusions described in our paper. Sneppen and Jensen \cite{SnepJens}, and Tang and Leschhorn \cite{Tang} find temporal intermittency in their model. We believe temporal structure function for the noisy KPZ and Burgers equation are important and will shed further insights into the dynamics of these systems. Krug \cite{Krug} and Kundagrami et al.~\cite{Chan} have investigated the existence of intermittency in the surface growth model of Das Sarma and Tamborenea (DT) \cite{DT} model and found multiscaling in it. Krug, however, finds absence of multiscaling in a variation of the DT model that was tilt independent. The DT model is related to linear Lengevin equation which is not expected to show intermittency, but Krug has attempted to relate intermittency in the DT model with the existence of relevant variables of renormalization groups. The connection of Krug's result to KPZ equation is not clear to us at this stage. Also, it is not clear whether the multiscaling in the DT model is an artifact of lattice effects, or it is due to appearance of relevant of nonlinearities in the continuum equation \cite{Krug}. Another important point to note is that $u$ of fluid turbulence should probably be mapped to $h'$, not $h$ as done in Krug's \cite{Krug} and Kundagrami et al.'s~\cite{Chan} papers. In the regime $0 \le \sigma \le 3/2$, both structures and noise coexist. Recently, Polyakov \cite{Poly}, Boldyrev \cite{Bold1,Bold2}, Gurarie and Migdal \cite{Gura}, E and Eijnden \cite{E}, and Kraichnan \cite{KraiBurg} have attempted to analytically solve for the spectral exponents and the probability density $P(u')$ for this regime. Many of the recent attempts are based on methods of quantum field theory. Considering the important role played by the shock structures in determination of the roughening exponents, we believe that a calculation which incorporates both structures and noise will be very useful for such problems. The roughening exponent will get contributions both from fluctuations and embedded structures (see Appendix B). However, a careful analysis is required to isolate the individual contributions. Krishnamurthy and Barma \cite{Supr1} have isolated a moving pattern in the surface growth phenomena in the presence of quenched disorder. Their method may be applied here to separate the fluctuations from the structures. Bouchaud et al.~\cite{Bouc} have calculated the structure function of the Burgers equation in higher dimensions using the connection of KPZ equation to directed polymers. They showed that $S_q(r) \propto r$. In Appendix C we argue that using the shock structures of the higher dimensional Burgers equation, one can obtain $S_q(r) \propto r$. In this paper we have demonstrated the usefulness of structures in calculating the dynamical exponents of the system. The role of structures in dynamics is being studied in fluid turbulence, intermittency, self organized criticality etc. For example, in fluid turbulence Hatakeyema and Kambe \cite{Hata} have used the vortex structures to calculate the scaling exponents. Therefore, discovery of the connections between the structures, fluctuations, and dynamics will yield interesting insights in the nonequilibrium phenomena around us. \acknowledgments The author thanks Mustansir Barma, Supriya Krishnamurthy, and Deepak Dhar for discussions, references, and their kind hospitality during his stay at TIFR, where part of this work was done. V. Subrahmanyam's ideas and criticisms are gratefully acknowledged. The author also thanks J. K. Bhattacharjee, Agha Afsar Ali, Prabal Maiti, S. A. Boldyrev, and S. D. Joglekar for discussions at various stages, and R. K. Ghosh for providing computer time on DEC workstation.
\subsubsection*{#1}}{\quad$\diamondsuit$\par\bigskip} \newcommand{\Span}[1]{\left< #1 \right>} \newcommand{\textstyle\frac12}{\textstyle\frac12} \newcommand{\mathrm{d}}{\mathrm{d}} \newcommand{\operatorname{id}}{\operatorname{id}} \newcommand{\operatorname{pr}}{\operatorname{pr}} \newcommand{\operatorname{ch}}{\operatorname{ch}} \newcommand{\operatorname{td}}{\operatorname{td}} \newcommand{\operatorname{CD}}{\operatorname{CD}} \newcommand{\operatorname{CS}}{\operatorname{CS}} \newcommand{\mathrm{BS}}{\mathrm{BS}} \newcommand{k\mathrm{\text{-}BS}}{k\mathrm{\text{-}BS}} \newcommand{\mathrm{HP}}{\mathrm{HP}} \newcommand{\mathrm{KH}}{\mathrm{KH}} \newcommand{\mathrm{LC}}{\mathrm{LC}} \newcommand{\mathrm{US}}{\mathrm{US}} \newcommand{\mathrm{SDAG}}{\mathrm{SDAG}} \newcommand{\mathrm{II}}{\mathrm{II}} \newcommand{\mathrm{gen}}{\mathrm{gen}} \newcommand{^{\mathrm{ss}}}{^{\mathrm{ss}}} \newcommand{^{\mathrm{irr}}}{^{\mathrm{irr}}} \newcommand{^{\mathrm{red}}}{^{\mathrm{red}}} \newcommand{\rest}[1]{{}_{{\textstyle{|}}#1}} \newcommand{\overline{\alpha}}{\overline{\alpha}} \newcommand{\overline{\partial}}{\overline{\partial}} \newcommand{\overline{D}}{\overline{D}} \newcommand{\overline{g}}{\overline{g}} \newcommand{\overline{u}}{\overline{u}} \newcommand{\overline{I}}{\overline{I}} \newcommand{\partial}{\partial} \newcommand{\widetilde}{\widetilde} \newcommand{\otimes}{\otimes} \newcommand{\overline}{\overline} \newcommand{\sA}{\mathcal A} \newcommand{\mathcal E}{\mathcal E} \newcommand{\sG}{\mathcal G} \newcommand{\mathcal F}{\mathcal F} \newcommand{\mathcal H}{\mathcal H} \newcommand{\mathcal L}{\mathcal L} \newcommand{\mathcal M}{\mathcal M} \newcommand{\mathcal O}{\mathcal O} \newcommand{\mathcal S}{\mathcal S} \newcommand{\mathcal W}{\mathcal W} \newcommand{\alpha}{\alpha} \newcommand{\beta}{\beta} \newcommand{\delta}{\delta} \newcommand{\varepsilon}{\varepsilon} \newcommand{\varphi}{\varphi} \newcommand{\gamma}{\gamma} \newcommand{\kappa}{\kappa} \newcommand{\omega}{\omega} \newcommand{\sigma}{\sigma} \newcommand{\Delta}{\Delta} \newcommand{\Gamma}{\Gamma} \newcommand{\Lambda}{\Lambda} \newcommand{\Omega}{\Omega} \newcommand{\lambda}{\lambda} \newcommand{\Sigma}{\Sigma} \newcommand{\nabla}{\nabla} \newcommand{\mathbb P}{\mathbb P} \newcommand{\mathbb C}{\mathbb C} \newcommand{\mathbb H}{\mathbb H} \newcommand{\mathbb Q}{\mathbb Q} \newcommand{\mathbb R}{\mathbb R} \newcommand{\mathbb Z}{\mathbb Z} \newcommand{\ad}{\operatorname{ad}} \newcommand{\operatorname{diag}}{\operatorname{diag}} \newcommand{\operatorname{rank}}{\operatorname{rank}} \newcommand{\operatorname{hcf}}{\operatorname{hcf}} \newcommand{\operatorname{mir}}{\operatorname{mir}} \newcommand{\codim}{\operatorname{codim}} \newcommand{\operatorname{top}}{\operatorname{top}} \newcommand{\operatorname{v.dim}}{\operatorname{v.dim}} \newcommand{\operatorname{Diff}}{\operatorname{Diff}} \newcommand{\operatorname{tr}}{\operatorname{tr}} \newcommand{\operatorname{End}}{\operatorname{End}} \newcommand{\operatorname{GFT}}{\operatorname{GFT}} \newcommand{\operatorname{sGFT}}{\operatorname{sGFT}} \newcommand{\operatorname{\La_{\uparrow}\!}}{\operatorname{\Lambda_{\uparrow}\!}} \newcommand{\operatorname{Hol}}{\operatorname{Hol}} \newcommand{\operatorname{Hom}}{\operatorname{Hom}} \newcommand{\operatorname{Lie}}{\operatorname{Lie}} \newcommand{\Pic}{\operatorname{Pic}} \renewcommand{\Im}{\operatorname{Im}} \renewcommand{\Re}{\operatorname{Re}} \newcommand{\operatorname{Sing}}{\operatorname{Sing}} \newcommand{\operatorname{Vol}}{\operatorname{Vol}} \newcommand{\operatorname{SO}}{\operatorname{SO}} \newcommand{\U}{\operatorname U} \newcommand{\operatorname{PU}}{\operatorname{PU}} \newcommand{\operatorname{SU}}{\operatorname{SU}} \newcommand{\operatorname{\mathfrak{su}}}{\operatorname{\mathfrak{su}}} \newcommand{\operatorname{\mathfrak{su}}}{\operatorname{\mathfrak{su}}} \newcommand{\mathrm{pt}}{\mathrm{pt}} \renewcommand{\labelenumi}{(\arabic{enumi})} \begin{document} \title{Quantization and ``theta functions''} \markright{\hfill Quantization and ``theta functions'' \quad} \author{Andrei Tyurin} \date{Apr 1999} \maketitle \begin{abstract} Geometric Quantization links holomorphic geometry with real geo\-metry, a relation that is a prototype for the modern development of mirror symmetry. We show how this treatment can be used to construct a special basis in every space of conformal blocks. This is a direct generalization of the basis of theta functions with characteristics in every complete linear system on an Abelian variety (see \cite{Mum}). The same construction generalizes the classical theory of theta functions to vector bundles of higher rank on Abelian varieties and K3 surfaces. We also discuss the geometry behind these constructions. \end{abstract} \section{Introduction} \markright{\hfill Quantization and ``theta functions'' \quad} It is a fruitful question to ask for some special basis of the complete linear systems $\mathbb P H^0(X, L^k)$, where $X$ is a smooth complete algebraic variety and $L$ a polarization. After this, following Mumford, we can ask for special equations defining $X$ under its embedding in $\mathbb P H^0(X,L^k)^*$. Of course, this is a priori impossible (for example, for $\mathbb P H^0(\mathbb P^n,\mathcal O_{\mathbb P^n}(k))$), but it can be done after ``rigidification'' -- that is, fixing some discrete structure on $X$. This is the subject of Invariant Theory in its pre-Hilbert form; however, any proposed ``geometric'' rigidification depends on the level $k$, and there is no universal way of doing it. The amazing fact is we can do it in many cases using the ``classical'' {\em Geometric Quantization Procedure} (GQP); but for this, we must leave algebraic geometry and go over to symplectic geometry instead. I would like to call this method the {\em general theory of theta functions}. The starting point is that, together with a complex structure $I$ on $X$, a polarization $L$ gives us a quadruple $(X,\omega,L,a_L)$, where $\omega$ is the K\"ahler form and $a_L$ a Hermitian connection on $L$ with curvature form $F_a=2\pi i\cdot\omega$ of Hodge type $(1,1)$, giving the holomorphic structure on $L$. The pair $(X,\omega)$ is a symplectic manifold; we can thus view it as the {\em phase space} of some classical mechanical system, and the pair $(L,a_L)$ as {\em prequantization data} of this system. We should start by recalling the construction of spaces of wave functions for a pair $(S,\omega)$, where $S$ is a smooth symplectic manifold of dimension $2n$ with a given symplectic form $\omega$. To switch on any quantization procedure, we suppose that the cohomology class $[\omega]$ of the symplectic form is integral, that is, there exists a complex line bundle $L$ with $c_1(L)=[\omega]$. Moreover, suppose that $L$ has a Hermitian connection $a$ with curvature form $F_a=2\pi i\cdot\omega$. Any quadruple of this type \begin{equation} (S,\omega,L,a) \label{eq1.1} \end{equation} is called a {\em prequantization} of the classical mechanical system with phase space $(S,\omega)$. There are two approaches to the geometric quantization of $(S,\omega,L,a)$ (\ref{eq1.1}) (see \cite{A}, \cite{S1} or \cite{W}). We discuss here the simplest version of these constructions, avoiding questions such as the choice of metaplectic structures, densities and half densities specifying geometric conditions on the manifold $S$. (Roughly speaking, $S$ should be a {\em Calabi--Yau manifold}\/). The usual slogan is that we have to choose ``one half'' of the set of all functions on $S$ using some ``polarization'' conditions. The first approach is as follows: \subsubsection*{Complex polarizations} To define a complex polarization, we give $S$ a complex structure $I$ such that $S_I=X$ is a K\"ahler manifold with K\"ahler form $\omega$. Then the curvature form of the Hermitian connection $a$ is of type $(1,1)$, hence for any {\em level} $k\in\mathbb Z^+$, the line bundle $L^k$ is a holomorphic line bundle on $S_I$. Complex quantization provides the space of {\em wave functions of level $k$}: \begin{equation} \mathcal H_{L^k}=H^0(S_I,L^k), \label{eq1.2} \end{equation} -- that is, the space of {\em holomorphic} sections of $L^k$. Thus a {\em complex polarization} of $(S,\omega,L,a)$ (\ref{eq1.1}) returns to the algebraic geometry $S=X$ we started from. In particular, the spaces of wave functions (\ref{eq1.2}) obtained in this way is the collection of {\em complete linear systems} in the usual sense. We will suppose $L$ to be an {\em ample} holomorphic line bundle, and in particular, \[ H^i(S_I,L)=0 \quad\text{for all $i>0$.} \] The second approach is the choice of a real polarization: \subsubsection*{Real polarizations} A real polarization of $(S,\omega,L,a)$ is a Lagrangian fibration \begin{equation} \pi\colon S\to B, \label{eq1.3} \end{equation} such that \[ \omega\rest{\pi^{-1}(b)}=0 \quad\text{for every point $b\in B$,} \] and the fibre $\pi^{-1}(b)$ is a smooth Lagrangian submanifold for generic $b$. Thus a mechanical system admits a real polarization if and only if it is {\em complete integrable}. \begin{rmk} Actually, for the ordinary technical tricks of the theory of geometric quantization to work, we should require that the fibration has regular geometric behavior (see, for example, \cite{S2}). But beginning with Guillemin and Sternberg's paper \cite{GS2}, it is reasonable to consider more general fibrations, namely, {\em real polarizations with singularities}. \end{rmk} Then restricting $L$ to a Lagrangian fibre gives a flat connection $a\rest{\text{fibre}}$ or equivalently, a character of the fundamental group \[ \chi\colon\pi_1(\text{fibre})\to\U(1). \] Let $\mathcal L_{\pi}$ be the sheaf of sections of $L$ that are covariant constant along fibres. Then we get the space \[ \mathcal H_{\pi}=\bigoplus_{i} H^i(S,\mathcal L_{\pi}). \] In the regular case, \'Sniatycki proved that \[ H^i(S,\mathcal L_{\pi})=0 \quad\text{for $i \ne n$.} \] \begin{dfn} \begin{enumerate} \item A fibre of $\pi$ is a {\em Bohr--Sommerfeld} cycle of $(S,\omega,L,a)$ if $\chi=1$. \item $\mathrm{BS}\subset B$ is the subset of Bohr--Sommerfeld fibres. \item $k$-BS${}\subset B$ is the subset of Bohr--Sommerfeld fibres for $(S,\omega,L^k, ka)$. \end{enumerate} \end{dfn} According to the general theory of real quantizations, we expect to get a finite number of Bohr--Sommerfeld fibres, and in the regular case, \[ H^n(S,\mathcal L_{\pi})=\bigoplus_{\mathrm{BS}}\mathbb C\cdot s_{i}, \] where $s_{i}$ is a nonzero covariant constant section of the restriction of $(L,a)$ to a Bohr--Sommerfeld fibre of the real polarization $\pi$. In the general case, we can use this to {\em define} the new collection of spaces of wave functions (of level $k$): \begin{equation} \mathcal H_{\pi}^k=\bigoplus_{\text{$k$-BS}}\mathbb C\cdot s_{i}, \label{eq1.4} \end{equation} and use special tricks to compare (1.4) with (\ref{eq1.2}). There is a canonical way of describing the Bohr--Sommerfeld subset. For this, we must choose special coordinates on $B$, the so-called {\em action coordinates}, which are part of the {\em action angle} coordinates (see \cite{A}, \cite{GS1}, \cite{GS2}). Locally around a point $b\in B$, the action coordinates $c_i$ are given as periods along 1-cycles of the fibre $\pi^{-1}(b)$ of a 1-form $\alpha$ such that \begin{equation} \mathrm{d} \alpha=\omega. \label{eq1.5} \end{equation} This system of coordinates $\{c_i\}$ is defined up to additive constants and an {\em integral} linear transformations. Thus, if $B$ is simply connected, the action coordinates map $B$ locally diffeomorphically to some open subset \begin{equation} B_c\subset \mathbb R^n_{(c_1, \dots, c_n)} \label{eq1.6} \end{equation} with coordinates $\{c_i\}$. If $(0,\dots, 0)$ is a Bohr--Sommerfeld point, then \begin{equation} \mathrm{BS}=B_c \cap \mathbb Z^n \label{eq1.7} \end{equation} is the {\em set of integral points} in $B_c$. Let us return to our collections of spaces of wave functions. \begin{rmk} An important observation, proved mathematically in a number of cases, is that the projectivization of the spaces (\ref{eq1.2}) are given purely by the symplectic prequantization data and do not depend on the choice of complex structure on $S$. The same is true for the projectivization of the spaces (\ref{eq1.4}). Moreover, these spaces do not depend on the real polarization $\pi$ (\ref{eq1.3}), provided that we extend our prequantization data $(S,\omega,L,a,\mathcal F)$ by adding some ``half density'' $\mathcal F$ (see \cite{GS1}). \end{rmk} Our {\em main problem} is to compare the spaces \[ \mathcal H_{L^k} \quad \text{and} \quad \mathcal H_{\pi}^k. \] If we are lucky enough to be able to construct a canonical isomorphism between these spaces, we get a special basis in the space of wave functions of a complex polarization, and in particular in any ample complete linear system. To distinguish this basis from others, we call it the {\em system of theta functions} of level $k$, with ``characteristics'' which are Bohr--Sommerfeld fibres. Actually, this generalization of the theory of theta functions requires the final ingredient of the quantization procedure -- the algebra of {\em observables} represented as an algebra of operators on spaces of wave functions (like the Heisenberg algebra on spaces of classical theta functions). We avoid using such algebras in this article, but they underlie our constructions, so it is reasonable to recall briefly the general shape of this ingredient. \subsubsection*{Algebra of observables and its space of states} As a result of any quantization procedure, we get a $\mathbb C^*$-algebra of observables represented as some algebra $A$ of operators on the spaces of wave functions (\ref{eq1.2}) or (\ref{eq1.4}). As usual, this algebra is a noncommutative extension of some commutative $\mathbb C^*$-algebra $A_0\subset A$. For example, if $S=T^*M$ for some mani\-fold $M$ then $A_0$ is the algebra of continuous complex valued functions, so that $M$ is the {\em space of maximal ideals} of $A_0$. A pair $A_0\subset A$ gives us a space $ \mathcal H$ of wave functions (\ref{eq1.2}) or (\ref{eq1.4}) as the subset of the {\em space of states}. Recall that a {\em state} is a map: \begin{equation} \psi\colon A\to\mathbb C \quad\text{such that} \quad \psi(a^*a)\ge 0 \quad \text{and} \quad \Vert \psi \Vert=1. \label{eq1.8} \end{equation} The set $\mathcal S(A)$ of all states of $A$ is a convex space and its {\em boundary elements} are called {\em pure states} (for example, in the previous example, delta functions of points are pure states). If our $\mathbb C^*$-algebra is represented on $\mathcal H$ by bounded operators then every vector $\left|\psi\right>$ defines the state as the {\em expectation value}. The known strategy to identify spaces (\ref{eq1.2}) and (\ref{eq1.4}) is to represent both as irreducible representation spaces of some algebra admitting a {\em unique irreducible representation}. The constructions of Berezin, Toeplitz and Rawnsley (see for example \cite{R}) are extremely useful for our geometric investigations, and we consider them in \S6. \section{Model for our theory: the classical theory of theta functions} Let $A$ be a principally polarized Abelian variety of complex dimension $g$ with flat metric $g$. Then the tangent bundle $TA$ has the standard constant Hermitian structure (that is, the Euclidean metric, symplectic form and complex structure $I$). The K\"ahler form $2\pi i\omega$ gives a polarization of degree 1. We fix a {\em smooth} Lagrangian decomposition of $A$ \begin{equation} A=T^g_+\times T^g_-, \label{eq2.1} \end{equation} such that both tori are Lagrangian with respect to $\omega$. (In the smooth category, $A$ is the standard torus $\mathbb R^{2g}/\mathbb Z^{2g}$ with the standard constant integral form $\omega$, and the decomposition (\ref{eq2.1}) just consists of putting $\omega$ in normal form.) Let $L$ be a holomorphic line bundle with holomorphic structure given by a Hermitian connection $a$ with curvature form $F_a=2\pi i\cdot\omega$, and $L=\mathcal O_A(\Theta)$, where $\Theta$ is the classical {\em symmetric} theta divisor. The decomposition (\ref{eq2.1}) induces a decomposition \begin{equation} H^1(A,\mathbb Z)=\mathbb Z^g_+\times\mathbb Z^g_-, \label{eq2.2} \end{equation} and a Lagrangian decomposition \begin{equation} A_k=(T^g_+)_k \times (T^g_-)_k \label{eq2.3} \end{equation} of the group of points of order $k$. Any smooth ``irreducible'' $g$-cycle in $A$ is the image $\varphi(T^g)$ of a smooth linear embedding $\varphi\colon T^g\to A$. \subsubsection*{Complex quantization} This is nothing other than the {\em classical theory of theta functions}. Indeed, the decomposition (\ref{eq2.2}) defines the collection of compatible {\em theta structures} of every level $k$: the decomposition (\ref{eq2.3}) defines a Lagrangian decomposition $A_k=(\mathbb Z^g)_k^+\times(\mathbb Z^g)_k^-$, and a decomposition of the spaces of wave functions \begin{equation} \mathcal H_{L^k}=H^0(A, L^k)=\bigoplus_{w\in (\mathbb Z^g)_k^-}\mathbb C\cdot\theta_w, \quad\text{with}\quad \operatorname{rank}\mathcal H_{L^k}=k^g, \label{eq2.4} \end{equation} where $\theta_c$ is the theta function with {\em characteristic} $c$ (see \cite{Mum}). The decomposition (\ref{eq2.4}) is given by the following recipe: we identify the torus $T^g_-$ with the dual torus, and consider vectors $w\in(T^g_-)_k$ as (periodic) linear differential forms on $T^g_-$. Applying the symplectic form $\omega$ gives a collections of linear vector fields $\xi_w$ on $A$ parallel to the fibration by the tori $T^g_+$. Finally, the translations $t_w$ on $A$ obtained as the exponentials of these vector fields give a finite subgroup of the translations group of $A$. Now by choosing $\theta_0\in H^0(A, L^k)$ to be a {\em very symmetric} section (actually, the section with divisor the sum of all the translates of the theta divisor $\Theta$ by points of $(T^g_+)_k)$), we get a basis of $H^0(A, L^k)$: \begin{equation} \{\theta_w=t_w^*(\theta_0)\}. \label{eq2.5} \end{equation} \subsubsection*{Real polarization} The projection of the direct product (\ref{eq2.1}) gives us a real polarization \begin{equation} \pi\colon A\to T^g_-=B. \label{eq2.6} \end{equation} Remark that in this case the action coordinates (\ref{eq1.6}) are just {\em flat} coordinates on $T^g_-=B$, and under this identification \begin{equation} k\mathrm{\text{-}BS}=(T^g_-)_k \label{eq2.7} \end{equation} is the subgroup of points of order $k$. Now we can consider the dual fibration \begin{equation} \pi'\colon A'=\Pic(A/T^g_-)\to T^g_-=B, \label{eq2.8} \end{equation} with fibres \[ (\pi')^{-1} (p)=\operatorname{Hom}(\pi_1(\pi^{-1}(p),\U(1)). \] This fibration admits the section \begin{equation} s_0\in A \quad\text{with}\quad s_0 \cap (\pi')^{-1} (p)=\operatorname{id}\in\operatorname{Hom}(\pi_1(\pi^{-1}(p),\U(1)), \label{eq2.9} \end{equation} so that we have a decomposition \begin{equation} A'=(T^g)'\times T^g_-=B. \label{eq2.10} \end{equation} \begin{rmk} An amazing fact recently proved by Golyshev, Lunts and Orlov \cite{GLO} is that the $2g$-torus $A'$ is canonically equipped with \begin{enumerate} \item a symplectic form $\omega'$; \item a complex structure $I'$. \end{enumerate} \end{rmk} Now we can apply geometric quantization to the real polarization (\ref{eq2.6}) of the phase space $(A,\omega, L^k,a_k)$, where $a_k$ is the Hermitian connection defining the holomorphic structure on $L^k$. Sending the line bundle $L^k$ to the character of the fundamental group of a fibre gives a section \begin{equation} s_{L^k}\subset A'=\Pic(A/T^g_-); \label{eq2.11} \end{equation} and the Bohr--Sommerfeld subset of $B=T^g_-$ is \[ s_0 \cap s_{L^k} \,\subset\, s_0=B=T^g_-\,. \] Under the identification $s_0=T^g_-=\U(1)^g$, the intersection points \[ s_0 \cap s_{L^k}=(\U(1)^g)_k \] are elements of order $k$ in $T^g=\U(1)^g$. We thus get a decomposition \begin{equation} \mathcal H_{\pi}^k=\bigoplus_{\rho\in\U(1)^g_k}\mathbb C\cdot s_{\rho}. \label{eq2.12} \end{equation} \begin{cor} \begin{enumerate} \item $\operatorname{rank} \mathcal H_{L^k}=\operatorname{rank} \mathcal H_{\pi}^k$. \item Moreover, there exists a canonical isomorphism \[ \mathcal H_{L^k}=\mathcal H_{\pi}^k, \] up to a scaling factor. \end{enumerate} \end{cor} We get already this isomorphism up to the action of the $k^g$-torus $(\mathbb C^*)^{k^g}$ (compare decompositions (\ref{eq2.5}) and (\ref{eq2.12})). But the canonical iso\-morphism is defined by the action of the Heisenberg group $H_k$ on holomorphic sections of the line bundle $L^k$ (= the theory of theta functions, see \cite{Mum}) and the natural extension of the action of $H_k$ on the collection of Bohr--Sommerfeld orbits. Each of these representations is irreducible; thus the uniqueness of the irreducible representation of $H_k$ gives a canonical identification of these spaces up to scaling. The functions making up the special bases of these spaces are called {\em classical theta functions with characteristics of level $k$}. A real polarization without degenerate fibres such as $\pi$ in (\ref{eq2.6}) is called {\em regular}. Using more sophisticated techniques (as in \cite{GS2}) we get a basis of the same type for real polarizations with degenerate fibres (see Remark after (\ref{eq1.3})). But if we start with {\em any} polarized K\"ahler manifold $X$, the main question is the following: \begin{quote} how to find a real polarization like (\ref{eq1.3}) on $X$ (possibly with degenerate fibres)? \end{quote} The amazing fact is that we can do it in many absolutely unpredictable cases. For example, we now show how to find a real polarization of complex projective space $\mathbb P^3$. {\bf Warning:} We construct some real polarization of $\mathbb P^3$, but not a special theta basis in $\mathbb P H^0(\mathbb P^3,\mathcal O_{\mathbb P^3}(k))$! For this, we consider a special presentation of the complex threefold $\mathbb P^3$ as a real 6-manifold: let $\Sigma_2$ be a Riemann surface of genus 2. Then as a 6-manifold, \begin{equation} \mathbb P^3=\operatorname{Hom} (\pi_1(\Sigma_2),\operatorname{SU}(2)) /\operatorname{PU}(2)=R_2 \label{eq2.13} \end{equation} is the space of classes of $\operatorname{SU}(2)$-representations of the fundamental group of a Riemann surface of genus 2. Thus $\mathbb P^3$ is the first manifold of the collection of manifolds $R_g$. If we solve the problem of real polarizations of these, we get in particular a real polarization of $\mathbb P^3$. We do this in the following section, but we first extend the direct approach by giving a description in terms of general theories giving rise to these constructions. \section{Chern--Simons quantizations of $R_g$} According to the general procedure, we must present $R_g$ as the classical phase space of some mechanical system. We begin by recalling the full steps of this procedure. A classical field theory on a manifold $M$ has three ingredients: \begin{enumerate} \item a collection $\sA$ of {\em fields} on $M$, which are geometric objects such as sections of vector bundles, connections on vector bundles, maps from $M$ to some auxiliary manifold (the target space) and so on; \item an {\em action functional} \[ S \colon \sA\to\mathbb C \] which is an integral of a function $L$ (the Lagrangian) of fields; \item a collection of observable functionals on the space of fields, \[ \mathcal W \colon \sA\to\mathbb C. \] \end{enumerate} Our case is the following. \subsubsection*{Example: Chern--Simons functional} Here $M$ is a 3-manifold, \[ \sA=\Omega^1(M) \otimes \operatorname{\mathfrak{su}} (2) \] and \begin{equation} S(a)=\frac{1}{8} \pi^2\int_M \operatorname{tr}(a\mathrm{d} a + \frac{2}{3} a^3). \label{eq3.1} \end{equation} As observable, we can consider a {\em Wilson loop}, given by some knot $K\subset M$: \[ \mathcal W_K(a)=\operatorname{tr}(\operatorname{Hol}_K(a)) \] -- the trace of the holonomy of a connection $a$ around the knot $K$. Now let \begin{equation} R_g=\operatorname{Hom} (\pi_1(\Sigma_g), \operatorname{SU}(2))/ \operatorname{PU}(2) \label{eq3.2} \end{equation} be the space of classes of $\operatorname{SU}(2)$-representations of the fundamental group of a Riemann surface of genus $g$. This space is stratified by the subspace of reducible representations \begin{equation} R_g^{\mathrm{red}}\subset R_g, \quad R_g^{\mathrm{irr}}=R_g-R_g^{\mathrm{red}}. \label{eq3.3} \end{equation} To get this space as the phase space of some mechanical system, consider a compact smooth Riemann surface $\Sigma$ of genus $g > 1$ and the trivial Hermitian vector bundle $E_h$ of rank 2 on it. As usual, let $\sA_h$ be the affine space (over the vector space $\Omega^1(\operatorname{End} E_h)$) of Hermitian connections and $\sG_h$ the Hermitian gauge group. This space admits a stratification: \[ \sA_h^{\mathrm{red}}\subset \sA_h \] where the left-hand side is the subset of reducible connections. As usual, let \[ \sA_h^{\mathrm{irr}}=\sA_h-\sA_h^{\mathrm{red}}. \] Sending a connection to its curvature tensor defines a $\sG_h$-equivariant map \begin{equation} F \colon \sA(E_h)\to\Omega^2(\operatorname{End} E_h)=\operatorname{Lie}(\sG_h)^* \label{eq3.4} \end{equation} to the coalgebra Lie of the gauge group. We can consider this map as the moment map with respect to the action of $\sG_h$. The subset \begin{equation} F^{-1}(0)=\sA_F \label{eq3.5} \end{equation} is the subset of flat connections and \[ \sA_F^{\mathrm{irr}}=\sA_F \cap \sA_h^{\mathrm{irr}} \] the subspace of irreducible flat connections. For a connection $a\in \sA_F$ and a tangent vector to $\sA_h$ at $a$ \[ \omega\in\Omega^1(\operatorname{End} E_h)=T \sA_h, \] we have \begin{equation}\omega\in T \sA_F \iff \nabla_a (\omega)=0. \label{eq3.6} \end{equation} The trivial vector bundle $E_h$ admits the trivial connection $\theta$, which is interesting and important from many points of view, and it provides in particular the possibility of identifying $\sA_h$ with $\Omega^1(\operatorname{End} E_h)$ by sending a connection $a$ to the form $a-\theta$. We will identify forms and connections in this way. The space $\sA_h=\Omega^1(\operatorname{End} E_h)$ is the {\em collection of fields} of YM-QFT with the {\em Yang--Mills functional} \[ S(a)=\int_{\Sigma} |F_a|^2. \] Thus $\sA_F$ is a {\em classical phase space}, that is, the space of solutions of an {\em Euler--Lagrange equation} $\delta S(a)=0$. There exists a symplectic structure on the affine space $\sA_h$, induced by the canonical 2-form given on the tangent space $\Omega^1(\operatorname{End} E_h)$ at a connection $a$ by the formula \begin{equation} \Omega_0 (\omega_1,\omega_2)=\int_{\Sigma} \operatorname{tr} (\omega_1 \wedge\omega_2). \label{eq3.7} \end{equation} This form is $\sG_h$-invariant, and its restriction to $\sA_F^{\mathrm{irr}}$ is degenerate along $\sG_h$-orbits: at a connection $a$, for a tangent vector $\omega\in\Omega^1(\operatorname{End} E_h)$, we have \[ \omega\in T\sG_h \iff \omega=\nabla_a \varphi \quad\text{for $ \varphi\in\Omega^0(\operatorname{End} E_h)=\operatorname{Lie}(\sG_h)^*$,} \] and \[ \int_{\Sigma} \operatorname{tr} (\nabla_a \varphi \wedge\omega)=\int_{\Sigma}\operatorname{tr}(\varphi\wedge\nabla_a\omega)=0. \] Hence \begin{equation} \omega\in T \sA_F \iff \Omega_0 (\nabla_a \varphi,\omega)=0. \label{eq3.8} \end{equation} Interpreting (\ref{eq3.4}) as a moment map and using symplectic reduction arguments, we get a nondegenerate closed symplectic form $\Omega$ on the space \[ \sA_F /\sG_h=R_g \] of classes of $\operatorname{SU}(2)$-representations of the fundamental group of the Riemann surface, and a stratification of this space. The form $\Omega$ defines a symplectic structure on $R_g^{\mathrm{irr}}$ and a symplectic orbifold structure on $R_g$. On the other hand, the form $\Omega_0$ on $\sA_h$ is the differential of the 1-form $D$ given by the formula \begin{equation} D(\omega)=\int_{\Sigma} \operatorname{tr}((a) \wedge\omega). \label{eq3.9} \end{equation} We consider this form as a unitary connection $A_0$ on the trivial principal \hbox{$\U(1)$-bundle} $L_0$ on $\sA_h$. To descend this Hermitian bundle and its connection to the orbit space, one defines the $\Theta$-cocycle (or $\Theta$-torsor) on the trivial line bundle (see \cite{RSW}). This cocycle is the $\U(1)$-valued function $\Theta$ on $\sA_h\times\sG_h$ defined as follow: for any triple $(\Sigma,a,g)$ where $(a,g)\in\sA_h\times\sG_h$, we can find a triple $(Y,A,G)$ where $Y$ is a smooth compact 3-manifold, $A$ a $\operatorname{SU}(2)$-connection on the trivial vector bundle $\mathcal E$ on $Y$ and $G$ a gauge transformation of it, such that \[ \partial Y=\Sigma, \quad a=A \rest{\Sigma} \quad \text{and} \quad g=G\rest{\Sigma}. \] Then \begin{equation} \Theta (a, g)=e^{i(\operatorname{CS}(A^G)-\operatorname{CS}(A))}. \label{eq3.10} \end{equation} Recall that the Chern--Simons functional on the space $\sA (\mathcal E_h)$ of unitary connections on the trivial vector bundle is given by the formula \begin{equation} \operatorname{CS}_Y (a_0 +\omega)=\int_Y \operatorname{tr}\left(\omega \wedge F_{a_0}-\frac{2}{3} \omega\wedge\omega\wedge\omega\right). \label{eq3.11} \end{equation} It can be checked that the function (\ref{eq3.10}) does not depend on the choice of the triple $(Y,A,G)$ (see \cite{RSW}, \S2). The differential of $\Theta$ at $(a, g)$ is given by the formula \begin{multline} \mathrm{d}\Theta(\omega,\varphi)= \\ \frac{\pi i}{4}\,\Theta\int_{\Sigma}\Bigl(\operatorname{tr}(g^{-1}\mathrm{d} g\wedge g^{-1}\omega g)-\operatorname{tr}(a\wedge\nabla_{a^g}\varphi)+2\operatorname{tr}(F_{a^g} \wedge \varphi) \Bigr), \label{eq3.12} \end{multline} where $\omega\in\Omega^1(\operatorname{End} E_h)$ and $\varphi\in\Omega^0(\operatorname{End} E_h)=\operatorname{Lie}(\sG_h)$. But the restriction of this differential to the subspace of flat connections is much simpler: \begin{equation} \mathrm{d} \Theta (\omega, \varphi)=\frac{\pi i}{4}\,\Theta\int_{\Sigma}\operatorname{tr}(g^{-1}\mathrm{d} g \wedge g^{-1}\omega g), \label{eq3.13} \end{equation} and is independent of the second coordinate. That this function is in fact a cocycle results from the functional equation \begin{equation} \Theta (a, g_1 g_2)=\Theta (a, g_1) \Theta (a^{g_1}, g_2). \label{eq3.14} \end{equation} Using this function as a torsor $\sA_h \times_{\Theta}\U(1)$ we get a principal $\U(1)$-bundle $S^1(L)$ on the orbit space $\sA_h/\sG_h$: \begin{equation} S^1(L)=(\sA_h\times S^1)/\sG_h, \label{eq3.15} \end{equation} where the gauge group $\sG_h$ acts by \[ g(a, z)=(a^g, \Theta(a,g) z), \] or the line bundle $L$ with a Hermitian structure. Following \cite{RSW}, let us restrict this bundle to the subspace of flat connections $\sA_F$. Then one can check that the restriction of the form $D$ (\ref{eq3.9}) to $\sA_F$ defines a $\U(1)$-connection $A_{\operatorname{CS}}$ on the line bundle $L$. By definition, the curvature form of this connection is \begin{equation} F_{A_{\operatorname{CS}}}=i\cdot\Omega. \label{eq3.16} \end{equation} Thus the quadruple \begin{equation} (R_g,\Omega, L, A_{\operatorname{CS}}) \label{eq3.17} \end{equation} is a {\em prequantum system} and we are ready to switch on the Geometric Quantization Procedure. \section{Complex polarization of $R_g$} The standard way of getting a complex polarization is to give a Riemann surface $\Sigma$ of genus $g$ a conformal structure $I$. We get a complex structure on the space of classes of representations $R_g$ as follows: let $E$ be our complex vector bundle and $\sA$ the space of all connections on it. Every connection $a\in\sA$ is given by a covariant derivative $\nabla_a\colon\Gamma(E)\to\Gamma(E\otimes T^*X)$, a first order differential operator with the ordinary derivative $\mathrm{d}$ as the principal symbol and a complex structure gives the decomposition $\mathrm{d}=\partial+\overline{\partial}$, so any covariant derivative can be decomposed as $\nabla_a=\partial_a+\overline{\partial}_a$, where $\partial_a\colon\Gamma(E)\to\Gamma(E\otimes\Omega^{1,0})$ and $\overline{\partial}_a\colon\Gamma(E)\to\Gamma(E\otimes\Omega^{0,1})$. Thus the space of connections admits a decomposition \begin{equation} \sA=\sA'\times \sA'', \label{eq4.1} \end{equation} where $\sA'$ is an affine space over $\Omega^{1,0}(\operatorname{End} E)$ and $\sA''$ an affine space over $\Omega^{0,1}(\operatorname{End} E)$. The group $\sG$ of all automorphisms of $E$ acts as the group of gauge transformations, and the projection $\operatorname{pr}\colon \sA\to\sA''$ to the space $\sA''$ of $\overline{\partial}$-operators on $E$ is equivariant with respect to the $\sG$-action. Giving $E$ a Hermitian structure $h$, we get the subspace $\sA_h\subset \sA$ of Hermitian connections, and the restriction of the projection $\operatorname{pr}$ to $\sA_h$ is one-to-one. Under this Hermitian metric $h$, every element $g\in\sG$ gives an element $\overline{g}=(g^*)^{-1}$ such that \[ \overline{g}=g \iff g\in\sG_h. \] Now for $g\in\sG$, the action of $\sG$ on the component $\sA''$ is standard: \[ \overline{\partial}_{g(a)}=g\cdot \overline{\partial}_a\cdot g^{-1}=\overline{\partial}_a-(\overline{\partial}_a g)\cdot g^{-1}; \] and the action on the first component $\sA'$ of $\partial$-operators is \[ \partial_{g(a)}=\overline{g}\cdot\partial_a\cdot\overline{g}^{-1}=\partial_a-((\overline{\partial}_a g)\cdot g^{-1})^*. \] It is easy to see directly that the action just described preserves unitary connections: \begin{equation} \sG(\sA_h)=\sA_h, \label{eq4.2} \end{equation} and that the identification $\sA_h=\sA$ is equivariant with respect to this action. It is easy to see that $\overline{\partial}_a^2\in\Omega^{0,2}(\operatorname{End} E)=0$. Thus the orbit space \begin{equation} \sA'' /\sG=\bigcup \mathcal M_i \label{eq4.3} \end{equation} is the union of all components of the moduli space of topologically trivial \hbox{$I$-holomorphic} bundles on $\Sigma_I$. (This union doesn't admit any good structure, as it contains all unstable vector bundles). Finally, the image of $\sA_F\in \sA_h$ is the component $\mathcal M^{\mathrm{ss}}$ of maximal dimension ($3g-3$) of s-classes of semistable vector bundles. Thus by classical technique of GIT of Kempf--Ness type we get: \begin{prop} {\em (Narasimhan--Seshadri)} \[ R_{\Sigma}=R_g=\mathcal M^{\mathrm{ss}}. \] \end{prop} \begin{prop} The form $F_{A_{\operatorname{CS}}}$ (\ref{eq3.16}) is a $(1,1)$-form and the line bundle $L$ admits a unique holomorphic structure compatible with the Hermitian connection $A_{\operatorname{CS}}$. \end{prop} On the other hand, a complex structure $I$ on $\Sigma$ defines a K\"ahler metric on $\mathcal M^{\mathrm{ss}}$ (the so-called Weyl--Petersson metric) with K\"ahler form \begin{equation}\omega_{\mathrm{WP}}=i F_{A_{\operatorname{CS}}}=i\cdot\Omega. \label{eq4.4} \end{equation} This metric defines the Levi-Civita connection on the complex tangent bundle $T \mathcal M^{\mathrm{ss}}$, and hence a Hermitian connection $A_{\mathrm{LC}}$ on the line bundle \begin{equation} \det T \mathcal M^{\mathrm{ss}}=L^{\otimes 4}, \label{eq4.5} \end{equation} and a Hermitian connection $\frac{1}{4}A_{\mathrm{LC}}$ on $L$ compatible with the holomorphic structure on $L$. Thus we have \begin{prop} \label{prop4.3} \[ \frac{1}{4} A_{\mathrm{LC}}=A_{\operatorname{CS}}. \] \end{prop} Finally, considering $\mathcal M^{\mathrm{ss}}$ as a family of $\overline{\partial}$-operators, we get the Quillen determinant line bundle $L$ having a Hermitian connection $A_Q$ with curvature form \begin{equation} F_{A_Q}=i\cdot\Omega. \label{eq4.6} \end{equation} Hence we can extend the equality of Proposition~\ref{prop4.3}: \begin{prop} \[ \frac{1}{4} A_{\mathrm{LC}}=A_{\operatorname{CS}}=A_Q. \] \end{prop} Summarizing, the result of the complex quantization procedure of the prequantum system (\ref{eq3.17}) can be considered to be the spaces of wave functions of level $k$, that is, the spaces of $I$-holomorphic sections \begin{equation} \mathcal H_{L^k}=H^0 (L^k) \label{eq4.7} \end{equation} One knows that this system of spaces and monomorphisms is related to the system of representations of $\operatorname{\mathfrak{su}}(2,\mathbb C)$ in the Weiss--Zumino--Novikov--Witten model of CQFT. Namely, for a half integer $i$, consider the irreducible representation $V_i$ of dimension $2i+1$ of $\operatorname{\mathfrak{su}}(2,\mathbb C)$. The tensor product of two such representations is given by the Clebsch--Gordan rule \begin{equation} V_i\otimes V_j=V_{i+j}\oplus V_{i+j-1}\oplus\dots\oplus V_{i-j}\quad \text{for $i\ge j$,} \label{eq4.8} \end{equation} and the level of $V_i$ is $2i$. Then the {\em fusion ring}\linebreak[3] $R_k(\operatorname{\mathfrak{su}}(2,\mathbb C))$ of level $k$ is the quotient \begin{equation} R_k(\operatorname{\mathfrak{su}}(2,\mathbb C))=R(\operatorname{\mathfrak{su}}(2,\mathbb C))/\Span{V_{(k+1)/2}} \label{eq4.9} \end{equation} of the {\em representation ring} $R(\operatorname{\mathfrak{su}}(2,\mathbb C))$ by the ideal generated by $V_{(k+1)/2}$. Moreover, every character of the ring $R(\operatorname{\mathfrak{su}}(2,\mathbb C))$ is given by a complex number $z\in\mathbb C$ which we can consider as a diagonal $2\times2$ matrix $\operatorname{diag}(iz,-iz)$. This matrix acts on $\operatorname{\mathfrak{su}}(2,\mathbb C)$ and $V_i$ and \begin{equation} \chi_z(V_i)=\operatorname{tr}(\exp(\operatorname{diag}(iz,-iz)))=\frac{\sin((2i+1)z)}{\sin z}. \label{eq4.10} \end{equation} Thus \begin{equation} \chi_z(V_i)=0 \iff z=\frac{n \pi}{k+2}\quad \text{for $1 \le n \le 2i+1$.} \label{eq4.11} \end{equation} In these terms we get: \begin{equation} \mathcal H_{L^k}=\frac{(k+2)^{g-1}}{2^{g-1}} \sum_{n=1}^{k+1} \frac{1}{(\sin(\frac{n\pi}{k+2}))^{2g-2}}\,. \label{eq4.12} \end{equation} See \cite{B} for a mathematical derivation of this formula. \section{Real polarization of $R_g$} The collection of real polarizations of the prequantum system \[ (R_g,\Omega, L, A_{\operatorname{CS}}) \] is given in a very geometric way in the set-up of perturbation theory of \hbox{3-dimensional} Chern--Simons theory. The crucial point is a {\em trinion decomposition} of a Riemann surfaces, given by a choice of a maximal collection of disjoint, noncontractible, pairwise nonisotopic smooth circles on $\Sigma$. An isotopy class of such a collection of circles is called a {\em marking} of the Riemann surface. It is easy to see (\cite{HT}) that any such system contains $3g-3$ simple closed circles \begin{equation} C_1, \dots, C_{3g-3}\subset \Sigma_g, \label{eq5.1} \end{equation} and the complement is the union \begin{equation} \Sigma_g-\{C_1, \dots, C_{3g-3}\}=\bigcup_{i=1}^{2g-2} P_i \label{eq5.2} \end{equation} of $2g-2$ trinions $P_i$, where every trinion is a 2-sphere with 3 disjoint discs deleted: \[ P_i=S^2 \setminus \bigl(D_1\cup D_2\cup D_3\bigr) \quad \text{with} \quad\overline{D}_i\cap\overline{D}_j=\emptyset \quad \text{for} \quad i \ne j. \] A collection $\{C_i\}$ with these conditions is called a {\em trinion decomposition} of $\Sigma$. The invariant of such a decomposition by marking class is given by its {\em $3$-valent dual graph} $\Gamma(\{C_i\})$, associating a vertex to each trinion $P_i$, and an edge linking $P_i$ and $P_j$ to a circle $C_l$ (\ref{eq5.1}) such that \[ C_l\,\subset\,\partial P_i \cap\partial P_j. \] If we fix the isotopy class of a trinion decomposition $\{C_i\}$, we get a map \begin{equation} \pi_{\{C_i\}} \colon R_g\to \mathbb R^{3g-3} \label{eq5.3} \end{equation} with fixed coordinates $(c_1, \dots, c_{3g-3})$ such that \[ c_i (\pi_{\{C_i\}} (\rho))=\frac{1}{\pi}\,\cos^{-1}\bigl(\frac{1}{2}\operatorname{tr} \rho([C_i])\bigr)\in [0, 1]. \] We see that \begin{prop} \begin{enumerate} \item The map $\pi_{\{C_i\}}$ is a real polarization of the system $(R_g,k\cdot\omega,L^k,k\cdot A_{\operatorname{CS}})$. \item The coordinates $c_i$ are {\em action coordinates} for this Hamiltonian system (see (\ref{eq1.6}) and \cite{D}). \end{enumerate} \end{prop} These functions $c_i$ are continuous on all $R_g$ and smooth over $(0,1)$. Moreover, Goldman \cite{G} constructed $\U(1)$-actions on the open dense set \[ U_i=c_i^{-1}(0,1) \] for which the function $c_i$ is the {\em moment map}, and all these $\U(1)$-actions commute with each other. Hence we get: \begin{prop} \begin{enumerate} \item The restriction \[ \pi_{\{C_i\}}\rest{\bigcap_i U_i}\colon\bigcap_i U_i\to (0, 1)^{3g-3} \] is the moment map for the $\U(1)^{3g-3}$-action on $\bigcap_i U_i$. \item The image of $R_g$ under $\pi_{\{C_i\}}$ is a convex polyhedron \begin{equation} I_{\{C_i\}}\subset [0, 1]^{3g-3}. \label{eq5.4} \end{equation} \item The symplectic volume of $R_g$ equals the Euclidean volume of $I_{\{C_i\}} $: \begin{equation}\int_{R_g}\omega^{3g-3}=\operatorname{Vol} I_{\{C_i\}}. \label{eq5.5} \end{equation} \item The expected number of Bohr--Sommerfeld orbits of the real polarization $\{C_i\}$ \begin{equation} N_{\mathrm{BS}}(\pi_{\{C_i\}},R_g,\omega,L,A_{\operatorname{CS}}) \label{eq5.6} \end{equation} equals the number of half integer points in the polyhedron $I_{\{C_i\}}$, and \begin{equation} \lim_{k\to\infty} k^{3-3g}\cdot N_{k\mathrm{\text{-}BS}}=\int_{R_g}\omega^{3g-3} =\operatorname{Vol} I_{\{C_i\}}. \label{eq5.7} \end{equation} \end{enumerate} \end{prop} {From} the combinatorial point of view, the number $N_{\mathrm{BS}}$, or more generally the numbers $N_{k\mathrm{\text{-}BS}}$ of $k$-BS fibres, is determined as follows: consider functions \begin{equation} w\colon\{C_i\}\to\frac{1}{2k}\{0,1,2,\dots,k\} \label{eq5.8} \end{equation} on the collection of edges of the 3-valent graph $\Gamma (\{C_i\})$ to the collection of $\frac{1}{2k}$ integers, such that, for any three edges $C_l, C_m, C_n$ meeting at a vertex $P_i$, the following 3 conditions hold: \begin{enumerate} \item $w(C_l) + w(C_m) + w(C_n)\in \frac{1}{k}\cdot \mathbb Z$; \item $w(C_l) + w(C_m) + w(C_n) \le 1$; \item for any ordering of the triple $C_l, C_m, C_n$, \begin{equation} |w(C_l)-w(C_m)|\le w (C_n) \le w(C_l) + w(C_m). \label{eq5.9} \end{equation} \end{enumerate} Such a function $w$ is called an {\em admissible integer weight of level} $k$ on the graph $\Gamma(\{C_i\})$. \begin{prop} \begin{enumerate} \item The number $|W_g^k|$ of admissible weights of level $k$ is independent of the graph $\Gamma (\{C_i\})$; \item \begin{equation} |W_g^k|=N_{k\mathrm{\text{-}BS}}. \label{eq5.10} \end{equation} \end{enumerate} \end{prop} The conditions (\ref{eq5.9}) are called {\em Clebsch--Gordan conditions} for $\operatorname{\mathfrak{su}}(2,\mathbb C)$, for obvious reasons. We can view the space of all real functions with values in $[0,1]$ subject to these conditions to get a complex $I_{\{C_i\}}$. \begin{rmk} Following this combinatorial approach, we can construct a two dimensional complex $Y_g$: the set of vertices is the set of all dual graphs associated with all types of markings of $\Sigma$. Two vertices are joined by an edge if and only if the two graphs are related by an {\em elementary fusion operation}. The $2$-cells correspond to {\em pentagons}, and so on (see \cite{MS}). The topology of this complex reflects the combinatorial properties of real polarizations of this type. The geometric meaning of this combinatorial description is as follows: consider the space $\mathbb R^{3g-3}$ with action coordinates $c_i$ (\ref{eq5.3}). This space contains the integer sublattice $\mathbb Z^{3g-3}\subset \mathbb R^{3g-3}$, and we can consider the {\em action torus}: \begin{equation} T^A=\mathbb R^{3g-3} / \mathbb Z^{3g-3}. \label{eq5.11} \end{equation} In particular, we get a map \begin{equation} \pi_A \colon R_g\to T^A \label{eq5.12} \end{equation} which glues at most points of the boundary of $I_{\{C_i\}}$. \end{rmk} Now every integer weight $w$ (\ref{eq5.8}) satisfying (1) and (2), but a priori without the Clebsch--Gordan conditions (\ref{eq5.9}), defines a point of order $2k$ on the action torus \[ w\in T^A_{2k}. \] In particular, the collection $W_g^k$ of admissible integer weights (subject to (\ref{eq5.9})) can be considered as a subset of points of order $2k$ on the action torus: \begin{equation} W_g^k\subset T^A_{2k}. \label{eq5.13} \end{equation} On the other hand, every vector $w\in \mathbb R^{3g-3}$ can be interpreted as a {\em differential $1$-form} on $\mathbb R^{3g-3}$, and by the usual construction using the symplectic form $\Omega$, this defines a vector field $\xi_w$ tangent to the fibres of $\pi$. Integrating such vector fields defines the collection of transformations \begin{equation} \{t_w\}=e^{\xi_w}\subset\operatorname{Diff}^+ (R_g). \label{eq5.14} \end{equation} These transformations preserve the curvature form $A_{\operatorname{CS}}$ of the connection. Thus (because $R_g$ is simple connected), there exists a collection of gauge transformations $\alpha_w\in\sG_L$ of $L$ such that \begin{equation} (t_w)^*(A_{\operatorname{CS}})=A_{\operatorname{CS}}^{\alpha_w}. \label{eq5.15} \end{equation} We can view such gauge transformations as $\U(1)$-{\em torsors}, just as in describing the formulas for classical theta functions for Abelian varieties in \S2. Moreover, if $R_{\Sigma}$ is given the K\"ahler structure induced from $\Sigma$ and $s\in H^0(R_{\Sigma}, L^k)$ is a holomorphic section, then we have the following. \begin{prop} \begin{equation} (t_w)^*(s)\in H^0(R_{\Sigma}, L^k) \label{eq5.16} \end{equation} is also a holomorphic section. \end{prop} \begin{cor} If $s_0$ is a {\em sufficiently symmetric} holomorphic section of $L^k$, then the system \begin{equation} \{s_w=t_w^*(s_0)\}\subset H^0(R_{\Sigma}, L^k) \label{eq5.17} \end{equation} is a special theta basis of\/ {\em some subspace} of $H^0(L^k)$. \end{cor} Comparing (\ref{eq5.17}) and (\ref{eq2.6}), we see that the recipe to construct the theta basis is the same as for Abelian varieties with the action space $T^A$ (see \S2) but instead of the full collection $T^A_{2k}$ of points of order $2k$, we only use the subset $W_g^k\subset T^A_{2k}$. In our realistic situation, the prequantum system $(R_g,\Omega,L,A_{\operatorname{CS}})$ is far from the regular ``theoretical'' case. But in the fundamental papers \cite{JW1} and \cite{JW2} there is a well-defined correction to the ``theoretical'' situation. Here we only explain what we must do at a maximally degenerate Bohr--Sommerfeld fibre. We get proofs of the central statements of Proposition~5.4 and Corollary~5.1 by a quite fruitful method: we give new definitions making the statements almost obvious. We do this in the following special section. \subsubsection*{Unitary Schottky representations} Every oriented trinion $P_i$ defines a {\em handle\/} $\mathrm{HP}_i$, and all these handles glue together to give a {\em handlebody\/} $H_{\{C_i\}}$, a compact 3-manifold such that \begin{equation}\partial H_{\{C_i\}}=\Sigma. \label{eq5.18} \end{equation} We get a surjection \begin{equation} \varphi \colon \pi_1(\Sigma)\to \pi_1(H_{\{C_i\}}), \label{eq5.19} \end{equation} which defines the subspace \begin{equation} B_{\{C_i\}}=\bigl\{\rho\in R_g \bigm| \rho \rest{\ker \varphi}=1\bigr\}. \label{eq5.20} \end{equation} \begin{prop} \begin{enumerate} \item $B_{\{C_i\}}$ is a Lagrangian subspace of $R_g$. \item More precisely, it is a fibre of the real polarization $\pi_{\{C_i\}}$ (\ref{eq5.3}): \begin{equation} B_{\{C_i\}}=\pi_{\{C_i\}}^{-1} (1,\dots,1). \label{eq5.21} \end{equation} \item Moreover, $B_{\{C_i\}}$ is a Bohr--Sommerfeld orbit of $\pi_{\{C_i\}}$. \end{enumerate} \end{prop} This Lagrangian subspace $B_{\{C_i\}}$ is singular: \[ B_{\{C_i\}}^{\mathrm{red}}=B_{\{C_i\}} \cap R_g^{\mathrm{red}}=\operatorname{Sing} B_{\{C_i\}}; \] and $B_{\{C_i\}}^{\mathrm{irr}}=B_{\{C_i\}} \cap R_g^{\mathrm{irr}}$ is a nonsingular Lagrangian subvariety. Under the identification of $R_g$ with the moduli space of s-classes of semi\-stable vector bundles on the algebraic curve $\Sigma$, the subspace $B_{\{C_i\}}$ is called the subset of {\em unitary Schottky subbundles}. Obviously for this Bohr--Sommerfeld fibre $w_{\mathrm{US}}\in T^A_k$, we must use a special description of the symplectomorphism $t_{w_{\mathrm{US}}}$. This was done in the papers \cite{JW1} and \cite{JW2}. Returning to the general geometric quantization procedure and summarizing these results, we get two spaces of wave functions: complex quantization gives the spaces \[ \mathcal H_{\Sigma}^k=H^0(L^k) \] of $I$-holomorphic sections of $L^k$, and real quantization gives the direct sum \begin{equation} \mathcal H_{\pi}^k\,=\sum_{k\text{-BS fibres}}\!\mathbb C\cdot s_{k\mathrm{\text{-}BS}} \label{eq5.22} \end{equation} of lines generated by covariant constant sections of restrictions of our prequantum line bundle $(L^k,k\cdot A_{\operatorname{CS}})$ to the Bohr--Sommerfeld fibres of $\pi$ of level $k$. The amazing fact is the following: \begin{prop} For any level $k$, any complex Riemann surface $\Sigma$, and any trinion decomposition $\{C_i\}$ with the real polarization $\pi$ of $R_g$ we have \begin{equation} \operatorname{rank} \mathcal H_{\Sigma}=H^0(L^k)=\operatorname{rank} \mathcal H_{\pi}, \label{eq5.23} \end{equation} and these ranks can be computed by the Verlinde calculus. \end{prop} \begin{cor} Our construction gives a distinguished theta basis of the first space $H^0(L^k)$. \end{cor} This isomorphism between spaces of wave functions underlies all the ``modular'' behavior of gauge theory invariants in dimensions 2, 3 and 4. The final ``classical'' question concerns the Fourier decomposition of our non-Abelian theta functions $s_w$ (5.17). It can be done using the Fourier decomposition along coordinates $\{c_i\}$ of the action torus $T^A$ (\ref{eq5.11}) twisting by the system of torsors $\{\alpha_w\}$ (5.15). Roughly speaking, the theta functions $s_w$ (5.17) are {\em truncated} theta functions on the $(6g-6)$-dimensional ``Fourier torus'' \[ T_F=\U(1)^{3g-3} \times T^A. \] Namely all coefficients of Fourier decompositions not satisfying the Clebsch--Gordan conditions (\ref{eq5.9}) must go to zero. Can this condition be interpreted in terms of the heat equation? \section{Other definition of a theta basis} We must first recall the main constructions of GQP. Let $h$ be the Hermitian form on $L$, and \begin{equation} \mu=\frac{1}{(3g-3)!}\omega^{3g-3} \label{eq6.1} \end{equation} the volume form on $R_g$. Then we have a scalar product and norm on the space $\Gamma (L^k)$ of global differentiable sections of $L^k$: \begin{equation} \Span{s_1,s_2}=\int_{R_g} h(s_1, s_2)\cdot \mu \quad \text{and} \quad \Vert s \Vert=\sqrt{\Span{s,s}}\,. \label{eq6.2} \end{equation} Let $L^2(L^k)$ be the $L^2$-completion of $\Gamma(L^k)$ and \begin{equation} P_k \colon L^2(L^k)\to H^0(L^k) \label{eq6.3} \end{equation} the orthogonal projection to the finite dimensional subspace of {\em holomorphic} sections $H^0(L^k)\subset L^2(L^k)$. The ring $C^{\infty}(R_g)$ of smooth functions on $R_g$ acts on $L^2(L^k)$ by multi\-plication $s\to f\cdot s$, and acts on the space $H^0(L)$ as a {\em Toeplitz operator}: \begin{equation} T_f=P \odot f\in\operatorname{End}(H^0(L^k)); \label{eq6.4} \end{equation} the map $C^{\infty}(R_g)\to\operatorname{End}(H^0(L^k))$ is called the {\em Berezin--Toeplitz map}. Now, let \begin{equation} p \colon L^*\to R_g \label{eq6.5} \end{equation} be the principal $\mathbb C^*$-bundle of $L$. Every point $x\in L^k$ defines a linear form \begin{equation} l_x \colon H^0(L^k)\to\mathbb C, \quad\text{given by}\quad s(p(x))=l_x (s)\cdot x, \label{eq6.6} \end{equation} and the {\em coherent state} vector $s_x\in H^0(L^k)$ associated to $x$, which is uniquely determined by the equation \begin{equation} \Span{s_x,s}=l_x(s). \label{eq6.7} \end{equation} Thus we get a map \begin{equation} \varphi_k \colon R_g\to \mathbb P H^0(L^k), \label{eq6.8} \end{equation} which is nothing other than the Hermitian conjugate of the standard algebraic geometric map by a complete linear system to the dual space, because of the equality \[ s_{\alpha\cdot x}=\overline{\alpha}^{-1}\cdot s_x \quad \text{for $\alpha\in\mathbb C^*$.} \] Now, following John Rawnsley, we can define {\em coherent projectors} $P_{p(x)}$ and the {\em Rawnsley epsilon function} $\varepsilon\colon R_g\to\mathbb R^+$ in such a way that: \begin{equation} \varepsilon(p(x))=|x|^2\cdot \Span{s_x,s_x} \quad \text{and} \quad h(s_1, s_2)_{p(x)}=\varepsilon (p(x))\cdot \Span{s_1, P_{p(x)} s_2}. \label{eq6.9} \end{equation} Since $\varepsilon > 0$ we can modify the old measure on $R_g$: \begin{equation} \mu_{\varepsilon}=\varepsilon\cdot \mu, \label{eq6.10} \end{equation} where $\mu$ is (6.1). This measure gives an integral representation of Toeplitz operators: \begin{equation} T_f(s)=\int_{R_g} f(p(x))\cdot P_{p(x)}(s)\cdot \mu_{\varepsilon}. \label{eq6.11} \end{equation} Up to now, we have been working with a {\em complex polarization}. Let us return to the real polarization $\pi$ (\ref{eq5.3}). We can identify the target real space $\mathbb R^{3g-3}$ of $\pi$ with the dual space \[ \mathbb R^{3g-3}=(\mathbb R^{3g-3})^* \] and we can consider our vectors $w\in W^k_g\subset T^A_{2k}$ as {\em linear functions} on the target space $\mathbb R^{3g-3}$. Thus we get a collection of functions \begin{equation} \pi^* w \colon R_g\to \mathbb R \label{eq6.12} \end{equation} and a collection of Toeplitz operators \begin{equation} T_{\pi^* w}\in\operatorname{End}(H^0(L^k)). \label{eq6.13} \end{equation} Let us choose one (very symmetric) section $s_0$ in the following way: for $k=1$, the space $H^0(L)$ is the space of ordinary theta functions (see, for example, \cite{BL}) and every semistable bundle $E$ defines a theta divisor \[ \Theta_E=\bigl\{L\in\Pic_{g-1}(\Sigma)\bigm| h^0(E\otimes L)>0\bigr\}. \] Let $s_E$ be the section with this divisor as its zero set. Then one has the section \begin{equation} s_0=s_{\mathcal O_{\Sigma}\oplus\mathcal O_{\Sigma}}^k\in H^0(R_{\Sigma},L^k), \label{eq6.14} \end{equation} and the collection of sections \begin{equation} \{T_{\pi^* w}(s_0)=s_w\}\subset H^0(R_{\Sigma},L^k). \label{eq6.15} \end{equation} Using the integral representation (\ref{eq6.11}), Rawnsley's localization, and Proposition 5.6, we get immediately \begin{thm} The sections $\{T_{\pi^* w}(s_0)=s_w\}$ form a basis of $H^0(R_\Sigma, L^k)$. \end{thm} The reader not wishing to check the following statement may take (\ref{eq6.15}) as the {\em definition} of the theta basis: \begin{prop} The basis (5.17) coincides with the basis (\ref{eq6.15}). \end{prop} Thus, we do indeed get a generalization of theta functions. \section{What next?} The theory of theta functions outlined above is just a small sample of the applications of the Geometric Quantization Procedure to algebraic geometry. Here we extend the list, mentioning applications which are natural generalizations of the above constructions. \subsubsection*{Generalization to vector bundles of higher rank} This construction is new even in the classical set-up. Let us return to a real polarization of an Abelian variety $\pi\colon A\to T^g_-=B$, and its dual fibration \[ \pi'\colon A'=\Pic(A/T^g_-)\to T^g_-=B, \] with fibres \[ (\pi')^{-1} (p)=\operatorname{Hom}(\pi_1(\pi^{-1}(p)),\U(1)) \] and section \[ s_0\in A' \quad\text{with}\quad s_0 \cap (\pi')^{-1} (p)=\operatorname{id}\in\operatorname{Hom}(\pi_1(\pi^{-1}(p)),\U(1)). \] Every stable holomorphic vector bundle $E$ on a {\em generic} principally polarized Abelian variety $A$ carries a Hermitian--Einstein connection $a_E$ that defines a holomorphic structure on $E$ with curvature \begin{equation} F_{a_E}=\Lambda i\cdot\omega, \label{eq7.1} \end{equation} where $\Lambda$ is any constant element of $\U(\operatorname{rank}(E))$, for example, $\operatorname{id}$. Hence the restriction of $a_E$ to every fibre of $\pi$ is a flat Hermitian connection on a $g$-torus, and thus \begin{equation} \begin{gathered} (a_E)\rest{\pi^{-1}(b)}=\chi_1\oplus\dots\oplus \chi_{\operatorname{rank} E}, \quad\text{where}\\ \chi_i\in\operatorname{Hom}(\pi_1(\pi^{-1}(b)),\U(1)))=\Pic(\pi^{-1}(b))=(\pi')^{-1}(b). \end{gathered} \label{eq7.2} \end{equation} \begin{dfn} For vector bundles of higher rank, an {\em $E$-{\rm BS} fibre} is a fibre $\pi^{-1}(b)$ such that the restriction $(E,a_E)\rest{\pi^{-1}(b)}$ admits a covariant constant section. \end{dfn} Suppose that $E$ is {\em ample}, and in particular that \[ H^i(A,E)=0\quad\text{for}\quad i>0. \] Then, alongside the complex ``space of wave functions'' $H^0(A,E)$, we get a new space of wave functions \begin{equation} \mathcal H_{\pi}^E=\bigoplus_{\text{$E$-BS}}\mathbb C\cdot s_{i}, \label{eq7.3} \end{equation} where $s_i$ is a covariant constant section of the restriction of $E$ to a $E$-BS fibre. We again have the problem of comparing the spaces \begin{equation} \mathcal H_{E}=H^0(A, E) \quad \text{and} \quad \mathcal H_{\pi}^E. \label{eq7.4} \end{equation} This problem can be solved by analogous (but more sophisticated) methods from GQP. In particular \begin{prop} The space $H^0(A,E)$ admits a canonical theta basis. \end{prop} Of course, if $X$ is any K\"ahler manifold with some real polarization (\ref{eq1.3}) and stable holomorphic vector bundle $E$ admitting an Hermitian connection with curvature of the form (\ref{eq7.1}), we get two spaces \begin{equation} \mathcal H_{E}=H^0(X, E) \quad \text{and} \quad \mathcal H_{\pi}^E \label{eq7.5} \end{equation} to compare. In particular if $X=R_{\Sigma_g}$ one has \begin{prop} For a stable vector bundle of higher rank $E$ on $R_{\Sigma}$, \[ H^0(R_{\Sigma_g}, E)=\bigoplus_{E\text{-}\mathrm{BS}}\mathbb C\cdot s_{i}. \] \end{prop} This holds in particular for all the symmetric powers of the Poincar\'e bundles. \subsection*{K3 surfaces} If the transcendental lattice $T_S$ of a K3 surface $S$ contains an even unimodular sublattice $H$ of rank 2, then $S$ admits a real polarization (see for example \cite{G1}, \cite{G2} or \cite{T}). Then every ample complete linear system on $S$ admits a special theta basis. \subsection*{Geometry behind these constructions: mirror reflection of holomorphic geometry} If our real polarization (\ref{eq1.3}) is regular, that is, the differential of the map $\pi\colon X\to B$ is surjective then all fibres are $n$-tori, and the second fibration $\pi'\colon X'\to B$ can be defined fibrewise in the usual way: \[ (\pi')^{-1}(b)=\operatorname{Hom}(\pi_1(\pi^{-1}(b)),\U(1)) \quad\text{for any point $b\in B$;} \] that is, the fibre of $\pi'$ is the space of classes of flat connections on the trivial line bundle on $\pi^{-1}(b)$. This is a fibration in groups, and we want to consider its zero section $s\colon B\to X'$ as a submanifold $s_0\subset X'$. The restriction of a pair $(E, a_E)$ to any fibre $\pi^{-1}(b)$ defines a finite set of points $(\pi')^{-1}(b)$, and hence a multisection \begin{equation} \operatorname{GFT}(E)\subset X', \label{eq8.1} \end{equation} which we again consider as a middle dimensional submanifold of $X'$. This cycle is called the {\em Geometric Fourier Transformation} of $E$. Under the identification $s_0=B$, the set of $E$-BS fibres is defined now as the set of intersection points \begin{equation} E\text{-BS}=s_0 \cap \operatorname{GFT}(E), \label{eq8.2} \end{equation} and under some geometric conditions, we expect that the number \begin{equation} \# E\text{-BS}=[s_0] \cap [\operatorname{GFT}(E)] \label{eq8.3} \end{equation} where $[\ ]$ is the cohomology class of a submanifold. In the general case of a polarization with degenerate fibres, this construction can be performed over the open subset $B_0\subset B$ of smooth tori and a number of questions arise: \begin{enumerate} \item to construct a smooth compactification $S'$; \item to construct a symplectic form $\omega'$ and extend it to $S'$, in such a way that $\pi'$ is a new real polarization; \item to construct a complex polarization of $S'$ such that the fibration $\pi$ is given by construction we have described, starting from $(S',\omega',L',a')$. \end{enumerate} In full generality these problems are very hard (see for example \cite{G1}, \cite{G2}). The ideal picture is described by the mirror diagram \begin{equation} \renewcommand{\arraystretch}{1.5} \begin{matrix} &&S &\longleftarrow & E \\ && \kern3mm \big\downarrow \pi &&\\ & & B&& \\ && \kern4.5mm \big\uparrow \pi'\\ \operatorname{GFT}(E)&\longrightarrow &S' &\longleftarrow & s_0 \\ \end{matrix} \label{eq8.4} \end{equation} with holomorphic objects (vector bundles) corresponding to the top of (\ref{eq8.4}) and special Lagrangian cycles to the bottom. \subsection*{The inverse problem} Every stable holomorphic vector bundle $E$ on an $S_I$ (top of (\ref{eq8.4})) is a point in the moduli space of stable holomorphic vector bundles \begin{equation} E\in \mathcal M_{[E]}^\mathrm{s} \label{eq8.5} \end{equation} of topological type $[E]$. But the cycle $\operatorname{GFT}(E)$ (bottom of (\ref{eq8.4})) is a point in the moduli space of special Lagrangian cycles (see \cite{HL}) \begin{equation} \operatorname{GFT}(E)\in \mathcal M^{[\operatorname{GFT}(E)]} \label{eq8.6} \end{equation} of topological type $[\operatorname{GFT}(E)]$. Thus we get a map \begin{equation} \operatorname{GFT} \colon \mathcal M_{[E]}^\mathrm{s}\to \mathcal M^{[\operatorname{GFT}(E)]} \label{eq8.7} \end{equation} sending $E$ to $\operatorname{GFT}(E)$. However, we have not used all the information contained in $E$. Namely, $\operatorname{GFT}(E)$ can be defined as a {\em supercycle} (or {\em brane}). It's easy to see that any cycle $\operatorname{GFT}(E)$ admits a tautological topologically trivial line subbundle $L$ with Hermitian connection $s_{\tau}$. A pair \begin{equation} (\operatorname{GFT}(E),a_{\tau})=\operatorname{sGFT}(E) \label{eq8.8} \end{equation} of this type is called a {\em supercycle} (or brane). The attempt to reconstruct the vector bundle $E$ (top of (\ref{eq8.4})) from the supercycle $\operatorname{sGFT}(E)$ (bottom of (\ref{eq8.4})) is called the {\em inverse problem}. More formally, let $S\mathcal M^{[\operatorname{GFT}(E)]}$ be the moduli space of supercycles of topological type $[\operatorname{GFT}(E)]$. Then in many special cases, one can prove that the map \[ \operatorname{sGFT}\colon\mathcal M^\mathrm{s}_{[E]}\to S\mathcal M^{([\operatorname{GFT}(E)])} \] is an embedding at the general point. That is, a general stable vector bundle $E$ (top of (\ref{eq8.4})) is uniquely determined by the supercycle $\operatorname{sGFT}(E)$ on $S'$ (bottom of (\ref{eq8.4})). For example, if the fibration $\pi \colon X\to B$ is the family of all deformations (with degenerations) of the general fibre $\pi^{-1}(b)=T^n$ as a torus with special structure inside $S$, then \[ X'=S\mathcal M^{[T^n]} \] is the family of all deformations (with degenerations) of the pair $(T^n, \tau_0)$, where $\tau_0$ is the trivial connection. At present this program is only realized in part (see, for example, \cite{T}). We must first use the experience of the geometric quantization procedure, and apply it in the Calabi--Yau realm of {\em simply connected K\"ahler manifolds with canonical class zero}. But in this paper, we want to emphasize that there exists the collection of singular Fano varieties $R_g$ for which these constructions are very important, although this is an extremely irregular case. \subsubsection*{Acknowledgments} I would like to express my gratitude to the Institut de Math\'ematiques de Jussieu and the Ecole Normale Sup\'erieure, and personally to Joseph Le Potier and Arnaud Beauville for support and hospitality. I wish to thank Yves Laslo and Christoph Sorger for many helpful discussions. Thanks are again due to Miles Reid for tidying up the English.
\section*{Introduction} It is well known that defects like quantum vortices, spin vortices, dislocations and disclinations play an essential role in physics of low-temperature phases of thin films. Berezinskii \cite{Ber} and then Kosterlitz and Thouless \cite{KT} recognized that there is a class of phase transitions in $2d$ systems related to the defects. The main idea of their approach is that in $2d$ the defects can be treated as point objects interacting like charged particles. It is usually called Coulomb gas analogy. The low-temperature phase corresponds to a fluid constituted of bound uncharged defect-antidefect pairs, which is an insulator, whereas the high-temperature phase contains free charged particles and can be treated as plasma. Correspondingly, in the low-temperature phase the correlation length is infinite whereas in the high-temperature phase it is finite. A huge number of works is devoted to different aspects of the problem, see, e.g., the surveys \cite{78KT,80Nel,83Nel,87Min,GG98}. The scheme proposed by Kosterlitz and Thouless can be applied to superfluid and hexatic films and planar $2d$ magnetics. It admits a generalization for crystalline films, see Refs. \cite{Solid,88Str}. There are also applications to superconductive materials, especially to high-$T_c$ superconductors, see, e.g., Ref. \cite{BFGLV}. The dynamics of the films in the presence of the defects was considered in the papers \cite{Dynamics1,Dynamics2}. In the works a complete set of equations is formulated describing both motion of the defects and hydrodynamic degrees of freedom. Then, to obtain macroscopic dynamic equations, an averaging over an intermediate scale was performed. At the procedure the ``current density'' related to the defects was substituted by an expression proportional to the average ``electric field'' and to gradients of the temperature and of the chemical potential. The resulting equations perfectly correspond to the problems solved in the works \cite{Dynamics1,Dynamics2}. Unfortunately, at the procedure an information concerning high-order correlations of the defect motion is lost. That is the motivation for the present work where these high-order correlations are examined. We start from the same ``microscopical'' equations of the defect dynamics as was accepted in Ref. \cite{Dynamics1}. Following the works we focus mainly on the case when the motion of the defects is determined by the Langevin equation describing an interplay between the Coulomb interaction and the thermal noise. We believe that the approach is correct for hexatic films (membranes, Langmuir films, freely suspended films). The situation is a bit more complicated for the vortices in superfluid films because of the Magnus force. Nevertheless, the equation for the vortices is close to the Langevin equation, see Ref. \cite{Dynamics1}. Similar equations can be formulated for the dislocations in crystalline films, see Ref. \cite{Dynamics2}, for the vortices in superconductors in some interval of scales, see, e.g., Ref. \cite{BFGLV}, and for the spin vortices in planar $2d$ magnetics. We will not consider the last cases here, though our scheme is, generally, applicable to the systems. Treating non-simultaneous correlation functions related to the defects one should take into account creation and annihilation processes also. For the purpose we use the Doi technique \cite{76Doi} who demonstrated that dynamics of classical particles involved into chemical reactions can be examined in terms of the creation and annihilation operators, like in the quantum field theory. We consider correlation functions $F_{2n}$ of the ``charge density'' $\rho$ (vorticity, disclinicity etc) provided that the so-called renormalized fugacity $y$ is small. The inequality $y\ll1$ is satisfied for large scales in the low-temperature phase and probably in some region of scales above $T_c$. In statics, the normal estimate $F_{2n}\sim F_2^n$ is valid at the condition. Surprisingly, the non-simultaneous high-order correlation functions $F_{2n}$ appear to be much larger than their normal estimate $F_2^n$. In the low-temperature phase the phenomenon reveals an anomalous scaling on large scales. The reason for such unusual behavior is that the main contribution to high-order non-simultaneous correlation functions is associated with rare single defect-antidefect pairs. The situation resembles the intermittency phenomenon in turbulence, see, e.g., Ref. \cite{Frish}. It can also be compared with non-trivial tails of probability distribution functions in the physics of disordered materials, see, e.g., Refs. \cite{LGP,efetov}. Some preliminary results were published in the paper \cite{99Leb}. Let us give a qualitative explanation of the phenomenon. To obtain a non-zero contribution to the correlation function $F_{2n}(t_1,\dots,t_n;{\bbox r}_1,\dots,{\bbox r}_n)$ one must consider trajectories of the particles passing through the points ${\bbox r}_1, \dots, {\bbox r}_n$ at the time moments $t_1, \dots, t_n$. The situation is illustrated in Fig. \ref{vort8}. The ``single-pair'' contribution has to be compared with a ``normal'' contribution associated with a number of defect-antidefect pairs. Though the normal contribution contains an additional large entropy factor it has also an additional small factor related to a small probability to observe a defect-antidefect pair with a separation larger than the core radius. As a result of the competition, the normal contribution appears to be smaller. To avoid a misunderstanding, let us stress that the arguments do not work for the simultaneous correlation functions. The reason is that trajectories of two defects cannot pass through $n>2$ points simultaneously, see Fig. \ref{vort7}. Our paper is organized as follows. In Section \ref{basic} we remind some basic facts concerning static properties of the $2d$ defects and their dynamics and then we shortly review the Doi technique \cite{76Doi} suitable for our problem. In Section \ref{quantum} we develop a diagrammatic representation for dynamical objects and examine the two-particle conditional probability which is extensively exploited in the subsequent consideration. In Section \ref{renorm} we demonstrate how renormalization of different parameters can be obtained in the framework of our dynamic approach. Actually, the renormalization is reduced to the well known static renormalization group equations. In Section \ref{corr} we consider correlation functions of the ``charge density'' and ground the properties announced above. In Section \ref{helium} we generalize our procedure for the case of superfluid films. In the Conclusion we discuss the main results of our work and their possible relations to other systems. Some calculations are placed into Appendices. \section{Basic Relations} \label{basic} Static properties of the system of the vortex-like defects in thin films can be described quite universally. The starting point of the description is the free energy associated with the defects \begin{equation} {\cal F}=-\sum_{i\neq j}{T\beta}\,n_in_j \ln\left(\frac{|{\bbox x}_i-{\bbox x}_j|}{a}\right)+\sum_j \mu(n_j) \,, \label{H04} \end{equation} where the subscripts $i,j$ label defects, ${\bbox x}_i$ are positions of the defects, $a$ is a cutoff parameter of the order of the size of the defect core, $n_i$ are integer numbers determining the ``strength'' of the defects, $\beta$ is a dimensionless $T$-dependent factor and $\mu$ is the energy associated with the core. The expression (\ref{H04}) is correct for quantum vortices in superfluid films, for disclinations in hexatic films, and for spin vortices in $2d$ planar magnets. For dislocations in crystalline films the expression (\ref{H04}) has to be slightly modified \cite{Solid}, but the main peculiarity of the free energy, the logarithmic dependence on the separation, remains the same. The Gibbs distribution $\exp(-{\cal F}/T)$ corresponding to the energy (\ref{H04}) can be treated as the partition function of two-dimensional point particles with charges $n_j$, $\beta$ playing a role of the ``inverse temperature''. The parameter $\beta$ can be considered also as the Coulomb coupling constant. Basing on the electrostatic analogy one can introduce the ``charge density'' \begin{equation} \rho({\bbox r})=\sum\limits_j n_j \delta\left({\bbox r}-{\bbox x}_j\right) \,. \label{den} \end{equation} The quantity $\rho$ is vorticity for superfluid films and disclinicity for hexatic films. We will treat the case when defects are produced by thermal fluctuations. Since both creation and annihilation processes conserve the ``charge'' we should accept that the total charge is zero: $\sum_j n_j=0$. It leads to the constraint \begin{eqnarray} && \int{\rm d}^2 r\,\rho({\bbox r})=0 \,, \label{Ch1} \end{eqnarray} where the integration is performed over the total area of the specimen. Below we assume that for $|n|>1$ the core energy $\mu(n)$ is so large that such defects are hardly created. Then only defects with the charges $n_i=\pm1$ should be taken into account. We will call the objects with the charges $n_i=1$ defects and the objects with the charges $n_i=-1$ antidefects. Because of the constraint $\sum_j n_j=0$ there can be simultaneously $N$ defects and $N$ antidefects in the system. Thus, the partition function of the system can be characterized via a set of probability distribution functions ${\cal P}_{2N}$ depending on coordinates of $2N$ ``particles''. In accordance with Eq. (\ref{H04}) the functions can be written as \begin{eqnarray} && {\cal P}_{2N}({\bbox x}_1, \dots, {\bbox x}_{2N}) \nonumber \\ && =Z^{-1}\left(\frac{y_0}{a^2}\right)^{2N} \exp\left\{\sum_{i\neq j}{\beta}\,n_in_j \ln\left(\frac{|{\bbox x}_i-{\bbox x}_j|}{a}\right)\right\} \,, \label{parti} \end{eqnarray} where $Z$ is the sum over states and the quantity $y_0=\exp(-\mu/T)$ is usually called fugacity. The possibility to neglect charges with $|n|>1$ implies that the fugacity is small. The low-temperature (insulator) phase can be treated as a system constituted of bound defect-antidefect pairs. In the high-temperature (plasma) phase there are unbound charges which essentially influence the system on scales larger than the correlation length $r_c$. We will treat the low-temperature phase and the region of scales between $a$ and $r_c$ in the high-temperature phase where one can neglect the role of the unbound charges and only the bound defect-antidefect pairs have to be taken into account. The presence of the pairs in the system leads to non-trivial ``dielectric'' properties of the medium. As a result the interaction between the charges is modified, the effect can be described in terms of a scale-dependent ``dielectric constant'' of the medium as is suggested in Ref. \cite{KT}. In other words, the effective coupling constant $\beta$ becomes dependent on the separation between the charges. The scale dependence of $\beta$ can be described in the framework of the scheme proposed by Kosterlitz, \cite{74Kos}. Namely, the partition function of the system can be integrated over separations of the defect-antidefect pairs between the core size ${a}$ and a scale $r$. After the procedure, that can be interpreted as shifting the core radius $a\to r$, the form of the probability distribution functions (\ref{parti}) is reproduced (with $r$ instead of $a$), but the parameters $\beta$ and $y$ are renormalized. The $r$-dependence of $\beta$ and $y$ is determined by the following renormalization group equations found in Ref. \cite{74Kos} \begin{eqnarray} && \frac{{\rm d}\beta}{{\rm d}\ln(r/a)}=-cy^2 \,, \qquad \frac{{\rm d}y}{{\rm d}\ln(r/a)}=(2-\beta)y \,, \label{rg} \end{eqnarray} where $c$ is a numerical factor of order unity. The $r$-dependent function $y$ is the renormalized fugacity. It determines a concentration of defects belonging to the bound pairs with separations of the order of $r$, the concentration can be estimated as $y/r^2$. The renormalized value of $\beta$ determines the dependence of the strength of the Coulomb interaction on the separation between the charges. In the low-temperature phase, the effective value of $\beta$ tends to a constant on large scales. The asymptotic value of $\beta$ is larger than $2$, the critical value $\beta=2$ corresponds to the transition temperature. In the asymptotic region, where $\beta$ can be treated as $r$-independent, the renormalized fugacity $y$ remains $r$-dependent. Its asymptotic behavior can easily be extracted from Eq. (\ref{rg}): \begin{eqnarray} && y\propto r^{2-\beta} \,. \label{fug} \end{eqnarray} Thus, in the low-temperature phase $y$ tends to zero as scale increases. Let us turn to simultaneous correlation functions of the charge density $\rho$ (\ref{den}). The odd correlation function are zero. Indeed, the system is symmetric under permuting defects and antidefects whereas the charge density (\ref{den}) changes its sign at the permutation. The pair correlation function can be written as (see, e.g., Ref. \cite{JKKN}) \begin{eqnarray} && \langle\rho({\bbox r})\rho(0)\rangle \sim y^2(r)/r^4 \,. \label{pair} \end{eqnarray} A generalization of the relation (\ref{pair}) can be obtained (see Ref. \cite{Kad}) which is \begin{eqnarray} && \langle\rho({\bbox r}_1)\dots\rho({\bbox r}_{2n})\rangle \sim \frac{y^{2n}(r_*)}{r_*^{4n}}\sim \langle\rho({\bbox r_*})\rho(0)\rangle^n \,, \label{simul1} \end{eqnarray} where all separation $|{\bbox r}_i-{\bbox r}_j|$ are assumed to be of the same order $r_*$. In the large-scale limit where $\beta$ is saturated we have \begin{eqnarray} && \langle\rho(X{\bbox r}_1)\dots\rho(X{\bbox r}_{2n})\rangle =X^{-2\beta n}\langle\rho({\bbox r}_1)\dots\rho({\bbox r}_{2n})\rangle \,, \label{high} \end{eqnarray} where $X$ is an arbitrary factor. The relation (\ref{high}) shows that the simultaneous statistics of $\rho$ has normal scaling, that is scaling exponents of the correlation functions of the order $2n$ are equal to $n$ times the scaling exponent of the pair correlation function (\ref{pair}). We will demonstrate that the behavior of non-simultaneous correlation functions of the charge density is quite different. \subsection{Dynamics} To examine dynamical characteristics of the system we should formulate a dynamical equation for a defect motion. Following Ref. \cite{Dynamics1} we accept the following stochastic equation \begin{equation} \frac{{\rm d} {\bbox x}_j}{{\rm d}t}= -\frac{D}{T}\frac{\partial {\cal F}} {\partial {\bbox x}_j}+{\bbox\xi}_j \,, \label{H51} \end{equation} determining the trajectory of the $j$-th defect. Here ${\cal F}$ is the free energy (\ref{H04}), $D$ is a diffusion coefficient, and ${\bbox\xi}_j$ are Langevin forces with the correlation function \begin{equation} \langle\xi_{i,\alpha}(t_1)\xi_{j,\beta}(t_2)\rangle= 2D\delta_{ij}\delta_{\alpha\beta}\delta(t_1-t_2) \,. \label{H52} \end{equation} The diffusion coefficient $D$ determines mobility of the defects. We believe that the equation (\ref{H51}) is applicable to the dynamics of disclinations in hexatic films like membranes, freely suspended films and Langmuir films. The equation for the vortices in superfluid films is a bit more complicated. It is written in Section \ref{helium} where the correlation functions of the vorticity are analyzed. The equations (\ref{H51},\ref{H52}) describe trajectories of separate defects. We should also take into account annihilation and creation processes. Remember that we neglect defects with $|n_j|>1$. Next, processes where a number of defect-antidefect pairs are created at the same point are suppressed since probability of such events is small due to the energy associated with the cores of defects. Then we have to take into account the creation processes of single pairs solely, they are characterized by the creation rate $\bar R(r)$ which is a probability density for a defect-antidefect pair with the separation $r$ to be created per unit time per unit area. The annihilation processes have to be characterized by the annihilation rate $R(r)$ which is a probability for a defect-antidefect pair to annihilate per unit time if the pair is separated by the distance $r$. Really, both $\bar R(r)$ and $R(r)$ are nonzero only if $r$ is of the order of the core size $a$. Let us introduce the integrals \begin{equation} \bar\lambda=\int{\rm d}^2 r\, \bar R(r) \,, \qquad \lambda=\int{\rm d}^2 r\, R(r) \,. \label{rate} \end{equation} Here, the creation constant $\bar\lambda$ is a probability for a defect-antidefect pair to be created per unit time per unit area and $\lambda$ is a constant having the same dimensionality as the diffusion coefficient $D$. Below, the diffusion coefficient $D$ is put to unity by rescaling time. Then the annihilation constant $\lambda$ is a dimensionless parameter of order unity and the creation constant $\bar\lambda$ can be estimated as \begin{equation} \bar\lambda\sim a^{-4} \exp(-2\mu/T) \,, \label{fuga} \end{equation} which is the second power of the defect concentration. As in statics, in dynamics the system of defects can be described in terms of probability distribution functions. In our case, when the total charge of the system is zero, only even probability densities ${\cal P}_{2N}(t,{\bbox x}_1,\dots,{\bbox x}_N,{\bbox z}_1,\dots,{\bbox z}_N)$ are non-zero, where ${\bbox x}_j$ and ${\bbox z}_j$ are positions of the defects and of the antidefects correspondingly. The total probability should be equal to unity which gives the normalization condition \begin{eqnarray} && \sum\limits_{N=0}^\infty \frac{1}{(N!)^2} \int{\rm d}^2x_1\dots{\rm d}^2x_N\, {\rm d}^2z_1\dots{\rm d}^2z_N\, \nonumber \\ && \times {\cal P}_{2N}(t,{\bbox x}_1,\dots,{\bbox x}_N, {\bbox z}_1,\dots,{\bbox z}_N)=1 \,. \label{norm} \end{eqnarray} Starting from the equation (\ref{H51}) and taking into account the creation and annihilation processes one can derive a system of master equations for the probability densities \end{multicols} \begin{eqnarray} && \frac{\partial{\cal P}_{2N}}{\partial t} =\sum\limits_j \left\{\frac{\partial}{\partial{\bbox x}_j} \left[\frac{1}{T}\frac{\partial{\cal F}}{\partial{\bbox x}_j} {\cal P}_{2N}\right] +\frac{\partial}{\partial{\bbox z}_j} \left[\frac{1}{T}\frac{\partial{\cal F}}{\partial{\bbox z}_j} {\cal P}_{2N}\right]\right\} +\sum\limits_j\left[\frac{\partial^2}{\partial{\bbox x}_j^2} +\frac{\partial^2}{\partial{\bbox z}_j^2}\right]{\cal P}_{2N} \nonumber \\ && -\sum\limits_{j,k}R({\bbox x}_j-{\bbox z}_k){\cal P}_{2N} +\int{\rm d}^2 x\,{\rm d}^2 z\,R({\bbox x}-{\bbox z}) {\cal P}_{2N+2}({\bbox x}_1,\dots,{\bbox x}_N,{\bbox x}, {\bbox z}_1,\dots,{\bbox z}_N,{\bbox z}) \label{master} \\ && +\sum\limits_{j,k}\bar R({\bbox x}_j-{\bbox z}_k) {\cal P}_{2N-2}({\bbox x}_1,\dots,{\bbox x}_{j-1}, {\bbox x}_{j+1},\dots,{\bbox x}_N, {\bbox z}_1,\dots,{\bbox z}_{k-1},{\bbox z}_{k+1},\dots,{\bbox z}_N) -\bar\lambda A {\cal P}_{2N} \,, \nonumber \end{eqnarray} \begin{multicols}{2} \noindent where $A$ is the area of the film. The Gibbs distribution (\ref{parti}) must be a solution of the master equations (\ref{master}). The condition imposes the following constraint on the creation and the annihilation rates \begin{eqnarray} && \bar R(r)=\frac{y_0^2}{a^4} \left(\frac{a}{r}\right)^{2\beta} R(r) \,, \label{fdt} \end{eqnarray} where we imply $r>a$. The constraint (\ref{fdt}) can be treated as the manifestation of the equilibrium state of the thermal bath which in our case is related to short-scale fluctuations. Thus the constraint (\ref{fdt}) has the same origin as Eq. (\ref{H52}). At deriving Eq. (\ref{fdt}) we assumed that the separation $|{\bbox x}-{\bbox z}|$ in the argument of the rate $R$ or $\bar R$ is smaller than separations between ${\bbox x}$ or ${\bbox z}$ and other points. That is accounted for the small characteristic value of the separation which are of the order of the core radius $a$. Note that for such $r$ the factor $y_0(r/a)^{2-\beta}$ entering Eq. (\ref{fdt}) can be treated as the renormalized value $y$ of the fugacity as follows from Eq. (\ref{rg}). Principally, the master equations (\ref{master}) enable one to find conditional probability densities, related to different time moments. Consequently, starting from the equations one can examine non-simultaneous correlations in the system. However, there are terms in the master equations (associated with the creation and annihilation processes) mixing the probability densities ${\cal P}_{2N}$ with different $N$. That makes the master equations hardly useful, that is a motivation to look for some more suitable technique. Such a technique was developed by Doi, Ref. \cite{76Doi}, we formulate it in the subsequent subsection. \subsection{Quantum Field Formulation} The Doi technique \cite{76Doi} enables one to treat systems of classical particles where creation and annihilation processes occur. The main idea introduced by Doi is that correlation functions of different quantities characterizing the particles can be written in the form close to the one known in the quantum field theory. Of course there are some peculiarities related to the fact that for classical particles one should deal directly with probabilities whereas in the quantum field theory one starts from the scattering matrix. Nevertheless the Doi technique enables, say, to formulate a diagrammatic expansion with the conventional rules. The technique was originally developed to describe systems of molecules involved into chemical reactions. But it is definitely applicable also to the system of point defects. The Doi technique is formulated in terms of the creation $\hat\psi$ and annihilation $\psi$ operators which satisfy the same commutation rules as ones for Bose-particles \begin{eqnarray} && [\psi({\bbox r}_1),\hat\psi({\bbox r}_2)]= \delta({\bbox r}_1-{\bbox r}_2)\,, \nonumber \\ && [\hat\psi({\bbox r}_1),\hat\psi({\bbox r}_2)]= [\psi({\bbox r}_1),\psi({\bbox r}_2)]=0 \,. \label{do01} \end{eqnarray} For our system of defects we should introduce annihilation and creation operators $\psi_\pm$ and $\hat\psi_\pm$ where the subscripts $+$ and $-$ label fields related to the defects and to the antidefects. The state of the system at a time moment $t$ can be written in terms of a ``quantum'' state \begin{eqnarray} && |t\rangle=\sum\limits_{N=0}^\infty\frac{1}{(N!)^2} \int{\rm d}^2\,x_1\dots{\rm d}^2\, x_N\, {\rm d}^2\,z_1\dots{\rm d}^2\, z_N\,{\cal P}_{2N} \nonumber \\ && \times \hat\psi_+({\bbox x}_1)\dots \hat\psi_+({\bbox x}_N) \hat\psi_-({\bbox z}_1)\dots \hat\psi_-({\bbox z}_N)|0\rangle \,, \label{do02} \end{eqnarray} where ${\cal P}_{2N}$ are the probability densities introduced above and $|0\rangle$ designates the vacuum state: $\psi_\pm|0\rangle=0$. In accordance with the expression (\ref{do02}) an evolution of the quantum state $|t\rangle$ is determined by the master equations. The evolution equation can be written as \begin{eqnarray} && \partial_t|t\rangle=-{\cal H}|t\rangle \,, \quad {\rm and} \quad |t_2\rangle=\exp\left[(t_1-t_2){\cal H}\right]|t_1\rangle \,, \label{do03} \end{eqnarray} where ${\cal H}$ is an operator expressed in terms of the fields $\psi_\pm$ and $\hat\psi_\pm$. By analogy with the quantum field formulation it can be called the Hamiltonian operator or simply the Hamiltonian. The total probability must be equal to unity, which leads to Eq. (\ref{norm}). The condition can be written in terms of the quantum state $|t\rangle$ as \begin{eqnarray} && \langle{\rm sum}|t\rangle=1 \,, \label{sum} \end{eqnarray} where \begin{eqnarray} && \langle{\rm sum}|=\langle 0| \exp\left[\int{\rm d}^2r\, \left(\psi_++\psi_-\right)\right] \,. \nonumber \end{eqnarray} Note the following identity \begin{eqnarray} && \langle{\rm sum}|\hat\psi_\pm({\bbox r})=\langle{\rm sum}|\,, \label{prop} \end{eqnarray} which can be easily checked using the commutation rules (\ref{do01}) and the equality $\langle 0|\hat\psi_\pm=0$. Since the evolution must conserve the total probability $\langle{\rm sum}|t\rangle$ the relations \begin{eqnarray} && \langle{\rm sum}|{\cal H}=0\,, \quad {\rm and} \quad \langle{\rm sum}|\exp\left(\tau{\cal H}\right) =\langle{\rm sum}| \,, \label{do04} \end{eqnarray} have to be satisfied, where $\tau$ is an arbitrary parameter. Quantities characterizing the system can be represented by corresponding operators, see Ref. \cite{76Doi}. Say, the operator of the charge density is \begin{eqnarray} && \tilde\rho=\hat\psi_+\psi_+-\hat\psi_-\psi_- \,. \label{do3} \end{eqnarray} If $\tilde A$ is such operator corresponding to a quantity $A$, then an average value of the quantity at a time moment $t$ can be expressed as \begin{eqnarray} && \langle A(t) \rangle= \langle{\rm sum}|\tilde A |t\rangle \,. \label{do05} \end{eqnarray} Note that \begin{eqnarray} && \langle\hat\psi_\pm\rangle=\langle{\rm sum}| \hat\psi_\pm|t\rangle=1 \,, \label{prop1} \end{eqnarray} which is a consequence of Eqs. (\ref{sum},\ref{prop}). The relation (\ref{prop1}) shows that it is natural to shift the creation operators $\hat\psi_\pm$ introducing new variables (see Ref. \cite{80GS}) \begin{eqnarray} && \hat\psi_\pm= 1+\bar\psi_\pm \,, \qquad \langle\bar\psi_\pm\rangle=0 \,. \label{shift} \end{eqnarray} Correlation functions of different quantities can be presented analogously to Eq. (\ref{do05}). For example, the pair correlation function of two quantities $A(t_1)$ and $B(t_2)$ (we assume $t_2>t_1$) can be written as \begin{eqnarray} && \langle B(t_2)A(t_1)\rangle= \langle{\rm sum}|\tilde B\exp\left[({t_1}-{t_2}) {\cal H}\right]\tilde A|t_1\rangle \,. \label{do1} \end{eqnarray} Using the relations (\ref{do03},\ref{do04}) we can rewrite the expression (\ref{do1}) as \begin{eqnarray} && \langle B(t_2)A(t_1)\rangle= \langle{\rm sum}|\tilde B(t_2)\tilde A(t_1) |{\rm in}\rangle \,, \label{do06} \end{eqnarray} where $\tilde A(t),\tilde B(t)$ in Eq. (\ref{do06}) are operators in the Heisenberg representation \begin{eqnarray} && \tilde A(t)=\exp\left[-({t_{\rm f}}-{t}){\cal H}\right] \tilde A \exp\left[-({t}-t_{\rm in}){\cal H}\right] \,, \label{heis} \end{eqnarray} satisfying the equation $\partial_t \tilde A(t)=\left[{\cal H},\tilde A(t)\right]$. In Eq. (\ref{do06}) $|{\rm in}\rangle$ is an initial state (realized at a time moment $t_{\rm in}$) and $t_{\rm f}$ is a ``final'' time, so that $t_{\rm f}>t_2>t_1>t_{\rm in}$. The expressions like (\ref{do06}) enable one to reformulate the problem of calculating correlation functions in terms of a functional integral, see Ref. \cite{85Pel}. Namely, we can write \begin{eqnarray} && \langle A_1(t_1)\dots A_n(t_n)\rangle =\int {\cal D}\hat\psi_\pm {\cal D}\psi_\pm \tilde A_1\dots \tilde A_n \nonumber \\ && \times\exp\biggl\{-\int_{-\infty}^{t_{\rm f}}{\rm d}t\, \left[{\cal H} +\int {\rm d}^2 r\,\left(\hat\psi_+\partial_t\psi_+ +\hat\psi_-\partial_t\psi_-\right)\right] \nonumber \\ && +\int {\rm d}^2 r\,\left[\psi_+(t_{\rm f},{\bbox r}) +\psi_-(t_{\rm f},{\bbox r})\right] \biggr\} \,, \label{path1} \end{eqnarray} where $\psi_\pm, \hat\psi_\pm$ are to be interpreted as functions of $t$ and ${\bbox r}$. We assume that $t_{\rm f}>t_1,\dots,t_n$ in Eq. (\ref{path1}). Deriving the expression one has taken the limit $t_{\rm in}\to-\infty$ and assumed $|{\rm in}\rangle=|0\rangle$. Because of the creation processes the vacuum has to be turned into a stationary state during the infinite time. To ensure convergence of the functional integral (\ref{path1}) the integration contour over the field $\hat\psi$ should go parallel to the imaginary axis. Note that the shift (\ref{shift}) kills the boundary term $\int{\rm d}^2r\,(\psi_++\psi_-)$: It is cancelled by a contribution originating from the derivatives $\partial_t\psi_\pm$ after integrating over time. Then we come to a conventional representation of the correlation functions in terms of a functional integral \begin{eqnarray} && \langle A_1(t_1)\dots A_n(t_n)\rangle =\int {\cal D}\bar\psi_\pm {\cal D}\psi_\pm \exp\biggl\{-\int{\rm d}t\, \biggl[{\cal H} \nonumber \\ && +\int {\rm d}^2 r\,\left(\bar\psi_+\partial_t\psi_+ +\bar\psi_-\partial_t\psi_-\right)\biggr]\biggr\} \tilde A_1\dots \tilde A_n \,. \label{path} \end{eqnarray} The relation (\ref{path1}) or (\ref{path}) is a convenient starting point for treating a system of particles involved into the creation and annihilation processes. \section{Diagrammatic Representation} \label{quantum} Below, we apply the Doi technique to our particular problem. The explicit expression for the Hamiltonian determining the evolution of the defect system in accordance with Eq. (\ref{do03}) can be found from the master equations (\ref{master}). Comparing the equations with Eqs. (\ref{do02},\ref{do03}) we get \begin{eqnarray} && {\cal H}={\cal H}_0+{\cal H}_R+{\cal H}_\beta \,. \label{ham} \end{eqnarray} The explicit expressions for the terms entering Eq. (\ref{ham}) are \begin{eqnarray} && {\cal H}_0=\int{\rm d}^2 r\,\left( \nabla\hat\psi_+\nabla\psi_+ +\nabla\hat\psi_-\nabla\psi_-\right) \label{ham0} \end{eqnarray} \begin{eqnarray} && {\cal H}_R=-\int{\rm d}^2 r_1\,{\rm d}^2 r_2\, \biggl[\bar R({\bbox r}_1-{\bbox r}_2) (\hat\psi_{+,1}\hat\psi_{-,2}-1) \nonumber \\ && +R({\bbox r}_1-{\bbox r}_2) (\psi_{+,1}\psi_{-,2} -\hat\psi_{+,1}\hat\psi_{-,2}\psi_{+,1}\psi_{-,2})\biggr] \label{hamr} \end{eqnarray} \begin{eqnarray} && {\cal H}_\beta=2\beta\int {\rm d}^2 r_1\,{\rm d}^2 r_2\, \left(\nabla\hat\psi_{+,1}\hat\psi_{-,2} -\hat\psi_{+,1}\nabla\hat\psi_{-,2}\right) \nonumber \\ && \times \frac{{\bbox r}_1-{\bbox r}_2}{|{\bbox r}_1-{\bbox r}_2|^2} \psi_{+,1}\psi_{-,2} \nonumber \\ && -2\beta\int {\rm d}^2 r_1\,{\rm d}^2 r_2\,\biggl[ \nabla\hat\psi_{+,1}\hat\psi_{+,2} \frac{{\bbox r}_1-{\bbox r}_2}{|{\bbox r}_1-{\bbox r}_2|^2} \psi_{+,1}\psi_{+,2} \nonumber \\ && +\nabla\hat\psi_{-,1}\hat\psi_{-,2} \frac{{\bbox r}_1-{\bbox r}_2}{|{\bbox r}_1-{\bbox r}_2|^2} \psi_{-,1}\psi_{-,2}\biggr] \,, \label{hamb} \end{eqnarray} where $\psi_{+,1}=\psi_+(t,{\bbox r}_1)$ and so further. The diffusive contribution (\ref{ham0}) is related to the Langevin forces, in Eq. (\ref{hamr}) $R$ is the annihilation rate and $\bar R$ is the creation rate for the defect-antidefect pairs (the quantities were introduced in Section \ref{basic}), and the term (\ref{hamb}) describe the Coulomb interaction. Using the property (\ref{prop}), one can easily check the conditions (\ref{do04}) for all contributions (\ref{ham0}-\ref{hamb}). Performing the substitution (\ref{shift}) we can express the Hamiltonian (\ref{ham}) in terms of the fields $\bar\psi_\pm$. Note that the shift (\ref{shift}) kills terms of the second order proportional to $\lambda$, $\bar\lambda$ and generates additional third-order vertices. Of course one can work in both representations. It is more convenient for us to use Eqs. (\ref{path1},\ref{ham}). We can easily convince ourselves that odd correlation functions of the charge density (\ref{do3}) are zero. Indeed, the exponent in Eq. (\ref{path1}) is invariant under permuting $\psi_+\leftrightarrow\psi_-$, $\hat\psi_+\leftrightarrow\hat\psi_-$, whereas the charge density changes its sign at the permutation. The constraint (\ref{Ch1}) shows that the symmetry is not spontaneously broken what could lead to non-zero odd correlation functions. It follows from Eq. (\ref{heis}) that the commutator $[{\cal H},\tilde\rho]$ should be equal to $-\nabla{\bbox j}$ where ${\bbox j}$ is the current density operator. Calculating the commutator with Eqs. (\ref{do3},\ref{ham}) we get \begin{eqnarray} && {\bbox j}=\nabla\hat\psi_+\psi_+-\hat\psi_+\nabla\psi_+ -\nabla\hat\psi_-\psi_-+\hat\psi_-\nabla\psi_- \nonumber \\ && -(\hat\psi_+\psi_++\hat\psi_-\psi_-)\nabla\phi \,, \label{curr} \end{eqnarray} where $\phi$ is an ``electrostatic potential'' \begin{eqnarray} && \phi({\bbox r})=-2\beta\int{\rm d}^2 x\, \ln\left(\frac{|{\bbox r}-{\bbox x}|}{a}\right) \tilde\rho({\bbox x}) \,. \label{poten} \end{eqnarray} Principally, besides the ``internal'' potential (\ref{poten}) an ``external'' potential $\phi_{\rm ext}$ can be imposed onto the system, satisfying the equation $\nabla^2\phi_{\rm ext}=0$. Then an ``external force'' should be added to the right-hand side of the equation (\ref{H51}). The force generates an ``external'' contribution to the Hamiltonian \begin{eqnarray} && {\cal H}_{\rm ext}=\int{\rm d}^2 r\,\left( \nabla\hat\psi_+\psi_+ -\nabla\hat\psi_-\psi_-\right)\nabla\phi_{\rm ext} \,. \label{ext} \end{eqnarray} The expression (\ref{curr}) for the current density has also to be corrected by substituting $\phi\to\phi+\phi_{\rm ext}$. With the term (\ref{ext}) we can examine susceptibilities describing a response of the system to the external influence. Substituting the expression (\ref{ham}) into (\ref{path1}) or (\ref{path}) and expanding the exponent over ${\cal H}_R$ and ${\cal H}_\beta$ one can obtain a conventional perturbation series for calculating different correlation functions of $\psi,\hat\psi$. The series is an expansion over $R$, $\bar R$ and $\beta$ in terms of the conventional diffusion propagators: \begin{eqnarray} && G(t,{\bbox r})=\langle \psi_+(t,{\bbox r})\hat\psi_+(0,0)\rangle_0 \nonumber \\ && = \langle \psi_-(t,{\bbox r})\hat\psi_-(0,0)\rangle_0 =\frac{\theta(t)}{4\pi t}\exp\left(-\frac{r^2}{4t}\right) \,, \label{diff} \end{eqnarray} where $\theta(t)$ is the step function. However, effects related to the Coulomb interaction and to the annihilation processes are not weak. Therefore one must take into account the Coulomb interaction and the annihilation processes exactly. By other words, at calculating the correlation functions one must consider the complete series over $\beta$ and $R$. Fortunately, the the expansion over $\bar R$ is equivalent to an expansion over the fugacity $y$ which is assumed to be a small parameter. Therefore we can take only principal terms in the expansion over $\bar R$. The perturbation expansion can be formulated as a diagrammatic series. We develop the diagrammatic technique starting from the representation (\ref{path1}), pushing the final time $t_{\rm f}$ to the far future. We depict the propagator (\ref{diff}) by a line directed from $\hat\psi$ to $\psi$. The term with the creation rate $\bar R$ in Eq. (\ref{ham}) generates vertices where two propagator lines start, the vertices correspond to the defect-antidefect creation processes. The Coulomb term in Eq. (\ref{ham}) generates two-point objects which we will designate by dashed lines, such line describes the Coulomb interaction of defects located in points connected by the line. And the term proportional to the annihilation rate $R$ in Eq. (\ref{ham}) produces two types of vertices. First, it produces vortices where two propagator lines finish, that corresponds to an annihilation process. Second, it produces fourth-order vertices which correspond to an effective interaction related to a finite probability for a defect-antidefect pair to annihilate, see Ref. \cite{76Doi}. A typical diagram block is presented in Fig. \ref{diag}. The block is drawn in real ${\bbox r}-t$ space-time. The curves constituted of the propagator lines can be interpreted as trajectories of defects and antidefects. Due to causality the particles always move forward in time. Note that the dashed lines corresponding to the Coulomb interaction are perpendicular to the $t$-axis since the interaction is simultaneous. \subsection{Pair Conditional Probability} In the subsection we examine an auxiliary object which will be needed for us at intermediate stages of subsequent calculations. The object is the following correlation function \begin{eqnarray} && M(t_2-t_1,{\bbox r}_1,{\bbox r}_2,{\bbox r}_3,{\bbox r}_4) \nonumber \\ && =\langle \psi_+(t_2,{\bbox r}_1)\psi_-(t_2,{\bbox r}_2) \hat\psi_+(t_1,{\bbox r}_3)\hat\psi_-(t_1,{\bbox r}_4)\rangle \,. \label{mmm} \end{eqnarray} For a stationary case the average (\ref{mmm}) depends on the difference $t=t_2-t_1$ only. Due to causality $M$ is equal to zero provided $t<0$. The quantity (\ref{mmm}) can be interpreted as a probability density to find a defect and an antidefect at the time moment $t_2$ in the points ${\bbox r}_1$ and ${\bbox r}_2$ provided they were located in the points ${\bbox r}_3$ and ${\bbox r}_4$ at the time moment $t_1$. It can be considered also as a two-particle matrix element of the evolution operator $\exp[-(t_2-t_1){\cal H}]$. As we explained above, the perturbation series in terms of the creation rate $\bar R$ is an expansion over a small parameter. Here we examine the principal contribution to the conditional probability (\ref{mmm}) which is of the zero order over $\bar R$. Then the average (\ref{mmm}) can be represented as a series of diagrams of the type depicted in Fig. \ref{diag1}. One can interpret the picture as trajectories of a defect and of an antidefect which are driven by the Langevin forces, and are influenced the Coulomb interaction (dashed lines) and the effective interaction associated with the annihilation processes (point vertex). Note that in this approximation direct annihilation events do not contribute to the conditional probability (\ref{mmm}) since they would lead to terminating the lines in the diagrams. It is of crucial importance that both the Coulomb interaction and the effective interaction associated with the annihilation processes are local in time. Therefore all the diagrams representing the conditional probability (\ref{mmm}) are ladder diagrams, like in Fig. \ref{diag1}. Summing up the ladder sequence we get an equation for $M$ which can be written in the differential form \begin{eqnarray} && \partial_t M=\left(\nabla_1^2+\nabla_2^2\right)M +2\beta\left(\nabla_1-\nabla_2\right)\left[ \frac{{\bbox r}_1-{\bbox r}_2}{|{\bbox r}_1-{\bbox r}_2|^2}M\right] \nonumber \\ && -R({\bbox r}_1-{\bbox r}_2)M +\delta(t)\delta\left({\bbox r}_1-{\bbox r}_3\right) \delta\left({\bbox r}_2-{\bbox r}_4\right) \,. \label{MM} \end{eqnarray} Since $M=0$ at $t<0$ we conclude from Eq. (\ref{MM}) that at $t\to+0$ \begin{equation} M(t,{\bbox r}_1,{\bbox r}_2,{\bbox r}_3,{\bbox r}_4) \to \delta\left({\bbox r}_1-{\bbox r}_3\right) \delta\left({\bbox r}_2-{\bbox r}_4\right) \,. \label{small} \end{equation} The total probability to find the defect-antidefect pair is determined by the integral $\int {\rm d}^2 r_1\,{\rm d}^2 r_2\,M$. Let us calculate the time derivative of this integral substituting $\partial_t M$ by the right-hand side of Eq. (\ref{MM}). Then the first two terms will give zero contributions (since they are total derivatives) and only the term with $R$ will produce a non-zero (negative) contribution. It is quite natural since the Langevin forces and the Coulomb interaction cannot change the total probability whereas the annihilation processes diminish it. Note that all the terms in the right-hand side of Eq. (\ref{MM}) proportional to $M$ have the same dimensionality. Therefore one could expect a simple scaling behavior when $t$ scales as $r^2$. The subsequent calculations confirm the expectation. In terms of the variables \begin{eqnarray} && {\bbox r}={\bbox r}_1-{\bbox r}_2 \,, \qquad {\bbox\varrho}=\frac{{\bbox r}_1+{\bbox r}_2}{2}\,, \qquad {\bbox r}_0={\bbox r}_3-{\bbox r}_4 \,, \label{J3} \end{eqnarray} the equation (\ref{MM}) for $M$ is rewritten as \begin{eqnarray} && \partial_t M=\left(\frac{1}{2}\nabla_{\varrho}^2 +2\nabla_r^2+4\beta\nabla_r\frac{\bbox r}{r^2}-R({\bbox r})\right)M \nonumber \\ && +\delta(t)\delta\left({\bbox r}-{\bbox r}_0\right) \delta\left({\bbox\varrho}-{\bbox r}_3/2-{\bbox r}_4/2\right) \,. \nonumber \end{eqnarray} We see that the differential operator in the right-hand side of the equation falls into two parts depending on ${\bbox\varrho}$ and ${\bbox r}$ only and that the ``source'' is a product of $\delta$-functions of the same variables. Therefore the solution of the equation can be written in a multiplicative form \begin{eqnarray} && M=\frac{1}{2\pi t} \exp\left\{-\frac{\left(2{\bbox\varrho}-{\bbox r}_3 -{\bbox r}_4\right)^2}{8t}\right\} S(t,{\bbox r},{\bbox r}_0) \,, \label{H65} \end{eqnarray} the function $S$ satisfies the following equation \begin{eqnarray} && \partial_t S=2\nabla^2S+4\beta\nabla\left(\frac{\bbox r}{r^2}S\right) \nonumber \\ && -R({r})S+\delta(t)\delta\left({\bbox r}-{\bbox r}_0\right) \,. \label{doo} \end{eqnarray} Note that \begin{eqnarray} && {\rm if} \quad t\to+0 \quad {\rm then} \quad S\to \delta\left({\bbox r}-{\bbox r}_0\right) \,. \label{small1} \end{eqnarray} The relation follows from Eq. (\ref{doo}) and causality (leading to $S=0$ for negative $t$). In accordance with Eq. (\ref{H65}), a motion of the mass center and the relative motion of the defects are separated. The motion of the mass center is purely diffusive whereas the relative motion is strongly influenced by the interaction. The function $S$ can be treated as the probability density for the relative motion of the defect-antidefect pair. It is natural to expand the function into the Fourier series over the angle $\varphi$ between the vectors ${\bbox r}$ and ${\bbox r}_0$: \begin{eqnarray} && S(t,{\bbox r},{\bbox r}_0) =\sum\limits_{-\infty}^{+\infty}S_m(t,r,r_0)\exp(im\varphi) \,. \label{angle} \end{eqnarray} Motions corresponding to different angular harmonics are separated. In terms of the angular harmonics Eq. (\ref{doo}) is rewritten as \begin{eqnarray} && \frac{1}{2}\partial_t S_m =\left[\partial_r^2+(1+2\beta)\frac{1}{r}\partial_r -\frac{m^2}{r^2}\right]S_m \nonumber \\ && -\frac{1}{2}R(r)S_m +\frac{1}{4\pi r_0}\delta(t)\delta(r-r_0) \,. \label{four} \end{eqnarray} It is possible to get equations for $S$ analogous to Eqs. (\ref{doo},\ref{four}) in terms of ${\bbox r}_0$. They have practically the same form as Eqs. (\ref{doo},\ref{four}). The only difference is in the sign of $\beta$ which is opposite. That leads to the relation \begin{eqnarray} && S_m(t,r,r_0)=\left(\frac{r_0}{r}\right)^{2\beta} S_m(t,r_0,r) \,. \label{rr0} \end{eqnarray} Let us stress that the relation (\ref{rr0}) is correct for an arbitrary function $R(r)$. Consider a behavior of the angular harmonics $S_m(t,r,r_0)$ at small $r$. More precisely, we assume $t\gg a^2$ and examine the region $\sqrt t\gg r\gg a$. Then it is possible to use the equation (\ref{four}) with the time derivative and the annihilation term neglected. As a result we get \begin{eqnarray} && S_m=C_{1,m}r^{\nu-\beta} +C_{2,m}r^{\nu-\beta}(r/a)^{-2\nu} \,, \label{att} \\ && \nu=\sqrt{\beta^2+m^2} \,, \label{att1} \end{eqnarray} where $C_{1,m},C_{2,m}$ are some factors dependent on $t$ and $r_0$. The ratio of the factors is determined by a concrete $r$-dependence of the annihilation rate $R$, one can assert only that $C_{1m}$ and $C_{2m}$ are of the same order. Therefore, if we consider the behavior of the function $S$ for $r\gg a$, then the second term in the right-hand side of Eq. (\ref{att}) can be neglected. By other words, being interested in the scales $r\gg{a}$, we can solve the equation (\ref{four}) neglecting the annihilation term and requiring a finite value of $S_m$ at $r\to 0$ instead. The requirement can be treated as the boundary condition for $S_m$ at small $r$. The other boundary condition is that $S_m$ tend to zero at $r\to\infty$. The equations for $S_m$ with the boundary conditions are solved in Appendix \ref{condi}. We present here only the answer \begin{equation} S_m(t,r,r_0)=\frac{1}{8\pi t}\left(\frac{r_0}{r}\right)^\beta \exp\left(-\frac{r^2+r_0^2}{8t}\right) I_\nu\left(\frac{rr_0}{4t}\right) \,, \label{H69} \end{equation} where $I$ is the modified Bessel function and $\nu$ is introduced by Eq. (\ref{att1}). Remind that the expression is correct provided $r,r_0,\sqrt t\gg a$. Extracting from Eq. (\ref{H69}) an asymptotics at small $r$ we get \begin{eqnarray} && C_{1,m}=\frac{r_0^\beta}{8\pi\Gamma(1+\nu)t} \left(\frac{r_0}{8t}\right)^\nu \exp\left(-\frac{r_0^2}{8t}\right) \,. \nonumber \end{eqnarray} Note that the Coulomb term in Eq. (\ref{doo}) produces a probability flux to the origin. To find it we should integrate the equation (\ref{doo}) over a disk of a radius $a\ll r\ll\sqrt t$ centered at the origin and single out the contribution to $\partial_t\int{\rm d}^2r S$ associated with the Coulomn term. Then we find the flux $\lambda_{\rm r}C_{1,0}$ where \begin{eqnarray} && \lambda_{\rm r}=8\pi\beta \,. \label{rrr} \end{eqnarray} One can treat the quantity (\ref{rrr}) as the renormalized (``dressed'') value of the annihilation constant. Now we understand why the solution (\ref{H69}) (realized at $r\gg{a}$) is insensitive to a particular form of the annihilation rate. The probability for a defect-antidefect pair with the separation $r\gg a$ to annihilate is determined by the Coulomb attraction. And only the behavior of the probability density at $r\sim{a}$ is sensitive to the particular form of the annihilation rate $R(r)$: The coefficients $C_{2,m}$ in Eq. (\ref{att}) are positive if $\lambda<\lambda_{\rm r}$ and are negative if $\lambda>\lambda_{\rm r}$. Returning to the conditional probability (\ref{mmm}) we obtain from Eqs. (\ref{H65},\ref{angle},\ref{H69}) \begin{eqnarray} && M=\frac{1}{(4\pi t)^2}\left(\frac{r_0}{r}\right)^\beta \exp\left\{-\frac{\left({\bbox r}_1+{\bbox r}_2-{\bbox r}_3 -{\bbox r}_4\right)^2}{8t}\right\} \nonumber \\ && \times\exp\left(-\frac{r^2+r_0^2}{8t}\right) \sum\limits_{m=-\infty}^{+\infty}\exp(im\varphi)\ I_\nu\left(\frac{rr_0}{4t}\right)\,, \label{H70} \end{eqnarray} where $\nu$ is introduced by Eq. (\ref{att1}). One can easily check that the expression (\ref{H70}) is reduced to Eq. (\ref{small}) at $t\to+0$. It is possible to calculate an explicit expression for the total probability to find a defect and an antidefect in any points at a fixed time separation $t$, see Appendix \ref{condi}. The asymptotic behavior of the expression (\ref{H70}) at the condition $rr_0/t\gg1$ is examined in Appendix \ref{asympt}, see Eq. (\ref{WW8}). If both $r,r_0$ are much greater than $\sqrt t$, it can be written as \begin{eqnarray} && M\approx\frac{1}{(4\pi t)^2} \exp\left\{-\frac{({\bbox r}-{\bbox r}_0)^2}{8t}\right\} \nonumber \\ && \times \exp\left\{-\frac{\left({\bbox r}_1+{\bbox r}_2-{\bbox r}_3 -{\bbox r}_4\right)^2}{8t}\right\} \,. \label{W8} \end{eqnarray} The answer is quite natural. The characteristic values of the separations ${\bbox r}_1-{\bbox r}_3$ and of ${\bbox r}_2-{\bbox r}_4$ are of the order of $\sqrt t$ and are consequently much smaller than $|{\bbox r}_1-{\bbox r}_2|$ (or $|{\bbox r}_3-{\bbox r}_4|$). Then $r\approx r_0$ and it is possible to neglect all terms, containing $|{\bbox r}_1-{\bbox r}_2|$ in the denominators, in the equation (\ref{MM}). Thus we come to a purely diffusive equation leading to the asymptotic law (\ref{W8}). The consideration presented above can be generalized for the case when an ``external electrostatic potential'' $\phi_{\rm ext}$ is imposed onto the system. Its influence is described by the contribution (\ref{ext}) to the Hamiltonian. Performing the same procedure as above we get a modified equation for the correlation function (\ref{mmm}) \begin{eqnarray} && \frac{\partial}{\partial t_2}M =\left(\nabla_1^2+\nabla_2^2\right)M +4\beta\nabla_r \left(\frac{{\bbox r}}{r^2}M\right)-R(r)M \nonumber \\ && +\nabla\phi_{\rm ext}(t_2,{\bbox r}_1)\nabla_1M -\nabla\phi_{\rm ext}(t_2,{\bbox r}_2)\nabla_2M \nonumber \\ && +\delta(t_2-t_1)\delta\left({\bbox r}_1-{\bbox r}_3\right) \delta\left({\bbox r}_2-{\bbox r}_4\right) \,. \label{MM1} \end{eqnarray} The expression (\ref{MM1}) shows that the gradient of the external potential has to be added to the gradient of the internal one. Of course, in the presence of the external field the correlation function (\ref{mmm}) depends on both time moments $t_1$ and $t_2$. Note that the operator in the right-hand side of Eq. (\ref{MM1}) is the same as that for the Fokker-Plank equation formulated in Ref. \cite{Dynamics1} (excluding for the annihilation term). \section{Renormalization} \label{renorm} In this section we are going to discuss effects related to high-order terms over the creation rate $\bar R$. The effects are relevant only near the transition point where $\beta$ is close to $2$. Then the influence of small-scale defect-antidefect pairs on larger scales becomes essential. In the situation the most natural language is the renormalization group approach. One can formulate a renormalization group procedure in the spirit of Kosterlitz, Ref. \cite{74Kos}. We will single out blocks corresponding to small separations of the pairs and treat them as renormalized quantities entering the Hamiltonian (\ref{ham}). \subsection{Creation and annihilation rates} Sizes of the pairs are small near creation and near annihilation points. Here, we consider vicinities of the points. Then it is possible to neglect the interaction of the defect and of the antidefect with the environment. Thus we turn to the situation when only a single pair can be treated. If this is the case then one should analyze diagram blocks of the type drawn in Fig. \ref{diag2}. The left part of the figure corresponds to a vicinity of the creation occurring at a time moment $t_1$ and the right part of the figure corresponds to a vicinity of the annihilation occurring at a time moment $t_4$. Consider processes occurring during a time interval $\tau$ from the creation time $t_1$. One can separately treat a block corresponding to the time interval from $t_1$ till $t_2=t_1+\tau$. For the purpose we use the well-known property of the propagators (\ref{diff}) \begin{eqnarray} && G(s_3-s_1,{\bbox r})=\int{\rm d}^2 x\, G(s_3-s_2,{\bbox r}-{\bbox x}) \nonumber \\ && \times G(s_2-s_1,{\bbox x}) \,, \label{conv} \end{eqnarray} where $s_3>s_2>s_1$. For each diagram we extract propagators $G$ containing $t_2$ inside their time interval and represent the propagators like in Eq. (\ref{conv}) believing $s_2=t_2$. The procedure is reflected in Fig. \ref{diag2} where the dotted line represents a plane $t=t_2$ in the ${\bbox r}-t$ space-time and the integration in Eq. (\ref{conv}) corresponds to the integration in the plane. As a result, the block to the left of the plane is separated, it is characterized by the time separation $\tau$ and by two points ${\bbox r}_1$ and ${\bbox r}_2$ lying in the plane, the points are intersections of the plane with the trajectories of the particles. The block has to be inserted into more complicated objects via a convolution over ${\bbox r}_1$ and ${\bbox r}_2$. The same is true for the vicinity of the annihilation point also. Let us take a time moment $t_3$ separated by a time interval $\tau\gg a^2$ from an annihilation time $t_4$. Then it is possible to introduce the block which is a sum of the diagrams where the trajectories of the annihilating particles start from two given points ${\bbox r}_3$ and ${\bbox r}_4$ at $t=t_3$. The block has to be inserted into more complicated objects via a convolution over the points. In the vicinity of the annihilation point we can take into account the interaction of the annihilating defect-antidefect pair solely. That leads to the same ladder diagrams treated in Section \ref{quantum}. Therefore we can write an expression for the block without an additional analysis \begin{eqnarray} && R_\tau({r}_0)=\int{\rm d}^2 r_1\,{\rm d}^2 r_2\,R({\bbox r}) M(\tau,{\bbox r}_1,{\bbox r}_2,{\bbox r}_3,{\bbox r}_4) \nonumber \\ && =\int{\rm d}^2r\, S(\tau,{\bbox r},{\bbox r}_0)R({r}) \,. \label{dress} \end{eqnarray} Here ${\bbox r}$ and ${\bbox r}_0$ are defined by Eq. (\ref{J3}), $M$ is the conditional probability (\ref{mmm}), $S$ is the conditional probability for the relative motion of the defects, see Eq. (\ref{H65}), it is the solution of the equation (\ref{doo}). The physical meaning of the quantity $R_\tau({r}_0)$ is a distribution of the annihilating particles over the separation $r_0$ between the particles at the time moment $t_3$. It is natural to name this distribution ``dressed'' annihilation rate since the quantity determines a probability for the particles to annihilate after the time interval $\tau$. Note that all processes occurring on scales larger than $\sqrt\tau$ are sensitive only to this dressed quantity. Let us substitute into Eq. (\ref{dress}) the product $RS$ expressed from Eq. (\ref{doo}). The terms with the total derivatives give zero contribution to the integral over ${\bbox r}$ and we get \begin{eqnarray} && R_\tau({r}_0)=-\partial_\tau \int{\rm d}^2r\,S(\tau,{\bbox r},{\bbox r}_0) \,, \label{dres1} \end{eqnarray} Since $S(\tau)$ tends to zero at $\tau\to+\infty$ and is zero for negative $\tau$ we get from (\ref{small1},\ref{dres1}) \begin{eqnarray} && \int{\rm d}\tau\,R_\tau({r}_0) =1 \,. \label{unity} \end{eqnarray} The relation means that the total probability for a given pair to annihilate is equal to unity. As is seen from Eq. (\ref{att}) at the condition $\tau\gg a^2$ the main contribution to the integral in the right-hand side of Eq. (\ref{dres1}) is associated with the region $r\sim\sqrt\tau$ and therefore the contribution to the integral associated with the region $r\sim{a}$ is negligible. Therefore we can use the expression (\ref{angle}) with Eq. (\ref{H69}). Substituting the expression (\ref{dres2}) into Eq. (\ref{dres1}) we get a universal expression for the dressed quantity $R_\tau({r}_0)$ which is insensitive to the bare quantity $R(r)$. In Sec. \ref{quantum} we established the renormalized value (\ref{rrr}) of the annihilation constant $\lambda$. This analysis concerned the fourth-order interaction term written in Eq. (\ref{ham}). Below we demonstrate that the renormalized coefficient at the second-order annihilation term has the same value, independent of the bare one. In accordance with Eq. (\ref{rate}), to find the renormalized value $\lambda_r$ we should calculate the integral of $R_\tau({r}_0)$. As is demonstrated in Appendix \ref{condi} at $\tau\gg a^2$ the value of the integral is independent of $\tau$ and coincides with the value written in Eq. (\ref{rrr}), as one anticipated: \begin{eqnarray} && \lambda_{\rm r}=\int {\rm d}^2r_0\, R_\tau({r}_0) =8\pi\beta \,. \label{dres3} \end{eqnarray} The phenomenon resembles the renormalization of the reaction rate due to diffusion, see Refs. \cite{76Doi,86Pel}. Analogously, one can introduce the renormalized creation rate $\bar R_\tau(r)$ which is determined by the block describing the vicinity of the creation point (see Fig. \ref{diag2}). Summing up the same ladder sequence of the diagrams we get \begin{eqnarray} && \bar R_\tau({\bbox r})= \int{\rm d}^2 r_3\,{\rm d}^2 r_4\,\bar R({\bbox r}_0) M(\tau,{\bbox r}_1,{\bbox r}_2,{\bbox r}_3,{\bbox r}_4) \nonumber \\ && =\int{\rm d}^2 r_0\,\bar R({\bbox r}_0) S(\tau,{\bbox r},{\bbox r}_0) \,. \label{J1} \end{eqnarray} Here ${\bbox r}$ and ${\bbox r}_0$ are defined by Eq. (\ref{J3}), $M$ is the conditional probability (\ref{mmm}), $S$ is the conditional probability for the relative motion of the defects, see Eq. (\ref{H65}). Using the relations (\ref{fdt},\ref{rr0}) we get from Eq. (\ref{J1}) \begin{eqnarray} && \bar R_\tau({\bbox r}) =\frac{y_0^2}{a^4}\left(\frac{a}{r}\right)^{2\beta} R_\tau(r) \,. \label{fdt1} \end{eqnarray} Thus we see that the relation (\ref{fdt}) is reproduced for the renormalized quantities $R_\tau$ and $\bar R_\tau$. The renormalized creation rate $\bar R_\tau({\bbox r})$ can be interpreted as a probability density to find a defect-antidefect pair with a space separation ${\bbox r}$ provided the pair was born on time separated by $\tau$ from the measurement. Let us calculate the total probability density $\bar\lambda_{\rm r}$ to find the defect-antidefect pair at a fixed time separation $\tau$ regarding $\tau\gg a^2$. The probability is determined by the integral of $\bar R_\tau({\bbox r})$ over ${\bbox r}$. We conclude from the expressions (\ref{H69},\ref{dres1}) that the integral is determined by the region $r\sim\sqrt\tau$. Taking into account Eq. (\ref{dres3}) we get \begin{eqnarray} && \bar\lambda_{\rm r}=\int{\rm d}^2 r\,\bar R_\tau({\bbox r}) \sim\frac{y_0^2}{a^4} \left(\frac{a^2}{\tau}\right)^{\beta} \,. \label{J4} \end{eqnarray} We see that due to annihilation of defects at collisions the total probability diminishes at increasing the time separation $\tau$ as a power of $\tau$. The property can be interpreted as follows: The majority of defect-antidefect pairs annihilate fast after their creation and only a minor part of the defects achieve a separation $r\gg{a}$. The probability of such event is proportional to $(r/a)^{-2\beta}$. The results obtained in the subsection are correct if the variation of the coupling constant $\beta$ on the scale interval $a<r<\sqrt\tau$ is small. The existence of such interval is justified by the assumed small value of the fugacity $y_0$. Near $T_c$ variations of $\beta$ on a wide region of scales can be relevant. Then the consideration needs a generalization made in the last subsection of this section. \subsection{Coulomb interaction and diffusion coefficient} Let us consider the renormalization of the Coulomb interaction related to small defect-antidefect pairs. It is known that the influence of such pairs can be described in terms of a contribution to the effective dielectric constant, see Ref. \cite{KT}. The picture is naturally generalized for the dynamics. Before proceeding to calculations, it will be convenient for us to express the Coulomb part (\ref{hamb}) of the Hamiltonian (\ref{ham}) in an alternative form. Namely, using the Hubbard-Stratonovich trick we rewrite the fourth-order term ${\cal H}_\beta$ as a functional integral over auxiliary fields $\sigma$ and $\phi$ \begin{eqnarray} && \exp\left(-\int{\rm d}t\,{\cal H}_\beta\right) \nonumber \\ && =\int{\cal D}\phi\,{\cal D}\sigma \exp\left[-\int{\rm d}t\, \left({\cal H}_1+{\cal H}_2\right)\right] \,, \label{B0} \\ && {\cal H}_1=\int{\rm d}^2 r\, \biggl[\left(\nabla\hat\psi_+\psi_+-\nabla\hat\psi_-\psi_-\right)\nabla\phi \nonumber \\ && -\left(\hat\psi_+\psi_+-\hat\psi_-\psi_-\right)\sigma\biggr] \,, \label{B1} \\ && {\cal H}_2=\frac{1}{4\pi\beta}\int{\rm d}^2 r\,\nabla\sigma\nabla\phi \,. \label{B3} \end{eqnarray} The relation (\ref{B0}) can be easily checked using the bare expression \begin{eqnarray} && \langle\nabla\phi(t_1,{\bbox r}_1) \sigma(t_2,{\bbox r}_2)\rangle_0= -2\beta\frac{{\bbox r}_1-{\bbox r}_2} {|{\bbox r}_1-{\bbox r}_2|^2} \,, \label{B4} \end{eqnarray} following from Eq. (\ref{B3}). We see that the correlation function (\ref{B4}) corresponds to the dashed line on the diagrams. Note that the field $\phi$ in the expressions is the electrostatic potential introduced by Eq. (\ref{poten}). Indeed, integrating over the field $\sigma$ in Eq. (\ref{B0}) we get the Poisson equation \begin{eqnarray} && \nabla^2\phi=-\frac{1}{4\pi\beta}\tilde\rho \,, \label{BB4} \end{eqnarray} leading to the expression (\ref{poten}). Now we can work in terms of the sum ${\cal H}_0+{\cal H}_R+{\cal H}_1+{\cal H}_2$ where two first terms are introduced by Eqs. (\ref{ham0},\ref{hamr}). Note that the sum ${\cal H}_0+{\cal H}_1$ is invariant under the following infinitesimal transformation \begin{eqnarray} && \delta\psi_+=\alpha\psi_+\, \quad \delta\psi_-=-\alpha\psi_-\, \quad \delta\hat\psi_+=-\alpha\psi_+ \nonumber \\ && \delta\hat\psi_-=\alpha\psi_-\, \quad \delta\phi=2\alpha \, \quad \delta\sigma=\nabla^2\alpha \,, \label{ward} \end{eqnarray} where $\alpha$ is a function of coordinates. If to substitute $R(r)\to\lambda\delta({\bbox r})$ and $\bar R(r)\to\bar\lambda\delta({\bbox r})$ into the expression (\ref{hamr}) then it will be invariant under the transformation (\ref{ward}) also. Deviations from the symmetry related to $r$-dependencies of $R$ and $\bar R$ are irrelevant. If the function $\alpha$ contains only terms linear and quadratic over ${\bbox r}$ then the contribution ${\cal H}_2$ (\ref{B3}) is invariant under the transformation (\ref{ward}) also. The symmetry leads to a number of the Ward identities. Particularly, they connect the renormalized triple vertices to the self-energy function of the propagator. A typical diagram contributing to renormalization of the effective ``dielectric constant'' is drawn in Fig. \ref{diag3}. There we see a loop composed of the trajectories of a defect and of an antidefect which annihilate after their creation. There are also two ``external'' dashed lines corresponding to the interaction of the defect-antidefect pair with an environment. Besides the diagrams of the type drawn in Fig. \ref{diag3} there are also diagrams with two external dashed lines attached to the same trajectory. We draw the external lines with arrows to remember that two sides of the dashed line representing the correlation function (\ref{B4}) are not equivalent. We imply that the dashed lines are directed from the field $\sigma$ to the field $\phi$. As previously, we can dissect the diagram into parts which can be treated separately. Then the answer can be found as a convolution of the corresponding expressions. We perform the dissection along the planes in the ${\bbox r}-t$ space-time perpendicular to the $t$-axis and corresponding to the time moments $t_2$ and $t_3$ of the external Coulomb lines. In Fig. \ref{diag3} the dissection is shown by the dotted lines. We see that the loop is divided into three parts. The left part of the loop implying the integration over the time $t_1$ (see Fig. \ref{diag3}) corresponds to \begin{eqnarray} && \int_0^\infty{\rm d}\tau\, \int{\rm d}^2r_3\,{\rm d}^2r_4\, \bar R({\bbox r}_3-{\bbox r}_4) M(\tau,{\bbox x}_1,{\bbox x}_2,{\bbox r}_3,{\bbox r}_4) \nonumber \\ && = \int_0^\infty{\rm d}\tau\,\bar R_\tau(|{\bbox x}_1-{\bbox x}_2|) \,. \label{J6} \end{eqnarray} Substituting here Eq. (\ref{fdt1}) and using Eq. (\ref{unity}) we get \begin{eqnarray} && \int_0^\infty{\rm d}\tau\,\bar R_\tau(r) =\frac{y_0^2}{a^4} \left(\frac{a}{r}\right)^{2\beta} \,. \label{J7} \end{eqnarray} The central part of the diagram depicted in Fig. \ref{diag3} corresponds to the conditional probability (\ref{mmm}) $M(t_3-t_2,{\bbox x}_3,{\bbox x}_4,{\bbox x}_1,{\bbox x}_2)$. And the right part of the diagram in Fig. \ref{diag3} corresponds to the integral (\ref{unity}). The relation can be recognized as a manifestation of independence of all results of the final time $t_{\rm f}$ in the relation (\ref{path1}). If we chose $t_{\rm f}=t_3$ then the right part of the diagram in Fig. \ref{diag3} disappears and we should substitute $1$ instead, in accordance with (\ref{unity}). The structure of the diagram depicted in Fig. \ref{diag3} shows that the block related to the defect-antidefect pair can be treated as a self-energy insertion to the line corresponding to the Coulomb interaction. Thus it is natural to expect that this insertion can be treated as a contribution to the ``dielectric constant'', leading to a renormalization of the Coulomb constant $\beta$. The corresponding quantitative analysis is presented in Appendix \ref{rebeta} giving the following expression for the correction to $\beta$ \begin{eqnarray} && \Delta\beta\sim -y_0^2 \int\frac{{\rm d}r}{r}\left(\frac{a}{r}\right)^{2\beta-4} \,. \label{B11} \end{eqnarray} The expression can be treated as an integral over the characteristic sizes of the defect-antidefect pairs. One may try to find more complicated blocks contributing to a renormalization of the Coulomb coupling constant $\beta$. An example of such block is depicted in Fig. \ref{diag4} where a number (three) ``external'' lines are attached to the loop corresponding to the trajectories of the defect-antidefect pair. One can easily check that the block depicted in Fig. \ref{diag4} gives a correction to the Coulomb force which diminishes faster than $r^{-1}$ at increasing the distance $r$ between the interacting particles. Therefore the contribution is irrelevant. The same is true for more complicated diagrams of the same type. We can also consider blocks which can be treated as contributions to the diffusion coefficient $D$ introduced by Eqs. (\ref{H51},\ref{H52}). An example is depicted in Fig. \ref{diag5}, where the block between two dotted lines is a self-energy insertion to the propagator (\ref{diff}) which gives the renormalization of the diffusion coefficient. We will assume that the fields $\psi_\pm$ are corrected to keep the term (\ref{ham0}) unchanged. Then the contribution (\ref{diff3}) has to be extracted from the renormalization of the coefficient in front of the time derivatives. To analyze the correction $\Delta D$ quantitatively one should know a three-particle conditional probability which is more complicated than the two-particle conditional probability (\ref{mmm}). Fortunately, one can estimate the value of $\Delta D$ without detailed calculations. The point is that the dependence of $\Delta D$ on the cutoff $a$ can be produced only by regions near the creation or near the annihilation point (which are designated by ovals in Fig. \ref{diag5}). The regions can be analyzed in terms of the two-particle conditional probability (\ref{mmm}) since only the interaction of the nearest ``particles'' is relevant there. We already know the answer: the region near the creation point produces the renormalized creation rate (\ref{J1}) whereas the region near the annihilation point produce no $a$-dependence. Then simple dimensional estimates give the answer similar to the expression (\ref{B11}) \begin{eqnarray} && \Delta D\sim -y_0^2 \int\frac{{\rm d}r}{r}\left(\frac{a}{r}\right)^{2\beta-4} \,. \label{diff3} \end{eqnarray} Remind that the bare value of $D$ is assumed to be equal to unity. Of course there are also blocks which can be interpreted as corrections to the triple vertices describing the interaction of the fields $\hat\psi_\pm$, $\psi_\pm$ with the Coulomb fields $\phi$ and $\sigma$. An example of such block can be imagined if to attach an external dashed line to a trajectory between the dotted lines in Fig. \ref{diag5}. However, in the renormalization scheme accepted the corrections are fully absorbed into the renormalization of the fields $\psi_\pm$. That is a consequence of the symmetry of the Hamiltonian under the transformation (\ref{ward}). Namely, the cubic term (\ref{B1}) is unchanged provided the term (\ref{ham0}) is unchanged. \subsection{Summary} In the previous subsection we obtained the expressions for the corrections (\ref{B11},\ref{diff3}) of the Coulomb constant and of the diffusion coefficient in the main order over the fugacity $y_0$. Now we are going to discuss high-order corrections over $y_0$ which in statics lead to the renormalization group equations (\ref{rg}). It will be more convenient for us to proceed in spirit of the Kosterlitz renormalization group scheme. Namely, we see that the expressions (\ref{B11},\ref{diff3}) are written as integrals over the space variable $r$ which can be treated as the size of a defect-antidefect pair. We can first perform the integration over a restricted interval of the sizes, what gives slightly renormalized values of the coupling constants. Then we can repeat the integration. In the limit this multi-step procedure gives the renormalization group equations for the coupling constants, as Kosterlitz suggested. On the diagrammatic language the procedure means that we gradually substitute blocks corresponding to small separations between the particles by their effective values relative to larger scales. The procedure can also be considered as increasing an effective size of the defects $a\to r$. Then the renormalization of the coupling constants can be described in terms of the differential renormalization group equations. At each step of the procedure we deal with correlation functions like (\ref{mmm}). For an interval of scales where a variation of the Coulomb constant $\beta$ is small one can use for the function the expression (\ref{H70}) where now one should substitute the renormalized value of the Coulomb constant $\beta$. Analogously, the renormalized annihilation rate is determined by Eq. (\ref{dres1}) where one should substitute the expression (\ref{H69}) with the renormalized value of the Coulomb constant $\beta$. Next, for the renormalized creation rate we should use the relation \begin{eqnarray} && \bar R_{\tau_1}=\int{\rm d}^2 r_0\, S(\tau_1-\tau_2,{\bbox r},{\bbox r}_0) \bar R_{\tau_2} \,, \label{JJ1} \end{eqnarray} where $\tau_1>\tau_2$. The relation (\ref{JJ1}) can be derived from Eq. (\ref{J1}) if to use the property of $S$ analogous to Eq. (\ref{conv}). For the renormalized creation rate the relation (\ref{JJ1}) is correct only if $\beta(r_1)$ weakly differs from $\beta(r_2)$ for characteristic values of the parameters $r_1\sim\sqrt{\tau_1}$ and $r_2\sim\sqrt{\tau_2}$. A relation analogous to Eq. (\ref{JJ1}) can be formulated for the annihilation rate $R$. The relation leads to the same expression (\ref{dres1}) where $\beta$ is now scale-dependent. Therefore the renormalized quantity of the annihilation constant $\lambda_{\rm r}$ flows together with $\beta$ in accordance with Eq. (\ref{rrr}). That is accounted for the non-logarithmic character of the integrals leading to the relation (\ref{rrr}). Then we should define the renormalized fugacity in dynamics. For the purpose let us generalize the expression (\ref{J7}) \begin{eqnarray} && \int_0^\infty{\rm d}\tau\,\bar R_\tau(r) =\frac{y^2}{r^4}\,. \label{JJ7} \end{eqnarray} Then Eq. (\ref{J7}) is rewritten as \begin{eqnarray} && y=\left(\frac{a}{r}\right)^{\beta-2}y_0 \,. \label{renoy} \end{eqnarray} The relation can be treated as an elementary step of the renormalization group procedure which is described by Eq. (\ref{rg} for $y$. Thus the renormalization group equation for $\beta$ in dynamics coincides with one in statics. The expression (\ref{B11}) for the correction to the Coulomb constant $\beta$ can be considered as arising at the elementary step of the renormalization group procedure. The corresponding renormalization group equation can be found if to pass to the differential form and to substitute $y_0$ by the renormalized value $y$ in accordance with Eq. (\ref{renoy}). Then we obtain the renormalization group equation coinciding with Eq. (\ref{rg}) for $\beta$. The expression (\ref{diff3}) for the correction to the diffusion coefficient leads to the following renormalization group equation \begin{eqnarray} && \frac{{\rm d}D}{{\rm d}\ln(r/a)}\sim -y^2 \,, \label{diff4} \end{eqnarray} analogous to the equation (\ref{rg}) for $\beta$. We conclude that the correction to $D$ is small due to the small value of the fugacity and is therefore irrelevant. To avoid a misunderstanding, remind that the variation of the Coulomb constant $\beta$ with increasing scale is also small. Nevertheless, as seen from Eq. (\ref{rg}), it is the difference $\beta-2$ that enters the renormalization group equations and the variation of the difference can be essential. \section{Correlation Functions} \label{corr} Here, we treat non-simultaneous correlation functions of the charge density $\rho$ (\ref{den}) \begin{eqnarray} && F_{2n}(t_1,\dots,t_{2n};{\bbox r}_1,\dots,{\bbox r}_{2n}) \nonumber \\ && =\left\langle \rho(t_1,{\bbox r}_1) \dots \rho(t_{2n},{\bbox r}_{2n})\right\rangle \,. \label{H30} \end{eqnarray} Note an obvious consequence of the constraint (\ref{Ch1}) \begin{eqnarray} && \int{\rm d}^2 r_1\, F_{2n}=0 \,. \label{Ch2} \end{eqnarray} To examine the correlation functions (\ref{H30}) we use the representation (\ref{do3}). We will assume that all the diagrammatic blocks corresponding to small defect-antidefect pairs are already included into the renormalization of the corresponding coupling constants as discussed in Section \ref{renorm}. Therefore the fugacity $y$ and the Coulomb constant $\beta$ entering all subsequent expressions should be taken at the current scale. First we will examine contributions to the functions associated with a single defect-antidefect pair and then we will consider contributions related to a number of defect-antidefect pairs. \subsection{Pair Correlation Function} We start with the pair correlation function \begin{eqnarray} && F_2(t_2-t_1,{\bbox r}_2-{\bbox r}_1) =\langle\rho(t_2,{\bbox r}_2) \rho(t_1,{\bbox r}_1)\rangle \,, \label{J22} \end{eqnarray} with $t_2>t_1$. The average (\ref{J22}) can be calculated in accordance with the relation (\ref{path1}) where one should substitute the expression (\ref{do3}). We assume $t_{\rm f}=t_2$. The contribution to the average (\ref{J22}) related to a single defect-antidefect pair can be represented by a series of the diagrams with two lines constituted of the defect and antidefect propagators. The lines start from the creation point. A half of the diagrams have the structure depicted in Fig. \ref{vort4}. Here we omitted lines and vertices corresponding to the interaction of the defects (which is implied) and keep only trajectories of the defects. The trajectories should pass through the points ${\bbox r}_1$ and ${\bbox r}_2$ at the time moments $t_1$ and $t_2$ (the events are designated by black circles). An additional contribution to the average (\ref{J22}) is determined by similar diagrams where both events $t_1,{\bbox r}_1$ and $t_2,{\bbox r}_2$ belong to the same trajectory. As above, we dissect the diagram into parts which can be treated separately. Let us make a cut along planes in ${\bbox r}-t$ space-time corresponding to the time moments $t_1$ and $t_2$, they are shown in Fig. \ref{vort4} by dotted lines. Intermediate points appearing in the convolution (\ref{conv}) are designated in Fig. \ref{vort4} as ${\bbox r}_3$ and ${\bbox r}_4$. After that the diagram is divided into two parts separated by the dotted line. The part of the diagram to the right from the dotted line corresponds to the conditional probability $M$ (\ref{mmm}) and the part of the diagram to the left from the dotted line corresponds to the correlation function \begin{eqnarray} && \Phi_2({\bbox r}_1-{\bbox r}_2) =\langle \psi_+(t,{\bbox r}_1)\psi_-(t,{\bbox r}_2)\rangle \,. \label{J21} \end{eqnarray} The quantity (\ref{J21}) can be treated as the probability density to find a defect-antidefect pair with a given separation. Correspondingly, the integral $\int{\rm d}^2r\,\Phi_2({\bbox r})$ determines the density of the defect-antidefect pairs. The correlation function (\ref{J21}) coincides with the integral in the left-hand side of Eq. (\ref{JJ7}). Hence \begin{eqnarray} && \Phi_2(r)=y^2/r^4 \,, \label{rela} \end{eqnarray} where $y$ is the renormalized fugacity. The equations (\ref{fug},\ref{rela}) show that asymptotically in the low-temperature phase \begin{eqnarray} && \Phi_2(r)\propto r^{-2\beta}\,. \label{new} \end{eqnarray} Note that the same behavior (\ref{new}) is observed up to a slowly varying factor in the whole region of scales. The diagram depicted in Fig. \ref{vort4} gives a convolution of $\Phi_2$ and $M$. Adding the contribution corresponding to the case where both events $t_1,{\bbox r}_1$ and $t_2,{\bbox r}_2$ belong to the same trajectory we get the following expression for the pair correlation function (\ref{J22}) \begin{eqnarray} && F_2(t,{\bbox r}_2-{\bbox r}_1)= -2\int{\rm d}^2 r_3\,{\rm d}^2 r_4\, \Phi_2({\bbox r}_4-{\bbox r}_1) \nonumber \\ && \times \left[M(t,{\bbox r}_3,{\bbox r}_2,{\bbox r}_1,{\bbox r}_4) -M(t,{\bbox r}_2,{\bbox r}_3,{\bbox r}_1,{\bbox r}_4)\right] \,, \label{H81} \end{eqnarray} where $t>0$. At small times $t$ we turn to the limit law (\ref{small}). Substituting the expression into Eq. (\ref{H81}) we get \begin{eqnarray} && F_2(t=0,{\bbox r}_1-{\bbox r}_2) \nonumber \\ && =2\delta({\bbox r}_1-{\bbox r}_2)\int{\rm d}^2r\,\Phi_2({\bbox r}) -2\Phi_2({\bbox r}_1-{\bbox r}_2) \,. \label{H83} \end{eqnarray} Here the second contribution corresponds to the law (\ref{pair}) and the term proportional to $\delta$-function is an autocorrelation contribution associated with a single defect. The factor in front of the $\delta$-function (which is the density of defects) is in accordance with the relation (\ref{Ch2}). Note that at small $t$ the $\delta$-function is converted into a narrow function of the width $\sim\sqrt t$. It follows from Eqs. (\ref{H70},\ref{rela},\ref{H81}) that for $t\sim r^2$ \begin{eqnarray} && F_2(t,{\bbox r})\sim\frac{y^2(r)}{r^4} \,, \label{HH83} \end{eqnarray} To justify Eq. (\ref{HH83}) one should check that there are no divergences in the integral (\ref{H81}). It can be done directly using Eq. (\ref{H70}). The convergence at small separations $r$ and $r_0$ is accounted for the behavior of the modified Bessel functions $I_\nu(x)\propto x^\nu$ at small values of the argument. The convergence at large separations $r$ and $r_0$ can be checked using the asymptotic law (\ref{W8}). If $|t|\gg r^2$ then \begin{eqnarray} && F_2\sim -\frac{y^2(r)}{r^{4-2\beta}|t|^\beta} \,. \label{H84} \end{eqnarray} The asymptotic law is established in Appendix \ref{pairco}. The behavior of the pair correlation function determined by the laws (\ref{HH83},\ref{H84}) corresponds to a conventional critical dynamics (see, e.g., Ref. \cite{HH}) with the dynamical critical index $z=2$. However, as we will see below, the behavior of the high-order correlation functions is beyond the conventional scheme. Besides, the scaling law $t\sim r^2$ is true for the high-order correlation functions as well. \subsection{High-Order Correlation functions} \label{high-o} Here we extend the procedure of the preceding subsection to the case of the high-order correlation functions $F_{2n}$ (\ref{H30}). We will assume that $t_1<t_2<\dots<t_{2n}$. Again, we examine the contribution to $F_{2n}$ associated with a single defect-antidefect pair. Corresponding diagrams contain two trajectories starting anywhere and passing through the points ${\bbox r}_1$, \dots, ${\bbox r}_{2n}$ at the time moments $t_1$, \dots, $t_{2n}$. We will designate the trajectories of the defect and of the antidefect as ${\bbox x}(t)$ and ${\bbox z}(t)$. Let us dissect the diagrams along planes in the ${\bbox r}-t$ space-time corresponding to the time moments $t_1$, \dots, $t_{2n}$. Then the diagram is divided into a number of blocks, see Fig. \ref{vort5}. The left block in Fig. \ref{vort5} corresponds to the object (\ref{J21}) and all the other blocks correspond to the correlation function (\ref{mmm}). Again, using Eq. (\ref{conv}) we can write the contribution to the correlation function (\ref{H30}) associated with a single defect-antidefect pair as the following convolution \begin{eqnarray} && F_{2n}(t_1,\dots,t_{2n};{\bbox r}_1,\dots{\bbox r}_{2n}) \nonumber \\ && =\prod\limits_{j=1}^{2n}\int{\rm d}^2 x_j\,{\rm d}^2z_j\, \Phi_{2}({\bbox x}_1-{\bbox z}_1) \left[\delta({\bbox r}_j-{\bbox x}_j) -\delta({\bbox r}_j-{\bbox z}_j)\right] \nonumber \\ && \times M(t_{j+1}-t_j,{\bbox x}_{j+1},{\bbox z}_{j+1}, {\bbox x}_{j},{\bbox z}_{j})\,. \label{J31} \end{eqnarray} Here, one must replace the last factor $M(t_{2n+1}-t_{2n})$ by unity. The relation (\ref{J31}) is a generalization of Eq. (\ref{H81}). Thus we got an expression for the correlation function which is a multiple integral of functions determined by explicit formulas. A recurrent procedure for calculating $F_{2n}$ is suggested in Appendix \ref{recurr}. Of course the expression (\ref{H81}) for the pair correlation function is reproduced by Eq. (\ref{J31}). Note also the following expression \end{multicols} \begin{eqnarray} && \langle\rho(t,{\bbox r}_1)\rho(t,{\bbox r}_2) \rho(0,{\bbox r}_3)\rho(0,{\bbox r}_4)\rangle =2\left[M(t,{\bbox r}_1,{\bbox r}_2,{\bbox r}_3,{\bbox r}_4) +M(t,{\bbox r}_1,{\bbox r}_2,{\bbox r}_4,{\bbox r}_3)\right] \Phi_2({\bbox r}_3-{\bbox r}_4) \nonumber \\ && -B({\bbox r}_1-{\bbox r}_2)\delta({\bbox r}_3-{\bbox r}_4) -B({\bbox r}_3-{\bbox r}_4)\delta({\bbox r}_1-{\bbox r}_2) +\int{\rm d}^2r\,B({\bbox r})\, \delta({\bbox r}_1-{\bbox r}_2) \delta({\bbox r}_3-{\bbox r}_4) \,, \label{fou} \end{eqnarray} \begin{multicols}{2} \noindent where \begin{eqnarray} && B({\bbox r}_3-{\bbox r}_4)=2\int{\rm d}^2r_1\, \bigl[M(t,{\bbox r}_1,{\bbox r}_2,{\bbox r}_3,{\bbox r}_4) \nonumber \\ && +M(t,{\bbox r}_1,{\bbox r}_2,{\bbox r}_4,{\bbox r}_3)\bigr] \Phi_2({\bbox r}_3-{\bbox r}_4) \,. \nonumber \end{eqnarray} The formula can be found from Eq. (\ref{J31}) using the relation (\ref{small}). Naturally, the expression (\ref{fou}) is symmetric under the permutation ${\bbox r}_1,{\bbox r}_2\leftrightarrow{\bbox r}_3,{\bbox r}_4$, which can be checked using the expressions (\ref{H70},\ref{new}). The formula (\ref{fou}) is in agreement with the general property (\ref{Ch2}). The recurrent procedure for calculating $F_{2n}$ suggested in Appendix \ref{recurr} shows that there are no divergences in the integrals at all steps of calculating $F_{2n}$. This means that we can evaluate the correlation functions from naive dimension estimates. Namely, if all space separations among $|{\bbox r}_i-{\bbox r}_j|$ are of the same order $r_*$ and all time intervals are of the order $r_*^2$ then \begin{eqnarray} && F_{2n}\sim y^2(r_*)r_*^{-4n} \,. \label{H91} \end{eqnarray} In the large-scale limit when $\beta$ is saturated we have in accordance with Eq. (\ref{fug}) \begin{eqnarray} && F_{2n}\propto r_*^{-4(n-1)-2\beta} \,. \label{H99} \end{eqnarray} If some space separations among $|{\bbox r}_i-{\bbox r}_j|$ and/or some time intervals differ strongly then one can formulate some simple rules following from Eqs. (\ref{H70},\ref{J31}). Let us give some examples. If one of the time intervals $\tau$ is much larger than all values of the squared separations $|{\bbox r}_i-{\bbox r}_j|^2$ then the correlation function behaves like $F_{2n}\propto\tau^{-\beta}$. It can be proved like it is done in Appendix \ref{pairco} for the pair correlation function. For small $\tau$ there appear contributions to $F_{2n}$ short-correlated in space (on scales $\sim\sqrt\tau$). In the limit $\tau\to0$ the contributions turn into $\delta$-functions, as it was for the pair correlation function, see Eq. (\ref{H83}), representing an autocorrelation of single defects. If the points ${\bbox r}_j$ can be divided into two ``clouds'' with a separation $r$ between the clouds much larger than their sizes (and all time intervals are much smaller than $r^2$) then the principal $r$-dependence of the correlation function $F_{2n}$ is the same as in the function $\Phi_2(r)$ (\ref{rela}). Remember that the charge density $\rho$ is related to the curl of the gradient of the hexatic angle $\varphi$ for hexatics and to the curl of the gradient of the order parameter phase for superfluid films. It is instructive to re-express our results in terms of the phase gradient circulations of $\nabla\varphi$ over a closed loop $C$ \begin{eqnarray} && \Gamma(t,C)=\oint{\rm d}{\bbox r}\,\nabla\varphi =2\pi\int {\rm d}^2 r\,\rho(t,{\bbox r}) \,, \label{circ} \end{eqnarray} where the second integral is taken over the area inside the loop. Correlation functions of $\Gamma$'s can be rewritten as integrals of the correlation functions $F_{2n}$. As an example, consider the following average \begin{eqnarray} && \Psi_{2n}=\bigl\langle \Gamma(t,C)\Gamma(t+\tau,C)\dots \nonumber \\ && \times \Gamma[t+(2n-1)\tau,C]\bigr\rangle \,. \label{aver} \end{eqnarray} Suppose that the characteristic size of the loop $r$ is large enough so that we can assume that $\beta$ is saturated, and that $\tau\sim r^2$. Then the following scaling law is satisfied: if $r\to Xr$ and $\tau\to X^2 \tau$ then \begin{eqnarray} && \Psi_{2n}\to X^{4-2\beta}\Psi_{2n} \,, \label{scal} \end{eqnarray} where $X$ is an arbitrary factor. The law (\ref{scal}) is a consequence of Eq. (\ref{H99}). It has two striking peculiarities. First, it possesses a clear critical dependence. Second, it is independent of the order $n$. The procedure described above can be generalized to include the external potential. Then one should take the solution of the equation (\ref{MM1}) for the function (\ref{mmm}) and the corresponding expression for the object (\ref{J21}). \subsection{Many-Pair Contributions} We have established the contributions to the charge density correlation functions $F_{2n}$ associated with a single defect-antidefect pair. Now we are going to discuss other contributions to the correlation functions related to an arbitrary number of defect-antidefect pairs. Correspondingly, we should take diagrams with a number of trajectories passing through the points ${\bbox r}_1$, \dots, ${\bbox r}_{2n}$ at the time moments $t_1$, \dots, $t_{2n}$. The picture illustrating the situation is drawn in Fig. \ref{vort6} where black circles correspond to the arguments of $F_{2n}$: ($t_1,{\bbox r}_1$), \dots, ($t_{2n},{\bbox r}_{2n}$). There we omitted blocks related to short-living defect-antidefect pairs regarding that the blocks are already included into the renormalization of the Coulomb constant $\beta$. As previously, we can dissect the diagrams along the planes in the ${\bbox r}-t$ space-time corresponding to the time moments $t_1$, \dots, $t_{2n}$. Then any diagram will be divided into a number of strips, see Fig. \ref{vort6}. The part of the diagram within each strip can be treated as the corresponding matrix element of the evolution operator $\exp(-\int{\rm d}t\,{\cal H})$. Then the contribution to $F_{2n}$ will be written like (\ref{J31}) as a convolution of the matrix elements. Generally, the matrix elements can be estimated like the function (\ref{mmm}): each trajectory segment gives a factor which scales as $1/t$ and $t$ scales as $r^2$. But there are obvious exceptions from the rule. Namely, going back in time we will come to a moment where a defect-antidefect pair was created. Again, when we consider small separations between the defect and the antidefect (which is correct for time moments close to the creation time) we can take into account only the Coulomb interaction between the two created defects. The corresponding regions in Fig. \ref{vort6} are inside the ovals. Each such region produces the factor $y^2$. Therefore generally $F_{2n}\propto y^{2k}$, where $k$ is the number of the pairs. Taking into account also a scale-dependent factor we get \begin{eqnarray} && F_{2n}\sim y^{2k}(r_*)r_*^{-4n} \,, \label{J41} \end{eqnarray} where we assume that all space separations are of the order $r_*$ and all time intervals are of the order $r_*^2$. The expression (\ref{J41}) is a generalization of Eq. (\ref{H91}). Comparing these two expressions we see that the ratio of the contribution (\ref{J41}) to the contribution (\ref{H91}) is the $(k-1)$-th power of a dimensionless small parameter $y^2(r_*)$. Thus we conclude that the leading contribution to $F_{2n}$ is related to a single defect-antidefect pair, that corresponds to $k=1$. Now we can explain the origin of the estimate (\ref{simul1}) for the simultaneous correlation functions which obviously does not coincide with Eq. (\ref{H91}). The estimate (\ref{simul1}) in terms of Eq. (\ref{J41}) corresponds to $k=n$. The reason is quite obvious: Two defects cannot pass simultaneously through $2n$ points and at least $k=n$ defect-antidefect pairs should be taken to obtain a non-zero contribution to the simultaneous correlation function $F_{2n}$. The situation is illustrated by Fig. \ref{vort7}. Note that the estimate (\ref{simul1}) is not correct for the autocorrelation contributions proportional to $\delta$-functions, as written in Eq. (\ref{H83}). Thus we have two different regimes: for simultaneous and for non-simultaneous correlation functions. Let us establish the boundary between the regimes. For the purpose we should consider small time intervals where the single-pair contribution is finite but small. The smallness is associated with diffusive exponents presented, e.g., in the expression (\ref{H70}). Therefore the characteristic time where the simultaneous regime passes into the non-simultaneous one can be estimated as \begin{eqnarray} && t\sim\frac{r^2}{|\ln[y(r)]|} \,, \label{simul} \end{eqnarray} where $r$ is a space separation corresponding to the small time interval. In the low-temperature phase on large scales (where $\beta$ is saturated) we have $|\ln y|\approx(\beta -2)\ln(r/a)$. \section{Superfluid films} \label{helium} Let us consider superfluid films. The equation of motion for the vortices contains an additional term (Magnus force). Thus instead of Eq. (\ref{H51}) we should write (see Ref. \cite{Dynamics1}) \begin{eqnarray} && \frac{{\rm d}x_{j,\alpha}}{{\rm d}t}=-\frac{D}{T} \left[\frac{\partial{\cal F}}{\partial x_{\alpha j}} +n_j\gamma\epsilon_{\alpha\beta} \frac{\partial{\cal F}}{\partial x_{\beta j}}\right] +{\xi}_{j,\alpha} \,, \label{H93} \end{eqnarray} where $\gamma$ is a new dimensionless parameter. The equation (\ref{H93}) can be derived in the spirit of the procedure proposed by Hall and Vinen for the $3d$ superfluid, see Ref. \cite{HV}. Huber \cite{82Hu} argued that the same equation is correct for spin vortices in planar $2d$ magnetics. For the superfluid films the ``charge density'' (\ref{den}) is proportional to the vorticity ${\rm curl}\,{\bbox v}_{\rm s}$. To calculate correlation functions $F_{2n}$ (\ref{H30}) one can use the scheme developed in the previous sections. The only difference is that instead of the expression (\ref{H70}) for the conditional probability $M$ (\ref{mmm}) one should use the solution of the equation \begin{eqnarray} && \partial_t M=\left(\frac{1}{2}\nabla_\varrho^2+ 2\nabla_r^2+4\beta\frac{\bbox r}{r^2}\nabla_r\right)M +2\gamma\beta\epsilon_{\alpha\beta}\frac{r_\beta}{r^2} \nabla_{\varrho\alpha}M \nonumber \\ && -R(r)M+\delta(t)\delta({\bbox r}-{\bbox r}_0) \delta\left({\bbox\varrho}-\frac{{\bbox r}_3}{2} -\frac{{\bbox r}_4}{2}\right) \,. \label{H95} \end{eqnarray} The variables ${\bbox r}$, ${\bbox r}_0$ and ${\bbox\varrho}$ are introduced by Eq. (\ref{J3}). Again, on scales $r\gg a$ one can omit the term with the annihilation rate $R$ in Eq. (\ref{H95}) demanding a finite value of $M$ at $r\to 0$ instead. Unfortunately, a cross term over ${\bbox r}$ and ${\bbox\varrho}$ appears in the operator in the right-hand side of Eq. (\ref{H95}). Thus one cannot obtain an explicit expression for $M$ of the type of Eq. (\ref{H70}). Nevertheless, this additional term has the same dimensionality as the other terms and does not change the scaling estimates $M\sim t^{-2}$, $t\sim r^2$ determining the function $M$. Moreover, the equation for the object \begin{eqnarray} && S(t,{\bbox r},{\bbox r}_0)=\int{\rm d}^2\varrho\, M(t,{\bbox r},{\bbox\varrho},{\bbox r}_0,{\bbox r}_3/2+{\bbox r}_4/2) \,, \label{H96} \end{eqnarray} following from (\ref{H95}) is identical to Eq. (\ref{four}). Therefore for the object (\ref{H96}) we have the same series (\ref{angle}) with the coefficients (\ref{H69}). Looking through the derivation presented in Section \ref{renorm} we see that just the function $S$ (\ref{H96}) enters all the relations. Therefore we can make the same assertions as previously. First, on large scales the annihilation coefficient $\lambda$ is equal to its universal value (\ref{rrr}). Second, we can write the same expression (\ref{rela}) for the average (\ref{J21}). Third, in dynamics we get the same renormalization group equation (\ref{rg}) for $\beta$, see Appendix \ref{rebeta}. Fourth, the renormalization of the diffusion coefficient is irrelevant. And finally, one can assert that a renormalization of the parameter $\gamma$ introduced by Eq. (\ref{H93}) is determined by the equation \begin{eqnarray} && \frac{{\rm d}\gamma}{{\rm d}\ln(r/a)}\sim y^2 \,, \nonumber \end{eqnarray} analogous to (\ref{diff4}). Therefore the renormalization of $\gamma$ is irrelevant. Again, the scheme can be generalized to include the ``external potential'', which now is the average value of the superfluid velocity. Next, we proceed to the correlation functions $F_{2n}$ (\ref{H30}). Formally they are determined by the same convolution (\ref{J31}) as previously. However, one should substitute there the solution of the equation (\ref{H95}). Therefore the concrete expressions for $F_{2n}$ will be different. Nevertheless, the estimates like (\ref{H91},\ref{H99},\ref{J41}) remains true because of the following reasons. First, due to the same dimensionality of all the terms in the right-hand side of Eq. (\ref{H95}) the function $M$ possesses the simple scaling properties noted above. Second, there are no divergences in the convolutions like (\ref{J31}) determining the objects. To prove the second property, we should analyze a behavior of $M$ at large and at small separations. In the case $rr_0/t\gg1$ the characteristic values of the separations ${\bbox r}_1-{\bbox r}_3$ and of ${\bbox r}_2-{\bbox r}_4$ are $\sim\sqrt t$ and are consequently much smaller than $|{\bbox r}_1-{\bbox r}_2|$ (or $|{\bbox r}_3-{\bbox r}_4|$). Then it is possible to neglect all terms containing $|{\bbox r}_1-{\bbox r}_2|$ in denominators in the equation (\ref{H95}) and we come to a purely diffusive equation leading to the asymptotics (\ref{W8}). It is possible to establish that the small-scale of the conditional probability $M$ for the vortices coincides with that examined above. The properties ensure convergence of all intermediate integrals appearing at calculating the correlation functions of vorticity $F_{2n}$. We have also the same scaling law (\ref{scal}) for the correlation function (\ref{aver}) of the integrals (\ref{circ}) which are now proportional to the circulations of the superfluid velocity. We conclude that all the scaling laws for the correlation functions of the vorticity and their asymptotic behavior remains the same as previously. \section{Discussion} \label{conclu} The main result of our consideration is the expression (\ref{H91}) for high-order correlation functions of the ``charge density'' (\ref{H30}) which is disclinicity for hexatic films and vorticity for superfluid films. We see from Eq. (\ref{H91}) that the high-order correlation functions are much larger than their normal estimates via the pair correlation function. Namely, in accordance with Eqs. (\ref{HH83},\ref{H91}) we have \begin{eqnarray} && {F_{2n}}/{F_2^n}\sim y^{-2n+2}\gg1 \,, \label{intem} \end{eqnarray} where $y$ is the renormalized fugacity. The asymptotic behavior of the ratio at large scales is determined by the law (\ref{fug}). Though at developing our scheme we accepted that the defect-antidefect pairs constitute a dilute solution we hope that the scaling law (\ref{intem}) is universal. The ground for the hope is the renormalization group procedure (formulated in Ref. \cite{74Kos}) which shows that on large scales we come to an effectively dilute solution of the pairs. We believe that the most interesting fact to be compared with experiment or numerics is the scaling law (\ref{scal}) which is a consequence of Eq. (\ref{intem}). The physics behind the inequality (\ref{intem}) is as follows. The main contribution to the correlation functions is associated with a single defect-antidefect pair. Though the contribution associated with a number of defect-antidefect pairs contains an additional huge entropy factor it has also an additional small factor associated with small probability to observe defect-antidefect pairs with separations larger than the core radius $a$. The smallness is accounted for the strong Coulomb attraction. The considered effect is a consequence of the competition of those two factors. The result of the competition manifests in the law (\ref{J41}) which gives the estimate for the contribution associated with $k$ defect-antidefect pairs. For simultaneous correlation functions nothing similar occurs and we have the conventional estimate (\ref{simul1}). That is the reason why the effect cannot be observed in statics. The property is directly related to causality since a defect-antidefect pair cannot simultaneously pass through $2n$ points and at least $n$ defect-antidefect pairs is needed to get a non-zero contribution to the simultaneous correlation function $F_{2n}$, see Fig. \ref{vort7}. That explains the estimate (\ref{simul1}). Thus we have two different regimes for simultaneous and non-simultaneous correlation functions. The characteristic boundary time separating those two regimes is written in Eq. (\ref{simul}). The considered effect resembles intermittency in turbulence (see, e.g., Ref. \cite{Frish}) leading to large $r$-dependent factors in the ratios like (\ref{intem}) in the velocity correlation functions of a turbulent flow. However, as is seen from Eq. (\ref{intem}), for the defects the large $r$-dependent factors are related to the ultraviolet cutoff parameter ${a}$ whereas for intermittency in turbulence the large $r$-dependent factors are related to the infrared (pumping) scale. Our situation is thus closer to the inverse cascade (see Ref. \cite{Kraich}) realized on scales much larger than the pumping length. There are experimental data \cite{98Tab} concerning the inverse cascade in $2d$ hydrodynamics and analytical observations concerning the inverse cascade for a compressible fluid \cite{CKGV} which indicate the absence of the intermittency in the inverse cascades. Note that only simultaneous objects were examined in the works \cite{98Tab,CKGV}, and there is no intermittency in our simultaneous correlation functions. So, based on the analogy, one may think that for the inverse cascades non-simultaneous objects reveal some intermittency. The consideration presented in our work is applicable to superfluid films. There exist also films and quasi-$2d$ systems of different symmetry. Huber \cite{82Hu} argued that the same equation as for quantum vortices is correct for spin vortices in planar $2d$ magnets. We believe that our approach based on the equation (\ref{H51}) is correct for the dynamics of disclinations in hexatic films like membranes, freely suspended films and Langmuir films. Next, the above scheme seems to work also for dislocations in solid films. The system needs a special treatment since a modification should be introduced into the procedure. Maybe some features of the presented picture can be observed also in superconductive materials, especially in high-$T_c$ superconductors. There are analytical and numerical indications that for a purely Langevin dynamics of the order parameter there are logarithmic corrections to the law (\ref{H51}), see Refs. \cite{87Cl,88Pl,91Pa,91Ry,92Mu,93Yu,94Ko}. We believe that the logarithms are destroyed if to switch on an interaction of the order parameter with other degrees of freedom. Nevertheless, it would be interesting to generalize our scheme including the logarithmic corrections. \acknowledgements I am grateful to G. Falkovich, K. Gawedzki, I. Kolokolov, and M. Vergassola for useful discussions and to E. Balkovsky, A. Kashuba and S. Korshunov for valuable remarks. This research was supported in part by a grant of Israel Science Foundation, by a grant of Minerva Foundation and by the Landau-Weizmann Prize program.
\section{Introduction} The high--momentum asymptotics of QCD and QED amplitudes is under active study for several decades [1,2]. The basic method used is the summing up of the series of dominant Feynman amplitudes keeping the leading asymptotic term in each of them. These terms appear to be of the type $\alpha^n_s ln^r Q^2$ and in case $r=2n$ one has the so-called double logarithmic asymptotics, which e.g. defines behaviour of the QED form--factor at large $Q^2$ [3]. An extension of the double--logarithmic asymptotics to other processes has been made both in QED [4] and QCD [2,5]. One of attractive features of these results is the exponentiation of leading logarithmic terms coming from each Feynman amplitude due to summation over $n$, which might be a hint, that there can exist an alternative, more direct method of deriving this result. A step in this direction is made in the present paper. We start with a most general n-point Green's function, depending on $n$ external momenta $p_k, k=1,...n$. Two kinds of theories are considered: QED and QCD, but the generalization to the EW case is straightforward. In this first study we confine ourselves to the simplest setting of the problem: all combinations of momenta are large compared to masses, external $M_i$ and internal $m_i$, \begin{equation} p^2_i\gg M_i^2, ~~m_k^2,~~i,~~k=1,...n \end{equation} Moreover the following condition is assumed to hold \begin{equation} |p_ip_k|\gg p^2_i,p^2_k; i,k=1,...n \end{equation} Our main object of the study will be the one--fermion--loop amplitudes which assumes the limit of large $N_c$ for QCD. However one may expect that the inclusion of additional fermion loops gives only small correction to the leading asymptotics, so that the one-- fermion--loop amplitudes are dominant in the asymtotics (1) and (2) both for QCD and QED. For QCD there is a special problem of nonperturbative (NP) contributions, which can be handled as in [6], i.e. dividing the total vector field $A_\mu=B_\mu+a_\mu$, where $B_\mu$ is the NP background and $a_\mu$ is the perturbative field. However with the conditions (1) and (2) the standard expectation is that the NP contribution dies out as a power, e.g. $\sigma/p^2_i,\frac{\sigma}{|p_ip_k|}$, where $\sigma $ is the string tension (or any equivalent combination like gluonic condensate) made of the NP field $B_\mu$. We shall argue in the last section of the paper that actually NP effects can be also exponentiated and may strongly modify the perturbative result. The plan of the paper is as follows. In section 2 we introduce FFSR for the $n$--point Green's function . In section 3 the approximate path integration is done in a way pertinent to the high momentum kinematics (1),(2) and double logarithmic terms in the asymptotics are calculated. In section 4 the Sudakov vertex asymptotics is identified and formfactors are discussed. In section 5 the influence of confinement is considered and as a result the confinement correction is calculated in addition to the DLA term. Discussion of other applications is given in the conclusions, and in the appendix the spin--dependent terms $(\sigma F)$ in the fermion propagators are estimated and shown to be subleading. Note, that the situation with NP fields changes, if one abandons condition (1), since the Green's function under consideration has bound state poles at $p^2_i=M^2_i$, and those are mostly due to the NP forces between quark and antiquark. As it is, however, both conditions ensure that to the leading approximation one can consider only perturbative (gluon or photon) exchanges between fermion lines and the main point is how to sum up those in the most effective way. To this end we shall exploit the Fock--Feynman--Schwinger representation (FFSR) for the fermion propagator in the external field [7,8], which was used previously for similar purposes in [9]. The path--integral method have been used for the calculation of the Green's function asymptotics in QED for a long time, see e.g. [10] and refs. therein. More recently the so-called world line method was introduced in [11]. Below we follow the formalism of [9], which allows for a more economical treatment of spin. \section{Integral representation for the fermion propagator} For the spin--1/2 particle propagator in the external vector field $A_\mu$ one has the FFSR (we start with the Euclidean metric for convenience) \begin{equation} S(x,y)=(m-\hat D)\int^\infty_0 ds(Dz)_{xy}e^{-K} \Phi_\Sigma(x,y) \end{equation} where notations used are \begin{equation} K=m^2s+\frac{1}{4}\int^s_0 d\tau(\frac{dz_\mu}{d\tau})^2;~~ D_\mu=\partial_\mu-igA_\mu, \end{equation} \begin{equation} (Dz)_{xy}=lim_{N\to \infty}\prod^N_{n=1}\frac{d^4\xi(n)} {(4\pi\varepsilon)^2}\frac{d^4q}{(2\pi)^4} e^{iq(\sum\xi(n)-(x-y))}, N\varepsilon=s \end{equation} \begin{equation} \Phi_\sigma(x,y)=P_AP_F exp ig \int^x_y A_\mu dz_\mu exp g\int^s_0 d\tau\sigma_{\mu\nu}F_{\mu\nu}, \end{equation} and $$ \sigma_{\mu\nu}=\frac{1}{4i}(\gamma_\mu\gamma_\nu-\gamma_\nu\gamma_\mu), ~~\xi(n)=z(n)-z(n-1). $$ Consider now the n-point Green's function with external momenta at $n$ vertices equal to $p_i$ (see Fig.1) \begin{equation} G(p_1,...p_n)=<J_1(p_i)...J_n(p_n)>, J_i(x)=\psi(x)\Gamma_i\psi(x) \end{equation} Insertion of (3) into (7) for the one--fermion loop yields \begin{equation} G(p_1,... p_n)=<tr\prod^n_{i=1}\Gamma_i(m_i-\hat D_i)\int^\infty_0 ds_i(Dz^{(i)})_{x^{(i)},x^{i-1}}e^{-K_i}\Phi^{(i)}_\sigma e^{ip^{(i)} x^{(i)}}dx^{(i)}>_A \end{equation} We shall disregard in what follows the factors $\Gamma_i(m_i-\hat D_i)$ since we shall be interested only in the exponentiated contributions; doing the $dx^{(i)}$ integrals, one obtains $$ G\to \bar G_n\delta(\sum^n_{i=1}p_i)(2\pi)^4, $$ where \begin{equation} \bar G_n=\int\frac{d^4q}{(2\pi)^4}\prod^n_{i=1} ds_i\prod^N_{k=1}\frac{d\xi^{(i)}(k)}{(4\pi\varepsilon)^2} e^{iq^{(i)}\sum_k\xi^{(i)}(k)}e^{-K_i}< \Gamma_i(m_i-\hat D_i)W_\sigma> \end{equation} and $$ <W_\sigma>=<\prod^n_{i=1}\Phi_\sigma^{(i)}>_A, $$ The integral $d^4q$ denotes the integral over one of $q_i$, all others being expressed through it and all $p_i$. We note that $<W_\sigma>$ is a gauge invariant quantity summing all the perturbative exchanges inside the fixed Wilson contour, defined by the set $\{\xi^{(i)}(k)\}$. In addition to the usual Wilson (charge) vertices, there are also magnetic moment vertices $\sigma F$, hence the notation $<W_\sigma>$. We concentrate now on the contribution of the $A_\mu$ field in (6), referring the reader for the discussion of the $\sigma F$ term to the Appendix. The crucial step for what follows is the use of the cluster expansion theorem [12], which yields for $<W_\sigma>\to <W>$ \begin{equation} <W>\equiv exp \{\sum^\infty_{r=1}\frac{(ig)^r}{r!} \sum_{k_i}\xi_{\mu_1}(k_1)\xi_{\mu_2} (k_2) ... \xi_{\mu r} (k_r)\ll A_{\mu_1} (z_{k_1})... A_{\mu_r} (z_{k_r})\gg\} \end{equation} Here double brackets denote cumulants [12], the lowest order contribution (in the exponent) is expressed through photon (gluon) propagator, which in the Feynman gauge is (the gauge is irrelevant since $<W>$ is gauge invariant) \begin{equation} <A_{\mu}(z)A_\nu(z')>= \frac{\delta_{\mu\nu}C_2(f)\hat 1}{4\pi^2 (z-z')^2} \end{equation} Here $C_2(f)$ is the quadratic Casimir operator for fundamental representation, $\hat 1$ is the unit color matrix, for QED one should replace $C_2\hat 1\to 1$. In what follows we confine ourselves to the lowest contribution (11) in (10) and show that it yields the double logarithmic asymptotics, leaving next order terms for the future. First of all one can persuade oneself that the approximation (11) yields in (9) all diagrams with exchanges of photon/gluon lines between fermion lines, all orderings of lines included. For QCD this means the following: all orderings, i.e. all intersection of gluon lines in space-time are included, except that color ordering of operators $t^a$ is kept fixed. Since commutator of any two generators $t^a$ is subleading at large $N_c$, it means that (11) sums up all exchanges including intersections of gluonic lines in the leading $N_c$ approximation ($cf$ the discussion of this point in [13], note that e.g. the nonplanar diagram $0(g^4)$ is given by the quartic cumulant in (10) and the latter is suppressed by the factor $1/N_c$). \section{The Gaussian integration} Our next point is the integration over $d\xi(k)$ in (9) which is Gaussian in the main term $K_i$, defining the measure of integration; therefore we shall do it expanding the exponent in (9) around the stable fixed point $\bar \xi$, which is obtained by differentiating the exponent in (9) with respect to $\xi^{(i)}(k)$. One has $$ \bar \xi^{(i)}(k)=2\varepsilon_i\{iq^{(i)}- \frac{g^2C_2(f)}{4\pi^2}\sum_{j,k'} \frac{\bar \xi^{(j)} (k')}{(\bar z^{(i)}(k)-\bar z^j(k'))^2}+ $$ \begin{equation} +\frac{2g^2C_2(f)}{4\pi^2}\sum_{j,k'} \sum_{m\geq k}\frac{( \bar \xi^{(i)} (m)\bar \xi^{(j)}(n'))(\bar z^{(i)}(m)-\bar z^{(j)}(n'))}{(\bar z^{(i)}(m)-\bar z^{(j)}(n'))^4}\}+ 0(g^4) \end{equation} Here e.g. $\bar z^{(i)}(k)=\sum^i_{j=1}\sum^k_{\nu=1}\bar \xi^{(j)}(\nu)$, where we have chosen as the origin the coordinate $x^{(1)}$ of the first vertex, and all other coordinates are calculated using the connection $x^{(i)}-x^{(i-1)}=\sum^N_{k=1}\xi^{(i)}(k)$ with the cyclic condition $x^{(n+1)}=x^{(1)}$. One can solve the system of equations (12) iteratively expanding in powers of $g^2$, the first two terms are given in (12), where one should replace $\bar \xi^{(i)}$ inside the curly brackets by $2i\varepsilon_i q^{(i)}$. If one represents the exponential appearing in (9) after insertion of (10) as $exp (-f(\xi, q))$, then one can write \begin{equation} f(\xi,q) = \sum_{i,k}\frac{(\xi^{(i)}(k))^2}{4\varepsilon_i}-i\sum_{i,k} q^{(i)} \xi^{(i)} (k)-\frac{g^2C_2(f)}{8\pi^2} \sum_{i,j,kk'}\frac{\xi^i(k) \xi^j(k')}{(z^i(k)-z^j(k'))^2}+0(g^4) \end{equation} The Gaussian integration in (9) finally yields \begin{equation} \bar G_n\sim \int\frac{d^4q}{(2\pi)^4} \prod^n_{i=1} ds_i e^{-f(\bar \xi,q)-\frac{1}{2} tr ln \varphi} \end{equation} where the matrix $\varphi$ is \begin{equation} \varphi^{ij}_{kn} = \frac{1}{2} \frac{\partial^2}{\partial \xi^{(i)}(k) \partial\xi^{(j)} (n)}f(\xi, q)\biggl | _{\xi=\bar \xi} \end{equation} The most important for what follows is the term $f(\bar \xi,q)$ which can be written as (at this point we reestablish Minkowskian metric) \begin{equation} f(\bar \xi, q) =\sum^n_{i=1}s_i(q^{(i)})^2+\frac{g^2C_2 (f)}{8\pi^2} \sum_{ij}\int^{s_i}_0 \int^{s_j}_0\frac{d\tau_id\tau_j(q^{(i)} q^{(j)})}{(\tau_iq^{(i)}-\tau_jq^{(j)} -\Delta_{ij})^2} \end{equation} where we have defined $\tau_i=k\varepsilon_i$, and \begin{equation} \Delta_{ij}=\sum^{j-1}_{k=i}s_kq^{(k)}, i<j \end{equation} The integral in the last term on the r.h.s. of (16) can be written as $s_is_j(q^{(i)}q^{(j)})I(s,q)$, where \begin{equation} I_{ij}(s,q)=\int^1_0\int^1_0\frac{d\alpha d\beta}{(\alpha s_iq^{(i)}-\beta s_jq^{(j)}-\Delta_{ij})^2} \end{equation} The diagonal terms, $I_{ij},$ with $i=j$ do not contribute to the asymptotics and contain only selfenergy divergencies, which are of no interest to us in what follows. Therefore we shall consider nondiagonal terms with $i\neq j$. Let us first study the term with $i=j-1$ ('the dressed vertex contribution"), and $\Delta_{i,i+1}=s_iq^{(i)}$, see Fig. 2. Then (18) is reduced to the form which will be studied below \begin{equation} I_i\equiv I_{i,i+1}(s,q)=\int^1_0\int^1_0 \frac{d\alpha d\beta}{(\alpha s_iq^{(i)}+\beta s_jq^{(j)})^2}= \int^1_0\int^1_0 \frac{d\alpha d\beta}{(a^2\alpha^2+\beta^2b^2+2\alpha\beta(ab)} \end{equation} with $a=s_iq^{(i)}, b=s_jq^{(j)}, j=i+1$. As it stands the integral (19) diverges at small $\alpha,\beta$ (or at small $\tau_i,\tau_j$ in (16)). The origin of this divergence becomes physically clear, when one expresses the distance $z^{(i)}$ from the vertex position, (we go over to the Minkowskian space--time) \begin{equation} z^{(i)}=2q^{(i)}\tau_i, z^{(j)}=2q^{(j)}\tau_j. \end{equation} The quasiclassical motion (20) cannot be true for small $\tau_i$ when quantum fluctuations wash out the straight--line trajectories, and the lower limit $\tau_{min}$ can be obtained from the quantum uncertainty principle \begin{equation} \Delta z\Delta q\sim (z^{(i)}-z^{(j)})(q^{(i)}-q^{(j)})\sim 1 \end{equation} We shall be interested in the kinematical region, where \begin{equation} |(q^{(i)}q^{(j)}|\gg(q^{(i)})^2, (q^{(j)})^2, \end{equation} and $\tau_{min}$ then is found from (21) to be \begin{equation} \tau_{min}\sim \frac{1}{2|q^{(i)}q^{(j)}|} \end{equation} Using (23) one can easily calculate the integral (19) since the term $2\alpha\beta(ab)$ in the denominator of the integrand in (19) always dominates. The result is \begin{equation} I_i=\frac{1}{2s_is_{i+1}(q^{(i)}q^{(i+1)})}ln(2(qq')s_i)ln(2(qq') s_{i+1}) \end{equation} The integration of the general term $I_{ij}$ with $j\neq i-1,i+1$ can be done using the expressions for the Spence functions. However in the general case the lower limit $\tau_{min}$ is inessential and the double logarithmic situation does not occur unless there is a large ratio, $|\frac{(q_iq_k)}{(q_lq_m)}|\gg 1$. We leave the detailed study of this point to the future. \section{The 3-- point Green's functions} We start with the open triangle, Fig.2 corresponding to the Sudakov vertex function asymptotics, \begin{equation} \bar G_3 = (-i\hat q +m)^{-1} \Gamma(q,q')(-i\hat q'+m)^{-1} \end{equation} In this case there is no integration over $dq$ in (9) and only one integral $I_{12}$ (18),(24) is present (we disregard as before the selfinteracting pieces $I_{ii}$, which do not contribute to the asymptotics). Insertion of the integral (24) into (16) and integration over $ds_1ds_2$ in (14) which yields in (24) (with logarithmic accuracy) replacement $s_i\to \frac{1}{q_i^2}$, finally leads one to the answer for QED $(C_2\equiv 1)$ \begin{equation} \Gamma(q,q')\sim exp (-\frac{\alpha}{2\pi}ln \frac{2|qq'|}{q^2} ln \frac{2|qq'|}{(q')^{2}}) \end{equation} which coincides with the known Sudakov asymptotics [3]. We turn now to the case of QCD, where the basic triangle diagram is closed due to color gauge invariance and try to find out whether the kinematical region (22) plays important role in the integral over $dq$ in (14). In the general case, when all $q_i$ are unconstrained and expressed through three external momenta $p_1,p_2,p_3$ and one integration variable, the region (22), yielding double logarithmic asymptotics (DLA) (24), is suppressed due to large values of $f(\bar \xi, q)$ in the exponent. As a result the integral over $dq$ does not lead to the DLA form for $\bar G_3$. The situation changes however, if one considers instead of $\bar G_3$ the formfactor, i.e. when the pole terms are factored out from the vertices 2 and 3 see Fig. 3, and vertex functions appear there. To simplify matter, one can consider for the formfactor the same representation (14), where under the integral one introduces vertex functions $\psi_i(k_i),i=1,3$, where $$ k_1= (q^{(1)}+q^{(3)})-p^{(1)}\frac{(q^{(1)}+q^{(3)})p^{(1)}}{(p^{(1)})^2} $$ \begin{equation} k_3= q^{(2)}+q^{(3)}-p^{(3)}\frac{(q^{(2)}+q^{(3)})p^{(3)}}{(p^{(2)}_3) } \end{equation} The definition (29) yields in the c.m.system of particle 1 or 3 the familiar relative momentum of two emitted fermions. The presence of $\psi_i$ imposes restriction on momenta $q_i$, namely \begin{equation} k^2_1, k^2_3 \mathrel{\mathpalette\fun <} \kappa^2 \end{equation} where $\kappa^2$ is some hadronic scale. We can define in the Breit system momenta as follows $p^{(1)}, p^{(2)}, p^{(3)}= (p_0,-\frac{\vec Q}{2}),(0,\vec Q), (p_0, \frac{\vec Q}{2})$ where $\vec Q^2\gg \kappa^2$ and $p_0^2=M^2+\frac{\vec Q^2}{4}$. One can easily see, that (30) constrains the region of integration over $dq\equiv d^4q_1$ to the region $|\vec q |\sim \kappa, |q_0-p_0|\sim \kappa$, and conditions (22) are satisfied. Hence in this case one can reproduce the Sudakov asymptotics (26), where \begin{equation} |qq'|\to |q^{(1)}q^{(2)}|\approx \frac{\vec Q^2}{4}, q^2\sim q^{\prime 2}\sim \kappa^2,~~ \alpha\to \alpha_sC_2 \end{equation} \section{Modifications due to confinement} As was discussed in the Introduction, in the kinematical region (1),(2) one usually assumes that all distances in the $n$--point function are small and therefore nonperturbative interaction is ineffective. Here we show that this is in general not true, and in particular for the DLA processes large distances appear naturally. Indeed, the distances between two straight--line trajectories (20), starting from some vertex, $l\equiv |z-z'|\sim 2 |q\tau -q'\tau'|$ may become large, $ l=0(\sqrt{\frac{|qq'|}{q^2q^{\prime 2}}})$, at the end points, $\tau=s\sim\frac{1}{q^2},$ $\tau'=s'\sim \frac{1}{q^{\prime 2}}$. The nonperturbative background $B_{\mu}$ produces the area--law term for $<W_\sigma>$ in (9), $<W_\sigma> \sim exp (-\sigma S_{\min})$, where $S_{\min}$ is the area of the minimal surface bounded by the line of trajectories, $\{ \xi^{(i)}(k)\}$, and the natural measure for influence of confinement is the value of $\sigma S_{min}$ for trajectories (20). For the triangle formed by the latter the area is $ S_{min}\sim \sqrt{(zz')^2}\sim |qq'|\tau\tau'$. It is clear that confinement starts to play role at the value $\tau\sim\tau'\sim\tau_0\equiv (\sigma|qq'|)^{-1/2}$ and at later proper time trajectories are no more straight lines. To estimate corrections to the trajectories one can add the term $\sigma S_{min}$ to the effective action (13). Differentiation of $\sigma S_{min}$ with respect to $\xi_\mu^{(i)}(k)$ yields an additional term on the r.h.s. of (12) $\Delta \xi \sim \sigma r^{\perp}$, where $r^{\perp}$ is the component of $r_\mu=2i(q_\mu\tau-\bar q_\mu\bar \tau)$ perpendicular to $\xi^{(i)}_\mu (k)$. Since $\tau_i\leq s_i,~~ s_i\sim \frac{1}{q^2_i}$, one obtains correction $O(\frac{\sigma}{q^2})$ to the leading term $q^{(i)}$ in the curly brackets of (12). Hence the condition of stability of Sudakov asymptotics with respect to the nonperturbative (confining) effects is $q^2_i\gg\sigma$. However there is nevertheless an additional contribution due to $\sigma S_{min}$ in the exponent, which obtains when one inserts in $S_{min}$ straight--line trajectories (20). With the conditions (22) one gets \begin{equation} <\exp(-\sigma S_{min}>_{eik} \cong exp (-\frac{2\sigma|q_iq_j|}{|q^2_iq^2_j|}) \end{equation} One can notice that the term in the exponent is not necessarily small and may substantially modify the double--logarithmic asymptotics. \section{Conclusion} The formalism described above is a direct development of the relativistic path integral, using the proper time, which we call Fock--Feynman--Schwinger Representation (FFSR) [7,8]. As in other versions of the path integral method [10,11] it is convenient to yield results already in the exponentiated form. We kept here the bispinor variables in the original Feynman form (the term $<W_\sigma>$ in (9)) which in principle allows to calculate spin--dependent effects also in an exponentiated form, however it was shown in the appendix that those effects are subdominant. One of the important advantages of the present formalism is that it allows to take into account NP effects also in the exponent, as it was demonstrated in section 5. There are possible many lines of development, including application to Drell--Yan and heavy--quark production, as well as finding connections with the factorization theorem, as it was done in a similar formalism in [14]. The author is grateful to A.B.Kaidalov for useful suggestions and to L.N.Lipatov for discussion of the results. \setcounter{equation}{0} \defA.\arabic{equation}{A.\arabic{equation}} \section{Appendix. Contribution of the magnetic moment term $\sigma_{\mu\nu}F_{\mu\nu}$.} We prove below that $(\sigma F)$ terms do not contribute to the DLA, but yield the subleading terms. The terms to be estimated are \begin{equation} <exp (g\int^s_0d\tau \sigma_{\rho\mu}F_{\rho\mu}(z(\tau))\Gamma_j exp (g\int^{s'}_0d\tau' \sigma_{\sigma\nu}F_{\sigma\nu}(z'(\tau')))> \end{equation} To illustrate the procedure it is enough to consider the $0(g^2)$ terms, namely \begin{equation} \sigma_{\rho\mu}\gamma_j\sigma_{\sigma\nu} \int^s_0d\tau \int^{s'}_0d\tau' D_{\rho\mu, \sigma\nu} \end{equation} where $D_{\rho\mu,\sigma\nu}\equiv <F_{\rho\mu}(z) F_{\sigma\nu}(z')>$ can be written perturbatively as $(u\equiv z-z')$ \begin{equation} D_{\rho\mu,\sigma\nu}=\frac{1}{2\pi^2} [\frac{\partial}{\partial u_\rho}(u_\sigma\delta_{\mu\nu}- u_\nu\delta_{\mu\sigma})+ \rho\sigma\leftrightarrow \mu\nu]\frac{1}{u^4} \end{equation} Commuting $\gamma$--matrices (A.2) can be rewritten as \begin{equation} \frac{2}{\pi^2}(4\gamma_\mu\delta_{\nu j}-\gamma_j\delta_{\mu\nu}) \int^s_0d\tau\int^{s'}_0d\tau'\frac{u_\mu u_\nu}{u^6} \end{equation} In the lowest saddle--point approximation for $\xi$, Eq. (12), one can replace according to (20) \begin{equation} u_\mu=2 i(q_\mu\tau+q'_\mu\tau') \end{equation} At the same time the lower limit of integration over $\tau,\tau'$ in (A.4) should be replaced by $\tau_{min}=\frac{1}{2|qq'|}$, and the integral can be estimated as \begin{equation} \frac{1}{2|qq'|}[\frac{q_\mu q_\nu}{2}I_1+ (q_\mu q'_\nu+q'_\mu q_\nu)I_2+ \frac{q'_\mu q'_\nu}{2}I_3] \end{equation} where $I_n, n=1,2,3$ grow at most logarithmically at large $|qq'|$ and all terms of the order $0(\frac{q^2}{|qq'|}, \frac{q'^2}{|qq'|})$ have been neglected. Thus the terms $\sigma F$ do not contribute to DLA in the order $0(g^2)$. One can follow the same procedure to higher orders and persuade oneself that leading powers of logarithms, $\alpha ln^2|qq'|$, do not appear.
\section{Introduction} The propagation of light through media with non-trivial optical properties is subject to three important effects: one is a variation in the light velocity with photon energy, namely a frequency-dependent refractive index, a second is a difference between the velocities of light with different polarizations, namely birefringence, and the third is a diffusive spread in the apparent velocity of light. We have argued previously that quantum-gravitational fluctuations in the space-time background may endow the conventional particle vacuum with such non-trivial optical properties, in particular a frequency-dependent refractive index~\cite{aemn}. We have also observed that such an effect may be severely constrained by careful observations of distant astrophysical objects whose emissions exhibit short time structures, such as Gamma-Ray Bursters (GRBs)~\cite{nature}. The possibility of quantum-gravitational birefringence has been raised within a loop approach to quantum gravity~\cite{pullin}. The purpose of this paper is to propose how quantum-gravitational diffusion may spread the arrival times of photons from distant sources, even if they have the same energies (frequencies), corresponding to stochastic fluctuations in the apparent velocity of light. An example of this phenomenon in a conventional optical medium, which illustrates clearly the intuition behind our proposal, has been discussed in~\cite{AMANDA}. Light propagating through ice may encounter air bubbles, which have a different refractive index and hence induce scattering and diffusion. We have argued previously~\cite{aemn} that foamy fluctuations in space-time generate a quantum-gravitational `aerogel' with an effective refractive index $\sim E/M_{QG}$ {\it in vacuo}, where $M_{QG}$ is some characteristic scale that may approach the Planck mass $M_P$. Now we argue that the quantum uncertainties in these foamy quantum-gravitational fluctuations in turn induce fluctuations in the refractive index that act as scattering centres analogous to the air bubbles of~\cite{AMANDA}. As illustrations of these quantum-gravitational fluctuations, one might consider foaming microscopic black holes that induce local fluctuations in the wave front close to their horizons, leading on larger distance scales to diffusive broadening of any light pulse. In this paper, we provide a mathematical formulation of this physical intuition, using the framework for quantum gravity adopted in~\cite{aemn}, namely that of Liouville strings~\cite{emn}, in which classical conformal string backgrounds are allowed to exhibit quantum space-time fluctuations. These cause departures from criticality and conformal symmetry that may be counterbalanced by introducing on the string world sheet a Liouville field with non-trivial dynamics. This Liouville field may be considered as a dynamical renormalization scale, and conformal invariance may be restored by dressing operators with appropriate Liouville factors~\cite{ddk}. Since the quantum-gravitational space-time fluctuations drive the string super-critical, the Liouville field has negative metric, and we identify it with target time~\cite{emn,kogan}. Interactions of the Liouville field with the conventional string degrees of freedom are controlled by the Zamolodchikov flow~\cite{zam} of the world-sheet renormalization-group equations. These incorporate the conventional Hamiltonian structure and $S$-matrix description of the scattering of low-energy particles, but also encode information on the energy-dependent interactions of the probe with non-local quantum-gravitational degrees of freedom. These fail to decouple~\cite{emn} from the low-energy matter in the presence of foamy singular metric fluctuations. Since a low-energy observer cannot detect such global modes by direct scattering experiments, integrating them out of the low-energy effective theory makes the low-energy system resemble an `open' quantum system interacting with an unobserved `environment' of these quantum-gravitational solitonic states. The outline of this paper is as follows. In Section 2, we review relevant features of this Liouville string framework~\cite{emn} for space-time foam, discussing in particular the formal basis for the appearance of stochastic fluctuations in theory space. Next, in section 3, we construct a specific realization of this non-critical string approach by considering quantum fluctuations of $D$-brane excitations in the vacuum. Then, in Section 4, we discuss the propagation of photons through this quantum $D$-brane foam. Section 5 shows how stochastic fluctuations occur in the context of $D$ branes, and exhibits the expected diffusion effect~\footnote{We do not find any evidence to support the suggestion of birefringence~\cite{pullin}.}. Finally, in Section 6, we discuss how this phenomenon may be constrained or detected by observations of distant astrophysical objects. \section{Relevant Aspects of Liouville String} In this Section we review briefly features of Liouville that play roles in our subsequent discussion. In this approach, one treats quantum fluctuations in the space-time background as deviations from a classical string background that is described by supercritical deformations of a conformal field theory on the world sheet. One may restore criticality by Liouville dressing of world-sheet model fields~\cite{ddk}. Specifically, consider a conformal $\sigma$ model, described by an action $S^*$ on the world sheet $\Sigma$, subject to non-conformal deformations $\int_{\Sigma}g^iV_id^2\sigma$, with $V_i$ appropriate vertex operators: \begin{equation} S_g = S^* + \int_{\Sigma}g^iV_id^2\sigma \label{sigma} \end{equation} The non-conformal nature of the couplings $g^i$ implies that their world-sheet renormal- \\ ization-group $\beta$ functions $\beta^i$ do not vanish. The following is the generic structure of such $\beta$ functions, close to such a fixed point $\{g^{i*} = 0\}$: \begin{equation} \beta^i = (h_i - 2)g^i + c^i_{jk}g^jg^k + o(g^3). \label{fixed} \end{equation} One cancels the deviation from criticality by world-sheet `gravitational' dressing that corresponds to defining renormalized couplings in a curved space. To ${\cal O}(g^2)$, one has~\cite{ddk,schmid,dorn}: \begin{equation} \lambda^i(t) = \lambda^ie^{\alpha_it} + \frac{\pi}{Q \pm 2\alpha_i}c^i_{jk}\lambda^j\lambda^kte^{\alpha_it} + O(\lambda^3), \qquad Q^2 = c-25 \nonumber \label{renorm} \end{equation} where $t$ is the zero mode of the Liouville field, $Q^2$ is the central charge deficit which is $\ge 0$ for the supercritical string case of interest here, and the $\alpha_i$ are gravitational anomalous dimensions: \begin{equation} \alpha_i(\alpha_i + Q) = h_i - 2 \qquad {\rm for} \qquad c \ge 25 \nonumber \label{anom} \end{equation} The supercriticality implies a Minkowskian signature for the Liouville field~\cite{aben}. enabling us to identify its zero mode with target time~\cite{emn,kogan}~\footnote{ We note for later use that~\cite{dorn}, in the gravitationally-dressed (Liouville) world-sheet theory, only the leading-order coefficients in the $\beta$ functions are renormalization-scheme independent. This implies that, when one identifies the Liouville field with target time, one loses general covariance in the foamy ground state. This should be thought of as a spontaneous breaking of the symmetry by quantum-gravitational fluctuations.}. After the renormalization (\ref{renorm}), the critical-string conformal invariance conditions corresponding to the vanishing of flat-space $\beta$ functions are replaced by~\cite{ddk,schmid,dorn,emn}: \begin{equation} {\ddot \lambda}^i + Q{\dot \lambda}^i = -\beta^i \qquad {\rm for} c \ge 25. \nonumber \label{neweq} \end{equation} The minus sign in front of the flat-world-sheet $\beta$ functions reflects the supercriticality of the string. \paragraph{} The propagation of non-relativistic light-particle modes was examined in~\cite{emn}, where a modification of the quantum Liouville equation for the density matrix was found, as proposed in~\cite{ehns}. The propagation of massless probes was examined in~\cite{aemn}, where it was found that the conventional relativistic energy-momentum dispersion relation is modified in non-critical Liouville strings, as a result of the interaction with the quantum-gravitational environment, as we now review briefly for the benefit of non-expert readers. \paragraph{} The first step is to observe that, in the case of interest, the non-criticality of the $\sigma$ model describing the effective theory is induced by the operator-product-expansion coefficients $c^i_{j{\hat k}}$ that express the interaction of a low-energy probe (latin indices) with quantum-gravitational modes (hatted latin indices). Hence, $Q^2 =0$ to lowest order, and the Liouville anomalous-dimension coefficients $\alpha _i$ are given simply by the magnitude of the spatial momentum of the massless probe: \begin{equation} \alpha _i = |{\overline k}| \label{spatial} \end{equation} We use this to rewrite (\ref{renorm}) approximately, to order $g^2$, as~\cite{aemn}: \begin{equation} \lambda ^i(t) \simeq g^i e^{(\alpha _i + \Delta \alpha _i )t} \label{shift2} \end{equation} where the shift $\Delta \alpha _i$ is given by: \begin{equation} \Delta \alpha _i \simeq \frac{\pi}{2\alpha _i}c^i_{j{\hat k}}g^{{\hat k}} \label{shift} \end{equation} We next make the generic hypothesis, which is supported by some explicit examples~\cite{kanti}, especially in the context of $D$ branes~\cite{dbranes}, as discussed in more detail below, that \begin{equation} \frac{1}{\alpha _i}c^i_{j{\hat k}} \sim \xi E/M_{QG}, \label{estimate} \end{equation} where $E$ is the energy scale of the low-energy probe, $\xi =\pm {\cal O}(1)$ and $M_{QG}$ is a characteristic of gravitational interactions, possibly to be identified with the Planck scale $M_P \sim 10^{19}$ GeV. We infer from (\ref{spatial}), (\ref{shift2}), (\ref{shift}) and (\ref{estimate}) a modified dispersion relation \begin{equation} E = |{\overline k}| \times \left( 1 + \xi E/M_{QG} \right) \label{dispersion} \end{equation} corresponding to an energy- (frequency-)dependent refractive index $\eta = 1 + {\cal O}(E / M_{QG})$. \paragraph{} It was observed in~\cite{nature} that such a quantum-gravitational effect could be probed by observations of distant astrophysical sources, such as GRBs, Active Galactic Nuclei (AGNs) or pulsars, and that these could place important constraints on models in which such a refractive index is suppressed only by a single power of $1/M_{QG}$, potentially with sensitivity to $M_{QG} \sim M_P$. Here we go further, remarking that quantum effects due to world-sheet topology fluctuations~\cite{emnquantum}, which arise in the context of first-quantized strings from the summation over genera, may cause observable stochastic effects that lead to a diffusive spread in apparent velocities even for photons of fixed energy (frequency). This would imply the spreading of an initial pulse, and place limitations on the resolutions of experimental measurements of distant astrophysical sources. \paragraph{} The basic motivation for this suggestion comes from the quantization of the target-space fields $\{ g^i \}$ in string theory, which arises from the summation over world-sheet topologies. In our case, this procedure leads to quantum fluctuations in the $\sigma$-model couplings $g^i$ felt by the propagating (low-energy) string particle modes. The probability density ${\cal P}[g]$ in the space of $\sigma$-model theories is given in leading approximation by: \begin{equation} {\cal P}[g] \sim {\rm exp}\left(-g^{\hat i} \frac{G_{{\hat i}{\hat j}}}{\Gamma}g^{\hat j} \right) \label{prob} \end{equation} where $G_{ij} \sim <V_iV_j>$ is the Zamolodchikov metric in theory space~\cite{zam}, for vertex operators $V_i$ corresponding to the couplings $g^i$. Formally, the non-trivial probability density (\ref{prob}) arises~\cite{emnquantum,lizzi} as a result of the requiring the cancellation of modular infinities against renormalization-group divergences associated with singular configurations in the sum over genera~\cite{emnquantum}. The width $\Gamma $ in (\ref{prob}) depends on some positive power of the string coupling constant $g_s$, and the precise form is model-dependent~\cite{lizzi,ms}. The quantum uncertainties $\delta g^{i}$ in the couplings are found by diagonalizing the basis in theory space $g^i$, for which knowledge of the Zamolodchikov metric is essential. \paragraph{} In the presence of fluctuations in the world-sheet topology of the string, which were not discussed in~\cite{aemn,nature}, the shifts (\ref{shift}) associated with the refractive index of the vacuum will fluctuate as: \begin{equation} |\delta (\Delta \alpha _i)| \sim |\frac{1}{2\alpha _i}c^i_{j{\hat k}}\delta g^{{\hat k}}| \label{dispersive} \end{equation} leading in turn to stochastic quantum fluctuations in the refractive index {\it in vacuo}. This in turn leads to a characteristic diffusive spread in the arrival times of photons identical energies, to which we return later. Before that, we first compute the Zamolodchikov metric and the related width $\Gamma$ (\ref{prob}) in a specific model for quantum-gravitational foam based on $D$ particles. \section{Fluctuations in $D$-particle Foam} We compute the quantum flucuations in (\ref{prob}) in the particular case where the gravitational degrees of freedom $g^{\hat k}$ are the collective coordinates and/or momenta of a system of $N$ $D$ particles~\cite{dbranes,ms}. As discussed in the literature~\cite{kmw,emndbrane,lizzi,ms} such collective coordinates can be described by operators characterizing the recoil induced by the scattering of string matter off the $D$-particle background. The corresponding $\sigma$-model deformation is: \begin{equation} Y_i^{ab}(x^0)= \ell_s(Y_i^{ab}\ell_s\epsilon+U_i^{ab}x^0)\epsilon\Theta_\epsilon(x^0) \label{gal} \end{equation} where the spatial coordinates of the $D$ particle are identified with the couplings $Y_i$ and the $U_i$ correspond to their Galilean recoil velocities. These coordinates are to be regarded as $\sigma$-model couplings $g^i$ in the sense of the previous section. The parameter $\epsilon\to0^+$ regulates the ambiguous value of $\Theta(s)$ at $s=0$, which ensures that the $D$-particle system starts moving only at the time $x^0=0$~\footnote{Throughout this paper, we consider the $\sigma$-model `renormalized' velocity~\cite{emndbrane,lizzi,ms}, which is related to the `bare' velocity $U_{Bi}$ by $U_i = U_{Bi}/\epsilon $. It is $U_i$ which is an exactly-marginal deformation in a world-sheet renormalization group sense, and therefore corresponds to uniform $D$-particle motion for times $t > 0$.}. It is related to the world-sheet ultraviolet cutoff scale $\Lambda$ (measured in units of the world-sheet size $\Sigma$) by $\epsilon^{-2}=2\ell_s^2\log\Lambda$, where $\ell _s$ is the fundamental string length. For finite $\epsilon$, the operators $Y_i, U_i$ each have an anomalous dimension $-\frac12|\epsilon|^2<0$~\cite{kmw}, and thus lead to a relevant deformation of $S^*$. The corresponding renormalization-group equations are~\cite{lizzi}: \begin{equation} dY_i^{ab}/dt=U_i^{ab}, \; dU_i^{ab}/dt=0, \label{rgemotion} \end{equation} which are the Galilean equations of motion for the $D$ particles, if we identify the time with the world-sheet scale: $t=\ell_s\log\Lambda$. \paragraph{} The natural geometry on the moduli space $\cal M$ of deformed conformal field theories described by the above recoil operators (\ref{gal}) is given by the following Zamolodchikov metric $G_{ab;cd}^{ij} = \langle V_{ab}^i V_{cd}^j \rangle$~\cite{ms}: \begin{eqnarray} G_{ab;cd}^{ij}&=&\frac{4g_s^2}{\ell_s^2}\left[\delta^{ij}\,I_N\otimes I_N-\frac{g_s^2}6\left\{I_N\otimes\left(U^iU^j+U^jU^i\right)+ U^i\otimes U^j\right.\right.\nonumber\\& &\biggl.\left.+\,U^j\otimes U^i+\left(U^iU^j+U^jU^i\right)\otimes I_N\right\}\biggr]_{db;ca}+{\cal O}\left(g_s^6\right) , \label{Gfinal}\end{eqnarray} where $I_N$ is the $N\times N$ identity matrix and we have renormalized $g_s$ to the time-independent coupling $g_s/|\epsilon|\ell_s$. The canonical momentum $P_{ab}^i$ of the $D$-particle system is given in the Schr\"odinger picture by the expectation value of $-i\delta/\delta Y_i^{ab}$ evaluated in a $\sigma$ model deformed by the operator $V_{ab}$, i.e., $P_{ab}^i \equiv \langle V_{ab}^i\rangle$~\cite{ms}: \begin{equation} P_{ab}^i=\frac{8g_s^2}{\ell_s}\left[U^i-\frac{g_s^2}6 \left(U_k^2U^i+U_kU^iU^k+U^iU_k^2\right)\right]_{ba}+{\cal O}\left(g_s^6\right) \label{canmom}\end{equation} which coincides with the contravariant velocity $P_{ab}^i=\ell_sG_{ab;cd}^{ij}\dot Y_j^{cd}$ on $\cal M$. \paragraph{} In agreement with general arguments~\cite{emn}, we note that the moduli space dynamics can be derived~\cite{ms} from a Lagrangian of the form \begin{equation} {\cal L} = -\frac{\ell_s}2\dot Y_i^{ab}G_{ab;cd}^{ij}\dot Y_j^{cd}-{\cal C} \label{lagrangian} \end{equation} which coincides~\cite{ms} to leading order with the non-Abelian Born-Infeld effective action~\cite{tseytlin} for the target-space $D$-particle dynamics: \begin{equation} {\cal L}_{\rm NBI}=\frac1{\ell_s\bar g_s}\,Tr~{\rm Sym}\,\sqrt{\det_{M,N}\left[\eta_{MN}\,I_N+\ell_s^2\bar g_s^2\,F_{MN}\right]} \label{NBIaction} \end{equation} where the trace $Tr$ is taken over $U(N)$ group indices, \begin{equation} {\rm Sym}(M_1,\dots,M_n) \equiv \frac1{n!}\sum_{\pi\in S_n}M_{\pi_1}\cdots M_{\pi_n} \label{product} \end{equation} is the symmetrized matrix product, and the components of the dimensionally-reduced field-strength tensor are given by $F_{0i}=\frac1{\ell_s^2}\dot{\bar Y}_{\!i}$ and $F_{ij}=\frac{\bar g_s}{\ell_s^4}[\bar Y_i,\bar Y_j]$. In the Abelian reduction to the case of a single $D$ particle, the Lagrangian (\ref{NBIaction}) reduces to the usual one describing the free relativistic motion of a massive particle. The leading order $F^2$ term in the expansion of (\ref{NBIaction}) is just the usual Yang-Mills Lagrangian. \paragraph{} The formalism described in~\cite{ms} is a non-trivial application of the theory of Liouville string and logarithmic operators. We note, moreover, that the derivation of the Lagrangian (\ref{NBIaction}) was made `off-shell', in other words we have compared generalized momenta in theory space with those derived from the Born-Infeld lagrangian. The equivalence between the two formalisms extends beyond the equations of motion, which are the conformal-invariance conditions of the $\sigma$ model. This is a central aspect of the recoil approach: there are deviations from the usual equations of motion for low-energy modes, because conformal invariance is violated by the recoil process~\cite{kmw,emndbrane}, and the Zamolodchikov world-sheet renormalization-group flow provides an off-shell treatment of this recoil problem. \paragraph{} The leading contributions to the quantum fluctuations in $\cal M$ arise from pinched annulus diagrams in the summation over world-sheet genera of the $\sigma$ model~\cite{emnquantum}. Symbolically, these lead to contributions in the genus expansion of the form \unitlength=1.00mm \linethickness{0.4pt} \begin{equation} \begin{picture}(70.00,10.00) \LARGE \put(5.00,2.00){\circle*{100.00}} \put(8.00,2.00){\circle{3.00}} \put(15.00,2.00){\makebox(0,0)[l]{$+$}} \put(28.00,2.00){\circle*{100.00}} \put(31.00,2.00){\circle{3.00}} \put(28.00,5.00){\circle{3.00}} \put(38.00,2.00){\makebox(0,0)[l]{$+$}} \put(48.00,2.00){\circle{3.00}} \put(51.00,2.00){\circle*{100.00}} \put(54.00,2.00){\circle{3.00}} \put(51.00,5.00){\circle{3.00}} \put(61.00,2.00){\makebox(0,0)[l]{$+~\dots$}} \end{picture} \label{pinch}\end{equation} consisting of thin tubes of width $\delta\rightarrow 0$ (world-sheet wormholes) attached to the world-sheet surface $\Sigma$. The attachment of each tube corresponds to inserting a bilocal pair $V_{ab}^i(s)V_{cd}^j(s')$ on the boundary $\partial\Sigma$, with interaction strength $g_s^2$, and computing the string propagator along the thin tubes. There are modular divergences of the form $\log\delta$, which should be identified with world-sheet divergences at lower genera~\cite{emn}, so we set \begin{equation} \log\delta=2g_s^\eta\log\Lambda=\frac1{\ell_s^2}g_s^\eta\epsilon^{-2} \label{fs} \end{equation} The exponent $\eta \neq 0$ in this Fischler-Susskind-like~\cite{fis} relation, which allows one to cancel logarithmic modular divergences by relating the strip widths $\delta$ to the worldsheet ultraviolet scale $\Lambda$, arises because the relation (\ref{fs}) is induced by the string loop expansion. As we have argued in~\cite{ms}, the case $\eta\neq0$ allows direct comparison of our results with other results in the string literature, based on alternative approaches~\cite{liyoneya}. However, we do not determine $\eta$ here, since we only consider string interactions between $D$ branes. Considering brane exchanges between the system of $D$ branes, may enable the value of $\eta$ to be fixed (see also below), but this point is not important for the present analysis. \paragraph{} One effect of the dilute gas of world-sheet wormholes is to exponentiate the bilocal operator, leading to a change in the $\sigma$-model action~\cite{emnquantum,lizzi}. This contribution can be cast into the form of a local action by rewriting it as a Gaussian functional integral over wormhole parameters $\rho_i^{ab}$, as described in the previous section, and we arrive finally at~\cite{ms}: \begin{equation} \sum_{\rm genera}Z_N[Y]\simeq\left\langle\int_{\cal M}D\rho~e^{-\rho_i^{ab}G_{ab;cd}^{ij}\rho_j^{cd}/2|\epsilon|^2\ell_s^2 g_s^2\log\delta}~W[\partial\Sigma;Y+\rho]\right\rangle_0 \label{genusexp}\end{equation} We see from (\ref{genusexp}) that the effect of this resummation over pinched genera is to induce quantum fluctuations of the solitonic background, giving a statistical Gaussian spread to the $D$-particle couplings. Note that the width of the Gaussian distribution in (\ref{genusexp}), which we identify as the wave function of the system of $D$ particles \cite{emn}, is time-independent, and represents not the spread in time of a wave packet on $\cal M$, but rather the true quantum fluctuations of the $D$-particle coordinates. \paragraph{} The corresponding spatial uncertainties can be found by diagonalizing the Zamolodchikov metric (\ref{Gfinal}), as was done in~\cite{ms} and will not be repeated here. For the case of a single $D$ particle: $a=b$, one arrives at the uncertainties: \begin{equation} \left|\Delta X^i_{aa}\right|\equiv \Delta Y^i =\ell_sg_s^{\eta/2}\left(1+\frac{g_s^2}{8\pi^3}\,u^2\, \delta^{i,1}\right)+{\cal O}\left(g_s^4\right)\geq\ell_s\,g_s^{\eta/2} \label{minimum}\end{equation} for the individual $D$-particle coordinates. For $\eta=0$, the minimal length in (\ref{minimum}) coincides with the standard string smearing~\cite{ven}, whereas for $\eta=\frac23$ it matches the form of $\ell_{\rm P}$ which arises from the kinematical properties of $D$ particles~\cite{liyoneya}. A value $\eta\neq0$ is more natural, because the modular divergences should be small for weakly-interacting strings. We note that the uncertainty (\ref{minimum}) is {\it time-independent}~\cite{ms}, which is important for experimental tests of the phenomenon, as we discuss below. The coordinate uncertainties for $a\ne b$ are responsible for the emergence of a true non-commutative structure of quantum space-time, and represent the genuine non-Abelian characteristics of multi-$D$-particle dynamics, but we are not concerned with such a case here. \section{Propagation of Photons in a $D$-Particle Foam Background} After this general discussion, we now discuss the propagation of photons in the background of multiple $D$ particles, taking into account the recoil of the latter due to the scattering of the photons. This is immediate in the context of~\cite{ms}: one simply adds to the argument of the determinant (\ref{NBIaction}) an Abelian $U(1)$ field strength $f_{MN}=\partial_M a_N - \partial _N a_M$, where $a_M$ denotes the $U(1)$ electromagnetic potential of the photon fields. The corresponding Born-Infeld effective action is considered in four-dimensional space time. We adopt the conventional view point that photons interacting with $D$ particles may be represented by adding to the $\sigma$-model action $S_\sigma$ on the (open) world sheet an electromagnetic potential background as a boundary term~\cite{abouel}: \begin{equation} S_\sigma \ni \frac{e}{c}\int_{\partial \Sigma} d\tau a_M (X) \partial _\tau X^M \label{abouel} \end{equation} in a standard (Neumann) open-string notation~\footnote{The target-space Born-Infeld action, including photons interacting with $D$ particles, may {\it alternatively } be considered as the target-space action of a three brane, i.e. a solitonic object in string theory with four coordinates obeying Neumann boundary conditions and the remainder Dirichlet boundary conditions~\cite{kmw}. It is possible that the above Neumann picture can be obtained from the Dirichlet one by a world-sheet $T$-duality transformation, but there are problems with this transformation at a quantum level~\cite{otto}. We restrict ourselves here to the Neumann string picture.}. \paragraph{} The non-relativistic heavy $D$-particle~\cite{ms} corresponds to a time-dependent background (\ref{gal}), whilst the photon propagator depends on the full four-dimensional space time. It is important to note, though, that in our approach the coupled $D$-particle-photon system is {\it out of equilibrium}, due to the recoil process and its associated distortion of the surrounding space time. This is reflected in the fact that the resulting backgrounds do not satisfy the classical equations of motion, as was mentioned earlier in the general context of the departure from conformal symmetry. Moreover~\cite{kanti}, the recoil curves the surrounding space time, since it induces - via Liouville dressing and the identification~\cite{emn,kanti} of the Liouville mode with the target time $t$ - graviton excitations for the string, which in a $\sigma$-model framework correspond to graviton backgrounds with non-trivial off-diagonal elements: \begin{equation} G_{0i} \sim \epsilon ^2 U_i t \Theta (t) \label{graviton} \end{equation} where $U_i$ is the velocity of the recoiling $D$ particle. \paragraph{} The element (\ref{graviton}) is obviously not Lorentz covariant, reflecting the spontaneous breaking of this symmetry by the ground state of the string. The splitting between the quantum-gravitational `medium' and the propagating particle subsystem is not possible in a Lorentz-invariant way, and there is no formal reason that this should be so. On the contrary, spontaneous violation of Lorentz symmetry is generic in Liouville strings~\cite{aben,emn}. Thus we may expect the interaction of such graviton modes with the photons to lead to a modification of the photon dispersion relation, analogous to non-Lorentz-covariant, e.g., thermal, effects in conventional media. A key difference between quantum-gravitational effects and those in conventional media~\cite{nature} is that the former increase with the energy of a probe, whilst the latter attenuate with increasing energy. \paragraph{} Gravitational interactions may be incorporated in the Born - Infeld lagrangian (\ref{NBIaction}) by replacing $\eta _{MN} \rightarrow G_{MN}$ and using the non-covariant expression (\ref{graviton}) when calculating amplitudes. Thus the complete off-shell Born - Infeld Lagrangian for the interaction of photonic matter with $D$ particles is: \begin{equation} {\cal L}_{\rm NBI}=\frac1{\ell_s\bar g_s}\,Tr~{\rm Sym}\,\sqrt{\det_{M,N}\left[G_{MN}\,I_N+ \ell _s^2 \frac{e^2}{c}f_{MN}(a)\,I_N + \ell_s^2\bar g_s^2\,F_{MN}\right]} \label{NBIaction2} \end{equation} which is the basis for our subsequent discussion. \paragraph{} We first re-examine the possible order of magnitude of the frequency-dependent refarctive index (\ref{shift}) induced by such effects, concentrating first on the case of Abelian (single $D$-particle) defects. The terms in the effective action (\ref{NBIaction}) that are relevant for the modification (\ref{shift}) of the photon dispersion relation arise from the three-point function terms in (\ref{renorm}) that involve two photon excitations and one induced-graviton excitation (\ref{graviton}). The appropriate term in a derivative expansion of the Born-Infeld action is: \begin{equation} {\cal L}^{photons,graviton}_{NBI} \ni f_{MN}(a) G_{NA} f_{AM} \sim f_{ij}(a) U_j f_{0i} \epsilon ^2 t \label{effact} \end{equation} where latin indices are spatial, and we used (\ref{graviton}). Terms that are similar in order of magnitude, but not in tensorial structure, can also be obtained by combining the $f_{MN}^2$ terms with the determinant $\sqrt{-{\rm det}(G_{MN})}$ of the target-space metric~\footnote{It is clear that Lorentz-invariant interactions of powers of $f^2$ terms with generic ${\rm Tr}(F^{2n})$ terms in the derivative expansion of (\ref{NBIaction2}) do not affect the photon dispersion relation. It is only {\it non-covariant} terms, such as the the induced-graviton-photon interactions considered above, that do so.}. An appropriate order-of-magnitude estimate of the recoil velocity is: \begin{equation} U_i \sim g_s |{\overline k}|/M_P \label{recoilvel} \end{equation} reflecting energy-momentum conservation in the scattering process~\cite{emndbrane,lizzi,ms}, and assuming that the momentum of the recoiling $D$ particle is of the same order as the energy-momentum of the photon. Working at times $t$ after the recoil that are sufficiently large, we may assume~\cite{kmw,emndbrane,lizzi,ms} that \begin{equation} \epsilon ^{-2} \sim t \label{timeapprx} \end{equation} on average. Using now for the graviton a linear approximation about flat Minkowski space $G_{MN} \sim \eta_{MN} + h_{MN}$, which allows one to consider a simple Fourier momentum-space decomposition of the pertinent amplitudes, we conclude that the effective photon-graviton interactions in (\ref{effact}) are of order ${\cal O}(g_s E |{\overline k}|^2/M_P )$. Such estimates apply to the string amplitudes $c^i_{j{\hat k}}g^{{\hat k}} $ in the shift (\ref{shift}) in the single-defect case, which therefore becomes, to leading order in a low-energy approximation: \begin{equation} \Delta \alpha _i ={\cal O}(-g_s \frac{E |{\overline k}|} {M_P})) \label{shiftabel} \end{equation} Using (\ref{shift2}), we see immediately that this leads to a modified dispersion relation for the photon: \begin{equation} E = |{\overline k}| + {\cal O}\left(-g_s\frac{E |{\overline k}|}{M_P} )\right) \label{disp} \end{equation} The refractive index is then determined from the photon group velocity~\cite{nature}: \begin{equation} v(|{\overline k}|) \equiv \frac{\partial E}{\partial |{\overline k}|} \simeq 1 - {\cal O}({2g_s |{\overline k}| \over M_P}) \simeq 1 - {\cal O}({2g_s E \over M_P}) \label{refind} \end{equation} The sign of the refractive index is determined by the fact that the Born-Infeld action underlying the above analysis prevents superluminal propagation. This sign was not determined in our previous discussion~\cite{nature}. As discussed there, such a variation in the velocity of light will cause a spread in the arrival times of pulses of photons, according to their energy~\cite{nature}: \begin{equation} \Delta t \sim {\xi L E \over M_P} \label{figuremerit} \end{equation} The possibility of observing experimentally such a shift using distant astrophysical sources appears conceivable, as dicussed in~\cite{nature} and reviewed in the last section of this paper. \section{Stochastic Fluctuations in the Apparent Velocity of Light} We now extend the above discussion to include an estimate of the stochastic fluctuations in the apparent velocity of light, and hence the refractive index, due to the summation over higher-genus world-sheet topologies. The general theory of this summation in the context of Liouville string, discussed in section 2, leads to stochastic fluctuations in the collective coordinates of the $D$ particles discussed in section 3~\cite{ms}: \begin{equation} \delta Y^i \ge \ell _s g_s^{\eta/2}, \label{deltay} \end{equation} The conventional Heisenberg uncertainty relation between the coordinates $Y^i$ and the corresponding canonical momenta $P_j$ has been shown to take the following form for $D$ particles, after the summation over genera~\cite{ms}: \begin{equation} \delta Y^i \delta P_j \ge 2 g_s^{1 + \eta/2} \delta^{i}_j \label{heisenberg} \end{equation} Saturating the bound (\ref{deltay}) in (\ref{heisenberg}), we obtain the following estimate of the uncertainty in the associated collective canonical momenta of the $D$ particles: \begin{equation} \delta P_i \simeq \frac{2g_s}{\ell _s} \label{momentumvar} \end{equation} We note that this estimate of the uncertainty in the collective momentum is independent of the exponent $\eta$, which, as already noted, would be required to take the value $\eta = 2/3$ in order to match results on $D$ particles in conformal string theory. \paragraph{} In our interpretation, the fluctuations arising from the summation over genera lead to a statistical superposition of theories with different values of the couplings $g^{{\hat j}}$. This `stochastic environment' is characterized by the Gaussian form (\ref{prob}) of the associated probability distribution in the space of $\sigma$-model backgrounds. A photon of given energy $E$ propagating through the fluctuating quantum-gravitational medium is subject to stochastic fluctuations in its velocity, which may be obtained from Liouville interactions of the form (\ref{dispersive}) by endowing the generalized $\sigma$-model coordinates $\{ g^{\hat k} \}$ with the fluctuations (\ref{deltay}) found for the collective coordinates $Y^i$ of the $D$-particle foam. \paragraph{} To see how this effect indeed induces stochastic fluctuations in the refractive index of the photon, one first calculates the amplitudes $c^i_{j{\hat k}}g^{{\hat k}}$ appearing in (\ref{dispersive}) using the Born-Infeld Lagrangian (\ref{NBIaction2}), calculated on world sheets with disc topology. Using (\ref{timeapprx}), one sees that the leading terms in a derivative expansion are the ones given in (\ref{effact}). The summation over world-sheet genera leads to stochastic fluctuations in $U_i$ (\ref{recoilvel}), which are given by (\ref{momentumvar}): \begin{equation} \delta U_i = {g_s \delta P_i \over M_P} \sim 2g_s^2 \label{uvel} \end{equation} These stochastic fluctuations in $U_i$ generate fluctuations in the overall proportionality coefficient of the Maxwell terms in the effective action for the photon: \begin{equation} \delta {\cal L}^{photons,graviton}_{NBI} \ni f_{MN}(a) \left( \delta G_{NA}\right) f_{AM} \sim f_{ij}(a) \left( \delta U_j \right) f_{0i} \epsilon ^2 t \label{effact2} \end{equation} whose Fourier transform is ${\cal O} (2 g_s^2 E |{\overline k}| /M_P) $. This does not itself affect the velocity of the photon, but simply renormalizes the energy scale. \paragraph{} However, when one proceeds to higher orders, one picks up contributions that lead to stochastic fluctuations in the refractive index. To see this, we concentrate on terms of quadratic order in $U_i$, stemming from terms in a derivative expansion of the Born-Infeld lagrangian (\ref{NBIaction2}) that are of the generic form $f_{MN}^2 {\rm Tr} F_{AB}^2 $, where the non-Abelian field strength $F_{MN}^{ab}$ is calculated by considering $Y_i^{ab}$ (\ref{gal}) formally as a `gauge potential'. For long times of order (\ref{timeapprx}), the only non-vanishing components of $F_{AB}$ are $F_{0i} \sim U_i$ which lead to stochastic fluctuations in the quadratic terms of order \begin{equation} f_{MN}^2 \left( \delta U_i^2\right) = 4 g_s^2 f_{MN}^2 U_i \label{stochvel} \end{equation} Thus, in order to discuss the leading-order effects induced by the stochastic fluctuations of our $D$-particle foam, as implied by the summation over world-sheet topologies, we should consider corrections to the Liouville-dressed couplings that go beyond quadratic order in the $\sigma$-model couplings $\{ g^{{\hat i}} \}$. Such corrections have been studied in~\cite{dorn}, and are not given here explicitly. It is sufficient for our purposes to point out that some are proportional to four-point amplitudes, $c^i_{jkl}$ divided by `energy denominators' $\alpha_j+\alpha _k$: $c^i_{jkl}/(\alpha_i + \alpha_j)$, in the limit of vanishingly small central-charge deficits $Q$ that are of interest to us, whereas others are of the form $c^i_{jm} c^m_{kl}$ divided by terms of order $\alpha_i^2$. The three- and four-point amplitudes are computed from the Born-Infeld lagrangian (\ref{NBIaction2}) above~\footnote{We again reminder the reader that, since such higher-order corrections are world-sheet renormalization-scheme dependent~\cite{dorn}, this reflects the (spontaneous) breaking of general covariance by our foamy ground-state when the Liouville field is identified with target time~\cite{emn}.}. Thus, for long times of order (\ref{timeapprx}), and using (\ref{recoilvel}) for $U_i$, one finds the following estimate for the corresponding stochastic fluctuations in (\ref{shift}): \begin{equation} \delta \left(\Delta \alpha _i\right) = {4 g_s^2 E^2 \over M_P} \label{stochfluct} \end{equation} Correspondingly, we find fluctuations in the velocity of light in the quantum-gravitational medium of order $\delta c \sim 8g_s^2E/M_P$, motivating the following parametrization in the stochastic spread in photon arrival times: \begin{equation} \left(\delta \Delta t \right) \sim {L E \over \Lambda_{QG}} \label{error} \end{equation} where $\Lambda_{QG} \sim M_P / 8 g_s^2$. We emphasize that, in contrast to the variation (\ref{figuremerit}) in the refractive index, which refers to photons of different energy, the fluctuation (\ref{error}) characterizes the statistical spread in the velocity of photons {\it of the same energy}. \section{Observational tests} The most important signatures of the refractive index and the stochastic fluctuation in the velocity of light that we find are that they increase {\it linearly} with the photon energy (frequency). This means that they can, in principle, easily be distinguished from more conventional medium effects, that attenuate with increasing energy. The effects scale inversely with some quantum-gravitational scale characteristic of strings and $D$ branes. We are not in a position to estimate it numerically, but we expect it to be within a few orders of magnitude of $M_P \sim 10^{19}$ GeV. In standard string theories one has $\eta=2/3$ and $g_s^2/4\pi \sim 1/20$, but the latter may well be modified in a more realistic theory. In principle, one could even envisage using upper limits on (measurements of) the rate of broadening of a radiation spike of definite energy (frequency) to constrain (measure) $g_s$. \paragraph{} We conclude this paper by mentioning some possible observational tests of these ideas. As has been emphasized previously~\cite{nature}, the figure of merit for constraining the possibility of an energy- (frequency-)dependent refractive index {\it in vacuo} is the combination $L \times \Delta E / \Delta t$, where $L$ is the distance of a source, $\Delta E$ is the range of photon energies studied, and $\Delta t$ is the observational sensitivity to differences in light-travel times. The latter is limited by the durations of pulses produced by the source as well as by the resolution of the detector. The corresponding figure of merit for testing the new possibility advanced in this paper, namely a stochastic spread in light-travel times for different photons with the same energy $E$, is simply $L \times E / \Delta t$. In practice, when comparing photons of different energies, $\Delta E$ is often dominated by the highest photon energy $E$ that is measured, so that $\Delta E \sim E$, and the two figures of merit are essentially equivalent. \paragraph{} Astrophysical sources offer the largest figures of merit, and the most promising that have been considered include GRBs~\cite{nature}, AGNs~\cite{biller} and pulsars~\cite{crab}. These have all been used already to constrain the refractive-index parameter $M_{QG}$, and offer similar sensitivities to the stochastic-spread parameter $\Lambda_{QG}$. As mentioned earlier in this paper, the $D$-brane Born-Infeld analysis indicates that higher-energy photons should be {\it retarded} relative to lower-energy photons, rather than advanced, and their arrival times should be more spread out. In the Table below, we list some of the sources that have been considered, and the sensitivities (limits) that have been obtained. For completeness, we have also indicated the sensitivity that might be obtainable from a detailed analysis of the recent GRB 990123. We see from the Table that $M_{QG}$ cannot be much smaller than the Planck scale $M_P \sim 10^{19}$ GeV, and that some of these astrophysical sources may already providing sensitivities to $M_{QG}, \Lambda_{QG} \sim 10^{19}$~GeV. This provides additional fundamental-physics motivation for such $\gamma$-ray observatories as AMS~\cite{AMS} and GLAST~\cite{GLAST}. \paragraph{} Other probes of the signatures of quantum gravity that might be provided by the unorthodox photon propagation proposed here might be possible using laboratory experiments, for example those testing quantum optics and searching for gravitational waves, but we do not explore these possibilities further in this paper~\footnote{However, we do observe that the considerations of~\cite{interfere} concerning interferometric signatures are not applicable in our framework.}. However, we think that the discussion given here demonstrates amply the possibility that at least some quantum-gravity ideas may be accessible to experimental test, and need not remain for ever in the realm of mathematical speculation. \begin{center} {\bf Table: Observational Sensitivities and Limits on $M_{QG}, \Lambda_{QG}$} \end{center} \begin{figure}[h] \begin{center} \begin{tabular}{|c||c|c|c|c|} \hline Source & Distance & $E$ & $\Delta t$ & Sensitivity (Limit) \\ \hline GRB 920229~\cite{nature} & 3000 Mpc (?) & 200 keV & $10^{-2}$ s & $10^{16}$ GeV (?) \\ \hline GRB 980425~\cite{nature} & 40 Mpc & 1.8 MeV & $10^{-3}$ s (?) & $10^{16}$ GeV (?) \\ \hline GRB 920925c~\cite{nature} & 40 Mpc (?) & 200 TeV (?) & 200 s & $10^{19}$ GeV (?) \\ \hline Mrk 421~\cite{biller} & 100 Mpc & 2 TeV & 280 s & $> 4 \times 10^{16}$ GeV \\ \hline Crab pulsar~\cite{crab} & 2.2 kpc & 2 GeV & 0.35 ms & $> 1.8 \times 10^{15}$ GeV \\ \hline GRB 990123 & 5000 Mpc & 4 MeV & 1 s (?) & $3 \times 10^{14}$ GeV (?) \\ \hline \end{tabular} \end{center} \end{figure} {\it The question marks in the Table indicate uncertain inputs. Hard limits are indicated by inequality signs.} \vspace{0.5cm} \noindent {\bf Acknowledgements} We thank Phil Allport, Gianni Amelino-Camelia, Kostas Farakos, Hans Hofer, Subir Sarkar and S.C.C. Ting for discussions and interest. The work of D.V.N. is supported in part by D.O.E. Grant DE-FG03-95-ER-40917. \vspace{0.05cm}
\section{Introduction} \noindent Optical Second Harmonic Generation (SHG) has been proven to be a very useful technique for the investigation of ferromagnetism at surfaces. The obvious question is if this technique can also yield some new information in the case of more general spin configurations, such as antiferromagnetic (AF) ordering. An experimental answer to this question has been provided by Fiebig {\em et al.} \cite {ref18}, who obtained a pronounced optical contrast from AF 180$^\circ$ domains of rhombohedral bulk Cr$_2$O$_3$. The authors attributed this contrast to the interference of magnetic and electric dipole contributions, the latter being present only below the N\'eel temperature. Since it is known that, in {\em cubic} materials, within the electric dipole approximation, optical SHG originates only from surfaces, interfaces, or thin films, an important question is if SHG is also sensitive to antiferromagnetism at surfaces of cubic antiferromagnets. In this paper, we will show that the surface of a cubic material can lower the symmetry of an AF fcc crystal (two-sublattice antiferromagnet) in a way similar to the trigonal distortion in a four sublattice antiferromagnet Cr$_2$O$_3$. Besides, even the imaging of AF {\em domains} is possible also for many cubic materials that exhibit unit-cell doubling. The first theoretical explanation of {\em linear} magneto-optic effects in ferromagnets has been given by Argyres \cite{ref25} in the 50s. He used linear response theory for current-current correlation functions. His microscopic explanation was already based on the combination of spin-orbit and exchange coupling. Experimental techniques for the detection of AF domain {\em walls} using linear optics in some special geometries were elaborated a few years later~\cite {ref7zref27}. The {\em interior} of the domains has been visualized in piezoelectric AF crystals using a linear magneto-optical effect \cite {ref4zref17}. However, linear optical experiments suffer from mixing the desired signal with a contribution from other linear effects, such as birefringence or dichroism. A review of linear optical experimental methods for the investigation of AF domains is given by Dillon~\cite{ref17}. The observation of domain structure in antiferromagnets is more complicated than in ferromagnetic materials since the reduction of the spatial symmetry is, unlike for ferromagnets, not linked to an imbalance in the occupation of majority and minority spin states. On the basis of group theoretical considerations, Brown {\em et al.} \cite {ref28} proposed the use of linear optical effects, namely gyrotropic birefringence, for the observation of AF domains related to each other by the space-inversion operation. A theoretical review of effects found by a group-theoretical approach is presented by Eremenko and Kharchenko \cite {ref13}. They performed a comprehensive study of linear optical effects for various AF materials. Another effect proposed recently by Dzyaloshinskii {\em et. al.} \cite{ref41} gives a possibility to detect antiferromagnetism taking advantage from optical path differences from antiferromagnetically coupled but intrinsically ferromagnetic planes. {\em Nonlinear} optics exhibits an additional degree of freedom, since its elementary process involves three photons instead of two in linear optics. For that reason, some authors, e.g. Fr\"ohlich \cite{ref40} suggested the application of nonlinear optics even for k-selective spectroscopy, since multi-photon phenomena allow for the ``scanning'' of a small part of the Brillouin zone, at least for semiconductors. Recently, non-linear optics has attracted more and more attention for the investigation of magnetism due to its enhanced sensitivity to twodimensional {\em ferromagnetism} \cite {ref53}. The magnetic effects are usually much stronger than in linear optics (rotations up to 90$^\circ$, pronounced spin polarized quantum well state oscillations \cite{ref64,ref65}, magnetic contrasts close to 100$\%$) \cite {ref22,pustu}. An example of ferromagnetic effects measurable only by SHG deals with the existence of surface magnetism in very thin films of Fe/Cu(001) and is given in Ref. \cite {ref31}. Nonlinear optical effects were invoked to explain the behavior of lasers in magnetic fields \cite{ref55}, to investigate high temperature superconductors \cite{ref67,ref30}, and to study structures composed from alternately ferro- and antiferromagnetically ordered thin films \cite{ref34}. One of the first theoretical investigations of the possibility to apply nonlinear optics to {\em antiferromagnetism} was performed by Kielich and Zawodny \cite{Kieliszek}. However, the first experiments concerning the detection of the AF domains in materials such as Cr$_2$O$_3$ were carried out only recently~\cite {ref4,ref12}. Already in the 70s, it has been proposed \cite{ref10} that experimental studies of dc magnetic and electric field-induced SHG could become an effective method of determining the crystal structure of solids, the symmetry of which cannot be investigated by other methods. Extending this idea towards surface crystallography provides us with a new technique for determining the spin configuration in a given surface structure. In turn, it permits to use a known magnetic configuration for the determination of the surface structure. All the mentioned effects are more difficult or even impossible to obtain in linear optics, and moreover other linear methods like neutron scattering have difficulties to probe AF spin configurations. The nonlinear magneto-optical susceptibility tensor $\chi^{(2\omega)}_{el}$ (the source for SHG within the electric dipole approximation) has predominantly been investigated from the symmetry point of view. A classification following this approach, with tensors of a rank up to six, has been performed by Lyubchanskii {\em et al.} \cite {ref19,ref20,ref21,ref22,ref23}. In Ref. \cite {ref22} the authors include the magnetization-gradient terms and apply the group-theoretical classification to higher-rank susceptibility tensors. This approach then allows them to study the thickness and the character (Bloch vs. N\'eel type) of domain walls. An attempt by Muthukumar {\em et. al.} \cite {ref27} to calculate the $\chi^{(2\omega)}_{el}$ tensor elements for the antiferromagnetic Cr$_2$O$_3$ both from group theory as well as {\em from the microscopic point of view} is rather unique. They implemented a (CrO$_6$)$_2$ cluster, thus taking into account only half of the spins present in the elementary magnetic cell. In this approximation they explained the SHG from Cr$_2$O$_3$ as observed by Fiebig {\em et al.} \cite{ref18} and they were able to give a quantitative estimate for that. Tanabe {\em et al.} \cite {ref56}, however, pointed out that the occurrence of purely real or imaginary values of the tensor elements plays a decisive role for the existence of SHG from this substance. They found that for a (CrO$_6$)$_2$ cluster SHG can take place only in the case where the tensor elements are imaginary, and thus should vanish in Muthukumar's approximation. They proposed to take into account the full unit cell with four inequivalent Cr ions including their ``twisting'' interaction with the environment. However Tanabe {\em et al.} neglected the dissipation in the process of SHG \cite {RemTanabe}, which is a rather crude approximation. In general, taking into account the dissipation makes the $\chi^{(2\omega)}_{el}$ tensor elements complex and invalidates their separation in purely real and imaginary ones \cite{unpublished}. Lifting the inversion symmetry of a crystal is the source for SHG. Lyubchanskii {\em et al.} \cite {ref19,ref21} suggested crystal lattice deformations and displacements as possible reasons for SHG from YIG films. In the case of Cr$_2$O$_3$ and YBa$_2$Cu$_3$O$_{6+\delta}$, described by Lyubchanskii {\em et al.} \cite {ref20,ref21}, AF ordering lowers the symmetry of an otherwise centrosymmetric crystal. In this paper, however, we rely on the idea that, rather than lowering the crystal symmetry in the bulk, SHG may also result from the breaking of inversion symmetry at the surface of a bulk inversion-symmetric system. Magnetically active oxide layers are of importance for the construction of TMR (tunneling magnetoresistance) devices, where a trilayer structure is commonly used. The central layer of TMR devices consists of an oxide sandwiched between a soft and a hard magnetic layer (these two layers are often composed from the same material but of different thicknesses). For these technological applications it is necessary to develop a technique to study buried oxide interfaces. Such a technique can be SHG. One of the most promising materials for the mentioned devices is NiO. However, to the best of our knowledge, the understanding of its detailed spin structure is scarce - even the spin orientation on the ferromagnetically ordered (111) surfaces is not known. The technique presented here can shed some light on that issue. Our paper is organized as follows: in Sec. II we present our methods for obtaining sets of nonvanishing $\chi^{(2\omega)}_{el}$ tensor elements. In Sec. III we present the results of our analysis, first for the nondistorted surface of a simple fcc structure (subsection III.A), then for the the distorted one (III.B). Subsequently, we discuss the influence of a second kind of magnetic atoms (III.C) and of oxygen sublattice distortion (III.D). The issue of domain imaging is addressed in subsection III.E. Possible experimental geometries allowing for the detection of the mentioned structures and effects are discussed in Sec. IV. The conclusions are presented in Sec. V. \section{Theory} \noindent Based on group theory, D\"ahn {\em et al.}~\cite{dahn} proposed a new nonlinear magneto-optic Kerr effect (NOLIMOKE) at the surface of cubic antiferromagnets. They also gave an example of an antiferromagnetic structure (NiO) and an optical configuration, where this new effect could be observed. Here, we perform a complete group-theory based analysis of collinear AF fcc low-index crystal surfaces. Surfaces of other crystal structures are as well described by our theory provided they are similar to fcc crystal surfaces, i.e. squares or hexagons. The results can be used to detect the magnetic order of a specific surface under investigation and allow for the determination of the surface spin configuration in some important cases. However, in order to calculate the SHG yield quantitatively, it is necessary to go beyond the present study and use electronic calculations of the nonlinear susceptibility. Group theory can give a unified picture of different experimental observations and predict new effects \cite {symmetry}, while the microscopic origins of the observed phenomena may remain unclear. In order to be clear with respect to the essential notion of time reversal we would like to emphasize the point of view taken in this paper in the beginning. Here, we do not divide $\chi^{(2\omega)}_{el}$ into even and odd parts in the magnetic order parameter. Instead, the behavior of $\chi^{(2\omega)}_{el}$ with respect to the magnetic order parameter (which for ferromagnetic materials corresponds to the dependence of $\chi^{(2\omega)}_{el}$ on magnetization) is fully taken into account by the considerations of the magnetic point group. At no stage of our consideration we invoke the notion of time reversal, consequently we do not apply the characterization of the susceptibility $\chi^{(2\omega)}$ as c-tensor (changing its sign in the time-reversal operation) or i-tensor (invariant under the time-reversal operation) \cite{unpublished}. Before we start our group theoretical classification of the nonlinear optical susceptibilities of AF surfaces we would like to emphasize the following four important points:\\ (i) We are not interested in effects resulting from the {\em optical path difference} from adjacent crystal planes which are ferromagnetically ordered but only antiferromagnetically coupled to each other. We do not consider this as an intrinsic AF effect. \\ (ii) Cubic crystals that we are interested in reveal a center of inversion in the para-, ferro-, and all antiferromagnetic phases. Thus, within the electric dipole approximation, the SHG signal from the bulk vanishes.\\ (iii) While in principle linear optical methods can be sensitive to the presence of a spin structure, in practice they are not useful because, within the group theoretical approach, they cannot distinguish the AF phase from either paramagnetic or ferromagnetic, nor can they distinguish different AF configurations from each other. They have to resort to methods like lineshape analysis, where no strong statements characteristic for symmetry analysis can be made. \\ (iv) Although the tensor elements for all the magnetic point groups are known and tabulated in the literature (e.g. \cite{birss}), the connection between the different spin configurations described by us and the mentioned symmetry groups has not been made, except for some easy cases \cite{dahn}. Thus, for SHG from antiferromagnetic surfaces there has been up to now no connection between the group theoretical classification and the real situations found in experiments. The following part of the text should explain the fundamentals of applying NOLIMOKE observations to investigate antiferromagnetism of surfaces. Now we turn to SHG, the source of which is the nonlinear electrical polarization $P^{(2\omega)}_{el}$ given by: \begin{equation} P^{(2\omega)}_{el}=\epsilon_0 \chi^{(2\omega)}_{el}:E^{(\omega)}E^{(\omega)} . \end{equation} Here, $E^{(\omega)}$ is the electric field of the incident light, while $\chi^{(2\omega)}_{el}$ denotes the nonlinear susceptibility within the electric dipole approximation, and $\epsilon_0$ is the vacuum permittivity. The intensity of the outgoing SHG light is \cite{Sipe}: \begin {multline} \label{eqGlown} I^{(2\omega)}\sim (I_0)^2 \Bigl [ F(\Theta ,\Phi , 2\omega) \times \\ \times \begin{pmatrix} \chi_{xxx}&\chi_{xyy}&\chi_{xzz}&\chi_{xyz}&\chi_{xzx}&\chi_{xxy}\\ \chi_{yxx}&\chi_{yyy}&\chi_{yzz}&\chi_{yyz}&\chi_{yzx}&\chi_{yxy}\\ \chi_{zxx}&\chi_{zyy}&\chi_{zzz}&\chi_{zyz}&\chi_{zzx}&\chi_{zxy} \end{pmatrix} \times \\ \times f(\vartheta ,\varphi ,\omega) \Bigr ]^2 \end{multline} where $I_0$ is the intensity of the incident light, $F()$ ($f()$) describe Fresnel and geometrical factors for the incident (reflected) light, $\vartheta$ and $\Theta$ angles of incidence and reflection, respectively ($\vartheta$=$\Theta$), and $\Phi$ ($\varphi$) is output (input) polarization angle. According to Neumann's principle, ``any type of symmetry which is exhibited by the crystal is possessed by every physical property of the crystal'' \cite{birss}. To examine these physical properties, we determine the magnetic point group of the crystal lattice, thus determine its symmetries. The same symmetries must leave the investigated property tensor (in our case the nonlinear electric susceptibility~$\chi^{(2\omega)}_{el}$) invariant. This fact is mathematically expressed by the following condition: \begin{equation} \label {eq2} \chi^{(2\omega)}_{el,i'j'k'}=l_{i'i}l_{j'j}l_{k'k} \chi^{(2\omega)}_{el,ijk},\,\, \,\, i,j,k,i',j',k'=x,y,z . \end{equation} Here, $\mathit{l_{n,n'}} (n = i,j,k, n'= i',j',k',)$ is a representation of an element of the magnetic point group describing the crystal. For symmetry operations including the time reversal there should be an additional ``$\pm$'' sign in Eq.(\ref {eq2}), but we do not use it here since we exclude the time reversal from our consideration. In particular, from Eq.(\ref {eq2}) it follows immediately that polar tensors of odd rank (such as~$\chi^{(2\omega)}_{el}$ ) vanish in inversion symmetric structures. This explains why SHG is possible only at surfaces and interfaces, where this symmetry is broken. For a given spin configuration we apply Eq. (\ref {eq2}) for every symmetry operation exhibited by the system. Thus, each of these symmetries gives rise to a set of 27 equations with 27 unknown elements of the tensor $\chi^{(2\omega)}_{el}$. This set can be reduced to 18 equations, since \begin {equation} \label {eq3} \chi^{(2\omega)}_{el,ijk}=\chi^{(2\omega)}_{el,ikj}, \end{equation} which expresses the equivalence of the incident photons of frequency $\omega$, see also the reduced notation in Eq. (\ref{eqGlown}). The analytic solution of even this reduced set of equations seems cumbersome, but the set can be split into several decoupled subsets. For example, an obvious subset in every case is the equation $\chi_{zzz} = \chi_{zzz}$, this tensor element occurs nowhere else. The rank of other subsets is, for our cases, never higher than six. In this manner, one may obtain a set of forbidden elements of the susceptibility tensor as well as relations between existing ones. \section{Results} First, we will define the notions of ``phase'', ``case'', and ``configuration'', used henceforth to classify our results. ``Phase'' describes the magnetic phase of the material, i.e. paramagnetic, ferromagnetic, or AF. Secondly, the word ``configuration'' is reserved for the description of the magnetic ordering of the surface. It describes various possibilities of the spin ordering, which are different in the sense of topology. We describe up to 18 AF configurations, denoted by little letters a) to r), as well as several ferromagnetic configurations, denoted as ``ferro1'', ``ferro2'', etc. The number of possible configurations varies depending on surface orientation. Thirdly, we describe different ``cases'', i.e. additional structural features superimposed on the symmetry analysis. ``Case A'' does not have such additional features. In ``case B'' we address distortions of the lattice. ``Case C'' deals with two kinds of magnetic atoms in an undistorted lattice. In ``case D'' we take into account a distorted sublattice of nonmagnetic atoms, keeping the magnetic sublattice undistorted. All the analysis concerns collinear antiferromagnets, with one easy axis. The tables show the SHG response types for each configuration. The various response types are encoded by a ``key'', which is then decoded in Tab. \ref{keytab}. This table presents the symmetries, domain operations, and nonvanishing tensor elements for each response type. This is done in order to shorten the overall length of tables, because a given response type can appear in several different cases. Several spin structures depicted in Fig. \ref{fig1} and Fig. \ref{fig3} are distinct configurations only in case B, and they are addressed in the tables that concern only this case. For the rest of the cases they are domains of other, fully described configurations, thus they are left out in these cases. The philosophy of the paper is that, to save some space, we show the spin structure in one figure for each surface (Fig. \ref{fig1}, \ref{fig2}, and \ref{fig3}) for all the four cases (A-D), and depict the effects taken into account in the cases B-D only for the paramagnetic phase (Fig \ref{fdis111}, \ref{fnierow}, and \ref{ftlen}). Table \ref{keytab} also contains the information on the parity of the nonvanishing tensor elements: the odd ones are printed in boldface. In some situations an even tensor element (shown in lightface) is equal to an odd element (shown in boldface), this means that this pair of tensor elements is equal in the domain which is depicted on the corresponding figure, but they are of opposite sign in the other domain. This happens in the structures where two pairs of domains are possible (two distinct entries in Table \ref{keytab}). The tensor elements that change their parity in the domain operation which is the inverse of the displayed one are shown in italic font. For example, the entry j) of Table \ref{keytab} shows a tensor element {\it xxx}, which is even under the operation $4_z$, this means that this tensor element is odd under $-4_z$. This strange at the first sight behavior of tensor elements is caused by the fact that under these operations, tensor elements are not mapped on themselves. In our example, after applying $4_z$ the tensor element $xxx$ becomes $yyy$, without changing its sign. If we now apply $-4_z$, $yyy$ (which is now even) becomes $xxx$, again without changing the sign. The parity of the elements has been checked in the operations $2_z$, $4_z$, and in the operation connecting mirror-domains to each other (for the definition of the mirror-domain structure see subsection E). The domain operation(s) on which the parity depends is (are), if applicable, also displayed in this Table. If two or more domain operations have the same effect, we display all of them together. To make the Table \ref{keytab} shorter and more easily readable some domain operations (and the corresponding parity information for the tensor elements) are not displayed, namely those that can be created by a superposition of the displayed domain operations. We also do not address the parity of tensor elements in the $6_z$ nor $3_z$ operations for (111) surfaces nor any other operation that ``splits'' tensor elements, although these operations also lead to a domain structure \cite{par111}. As will be discussed later (subsection E) it is possible to define a parity of the tensor elements for the $3_z$ and $6_z$ operations, however the tensor elements then undergo more complicated changes. The situations where the parity of the tensor elements is too complicated to be displayed in the Table are indicated by a hyphen in the column ``domain operation''. For some configurations, none of the operations leads to a domain structure - in those configurations we display the information ``one domain''. The reader is referred to the Appendix for the particularities of the parity check. As far as the first layer is concerned, we address all the spin configurations of the low index surfaces of fcc antiferromagnets, with magnetic order vector lying in plane or perpendicular to it and antiferromagnetic coupling between nearest neighbors. For the (001) surfaces we also discuss the configurations, where the antiferromagnetic coupling exists between the second-nearest neighbors (configurations a), b), c), f), and o), along with d), g), and h) for case B.). We do not consider the coupling to the third and further neighbors. This would not give rise to configurations of different symmetries in two dimensions. It may at most replace spins by grains (blocks) of spins in the configurations described by us. Throughout this paper we take into account the spin structure only of the first (uppermost) atomic layer. This is sufficient to study all the symmetries of (001) and (110) surfaces both in the paramagnetic and ferromagnetic phases. For the (111) surface it is necessary to recognize the atomic positions (but not the spins) in the second layer for the same purpose. For the sake of completeness we also present a study of (111) surfaces without this extension. However, in the antiferromagnetic phase, the spin structure of the second and deeper layers plays a role in determining the symmetry of the surface. This is presented in this paper using the (001) surface as an example. For the (110) and (111) surfaces it will be published elsewhere \cite{Noptipap}. These structures can serve as simple models for deriving predictions for more complicated cases, while the full consideration of the second layer would not bring any new interesting results. Taking into account the spin structure of the second layer (deeper layers do not bring up anything new to the analysis) results in creating several (up to two for the (001) surface and three for the (111) surface) configurations out of each one addressed here by us. The symmetry of these configurations may remain the same or be lowered (sometimes even below the symmetry of the ferromagnetic phase) with respect to the ``two-dimensional'' configurations they are generated from. Consequently the distinction of the configurations from each other may be limited, but the possibility to detect the magnetic phase is not severely affected. Also our remarks on domain imaging remain valid. However the number of domains is increased, thus the possibility to identify each of them might be hampered. Consequently, one can state that the symmetry of an AF surface depends on two atomic layers. They are also necessary (and sufficient) to define AF bulk domains. As will be presented in our results, SHG can probe both these layers on AF surfaces. \subsection{Equivalent atoms} \noindent The predicted new nonlinear magneto-optical effects result from the fact that the magnetic point groups of antiferromagnetic configurations are different from those describing paramagnetic or ferromagnetic phases of the same surface. Since, depending on the magnetic phase, different tensor elements vanish, it is possible to detect antiferromagnetism optically by varying the polarization of the incoming light. The current subsection discusses nonvanishing elements of the nonlinear susceptibility tensor for an fcc crystal consisting of only one kind of magnetic atoms. The influence of nonmagnetic atoms in the material will be discussed later. The configurations considered here are ``ferro1", ``ferro2", ``ferro4", a), b), c), e), f), i), k), m), o), p), and r) for the (001) surface (see Fig. \ref{fig1}), ``ferro1", ferro3", ``ferro5", a), c), f), i), and k) for the (111) surface (see Fig. \ref{fig3}), and all configurations depicted in Fig. \ref{fig2} for the (110) surface. Other depicted spin structures form domains of these configurations and are not referred to in this subsection nor in the tables concerning the current subsection \cite{leftout}. All possible configurations (confs.) of a fcc (001) surface are shown in Fig.~\ref{fig1}, which displays the conventional rather than magnetic unit cells. However, these are sufficient to fix the spin configuration of the whole surface imposing of the following ``convention'': the fcc surface is constructed from the depicted plaquette in the way that neighboring spins along the $x$ and $y$ directions point the same way (alternate) if they are parallel (antiparallel) on the plaquette in these two directions. The spins in rows and columns where only one spin is presented are continued in the same way as the corner spins. For instance in the configuration a) of the (001) surface, both the right-hand side and left-hand side neighbors of the ``central'' spin will point upwards, while the spin direction will be alternated along the $x$ axis. This convention will be maintained henceforth (for a (111) surface one has to alter or keep the spins along three axes, instead of two). The smallest set that gives a complete idea about the spin structure is presented in Fig.~\ref{fprim} \cite{4atoms}; this ``magnetic primitive cell'' does not give a clear picture of the crystal symmetries, however. The whole crystal lattice can be reproduced by translations of this cell, without performing other operations such as reflections or rotations. The SHG response types for the (001) monolayer are given in Table~\ref{tab1}, for the paramagnetic, ferromagnetic, and all AF phases. We can observe several sets of allowed tensor elements. The Conf. r) will produce the same signal as the paramagnetic phase. The Conf. ``ferro1'' reveals a completely different, distinguishable set of tensor elements. In addition, the conf. ``ferro2'' produces another set of tensor elements, different from any other configuration. It is equivalent to the conf. ``ferro1'' rotated by 45$^\circ$. In the confs. a), b), e), and o) we find the same tensor elements as for the paramagnetic phase. However, due to the lower symmetry, their values are no longer related to each other. Confs. c) and f) bring new tensor elements, thus allowing for the distinction of these confs. from the previous ones. Confs. i), k), m), p) reveal the same tensor elements as c) and f) but some of these elements are related. Thus one may possibly distinguish these two sets of configurations. Conf. ``ferro4'' presents a completely different, distinguishable set of the nonvanishing tensor elements. Consequently, in six configurations (i.e. c), f), i), k), m), and p)) some susceptibility tensor elements appear only in the AF phase, allowing for the detection of this phase by varying the incident light polarization, as will be outlined in Sec. IV. In addition, all other antiferromagnetic configurations but r) reveal the breakdown of some of the relations between the different tensor elements, compared to the paramagnetic phase, and thus can be detected as well. Generally, all the phases can be distinguished from each other. There exists as well a possibility to distinguish different AF configurations provided the corresponding tensor elements can be singled out by the proper choice of the experimental geometry. For the sake of completeness, we now present a short study of the (001) surface where the spin structure of the two topmost atomic layers is taken into account. The paramagnetic phase and all the ferromagnetic configurations remain unchanged with respect to the results of the previous paragraph (for the (001) monolayer). However, most of the AF configurations previously addressed break up into two different configurations (sometimes even with a different symmetry). These configurations are constructed from the ones of the previous paragraph by assuming that the structure of the second atomic layer is identical with that of the topmost one but shifted along the positive $x$ axis (indicated by x after the name of the original configuration) or positive $y$ axis (indicated by y after the name of the ``parent'' configuration) in a proper way to form a fcc structure; if only one configuration can be produced in this way we use the name of the original one. This construction is depicted in Fig. \ref{seclay}, along with the corresponding conventional unit cells for the two topmost layers of the AF fcc (001) surface. The resulting SHG response types are presented in Table \ref{t002}. In general, seven types of response are possible. Firstly, the paramagnetic phase reveals a characteristic set of tensor elements. Thus it can be unambiguously distinguished from any other magnetic phase. Secondly, confs. ``ferro1'', ax), ox), bx), by), ex), and ey) bring some additional tensor elements into play. The symmetry of confs. ax) and ox) is slightly different from the one of the rest of this group, since the mirror plane is rotated by $90^\circ$ around the $z$ axis. A different set of tensor elements is brought up by confs. ``ferro2", i), m), and p). The difference between the response yielded by conf. i) and the other confs. in this group, due to a slightly different symmetry, can be compensated by rotating the sample by $90^\circ$ around the $z$ axis. Another, characteristic set of tensor elements is presented by conf. ``ferro4" alone. The fifth type of SHG response is given by confs. ay), oy), and r). Tensor elements, that do not vanish in these configurations, are the same as for the paramagnetic phase but some relations between them are broken due to a lower symmetry in the AF phase. Confs. cx), fx), and fy) yield all tensor elements in an unrelated way. The last, characteristic type of response is presented by conf. k) alone. Consequently, the detection possibilities of an antiferromagnetic bilayer are slightly worse than those for a monolayer. Especially, a difficulty in distinguishing the ferromagnetic phase from the antiferromagnetic one may arise for some configurations where then the combination of SHG with linear magneto-optics is definitly required. There exists a possibility to distinguish AF configurations from each other, similarly to the previous situation. In most configurations, the difference (in terms of the SHG response) between the bilayer structure described here and the previously addressed (001) monolayer can be detected. We now turn to the (110) surface (Fig.~\ref{fig2}), which, in the paramagnetic phase, reveals a lower symmetry than the (001) surface. On the other hand, the number of symmetry operations in the AF configurations is comparable to the (001) surface. In addition, as shown in Table~\ref{tab2}, the resultin SHG response types are not very characteristic, so the detection possibilities for this surface are very limited. In particular, confs. a), b), c), g), h), i), j), k), and l) give the same tensor elements as the paramagnetic phase. Confs. d), e), f), and ``ferro3'' bring new tensor elements. Other ferromagnetic configurations (``ferro1'' and ``ferro2'') present different sets of new tensor elements, making these configurations distinguishable from the others as well as from each other. Conf. ``ferro4'' yields a completely different set of tensor elements, however this set is related to the one of conf. ``ferro1'' by $90^\circ$ rotation. The study of the (111) surface (see Fig.~\ref{fig3}) has to be separated in two subcases, according to whether we take into account only one atomic monolayer or more. In both subcases, we consider the same configurations. The SHG response types for the first subcase are listed in Table~\ref{tab3}, and for the second subcase in Table ~\ref{tab4}. For the {\em first} subcase, confs. a), i), and k) reveal the same tensor elements as the paramagnetic phase, however due to the lower symmetry their values are not related to each other. Configurations c) and f) present new tensor elements. As for the previous surfaces, the ferromagnetic phase reveals completely different sets of tensor elements, and the three ferromagnetic configurations can be distinguished from each other since they bring different tensor elements into play. Unlike for the (110) surface, the axes $x$ and $y$ are not topologically equivalent, and thus the fact that tensor elements of ``ferro1'' are related to those of ``ferro3'' by $90^\circ$ rotation does not affect the possibility to distinguish these two configurations. The ferromagnetic conf. ``ferro5'' brings up the same tensor elements as AF confs. c) and f), but the relations between the elements are different. The {\em second} subcase (more layers taken into account) gives different sets of allowed tensor elements (compared to the first subcase) for each but the ``ferro3'' configuration. Confs. a), i), k), and ``ferro3'' share the same set of allowed tensor elements and can be easily distinguished from the paramagnetic phase. Confs. c), f), and ``ferro1'' reveal all tensor elements, with their values unrelated. Similarly, conf. ``ferro5'' presents another, distinguishable set of tensor elements. The possibility to distinguish the magnetic phases is rather limited. The symmetry analysis of nonvanishing tensor elements for ferromagnetic surfaces in the case A have been performed by Pan {\em et. al.} \cite {ref53}. Our analysis yields the same results, taking into account the corrections made by H\"ubner and Bennemann \cite {ref80}. \subsection{Distortions of monoatomic lattice} \noindent The rhombohedral distortion of the atomic lattice, described here and shown in Fig. \ref{fdis111}, makes the $x$ and $y$ axes of the (001) surface inequivalent, even in the paramagnetic phase. On the (111) surface, the $y$ axis is not equivalent any longer to other axes connecting the nearest neighbors, namely $S_{(xy)}$ and $S_{(-xy)}$ (for the definition of the ``S'' and ``H'' axes see Fig. \ref{fig3}, the paramagnetic conf.). These inequivalences of axes are the reasons for the reduction of the number of symmetry operations in the paramagnetic phase. Because of this reduction some spin structures that previously formed different domains of a single configuration now cannot be transformed into each other and become ``independent" configurations. This happens for almost every of the previously addressed configurations of the (001) and (111) surfaces. Consequently, all the depicted spin structures are in fact configurations, and are addressed in this subsection. The resulting SHG response types for the (001) surface are listed in Table \ref {tdis100}. For this surface, only two of the ferromagnetic configurations, namely ``ferro1'' and ``ferro2'' can be easily distinguished from both the paramagnetic as well as the antiferromagnetic phases. These ferromagnetic configurations can be also distinguished from each other. On the contrary, all the AF configurations yield only two types of response, and in addition one of them is equivalent to the response of the paramagnetic phase. Consequently, it will not be possible to determine the surface spin structure, and the distinction of the AF phase from the paramagnetic one can be successfully performed only in confs. a)-h) and o). Compared to the case A, there is an important symmetry breaking for most configurations. Thus, the distinction between the two cases (A and B) is possible (compare Tabs. \ref{tab1} and \ref{tdis100}). All the (110) surfaces of an fcc crystal with a rhombohedral distortion are topographically equivalent to the (110) surface of the case A. The distortion only stretches the $x$ or $y$ axis, so the structure remains rectangular. The analysis of the (111) surface (depicted in Fig. \ref{fdis111}) in the subcase of only one monolayer reveals sets of symmetries very similar to the (110) surface, as it follows from the Table \ref{tdis1111}. In fact, the (111) surface of a fcc crystal with a rhombohedral distortion can be treated as two rectangular lattices superimposed on each other. In turn, due to the distortion, it is not convenient any longer to describe the spin structures using ``S" and ``H" axes. The possibility to distinguish AF configurations is very poor, and two of the AF configurations (a) and k)) yield the same signal as the paramagnetic surface. In confs. b) - j), l), and m) the AF phase can be distinguished from the paramagnetic one, but they give the same signal as conf ``ferro5". Conf. ``ferro2" can be easily distinguished since it reveals a characteristic set of (all) tensor elements. Confs. ``ferro1" and ``ferro3" yield different sets of tensor elements, but they are related to each other by $90^\circ$ rotation. Most of the configurations allow for the distinction of the cases A and B (compare Tabs. \ref{tab3} and \ref{tdis1111}). In the subcase of two monolayers of the (111) surface, the symmetry is dramatically reduced (see Tab. \ref{tdis1112}). Even in the paramagnetic phase the group of symmetries consists of only one nontrivial operation, and this appears to occur also in the AF configurations a), i), k), and ``ferro3". In all the other configurations all tensor elements are allowed due to the lack of any symmetry. Only confs. paramagnetic and ``ferro5" allow for the unambiguous distinction of the cases A and B (compare Tabs. \ref{tab4} and \ref{tdis1112}). Consequently, this surface is not very useful to an analysis of the magnetic structure, with the exception of stating the distortion itself. As the conclusion of the case of the distorted sublattice of magnetic atoms, the surfaces give extremely limited possibilities to investigate the magnetic properties. In our further study, we will limit ourselves to lattices of undistorted magnetic atoms. \subsection{Structure with nonequivalent magnetic atoms} \noindent We assume now that not all the magnetic atoms in the cell are equivalent. An example of such a structure is a material composed of two magnetic elements, but also a situation when the magnetic lattice sites are inequivalent due to different bonds to a nonmagnetic sublattice; distortions of the sublattice of nonmagnetic atoms that preserve the center of twodimensional inversion produce the same effect. Other distortions of the sublattice of nonmagnetic atoms will be discussed in subsection D. The magnetic moment at the distinguished positions can be changed or not - this does not affect the results obtained by symmetry analysis. The configurations considered here are ``ferro1", ``ferro2", ``ferro4", a), b), c), e), f), i), k), m), o), p), and r) for the (001) surface (see Fig. \ref{fig1}), ``ferro1", ferro3", ``ferro5", a), c), f), i), and k) for the (111) surface (see Fig. \ref{fig3}), and all configurations depicted in Fig. \ref{fig2} for the (110) surface. Other depicted spin structures form domains of these configurations and are not referred to in this subsection nor in the tables concerning the current subsection. The structure is depicted in Fig.~\ref{fnierow}. For the sake of brevity, we show the structure of the distinguished atoms only for the paramagnetic phase. All the configurations are the same as in case A, for all surface orientations. The already mentioned ``convention'' of alternating (or not) spin directions along certain axes is applied regardless of the atom type. This allows us to obtain the whole crystal surface from the small displayed fragment. Our analysis starts with the (001) surface of an fcc crystal. The SHG response types for each configuration are listed in Table~\ref{tnierow100}. In general, we can observe seven types of response. The first of them is represented by the paramagnetic phase alone. The second type of response, exhibited by the ferromagnetic ``ferro1'' and the AF a), b), e), o) confs., differs from any other type by some tensor elements. The confs. a) and o) reveal different tensor elements than the other configurations from the mentioned group. However, the signal from confs. a) and o) is the same as for the confs. b), e), and ``ferro1'' if one exchanges the axes $x$ and $y$. Thus, if the directions of the spins cannot be determined by another method, confs. a) and o) cannot be distinguished from b), e), and ``ferro1''. The next type consists of conf. f) and reveals all tensor elements, while no relations between them are enforced by the symmetry analysis. A completely different type of response is presented by conf. c) alone. Another type, where confs. i), m) and p) belong to brings the same tensor elements as conf. c), but there exist more relations between the elements due to a higher symmetry in these configurations. The next type is given by confs. ``ferro2'' and k). As in conf. f) all the tensor elements are present but this time there are some relations between them. In addition, confs. r) and ``ferro4'' yield a completely new set of tensor elements due to the preserved fourfold rotational symmetry. Thus, assuming one atom as distinguished may reduce the symmetry. New tensor elements appear in confs. a), b), e), f), k), o), and r) compared to case A (compare Tabs. \ref{tab1} and \ref{tnierow100}). In these configurations it is therefore possible to distinguish the cases of equivalent and nonequivalent magnetic atoms, provided the tensor elements that make the cases different can by singled out by the experimental geometry. There exists also a possibility to distinguish different AF configurations in case C. The antiferromagnetic {\em phase} can be undoubtelly detected in the surface configurations c), f), i), m), and p). For the (110) surface, there are more possibilities to distinguish the configurations with nonequivalent magnetic atoms than in the case A. However, the configurations still produce ambiguous signals (see Tab. \ref{tnierow110}). Confs. b), c), h), i), k), and l) are equivalent to the paramagnetic phase. Conf. a) is equivalent to the ferromagnetic ``ferro1'' configuration, and conf. d) to ``ferro2''. In addition, the confs. e), f), and g) are equivalent to the conf. ``ferro3'' and conf. j) gives the same signal as conf. ``ferro4''. Even the presence of nonequivalent atomic sites in the lattice cannot be detected by SHG on this surface, since the symmetry of the (110) surface is usually not lowered further by the existence of equivalent magnetic sites (compare Tables \ref{tab2} and \ref{tnierow110}. The only exception are the confs. a), d), g), and j) which give different tensor elements in the two cases. As in the case of equivalent atoms, the (110) surface is not very useful for the analysis. The study of the (111) surface must again be divided in the two subcases of one or more monolayers, respectively. Fig.~\ref{fnierow} depicts the situation in the paramagnetic phase. The SHG response types are listed in Tables ~\ref{tnierow1111} and~\ref{tnierow1112} for the first and the second subcase respectively. In the first subcase (one monolayer) the symmetry establishes six different types of nonlinear response. The ``paramagnetic'' type (for the paramagnetic configuration only) is characteristic - all the other configurations have additional tensor elements. The next type of response (the ferromagnetic conf. ``ferro1'' and the antiferromagnetic conf. a)) brings some new tensor elements. Other tensor elements appear in the conf. k). Configurations ``ferro3'' and i) show another set of nonvanishing tensor elements. The confs. c) and f) reveal all tensor elements in an unrelated way. In addition, conf. ``ferro5'' presents a characteristic set of tensor elements. In the second subcase, only four different SHG responses are possible. Firstly, the paramagnetic phase is characteristic - all the other configurations bring additional tensor elements into play. The next type of response is presented by confs. ``ferro3'' and i) - they yield some additional tensor elements. Confs. ``ferro1'', a), c), f), and k) reveal all tensor elements and no relations between them appear from our symmetry analysis. Again, the conf. ``ferro5'' presents a unique set of nonvanishing tensor elements. Consequently, for the (111) surface, the symmetry breaking due to the presence of a second kind of magnetic atoms has even more important consequences than for the (001) surface. In the situation of only one monolayer, the distinction between the cases may be possible for all the AF configurations (compare Tables \ref{tab3} and \ref{tnierow1111}). Considering additional layers leads to further symmetry breaking and renders the distinction between the configurations impossible. The distinction between the cases A and C is possible in confs. a) and k) (compare Tables \ref{tab4} and \ref{tnierow1112}). Besides, in most configurations it is possible to decide if these additional layers play any role (compare Tables \ref{tnierow1111} and \ref{tnierow1112}). \subsection{Distorted oxygen sublattice} \noindent Due to the strong charge-transfer between nickel and oxygen in NiO the sublattices may be distorted. This effect can lower the symmetry of the surface. A point-charge model calculation by Iguchi and Nakatsugawa \cite{ref42} presented a shift of the oxygen sublattice (``rumpling'') in the direction perpendicular to the surface. Their method did not show any in-plane displacement and thus no change of the surface symmetry. However, if the ``rumpling'' also has an in-plane component, i.e. if the oxygen atoms are displaced also in the $x$ and $y$ directions, it will also have a considerable effect on the symmetry of the crystal surface. For this paper, we have chosen a distortion that can lower the symmetry of the surface and besides can be represented within one conventional unit cell. The configurations considered here are ``ferro1", ``ferro2", ``ferro4", a), b), c), e), f), i), k), m), o), p), and r) for the (001) surface (see Fig. \ref{fig1}), ``ferro1", ferro3", ``ferro5", a), c), f), i), and k) for the (111) surface (see Fig. \ref{fig3}), and all configurations depicted in Fig. \ref{fig2} for the (110) surface. Other depicted spin structures form domains of these configurations and are not referred to in this subsection nor in the tables concerning the current subsection. As will be shown later, the best conditions for the detection of this kind of distortion are presented by the (110) surface. The (111) surface could show equally good possibilities if only a monolayer of magnetic atoms is present. In the presence of an oxygen sublattice distortion, the chemical unit cell is also doubled. This effectively means that magnetic unit-cell-doubling (describing the fact that the magnetic unit cell is twice as big as the chemical one) is lifted. In general, taking into account distorted oxygen atoms in the paramagnetic phase does not lower the symmetry of the problem. The exception is the (111) surface, where the six-fold axis is replaced by the three-fold one. In the case of the distorted oxygen sublattice, the symmetry group for each configuration is a subgroup of the corresponding ``non-distorted'' configuration, i.e. of the corresponding spin configuration in the case where the oxygen atoms are not considered. As in case C we display only the paramagnetic phase in Fig. \ref {ftlen} to depict the atom positions. All the spin configurations are the same as for the corresponding surfaces in case A, and the spins are assumed to be equivalent. As Table~\ref {ttlen100} shows, six different responses can be expected from the (001) surface. The paramagnetic surface will give a characteristic response. The second group is formed by the confs.: a), b), e), o), and ``ferro1''. Although confs. a) and o) have elements different from the remaining configurations in this group, this fact corresponds simply to rotating the sample by 90$^\circ$ with respect to the $z$ axis. Confs. c) and f) reveal all tensor elements without relations between them. Confs. ``ferro2'', i), k), and m) reveal all tensor elements with some relations. The only difference between conf. m) and others from this group is like for the previous group a 90$^\circ$ rotation with respect to the $z$ axis. Another group consists of conf. p) alone. It reveals the same tensor elements as the paramagnetic phase, but certain relations between tensor elements are broken due to a lower symmetry of the conf. p). The confs. r) and ``ferro3'' form the last group. All the configurations but k) and ``ferro3'' can be distinguished from those of case A (compare Tabs. \ref{tab1} and \ref{ttlen100}). However only confs. c) and g) can be distinguished from case C (compare Tabs. \ref{tnierow100} and \ref{ttlen100}). Thus, only in these configurations it will be possible to detect oxygen sublattice distortions by SHG. The SHG response types for the (110) surface are presented in Table \ref {ttlen110}. One can observe that only configurations c), f) and i) give rise to new (compared to case A, Table \ref{tab2}) tensor elements. Compared to case C (Table \ref{tnierow110}), confs. c), f), and i) bring new tensor elements, and, surprisingly, confs. a) and g) have less tensor elements, due to higher symmetries in the case D. Consequently, the confs. a), c), f), g), and i) allow for an unambiguous determination of the oxygen sublattice distortion from the (110) surface. The possibility to distinguish different configurations is rather limited. Oxygen sublattice distortion similar to the one presented in Fig. \ref{ftlen} for a (111) surface was found by Renaud {\em et al.} \cite{ref68} and calculated by Gillan \cite{refGillan} in M$_2$O$_3$ materials (M = Al, Fe). Since the nonmagnetic sublattice symmetry group has an influence on SHG this distortion can be detected also on surfaces of fcc crystals. In the previous cases A and C we divided the study of (111) surfaces in two subcases, considering either one or more atomic layers. Taking into account a distorted oxygen sublattice leads us immediately to the subcase of ``more atomic layers''. It is caused by the fact that the oxygen and magnetic atoms belong to mutually exclusive planes. The resulting SHG response types are listed in Table \ref {ttlen111}. For the AF and ferromagnetic phases, all tensor elements are allowed for every configuration. Thus SHG cannot detect the magnetic phase of the surface nor distinguish different configurations. Only confs. paramagnetic, ``ferro3'', ``ferro5", and d) allow to decide unambiguously whether the oxygen sublattice is distorted or not (compare Tabs. \ref{tab4}, \ref{tnierow1112}, and \ref{ttlen111}). For both the (001) and (111) surfaces, the symmetry groups of case D appear to be the subgroups of the corresponding configurations of case C. This means that the oxygen sublattice distortion makes some (one half of all) magnetic atoms distinguished as in case C, even though we did not apply this distinction explicitly in case D. On the other hand, the symmetry groups of the case D differ essentially from those of case B. This is caused by the difference in distortions assumed in these cases: the rhombohedral one in case B and rotation-like in case D. \subsection{Domain imaging} \noindent For simplicity, we will consider here only surfaces described hitherto by the case A of our analysis. In this case, for AF surfaces, no $180^\circ$ domains can be expected due to the presence of magnetic unit-cell doubling. The allowed domains can be detected by surface-sensitive SHG under the following two conditions. First, domains can be imaged by our method only if they manifest themselves at the surface, i.e. if the surface spin ordering changes while passing from one domain to another \cite {domeny}. It is necessary to note, however, that the spin orderings for different domains must belong to the same {\em configuration} in the sense of our classification. We do not consider it as a domain structure if one portion of the surface is in one configuration and another portion is in a different configuration. Under such conditions, we can encounter two different types of domains: $90^\circ$ domains (for the (111) surface they are rather $60^\circ$ domains), resulting from the rotations around $z$ axis, and the second type (called by us mirror-domains, characteristic for antiferromagnets), where spins point along the same axis in all domains, but the ordering is still different (they are no $180^\circ$ domains!). The tables contain the complete information about the parity of tensor elements in mirror-domain operations, and also for $90^\circ$ type domains, but not for $60^\circ$ domains. The $90^\circ$ type domains will be addressed later on. In the mirror-domain structure, the magnetic point group describing the configuration must lack an operation that, while belonging to the (nonmagnetic) point group of the system {\em and} leaving the spin axes invariant, only flips some of the spins. Note, the flipped subset of the spins must be antiferromagnetically ordered in itself. Configurations, the symmetry groups of which {\em lack} one of these operations can reveal surface domains, related to each other by this operation. For an illustration we choose the configuration c) of the (001) surface (see Fig. \ref{fig1}). The spins point along the $x$ axis. Thus operations leaving the axis invariant are $\overline 2_x$, $\overline 2_y$ and $2_z$. Of them, $\overline 2_x$ and $\overline 2_y$ are absent in the magnetic point group of the considered configuration (see Tab. \ref{tab1}, conf. c), and Tab. \ref{keytab}). The flipped subset of spins consists of the four outer spins for the $\overline 2_x$ operation, and of the central spin for $\overline 2_y$ (see Fig. \ref{fimage} b) and c), respectively). In fact, there are two domains possible in this configuration: one with the spins kept invariant under translations by the vector ($-{a\over 2},{a\over 2},0$) (this domain is shown) and the other with the spins kept invariant under translations by the vector (${a\over 2},{a\over 2},0$). Here, $a$ denotes the lattice constant. These domains are depicted in Fig. \ref{fimage}. The second condition for domain imaging is an interference. It can be created internally by different elements of the tensor $\chi^{(2\omega)}$ or by external reference \cite {external1,external2}. The interfering elements should be of a similar magnitude for the largest possible image contrast. Group theory, however cannot account for the amplitudes. With external as well as internal reference, a tensor element that changes its sign under the reversal of the antiferromagnetic order parameter {\bf L} is necessary. Actually, every {\bf L} dependence of $\chi^{(2\omega)}$ can be represented by splitting the tensor elements into odd and even ones in {\bf L}; even if a tensor element is not purely odd or even we can always decompose it according to: \begin {equation} \label{element} \chi^{(2\omega)}_{ijk}=\chi^{(2\omega ),odd}_{ijk}+\chi^{(2\omega),even}_{ijk} \end{equation} i.e. a tensor element consists of parts which are odd and even in {\bf L}, respectively. In a system with many terms of that kind the possibility of detecting domains may be limited, since they can influence the signal with opposite sign, thus diminishing the interference. In highly symmetric structures, such as an fcc crystal, the situation is more comfortable: every tensor element is either odd or even in {\bf L} (see Appendix). By the appropriate set of experiments an element can be singled out and give a clear image of AF domains. As an example we consider tensor elements that are present in all the phases, e.g. $\chi^{(2\omega)}_{zzz}$: they are even in the magnetic order parameters {\bf L} and {\bf M}, for the AF and ferromagnetic phases, respectively. The tensor element $\chi^{(2\omega)}_{zxy}$, present for example in the previously discussed conf c) of the (001) surface (see Fig. \ref{fimage}), is odd, since it changes its sign under the operation $\overline{2}_x$ transforming one domain into another. For other configurations other tensor elements and operations can be found. In the discussed configuration both these elements are present, we have intensity contributions proportional to $(\chi^{(2\omega)}_{zzz})^2$, $(\chi^{(2\omega)}_{zxy})^2$ and $\chi^{(2\omega)}_{zzz}\cdot \chi^{(2\omega)}_{zxy}$, due to the square in Eq. (\ref{eqGlown}). As a result, one obtains an interference: \begin{equation} I_p\sim ... +(\chi^{(2\omega)}_{zzz})^2+(\chi^{(2\omega)}_{zxy})^2\pm 2\chi^{(2\omega)}_{zzz}\cdot \chi^{(2\omega)}_{zxy}+... \end{equation} where ``+'' stands for one domain, ``-'' for a different one. Now, we turn to the $90^\circ$ domain structure. Again, we take the conf. c) of the (001) surface as an example. The operation connecting the domains is $4_z$. Under this operation, the tensor element $\chi^{(2\omega)}_{zxy}$ changes its sign, thus again we have an interference which renders the domain imaging possible. This tensor element is even in the domain operation $\overline 2_{xy}$ (which is equivalent to the superposition of $\overline 2_x$ and $4_z$), which means that domains related to each other by this operation cannot be imaged using this particular tensor element. Similarly, if a tensor element is odd in one domain operation and even in another, it must be odd in their superposition. Concerning the $60^\circ$ domains for (111) surfaces, the parity of the tensor elements must be treated more carefully, as indicated already in \cite{par111}. We can still define three ``twofold'' operations, and each of them has its own set of odd and even tensor elements. The sets corresponding to different of those operations are not mutually exclusive, i.e. a tensor element is usually shared among different parities. In this way, this tensor can be positive in one domain, negative in the second, and zero in the third one. Thus, the existence of a well defined parity of tensor elements is necessary for domain imaging, but not sufficient for $60^\circ$ and $120^\circ$ domain structure. This unleashes an interesting question of the antiferromagnetic order parameter. There are as many order parameters as different domain structures for a given configuration. For $60^\circ$ and $120^\circ$ domain structures, the AF order parameter must be a vector, while for mirror domains it is a number. For $90^\circ$ domains it is can be also a number, since there are only {\em two} $90^\circ$ domains. The vectorial order parameter transforms itself under the domain operation like a usual vector. It is necessary to mention at this point that taking into account the spin structure in the {\em second layer} would not change the validity of the analysis presented in this subsection. The only modifications would result from addressing bulk domains rather then surface domains, and the symmetry of the AF configurations would be changed. Yet it would still be possible to find domain operations as well as odd and even tensor elements leading to interference and AF domain contrast. However the possibility to identify each of the domains may be limited in some cases due to the increased number of domains. \section{Possible experimental setups} \noindent In this section, we propose and discuss possible experimental setups for the detection of AF configuration and the imaging of AF domains from low-index surfaces of NiO that exhibit magnetic unit-cell doubling in contrast to bulk $Cr_2O_3$ \cite {ref18,ref4}. We propose an experimental setup for the {\em detection} of antiferromagnetism in the following way: both the incident and reflected beams may lie in the $xz$ plane (optical plane), and form the angle $\vartheta$ with the $z$-axis (normal to the sample surface). In the plane perpendicular to the outgoing beam axis, the electric field of the second-harmonic generated light has two components, $E^{(2\omega)}_p$ and $E^{(2\omega)}_s$, given by the formulae \begin{equation} \begin{split} \label{electric} |E^{(2\omega)}_p| &=|\cos \vartheta E^{(2\omega)}_x-\sin \vartheta E^{(2\omega)}_z| \\ |E^{(2\omega)}_s| &=|E^{(2\omega)}_y| \end{split} \end{equation} $E^{(2\omega)}_x$, $E^{(2\omega)}_y$, and $E^{(2\omega)}_z$ are the components of the electric field resulting from SHG in the coordinate system of the sample. The dependence of these components on the input electric field is indicated by the tensor $\chi^{(2\omega)}$. The aim of the experiment is the determination of vanishing and nonvanishing tensor elements. The easiest way to do this is to analyze the output signal intensity as a function of the input polarization in both output polarizations $s$ and $p$, for a fixed angle of incidence and reflection. The dependence of the output second-harmonic electric field on the input polarization is schematically displayed in Fig. \ref{fEl}a)-c) for all tensor elements. The intensity of SHG light is the square of the linear combination of these partial responses. An example of the intensity dependence on the input polarization is presented in Fig. \ref{fEl}d). The intensity need not be symmetric with respect to $\varphi = 90^{\circ}$, this results from the influence of the electric field depicted in Fig. \ref{fEl}c). The coefficients of the mentioned combination are the products of the $\chi^{(2\omega)}$ tensor elements and the corresponding Fresnel coefficients, according to Eq.~ (\ref{eqGlown}). Thus performing a best fit of these coefficients to the experimental results will give (after taking into account the Fresnel and geometrical coefficients, known for the given experimental geometry and material \cite{Sipe}) a set of non-vanishing elements of the $\chi^{(2\omega)}$ tensor. Thus for instance, the magnetic phase can be determined. Concerning another experimental geometry, with input polarization fixed and intensity measured as a function of the output polarization, it is possible to determine whether the nonlinear Kerr effect takes place. For instance, with the input polarization $\varphi=90^\circ$, the output electric field is given as follows \cite{Sipe}: \begin{multline} \label{varout} E^{(2\omega )}=\sin \Phi (A_2(\Theta)\chi^{(2\omega)}_{yyy}B_2(\vartheta))+\\ \cos \Phi (A_1(\Theta)\chi^{(2\omega)}_{xyy}B_2(\vartheta) + A_3(\Theta)\chi^{(2\omega)}_{zyy}B_2(\vartheta)) \end{multline} As the result, maximum of the intensity is for $\Phi \neq 90^\circ$, if at least one of the tensor elements $\chi^{(2\omega)}_{xyy}$ or $\chi^{(2\omega)}_{zyy}$ does not vanish. Actually, tensor element $\chi^{(2\omega)}_{zyy}$ is even in all the investigated order parameters, but the tensor element $\chi^{(2\omega)}_{xyy}$ can be odd. For such configurations the Kerr effect (change of polarization caused by inversion of the magnetic order parameter) takes place. Thus, it is possible to determine which tensor elements are associated with the spin-orbit coupling. The geometry with p polarization of the reflected SHG light seems to be less useful, since there the tensor element $\chi^{(2\omega)}_{zzz}$ is always present, regardless of the configuration. Besides, this polarization mixes the $\chi^{(2\omega)}_{x..}$ and $\chi^{(2\omega)}_{z..}$ tensor elements. This mixing, however, can be tuned by varying the angle of incidence $\vartheta$ and taking into account the influence of the Fresnel coefficients. For smaller $\vartheta$ only the $\chi^{(2\omega)}_{x..}$ elements are important, while for larger $\vartheta$ the $\chi^{(2\omega)}_{z..}$ dominate. If the experiment does not show any difference for these two situations, the tensor elements must be related. This is the possibility to distinguish the configurations with some relations between the tensor elements from those without such relations. On the other hand, the p polarization is useful for AF domain {\em imaging}. Thus one of the experimental possibilities is to carry out the measurements first in s polarized outgoing SHG light to make sure that the material is in the AF phase and determine its spin configuration. Then a second measurement in p polarization can be performed for the domain imaging. \section{Conclusions} \noindent Already a short look at the presented tables shows that our method works best if the paramagnetic phase is of high symmetry, since then a wide variety of different symmetries exists which may be broken by different spin configurations. In other words, there is enough room for different new tensor elements to appear along with different spin ordering under these circumstances. In general, this is the main reason why only nonlinear optics is suited for the detection of antiferromagnetism and the imaging of AF domains. The linear susceptibility tensor has too low a number of elements for these purposes in order to produce unambiguous results. Similarly, among the considered surfaces, the (110) surface is the least useful for the analysis as it yields ambiguous signal interpretations due to its low symmetry in the paramagnetic phase, and, on the other hand, very similar symmetries in all the AF configurations. The (001) and (111) surfaces present alike possibilities of distinction between the cases. If more than one monolayer is involved, however, the (111) surface will give the same response in the cases A (all atoms equivalent) and C (two kinds of magnetic atoms). Both the (001) and (111) surfaces also allow for the determination of the spin structure, provided the case is known. The (111) surface in the case D (oxygen sublattice distortion) is an exception - all the AF configurations produce the same response. It is possible, however, to determine the phase of the material. The case D appears to be a subgroup of the case C, i.e. all the magnetic point groups describing the configurations of the case D are subgroups of the corresponding ones in the case C. The only exception is the (110) surface. This inclusion means that the oxygen sublattice distortion makes some (one half of all) magnetic atoms distinguished as in case C, even though we did not apply this distinction explicitly in case D. From the fact, that the influence of oxygen sublattice distortion (case D) is not detectable in the paramagnetic and ferromagnetic phases it follows that only antiferromagnetic ordering can give an extensive information about the structure of the surface. It is the magnetic atoms and their magnetism which reveal the presence and position of oxygen. Our short analysis of an AF bilayer structure (surface (001)) indicates very similar features to the (001) monolayer. There exists a possibility to distinguish AF configurations from each other, and a certain possibility to detect the magnetic phases. Furthermore, introducing the second atomic layer does not affect the possibility to image AF domains. Concerning the {\em magnetic} phases, configurations and cases considered in this paper, some {\em a priori} information about the {\em structure} is needed in order to draw unique conclusions from the experimental results. For the detection of the phase and the spin configuration this additional information is the case (A, B, C, D). Vice versa, the case (for instance a possible distortion of the oxygen sublattice) can be determined if one knows the configuration (and if it had been previously deduced that the investigated material is antiferromagnetic). Actually, in most measurements of AF spin structures some {\em a priori} knowledge is required. For example, in experiments by Fiebig {\em et. al.} \cite{ref4} such a prerequisite is the assumption of the AF spin-flop phase of the material. In both experimental approaches mentioned here the (001) surface seems to provide the best possibilities of drawing valuable conclusions, while the (110) surface is the least suitable in that respect. Finally, our paper demonstrates that the AF domain imaging is possible even in the presence of magnetic unit-cell doubling. Thus optical SHG, unlike linear optics, is able to image AF {\em surface} domains. For most AF configurations, there are more than one surface domain structures. The rule stating that the number of domains is equal to the number of symmetry operations in the paramagnetic phase divided by the number of symmetry operations in the magnetic phase is applicable also for antiferromagnets (thus, with unit-cell doubling the number of domains is reduced by a half). However, not all the domains can be imaged at the same time. \section*{ Appendix \\ On the group-theoretical analysis of magnetic systems} \noindent In this Appendix, we would like to address some particularities of our group-theoretical analysis. The first general remark is that although symmetry analysis can provide us with a set of nonvanishing tensor elements for a given configuration, but cannot give any information about their magnitude. This equally applies to the distortion effects, as treated e.g. in Ref. \cite{ref56}. Another interesting issue is the behavior of the tensor elements with respect to the AF order parameter {\bf L} (for ferromagnetic phases {\bf L} should be replaced by the magnetization {\bf M}), i.e. the parity of tensor elements. In general, a tensor element consists of even and odd parts with respect to {\bf L}, as shown in Eq. (\ref{element}). In systems with high symmetry, it is possible to describe an operation which reverses {\bf L} (or {\bf M}) by a spatial operation $\Hat{l}$. The operation $\Hat{l}$ belongs to the point group of the system, but not to its magnetic point group. The application of this operation to a tensor element will change its sign (keep it invariant) if this element is odd (even) in {\bf L}. Consequently, each tensor element can be either odd or even in {\bf L}, a mixed behavior is forbidden. Actually, the parity of a given tensor element is a function of the chosen operation $\Hat{l}$. In most antiferromagnetic configurations more than one operation leading to different domain structures are possible (this means, more than one order parameter can be defined). For example, for (001) surface one has $4_z$ rotations leading to different domains {\em in addition} to the eventual mirror-structure. For the (111) surface, there are three domains resulting from the rotations with respect to the $z$ axis alone. For some configurations, they exist in addition to the mirror-domains. This whole analysis of the parity of the tensor elements cannot be performed for the systems with a lower symmetry, where it is impossible to find an operation $\Hat{l}$ describing the inversion of {\bf L} or {\bf M}, a mixed behavior is then allowed. Note, the presence of dissipation (redistribution of response frequencies) does not influence the above consideration. In general, dissipation in frequency space is responsible for the mixing of the real and imaginary parts in the tensor elements, while point-group symmetry governs the (non)existence of tensor elements purely odd or even in the magnetic order parameters {\bf L} or {\bf M}. We acknowledge financial support by TMR network through NOMOKE contract no. FMRX-CT96-0015. We also acknowledge numerous fruitful discussions with Dr. R. Vollmer.
\section{Introduction} \label{SEC:INT} The normal state of underdoped cuprates exhibits unusual magnetic properties which are believed to be intimately related to the mechanism of high-$T_c$ superconductivity. Most peculiar in this respect are the simultaneous occurrence of a magnetic pseudogap and the persistence of antiferromagnetic (AF) correlations as holes are doped into the antiferromagnetic insulator and the system becomes metallic. It is one of the most challenging theoretical problems in the physics of high-$T_c$ cuprates to reconcile the gaplike features reminiscent of a spin liquid with the presence of antiferromagnetic correlations signaling the closeness of the system to a spin-ordered N\'eel state. Experimentally, insight into the nature of these anomalous features can be gained by introducing impurities into the magnetically active Cu sites. A subsequent NMR probe on nuclei coupled to the CuO$_2$ planes yields information on the local magnetic structure. In this paper we present a microscopic theory of the impurity-induced local spin polarization of CuO$_2$ planes and its impact on the NMR linewidth. Introducing magnetically active or inert impurities into underdoped cuprates leads, in both cases, to the formation of local magnetic moments. Specifically, Cu ($d^9$) with an effective in-plane spin $S=\case{1}{2}$ can be replaced by Ni ($d^{8}$) with $S=1$ or Zn ($d^{10}$) with $S=0$. Superconducting quantum interference device (SQUID) measurements of the macroscopic susceptibility \cite{MEN94} reveal an almost perfect $1/T$ Curie behavior. Recently, Bobroff {\it et al.} \cite{BOB97} presented NMR measurements on $^{17}$O for the underdoped compound $\text{YBa}_2(\text{Cu}_{1-x}M_x)_3\text{O}_{6.6}$, with $M$ = Zn or Ni. The polarization of Cu spins in the presence of impurities leads to a broadening of the NMR line. In contrast to the aforementioned SQUID measurement, the linewidth displays a marked non-Curie behavior, indicating an inherent temperature dependence of the polarizability of CuO$_2$ planes. This was suggested by Morr {\it et al.} \cite{MOR98} to be a clear indication for a temperature dependence of the AF correlation length. Still another interesting observation can be made by comparing the two experiments: While the NMR study shows nonmagnetic Zn to have a more pronounced effect on the linewidth than Ni, measurements of the macroscopic susceptibility reveal a reversed effect. Since only the NMR experiment is sensitive to a spatial variation of the spin polarization, a very different shape of the spin density induced by the two types of impurities can be inferred. In the following, we present a microscopic theory of moments induced by magnetic and nonmagnetic impurities in the spin gap phase of underdoped cuprates. We analyze the different nature of coupling between Cu and impurity spins and derive expressions for the local spin polarization of CuO$_2$ planes. The presence of the spin gap and of short-range AF correlations is shown to strongly modify the conventional Ruderman-Kittel-Kasiya-Yosida (RKKY) picture. Finally, we derive expressions for the NMR line broadening which account well for the peculiarities of the experimental data. \section{Impurity Model} \label{SEC:IMP} The relevant physics of the CuO$_2$ planes of high-$T_c$ cuprates is believed to be described by the large-$U$ Hubbard or $t$-$J$ model. The dualism between itinerant charge motion and local electron interaction that is inherent to these models can, in an approximate way, be captured by introducing separate quasiparticles for spin and charge degrees of freedom. Within this picture, the normal state of underdoped cuprates is viewed as a phase in which spins form singlet pairs while coherence between holes that would eventually lead to superconductivity has not been established. We follow this line of thinking but restrict ourselves to the magnetic sector of the Hilbert space. Our starting point is the spin-$\case{1}{2}$ AF Heisenberg model $H_J = J\sum_{\langle ij \rangle} \bbox{s}_i \bbox{s}_j$. Keeping in mind the presence of itinerant holes which prevent the system from developing long-range magnetic order, we treat this Hamiltonian within resonance valence bond (RVB) mean-field theory \cite{AND87}~-- this accounts well for the spin-liquid features of cuprates. The mean-field Hamiltonian is \begin{equation} H_{\text{RVB}} = -\sum_{\langle ij \rangle \sigma} \left( \Delta_{ij} f^{\dagger}_{i \sigma} f_{j \sigma}+ \text{H.c.} \right). \label{HRVB} \end{equation} Original spin operators $\bbox{s}_i$ have been expressed in terms of fermionic operators by $\bbox{s}_i = \case{1}{2} \sum_{\sigma\sigma'} \bbox{\tau}_{\sigma\sigma'} f^{\dagger}_{i\sigma} f_{i\sigma'}$ with Pauli matrix vector $\bbox{\tau} = (\tau^x,\tau^y,\tau^z)$. The local constraint prohibiting a double occupancy of sites has been relaxed to a global one. The mean-field bond parameter is $\Delta_{ij}=\Delta_{\bbox{\delta}}= J \sum_{\sigma}\langle f^{\dagger}_{i+\delta,\sigma} f_{i\sigma}\rangle^0$, where $\langle\cdots\rangle^0$ is the expectation value that corresponds to Hamiltonian (\ref{HRVB}). The phase of this mean-field parameter is yet undetermined and has to be chosen such as to resemble the experimental situation most closely. An appropriate choice for the spin gap regime is the flux phase\cite{AFF88} $\Delta_{\pm x} = i\Delta_{\pm y} \equiv \Delta$. Dividing the lattice into two sublattices $A$ and $B$ and going to the momentum representation, Hamiltonian (\ref{HRVB}) can be diagonalized \begin{equation} H_{\text{RVB}} = \sum_{\bbox{k}\nu} \xi^{\nu}_{\bbox{k}} f^{\dagger}_{\bbox{k}\nu} f_{\bbox{k}\nu}, \label{HRVB2} \end{equation} with index $\nu=\pm$. The spectrum of spin excitations or spinons is \[ \xi^{\pm}_{\bbox{k}} = \pm 2\Delta\left(\cos^2k_x+\cos^2 k_y\right)^{1/2}. \] It has nodes at $(\pm\pi/2,\pm\pi/2)$, yielding a V-shaped pseudogap in the density of states centered at the spinon chemical potential $\mu_s=0$: $\rho^{(0)}(\omega) = |\omega|/D^2$ (defined per spin up/down state), where $D=2\sqrt{\pi}\Delta$ is the spinon half-band-width. To simulate first a nonmagnetic Zn impurity we introduce into Hamiltonian (\ref{HRVB}) a local chemical potential $\lambda$ acting on site $\bbox{R}=0$, which by convention lies on sublattice $A$. In the limit $\lambda\rightarrow\infty$ spinons are expelled from that site, creating a vacancy. The Hamiltonian is then \begin{equation} H_{\text{Zn}} = H_{\text{RVB}} + \lambda \sum_{\sigma} f^{\dagger}_{0\sigma} f_{0\sigma}\big|_{\lambda\rightarrow \infty}. \label{HZN} \end{equation} To describe a magnetic Ni impurity, into the empty site we insert an impurity spin $\bbox{S}_0$ with $S=1$ which is coupled antiferromagnetically to the surrounding Cu spins $\bbox{s}_{\bbox{\delta}}$. The corresponding Hamiltonian is \begin{equation} H_{\text{Ni}} = H_{\text{RVB}} + \lambda \sum_{\sigma} f^{\dagger}_{0\sigma} f_{0\sigma}\big|_{\lambda\rightarrow \infty} + H_{\text{imp}} \label{HNI} \end{equation} with the exchange interaction term \[ H_{\text{imp}} = J'\sum_{\bbox{\delta}} \bbox{S}_0 \bbox{s}_{\bbox{\delta}}. \] Formally, Hamiltonian (\ref{HNI}) differs from Eq.\ (\ref{HZN}) only in the presence of an additional term $H_{\text{imp}} \propto J'$. In the following, we put emphasis on the case of a magnetic impurity with $J'>0$. A nonmagnetic impurity can be simulated by setting $J'=0$, which decouples the impurity site from the rest of the system. The $S=1$ impurity spin is then free and can easily be disregarded. We discuss this limit in the following, but only shortly. More detailed treatments on nonmagnetic impurities are given in Ref.\ \onlinecite{KKK97} as well as in Refs.\ \onlinecite{NAG95}-\onlinecite{PEP98}. \section{Local Magnetic Moments} \label{SEC:LOC} We analyze an impurity spin $S=1$ embedded in a spin gap system as described by Hamiltonian (\ref{HNI}). Spinons stemming from the initial Cu spin at site $\bbox{R}=0$ are ejected by the local potential $\lambda$. The impurity spin, which is placed in the vacant site, is conveniently represented by two spins $\case{1}{2}$, i.e., $\bbox{S} = \bbox{S}_a + \bbox{S}_b$. An infinitely strong ferromagnetic interaction $H_{c} = -J_c\bbox{S}_a\bbox{S}_b$ between these two spins is assumed. Expressing $\bbox{S}_a$ and $\bbox{S}_b$ in terms of fermionic operators $a_{\sigma}$ and $b_{\sigma}$, respectively, a mean-field decoupling can be performed: \begin{equation} H_{\text{imp}} = -\sum_{\bbox{\delta} \sigma} \left( \Delta'_{\bbox{\delta}} \frac{a^{\dagger}_{\sigma} + b^{\dagger}_{\sigma}}{\sqrt{2}} f_{\bbox{\delta} \sigma} + \text{H.c.} \right) - J_c \bbox{S}_{a} \bbox{S}_{b}. \label{HIM1} \end{equation} Introducing operators $f_{0\sigma} = \left(a_{\sigma} + b_{\sigma}\right)/\sqrt{2}$ and $d_{\sigma} = \left(a_{\sigma} - b_{\sigma}\right)/\sqrt{2}$, one obtains \begin{equation} H_{\text{imp}} = -\sum_{\bbox{\delta} \sigma} \left( \Delta'_{\bbox{\delta}} f^{\dagger}_{0\sigma} f_{\bbox{\delta}\sigma} + \text{H.c.} \right) - J_{c} \bbox{S}_{\text{eff}} \bbox{s}_0 . \label{HIM2} \end{equation} The impurity spin has thus been decomposed into two $S=\case{1}{2}$ effective spins $\bbox{S}_{\text{eff}}$ and $\bbox{s}_0$. The former is represented by operators $d_{\sigma}$, the latter by operators $f_{0\sigma}$. Due to the first term in Eq.\ (\ref{HIM2}), the $f$ spinons on the impurity site hybridize with the ones on adjacent Cu sites. This process is controlled by the local mean-field parameter $\Delta'_{\bbox{\delta}} = J'\sum_{\sigma}\langle f_{\bbox{\delta}\sigma}^ {\dagger} f_{0\sigma}\rangle$ replacing $\Delta_{\bbox{\delta}}$ on bonds connecting to the impurity. A system of itinerant spinons extending over the whole lattice including the impurity site is formed. These itinerant spinons couple ferromagnetically to the localized spin $\bbox{S}_{\text{eff}}$. In the presence of a magnetic field this coupling is responsible for a polarization of the spinon system to be discussed in Sec.\ \ref{SEC:SPI}. The $T$ matrix that describes scattering of spinons on the localized spin vanishes as $T(\omega) \propto \omega\ln|\omega|$ in the flux phase. \cite{KKK97} This means that the effective local spin $\bbox{S}_{\text{eff}}$ becomes asymptotically free in the limit of low energies. In the remainder of the present section we analyze this low-energy fixed point, emphasizing the role of bond parameters $\Delta'_{\bbox{\delta}}$ that induce an inhomogeneity in the spinon sector. First we consider the special case of equal exchange integrals $J'=J$. Regarding the spinon sector, the impurity site becomes indistinguishable from the rest of the system as $\Delta'_{\bbox{\delta}} = \Delta_{\bbox{\delta}}$. The spin $\bbox{s}_0$ takes the role of the original Cu spin at $\bbox{R}=0$, and a homogeneous spin liquid, as described by $H_{\text{RVB}}$, Eq.\ (\ref{HRVB}), is formed. Generally, the two exchange integrals differ, $J' < J$, and translational invariance of the spinon system is broken. The bond parameter then acquires an additional spatial dependence which has to be treated self-consistently. To simplify the discussion, however, we distinguish only between bonds that do and do not connect to the impurity (see Fig.\ \ref{FIG:BON}), respectively: \[ \Delta_{ij} = \left\{ \begin{array}{l} \Delta'_{\bbox{\delta}}\quad\text{for}\quad i=0 \; \text{or} \; j=0\\ \Delta_{\bbox{\delta}}\quad\text{for}\quad i,j\ne 0, \end{array}\right. \] where $\Delta_{\bbox{\delta}}$ is the mean-field parameter of the impurity-free system. \begin{figure} \noindent \centering \epsfxsize=0.4\linewidth \epsffile{bond.eps}\\[6pt] \caption{Mean-field parameters $\Delta'_{\bbox{\delta}}$ and $\Delta_{\bbox{\delta}}$ are assigned to bonds that do (dashed line) or do not (solid line) connect to the impurity site denoted by a large dot.} \label{FIG:BON} \end{figure} The two parameters $\Delta'_{\bbox{\delta}}$ and $\Delta_{\bbox{\delta}}$ are assumed to exhibit the same phase relation, but in general their amplitudes differ. As a result, spinons scatter on the impurity bonds. To study this effect we write the spinon part of Hamiltonian (\ref{HNI}) as \begin{equation} H^{\text{sp}}_{\text{Ni}} = H_{\text{RVB}} + (1-x) \sum_{\bbox{\delta} \sigma} \left(\Delta_{\bbox{\delta}} f^{\dagger}_{0 \sigma} f_{\bbox{\delta} \sigma} + \text{H.c.} \right), \label{HSC} \end{equation} where $H_{\text{RVB}}$ represents the impurity-free system. The scattering amplitude $(1-x)$ with $x=|\Delta'_{\bbox{\delta}}/ \Delta_{\bbox{\delta}}|$ is controlled by the ratio of $J'$ to $J$. It vanishes for $J'=J$, and has to be treated self-consistently for $J'<J$. Approximately, we find $x=J'/J$. At this point, we introduce spinon propagators $g^{(0)}_{\lambda}(i\omega) = -\langle T_{\tau} f_{\lambda}(\tau) f^{\dagger}_{\lambda}(0) \rangle^0_{i\omega} = (i\omega-\xi_{\lambda})^{-1}$ and $g_{\lambda\lambda'}(i\omega) = -\langle T_{\tau} f_{\lambda}(\tau) f^{\dagger}_{\lambda'}(0)\rangle_{i\omega}$ for the pure and impurity-doped system. These can be related by a scattering matrix $T_{\lambda \lambda'}(i \omega)$: \begin{equation} g_{\lambda\lambda'}(i \omega) = g^{(0)}_{\lambda}(i \omega) \delta_{\lambda \lambda'} + g^{(0)}_{\lambda}(i \omega) T_{\lambda \lambda'} (i\omega)g^{(0)}_{\lambda'}(i \omega). \label{PRO} \end{equation} A simplified notation $\lambda = (\bbox{k},\nu)$ and Matsubara frequencies $i\omega = i(2n+1)\pi T$, where $T$ denotes temperature and $n$ integer numbers, are employed. The $T_{\lambda\lambda'}$ matrix in Eq.\ (\ref{PRO}) describes scattering of spinons on the four bonds that connect the impurity site to its nearest neighbors. We find it to be given by the expression \begin{equation} T_{\lambda \lambda'}(i \omega) = \frac{t_{\lambda \lambda'} (i\omega)}{i \omega G^{(0)}(i \omega) + p^2}, \label{TMA} \end{equation} with \begin{eqnarray*} t_{\lambda \lambda'}(i \omega)&=& \frac{1-x}{1+x} G^{(0)}(i \omega)(i \omega - \xi_{\lambda})(i \omega - \xi_{\lambda'}) \nonumber\\ &&+\frac{x}{1+x}(2i \omega - \xi_{\lambda} - \xi_{\lambda'}) - i\omega. \end{eqnarray*} Here $G^{(0)}(i\omega) = \sum_{\lambda} g^{(0)}_{\lambda}(i\omega) = -(2i\omega/D^2) \ln(D/|\omega|)$, and $p^2 = x^2/(1-x^2)$. The important point is that in the flux phase the scattering matrix of Eq.\ (\ref{TMA}) has two poles that are determined by the roots of \begin{equation} \omega G^{(0)}(i \omega \rightarrow \omega + i0^+) + p^2 = 0. \label{BOU} \end{equation} One of the poles lies below the spinon chemical potential which signals the formation of a spinon bound state. This can be interpreted as follows: Due to impurity substitution, one Cu spin loses its RVB singlet partner. In a spin gap system in which short-range spin-singlet correlations dominate, this unpaired spin does not dissolve into the RVB ground state but rather forms a local moment distributed over Cu sites in the proximity of the impurity. At finite coupling $J'$ this moment forms a local singlet with the impurity-site spinon $f_{0\sigma}$. The characteristic binding energy $\omega_K$ and lifetime $\delta_K$ of the resulting bound state are given by the real and imaginary part of the pole, respectively. For $J'\ll J$, one obtains \begin{equation} \omega_K = \frac{\pi}{4} \frac{J'}{\ln D/J'}, \quad \delta_K = \frac{\pi}{4} \frac{\omega_K}{\ln D/\omega_K}. \label{WDK} \end{equation} In the following, two different energy scales are distinguished: $\omega < \omega_K$ and $\omega>\omega_K$. These control the physics at large and short distance from the impurity as compared to $R_K = D/\omega_K$, respectively, where $R_K$ is measured in units of lattice spacing. First we analyze the low-energy fixed point of the system with a magnetic impurity for which $J'$ and hence $\omega_K$ are finite. It is determined by the regime $\omega<\omega_K$ and applies to distances $R>R_K$ from the impurity site. We calculate the impurity contribution $\delta\rho(\omega)$ to the density of states from the Green's function $\delta G(i\omega) = \sum_{\lambda\lambda'} g_{\lambda}^{(0)}(i\omega) T_{\lambda\lambda'} (i\omega)g_{\lambda'}^{(0)}(i\omega) = (\partial/\partial i\omega)\ln [i\omega G^{(0)}(i\omega)+p^2]$. For $\omega\ll D$, the latter is \begin{equation} \delta G(i\omega) = \frac{2 G^{(0)}(i\omega)}{i\omega G^{(0)}(i\omega)+p^2}, \end{equation} which yields \begin{equation} \delta \rho (\omega) = \frac{2}{\pi}\omega_K\delta_K \frac{|\omega|}{(\omega^2-\omega^2_K)^2+(2\omega_K\delta_K)^2}. \label{DOS1} \end{equation} Figure \ref{FIG:DOS}(a) schematically shows the spinon density \begin{figure} \noindent \centering \epsfxsize=0.42\linewidth \epsffile{plot1.eps} \hspace{0.05\linewidth} \epsfxsize=0.42\linewidth \epsffile{plot2.eps}\\[6pt] \caption{Schematic plot of the spinon density of states for (a) $J'>0$ and (b) $J'=0$ corresponding to magnetic and nonmagnetic impurities, respectively. Solid lines represent the impurity-doped system, dashed lines the pure system. The spinon chemical potential $\mu_s$ lies in the center of the gap, and the $\delta$ function in (b) is artificially broadened.} \label{FIG:DOS} \end{figure} of states $\rho^{(0)}(\omega)$ and $\rho(\omega)=\rho^{(0)}(\omega) +\delta\rho(\omega)$ for the pure and impurity-doped system. The very existence of a magnetic pseudogap is found to be unaffected by the presence of the impurity~-- $\rho^{(0)}(\omega)$ as well as $\rho(\omega)$ vanish linearly in the limit $\omega\rightarrow 0$. As a consequence, the static spin susceptibility, which is related to the spinon density of states by \cite{REM1} \begin{equation} \chi(T) = \frac{1}{4T} \int_{-\infty}^{\infty} dx \frac{\rho(x)}{\cosh^2(x/2T)}, \end{equation} vanishes as $\propto T$ at low temperatures. This indicates that in the low-energy limit all spins (except $\bbox{S}_{\text{eff}}$ which is not part of the spinon system) participate in the formation of singlets. The spinon bound state discussed above hence partially screens the impurity spin by forming a Kondo singlet with $\bbox{s}_0$. An effective $S=\case{1}{2}$ impurity spin $\bbox{S}_{\text{eff}}$ remains. In this underscreened Kondo problem, the spinon binding energy $\omega_K$ of Eq.\ (\ref{WDK}) plays the role of the Kondo temperature: $T_K=\omega_K$. For temperatures $T\gg T_K$, the susceptibility associated with the spinon bound state is that of a free spin $\case{1}{2}$, i.e., $\chi(T)=1/(4T)$; simultaniously, the original $S=1$ impurity spin is recovered. We note that the Kondo temperature exhibits an unconventional power-law dependence on the coupling parameter $J'$, contrasting the conventional exponential behavior. This peculiarity is ascribed to the fact that the impurity spin couples to bound spinons which are predomenantly in localized rather than bandlike states. Finally, we shortly discuss how the presence of a Kondo singlet affects the properties of the spinon system at $T\ll T_K$. Although the impurity does not fill the magnetic pseudogap, it nevertheless renormalizes its slope. The leading term in a low-energy expansion of Eq.\ (\ref{DOS1}) is related to the density of states of the pure system by \begin{equation} \delta\rho(\omega) = \frac{1}{p^2} \rho^{(0)}(\omega), \label{DOS2} \end{equation} valid for $J'\ll J$. At low energy and large distance from the impurity, the spinon system hence behaves qualitatively as in the impurity-free case. To finish the discussion of magnetic moments, we turn to the case of a nonmagnetic impurity. The relevant physics is modelled by decoupling the spinon sector from the impurity site, setting $J'=0$, and by discarding contributions stemming from the impurity spin which is now free. Since $\omega_K$ consequently vanishes, one is always in the regime $\omega>\omega_K$. A Kondo singlet cannot form even in the zero-energy limit as the impurity carries no inherent spin. The spinon bound state induced by the impurity lies at the spinon chemical potential in the center of the pseudogap [see Fig.\ \ref{FIG:DOS}(b)]: \begin{equation} \delta\rho(\omega) = \delta(\omega). \end{equation} This is associated with the magnetic susceptibility $1/(4T)$ of a free spin $\case{1}{2}$ which holds down to zero temperature. We note that the impurity-induced moment is broadly distributed over planar Cu sites on sublattice $B$ that does not contain the impurity site, its density falling off as $R^{-2}$ with distance from the impurity. \cite{KKK97} To summarize, magnetic Ni and nonmagnetic Zn impurities are both associated with $S=\case{1}{2}$ magnetic moments. These are, however, of very different natures (see Fig.\ \ref{FIG:RVB}): \begin{figure} \noindent \centering \hfill \epsfxsize=0.4\linewidth \epsffile{rvb1.eps} \hfill \epsfxsize=0.4\linewidth \epsffile{rvb2.eps} \hfill\\[6pt] \caption{``Snapshot'' of the low-energy fixed point of a RVB liquid state with (a) the $S=1$ magnetic impurity and (b) the nonmagnetic impurity denoted by a dot. In the former case, the impurity spin is partially screened by moments induced on Cu sites. Effectively, a local impurity spin $\case{1}{2}$ and a ``healed'' spin liquid results. In the latter case, the impurity induces a broadly distributed moment that resides on Cu sites in the proximity of the impurity.} \label{FIG:RVB} \end{figure} In the former case, the spinon bound state partially screens the original $S=1$ impurity spin. One is left with an effective impurity spin $\case{1}{2}$ ferromagnetically coupled to an ensemble of inherent spinons that, in the absence of a magnetic field, behaves qualitatively the same as an impurity-free system. In the latter case, the moment is carried by the spinon bound state itself, and is broadly distributed over Cu sites. \section{Spin Polarization} \label{SEC:SPI} The effective impurity moments discussed in Sec.\ \ref{SEC:LOC} can be polarized by applying an external magnetic field. In this section we analyze the incidental local response of planar Cu spins. In the case of a magnetic impurity, the applied field acts on a localized impurity spin $\case{1}{2}$ ferromagnetically coupled to the spin liquid. Cu spins respond via a RKKY-type interaction. In the case of a nonmagnetic impurity, the moment itself resides on Cu sites. Applying a magnetic field therefore directly polarizes the Cu spins. We first discuss the situation of a magnetic impurity. The static polarizability is defined by $K_{\text{Ni}}(T,\bbox{R}) = \langle T_{\tau} s^z_{\bbox{R}}(\tau) S_{\text{eff}}^z(0) \rangle_{\omega=0}$, where $s^z_{\bbox{R}}$ and $S_{\text{eff}}^z$ denote the $z$ components of a given Cu spin at site $\bbox{R}$ and of the effective impurity spin, respectively. It is expressed in terms of Green's functions as (see Fig.\ \ref{FIG:BUB}) \begin{figure} \noindent \centering \epsfxsize=0.8\linewidth \epsffile{bubble.eps}\\[12pt] \caption{Diagrammatic representation of the static polarization $K_{\text{Ni}}(T,\bbox{R})$ of a Cu spin at site $\bbox{R}$ due to RKKY coupling to the localized impurity moment. Dashed and solid ovals represent particle-hole convolution functions $\Pi_d(i\varepsilon')$ for the local moment and $\Pi_f(i\varepsilon,\bbox{R})$ for itinerant spinons, respectively. The effective coupling is described by the vertex function $J_c(i\varepsilon-i\varepsilon')$ denoted by a circle.} \label{FIG:BUB} \end{figure} \begin{equation} K_{\text{Ni}}(T,\bbox{R}) = -T^2\sum_{\varepsilon,\varepsilon'} \Pi_d(i\varepsilon') J_c(i\varepsilon-i\varepsilon') \Pi_f(i\varepsilon,\bbox{R}) \label{POL} \end{equation} with particle-hole convolution functions \begin{eqnarray*} \Pi_d(i\varepsilon) &=& D^2(i\varepsilon),\\ \Pi_f(i\varepsilon,\bbox{R}) &=& G(i\varepsilon,-\bbox{R}) G(i\varepsilon,\bbox{R}). \end{eqnarray*} Here the impurity Green's function is $D(i\omega) = -\langle T_{\tau} d_{\sigma}(\tau) d^{\dagger}_{\sigma}(0)\rangle_{i\omega} = 1/(i\omega)$, and the intersite spinon Green's function $G(i\omega,\bbox{R}) = -\langle T_{\tau} f_{0\sigma}(\tau) f^{\dagger}_ {\bbox{R}\sigma}(0)\rangle_{i\omega}$. Operators $d$ and $f$ act on separated sectors of the Hilbert space. At site $\bbox{R}=0$, however, $f$ spinons are polarized by the local spin $\bbox{S}_{\text{eff}}$ due to the ferromagnetic interaction of bare strength $J_c$. This coupling is accounted for by the vertex function $J_c(i\omega)$. Employing a ladder approximation it is \begin{equation} J_c(i\omega) = \frac{J_c}{1+J_c \Pi_c^{i\omega}} = \frac{1} {\Pi_c^{i\omega}}, \label{LAD} \end{equation} with \[ \Pi_c^{i\omega} = -T\sum_{\varepsilon} D(i\varepsilon+i\omega) G(i\varepsilon,\bbox{R}=0). \] The second equality in Eq.\ (\ref{LAD}) holds due to $J_c$ being infinitely large. Replacing the vertex function by its zero-frequency limit, $J(i\omega)\rightarrow J(0)$, the polarizability of Eq.\ (\ref{POL}) can be factorized. Within this approximation, which is valid at low temperatures, one obtains \begin{equation} K_{\text{Ni}}(T,\bbox{R}) = \chi_{\text{eff}}(T) J_c(0) \chi_{\text{pl}}(T,\bbox{R}). \label{POL2} \end{equation} The polarizability has thus been decomposed into the magnetic susceptibility of the effectively free $\case{1}{2}$ impurity spin, $\chi_{\text{eff}}(T)=1/(4T)$, the nonlocal magnetic susceptibility of CuO$_2$ planes, $\chi_{\text{pl}}(T,\bbox{R})$, and an effective coupling parameter $J_c(0)$. The susceptibilities are defined as $\chi_{\text{eff}}(T) = \langle T_{\tau} S^z_{\text{eff}}(\tau) S^z_{\text{eff}}(0) \rangle_{\omega=0} = -T\sum_{\varepsilon} \Pi_d(i\varepsilon)$ and $\chi_{\text{pl}}(T,\bbox{R}) = \langle T_{\tau} s^z_{\bbox{R}}(\tau) s^z_0(0)\rangle_{\omega=0} = -T\sum_{\varepsilon} \Pi_f(i\varepsilon,\bbox{R})$. To further analyze the polarizability in Eq.\ (\ref{POL2}), $J_c(0)$ and $\chi_{\text{pl}}(\bbox{R})$ have to be evaluated. This requires the on-site and intersite spinon Green's functions \begin{equation} G(i\omega,\bbox{R}) = \left\{ \begin{array}{l} \displaystyle \left(\frac{p}{x}\right)^2 \frac{G^{(0)}(i\omega)}{i\omega G^{(0)}(i\omega)+p^2} \quad\text{for}\quad R=0\\ \displaystyle \frac{1}{x} G^{(0)}(i\omega,\bbox{R})\quad\text{for}\quad R>R_K, \end{array}\right. \label{GRE} \end{equation} where $G^{(0)}(i\omega,\bbox{R})$ is defined for the impurity-free system \begin{equation} G^{(0)}(i\omega,\bbox{R}) = -\frac{2i|\omega|}{D^2} \varphi(\bbox{R}) K_1\left(\frac{R |\omega|}{D}\right), \label{G0R} \end{equation} with a modified Bessel function of the second kind, $K_{\nu}(x)$. Equation (\ref{G0R}) holds for sites on sublattice $B$; contributions from sublattice $A$ containing the impurity are found to be negligible. The angular dependence is determined by the phase factor \begin{equation} \varphi(\bbox{R}) = \frac{1}{2}\left(\tilde{R}^+e^{i\pi R^+/2} + \tilde{R}^-e^{i\pi R^-/2}\right) \end{equation} with $R^{\pm} = R_x\pm R_y$ and $\tilde{R}^{\pm} = (R_x\pm iR_y)/R$. We are now in the position to calculate the effective coupling parameter from the zero-frequency limit of Eq.\ (\ref{LAD}), \begin{equation} J_c(0) = \left\{ \begin{array}{l} D \quad \text{for} \quad J'=J\\ 2\omega_K \quad \text{for} \quad J'\ll J, \end{array}\right. \label{JC0} \end{equation} and the nonlocal spin susceptibility of CuO$_2$ planes in the presence of the impurity, \begin{equation} \chi_{\text{pl}}(\bbox{R}) = -\frac{3}{4\pi} \frac{1}{Jx^2} \frac{\Phi(\bbox{R})}{R^3}, \label{SPL} \end{equation} the latter being valid for $\bbox{R}\in B$ with $R>R_K$. The phase factor in Eq.\ (\ref{SPL}) is defined by $\Phi(\bbox{R}) = |\varphi(\bbox{R})|^2$. Finally, combining these results, we obtain \begin{equation} K_{\text{Ni}}(T,\bbox{R}) = -\frac{3}{16\pi} \frac{J_c(0)}{Jx^2} \frac{\Phi(\bbox{R})}{R^3}\frac{1}{T} , \label{PNI} \end{equation} which describes the polarizability of a Cu spin at site $\bbox{R}\in B$ responding to a magnetic field that acts on the effective impurity spin $\bbox{S}_{\text{eff}}$; contributions from sublattice $A$ are found to be small. We note that the $T^{-1}$ Curie behavior displayed by Eq.\ (\ref{PNI}) stems solely from the susceptibility $\chi_{\text{eff}}(T)$ of the effective impurity spin. Within the present mean-field treatment, the planar susceptibility is independent of temperature: $\chi_{\text{pl}} (T,\bbox{R}) = \chi_{\text{pl}}(\bbox{R})$. We now briefly review the result for a nonmagnetic impurity which was derived in Ref.\ \onlinecite{KKK97}. Here, the Cu spins carry the impurity-induced moment, and can therefore be directly polarized by the magnetic field. The polarizability is given by the local susceptibility of the impurity-induced moment, $K_{\text{Zn}}(T,\bbox{R}) = \delta\chi_{\text{pl}}(T,\bbox{R}) = \sum_{\bbox{R}'}\left(\langle T_{\tau} s^z_{\bbox{R}}(\tau) s^z_{\bbox{R}'}(0) \rangle_{\omega=0} - \langle T_{\tau} s^z_{\bbox{R}}(\tau) s^z_{\bbox{R}'}(0) \rangle_{\omega=0}^0\right)$, yielding \begin{equation} K_{\text{Zn}}(T,\bbox{R}) = \frac{1}{2\pi} \frac{\Phi(\bbox{R})}{R^2} \frac{1}{T\ln D/T}. \label{PZN} \end{equation} Equation (\ref{PZN}) is valid for $\bbox{R}\in B$, while contributions from $\bbox{R}\in A$ are again negligible. The polarizability is found to decay slowly as $R^{-2}$ with distance from the impurity which compares to a $R^{-3}$ behavior in the case of Ni, reflecting the delocalized nature of the moment induced by a Zn impurity. Further, a logarithmic correction to the Curie-like temperature behavior is to be marked. In deriving Eqs.\ (\ref{PNI}) and (\ref{PZN}) for the polarizability of Cu spins, we have, up to this point, built upon RVB mean-field theory. This picture accounts well for the spin liquid features of underdoped cuprates including the presence of a magnetic pseudogap. Its strength lies on the description of long-range properties controlled by low-energy excitations. The mean-field treatment does, however, severely underestimate local AF correlations which reflect the proximity of a critical instability towards AF spin ordering. As a consequence, the above expressions contain no reference to the AF correlation length which was suggested to introduce a temperature dependence beyond the Curie behavior of free moments. \cite{MOR98} Furthermore, mean-field theory yields a polarizability of Cu spins on one sublattice only, undervaluing the staggered magnetization of spins on the opposite sublattice. This is in disaccord with NMR measurements \cite{BOB97} that yield no overall shift of the $^{17}$O line, as would be expected from the polarization of only one sublattice as well as with numerical studies. \cite{MAR98} To compensate for these deficiencies of the mean-field treatment, we simulate the closeness of the spin system towards an antiferromagnetically ordered state by performing a random-phase approximation (RPA) in the magnetic susceptibility. In the momentum representation, the susceptibility of planar Cu spins then becomes \begin{equation} \chi_{\text{pl}}^{\text{RPA}}(T,\bbox{q})=\chi_{\text{pl}}(\bbox{q}) S(T,\bbox{q}) \label{RPA} \end{equation} with the Stoner enhancement factor \begin{equation} S(T,\bbox{q}) = \frac{1}{1+J_{\bbox{q}} \chi_{\text{pl}}(T,\bbox{q})}, \label{STO1} \end{equation} where $J_{\bbox{q}} = 2J(\cos q_x+\cos q_y)$. We closely follow the theory of a nearly AF Fermi liquid,\cite{MON94} which maps Eq.\ (\ref{RPA}) onto a phenomenological expression involving the AF correlation length $\xi(T)$. Within this picture, $\chi_{\text{pl}}^{\text{RPA}}(T,\bbox{q})$ is assumed to be controlled solely by the momentum region close to the AF wave vector $\bbox{Q}=(\pi,\pi)$. However, we do take a slightly different point of view in this respect: The momentum dependence of the bare susceptibility $\chi_{\text{pl}}(\bbox{q})$ in Eq.\ (\ref{RPA}), which describes the long-range characteristics of spin correlations in the presence of a magnetic pseudogap, is explicitly kept. Only the scaling function $S(T,\bbox{q})$, which controls short-range AF correlations, is expanded around $\bbox{Q}=(\pi,\pi)$. Identifying $J\chi_{\text{pl}}(T,\bbox{Q})/[1-4J\chi_{\text{pl}}(T,\bbox{Q})]=\xi^2(T)$ and $1/[J\chi_{\text{pl}}(T,\bbox{Q})]=\alpha$, Eq.\ (\ref{STO1}) can be written in phenomenological form \begin{equation} S(T,\bbox{q})=\frac{\alpha\xi^2(T)}{1+(\bbox{q}-\bbox{Q})^2\xi^2(T)}, \label{STO} \end{equation} where $\alpha\approx 1$ on a mean-field level. We note that the explicit form of $\xi(T)$ lies beyond the accessibility of a mean-field treatment, and has to be chosen according to general physical considerations. Turning back to real space, the nonlocal susceptibility is \begin{equation} \chi^{\text{RPA}}_{\text{pl}}(T,\bbox{R}) = \sum_{\bbox{R}'\in B} \chi_{\text{pl}}(\bbox{R}') S(T,\bbox{R}-\bbox{R}'). \end{equation} For distances $R\gg\xi(T)$, it can be approximated by $\chi_{\text{pl}}^{\text{RPA}}(T,\bbox{R}) = \chi_{\text{pl}} (\bbox{R})\xi^2(T)/2$ with interpolation formula for the $A$ sublattice $\chi_{\text{pl}}(\bbox{R}\in A) = -(1/z)\sum_{\bbox{\delta}} \chi_{\text{pl}}(\bbox{R}+\bbox{\delta})$. Analogous expressions are obtained for the local susceptibility $\delta\chi_{\text{pl}} (T,\bbox{R})$ induced by a nonmagnetic impurity. Combining these results with Eqs.\ (\ref{PNI}) and (\ref{PZN}) and performing an angular average over phase factors $\Phi(\bbox{R})$, one finally arrives at the following expressions for the polarizability of Cu spins in the impurity-doped system: \begin{eqnarray} K_{\text{Ni}}(T,\bbox{R}) &=& \cos(\bbox{Q}\bbox{R}) \frac{3}{64\pi} \frac{J_c(0)}{Jx^2}\frac{1}{R^3} \frac{\xi^2(T)}{T}, \label{PNI2}\\ K_{\text{Zn}}(T,\bbox{R}) &=& -\cos(\bbox{Q}\bbox{R}) \frac{1}{8\pi}\frac{1}{R^2}\frac{\xi^2(T)}{T\ln D/T}. \label{PZN2} \end{eqnarray} These equations now hold for both sublattices, $\bbox{R} \in \{A,B\}$, the staggered nature of spin correlations being manifested in the alternating sign implied by $\cos(\bbox{Q}\bbox{R})$. Further, the dependence upon the AF correlation length $\xi(T)$ is now explicitly accounted for. \section{NMR Line Broadening} \label{SEC:NMR} The impurity-induced polarization of Cu spins in a magnetic field affects the energy levels of nuclear spins via supertransferred hyperfine interaction. The coupling of a given nuclear spin $\bbox{I}$ to electron spins $\bbox{s}_i$ on close by Cu sites is described by \begin{equation} H_{\text{hf}} = \gamma_n\gamma_e C_{\text{hf}}\sum_i\bbox{s}_i\bbox{I}, \end{equation} where $\gamma_n$ and $\gamma_e$ denote the nuclear and electron gyromagnetic ratios, respectively, and $C_{\text{hf}}$ is the supertransferred hyperfine coupling constant. In the following, we restrict ourselves to NMR measurements on $^{17}$O nuclei ($I=\case{5}{2}$). On a mean-field level, $\bbox{s}_i$ can be replaced by its average value $\langle \bbox{s}_i\rangle = K(T,\bbox{R}_i) \bbox{H}_0$ with external magnetic field $\bbox{H}_0$ and polarizability $K(T,\bbox{R}_i)$ given by either one of Eqs.\ (\ref{PNI2}) and (\ref{PZN2}) for the two types of impurities. Since each $^{17}$O nucleus lies symmetrically in between two Cu sites that belong to different sublattices with spins polarized in opposite directions, the impurity-induced energy shift partially cancels (see Fig.\ \ref{FIG:SPIN}). \begin{figure} \noindent \centering \epsfxsize=0.55\linewidth \epsffile{spin.eps}\\[6pt] \caption{Schematic cut through a CuO$_2$ plane, showing the position of O ions (diamonds) placed in between successive Cu sites with staggered spin polarization (arrows). The magnitude of the polarization falls off with distance from the impurity (circle) as $R^{-3}$ in the case of Ni, and as $R^{-2}$ in the case of Zn.} \label{FIG:SPIN} \end{figure} At large enough distance from the impurity, the shift is then effectively determined by the spatial derivative of the polarizability, \begin{equation} \omega(\bbox{R}) = \kappa \frac{\partial |K(T,\bbox{R})|}{\partial R} \cos\phi, \label{WRR} \end{equation} where $\phi$ denotes the angle enclosed by $\bbox{R}$ and the $x$ or $y$ axis and $\kappa = \gamma_n\gamma_e C_{\text{hf}}H_0$. In a system with randomly distributed impurities of concentration $c$, the superposition of energy shifts induced by different impurities leads to a broadening of the NMR line. We calculate the line shape that follows from Eq.\ (\ref{WRR}), employing the formalism of Ref.\ \onlinecite{WAL74}. The line shape function $g(\nu)$ is defined as the Fourier transform of the characteristic or free-induction function \begin{equation} f(t) = \exp\Big[-c\sum_{\bbox{R}} \left(1-e^{i\omega(\bbox{R})t} \right)\Big]. \end{equation} Integrating over lattice sites, for Ni and Zn, respectively, one obtains \begin{equation} \ln f(t) = \left\{ \begin{array}{l} \displaystyle -\Lambda_{\text{Ni}}|t|^{1/2}\\ \displaystyle -\Lambda_{\text{Zn}}|t|^{2/3}, \end{array}\right. \label{LNF} \end{equation} with \begin{eqnarray*} \Lambda_{\text{Ni}} &=& \frac{2\sqrt{6}\pi\Gamma(3/4)}{\Gamma(1/4)}\left[\kappa\frac{3}{64\pi} \frac{J_c(0)}{Jx^2} \frac{\xi^2(T)}{T}\right]^{1/2}c,\\ \Lambda_{\text{Zn}} &=& \frac{2\sqrt{3}\pi^2}{\Gamma^2(1/3)} \left[\kappa \frac{1}{8\pi} \frac{\xi^2(T)}{T\ln D/T}\right]^{2/3}c. \end{eqnarray*} Figure \ref{FIG:POL} shows the different line shapes induced by Ni and Zn impurities as obtained by performing a Fourier transformation on $f(t)$. \begin{figure} \noindent \centering \setlength{\unitlength}{0.85\linewidth} \begin{picture}(1.0,0.84) \put(0,0){ \epsfxsize=\unitlength \epsffile{pol.eps}} \put(0.2,0.75){ \fontsize{0.04\unitlength}{0.045\unitlength}\selectfont \begin{tabular}[t]{c|@{\protect\rule{0.03\unitlength}{0\unitlength}}l} $\eta$ & HWHH\\ \hline\rule{0\unitlength}{0.05\unitlength} $1/2$ & $0.22\,\Lambda^{2}$\\ $2/3$ & $0.51\,\Lambda^{3/2}$ \\ $1$ & $1.00\,\Lambda$ \end{tabular}} \end{picture}\rule{0.1\linewidth}{0\linewidth} \caption{Line-shape function $g(\nu)$ obtained by performing a Fourier transformation on the characteristic function $f(t) = \exp[-\Lambda |t|^{\eta}]$ with $\eta = \case{1}{2}$ for Ni and $\eta = \case{2}{3}$ for Zn. For comparison, the Lorentzian line shape of conventional RKKY theory with $\eta=1$ is indicated by a dashed line. Numerical values for the half width at half height (HWHH) are given in the inset.} \label{FIG:POL} \end{figure} Comparing to the Lorentzian shape that results from $\ln f(t)\propto -|t|$ in conventional RKKY theory, one finds a marked difference in both shape and width. Using the numerical values shown in the inset of Fig.\ \ref{FIG:POL}, we finally arrive at the following expressions for the full linewidth induced by magnetic and nonmagnetic impurities: \begin{eqnarray} \Delta\nu_{\text{Ni}} &=& 2 \times 0.22 \left(\Lambda_{\text{Ni}}\right)^2, \label{NNI}\\ \Delta\nu_{\text{Zn}} &=& 2 \times 0.51 \left(\Lambda_{\text{Zn}}\right)^{2/3}. \label{NZN} \end{eqnarray} \section{Comparison with Experiment} \label{SEC:COM} In this section, we compare the impurity-induced $^{17}$O NMR line broadening as described by Eqs.\ (\ref{NNI}) and (\ref{NZN}) with experimental data of Bobroff {\it et al.} \cite{BOB97} obtained on $\text{YBa}_2(\text{Cu}_{1-x}M_x)_3\text{O}_{6.6}$ with $M$ = Zn or Ni. The following constants are chosen: The superexchange parameters are specified by $J=0.13$~eV for Cu-Cu interaction and $J'=J/2$ for Cu-Ni. \cite{REM2} A self-consistent treatment yields a scattering amplitude $(1-x)=0.5$, where $\Delta=\frac{1}{4}$ has been used. The Kondo temperature is obtained by numerically solving Eq.\ (\ref{BOU}) which gives $T_K = 560$~K. Below this temperature, the Ni spin is partially screened and behaves as a spin $\case{1}{2}$ ferromagnetically coupled to the CuO$_2$ plane. The effective coupling constant of this interaction given by Eq.\ (\ref{JC0}) is $J_c(0) = 0.1$~eV. The hyperfine coupling constant between $^{17}$O nuclear and Cu electron spins is $C_{\text{hf}}=3.3$~T/$\mu_B$. \cite{MOR98} The magnetic-field strength used in the experiment is $H_0=7.5$~T, and the concentration of Ni and Zn impurities is $1\%$. The effective impurity concentration within CuO$_2$ planes, which is larger by a factor of $\case{3}{2}$, is finally $c=1.5\%$. Next, an expression for the AF correlation length $\xi(T)$ has to be specified. It is argued in Ref.\ \onlinecite{STO97} that below a critical temperature $T_{\text{cr}}$ specified by $\xi(T_{\text{cr}})\approx 2$, the correlation length assumes the form \begin{equation} \xi(T) = \frac{1}{a+bT}, \label{XIT} \end{equation} where $a$ and $b$ are fitting constants of the theory. Saturation of $\xi(T)$ at low temperatures is neglected here. Figure \ref{FIG:EXP} shows the impurity-induced line broadening $\Delta \nu_{\text{imp}}$ scaled with temperature. \begin{figure} \noindent \centering \epsfxsize=0.85\linewidth \epsffile{exp.eps} \caption{Impurity-induced NMR line broadening $\Delta\nu_{\text{imp}}$ multiplied by temperature. The theoretical result is indicated by solid lines fitted to experimental $^{17}$O data for $1\%$ Ni-doped (triangles) and $1\%$ Zn-doped (diamonds) $\text{YBa}_2(\text{Cu}_{1-x}M_x)_3 \text{O}_{6.6}$, $M$ = Ni or Zn.} \label{FIG:EXP} \end{figure} The curves are fitted to the experimental data by setting $a=0.07$ and $b=0.0007$, which correspond to an AF correlation length of $\xi=4.8$ in units of lattice spacings at $T=200$~K. This compares well to $\xi=5.9$ obtained in Ref.\ \onlinecite{BAR95}. No further fitting parameters are needed. The theory correctly accounts for the peculiar experimental observation of Zn having a more pronounced effect on the NMR signal than Ni~-- this seems to be in contradiction to SQUID measurements on the macroscopic susceptibility. \cite{MEN94} We are able to ascribe this behavior to the different spatial dependence of the polarizability: $K(T,\bbox{R})$ decays as $R^{-3}$ in the case of Ni, but only as $R^{-2}$ in the case of Zn. Averaging over all impurity site, this leads to an enhanced line-broadening effect of Zn (see Fig.\ \ref{FIG:POL}). Our theory further correctly describes the anomalous non-Curie temperature dependence exhibited by the NMR linewidth~-- this seems to be in disaccord with an almost perfect $T^{-1}$ behavior exhibited by the macroscopic susceptibility. \cite{MEN94} One can resolve this disagreement by assuming a temperature dependence of the AF correlation length $\xi(T)$ which enters the polarizability of Cu spins. Good agreement with experiment is obtained by employing $\xi(T)$ of the form given in Eq.\ (\ref{XIT}). \section{Conclusion} \label{SEC:CON} In summary, we have studied local moments induced in underdoped cuprates by doping with magnetic ($S=1$) Ni and nonmagnetic Zn impurities. In the presence of a spin gap, both types of impurities are associated with $S=\case{1}{2}$ magnetic moments in the CuO$_2$ planes. These are, however, of very different natures. Ni as well as Zn disturb the spin liquid formed by planar Cu spins, resulting in a magnetic moment residing on Cu sites in the proximity of the impurity. In the case of Ni, this moment partially shields the impurity spin below a critical temperature $T_K$ in what resembles an underscreened Kondo model; an effective impurity spin $\case{1}{2}$ results. Since predominantly localized rather than bandlike states are involved in the screening of the impurity spin, the Kondo temperature exhibits an unconventional power-law dependence on the coupling constant. In the case of Zn, on the other hand, one deals with a $S=\case{1}{2}$ moment broadly distributed over Cu sites. We have further investigated the RKKY-type response of Cu spins in a magnetic field. The spin polarization is found to decay as $R^{-3}$ with distance from a Ni impurity, but only as $R^{-2}$ in the case of Zn. This different behavior reflects the delocalized character of Zn moments, and explains why Zn has a stronger impact on the NMR linewidth than Ni. Further, accounting for the presence of temperature-dependent AF correlations in underdoped cuprates, we can successfully describe the non-Curie behavior of the impurity effect on the NMR linewidth. In general, it can be concluded that the anomalous impurity properties of underdoped cuprates are a clear manifestation of the peculiar mixture of spin-singlet and antiferromagnetic correlations present in these compounds.
\section{Introduction} Top Quark Physics will be one of the main physics cases for future collider physics. Whereas the first direct discovery of top was one of the main successes of the proton collider at Fermilab, the precise measurement of the top quark mass and its couplings will remain the task of a future lepton collider. But why should we be interested in such high precision measurements in the top sector? The top quark is the heaviest elementary particle observed up to now. Because of its very high mass $m_t \approx 175$ GeV it plays a prominent role for our understanding of the Standard Model (SM) and the physics beyond. Already before its direct observation there was indirect evidence of the large top quark mass: through radiative corrections $m_t$ enters quadratically into the $\rho$ parameter. From precision measurements of the electroweak parameters $M_Z$, $M_W$, $\sin^2\theta_W$ and $G_F$ a top quark mass was predicted in striking agreement with the value measured at Fermilab. Within the framework of the SM the mass of the Higgs boson can be constrained from the weak boson masses $M_W$ and $M_Z$ together with $m_t$: $M_H = f(M_Z, M_W, m_t)$. As the Higgs mass enters in logarithmic form, stringent mass bounds can be derived only once the other parameters are known with high accuracy. With an absolute uncertainty of the top quark mass $\Delta m_t \stackrel{\scriptstyle <}{\scriptstyle \sim} 200$ MeV the Higgs mass will be extracted with an accuracy better than 17\%. This will constitute one of the strongest tests of the mechanism of electroweak symmetry breaking at the quantum level and therefore of our understanding of the structure of the SM. At the starting time of a future Linear Collider (LC) Higgs boson(s) may hopefully already have been discovered with the hadron machine at Fermilab or at the LHC (assuming LEP2 is not the lucky one in the next future). Still, to pin down parameters precisely and to learn which sort of physics beyond the SM is realized in nature, many detailed studies will be required. With an expected accuracy of $\Delta m_t/m_t \approx 1\cdot 10^{-3}$ ($\Delta m_b/m_b \simeq {\cal O}(\%)$) and the large Yukawa coupling $\lambda_t^2 \approx 0.5$ ($\lambda_b^2 \approx 4\cdot 10^{-4}$) the top quark will play a key role in finding the theory that gives the link between masses and mixings and quarks and leptons. Apart from that the large top quark mass has another important consequence: being much heavier than the $W$ boson the top decays predominantly into the $W$ and a bottom quark with the large (Born) decay rate \begin{equation} \Gamma_t^{(0)} = \frac{G_F}{\sqrt{2}}\frac{m_t^3}{8\pi} \approx 1.5\ {\rm GeV} \gg \Lambda_{\rm QCD}\,. \end{equation} Therefore top is the only quark that lives too short to hadronize. The large width $\Gamma_t$ serves as a welcome cut-off of non-perturbative effects \cite{K} and the top quark behaves like a {\em free} quark. In this way top quark physics is an ideal test-laboratory for QCD at high scales, where predictions within perturbation theory are reliable. Having these goals in mind a future $e^+ e^-$ Linear Collider (see e.g.~\cite{Designrep, PRep}) will be the ideal machine to study the top quark in detail. (The same will be true for a $\mu^+ \mu^-$ collider, once technologically feasible.) The clean environment and generally small backgrounds make it complementary to hadron machines, where higher energies can be achieved more easily. In addition the collision of point-like, colourless leptons guarantees very good control of the systematic uncertainties. Operation with highly polarized electrons (and to a smaller extent also positrons) is realizable and will open new possibilities. Another option is the use of Compton back-scattered photons of intense lasers from the electron and positron bunches, allowing for operation of the $e^+ e^-$ collider in the $\gamma \gamma$ (or $e \gamma$) mode. These modes can be very useful for certain studies of the Higgs sector and other areas of electroweak physics, but will be less important for top quark physics. Therefore the following discussion will be limited to $e^+ e^-$ collisions.\footnote{Reader interested in the physics of $\gamma \gamma$ collisions are referred to \cite{Designrep} (and references therein) for a general discussion and to \cite{gammagamma} especially for $\gamma \gamma \to t \bar t$ at threshold.} The article is organized as follows: In Section 2 the scenario of top quark pair production at threshold is described in some detail. I discuss the important parameters, accessible observables and their sensitivity, and the corresponding theoretical predictions. Recent higher order calculations are reviewed. It will be shown how the large theoretical uncertainties in the shape of the cross section near threshold can be avoided by using a mass definition different from the pole mass scheme. In Section 3 a brief discussion of some important issues in top quark production above threshold is given. Section 4 contains the conclusions. For a comprehensive review of top quark physics (including top at hadron colliders) see also \cite{Kslacrep}. Clearly the rich field of top quark physics cannot be completely covered in this contribution, which is somewhat biased towards $t\bar t$ at threshold. This is also partly due to the author's experience. I would like to apologize to those who miss important information or feel own contributions to top physics not covered properly or not mentioned at all. \section{The $t\bar t$ Threshold} \subsection{What's so special about the top threshold?} Close to the nominal production threshold $\sqrt{s} = 2 m_t$ top and antitop are produced with non-relativistic velocities $v = \sqrt{1 - 4 m_t^2/s} \ll 1$. The exchange of (multiple, ladder-like) Coulombic gluons leads to a strong attractive interaction, proportional to $(\alpha_s/v)^n$. These terms are not suppressed and the usual expansion in $\alpha_s$ breaks down. Summation leads to the well known Coulomb enhancement factor at threshold, giving a smooth transition to the regime of bound state formation below threshold, which cannot be described using ordinary perturbation theory. In principle we would expect a picture like this with ``Toponium'' resonances similar to the case of bottom quarks which form the $\Upsilon(nS)$ mesons at threshold. However, in the case of top quarks, the rapid decay makes a formation of real $t\bar t$ bound states impossible. The width of the $t\bar t$ system is saturated by the decay of its constituents: $\Gamma_{t-\bar t} \approx 2 \Gamma_t \approx 3$ GeV. This is much larger than the expected level spacing and leads to a smearing of any sharp resonance structure, leaving only a remnant of the $1S$ peak visible in the excitation curve. Therefore there will be nothing like $t\bar t$-spectroscopy to study at the top threshold. Nevertheless the short life-time of the top quarks also has a remarkable advantage: non-perturbative effects, hadronization and real (soft) gluon emission are suppressed by $\Gamma_t$, $m_t$.\footnote{For studies concerning the effects of real gluon emission see also Ref.~\cite{Orr}.} Therefore, in contrast to the bottom (let alone the charm) quark sector, top quark production becomes calculable in perturbative QCD \cite{FK}. $t\bar t$ is, from the theoretical point of view, much ``cleaner'' than $c\bar c$ and $b\bar b$ and will allow for more detailed tests of the underlying theory and a more precise determination of the basic parameters $m_t$, $\alpha_s$ (and $\Gamma_t$). In this sense $t\bar t$ at threshold is a unique system, which deserves to be studied in detail at a future $e^+ e^-$ collider. \subsection{Parameters to be determined} $\bullet$ As mentioned already above the main goal will be a precise measurement of the top quark mass. Current analyses from CDF and D0 at the Tevatron at Fermilab determine $m_t$ by reconstructing the mass event by event. Current values are \begin{eqnarray} m_t^{\rm pole} &=& 176.0 \pm 6.5\ {\rm GeV}\qquad ({\rm CDF}\ \cite{CDF})\,, \nonumber\\ m_t^{\rm pole} &=& 172.1 \pm 7.1\ {\rm GeV}\qquad ({\rm D0}\ \cite{D0})\,. \end{eqnarray} The Run II at the Tevatron is expected to improve the accuracy down to maybe $\Delta m_t = 2$ GeV. It looks impossible to reach a higher accuracy at hadron colliders. In contrast, with a threshold scan of the cross section at a future $e^+ e^-$ Linear Collider one will be able to reach $\Delta m_t = 200$ MeV or even better \cite{PRep, MM, FMY}. High luminosity will allow for very small statistical errors so that the accuracy will be limited mainly by systematic errors and theoretical uncertainties. $\bullet$ The strong coupling $\alpha_s$ governs the interaction of $t$ and $\bar t$. It enters the Coulombic potential $V(r) = -C_F\,\alpha_s/r$ which dominates close to threshold, as well as other corrections which get important at higher orders of perturbation theory (see below). $\alpha_s$ may either be taken as an input (with some error) measured independently at other experiments or, alternatively, can be determined simultaneously with $m_t$ in a combined fit. $\bullet$ The (free) top quark width $\Gamma_t$ leads to the smearing of the resonances and strongly influences the shape of the cross section at threshold. As will be discussed below, $\Gamma_t$ can be measured with good precision near threshold either in the $t\bar t$ production process or by help of observables specific to the decay.\footnote{For a detailed discussion of top quark decays see also \cite{Jnpbproc}.} In the framework of the SM $\Gamma_t$ can be predicted reliably: the first order $\alpha_s$ \cite{JK} and electroweak \cite{DSao} corrections are known for some time (see also \cite{JK2}), and recently even corrections of order $\alpha_s^2$ became available \cite{CM}. The ${\cal O}(\alpha_s)$ corrections lower the Born result by about 10\%, whereas ${\cal O}(\alpha_s^2)$ and electroweak contributions effectively cancel each other, with corrections of about $-2\%$ and $+2\%$, respectively. In extensions of the SM the top quark decay rate can be significantly different from the SM value: new channels like the decay in a charged Higgs ($t \to b H^+$) in supersymmetric theories will lead to an increase of $\Gamma_t$. In models with a forth generation the Cabibbo-Kobayashi-Maskawa (CKM) quark-mixing-matrix element $V_{tb}$ will be smaller than the SM value $V_{tb}^{({\rm SM})} \simeq 1$ and lead to a suppression of $\Gamma_t^{({\rm SM})}$. $\bullet$ The electroweak couplings of the top quark enter both in production and decay. Especially in angular distributions (of the decay products) and in observables sensitive to the polarization of the top quarks deviations from the SM may be found. In principle even the influence of the Higgs on the $t\bar t$ production vertex should be visible \cite{HJK}. Unfortunately, for the currently allowed range of Higgs-masses, effects due to (heavy) Higgs exchange mainly result in a ``hard'' vertex correction which changes the overall normalization of the cross section. As will be discussed below, contributions of this sort are in competition with uncertainties from other higher order corrections and therefore difficult to disentangle at the $t\bar t$ threshold.\\ Therefore, to determine the parameters with high precision and to eventually become sensitive to new physics, a thorough understanding of the SM physics, in particular the QCD dynamics, is mandatory. \subsection{Theory's tools to make predictions} \label{theorytools} How to predict the cross section close to threshold? In principle one could write the cross section as a sum over many overlapping resonances \cite{Kwong}: \begin{equation} \sigma(e^+ e^- \to t\bar t\,) \sim -{\rm Im}\, \sum_n \frac{|\psi_n(r=0)|^2}{E-E_n+i \Gamma_t}\,, \end{equation} where $\psi_n$ are the wave functions of the $nS$ states with the corresponding Eigenenergies $E_n$. (Close to threshold $S$ wave production is dominating with the contributions from $P$ waves being suppressed by two powers of the velocity $v$. With $v \approx \alpha_s$ these contributions have to be considered only at next-to-next-to-leading order.) However, this explicit summation is not very convenient, as the sum does not converge fast, especially for positive energies $E = \sqrt{s} - 2 m_t$. As shown by Fadin and Khoze \cite{FK}, the problem can be solved within the formalism of non-relativistic Green functions: \begin{equation} \sigma(e^+ e^- \to t\bar t\,) \sim -{\rm Im}\, G(r=0, E+i\Gamma_t)\,. \end{equation} The Green function $G$ is the solution of the Schr\"odinger equation \begin{equation} \left[ \left( -\frac{\vec\nabla^2}{m_t} + V\left(\vec r\,\right) \right) - \left(E+i\Gamma_t\right) \right] G\left(\vec r, E+i\Gamma_t\right) = \delta^{(3)}\left(\vec r\,\right) \end{equation} or, equivalently, the Lippmann-Schwinger equation in momentum space \begin{equation} \tilde G\left(\vec p, E+i\Gamma_t\right) = \tilde G_0 + \tilde G_0 \int \frac{{\rm d}^3 q}{(2\pi)^3} \tilde V\left(\vec p - \vec q\,\right) \tilde G\left(\vec q, E+i\Gamma_t\right) \,, \end{equation} where $\tilde G_0 \equiv \left(E+i\Gamma_t-p^2/m_t\right)^{-1}$ is the free Green function. At leading and next-to-leading order the continuation of the energy in the complex plane $E+i\Gamma_t$ is all that is needed to take care of the finite decay width of the top quarks. These equations can be solved numerically using a realistic QCD potential $V(r) = -C_F \alpha_s(r)/r$ or $\tilde V\left(q^2\right) = - 4 \pi C_F \alpha_s(q^2)/q^2$ to give the total cross section \cite{StrasslerPeskin, Sumino, JKT} \begin{equation} \sigma\left(e^+ e^- \to \gamma^* \to t\bar t\,\right) = \frac{32\,\pi^2\,\alpha^2}{3\,m_t^2\,s}\, {\rm Im}\, G\left(r=0, E+i\Gamma_t\right)\,. \label{eqsigmatot} \end{equation} The top quark momentum distribution (differential with respect to the modulus of the top quark three momentum $p$), which reflects the Fermi motion in the would-be bound state and the instability of the top quarks, is obtained by \begin{equation} \frac{{\rm d}\sigma(p, E+i\Gamma_t)}{{\rm d}p} = \frac{16\,\alpha^2}{3\,s\,m_t^2}\, \Gamma_t \, p^2 \left| \tilde G\left(p, E+i\Gamma_t\right)\right|^2\,. \label{eqsigmadif} \end{equation} Eqs.~(\ref{eqsigmatot}, \ref{eqsigmadif}) are correct at leading order in $\alpha_s$, $v$. At next-to-leading order (NLO) various new effects have to be taken into account. Apart from the well known ${\cal O}(\alpha_s)$ corrections to the static QCD potential \cite{FandB} the exchange of ``hard'' gluons results in the vertex correction factor $\left( 1 - 16 \alpha_s/(3\pi)\right)$ in the (total and differential) cross section \cite{Barbierietal}. Interference of the production through a virtual photon and a virtual $Z$ boson leads to the interference of the vector current induced $S$ wave with the axial-vector current induced $P$ wave contributions. This $S$-$P$ wave interference is suppressed by order $v$ and drops out in the total cross section after the angular ($\cos\theta$) integration. However, it contributes to the differential rate and will be measured in observables like the forward-backward asymmetry ${\cal A}_{\rm FB}$ \cite{Sumino2, HJKT}. In addition, at order $\alpha_s$, there are final state corrections coming from gluon exchange between the produced $t$ and $\bar t$ and their strong interacting decay products $b$ and $\bar b$. The final state interactions in the $t b$ and $\bar t \bar b$ systems factorize and are easily taken into account by using the (order $\alpha_s$) corrected free top quark width $\Gamma_t$, with no other corrections at ${\cal O}(\alpha_s)$ \cite{JT, MK}. However, the ``crosstalk'' between $t$ $\bar b$, $\bar t$ $b$ and $b$ $\bar b$ leads to non-factorizable corrections which have to be considered in addition.\footnote{In principle hadronically decaying $W$ bosons also take part in these final state interactions.} These corrections are suppressed in the total cross section \cite{FKM, MY}, but contribute in differential distributions and hence in ${\cal A}_{\rm FB}$.\footnote{See also Ref.~\cite{khoze} and references therein for a discussion of the possible impact of colour reconnection effects on the top quark mass determination.} Results obtained in the framework of the non-relativistic Green function approach are available at ${\cal O}(\alpha_s)$, see \cite{HJKP, PS}. At the same accuracy polarization of the produced $t$ and $\bar t$, depending on the polarization of the $e^+$ and $e^-$ beams, has been studied in \cite{HJKT, HJKP, PS}. Therefore, at order $\left(\alpha_s,\,v\right)$, theoretical predictions are available for a variety of observables at the top quark threshold, as will be discussed in the next paragraph. Electroweak corrections to the $t \bar t$ production vertex have been calculated for the threshold region \cite{GK} as well as for general energies \cite{BHMandBH} in the SM and even in the Minimal Supersymmetric SM (MSSM), see \cite{HS}. \subsection{Observables and their sensitivity} \begin{figure}[htb] \vspace{-0.5cm} \begin{center} \leavevmode \epsfxsize=10.0cm \epsffile[105 275 475 555]{top_sig_thr.ps} \end{center} \vspace{-0.5cm} \caption[]{\label{fig1} Total cross section $\sigma(e^+ e^- \to t\bar t\,)$ (in pb) as a function of the total centre of mass energy for two different values of $m_t$ and $\alpha_s$. The upper curves correspond to $\alpha_s(M_Z) = 0.121$, the lower ones to $\alpha_s(M_Z) = 0.115$. (Figure taken from \cite{PRep}.)} \end{figure} $\bullet$ The (from the theoretical as well as from the experimatal point of view) cleanest observable is the total cross section $\sigma_{\rm tot}$. Depending on the decays of the $W^+$ and $W^-$ from the $t$ and the $\bar t$ quarks, $t\bar t$ decays into six jets (46\%), four jets + $l + \nu_l$ (44\%) or two jets + $l\,l'\,\nu\,\nu'$ (10\%) (60, 35 and 5\%, respectively, if $l = e,\,\mu$ only and $\tau$-leptons are excluded). The main backgrounds are from $e^+ e^- \to W^+ W^-$, $Z^0 Z^0$ and $f \bar f\,$(plus gluons and photons). These processes are well under control as distinguishable from the signal (e.g.\ by higher Thrust or less jets) and constitute no big problem for the experimental analysis. The total cross section is mainly sensitive to $m_t$ and $\alpha_s$. Fig.~\ref{fig1} shows the cross section for two different values of the top quark mass and two values of the strong coupling, plotted over the total centre of mass energy. Note the correlation between $m_t$ and $\alpha_s$: higher top-masses lead to a shift of the remainder of the $1S$ peak to larger energies. In a similar way an increase of $\alpha_s$ is equivalent to a stronger potential (a larger negative binding energy) and hence lowers the peak position. I will come back to this point later. In practice, the shape of the cross section will not look as pronounced as in Fig.~\ref{fig1}. \begin{figure}[htb] \begin{center} \epsfig{file=top_exp_thr.eps,width=7cm,angle=89.5} \end{center} \vspace{-.2cm} \caption[]{\label{fig2} Total cross section in the threshold region including initial-state and beam-strahlung. The errors of the data points correspond to an integrated luminosity of $\int{\cal L} = 50\ {\rm fb}^{-1}$ in total. The dotted curves indicate shifts of the top mass by 200 and 400 MeV. (Figure taken from \cite{PRep}.)} \end{figure} Initial state radiation (of photons from the $e^+$ and $e^-$ beams) as well as the beamstrahlung-effects from the interaction of the $e^+$ and $e^-$ bunches lead to a distortion of the original shape. Fig.~\ref{fig2} displays how the total cross section is expected to look under realistic conditions. The dots in the plot are Monte-Carlo generated ``data points'' of a typical planned threshold scan. In addition $\sigma_{\rm tot}$ also depends on the Higgs mass $M_H$ and the top quark width $\Gamma_t$. As mentioned already above, the Higgs mainly influences the normalization of the cross section which will probably not allow for a high sensitivity to $M_H$ once other uncertainties are taken into account. $\Gamma_t$, on the other hand, influences the shape: the smaller the width the more pronounced the peak. This will be used together with the sensitivity of other observables to measure $\Gamma_t$.\\ $\bullet$ Another observable is the momentum distribution ${\rm d}\sigma/{\rm d}p$, obtained from the reconstruction of the three momentum of the top (and antitop) quark. With the possible high statistics at a future Linear Collider the distribution can be well measured. As shown in Fig.~\ref{fig3}, the peak position strongly depends on $m_t$ but less on the QCD coupling: for higher values of $m_t$ the distribution is peaked at much lower momenta, whereas the coupling strength mainly changes the normalization. \begin{figure}[htb] \begin{center} \leavevmode \epsfxsize=10.0cm \epsffile[105 275 475 555]{top_dsdpt_thr.ps} \end{center} \vspace{-0.5cm} \caption[]{\label{fig3} The differential cross section ${\rm d}\sigma/{\rm d}p$ as a function of the top quark momentum $p$ for a fixed value of the centre of mass energy (349 GeV). $m_t$ and $\alpha_s$ are chosen as indicated. (Figure taken from \cite{PRep}.)} \end{figure} Therefore a measurement of the momentum distribution can help to disentangle the strong correlation of $m_t$ and $\alpha_s$ in the total cross section (see \cite{MM}). There is also a less pronounced dependence on $\Gamma_t$.\\ $\bullet$ As mentioned above, $S$-$P$ wave interference leads to a nontrivial $\cos\theta$ ($\theta$ being the angle between the $e-$ beam and the $t$ direction) dependence of the cross section. The resulting forward-backward asymmetry \begin{equation} {\cal A}_{\rm FB} = \frac{1}{\sigma_{\rm tot}} \left[ \int_0^1 {\rm d}\cos\theta - \int_{-1}^0 {\rm d}\cos\theta \right] \frac{{\rm d}\sigma}{{\rm d}\cos\theta} \label{defafb} \end{equation} shows a considerable dependence on $\Gamma_t$ and $\alpha_s$, but is not very sensitive to $m_t$. \begin{figure}[htb] \begin{center} \leavevmode \epsfxsize=9.0cm \epsffile[90 150 470 655]{afb.ps} \end{center} \vspace{-0.5cm} \caption[]{\label{fig4} Forward-backward asymmetry ${\cal A}_{\rm FB}$ as a function of $E = \sqrt{s} - 2 m_t$ for three different values of the top quark width and the strong coupling. Upper plot: variation of $\Gamma_t$ by $\pm 20\%$ around the SM value $\Gamma_t^{\rm SM} = 1.43$ GeV and $\alpha_s(M_Z) = 0.118$. Lower plot: $\alpha_s(M_Z) = 0.115, 0.118, 0.121$ and $\Gamma_t = 1.43$ GeV. ($m_t = 175$ GeV.)} \end{figure} \begin{figure}[htb] \vspace{-0.5cm} \begin{center} \leavevmode \epsfxsize=8.0cm \epsffile{mor_3.eps} \end{center} \vspace{-0.5cm} \caption[]{\label{fig5} Increase of the $\chi^2$ of the fit as a function of the top quark width using the forward-backward asymmetry ${\cal A}_{\rm FB}$ (solid line), adding the top quark momentum distribution (dashed line) and the total cross section (dotted line). (Figure taken from \cite{MM}.)} \end{figure} In Fig.~\ref{fig4} ${\cal A}_{\rm FB}$ is plotted as a function of $\sqrt{s}$ for three different choices of $\Gamma_t$ and $\alpha_s$. With increasing width the overlap of $S$ and $P$ waves becomes bigger and hence the asymmetry is enhanced. Together with the total and differential cross section the measurement of ${\cal A}_{\rm FB}$ can be used to determine $\Gamma_t$ by a fit. The sensitivity of such a fit to the different observables is demonstrated in Fig.~\ref{fig5}.\\ Please note that the figures for the cross section and the asymmetry do not contain the (nonfactorizable) ${\cal O}(\alpha_s)$ rescattering corrections discussed in Section \ref{theorytools}. They are absent in the total cross section but slightly change ${\rm d}\sigma/{\rm d}p$ and ${\cal A}_{\rm FB}$, see \cite{HJKP, PS} for a detailed discussion.\\ $\bullet$ Top Quark Polarization: Near threshold $S$ wave production dominates ($\vec L = 0$) and the total spin consists of the spins of the top and antitop quarks, $\vec J_{\gamma^*,\,Z^*} = \vec S_t + \vec S_{\bar t}$. In leading order the top spin is aligned with the $e^+ e^-$ beam direction. Even without polarization of the initial $e^+$ and $e^-$ beams, the top quarks are produced with $-40\%$ (longitudinal) polarization. For a realistic (longitudinal) $e^-$ polarization of $P_{e^-} = +80\%$ ($-80\%$) and an unpolarized $e^+$ beam ($P_{e^+} = 0$) the top polarization amounts to $+60\%$ ($-90\%$). This picture is changed only slightly due to $S$-$P$ wave interference effects of ${\cal O}(v)$ and rescattering effects of ${\cal O}(\alpha_s)$, which lead to top polarizations perpendicular to the beam direction (transverse) and normal to the production plane. Normal polarization could also be induced by time reversal odd components of the $\gamma t \bar t$- or $Z t \bar t$-couplings, e.g.\ by an electric dipole moment, signalling physics beyond the SM. The influence of the bound state dynamics near threshold was calculated in the Green function formalism, including the polarization of the initial beams, the $S$-$P$ wave interference contributions and the ${\cal O}(\alpha_s)$ rescattering effects \cite{HJKT, HJKP, PS}. Neglecting contributions due to rescattering, the three polarizations can be written as \begin{eqnarray} \left| \vec S_{\|} \right| &=& C^0_{\|} + C^1_{\|}\,\varphi_{\rm R}(p, E) \,\cos\theta\,,\nonumber\\ \left| \vec S_{\bot} \right| &=& C_{\bot}\,\varphi_{\rm R}(p, E)\, \sin\theta\,,\nonumber\\ \left| \vec S_{\rm N} \right| &=& C_{\rm N}\,\varphi_{\rm I}(p, E)\, \sin\theta\,. \end{eqnarray} The functions $\varphi_{\rm R, I}$ contain all information about the threshold dynamics, whereas the coefficients $C$ are dependent on the electroweak couplings and the $e^+ e^-$ polarization (see e.g.\ Ref.~\cite{HJKP} for complete formulae). \begin{figure}[htb] \begin{center} \leavevmode \epsfxsize=12.0cm \epsffile[30 285 540 530]{pol_coefs.ps} \end{center} \vspace{-0.5cm} \caption[]{\label{fig6} Coefficients $C$ as functions of the polarization $\chi$ as described in the text, for $\sqrt{s} = 180$ GeV and $\sin^2\theta_W = 0.2317$.} \end{figure} Fig.~\ref{fig6} shows the coefficients $C^0_{\|}$, $C^0_{\bot}$, $C_{\bot}$ and $C_{\rm N}$ as functions of the effective polarization $\chi = (P_{e^+} - P_{e^-})/(1 - P_{e^+}P_{e^-})$. From Fig.~\ref{fig6} it becomes clear that by choosing the appropriate longitudinal polarization of the $e^-$ beam one can tune the normal polarization of the top quarks $\vec S_{\rm N}$ to dominate. The functions $\varphi_{\rm R, I}(p, E)$ are displayed in Fig.~\ref{fig7} for four different energies $E$ around the threshold. \begin{figure}[htb] \begin{center} \leavevmode \epsfxsize=11.0cm \epsffile[30 170 540 630]{fig7.ps} \end{center} \vspace{-0.5cm} \caption[]{\label{fig7} Functions $\varphi_{\rm R}(p, E)$ (solid curves) and $\varphi_{\rm I}(p, E)$ (dashed) for four different energies close to threshold ($m_t = 180$ GeV, $\alpha_s = 0.125$). The dotted lines show the free particle result $\varphi_{\rm R} = p/m_t$. (Figure taken from \cite{HJKP}.)} \end{figure} Also shown is the result for free quarks, $\varphi_{\rm R} = p/m_t$. The normal polarization depends basically on the parameters $\Gamma_t$, $\alpha_s$ and is relatively stable against rescattering corrections. The $\alpha_s$ dependence can be understood from the case of stable quarks and a pure Coulomb potential, where the analytical solution exists \cite{FKK}: $\varphi_{\rm I} \to \frac{2}{3}\,\alpha_s$. In contrast, the subleading (angular dependent) part of the longitudinal polarization and the transverse polarization both are (strongly) changed by rescattering corrections, but vanish after angular integration. For a detailed discussion of the rescattering corrections and the construction of inclusive and exclusive observables which are sensitive to the top quark polarization, see \cite{HJKP, PS, P}. Let me just note here that the rescattering corrections destroy the factorization of the production and decay of the polarized top quarks. Nevertheless, observables can be constructed which depend neither on the subtleties of the $t\bar t$ production process nor on rescattering corrections, but only on the decay of free polarized quarks, even in the presence of anomalous top-decay vertices (see \cite{PS, SandJS}).\\ $\bullet$ Axial contributions to the angular integrated cross section: $P$ wave contributions arise not only at ${\cal O}(v)$ due to $S$-$P$ wave interference but also as $P^2$-terms at next-to-next-to-leading order (NNLO). These contributions are suppressed by $v^2$ close to threshold. Still, they contribute at the percent level and have to be taken into account at the NNLO-accuracy discussed below. In addition these axial current induced corrections are an independent observable and strongly depend on the polarization of the $e^+ e^-$ beams. Numerical results for the total and differential cross section were obtained recently within the formalism of non-relativistic Green functions \cite{KT}. Fig.~\ref{fig8} shows the total cross section as a function of the energy with and without these contributions and their size relative to the pure $S$ wave result for three different values of the $e^-$ polarization. \begin{figure}[htb] \begin{center} \vspace{-0.5cm} \leavevmode \epsfxsize=9.5cm \epsffile[100 120 460 700]{tot.ps} \end{center} \vspace{-0.5cm} \caption[]{\label{fig8} a) The total cross section $\sigma(e^+ e^- \to t\bar t\,)$ as a function of $E$ for three different choices of the $e^-$ polarization: the continuous, dashed and dash-dotted lines correspond to $P_- = -1$, $0$ and $1$, respectively, where only $S$ wave production is taken into account. The dotted lines show the corresponding total cross sections including the $P$ wave contributions. b) Ratio of the $P$ to the $S$ wave contribution $\sigma_{\rm tot}^{\rm AA}/\sigma_{\rm tot}^{\rm VV}$ for the three different $e^-$ polarizations. (Figure taken from \cite{KT}.)} \end{figure} A cut-off $p_{\rm max} = m_t/2$ has been applied to cure the divergence of the integrated $P$ wave Green function coming from the large momentum region, where the non-relativistic approximation breaks down. \subsection{Large next-to-next-to-leading order corrections} In view of the size of the NLO corrections one may ask how accurate the theoretical predictions are. To answer this question within perturbation theory convincingly one has to go to the next order, in our case to the NNLO. The first step in this direction was done by M.~Peter who calculated the ${\cal O}(\alpha_s^2)$ corrections to the static potential \cite{PeterSchroeder}. They turned out to be sizeable and, furthermore, indicate limitations of the accuracy achievable due to the asymptoticness of the perturbative series. As was studied in \cite{JKPST}, the series for the effective coupling in the Coulomb potential behaves differently in the position and in the momentum space. Although potentials formally may differ only in N$^3$LO, the resulting theoretical uncertainty of the total cross section in the $1S$ peak region is estimated to be of the order 6\% \cite{JKPST}. Recently results of the complete NNLO relativistic corrections\footnote{Here NNLO means corrections of the order ${\cal O}(\alpha_s^2,\,\alpha_s v,\,v^2)$ relative to the Born result which contains the resummation of the leading $(\alpha_s/v)^n$ terms.} to $t \bar t$ production near threshold became available \cite{HT2, MYel, Yakovlev, BSS}. The results are in fair agreement and modify the NLO prediction considerably. In the following I will briefly describe the calculation and results.\\ \noindent {\bf Calculation and results.} The problem can be formulated most transparently in the framework of effective field theories. There one makes use of the strong hierarchy of the physical scales top mass, momentum, kinetic energy and $\Lambda_{\rm QCD}$ with $m_t \gg m_t v \gg m_t v^2 \gg \Lambda_{\rm QCD}$ by integrating out ``hard'' gluons with momenta large compared to the scales relevant for the nonrelativistic $t \bar t$ dynamics. This leads to non-relativistic QCD (NRQCD) \cite{CLandBBL}. With $m_t v \gg \Lambda_{\rm QCD}$ one can go one step further and integrate out gluonic (and light quark) momenta of order $m_t v$. Doing so one arrives at the so-called potential NRQCD \cite{PinedaSoto}, and the dynamics of the $t \bar t$ system can be described by the NNLO Schr\"odinger equation \begin{eqnarray} \left[ -\frac{\vec\nabla^2}{m_t} - \frac{\vec\nabla^4}{4m_t^3} + V_{\rm C}(\vec r\,) + V_{\rm BF}(\vec r\,) + V_{\rm NA}(\vec r\,) -\left(E+i\Gamma_t\right)\,\right]\,G(\vec r,E + i \Gamma_t) \, & = & \, \nonumber\\ = \delta^{(3)}(\vec r\,)\,. \qquad & & \label{schroedingerfull} \end{eqnarray} Note the appearance of the operator $-\vec\nabla^4/(4m_t^3)$ which is a correction to the kinetic energy. The instantaneous potentials are the two-loop corrected Coulomb potential $V_{\rm C}$ \cite{PeterSchroeder}, the Breit-Fermi potential $V_{\rm BF}$ known from positronium, and $V_{\rm NA}$ is an additional purely non-Abelian potential. The cross section is again related to the imaginary part of the Green function at $\vec r = 0$. In contrast to the NLO calculation the additional potentials lead to ultraviolet divergencies in Eq.~(\ref{schroedingerfull}) which have to be regularized. This can be done by introducing a factorization scale $\mu_{\rm fac}$ which serves as a cut-off in the effective field theory. The complete renormalization also requires the matching of the effective field theory to full QCD. This involves the determination of (energy independent) short distance coefficients. They contain all information from the ``hard'' momenta integrated out before and also depend on the cut-off $\mu_{\rm fac}$, so that in the final result the biggest part of the factorization scale dependence cancels. In order to perform this matching the knowledge of the corresponding NNLO results of the $t\bar t$ cross section in full QCD above threshold is essential \cite{CM2}. Let me skip further details and immediately discuss the results of the NNLO calculation\footnote{A more detailed discussion and complete formulae can be found in \cite{HT2} (see also \cite{Andre}).}: Fig.~\ref{fig9}a shows the total cross section $e^+ e^- \to \gamma^* \to t \bar t$ in units of $\sigma_{\rm point} = 4\pi\alpha^2/(3s)$ in LO, NLO and NNLO (dotted, dashed and solid lines, respectively), where in each case the three curves correspond to three values of the scale $\mu_{\rm soft}$ governing the strong coupling in the potential(s). In Fig.~\ref{fig9}b the dependence of the NNLO prediction on the input parameter $\alpha_s(M_Z)$ is demonstrated. \begin{figure}[htb] \begin{center} \vspace{0.5cm} \leavevmode \epsfxsize=3.5cm \epsffile[220 420 420 550]{fig9a.ps}\\ \vspace{3.5cm} \leavevmode \epsfxsize=3.5cm \epsffile[220 420 420 550]{fig9b.ps} \vskip 3.5cm \caption[]{\label{fig9} (a) The total normalized photon-mediated $t\bar t$ cross section at LO (dotted lines), NLO (dashed lines) and NNLO (solid lines) for the scales $\mu_{\rm soft}=50$ (upper lines), $75$ and $100$~GeV (lower lines). (b) The NNLO cross section for $\alpha_s(M_Z)=0.115$ (solid line), $0.118$ (dashed line) and $0.121$ (dotted line). ($m_t = 175$ GeV, $\Gamma_t = 1.43$ GeV. Figures taken from \cite{HT2}.)} \end{center} \end{figure} These results are somewhat surprising: whereas large corrections are not unusual for NLO calculations, the large corrections arising at NNLO were unexpected. It is well visible from Fig.~\ref{fig9}a that from leading to NLO the $1S$ peak is shifted to lower energies by about 1 GeV and again moves by about 300 MeV if one includes the NNLO corrections. Moreover, the large negative correction in the normalization from leading to NLO is partly compensated by the big positive correction at NNLO. In addition the scale uncertainty, which is often used as an estimate of the uncertainty of a (fixed order) perturbative calculation from higher orders, seems to be artificially small at NLO but fairly big again at NNLO. This will make studies which mainly depend on the normalization of the $t\bar t$ cross section (like the extraction of the Higgs mass) very difficult. \begin{figure}[htb] \begin{center} \leavevmode \epsfxsize=8.cm \epsffile[72 235 540 555]{fig10.ps} \caption[]{\label{fig10} Total cross section at NNLO as a function of the energy relative to threshold with parameters as in Fig.~\ref{fig9}a. The solid lines give the complete result of \cite{HT2} whereas the dashed lines contain only the NNLO corrections to the static Coulomb potential $V_{\rm C}$ \cite{PeterSchroeder}.} \end{center} \end{figure} In Fig.~\ref{fig10} the importance of the NNLO relativistic corrections to the kinetic energy and through the additional potentials $V_{\rm BF}$ and $V_{\rm NA}$ in Eq.~(\ref{schroedingerfull}) are demonstrated: the dashed lines show the result where only the NNLO corrections to the static Coulomb potential $V_{\rm C}$ \cite{PeterSchroeder} are applied, the solid lines show the complete NNLO result from \cite{HT2}. We have argued above that the total cross section with its steep rise in the threshold region (the remainder of the $1S$ peak as shown in Fig.~\ref{fig2}) is the ``cleanest'' observable to determine $m_t$. From Fig.~\ref{fig9} it now becomes clear that the problem of the strong correlation between $m_t$ and $\alpha_s$, which was already discussed above, also appears through the different orders of perturbation theory: a fit of experimental data from a threshold scan to theoretical predictions (like indicated in Fig.~\ref{fig2}) at a given order will result in a determination of $m_t$ depending on the order. This is in principle nothing wrong and is easily understood, as in higher orders the corrections to the potential lead to a stronger effective coupling. Nevertheless now the question arises:\\ \noindent {\bf Are there large theoretical uncertainties in the determination of $m_t$?} First I would like to point out that the $1S$ peak shift from NLO to NNLO is actually not too dramatic. Taking this shift as an estimate of unknown effects in even higher orders would indicate a theoretical uncertainty $\Delta m_t \stackrel{\scriptstyle <}{\scriptstyle \sim} \Lambda_{\rm QCD}$, which still leads to a relative accuracy of $\Delta m_t /m_t \sim {\cal O}(10^{-3})$ for the top mass. Still, having argued that due to the large width $\Gamma_t > \Lambda_{\rm QCD}$ non-perturbative effects should be suppressed, an even smaller theoretical uncertainty should be achievable. Concerning the large NNLO corrections to the normalization and the large scale uncertainty I would like to comment that there is reason to believe that the NNLO result is a much better approximation than the NLO one and that corrections in even higher orders should not spoil this picture \cite{HT3}. But how can the stability of the prediction be improved? The key point here is to remember that in all formulae and results discussed up to now $m_t$ is defined as the pole mass. This scheme seems, at first glance, to be the most intuitive one and to be suited for the non-relativistic regime. Nevertheless we know that $m^{\rm pole}$ is {\em not} an observable. It is defined only up to uncertainties of ${\cal O}(\Lambda_{\rm QCD})$, and the large top quark width $\Gamma_t$ does not protect the pole mass $m_t^{\rm pole}$ \cite{SW}. By performing a renormalon analysis it was recently shown in \cite{B, HSSW} that the leading long-distance behaviour which affects the pole mass in higher orders also appears in the static potential. However, in the sum $E_{\rm static} = 2 m^{\rm pole} + E_{\rm binding}$ these contributions cancel and $E_{\rm static}$ is free from renormalon ambiguities. The separate quantities, mass and potential, suffer from a scheme ambiguity which is not present in the sum. Therefore one should make use of a ``short distance'' mass definition different from the pole mass scheme, which avoids these large distance ambiguities.\\ \noindent {\bf Short distance mass definitions: Curing the problem.} In principle there exist infinitely many mass definitions which subtract the renormalon ambiguities. In practice, however, this is not enough. On the one hand, any new short distance mass $m^{\rm SD}$ has to be related with high accuracy to a mass in a more general scheme like the (modified) Minimal Subtraction scheme ($\overline{\rm MS}$).\footnote{This is possible because of the short distance characteristics of $m^{\rm SD}$ and $m^{\overline{\rm MS}}$ which makes the perturbative relation between the masses well behaved. The $\overline{\rm MS}$ mass itself cannot be used directly for the calculation of the $t\bar t$ threshold, see \cite{B}.} Otherwise the extraction of $m^{\rm SD}$ would be more or less useless. On the other hand, the subtraction of renormalon contributions, which become important at high orders of perturbation theory, will not be enough to compensate the large shifts of the $1S$ peak observed at NLO and NNLO. Recently different mass definitions were proposed which can fulfill all the requirements: in Ref.~\cite{B} Beneke defined the ``Potential Subtracted'' mass by \begin{equation} m^{\rm PS}(\mu_f) = m^{\rm pole} - \delta m(\mu_f) \end{equation} where the subtraction is given by \begin{equation} \delta m(\mu_f) = -\frac{1}{2}\,\int_{|\vec q\,|<\mu_f} \frac{{\rm d}^3 q}{(2\pi)^3}\,\tilde V(q)\,. \end{equation} The subtracted potential in position space then reads \begin{equation} V(r,\,\mu_f) = V(r) + 2 \delta m(\mu_f)\,. \end{equation} This is equivalent to suppressing contributions from momenta $q$ below the scale $\mu_f$ in the potential. For $\mu_f \to 0$ one recovers the pole mass $m^{\rm PS} \to m^{\rm pole}$. By choosing $\mu_f$ larger, say 20 GeV, one can achieve a compensation of the $1S$ peak shifts. Another mass definition is the $1S$ mass, originally introduced in $B$ meson physics \cite{HLM}, which defines the $1S$ mass as half of the perturbatively defined $1S$ energy. This $m^{1S}$ mass can be related reliably to the $\overline{\rm MS}$ mass. There are also other mass definition in the literature, see e.g.\ the ``low scale running mass'' \cite{Uraltsevetal}, which is similar to the concept of the PS mass but differs in the actual $\mu_f$-dependent subtraction. Studies about the application of different mass definitions are underway and I can only present preliminary results here: Fig.~\ref{fig11} shows our best prediction \cite{HT3} for the NNLO $t\bar t$ cross section together with the NLO and LO results for two different values of the renormalization scale $\mu_{\rm soft}$ governing the strong coupling $\alpha_s$. \begin{figure}[htb] \begin{center} \leavevmode \epsfxsize=9.cm \epsffile[90 95 470 705]{fig11.ps} \caption[]{\label{fig11} Total cross section $e^+ e^- \to \gamma^* \to t \bar t$ in units of $\sigma_{\rm point} = 4\pi\alpha^2/(3s)$ as a function of $\sqrt{s}$. Dotted, dashed and solid lines correspond to the LO, NLO and NNLO results. The upper curves are obtained with the renormalization scale $\mu = 20$ GeV, the lower ones with $\mu = 60$ GeV. a) $1S$ mass scheme, and b) PS mass scheme with $\mu_f = 20$ GeV. ($m_t = 175$ GeV, $\Gamma_t = 1.43$ GeV and $\alpha_s(M_Z) = 0.118$.)} \end{center} \end{figure} In the upper plot the $1S$ mass scheme is used, whereas for the lower plot the PS mass scheme is adopted. It is clear from these curves that both mass definitions work well. The shift of the $1S$ peak is nearly completely compensated. Differences in the normalization remain, but they will not spoil the mass determination from the shape of the total cross section near threshold. Of course more detailed studies are needed to find the best strategy for a precise determination of $m_t^{\overline{\rm MS}}$, which is needed in electroweak calculations. \section{Studies above Threshold} In the continuum top quarks are produced through the same annihilation process as near threshold: $e^+ e^- \to \gamma^*,\, Z^* \to t \bar t\,$. Other (gauge boson fusion) channels like $\ e^+ e^- \to \nu_e \bar\nu_e t\bar t\ $ or $\ e^+ e^- \to e^+\bar\nu_e t\bar b\ $ are negligible, except for $e^+ e^- \to e^+ e^- t\bar t\,$, where the contribution from $\gamma\gamma$ fusion becomes important at TeV energies. Formulae for the (polarized) production cross section and subsequent decay are well known (see e.g.~\cite{report123A} and references therein). Similar to the top quark analyses at Fermilab $t\bar t$ events will be reconstructed at an event by event basis and allow for a determination of the top quark mass and its couplings. Due to the clean environment and the large statistics (at $\sqrt{s} = 500$ GeV and with an integrated luminosity of $\int{\cal L} = 50\ {\rm fb}^{-1}$ there will be $\stackrel{\scriptstyle >}{\scriptstyle \sim} 30000\ t\bar t$ pairs!) high precision will be reached at a future Linear Collider. In the following I will briefly outline a few important cases of top physics above threshold.\\ \noindent $\bullet$ {\bf Kinematical reconstruction of $m_t$ above threshold.} The top can be reconstructed from 6 jet and 4jet+$l$+$\nu$ events. For centre of mass energies far above threshold the top and antitop signals will be in different hemispheres and $t$ and $\bar t$ may be reconstructed separately. Constraints from energy and momentum conservation in the fitting procedure can improve the mass resolution considerably. Experimental studies \cite{report123A} (see also \cite{report123E}) have demonstrated that a high statistical accuracy of the order of $\Delta m_t ({\rm stat.}) \sim 150$ MeV can be achieved at a future Linear Collider. But in contrast to the analysis at threshold many experimental uncertainties and not very well known hadronization effects will limit the total expected accuracy to $\Delta m_t \sim 0.5$ GeV.\\ \noindent $\bullet$ {\bf Top formfactors.} Top quarks are produced with a high longitudinal polarization. Due to the large top width $\Gamma_t$ hadronization is suppressed and the initial helicity is transmitted to the final state without depolarization. Therefore, in contrast to the case of light quarks, $t$ helicities can be determined from the (energy-angular) distributions of jets and leptons in the decay $t \to b W^+ \to b f \bar f'\,$, similar to the case of $Z$ polarization analyses at LEP and SLC. This will allow to measure the formfactors of the top quark in detail \cite{report123D}. The relevant current can be written as \begin{equation} j_{\mu}^a \propto \gamma_{\mu} \left( F_{1, L}^a P_L + F_{1, R}^a P_R \right) + \frac{i \sigma_{\mu\nu} q^{\nu}}{2m_t} \left( F_{2, L}^a P_L + F_{2, R}^a P_R \right)\,, \end{equation} with the form factors $F^a$ ($a = \gamma,\,Z,\,W$). At lowest order in the SM, $F_{1, L}^{\gamma} = F_{1, R}^{\gamma} = F_{1, L}^W = 1\,$, $F_{2, L}^{\gamma} = F_{2, R}^{\gamma} = F_{1, R}^W = 0\,$ and $F_{1, L}^Z = g_L$, $F_{1, R}^Z = g_R$. A non-zero value for $(F_{2, L}^{\gamma,Z} + F_{2, R}^{\gamma,Z})$ is caused by a magnetic ($\gamma$) or weak ($Z$) dipole moment, whereas a non-zero value for the CP-violating combination $(F_{2, L}^{\gamma,Z} - F_{2, R}^{\gamma,Z})$ by an electric (weak) dipole moment. These moments would influence distributions for the top production process, e.g.\ by inducing an extra contribution proportional to $\sin^2\theta$ in the differential cross section: \begin{equation} \frac{{\rm d}\sigma}{{\rm d}\cos\theta} \propto \left[ \frac{m_t}{E} \left( F_{1,L}+F_{1,R} \right) + \frac{E}{m_t} 2 \left( F_{2,L} +F_{2,R} \right) \right]^2 \sin^2\theta\,. \end{equation} The extra ($F_{2,L} +F_{2,R}$) term leads to an additional spin-flip contribution and therefore changes the total and differential cross section. At a future Linear Collider such an anomalous magnetic moment of the top quark $(g-2)_t$ could be seen up to a limit of $\Delta\delta \stackrel{\scriptstyle <}{\scriptstyle \sim}4\%$ ($\delta \equiv F_{2,L}^{\gamma} + F_{2,R}^{\gamma}$) \cite{report123D}, for $\int{\cal L} = 50$ fb$^{-1}$ at $\sqrt{s} = 500$ GeV. With especially defined observables an anomalous electric and weak dipole moment due to CP violating formfactors $\delta_t^{\gamma, Z} \propto (F_{2, L}^{\gamma,Z} - F_{2, R}^{\gamma,Z})$ could be observed up to a limit of $\Delta d_t^{\gamma, Z} \stackrel{\scriptstyle <}{\scriptstyle \sim} 5 \cdot 10^{-18} {\rm e cm}$ (for $\int{\cal L} = 10$ fb$^{-1}$ at $\sqrt{s} = 500$ GeV). A measurement of $F_{1,R}^W \neq 0$ would signal non-SM physics like a ($V$+$A$) admixture to the top charged current, a $W_R$ boson or the existence of a charged Higgs boson. $F_{1,R}^W$ can be studied by help of the energy and angular distributions of the top quark decay leptons \cite{JK3}. It could be constrained up to $\Delta \kappa^2 \stackrel{\scriptstyle <}{\scriptstyle \sim} 0.02$ ($\kappa^2 \sim |F_{1,R}^W|^2$) with a luminosity of $\int{\cal L} = 50$ fb$^{-1}$ in the threshold regime, which is best suited for such a measurement.\\ \noindent $\bullet$ {\bf Rare top decays.} In the SM top quark decays different from $t \to b W^+$ are strongly suppressed. On one hand, the unitarity of the CKM matrix constrains $V_{tb} \simeq 0.999$, giving not enough room for top decays to the $s$ or $d$ quark at an observable rate. On the other hand, due to the GIM mechanism \cite{GIM}, flavour-changing one-loop transitions like $t \to c g$, $t \to c \gamma$, $t \to c Z$ or $t \to c H$ are also extremely small \cite{EHS, MPS}. However, in extensions of the SM like the MSSM extra top quark decay channels like $t \to b H^+$, $t \to \tilde t \tilde\chi^0,\ \tilde b \tilde \chi_1^+$ may be open. In general branching fractions of up to 30\% are possible. The experimental signatures are clear and will be easily detectable \cite{VB, PRep}. With an integrated luminosity of $\int{\cal L} = 50$ fb$^{-1}$ it will be possible to observe $t \to b H^+$ up to $m_{H^+} \stackrel{\scriptstyle <}{\scriptstyle \sim} m_t - 15$ GeV, and $t \to \tilde t \tilde \chi^0$ down to a branching fraction of $\sim 1\%$ at the 3$\sigma$ level.\\ \noindent $\bullet$ {\bf Direct observation of the top Yukawa coupling.} Although the Higgs boson will hopefully be discovered before the future Linear Collider starts operation, the detailed study of the Higgs and its couplings will remain one of the main tasks of the LC. There one will be able to test if the Higgs Yukawa coupling to the top quark deviates from the SM value $\lambda_t^2 = \sqrt{2}\,G_F\,m_t^2 \sim 0.5$. Studies at threshold will be difficult (see above), but due to this large coupling (in comparison to $\lambda_b^2 \sim 4\cdot 10^{-4}$) the $t\bar t H^0$ vertex will be accessible through Higgs-strahlung at high energies. For $M_H \leq 2m_t$ one will measure $\lambda_t^2$ through the process $e^+ e^- \to t\bar t H$ with the Higgs subsequently decaying into a pair of $b$ quarks. For $M_H \geq 2m_t$ two different processes will be dominant: Higgs radiation from $Z$ (in $e^+ e^- \to Z H$) with subsequent decay of the Higgs into $t\bar t$, and the fusion of $W^+ W^-$ (in $e^+ e^- \to \nu \bar\nu H$) into the Higgs which then decays into $t\bar t$. With eight jets in the final state of the fully hadronic decay channels, which satisfy many constraints, these processes will have clear signatures. Still, even despite the large Yukawa coupling, the cross sections are quite small, amounting only to a few fb. Here the planned high luminosity of the latest TESLA design will be most welcome. Extensive studies were performed and come to the conclusion that at high energy and with high luminosity $\lambda_t^2$ may finally be measurable with an accuracy of 5\% at a future LC \cite{ttbarhiggs}.\\ \section{Conclusions} I have reviewed the subject of top quark physics at a future $e^+ e^-$ Linear Collider, emphasizing top quark physics at threshold. Threshold studies will determine the SM parameters $m_t$, $\alpha_s$ and $\Gamma_t$ with very high accuracy: $\Delta m_t / m_t \stackrel{\scriptstyle <}{\scriptstyle \sim} 10^{-3}\,$, $\,\Delta\alpha_s \stackrel{\scriptstyle <}{\scriptstyle \sim} 0.003\,$ and $\,\Delta\Gamma_t/\Gamma_t \stackrel{\scriptstyle <}{\scriptstyle \sim} 0.05\,$ seem to be possible from experimental point of view. Recent theoretical progress shows, that in order to achieve such a high accuracy also in the theoretical predictions, mass schemes different from the pole mass should be employed to disentangle correlations between $m_t$ and $\alpha_s$ as well as infrared ambiguities in the definition of $m_t^{\rm pole}$. In addition to the total cross section and the momentum distribution of top quarks also observables like the forward-backward asymmetry, polarization and axial contributions are calculated. These observables will be accessible by help of large statistics due to the high luminosity and by the possibility to have polarized $e^+ e^-$ beams. Above threshold formfactors of the top quark and the top Yukawa coupling will be measured. One may study rare top decays and get sensitive to non-SM physics. The future Linear Collider will therefore be {\em the} machine to study top quark physics in detail, to understand the SM better and eventually to learn more about what comes beyond it. I hope to have shown that top quark physics is an interesting field both for Theory and Experiment. Further work will be needed to understand the heaviest known particle better, before data become available.\\ \noindent {\bf Acknowledgements}\\[1mm] It is my great pleasure to thank the organizers for having made the {\em Cracow Epiphany Conference '99}\ such an enjoyable and stimulating event. I would also like to express my gratitude to all friends and colleagues I have worked with on various topics about toppik and top peaks reported here for their fruitful collaboration. While writing this contribution I was hit by the shock of the tragic death of Bj{\o}rn H.\ Wiik. His outstanding efforts for the future Linear Collider and his fascinating personality will be missed.
\section{Introduction} Quarks appear to be confined in nature. This means that free quarks have not been detected so far but only hadrons, their bound states. Therefore, predictions of the hadron spectrum are an explicit and direct test of our understanding of the confinement mechanism as a result of the low energy dynamics of QCD. In these lectures we will focus on heavy quark bound states and for simplicity we will only treat the quark-antiquark case i.e. the mesons. Indeed, also in order to extract the Cabibbo--Kobayashi--Maskawa matrix elements and to study CP violation from the experimental decay rate of heavy mesons, hadronic matrix elements are needed. It is reasonable to expect that to this aim the somewhat simpler matching of the theoretical prediction for the spectrum to experiment has to be previously established. Moreover, we need to achieve some understanding of the bound state dynamics in QCD in order to make reliable identifications of the gluonic degrees of freedom in the spectrum (hybrids, glueballs). All this is relevant to the programme of most of the accelerators machines, Godfrey et al. (1998). From a more general point of view, the issue about the consequences of a nontrivial vacuum structure and the nonperturbative definition of a field theory are questions that overlap with the domain of supersymmetric and string theories. These lectures are quite pedagogical and introductory and contain several illustrative exercises. The interested reader is referred for details to the quoted references. The plan of the lectures is the following one. In Secs. 2 and 3 we give a brief overview on the hadron spectrum and on the phenomenological models devised to explain it. The nonperturbative phenomenological parameter $\sigma$ is introduced and connected to the Regge trajectories as well as to the string models. In Sec. 4 we evaluate perturbatively and nonperturbatively (via lattice simulations) the static Wilson loop. We discuss the area law behaviour and the flux tube formation as signals of confinement. In Sec. 5 we summarize some existing model independent results on the heavy quark interaction. QCD effective field theories are introduced. In Sec. 6 we connect confinement to the structure of the QCD vacuum. We study with some detail the Minimal Area Law model. We discuss the Abelian Higgs model and the Dual Meissner effect and introduce the idea of 't Hooft Abelian projection. We list some of the results obtained on the lattice with partial gauge fixing. Finally we briefly review two models of the QCD vacuum: Dual QCD and the Stochastic Vacuum Model. Each section is supplemented with some exercises. We tried to be as self-contained as possible reporting all the relevant definitions and the basic concepts. \section{The Hadron Spectrum} The meson and the baryon resonances together with an introduction to the quark model have been discussed at this school by Jim Napolitano. Since these lectures are mainly concerned with the quark confinement mechanism, they contain only a general overview on the spectrum pointing out its relevant features. \begin{figure}[thb] \makebox[3.0truecm]{\phantom b} \epsfxsize=5truecm \epsffile{verylight.eps} \caption{\it The spectrum of the lightest mesons labeled by $({\rm spin})^{\rm parity}$.} \label{pluno} \end{figure} Let us concentrate on the meson spectrum as given in Figs. \ref{pluno}-\ref{pltre}. In principle, one should be able to explain and predict it only by means of the QCD Lagrangian \begin{equation} L= -{1\over 4} F^{(a)}_{\mu\nu} F^{(a)\mu \nu} + i\sum_{q=1}^{N_f} \bar{\psi}_q^i\gamma^\mu (D_\mu)_{ij}\psi_q^j - \sum_{q=1}^{N_f} m_q \bar{\psi}_q^i\psi_q^i \label{lagr} \end{equation} where \begin{eqnarray} F_{\mu \nu} & & \equiv \partial_\mu A_\nu-\partial_\nu A_\mu +ig [A_\mu, A_\nu] \nonumber \\ (D_\mu)_{ij} & & \equiv \delta_{ij}\partial_\mu + ig {\lambda^a_{ij}\over 2}A^a_\mu \, ; \quad \quad \qquad A_\mu\equiv A_\mu^a {\lambda_a\over 2} \quad a=1, \dots ,8. \label{def} \end{eqnarray} $\psi^j_q$ are the quark fields of flavour $q=1,\dots, 6=N_f$ and colour $j =1,\dots , 3$, $m_q$ the (current) quark masses and $\lambda^a $ the Gell-Mann matrices of $SU(3)$. However, if we proceed to calculate the spectrum from this Lagrangian using the familiar tool of a perturbative expansion in the coupling constant $g$, we get no match with the experimental data presented in Figs. \ref{pluno}-\ref{pltre}. This is a consequence of the most relevant feature of the Lagrangian (\ref{lagr}): asymptotic freedom (Gross and Wilczek (1973), Politzer (1973)). The coupling constant ($\alpha_{\rm s} \equiv g^2/4 \pi$) vanishes in the infinitely high energy region and grows uncontrolled in the infrared energy region: \begin{equation} {d \, \alpha_{\rm s}(\mu) \over d \log \mu^2}\equiv \beta(\alpha_{\rm s}) = -\alpha_{\rm s} \left( \beta_0 {\alpha_{\rm s}\over 4\pi} + \beta_1 \left({\alpha_{\rm s}\over 4\pi}\right)^2 + \cdots \right) \label{betaqcd} \end{equation} where $\beta_0 = 11 - 2/3 N_f > 0$. As a consequence, a perturbative treatment is expected to be reliable in QCD only when the energy $\mu$ is large compared to $\Lambda_{QCD}$, which is the infrared energy scale defined by Eq. (\ref{betaqcd})\footnote{ $\Lambda_{\rm \overline{MS}}^{N_f=4}= 305\pm 25\pm 50$ MeV; this value corresponds to $\alpha_s(M_Z)= 0.117 \pm 0.002\pm 0.004 $, Particle Data Group (PDG) (1998).}. The point is that the relevant energies in a quark bound state are in most cases of the order of the naturally occurring scale of 1 fm ($\simeq \Lambda_{QCD}^{-1}$), the average hadron size. The only exception are the mesons made up with top quarks. Unfortunately, they have not enough time to exist!\footnote{ The top quark decays into a real $W$ and a $b$ with a large width. The toponium life-time would be even smaller than the revolution time, thereby precluding the formation of a mesonic bound state, Quigg (1997), Bigi et al. (1986).} \begin{figure}[thb] \makebox[2.0truecm]{\phantom b} \epsfxsize=8truecm\epsffile{heavy.eps} \caption{\it The experimental heavy meson spectrum ($b\bar{b}$ and $c\bar{c}$) relative to the spin-average of the $\chi_b(1P)$ and $\chi_c(1P) $ states.} \label{pldue} \end{figure} The masses appearing in the Lagrangian are the so-called ``current mas\-ses'' and fall into two categories: light quark masses $m_u=1.5 \div 5$ MeV, $m_d= 3 \div 9$ MeV, $m_s= 60 \div 170$ MeV ($m_u$, $m_d\ll \Lambda_{QCD}$ and $m_s \sim \Lambda_{QCD}$) and heavy quark masses $m_c=1.1 \div 1.4$ GeV, $m_b= 4.1 \div 4.4$ GeV, $m_t=173.8\pm 5.2$ GeV ($m_c, m_b, m_t \gg \Lambda_{QCD}$)\footnote{The masses are scale dependent objects. For what concerns the experimental values quoted above (PDG (1998)), the $u$, $d$, $s$ quark masses are estimates of the so-called current quark masses in a mass independent subtraction scheme such as $\overline{\rm MS}$ at a scale $\mu \simeq 2$ GeV. The $c$ and $b$ quark masses are estimated from charmonium, bottomonium, $D$ and $B$ masses. They are the running masses in the $\overline{\rm MS}$ scheme. These can be different from the constituent masses obtained in potential models, see below. We remark that it exists a definition of the mass, the pole quark mass (appropriate only for very heavy quarks) which does not depend of the renormalization scale $\mu$. The quark masses are calculated in lattice QCD, QCD sum rules, chiral perturbation theory. For further explanations see Dosch and Narison (1998), Jamin et al. (1998), Kenway (1998), Leutwyler (1996).}. The spectrum should be obtained from the QCD Lagrangian with these values of the masses. However, in the light quark sector, spontaneous chiral symmetry breaking and non-linear strongly coupled effects cooperate in a highly non trivial way. Indeed, it is peculiar of QCD that, due to confinement, the quark masses are not physical, i.e. directly measurable quantities. Therefore, it turns out to be useful, in order to make phenomenological predictions, to introduce the so-called constituent quark masses, containing the current masses as well as mass corrections also due to confinement effects. These constituent masses can be defined, for example, using the additivity of the quark magnetic moments inside a hadron or using phenomenological potential models to fit the spectrum. For quarks heavier than $\Lambda_{QCD}$ the difference between current and constituent masses is not quite relevant. In the framework of the constituent quark model the meson states are classified as follows. For equal masses the quark and the antiquark spins combine to give the total spin ${\bf S}={\bf S}_1 + {\bf S}_2$ which combines with the orbital angular momentum ${ \bf L}$ to give the total angular momentum ${ \bf J}$. The resulting state is denoted by $n ^{2 S+1}L_J$ where $n-1$ is the number of radial nodes. As usual, to $L=0$ is given the name $S$, to $L=1$ the name $P$, to $L=2$ the name D and so on. The resonances are classified via the $J^{PC}$ quantum numbers, $P= (-1)^{L+1} $ being the parity number and $C= (-1)^{L+S}$ the C-parity. \begin{figure}[thb] \vskip -2truecm \makebox[1.0truecm]{\phantom b} \centerline{ \epsfxsize=8truecm \epsffile{regge.ps}} \vskip -3truecm \caption{\it The Regge trajectories for the $\rho$, $K^*$ and $\phi$. From Godfrey et al. (1985).} \label{pltre} \end{figure} Light mesons (as well as baryons) of a given internal symmetry quantum number but with different spins obey a simple spin ($J$)-mass ($M$) relation. They lie on a Regge trajectory \begin{equation} J(M^2)=\alpha_0 +\alpha^\prime M^2 \label{reg} \end{equation} with $\alpha^\prime \simeq 0.8 -0.9 \, {\rm GeV}^{-2}$, see Fig. \ref{pltre}. Up to now, free quarks have not been detected. The upper limit on the cosmic abundance of relic quarks, $n_q$, is $n_q/n_p< 10^{-27}$, $n_p$ being the abundance of nucleons, while cosmological models predict $n_q/n_p< 10^{-12}$ for unconfined quarks. The fact that no free quarks have been ever detected hints to the property of quark confinement. Hence, the interaction among quarks has to be so strong at large distances that a $q\bar{q}$ pair is always created when the quarks are widely separated. From the data it is reasonable to expect that a quark typically comes accompanied by an antiquark in a hadron of mass $ 1 \, {\rm GeV}$ at a separation of 1 {\rm fm} ($\simeq \Lambda_{QCD}^{-1}$). This suggests that between the quark and the antiquark there is a linear energy density (called string tension) of order \begin{equation} \sigma = {\Delta E\over \Delta r} \simeq 1 {{\rm GeV}\over {\rm fm}}\simeq 0.2 \, {\rm GeV}^2 . \label{sigma1} \end{equation} The evidence for linear Regge trajectories (see Fig. \ref{pltre}) supports this picture. A theoretical framework is provided by the string model, Nambu (1974). In this model the hadron is represented as a rotating string with the two quarks at the ends. The string is formed by the chromoelectric field responsible for the flux tube configuration and for the quark confinement (see Fig. \ref{pltube}), Buchm\"uller (1982). Upon solution of the Exercise 2.1, the reader can verify that it is possible to establish the relation \begin{equation} \alpha^\prime = {1\over 2 \pi \sigma } \label{rel} \end{equation} between the slope of the Regge trajectories and the string tension. The string tension $\sigma$ emerges as a key phenomenological parameter of the confinement physics. \begin{figure}[t] \vskip -1.5truecm \makebox[3.0truecm]{\phantom b} \epsfxsize=6truecm \epsffile{tube.eps} \caption{\it Picture of the quark-antiquark bound state in the string model.} \label{pltube} \end{figure} From the light mesons spectrum of Fig. \ref{pluno} it is evident that \begin{itemize} \item{} the separations between levels are considerably larger than the mesons masses $\Longrightarrow$ they are truly relativistic bound states; \item{} the splitting between the pseudoscalar $\pi$ and the vector $\rho$ mesons is so large to be anomalous $\Longrightarrow$ it is due to the Goldstone boson nature of the $\pi$. \end{itemize} Therefore, understanding the light meson spectrum means to solve a relativistic many-body bound state problem where confinement is strongly related with the spontaneous breaking of chiral symmetry. Since the target of these lectures is to gain some understanding of the confinement mechanism in relation to the spectrum, we will try to separate problems and to consider first the spectrum of mesons built by heavy valence quarks only. In this case we have still bound states of confined quarks, however, due to the large mass of the quarks involved, we can hope to treat relativistic and many-body (i.e. quark pair creation effects) contributions as corrections. The problem simplifies remarkably if we consider mesons made up by two valence heavy quarks ($b\bar{b}$, $c\bar{c}$, $b\bar{c}$, ...) i.e. quarkonium. \vskip 1truecm \leftline{\bf 2. Exercises} \begin{itemize} \item[2.1]{Consider two massless and spinless quarks connected by a string of length $R$ rotating with the endpoints at the speed of light, so that each point at distance $r$ from the centre has the local velocity $v/c = 2 r/R$. Recalling that the string tension $\sigma$ is the linear energy density of the string between the quarks, calculate the total mass and the total angular momentum and demonstrate that they lie on a Regge trajectory with $\alpha^\prime = 1 / (2 \pi \sigma)$.} \item[2.2]{Consider the spin-independent Lagrangian \begin{eqnarray} L & & = - m_1 \sqrt{1-{\bf v}^2_1}- m_2 \sqrt{1-{\bf v}^2_2}-U_0(r)\nonumber \\ & & -{U_+(r)\over 4 } ({\bf v}_1 -{\bf v}_2)^2 -{U_-(r)\over 4 }({\bf v}_1 +{\bf v}_2)^2 \nonumber \end{eqnarray} where ${\bf x}_1, {\bf v}_1$ and $ {\bf x}_2, {\bf v}_2 $ are the positions and velocities respectively, of the quark and the antiquark and ${\bf r}={\bf x}_1 -{\bf x}_2$. $U_0$ is the static potential and $U_+ $ and $U_-$ are the coefficients of the velocity dependent terms in the potential. Velocity dependent terms of this type are obtained in Baker et al. (1995), Brambilla et al. (1994). Making some simplifications (circular orbits, ${v}_j\equiv {\bf \omega} \times {\bf r}_j$ with ${\bf r}_j=(-1)^{j+1} r_j \hat{{\bf r}}$, $r_1+r_2=r$, $m_1=m_2$), calculate the energy $E$ and the angular momentum ${\bf J}$. Determine the moment of inertia of the colour field produced by the rotating quarks and establish the physical meaning of $U_+$. Then, obtain the slope of the Regge trajectories in the case in which $U_0=\sigma r$ and $U_+ = -A r$.} \end{itemize} \section{Quarkonium and Confining Phenomenological Potentials} The relevant features of the quarkonium spectrum are: the pattern of the levels, the spin separation between pseudoscalar mesons $n^1S_0 (0^{-+})$ and vector mesons $n^3S_1 (1^{--})$ (called hyperfine splitting), the spin separations between states within the same $L\neq 0$ and $S$ multiplets (e.g. the splitting in the $ 1^3P_J$ multiplet $\chi_c(1P)$ in charmonium cf. Fig. \ref{pldue}) (called fine splitting), and the transition and decay rates, see PDG. To separate the sub-structure from the radial and orbital splittings it is convenient to work in terms of spin-averaged splittings. Spin-averaged states are obtained summing over masses of given $L$ and $n$, and weighting by $2 J+1$. The hyperfine spin splittings appear to scale roughly with a $1/m_Q$ dependence. We note that states below threshold are considerably narrow since they can decay only by annihilation. The fact that all the splittings are considerably smaller than the masses implies that all the dynamical scales of the bound state, such as the kinetic energy or the momentum of the heavy quarks, are considerably smaller than the quark masses. Therefore, the quark velocities are nonrelativistic: $v\ll 1$. The energy scales in quarkonium are the typical scales of a nonrelativistic bound state: the momentum scale $m_Q v$ and the energy scale $m_Q v^2$. Being the time scale $T_g \sim 1/m_Q v$ associated with the binding gluons smaller than the time scale $T_Q \sim 1/m_Q v^2$ associated with the quark motion, the gluon interaction between heavy quarks appears ``instantaneous''. Therefore, it can be modelled with a potential and the energy can be obtained solving the corresponding Schr\"odinger equation. In the extreme nonrelativistic limit of very heavy quarks the spin splittings vanish and the spin-averaged spectrum is described by a single static central potential. \begin{figure}[htb] \vskip -0.2truecm \makebox[1.0truecm]{\phantom b} \centerline{ \epsfxsize=8truecm \epsffile{levelqq.eps}} \vskip -0.2truecm \caption{\it $\rho = (E_n- E_1)/(E_2-E_1)$, where $E_n$ is either the $n$th energy level of the physical system or the $n$th eigenvalue of the Schr\"odinger equation corresponding to the indicated potential.} \label{pllevel} \end{figure} A lot of work has been done to find the phenomenological form of the static potential (see the report of Gromes, Lucha and Sch\"oberl (1991)). Lowest order perturbation theory for QCD gives a flavour-independent central potential based on the one-gluon exchange, which has a Coulomb-like form (see Exercise 3.1) \begin{equation} V_0(r)= - {4\over 3} {\alpha_{\rm s} \over r} \label{coulomb} \end{equation} where $r$ is the distance between the two quarks and $\alpha_{\rm s}$ is the strong coupling constant \footnote{To compute Eq. (\ref{coulomb}), cf. Exercise 3.1, it is necessary to evaluate products of the type $\lambda^{(1)}\cdot \lambda^{(2)}$ in the representation of interest. It turns out that, out of all two-body channels, the colour singlet ($q\bar{q}$) is the most attractive. This hints to the fact that coloured mesons should not exist.} \footnote{Of course, $\alpha_{\rm s}$ in Eq. (\ref{coulomb}) depends on a scale. If we work in a physical gauge, e.g. in a lightlike gauge, the quarks in the bound state interact through gluon exchange and these interactions are renormalized by loop corrections. In the extreme nonrelativistic limit, only the corrections to the gluon propagator survive and this give $\alpha_{\rm s}$ evaluated at the square momentum of the exchanged gluon $Q^2\simeq {\bf Q}^2$.}. This cannot be the final answer since it does not confine quarks and gives a spectrum incompatible with the data (see Fig. \ref{pllevel}). Nevertheless we can regard the one-gluon exchange formula to be valid for $r \simeq 0.1$ fm. It was found that the addition of some positive power of $r$ to Eq. (\ref{coulomb}) rescues the phenomenology. The intuitive argument of a constant energy density (string tension), exposed in Sec. 2, led to the flavor-independent Cornell potential (Eichten et al. (1978)) \begin{equation} V_0(r) = -{4\over 3} {\alpha_{\rm s}\over r } +\sigma r + {\rm const.} \label{cornell} \end{equation} Here, $\alpha_{\rm s}$ and $\sigma$ are regarded as free parameters to be fitted on the spectrum. The Schr\"odinger equation with the potential (\ref{cornell}) and parameters $\alpha_{\rm s} = 0.39$ and $\sigma=0.182$ GeV$^2$ gives quite a satisfactory agreement with the data. In Eichten et al. (1980), the coupling to charmed meson decay-channels was also taken into account. It was found that the mass shifts due to the coupled channel effects are indeed large also below threshold and yet they do not spoil the predictions of the naive potential model. Indeed these effects can essentially be absorbed into a redefinition of the effective parameters. \begin{figure}[htb] \vskip -2truecm \makebox[1.0truecm]{\phantom b} \centerline{ \epsfxsize=8truecm \epsffile{physpot.ps}} \vskip -2truecm \caption{\it The rms $q\bar{q}$ separations in some representative mesons is shown with respect to the Cornell potential. All the phenomenological potentials agree in the range of $0.1-1$ fm which is the physical range for quarkonia. From Godfrey and Isgur (1985).} \label{plpot} \end{figure} Since then several different phenomenological forms of the static potential have been exploited, e.g. the Richardson potential, $${\displaystyle V_0(r)= \int {d^3 {\bf Q}\over (2 \pi)^3} \exp{\{i {\bf Q}\cdot {\bf r}\}} {{\rm const.}\over Q^2\log(1+ Q^2/\Lambda^2)}}$$ (Richardson (1979)), the logarithmic potential, $V_0(r)= A\log (r/ r_0)$ (Quigg and Rosner (1977)), the Martin potential, $V_0(r)= A(r/ r_0)^\alpha$ (Martin (1980)). By fitting the parameters, all these potentials can reproduce the spectrum. This is not surprising if we look at Fig. \ref{plpot}: the potentials essentially agree in the region $r \sim 0.1\div0.8$ fm in which the $\sqrt{\langle r^2\rangle}$ for quarkonia sits (Buchm\"uller and Tye (1981)). On the other hand, it is a considerable limit of the potential approach that the connection with the true QCD parameters of Eq. (\ref{lagr}) remains totally hidden and mysterious. However, the existence of a spin substructure in the spectrum indicates that relativistic corrections to the static central potential $V_0$ have to be taken into account. From the radial level splitting (see Exercise 3.2) as well as from the fits with the phenomenological potential, we find for \begin{eqnarray} c \quad {\rm in } \quad \psi, & & v^2 \simeq 0.3 \nonumber \\ b \quad {\rm in }\quad \Upsilon, & & v^2 \simeq 0.1 \quad , \label{modvel} \end{eqnarray} and therefore we expect relativistic corrections of order $20\div 30 \%$ for the charmonium spectrum and up to $10 \%$ for the bottomonium spectrum. This determination of the heavy quark velocity in the bound states is confirmed by lattice calculation, see Tab.1. Notice that relativistic corrections are of critical importance for the observables sensitive to the details of the wave functions (e.g. the radiative transitions ) (Mc Clary et al. (1983)). \begin{center} \begin{table} \begin{center} \begin{tabular}{|c|c|c|c|c|} \hline $nL$&$\langle { v_b^2}\rangle$&$\langle{ v_c^2}\rangle$& $\sqrt{\langle { r_b^2}\rangle}/$fm&$\sqrt{\langle { r_c^2}\rangle}/$fm\\\hline $1S$&0.080&0.27&0.24&0.43\\ $2S$&0.081&0.35&0.51&0.85\\ $3S$&0.096&0.44&0.73&1.18\\ $1P$&0.068&0.29&0.41&0.67\\ $2P$&0.085&0.39&0.65&1.04\\ $1D$&0.075&0.34&0.54&0.87\\ \hline \end{tabular} \end{center} \caption{\it From Bali et al. (1997). Notice that for a Coulombic system $ v \sim \alpha/ n$ and for a confined system $ v$ grows with $ n$.} \label{pltab} \end{table} \end{center} The phenomenological potential model predictions of the relativistic corrections are calculated by means of a Breit--Fermi Hamiltonian of the type \begin{equation} H= \sum_{j=1,2} \left(m_j + {p_j^2 \over 2 m_j} -{p_j^4 \over 8 m_j^3}\right) + V_0 + V_{\rm SD} + V_{\rm VD}. \label{ham} \end{equation} The $1/m^2$ spin-dependent $V_{\rm SD} $ and velocity-dependent $V_{\rm VD} $ potentials are derived from the semirelativistic reduction of the Bethe--Salpeter (BS) equation for the quark-antiquark connected amputated Green function or, equivalently at this level, from the semirelativistic reduction of the quark-antiquark scattering amplitude with an effective exchange equal to the BS kernel. Several ambiguities are involved in this procedure, due on one hand to the fact that we do not know the relevant confining Bethe--Salpeter kernel, on the other hand due to the fact that we have to get rid of the temporal (or energy $Q_0$, $Q=p_1-p_1^\prime$ being the momentum transfer) dependence of the kernel to recover a potential (instantaneous) description. It turns out that, at the level of the approximation involved, the spin-independent relativistic corrections at the order $1/ m^2$ depend on the way in which $Q_0$ is fixed together with the gauge choice of the kernel. The Lorentz structure of the kernel is also not known. On a phenomenological basis, the following ansatz for the kernel was intensively studied \begin{eqnarray} I(Q^2)= (2 \pi)^3 \left[ \gamma_1^\mu \gamma_2^\nu P_{\mu\nu} J_v(Q) + J_s(Q) \right] \label{kernel} \end{eqnarray} in the instantaneous approximation $Q_0=0$. Notice that the effective kernel above was taken with a pure dependence on the momentum transfer $Q$. But, of course, the dependence on the quark and antiquark momenta could have been more complicated. The vector kernel $\gamma_1^\mu \gamma_2^\nu P_{\mu\nu} J_v(Q)$ corresponds to the one gluon exchange, and e.g. in the Coulomb gauge $P_{\mu\nu}$ has the structure $P_{\mu\nu}= g_{\mu\nu}+\displaystyle{J_v^\prime(-{\bf Q}^2)\over J_v(-{\bf Q}^2)} Q_\mu Q_\nu - \displaystyle{J_v^\prime(-{\bf Q}^2)\over J_v(-{\bf Q}^2)} Q_0(Q_\mu n_\nu+Q_\nu n_\mu)$ where the prime indicates the derivative and $n_\mu$ is the unit vector in the time direction. The scalar kernel $J_s(Q)$ accounts for the nonperturbative interaction. The semirelativistic reduction of the kernel (\ref{kernel}) (in the instantaneous approximation, with the Coulomb gauge fixed for the vectorial part, in the centre of mass frame ${\bf p}_1= -{\bf p}_2 ={\bf q}, {\bf p}^\prime_1= -{\bf p}^\prime_2 ={\bf q}^\prime, {\bf Q} \equiv {\bf q} -{\bf q}^\prime$ and in the equal mass case), gives \begin{eqnarray} V_0 &=& \tilde{J}_s(r) + \tilde{J}_v(r) \label{vfenum}\\ V_{\rm SD} &=& {3\over m^2} {1\over r} \left({d\tilde{J}_v\over dr} -{1\over 3} {d\tilde{J}_s\over dr}\right) {\bf S} \cdot {\bf L} \label{vspin} \\ &+&{1\over m^2} \left({1\over r} {d\tilde{J}_v \over dr} - {d^2 \tilde{J}_v\over dr^2}\right) S_1^h \left({r^hr^k\over r^2}-{\delta^{hk}\over 3}\right) S^k_2 + {2\over 3 m^2} \Delta \tilde{J}_v(r) {\bf S}_1 \cdot {\bf S}_2 \nonumber \\ V_{\rm VD} &=& {1\over 4 m^2 } \Delta(\tilde{J}_s +\tilde{J}_v) -{1\over 4 m^2} q^h \tilde{J}_s q^h +{1\over 2 m^2} q^h \left(\delta^{hk} -{\partial_h \partial_k \over \partial^2} \tilde{J}_v\right) q^k \label{vvel} \end{eqnarray} with $\tilde{J}_{v,s}({\bf r}) \equiv \displaystyle\int {d^3 {\bf Q}} \, e^{i{\bf Q} \cdot {\bf r} } J({\bf Q})_{v,s}$. Taking $J_v \! = \!\displaystyle-{1\over 2 \pi^2} {4\over 3} {1\over {\bf Q}^2}$ and $J_s \! = \! \displaystyle -{\sigma\over \pi^2} {1\over {\bf Q}^4}$, $V_0$ reproduces the Cornell potential. The confining part of the kernel is usually chosen to be a Lorentz scalar in order to match the data on the fine separation. Indeed, the ratio $\rho_{FS}$ of the fine structure splitting \begin{equation} \rho_{FS} = {M( ^3P_2)- M( ^3P_1)\over M( ^3P_1)-M( ^3P_0)} \end{equation} is $\rho_{FS}= 0.8 $ for a pure vector Coulomb exchange while the data give $\rho_{FS} \simeq 0.49, 0.66, 0.57$ for the $c\bar{c}(1P), b \bar{b}(1P)$ and $ b\bar{b}(2P)$ respectively. Adding a scalar exchange gives a contribution to the spin orbit interaction that reduces the vector part and thus also the value of the ratio $\rho_{FS}$ (Schnitzer (1978)). Moreover, since the two terms, vectorial short range and scalar long range exchange, contribute with opposite signs, we expect that at high orbital excitations, where the $Q\bar{Q}$ pair probes large average distances (see Fig. \ref{plpot}), the triplet multiplet inverts with reference to the ordering of the low excitation multiplets (i.e. $M( ^3L_{L+1})\ge M( ^3L_{L}) \ge M( ^3L_{L-1})$ at low $L$ and the reverse at high $L$). The fine structure turns out to be a nice test for the form of the confining interaction (cf. e.g. Isgur (1998)). We have presented a way to obtain phenomenologically the ${1/ m^2}$ relativistic corrections. However, this is not really rewarding. In a confining interaction the average $\langle p^2\rangle $ increases with the excited states (see Tab. 1) and so the pattern of the excited levels is likely to be considerably distorted if the nonperturbative relativistic corrections are not the appropriate ones. Indeed, fits made with only the static potential turn out to be better than fits made with the interaction (\ref{vfenum})-(\ref{vvel}) (Brambilla et al. (1990)). Moreover, there are characteristics of the spectrum, like the hyperfine separation as well as the leptonic decays, that are due to processes taking place at very short scale. In this case it is important to add higher order perturbative corrections to the one gluon exchange as well as the running of $\alpha_{\rm s}$. In the phenomenological potential framework it is somehow ambiguous how to take into account the running of $\alpha_{\rm s}(\mu)$ as well as the scale $\mu$. In this section we realized that the description of the heavy meson spectrum turns out to be a priori a quite complicate problem with an interplay of different relevant scales as well as of perturbative and nonperturbative effects. The conclusion is that we badly need, on one hand a framework in which relativistic as well as perturbative corrections to the quark-antiquark interaction can be evaluated unambiguously and systematically and, on the other hand a clear, well founded and eventually computable approach to the long range quark-antiquark interaction which is essentially nonperturbative. We need this both at a concrete level, in order to make quantitative predictions in which the size of the neglected terms can be estimated, and both at a fundamental level, in order to use the spectrum to get some insight into the confinement mechanism. In the next section we address the problem of how to study quark confinement beyond phenomenological models i.e. in a QCD based framework and we give a criterium that decides whether a gauge theory is confined or not. \vskip 1truecm \leftline{\bf 3. Exercises} \begin{itemize} \item[3.1]{Consider the quark-antiquark scattering \begin{equation} q_i(p_1,\sigma_1) + \bar{q}_j(p_2,\sigma_2) \to q_k(q_1,\tau_1) +\bar{q}_l(q_2,\tau_2) \label{scat} \end{equation} where $i,j,\cdots =1,2,3$ label the colour indices. Remembering that the quarks in the meson are in a colour-singlet state and introducing the meson colour wavefunctions $\delta_{ij}/\sqrt{3}$, calculate the $T$-matrix element, extract the first contribution in the nonrelativistic limit and obtain, via Fourier transform, the perturbative one gluon exchange potential of Eq. (\ref{coulomb}). Show that the contribution of the annihilation graph vanishes.} \item[3.2]{Consider the average excitation energy in charmonium and bottomonium (e.g. $M_\Upsilon^\prime -M_\Upsilon $). This should be of the order of the average kinetic energy $E\simeq mv^2$. Taking $m$ to be roughly half of the ground state mass obtain the estimates (\ref{modvel}) for the quark velocities.} \item[3.3]{Consider the same scattering of (\ref{scat}) with the exchange given in Eq. (\ref{kernel}). Obtain the result (\ref{vfenum})-(\ref{vvel}) by computing the scattering matrix element of this process, expanding up to the ${1/ m^2}$ order and Fourier transforming.} \item[3.4]{Consider the same scattering matrix of Exs. 3.1 and 3.3 but with a kernel of the type $\displaystyle I= \gamma^1_5 \gamma^2_5 V_p(Q) $. Show that there is no static potential in the nonrelativistic limit of the matrix element. Discuss the result in relation to the deuteron.} \end{itemize} \section{The Wilson Loop: Confinement and Flux Tube Formation} The most powerful technique in order to extract nonperturbative information from QCD is the lattice gauge theory approach. This has been undoubtedly successful and rewarding and has produced over the last years an impressive amount of results. Yet, in spite of almost two decades of intensive efforts the characteristics of QCD associated with colour confinement are still not understood. It is our belief that some insight in the mechanism of confinement cannot be obtained without developing, in strict connection with lattice QCD, also analytic methods. In this way information coming from the lattice can be inserted inside analytic models or vice versa lattice calculations can be used in order to interpret analytic models. To this aim we need an unambiguous way of establishing a gauge invariant and systematic procedure to calculate the quark dynamics. In Sec. 5, we will show that this is feasible in the case of heavy quarks in which the whole dynamics can be reduced to few expectation values of chromoelectric and chromomagnetic fields that can be calculated analytically (once a model for the QCD vacuum is assumed) or numerically on the lattice. Then, the comparison between the two results supply us with hints about the mechanism of confinement. The simplest manifestation of confinement in quenched QCD is the linear rising of the potential $V_0(r)$ between static colour sources in the fundamental representation. In this section we will show how this has been clearly proved and connected to the formation of a chromoelectric flux tube between the quarks. The question of the nature and the origin of these nonperturbative field configurations will be addressed in Sec. 6. \subsection{The QCD static potential and the Wilson loop} Let us consider a locally gauge invariant quark-antiquark singlet state\footnote{The contribution of the string to the potential vanishes in the limit $T\to \infty$, Eichten et al. (1981) and Brambilla et al. (1999).} \begin{equation} \vert \phi_{\alpha \beta}^{lj} \rangle \equiv {\delta_{lj}\over \sqrt{3}} \bar{\psi}^i_\alpha(x) U^{ik}(x,y,C) \psi^k_\beta(y) \vert 0\rangle \label{statgaug} \end{equation} where $i,j,k,l$ are colour indices (that will be suppressed in the following), $\vert 0\rangle$ denotes the ground state and the Schwinger string line has the form \begin{equation} U(x,y;C) =P \exp \left\{ i g \int_y^x A_\mu(z) \, dz^\mu \right\} \, , \label{string} \end{equation} where $A_\mu$ is the gauge potential of Eq. (\ref{def}), $g$ the QCD coupling constant, and the integral is extended along the path $C$. The operator $P$ denotes the path-ordering prescription\footnote{ Path ordering prescription means operatively that one has to decompose the path $C$ connecting $y$ with $x$ into infinitesimal pieces, then take the exponential along the infinitesimal pieces, expand at the first order and order the factors according to their appearance along the path.} which is necessary due to the fact that $A_\mu$ are non-commuting matrices. Remembering that under a $SU(3)$ gauge transformation ${\cal V}(\theta)$ $=$ $\exp({-i \theta})$ $\simeq 1-i\theta$, the gauge potential undergoes the transformation $A_\mu$ $\to$ $A_\mu$ $+$ $i [\theta, A_\mu]$ $+$ $g^{-1} \partial_\mu \theta$, we ob\-tain the tran\-sfor\-ma\-tion law of the string \begin{equation} U^\prime(x,y;C) =\exp{ \{i\theta(x)}\} U(x,y;C) \exp{\{-i \theta(y)\} } \label{stringtrasf} \end{equation} and then it is clear that (\ref{statgaug}) is a gauge-invariant state. Actually, it is a colour singlet and we are interested only in colour singlet being the only existing initial and final states. The quark-antiquark potential can be extracted from the quark-antiquark Green function. A simple example clarifies in which way. Let us consider the following two-particle Green function \begin{equation} G(T) = \langle \phi({\bf x},0) \vert \phi({\bf y}, T)\rangle = \langle \phi({\bf x},0) \vert \exp{(-iH T)}\vert \phi({\bf y}, 0)\rangle . \label{green} \end{equation} Inserting a complete set of energy eigenstates $\psi_n$ with eigenvalues $E_n$ and making a Wick rotation we find \begin{eqnarray} G(-i T) & = & \sum_n \langle \phi({\bf x},0) \vert \psi_n \rangle \langle \psi_n \vert \phi({\bf y}, 0)\rangle \exp{(-E_n T)} \nonumber \\ &\to & \langle \phi({\bf x},0) \vert \psi_0 \rangle \langle \psi_0 \vert \phi({\bf y}, 0)\rangle \exp{(-E_0 T)} \quad {\rm for }\quad T\to \infty \label{inf} \end{eqnarray} which gives the Feynman--Kac formula for the ground state energy \begin{equation} E_0 = - \lim_{T \to \infty} {\log G(-iT) \over T}. \label{fey} \end{equation} The only condition for the validity of Eq. (\ref{fey}) is that the $\phi$ states have a non-vanishing component over the ground state. The same is still true for finite $T$ if the overlap with the ground state is not too small. This is precisely the way in which hadron masses are computed on the lattice. Of course, many tricks are used in order to maximize the overlap with the ground state in consideration. If the $\phi$ state denotes a state of two exactly static particles interacting at a distance $r$, then the ground state energy is a function of the particle separation, $E_0\equiv E_0(r)$, and gives the potential of the first adiabatic surface. With this in mind, we will perform in the remaining of this section an explicit evaluation of the quark-antiquark Green function for infinitely heavy quarks ($m_j\to \infty$) and for large temporal intervals ($T\to \infty$). In the following we will be working in the Euclidean space taking advantage of the usual relation between Euclidean position $x^E_\mu$, momentum $k^E_\mu$, field $A^E_\mu$, gamma matrices $\gamma^E$ and the corresponding quantities in Minkowski space \begin{eqnarray} & & t^E= it^M \qquad\quad~ x_i^E=x_i^M \nonumber \\ & & k_4^E = -i k_0^M \qquad\> k_i^E= k_i \nonumber \\ & & A_4^E= -i A_0^M \qquad\! A_i^E=A_i^M \nonumber \\ & & \gamma_4^E=\gamma_0 \qquad\qquad\! \gamma_i^E= -i \gamma^i . \label{euclidep} \end{eqnarray} Let us assume that at a time $t=0$ a quark and an antiquark are created and that they interact while propagating for a time $t=T$ at which they are annihilated. Then ($x_j=({\bf x}_j, T), y_j=({\bf y}_j, 0)$) \begin{eqnarray} & & G_{\beta_1\beta_2\alpha_1\alpha_2}(T) \nonumber\\ & & = \langle 0 \vert \bar{\psi}_{\beta_2}({\bf y}_2, 0)U(y_2,y_1) \psi_{\beta_1}({\bf y}_1,0)\bar{\psi}_{\alpha_1} ({\bf x}_1, T) U(x_1,x_2)\psi_{\alpha_2}({\bf x}_2,T)\vert 0\rangle \nonumber \\ & & = {1\over Z}\int {\cal D}\psi {\cal D}\bar{\psi} {\cal D}A \, \bar{\psi}_{\beta_2}({\bf y}_2, 0)U(y_2,y_1)\psi_{\beta_1}({\bf y}_1,0) \nonumber \\ & &\qquad\qquad \times \bar{\psi}_{\alpha_1}({\bf x}_1, T) U(x_1,x_2)\psi_{\alpha_2}({\bf x}_2,T) e^{-\int L \,\,d^4x} \label{green2} \end{eqnarray} where $L$ is the Euclidean version of Eq. (\ref{lagr}), $L=L_{YM} + L_F= {1\over 4} F_{\mu\nu} F_{\mu \nu} +\bar{\psi} (\gamma_\mu D_\mu + m) \psi $. The indices $\alpha, \beta$ are spinor indices, while the detailed structure of the colour indices is not displayed. Since the action is quadratic in the quark fields, it is possible to perform the fermion integration \begin{eqnarray} & & G_{\beta_1\beta_2\alpha_1\alpha_2}(T) \nonumber \\ & & = {1\over Z}\int {\cal D}A \, \big( {\rm Tr} \{ S_{\alpha_2 \beta_2}(x_2,y_2; A) U(y_2,y_1) S_{\beta_1\alpha_1}(y_1,x_1;A) U(x_1,x_2) \}\nonumber \\ & & \qquad - {\rm Tr} \{ S_{\beta_1 \beta_2} (y_1,y_2;A) U(y_2,y_1) \} {\rm Tr} \{ S_{\alpha_2 \alpha_1}(x_2,x_1;A) U(x_1,x_2) \} \big) \nonumber \\ & & \qquad \times \det{K(A)} e^{-\int L_{YM} d^4x} \label{green3} \end{eqnarray} where the trace is over the colour indices and $K$ is the fermionic determinant of the matrix $K_{\alpha x \beta y}(A) \equiv [\gamma_\mu D_\mu + m]_{\alpha \beta} \delta^4(x-y)$. In the following, we will assume the quenched approximation\footnote{In perturbation theory the logarithm of this determinant is given by the sum of Feynman diagrams consisting of fermion loops with an arbitrary number of fields $A_\mu$ attached to it. In the limit $m\to \infty$ this determinant approaches a constant (infinite but canceled by a factor in $Z$) and then $\det K= 1 \> +$ corrections of order $O(1/m^n)$. Therefore for heavy quarks the quenched approximation $\det K=1$ makes sense. On the contrary the fermionic determinant associated to light quarks in principle cannot be neglected. This determinant will eventually be responsible for the breaking of the string between the quarks.}, $\det K=1$. The second term in Eq. (\ref{green3}) describes quark-antiquark annihilation and hence appears only for quarks of the same flavour. Since this effect is dominated by the perturbative two or three gluons exchange in the s channel we will not consider this term any more here. Then we obtain \begin{equation} G(T) \simeq {1\over Z}\int {\cal D}A \, {\rm Tr} \{ S(x_2,y_2; A) U(y_2,y_1) S(y_1,x_1;A) U(x_1,x_2) \} e^{-\int L_{YM} d^4x}. \label{utile} \end{equation} In Eq. (\ref{utile}) $S(x,y;A)$ denotes the quark propagator in the presence of the gluon field $A_\mu$. It obeys the equation \begin{equation} (\gamma_\mu D_\mu + m)S(x,y;A)= \delta^4(x-y). \label{direq} \end{equation} This is in principle a system of coupled partial differential equations that cannot be solved in a closed form for an arbitrary $A_\mu$. However, we are interested in the limit $m\to \infty$. In this approximation (Wilson (1974), Brown and Weisberger (1979)) we can replace $S$ by the static solution $S_0$ obtained dropping the spatial part of the gauge-covariant derivative in (\ref{direq}) while maintaining the time component. This approximation maintains the manifest gauge invariance. Then we have \begin{equation} (\gamma_4 D_4 + m)S_0(x,y;A)= \delta^4(x-y) \label{eqstat} \end{equation} which is an ordinary differential equation solvable in a closed form. Indeed, we can get rid of the $A_4$ in the equation making the ansatz \begin{equation} S_0(x,y;A) = {\rm P} \exp{\left\{ ig \int_{x_4}^{y_4} dt A_4({\bf x},t)\right\} } \hat{S}_0(x-y) \label{ans} \end{equation} with $\hat{S}_0$ satisfying \begin{equation} (\gamma_4\partial_4 + m) \hat{S}(x-y)=\delta^4(x-y). \label{rideq} \end{equation} Therefore the solution has the form \begin{eqnarray} S_0(x,y;A) &=& \delta^3({\bf x}-{\bf y}) {\rm P} e^{ ig \int_{x_4}^{y_4} dt A_4({\bf x},t)} \left\{\theta(x_4-y_4) {1+\gamma_4\over 2} e^{ -m (x_4-y_4) }\right.\nonumber \\ & & \qquad\qquad \left. + \theta(y_4-x_4) {1-\gamma_4\over 2} e^{- m (y_4-x_4) } \right\}. \label{soles} \end{eqnarray} This expression shows that the time evolution of a (infinitely) heavy quark field consists purely in the accumulation of phase determined by $A_4$ and the quark mass (cf. Ex. 4.1.1). The spatial delta function says that the infinitely heavy quark cannot propagate in space\footnote{ For this solution the annihilation term in Eq. (\ref{green2}) does not give contribution since the ${\bf x} = {\bf y}$ condition is not satisfied.}. \begin{figure}[htb] \vskip -0.1truecm \makebox[4.0truecm]{\phantom b} \epsfxsize=4truecm\epsffile{gre0.eps} \vskip 0.1truecm \caption{\it Static Wilson loop with contour $\Gamma_0$.} \label{plwil} \end{figure} The quark-antiquark Green function is given by \begin{eqnarray} G_{\beta_1\beta_2\alpha_1\alpha_2}(T) &{\buildrel {m_j\to\infty}\over \longrightarrow }& \, \delta^3({\bf x}_1-{\bf y}_1) \delta^3({\bf x}_2-{\bf y}_2) (P_+)_{\beta_1\alpha_1} (P_-)_{\alpha_2\beta_2} \nonumber \\ & & \times e^{-(m_1+m_2) T}\langle {\rm Tr}\, {\rm P} e^{ig \oint_{\Gamma_0} dz_\mu A_\mu (z) }\rangle \label{solgreen} \end{eqnarray} with $P_\pm\equiv (1\pm \gamma_4)/2$. The integral in Eq. (\ref{solgreen}) extends over the circuit $\Gamma_0$ which is a closed rectangular path with spatial and temporal extension $r=\vert {\bf x}_1 -{\bf x_2}\vert$ and $T$ respectively, and has been formed by the combination of the path-ordered exponentials along the horizontal (=time fixed) lines, coming from the Schwinger strings, and those along the vertical lines coming from the static propagators (see Fig. \ref{plwil}). The brackets in (\ref{solgreen}) denote the pure gauge vacuum expectation value. In Euclidean space \begin{equation} \langle f[A]\rangle \equiv {1\over Z} \int {\cal D} A f [A] e^{-\int d^4 x L^E_{YM}}. \label{expece} \end{equation} From Eq. (\ref{solgreen}) it is clear that the dynamics of the quark-antiquark interaction is contained in \begin{equation} W(\Gamma_0) = {\rm Tr\, P} e^{\displaystyle i g \oint_{\Gamma_0} dz_\mu A_\mu (z) }. \label{wilsstat} \end{equation} This is the fa\-mous ({\it static}) Wegner--Wilson loop (Wegner (1971) and Wil\-son (1974)). In the limit of infinite quark mass considered, the kinetic energies of the quarks drop out of the theory, the quark Hamiltonian becomes identical with the potential (see Ex. 4.1.1) while the full Hamiltonian contains also all types of gluonic excitations. According to the Feynman--Kac formula the limit $T\to \infty $ projects out the lowest state i.e. the one with the ``glue'' in the ground state. This has the role of the quark-antiquark potential for pure mesonic states. Now, comparing Eq. (\ref{solgreen}) with Eq. (\ref{inf}) and considering that the exponential factor $\exp{(- (m_1+m_2) T)}$ just accounts for the the fact that the energy of the quark-antiquark system includes the rest mass of the pair\footnote{Moreover $E(R)$ includes also self-energy effects which need to be subtracted when calculating the quark-antiquark potential.}, we obtain \begin{equation} V_0(r) \equiv E(r) = -\lim_{T\to \infty} {1\over T } \log \langle W(\Gamma_0) \rangle. \label{potfond} \end{equation} The quark degrees of freedom have now completely disappeared and the expectation value in (\ref{potfond}) has to be evaluated in the pure Yang--Mills theory ((Wilson (1974), Brown and Weisberger (1979)). Notice that the potential is given purely in terms of a gauge invariant quantity (the Wilson loop precisely). In this way we have reduced the calculation of the static potential to a well posed problem in field theory: to obtain the actual form of $V_0$ we need to calculate the QCD expectation value of the static Wilson loop. We conclude this section pointing out that Eq. (\ref{potfond}) is rigorously true for static sources. For (realistic) heavy quarks with finite mass, the static potential, interpreted as the static limit of the potential appearing in the Schr\"odinger equation, could in principle not coincide exactly with Eq. (\ref{potfond}). Actually it does not. The reason is that in QCD quarks in the static limit can still change colour by emission of gluons. This introduces a new dynamical scale in the evaluation of the potential from Eq. (\ref{potfond}) of the order of the kinetic energy which is finite if the quarks have finite mass. In perturbative QCD this new scale is given by the difference between the singlet and the octet potential. Contributions of the same order of the kinetic energy are not of potential type and have to be explicitly subtracted out from Eq. (\ref{potfond}). In the next section we will give the leading effect of this subtraction on the static Wilson loop. For an extended analysis we refer to Brambilla et al. (1999). \vskip 1truecm \leftline{\bf 4.1 Exercises} \begin{itemize} \item[4.1.1]{ Consider a particle of mass $m$ moving in a potential $V(x)$ in one space dimension. The propagator is given by $\displaystyle K(x^\prime, t; x, 0)= \langle x^\prime \vert \exp{(-i Ht)}\vert x\rangle$ with $H= \displaystyle{p^2\over 2 m} +V(x)$. Obtain the form of the propagator $K$ in the static limit $m\to \infty$ and compare it with Eq. (\ref{soles}).} \item[4.1.2]{ Consider quenched QED. Show that the functional generator for the QED Lagrangian with a source term added of the form $J_\mu(x) A_\mu(x) $ with $J_\mu(x) =e \delta_{\mu 4} (\delta^3({\bf x} -{\bf x}_1)-\delta^3({\bf x} -{\bf x}_2)) $, coincides with the vacuum expectation value of the static Wilson loop in the limit of infinite interaction time.} \end{itemize} \subsection{The Wilson loop in perturbative QCD} The Wegner--Wilson loop of contour $\Gamma$ is defined in QCD as \begin{equation} W(\Gamma)\equiv {\rm Tr\, P}\, e^{ig \displaystyle\oint_{\Gamma} d z_\mu A_\mu(z) }. \label{wilver} \end{equation} Due to the presence of the colour trace, it is a manifestly gauge invariant object. The field $A_\mu$ can be taken in any representation of $SU(3)$. When describing the quark-antiquark interaction, as in the present case, $A_\mu\equiv A_\mu^a \lambda^a /2$. The Wilson loop is called static when the integral is extended to a rectangular $\Gamma_0$ as in Fig. \ref{plwil}. Physically, in the static Wilson loop only the time component $A_4$ is relevant. We know from the previous section that the vacuum expectation value of $W$ on the QCD measure gives the static potential. However, we are in trouble when we step in to calculate it. Indeed, if we want to describe the long range quark-antiquark interaction, we should be able to calculate the Wilson loop in the region in which the running coupling constant is no longer small and we should be able to sum up all the relevant diagrams. Unfortunately, we have no methods at hand to sum up such contributions. Worse enough, also usual semiclassical approaches are not doomed to work in this case. We do not know of any dominant and confining configurations in the QCD measure and in the path integral that can make the work\footnote{Instantons do not confine directly, i.e. give a zero string tension. Monopoles arise in the Abelian projection or after a dual transformation, see Sec.6.}! Hence, to obtain information on the behaviour of the Wilson loop in the nonperturbative region we have to resort either to strong coupling expansion and to lattice simulations, see Sec. 4.3 and Sec. 4.5, or to analytic models of the QCD vacuum, see Sec. 6. However, in the weak coupling region the Wilson loop can be calculated perturbatively. Recently, the fully analytic calculation of the two-loop diagrams contributing to the static potential has been performed (Peter (1997), Schr\"oder (1999)). We refer to the original papers for the details of the calculation. However, due to the high interest, we report here the final result ($\alpha_{\rm s}$ is in the $\overline{MS}$ scheme): \begin{eqnarray} && V_0(r) = -C_F{\alpha_V(r) \over r}\label{vr}\\ & & \alpha_V (r) = \alpha_{\rm s}(r) \left\{1+\left(a_1+ {\gamma_E \beta_0 \over 2}\right) {\alpha_{\rm s}(r) \over \pi}\right. \nonumber\\ &&+\left.\left[\gamma_E\left(a_1\beta_0+ {\beta_1 \over 8}\right)+\left( {\pi^2 \over 12}+\gamma_E^2\right) {\beta_0^2 \over 4}+b_1\right] {\alpha_{\rm s}^2(r) \over \pi^2} \right \}, \nonumber \end{eqnarray} where $\gamma_E$ is the Euler constant, $a_0=1$, $$ a_1 = \frac{31}{9}C_A - \frac{20}{9}T_FN_f $$ and \begin{eqnarray*} a_2 &=& \Big(\frac{4343}{162}+4\pi^2-\frac{\pi^4}{4}+\frac{22}{3}\zeta_3\Big)C_A^2 -\Big(\frac{1798}{81}+\frac{56}{3}\zeta_3\Big)C_AT_FN_f\nonumber \\ && -\Big(\frac{55}{3}-16\zeta_3\Big)C_FT_FN_f + \Big(\frac{20}{9}T_FN_f\Big)^2. \end{eqnarray*} $C_F = 4/3$ and $C_A=3$ are the Casimir of the fundamental and of the adjoint representation respectively. Moreover in QCD we have $T_F = 1/2$. From this result we learn that: the two-loop contribution is nearly as large as the one-loop term, both make the potential more attractive and eventually the perturbative potential seems to be reliable up to a distance $r\Lambda_{QCD} < 0.07$, which is considerably smaller than the average radius in quarkonia (see Fig. \ref{plpot}). For larger values a strong scale-dependence remains and the perturbation series breaks down above 0.1 fm. Hence, even the pure perturbative calculation indicates the need of a different long range approach\footnote{In Pineda and Yndurain (1998) the two-loop static potential and the one-loop relativistic perturbative corrections to the potential were used in order to calculate the ground state energies of bottomonium and charmonium and thus to obtain a value for the bottom and charm masses. Nonperturbative effects were encoded in the local gluon condensate. In the next section, we will show that nonperturbative contributions are actually carried by non-local quantities, Gromes (1982), like the Wilson loop, which can be approximated by local condensates only if the involved physical scales enable a local expansion. This is indeed the case of the bottomonium ground state.}. We mention that the next perturbative correction to the static potential can be obtained only in an effective theory framework (pNRQCD, see Sec. 5). In that framework the leading log three-loop term has been very recently calculated in Brambilla et al. (1999). It amounts to a correction $C_A^3 \alpha_{\rm s}^4(r) / 12 \pi \log{ r \mu^\prime}$ to $\alpha_V$ in Eq. (\ref{vr}), $\mu^\prime$ being the scale of the matching. Notice that the potential comes to depend on the infrared scale $\mu^\prime$. This signals the appearance at three-loop of a nonpotential type of contribution to the static Wilson loop which has been subtracted out explicitly at $\mu^\prime$. In QED in the quenched approximation (i.e. neglecting light fermions) the (non-static) Wilson loop can be calculated analytically in a closed form. We sketch here the derivation. We have (in Euclidean space) \begin{equation} \langle W(\Gamma)\rangle = { \displaystyle\int {\cal D} A \, e^{\displaystyle {1\over 2} \displaystyle\int d^4 x A_\mu\, M_{\mu \nu} A_\nu +i e \oint dz_\mu A_\mu }\over \displaystyle\int {\cal D} A \, e^{\displaystyle{1\over 2} \displaystyle\int d^4 x A_\mu\, M_{\mu \nu} A_\nu}} \label{wilqed} \end{equation} where $M_{\mu\nu} \equiv \delta_{\mu \nu } \partial^2 -\partial_\mu \partial_\nu $ has been obtained by integrating by parts the original QED action. The integral over the $A$ field in Eq. (\ref{wilqed}) is Gaussian and therefore can be performed provided that a gauge condition is imposed. However, since the Wilson loop is a gauge-invariant quantity, the choice of the gauge is immaterial. We obtain \begin{equation} \langle W(\Gamma) \rangle = \exp\left\{ {e^2\over 2} \oint_\Gamma dx_\mu \oint_\Gamma dx^\prime_{\nu} D_{\mu \nu}(x-x^\prime) \right\} \label{solqed} \end{equation} where $D_{\mu \nu}(x-x^\prime) \equiv \langle A_\mu(x) A_\nu(x^\prime)\rangle$ is the photon propagator. E.g. in Feynman gauge we have $D_{\mu \nu}(x) = - \displaystyle{\delta_{\mu\nu} \over 4\pi^2}{1\over x^2}$. On a rectangular Wilson loop Eq. (\ref{solqed}) becomes \begin{equation} \langle W(\Gamma_0) \rangle = \exp\left\{{e^2\over 4 \pi r} T f(r,T)\right\} \label{intwil} \end{equation} with $f(T,r) =\displaystyle{2\over \pi} \left[{\rm arctan}{T\over r} -{r\over 2 T}\log(1+{T^2\over r^2})\right] $, $f\to 1 $ for $T\to \infty$ getting back the Coulomb potential. \begin{figure}[htb] \vskip -0.1truecm \makebox[4.0truecm]{\phantom b} \centerline{ \epsfxsize=5truecm\epsffile{gre2.eps}} \vskip 0.1truecm \caption{\it Generalized Wilson loop with contour $\Gamma$.} \label{plwil2} \end{figure} In the weak coupling region the vacuum expectation value of the (non-static) Wilson loop in QCD can be obtained by making a Gaussian approximation on the functional integral (i.e. by neglecting non-Abelian contributions). In this case, since the contribution coming from the Schwinger strings vanish in the limit $T \to \infty$ and taking advantage of the notation given in Fig. \ref{plwil2}, we have \begin{eqnarray} \langle W(\Gamma) \rangle \! &=& \! e^{\displaystyle{4\over 3} g^2 \displaystyle\oint_{\Gamma} dz^1_\mu \oint_{\Gamma} dz^2_\nu D_{\mu\nu}(z_1 - z_2)}\nonumber \\ &{\buildrel { T\to \infty} \over \longrightarrow}& \! e^{\displaystyle{4\over 3} g^2 \displaystyle \int_{t_i}^{t_f} dt_1 \int_{t_i}^{t_f} dt_2 \dot{z}_\mu^1(t_1) \dot{z}_\nu^2(t_2) D_{\mu\nu}(z_1(t_1) - z_2(t_2))}\, , \label{propp} \end{eqnarray} $D_{\mu\nu}$ being the gluon propagator. \vskip 1truecm \leftline{\bf 4.2. Exercises} \begin{itemize} \item [4.2.1]{ From Eqs. (\ref{potfond}) and (\ref{propp}) obtain the QCD one gluon exchange contribution to the static QCD potential.} \item [4.2.2]{ Calculate Eq. (\ref{intwil}) from Eq. (\ref{solqed}).} \item [4.2.3]{ From the correspondent of Eq. (\ref{willexpp}) (Sec. 5) in Euclidean space, obtain the QCD $V_0$ and $V_{\rm VD}$ potentials using the weak coupling behaviour of the Wilson loop given in (\ref{propp}) with the gluon propagator first in the Coulomb gauge and then in the Feynman gauge. Demonstrate that, due to the gauge invariance of the Wilson loop, the two expression coincide. [Hint: perform the change of variables $t=(t_1+t_2)/2; \tau=t_1-t_2$ in the integrals in (\ref{propp}), expand ${\bf z}_j$ around $t$ and integrate over $\tau^{+\infty}_{-\infty}$.]} \end{itemize} \subsection{Lattice formulation, strong coupling expansion and area law} Equation (\ref{potfond}) is particularly useful on the lattice where the dynamical variables are unitary matrices associated with the links. We do not want to give here an introduction to lattice QCD (for this we refer the reader to Rothe (1992) and Montvay and Munster (1994)). However, in order to illustrate some interesting results, we recall the basic definitions and concepts. \begin{figure}[htb] \makebox[4.0truecm]{\phantom b} \epsfxsize=4truecm\epsffile{plaquette.eps} \caption{\it Elementary plaquette on the lattice and tiling of the Wilson loop in strong coupling.} \label{plpla} \end{figure} Let us consider QCD in the pure gauge sector on a four-dimensional Euclidean discretized space-time, the ``lattice'' of step $a$. Lattice sites are denoted by $n $ and lattice directions are denoted by $\mu, \nu$. We define the group element associated with a link, from the lattice site $n$ to $n+\hat{\mu}$ ($\hat{\mu} $ being a unit vector along the axis $\mu$) as \begin{equation} U_\mu(n)\equiv U(n,n+\hat{\mu})\simeq e^{ig a A_\mu(n+{\hat{\mu}\over 2 })} \quad \quad U^\dagger_\mu(n) \equiv U(n+\hat{\mu},n) . \label{link} \end{equation} These are the dynamical variables, the gluonic colour fields that relate the colour coordinate system at different space-time points. From (\ref{link}) we see that they are represented as a $3\times 3$ colour matrix that can be interpreted as the path-ordered exponential of the continuum colour fields $A_\mu(x)$. The gauge transformation on the lattice, ${\cal V}(x_n)$, acts directly on the link elements \begin{equation} U_\mu(n) \to {\cal V}^\dagger(n+\hat{\mu}) U_\mu(n) {\cal V}(n) . \label{trasf} \end{equation} The Yang--Mills action is \begin{equation} S= -{\beta\over 3} \sum_{\mu > \nu} \sum_n \, {\rm Re}\, {\rm Tr}\, U_{\mu\nu} \label{latac} \end{equation} with $\beta= 6/g^2$. The elementary plaquette is the trace of the path-ordered product of links around a unit square (see Fig. \ref{plpla}) \begin{equation} U_{\mu\nu}(n)\equiv U_\mu(n) U_\nu(n+\hat{\mu})U^\dagger_\mu(n+\hat{\nu})U^\dagger_\nu(n); \quad \quad U_P\equiv {1\over 3} {\rm Re}\, {\rm Tr} \, U_{\mu\nu} . \label{defplaq} \end{equation} In the continuum limit ($a\to 0$) expression (\ref{latac}) reduces to the usual Yang--Mills action\footnote{ Of course there exists an infinite number of lattice actions that have the same naive continuum limit, in particular when one considers also the fermion contribution. To be sure that the given lattice action reproduces really QCD for $a\to 0$, the lattice theory should exhibit a critical region in parameter space where the correlation length diverges. See also Sec. 4.5.}. The corresponding partition function is \begin{equation} Z= \int \prod_{n,\mu} {\cal D}U e^{-{\beta} \sum_P U_P } \label{part} \end{equation} where the integral is over the group manifold of colour $SU(3)$ for each link matrix $U$. The Wilson loop is simply the trace of the product of the matrices $U(n,\mu) $ along the contour $\Gamma$ which here is a rectangle. Then \begin{equation} \langle W(\Gamma) \rangle =\langle {\rm Tr} \, \prod_{l\in \Gamma} U(l,\mu_l) \rangle= {1\over Z} \int \prod_{n,\mu} DU(n,n+\hat{\mu}) {\rm Tr}\, \prod_{l\in \Gamma} U(l,\mu_l) e^{-{\beta} \sum_P U_P} . \label{will} \end{equation} Of particular interest is the strong coupling expansion on the lattice, which means to expand for large $g$ (small $\beta$). This bears a relation to the string picture that, as we explained, characterizes the long range quark-antiquark interaction. In the strong coupling, we can expand the exponential of the action in (\ref{will}) \begin{eqnarray} & & \langle W(\Gamma) \rangle = {1\over Z} \int \prod_{n, \mu} DU(n,n+\hat{\mu}) {\rm Tr}\, \prod_{l\in \Gamma} U(l,\mu_l) \nonumber \\ & & \quad \times \left[1-{\beta} \sum_P {\rm Tr}\, U_P +{1\over 2} {\beta }^2 \sum_P\sum_{P^\prime} {\rm Tr}\, U_P\, {\rm Tr}\, U_{P^\prime} +\cdots \right]. \label{exp} \end{eqnarray} Since each plaquette in the expansion costs a factor $\beta$, the leading contribution in the limit $\beta \to 0$ is obtained by paving the inside of the Wilson loop with the smallest number of elementary plaquettes yielding a non-vanishing value for the integral. Using the orthogonality relations supplied in Exercise 4.3.1, it is possible to show that the relevant configuration is the one presented in Fig. \ref{plpla} and that \begin{equation} \langle W(\Gamma) \rangle \simeq \left({1\over g^2}\right)^{N_P} \, , \label{area} \end{equation} $N_P$ being the minimal number of plaquettes required to cover the area enclosed by the path $\Gamma$ (for more details see Creutz (1983)). This corresponds to the area law (Wilson (1974)) since the area enclosed by the path $\Gamma$ is given by $A(\Gamma)=a^2 N_P$. Furthermore, it is possible to demonstrate that the strong coupling expansion (\ref{exp}) has a finite radius of convergence. Hence, for $g^2$ large enough, the vacuum expectation value of the Wilson loop has the behaviour \begin{equation} \langle W(\Gamma) \rangle \simeq (g^2)^{-A(\Gamma)/a^2} = e^{-\displaystyle{r\, T \log g^2 \over a^2}} \label{behav} \end{equation} where the last equality holds for a rectangular path $r \times T$. The behaviour of the Wilson loop given by Eq. (\ref{behav}) leads to a linear static potential with \begin{equation} \sigma \simeq {\log g^2\over a^2}. \label{sigimp} \end{equation} The layer of plaquettes giving this area contribution corresponds to a constant (chromo)electric field along the string connecting the quark and the antiquark. This suggests once again the relevance of a flux tube description of the nonperturbative interaction and is at the origin of the formulation of the flux tube model of Isgur et al. (1983). However, we have to keep in mind that the continuum limit of lattice QCD is reached in the weak coupling limit, $\beta \to\infty$. Therefore it is impossible to extrapolate the strong coupling results directly to the continuum physics. Still, we can argue that a rather coarse lattice with large lattice spacing should already give some indicative results for a theory {\it without} phase transition. To such a lattice corresponds a large bare coupling $g$ and hence the strong coupling result (\ref{behav}) should give the correct qualitative picture. Our philosophy is to take the behaviour (\ref{behav}) as a reliable suggestion to be used in the continuum physics. The leading order strong coupling expansion of the Wilson loop in QED is identical to Eq. (\ref{behav}) and therefore produces a linear potential too. However, in the case of QED on the lattice a phase transition is clearly seen (Kogut et al. (1981)) when going from strong coupling to weak coupling. In QCD the analytic proof that there cannot be such a phase transition together with the finite radius of convergence of the strong coupling expansion would be equivalent to {\it a proof of confinement}. Such a proof does not exist up to now. However, the numerical lattice simulations present no hint of such a transition in the intermediate coupling region. On the contrary, the strong coupled behaviour $g^2(a) \sim e^{\sigma a^2}$ continuously goes into the weak coupling $g^2(a) \sim 1 / \log{a^{-1}}$ as $a \to 0$. Moreover, within the coupling regions accessible to present day computers, there are already overlaps between lattice results and weak coupling expansion, the most impressive being the result of the alpha collaboration, see Capitani et al. (1998). These results hint to the fact that QCD possesses both the property of asymptotic freedom and colour confinement. In other words the Wilson loop in QCD displays a perimeter law in weak coupling and an area law in strong coupling. \vskip 1truecm \leftline{\bf 4.3 Exercises} \begin{itemize} \item[4.3.1]{ The orthogonality properties of the group integral in $SU(3)$ are given by \begin{eqnarray*} & & \int dU(n, n+\hat{\mu}) [U(n,n+\hat{\mu})]_{ij} =0 \nonumber \\ & & \int dU(n, n+\hat{\mu}) [U(n,n+\hat{\mu})]_{ij} [U^\dagger(n,n+\hat{\mu})]_{kl} = {1\over 3} \delta_{il} \delta_{jk} \nonumber \\ & & \int dU(n, n+\hat{\mu}) [U(n,n+\hat{\mu})]_{ij} [U(n,n+\hat{\mu})]_{kl} = 0 \end{eqnarray*} Using these relations justify the result (\ref{area}).} \end{itemize} \subsection{Area law as a criterium for confinement, \\ duality and the Wilson loop as an order parameter} To see whether QCD shows confinement, one can study the energy of a system composed of a quark and an antiquark along the lines exposed in Sec. 4.1. Then Eq. (\ref{potfond}) tells us that it is the Wilson loop and its behaviour that determines the confinement property of the theory. In the previous section we have seen that in QCD in strong coupling expansion the Wilson loop in the fundamental representation obeys an area law behaviour and this in turn via Eq. (\ref{potfond}) confirms the property of quark confinement. For very large loops $\log \langle W(\Gamma) \rangle$ generally exhibits these two types of behaviour: it decreases either as the perimeter or as the area of $\Gamma$. In the first case, expansive loops are allowed and quark and antiquark can be far apart from each other. In the second case quark and antiquark propagate as a bound state. This is the Wilson criterion for confinement of electric charges (Wilson (1974)) \begin{eqnarray} \langle W(\Gamma) \rangle & & \sim e^{- K L(\Gamma)} \quad \quad~ {\rm no} \> {\rm confinement}\\ \langle W(\Gamma) \rangle & & \sim e^{- K^\prime A(\Gamma)} \quad \quad {\rm confinement} \label{order} \end{eqnarray} with $L(\Gamma)$ the perimeter of $\Gamma$, $A(\Gamma)$ the minimal area enclosed by $\Gamma$, and $K$ and $K^\prime$ dimensionful constants. These are statements about the response of the pure gauge vacuum to external perturbations. This criterion inspires physical pictures of the QCD vacuum (see Sec. 6). In particular it was suggested by 't Hooft (see e.g. 't Hooft (1994)) that a pure non-Abelian gauge theory could display quite a complicate pattern of vacuum phases: 1) The Higgs mode. Only colour magnetic charges are confined. 2) The Coulomb mode, featuring ordinary massless ``photons'' and no superconductivity or confinement. 3) The confinement mode or ``magnetic superconductor''. Quarks (i.e. electric charges) are permanently confined. The properties of these phases are then expressed in terms of operators that create vortices of electric flux (Wilson loops $W_A(\Gamma)$, evaluated on the gauge fields $A_\mu$) and operators that create vortices of magnetic flux ('t Hooft operators $W_C(\Gamma)$, evaluated on the dual potentials $C_\mu$). Using the definition of $W_C(\Gamma)$ via its commutation relation with $W_A(\Gamma)$, 't Hooft (1979) showed that if $W_A(\Gamma)$ obeys an area law, then $W_C(\Gamma)$ necessarily satisfies a perimeter law. The phase in which the 't Hooft operator obeys an area law is the Higgs phase, while that one in which the Wilson loop obeys an area law is the confinement phase. 't Hooft result is a precise way of saying that the vacuum of a confining theory has the properties of a magnetic superconductor. The key ideas are the fact that a non-Abelian gauge theory can be viewed as an Abelian gauge theory enriched with Dirac magnetic monopoles and the concept of electric-magnetic duality. For QCD this implies that the vacuum behaves as a dual superconductor and confinement is explained through the monopole condensation, see Sec. 6. However, if in an Abelian theory it is possible to introduce dual field strengths and electric potentials $C^\mu$ dual to the ordinary (magnetic) potentials $A^\mu$ in a straightforward manner, in a non-Abelian theory it is not possible to express the dual potentials explicitly in terms of ordinary potentials. The explicit form of the exact Yang--Mills Lagrangian as a function of the $C^a_\mu$ fields is unknown. The construction of the long distance limit of the Lagrangian is based on the fact that the dual potentials are weakly coupled since the dual Wilson loop obeys a perimeter law. The quadratic part of the dual Lagrangian is thereby determined. The minimal extension can be constructed under requisite of dual gauge invariance. For a concrete construction of a QCD dual Lagrangian see Baker et al. (1985)-(1995); Maedan et al. (1989) and Sec. 6. These ideas have originated a quite intensive research activity in the ``nonperturbative'' physics community in the last few years (in QCD mainly on the lattice while in supersymmetric field theories very promising analytic results are obtained just now, see e.g. Alvarez-Gaum\'e et al. (1997)). We will come back to this point in Sec. 6.2. Here we refer the reader to the historical papers of Nambu (1974), Mandelstam (1979) and 't Hooft (1982) and references therein. \subsection{Lattice simulations and lattice results} With the action of Eq. (\ref{latac}), using periodic boundary conditions in space and time and taking a lattice of finite volume and finite spacing, the system has a finite number of degrees of freedom: the gluon fields on the link and (if we add them) the quark fields at the lattice sites. The functional integral over the gauge fields is converted to a multiple integral with a positive definite integrand (in Euclidean space): Eq. (\ref{part}). Unlike in continuum perturbation theory the calculation of averaged values of gauge invariant observables is done without any gauge-fixing. For a lattice of $L^4$ sites with colour group $SU(3)$ the integral in (\ref{part}) would be a $8\times 4 \times L^4$ dimensional integral: the standard approach is to use a Monte Carlo approximation to the integrand. Precisely, a stochastic estimate of the integral is made from a finite number of samples (the ``configurations'') of equal weight. For more details see Rothe (1992). Here we are interested only in explaining that in this way we have at our disposal a set of samples of the vacuum, then, it is possible to evaluate the average of fields over these samples and obtain a nonperturbative evaluation of Green functions as well as of any field vacuum correlator\footnote{ Analytic continuation from Euclidean to Minkowski time of Green functions that are only available on a finite set of points is not possible. Therefore, on the lattice one can only determine masses and on-shell matrix elements in a straightforward way. Real time processes like scattering, hadronic decays, are not directly accessible.}. In this section we present some lattice results on the calculation of the Wilson loop expectation value, the static potential, the flux tube configuration and the determination of the string tension $\sigma$. However, before, we comment briefly on the validity of the lattice results and give few concepts to enable the ``reading'' of those results. The lattice is not the real world so that before extracting the physics we have to be sure that: 1) the Green functions have been extracted without contamination (the ground state mass should not be contaminated by pieces coming from the excited states); 2) the lattice size is big enough, we mean that the relevant physical distance should fit in! (finite size error); 3) the statistical errors of the calculation is under control (statistical error); 4) the lattice spacing is small enough, since we have to approach the continuum limit (discretization error). On the other hand the lattice spacing supplies an explicit ultraviolet cut-off. The last condition is the most subtle one. Let us just sketch the idea. In order to extract the continuum limit from the lattice, it has to be shown that the results do not change if the lattice spacing is decreased further. However, the lattice spacing in physical units is not known directly. Actually it is measured. The lattice simulations are performed at a fixed value of $\beta$ (cf. Eq. (\ref{latac})). We recall that $\beta= 6/g^2$ and therefore a large $\beta$ corresponds to a small $g^2$. Ensembles at different values of $a$ are obtained by using different bare coupling constants $g$ in the action. However, the value of $a$ for a given value of $g$ is not known {\it a priori} but has to be obtained calculating a dimensionful parameter and comparing it to an experiment. As an example let us consider the string tension $\sigma$. Measured in lattice units it is only a function of the bare coupling: $\hat\sigma(g)$, the $\hat{}$ denoting a dimensionless quantity. In physical units, however, it has the dimension of a (mass)$^2$, so that the physical string tension is given by \begin{equation} \sigma = \lim_{a\to 0} {1\over a^2} \hat{\sigma}(g(a)). \label{sigfis} \end{equation} It approaches a finite limit for $a\to 0$, if $g$ is tuned with $a$ in an appropriate way. By requiring that in this limit \begin{equation} a{\partial \over \partial a}\sigma =0 \label{eqdif} \end{equation} we obtain \begin{eqnarray} \sqrt{\sigma}&=& c_\sigma \Lambda_L \nonumber\\ \Lambda_L &=& a^{-1} e^{-1\over 2 \beta_0 g^2} (\beta_0 g^2)^{-\beta_1\over 2 \beta_0^2} (1+ O(g^2)) \label{el} \end{eqnarray} where we have used the usual lattice convention for the beta function: $\beta(g)$ $=$ $-\beta_0 g^3$ $-\beta_1 g^5$ $+ O(g^7)$ (which differs for a factor $1/(4\pi)^2$ from Eq. (\ref{betaqcd})) and $\Lambda_L$ is independent of $a$ and fixes the mass scale of the theory\footnote{The appearance of a scale as $\Lambda_L$ is well known from perturbative continuum QCD, where the necessity of renormalizing the theory also requires the introduction of a scale. Usually the relation between $\Lambda_{QCD}$ in different regularization schemes and $\Lambda_L$ is known in perturbation theory in the bare coupling. See however S. Capitani et al. (1998).}. This equation tells us again that a perturbative calculation of $\sigma$ is a priori impossible due to the non-analytic behaviour in $g$ of Eq. (\ref{el}). It is possible to show that all dimensionful physical quantities can be expressed in the form \begin{equation} \theta_{phys}= \lim_{a\to 0} {1\over a^{d_\theta}}\hat{\theta}(g(a),a)=c_{\theta} (\Lambda_L)^{d_\theta} \label{thet} \end{equation} $d_\theta$ being the naive dimension. Then, for small lattice spacing we can determine the universal function $g=g(a) $ by fixing the l.h.s. of (\ref{el}) at the physical value of the string tension. This gives $g$ as a function of $a \sqrt{\sigma}$ \begin{equation} g^2(a)= - {1\over 2 \beta_0 \log(a \Lambda_L)} \label{behavg} \end{equation} (where $\Lambda_L$ is now determined by the physical condition imposed) which ensures the finiteness of any observable and allows us to convert them to physical units\footnote{A corresponding statement is expected to hold if the action depends on several parameters, e.g. coupling constant and quark masses.}. The continuum limit has to be taken at a constant physical volume $L^4$. As $a$ is decreased the number of lattice points $\hat{L}={L/a}$ increases. Therefore, a finite computer limits the lattice spacing to $a \geq a_{\rm min} > 0$. In practice one is looking for scaling of the results within this window or at least checking that the results follow the expected leading order $a$ dependence in order to safely extrapolate them to $a=0$. A typical value for lattice simulations of QCD is $\beta \simeq 6$ and hence the bare coupling constant is $\alpha_{\rm s}= g^2/ (4 \pi) \simeq 0.08$. In the early years of lattice QCD it was widely assumed that at least for $\beta \simeq 6$, two-loop perturbation theory can be applied to $a(g)$. Therefore, after having determined $a$ at a coupling $g$, the $\Lambda$ parameter was extracted and $a(g\prime)$ for $g\prime \neq g$ computed via perturbation theory. Nowadays, the lattice spacing is determined separately for each simulation point $\beta$ by inputting one experimental value. In doing so, $a(g)$ is obtained as a function of the bare lattice $g$. One finds big deviations from perturbation theory (asymptotic scaling) that are related to the importance of the so called tadpole diagrams in lattice perturbation theory (see e.g. Michael (1997), Davies (1997) and Lepage et al. (1993) for these developments)\footnote{One might ask if it was nonetheless possible to invert the $a(g)$ relation to obtain an $\alpha(q)$ and run the $q$ to high momenta subsequently. For this purpose effective couplings other than $g$ have been suggested and determined which show an improved asymptotic scaling behaviour.}. The actual methods of simulating lattice QCD and extracting physical information have reached quite a high level of sophistication and we refer the interested reader to the reviews (e.g. Rothe (1992), Davies (1997), Montvay and Munster (1994)). The above discussion should be sufficient to make clear that: 1) lattice simulations are performed at a fixed value of $\beta$, 2) the value of $\beta$ is connected with the lattice spacing $a$, 3) results relevant for continuum physics are effectively independent of $a$ (scaling) 4) the extraction of the physical results requires to fix $a$ and in the quenched approximation it may be dependent on the experimental quantity chosen to fix it. It is, therefore, important to know what quantity was chosen when looking at the lattice data. \begin{figure}[htb] \makebox[1.0truecm]{\phantom b} \centerline{ \epsfxsize=8truecm \epsffile{fig1.ps}} \vskip 0.1truecm \caption{\it Static potential at $\beta= 6.0$ and $\beta=6.2$ in the quenched approximation. The solid line is a fit of the form of Eq. (\ref{paramm}). Bali et al. (1997).} \label{plpotlat} \end{figure} Now let us come to some results. First, we discuss the lattice measure of $\sigma$. The lattice counterpart of Eq. (\ref{potfond}) is \begin{equation} \hat{V}(\hat{r})= -\lim_{\hat{T}\to \infty} {1\over \hat{T}} \log W(\hat{r},\hat{T}) \label{latwilexp} \end{equation} where $ W(\hat{r},\hat{T})$ denotes the expectation value of a Wilson loop with spatial and temporal extension $\hat{r}$ and $\hat{T}$ respectively. Assuming that $ W(\hat{r},\hat{T})= \exp\{ \hat{\sigma} \hat{r} \hat{T} -\hat{\delta} (\hat{r}+\hat{T}) +\hat{\gamma}\} $, the famous Creutz ratio \begin{equation} \chi(\hat{r},\hat{T}) =-\log \left({W(\hat{r},\hat{T}) W({\hat{r}} - 1,{\hat{T}}-1)\over W(\hat{r},{\hat{T}}-1) W({\hat{r}}-1,\hat{T})}\right) \end{equation} coincides with the string tension, since the parameters $\hat{\delta}$ and $\hat{\gamma}$ drop out. The original calculation (Creutz et al. (1982)) was performed on a $6^4$ lattice and the asymptotic scaling seemed to show up for values of $\beta$ slightly below 6.0. This measurement of a string tension different from zero from the strong coupling region to the asymptotic scale region, without any indication of a phase transition in the intermediate region, constituted the first {\it evidence} of quark confinement in QCD. In Fig. \ref{plpotlat} we show the most recent lattice measurement of the quenched static potential from Eq. (\ref{latwilexp}) on a hypercubic lattice $V= 16^4$ at $\beta=6.0$ and $V=32^4$ at $\beta=6.2$. These values correspond to inverse lattice spacings $a^{-1}\simeq 2.1$ GeV and $a^{-1}\simeq 2.9$ GeV respectively. The scale is adjusted to optimally reproduce the bottomonium level splittings (Bali et al. (1997)). The fit curve corresponds to the parameterization \begin{equation} V_0(r)= -{e\over r} +\sigma r + {f\over r^2} \label{paramm} \end{equation} which clearly confirms the Cornell potential (\ref{cornell}) (the $1/r^2$ correction, that accounts for the running of the coupling, is not meant to be physical but has been introduced to effectively parameterize the data within the given range of $r$). The parameters take the value: $e= 0.321$ and $\sigma= (468$ MeV$)^2$. The coefficient of the Coulomb term is quite far from the effective value for $\alpha_{\rm s}$ coming from the potential models\footnote{Part of this discrepancy can be traced back to the quenched approximation.}. Notice that the lattice potential becomes clearly linear around $0.2$ fm. Similar lattice measurements exist for the unquenched static potential (see e.g. Bali (1998)): the string is found to break down around a quark-antiquark distance of about 1.2 fm. The static potential in Fig. \ref{plpotlat} has been extracted as the ground state energy of the quark-antiquark configuration (cf. Eq. (\ref{fey})), which in turn corresponds to the lowest energy configuration of the ``glue'' between the quarks. Yet, also the excited gluonic modes have been measured on the lattice and the corresponding potentials have been adiabatically extracted. This should corresponds to the potential of heavy hybrids. For further details of this confirmation of the excited structure of the flux tube we refer the reader to Michael (1997), Morningstar et al. (1998). \begin{figure}[htb] \vskip -0.2truecm \makebox[1.0truecm]{\phantom b} \centerline{ \epsfxsize=10truecm \epsffile{act14.ps}} \vskip 0.5truecm \caption{\it Action density distribution between two static quarks in $SU(2)$ measured on a lattice $V=32^4$ at $\beta=2.5$. The physical quark-antiquark distance is 1.2 fm. Bali et al. (1995).} \label{plac} \end{figure} Lattice studies have been undertaken to probe the energy momentum tensor of the colour fields. The probe used is the gauge-invariant insertion of a plaquette in the presence of a static Wilson loop. Depending on the Lorentz orientation of the plaquette, this corresponds to the average of a (chromo)electric or a (chromo)magnetic field in the presence of quark sources. Lattice sum rules can be used to normalize these distributions and to relate them to the $\beta$ function. For separations $r> 0.7$ fm, a string-like spatial distribution is found with a transverse rms width increasing very slowly with $r$ and reaching a rather constant value between 1 and 2 fm. The physical value for this constant ranges between 0.5 and 0.75 fm. The averages of the squared components of the colour fields (which are gauge invariant quantities) are found to be roughly equal (i.e. $\langle E^2_i\rangle \sim \langle B^2_j\rangle$ in the presence of the Wilson loop, for $i,j=1,2,3$). This implies that the energy density is much smaller than the action density (in Euclidean space). (See Sec. 6.3 for a discussion.) The result for the action density is presented in Fig. \ref{plac} and is quite impressive! The formation of the interquark flux tube is evident. The interpretation of this phenomenon is the following. When the distance between the quarks becomes larger than some critical value (connected to $\Lambda_{QCD}$) the branching of the gluons, due to the non-Abelian nature of QCD, becomes so intensive that it makes no sense to speak about individual gluons. A coherent effect develops with the subsequent formation of the flux tube. It is conjectured that a specific organization of the QCD vacuum makes this kind of configuration energetically favorable (see Sec. 6). However, lattice QCD seems more suitable to ask ``what'' and not ``why''. Here, we have shown some model independent results on the quark interaction that provide evidence of quark confinement but do not explain {\it why} quarks are confined. Eventually, to get some insight in the quark confinement mechanism, it is necessary to build up and use some analytic models. The role of the lattice measurements will then be to validate these models. The combination of analytic and lattice techniques will give us some information on the nature of the nonperturbative quark interaction. Finally, we mention that there are a lot of measurements of hadron masses performed in lattice QCD as well as in lattice nonrelativistic QCD (NRQCD). However, since in these lectures we are more interested in the mechanism of confinement as well as in developing analytic approaches to it, we refer the reader to the literature (see e.g. Davies (1997)). \section{The Heavy Quark Interaction} In this part of the lectures we summarize some nonperturbative analytic results on the heavy quark interaction. With nonperturbative we mean the fact that they do not rely necessarily on a perturbative expansion in $g$. An analytic result of this type was already obtained for the static potential in Sec. 4 with Eq. (\ref{potfond}). Here, we want to establish a systematic procedure to obtain relativistic corrections. In order to take full advantage from lattice calculations, our analytic approach will be manifestly gauge-invariant. In the next sections we present some model independent results: 1) the effective theories of QCD that enable the performing of a systematic expansion of the heavy quarks dynamics in some small parameters which are nonperturbative and 2) the exact and physically transparent expression for the quark-antiquark interaction at order $1/m^2$. However, at the very moment we want to calculate the quark dynamics we have to resort to lattice evaluations or to models of the QCD vacuum (see Sec. 6). Nevertheless, the present approach allows us to gain something with respect to the standard lattice formulation. In fact, here the lattice simulations come in at an intermediate step for the evaluation of some definite expectation values of fields inserted in the static Wilson loop. These can be directly tested with the analytic (model dependent) results. From this comparison, in a process of model validation, we gain a deeper understanding of the confinement physics. A remark: we have performed all the calculations of the last section in Euclidean space since lattice simulations (as all numerical evaluations) are done in Euclidean space. We no longer have this restriction in the next sections where we come back to the Minkowski space. We use the following notation for the vacuum expectation value (to be compared with Eq. (\ref{expece})) \begin{equation} \langle f[A]\rangle \equiv {1\over Z} \int {\cal D} A f [A] e^{i \int d^4 x L_{YM}}. \label{expecm} \end{equation} In most of the cases it will be straightforward to switch from Euclidean to Minkowski space by simply using the relations (\ref{euclidep}). \subsection{Effective theories for heavy quarks} We have seen that the physics of heavy quark bound states is complicated by the interplay of different characteristic scales like the heavy quark mass $m$, the momentum of the bound state $mv$, the energy of the bound state $mv^2$ (and in principle also $\Lambda_{QCD}$), being $v$ the heavy quark velocity. These scales can be disentangled using effective theories of QCD. From the technical point of view, this simplifies considerably the calculation. From the conceptual point of view, this enables us to factorize the part of the interaction that we know (and we are able to calculate in perturbation theory) from the low energy part which is dominated by nonperturbative physics. Nonrelativistic QCD (NRQCD) is an effective theory equivalent to QCD and constructed integrating out the high energy ($E>m$) degrees of freedom, thus making explicit the mass parameter. NRQCD has been extensively discussed at this school by Peter Lepage. We recall only few points useful to prepare the developments of the next sections. The interested reader is referred to Lepage (1996) and Thacker et al. (1991). NRQCD was devised to be applied to lattice simulations, however, here we are interested in the definition of NRQCD in the continuum. In Sec. 4 we have considered the static limit of an infinitely massive quark. In that case the Dirac equation for the quark propagator in an external field is exactly solvable. Now, we want to calculate the subsequent corrections showing up when the quark mass is large but finite. In order to disentangle the dynamical scales it is convenient to use the heavy quark velocity, $v_Q\equiv v$, as an expansion parameter. Notice that, even if this seems to be analogous to what happens in QED where, e.g. in positronium, the expansion parameter is $v_e \simeq \alpha$, the quark velocity is also sensitive to the nonperturbative quark interaction and turns out to be a function of both $\alpha_{\rm s}$ and $\Lambda_{QCD}$ (or $\sigma$). Let us consider Eq. (\ref{direq}) (now in Minkowski space) or equivalently the Lagrangian \begin{equation} L= \bar{\psi} (i \gamma^\mu D_\mu -m) \psi\, . \label{lcri} \end{equation} Applying a Foldy--Wouthuysen transformation and expanding in the inverse of the quark mass $m$, we obtain\footnote{We neglect here operators involving more than 2 fermions. These are irrelevant to our discussion here, see Brambilla et al. (1998).} up to order $1/m^2$ \begin{eqnarray} L &=& \psi^{\dag} \left( m + iD_0 +\frac {{\bf D}^2} {2m} + c_1(m/\mu) \frac {({\bf D}^2)^2} {8 m^3} + c_2(m/\mu) \frac {g} {8 m^2} ({\bf D}\cdot {\bf E} - {\bf E}\cdot {\bf D}) \right.\nonumber\\ && \left. + i c_3(m/\mu) \frac {g} {8m^2} {\mbox{\boldmath $\sigma$}} \cdot ( {\bf D} \times {\bf E} - {\bf E} \times {\bf D}) + c_4(m/\mu) \frac {g} {2m} {\mbox{\boldmath $\sigma$}}\cdot {\bf B} + \dots\right)\psi \nonumber \\ & & + \hbox{antiquark terms} \label{nrqcd} \end{eqnarray} In expanding the QCD Lagrangian in the inverse of the mass we have lost explicit renormalizability. The NRQCD Lagrangian (\ref{nrqcd}) must be regularized. $\mu$ is an ultraviolet cut-off which restricts the momenta to the region $p \sim m v < \mu < m$. The effect of the excluded momenta, e.g. in gluon loops, is factorized in the (matching) coefficients which multiply the nonrelativistic operators. In Eq. (\ref{nrqcd}) we call them $c_i(m,\mu)$. Such coefficients are straightforwardly calculated in perturbation theory by imposing that scattering amplitudes evaluated with (\ref{nrqcd}) are equal to the same amplitudes evaluated in QCD order by order in $\alpha_{\rm s}$ and in $1/m$ (see Manohar (1997)). This procedure is called {\it matching}. The coefficients $c_i$ are normalized in such a way to be one (or zero) at tree-level. The terms in the Lagrangian can be ordered in powers of the squared velocity of the heavy quark using the power counting rules for momentum and kinetic energy of Thacker et al. (1991): ${\bf D} \sim p \sim m v$, $K \sim m v^2$. From the lowest order field equation $$ \left( i\partial_0 - g A_0 - \frac {{\bf D}^2} {2 m}\right) \psi = 0 $$ we get \begin{eqnarray*} g A _0 \sim \partial_0 \sim K = m v^2\\ g {\bf E} = [ D_0, {\bf D}] \sim pK = m^2v^3\\ -ig\epsilon_{ijk}B^{k} = [D_i, D_j] \sim K^2 = m^2v^4. \end{eqnarray*} Following these power-counting rules the number of operators to be included in $L$ can be truncated at a fixed order in $v^2$ depending on the precision we require on the calculation of the energy. This power counting is not exact. It accounts only for the leading binding contributions. The reason is that two scales, the soft one $\sim mv$ and the ultrasoft $\sim m v^2$, are still mixed up in the NRQCD Lagrangian. An exact power counting can be achieved by integrating out from NRQCD the soft scale with the same procedure as the mass scale was integrated out from QCD in order to get NRQCD. In this way one obtains a further effective theory, called potential nonrelativistic QCD (pNRQCD) (Pineda and Soto (1998)) where only dynamical ultrasoft degrees of freedom are present. These are the heavy quark bound states (explicitly projected in colour singlet and octet states) and gluons propagating at the ultrasoft scale. Since nonrelativistic potentials get contributions only from the soft scale, in the pNRQCD Lagrangian potential and nonpotential contributions are explicitly disentangled. More precisely, the pNRQCD Lagrangian density with leading nonpotential corrections is given by (Brambilla et al. (1999)): \begin{eqnarray} L &=& {\rm Tr} \left\{ S^\dagger \left( i\partial_0 - {{\bf p}^2\over m} + \sum_n {V^{(n)}_s({\bf r},{\bf p},\mbox{\boldmath $\sigma$}; \mu^\prime)\over m^n} \right) S \right\} \nonumber\\ & & + {\rm Tr} \left\{ O^\dagger \left( iD_0 - {{\bf p}^2\over m} + \sum_n {V^{(n)}_o({\bf r},{\bf p},\mbox{\boldmath $\sigma$}; \mu^\prime)\over m^n} \right) O \right\} \nonumber\\ & & + g V_A ({\bf r};\mu^\prime) {\rm Tr} \left\{ O^\dagger {\bf r} \cdot {\bf E} \,S + S^\dagger {\bf r} \cdot {\bf E} \,O \right\} \nonumber\\ & & + g {V_B ({\bf r};\mu^\prime) \over 2} {\rm Tr} \left\{ O^\dagger {\bf x} \cdot {\bf E} \,O + O^\dagger O {\bf r} \cdot {\bf E} \right\} \label{pnrqcd0} \end{eqnarray} where ${\bf r} \equiv {\bf x}_1-{\bf x}_2$ and ${\bf X} \equiv ({\bf x}_1+{\bf x}_2)/2$, $S = S({\bf r},{\bf X},t)$ and $O = O({\bf r},{\bf X},t)$ are the singlet and octet wave function respectively. All the gauge fields in Eq. (\ref {pnrqcd0}) are evaluated in ${\bf X}$ and $t$. In particular ${\bf E} \equiv {\bf E}({\bf X},t)$ and $iD_0 O \equiv i \partial_0 O - g [A_0({\bf X},t),O]$. The matching coefficients $V_A$ and $V_B$ are normalized in such a way that at the leading perturbative order they are one. $\mu^\prime$ is the ultrasoft cut-off corresponding to the regularization of the Lagrangian (\ref{pnrqcd0}): $m v$ $> \mu^\prime >$ $m v^2$. At the leading order in the soft scale the equations of motion associated with the pNRQCD Lagrangian (\ref{pnrqcd0}) are two decoupled nonrelativistic Schr\"odinger equations describing the propagation of a singlet and an octet bound state defined by the potentials $\sum V_s^{n}/m^n$ and $\sum V_o^{n}/m^n$ respectively. Next-to-leading nonpotential corrections involve the coupling of octet and singlet states via ultrasoft chromoelectric fields. We point out that: 1) The potentials have now the status of matching coefficients (in the matching from NRQCD to pNRQCD) and depend in general on the scale $\mu^\prime$ (see Sec. 4.2). They can be calculated comparing Wilson loop functions in NRQCD and singlet/octet propagators in pNRQCD. The novel feature is that the matching takes place now in the low energy region and thus it can be even nonperturbative. In this case the potentials are given by expressions to be evaluated on the lattice. Notice that the potentials contain in their definition also the NRQCD coefficients $c_i(m/\mu)$. 2) the pNRQCD Lagrangian contains both potential and nonpotential contributions. In some kinematic regions they coincide with the Voloshin (1979) and Leutwyler (1981) corrections. 3) Only the ultrasoft scale is left. Hence each term in (\ref{pnrqcd0}) has a definite power counting in $v$! In particular: $1/r \sim p \sim m v$, $V^{(n)}_{s,o} \sim m v^2$ and $F_{\mu\nu} \sim m^2 v^4$. In the lattice NRQCD approach the NRQCD Lagrangian of Eq. (\ref{nrqcd}) is discretized on the lattice and the heavy hadron masses are obtained by performing lattice simulations. These are considerably less time consuming than the traditional QCD lattice simulations because the cut-off scale $\mu$ (smaller than $m$) makes possible to use coarse lattices. Since in the following we will not deal explicitly with nonpotential contributions we will not perform explicitly the matching from NRQCD to pNRQCD. As far as ultrasoft degrees of freedom are not considered, this corresponds only to a choice of language. Actually, getting the heavy quark potential from the NRQCD Lagrangian, in any way one is doing it, if done properly, is nothing else than performing the leading order matching with the pNRQCD Lagrangian. In particular we will show how to obtain the spin and velocity dependent singlet potentials up to order $1/m^2$. A possible method is to calculate these as corrections to the static propagator from the Lagrangian (\ref{nrqcd}) (see Eichten et al. (1981), Tafelmayer (1986)). We will use a path integral formalism (Peskin (1983)). This approach has the advantage that the result, expressed in terms of deformed Wilson loops, is suitable to be used also in QCD vacuum models (see Sec. 6). One remark at the end. The matching coefficients $c_i$ contain all the fermionic loop contributions ~at the hard ~scale. If the matching ~between NRQCD and pNRQCD is nonperturbative, then the quark loop contributions at the low energy scale are contained in the functional measure of the object to be evaluated on the lattice (and it is indeed a hard task even for the lattice up to now). In the following section for simplicity we work in the quenched approximation. \subsection{The QCD spin-dependent and velocity-dependent potentials} Let us consider Eq. (\ref{nrqcd}), taking the matching coefficients at tree level\footnote{ The matching coefficients can be simply included in the calculation (see Chen et al. (1995), Bali et al. (1997)). They are needed for example to obtain the one-loop perturbative behaviour of the potentials, see Brambilla et al. (1998)).}. From Eq. (\ref{nrqcd}) it follows that the heavy quark (of mass $m_j$) propagator $K_j$ in external field (that now is reduced to a Pauli quark propagator, i.e. a $2\times 2$ matrix in the spin indices) satisfies the Schr\"odinger equation \begin{eqnarray} & &i \frac{\partial}{\partial x^0} K_j(x,y;A) = H_{{\rm FW}} K_j(x,y;A) \nonumber\\ & & \equiv \left[ m_j +\frac{1}{2m_j} ({\bf p}_j - g{\bf A})^2 - \frac{1}{8m_j^3} ({\bf p}_j - g{\bf A})^4 - \frac{g}{m_j} {\bf S}_j \cdot {\bf B} + gA^0 \right. \label{simpeq}\\ & & - \left. \frac{g}{8m_j^2} (\partial_i E^i - ig [A^i,E^i]) + \frac{g}{4m_j^2} \varepsilon^{ihk} S_j^k \{(p_j -gA)^i,E^h\} \right] K_j(x,y;A) \nonumber \end{eqnarray} with the Cauchy condition \begin{equation} K_j(x,y;A) |_{x^0=y^0} = \delta^3({\bf x}-{\bf y}) \label{incon} \end{equation} where $\varepsilon^{ihk}$ is the three-dimensional Ricci symbol. By standard techniques the solution of Eq. (\ref{simpeq}) (see Sakurai (1985) and Exercise 5.2.1) with the initial condition (\ref{incon}), can be expressed as a path integral in the phase space \begin{equation} K_j(x,y;A)= \int_{{\bf z}_j(y^0)={\bf y}}^{{\bf z}_j(x^0)={\bf x}} {\cal D} [{\bf z}_j, {\bf p}_j] \, {\rm T} \exp \left\{ i \int_{y^0}^{x^0} dt \, [{\bf p}_j \cdot \dot{{\bf z}}_j - H_{{\rm FW}}] \right\}. \label{pathkappa} \end{equation} Here, the time-ordering prescription T acts both on spin and gauge matrices. The trajectory of the quark $j$ in coordinate space is denoted by ${\bf z}_j = {\bf z}_j(t)$, the trajectory in momentum space by ${\bf p}_j = {\bf p}_j(t)$ and the spin by ${\bf S}_j$ (See Fig. \ref{plwil2}). Standard path integral manipulations on Eq. (\ref{pathkappa}) give \begin{eqnarray} K_j(x,y;A) \!\! &=& \!\! \int_{{\bf z}_j(y^0)={\bf y}}^{{\bf z}_j(x^0)={\bf x}}\!\!\!\! {\cal D} [{\bf z}_j,{\bf p}_j] \, {\rm T} \exp \left\{ i \int_{y^0}^{x^0} \!\! dt \, \left[ {\bf p}_j \cdot \dot{{\bf z}}_j - m_j - \frac{{\bf p}_j^2}{2m_j} + \frac{{\bf p}_j^4}{8m_j^3} \right. \right. \nonumber\\ & &\qquad - gA^0 + \frac{g}{m_j} {\bf S}_j \cdot {\bf B} + \frac{g}{2m_j^2} {\bf S}_j \cdot ({\bf p}_j \times {\bf E}) - \frac{g}{m_j} {\bf S}_j \cdot (\dot{{\bf z}}_j \times {\bf E}) \nonumber\\ & & \qquad \left. \left. + g \dot{{\bf z}}_j \cdot {\bf A} + \frac{g}{8m_j^2} (\partial_i E^i -ig[A^i,E^i]) \right] \right\} . \label{stran} \end{eqnarray} Inserting Eq. (\ref{stran}) into expression (\ref{utile}) of the quark-antiquark Green function and taking $x_1^0=x_2^0=t_{\rm f}$, $y_1^0=y_2^0=t_{\rm i}$ with $T\equiv t_{\rm f}-t_{\rm i} >0$, one obtains the two-particle Pauli-type propagator $K$ in the form of a path integral on the world lines of the two quarks \begin{eqnarray} K({\bf x}_1, {\bf x}_2, {\bf y}_1, {\bf y}_2;T)= \int_{{\bf z}_1(t_{\rm i})={\bf y}_1}^{{\bf z}_1(t_{\rm f})={\bf x}_1} {\cal D} [{\bf z}_1, {\bf p}_1] \int_{{\bf z}_2(t_{\rm i})={\bf y}_2}^{{\bf z}_2(t_{\rm f})={\bf x}_2} {\cal D} [{\bf z}_2, {\bf p}_2] \nonumber\\ \times \exp\left\{i\int_{t_{\rm i}}^{t_{\rm f}} dt\, \sum_{j=1}^2 \left[{\bf p}_j \cdot \dot{{\bf z}}_j -m_j- \frac{{\bf p}^2_j}{2m_j}+\frac{{\bf p}^4_j}{8m_j^3}\right] \right\} \nonumber\\ \times \left\langle \frac{1}{3} {\rm Tr \, T_s \, P} \exp\left\{ig\oint_{\Gamma} dz^{\mu} \, A_{\mu}(z) +\sum_{j=1}^2\frac{ig}{m_j} \int_{{\Gamma}_j} dz^{\mu} \right. \right. \nonumber\\ \left. \left. \times \left(S_j^l \hat{F}_{l{\mu}}(z) -\frac{1}{2m_j} S_j^l\varepsilon^{lkr}p_j^k F_{{\mu}r}(z)- \frac{1}{8m_j} D^{\nu}F_{{\nu}{\mu}}(z) \right) \right\} \right\rangle, \label{risfin} \end{eqnarray} where the dual tensor field is defined to be $\hat{F}^{\mu\nu} \equiv \varepsilon^{\mu \nu\rho\sigma} F_{\rho\sigma}/2$ Here ${\rm T_s}$ is the time-ordering prescription for spin matrices, P is the path-ordering prescription for gauge matrices along the loop $\Gamma$, $\Gamma_1$ denotes the path going from $(t_{\rm i}, {\bf y}_1)$ to $(t_{\rm f},{\bf x}_1)$ along the quark trajectory $(t,{\bf z}_1(t))$, $\Gamma_2$ the path going from $(t_{\rm f}, {\bf x}_2)$ to $(t_{\rm i}, {\bf y}_2)$ along the antiquark trajectory $(t, {\bf z}_2(t))$ and $\Gamma$ is the path made by $\Gamma_1$ and $\Gamma_2$ closed by the two straight lines joining $(t_{\rm i}, {\bf y}_2)$ with $(t_{\rm i}, {\bf y}_1)$ and $(t_{\rm f}, {\bf x}_1)$ with $(t_{\rm f}, {\bf x}_2)$ (see Fig. \ref{plwil2}). Finally Tr denotes the trace over the gauge matrices. Note that the right-hand side of (\ref{risfin}) is manifestly gauge invariant. Defining the angular bracket term in Eq. (\ref{risfin}) as (more rigorously this should correspond to the leading matching between NRQCD and pNRQCD) \begin{equation} \left\langle \frac{1}{3} {\rm Tr \, T_s \, P} \exp \ldots\right \rangle = {\rm T_s} \exp\left[ -i\int_{t_{\rm i}}^{t_{\rm f}} dt \, V_{{\rm Q}\bar{{\rm Q}}}({\bf z}_1,{\bf z}_2,{\bf p}_1,{\bf p}_2,{\bf S}_1,{\bf S}_2) \right]; \label{risfin2} \end{equation} we obtain that \begin{equation} i\frac{\partial}{\partial T} K= \left[ \sum_{j=1}^2 \left( m_j+ \frac{{\bf p}_j^2}{2m_j}-\frac{{\bf p}_j^4}{8m_j^3}\right) +V_{{\rm Q}\bar{{\rm Q}}} \right] K, \end{equation} where $V_{{\rm Q}\bar{{\rm Q}}}$ is {\it the complete QCD (quenched) quark-antiquark potential at the order $v^4$}. Expanding the logarithm of the left-hand side of (\ref{risfin2}) up to $1/m^2$, we find\footnote{ Notice that $\int_{\Gamma_j} dz^{\mu}f_{\mu}(z) = (-1)^{j+1} \int_{t_{\rm i}}^{t_{\rm f}} dt ( f_0(z_j) - {\dot{{\bf z}}}_j \cdot {\bf f} (z_j))$, where $z_j=(t,{\bf z}_j(t))$. The factor $(-1)^{j+1}$ accounts for the fact that world line $\Gamma_2 $ runs from $t_{\rm f}$ to $t_{\rm i}$. We also use the notation $z_j^{\prime}=(t^{\prime},{\bf z}_j(t^{\prime}))$.} \begin{eqnarray} \int_{t_{\rm i}}^{t_{\rm f}} dt \, V_{{\rm Q} \bar{{\rm Q}}} &=& i \log \langle W(\Gamma) \rangle - \sum_{j=1}^2 \frac{g}{m_j} \int_{{\Gamma}_j}dz^{\mu} \left( S_j^l \, \langle\!\langle \hat{F}_{l\mu}(z) \rangle\!\rangle \right. \nonumber\\ &~& \quad\quad -\frac{1}{2m_j} S_j^l \varepsilon^{lkr} p_j^k \, \langle\!\langle F_{\mu r}(z) \rangle\!\rangle - \left. \frac{1}{8m_j} \, \langle\!\langle D^{\nu} F_{\nu\mu}(z) \rangle\!\rangle \right) \nonumber\\ &~& \quad\quad - \frac{1}{2} \sum_{j,j^{\prime}=1}^2 \frac{ig^2}{m_jm_{j^{\prime}}} {\rm T_s} \int_{{\Gamma}_j} dx^{\mu} \, \int_{{\Gamma}_{j^{\prime}}} dx^{\prime\sigma} \, S_j^l \, S_{j^{\prime}}^k \nonumber\\ &~& \quad\quad \times \left( \, \langle\!\langle \hat{F}_{l \mu}(z) \hat{F}_{k \sigma}(z^{\prime}) \rangle\!\rangle - \, \langle\!\langle \hat{F}_{l \mu}(z) \rangle\!\rangle \, \langle\!\langle \hat{F}_{k \sigma}(z^{\prime}) \rangle\!\rangle \right), \label{potential} \end{eqnarray} where we recall that $$ \langle f(A) \rangle \equiv {1\over 3}{\rm Tr \>}{\rm P \>} {\displaystyle\int {\cal D} A \, e^{iS_{\rm YM} (A)} f(A) \over \displaystyle\int {\cal D} A\, e^{iS_{\rm YM} (A)}} $$ and we have introduced the vacuum expectation value in presence of quarks (i.e. of the Wilson loop) $$ \langle\!\langle f(A) \rangle\!\rangle \equiv{\displaystyle \int {\cal D} A \, e^{iS_{\rm YM} (A)} {\rm Tr \>}{\rm P \>} f(A) \exp \left[i g \displaystyle\oint_{\Gamma} dz^\mu A_\mu (z) \right] \over \displaystyle \int {\cal D} A\, e^{iS_{\rm YM} (A)} {\rm Tr \>}{\rm P \>} {\exp \left[i g \displaystyle\oint_{\Gamma} dz^\mu A_\mu (z) \right] } }. $$ In this way one obtains from QCD the static, spin-dependent and velocity dependent terms that control the quarkonium spectrum and that were introduced on a pure phenomenological basis in Sec. 3, cf Eqs. (\ref{vfenum})--(\ref{vvel}): \begin{equation} V_{{\rm Q} \bar {\rm Q}} = V_0 + V_{\rm VD} + V_{\rm SD} \> . \end{equation} These terms have a physical direct interpretation. The spin independent part of the potential, $V_0 + V_{\rm VD}$, is obtained in (\ref{potential}) from the expansion of $\log \langle W(\Gamma) \rangle $ for small velocities ${\dot{\bf z}}_1(t) = {\bf p}_1/m_1$ and ${\dot{\bf z}}_2(t) = {\bf p}_2/m_2$: \begin{equation} i \log \langle W(\Gamma) \rangle = \int_{t_{i}}^{t_{f}} dt \, \left (V_0 (r(t)) + V_{\rm VD}({\bf r}(t)) \right ) , \label{willexpp} \end{equation} where $V_0$ is the static part and ${\bf r}(t) \equiv {\bf z}_1(t) - {\bf z}_2 (t)$. The spin-dependent part, $V_{\rm SD}$, contains for each quark terms analogous to those one would obtained by making a Foldy--Wouthuysen transformation of a Dirac equation in an external field $\langle\!\langle F_{\mu \nu} \rangle\!\rangle$, along with an additional term $V_{\rm SS}$ having the structure of a spin-spin interaction. Therefore we can write \begin{equation} V_{\rm SD} = V_{\rm LS}^{\rm MAG} + V_{\rm Thomas} + V_{\rm Darwin} + V_{\rm SS} \label{vsd} \end{equation} using a notation which indicates the physical significance of the individual terms (MAG denotes Magnetic). The correspondence between (\ref{vsd}) and (\ref{potential}) is given by \begin{eqnarray} \int_{t_{\rm i}}^{t_{\rm f}} dt V_{\rm LS}^{\rm MAG} \!\!&=& \!\!- \sum_{j=1}^2 \frac{g}{m_j} \int_{{\Gamma}_j}\!dz^{\mu} S_j^l \, \langle\!\langle \hat{F}_{l\mu}(z) \rangle\!\rangle \>, \label{vmag} \\ \int_{t_{\rm i}}^{t_{\rm f}} dt V_{\rm Thomas} &=& \!\! \sum_{j=1}^2 \frac{g}{2 m^2_j} \int_{{\Gamma}_j}\!dz^{\mu} S_j^l \varepsilon^{lkr} p_j^k \, \langle\!\langle F_{\mu r}(z) \rangle\!\rangle \>, \label{vthomas} \\ \int_{t_{\rm i}}^{t_{\rm f}} dt V_{\rm Darwin} &=& \!\! \sum_{j=1}^2 \frac{g}{8 m^2_j} \int_{{\Gamma}_j}\!dz^{\mu} \langle\!\langle D^{\nu} F_{\nu\mu}(z) \rangle\!\rangle \>, \label{vdarwin} \\ \int_{t_{\rm i}}^{t_{\rm f}} dt V_{\rm SS} &=& \!\! - \frac{1}{2} \sum_{j,j^{\prime}} \frac{ig^2}{m_jm_{j^{\prime}}} {\rm T_s} \int_{{\Gamma}_j} \!dz^{\mu} \, \int_{{\Gamma}_{j^{\prime}}} \!dz^{\prime\sigma} \, S_j^l \, S_{j^{\prime}}^k \left( \langle\!\langle \hat{F}_{l \mu}(z) \hat{F}_{k \sigma}(z^{\prime})\rangle\!\rangle \right. \nonumber\\ &~& \quad\quad\quad\quad\quad\quad \left. - \langle\!\langle \hat{F}_{l \mu}(z) \rangle\!\rangle \, \langle\!\langle \hat{F}_{k \sigma}(z^{\prime}) \rangle\!\rangle \right) \> . \label{vss} \end{eqnarray} If we had worked in QED, we would have obtained the same formal result. The point is that in QED one can calculate perturbatively the field strength expectation values in the presence of the Wilson loop, here one can rely on a perturbative calculation only for very short interquark distances, shorter than the typical radius of the bound system. Therefore, we have to obtain a nonperturbative evaluation of $\langle \!\langle F \rangle \!\rangle $ and $\langle \!\langle FF\rangle\! \rangle$ that, together with the Wilson loop contain, all the relevant information on the heavy quark dynamics. Notice that the expression for the potential contains only manifestly gauge-invariant quantities. We have at our disposal two ways of obtaining the nonperturbative quark interaction. First, we can exactly translate all our results in terms of field strength expectation values in presence of a {\it static} Wilson loop. These are plaquette insertions in the static Wilson loop and have been evaluated on the lattice (Bali et al. (1997)). The interesting fact is that we can also perform an analytic evaluation making only an assumption on the nonperturbative behaviour of the Wilson loop. This is due to the fact that all the expectation values of Eq. (\ref{vmag})-(\ref{vss}) can be obtained as functional derivatives of $\log \langle W(\Gamma) \rangle $ with respect to the path, i.e. with respect to the quark trajectories ${\bf z}_1 (t)$ or ${\bf z}_2 (t)$. In fact let us consider the change in $\langle W(\Gamma) \rangle$ induced by letting $ z_j^\mu (t) \rightarrow z_j^\mu (t) + \delta z_j^\mu (t)$ where $\delta z_j^\mu (t_{\rm i}) = \delta z_j^\mu (t_{\rm f}) = 0$, then we have \begin{equation} g \langle\!\langle F_{\mu\nu}(z_j) \rangle\!\rangle = (-1)^{j+1} {\delta i \log \langle W(\Gamma) \rangle \over \delta S^{\mu\nu} (z_j)} , \label{e20} \end{equation} $$ \delta S^{\mu\nu} (z_j) = dz_j^\mu \delta z_j^\nu - dz_j^\nu \delta z_j^\mu , $$ and varying again the path \begin{equation} g^2 \left(\langle\!\langle F_{\mu\nu}(z_1) F_{\lambda\rho}(z_2) \rangle\!\rangle - \langle\!\langle F_{\mu\nu}(z_1) \rangle\!\rangle \langle\!\langle F_{\lambda\rho}(z_2) \rangle\!\rangle \right) = - i g {\delta\over \delta S^{\lambda\rho}(z_2)} \langle\!\langle F_{\mu\nu}(z_1) \rangle\!\rangle. \label{e21} \end{equation} {\it Therefore, to obtain the whole quark-antiquark potential no other assumptions are needed than the behaviour of $\langle W(\Gamma) \rangle$. In particular all contributions to the spin dependent part of the potential can be expressed as first and second variational derivatives of $\log \langle W(\Gamma) \rangle$. The obtained expressions are correct up to order $v^4$. Higher order corrections can in principle be included systematically in the same way.} We can use this result as a laboratory to understand confinement. In fact any assumption on the QCD vacuum, i.e. on the nonperturbative behaviour of the Wilson loop, is put in direct connection, on one hand with the lattice evaluation and on the other hand with the phenomenological data. We address this issue in Sec. 6. We conclude this section presenting one of the most common representations of the $1/m^2$ potentials (Eichten et al. (1981) and Barchielli et al. (1988)) which we will use in Sec. 6: \begin{eqnarray} V_{\rm SD} &=& {1\over 8} \left( {1\over m_1^2} + {1\over m_2^2} \right) \Delta \left[ V_0(r) +V_{\rm a}(r) \right] \nonumber\\ &+& \left( {1\over 2 m_1^2} {\bf L}_1 \cdot {\bf S}_1 - {1\over 2 m_2^2} {\bf L}_2 \cdot {\bf S}_2 \right) {1\over r} {d \over dr} \left[ V_0(r)+ 2 V_1(r) \right] \nonumber \\ &+& {1\over m_1 m_2}\left( {\bf L}_1 \cdot {\bf S}_2 - {\bf L}_2 \cdot {\bf S}_1 \right) {1\over r} {d \over dr} V_2(r) \nonumber\\ &+&{1\over m_1 m_2} \left( { {\bf S}_1\cdot{\bf r} \> {\bf S}_2\cdot{\bf r}\over r^2} - {{\bf S}_1 \cdot {\bf S}_2 \over 3} \right) V_3(r) + {1\over 3 m_1 m_2} {\bf S}_1 \cdot {\bf S}_2 \, V_4(r) \label{sd} \end{eqnarray} with ${\bf L}_j = {\bf r} \times {\bf p}_j$ and \begin{eqnarray} V_{\rm VD} &=& \sum_{j=1}^2 {V_{\prime} (r)\over m_j} \nonumber\\ &+& {1\over m_1 m_2} \bigg\{ V_{\prime\prime}(r) + {\bf p}_1\cdot{\bf p}_2 V_{\rm b}(r) + \left( {{\bf p}_1\cdot{\bf p}_2\over 3} - {{\bf p}_1\cdot {\bf r} \>~ {\bf p}_2 \cdot {\bf r} \over r^2}\right) V_{\rm c}(r) \bigg\} \nonumber\\ &+& \sum_{j=1}^2 {1\over m_j^2}\left\{ V_{\prime\prime\prime}(r) + p^2_j V_{\rm d}(r) + \left( { p^2_j\over 3} - {{\bf p}_j\cdot {\bf r} \>~ {\bf p}_j \cdot {\bf r} \over r^2}\right) V_{\rm e}(r) \right\} \, . \label{vd} \end{eqnarray} The brackets $\{ \cdots \}$ mean here an ordering prescription between position and momentum operators. The functions $V_i(r)$ contain all the dynamics and are given by expectation values of electric and magnetic field insertion in the static Wilson loop. Explicit expressions can be found in Eichten et al. (1981), Gromes (1984), Barchielli et al. (1988), (1990), Brambilla et al. (1990), (1993), (1997a). \vskip 1truecm \leftline{\bf Exercises} \begin{itemize} \item[5.2.1]{ Consider a free particle in one dimension. Using Eq. (\ref{pathkappa}) calculate the free particle propagator. [Hint: use the discretized version of the path integral. See e.g. Sakurai (1985) for details].} \item[5.2.2]{ Demonstrate Eqs. (\ref{e20}) and (\ref{e21}) first in QED and then in QCD.} \end{itemize} \section{Modelling the QCD vacuum} The limitations of a perturbative approach are appreciated when considering the vacuum. In perturbation theory the vacuum is approximated as an empty state with rare quark or gluon loop fluctuations. This in turn means that quarks and gluons are allowed to propagate freely. We have seen that experimentally this is not the case. The true, nonperturbative vacuum could be better imagined as a disordered medium with whirlpools of colour on different scales, thus densely populated by fluctuating fields whose amplitude is so large that they cannot be described by perturbation theory. Such a vacuum would be responsible for the fact that quark and gluons are confined. Such a vacuum would be responsible for the area law behaviour of the Wilson loop. It is established that the QCD vacuum is (phenomenologically) characterized by various nonperturbative condensates, for a recent review see Shifman (1998). (An introduction to the topic has been given at this school by Anatoly Radyushkin.) Half-dozen of them are known: the gluon condensate $F_2 \equiv\langle {\alpha_s\over \pi}F_{\mu\nu}^a(0) F^a_{\mu\nu}(0) \rangle $, the quark condensate $\langle \bar{q} q\rangle$, the mixed condensate $\langle \bar{q} \sigma_{\mu\nu} F_{\mu\nu}q\rangle $ and so on. Physically, the gluon condensate measures the vacuum energy density $\varepsilon_{vac}$. Indeed, due to the scale anomaly of QCD, the trace of the energy-momentum tensor is given by \begin{equation} \theta_\mu^\mu = {\beta(\alpha_s) \over 2 \alpha_s} F_{\mu\nu}^a(0) F^a_{\mu\nu}(0). \label{tens} \end{equation} Then, in the lowest order expansion of the beta function, we get \begin{equation} \varepsilon_{vac} \equiv {1\over 4} \langle \theta_\mu^\mu \rangle \simeq -{\beta_0\over 32} \langle {\alpha_s\over \pi} F_{\mu\nu}^a(0) F^a_{\mu\nu}(0) \rangle . \label{vac} \end{equation} Therefore, the nonperturbative gluon condensate shifts the vacuum energy downwards, making it advantageous. We remark that the negative sign in (\ref{vac}) is due to the asymptotic freedom! The vacuum fields fluctuate and these fluctuations contribute to the vacuum energy density. High energy modes of the fluctuating fields are in the weak coupling regime and can be dealt with perturbation theory as usual. The low frequency modes are responsible for the peculiar properties of the vacuum medium. In the sum rule approach of Shifman, Vainshtein and Zakharov (1979) the effects caused by the vacuum fields are parameterized into few local vacuum condensates. This approach describes quite successfully the low-lying hadrons where the QCD string is not so relevant. The underlying assumption is that the characteristic frequencies of the valence quarks in the bound state are larger than the characteristic scale parameter of the vacuum medium. In other words, the valence quark pair injected in the vacuum is assumed to perturb it only slightly. However, sum rules cannot tell us anything about confinement. This is due to the fact that there is no local order parameter for confinement. But, as discussed in particular in Sec. 4.4, confinement manifests itself in the area law of the Wilson loop which is a nonlocal quantity\footnote{ We remark that instead the local quark condensate is the order parameter of chiral symmetry breaking.}. In general {\it supplying an analytic form for the nonperturbative Wilson loop average amounts to defining a model of the QCD vacuum.} In the previous section we have seen that the static and the semirelativistic quark-antiquark interaction can be expressed in terms of the Wilson loop only. Therefore, it is tempting to explore the confinement mechanism using the simplest case of the heavy quark interaction and building analytic models of the QCD vacuum to be tested directly on the lattice and on the phenomenology. In Sec. 6.1 we present a pedagogical model, the so-called Minimal Area Law model, which takes seriously the lattice results and assumes that the logarithm of the Wilson loop is simply given by a constant (the string tension) times the minimal area enclosed by the Wilson loop. This model will turn out to be very instructive for three reasons: 1) it gives the feeling of how a real calculation is done in practice, 2) it gives at the end a concrete and non-trivial form for the QCD potential, 3) it shows how the flux tube degrees of freedom emerge. In Sec. 6.2 we briefly summarize the main points underlying the issue of the dual Meissner effect as the mechanism of confinement. In order to give a concrete idea of how this mechanism could lead to quark confinement, we present the ordinary Abelian Higgs model of superconductivity. We also briefly explain how a mechanism similar to the Abelian Higgs model arises in QCD using the 't Hooft Abelian projection idea. We present some lattice results obtained in Abelian projection on the interquark flux tube distribution and on the static potential. In Sec. 6.3 we discuss the flux tube structure at the light of some general low energy theorems. Finally in Sec. 6.4 we only summarize the main ingredients of some other analytic models of the QCD vacuum, in particular Dual QCD and the Stochastic Vacuum Model. \subsection{Minimal Area Law Model (MAL)} In this model (Brambilla et al. (1993)) $\langle W(\Gamma) \rangle$ is approximated by the sum of a short range part given at the leading order by the perturbative gluon propagator $D_{\mu\nu}$ and a long-range part given by the value of the minimal area enclosed by the deformed Wilson loop of fixed contour $\Gamma$ (see Fig. \ref{plwil2}) plus a perimeter contribution $\cal P$: \begin{eqnarray} i \log \langle W (\Gamma) \rangle &=& i\log \langle W (\Gamma) \rangle^{\rm SR} + i \log \langle W (\Gamma) \rangle^{\rm LR} \nonumber\\ &=& - \frac{4}{3} g^2 \oint_{\Gamma}dx^{\mu}_1 \oint_{\Gamma} dx^{\nu}_2 ~iD_{\mu \nu} (x_1-x_2) + \sigma S_{\rm min} + {C\over 2} {\cal P}. \label{mal} \end{eqnarray} Denoting by $u^{\mu}=u^{\mu}(s,t)$ the equation of a typical surface of contour $\Gamma$ ($s \in [0,1],\, t \in [t_{\rm i},t_{\rm f}], \, u^0(s,t)=t, \, {\bf u}(1,t)= {\bf z}_1(t), \, {\bf u}(0,t)= {\bf z}_2(t) \,$) and defining ${\bf u}_{\rm T} \equiv {\bf u} - ({\bf u}\cdot {\bf n})~{\bf n}$ with ${\bf n} = (\partial {\bf u}/\partial s) |\partial {\bf u}/\partial s|^{-1}$, we can write: \begin{eqnarray} S_{\rm min} & = & \min \int_{t_{\rm i}}^{t_{\rm f}} dt \, \int_0^1 ds\, \left[-\left( \frac{\partial u^{\mu}}{\partial t} \frac{\partial u_{\mu}} {\partial t} \right) \left( \frac{\partial u^{\mu}}{\partial s} \frac{\partial u_{\mu}}{\partial s} \right) + \left( \frac{\partial u^{\mu}}{\partial t} \frac{\partial u_{\mu}} {\partial s} \right)^2 \right]^{\frac{1}{2}} \nonumber\\ & = & \min \int_{t_{\rm i}}^{t_{\rm f}} dt \, \int_0^1 ds \, \left|\frac{\partial {\bf u}}{\partial s } \right| \left\{ 1-\left[ \left(\frac{\partial {\bf u}}{\partial t} \right)_{\rm T} \right]^2 \right\}^{\frac{1}{2}}, \end{eqnarray} which coincides with the Nambu--Goto action. Up to the relative order $1/m^2$ ($v^2$ in the velocity) the minimal surface can be identified exactly with the surface spanned by the straight-line joining $(t,{\bf z}_1(t))$ to $(t,{\bf z}_2(t))$ with $t_{\rm i} \le t \le t_{\rm f}$. The generic point of this surface is \begin{equation} u^0_{\min}=t \quad \quad \quad {\bf u}_{\min} = s~{\bf z}_1(t) + (1-s)~ {\bf z}_2(t) \>, \label{straight} \end{equation} with $0\leq s \leq 1$ and ${\bf z}_1(t)$ and ${\bf z}_2(t)$ being the positions of the quark and the antiquark at the time $t$. Then, the expression for the minimal area at order $1 /m^2$ in the MAL turns out to be \begin{eqnarray} S_{\min} &=& \int_{t_{\rm i}}^{t_{\rm f}} dt \, r \int_0^1 ds \, [1-(s~\dot{{\bf z}}_{1 \rm T} + (1-s)~\dot{{\bf z}}_{2 \rm T} )^2]^{\frac{1}{2}} \nonumber\\ &=& \int_{t_{\rm i}}^{t_{\rm f}} dt \, r \, \left[ 1-\frac{1}{6} \left(\dot{{\bf z}}_{1 \rm T}^2 + \dot{{\bf z}}_{2 \rm T}^2+ \dot{{\bf z}}_{1 \rm T} \cdot \dot{{\bf z}}_{2 \rm T} \right) + \cdots \, \, \right], \end{eqnarray} where $r=|{\bf z}_1- {\bf z}_2|$. The perimeter term is given by \begin{equation} {\cal P} = \vert {\bf x}_1 - {\bf x}_2 \vert + \vert {\bf y}_1 - {\bf y}_2 \vert + \sum_{j=1}^2 \int_{t_{\rm i}}^{t_{\rm f}} dt \sqrt {\dot{z}_j^\mu \dot{z}_{j\mu}} \>. \label{per} \end {equation} In the limit of large time interval $t_{\rm f} - t_{\rm i}$ the first two terms in the right-hand side of Eq. (\ref{per}) can be neglected. By expanding also Eq. (\ref{per}) up to $1/m^2$, we obtain \begin{eqnarray} i \log \langle W (\Gamma) \rangle^{\rm LR} &=& \int_{t_{\rm i}}^{t_{\rm f}} dt \, \sigma r \, \left[ 1-\frac{1}{6} \left(\dot{{\bf z}}_{1 \rm T}^2 + \dot{{\bf z}}_{2 \rm T}^2 + \dot{{\bf z}}_{1 \rm T} \cdot \dot{{\bf z}}_{2 \rm T} \right) \right] \nonumber\\ &+& {C\over 2} \sum_{j=1}^2 \int_{t_{\rm i}}^{t_{\rm f}} dt \left( 1-{1\over 2} \dot{\bf z}_j\cdot \dot{\bf z}_j \right). \end{eqnarray} For what concerns the perturbative part in the limit of large $t_{\rm f} - t_{\rm i}$ the only non-vanishing contribution to the Wilson loop is given by \begin{equation} i\log \langle W (\Gamma) \rangle^{\rm SR} = - \frac{4}{3} g^2 \int_{t_{\rm i}}^{t_{\rm f}} dt_1 \int_{t_{\rm i}}^{t_{\rm f}} dt_2 ~{\dot z}_1^\mu(t_1) ~ {\dot z}_2^\nu(t_2) ~iD_{\mu \nu} (z_1-z_2). \label{sr} \end{equation} In the infinite time limit this expression is still gauge invariant. Expanding $z_2(t_2)$ around $t_1$ it is possible to evaluate explicitly from Eq. (\ref{sr}) the short-range potential up to a given order in the inverse of the mass (this was the task of Exercise 4.2.3). Self-energy terms are neglected. Eventually, in the MAL model the following static and velocity dependent potentials are obtained: \begin{eqnarray} V_0 &=& -{4\over 3} {\alpha_{\rm s} \over r} + \sigma r + C \>, \label{v0mal}\\ V_{\rm b}(r)&=&{8\over 9} {\alpha_{\rm s} \over r} - {1\over 9}\sigma r \>, \quad \quad \,~ V_{\rm c}(r) = -{2\over 3} {\alpha_{\rm s} \over r} - {1\over 6}\sigma r \>, \nonumber \\ V_{\rm d}(r)&=& -{1\over 9} \sigma r -{1\over 4} C \>, \quad \quad V_{\rm e}(r) = -{1 \over 6}\sigma r \>, \label{vdmal} \end{eqnarray} which fulfill the relations (Barchielli et al. (1990)) \begin{eqnarray} & & V_{\rm d}(r) +{1\over 2} V_{\rm b}(r) +{1\over 4} V_0(r) - {r\over 12} {d V_0(r)\over dr}=0 , \label{relvel1}\\ & &V_{\rm e}(r) +{1\over 2} V_{\rm c}(r) + {r\over 4} {dV_0(r)\over dr}=0 . \label{relvel2} \end{eqnarray} By evaluating the functional derivatives of the Wilson loop, as given by Eqs. (\ref{e20})-(\ref{e21}), we obtain also the spin-dependent potentials \begin{eqnarray} \Delta V_{\rm a}(r) &=& 0, \quad {d\over dr} V_1(r) = -\sigma, \quad {d\over dr} V_2(r) = {4\over 3} {\alpha_{\rm s} \over r^2}, \nonumber\\ V_3(r) &=& 4 {\alpha_{\rm s}\over r^3}, \quad V_4(r) = {32\over 3}\pi \alpha_{\rm s} \delta^3({\bf r}). \label{vsmal} \end{eqnarray} These potentials reproduce the Eichten--Feinberg--Gromes results (Eichten and Feinberg (1981), Gromes (1984)) and fulfill the Gromes relation \begin{equation} {d\over dr} \left[ V_0(r) +V_1(r)-V_2(r) \right] = 0 . \label{grom} \end{equation} Notice that, as a consequence of the vanishing of the long-range behaviour of the spin-spin potential $V_{\rm SS}$ and of the spin-orbit magnetic potential $V_{\rm LS}^{\rm MAG}$ in this model, there is no long-range contribution to $V_2$, $V_3$ and $V_4$. Furthermore, $V_1$ has only a nonperturbative long-range contribution, which comes from the Thomas precession potential (\ref{vthomas}). The MAL model strictly corresponds to the Buchm\"uller picture (see Fig. \ref{pltube}) (Buchm\"uller (1982)) where the magnetic field in the comoving system is taken to be equal to zero. Let us first notice that the perimeter contributions at the $1 /m^2$ order can be simply absorbed in a redefinition of the quark masses $m_j \to m_j + C/2$. Then, let us consider the moving quark and antiquark connected by a chromoelectric flux tube and let us describe the flux tube as a string with transverse velocity ${\bf v}_{\rm T}$. At the classical relativistic level the system is described by the flux tube Lagrangian (Olsson et al. (1993)) \begin{equation} L = - \sum_{j=1}^2 m_j \sqrt{1- {\bf v}_j^2} - \sigma \int_0^r dr^\prime \sqrt{1- {\bf v}_{\rm T}^{\prime 2}} \>, \end{equation} with $ {\bf v}_{\rm T}^\prime = {\bf v}_{1{\rm T}}~ {r^\prime / r} + {\bf v}_{2{\rm T}} (1- {r^\prime / r} ), \quad 0< r^\prime < r $. The semirelativistic limit of this Lagrangian gives back the nonperturbative part of the $V_0$ and $V_{\rm VD}$ potentials in the MAL model (\ref{v0mal}) (\ref{vdmal}) (notice that the minimal area law in the straight-line approximation is the configuration given by a straight flux tube). The remarkable characteristics of the obtained $V_{\rm VD}$ potential is the fact that it is proportional to the square of the angular momentum and so takes into account the energy and the angular momentum of the string \begin{equation} V_{\rm VD}^{\rm LR} = - {1\over 12 m_1 m_2} {\sigma \over r} ({\bf L}_1 \cdot {\bf L}_2 + {\bf L}_2 \cdot {\bf L}_1 ) - \sum_{j=1}^2 { 1\over 6 m_j^2 } {\sigma \over r} {\bf L}_j^2 \>. \label{vdflux} \end{equation} Finally, the nonperturbative spin-dependent part of the potential in this intuitive flux tube picture simply comes from the Buchm\"uller ansatz that the chromomagnetic field is zero in the comoving framework of the flux tube. We notice that even if $V_1$ seems to arise from an effective Bethe--Salpeter kernel which is a scalar and depends only on the momentum transfer, a simple convolution kernel (see Eq. (\ref{kernel})) cannot reproduce the correct velocity dependent potential (\ref{vdflux}) or equivalently (\ref{vdmal}). Nevertheless the behaviour (\ref{vdflux}) is important to reproduce the spectrum\footnote{The extension of the Wilson loop approach to the relativistic treatment of heavy-light bound states supports the fact that the nonperturbative kernel {\it is not} a scalar convolution kernel, cf. Brambilla (1998) and Brambilla et al. (1997b).}. The static, spin-dependent and velocity-dependent potentials have been recently evaluated on the lattice (Bali et al. (1997)) and, at the present level of accuracy, confirm the prediction of this simple model. \subsection{Dual Meissner effect: a simple example} In the previous sections we have seen that the main characteristic of the quark nonperturbative interaction is the chromoelectric flux tube formation (see Fig. \ref{plac}). The formation of a flux tube is reminiscent of the Meissner effect in (ordinary) superconductivity (for a review see Weinberg (1986)). As it is well-known, the superconducting media do not tolerate the magnetic field. If one imposes a certain flux of magnetic field through such a medium, the magnetic field will be squeezed into a thin tube carrying all the magnetic flux. The superconducting phase is destroyed inside the tube. A well-known example of the string-like solution of the classical equations of motion is given by the Abrikosov string (Abrikosov (1957)). Superconductivity is caused by condensation of the Cooper pairs (pairs of electric charges). If there were monopoles and antimonopoles in the superconducting medium, then a string would be formed and confine them. Therefore, to explain the confinement of electric charges (the quarks), we need a condensate of (chromo)magnetic monopoles (see Fig. \ref{pldual}). This is the simple qualitative idea suggested by Nambu (1974) and 't Hooft (1976) and Mandelstam (1976). Then, the vacuum of QCD behaves like a dual superconductor, where the word ``dual'' here means that the role of electric and magnetic quantities is interchanged with respect to an ordinary superconductor (see Fig. \ref{pldual}). In the next section we present a concrete model for a superconductor. \begin{figure}[htb] \vskip -0.2truecm \makebox[1.0truecm]{\phantom b} \centerline{\epsfxsize=8truecm\epsffile{dual1.eps}} \vskip 0.2truecm \caption{\it The QCD vacuum as a ``dual'' superconductor .} \label{pldual} \end{figure} \subsubsection{Abelian Higgs Model} A relativistic version of the Ginzburg--Landau model for superconductivity is the Abelian Higgs model: \begin{equation} L = -\frac{1}{4} F^{\mu\nu} F_{\mu\nu} + {1\over 2}(D^\mu\Phi)^\dagger(D_\mu\Phi) - U(\Phi) \label{eq:2.1} \end{equation} where $\Phi$ is the Higgs field, in this case a complex (charged) scalar field describing Cooper pairs (i.e. condensate of electrons pairs in a lattice of positive ions, a superconducting solid) of charge $q= 2 e$, $e$ is the electron charge. $D_\mu\Phi = (\partial_\mu - i q A_\mu)\Phi$ is the covariant derivative, and $U(\Phi) = \displaystyle\frac{\lambda^\prime}{4} (\Phi^\dagger\Phi -\mu^2)^2$ is the Higgs potential. In Nielsen and Olesen (1973) it was shown that the Lagrangian (\ref{eq:2.1}) allows for vortex-line solutions. These vortex-lines solutions were approximately identified with the Nambu string. The Lagrangian $L$ is invariant under local $U(1)$ transformations. However, if $\mu^2 >0$, there is a condensation of Cooper pairs signalled by the vacuum expectation value $\langle\Phi\rangle\neq 0$. This corresponds to the spontaneous breaking of the U(1) electric symmetry\footnote{This is often referred as spontaneous breaking of the local gauge symmetry. However, this statement is somewhat inaccurate. In this kind of theories the vacuum {\it does not} break local gauge invariance. Any state in the Hilbert space that fails to be invariant under local gauge transformation is an unphysical state. The vacuum is entirely gauge invariant at variance with what happens in theories with a global symmetry. See 't Hooft (1994).}. As a consequence the photon becomes massive inside a superconducting body and an external magnetic field can penetrate the body only to a finite depth $\lambda$ equal to the inverse of the photon mass (Meissner effect). In fact putting $\Phi = \rho e^{i\theta}$, $\rho>0$, and $\tilde A_\mu = (A_\mu - \partial_\mu\theta)$, $L$ can be rewritten in the gauge-invariant form \begin{equation} L = -\frac{1}{4} F^{\mu\nu} F_{\mu\nu} + \frac{M^2}{2}\tilde A^\mu\tilde A_\mu + {\tilde{L}} [\rho] \label{eq:2.4} \end{equation} where ${\tilde{L}}$ is the sector of the Lagrangian describing the propagation of the massive field $\rho$ (mass $M_\phi$). The equations of motion for the electromagnetic field read \begin{equation} \partial^\mu F_{\mu\nu} + M^2 \tilde A_\nu = 0 \label{eq:2.5} \end{equation} with $M^2 = q^2 \langle\Phi\rangle^2$, the mass acquired by the photon. In a stationary state with no charges $A_0=0$, $\partial_0 {\bf A }= 0$, Eq. (\ref{eq:2.5}) gives $({\bf H} = \nabla\wedge{\bf A})$ \begin{eqnarray} \nabla\wedge{\bf H} + M^2 {\tilde {\bf A}} &=& 0 \label{eq:2.6a}\\ \nabla^2 {\bf H} + M^2{\bf H} &=& 0 \label{eq:2.6b} \end{eqnarray} Eq. (\ref{eq:2.6a}) means that a permanent current (London current) ${\bf j} = M^2 {\tilde {\bf A}}$ exists, and since ${\bf E} = 0$ and ${\bf E} = \sigma_c \,{\bf j}$ ($\sigma_c$ is the conductivity), $\sigma_c=0$. The key parameter is the order parameter $\langle\Phi\rangle$ which signals the Higgs phenomenon. There are two characteristic lengths in the system related to $\langle\Phi\rangle$: the correlation length of the $\Phi$ field (or the inverse Higgs mass) $\Lambda = 1/M_\phi$, $M_\phi=\lambda^\prime \langle \Phi \rangle$, and the penetration depth of the photon, $\lambda = 1/M$. If $\lambda > \Lambda$ the superconductor is called of type II, and the formation of Abrikosov flux tubes is favored in the process of penetrating the material with a magnetic field. If the opposite inequality holds, $\lambda < \Lambda$, it happens that, when the magnetic field is increased, there is an abrupt penetration of it at some value and superconductivity is destroyed. The superconductor is of type I. Summarizing, in this model the condensation of electric charges leads to the formation of a quantized flux tubes (see Exercise 6.2.1.1) whose radius and shape are controlled by $\Lambda$ and $\lambda$. If a magnetic monopole and antimonopole were introduced into such a superconducting medium they would be connected by a flux tube of finite energy per unit length (finite string tension). Thus magnetic monopoles would be confined due to a linear potential. \vskip 1truecm \leftline{\bf Exercises} \begin{itemize} \item[6.2.1.1]{ Show that a side consequence of the Meissner effect is the flux quantization. [Hint: Calculate the integral $\displaystyle \oint \tilde{\bf A} \cdot d {\bf x}$ around a large circle centered on the flux tube.] } \item[6.2.1.2]{ Consider the Maxwell equations in a relativistic medium without sour\-ces. Instead of expressing as usual the ${\bf E}$ and ${\bf B}$ fields in terms of the magnetic potential $A_\mu$, express the ${\bf D}= \varepsilon {\bf E} $ and ${\bf H}={\bf B}/\mu$ fields in terms of a (dual) electric potential $C_\mu$. Write down the Maxwell equations in a covariant form in terms of the field strength tensor of $C_\mu$. How do you have to modify the definition of this field strength tensor in the presence of sources?} \end{itemize} \begin{figure}[htb] \vskip 0.1truecm \makebox[2.5truecm]{\phantom b} \epsfxsize=8truecm\epsffile{dual2.eps} \vskip 0.2truecm \caption{\it Schematic diagram illustrating how a Non-Abelian gauge theory reduces to an Abelian theory with electric charges and monopoles and the pattern of vacuum phases (see Sec. 4.4).} \label{plhooft} \end{figure} \subsubsection{The QCD Vacuum} We have seen in the previous section how in ordinary superconductivity the condensation of electric charges gives origin to monopole confinement. On the other hand, in all the theories allowing for an analytic proof of electric charge confinement, like compact electrodynamics, the Georgi--Glashow model or some supersymmetric Yang--Mills theories (see Alvarez-Gaum\'e (1997)), this is due to the condensation of monopoles. However, if we want to apply this idea to QCD we have first to understand in which way we can obtain Abelian degrees of freedom and monopoles in QCD, since {\it in the QCD Lagrangian there are no Higgs fields!}. Subsequently we have to prove that the monopoles actually do condense. The application of this idea to non-Abelian gauge theories is based on the so-called Abelian projection, see 't Hooft (1981). The Abelian projection is a partial gauge fixing (of the off-diagonal components of the gauge field) under which the Abelian degrees of freedom remain unfixed. In QCD the Abelian projection reduces the $SU(3)$ gauge symmetry to a $U(1)^2$ gauge symmetry. The Abelian projection monopoles appear as topological quantities in the residual Abelian channel. QCD is then reduced to an Abelian theory with electric charges and monopoles, see Fig. \ref{plhooft}. Precisely, it can be regarded as an Abelian gauge theory with magnetic monopoles and charged matter fields (quarks and off-diagonal gluons). Then, the dual superconductor picture is realized if these Abelian monopoles condense: this causes confinement of the particles that are electrically charged with respect to the above ``photons''. In this scenario large distance (low momentum) properties of QCD are carried by the Abelian degrees of freedom (Abelian dominance) and specifically by the monopole configurations (monopole dominance), see Suzuki (1993). We do not have enough space to give further details. The reader is referred to Ex. 6.2.2.1 and to the reviews e.g. of Bali (1998), Chernodub et al. (1997), Di Giacomo (1998), Haymaker (1998). In the last years intensive work has been done to collect information on the quark confinement mechanism via lattice measurements. Depending on the picture, the excitations giving rise to confinement are thought to be magnetic monopoles, instantons, dyons, centre vortices (see Faber (1998)), etc. The above ideas are not completely disjoint and do not necessarily exclude each other. The above mentioned topological excitations indeed, are found to be correlated with each other in the lattice studies (e.g. correlations between instantons and monopoles). Many questions are still not completely settled. Here, we would like to summarize briefly the present understanding and to show some of the lattice measurements made in Abelian projection. \begin{figure}[htb] \vskip 0.2truecm \makebox[1.0truecm]{\phantom b} \centerline{ \epsfxsize=8truecm \epsffile{sig16.ps}} \vskip 0.2truecm \caption{\it Action density distribution in Maximal Abelian Projection of $SU(2)$ measured on a lattice $V=32^4$ at $\beta=2.5115$. The physical quark-antiquark distance is $1.2$ fm. Figure provided by G. Bali.} \label{plpac} \end{figure} First, we mention some general problems one encounters in this approach. 1) The identification of ``photon fields'' and monopoles is a gauge invariant process. However, the choice of the operator that defines the Abelian projection is somehow ambiguous. There are many different ways of making the Abelian projection. The physics, e.g. the monopole condensation, should be independent of the gauge fixing. 2) The actual composition of the condensate should be known. 3) The origin of the monopole potential that yields to a non-vanishing vacuum expectation value of the magnetic condensate should be understood. On the other hand it is clear that: 1) monopoles are condensed in the confinement phase and this independently of the choice of the Abelian projection. 2) Monopoles are responsible for the main part of the nonperturbative dynamics. In particular Abelian and monopoles dominance seem to hold (in the Maximal Abelian Projection (MAP)): expectation values of physical quantities in the non-Abelian theory coincide with (or are very close to) the expectation values of the corresponding Abelian operators in the Abelian theory obtained via Abelian projection. In other words, disregarding the off-diagonal gluons, the long range features of a $SU(N)$ gauge theory are reproduced while short range features may be altered. Monopole dominance means that the same result can be approximately calculated in terms only of the monopole currents extracted from the Abelian fields. \begin{figure}[htb] \vskip 0.2truecm \makebox[1.0truecm]{\phantom b} \centerline{ \epsfxsize=8truecm \epsffile{mon16.ps}} \vskip 0.2truecm \caption{\it Monopole contribution to the action density distribution in Maximal Abelian Projection of $SU(2)$. Parameters as in Fig. \ref{plpac}. } \label{plpmon} \end{figure} As a general conclusion one can say that {\it the lattice data in MAP show that the lattice gluodynamics is in a sense equivalent at large distances to the Dual Abelian Higgs model at the border between superconductor of type I and superconductor of type II}. More precisely the Abelian monopole action has been extracted from $SU(2)$ lattice data in MAP and mapped into an Abelian Higgs model. From this the effective string theory has even been reconstructed and it occurs that the classical string tension of the string model is close to the quantum string tension of $SU(2)$ lattice gluodynamics, cf. Chernodub et al. (1999). The field distributions in MAP are found to satisfy the dual Ginzburg Landau equations and the coherence and the penetration lengths have been measured, cf. Bali (1998). In Fig. \ref{plpac} we present the action density distribution between two static quarks in $SU(2)$ in MAP to be compared on one hand with the action density in Fig. \ref{plac} and on the other hand with the monopole contribution in Fig. \ref{plpmon}. We see that still the ``photon'' (neutral gluon) contributes to the self-energy of charges: the monopole part is free of self-energy while the ``photon'' part is free of string tension. Electric flux tubes (cf. Fig. \ref{ple}) are found to be significantly thinner after Abelian projection. \begin{figure}[htb] \vskip 0.1truecm \makebox[1.0truecm]{\phantom b} \centerline{ \epsfxsize=8truecm \epsffile{e12.ps}} \vskip 0.5truecm \caption{\it Electric field at 12 lattice units between two static quarks in $SU(2)$ measured on a lattice $V=32^4$ at $\beta=2.5115$. The lattice spacing is fixed on $\sigma= (440$ MeV$)^2$. The physical quark-antiquark distance is 1 fm. Figure provided by G. Bali.} \label{ple} \end{figure} At the end of this section we present two plots of the static potential. Fig. \ref{plppot1} is a plot of the static potential calculated by means of either non-Abelian or Abelian projected Wilson loops. If one takes into account that in the Abelian projected case all the ``charged components'' of $A_\mu$ are neglected, one finds an impressive agreement between the two curves. Fig. \ref{plppot2} shows the static potential calculated in the Abelian projection. The photon contribution and the monopole contribution have been calculated separately. This plot confirms the action density result discussed above. Moreover, it is possible to quantify $\sigma_{U(1)}\simeq 92 \% \, \sigma_{SU(2)}$ and $\sigma_{mon} \simeq 95 \% \, \sigma_{U(1)}$. Notice that in the Abelian projection the coefficient of the Coulombic part comes out to be smaller by more than a factor of two. {\it The original phenomenological Cornell potential that was confirmed by the lattice simulations on the static Wilson loop in Sec. 4.5, is now understood in terms of monopole contributions while the monopole currents are the origin of the chromoelectric flux tube.} It is also possible to construct infrared effective ``analytic'' models based upon the assumptions and on the results presented above and to use them to explain the QCD low energy physics. As already discussed, in the Wilson loop formalism only an assumption on the Wilson loop behaviour is necessary. This is outlined in the next sections. \begin{figure}[htb] \vskip 0.2truecm \makebox[1.0truecm]{\phantom b} \centerline{\epsfxsize=8truecm\epsffile{bornfull.ps}} \vskip 0.2truecm \caption{\it Static potential in $SU(2)$ (diamonds) and in the Abelian projection of $SU(2)$ (squares) (in lattice units, $a\simeq 0.081$ fm). Bali et al. (1996).} \label{plppot1} \end{figure} \begin{figure}[htb] \vskip 0.2truecm \makebox[1.0truecm]{\phantom b} \centerline{\epsfxsize=8truecm\epsffile{bornfin.ps}} \vskip 0.2truecm \caption{\it The Abelian projected $SU(2)$ potential (diamonds) in comparison with the ``photon'' contribution (squares), the monopole contribution (crosses) and the sum of the two parts (triangles) (in lattice units, $a=0.081$ fm). No self-energy constant have been subtracted. Bali et al. (1996).} \label{plppot2} \end{figure} \vskip 1truecm \leftline{\bf Exercises} \begin{itemize} \item[6.2.2.1]{ This is a simple example of Abelian projection for SU(2). Consider the following condition $F_{12}(x) = \hbox{diagonal matrix}$. Show that $F_{12}$ is still invariant under the U(1) gauge transformation $\Omega(x)_{ij}$ defined as $$ \Omega_{12}=\Omega_{21}=0 \qquad \Omega_{11}=\Omega_{22}^* = e^{i \alpha(x)} \qquad \alpha \in [0,2\pi). $$ Define $A_\mu^{\pm}\equiv A_\mu^1\pm A_\mu^2$, $1,2,3=$ colour indices. Show that $A_\mu^3 $ transforms as a photon under $\Omega$ while $A_\mu^\pm$ transform as charged fields. Define an Abelian field strength tensor in terms of $A_\mu^3$ and show that, if the matrix of the gauge transformation $\Omega$ contains singularities, also the Abelian field strength tensor may contain singularities (monopoles).} \end{itemize} \subsection{Low energy theorems and flux tube} We have seen that one of the relevant features of confinement is the interquark flux tube formation, see Figs. \ref{plac}, \ref{ple} in $SU(2)$ and Figs. \ref{plpac}, \ref{plpmon} in the MAP of $SU(2)$. Typical quantities of interests are the transverse extent of the flux tube and the nature of the colour fields (i.e. electric or magnetic), see Bali et al. (1995) and Green et al. (1997). There are two low energy theorems relating the potential of a static quark-antiquark pair with the total energy and the action stored in the flux tube between the sources (in lattice QCD these are know as Michael's sum rules (Michael (1987)), Dosch et al. (1995), Novikov et al. (1981), Shifman (1998), \begin{eqnarray} V_0(r) &=& {1\over 2} \langle \int d^3x (-{\bf E}(x)^2 +{\bf B}(x)^2)\rangle_{r\times T}\nonumber\\ V_0(r)+ r{\partial V_0(r)\over \partial r} &=& {1\over 2} {\beta({\alpha_{\rm s}})\over \alpha_{\rm s}} \langle \int d^3x ({\bf E}(x)^2 +{\bf B}(x)^2)\rangle_{r\times T}. \label{teor} \end{eqnarray} $\langle \cdots \rangle_{r\times T}$ denotes the expectation value in the presence of the static Wilson loop where the expectation value in the absence of the sources has been subtracted. The formulas are in Euclidean space and renormalized composite operators are used. The squared electric and magnetic field strengths are separately not renormalization group invariant. Thus statements like $\langle g^2 {\bf B}^2\rangle_{r\times T} \simeq 0$ or $\langle g^2 {\bf E}_\perp^2\rangle_{r\times T} \simeq 0$ are scale dependent. As we already pointed out, on the lattice the averages of the squared components of the colour fields are found to be roughly equal (i.e. $\langle E_i^2\rangle_{r\times T}$ $\simeq$ $\langle B^2_j\rangle_{r\times T}$, $i,j=1,2,3$), while in the greater part of the models the flux tube is mainly made of the longitudinal electric field. Eqs. (\ref{teor}) fixes the scale at which each of these situations can be fulfilled. In particular for models, typically describing the long range behaviour with $V_0 \simeq \sigma r$, one has $\beta/\alpha_{\rm s} \simeq -2$. Taking the three-loop beta function, this gives $\alpha_{\rm s} \simeq 0.6$. \subsection{Models of the QCD Vacuum} We briefly mention some analytic models of the QCD vacuum. We refer the reader to the papers quoted below for further details. \begin{figure}[htb] \vskip 0.2truecm \makebox[1.0truecm]{\phantom b} \centerline{\epsfxsize=7truecm\epsffile{bbzact.eps}} \vskip -0.4truecm \caption{\it On a semi-log scale, a comparison of the total energy profile for dual QCD (solid line) and in the lattice predictions (solid circles) for $R$ (=$q\bar{q}$ distance) ranging from 8 down to 1 lattice units as a function of the transverse coordinate $r_{\rm T}$ (with respect to the interquark string). All profiles are in lattice units with $a\simeq 0.6$ $Gev^{-1}$. From Green et al. (1997).} \label{bbzact} \end{figure} \subsubsection{Dual QCD (DQCD)} Impressive progresses have been done recently towards an understanding of confinement via duality and monopoles condensation in supersymmetric theories, see e.g. Alvarez-Gaum\'e et al. (1997). However, it is also possible to construct a dual effective theory of long distance Yang--Mills theory, Baker et al. (1986-1996), see also Maedan et al. (1989). We call this theory Dual QCD (DQCD). DQCD is a concrete realization of the Mandelstam and 't Hooft dual superconductor mechanism of confinement. It describes the QCD vacuum as a dual superconductor on the border between type I and II. The Wilson loop approach supplies a simple method to connect averaged local quantities in QCD and in dual QCD. For large loops we assume \begin{equation} \langle W(\Gamma) \rangle \simeq {\displaystyle{\int \!{\cal D} { C} \, {\cal D} {B} \, \exp \left[ i \int dx L(G_{\mu\nu}^{\rm S}) \right] } \over \displaystyle{\int \!{\cal D} { C}\, {\cal D} { B} \, \exp \left[ i \int dx L(0) \right] }} \label{deq} \end{equation} where $L$ is the effective dual Lagrangian. The fundamental variables are an octet of dual potential $C_\mu$ coupled minimally to three octets of scalar Higgs field $B_i$ carrying magnetic colour charge, the dual coupling constant being $e = 2\pi/g$. Notice that the dual transformation exchange the strong coupling limit of QCD with the weak coupling limit of DQCD. The monopole fields $B_i$ develop nonvanishing expectation values $B_{0i}$ giving rise to massive $C_\mu$ and to a dual Meissner effect. Dual potentials couple to electric colour charges like ordinary potentials couple to monopoles, i.e. $C_\mu$ couple to the quark-antiquark pair via a Dirac string connecting the pair. It turns out that, for the description of a quark-antiquark state, the relevant subset of the theory is a (dual) Abelian Higgs model. In this case indeed, the effective Lagrangian is explicitly given by ${ L}(G_{\mu\nu}^{\rm S}) = 2~{\rm Tr} \{ - {1\over 4} {G}^{\mu\nu} {G}_{\mu\nu}$ $+ {1\over 2} ({D}_\mu {B}_i)^2 \} - U({B}_i)$, where $G_{\mu\nu}=(\partial_\mu C_\nu-\partial_\nu C_\mu + G^S_{\mu\nu})$. The dual field is directed along the hypercharge matrix $Y$ in the colour space, $G_{\mu\nu}^{\rm S}(x) = g \,\epsilon_{\mu\nu\alpha\beta} \displaystyle\int ds \int d\tau\, \displaystyle{\partial y^\alpha \over \partial s} {\partial y^\beta\over \partial \tau} \delta(x-y(s,\tau))\, Y$, ($y(s,\tau)$ is a world sheet with boundary $\Gamma$ swept out by the Dirac string) and $U({B}_i)$ is the Higgs potential with the minimum values $B_{0i}$ chosen in order to completely break the dual $SU(3)$ symmetry. It essentially coincides with the Abelian Higgs Lagrangian (plus sources) of Eq. (\ref{eq:2.1}) where the fields $A_\mu$ play the role of the dual potentials $C_\mu$ and a combination of fields $B_i$ plays the role of the Higgs field $\Phi$. From (\ref{deq}) it follows that (see Baker et al. (1996)) \begin{equation} \langle\!\langle F_{\mu \nu} (z_j)\rangle\!\rangle = {2\over 3} \, \varepsilon_{\mu\nu \rho\sigma}\langle\!\langle G^{\rho\sigma} (z_j) \rangle\!\rangle_{Dual}\, . \label{valdual} \end{equation} Therefore, it is possible to relate averaged values of local quantities in QCD and in the dual theory. The nonperturbative parameters are the v.e.v. of the Higgs field and the coupling constant of the Higgs potential (from these the penetration length and the correlation length can be constructed). Flux tube configurations with finite $r_{MS} \sim {1/ M}$, $M=$ mass of the dual gluon, arise (Baker et al. (1996)) from the numerical solution of the classical dual Ginzburg--Landau type of equations obtained from $L$. The flux tube profile agrees well with lattice data (Green et al. (1996)), see Fig. \ref{bbzact}. It is the presence of the Higgs field which confines the transverse energy distribution in a flux tube. The string tension $\sigma$ comes from the integral of the exponentially decreasing energy distribution. \begin{figure}[htb] \vskip 0.2truecm \makebox[1.0truecm]{\phantom b} \centerline{\epsfxsize=6truecm\epsffile{svmdosch.eps}} \vskip -0.4truecm \caption{\it The energy density distribution in GeV/$fm^3$ caused by the nonperturbative correlator $D$ and the colour Coulomb contribution $D_1$ with $\alpha_s=0.57$. In the figure $a=T_g=0.35 fm$, the $q\bar{q}$ distance is 9 $T_g$. From R\"uter et al. (1995).} \label{rut} \end{figure} \subsubsection{Stochastic Vacuum Model} The Stochastic Vacuum Model (Dosch (1987), Dosch et al. (1988)) is based on the idea that low frequency contributions in the functional integral can be taken into account by a simple stochastic process with a converging cluster expansion. Under this assumption the Wilson loop manifests an area law behaviour and therefore linear confinement. The model does not give rise to confinement for an Abelian gauge theory. In order to make quantitative predictions it is convenient to make a more radical assumption, namely that all higher cumulants can be neglected as compared to the two point function. This assumption appears to be in agreement with a recent lattice analysis (Bali et al. (1998)). Then, the stochastic process is Gaussian and all fields correlators are reduced to products of two point functions by factorization. This means that the Wilson loop can be approximated simply by (in Euclidean space) \begin{equation} \langle W(\Gamma) \rangle \simeq \exp\left[ - \displaystyle {g^2\over 2} \int_{S(\Gamma)} \!\!\!\!d S_{\mu\nu}(x) \int_{S(\Gamma)} \!\!\!\!d S_{\lambda\rho}(y) \langle F_{\mu\nu}(x) U(x,y) F_{\lambda\rho}(y) U(y,x) \rangle \right] \label{Wsvm} \end{equation} where $S(\Gamma)$ is a surface with contour $\Gamma$, typically the minimal area surface. The nonperturbative dynamics is given in terms of one unknown function only: the non-local gluon condensate \begin{eqnarray} && g^2 \langle U(0,x) F_{\mu\nu}(x) U(x,0) F_{\lambda\rho} (0) \rangle = \bigg\{ (\delta_{\mu\lambda}\delta_{\nu\rho} - \delta_{\mu\rho}\delta_{\nu\lambda})(D(x^2) + D_1(x^2)) \nonumber \\ && + (x_\mu x_\lambda \delta_{\nu\rho} - x_\mu x_\rho \delta_{\nu\lambda} + x_\nu x_\rho \delta_{\mu\lambda} - x_\nu x_\lambda \delta_{\mu\rho}) {d\over dx^2}D_1(x^2) \bigg\} . \label{due} \end{eqnarray} Eq. (\ref{due}) is the most general Lorentz decomposition of the two-point correlator. The dynamics is contained in the form factors $D$ and $D_1$. The function $D$ is responsible for the area law and confinement. Lattice simulations (for a sum-rule analysis see Dosch, Eidem\"uller and Jamin (1998)) show that the $D$ and $D_1$ functions exhibit a long-range exponential fall off $\simeq F_2 \exp\{-\vert x\vert/T_g\}$ where the correlation length $T_g$ is about 0.2 $\div$ 0.3 fm (D'Elia et al. (1997)). The model has thus two nonperturbative parameters $F_2$ and $T_g$. The static potential is given by \begin{equation} V_0(r) \simeq F_2 \int_0^\infty d\tau \int_0^r d \lambda\, (r-\lambda) D(\tau^2 +\lambda^2) \label{svpot} \end{equation} and the string tension $\sigma$ emerges as an integral on the function $D$, $\sigma \simeq F_2 \displaystyle \int_0^\infty d\tau $ $\displaystyle \int_0^\infty d\lambda$ $D(\tau^2 +\lambda^2)$, in the limit ${T_g/ r}\to 0$. The field distribution between the quark and the antiquark is a flux tube (R\"uter et al. (1995)) with $r_{MS} \simeq 1.8\, T_g$, see Fig. \ref{rut}. We refer to Dosch (1994), Nachtmann (1996) and Simonov (1996) for a complete description of the model and its applications. \vskip 0.7truecm Finally, we mention in this neither exhaustive nor complete list, the flux tube model of Isgur et al. (1983). This model is extracted from the strong coupling limit of the QCD lattice Hamiltonian. A $N$-body discrete string-like Hamiltonian describes the gluonic degrees of freedom. The limit $N\to \infty$ corresponds to a localized string with an infinite number of degrees of freedom. All the mentioned models allow to obtain via Eqs. (\ref{potential}), (\ref{e20}), (\ref{e21}), the complete semirelativistic quark interaction. For a discussion and a comparison of the result see Baker et al. (1997), Brambilla et al. (1997a) and Brambilla (1998). A relation between DQCD and SVM has been established in Baker et al. (1998). Here, we only stress that we need two parameters like $T_g$ and $F_2$ to control the structure of the flux tube. Had we only one parameter, like the string tension $\sigma$, we could only encode the information of a constant energy density in the flux tube. However, the whole structure is important and also the information about the width of the flux tube (typically proportional to $T_g$) has to be considered. \vskip 1truecm \leftline{\bf Exercises} \begin{itemize} \item[6.4.1]{ Show that in the limit $T_g \to 0$ the Wilson loop in Eq. (\ref{Wsvm}) is given by $\exp \{-\sigma S\}$ with $\sigma$ as defined above and $S$ the surface enclosed by the loop $\Gamma$.} \item[6.4.2]{ Show that using the lattice parameterization of $D$ given above, with $T_g \simeq 0.3$ fm and $F_2 \simeq 0.048$ GeV$^4$, the Stochastic Vacuum Model gives a string tension $\sigma$ compatible with the phenomenological value of Eq. (\ref{sigma1}).} \end{itemize} \subsection{String description of QCD} An effective string theory of strong interaction has been proposed first by Nambu and Goto in 1960 to explain the fact that the hadrons lie on Regge trajectories (see Sec. 2). Nambu was the first who explicitly constructed an effective theory of the Abelian Higgs model working in the limit of infinite Higgs mass which turned out to be coincident with the Nambu--Goto string Lagrangian. Recent works have shown that properly summing over all the string positions to account for the fluctuation of the string generates an effective string in four dimension free of the conformal anomaly. Results exist both in the limit of infinite Higgs mass (Akhmedov et al. (1996)) and without taking that singular limit (Baker et al. (1999)). A phenomenological description of hybrids has been obtained using a Nambu--Goto action in Allen et al. (1998). \section{Summary} These lectures have been devoted to a summary of our understanding of the heavy hadron spectrum in QCD. We started from some phenomenological guess of the interquark static potential inspired from a naive flux tube picture. Then, we managed to give a well founded definition of such a potential in field theory and to ``derive'' the quark confinement property in strong coupling expansion. To obtain the form of the potential we resorted to the lattice formulation and to numerical techniques that eventually confirmed the linear rising of the phenomenological static potential. The relativistic corrections were added in the framework of effective field theories establishing a systematic way of estimating the neglected terms. This allowed us to obtain a gauge-invariant field theory based definition of the semirelativistic quark interaction devised for lattice simulations as well as for analytic calculations. Eventually this procedure enabled us to connect directly the QCD vacuum with the spectrum. At the end we examined the Wilson loop in the Abelian projection suggested by 't Hooft, i.e. under the assumption that the QCD vacuum is a dual superconductor, and we found that our initial Cornell potential seems to be the output of dominant monopoles configurations. The aim of these lectures has been to show how heavy quark bound states offer an ideal situation where exact results from QCD (effective field theories like NRQCD and pNRQCD) can be used to explore the QCD vacuum either with lattice numerical tools or by means of analytic models. Therefore many theoretical ideas and techniques can be and have been used in order to explain that kind of systems. All of them cooperate to enlarge our predictability and to open new perspectives. It is not surprising that, in spite of several still open problems, some remarkable progress have been achieved. \section*{Acknowledgments} We gratefully acknowledge interesting discussions with Marshall Baker, Manfried Faber, Dieter Gromes, Khin Maung, Martin Zach and the participants at the HUGS'98 School. We thank Gunnar Bali for making available to us his lattice data, for many enlightening discussions and for making many useful comments and suggestions especially on the lattice part of these lectures. We acknowledge M. Baker and D. Gromes for reading the manuscript and making useful comments. It is a pleasure for N. B. to thank Jos\'e Goity for the invitation to give these lectures and for the perfect organization of the school. N. B. acknowledges the support of the European Community, Marie Curie fellowship, TMR contract No. ERBFMBICT961714; A. V. acknowledges the FWF contract No. 9013. \section*{References}
\section{General Appearance} \vspace*{-0.5pt} \noindent {\em PACS}: 03.65.Ca, 03.65Fd {\em Keywords}: Intertwining; Factorization energy; q deformation \bigskip {\footnotesize {\bf To arXive or Not to arXive ? That's a good question !}} {\footnotesize except for the title and minor changes in the text, version of 3/8/99} {\footnotesize \copyright 1999, by H.C. Rosu} \noindent \newpage \noindent I present herein a simple $q$-deformed procedure [\refcite{Jac}] for the basic case of the one-dimensional quantum harmonic oscillator, by which I build a `supersymmetric' pair of q nonlocal operators possessing terms whose $q\rightarrow 1$ limits belong to either Hermite polynomial operator or Schroedinger quantum oscillator operator. There are $\pm x$ extra terms as well. The procedure is based on the idea of using, as fundamental tools, a sort of deformed counterparts of the intertwining operators encountered in the area of Darboux transformations [\refcite{darb}]. I shall use their factorization property to get the q-deformed second-order operators which, being $q$ non-local, may be considered as more general than both the usual Hermite one and the quantum mechanical harmonic oscillator operator. The standard Hermite operator ${\hat{O}} _{H}$ reads ($D=d/dx$) \begin{equation} {\hat{O}}_{H}=D^2-2xD+2n~, \end{equation} and gives rise to the equation for the Hermite polynomials $H_{n}(x)$, ${\hat{O}} _{H}H_{n}(x)= 0$. Writing ${\hat{O}}_{H}=D^2-2xD-2+2(n+1)$, I shall treat $D^2-2xD-2$ as the Fokker-Planck (FP) part of the Hermite operator for a stationary transition-probability density, since $-2xD$ corresponds to the $(\frac{dU}{dx})D$ drift contribution in the FP stationary operator (drift potential $U=-x^2$), whereas $-2$ stands for the $d^2U/dx^2$ contribution of the FP drift. The last term $2(n+1)$ gives the departure of the Hermite operator from the corresponding FP stationary operator for which polynomial oscillations are not allowed, and in fact is responsible for turning the FP interpretation into a formal one and not a physical one. As well known for this basic case, by means of the functions $\phi _{n}= e^{-x^2/2}H_{n}(x)$ one can go to the operator \begin{equation} {\hat{O}}_{\phi}=-D^2+[x^2- (2n+1)]~, \end{equation} which, in the $\phi _{n}$ space, is essentially the Schroedinger quantum harmonic oscillator operator up to a scaling, choose-of-units factor. One should notice that this usage of the $\phi _{n}$ functions leads to the loss of one half of the $\frac{d^2U}{dx^2}$ drift contribution. The remaining half gets the famous zero-point energy interpretation when the scaling $\frac{1}{2}{\hat{O}}_{\phi}$ is performed. In the FP interpretation, the latter scaling corresponds to setting the diffusion constant equal to $1/2$ and provides the usual quantum mechanical harmonic oscillator wavefunctions $N_{n}\phi _{n}$, where $N_{n}=(2^n n!\sqrt{\pi})^{-1/2}$ is the normalization factor. I now briefly recall that in the case of the one-dimensional Schroedinger operator within the context of supersymmetric quantum mechanics (SUSYQM) [\refcite{rsqm}] the standard Darboux transformation operator reads \begin{equation} \label{T} {T}=-t_{u}(x)+D=-u^{\prime }(x)/u(x)+D~, \end{equation} where the prime denotes the derivative with respect to $x$. When acting on the solutions $\psi _n(x)$ of the initial Schroedinger equation $ h_0\psi _n(x)=E_n\psi _n(x), $ it transforms them into the solutions of another Schroedinger equation $h_1\varphi _n(x)=E_n\varphi _n(x)$, $\varphi _n(x)=N_n{T}\psi _n(x)$, with the same eigenvalues $E_n$. Henceforth, I will put the ground state energy equal to zero, $E_0=0$, since this does not affect in any way the results. The new exactly solvable Hamiltonian has the form $h_1=h_0+\Delta V(x)$, where the potential difference is of Darboux type $\Delta V(x)=-2(\ln u)^{\prime \prime }$. The function $u=u(x)$ is a so-called transformation function, being a solution of the initial Schroedinger equation $ h_0u(x)=\epsilon u(x), $ with $\epsilon \leq 0$ usually known as the factorization energy. It is well established that when $\epsilon <0$ one can work with a nodeless transformation function by performing an analytic continuation [\refcite{bsam}]. Thus, $u(x)\ne 0$ for any value of the variable and $1/u(x)$ is not a square integrable function. In this case $u\notin {\cal H}_{1}$ and the set $\{\mid \varphi _n\rangle \}$ is a complete basis in the Hilbert space ${\cal H}_{1}$ provided the initial system $\{\mid \psi _n\rangle \}$ is complete. The operator ${T}^{+}=-t_{u}(x)-D$ provides the backward transformation $ \label{psin}|\psi _n\rangle =N_n {T}^{+}|\varphi _n\rangle , $ and together with ${T}$ allows for the following factorizations \begin{equation} {T}^{+}{T}=h_0-\epsilon ,\quad {T}{T}^{+}=h_1-\epsilon \ . \end{equation} The operators ${T}$ and ${T}^{+}$ are well defined $\forall \psi \in {\cal H}_{1}$ and are conjugated to each other with respect to the inner product in the ${\cal H}_{1}$ space. My purpose now is to get $q$-deformed second-order operators by means of deformed counterparts of the aforementioned intertwining operators. I still have to present some definitions and rules of the deformed calculus. Since the independent variable is maintained commutative, the employed version of the deformed calculus is similar to that previously used by some authors to deform the Coulomb problem [\refcite{coul}]. Symmetric definitions of the q-number $[x]_q=\frac{q^{x}-q^{-x}}{q-q^{-1}}$ and $q$-derivative \begin{equation} D_{q}f(x)=\frac{f(qx)-f(q^{-1}x)}{x(q-q^{-1})}~ \end{equation} are used together with some basic rules of Jackson's calculus [\refcite{Jac}] such as $D_{q} x^{n}=[n]_{q}x^{n-1}$, $D_{q}^{2}x^n=[n]_q[n-1]_qx^{n-2}$, $D_{q}(FG)=(D_{q}F)G(qx)+F(q^{-1}x)(D_{q}G)$ for any two functions $F$ and $G$, respectively. The definition of the $q$-exponential is \begin{equation} e_{q}(x)= \sum _{n=0}^{\infty}\frac{x^{n}}{[n]_{q}!}~, \end{equation} which reduces to the usual exponential function as $q\rightarrow 1$, and moreover is invariant under $q\rightarrow q^{-1}$. The main idea of this work is based on the following scheme. First, to employ as Darboux transformation functions deformed counterparts of the oscillator vacua $\psi_{q}\propto e_q(\beta x^2)$, where $\beta =\pm 1/2$ for the irregular and regular vacuum, respectively. Second, to exploit the factorization property of first-order deformed operators of the form \begin{equation} T_{+}^{q}=D_{q}-\frac{D_{q}\psi_{q}}{\psi_{q}}= D_q -\beta _{q}(x^2) x~, \end{equation} \begin{equation} T_{-}^{q}=-D_{q}-\frac{D_{q}\psi_{q}}{\psi_{q}}= -D_ -\beta _{q}(x^2)x~, \end{equation} where \begin{equation} \beta _{q}(x^2)=\beta \left(\frac{qe_{q}(q\beta x^{2})+ q^{-1}e_{q}(q^{-1}\beta x^{2})}{e_{q}(\beta x^2)}\right)~. \end{equation} The form of $\beta _{q}(x^2)$ is a result of Jackson's calculus rules. As one can see, the $T_{+}^{q}$ and $T_{-}^{q}$ operators have been written by analogy to the continuous intertwining operators. A straightforward calculation gives the second-order deformed operators that can be obtained from the products $T_{-}^{q}T_{+}^{q}$ and $T_{+}^{q}T_{-}^{q}$, respectively. One gets \begin{equation} {\hat{O}}_{b}^{q}\equiv T_{-}^{q}T_{+}^{q}= -D_{q}^{2}- [(\Delta \beta _{q})xD_{q} +[ \beta _{q}^{2}(x^2)x^2 ]_{\rightarrow x} +[q(D_{q}\beta _{q}(x^2))x]_{\rightarrow qx +[\beta _{q}(q^{-2}x^2)]_{\rightarrow qx \end{equation} and \begin{equation} {\hat{O}} _{f}^{q}\equiv T_{+}^{q}T_{-}^{q}= -D_{q}^{2}+ [(\Delta \beta _{q}) D_{q} +[\beta _{q}^{2}(x^2)x^2] _{\rightarrow x} -[q(D_{q}\beta _{q}(x^2))x]_{\rightarrow qx -[\beta _{q}(q^{-2}x^2)]_{\rightarrow qx} \end{equation} where \begin{equation} \Delta \beta _{q}=\beta _{q}(x^2)- q^{-1}\beta _{q}(q^{-2}x^2)~. \end{equation} \noindent The operators ${\hat{O}} _{b}^{q}$ and ${\hat{O}} _{f}^{q}$ may be considered as supersymmetric partners since they have been built according to the well-known SUSYQM method. At this point one should notice the interesting mixed structure of the two $q$ non-local operators that entail parts of both ${\hat{O}}_{H}$ and ${\hat{O}}_{\phi}$. The directional subindices indicate the argument of the solution on which the nonoperatorial parts act. The two operators are nonlocal operators whose space of solutions are the functions $\phi ^{(q)}_{n}(x)\propto e_{q}(-x^2/2)H^{(q)}_{n}(x)$, where the deformed Hermite polynomials can be defined by a $q$ deformed Rodrigues representation \begin{equation} H^{(q)}_{n}=(-1)^{n}e_{q}(x^2)D_{q}^{n}(e_{q}(-x^2))~. \end{equation} Notice that the $D_{q}$ terms in (10) and (11) are identical but opposite in sign and correspond to the first derivative drift term in the Hermite differential operator. Of course, if one prefers the FP interpretation the two operators should be multiplied by (-1). Writing the finite difference $x(q-q^{-1})=x\Delta q=\Delta _{q} x$, which for $q\rightarrow 1$ is assumed to be a $q$ scaling way of going to the infinitesimal limit $dx$, the $q$ drift parts go to zero, whereas in the same limit the potential and zero point sectors take forms identical to those of the undeformed case. More precisely, the undeformed limits read \begin{equation} {\hat{O}} _{b}^{1}\equiv h_{0}=-D^2 +\beta _{1}^{2}x^2+\beta _{1 \end{equation} and \begin{equation} {\hat{O}} _{f}^{1}\equiv h_{1}=-D^2 +\beta _{1}^{2}x^2-\beta _{1}~ \end{equation} (14) and (15) are the usual quantum mechanical supersymmetric partner Hamiltonians for this case. In SUSYQM terminology, only the case of zero factorization energy $\epsilon =0$ has been tackled up to now, but following a suggestion of Bagrov and Samsonov [\refcite{bsam}], there is no difficulty to sketch the procedure for the more general case $\epsilon <0$. First, the deformed Schroedinger solution corresponding to the excited harmonic oscillator states can be sought in the form \begin{equation} \psi _{n}^{(q)}(x)\propto H_{n}^{(q)}(x)e_{q}(-x^2/2)~. \end{equation} \noindent Next, in order to avoid any singularities, it is convenient to perform an $i$-rotation $x\rightarrow ix$, leading to \begin{equation} u_{p}^{(q)}(x)\propto H_{p}^{(q)}\left( ix\right) e _{q} \left( x^2/2\right) \quad p=0,1,2,3\ldots \end{equation} The undeformed functions $u_{p}^{(1)}$ are solutions of $h_{0}u_{p}^{(1)}=-(p+1)u_{p}^{(1)}$ and are nodeless on the full line for even $p=2k$. Therefore, they have been used by Bagrov and Samsonov as Darboux transformation functions to generate a family of regular potentials, which, according to an interpretation due to Veselov and Shabat [\refcite{vs}], has a spectrum made up of $2k+1$ segments with equidistant levels. This immediately suggests using $u_{p}^{(q)}(x)$ for even $p=2k$ as Darboux transformation functions in the deformed case. Thus, the intertwining operators can be calculated according to $T_{2k,+}^{q}=D_{q}-\frac{D_q u_{2k}^{(q)}}{u_{2k}^{(q)}}$ and $T_{2k,-}^{q}=-D_{q}-\frac{D_q u_{2k}^{(q)}}{u_{2k}^{(q)}}$, and again by exploiting the factorization property one is led to second-order deformed, non-local operators of more complicated formulas than (10) and (11) for which they are not written down here. In conclusion, a pair of $q$ non-local second-order $q$-differential ($q$-difference) operators have been introduced in this work by means of a particular $q$-deformed intertwining based on $q$-deformed oscillator vacua as Darboux transformation functions. These operators display a mixed structure between the Hermite operator, to which they are similar as regards the first derivative term, and the quantum mechanical oscillator operator, to which they are similar as regards the $x^2$ potential and zero-energy contributions. On the other hand, they present a supplementary $q$ non-local $\pm x$ potential contribution with no counterpart in either Hermite polynomial operator or Schroedinger $x^2$ oscillator operator. All these features suggest many possible applications, e.g., in mesoscopic physics. A more general case corresponding to negative factorization energies of the type $\epsilon _{m}=-(m+1)$, where $m$ is an even positive integer, has also been briefly described. \vspace{0.5cm} \noindent This work has been supported in part by CONACYT project 458100-5-25844E . \vspace{3mm}
\section{Introduction} The final state strong interactions are known to play an important role in weak nonleptonic decays. In particular, the interplay between the strong and the weak phases is required by many signals of direct $CP$ violation in $B$ decays. The presence of final state rescattering is relevant also for the extraction of the $CKM$ phases from time dependent asymmetries, and can modify the magnitude of some processes suppressed in the Standard Model such as $B\to \pi K$ transitions, thus reducing their sensitivity to New Physics. The effect of the final state interactions (FSI) was assumed until recently to be small in the very energetic decays like those of $B$ meson to light pseudoscalar mesons, where the final particles get apart very quickly and have no time to interact strongly by soft multigluon exchanges. A related argument was based on the fact that at the high energy scale imposed by the mass of the $B$ meson there is a suppression of rescattering due to the specific form of the Regge amplitudes dominant at this scale. Recently, this qualitative picture was challenged by a more detailed dynamical approach \cite{Dono1}-\cite{Falk}. The crucial remark made in \cite{Dono1} is that, contrary to conventional expectations, the soft final state interactions do not disappear at the center of mass energy set by $m_B$. The analysis made in \cite{Dono1}-\cite{Falk} is based on hadronic unitarity and very general features of high energy soft interactions. Consider the weak decay $B \to P_1 P_2$, where $P_i$ are pseudoscalar mesons, and denote by $A_{B\to P_1 P_2}$ the amplitude of this process. In the most general way, the unitarity of the $S$-matrix allows one to express the discontinuity of $A_{B\to P_1 P_2}$ as \begin{equation}\label{unit} {\mathcal D}isc A_{B\to P_1 P_2}={1\over 2i}\left[ \langle P_1 P_2| {\mathcal T}|B\rangle - \langle P_1 P_2| {\mathcal T}^\dagger|B\rangle \right]= {1\over 2}\sum_{I}\langle P_1 P_2| {\mathcal T}^\dagger| I\rangle\langle I|{\mathcal T}|B\rangle \,, \end{equation} where ${\mathcal T}$ is the transition operator ($S=1-i {\mathcal T}$), which describes both the weak and strong interactions. To first order, the weak hamiltonian ${\mathcal H}_w$ can appear either in the first matrix element of the product in the right hand side of (\ref{unit}), or in the second. The intermediate hadronic states $\{I\}$ depend of course on the place of ${\mathcal H}_w$. If ${\mathcal H}_w$ is acting in the matrix element containing $B$, $\{I\}$ denote hadronic states produced by the weak decay of $B$ and connected to the final state $\{P_1 P_2\}$ by a strong rescattering. It is this configuration of the sum (\ref{unit}) which describes the final state interactions in $B$ nonleptonic decays. Alternatively, the operator ${\mathcal H}_w$ can be located in the first matrix element in the unitarity sum, in which case $\{I\}$ are states produced by the strong decay of $B$, and connected to $\{P_1 P_2\}$ through a weak interaction. When all the particles are on-shell these terms vanish, since $B$ is stable with respect to strong interactions. The above remarks, though rather trivial, will be useful below for the discussion of the dispersion relations. According to general principles, the decay amplitude $A_{B\to P_1 P_2}$ can be obtained from its discontinuity by means of a dispersion relation. This approach was considered in \cite{Blok}-\cite{Falk}, where, neglecting possible subtractions, a dispersion relation of the following form was used \begin{equation}\label{drel} A(m_B^2, m_1^2, m_2^2) ={1\over \pi }\int _{s_0}^\infty {\mathrm d}s{{\mathcal D}isc A(s+i\epsilon, m_1^2, m_2^2) \over s-m_B^2-i\epsilon}\,. \end{equation} In this relation and the similar ones written below, the limit $\epsilon\to0$ is implicitly assumed. We use here the notation $A_{B\to P_1 P_2}= A(m_B^2, m_1^2, m_2^2)$ to show explicitely the dependence of the decay amplitude on the external masses. As for the discontinuity (already divided by $2i$), it was taken in \cite{Blok}-\cite{Falk} as the "rescattering part" of the unitarity sum (\ref{unit}) discussed above, evaluated for an off-shell $B$ meson of mass squared equal to $s$. The dispersion relation (\ref{drel}) is based obviously on the analytic continuation of the decay matrix element with respect to the mass of the initial meson $B$. We recall that dispersion relations in the external mass variables were derived in the frame of axiomatic field theory \cite{Oehme}-\cite{LSZ}, and were used in phenomenological calculations of form factors \cite{Binc},\cite{Omnes} (see also \cite{Bart}). Heuristic derivations of such dispersion relations are based on the Lehmann, Symanzik, Zimmermann (LSZ) reduction formalism \cite{LSZ}, combined with causality and hadronic unitarity. As emphasized in \cite{Bart} one must be cautious in using such dispersion relations, since the heuristic conjectures might be violated, in particular through the appearance of anomalous thresholds. Of interest to the present work is the fact that the specific form of the dispersion relation and of its discontinuity depends on which of the external particles is reduced in the LSZ formula. By treating the matrix element of the decay $B\to P_1P_2$ in the frame of this formalism, one can prove that the spectral function of a dispersion relation with respect to $s$ (the mass squared of the $B$ meson), like Eq.~(\ref{drel}), is given {\it not} by the ``rescattering'' part of (\ref{unit}), as it was assumed in \cite{Blok}-\cite{Falk}, but by the second class terms in this sum. Moreover, one can also prove that the terms describing the final state interaction in the unitarity sum (\ref{unit}) appear as spectral function in a dispersion relation in terms of the mass squared of one of the final mesons ($P_1$ or $P_2$). Therefore, the calculation of the whole decay amplitude from its discontinuity proceeds along a different line than that applied in Refs.\cite{Blok}-\cite{Falk}. In the present paper, we provide arguments for the assertions made above and consider some applications. In the next Section, using the LSZ formalism, we discuss the heuristic derivation of dispersion relations for the decay amplitude $A_{B\to P_1 P_2}$, when either $B$ or one of the final mesons $P_1$ or $P_2$ are off-mass shell. We do not attempt to give rigorous proofs, but only to establish the correspondence between the dispersion variable and the expression of the absorbtive part. In Section 3 we consider in more details the approximation of two particle unitarity combined with Regge theory for strong interactions, and in Section 4 we discuss some applications: first we briefly indicate how the conclusions of Ref. \cite{Falk} are modified by the use of the adequate dispersion relation. Then, by combining the formalism with $SU(3)$ flavour symmetry we derive constraints on the amplitudes of the $B^0\to \pi^+\pi^-$ decay. \section{ Dispersion relations in the external mass variable} We consider the weak decay amplitude $A_{B\to P_1 P_2}$ defined as \begin{equation}\label{def} A (p^2, k_1^2, k_2^2)= \langle P_1(k_1) P_2(k_2), out| {\mathcal H}_w(0)|B(p)\rangle\,, \end{equation} where we indicated the dependence on the Lorentz invariants $p^2, k_1^2$ and $k_2^2$ (for the physical amplitude $p^2=m_B^2,\, k_1^2=m_1^2,\, k_2^2=m_2^2$ ). By applying the well known LSZ formalism \cite{LSZ} we "reduce" the particle $P_1$, which gives \begin{equation}\label{lsz1} A (m_B^2, k_1^2, m_2^2)= {i\over \sqrt{2 \omega_1}}\int {\mathrm d}x {\mathrm e}^{ik_1 x} \theta (x_0) \langle P_2(k_2)|[\eta _1(x), {\mathcal H}_w(0)]|B(p)\rangle\,. \end{equation} In this relation, $\eta_1(x)$ denotes the source of the meson $P_1$, defined as $ {\mathcal K}_x \phi_1(x)= \eta_1(x)$, where ${\mathcal K}_x$ is the Klein-Gordon operator and $\phi_1(x)$ the interpolating field ($\omega_1=\sqrt{{\mathbf k}_1^2+m_1^2}$, is the energy of the on shell particle $P_1$). We use here and in what follows the fact that the single particle states $in$ and $out$ are identical. As shown in \cite{Oehme}-\cite{LSZ}, due to the factor $\theta(x_0)$ the amplitude (\ref{lsz1}) can be extended as an analytic function in the upper half of the complex plane of the time component $k_{10}$. A more detailed analysis \cite{KaWi}, exploiting also the causality properties of the retarded commutator and the relation $k_1^2=k_{10}^2 +{\mathbf k}^2_1$, shows that the amplitude $A (m_B^2, k_1^2, m_2^2)$ can be extended as an analytic function in the whole complex plane $k_1^2$, cut along a part of the real axis, where its discontinuity is \cite{Oehme} \begin{equation}\label{spectr} {\mathcal D}isc A(m_B^2, k_1^2, m_2^2)= {1\over 2\sqrt{2 \omega_1}}\int {\mathrm d}x {\mathrm e}^{ik_1 x} \langle P_2(k_2)|[\eta _1(x), {\mathcal H}_w(0)]|B(p)\rangle\,. \end{equation} This discontinuity coincides actually with the imaginary part $\hbox {Im} A(m_B^2, k_1^2+i\epsilon, m_2^2)$ of the decay amplitude on the upper edge of the cut (it will be shown below that this spectral function is real). The r.h.s. of (\ref{spectr}) is treated in the standard way by inserting a complete set of states in the two terms of the commutator. By performing the translation \begin{equation}\label{trans} \eta_1(x)={\mathrm e}^{iPx} \eta_1(0){\mathrm e}^{-iPx}\,, \end{equation} we write (\ref{spectr}) as \begin{eqnarray}\label{spectr1} &&\hbox{Im} A(m_B^2, k_1^2+i\epsilon, m_2^2)= {1\over 2 \sqrt{2 \omega_1}} \int {\mathrm d}x {\mathrm e}^{ik_1 x}\times \nonumber \\ &&\sum_{n}\left[{\mathrm e}^{i(k_2-p_n)x}\langle P_2(k_2)| \eta _1|n\rangle\langle n| {\mathcal H}_w|B(p)\rangle- {\mathrm e}^{i(p_n-p)x}\langle P_2(k_2)| {\mathcal H}_w |n\rangle\langle n|\eta _1|B(p)\rangle\right]\,, \end{eqnarray} where we denoted ${\mathcal H}_w={\mathcal H}_w(0)$ and $\eta _1=\eta _1(0)$. The trivial integral with respect to $x$ gives \begin{eqnarray}\label{spectr2} \hbox{Im} A(m_B^2, k_1^2+i\epsilon, m_2^2)&=& {1\over 2 \sqrt{2 \omega_1}} \sum_{n}\left[\delta (k_1+k_2-p_n) \langle P_2(k_2)| \eta _1|n\rangle\langle n| {\mathcal H}_w|B(p)\rangle \right. \nonumber \\ &&-\left.\delta (k_1+p_n-p)\langle P_2(k_2)| {\mathcal H}_w |n\rangle\langle n|\eta _1|B(p)\rangle\right]\,. \end{eqnarray} The states contributing to the first sum have the 4-momentum $p_n=k_1 +k_2=p$ and the invariant mass $p_n^2=m_B^2$, they correspond to what we called above the "rescattering" part of (\ref{unit}). It is easy to see that this sum is nonzero for $k_1^2$ in the allowed interval $0< k_1^2 < (m_B-m_2)^2$ (as we mentioned anomalous thresholds might be present). As the second sum in (\ref{spectr2}) is concerned, it gives a vanishing contribution, since $B$ is stable with respect to the strong interactions. We recall that by the reduction formula we obtained the analytic continuation of the physical matrix element with respect to the variable $k_1^2$ (the mass squared of an off-shell meson $P_1$). Therefore, the decay amplitude can be calculated from its discontinuity by means of a dispersion relation in this external mass variable. As discussed above, the integral extends along a finite interval, so the dispersion relation reads \begin{equation}\label{direl} A(m_B^2, m_1^2, m_2^2)= {1\over \pi} \int_0^{(m_B-m_2)^2}{\mathrm d}z {\hbox{Im} A(m_B^2, z+i\epsilon, m_2^2)\over z-m_1^2-i\epsilon}\,, \end{equation} with the discontinuity given by the first sum in Eq.~(\ref{spectr2}). Let us see now what is the form of the dispersion relation obtained in the LSZ formalism, when the analytic continuation is done with respect to the mass squared of the $B$ meson. To this end, we start again from the matrix element $\langle P_1(k_1) P_2(k_2), out|{\mathcal H}_w(0)|B(p)\rangle$, and apply the LSZ formula, reducing this time the initial meson $B$. We obtain \begin{equation}\label{lsz2} A (p^2, m_1^2, m_2^2)= {i\over \sqrt{2 \omega_B}}\int {\mathrm d}x {\mathrm e}^{-ipx} \theta (-x_0) \langle P_1(k_1)P_2(k_2), out|[ {\mathcal H}_w(0), \eta _B(x)]|0\rangle\,. \end{equation} By exploiting the causality properties of the retarded commutator one can prove \cite{KaWi} that the amplitude can be extended as a real analytic function in the complex plane $s=p^2$, with the discontinuity across the real axis given by \begin{eqnarray}\label{spectrs} \hbox{Im} A(p^2+i\epsilon, m_1^2, m_2^2)&=& {1\over 2 \sqrt{2 \omega_B}} \sum_{n}\left[\delta (p-p_n) \langle P_1(k_1)P_2(k_2),out| {\mathcal H}_w|n\rangle\langle n| \eta_B|0\rangle \right. \nonumber \\ &&-\left.\delta (p_n)\langle P_1(k_1) P_2(k_2),out| \eta_B|n\rangle\langle n|{\mathcal H}_w|0\rangle\right]\,. \end{eqnarray} We obtained this result in the standard way \cite{Oehme}, replacing $i\theta(-x_0)$ in (\ref{lsz2}) by $1/2$, and inserting a complete set of states in the commutator. Actually, the second sum in (\ref{spectrs}) vanishes because the only intermediate states allowed have zero 4-momentum. In the first term, the allowed particles are those connected to the final state through a weak process and to $B$ through a strong transition, (we recall that the last matrix element vanishes when $B$ is on the mass shell). The lowest two particle state entering the unitarity sum is $B^*\pi$ which defines the normal unitarity threshold. The spectral function (\ref{spectrs}) enters a dispersion relation of the form (\ref{drel}) with respect to $s=p^2$. However, it is obvious that such a dispersion relation is not useful for estimating the rescattering effects in nonleptonic $B$ decays. As discussed above, in this way one does not describe the strong interactions in the final state, but rather the strong interactions in the initial state. The fact that a dispersion relation in the mass squared of the $B$ meson cannot describe final state rescattering effects is understood by simple qualitative arguments: in order to make the analytic continuation in the variable $s$ we must reduce the $B$ meson. Hence, the source $\eta_B$ and the weak hamiltonian ${\cal H}_w$ enter different matrix elements in the unitarity sum, and terms describing the weak decay of $B$ multiplied by strong scattering amplitudes cannot appear. Therefore the procedure applied in \cite{Blok}-\cite{Falk}, based on the analytic continuation in $s$ combined with a discontinuity containing rescattering terms is not consistent. As we pointed out above the unitarity sum (\ref{unit}) defines the spectral function in a dispersion relation with respect to the mass variable of one of the final mesons. The relations (\ref{spectr2}) and (\ref{direl}) are the main results we obtained in the frame of the standard LSZ formalism. A few comments about the above formulae are of interest. First, it is clear that one can repeat the procedure by reducing the meson $P_2$ instead of $P_1$. The corresponding expressions can be obtained easily from those given above by permutting the indices 1 and 2. The expressions seem different, but of course the results should be the same when a complete set of states is inserted in the unitarity sum. A more subtle question, which is also connected to the completeness of the set inserted in the unitarity sum is whether the discontinuity defined in (\ref{spectr2}) is real or complex. For $T$ (or $CP$) conserving interactions, the reality of the spectral function was proved a long time ago \cite{GoTr}, \cite{GoWa}. It turns out that the absorbtive part remains real even if the relevant terms in the weak hamiltonian are not $CP$ conserving. We take into account the fact that these terms have the form \begin{equation}\label{hweak} {\mathcal H}_w=\sum _{j}c_j {\cal O}_j\,, \end{equation} where $c_j$ are complex numbers and ${\cal O}_j$ are products of $V$ and $A$ currents. Consider the spectral function \begin{equation}\label{sigma} \sigma (z)= \hbox{Im} A(m_B^2, z+i\epsilon, m_2^2)\,, \end{equation} defined by the first sum in (\ref{spectr2}), and assume that a complete set of $in$ states is inserted in the unitarity sum. Following \cite{GoTr}, \cite{GoWa} (see also \cite{Bart}) we can express the two matrix elements in this sum as \begin{equation}\label{etapt} \langle P_2(k_2) |\eta_1| n, in\rangle = \langle P_2(k_2)|(PT)^{-1}(PT) \eta_1 (PT)^{-1}(PT)|n, in\rangle = \langle P_2(k_2)|\eta_1|n, out\rangle ^*\,, \end{equation} and \begin{equation}\label{hweakpt} \langle n, in | {\mathcal H}_w | B(p)\rangle= \langle n, in |(PT)^{-1}(PT) {\mathcal H}_w (PT)^{-1} (PT)| B(p)\rangle \label{hpt1}\, =\! \langle n, out|{\mathcal H}_w| B(p)\rangle^*\,. \end{equation} We used here the transformation properties of the $V$ and $A$ currents under $P$ and $T$ transformations and the fact that under space-time reversal the particles conserve their momenta, and the $in (out)$ states become $out (in)$ states, respectively. Moreover, the intrinsic parities of the states and the operators have a product equal to $+1$, and the matrix element are replaced by their complex conjugates, given the antiunitary character of the operator $T$. By using the relations (\ref{etapt}) and (\ref{hweakpt}) in (\ref{spectr2}) we obtain \begin{eqnarray}\label{reality} \sigma (z) &=& {1\over 2\sqrt{2\omega_1}}\sum_{n}\delta (k_1+k_2-p_n) \langle P_2(k_2) | \eta_1|n, in\rangle\langle n, in| {\mathcal H}_w| B(p)\rangle \,\nonumber\\ &=&{1\over 2 \sqrt{2\omega_1}}\left [\sum_{n}\delta (k_1+k_2-p_n) \langle P_2(k_2) | \eta_1|n, out\rangle\langle n, out| {\mathcal H}_w|B(p)\rangle \right ]^*= \sigma^*(z)\, \end{eqnarray} where the equivalence between the complete sets of $in$ and $out$ states in the definition of $\sigma (z)$ was taken into account. From (\ref{reality}) it follows that the discontinuity is manifestly real only if the intermediate states form a complete set. If the unitarity sum is truncated, this property is lost, since various terms have complex phases which do not compensate each other in an obvious way. As noticed in \cite{GoTr}, in order to maintain the proper reality condition at all stages of approximation, it is convenient to write the sum over the complete set of states $|n\rangle$ as a combination $1/2|n, in\rangle +1/2|n, out\rangle$. This prescription will be applied in Section 4 when discussing the $B\to \pi\pi$ decay. \section{Two-particle unitarity and Regge amplitudes} In this section we write down the dispersion relation (\ref{direl}) in the approximation that only two particle states are kept in the unitarity sum (\ref{spectr2}). Denoting by $\{P_3P_4\}$ the two meson intermediate states in this sum, the off-shell imaginary part of the decay amplitude $A_{B\to P_1P_2}$ required in the dispersion relation (\ref{direl}) reads \begin{eqnarray}\label{disc} \hbox{Im} A_{B\to P_1P_2}(m_B^2, z+i\epsilon, m_2^2) &=&{1\over 2}\sum _{\{P_3P_4\}}\int {{\mathrm d}^3{\mathbf k}_3\over(2\pi)^{3} 2\omega_3}{{\mathrm d}^3{\mathbf k}_4\over(2\pi)^{3}2\omega_4} (2\pi)^4\delta^{(4)}(p-k_3-k_4)\times\nonumber\\ &&A_{B\to P_3 P_4}(m_B^2, m_3^2, m_4^2)\, {\mathcal M}^*_{P_3 P_4 \to P_1 P_2}(s,t)\,. \end{eqnarray} In the sum we include the two particle states ${P_3 P_4}= { P_1P_2}$ defining the elastic channel, as well as ${P_3 P_4}\ne { P_1P_2}$ responsible for the inelastic scattering. Let us note that the {\it c.m.} energy is set up by the mass of the $B$ meson ($\sqrt{s}= m_B= 5.2\, \hbox{GeV})$, and the weak decay amplitudes $ A_{B\to P_3 P_4}$ are on shell, and independent on the Mandelstam variable $t$ (or the rescattering angle $\theta$). Therefore Eq.(\ref{disc}) can be written as \begin{equation}\label{sum} \hbox {Im} A_{B\to P_1P_2}(m_B^2,z+i\epsilon, m_2^2) =\sum _{\{P_3P_4\}} C^*_{P_3P_4;P_1P_2}(z) A_{B\to P_3 P_4}(m_B^2, m_3^2, m_4^2)\,, \end{equation} where \begin{equation}\label{coef} C_{P_3P_4;P_1P_2}(z) ={1\over 2}\int {{\mathrm d}^3{\mathbf k}_3\over(2\pi)^{3} 2\omega_3}{{\mathrm d}^3{\mathbf k}_4\over(2\pi)^{3}2\omega_4} (2\pi)^4\delta^{(4)}(p-k_3-k_4) {\mathcal M}_{P_3 P_4 \to P_1 P_2}(s,t)\,. \end{equation} These coefficients depend on the masses of all the particles participating in the rescattering process. To simplify the notation we indicate explicitly only the dependence on the off shell mass squared $z$ of the particle $P_1$. Following \cite{Dono1}-\cite{Falk} we adopt for the strong amplitudes ${\mathcal M}$ the parametrizations obtained from the Regge theory \cite{Coll} \begin{eqnarray}\label{regge} {\mathcal M}_{P_3 P_4;P_1 P_2}(s,t)=-\sum_{V=P, f, A_2, K_2^*...} \gamma^V_{P_3 P_4;P_1 P_2} (t){{\mathrm e}^{-i{\pi\alpha_V(t)\over 2}}\over \sin{\pi\alpha_V (t)\over 2}} \left({s\over s_0}\right)^{\alpha_V(t)}\,+\nonumber\\ \sum_{V=\rho, K^*...}i \gamma ^V_{P_3 P_4\to P_1 P_2}(t){{\mathrm e}^{-i{\pi \alpha_V(t)\over 2}}\over \cos{\pi\alpha_V(t)\over 2}}\left({s\over s_0} \right)^{\alpha_V(t)}\,, \end{eqnarray} where the first sum includes $C=1$ trajectories and the second one $C=-1$ trajectories. As usual we take $s_0\approx 1\,\hbox{GeV}^2$ and linear trajectories \begin{equation}\label{alpha} \alpha_V(t)=\alpha_0+\alpha' t\,, \end{equation} with the standard choices \cite{Coll} \begin{equation}\label{alphapom} \alpha_0=1.08 \,,\quad \alpha'=0.25~{\rm GeV}^{-2} \, \end{equation} for the Pomeron, and \begin{equation}\label{alphapart} \alpha_0=0.45\,,\quad \alpha'=0.94\,\hbox{GeV}^{-2}\, \end{equation} for all the other trajectories. The possible divergences occuring in the expression (\ref{regge}) for $t\ne0$ are avoided by taking $\alpha(t)\approx\alpha_0$ in the denominators. We shall therefore obtain \begin{eqnarray}\label{aprox} {\mathrm sin}{\pi\alpha_P(t)\over2}&\approx&1\nonumber\\ {\mathrm sin}{\pi\alpha_V(t)\over2}&\approx&{1\over\sqrt{2}},~~~ V=f,~A_2~K^*_2\nonumber\\ {\mathrm cos}{\pi\alpha_V(t)\over2}&\approx&{1\over\sqrt{2}},~~~ V=\rho,~K^*. \end{eqnarray} As far as the Regge residua $\gamma^V_{P_3 P_4;P_1 P_2} (t)$ are concerned, they are supposed to satisfy the factorization relation \begin{equation}\label{resid} \gamma^V_{P_3 P_4;P_1 P_2} (t)=\gamma_{P_3 P_1 V} (t)\gamma_{P_4 P_2 V} (t)\, , \end{equation} and their values at $t=0$ will be determined using the optical theorem and the phenomenological Regge-like parametrizations of the total cross sections \cite{DoLa},\cite{Part} (details will be given in the next Section). The $t$-dependence of these functions is, however, poorly known and we assume they are simply constants. In order to perform the integral (\ref{coef}), the Mandelstam variable $t$ is expressed in terms of the scattering angle $\theta$ \begin{equation}\label{t} t(z)=t_0(z)+2k_{12}(z)k_{34}\cos\theta\,, \end{equation} with \begin{eqnarray}\label{k12k34} t_0(z)&=&z+m_3^2-{(m_B^2+m_3^2-m_4^2)(m_B^2+z-m_2^2)\over2m_B^2} \nonumber\\ k_{12}(z)&=&{1\over2m_B}\sqrt{(m_B^2-z-m_2^2)^2- 4z m_2^2}\nonumber\\ k_{34}&=&{1\over2m_B}\sqrt{(m_B^2-m_3^2-m_4^2)^2- 4m_3^2m_4^2}\,, \end{eqnarray} where we indicated explicitely only the dependence on the variable $z$. With these kinematic variables the integration over the momenta ${\mathbf k}_3$ and ${\mathbf k}_4$ in (\ref{coef}) is straightforward, and the coefficients $C_{ P_3P_4; P_1P_2}$ can be expressed as \begin{equation}\label{coef1} C_{ P_3P_4; P_1P_2}(z)= \sum_{V} \gamma^V_{P_3 P_4;P_1 P_2} (0) \kappa^V_{P_3 P_4;P_1 P_2}\,, \end{equation} where \begin{equation}\label{kappa} \kappa^V_{P_3 P_4;P_1 P_2}(z)=\xi_V { k_{34} \over 16 \pi m_B} {\mathcal R}_V^{-1}(z) \left[{\mathrm e}^{{\mathcal R}_V(z)}-{\mathrm e}^{-{\mathcal R}_V(z)}\right] \exp\left[(\alpha_{0,V}+\alpha'_Vt_0) \left(\ln {m_B^2\over s_0}-i{\pi\over2}\right)\right], \end{equation} \begin{equation}\label{K} {\mathcal R}_V(z)= 2\alpha'_Vk_{12}(z) k_{34}\left(\ln{m_B^2\over s_0}-i{\pi\over2}\right)\, , \end{equation} and $\xi_V$ is a numerical factor equal to $-1$ for the Pomeron, $i\sqrt{2}$ for $C=-1$ trajectories and $-\sqrt{2}$ for $C=1$ physical trajectories. By inserting the expression (\ref{sum}) of the spectral function in the dispersion relation (\ref{direl}) and recalling that the decay amplitudes $A_{B\to P_3P_4}$ do not depend on $z$, we obtain (for simplicity we omit now the mass arguments when the amplitudes are on-shell) \begin{equation}\label{sumint} A_{B\to P_1P_2}=\sum_{\{P_3 P_4\}}\overline\Gamma_{P_3 P_4; P_1 P_2}A_{B\to P_3 P_4}\,, \end{equation} where \begin{equation}\label{bareta} \overline\Gamma_{P_3 P_4;P_1 P_2}=\sum_{V} \gamma^V_{P_3 P_4; P_1 P_2} (0) \bar\eta^V_{P_3 P_4;P_1 P_2}\, \end{equation} and \begin{equation}\label{coefint} \bar\eta^V_{ P_3P_4;P_1P_2}={1\over \pi} \int_0^{(m_B-m_2)^2}{\mathrm d}z{\kappa^{V*}_{ P_3P_4; P_1P_2}(z)\over z-m_1^2-i \epsilon}\,. \end{equation} For further use we also define \begin{equation}\label{eta} \Gamma_{P_3 P_4; P_1 P_2}=\sum_{V} \gamma^V_{P_3 P_4;P_1 P_2} (0) \eta^V_{P_3 P_4;P_1 P_2}\, \end{equation} and \begin{equation}\label{coefbar} \eta^V_{ P_3P_4;P_1P_2}={1\over \pi} \int_0^{(m_B-m_2)^2}{\mathrm d}z{\kappa^V_{ P_3P_4; P_1P_2}(z)\over z-m_1^2-i \epsilon}\,. \end{equation} Before considering applications, let us make a few comments about the approximations adopted when deriving the above formulae. First, we notice that the Regge expression (\ref{regge}) is valid for large $s$ and $t$ close to 0. The value $s=m_B^2$ satisfies this condition, but the values of $t$ appearing in the integral upon the scattering angle in (\ref{coef}) can be large, outside the range of validity of the Regge theory. However, the hadronic amplitudes decrease at large $|t|$, so the contribution of the large scattering angles in the unitarity integral is expected to be small and not very sensitive to the inaccuracy of the dynamical model. Another difficulty is related to the fact that the particle $P_1$ is off shell, and its mass $k_1^2=z$ becomes very large at the upper limit of integration in the dispersion relation (\ref{direl}). Here again, one of the assumptions for the validity of the Regge expression, namely $\sqrt{s}>> m_i$ \cite{Coll} is not met. However, this part of the integral, which is not correctly evaluated, brings a small contribution in the dispersion integral due to the denominator in (\ref{direl}) (this statement is true when the masses of the intermediate particles $P_3$ and $P_4$ are not too large). We emphasize that the main advantage of the formalism is that it provides, with no approximation, an algebraic relation involving only physical decay amplitudes. Indeed, as we mentioned, all the quantities $A_{B\to P_3 P_4}$ appearing in (\ref{sumint}) are on shell, the dynamical approximations affecting only the coefficients $\Gamma_{P_3P_4;P_1P_2}$. \section{Constraints on the amplitudes of $B^0\to\pi^+\pi^-$ decays} Unitarity and the dispersion relations were used in previous works \cite{Dono1}-\cite{Falk} in order to estimate the FSI corrections to the decay amplitudes calculated in an approximation which does not include strong rescattering (like, for instance, factorization). Such an evaluation was made in \cite{Falk} for the magnitude of final state interactions in $B^-\to \pi^-\bar K^0$. In the notations used in Section 3, it corresponds to $P_1P_2= \pi^-\bar K^0$, with the intermediate states $P_3P_4=\pi^0 K^-$ and $\eta K^0$ inserted in the unitarity sum. Then only physical trajectories ($V=\rho,~K^*$) contribute to (\ref{regge}), for which the intercept and the slope are given in (\ref{alphapart}). With values of the residua extracted from the phenomenological parametrization of the cross sections the authors of Ref. \cite{Falk} suggested a modification of the magnitude of the decay amplitude of the decay $B^-\to \pi^-\bar K^0$ by a factor of $10\%$. However, the coefficients $\Gamma_{P_3P_4;P_1P_2}$ appearing in a relation of the type (\ref{sumint}) were calculated in \cite{Falk} by using the rescattering absorbtive function in a dispersion relation with respect to the variable $s$ (see Ref. \cite{Falk}, Eq. 2.17). By performing the correct calculations with the same values of the parameters, we find that the coefficient $\Gamma_{\pi^0 K^-; \pi^-\bar K^0}$, for instance, is larger by a factor of about 2.5 compared to the value reported in \cite{Falk}. This shows that the correct treatment can modify the conclusions about the magnitude of FSI in $B\to\pi K$ decays. In the present paper we consider an other application of the dispersive formalism to the decay $B^0\to \pi^+\pi^-$. The time dependent $CP$ asymmetry in this decay is considered as one of the ways of extracting the angle $\alpha$ of the unitarity triangle \cite{GrLo}. However, the unknown strong phase difference between the tree and the penguin amplitudes of the process affects the accuracy of this determination. Additional theoretical constraints on these amplitudes would be very helpful for reducing the uncertainty of the method. As we shall show below, the dispersion relations can provide such a constraint. We investigate the problem by combining the relations (\ref{sum}) and (\ref{sumint}) derived above with isospin or $SU(3)$ symmetry \cite{Gron}-\cite{Char}. The idea is that by unitarity and dispersion relations we obtain a set of correlations between exact decay amplitudes, containing both weak and strong phases. By imposing in addition $SU(3)$ flavour symmetry, all the amplitudes can be expressed in terms of a small number of parameters, for which unitarity and the dispersion relations provide nontrivial constraints. Following \cite{Gron} we write most generally the amplitude of the decay $B^0\to \pi^+\pi^-$ as a sum of diagram contributions \begin{equation}\label{notation} A_{B^0\to \pi^+\pi^-}= -(A_T~{\mathrm e}^{i\gamma}+A_P~{\mathrm e}^{-i\beta}+A_P'~{\mathrm e}^{i\gamma}+A_E~{\mathrm e}^{i\gamma}+ A_{PA}~{\mathrm e}^{-i\beta})\,, \end{equation} where $A_T$, $A_P (A_P')$, $A_E$ and $A_{PA}$ denote the amplitudes of the tree, penguin, exchange and penguin annihilation diagrams, respectively. We indicated explicitely the weak phases defined as $\beta= {\mathrm Arg}(-V^*_{td}$) and $\gamma= {\mathrm Arg}(-V^*_{ub})$ \cite{Part}. As intermediate states ${P_3P_4}$ in the equations (\ref{disc}) and (\ref{sumint}) we keep $\pi^+\pi^-$ giving the elastic channel, as well as two meson states responsible for the soft inelastic scattering, e.g., $\pi^0\pi^0$, $ K^+ K^-$, $ K^0\bar K^0$, $\pi^0\eta_8$ and $\eta_8 \eta_8$ ($\eta_8$ is the $\eta,~\eta'$ superposition belonging to the $SU(3)$ octet). We notice that $\pi^0\eta_8$ will not contribute finally due to isospin conservation in the strong rescattering. Of course, besides these states, other inelastic channels, like for instance multipion states can contribute. The states $D^+ D^-$ and $D^0\bar D^0$ are not included because they contribute to the hard scattering \cite{Dono1}. Assuming SU(3) flavour symmetry we express the decay amplitudes $B\to P_3P_4$ of interest as \cite{Gron}-\cite{Char} \begin{eqnarray}\label{notations} &&A_{B^0\to \pi^0\pi^0}={1\over\sqrt{2}} (-A_C~{\mathrm e}^{i\gamma}+ A_P~{\mathrm e}^{-i\beta}+A_P'~{\mathrm e}^{i\gamma}+A_E~{\mathrm e}^{i\gamma})\,,\nonumber\\ &&A_{B^0\to K^+ K^-}= -A_E~{\mathrm e}^{i\gamma}\,,\nonumber\\ &&A_{B^0\to K^0\bar K^0}=A_P~{\mathrm e}^{-i\beta}+A_P'~{\mathrm e}^{i\gamma}\,,\nonumber\\ &&A_{B^0\to\pi^0\eta_{8}}=-{1\over\sqrt{3}}(A_P~{\mathrm e}^{-i\beta}+A_P'~{\mathrm e}^{i\gamma}-A_E~{\mathrm e}^{i\gamma}) \,,\nonumber\\ &&A_{B^0\to\eta_{8}\eta_{8}}={1\over3\sqrt{2}}(A_C~{\mathrm e}^{i\gamma} +A_P~{\mathrm e}^{-i\beta}+A_P'~{\mathrm e}^{i\gamma}+A_E~{\mathrm e}^{i\gamma} )\,, \end{eqnarray} where $A_T$, $A_P$, $A_P'$ and $A_E$ are the same as in (\ref{notation}) and $A_C$ denotes the amplitude of the tree colour suppressed diagrams. Due to the lack of detailed dynamical calculations, various phenomenological assumptions are made in the literature about the above amplitudes. The conservative bound $\vert A_P/A_T \vert <1$ for the ratio of the penguin and tree amplitudes is mentioned in \cite{Char} (a more specific estimate $\vert A_P/A_T \vert \approx 0.2$ is also quoted in this reference). The penguin annihilation amplitude $A_{PA}$ correspond to OZI-suppressed diagrams \cite{Char}, while $A_C$ and $A_E$ are colour suppresed by a factor of about $0.25$ with respect to the corresponding colour favoured amplitude. Finally, the two penguin amplitudes $A_P$ and $A_P'$ are assumed to satisfy $\vert A_P'/A_P \vert \approx 0.4$ \cite{Char}. Using these estimates, we assume as a first approximation that we can neglect in (\ref{notation}) and (\ref{notations}) the suppressed terms $A_P'$, $A_C$, $A_E$ and $A_{PA}$, keeping only the dominant amplitudes $A_T$ and $A_P$. The next step is to introduce the decay amplitudes $A_{B\to P_iP_j}$ discussed above in the dispersion relation (\ref{sumint}). We should recall however that, due to the truncation of the unitarity sum, the imaginary part of the decay amplitude $A_{B\to P_1P_2}$ obtained from (\ref{sumint}) (or equivalently from the unitarity relation (\ref{sum}) evaluated on-shell) might be not real. In order to avoid this situation we apply the procedure suggested in Ref.\cite{GoTr} which maintains the proper reality condition of the spectral function and simulates the effect of other inelastic channels. As discussed at the end of Section 2, this method amounts to insert in the unitarity sum the complete set of states $1/2|in\rangle+1/2|out\rangle$. We recall that this method was applied to include inelastic effects through complex phases in the dispersive analysis of the electromagnetic form factors \cite{Omnes}, \cite{Bart}. In our case this procedure yields, instead of (\ref{sumint}), the modified dispersion relation \begin{equation}\label{sumint1} A_{B\to P_1P_2} ={1\over 2}\sum _{\{P_3P_4\}} \Gamma_{P_3P_4;P_1P_2} A^*_{B\to P_3 P_4}+ {1\over 2}\sum_{\{P_3P_4\}} \bar\Gamma_{P_3P_4;P_1P_2} A_{B\to P_3 P_4}\,, \end{equation} where $\Gamma_{P_3P_4;P_1P_2}$ and $\overline\Gamma_{P_3P_4;P_1P_2}$ are defined in (\ref{eta}) and (\ref{bareta}), respectively. We notice that Eq.(\ref{sumint1}) can be splitted in two relations, one for the real part and another for the imaginary part of the decay amplitude. In particular, the relation giving the imaginary part is \begin{equation}\label{sum1} i A^*_{B\to P_1P_2}- iA_{B\to P_1P_2} =\sum _{\{P_3P_4\}} C_{P_3P_4;P_1P_2}(m_\pi^2)A^*_{B\to P_3 P_4}+ \sum_{\{P_3P_4\}} C^*_{P_3P_4;P_1P_2}(m_\pi^2) A_{B\to P_3 P_4}\,, \end{equation} and can be obtained also directly from the unitarity relation (\ref{sum}) evaluated on shell. Concerning the real part, it is obtained by taking the principal value of the dispersion integrals appearing in (\ref{coefint}) and (\ref{coefbar}). We describe now briefly the determination of the Regge residua $\gamma^V_{P_3 P_4;P_1 P_2}(0)$ which enters the expressions (\ref{bareta}) and (\ref{eta}) of the coefficients of the dispersion relation (\ref{sumint1}). We use the optical theorem and the Regge parametrization of the total hadronic cross-sections \cite{DoLa},\cite{Part}, which gives \begin{equation}\label{optical} {s_0\over s}~{\rm Im}{\mathcal M}_{f\to f}(s,0)\approx s_0\sigma_{tot} =s_0X\left({s\over s_0}\right)^{\alpha_P(0)-1}+s_0Y \left({s\over s_0}\right)^{\alpha(0)-1} \end{equation} where $s_0\approx 1 {\rm GeV}^2\approx{1\over0.38}{\rm mb}^{-1}$. The first term represents the Pomeron contribution, the second the contributions of all the other trajectories. By comparing (\ref{optical}) with the Regge parametrization (\ref{regge}) we obtain $$s_0X=\gamma^P_{f;f}\,,\,\,\,\, s_0Y=\sum_{V\ne P}\gamma^V_{f;f}\,.$$ For the Pomeron, which contributes to the elastic $\pi^+\pi^-$ channel, we assumed that the coupling constant is proportional to the number of quarks, taking as in \cite{Dono1} $\gamma^P_{\pi^+\pi^-;\pi^+\pi^-}=\left({2\over3}\right)^2s_0X_{NN}$. The residua of the physical trajectories (which in our case are: $\rho, f,f'K^*, K_2^*$ and $A_2$) were estimated by taking into account the factorization property (\ref{resid}) combined with the experimental data on several hadron-hadron scattering processes. We used the processes given in Table~\ref{t:table1}, for which we wrote the contributions of various trajectories as in \cite{Coll}. \begin{table} \begin{center} \begin{tabular}{c|l} \hline &\\ Process&Contributing trajectories\\ &\\ \hline &\\ $\pi^-p$&$P+f+f'-\rho$\\ $\pi^+p$&$P+f+f'+\rho$\\ $\bar p p$&$P+f+f'+\rho+\omega+\phi+A_2$\\ $\bar p n$&$P+f+f'-\rho+\omega+\phi-A_2$\\ $ p p$&$P+f+f'-\rho-\omega-\phi+A_2$\\ $p n$&$P+f+f'+\rho-\omega-\phi-A_2$\, \end{tabular} \end{center} \caption{Contributing trajectories to various hadronic processes.} \label{t:table1} \end{table} \noindent Using the $NN$ channels we write in particular \begin{eqnarray} &&s_0Y_{pp}=\gamma^2_{N\bar Nf}-\gamma^2_{N\bar N\rho}-\gamma^2_{N\bar N\omega} -\gamma^2_{N\bar N\phi}+\gamma^2_{N\bar NA_2}\nonumber\\ &&s_0Y_{pn}=\gamma^2_{N\bar Nf}+\gamma^2_{N\bar N\rho}-\gamma^2_{N\bar N\omega} -\gamma^2_{N\bar N\phi}-\gamma^2_{N\bar NA_2}. \end{eqnarray} Noticing that experimentally $Y_{pp}\approx Y_{pn}$ \cite{Part} we obtain \begin{equation} \gamma^2_{N\bar NA_2}\approx \gamma^2_{N\bar N\rho}\, \end{equation} and further \begin{equation} s_0(Y_{\bar pp}-Y_{\bar pn})=2\gamma^2_{N\bar N\rho}+2\gamma^2_{N\bar NA_2}\approx4\gamma^2_{N\bar N\rho}\,. \end{equation} Also, by replacing the contributions of $f$ and $f'$ with the octet member $f_8$, we obtain \begin{eqnarray}\label{Y} &&s_0(Y_{\pi^-p}-Y_{\pi^+p})=2\gamma_{\pi^+\pi^-\rho}~\gamma_{N\bar N\rho}\nonumber\\ &&s_0(Y_{\pi^-p}+Y_{\pi^+p})=2\gamma_{\pi^+\pi^-f_8}~\gamma_{N\bar Nf_8}\nonumber\\ &&s_0(Y_{pn}+Y_{pp}+Y_{\bar pn}+Y_{\bar pp})=4\gamma^2_{N\bar Nf_8}. \end{eqnarray} The coupling constants $\gamma^2_{\pi^+\pi^-\rho^0}$ and $\gamma^2_{\pi^+\pi^-f_8}$ can be easily calculated from these equations using the experimental values of the parameters $X$ and $Y$ for $\pi N$ and $NN$ scattering \cite{Part}. Other coupling constants we need are obtained from the previous ones by using $SU(3)$ symmetry, namely \begin{eqnarray} &&\gamma^2_{\pi^0\pi^-\rho^-}=\gamma^2_{\pi^+\pi^-\rho^0} \nonumber\\ &&\gamma^2_{\bar K^0\pi^-K^{*-}}={1\over2}\gamma^2_{\pi\pi\rho}\nonumber\\ &&\gamma^2_{\bar K^0\pi^-K_2^{*-}}={3\over2}\gamma^2_{\pi^+\pi^-f_8} \nonumber\\ &&\gamma^2_{\pi\eta_8A_2}=\gamma^2_{\pi^+\pi^-f_8}\,. \end{eqnarray} In Table~\ref{t:table2} we give the values of the Regge residua calculated with the aid of the above relations using the parameters of the total cross sections quoted in \cite{Part}, the coefficients $\kappa^V_{P_3P_4;P_1P_2}$ are defined in (\ref{kappa}), while the coefficients $\bar\eta^V_{P_3P_4;P_1P_2},~\eta^V_{P_3P_4;P_1P_2}$ are given by Eqs. (\ref{coefint}), and (\ref{coefbar}) respectively. We notice that the dominant contribution is brought by the elastic channel, and in particular by the Pomeron exchange. \begin{table} \begin{tabular}{c|c|c|c|c|c} \hline &&&&&\\ $P_3~P_4$&$V$&$\gamma^V_{P_3P_4}$&$\kappa^V_{P_3P_4;P_1P_2}$& $\eta^V_{P_3P_4;P_1P_2}$&$\bar\eta^V_{P_3P_4;P_1P_2}$\\ &&&&&\\ \hline $\pi^+\pi^-$&$P$&25.6&$-0.0089+0.0270 i$&$-0.0547+0.0768 i$& $-0.0007-0.0946 i$\\ $\pi^+\pi^-$&$\rho$ &31.4&$-0.0005-0.0015 i$&$0.0001-0.0051 i$&$-0.0029+0.0040 i$\\ $\pi^+\pi^-$&$f_8$&35.3&$-0.0015+0.0005 i$&$-0.0051-0.0001 i$&$-0.0040-0.0029 i$\\ $\pi^0\pi^0$&$\rho$& 31.4&$-0.0005-0.0015 i$&$0.0001-0.0051 i$&$-0.0029+0.0040 i$\\ $\bar K^0~K^0$&$K^*$&15.7&$-0.0003-0.0007 i$&$-0.0004-0.0030 i$&$-0.0019+0.0024 i$\\ $\bar K^0~K^0$&$K^*_2$&52.9&$-0.0007+0.0003 i$& $-0.0030+0.0004 i$&$0.0024-0.0019 i$\\ $\eta_8\eta_8$&$A_2$&$35.3$&$-0.0015+0.0005i$& $-0.0051-0.0001i$&$-0.0040-0.0029i$\\ \hline \end{tabular} \caption{The values of the Regge residua and of the coefficients entering the dispersion relation.} \label{t:table2} \end{table} With the numbers given in Table~\ref{t:table2}, the real and imaginary parts of the dispersion relation (\ref{sumint1}) can be completely evaluated, yielding two algebraic relations for the decay amplitudes written in (\ref{notation}) and (\ref{notations}). According to the discussion following these equations, we shall neglect in the first approximation the suppressed diagrams, keeping only the contribution of the tree and the penguin diagrams $A_T$ and $A_P$. Both these amplitudes have in principle strong rescattering phases, which we denote as $\delta_T$ and $\delta_P$, respectively. Let us write \begin{equation}\label{phase} {A_P\over A_T}=R{\mathrm e}^{i\delta}\,, \end{equation} where $R = \vert A_P/ A_T \vert$ and $\delta=\delta_P-\delta_T$. For simplicity, let us make also the additional assumption that the tree amplitude $A_T$ is real. Then this amplitude can be factored out from the dispersion relation (\ref{sumint1}), which gives the following two constraints for the ratio $R$ and the phase difference $\delta$ of the penguin and tree amplitudes describing the $B\to \pi^+\pi^-$ decay \begin{eqnarray}\label{final1} 1.1189 {\rm sin}(\gamma+0.721)+ R{\rm sin}(\delta-\beta+0.686)=0\,, \nonumber \\ -1.239{\rm sin}(\gamma-0.0403)+R{\mathrm sin}(\delta-\beta+1.452)= 0\, . \end{eqnarray} These equations can be explicitely solved as \begin{eqnarray}\label{final2} \delta &=& \beta +\epsilon(\gamma)\,,\nonumber\\ ~\nonumber \\ R &=& -1.1189{\sin{(\gamma + 0.721)} \over \sin{(\epsilon(\gamma) + 0.686)}}\,, \end{eqnarray} where \begin{equation}\label{final3} \epsilon(\gamma) = -\arctan \left[ {\sin{(\gamma - 0.0403)}\sin{0.686} +0.9027\sin{(\gamma + 0.721)}\sin{1.452} \over \sin{(\gamma -0.0403)}\cos{0.686} +0.9027\sin{(\gamma + 0.721)}\cos{1.452}} \right]\,. \end{equation} We recall that in these relations $\beta$ and $\gamma$ are the angles of the unitarity triangle which are expected \cite{Fleisch}-\cite{Parodi} to be in the ranges $0.17\leq\beta\leq0.52$ and $0.349\leq\gamma\leq2.79$. In Fig.\ref{fi:phase} and Fig.\ref{fi:ratio} we represent the strong phase difference $\delta$ as a function of $\gamma$ for two values of $\beta$ at the limits of the allowed intervals mentioned above, and the ratio $R$ as a function of $\gamma$, according to (\ref{final2}). One can see that despite the crude approximations we made, the results are qualitatively reasonable. We notice that the equation for $\delta$ is not restrictive for the weak angles, while the expected condition $R<1$ is satisfied only for a small range above $\gamma=2.4$. However, this somewhat intriguing limitation disappears if we relax the last approximation made above, namely that the tree amplitude $A_T$ is real. It can be easily seen that by allowing a nonzero $\delta_T$ in the dispersion relation we get two equations of the form (\ref{final1}), with $\gamma$ replaced by $\gamma+\delta_T$ and $\delta$ replaced by $\delta_P$. The two constraints similar to (\ref{final1}) involve now three parameters, $R$, $\delta_T$ and $\delta_P$. \section{Conclusions} In the present paper we investigated a recent treatment of the final state interactions in the nonleptonic $B$ decays, based on unitarity and dispersion relations \cite{Dono1}-\cite{Falk}. By considering the analytic continuation in the external mass variable in the frame of LSZ formalism, we established the connection between the dispersion variable and the part of the unitarity sum defining the spectral function. The strong rescattering part is shown to appear as a discontinuity in a dispersion relation in terms of the mass of one final particle. Our results prove that the dispersion relations written in \cite{Dono1}-\cite{Falk}, based on the analytic continuation in the mass of $B$, are not consistent. We derived the correct dispersion relation, and showed that it modifies the conclusions of \cite{Falk} on the magnitude of FSI effects in $B\to \pi K$ decay by a factor of approximately 2.5. We also applied the formalism to derive a theoretical constraint for the amplitudes of the $B^0\to\pi^+\pi^-$ decay. We included in the unitarity sum a few channels, connecting them by the $SU(3)$ symmetry \cite{GrLo},\cite{Gron} and took into account qualitatively the effect of higher inelastic channels by a procedure applied in the study of the electromagnetic form factors \cite{GoWa}, \cite{Omnes}. In spite of the various dynamical assumptions mentioned in the text our results (\ref{final2}) for the ratio $R$ and the strong relative phase $\delta$ of the penguin and tree amplitudes in terms of the weak angles (represented in Figs. \ref{fi:phase} and \ref{fi:ratio}) are qualitatively reasonable. As we mentioned, these results can be immediately modified to incorporate a nonzero strong phase for the tree amplitude. Also, the contributions of the suppresed diagrams neglected in the present analysis, as well as corrections to the exact $SU(3)$ symmetry, can be easily included in the dispersion formalism. A more complete analysis will be made in a future work. The results might be useful as additional constraints in the extraction of the angles of the unitarity triangle from the time dependent $CP$ asymmetry in the decay $B^0\to\pi^+\pi^-$. \vskip 0.5cm {\bf Acknowledgements:} Two of the authors (I. C. and L. M.) express their thanks to the Center of Particle Physics (CPPM) and the Center of Theoretical Physics of Marseille for hospitality. Useful discussions with the members of the ATLAS group of CPPM are gratefully acknowledged. This work was partly realized in the frame of the Cooperation Agreement between IN2P3 and NIPNE-Bucharest and the Agreement between CNRS and the Romanian Academy.
\section{Introduction} Localized solutions, often called solitons, play an increasingly important role in nonlinear field theories in two dimensions. Topological structures exist in particular in magnetic systems and have been studied extensively, both theoretically and experimentally \cite{slonc,baryak}. An easy-plane ferromagnet is described by the Landau-Lifshitz equation \begin{equation} \label{eq:lle} \dot{\nb n} = {\nb n} \times {\nb f}, \end{equation} $$ {\nb f} = \Delta {\nb n} - n_3\, \hat{\nb e}, \qquad \nb n^2 = 1. $$ The field $\nb n$ represents the local magnetization of the material. We shall study here the case of a two-dimensional medium so we assume $\nb n \!=\! \nb n(x,y,t)$. The dot denotes a time derivative, $n_3$ is the third component of $\nb n$, $\Delta$ is the Laplace operator and $\hat{\nb e} \!=\! (0,0,1)$ is the unit vector in the third direction. The normalization condition $\nb n\!=\!1$ imposed in the initial condition is preserved by the equations of motion. Static solutions of model (\ref{eq:lle}) are the well-studied vortices. Isolated vortices are spontaneously pinned objects that is no vortex in free translational motion can be found in a 2D ferromagnet. The same is true for any isolated object with nontrivial topology in a 2D ferromagnet \cite{pt}, the most well-known example being the magnetic bubbles in easy-axis ferromagnetic films \cite{slonc}. Coherently traveling solutions of (\ref{eq:lle}) have been found in \cite{semi}. They have the form of a vortex-antivortex pair and their velocity may take any value between zero and unity, which is the velocity of magnons in the system. We now turn to a different class of systems, namely antiferromagnets. The dynamics of the staggered magnetization in the antiferromagnetic continuum is given by the nonlinear $\sigma$-model \cite{baryak,halperin,afm}: \begin{equation} \label{eq:sigma} \nb n \times [\,\ddot{\nb n} - {\nb f}\,] = 0, \end{equation} $$ {\nb f} = \Delta \nb n - n_3\, \hat{\nb e}, \qquad \nb n^2 = 1, $$ where the double dot denotes a second time derivative. The above model has the same static vortex solutions as (\ref{eq:lle}). On the other hand, vortices in model (\ref{eq:sigma}) can be found in free translational motion. This is due to the fact that the model is invariant under Lorentz transformations. Our main purpose is to study collisions of solitons in models (\ref{eq:lle}) and (\ref{eq:sigma}). In the ferromagnet no collision between two vortices can take place. Two vortices with the same topological charge will rotate around each other while a vortex and an antivortex undergo Kelvin motion perpendicular to the line connecting them. However, collisions can occur between two vortex-antivortex pairs. We elaborate on the arguments of \cite{semi,cooper} and argue that head-on collisions between vortex-antivortex pair solitons give a right angle scattering pattern. We also study collisions between vortices in antiferromagnets in Eq.~(\ref{eq:sigma}). They scatter at right angles as found in numerical simulations. In fact, the right angle scattering phenomenon seems to be a robust feature in various two-dimensional models which have soliton solutions \cite{zakr,theodora}. However, it is a nontrivial and strange behaviour at least from the point of view of scattering of ordinary particles. The colliding objects in the two models that we study are essentially different from each other. While vortices within model (\ref{eq:sigma}) are topologically nontrivial objects, the colliding vortex-antivortex pairs in (\ref{eq:lle}) have a vanishing topological charge. However, we argue that the underlying Hamiltonian structure allows to study the soliton interaction in the two models in close analogy. The outline of the paper is as follows. In Section II we simulate head-on collisions of vortex-antivortex pairs in (\ref{eq:lle}) and give a theoretical description which exploits the form of the linear momentum. In Section III the results of head-on collision simulations of vortices in model (\ref{eq:sigma}) are given together with arguments for the understanding of this behaviour. The conclusions are given in Section IV. \section{Head-on collisions of vortex-antivortex pairs in planar ferromagnets} A ferromagnet can be described in terms of a magnetization vector which satisfies the Landau-Lifshitz equation (\ref{eq:lle}). The constraint on the field $\nb n$ can be resolved and the theory can be formulated in terms of a complex variable \begin{equation}\label{eq:omegadef} \Omega = \Omega(x,y,t) = {n_1 + i\, n_2 \over 1 + n_3} \end{equation} which satisfies the equation \begin{equation}\label{eq:omegaeq} i\, \dot{\Omega} = - \Delta \Omega + {2\, \overline{\Omega} \over 1 + \Omega \bOmega}\; \partial_\mu \Omega \; \partial_\mu \Omega - {1-\Omega \bOmega \over 1 + \Omega \bOmega}\; \Omega. \end{equation} $\overline{\Omega}$ denotes the complex conjugate of $\Omega$. We use the formulation through the complex variable $\Omega$ in all numerical simulations. We avoid the formulation through the vector variable $\nb n$ since the constraint on it makes an accurate computer calculation of the time derivatives of the field rather cumbersome. The model has some interesting static vortex solutions of the form \begin{equation} \label{eq:staticvortex} \Omega^o = f(\rho)\, e^{i\kappa\phi}, \qquad \kappa = \pm 1, \end{equation} where $\rho, \phi$ are polar coordinates and $f(\rho=0)=0, f(\rho \rightarrow \infty) \rightarrow 1$. We call the configuration with $\kappa\!=\!1$ a vortex and the one with $\kappa\!=\!-1$ an antivortex. Vortex solutions have infinite energy and it has been argued that they are physically relevant \cite{gross,afm}. In the study of the dynamics of magnetic vortices the central role is played by a scalar quantity called the local vorticity \cite{pt,afm} \begin{equation} \label{eq:vorticitydef} \gamma = \varepsilon_{\mu\nu}\, \partial_\mu \pi \, \partial_\nu \psi, \end{equation} where $\varepsilon_{\mu\nu}$ is the two-dimensional totally antisymmetric tensor. The two components of the linear momentum are then expressed as \begin{equation}\label{eq:momentum} p_x = - \int{y\, \gamma\; dx dy}, \qquad p_y = \int{x\, \gamma\; dx dy}. \end{equation} Of fundamental importance is the Poisson bracket relation between the two components of the linear momentum. This reads \begin{equation}\label{eq:poissonmomentum} \{ p_x, p_y \} = \Gamma, \end{equation} \begin{figure} \begin{center} \psfig{file=fig1.ps,width=7.0cm} \end{center} \vspace{-5pt} \caption{ A simple vorticity distribution of a soliton consists of two lumps with opposite sign. They are here symmetrically placed on either side of the $x$-axis. The lower shaded area represents positive vorticity while the upper shaded area represents negative vorticity. P denotes the linear momentum of the pair. } \label{fig:prototype} \end{figure} \vspace{10pt} \noindent where \begin{equation}\label{eq:totalvorticity} \Gamma = \int{\gamma \,dx dy} \end{equation} is the total vorticity. In the present model $\pi\!=\!\cos\!\Theta$ and $\psi\!=\!\Phi$ have been used \cite{pt} as the canonical fields. They are defined through $n_1\!=\!\cos\!\Theta \sin\!\Phi,\, n_2\!=\!\cos\!\Theta \sin\!\Phi, \, n_3\!=\!\sin\!\Theta$. The explicit form of the vorticity is \begin{equation} \label{eq:vorticitylle} \gamma = \varepsilon_{\mu\nu}\, \sin\!\Theta\;\partial_\nu\Theta \, \partial_\mu\Phi. \end{equation} The total vorticity of a vortex is \begin{equation} \label{eq:totalvorticitylle} \Gamma = \int{\gamma \,dx dy} = -2\,\pi\kappa, \end{equation} that is, $\Gamma\!=\!-2\,\pi$ for vortices and $\Gamma\!=\!2\,\pi$ for antivortices. The implications of a nonvanishing total vorticity to the dynamics is an issue which has been thoroughly studied in the case of magnetic vortices and bubbles \cite{pt,afm,thiele} as well as in the case of vortices in other interesting models \cite{gle,manton}. The most striking result is that it leads to spontaneous pinning of these topological objects. Since a nonvanishing total vorticity $\Gamma$ implies pinning of an object, we infer that a solution which moves freely should have a vanishing $\Gamma$. In this respect, the vortex-antivortex ansatz offers the simplest possibility. Fig.~\ref{fig:prototype} gives a schematic representation of it. This consists of two lumps, one having negative and the other one positive sign. We suppose that the vortex is roughly laying in the shaded area with the negative sign and the antivortex in the shaded area with the positive sign. This figure is supposed to act only as a guide for our discussion and there is no strict way to distinguish the two vortices and define their positions. However, if relation (\ref{eq:poissonmomentum}) is applied to each vortex separately then the quantity on the right hand side is nonvanishing. It is then implied that each vortex will propagate in the horizontal direction under the influence of the other vortex. The picture is consistent with linear momentum considerations. That is, an application of Eq.~(\ref{eq:momentum}) to the full ansatz gives a nonvanishing $x$-component of the linear momentum. Fig.~\ref{fig:prototype} will serve in the following discussion as a prototype and will motivate our theoretical arguments. The Kelvin motion of a vortex-antivortex pair in a ferromagnet has been investigated in \cite{mertens}. The motion of a bubble-antibubble ansatz has also been studied \cite{papzakr}. The situation is found to be similar in some other systems such as an antiferromagnet immersed in a uniform magnetic field \cite{afm}, a model for superconductors \cite{stratos}, for superfluid helium \cite{roberts} and the nonlinear Schr\"odinger equation \cite{staliunas}. It has been pointed out that the gross features of this dynamical behaviour are analogous to the planar motion of charges under the influence of a magnetic field perpendicular to the plane. In particular, two oppositely charged particles undergo Kelvin motion traveling along parallel trajectories. The analogy has been made precise by use of relation (\ref{eq:poissonmomentum}) and an analogous relation in the charge motion problem \cite{pt}. The calculation of steadily moving coherent structures in a 2D ferromagnet, which have the form of a vortex-antivortex pair has been done in \cite{semi}. We have used here the numerical code of \cite{semi} to reproduce them since there is no available analytical formula. Fig.~\ref{fig:fmsoliton} is an example contour plot for a soliton with velocity $v\!=\! 0.5$. The upper entry gives contour plots of the quantity 10 $|\Omega|$ and the two-vortex character of the configuration is rather obvious. The lower entry is a contour plot for the local vorticity (\ref{eq:vorticitylle}). An important result of the analysis in \cite{semi} is that the velocity is collinear with the linear momentum. We are now sufficiently motivated to explore the possibility of scattering of vortex-antivortex pair solitons. We denote by $\Lambda_v(x,y)$ the solution with velocity $v$ along the x-axis (set $t\!=\!0$). The product ansatz \begin{equation}\label{eq:ansatzlle} \Omega (x,y) = \Lambda_v \left(x\!+\!{\delta \over 2},y\right) \; \Lambda_{-v} \left(x\!-\!{\delta \over 2},y \right) \end{equation} represents two vortex-antivortex pair solitons at a distance $\delta$ apart which are in a head-on collision course. The ansatz (\ref{eq:ansatzlle}) is used as an initial condition in a straightforward numerical integration of Eq.~(\ref{eq:omegaeq}). We typically set $v\!=\!0.5$, $\delta \!=\! 10$. We have set up a numerical mesh as large as 600$\times$600 with uniform lattice spacing $h\!=\!0.1$. The space derivatives are calculated by finite differences and the time integration is performed by a fourth order Runge-Kutta method. The results are presented in Figs.~\ref{fig:fmomega} and \ref{fig:fmvorticity}. Fig.~\ref{fig:fmomega} presents a contour plot for the field 10 $|\Omega|$ at three characteristic snapshots. In the first entry the initial ansatz (\ref{eq:ansatzlle}) is shown. The second snapshot is taken when the solitons are more or less at a minimum separation. No vortex-antivortex annihilation process takes place. This behaviour should be expected since the vortex-antivortex pair solitons are stable solutions of the equation. The argument is supported by numerical simulations showing that a vortex-antivortex ansatz preserves its character when traveling, provided that the vortex and antivortex are not very close to each other \cite{mertens,afm}. The last snapshot shows the system after the collision. A right angle scattering pattern has been produced. \vspace{10pt} \begin{figure} \begin{center} \psfig{file=fig2.ps,width=6.5cm,bbllx=170bp,bblly=236bp,bburx=390bp,bbury=690bp} \end{center} \caption{ Contour plot for a vortex-antivortex pair soliton in a ferromagnet with velocity $v\!=\! 0.5$. The upper entry gives contour plots of the quantity $10\,|\Omega|$. We plot the levels 1, 3, 5, 7, 9. The lower entry is a contour plot of the local vorticity for the same soliton. Solid lines represent positive values and dashed-dotted lines negative values of vorticity. We plot the levels $\pm 0.1, \pm 0.2, \pm 0.4, \pm 0.8, \pm 1.2$. } \label{fig:fmsoliton} \end{figure} \begin{figure} \begin{center} \psfig{file=fig3.ps,width=6.0cm,bbllx=170bp,bblly=65bp,bburx=390bp,bbury=775bp} \vspace{30pt} \caption{ Contour plot of the field $10\,|\Omega|$ at three characteristic snapshots of the head-on collision simulation of vortex-antivortex pair solitons in ferromagnets. It is shown: the initial ansatz (upper entry, time $t\!=\!0$), a snapshot at the time of collision (middle entry, $t\!=\!6.6$), and well after collision (lower entry, $t\!=\!13.2$). Contour levels as in Fig.~\ref{fig:fmsoliton}. } \label{fig:fmomega} \end{center} \end{figure} \begin{figure} \begin{center} \psfig{file=fig4.ps,width=6.0cm,bbllx=170bp,bblly=65bp,bburx=390bp,bbury=775bp} \vspace{30pt} \caption{ Contour plot of the local vorticity $\gamma$ of Eq.~(\ref{eq:vorticitylle}) for the solitons of Fig.~\ref{fig:fmomega}. Contour levels as in Fig.~\ref{fig:fmsoliton}. } \label{fig:fmvorticity} \end{center} \end{figure} \vfill \eject \begin{figure} \begin{center} \psfig{file=fig5.ps,width=7.0cm,angle=-90,bbllx=160bp,bblly=230bp,bburx=420bp,bbury=505bp} \vspace{25pt} \caption{Stars denote the zeros of $\Omega$ during the numerical simulation of Fig.~\ref{fig:fmomega}. We also trace the maximums (circles) and minimums (diamonds) of the vorticity. The solitons are initially located at $AB$ and $CD$ respectively. Symbols are plotted every 0.4 time units. } \label{fig:fmorbit} \end{center} \end{figure} The situation becomes clearer in Fig~\ref{fig:fmvorticity} where we represent the solitons in terms of their local vorticity distribution. The soliton on the left half plane should be compared directly with that in the lower entry of Fig.~\ref{fig:fmsoliton}. It obviously has a linear momentum and velocity pointing to the positive $x$-direction. The other soliton has the opposite linear momentum and velocity. It is rather clear from the picture that the solitons will not bounce back after collision. This is precluded by the form of the local vorticity distribution. This possibility would require that the vortex and antivortex interchange their position. Instead, the possibility appears that, at collision time, the two pairs of vorticity lumps in the upper and lower half-planes will form two new vortex-antivortex pairs. One has to apply Eq.~(\ref{eq:poissonmomentum}) for each of the vorticity lumps separately. Alternatively, one can consider pairs of lumps which tend to travel parallel to each other undergoing Kelvin motion. Application of this idea to Fig.~\ref{fig:fmvorticity}, determine the time evolution of the system. Finally, the two pairs on the upper and lower half planes tend to travel parallel to each other along the $y$-axis and form bound states. An equivalent point of view is to follow the linear momentum of each soliton separately. The linear momentums of the outgoing solitons clearly lay on the $y$-axis and have opposite sign. A subtle but important question is whether we can apply Eq.~(\ref{eq:poissonmomentum}) separately for each of the vortices which consist the vortex-antivortex pair. A rigorous answer can not be given here. On the other hand the construction of solitary waves in \cite{semi,cooper,roberts} suggests an affirmative answer whose range of validity is interesting to study. The two solitons emerging after collision are very similar to the initial ones though not exactly the same. In fact the drift velocity of the outgoing solitons is somewhat larger. In Fig.~\ref{fig:fmorbit} we have traced, during the time evolution, the points where the complex field $\Omega$ vanishes and also the points where the vorticity $\gamma$ attains its maximum and minimum values. The two kinds of extrema are close during the whole period of time evolution. This is because the solitons used in the simulations of this chapter have a pronounced vortex-antivortex character. In \cite{piette} traveling solutions of the Landau-Lifshitz equation have been studied which are different than the ones used here. Numerical simulations of scattering of these solitons also produce a right angle pattern. It is possible to give a picture, corresponding to the scattering of vortex-antivortex pairs, in terms of 2D motion of charged particles interacting via their electric field and placed in a magnetic field perpendicular to the plane. In fact, we have to consider two electron-positron pairs. Consider the first electron-positron pair located at points $A, B$ of Fig.~\ref{fig:fmorbit}. and the second pair at $C, D$. The charges move similar to the vortices. Their actual orbits resemble those for the solitons shown in Fig.~\ref{fig:fmorbit}. Our last remark in this section goes to some related work in hydrodynamics. There are solutions of the two-dimensional Euler equations which describe vorticity dipoles. Note that, in this context, vorticity has its ordinary hydrodynamic meaning. The best-known such solution seems to be the Lamb dipole \cite{lamb}. Another one has been found in \cite{rasmussen}. A head-on collision between two dipoles produces a pattern analogous to that in Fig.~\ref{fig:fmvorticity} of the present paper \cite{rasmussen,orlandi}. Furthermore, a simple construction is given in \cite{lamb} page 223, to which our Fig.~\ref{fig:fmorbit} can be compared. Further interesting cases of scattering between pairs of objects in hydrodynamics have been studied. The most complex behaviour has been observed in \cite{shchur} and includes stochastic and quasiperiodic motion of vortices. \section{Head-on collisions of vortices in antiferromagnets} Our main objective is to study scattering of vortices within model (\ref{eq:sigma}) and to show that the process can be studied in close analogy to the corresponding phenomenon in the ferromagnet. A right angle scattering behaviour of solitons has been observed in the isotropic $\sigma$-model \cite{zakr}, that is model (\ref{eq:sigma}) without the anisotropy term. The examination of the local vorticity $\gamma$ has led to a successful approach for the collision of vortex-antivortex pairs in ferromagnets in Section II. We find it instructive to look at the collision process in terms of the vorticity also in the present model. A simple generalization of definition (\ref{eq:vorticitydef}) can be used \cite{afm}. The vorticity attains a simple form when it is expressed in terms of the vector field $\nb n$: \begin{equation}\label{eq:vorticitysigma} \gamma = \varepsilon_{\mu\nu}\; \partial_\mu \dot{\nb n} \cdot \partial_\nu \nb n = \varepsilon_{\mu\nu}\; \partial_\mu (\dot{\nb n} \cdot \partial_\nu \nb n). \end{equation} In \cite{afm} the equation for an antiferromagnet in a uniform magnetic field was studied. Vortices in this system are spontaneously pinned, thus their dynamics is analogous to that of ferromagnetic vortices. This unexpected behaviour is probed by a topological term which enters the vorticity. However, such a term is absent in the model studied in this section. The vorticity (\ref{eq:vorticitysigma}) has the form of a total divergence and can be integrated in all space to show that the total vorticity vanishes for solutions with reasonable behaviour at infinity: \begin{equation}\label{eq:totalvorticitysigma} \Gamma = \int{\gamma\; dx dy} = 0. \end{equation} In particular, it vanishes for the vortex solutions. Relations (\ref{eq:momentum}) - (\ref{eq:totalvorticity}) apply in the present context without modification and they will be the fundamental relations to be used in the following analysis. Eq.~(\ref{eq:vorticitysigma}) shows that for a static vortex $\gamma$ vanishes identically. On the other hand, we can obtain a steadily traveling vortex by applying a Lorentz transformation to the static vortex solution (\ref{eq:staticvortex}). We denote the traveling vortex by $\Omega_v^o$ and the velocity is $0< v <1$. The distribution of $\gamma$ for a Lorentz boosted vortex is nonvanishing and can be calculated numerically. The vortex with velocity $v\!=\!0.7$ is represented by a contour plot of the field $10 |\Omega|$ in the upper entry of Fig.~\ref{fig:vortex}. A corresponding plot for $\gamma$ is given in the lower entry of the figure. The vorticity distribution has the form of two lumps, thus it resembles the sketch of Fig.~\ref{fig:prototype}. This is no surprise. In fact the following two remarks make it plausible. Firstly, we see that the total vorticity vanish according to relation (\ref{eq:totalvorticitysigma}). Secondly, an inspection of the form (\ref{eq:momentum}) of the linear momentum makes it clear that a nonvanishing component is furnished by two lumps of vorticity with opposite signs. This is not the only form of local vorticity that furnishes a nonvanishing linear momentum but it is certainly the simplest. Since the vortex solution with $\kappa\!=\!1$ is indeed the one with the simplest topological complexity, we expect its vorticity distribution to have the simplest possible form. We calculate numerically the points where the maximum and minimum of the vorticity lumps are located. It turns out that these points are the $(0,\pm 0.59)$ for any value of the velocity $v$. A further example on the present ideas is offered by the Belavin-Polyakov solutions \cite{bp}. We apply a Lorentz transformation, with velocity $v$, to the simplest one: $ \Omega = {(x - v t)/ \sqrt{1-v^2}} + i y. $ Its local vorticity is \begin{equation}\label{eq:bpvorticity} \gamma = - {16\; v \over 1- v^2}\; \; {y \over \left( 1 + {(x-v t)^2 \over 1-v^2} +y^2 \right)^3}. \end{equation} In accordance with the above remarks, it has the shape of two lumps with opposite sign, located on either side of the $x$-axis and traveling along the $x$-axis. \begin{figure} \begin{center} \psfig{file=fig6.ps,width=6.5cm,bbllx=170bp,bblly=236bp,bburx=390bp,bbury=690bp} \end{center} \vspace{5pt} \caption{Contour plot for a traveling vortex with velocity $v\!=\!0.7$. The upper entry gives contours of the field $10\,|\Omega|$. The lower entry is a contour plot of the local vorticity $\gamma$. Solid lines represent positive values and dashed-dotted lines represent negative values of vorticity. Contour levels as in Fig.~\ref{fig:fmsoliton}. } \label{fig:vortex} \end{figure} \vspace{10pt} We are now ready to present numerical simulations of head-on collisions of vortices. We make an ansatz representing two vortices. The choice is not unique and the simplest one seems to be the product ansatz: \begin{equation} \label{eq:ansatzsigma} \Omega(x,y) = \Omega^o \left( x\!+\!{\delta \over 2},y \right)\;\; \Omega^o \left( x\!-\!{\delta \over 2},y \right), \end{equation} where $\Omega^o$ is the single vortex solution given in Eq.~(\ref{eq:staticvortex}). The two vortices are a distance $\delta$ apart. The numerical mesh as well as the details of the algorithm that we use here are similar to those of Section II. We use vortices with $\kappa\!=\!1$. They are initially at rest but immediately start to drift away from each other due to their mutual repulsion and escape to infinity. In order to invoke a head-on collision between vortices we consider the product ansatz of two vortices which have opposite velocities: \begin{equation}\label{eq:ansatzsigma2} \Omega(x,y) = \Omega_v^o \left(x+{\delta \over 2},y \right)\;\; \Omega_{-v}^o \left(x-{\delta \over 2},y \right). \end{equation} $\Omega_v^o(x,y)$ denotes the Lorentz transformed vortex solution with velocity $v$, at $t\!=\!0$. We typically use $\delta\!=\!6$. In all simulations the vortices start to move against each other with velocities close to the value $v$ but they immediately begin to decelerate due to their mutual repulsion. The future of the process depends crucially on the magnitude of the velocity. At low velocities the two vortices approach to a minimum distance at which they come to rest and then turn round and move off in opposite directions. When the velocity exceeds a critical value (which is $v_c \!\approx\! 0.65$ for $\delta\!=\!6$) the two vortices collide and scatter at right angles. This result does not depend on the details of the initial ansatz or on the initial velocity of the vortices, as long as this exceeds the critical value. We have tested our algorithm for velocities up to the value $v\!=\!0.9$ in ansatz (\ref{eq:ansatzsigma2}). However, one must keep in mind that the velocity of the vortices at the time of collision is smaller than the velocity in the initial ansatz. In Fig.~\ref{fig:sigmaomega} we give a contour plot for the norm of the field $\Omega$ at three characteristic snapshots. The first entry presents the initial configuration (\ref{eq:ansatzsigma2}). In the middle snapshot, taken at collision time, it is clear that the two vortices come on top of each other. There is no topological reason, related to the field $\Omega$, that could prevent this double vortex to form and there is also no such reason that could prevent the vortices either to continue traveling in the horizontal direction or to reemerge traveling in the vertical direction. We add that, at the present level of description, we can find no reason that would enforce them to follow either of the two possibilities. In the last snapshot the two new vortices that emerge after the collision, are drifting away from each other along the $y$-axis. We find it instructive to look at the collision process using the vorticity. Our description will closely follow that in Section II in connection with the scattering of vortex-antivortex pairs. The dynamics in both systems is determined by the corresponding vorticity distribution. A comparison of the lower entries of Figs.~\ref{fig:fmsoliton} and \ref{fig:vortex} gives a hint that the underlying dynamics should be of a similar nature in both models. In Fig.~\ref{fig:sigmavorticity} we give the vorticity at three snapshots which correspond to those of Fig.~\ref{fig:sigmaomega}. Fig.~\ref{fig:sigmavorticity} should be compared directly with Fig.~\ref{fig:fmvorticity}. An examination of these results shows that the arguments of Section II for the soliton scattering which rely upon the linear momentum relations (\ref{eq:momentum}), (\ref{eq:poissonmomentum}) are applicable here, too. \begin{figure} \begin{center} \psfig{file=fig7.ps,width=6.0cm,bbllx=170bp,bblly=65bp,bburx=390bp,bbury=775bp} \end{center} \caption{Contour plot of the field $10\,|\Omega|$ at three characteristic snapshots for the head-on collision simulation of vortices in an antiferromagnet. It is shown: the initial ansatz (upper entry, time $t\!=\!0$), a snapshot at the time of collision (middle entry, $t\!=\!4.2$), and well after collision (lower entry, $t\!=\!8.4$). Contour levels as in Fig.~\ref{fig:fmsoliton}. } \label{fig:sigmaomega} \end{figure} \begin{figure} \begin{center} \psfig{file=fig8.ps,width=6.0cm,bbllx=170bp,bblly=65bp,bburx=390bp,bbury=775bp} \end{center} \caption{Contour plot of the local vorticity $\gamma$ of Eq.~(\ref{eq:vorticitysigma}) for the vortices of Fig.~\ref{fig:sigmaomega}. Contour levels as in Fig.~\ref{fig:fmsoliton}. } \label{fig:sigmavorticity} \end{figure} The upper entry in the figure corresponds to the initial ansatz and each of the two vorticity dipoles should be compared to that given in the lower entry of Fig.~\ref{fig:vortex}. The two vortices in the first entry of Fig.~\ref{fig:sigmavorticity} are approaching each other while their dynamical features, as described by the vorticity, are not substantially modified. The repulsion which could decelerate them and make them turn round, is overcompensated by the large enough initial velocity. When the two vortices come close to each other (middle entry) the two pairs of vorticity lumps, lying in the upper and lower half plane, interact. The subsequent evolution of the vorticity lumps is governed by relation (\ref{eq:poissonmomentum}). In particular, this relation has to be applied to each of the vorticity lumps separately. As is indicated by the simulation, the lumps survive throughout the process and the simple dynamics implied by (\ref{eq:poissonmomentum}) is sustained during the scattering process. Following the discussion in Section II we examine the dynamics of lumps in a pairwise manner. The two pairs in the upper and lower plane tend to move along the $y$-axis. Consequently, the initial partners separate and two new vortices are formed which travel in opposite directions along the $y$-axis, as shown in the lower entry of the figure. An equivalent approach is to follow the linear momentum of the solitons. The two pairs in the lower and upper half-plane have their linear momentums lying along the $y$-axis but with opposite signs. They subsequently tend to go off along the $y$-axis. The important numerical result is that the vorticity lumps roughly preserve their shape throughout the process. This is due to the fact that traveling vortices are stable solutions of the model. We note here that our result is obtained when Eq.~(\ref{eq:poissonmomentum}) is applied to each vorticity lump separately. We have been motivated to follow this approach because of its success with respect to studying the dynamics of vortex-antivortex pairs in a ferromagnet and because of the consistency of the picture with the linear momentum considerations. On the other hand a rigorous proof of its validity is lacking. \begin{figure} \begin{center} \psfig{file=fig9.ps,width=7.0cm,angle=-90,bbllx=160bp,bblly=230bp,bburx=420bp,bbury=505bp} \end{center} \caption{Stars denote the zeros of $\Omega$ during the numerical simulation of Fig.~\ref{fig:sigmaomega}. We also trace the maximums (circles) and minimums (diamonds) of the vorticity. The vortices are initially located at points $A$ and $B$. Symbols are plotted every 0.4 time units. } \label{fig:sigmaorbit} \end{figure} \vspace{10pt} In Fig.~\ref{fig:sigmaorbit} we track some characteristic points of the vortices throughout the simulation time. The stars denote successive points where the centers of the two vortices lie during the simulation, that is, where the complex function $\Omega$ vanishes. The circles denote successive locations of the maximum and the diamonds the locations of the minimum of the vorticity distribution. We plot two circles and two diamonds at every time instant. At the beginning of the simulation the two vortices are centered at points $A$ and $B$ respectively. They immediately start to decelerate but when their centers reach a distance $\approx 2.4$ they seem to accelerate considerably, they eventually merge at the origin and then they separate along the $y$-axis. The trajectories of the extrema of the vorticity show that after the collision it takes some time until the two new vortices are organized again. The velocity at the end of the numerical simulation is somewhat lower than that in the initial ansatz. This should be due to spin waves emitted during the process. The remarks of the last paragraph on the motion of the vortex centers are in agreement with an analytical solution obtained in \cite{wardomega} representing scattering of solitons in an integrable chiral model. Close to collision time, the centers move according to the law $x \approx \pm \sqrt{-t}$ $(t < 0)$. This gives a velocity $dx/dt \rightarrow \infty$ as $t \rightarrow 0$. They collide at $t\!=\!0$ and the centers of the two new solitons, which emerge along the $y$ axis, follow $y \approx \pm \sqrt{t}\; (t>0)$. The arguments presented in this section are certainly not sufficient to exclude, other than right angle, interesting possibilities of interaction and scattering of solitons. Nevertheless, they imply that the right angle scattering process is expected to be generic for solitons in two-dimensional Hamiltonian models. In the present work we have found no relation of the topology of the solitons to the right-angle scattering phenomenon, Therefore right-angle scattering is also expected to occur among non-topological solitons \cite{leandros}. \section{Conclusions} We have given a description of the right angle phenomenon of solitons through numerical simulations, as well as arguments which suggest that it should be generic in two dimensions. Two systems have been examined. Vortex-antivortex pairs in planar ferromagnets and vortices in antiferromagnets. The peculiar scattering behaviour is mainly attributed to the fact that solitons are extended structures rather than point like particles. This point is accounted for by the representation of a traveling soliton through a pair of lumps. Furthermore, we find that these two lumps act as independent physical entities at the time of collision. It is desirable to observe experimentally scattering of solitons in ferromagnetic and antiferromagnetic films. We expect that the present theoretical analysis will be useful in studies of systems of a lot of vortices \cite{mertens} and in particular in studies of the thermodynamics of magnetic systems. Suffice it to say that, in a magnetic material, vortices are expected to appear in pairs. In the study of the thermodynamics of layered antiferromagnets one expects to find the signature of topological excitations. The remark may prove important especially in view of the difficulty to observe isolated antiferromagnetic solitons due to the lack of a significant total net magnetization. The right angle scattering pattern appears also and has been understood from the point of view of the geometry of the moduli space for BPS monopoles in a three-dimensional Yang-Mills-Higgs theory \cite{hitchin}. We have proceeded in the simulation of scattering of a vortex and an antivortex in the $\sigma$-model (\ref{eq:sigma}). We use an initial ansatz similar to that of Eq.~(\ref{eq:ansatzsigma}). We simulate the time evolution of the system numerically and find that the vortices attract each other. They eventually collide at the origin and annihilate. It is quite interesting that the energy is dissipated at right angles. The phenomenon is presumably closely related to the present study. A similar simulation with corresponding results has been done in \cite{shellard} for non-gauged cosmic strings. The basic dynamical structure which leads to right angle scattering in Hamiltonian models is also present in the complex Ginzburg-Landau equation (CGLE) which describes nonlinear oscillatory media \cite{kramer}. There is actually a variety of nonconservative systems where scattering behaviour analogous to that studied in the present paper has been observed. An interesting example is a 2D fluid layer subjected to externally imposed oscillations. Coherent structures are formed which are experimentally observed to scatter at an almost right angle \cite{lioubashevski}. \section{Acknowledgments} I am grateful to N. Papanicolaou and P.N. Spathis for providing me their numerical code which calculates the vortex-antivortex pairs used in the simulations of Section II and for useful discussions. I thank F.G. Mertens for discussion of the issues presented in this paper. I also thank L. Kramer for beneficial remarks on part of the text and L. Perivolaropoulos for his help with the numerical algorithm.
\section{Introduction} The cross-section of the process $e^{+}e^{-}\rightarrow \pi ^{+}\pi ^{-}$ is given by \[ \sigma = \frac{\pi\alpha^2}{3s}\beta^3_\pi \left| F_\pi(s) \right|^2, \] where \( F_\pi(s) \) is the pion form factor at the center-of-mass energy squared $s$ and $\beta_\pi$ is the pion velocity. The pion form factor measurement is important for a number of physics problems. Detailed experimental data in the time-like region allows measurement of the parameters of the $\rho(770)$ meson and its radial excitations. Extrapolation of the energy dependence of the pion form factor to the point $s=0$ gives the value of the pion electromagnetic radius. Exact data on the pion form factor is necessary for precise determination of the ratio \[ R=\sigma (e^{+}e^{-}\rightarrow hadrons)/ \sigma (e^{+}e^{-}\rightarrow \mu ^{+}\mu ^{-}). \] Knowledge of R with high accuracy is required to evaluate the hadronic contribution $a_{\mu}^{had}$ to the anomalous magnetic moment of the muon $(g-2)_{\mu}$ \cite{kino}. About 87\% of the hadronic contribution in this case comes from $s<2\;\mathrm{GeV^2/c^2}$ (VEPP-2M range), and about 72\% --- from \( e^{+}e^{-}\rightarrow \pi ^{+}\pi ^{-} \) channel with \( s<2\; \mathrm{GeV^{2}/c^{2}} \) \cite{ej,bw}. The E821 experiment at BNL \cite{E821} has collected its first data in 1997 and will ultimately measure $(g-2)_{\mu}$ with a 0.35 ppm accuracy. To calculate the hadronic contribution with the desired precision the systematic error in R should be below 0.5\%. Therefore a new measurement of the pion form factor with a low systematic error is required. Experiments at the VEPP-2M collider\cite{VEPP}, which started in the early 70s, yielded a number of important results in $e^{+}e^{-}$ physics at low center-of-mass energies from 360 to 1400 MeV. The high precision measurement of the pion form factor at VEPP-2M was done in the late 70s -- early 80s by OLYA and CMD groups \cite{OLYACMD}. In the CMD experiment, 24 points from 360 to 820 MeV were studied with a systematic uncertainty of about 2\%. In the OLYA experiment, the energy range from 640 to 1400 MeV was scanned with small energy steps and the systematic uncertainty varied from 4\% at the \( \rho \)-meson peak to 15\% at 1400 MeV. \begin{table} \begin{center} \begin{tabular}{|c|r|c|c|r|} \hline ID & \multicolumn{1}{|c|}{Date} & $2E$, GeV& \parbox[t]{2.1cm}{Number of energy points} & $N_{e^{+}e^{-}\rightarrow \pi ^{+}\pi ^{-}}$\\ \hline \hline 1 & Jan--Feb 1994 & 0.81--1.02 & 14 & 35000 \\ \hline 2 & Nov--Dec 1994 & 0.78--0.81 & 10 & 66000 \\ \hline 3 & Mar--Jun 1995 & 0.61--0.79 & 20 & 85000 \\ \hline 4 & Oct--Nov 1996 & 0.37--0.52 & 10 & 4500 \\ \hline 5 & Feb--Jun 1997 & 0.98--1.38 & 37 & 75000 \\ \hline 6 & Mar--Jun 1998 & 0.36--0.97 & 37 & 1900000 \\ \hline \end{tabular} \end{center} \caption{\label{runstable}CMD-2 runs dedicated to R measurement} \end{table} During 1988-92 a new booster was installed to allow higher positron currents and injection of the electron and positron beams directly at the desired energy. During 1991-92 a new detector CMD-2 was installed at VEPP-2M, and in 1992 it started data taking. The pion form factor measurement was one of the major experiments planned at CMD-2. The energy scan of the whole VEPP-2M 0.36-1.38 GeV energy range was performed in six separate runs listed in Table \ref{runstable}. The energy range below the $\phi$ meson was scanned twice (in 94-96 and in 98). In this article we present results of the analysis of the data from runs 1--3. The data were taken at 43 energy points with the center-of-mass energy from 0.61 GeV up to 0.96 GeV with a 0.01 GeV energy step. The small energy step allows calculation of hadronic contributions in model-independent way. In the narrow energy region near the $\omega$-meson the energy steps were $0.002\div 0.006$ GeV in order to study the $\omega$-meson parameters and the $\rho-\omega$ interference. Since the form factor is changing relatively fast in this energy region, it was important that the beam energy was measured with the help of the resonance depolarization technique at almost all energy points. That allowed a significant decrease of the systematic error coming from the energy uncertainty. \begin{figure} \begin{center} \begin{tabular}{cc} \includegraphics[width=0.45\textwidth]{pic1d.eps} & \includegraphics[width=0.5\textwidth]{pic2p.eps}\\ \end{tabular} \end{center} \caption{\label{cmd2p}Detector CMD-2. 1 --- beam pipe, 2 --- drift chamber, 3 --- Z-chamber, 4 --- main solenoid, 5 --- compensating solenoid, 6 --- endcap (BGO) calorimeter, 7 --- barrel (CsI) calorimeter, 8 --- muon range system, 9 --- magnet yoke, 10 --- storage ring lenses} \end{figure} The CMD-2 (Fig.\ \ref{cmd2p}) is a general purpose detector consisting of the drift chamber, the proportional Z-chamber, the barrel (CsI) and the endcap (BGO) electromagnetic calorimeters and the muon range system. The drift chamber, Z-chamber and the endcap calorimeters are installed inside a thin superconducting solenoid with a field of 10 kGs. More details on the detector can be found elsewhere \cite{ICFA,CMD2,PREP}. The data described here was taken before the endcap calorimeter was installed. Two independent triggers were used during data taking. The first one, {\em charged trigger}, analyses information from the drift chamber and the Z-chamber and triggers the detector if at least one track was found. For 0.81--0.96 GeV energy points there was an additional requirement for the total energy deposition in the calorimeter to be greater than a 20--30 MeV threshold. The second one, {\em neutral trigger}, triggers the detector according to information from the calorimeter only. Events triggered by the charged trigger were used for analysis, while events triggered by the neutral trigger were used for trigger efficiency monitoring. \section{Data analysis} \subsection{Selection of collinear events} \begin{figure} \begin{center} \includegraphics[width=0.9\textwidth]{colle.eps} \end{center} \caption{\label{colle}Definition of parameters for two track events} \end{figure} The data was collected at 43 energy points. At the beam energy of 405 MeV the data was collected in two different runs with different detector and trigger conditions. Therefore, two independent results for this energy point are presented. From more than \( 4\cdot 10^{7} \) triggers about \( 4\cdot 10^{5} \) were selected as collinear events. The selection criteria were as follows.\label{collsel} \begin{enumerate} \item The event was triggered by the charged trigger. Other triggers may be present as well. \item Only one vertex with two oppositely charged tracks was found in the drift chamber. \item Distance from the vertex to the beam axis $\rho$ is less than 0.3 cm. \item Z-coordinate of the vertex (distance to the interaction point along the beam axis) $|Z|$ is less than 8 cm. \item Average momentum of two tracks $(p_1+p_2)/2$ is between 200 and 600 MeV/c. \item Acollinearity of two tracks in the plane transverse to the beam axis $ |\Delta \varphi |=|\pi -|\varphi _{1}-\varphi _{2}|| $ is less than 0.15 radians. \item Acollinearity of two tracks in the plane that contains the beam axis $|\Delta \Theta |=|\Theta _{1}-(\pi -\Theta _{2})|$ is less than 0.25 radians. \item Average polar angle of two tracks $ [\Theta _{1}+(\pi -\Theta _{2})]/2 $ is between $\Theta _{min}$ and $(\pi -\Theta _{min})$. All analysis was done separately for $\Theta _{min}=1.0$ and $\Theta _{min}=1.1$ radian. \end{enumerate} Because of the unfortunate accident the high voltage was off for two neighbouring lines of the CsI calorimeter during the data taking in the 0.61--0.784 GeV energy region. In order to avoid additional calorimeter inefficiency, events with $4.35<\varphi_{+}<4.95$ or $3.90<\varphi _{-}<4.50$ have been removed from analysis for corresponding energy points (angles are measured in radians). Events with $1.05<\varphi _{+}<1.65$ or $0.55<\varphi _{-}<1.15$ were additionally removed for 0.65--0.70 GeV energy points. Definition of the collinear event parameters, such as \( \rho \), \( \Theta _{1,2} \), \( \varphi _{1,2} \), is illustrated in Fig.\ \ref{colle}. \subsection{Event separation} \subsubsection{Likelihood function} \begin{floatingfigure}{0.46\textwidth} \begin{center} \includegraphics*[width=0.45\textwidth]{e1e2distr1.eps} \end{center} \caption{\label{distr}$E^+$ versus $E^-$ distribution for collinear events for 0.8 GeV energy point\vspace{2mm}} \end{floatingfigure} The selected sample of events consists of collinear events \( e^{+}e^{_{-}}\rightarrow e^{+}e^{-}, \) \( e^{+}e^{_{-}}\rightarrow \pi ^{+}\pi ^{-} \), \( e^{+}e^{_{-}}\rightarrow \mu ^{+}\mu ^{_{-}} \) and background events mainly due to cosmic muons. The energy deposition in the barrel CsI calorimeter by negati\-vely $(E^-)$ and positively $(E^+)$ charged particles were used for event separation. The distribution of \( E^{+} \) versus \( E^{-} \) is shown in Fig.\ \ref{distr}. Electrons and positrons usually have large energy deposition since they produce electromagnetic showers. Muons usually have small energy deposition since they are minimum ionising particles. Pions can interact as minimum ionising particles producing small energy deposition or have nuclear interactions inside the calo\-ri\-meter, resulting in long tails to a higher value of the energy deposition. The separation was based on the minimization of the following unbinned likelihood function: \begin{equation} \label{lfunc1} L=-\sum _{events}\ln \left( \sum _{a}N_{a}\cdot f_{a}(E^{+},E^{-})\right) +\sum _{a}N_{a}, \end{equation} where \( a \) is the event type ($a=ee$, $\mu\mu$, $\pi\pi$, $cosmic$), \( N_{a} \) is the number of events of the type \( a \) and \( f_{a}(E^{+},E^{-}) \) is the probability density for a type \( a \) event to have energy depositions \( E^{+} \) and \( E^{-} \). It was assumed that \( E^{+} \) is uncorrelated with \( E^{-} \) for events of the same type, so we can factorize the probability density: \[ f_{a}(E^{+},E^{-})=f_{a}^{+}(E^{+})\cdot f^{-}_{a}(E^{-}).\] For \( e^{+}e^{-} \), \( \mu ^{+}\mu ^{-} \) pairs and cosmic events the energy deposition distribution is the same for negatively and positively charged particles, while these distributions are significantly different for \( \pi ^{+} \) and \( \pi ^{-} \). Therefore \( f_{a}^{+}\equiv f^{-}_{a} \) for \( a=ee,\; \mu \mu \) and \( cosmic \), but for pions these functions are different. The probability density functions are described in detail in the following sections. In minimization the ratio \( N_{\mu \mu }/N_{ee} \) was fixed according to the QED calculation \[ \frac{N_{\mu\mu}}{N_{ee}}= \frac{\sigma_{\mu\mu}\cdot (1+\delta_{\mu\mu}) \left( 1+\alpha_{\mu\mu} \right) \varepsilon_{\mu\mu}} {\sigma_{ee}\cdot (1+\delta_{ee}) \left( 1+\alpha_{ee} \right) \varepsilon_{ee}}, \] where $\sigma$ is the Born cross-section, $\delta$ is a radiative correction, \( \alpha \) is the correction for the experimental resolution of $\Theta$ angle measurement and \( \varepsilon \) is the reconstruction efficiency (see sec. \ref{piformsec} for details). The likelihood function (\ref{lfunc1}) was rewritten to have the following global fit parameters: \[ (N_{ee}+N_{\mu \mu }),\quad \frac{N_{\pi \pi }}{N_{ee}+N_{\mu \mu }},\quad N_{cosmic}\] instead of \( N_{ee} \), \( N_{\mu \mu } \), \( N_{\pi \pi } \) and \( N_{cosmic} \). The number of cosmic events $N_{cosmic}$ was determined before the fit as described below, and was fixed during the fit. \vspace{2mm} \subsubsection{Rejection of background events} \begin{floatingfigure}{0.47\textwidth} \begin{center} \includegraphics*[width=0.4\textwidth]{rho9495_both_rho_z.eps} \end{center} \caption{\label{rhoz} To determination of number of cosmic background events} \end{floatingfigure} \begin{figure}[tb] \begin{center} \begin{tabular}{cc} \subfigure[\label{rhoz1}$\rho$-distribution] {\includegraphics[width=0.45\textwidth]{rho9495_both_rho_z_1.eps}} & \subfigure[\label{rhoz2}$Z$-distribution] {\includegraphics[width=0.45\textwidth]{rho9495_both_rho_z_2.eps}} \\ \end{tabular} \end{center} \caption{\label{rhoz12} $\rho$- and $Z$-distributions for background and collinear events. Empty histograms correspond to all selected events, filled histograms correspond to background events only} \end{figure} Cosmic background events can be separated from the beam-produced col\-linear events by the distance $\rho$ from the vertex to beam axis and (or) by the $Z$-coordinate of the vertex. The $\rho$-distribution for collinear events has a narrow peak around 0, while the same distribution for the background events is almost flat (Fig.\ \ref{rhoz1}). Similarly, the $Z$-distribution for collinear events is Gaussian like, while it is almost flat for the background events (Fig.\ \ref{rhoz2}). Assuming that $\rho$- and $Z$-distributions for cosmic events are not correlated and that all events with $\rho>0.5$ cm or $|Z|>8$ cm are cosmic events, one can calculate the number of background events in the signal region from the number of events out of the signal region. Let us define $N_X$ as the number of cosmic events in the signal region ($\rho<0.3$ cm, $|Z|\leq 8$ cm) and $N_{1,2,3}$ as the number of events in different out-of-signal $\rho-Z$ regions (see Fig.\ \ref{rhoz}): \[ \left\{ \begin{array}{l} N_1=N\bigm|_{\rho<0.3;\;8<|Z|<12} \\ N_2=N\bigm|_{0.5<\rho<1.0;\;|Z|\leq 8} \\ N_3=N\bigm|_{0.5<\rho<1.0;\;8<|Z|<12} \end{array} \right. . \] Then under assumptions mentioned above we can derive: \[ N_X = \frac{N_2}{N_3}\cdot N_1 = \frac{N_1}{N_3}\cdot N_2. \] $N_X$ is equal to $N_{cosmic}$ only statistically. The confidence interval for $N_{cosmic}$, corresponding to one standard deviation, is \[ N_{cosmic} \approx N_X \pm \sqrt{5 N_X}. \] In order to take that into account, the $N_{cosmic}$ value was fixed to the $N_X$ value during fit, but $\Delta N_X$ was added to $\Delta N_{\pi\pi}$ at the end. \vspace{2mm} \subsubsection{Energy deposition parametrization} In order to construct the likelihood function (\ref{lfunc1}) it is very convenient to have an analytical form for energy deposition distributions. We parametrize all energy depositions by the linear combination of two types of Gaussian-like functions, described below. \begin{description} \item [Normal distribution]\( g(x;x_{0},\sigma ) \) \[ g(x;x_{0},\sigma )=\frac{1}{\sqrt{2\pi }\sigma }\exp \left[ -(x-x_{0})^{2}/2\sigma ^{2}\right] ,\] where the parameters are the mean value $x_0$ and the standard deviation $\sigma$. The Gaussian is normalized at the $(-\infty, +\infty)$ interval, but energy deposition values belong to the $[0,+\infty)$ range. Therefore for parametrization the Gaussian normalized at the $[0,+\infty)$ range is used: \[ g(x;x_{0},\sigma )/\left[ 1-I\left( -\frac{x_{0}}{\sigma }\right) \right] , \] where \[ I(x)=\frac{1}{\sqrt{2\pi }}\int _{-\infty }^{x}\exp \left( -\frac{t^{2}}{2}\right) dt. \] \item [Normal logarithmic distribution\cite{lgs}] \( g_{l}(x;x_{0},\sigma ,\eta ) \) \[ g_{l}(x;x_{0},\sigma ,\eta )=\frac{1}{\sqrt{2\pi }\sigma }\cdot \frac{\eta }{\sigma _{0}}\cdot \exp \left\{ -\frac{1}{2}\left[ \frac{\ln ^{2}\left( 1-\frac{x-x_{0}}{\sigma }\eta \right) }{\sigma _{0}^{2}}+\sigma _{0}^{2}\right] \right\} ,\] where \[ \sigma_0=\ln\left(\eta\sqrt{2\ln 2}+\sqrt{1+\eta^2\cdot 2\ln 2} \right) \left/ \sqrt{2\ln 2} \right. . \] Parameters of this function are the most probable energy $x_0$, the $\sigma=\mathrm{FWHM}/2.35$ and the asymmetry $\eta$. If \( \eta =0 \), the logarithmic Gaussian is equal to the plain Gaussian: \[ g_{l}(x;x_{0,}\sigma ,\eta )\left| _{\eta =0}\equiv g(x;x_{0},\sigma ).\right. \] The function is normalized at the \( [-\infty ,x_{0}+\sigma /\eta ] \) interval: \[ \int _{-\infty }^{x_{0}+\sigma /\eta }g_{l}(x)dx=1,\] but in our case its integral over the range \( (-\infty ,0] \) is negligible and it is assumed that this function is normalized at the \( [0,x_{0}+\sigma /\eta ] \) interval. \end{description} Since both functions are normalized, it is easy to construct the normalized energy deposition distribution. \vspace{2mm} \subsubsection{\label{ecorsec}Correction of energy deposition for polar angle} \begin{figure} \begin{center} \begin{tabular}{cc} \subfigure[Energy deposition versus $\Theta$ for collinear events before correction]{\resizebox*{0.4\textwidth}{!}{\includegraphics{ecor_2D_no.eps}}} & \subfigure[Average energy deposition versus $\Theta$ for electrons and cosmic background events before correction]{\resizebox*{0.4\textwidth}{!}{\includegraphics{ecor_prof_no.eps}}} \\ \subfigure[Energy deposition versus $\Theta$ for collinear events after correction]{\resizebox*{0.4\textwidth}{!}{\includegraphics{ecor_2D_yes.eps}}} & \subfigure[Average energy deposition versus $\Theta$ for electrons and cosmic background events after correction] {\resizebox*{0.4\textwidth}{!}{\includegraphics{ecor_prof_yes.eps}}} \\ \end{tabular} \end{center} \caption{\label{ecorfig}Correction of energy deposition for the polar angle of the track} \end{figure} Energy deposition in the calorimeter is correlated with the polar angle $\Theta$ of a track due to the dependence of calorimeter thickness on $\Theta$. So the probability density for an event of type $a$ to have energy deposition $E$ depends not only on $E$, but also on $\Theta$. In order to simplify the likelihood function, the energy deposition was corrected to make dependence of energy deposition on \( \Theta \) negligible. In other words, instead of \( f_{a}(E,\Theta ) \) we are using \( f_{a}(\overline{E}) \), where \( \overline{E} \) is the corrected energy deposition: \[ E\longrightarrow \overline{E}\quad to\; have\quad f_{a}(E,\Theta )\longrightarrow f_{a}(\overline{E}).\] The correction procedure was performed as follows. First, two groups of events were selected: cosmic events and electrons. Selection criteria for both types of events are described below in sections \ref{cosesec} and \ref{elecesec}. The scatter-plot of the energy deposition versus \( \Theta \) for collinear events is shown in Fig.\ \ref{ecorfig}a. The full range of possible \( \Theta \) values \( \left[ 1,\pi -1\right] \) was divided into 50 equal bins and for each bin an average energy deposition was calculated for electrons and cosmic events. The energy deposition of electrons depends on $\Theta$ because the CMD-2 calorimeter has only about $8X_{0}$ and therefore the shower leakage is not negligible. The effective calorimeter thickness increases for incident track angle away from normal. The corresponding small variation of the average energy deposition of electrons was fit to the parabolic function \[E_e(\Theta)=E_1\cdot \left[ 1+\alpha (\Theta-\pi/2)^2\right],\] where $E_{1}$ and $\alpha$ are the fit parameters. Cosmic events contain minimum ionising particles and their energy deposition is proportional to the length in the calorimeter through which the particle passes. This length is proportional to \( 1/\sin \Theta \) , and the fit function in this case was \[E_{c}(\Theta )=E_{0}/\sin \Theta ,\] where $E_{0}$ is a fit parameter. An example of the fit of the average energy depositions for electrons and cosmic tracks is shown in Fig.~\ref{ecorfig}b. The corrected energy was calculated as \[ \overline{E}(E,\Theta )=E\cdot \left[ k_{c}(\Theta )+\frac{k_{e}(\Theta )-k_{c}(\Theta )}{E_{e}(\Theta )-E_{c}(\Theta )}\left( E-E_{c}(\Theta )\right) \right] ,\] where \( k_{c}(\Theta )=E_{0}/E_{c}(\Theta ) \), \( k_{e}(\Theta )=E_{1}/E_{e}(\Theta ) \), \( E_{c}(\Theta ) \) and \( E_{e}(\Theta ) \) are the measured average energy deposition for cosmic and Bhabha events respectively. After such a correction the average energy deposition for both cosmic and Bhabha events does not depend on \( \Theta \). In fact, if \( E=E_{e}(\Theta ) \), then \( \overline{E}=E_{1} \); if \( E=E_{c}(\Theta ) \), then \( \overline{E}=E_{0} \). Since the energy deposition of cosmic events does not depend on the beam energy, the variations of $E_0$ reflects residual variations in the calibration of the CsI calorimeter. To correct for these variations, the energy deposition was normalized to have the average energy deposition of the cosmic events equal to 85 MeV (the arbitrary constant close to the experimental value) for all beam energies: \[ \overline{E}\longrightarrow \overline{E}\cdot \frac{85}{E_{0}}.\] The scatter-plot of the corrected energy deposition versus \( \Theta \) for collinear events is shown in Fig.\ \ref{ecorfig}c. The average energy deposition for electrons and cosmic events versus \( \Theta \) after the correction is shown in Fig.\ \ref{ecorfig}d. The described energy deposition correction was calculated independently for each energy point. In the all following sections \( E \) actually means the corrected energy $\overline{E}$. \vspace{2mm} \subsubsection{\label{cosesec}Energy deposition of cosmic events} As mentioned above, it is possible to select a pure sample of background events from the real data. In this case the selection criteria are the same as standard cuts for the collinear events, except that the distance $\rho$ from the vertex to the beam axis should be from 5 to 10 mm (instead of less than 3 mm). In order to make the selected sample of background events even cleaner, the fact that there are two clusters in the event was used. Analysing the energy deposition of one cluster, the energy deposition of the other one was required to be less than 150 MeV. Since the clusters are independent, the strict limit on one cluster energy does not affect the energy deposition distribution for the other cluster. The data for all energy points was combined together and the energy deposition of cosmic events was parametrized using the following function: \begin{equation} \label{fcosmic1} f(E)=p\cdot g_{l}(E;E_{0},\sigma ,\eta )+(1-p)\cdot g_{l}(E;E_{0}+\alpha \sigma ,\beta \sigma ,\gamma \eta ). \end{equation} The fit result is presented in Fig.\ \ref{cosfit}. The obtained function was used as the energy deposition for background events: \begin{equation} \label{fcosmic0} \begin{array}{c} f^{+}_{cosmic}(E)\equiv f^{-}_{cosmic}(E)\equiv f_{cosmic}(E) , \vphantom{\Big /} \\ \multicolumn{1}{l}{ f_{cosmic}(E) = 0.69\cdot g_{l}(E;81.2,9.0,-0.175)+ \vphantom{\Big /} } \\ \multicolumn{1}{r}{ \qquad \qquad \qquad \qquad \qquad \qquad + 0.31\cdot g_{l}(E;91.7,22.8,-0.186). \vphantom{\Big /} } \end{array} \end{equation} In order to check the stability of the energy deposition for cosmic events, all energy points were combined into 4 groups: 0.61--0.70 GeV, 0.71--0.778 GeV, 0.780--0.810 GeV and 0.810--0.960 GeV. The energy deposition of the selected background events for each group was fitted with the function (\ref{fcosmic0}). Results of the fit are demonstrated in Fig.\ \ref{cosefitg}. \begin{figure}[tbp] \begin{center} \includegraphics[width=0.9\textwidth]{cosefit.eps} \end{center} \caption{\label{cosfit}Fit of the energy deposition for cosmic events with the function (\ref{fcosmic1}). Fit parameters: P1 -- number of events, P2 -- $p$, P3 -- $E_0$, P4 -- $\sigma$, P5 -- $\eta$, P6 -- $\alpha$, P7 -- $\beta$, P8 -- $\gamma$} \end{figure} \begin{figure}[p] \begin{center} \begin{tabular}{cc} \subfigure[Group 0.61--0.70 GeV] {\includegraphics[width=0.45\textwidth]{coseg1.eps}} & \subfigure[Group 0.71--0.778 GeV] {\includegraphics[width=0.45\textwidth]{coseg2.eps}} \\ \subfigure[Group 0.780--0.810 GeV] {\includegraphics[width=0.45\textwidth]{coseg3.eps}} & \subfigure[Group 0.810--0.960 GeV] {\includegraphics[width=0.45\textwidth]{coseg4.eps}} \\ \end{tabular} \end{center} \caption{\label{cosefitg}Fit of energy deposition of cosmic events with the function (\ref{fcosmic0}). Data is combined into 4 groups of energy points. Fit parameter: P1 -- number of events} \end{figure} The probability to have no cluster in the calorimeter (zero energy deposition) for cosmic event tracks was also measured and found to be negligible --- less than 0.1\%. \vspace{2mm} \subsubsection{\label{elecesec}Energy deposition of electrons and positrons} \begin{figure}[tb] {\centering \begin{tabular}{cc} \subfigure[Fit of the positron energy deposition]{\resizebox*{0.46\textwidth}{!}{\includegraphics{elecfitp.eps}}} & \subfigure[Fit of the electron energy deposition]{\resizebox*{0.46\textwidth}{!}{\includegraphics{elecfitm.eps}}} \\ \end{tabular}\par} \caption{\label{elecefit}Fit of the energy deposition of electrons and positrons at center-of-mass energy 0.9 GeV. Fit parameters: P1 -- $p_{w}$, P2 -- $E_{0}$, P3 -- $\sigma$, P4 -- $\eta$, P5 -- number of events, P6 -- $\alpha$, P7 -- $\beta$, P8 -- $\gamma$} \end{figure} The energy deposition for electrons and positrons was obtained from the data. Electrons and positrons can be separated from mesons by their relatively high energy deposition in the calorimeter. If we have one ``clean'' electron (positron) in the event (a particle with a proper momentum and high energy deposition), then we know that the second particle is a positron (electron), and therefore may be used for the energy deposition study. The test $e^+e^-\rightarrow e^+e^-$ events were selected for the energy deposition study with the same selection criteria as for collinear events except for the following. \begin{enumerate} \item The distance $\rho$ from the vertex to the beam axis is less than 0.15 cm (instead of 0.3 cm). \item The average momentum of the particles is within the 10 MeV/c range around the beam energy. This requirement allows to decrease significantly a small admixture of $e^+e^- \rightarrow \pi^+\pi^-$ events. \end{enumerate} Analysing the energy deposition in one cluster, the energy deposition in the other one is required to be between $0.92E_B-100$ MeV and $0.92E_B$, where $E_B$ is the beam energy. As for cosmic rays since the clusters are independent, the strict limit on one cluster energy does not affect the energy deposition distribution for the other cluster. For each energy point the energy depositions of the selected electrons and positrons were fitted with the following function: \[ f(E)=(1-p_{w})\cdot g_{l}(E;E_{0},\sigma ,\eta )+p_{w}\cdot g_{l}(E;E_{0}+\alpha \sigma ,\beta \sigma ,\gamma \eta ).\] After analysis of the fit results the following values of the parameters were found: \[ \alpha =-0.8,\quad \beta =3.0,\quad \gamma =1.4.\] Again, as for cosmic rays, the probability to have no cluster was found to be negligible (less than 0.1\%) for all energy points. Finally, the energy deposition of electrons and positrons was parametrized as: \begin{equation} \label{fee0} \begin{array}{c} f^{+}_{ee}(E)\equiv f^{-}_{ee}(E)\equiv f_{ee}(E),\\ \\ f_{ee}(E)=(1-p_{w})\cdot g_{l}(E;E_{0},\sigma ,\eta)+p_{w}\cdot g_{l}(E;E_{0}-0.8\sigma ,3\sigma ,1.4\eta), \end{array} \end{equation} where \( p_{w} \), \( E_{0} \), \( \sigma \) and $\eta$ are free parameters of the fit. An example of the electron and positron energy depositions is shown in Fig.\ \ref{elecefit}. \vspace{2mm} \subsubsection{Energy deposition of muons} \begin{figure} \begin{center} \begin{tabular}{cc} \multicolumn{2}{c} {\subfigure[Energy deposition of 375 MeV muons] {\includegraphics[width=0.95\textwidth]{mip_edep_mu_1.eps}}} \\ \subfigure[Energy dependence of $E_0$] {\includegraphics[width=0.46\textwidth]{mip_edep_mu_2.eps}} & \subfigure[Energy dependence of $\sigma_0$] {\includegraphics[width=0.46\textwidth]{mip_edep_mu_3.eps}} \\ \end{tabular} \end{center} \caption{\label{mip_mu_sim} Simulated energy deposition of muons} \end{figure} In the center-of-mass energy range above 0.6 GeV, muons produced in the reaction $e^{+}e^{-}\rightarrow \mu ^{+}\mu ^{-}$ interact with the calorimeter as the minimum ionising particles, and therefore their energy deposition can be obtained from the detector simulation. The simulation of the $e^+e^-\rightarrow \mu^+\mu^-$ events was performed for center-of-mass energies 0.6, 0.62, 0.64, 0.66, 0.68, 0.70, 0.74, 0.75, 0.76, 0.77, 0.78, 0.79, 0.80, 0.82, 0.85, 0.90, 0.95, 1.0 GeV. 10000 events were simulated for each energy point. Simulated events were processed by the standard offline reconstruction program and the standard selection criteria for collinear events were applied. Then the energy deposition of the simulated muons was corrected as described in sec.\ \ref{ecorsec}. The resulting energy deposition for each energy point was parametrized by the following function: \begin{eqnarray*} \lefteqn{ f^\mu_{SIM}(E) = p_1 \cdot \left[ \, p_2 \cdot g_l(E; E_0, \sigma_0, \eta) + \right. } \qquad & & \\ & & \left. +(1-p_2)\cdot g_l(E; E_0, \beta\sigma_0, \gamma\eta) \right] + (1-p_1)\cdot g(E; E_1, \sigma_1) , \end{eqnarray*} where $p_1$, $p_2$, $E_0$, $\sigma_0$, $E_1$, $\sigma_1$, $\eta$, $\beta$ and $\gamma$ are the free fit parameters. In the simulation we have a perfectly calibrated calorimeter while the real data have an additional spread due to the uncertainty of the calibration coefficients. As a result, the real energy distribution is wider than that in the simulation. It could also be slightly shifted. To take that into account, the energy deposition distribution obtained from the simulation was modified to reflect the energy deposition of real muons: \[ \begin{array}{c} f_{\mu }^{+}(E)\equiv f_{\mu }^{-}(E)\equiv f_{\mu }(E), \vphantom{\Big /} \\ f_\mu (E) = f^\mu_{MIP}(E;k,\sigma_x), \vphantom{\bigg /} \\ \multicolumn{1}{l}{ f^\mu_{MIP}(E; k, \sigma_x) = p_1 \cdot \left[ \, p_2 \cdot g_l(\tilde{E}; E_0, \tilde{\sigma}_0, \eta) + \right. \vphantom{\bigg /} } \\ \multicolumn{1}{r}{ \qquad \qquad \qquad \left. +(1-p_2)\cdot g_l(\tilde{E}; E_0, \beta\tilde{\sigma}_0, \gamma\eta) \right] + (1-p_1)\cdot g(\tilde{E}; E_1, \tilde{\sigma_1}), \vphantom{\bigg /} } \\ \tilde{E}=E/k, \quad \tilde{\sigma}_0=\sqrt{\sigma_0^2+\sigma_x^2}, \quad \tilde{\sigma}_1=\sqrt{\sigma_1^2+\sigma_x^2}. \vphantom{\Big /} \end{array} \] The values of $p_1$, $p_2$, $E_0$, $\sigma_0$, $E_1$, $\sigma_1$, $\eta$, $\beta$ and $\gamma$ are fixed according to the simulation, so this function has only two free parameters: $k$ is the scale coefficient and $\sigma_x$ is the additional energy spread. An example of the fit for one energy point as well as the energy dependence of $E_0$ and $\sigma_0$ parameters are shown in Fig.\ \ref{mip_mu_sim}. The $k$ and $\sigma_x$ parameters for all energy points are shown in Fig.~\ref{pipar}a and \ref{pipar}b respectively. \vspace{2mm} \subsubsection{Energy deposition of pions} Pions produced in the reaction \( e^{+}e^{-}\rightarrow \pi ^{+}\pi ^{-} \) may interact with the calorimeter in two ways: as minimum ionising particles or through nuclear interaction. \begin{figure} \begin{center} \begin{tabular}{cc} \multicolumn{2}{c} {\subfigure[Energy deposition of 400 MeV pions] {\includegraphics[width=0.95\textwidth]{mip_edep_pi_1.eps}}} \\ \subfigure[Energy dependence of $E_0$] {\includegraphics[width=0.46\textwidth]{mip_edep_pi_2.eps}} & \subfigure[Energy dependence of $\sigma_0$] {\includegraphics[width=0.46\textwidth]{mip_edep_pi_3.eps}} \\ \end{tabular} \end{center} \caption{\label{mip_pi_sim} Simulated energy deposition of minimum ionising pions} \end{figure} The energy deposition of the minimum ionising pions was obtained from the detector simulation in the same way as for muons. The simulation was performed for the center-of-mass energies 0.6, 0.62, 0.64, 0.66, 0.68, 0.70, 0.75, 0.80, 0.85, 0.90, 0.95, 1.0 GeV. 10000 events were simulated for each energy point. Simulated events were processed by the standard offline reconstruction program and the standard selection criteria for collinear events were applied. Then the energy deposition of the simulated pions was corrected as described in sec.\ \ref{ecorsec}. The resulting energy deposition for each energy point was parametrized by the same function as for muons: \begin{eqnarray*} \lefteqn{ f^\pi_{SIM}(E) = p_1 \cdot \left[ \, p_2 \cdot g_l(E; E_0, \sigma_0, \eta) + \right. } \qquad & & \\ & & \left. +(1-p_2)\cdot g_l(E; E_0, \beta\sigma_0, \gamma\eta) \right] + (1-p_1)\cdot g(E; E_1, \sigma_1) , \end{eqnarray*} where $p_1$, $p_2$, $E_0$, $\sigma_0$, $E_1$, $\sigma_1$, $\eta$, $\beta$ and $\gamma$ are the free fit parameters depending on the beam energy. This function describes well the energy deposition only for those pions that do not stop in the calorimeter. As was done for muons, the modified simulated energy deposition was used for the parametrization of the energy deposition of real pions: \[ \begin{array}{c} \multicolumn{1}{l}{ f^\pi_{MIP}(E; k, \sigma_x) = p_1 \cdot \left[ \, p_2 \cdot g_l(\tilde{E}; E_0, \tilde{\sigma}_0, \eta) + \right. \vphantom{\bigg /} } \\ \multicolumn{1}{r}{ \qquad \qquad \qquad \left. + (1-p_2)\cdot g_l(\tilde{E}; E_0, \beta\tilde{\sigma}_0, \gamma\eta) \right] + (1-p_1)\cdot g(\tilde{E}; E_1, \tilde{\sigma_1}), \vphantom{\bigg /} } \\ \tilde{E}=E/k, \quad \tilde{\sigma}_0=\sqrt{\sigma_0^2+\sigma_x^2}, \quad \tilde{\sigma}_1=\sqrt{\sigma_1^2+\sigma_x^2}. \vphantom{\Big /} \end{array} \] The values of $p_1$, $p_2$, $E_0$, $\sigma_0$, $E_1$, $\sigma_1$, $\eta$, $\beta$ and $\gamma$ were fixed according to the simulation, so this function has only two free parameters: the scale coefficient $k$ and the additional energy spread $\sigma_x$. An example of the energy deposition of minimum ionising pions and the energy dependence of $E_0$ and $\sigma_0$ parameters are shown in Fig.\ \ref{mip_pi_sim}. The $k$ and $\sigma_x$ parameters for all energy points are shown in Fig.~\ref{pipar}a and \ref{pipar}b respectively. The energy deposition of the nuclear interacted pions cannot be simulated well. Therefore, in this case we have chosen to use an empirical parametrization, which fits experimental data well: \[ f^\pi_{NI} (E; \sigma, p_{1\ldots N}) = \frac{g(E; 0,\sigma)+ \sum\limits_{i=1}^{N} g(E; \frac{E_0}{N+1}i, \sigma)\cdot p_i^2}{ 1+\sum\limits_{i=1}^{N} p_i^2}, \] where $E_0$ is the beam energy, N is the number of Gaussian and $\sigma$ and $p_{1\ldots N}$ are the free fit parameters. All Gaussian are normalized at the $[0,+\infty]$, therefore $f^\pi_{NI}$ is also normalized. This function appears as a set of equidistant Gaussians of the same width but different area. The parametrization with $N=5$ or $6$ describes well the experimental data. The energy depositions of the minimum ionising $\pi^+$ and $\pi^-$ are the same while they are quite different for nuclear interacted $\pi^+$ and $\pi^-$. Unlike electrons and muons, pions have a significant probability to have no cluster in the calorimeter and this probability is different for $\pi^+$ and $\pi^-$. According to this, the overall pion energy deposition was parametrized with the following functions: \begin{eqnarray} \label{fpipip} \lefteqn{ f_{\pi }^{+}(E) = p_{0}^{+}\cdot\delta(E)+(1-p_{0}^{+})\times } & & \\ \nonumber & & \times \left[ \, p^{+}_{MIP}\cdot f^\pi_{MIP}(E; k, \sigma_x) +(1-p^{+}_{MIP})\cdot f^\pi_{NI}(E;\sigma^+, p^+_{1\ldots N})\right], \vphantom{\bigg /} \\ \label{fpipim} \lefteqn{ f_{\pi }^{-}(E) = p_{0}^{-}\cdot\delta(E)+(1-p_{0}^{-})\times } \\ \nonumber & & \times \left[ \, p^{-}_{MIP}\cdot f^\pi_{MIP}(E; k, \sigma_x) +(1-p^{-}_{MIP})\cdot f^\pi_{NI}(E;\sigma^-, p^-_{1\ldots N})\right], \vphantom{\bigg /} \end{eqnarray} where $p_0^\pm$, $p_{MIP}^\pm$, $k$, $\sigma_x$, $\sigma^\pm$, $p^\pm_{1\ldots N}$ are free fit parameters. Pions produced in the $\phi(1020)$ decay to $3\pi$ were used to test whether the functions (\ref{fpipip}) and (\ref{fpipim}) describe the energy deposition of the real pions well enough. The CMD-2 has collected about 20 pb$^{-1}$ in the $\varphi$-meson energy range. About 100000 ``clean'' $\pi^+$ and $\pi^-$ were selected from the completely reconstructed $\varphi\rightarrow\pi^+\pi^-\pi^0$ events. The energy deposition of $\pi^+$ ($\pi^-$) with energies $345<E_\pi<355$ MeV and $425<E_\pi<435$ MeV together with the fit are shown in Fig.\ \ref{piedepex}a and \ref{piedepex}c (\ref{piedepex}b and \ref{piedepex}d). \begin{figure} \begin{center} \begin{tabular}{cc} \subfigure[$\pi^+$, $345<E_\pi<355$ MeV] {\includegraphics[width=0.45\textwidth]{pi_3pi_350p.eps}} & \subfigure[$\pi^-$, $345<E_\pi<355$ MeV] {\includegraphics[width=0.45\textwidth]{pi_3pi_350m.eps}} \\ \subfigure[$\pi^+$, $425<E_\pi<435$ MeV] {\includegraphics[width=0.45\textwidth]{pi_3pi_430p.eps}} & \subfigure[$\pi^-$, $425<E_\pi<435$ MeV] {\includegraphics[width=0.45\textwidth]{pi_3pi_430m.eps}} \end{tabular} \end{center} \caption{\label{piedepex}An example of the experimental energy deposition for pions of different energy. Pions were extracted from $\varphi\rightarrow\pi^+\pi^-\pi^0$ data. The energy depositions of $\pi^+$ and $\pi^-$ were fitted with functions (\ref{fpipip}) and (\ref{fpipim}) respectively} \end{figure} \subsubsection{Fit results} The minimization of the likelihood function (\ref{lfunc1}) was carried out for all energy points with the typical number of 10000 events per energy point with the following 27 free fit parameters: \begin{itemize} \item global parameters \[ (N_{ee}+N_{\mu \mu }),\quad \frac{N_{\pi \pi }}{N_{ee}+N_{\mu \mu }} \] \item Electron energy deposition parameters \[ p_{W}, \quad E_{0}, \quad \sigma, \quad \eta \] \item Muon energy deposition parameters \[ k_{MIP}, \quad \sigma_x^\mu \] \item Pion energy deposition parameters \begin{eqnarray*} & \sigma_x^\pi, \quad p_{0}^{+}, \quad p_{0}^{-}, \quad p_{MIP}^{+}, \quad p_{MIP}^{-}, & \\ & \sigma _{NI}^{+},\quad \sigma _{NI}^{-}, \quad p^+_{1\ldots N}, \quad p^-_{1\ldots N}, \quad N=6 & \end{eqnarray*} \end{itemize} The number of background events was fixed during the fit. The scale parameter for minimum ionising muons and pions was the same (listed about as $k_{MIP}$). The results of the minimization are shown in Fig.~\ref{gpar}--\ref{pipar}: the global parameters are shown in Fig.~\ref{gpar}, the electron energy deposition parameters are shown in Fig.~\ref{epar} and the muon and pion energy deposition parameters are shown in Fig.~\ref{pipar}. \begin{figure} \begin{center} \begin{tabular}{cc} \subfigure[$N_{ee}+N_{\mu\mu}$] {\includegraphics[width=0.45\textwidth]{fitpar_nel.eps}} & \subfigure[$N_{\pi\pi}/(N_{ee}+N_{\mu\mu})$] {\includegraphics[width=0.45\textwidth]{fitpar_k.eps}} \end{tabular} \end{center} \caption{\label{gpar} Results of the minimization of the likelihood function (\ref{lfunc1}). The energy dependences of the global parameters are shown} \end{figure} \begin{figure} \begin{center} \begin{tabular}{cc} \subfigure[$E_{0}$] {\includegraphics[width=0.45\textwidth]{fitpar_e_mean.eps}} & \subfigure[$\sigma$] {\includegraphics[width=0.45\textwidth]{fitpar_e_sig.eps}} \\ \subfigure[$p_W$] {\includegraphics[width=0.45\textwidth]{fitpar_e_pw.eps}} & \subfigure[$\eta$] {\includegraphics[width=0.45\textwidth]{fitpar_e_ass.eps}} \end{tabular} \end{center} \caption{\label{epar} Results of the minimization of the likelihood function (\ref{lfunc1}). The electron energy deposition parameters are shown} \end{figure} \begin{figure} \begin{center} \begin{tabular}{cc} \subfigure[$k_{MIP}$] {\includegraphics[width=0.45\textwidth]{fitpar_kmip.eps}} & \subfigure[$\sigma_x^\mu$ and $\sigma_x^\pi$] {\includegraphics[width=0.45\textwidth]{fitpar_sigx.eps}} \\ \subfigure[$p_0^+$ and $p_0^-$] {\includegraphics[width=0.45\textwidth]{fitpar_pi_p0.eps}} & \subfigure[$p_{MIP}^+$ and $p_{MIP}^-$] {\includegraphics[width=0.45\textwidth]{fitpar_pi_pmip.eps}} \end{tabular} \end{center} \caption{\label{pipar} Results of the minimization of the likelihood function (\ref{lfunc1}). The muon and pion energy deposition parameters are shown} \end{figure} \newpage \subsection{\label{piformsec}Form factor calculation} After the minimization of the likelihood function, the pion form factor was calculated as: \begin{eqnarray} \label{piform} \lefteqn{ \left| F_{\pi }\right| ^{2}= \frac{N_{\pi \pi }}{N_{ee}+N_{\mu \mu }}\times } & & \\ \nonumber & & \times \frac{\sigma ^{B}_{ee}\cdot \left( 1+\delta _{ee}\right) \left( 1+\alpha _{ee}\right) \varepsilon _{ee}+ \sigma ^{B}_{\mu \mu }\cdot \left( 1+\delta _{\mu \mu }\right) \left( 1+\alpha _{\mu \mu }\right) \varepsilon _{\mu \mu }} {\sigma ^{B}_{\pi \pi }\cdot \left( 1+\delta _{\pi \pi }\right) \left( 1+\alpha _{\pi \pi }\right) \left( 1-\Delta _{H}\right) \left( 1-\Delta _{D}\right) \varepsilon _{\pi \pi }}- \\ \nonumber & & - \Delta _{3\pi }, \end{eqnarray} where \( N_{\pi \pi }/(N_{ee}+N_{\mu \mu }) \) was obtained as a result of minimization, \( \sigma ^{B}_{a} \) are the Born cross-sections, \( \delta _{a} \) are the radiative corrections, \( \varepsilon _{a} \) are the efficiencies, including trigger and reconstruction efficiencies, \( \alpha _{a} \) are the corrections for the experimental resolution of the \( \Theta \) angle measurement, \( \Delta _{H} \) is the correction for pion losses due to nuclear interactions, \( \Delta _{D} \) is the correction for pion losses due to decays in flight and \( \Delta _{3\pi } \) is the correction for the misidentification of \( \pi ^{+}\pi ^{-}\pi ^{0} \) events as a \( \pi ^{+}\pi ^{-} \) pair. Together with the pion form factor, the cross-section of $e^{+}e^{-}$ annihilation to $\pi^+\pi^-$ was calculated as \[ \sigma _{\pi \pi }=\frac{\pi \alpha ^{2}}{3s}\left( 1-\frac{4m_{\pi }^{2}}{s}\right) ^{3/2}\cdot \left| F_{\pi }\right| ^{2}, \] and the luminosity integral as \[ \int Ldt=\frac{N_{ee}+N_{\mu \mu }}{\sigma _{ee}^{B}\cdot (1+\delta _{ee})(1+\alpha _{ee})\varepsilon _{ee}+\sigma _{\mu \mu }^{B}\cdot (1+\delta _{\mu \mu })(1+\alpha _{\mu \mu })\varepsilon _{\mu \mu }}.\] \subsubsection{Radiative corrections} \begin{figure}[tb] \begin{center} \includegraphics[width=0.9\textwidth]{radcor.eps} \end{center} \caption{\label{rcorfig} Radiative corrections for $e^+e^-\rightarrow e^+e^-, \pi^+\pi^-, \mu^+\mu^-$ events} \end{figure} The radiative corrections are shown in Fig.~\ref{rcorfig}. The calculation of the radiative corrections for $e^+e^-\rightarrow e^+e^-$ events was based on \cite{radee}. The estimated systematic error is $\sim 1\%$. Radiative corrections for \( e^{+}e^{-}\rightarrow \mu ^{+}\mu ^{-},\pi ^{+}\pi ^{-} \) were calculated according to \cite{rc1} and \cite{rc2}. The estimated systematic error is $0.2-0.5\%$. The contribution from the lepton and hadron vacuum polarization is included in the radiative corrections for \( e^{+}e^{-}\rightarrow e^{+}e^{-},\mu ^{+}\mu ^{-} \), but excluded from the radiative correction for \( e^{+}e^{-}\rightarrow \pi ^{+}\pi ^{-} \). Some details about the radiative correction calculation could be found in \cite{PREP}. The radiative correction for $e^+e^-\rightarrow \pi^+\pi^-$ depends on the energy behaviour of the $e^+e^-\rightarrow \pi^+\pi^-$ cross-section itself. In order to take that into account, the calculation of this correction was done via a few iterations. At the first iteration, the existing $\left|F_\pi(s)\right|^2$ data were used for the calculation of the radiative correction. Using the calculated correction, we obtain $\left|F_\pi(s)\right|^2$. At subsequent iterations, $\left|F_\pi(s)\right|^2$ data obtained from the previous step, were used. It was found that after 3 iterations $\left|F_\pi(s)\right|^2$ values become stable. \subsubsection{Trigger efficiency} The trackfinder (TF) signal, neutral trigger (NT) signal and the calorimeter ``OR'' (CSI) trigger signal have been used for detector triggering during 1994-1995 runs. The trackfinder uses only the drift chamber and the Z-chamber data for a fast decision whether there is at least one track in the event. The neutral trigger analyses the geometry and energy deposition of the fired calorimeter rows. The calorimeter ``OR'' signal appears when there is at least one triggered calorimeter row. Two different trigger settings have been used during 1994-1995 runs. For 810(2)-960 MeV energy points the trigger was ``\textbf{(TF.and.CSI).or.NT}'' while for other energy points the trigger was ``\textbf{TF.or.NT}''. For analysis we use only those events where the trackfinder has found the track. If the trackfinder efficiency is $\varepsilon_{TF}$ and the CSI signal efficiency is $\varepsilon_{CSI}$, the overall trigger efficiency is $\varepsilon_{TF}\cdot\varepsilon_{CSI}$ for 810(2)-960 MeV energy points and $\varepsilon_{TF}$ for 610-810(1) MeV energy points. For the measurement of the trackfinder efficiency $e^+e^-\rightarrow e^+e^-$ events have been selected using only the calorimeter data (see the next section for details). Additionally, the positive decision of the neutral trigger was required. Analysing the probability to have also a positive decision of the trackfinder, its efficiency for $e^+e^-\rightarrow e^+e^-$ events can be calculated. The result is shown in Fig.\ \ref{tfefffig}. All types of collinear events are similar from the trackfinder point of view --- they have two collinear tracks and the track momenta are very close for different types of events. It is seen from the picture that the TF efficiency is high enough and that there is no visible energy dependence of the efficiency. Therefore we assume that the trackfinder efficiency is the same for all types of collinear events and hence cancels in (\ref{piform}). For the measurement of the CSI signal efficiency the fact that there are two clusters in the collinear event have been used: the CSI efficiency for a single cluster has been measured and was used for the calculation of the CSI efficiency for the event. The collinear events have been selected using the standard cuts. Then, the probability to trigger the CSI signal was measured for one cluster as a function of its energy. Using the measured probability and our knowledge about the energy deposition of particles of different types, the efficiency to trigger the CSI signal was calculated for different types of events. Results for 810(2)-960 MeV energy points are shown in Fig.\ \ref{csieff}. For simplicity, the values of 99.5\% for pions and 99.2\% for muons for the CSI efficiency were used for all 810-960 MeV energy points. \begin{figure} \begin{center} \begin{tabular}{c} \subfigure[\label{tfefffig} Trackfinder efficiency for $e^+e^-\rightarrow e^+e^-$ events] {\includegraphics[width=0.75\textwidth]{tfeff.eps}} \\ \subfigure[\label{csieff}CSI signal efficiency for collinear events] {\includegraphics[width=0.75\textwidth]{csieff2.eps}} \end{tabular} \end{center} \caption{\label{trgeff}Trigger efficiency for collinear events} \end{figure} \subsubsection{\label{receffsec}Reconstruction efficiency} \begin{figure} {\centering \resizebox*{0.9\textwidth}{!}{\includegraphics{receff.eps}} \par} \caption{\label{receff}Reconstruction efficiency for all energy points} \end{figure} The reconstruction efficiency was measured for $e^+e^-\rightarrow e^+e^-$ events using the experimental data: events were selected using only the calorimeter data, and then the probability to reconstruct two collinear tracks was measured. Since the selection criteria for collinear events are based on tracking data only, the measured probability is the reconstruction efficiency. The following selection criteria based only on the calorimeter data, were used. \begin{enumerate} \item There are exactly two clusters in the calorimeter. \item There is a hit in the Z-chamber near each cluster. This requirement selects the clusters produced by the charged particle. \item The energy deposition of both clusters is between \( (0.82\cdot E_{B}-40) \) and \( (0.82\cdot E_{B}+50) \) MeV. This is typical for $e^+$ and $e^-$, but not for other particles. \item The clusters are collinear if one takes into account the particle motion in the detector magnetic field: \begin{eqnarray*} & |\pi -(\Theta _{1}+\Theta _{2})|<0.1, & \\ & \left| \left| \pi -|\varphi _{1}-\varphi _{2}|\right| -2\arcsin \left( \frac{R\cdot 0.3B}{2\cdot E_{B} \cdot \sin \Theta }\right) \right| <0.1, & \end{eqnarray*} where \( \Theta \) and \( \varphi \) are the polar and azimuthal angles of the cluster, \( R \) is the effective calorimeter radius (45 cm) and \( B \) is the magnetic field (typically 10 kGs). \item The event was triggered by the trackfinder. Other triggers may be present. \end{enumerate} The selected sample contain only \( e^{+}e^{-}\rightarrow e^{+}e^{-} \) events with negligible ($<0.5\%$) background. If the number of selected events is \( N_{0} \) and the number of events, which in addition satisfy the standard selection criteria for collinear events is $N_1$, then the reconstruction efficiency is \[ \varepsilon _{rec}={N_{1}}/{N_{0}}.\] The measured reconstruction efficiency is shown in Fig.\ \ref{receff} (filled dots). The efficiency drops down for beam energies below 400 MeV. This additional inefficiency is completely explained by the lower precision of the \( \Theta \) angle measurement for these energy points. The empty dots in the same plot represent the reconstruction efficiency for the same selection criteria for collinear events except for \( \Delta \Theta \). As in studying the trigger efficiency, all types of collinear events are very similar from the point of view of the tracking reconstruction code. Therefore we assume that the reconstruction efficiency is the same for all types of collinear events and hence it cancels in (\ref{piform}). \subsubsection{Correction for the limited accuracy of the polar angle measurement} There is a special type of correction that arises since there is finite experimental resolution in the measurement of the $\Theta$ angles of tracks. Since we require the measured, rather than real, $\;\Theta_{avr}=[\Theta_1+(\pi-\Theta_2)]/2\;$ to be within $[\Theta_{min}, \pi-\Theta_{min}]$ interval, the visible cross-section is slightly different from \[ \int_{\Theta_{min}}^{\pi-\Theta_{min}} \frac{d\sigma}{d\Omega} \cdot 2\pi\sin\Theta \, d\Theta. \] The corresponding correction, denoted $\alpha$ in (\ref{piform}), was calculated as \[ 1+\alpha =\int\limits_{\Theta_{min}}^{\pi-\Theta_{min}}d\Theta \int\limits_{0}^{\pi}d\Theta^\prime \cdot f(\Theta^\prime) p(\Theta^\prime,\Theta) \left/ \int\limits_{\Theta_{min}}^{\pi-\Theta_{min}}d\Theta \cdot f(\Theta )\right. ,\] where $f(\Theta)$ is the theoretical $\Theta$-distribution and $p(\Theta^\prime,\Theta)$ is the detector response function taken as \[ p(\Theta ^{\prime },\Theta )=\frac{1}{\sqrt{2\pi }\sigma _{\Theta }}\exp \left[ -\frac{(\Theta ^{\prime }-\Theta )^{2}}{2\sigma _{\Theta }^{2}}\right] \] where \( \sigma _{\Theta } \) is the experimental resolution of \( \Theta_{avr} \) measurement. The resolution and the corresponding corrections for different types of collinear events for all energy points are shown in Fig.\ \ref{dthcor}. \begin{figure} {\centering \begin{tabular}{cc} \subfigure[$\Theta$-resolution]{\resizebox*{0.45\textwidth}{!}{\includegraphics{dthcor1.eps}}} & \subfigure[Correction to cross-section]{\resizebox*{0.45\textwidth}{!}{\includegraphics{dthcor2.eps}}} \\ \end{tabular}\par} \caption{\label{dthcor}The experimental resolution of the average polar angle measurement and the corresponding correction to cross-sections} \end{figure} \subsubsection{Correction for pion losses due to nuclear interactions} \begin{figure} \begin{center} \includegraphics[width=0.8\textwidth]{nuclcor.eps} \end{center} \caption{\label{nuccorr}The probability $\Delta_H$ to lose the \protect\( e^{+}e^{-}\rightarrow \pi ^{+}\pi ^{-}\protect \) event from the sample of collinear events due to nuclear interactions of pions with the beam pipe or drift chamber inner support} \end{figure} Unlike electrons and muons, a small fraction of pions has nuclear interactions in the beam pipe and therefore is lost from the selected samples of collinear events. The corresponding correction was calculated under the assumption that the event is lost from the collinear event sample if at least one pion had an inelastic nuclear interaction in the beam pipe or drift chamber inner support. The calculations were done separately for cross-sections obtained from two packages available in GEANT\cite{Geant} for simulation of nuclear interactions --- FLUKA\cite{FLUKA} and GHEISHA\cite{GHEISHA}. The resulting correction is shown in Fig.\ \ref{nuccorr}. It has been shown \cite{Krokovny} that FLUKA is more successful in describing experimental data about nuclear interactions of pions in our energy range, and therefore the results obtained from FLUKA were used for the cross-section correction while the difference between two calculations was added to the systematic error. \subsubsection{Correction for pion losses due to decays in flight} \begin{figure} \begin{center} \includegraphics[width=0.8\textwidth]{decaycorr.eps} \end{center} \caption{\label{decaycorr}The probability $\Delta_D$ to lose the \protect\( e^{+}e^{-}\rightarrow \pi ^{+}\pi ^{-}\protect \) event from the sample of collinear events due to pion decay in flight} \end{figure} A small fraction of pions decays in flight inside the drift chamber and the corresponding event may not be recognized as collinear if the decay angle between the pion and secondary muon is large enough. Depending on the beam energy 2-4\% of the $e^{+}e^{-}\rightarrow \pi ^{+}\pi ^{-}$ events have a decayed pion. But the maximum decay angle for the analysed energy range is small --- 80-150 mrad depending on the pion energy --- and therefore more than 90\% of such events are reconstructed as a pion pair. The probability to lose a pion pair due to the pion decay in flight was found from simulation and is shown in Fig.~\ref{decaycorr}. \subsubsection{Correction for the background from $\omega \rightarrow 3\pi$} \begin{figure} \begin{center} \includegraphics[width=0.8\textwidth]{3picorr.eps} \end{center} \caption{\label{3picorr}The $\Delta_{3\pi}$ contribution to the pion form factor from $\omega \rightarrow 3\pi$} \end{figure} There is a small probability to identify the $\pi^+\pi^-\pi^0$ final state as a $\pi^+\pi^-$ pair. Therefore the measured pion form factor is slightly bigger than it should be. This effect is significant only in the narrow energy range near the $\omega$-meson. The corresponding probability was estimated from simulation and was found to be \( 5\cdot 10^{-3} \) for \( \Theta _{min}=1.0 \) and \( 4\cdot 10^{-3} \) for \( \Theta _{min}=1.1 \) rad. The contribution to the pion form factor \( \Delta _{3\pi } \) was calculated assuming the $\omega$-meson parameters from \cite{PDG98} and is presented in Fig.\ \ref{3picorr}. \newpage \subsection{Systematic errors} The overall systematic error is estimated to be 1.5\% for 0.780 and 0.784 GeV, 1.7\% for 0.782, 0.84-0.87 and 0.94 GeV and 1.4\% for other energy points. The main sources of systematic error are summarized in Table \ref{systerr} and discussed in detail below. \begin{table}[h] \begin{center} \begin{tabular}{|c|c|} \hline Source & Estimated value\\ \hline \hline Events separation & 0.6\% \\ \hline Energy calibration of collider & 0.1\% (0.5\% for 390,392; 1\% for 391)\\ \hline Fiducial volume & 0.5\% \\ \hline Trigger efficiency & 0.2\% (1\% for 420-435, 470) \\ \hline Reconstruction efficiency & 0.3\% \\ \hline Hadronic interactions of pions & 0.4\% \\ \hline Pions decays in flight & 0.1\% \\ \hline Radiative corrections & 1\% \\ \hline \hline Total & 1.4\% \\ & (1.5\% for 390,392)\\ & (1.7\% for 391,420-435,470)\\ \hline \end{tabular} \end{center} \caption{\label{systerr}Main sources of systematic errors} \end{table} \paragraph{Event separation.} \begin{figure} \begin{center} \begin{tabular}{c} \subfigure[The relative difference between results of minimization of two different likelihood functions] {\includegraphics[height=0.3\textheight]{fpisim1.eps}} \\ \subfigure[The relative difference between simulated and reconstructed form factor values for the complete detector simulation of collinear events] {\includegraphics[height=0.3\textheight]{fpisim2.eps}} \end{tabular} \end{center} \caption{\label{fpisim} Estimation of the systematic error of events separation procedure. The relative difference $\Delta |F_\pi|^2 / |F_\pi|^2$ is shown with the error bars representing statistical error} \end{figure} Special studies were performed to estimate the systematic error due to event separation. \begin{enumerate} \item The separation was done by minimization of two independent likelihood functions. The ``standard'' likelihood function was described in detail in the previous sections. The second likelihood function was different in the following ways: different separation of background events; different energy deposition functions for minimum ionising muons and pions and for the nuclear interactions of pions (a single Gaussian was used instead of the set of Gaussians). The relative difference between two minimizations results is presented in Fig.\ \ref{fpisim}a. The error bars represent the statistical error. It is seen that there is no systematic shift and the difference is well below the statistical error. \item The complete detector simulation of $e^+e^-\rightarrow e^+e^-(\gamma),\mu^+\mu^-,\pi^+\pi^-$ events in the GEANT environment was done for 10 energy points: 0.61, 0.62, 0.64, 0.66, 0.70, 0.778, 0.782, 0.786, 0.860 and 0.960 GeV. 100000 of $e^+e^-\rightarrow e^+e^-(\gamma)$ events and the corresponding number of $e^+e^-\rightarrow\mu^+\mu^-,\pi^+\pi^-$ events were simulated for each energy point. The event separation was done by minimization of the same likelihood function as for real events. The relative difference between simulated and reconstructed form factor values is shown in Fig.\ \ref{fpisim}b. There is no systematic shift for all energy points except for the first three (0.61, 0.62 and 0.63 GeV). \item For some energy points there is a small difference between energy depositions of electrons and positrons, which comes mainly from a difference in the energy calibration of different parts of the calorimeter. The corresponding systematic error was estimated from the simulation. 100000 of $e^+e^-\rightarrow e^+e^-$ events and the corresponding number of $e^+e^-\rightarrow\mu^+\mu^-,\pi^+\pi^-$ events were simulated for 7 energy points. The energy deposition of particles was generated according to the energy deposition functions described in the previous sections, but with different parameters for $e^+$ and $e^-$. Then the event separation was done by minimization of the same likelihood function as for the real events. The systematic shift of about $0.5-0.8\%$ was observed between simulated and reconstructed form factor values. \end{enumerate} We estimate the average systematic error of event separation to be less than 0.6\%. For the first three energy points the systematic error is larger (4\%, 2\% and 1\% for 0.61, 0.62 and 0.63 GeV energy points respectively), but since the statistical error is a few times larger we neglect this fact. \paragraph{Energy calibration of the collider.} For almost all energy points the beam energy was measured using the resonance depolarization technique \cite{depol}. The systematic error in beam energy measurement does not exceed 50 keV. The corresponding systematic error in the pion form factor does not exceed 0.1\% for all energy points, except for the narrow \( \rho -\omega \) interference energy range, where it is estimated to be less than 0.5\% for 390 and 392 MeV energy points and less than 1\% for 391 MeV energy point. \paragraph{Fiducial volume (accuracy of \protect\(\Theta_{min}\protect\) measurement).} The Z coordinate of the track intersection with the Z-chamber is measured with better than 1 mm systematic accuracy. That allows, after the calibration of the drift chamber by the data from the Z-chamber \cite{dccal}, to have systematic accuracy in polar angle measurement better than 2.5 mrad. The corresponding systematic error in the pion form factor is 0.5\%. Since the pion form factor was measured independently for two \( \Theta _{min} \) values , we can estimate this systematic error by the average form factor deviation \( \left( 1-|F_{\pi }|^{2}\left| _{\Theta _{min}=1.0}\right. \left/ |F_{\pi }|^{2}\left| _{\Theta _{min}=1.1}\right. \right. \right) \). This deviation is shown in Fig.\ \ref{dfpith}. It is equal on the average to \( -0.1\%\pm 0.3\% \) and is consistent with the expected statistical fluctuation. \begin{figure}[h] {\centering \resizebox*{0.8\textwidth}{!}{\includegraphics{dfpith.eps}} \par} \caption{\label{dfpith}The relative difference in the form factor values, measured for $\Theta_{min}=1.0$ and $\Theta_{min}=1.1$ rad. The errors correspond to statistical fluctuation of the number of events in $1.1<\Theta<1.0$ region. In other words, the errors show the expected statistical difference for the form factor values measured for $\Theta_{min}=1.0$ and $\Theta_{min}=1.1$ } \end{figure} \paragraph{Trigger efficiency.} For most energy points the trackfinder efficiency is high ($>98\%$), and as it was discussed above, all types of collinear events are similar from the trackfinder point of view. Therefore it is assumed that the corresponding systematic error is negligible. This statement is questionable for 420-435 and 470 MeV energy points where the TF efficiency is lower than usual. The CsI efficiency is high for all energy points ($>99\%$) and was measured for all types of collinear events. Therefore the corresponding systematic error is negligible. The combined systematic error is estimated to be less than 0.2\% for all energy points except 420-435 and 470 MeV, where it is estimated to be less than 1\%. \paragraph{Reconstruction efficiency. } The reconstruction efficiency was measured only for Bhabha events, but it was found that it is uniform over \( \varphi \) and \( \Theta \) and that it comes mostly from errors in the track recognition algorithm. Therefore it is reasonable to assume that the reconstruction efficiency is the same for all types of collinear events. The efficiency drop for energy points below 800 MeV caused by the lower precision of $\Theta$ angle measurements is also the same for all types of collinear events. The corresponding systematic error is estimated to be less than 0.3\%. \paragraph{Correction for the nuclear interaction of pions.} The estimated accuracy of the correction calculation is about 20\%, determined by the cross-sections precision. The estimated contribution to this correction from nuclear interaction of pions inside the drift chamber and from elastic pion scattering by a large angle is about 30\% of the correction. Therefore the corresponding systematic error is less than 0.4\%. \paragraph{Correction for the pions decays in flight.} The systematic error of this correction is estimated to be less than 0.1\%. \paragraph{Radiative corrections.} The systematic error in radiative corrections is dominated by the error in radiative corrections for \( e^{+}e^{-}\rightarrow e^{+}e^{-} \), where it is estimated to be less than 1\%. \section{A fit of the pion form factor} \afterpage{ \clearpage \begin{longtable}{|c|c|c|c|} \hline 2E (MeV) & $N_{\pi\pi}$ & $\sigma_{\pi\pi}$ (nb) & $|F_\pi|^2$ \\ \hline \endhead \multicolumn{4}{|r|}{{\it continued on the next page}} \\ \hline \multicolumn{4}{c}{ } \\ \caption[]{The experimental data from the CMD-2 detector} \endfoot \multicolumn{4}{c}{ } \\ \caption{\label{data}The experimental data from the CMD-2 detector} \endlastfoot \hline \input{rho9495.tab1} \end{longtable} } The pion form factor data, obtained from the described analysis, is summarized in Table \ref{data}. Only statistical errors are shown. It is well known that to describe the pion form factor the higher resonances $\rho(1450)$ and $\rho(1700)$ should be taken into account in addition to leading contribution from $\rho(770)$ and $\omega(782)$ \cite{OLYACMD,tau}. At the same time it was shown \cite{Newrho} that the experimental data below 1 GeV is well described by the model based on the hidden local symmetry, which predicts a point-like coupling $\gamma\pi^+\pi^-$. Here we use both approaches to fit the data. Since we fit the pion form factor in the relatively narrow energy region 0.61-0.96 GeV, only one higher resonance $\rho(1450)$ is taken into account. \subsection{The Gounaris-Sakurai (GS) parametrization} This model is based on the Gounaris-Sakurai parametrization of the $\rho$-reso\-nance. Its modifications were used for parametrization of all previous $e^+e^-$ data \cite{OLYACMD} and $\tau$ decay data \cite{tau}. To describe the data, the $\rho(770)$, $\rho(1450)$ contributions and $\rho-\omega$ interference were taken into account: \begin{equation} \label{GS} F_\pi(s)=\frac{\BW^{\mathrm{GS}}_{\rho(770)}(s)\cdot \displaystyle\frac{\mathstrut 1+\delta \, \mathrm{BW}_{\omega}(s)}{1+\delta} + \beta \, \BW^{\mathrm{GS}}_{\rho(1450)}(s) }{1+\beta}. \end{equation} The $\rho(770)$ and $\rho(1450)$ contributions are taken according to the Gouna\-ris-Sakurai model \cite{gunsac}: \begin{equation} \BW^{\mathrm{GS}}_{\rho(\mathrm{M}_\rho)} = \frac{ \mathrm{M}_\rho^2 \left( 1 + d \cdot \Gamma_\rho / \mathrm{M}_\rho \right) }{ \mathrm{M}_\rho^2 - s + f(s) - i \mathrm{M}_\rho \Gamma_\rho(s) }, \end{equation} where \begin{equation} f(s) = \Gamma_\rho \frac{\mathrm{M}_\rho^2}{p_\pi^3(\mathrm{M}_\rho^2)} \left[ p_\pi^2(s) \left( h(s) - h(\mathrm{M}_\rho^2) \right) + (\mathrm{M}_\rho^2-s) \, p_\pi^2(\mathrm{M}_\rho^2) \left. \frac{dh}{ds} \right|_{s=\mathrm{M}_\rho^2} \right] , \end{equation} \begin{equation} h(s) = \frac{2}{\pi} \frac{p_\pi(s)}{\sqrt{s}} \ln \frac{\sqrt{s}+2 p_\pi(s)}{2m_\pi} , \end{equation} \begin{equation} \left. \frac{dh}{ds} \right|_{s=\mathrm{M}_\rho^2} = h(\mathrm{M}_\rho^2) \left[ \frac{1}{8p_\pi^2(\mathrm{M}_\rho^2)} - \frac{1}{2\mathrm{M}_\rho^2} \right] + \frac{1}{2\pi\mathrm{M}_\rho^2} , \end{equation} \begin{equation} p_\pi (s) = \frac{1}{2} \sqrt{s-4m_\pi^2}. \end{equation} The energy dependent width is \begin{equation} \label{rhowidth} \Gamma_\rho(s) = \Gamma_\rho \left[ \frac{p_\pi(s)}{p_\pi(\mathrm{M}_\rho^2)} \right]^3 \left[ \frac{\mathrm{M}_\rho^2}{s} \right]^{1/2}. \end{equation} The normalization $\BW^{\mathrm{GS}}_{\rho(\mathrm{M}_\rho^2)}(0)=1$ fixes the parameter $d$: \begin{equation} d = \frac{3}{\pi} \frac{m_\pi^2}{p_\pi^2(\mathrm{M}_\rho^2)} \ln \frac{\mathrm{M}_\rho + 2 p_\pi(\mathrm{M}_\rho^2)}{2m_\pi} + \frac{\mathrm{M}_\rho}{2\pi p_\pi(\mathrm{M}_\rho^2)} - \frac{m_\pi^2 \mathrm{M}_\rho}{\pi p_\pi^3(\mathrm{M}_\rho^2)}. \end{equation} The simple Breit-Wigner parametrization (\ref{BW}) is used for the $\omega(782)$ contribution. In order to extract $\Gamma(\rho\rightarrow e^+ e^-)$ we relate the pion form factor at the $\rho$ mass to the VMD form factor: \begin{equation} \left. F_\pi(s) \right|_{s=\mathrm{M}_\rho^2} = \frac{ g_{\rho\gamma} g_{\rho\pi\pi} }{\mathrm{M}_\rho^2 - s - i \mathrm{M}_\rho \Gamma_\rho} + \left( \begin{array}{c} \mathrm{non-resonant} \\ \mathrm{contribution} \end{array} \right) . \end{equation} From this one can obtain \begin{equation} g_{\rho\gamma} g_{\rho\pi\pi} = \frac{ \mathrm{M}_\rho^2 \left( 1 + d \cdot \Gamma_\rho / \mathrm{M}_\rho \right) } {(1+\delta)(1+\beta)}. \end{equation} Using well known VMD relations \cite{Newrho} \begin{equation} \label{Gvee} \Gamma_{V\rightarrow e^+e^-} = \frac{4\pi\alpha^2}{3\mathrm{M}_V^3} g_{V\gamma}^2, \end{equation} \begin{equation} \label{Gvpp} \Gamma_{V\rightarrow\pi^+\pi^-} = \frac{g_{V\pi\pi}^2}{6\pi} \frac{p_\pi^3 \left( \mathrm{M}_V^2 \right)}{\mathrm{M}_V^2} \end{equation} and assuming that $\Gamma_{\rho\rightarrow\pi^+\pi^-}=\Gamma_\rho$ one obtains \begin{equation} \Gamma_{\rho\rightarrow e^+e^-} = \frac{2\alpha^2 p_\pi^3 \left( \mathrm{M}_\rho^2 \right)}{9 \mathrm{M}_\rho \Gamma_\rho} \frac{(1+ d \cdot \Gamma_\rho / \mathrm{M}_\rho )^2}{(1+\delta)^2(1+\beta)^2}. \end{equation} A similar approach is used for the calculation of $Br(\omega\rightarrow\pi^+\pi^-)$. One can relate the pion form factor at the $\omega$ mass to the VMD form factor: \begin{equation} \left. F_\pi(s) \right|_{s=\mathrm{M}_\omega^2} = \frac{ g_{\omega\gamma} g_{\omega\pi\pi} }{\mathrm{M}_\omega^2 - s - i \mathrm{M}_\omega \Gamma_\omega} + \left( \begin{array}{c} \mathrm{non-resonant} \\ \mathrm{contribution} \end{array} \right) . \end{equation} From this one determines $g_{\omega\pi\pi}$: \begin{equation} g_{\omega\gamma} g_{\omega\pi\pi} = \frac{ \delta \cdot \mathrm{M}_\omega^2 \cdot |\BW^{\mathrm{GS}}_{\rho(770)}(\mathrm{M}_\omega^2)| } {(1+\delta)(1+\beta)}. \end{equation} Using (\ref{Gvee}) and (\ref{Gvpp}) one derives \begin{equation} Br(\omega\rightarrow\pi^+\pi^-) = \frac{2\alpha^2 p_\pi^3(\mathrm{M}_\omega^2)} {9\mathrm{M}_\omega\Gamma_{\omega\rightarrow e^+e^-}\Gamma_\omega} \left| \BW^{\mathrm{GS}}_{\rho(770)}(\mathrm{M}_\omega^2) \right|^2 \frac{\delta^2}{(1+\delta)^2 (1+\beta)^2}. \end{equation} \subsection{The Hidden Local Symmetry (HLS) parametrization} In the Hidden Local Symmetry (HLS) model \cite{Newrho,HLS,rhoomega} the $\rho$-meson appears as a dynamical gauge boson of a hidden local symmetry in the non-linear chiral Lagrangian. This model introduces a real parameter $a$ related to the non-resonant coupling $\gamma\pi^+\pi^-$. The resulting pion form factor for the HLS model is \begin{equation} \label{HLS} F_\pi (s)= -\frac{a}{2}+1+\frac{a}{2} \cdot \mathrm{BW}_\rho(s) \cdot \frac{ 1 + \delta\,\mathrm{BW}_{\omega}(s)}{1+\delta}, \end{equation} where \begin{equation} \label{BW} \mathrm{BW}_{V}(s) = \frac{\mathrm{M}_V^2}{\mathrm{M}_V^2-s-i\mathrm{M}_V\Gamma_V(s)}. \end{equation} The energy dependent width $\Gamma_\rho(s)$ is the same as for the GS parametrization and is given by (\ref{rhowidth}). Using the similar approach as for the GS parametrization, we obtain \begin{equation} \Gamma_{\rho\rightarrow e^+e^-} = \frac{2\alpha^2 p_\pi^3 \left( \mathrm{M}_\rho^2 \right)}{9 \mathrm{M}_\rho \Gamma_\rho} \left[ \frac{a}{2(1+\delta)} \right]^2, \end{equation} \begin{equation} Br(\omega\rightarrow\pi^+\pi^-) = \frac{2\alpha^2 p_\pi^3(\mathrm{M}_\omega^2)} {9\mathrm{M}_\omega\Gamma_{\omega\rightarrow e^+e^-}\Gamma_\omega} \left| \frac{a}{2}\mathrm{BW}_\rho(\mathrm{M}_\omega^2) \right|^2 \frac{\delta^2}{(1+\delta)^2}. \end{equation} \subsection{Result of fit} The fit to data was done by the minimization of the following $\chi^2$ function: \begin{equation} \chi^2 = \sum_{i=1\ldots 44} \frac{ \left( |F_\pi|^2_{exp}(s_i) - |F_\pi|^2_{theor}(s_i) \right)^2 } { \Delta_i^2 } + \sum_{j=1\ldots N_p} \frac{ \left( p_j - p_{0j} \right)^2 }{\sigma_j^2}, \end{equation} where $|F_\pi|^2_{exp}(s_i)$ and $|F_\pi|^2_{theor}(s_i)$ are the experimental and theoretical values of the pion form factor at the $i$-th energy point, $\Delta_i$ is the experimental error at the $i$-th energy point, $N_p$ is the number of model parameters, $p_j$ is the $j$-th model parameter, $p_{0j}$ is the expected value of $j$-th model parameter and $\sigma_j$ is the estimated error of $j$-th model parameter. The model parameters are the following (all values are taken from PDG'98 \cite{PDG98}). \begin{enumerate} \item The $\omega$ meson mass: $p_{01}=781.94$ MeV, $\sigma_1=0.12$ MeV. \item The $\omega$ meson width: $p_{02}=8.41$ MeV, $\sigma_2=0.09$ MeV. \item The $\omega$ meson leptonic width: $p_{03}=0.60$ keV, $\sigma_3=0.02$ keV. \item The $\rho(1450)$ meson mass: $p_{04}=1465$ MeV, $\sigma_4=25$ MeV. \item The $\rho(1450)$ meson width: $p_{05}=310$ MeV, $\sigma_5=60$ MeV. \end{enumerate} For the GS fit the number of model parameters is $N_p=5$, while for the HLS fit $N_p=3$. The parameter $\beta$ in (\ref{GS}) was assumed to be real. To determine a phase of $\rho-\omega$ mixing, the parameter $\delta$ in (\ref{GS}) and (\ref{HLS}) was assumed to be complex during the fit. However, it turns out that the phase was always consistent with zero. Therefore $\delta$ was assumed to be real for the final fit. The results of the fit of the CMD-2 experimental data are shown in Fig.\ \ref{fpifit} and are summarized in Table \ref{rhopar}. The first error is statistical, the second error is systematic. In order to estimate a systematic error, the fit was repeated with all data points shifted up and down by one systematic error. \begin{figure} {\centering \resizebox*{\textwidth}{!}{\includegraphics{fpi.eps}} \par} \caption{\label{fpifit}Fit of the CMD-2 (94, 95) pion form factor data according to GS and HLS models. Both theoretical curves are presented, but they are indistinguishable} \end{figure} \begin{table} \begin{center} \begin{tabular}{|l|c|c|} \hline & GS model & HLS model \\ \hline $\mathrm{M}_\rho$, MeV & $775.28\pm 0.61\pm 0.20$ & $774.57\pm 0.60\pm 0.20$ \\ \hline $\Gamma_\rho$, MeV & $147.70\pm 1.29 \pm 0.40$ & $147.65\pm 1.38\pm 0.20$ \\ \hline $Br(\omega\rightarrow\pi^+\pi^-)$, \% & $1.31\pm 0.23\pm 0.02$ & $1.32\pm 0.23\pm 0.02$ \\ \hline $\Gamma(\rho\rightarrow e^+e^-)$, keV & $6.93\pm 0.11\pm 0.10$ & $6.89\pm 0.12\pm 0.10$ \\ \hline $\beta$ (GS) & $-0.0849\pm 0.0053\pm 0.0050$ & --- \\ \hline $a$ (HLS) & --- & $2.381\pm 0.016\pm 0.016$ \\ \hline $\chi^2/n$ & 0.77 & 0.78 \\ \hline \end{tabular} \end{center} \caption{\label{rhopar}The results of fit of the CMD-2 (94,95) pion form factor data by GS and HLS models} \end{table} \subsection{Hadronic Contribution to (g-2)$_{\mu}$} We'll now estimate the implication of our results for the value and uncertainty of $a_{\mu}^{had}$ - the hadronic contribution to (g-2)$_{\mu}$. To this end we'll choose the c.m.energy range near the $\rho$-meson peak (from 630 to 810 MeV) which gives the dominant contribution to the muon anomaly and where good data are available from previous measurements by the OLYA and CMD groups \cite{OLYACMD} as well as by the DM1 detector \cite{DM1}. Table \ref{g-2} presents results of our calculations of $a_{\mu}^{had}$ performed by direct integration of the experimental data over the energy range above. The calculation used the following procedure: first we integrated the data of each group over the chosen energy range. Since measurements at different energy points are independent, the squared statistical error of the integral can be obtained by summing statistical errors squared of each energy point. After that the systematic uncertainty which is believed to be an overall normalization uncertainty is added quadratically. Contributions of separate groups are subject to weighted averaging taking into account if necessary a scale factor \cite{PDG98}. Such a procedure implies an assumption that systematical uncertainties are uncorrelated since they refer to different measurements. \begin{table}[h] \begin{center} \begin{tabular}{|l|c|c|} \hline Data & a$_{\mu}^{had}$, 10$^{-10}$ & Total error, 10$^{-10}$ \\ \hline Old & 284.2 $\pm$ 3.8 $\pm$ 7.1 & 8.1 \\ \hline New & 292.5 $\pm$ 2.2 $\pm$ 4.1 & 4.7 \\ \hline Old+New & 290.3 $\pm$ 1.9 $\pm$ 3.6 & 4.1 \\ \hline \end{tabular} \end{center} \caption{\label{g-2}Hadronic contributions to (g-2)$_{\mu}$, coming from the c.m.energy range near the $\rho$-meson peak.} \end{table} The first line gives the average of three independent estimates based on the data of OLYA, CMD and MD1 while the second one presents the corresponding value for the CMD-2 data. The third line of the Table is the average based on the four independent estimates above. For convenience, we list separately the statistical and systematic uncertainties in the second column while the third one gives the total error obtained by adding them quadratically. One can see that the estimate based on the CMD-2 data is in good agreement with that coming from the old data. Note also the significant improvement of both statistical and systematic uncertainties compared to the previous measurements. The combined result (old plus new data) is a factor of two more precise than before. \section{Conclusion} A new measurement of the pion form factor in the center-of-mass energy range $0.61-0.96$ GeV is presented. The statistical error is about the same as for all previous $e^+e^-$ data, but the systematic error is a few times smaller. The results are based only on part of the experimental data taken by CMD-2. There are several reasons why this data set was treated as independent. First, a different approach in data analysis is required for the data taken in the center-of-mass energy range below $0.6$ GeV and above $1.0$ GeV (runs 4 and 5 in table \ref{runstable}). Second, the data taking conditions were significantly different in 1998 (run 6 in table \ref{runstable}), when the data was taken in the same $0.61-0.96$ GeV energy range. As a result, the systematic errors for these runs will be different than for the data presented here. Therefore we prefer to present separate results for different groups of runs. The fit of the pion form factor based on the Gounaris-Sakurai and the Hidden Local Symmetry parametrizations was performed. For the final values we prefer the Gounaris-Sakurai parametrization as it was traditionally used for the determination of the $\rho$-meson parameters (see \cite{OLYACMD} and \cite{tau}). The difference between the results of the fit in two models is always smaller than the experimental error (Table \ref{rhopar}), therefore the ``model'' error is not specified. We give as the final results the following, where the first error is statistical and the second one is systematic: \begin{equation} \left\{ \begin{array}{ll} \mathrm{M}_\rho \, (\mathrm{MeV}) & = 775.28\pm 0.61\pm 0.20 , \vphantom{\Bigl( \Bigr)} \\ \Gamma_\rho \, (\mathrm{MeV}) & = 147.70\pm 1.29 \pm 0.40 , \vphantom{\Bigl( \Bigr)} \\ \Gamma(\rho\rightarrow e^+e^-) \, (\mathrm{keV}) & = 6.93\pm 0.11\pm 0.10 , \vphantom{\Bigl( \Bigr)} \\ Br(\omega\rightarrow\pi^+\pi^-) & = ( 1.31\pm 0.23 ) \% . \vphantom{\Bigl( \Bigr)} \end{array} \right. \end{equation} It has also been shown that the improvement of the experimental precision, particularly of the systematic uncertainties, can be crucial for the high precision calculation of the hadronic contribution to the muon anomaly. With the progress of analysis of the rest available data we hope to reduce the systematic error for the data presented here by additional factor of two. Important part of this improvement is the development of the new approach to the radiative corrections calculation with the ultimate goal to reduce the corresponding systematic error to $(0.3-0.5)\%$ level. Using the high statistics taken in 1998, it will be possible to perform the detailed analysis of small systematic effects. 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. We acknowledge the contribution to the experiment at the earlier stages of our colleagues V.A.Monich, A.E.Sher, V.G.Zavarzin and W.A.Worstell. Special thanks are due to N.N.Achasov, A.B.Arbuzov, M.Benayoun, V.L.Chernyak, V.S.Fadin, F.Jegerlehner, E.A.Kuraev, A.I.Milstein and G.N.Shestakov for useful discussions and permanent interest. This work is supported in part by grants RFBR-98-02-17851, INTAS 96-0624 and DOE DEFG0291ER40646.
\section{Deformations of cohomology algebras of DGBV algebras} \label{sec:deform} Let ${\bf k}$ be a graded commuative associative algebra with unit over ${\Bbb Q}$, $({\cal A}, \wedge)$ a graded commutative associative algebra with unit $1$ over ${{\bf k}}$. For any linear operator $\Delta$ of odd degree, define \begin{eqnarray} \label{eqn:bracket} [a \bullet b]_{\Delta} = (-1)^{|a|} (\Delta (a \wedge b) - (\Delta a) \wedge b - (-1)^{|a|} a \wedge \Delta b), \end{eqnarray} for homogeneous elements $a, b \in {\cal A}$. When there is no risk of confusion, we will simply use $[\cdot \bullet \cdot]$ to denote $[\cdot\bullet\cdot]_{\Delta}$. If $\Delta^2 = 0$ and \begin{eqnarray} \label{eqn:derivation} [a \bullet (b \wedge c)] = [a \bullet b] \wedge c + (-1)^{(|a|-1)|b|} b \wedge [a \bullet c], \end{eqnarray} for all homogeneous $a, b, c \in {\cal A}$, then $({\cal A}, \wedge, \Delta, [\cdot\bullet\cdot])$ is a {\em Gerstenhaber-Batalin-Vilkovisky (GBV) algebra}. Under the above conditions, it is straightforward to see that \begin{eqnarray*} && [a \bullet b] = - (-1)^{(|a|-1)(|b|-1)}[b \bullet a], \\ && [a \bullet [b \bullet c]] = [[a \bullet b] \bullet c] + (-1)^{(|a| -1)(|b| -1)} [b \bullet [a \bullet c]], \\ && \Delta [a\bullet b] = [\Delta a \bullet b] + (-1)^{|a|+1} [a \bullet \Delta b]. \end{eqnarray*} A {\em DGBV (differential Gerstenhaber-Batalin-Vilkovisky) algebra} is a GBV algebra with a ${\bf k}$-linear derivation $\delta$ of odd degree with respect to $\wedge$, such that $$ \delta^2 = \delta\Delta + \Delta \delta = 0.$$ It is easy to see that \begin{eqnarray*} && \delta [a\bullet b] = [\delta a \bullet b] + (-1)^{|a|+1} [a \bullet \delta b]. \end{eqnarray*} It is routine to define homomorphisms of DGBV algebras. Denote by ${\cal D}{\cal G}{\cal B}{\cal V}$ the category of DGBV algebra. One can define direct sum and tensor product in this category. Let $\delta_{x(t)} = \delta + [x(t) \bullet \cdot]$, where $t$ is an indeterminate and $x(t) \in {\cal A}[[t]]$ is a formal power series in $t$ with coefficients even elements of ${\cal A}$. Because of (\ref{eqn:derivation}), $\delta_x$ is a derivation of ${\cal A}[[t]]$. Now $$ \delta_{x(t)}^2 = [(\delta x(t) + \frac{1}{2} [x(t) \bullet x(t)]) \bullet \cdot]. $$ Hence if \begin{eqnarray} \label{eqn:MC} \delta x(t) + \frac{1}{2} [x(t) \bullet x(t)] = 0, \end{eqnarray} then $\delta_{x(t)}$ is a differential. It is easy to see that if $\Delta x(t) = 0$, then $\delta_{x(t)} \Delta = - \Delta \delta_{x(t)}$. As a consequence, $({\cal A}[[t]], \wedge, \delta_{x(t)}, \Delta, [\cdot \bullet \cdot])$ is a DGBV algebra. We will be concerned with the deformations of the multiplicative structure on the cohomology $H = H({\cal A}, \delta)$. Since $\delta$ is a derivation, i.e., $$\delta (a \wedge b) = \delta a \wedge b + (-1)^{|a|} a \wedge \delta b,$$ for homogeneous $a, b \in {\cal A}$, $\wedge$ induces an associative product (also denoted by $\wedge$) on $H$. Clearly, $(H, \wedge)$ is graded commutative. Since we have $\delta 1 = 0$, the class of $1$ gives a unit of $(H, \wedge)$. To summarize, $(H, \wedge)$ is a graded commutative associative algebra with unit over ${\bf k}$. Note that $({\cal A}[[t]], \wedge, \delta_{x(t)}, \Delta, [\cdot \bullet \cdot])$ is a formal deformation of $({\cal A}, \wedge, \delta, \Delta, [\cdot \bullet \cdot])$. An important idea in \cite{Bar-Kon} and \cite{Man2} is to obtain formal deformations of $(H, \wedge)$ by considering $H({\cal A}[[t]], \delta_{x(t)})$. We will need the following \begin{lemma} \label{lm:quasi} Let ${\cal A}$ be a vector space with two endomorphisms $\delta$ and $\Delta$ satisfying $\delta^2 = \Delta^2 = \delta \Delta + \Delta \delta = 0$. Then $\Img \delta \Delta = \Img \Delta \delta \subset \Img \delta \cap \Ker \Delta$ and $\Img \Delta \delta = \Img \delta \Delta \subset \Img \Delta. \cap \Ker \delta$. The following conditions are equivalent \begin{itemize} \item[(i)] The inclusions $i: (\Ker \Delta, \delta) \hookrightarrow ({\cal A}, \delta)$ and $j: (\Ker \delta, \Delta) \hookrightarrow ({\cal A}, \Delta)$ induce isomorphisms of homology. \item[(ii)] We have equalities: \begin{align*} &&\Img \delta \Delta = \Img \Delta \delta = \Img \delta \cap \Ker \Delta, \\ &&\Img \delta \Delta = \Img \Delta \delta = \Img \Delta \cap \Ker \delta. \end{align*} \end{itemize} When the above conditions hold, then both the cohomology groups in (i) are naturally isomorphic to $(\Ker \Delta \cap \Ker \delta)/\Img \delta \Delta$. \end{lemma} From now on, all the DGBV algebras will be assumed to satisfy the conditions in Lemma \ref{lm:quasi}. Denote by ${\cal D}{\cal G}{\cal B}{\cal V}_q$ the subcategory of all DGBV algebras which satisfies the conditions in Lemma \ref{lm:quasi}. \begin{lemma} \label{lm:tech} If $z \in {\cal A}$ satisfies $\delta \Delta z = 0$, then $$z = h + \Delta u + \delta v$$ for some $h \in \Ker \delta \cap \Ker \Delta$, $u, v \in {\cal A}$. \end{lemma} \begin{proof} Since $\Delta z \in \Ker \delta \cap \Img \Delta = \Img \delta \Delta$, so $\Delta z = \Delta \delta v$ for some $v \in {\cal A}$. Now $\Delta (z - \delta v) = 0$, so $z - \delta v$ determines a class in $H({\cal A}, \Delta)$. Since $H({\cal A}, \Delta) \cong (\Ker \delta \cap \Delta)/\Img \delta \Delta$, there exist $u \in {\cal A}$, such that $h = z - \delta v - \Delta u \in Ker \delta \cap \Ker \Delta$. This completes the proof. \end{proof} \begin{proposition} \label{prop:MC} Assume that $({\cal A}, \wedge, \delta, \Delta, [\cdot\bullet\cdot])$ is in ${\cal D}{\cal G}{\cal B}{\cal V}_q$. Then for any even cohomology class $[x] \in H({\cal A}, \delta)$, there is a formal power series solution $x(t) = x_1 t + \cdots + x_n t^n + \cdots$ to (\ref{eqn:MC}), such that $x_1 \in \Ker \Delta \cap \Ker \delta$ represents $[x]$, and $x_n \in \Img \Delta$ for $n > 1$. When $[x] = [1]$, we can take $x_1 = 1$, $x_n = 0$, for $n > 1$. \end{proposition} \begin{proof} The second statement is trivial. The first can be proved by a standard argument modeled on the method of Tian \cite{Tia} and Todorov \cite{Tod}. Rewrite (\ref{eqn:MC}) as a sequence of equations \begin{eqnarray*} && \delta x_1 = 0, \\ && \delta x_2 = - \frac{1}{2} [x_1 \bullet x_1], \\ && \cdots \cdots \\ && \delta x_n = - \frac{1}{2} \sum_{i + j = n} [x_i \bullet x_j], \\ && \cdots \cdots \end{eqnarray*} By Lemma \ref{lm:quasi}, we can take $x_1 \in \Ker \delta \cap \Ker \Delta$ to represent $[x]$. Suppose now we have found $x_1, \cdots, x_n$ with $x_j \in \Img \Delta$ for $1 < j \leq n$. By (\ref{eqn:bracket}), $\sum_{i + j = n + 1} [x_i \bullet x_j] \in \Img \Delta$. Also we have the following standard calculation \begin{eqnarray*} && \delta \sum_{i + j = n + 1} [x_i \bullet x_j] = \sum_{i + j = n + 1} [\delta x_i \bullet x_j] - \sum_{i + j = n + 1} [x_i \bullet \delta x_j] \\ & = & -\frac{1}{2} \sum_{i + j = n + 1} \sum_{p + q = i} [[x_p \bullet x_q] \bullet x_j] + \frac{1}{2} \sum_{i + j = n + 1} \sum_{p + q = j} [x_i \bullet [x_p \bullet x_q]] \\ & = & -\frac{1}{2} \sum_{i + j = n + 1} \sum_{p + q = i} [[x_p \bullet x_q] \bullet x_j] + \frac{1}{2} \sum_{i + j = n + 1} \sum_{p + q = j} [[x_p \bullet x_q] \bullet x_i] = 0. \end{eqnarray*} Hence $\sum_{i + j = n + 1} [x_i \bullet x_j] \in \Ker \delta \cap \Img \Delta = \Img \delta \Delta$, and so there exist $x_{n+1} \in \Img \Delta$, such that $$\delta x_{n+1} = - \frac{1}{2} \sum_{i + j = n+1} [x_i \bullet x_j].$$ \end{proof} \begin{proposition} \label{prop:quasi2} The triple $({\cal A}[[t]], \delta_{x(t)}, \Delta)$ is in ${\cal D}{\cal G}{\cal B}{\cal V}_q$. \end{proposition} \begin{proof} We need to show \begin{eqnarray*} && \Img \delta_{x(t)} \cap \Ker \Delta \subset \Img \delta \Delta = \Img \Delta \delta, \\ && \Img \Delta \cap \Ker \delta_{x(t)} \subset \Img \delta \Delta = \Img \Delta \delta. \end{eqnarray*} We prove the second inclusion first. Assume that $y(t) = \Delta z(t)$, $\delta_{x(t)}y(t) = 0$. We get a sequence of equations \begin{eqnarray*} && \delta y_0 = 0, \\ && \delta y_1 = - [x_1 \bullet y_0], \\ && \cdots \cdots \\ && \delta y_n = - \sum_{i + j = n} [x_i \bullet y_j], \\ && \cdots \cdots \end{eqnarray*} where $y_n = \Delta z_n$. Now $\delta \Delta z_0 = \delta y_0 = 0$, by Lemma \ref{lm:tech}, $z_0 = h_0 + \Delta u_0 + \delta v_0$, where $h_0 \in \Ker \delta \cap \Ker \Delta$. Hence $$y_0 = \Delta z_0 = \Delta \delta v_0 = \delta (- \Delta v_0).$$ And from \begin{eqnarray*} \delta\Delta z_1 = \delta y_1 = - [x_1 \bullet y_0] = [x_1 \bullet \delta \Delta v_0] = \delta \Delta [x_1 \bullet v_0], \end{eqnarray*} we get $z_1 = h_1 + \Delta u_1 + \delta v_1 + [x_1 \bullet v_0] $. Hence we have \begin{eqnarray*} y_1 = \Delta z_1 = \Delta \delta v_1 + \Delta [x_1 \bullet v_0] = - \delta \Delta v_1 - [x_1 \bullet \Delta v_0]. \end{eqnarray*} By induction, we find that $$y_n = - \delta \Delta v_n - \sum_{i + j = n} [x_i \bullet v_j],$$ in other word, $y(t) = - \delta_{x(t)} \Delta v(t) \in \Img \delta_{x(t)} \Delta$, where $v(t) = v_0 + v_1 t + \cdots$. Now assume $y(t) = \delta_{x(t)} z(t)$ and $\Delta y(t) = 0$, where $y(t) = y_0 + y_1 t + \cdots$ and $z(t) = z_0 + z_1 t+ \cdots$ are elments of ${\cal A}[[t]]$. Equivalently, we have a sequence of equations \begin{eqnarray*} && y_0 = \delta z_0, \\ && y_1 = \delta z_1 + [x_1 \bullet z_0], \\ && \cdots \cdots \\ && y_n = \delta z_n + \sum_{i + j = n} [x_i \bullet z_j], \\ && \cdots \cdots \end{eqnarray*} and $\Delta y_n = 0$. Now $\Delta \delta z_0 = \Delta y_0 = 0$, $z_0 = h_0 + \Delta u_0 + \delta v_0$ for some $h_0 \in \Ker \delta \cap \Ker \Delta$. Hence $$y_0 = \delta z_0 = \delta \Delta v_0.$$ From \begin{eqnarray*} && \Delta \delta z_1 = \Delta (y_1 - [x_1 \bullet z_0]) = [x_1 \bullet \Delta z_0] = [x_1 \bullet \Delta \delta v_0] = \Delta \delta [x_1 \bullet v_0], \end{eqnarray*} we get $z_1 = h_1 + \Delta u_1 + \delta v_1 + [x_1 \bullet v_0]$. Hence \begin{eqnarray*} y_1 & = & \delta z_1 + [x_1 \bullet z_0] = \delta \Delta u_1 + \delta [x_1 \bullet v_0] + [x_1 \bullet (h_0 + \Delta u_0 + \delta v_0)] \\ & = & \delta \Delta u_1 + [x_1 \bullet h_0] + [x_1 \bullet \Delta u_0]. \end{eqnarray*} By induction, we can show that \begin{eqnarray*} z_n = h_n + \Delta u_n + \delta v_n + \sum_{i+j= n} [x_i \cdot v_j], \end{eqnarray*} where each $h_n$ lies in $\Ker \delta \cap \Ker \Delta$. Consequently, \begin{eqnarray*} y_n & = & \delta \Delta u_n + \sum_{i + j = n} [x_i \bullet h_j] + \sum_{i + j = n} [x_i \bullet \Delta u_j] \end{eqnarray*} Define $w_n$ as follows. Set $w_0 = 0$. For $n > 0$, assume that $w_1, \cdots w_{n-1}$ have been defined such that $$\delta \Delta w_j = \sum_{q + r = j} [x_p \bullet (h_q - \Delta w_q)],$$ for $j < n$. Then it is straightforward to see that $\sum_{i + j = n} [x_i \bullet (h_j - \Delta w_j)] = \sum_{i + j = n} \Delta (x_i \wedge (h_j - \Delta w_j)) \in \Ker \delta \cap \Img \Delta = \Img \delta \Delta$, so $$ \sum_{i + j = n} [x_i \bullet (h_j - \Delta w_j)]= \delta \Delta w_n$$ for some $w_n \in {\cal A}$. Finally, we have \begin{eqnarray*} y_n & = & \delta \Delta u_n + \sum_{i + j = n} [x_i \bullet (h_j - \Delta w_j] + \sum_{i + j = n} [x_i \bullet \Delta (u_j + w_j)] \\ & = & \delta \Delta u_n + \delta \Delta w_n + \sum_{i + j = n} [x_i \bullet \Delta (u_j + w_j)] \\ & = & \delta \Delta (u_n + w_n) + \sum_{i + j = n} [x_i \bullet \Delta (u_j + w_j)]. \end{eqnarray*} \end{proof} \begin{proposition} \label{prop:extension} Any element $y_0 \in \Ker \delta \cap \Ker \Delta$ can be extended to a formal power series $y(t) = y_0 + y_1 t + \cdots y_n t^n + \cdots$, such that $\delta_{x(t)} y(t) = 0$ and $y_n \in \Img \Delta$ for $n \geq 1$. Furthermore, if $\bar{y}(t) = \bar{y}_0 + \bar{y}_1 t + \cdots$ satisfies $\delta_{x(t)} \bar{y}(t) = 0$, where $\bar{y}_0 \in \Ker \delta \cap \Ker \Delta$ represents the same class as $y_0$ in $H({\cal A}, \delta)$, $\bar{y}_n \in \Img \Delta$ for $n > 0$, then there exists $z(t) \in (\Img \Delta)[[t]]$ such that $\bar{y}(t) - y(t) = \delta_{x(t)} z(t)$. \end{proposition} \begin{proof} When expanded into the formal power series in $t$, we can rewrite $\delta_{x(t)} y(t) = 0$ as a sequence of equations \begin{eqnarray*} && \delta y_0 = 0, \\ && \delta y_1 = - [x_1 \bullet y_0], \\ && \cdots\cdots \\ && \delta y_n = - \sum_{i + j = n} [x_i \bullet y_j], \\ && \cdots \cdots \end{eqnarray*} To find $y_1$, notice that \begin{eqnarray*} && [x_1 \bullet y_0] = \Delta (x_1 \wedge y_0) \in \Img \Delta, \\ && \delta [x_1 \bullet y_0] = [\delta x_1 \bullet y_0] + [x_1 \bullet \delta y_0] = 0, \end{eqnarray*} i.e. $[x_1 \bullet y_0] \in \Ker \delta \cap \Img \Delta = \Img \delta \Delta$, hence $y_1 \in \Img \Delta$ can be found. Suppose now that we have found $y_1, \cdots, y_n \in \Img \Delta$, we have \begin{eqnarray*} && \sum_{i+j=n+1} [x_i \bullet y_j] = \sum_{i+j=n+1}\Delta (x_i \wedge y_j) \in \Img \Delta, \\ && \delta \sum_{i+j=n+1} [x_i \bullet y_j] = \sum_{i+j=n+1} [\delta x_i \bullet y_j] - \sum_{i+j=n+1} [x_i \bullet \delta y_j] \\ & = & -\frac{1}{2}\sum_{i+j=n+1} \sum_{p+q = i} [[x_p \bullet x_q]\bullet y_j] + \sum_{i+j=n+1} \sum_{q+ r = j} [x_i \bullet [x_q \bullet y_r]] \\ & = & -\frac{1}{2} \sum_{p + q + r = n + 1} ([[x_p \bullet x_q]\bullet y_r] - 2 [x_p \bullet [x_q \bullet y_r]]) = 0. \end{eqnarray*} I.e., $\sum_{i+j=n+1} [x_i \bullet y_j] \in \Ker \delta \Img \Delta = \Img \delta \Delta$, hence we can find $y_{n+1}$. By Lemma \ref{lm:quasi}, $H({\cal A}, \delta) \cong (Ker \Delta \cap \Ker \delta) / \Img \delta \Delta$, since $\bar{y}_0$ and $y_0$ represents the same class in $H({\cal A}., \delta)$, we have $\bar{y}_0 - y_0 = \delta z_0$ for some $z_0 \in \Img \Delta$. Now as above, we solve $\delta_{x(t)} z(t) = \bar{y}(t) - y(t)$ by induction: first expand in power series to get a sequence of equations \begin{eqnarray*} && \delta z_0 = \bar{y}_0 - y_0, \\ && \delta z_1 = \bar{y}_1 - y_1 - [x_1 \bullet z_0], \\ && \cdots \cdots \\ && \delta z_n = \bar{y}_n - y_n - \sum_{i + j = n} [x_i \bullet z_j], \\ && \cdots \cdots \end{eqnarray*} then inductively check the right hand side of each equation lies in $\Ker \delta \cap \Img \Delta = \Img \delta \Delta$ as above, hence one can find a solution $z_n$ in $\Img \Delta$. \end{proof} We now define a homomorphism $\phi_{x(t)}: H({\cal A}, \delta)[[t]] \to H({\cal A}[[t]], \delta_{x(t)})$ as follows: for any $[y] \in H({\cal A}, \delta)$, represent it by an element $y_0 \in \Ker \delta \cap \Ker \Delta$. Extend $y_0$ to $y(t)$ as in Proposition \ref{prop:extension}, then set $\phi_{x(t)}([y]) = [y(t)]$. This is well-defined by Proposition \ref{prop:extension}. Extend $\phi_{x(t)}$ to $H({\cal A}, \delta)[[t]]$ and still denote it by $\phi_{x(t)}$. We have the following \begin{proposition} \label{prop:isomorphism} The homomorphism $\phi_{x(t)}: H({\cal A}, \delta)[[t]] \to H({\cal A}[[t]], \delta_{x(t)})$ is an isomorphism of ${\bf k}[[t]]$-modules. \end{proposition} \begin{proof} We break the proof into two steps. \noindent {\em Step 1. $\phi_{x(t)}$ is injective.} By Lemma \ref{lm:quasi}, we can represent any class of $H({\cal A}, \delta)[[t]]$ by an element $y(t) = y^{(k)} t^k + y^{(k+1)} t^{k+1} + \cdots \in (\Ker \delta \cap \Ker \Delta)[[t]]$. Without loss of generality, we assume that $y^{(k)} \notin \Img \Delta \delta$ and hence $[y^{(k)}] \neq 0 \in H(\Ker \Delta, \delta)$. In fact, if $y^{(k)} = \Delta \delta u_k$ for some $u_k \in {\cal A}$, we replace $y(t)$ by $y(t) - t^k \delta\Delta u_k$ and consider the $(k+1)$-th term. Now if $\phi_{x(t)}(y(t)) = \delta_{x(t)} z(t)$ for some $z(t) = z_0 + z_1 t + \cdots \in (\Ker \Delta)[[t]]$, then by noticing that $\phi_{x(t)}(y(t))$ has leading term $y^{(k)}t^k$, we get a sequence of equations \begin{eqnarray*} && \delta z_0 = 0, \\ && \delta z_1 + [x_1 \bullet z_0 ] = 0, \\ && \cdots \cdots \\ && \delta z_{k -1} + \sum_{i + j = k - 1} [x_i \bullet z_j] = 0, \\ && \delta z_k + \sum_{i + j = k} [x_i \bullet z_j] = y^{(k)}. \end{eqnarray*} Now $\sum_{i + j = k} [x_i \bullet z_j] = \sum_{i + j = k} \Delta (x_i \wedge z_j) \in \Img \Delta$, and \begin{eqnarray*} && \delta \sum_{i + j = k} [x_i \bullet z_j] = \sum_{i + j = k} [\delta x_i \bullet z_j] - \sum_{i + j = k} [x_i \bullet \delta z_j] \\ & = & \sum_{i + j = k} [\delta x_i \bullet z_j] + \sum_{i + j = k} [x_i \bullet \sum_{q + l = j}[x_q \bullet z_l]] \\ & = & \sum_{i + j = k} [\delta x_i \bullet z_j] + \sum_{i + j = k} \frac{1}{2} \sum_{p+q = i} [[x_p \bullet x_q] \bullet z_j] = 0. \end{eqnarray*} Therefore $\sum_{i + j = k} [x_i \bullet z_j] \in \Img \delta \Delta$, and so $y^{(k)} \in \delta \Ker \Delta$, a contradiction to the assumption that $[y^{(k)}] \neq 0 \in H(\Ker \Delta, \delta)$. \noindent {\em Step 2. $\phi_{x(t)}$ is surjective.} By Lemma \ref{prop:quasi2}, any element of $H({\cal A}[[t]], \delta_{x(t)})$ can be represented by an element $y(t) = y^{(0)} + y^{(1)} t + \cdots\in \Ker \Delta \cap \Ker \delta_{x(t)}$. Then we have $y^{(0)} \in \Ker \Delta \cap \Ker \delta$, hence it can be extended to an element $\phi_{x(t)} (y^{(0)}) = y^{(0)} + y^{(0)}_1 t + \cdots \in \Ker \Delta \cap \Ker \delta_{x(t)}$. Consider now $y(t) - \phi_{x(t)} (y^{(0)})$, it can be written as $t y'(t)$, where $\tilde{y}(t) \in \Ker \Delta \cap \Ker \delta_{x(t)}$. Hence by induction, $y(t)$ lies in the image of $\phi_{x(t)}$. \end{proof} \begin{theorem} Assume that $({\cal A}, \wedge, \delta, \Delta, [\cdot \bullet \cdot])$ is a DGBV algebra in ${\cal D}{\cal G}{\cal B}{\cal V}_q$ and $x(t)$ as in Proposition \ref{prop:MC}. Then there is a naturally defined formal deformation $$\wedge_{x(t)}: H({\cal A}, \delta)[[t]] \otimes H({\cal A}, \delta)[[t]] \to H({\cal A}, \delta)[[t]]$$ of $\wedge$ defined by $a \wedge_{x(t)} b = \phi^{-1}_{x(t)}(\phi_{x(t)}(a) \wedge \phi_{x(t)}(b))$. \end{theorem} \section{Nice integrals and formal Frobenius manifold structures} In the above discussion, we have only considered one-parameter deformations such that the parameter $t$ commutes with the elements of the relevant algebras. In this section, we need to make some generalizations. First we will consider odd deformations. Let $\epsilon$ be an indeterminate, denote by ${\bf k}[[\epsilon]]$ the exterior algebra $\Lambda^*({\bf k} \oplus {\bf k} \epsilon)$. We regard ${\bf k}[[\epsilon]]$ as the super formal power series with an indeterminate of odd degree. If $({\cal A}, \wedge)$ is a ${\Bbb Z}_2$-graded associative algebra over ${\bf k}$, let ${\cal A}[[\epsilon]] = {\cal A} \otimes_{{\bf k}} {\bf k}[[\epsilon]]$. Any element of ${\cal A}[[t]]$ can be written as $a_0 + a_1 \epsilon$, and we can extend $\wedge$ as follows: $$(a_0 + a_1 \epsilon) \wedge (b_0 + b_1 \epsilon) = a_0 \wedge b_0 + (a_0 \wedge b_1 + a_1 \sigma(b_0) \epsilon,$$ where $\sigma(a) = (-1)^{|a|} a$ for a homogeneous element $a \in {\cal A}$. Here we have used the Koszul sign convention. It is straightforward to verify that this multiplication is associative. Furthermore, if $1$ is the unit for $({\cal A}, \wedge)$, then so is it for $({\cal A}[[\epsilon]], \wedge)$; and if $({\cal A}, \wedge)$ is graded commutative, then so is $({\cal A}[[\epsilon]], \wedge)$. Now if $({\cal A}, \wedge, \delta, \Delta, [\cdot \bullet\cdot])$ is a DGBV algebra, we extend $\Delta$ and $[\cdot \bullet \cdot]$ to ${\cal A}[[\epsilon]]$ by the Koszul sign convention. We have $\epsilon a = \sigma(a) \epsilon$, $\epsilon \delta = - \delta \epsilon$, $\epsilon \Delta = -\Delta \epsilon$ and \begin{eqnarray*} [a \epsilon \bullet b] = -[a \bullet \sigma(b)] \epsilon, & [a \bullet b \epsilon] = [a \bullet b] \epsilon. \end{eqnarray*} Given a cohomology class of $H({\cal A}, \delta)$ of odd degree, represent it by an element in $\Ker \delta \cap \Ker \Delta$. Then $x(\epsilon) = x_1 \epsilon$ satisfies the Maurer-Cartan equation $$\delta \omega + \frac{1}{2}[\omega \bullet \omega] = 0$$ over ${\cal A}[[\epsilon]]$. This is the analogue of Proposition \ref{prop:MC}. The simpler versions of the proofs (without inductions) of Propositions \ref{prop:quasi2} - \ref{prop:isomorphism} prove the corresponding statements for odd cohomology classes of $H({\cal A}, \delta)$. As a result, we obtain an odd deformation of $(H({\cal A}, \delta), \wedge)$. Secondly, we will also be interested in multi-parameter formal deformations which now turn to. From now on, assume that $H$ is a rank $n$ free ${\bf k}$-module. Let $\{e_{\alpha}: \alpha = 0, \cdots, n-1\}$ be a set of free homogeneous generators of $H$, such that $e_0 = 1$. Let $\{t^{\alpha} \}$ be the dual set of generators of $H^t = \Hom_{{\bf k}}(H, {\bf k})$. Denote by ${\bf k}[[{\bf t}]]$ the space of super power series in $\{t^0, \cdots, t^{n-1}\}$, and consider ${\cal A}[{\bf t}]] = {\cal A} \otimes_{{\bf k}} {\bf k}[[{\bf t}]]$. Again, modifying the proofs, the analogues of Propositions \ref{prop:quasi2} - \ref{prop:isomorphism} can be proved for ${\cal A}[[{\bf t}]]$. We give the corresponding statements below, and omit the proofs. Also, see Manin \cite{Man2} for different proofs. \begin{proposition} \label{prop:MC'} There is a formal power series solution $\Gamma = \Gamma_1 + \cdots + \Gamma_n + \cdots$ to (\ref{eqn:MC}), such that $\Gamma_1 = x_{\alpha} t^{\alpha}$, where $x_{\alpha} \in \Ker \Delta \cap \Ker \delta$ represents $e_{\alpha}$, and for $n > 1$, $\Gamma_n$ is a super homogeneous polynomial of order $n$ in $\{t^{\alpha}\}$ with coefficients in $\Img \Delta$ . Furthermore, $t^0$ only appears in $\Gamma_1$. (Such a solutions is called a {\em universal normalized solution}.) \end{proposition} \begin{proposition} \label{prop:quasi2'} Let $\Gamma$ be a universal normalized solution. Then the deformed triple $({\cal A}[[{\bf t}]], \delta_{\Gamma}, \Delta)$ is in ${\cal D}{\cal G}{\cal B}{\cal V}_q$. \end{proposition} \begin{proposition} \label{prop:extension'} Let $\Gamma$ be a universal normalized solution. Then any element $y_0 \in \Ker \delta \cap \Ker \Delta$ can be extended to $y({\bf t}) \in {\cal A}[[{\bf t}]]$ with leading term $y_0$, such that $\delta_{\Gamma} y(t) = 0$ and all the higher order terms lies in $\Img \Delta$. Furthermore, if $\bar{y}({\bf t}) \in {\cal A}[[{\bf t}]]$ has the same properties and its leading term $\bar{y}_0 \in \Ker \delta \cap \Ker \Delta$ represents the same class as $y_0$ in $H({\cal A}, \delta)$, then there exists $z({\bf t}) \in (\Img \Delta)[[{\bf t}]]$ such that $\bar{y}(t) - y(t) = \delta_{\Gamma} z(t)$. \end{proposition} Similar to the definition of $\phi_{x(t)}$, we define a homomorphism $\phi_{\Gamma}: H({\cal A}, \delta)[[{\bf t}]] \to H({\cal A}[[t]], \delta_{\Gamma})$. \begin{proposition} \label{prop:isomorphism'} The homomorphism $\phi_{\Gamma}: H({\cal A}, \delta)[[{\bf t}]] \to H({\cal A}[[{\bf t}]], \delta_{\Gamma})$ is an isomorphism of ${\bf k}[[{\bf t}]]$-modules. \end{proposition} We can now define an $n$-parameter super formal deformation $$\wedge_{\Gamma}: H({\cal A}, \delta)[[{\bf t}]] \otimes H({\cal A}, \delta)[[{\bf t}]] \to H({\cal A}, \delta)[[{\bf t}]]$$ by $a \wedge_{\Gamma} b = \phi^{-1}_{\Gamma}(\phi_{\Gamma}(a) \wedge \phi_{\Gamma}(b))$. Recall the following \begin{definition} An {\em integral} of a DGBV algebra $({\cal A}, \wedge, \delta, \Delta, [\cdot \bullet \cdot])$ is a ${\bf k}$-linear $\int: {\cal A} \to {\bf k}$, such that \begin{eqnarray*} && \int \delta a \wedge b = (-1)^{|a| + 1} \int a \wedge \delta b, \\ && \int \Delta a \wedge b = (-1)^{|a|} \int a \wedge \Delta b, \end{eqnarray*} for any homogeneous $a, b \in {\cal A}$. Obviously, an integral on ${\cal A}$ induces a well-defined supersymmetric bilinear form, $g: H \otimes H \to {\bf k}$ by $$g([a], [b]) = \int a \wedge b.$$ An integral is called {\em nice} if $g$ induces an isomorphism $H \to H^t$ by $[a] \in H \mapsto g([a], \cdot) \in H^t$. \end{definition} In the context of \S \ref{sec:deform}, we extend $\int$ to ${\cal A}[[t]]$. Then it is straightforward to see that $\int$ is an integral of $({\cal A}[[t]], \wedge, \delta_{x(t)}, \Delta, [\cdot\bullet\cdot])$ (Manin \cite{Man2}, Proposition 5.5.1). If for $j = 1, 2$, $y_j \in \Ker \delta \cap \Ker \Delta$ such that $\delta_{x(t)} (y_j + \Delta z_j(t)) = 0$ for some $z_j(t) \in {\cal A}[[t]]$, then we have \begin{eqnarray*} && \int (y_1 + \Delta z_1(t)) \wedge (y_2 + \Delta z_2(t)) \\ & = & \int y_1 \wedge y_2 + \int y_1 \wedge \Delta z_2(t) + \int \Delta z_1(t) \wedge (y_2 + \Delta z_2(t)) \\ & = & \int y_1 \wedge y_2 + \int \Delta \sigma(y_1) \wedge z_2(t) + \int \sigma(z_2(t)) \wedge \Delta (y_2 + \Delta z_2(t)) \\ & = & \int y_1 \wedge y_2. \end{eqnarray*} So if the integral $\int$ is nice for ${\cal A}$, so is it for ${\cal A}[[t]]$. Given a nice integral $\int$ on a DGBV algebra ${\cal A}$, since we have $$g([a] \wedge [b], [c]) = \int ([a] \wedge [b]) \wedge [c] = \int [a] \wedge ([b] \wedge [c]) = g([a], [b] \wedge [c]),$$ for $[a], [b], [c] \in H$, $(H, \wedge, g)$ is a {\em (graded) Frobenius algebra}. Therefore, given a nice integral of ${\cal A}$, $(H[[t]], \wedge_{x(t)}, g)$ is a formal deformation of the Frobenius algebra $(H, \wedge, g)$. Similarly, given a universal normalized solution $\Gamma$, the above discussion can be carried out for $({\cal A}[[{\bf t}]], \delta_{\Gamma}, \wedge, \delta, [\cdot\bullet\cdot])$, a nice integral on ${\cal A}$ then gives a multi-parameter formal deformation $(H[[t]], \wedge_{\Gamma}, g)$ of $(H, \wedge, g)$. We now state the following result in Barannikov-Kontsevich \cite{Bar-Kon} and Manin \cite{Man2}: \begin{proposition} \label{prop:Frobenius} Let $({\cal A}, \wedge, \delta, \Delta, [\cdot\bullet\cdot])$ be a DGBV algebra over ${\bf k}$ in ${\cal D}{\cal G}{\cal B}{\cal V}_q$. Assume that ${\bf k}$-module $H = H({\cal A}, \delta)$ has free homogeneous generators $\{e_{\alpha}: \alpha = 0, \cdots, n -1\}$ with $e_0 = 1$, and $\{ t^{\alpha}\}$ the dual generators of the dual module $H^t$. Also assume that there is a nice integral $\int$ on ${\cal A}$. Then given a universal normalized solution $\Gamma = \Gamma_1 + \Delta B$ with $\Gamma_1 = e_{\alpha} t^{\alpha}$, the super formal power series \begin{eqnarray} \label{eqn:potential} \Phi = \int (\frac{1}{6} \Gamma^3 - \frac{1}{2}\delta B \wedge \Delta B) = \int (\frac{1}{6} \Gamma^3 - \frac{1}{4} (\Gamma - \Gamma_1) \wedge \Gamma^2) \end{eqnarray} satisfies \begin{eqnarray*} \frac{\partial^3 \Phi}{\partial t^0 \partial t^{\alpha} t^{\beta}} = g_{\alpha\beta}, \end{eqnarray*} and the WDVV (Witten-Dijkgraaf-Verlinde-Verlinde) equations \begin{eqnarray*} \frac{\partial^3 \Phi}{\partial t^{\alpha} \partial t^{\beta} \partial t^{\mu}} g^{\mu \nu} \frac{\partial^3 \Phi}{\partial t^{\nu} \partial t^{\gamma} \partial t^{\delta}} = (-1)^{|e_{\alpha}|(|e_{\beta}| + |e_{\gamma}|)} \frac{\partial^3 \Phi}{ \partial t^{\beta} \partial t^{\gamma} \partial t^{\mu}} g^{\mu \nu} \frac{\partial^3 \Phi}{\partial t^{\nu} \partial t^{\alpha} \partial t^{\delta}}, \end{eqnarray*} where $g_{\alpha\beta} = g(e_{\alpha}, e_{\beta}) = \int e_{\alpha} \wedge e_{\beta}$ and $(g^{\alpha\beta})$ is the inverse matrix of $(g_{\alpha\beta})$. \end{proposition} \section{Gauge invariance and functorial property} In this section, we will discuss the gauge invariance of the constructions in previous sections. As mentioned in the introduction, special cases have been treated in the Kodaira-Spencer theory gravity \cite{Ber-Cec-Oog-Vaf} and the K\"{a}hler gravity \cite{Ber-Sad}. Barannikov and Kontsevich \cite{Bar-Kon} remarked that all universal normalized solutions are gauge equivalent but offered no proof. Let $(L = \oplus_{n \in {\Bbb Z}} L_n, [\cdot, \cdot], \delta)$ be a differential graded Lie algebra (DGLA). Denote by $L^e$ and $L^o$ the subspaces of elements with even and odd degrees respectively. For any element $A \in L$, denote by $\ad_A$ the automorphism $[A, \cdot]: L \to L$. Let $R = {\bf k}[t]$ or ${\bf k}[[{\bf t}]]$, and ${\frak m}$ the maximal ideal of $R$. Consider the group \begin{eqnarray*} {\cal G} = G(R) = \exp (L \otimes {\frak m})^e, \end{eqnarray*} with the multiplication $e^A e^B = e^C$ defined by the Campbell-Baker-Hausdorff formula: $C = \sum_{n \geq 1} C_n$, where $C_n = \frac{1}{n} \sum_{p + q = n} (C_{p, q}' + C_{p, q}'')$, and \begin{eqnarray*} && C_{p, q}' = \sum_{\substack{p_1 + \cdots p_m = p \\ q_1 + \cdots + q_{m-1} = q - 1 \\ p_i + q_i \geq 1 \\ p_m \geq 1}} \frac{(-1)^{m+1}}{m} \frac{\ad_A^{p_1}\ad_B^{q_1} \cdots \ad_A^{p_m}B} {(p_1)!(q_1)!\cdots (p_m)!} \\ && C_{p, q}'' = \sum_{\substack{p_1 + \cdots p_{m-1} = p -1 \\ q_1 + \cdots + q_m = q \\ p_i + q_i \geq 1}} \frac{(-1)^{m+1}}{m} \frac{\ad_A^{p_1}\ad_B^{q_1} \cdots \ad_B^{q_{m-1}}A} {(p_1)!(q_1)!\cdots (p_m)!} \end{eqnarray*} These are the explicit formulas in Dynkin's form (Serre \cite{Ser}, p. 29). (This is similar to a construction in Goldman and Millson \cite{Gol-Mil1, Gol-Mil2} where they use Artin local ${\bf k}$-algebras.) There is a natural action of ${\cal G}$ on $L \otimes R$ by \begin{eqnarray} \label{eqn:action} e^A \cdot \alpha = e^{\ad_A} \alpha. \end{eqnarray} It is clear that $e^{\ad_A} e^{-\ad_A} = 1$. \begin{lemma} \label{lm:commutator} Let $A \in (L \otimes {\frak m})^e$, $B \in L \otimes R$, then for $n \geq 1$, we have \begin{eqnarray*} && \ad_B \ad_A^n = \sum_{j = 0}^n \left( \begin{array}{c} n \\ j \end{array} \right) \ad_A^j \ad_{(-\ad_A)^{n-j}B}. \end{eqnarray*} Furthermore, we have \begin{eqnarray*} e^{\ad_A}\ad_B e^{-\ad_A} = \ad_{e^{\ad_A} B}. \end{eqnarray*} \end{lemma} \begin{proof} The first equality can be proved elementarily by induction. For the second equality, we have \begin{eqnarray*} && \ad_B e^{\ad_A} = \sum_{n \geq 0} \frac{1}{n!} \ad_B \ad_A^n = \sum_{n \geq 0} \frac{1}{n!} \sum_{j = 0}^n \left( \begin{array}{c} n \\ j \end{array} \right) \ad_A^j \ad_{(-\ad_A)^{n-j}B} \\ & = & \sum_{n \geq 0} \sum_{j = 0}^n \frac{1}{j!} \ad_A^j \frac{1}{(n-j)!} \ad_{(-\ad_A)^{n-j}B} = \sum_{j, k \geq 0} \frac{1}{j!} \ad_A^j \frac{1}{k!} \ad_{(-\ad_A)^k B} \\ & = & e^{\ad_A} \ad_{e^{-\ad_A}B}. \end{eqnarray*} \end{proof} \begin{lemma} \label{lm:commutator2} For $A \in (L \otimes {\frak m})^e$, we have $$e^{\ad_A} d e^{- \ad_A} = d + \sum_{q \geq 0} \frac{1}{(q+1)!} \ad_{\ad_A^q \delta A} =d + \ad_{\frac{1 - e^{\ad_A}}{\ad_A} \delta A}.$$ \end{lemma} \begin{proof} It is easy to see that \begin{eqnarray*} d \ad_A = \ad_A d + \ad_{\delta A}. \end{eqnarray*} Then by induction, it is easy to show that \begin{eqnarray*} && d \ad_A^n = \ad_A^n d + \sum_{p=0}^{n-1} \ad_A^{n - p - 1} \ad_{\delta A} \ad_A^p \end{eqnarray*} Now $\delta A \in L^o$, so we can use Lemma \ref{lm:commutator} to handle $\ad_{\delta A} \ad_A^p$ as follows: \begin{eqnarray*} && d \ad_A^n = \ad_A^n d + \sum_{p=0}^{n-1} \ad_A^{n - p - 1} \ad_{\delta A} \ad_A^p \\ & = & \ad_A^n d + \ad_A^{n - 1} \ad_{\delta A} + \sum_{p=1}^{n-1} \ad_A^{n - p - 1} \ad_{\delta A} \ad_A^p \\ & = & \ad_A^n d + \ad_A^{n - 1} \ad_{\delta A} + \sum_{p=1}^{n-1} \ad_A^{n - p - 1} \sum_{j = 0}^p \left( \begin{array}{c} p \\ j \end{array} \right) \ad_A^j \ad_{(-\ad_A)^{p-j}\delta A} \\ & = & \ad_A^n d + \ad_A^{n - 1} \ad_{\delta A} + \sum_{p=1}^{n-1} \sum_{j = 0}^p \left( \begin{array}{c} p \\ j \end{array} \right) \ad_A^{n - p - 1 + j} \ad_{(-\ad_A)^{p-j}\delta A} \\ & = & \ad_A^n d + n \ad_A^{n - 1} \ad_{\delta A} + \sum_{p=1}^{n-1} \sum_{j = 0}^{p-1} \left( \begin{array}{c} p \\ j \end{array} \right) \ad_A^{n - p - 1 + j} \ad_{(-\ad_A)^{p-j}\delta A} \\ & = & \ad_A^n d + n \ad_A^{n - 1} \ad_{\delta A} + \sum_{q=1}^{n-1} \sum_{j = 0}^{n-1-q} \left( \begin{array}{c} q+j \\ j \end{array} \right) \ad_A^{n - 1 - q} \ad_{(-\ad_A)^q \delta A} \\ & = & \ad_A^n d + n \ad_A^{n - 1} \ad_{\delta A} + \sum_{q=1}^{n-1} \left( \begin{array}{c} n \\ n-1-q \end{array} \right) \ad_A^{n - 1 - q} \ad_{(-\ad_A)^q \delta A} \\ & = & \ad_A^n d + \sum_{q=0}^{n-1} \left( \begin{array}{c} n \\ n-1-q \end{array} \right) \ad_A^{n - 1 - q} \ad_{(-\ad_A)^q \delta A} \end{eqnarray*} Hence we have \begin{eqnarray*} && d e^{\ad_A} = \sum_{n \geq 0} \frac{1}{n!} d \ad_A^n \\ & = & \sum_{n \geq 0} \frac{1}{n!} \ad_A^n d + \sum_{n \geq 0} \frac{1}{n!} \sum_{q=0}^{n-1} \left( \begin{array}{c} n \\ n-1-q \end{array} \right) \ad_A^{n - 1 - q} \ad_{(-\ad_A)^q \delta A} \\ & = & e^{\ad_A} d + \sum_{n \geq 0} \sum_{q=0}^{n-1} \frac{1}{(n-1-q)!}\ad_A^{n - 1 - q} \frac{1}{(q+1)!} \ad_{(-\ad_A)^q \delta A} \\ & = & e^{\ad_A} (d + \sum_{q \geq 0} \frac{1}{(q+1)!} \ad_{(-\ad_A)^q \delta A}). \end{eqnarray*} Replacing $A$ by $-A$ then completes the proof. \end{proof} Given an element $\omega \in (L \otimes {\frak m})^o$, set $\delta_\omega = \delta + \ad_{\omega}$. From Lemma \ref{lm:commutator} and Lemma \ref{lm:commutator2}, we see that for any $A \in (L\otimes {\frak m})^e$, we have \begin{eqnarray*} e^A \delta_{\omega} e^{-A} = \delta_{e^A \cdot \omega}, \end{eqnarray*} where \begin{eqnarray*} e^A \cdot \omega = e^{\ad_A} \omega + \frac{1 - e^{\ad_A}}{\ad_A} \delta A \end{eqnarray*} is called the gauge transformation of $\omega$. This formula is well-known. See e.g. Goldman-Millson \cite{Gol-Mil1, Gol-Mil2}. Let $X = \{ \omega \in (L \otimes {\frak m})^o: \delta\omega + \frac{1}{2}[\omega, \omega] =0 \}$, then the ${\cal G}$-action preserves $X$. Given any DGBV algebra $({\cal A}, \wedge, \delta, \Delta, [\cdot \bullet \cdot])$, denote by ${\cal A}[-1]$ the graded vector space with $({\cal A}[-1])_p = {\cal A}_{p+1}$. Then $({\cal A}[-1], \delta, [\cdot \bullet \cdot])$ is a DGLA. When $A = {\bf k}[[t]]$, we get the group ${\cal G}[[t]] = \{\exp y(t): y(t) \in t{\cal A}^o[[t]] \}$ and its natural action on ${\cal A}[[t]]$. \begin{proposition} \label{prop:gauge1} Assume that $x(t) = x_1 t + \cdots + x_n t^n + \cdots$ and $\bar{x}(t) = \bar{x}_1 t + \cdots + \bar{x}_n t^n + \cdots$ are two solutions of the Maurer-Cartan equation $$\delta \omega + \frac{1}{2} [\omega \bullet \omega] = 0,$$ such that $x_1, \bar{x}_1 \in \Ker \delta \cap \Ker \Delta$, and $x_n, \bar{x}_n \in \Img \Delta$. If $[x_1] = [\bar{x}_1]$, then there exists $w(t) = w_1 t + \cdots + w_n t^n + \cdots \in {\cal A}[[t]]$ such that $\exp (\Delta w(t)) \cdot x(t) = \bar{x}(t)$. \end{proposition} \begin{proof} It suffices to prove the following: if $x(t) = \bar{x}(t) \pmod {t^n}$, $n \geq 1$, then there exist $z_n \in {\cal A}$, such that $\exp (t^n \Delta z_n) \cdot x(t) = \bar{x}(t) \pmod {t^{n+1}}$. For $n = 1$, we clear have $x(t) = \bar{x}(t) \pmod t$. Since $[\bar{x}_1] = [x_1]$, there exists $z_1$ such that $\bar{x}_1 = x_1 - \delta \Delta z_1$. Now modulo $t^2$, we have \begin{eqnarray*} \exp (t \Delta z_1) \cdot x(t) & = & x(t) + \sum_{j > 0} \frac{1}{j!} \ad_{t \Delta z_1}^j x(t) - \delta \Delta z_1 - \sum_{j > 1} \frac{1}{j!} \ad_{t \Delta z_1}^{j-1} x(t) \\ & = & x_1 t - \delta \Delta z_1 = \bar{x}_1 t =\bar{x}(t). \end{eqnarray*} For $n > 1$, we have $\bar{x}_n - x_n \in \Img \Delta$, and \begin{eqnarray*} && \delta (\bar{x}_n - x_n) = \delta \bar{x}_n - \delta x_n \\ & = & \sum_{i + j = n} [\bar{x}_i \bullet \bar{x}_j] - \sum_{i+j =n} [x_i \bullet x_j] \\ & = & \sum_{i+j =n} [x_i \bullet x_j] - \sum_{i+j =n} [x_i \bullet x_j] = 0, \end{eqnarray*} hence there exists $z_n$ such that $\bar{x}_n - x_n = - \delta \Delta z_n$. Now modulo $t^{n+1}$, we have \begin{eqnarray*} && \exp (t^n \Delta z_n) \cdot x(t) \\ & = & x(t) + \sum_{j > 0} \frac{1}{j!} \ad_{t^n \Delta z_n}^j x(t) - t^n \delta \Delta z_n - \sum_{j > 1} \frac{1}{j!} \ad_{t^n \Delta z_n}^{j-1} x(t) \\ & = & x_1 t + \cdots + x_{n-1}t^{n-1} + (x_n - \delta \Delta z_n) t^n = \bar{x}_1 t + \cdots \bar{x}_n t^n = \bar{x}(t). \end{eqnarray*} \end{proof} As an easy corollary, we have \begin{proposition} The equivalence class of $\wedge_{x(t)}$ depends only on the class $[x]$. \end{proposition} Consequently, we have the following \begin{theorem} If $f: ({\cal A}_1, \wedge_1, \delta_1, \Delta_1, [\cdot\bullet\cdot]_{\Delta_1}) \to ({\cal A}_2, \wedge_2, \delta_2, \Delta_2, [\cdot\bullet\cdot]_{\Delta_2})$ is a homomorphism of two DGBV algebras in ${\cal D}{\cal G}{\cal B}{\cal V}_q$, then we have $$f_*(\alpha \wedge_x \beta) = f_*(\alpha) \wedge_{f_*(x)} f_*(\beta),$$ where $\alpha, \beta, x \in H({\cal A}_1, \delta_1)$, and $f_*: H({\cal A}_1, \delta_1) \to H({\cal A}_2, \delta_2)$ is the homomorphism on cohomology induced by $f$. \end{theorem} \begin{proof} Given any $x \in H({\cal A}_1, \delta_1)$, represent it by an element $x_1 \in \Ker \delta_1 \cap \Ker \Delta_1$ and extend it to a power series $x(t) = x_1 t+ \cdots + x_n t^n + \cdots$, such that $x_n \in \Img \Delta_1$ and $\delta x(t) + \frac{1}{2}[x(t)\bullet x(t)]_{\Delta_1} = 0$. Then $f(x(t)) = f(x_1) t + \cdots + f(x_n) t^n + \cdots$ satisfies $f(x_1) \in \Ker \delta_2 \cap \Ker \Delta_2$, $f(x_n) \in \Img \Delta_2$ and $\delta f(x(t)) + \frac{1}{2}[f(x(t))\bullet f(x(t))]_{\Delta_2} = 0$. \end{proof} \begin{proposition} \label{prop:gauge2} Assume $({\cal A}, \wedge, \delta, \Delta, [\cdot\bullet\cdot])$ is a DGBV algebra which satisfies the conditions in Proposition \ref{prop:Frobenius}. Then given any two universal normalized solutions $\Gamma$ and $\overline{\Gamma}$, there exists an odd element $A \in (\Img \Delta)[[{\bf t}]]$ whose zero order term vanishes, such that $e^A \cdot \Gamma = \overline{\Gamma}$. Furthermore, the potential function $\Phi$ is gauge invariant. \end{proposition} \begin{proof} The proof of the first statement is an easy modification of the proof of Proposition \ref{prop:gauge1}. To prove the second statement, we first linearize the gauge transformation: \begin{eqnarray*} \left.\frac{d}{d \lambda}\right|_{\lambda =0}e^{\lambda A} \cdot \Gamma = [A \bullet \Gamma] - \delta A. \end{eqnarray*} Then we have \begin{eqnarray*} && \left.\frac{d}{d \lambda}\right|_{\lambda =0} \Phi(e^{\lambda A} \cdot \Gamma) \\ & = & \left.\frac{d}{d \lambda}\right|_{\lambda =0} \int \left(\frac{1}{6} (e^{\lambda A} \cdot \Gamma)^3 - \frac{1}{4} (e^{\lambda A} \cdot \Gamma - (e^{\lambda A} \cdot \Gamma)_1) \wedge (e^{\lambda A} \cdot \Gamma)^2 \right)\\ & = & \int ( \frac{1}{2} \Gamma^2 \wedge ([A \bullet \Gamma] - \delta A) - \frac{1}{4}([A \bullet \Gamma] - \delta A - ([A \bullet \Gamma] - \delta A)_1) \wedge \Gamma^2 \\ && - \frac{1}{2}(\Gamma - \Gamma_1) \wedge \Gamma \wedge ([A \bullet \Gamma] - \delta A)) \\ & = & \int ( -\frac{1}{4} \Gamma^2 \wedge ([A \bullet \Gamma] - \delta A) + \frac{1}{4} \Gamma^2 \wedge ([A \bullet \Gamma] - \delta A)_1 + \frac{1}{2} \Gamma_1 \wedge \Gamma \wedge ([A \bullet \Gamma] - \delta A)), \end{eqnarray*} where the subscript means the the first order term. Since $A \in (\Img \Delta)[[{\bf t}]]$ is odd, we have \begin{eqnarray*} && \int \Gamma^2 \wedge [A \bullet \Gamma] = \frac{1}{3} \int [A \bullet \Gamma^3] = - \frac{1}{3} \int (\Delta (A \wedge \Gamma^3) - \Delta A \wedge \Gamma^3 + A \wedge \Delta \Gamma^3) \\ & = & \frac{2}{3} \int \Delta A \wedge \Gamma^3 = 0. \end{eqnarray*} Notice that \begin{eqnarray*} \int \Gamma^2 \wedge \delta A = - \int \delta \Gamma^2 \wedge A = - 2 \int \Gamma \wedge \delta \Gamma \wedge A = \int \Gamma \wedge [\Gamma \bullet \Gamma] \wedge A \end{eqnarray*} Now \begin{eqnarray*} && \int \Gamma \wedge [\Gamma \bullet \Gamma] \wedge A = \frac{1}{2} \int [\Gamma \bullet \Gamma^2] \wedge A \\ & = & \frac{1}{2} \int (\Delta \Gamma^3 - \Delta \Gamma \wedge \Gamma^2 - \Gamma \wedge \Delta \Gamma^2) \wedge A \\ & = & \frac{1}{2} \int (\Gamma^3 \wedge \Delta A - \Gamma \wedge [\Gamma \bullet \Gamma] \wedge A) = - \frac{1}{2} \int \Gamma \wedge [\Gamma \bullet \Gamma] \wedge A, \end{eqnarray*} hence $\int \Gamma \wedge [\Gamma \bullet \Gamma] \wedge A = 0$, and so $$\int \Gamma^2 \wedge \delta A = 0.$$ Replacing $A$ by $A_1$, we get $$\int \Gamma^2 \wedge \delta A_1 = 0.$$ Similarly, we have \begin{eqnarray*} && \int \Gamma_1 \wedge \Gamma \wedge [A \bullet \Gamma] = \frac{1}{2} \int \Gamma_1 \wedge [A \bullet \Gamma^2] \\ & = & - \frac{1}{2} \int \Gamma_1 \wedge (\Delta (A \wedge \Gamma^2) - \Delta A \wedge \Gamma^2 + A \wedge \Delta \Gamma^2) \\ & = & - \frac{1}{2} \int \Delta \Gamma_1 \wedge (A \wedge \Gamma^2) - \frac{1}{2} \int \Gamma_1 \wedge A \wedge [\Gamma \bullet \Gamma] \\ & = & \int \Gamma_1 \wedge A \wedge \delta \Gamma = \int \delta (\Gamma_1 \wedge A) \wedge \Gamma \\ & = & \int (\delta \Gamma_1 \wedge A \wedge \Gamma + \Gamma_1 \wedge \delta A \wedge \Gamma) = \int \Gamma_1 \wedge \Gamma \wedge \delta A, \end{eqnarray*} and so $$\int \Gamma_1 \wedge \Gamma \wedge ([A \bullet \Gamma] - \delta \Gamma) = 0.$$ Therefore, we have \begin{eqnarray*} \left.\frac{d}{d \lambda}\right|_{\lambda =0} \Phi(e^{\lambda A} \cdot \Gamma) = \int \Gamma^2 \wedge [A \bullet \Gamma]_1 = 0. \end{eqnarray*} Here we use the observation that since both $A$ and $\Gamma$ has no zeroth order term, $[A \bullet \Gamma]$ has no first order term. \end{proof} \begin{definition} Denote by ${\cal D}{\cal G}{\cal B}{\cal V}_{qi}$ the subcategory of ${\cal D}{\cal G}{\cal B}{\cal V}_q$ consists of DGBV elements in ${\cal D}{\cal G}{\cal B}{\cal V}$ which admits nice integrals. A {\em quasi-isomorphism} between two elements in ${\cal D}{\cal G}{\cal B}{\cal V}_q$ (${\cal D}{\cal G}{\cal B}{\cal V}_{qi}$) is a morphism in ${\cal D}{\cal G}{\cal B}{\cal V}_q$ (${\cal D}{\cal G}{\cal B}{\cal V}_{qi}$) which induces isomorphism on cohomology groups. \end{definition} \begin{theorem} Formal Frobenius manifolds obtained from quasi-isomorphic DGBV algebras in ${\cal D}{\cal G}{\cal B}{\cal V}_{qi}$ by Proposition 2.5 can be identified with each other. \end{theorem} \begin{proof} Just observe that a universal normalized solution is mapped to a universal normalized solution under quasi-isomorphism. \end{proof}
\section{Introduction} Strongly coupled quantum field theories play a central role in the description of nature at the most fundamental level. Quantum chromodynamics describes the observed strong interactions, and other strongly interacting theories are part of many efforts to extend the standard model. But strongly coupled field theories are notoriously difficult to analyze directly. Because of this, general constraints on their behavior can be very useful. In a recent paper \cite{acs}, a new general constraint on strongly coupled field theories was proposed. It takes the form of an inequality limiting the number of massless degrees of freedom in the infrared description of a field theory to be no larger than the number of ultraviolet degrees of freedom. This inequality was conjectured to apply to all asymptotically free theories (those governed by a free ultraviolet fixed point), whose infrared behavior is also governed by a fixed point, not necessarily free. It was noted that the inequality can also apply to certain theories with interacting ultraviolet fixed points. The inequality is formulated using finite temperature as a device to probe all energy scales, with the degree-of-freedom count defined using the free energy of the field theory. The zero-temperature theory of interest is characterized using the quantity $f_{IR}$, related to the free energy by \begin{equation} \label{eq:firdef} f_{IR} \equiv -\frac{90}{\pi^2} \lim_{T\to 0} \frac{\cal F}{T^4} \ , \end{equation} where $T$ is the temperature and $\cal F$ is the conventionally defined free energy per unit volume (which is equal to minus the pressure). This limit will be well defined if the theory has an infrared fixed point. For the special case of an infrared-free theory, $f_{IR}$ is simply the number of massless bosonic degrees of freedom plus $7/8$ times the number of massless fermionic degrees of freedom. The corresponding expression in the large $T$ limit is \begin{equation} \label{eq:fuvdef} f_{UV} \equiv - \frac{90}{\pi^2}\lim_{T\to \infty} \frac{\cal F}{T^4} \ . \end{equation} Just as in the infrared, this limit will be well defined if the theory has an ultraviolet fixed point. For an asymptotically free theory, $f_{UV}$ counts the underlying, ultraviolet degrees of freedom in a similar way. In terms of these quantities, the conjectured inequality for asymptotically-free theories is \begin{equation} \label{eq:ineq} f_{IR} \le f_{UV} \ . \end{equation} In Ref. \cite{acs}, the inequality was compared to known results and then used to derive new results for several strongly coupled, vector-like gauge theories. \footnote{\footnotesize{While the fixed points are necessary to ensure the existence of the limits Eqs. (\ref{eq:firdef},\ref{eq:fuvdef}), one can imagine applying the inequality when the ultraviolet fixed point is only approximate, that is when it is a feature of an effective low energy theory. $f_{UV}$ would then count degrees of freedom for energy scales much larger than masses and confinement or symmetry breaking scales in the effective theory, but much smaller than the scale of new physics.}} In the present paper, we extend this study to chiral gauge theories. These theories, in which the left- and right- handed fermions couple differently to the gauge fields, are potentially important examples of strongly coupled gauge field theories. The motivation for the effort to understand better the nonperturbative behavior of chiral gauge theories stems not just from their field-theoretic interest, but also from their possible application to physics beyond the standard model, including (i) models of particle substructure that can produce massless composite fermions and hence account for the fact that the observed fermions have masses much smaller than the lower bound on a hypothetical compositeness scale, and (ii) dynamical symmetry breaking of electroweak and higher symmetries. A key feature is that the fermion content is subject to an additional constraint not present in vectorial gauge theories, namely the absence of gauge and global $\pi_4$ (Witten) anomalies (and the absence of mixed gauge-gravitational anomalies, if one includes gravity). This greatly reduces the number of models that one can consider. On the other hand, a number of powerful techniques that one can use for vectorial gauge theories are absent for chiral gauge theories; these include correlation function inequalities (since the fermion measure is not positive). There are also complications in trying to formulate chiral gauge theories on the lattice because of fermion doubling. Nevertheless, one can still use the 't Hooft global anomaly matching conditions \cite{thooft} as well as large-$N$ methods. For asymptotically free chiral theories, a variety of infrared phenomena can be consistent with global anomaly matching. One possibility is that as in QCD the gauge symmetry remains intact but the theory confines and the global flavor symmetries break spontaneously. Another possibility is that confinement sets in but that the global symmetries are unbroken. This is realized by the formation of gauge singlet, massless composite fermions, along with other possible degrees of freedom. It is also possible that the theory does not confine but is governed in the infrared by an interacting fixed point, possibly weak. The symmetries will again remain unbroken. Yet another well-known possibility is that these theories dynamically break their own gauge symmetries, e.g., by the formation of fermion condensates. Each of these possibilities except the last will play a role here. We examine several anomaly-free chiral gauge theories. Some of these are automatically asymptotically free. For the others, we restrict the number of fermion representations so that they are asymptotically free. The quantity $f_{UV}$ may then be computed perturbatively. We consider the possible infrared realizations of the theories as allowed by global anomaly matching and large-$N$ methods, and compute $f_{IR}$. For some models, the inequality is automatically satisfied, while for others the assumption of the validity of the inequality provides a strong restriction on the infrared realization. \section{SU($N$) Models} We begin with models based on the gauge group SU($N$) \cite{drs,eppz}. For all the models considered in this paper, the beta function is generically written as \begin{equation} \beta = \mu \frac{d \alpha}{d\mu} = -\beta_1 \Bigl ( \frac{\alpha^2}{2\pi} \Bigr ) - \beta_2 \Bigl ( \frac{\alpha^3}{4\pi^2} \Bigr ) + O(\alpha^4) \ , \label{betafun} \end{equation} where the terms of order $\alpha^4$ and higher are scheme-dependent. The first model contains massless fermions transforming according to (i) a symmetric tensor representation of SU($N$), $S = \psi^{(ab)}_L$, and (ii) $N+4$ conjugate fundamental representations $F^c_{iL} =\psi^a_{iL}$, where $a,b$ are SU($N$) group indices, $i=1,..,N+4$ is a flavor index, and all fermion fields are written as left-handed Weyl fields. The $\beta$-function coefficients are $\beta_1 = 3N-2$, so that the theory is asymptotically free, and $\beta_2 = (1/4)(13N^2 - 30N + 1 + 12N^{-1})$. Since the theory is asymptotically free, $f_{UV}$ is given by the free-field count of the thermodynamic degrees of freedom: \begin{equation} f_{UV} = 2(N^2-1) + (7/4) [ (1/2)N(N+1) + (N+4)N ] \ . \label{finfsym} \end{equation} To determine $f_{IR}$, we assume that the theory confines and apply the 't Hooft anomaly matching conditions. The global flavor symmetry group is $G_f =$ SU($N+4$) $\times$ U(1), where the SU($N+4$) mixes the $(N+4)$ $F^c_i$ fields and the U(1) is the linear combination of the original U(1)'s generated by $S \to e^{i\theta_S}S$ and $F^c \to e^{i\theta_{F^c}}F^c$ that is left invariant by instantons. The anomaly matching conditions are consistent with the hypothesis that the global flavor symmetry group $G_f$ is unbroken and the massless spectrum is comprised of gauge-singlet composite fermions transforming according to the antisymmetric second-rank tensor representation of $G_f$. In the large-$N$ limit, it has in fact been argued \cite{eppz} that this {\it is} the infrared spectrum of the theory. From this spectrum one can determine that \begin{equation} f_{IR} = \frac{7}{4} \frac{(N+4)(N+3)}{2} \ . \label{fzerosym} \end{equation} whence \begin{equation} \Delta f \equiv f_{UV} - f_{IR} = (1/4)[ 15N^2 + 7N - 50 ] \ , \label{deltafsym} \end{equation} which is positive for all $N \ge 2$. The inequality is automatically satisfied. The second model contains massless fermions in (i) the antisymmetric tensor representation of SU($N$), $A = \psi^{[ab]}_L$ and, if $N \ge 5$, also (ii) $N-4$ conjugate fundamental representations $F^c_{iL} =\psi^a_{iL}$, where $i=1,...,N-4$. For the beta function, we have $\beta_1=3N+2 > 0$, and $\beta_2 = (1/4)(13N^2 + 30N + 1 - 12N^{-1})$. Since the theory is asymptotically free, we find \begin{equation} f_{UV} = 2(N^2-1) + \frac{7}{4}\biggl [ \frac{N(N-1)}{2} + (N-4)N \biggr ] \ . \label{finfantisym} \end{equation} For $N \ge 6$, the global symmetry group is $G_f =$ SU($N-4$) $\times$ U(1), where the SU($N-4$) mixes the $N-4$ fields $F^c_i$ and the U(1) is the linear combination of the original U(1)'s generated by $A \to e^{i\theta_A}A$ and $F^c \to e^{i\theta_F^c}F^c$ that is left invariant by instantons. The anomaly matching conditions are consistent with the conclusion that the massless spectrum consists of gauge-singlet composite fermions transforming according to the symmetric second-rank tensor representation of $G_f$. (In the degenerate case, $N=4$, there are no massless fermions.) From this spectrum, we deduce that \begin{equation} f_{IR} = \frac{7}{4} \frac{(N-4)(N-3)}{2} \ . \label{fzeroantisym} \end{equation} Whence \begin{equation} \Delta f = (1/4)[ 15N^2 - 7N - 50 ] \ , \label{deltafantisym} \end{equation} which is positive for the relevant range $N \ge 4$. The inequality is again satisfied. The third model is an extension of model 1 with the same gauge group and fermions transforming as (i) a symmetric tensor representation $S = \psi^{(ab)}_L$; (ii) $N+4$ conjugate fundamental representations: $F^c_i = \psi^c_{iL}$, where $i=1,...,N+4$; and (iii) $p$ pairs of fundamental and conjugate fundamental representations $F_{iL}, \ F^c_{iL}, i=1,...,p$. We have $\beta_1=3N-2-(2/3)p$ and $\beta_2=(1/4)\{13N^2-30N+1+12/N -2p((13/3)N-1/N)\}$. Hence, the theory is asymptotically free if \begin{equation} p < (9/2)N - 3 \ . \label{afby} \end{equation} We shall restrict $p$ so that this condition is satisfied. We then find \begin{equation} f_{UV} = 2(N^2-1) + \frac{7}{4}\biggl [ \frac{N(N+1)}{2} + (N+4)N + 2pN \biggr ] \ . \label{finfby} \end{equation} The global symmetry group is \begin{equation} G_f = SU(r) \times SU(p) \times U(1) \times U(1)' \label{gglobal3} \end{equation} where \begin{equation} r = N+4+p \ . \label{rdef} \end{equation} The first U(1) in (\ref{gglobal3}) is generated by $F_{iL} \to e^{i\omega}F_{iL}$ and $F^c_{iL} \to e^{-i\omega}F^c_{iL}$, which is a vectorial symmetry and hence is not affected by instantons; the U(1)$^{\prime}$ is the one left invariant by instantons. The anomaly matching conditions are consistent with the suggestion \cite{by} that the spectrum of the model consists of gauge-singlet massless composite fermions transforming according to the representations \begin{equation} (\ \asym \ , \ 1) + (\ \overline{\fund} \ , \ \fund \ ) + (1, \ \overline{\sym} \ ) \label{model3comp} \end{equation} of $SU(r) \times SU(p)$ in $G_f$. For simplicity, we will restrict our discussion to the large-$N$ limit. It was noted in Ref. \cite{eppz} that the above composite-fermion realization of $G_f$ is not possible if this limit is taken with fixed $p$. Assuming confinement, $G_f$ would have to be broken to a smaller group. If, on the other hand, we take the limit \begin{equation} N \rightarrow \infty, \quad p \rightarrow \infty, \quad \frac{p}{N} \sim O(1) \label{pnlimit} \end{equation} (so that loops of the $p$ fermions are not suppressed), the above realization with unbroken $G_f$ is possible. We therefore assume the limits in Eq. (\ref{pnlimit}), with \begin{equation} \frac{p}{N} \equiv \lambda < \frac{9}{2} \label{pnaf} \end{equation} for asymptotic freedom, and calculate that \begin{equation} f_{IR} = \frac{7}{4}\biggl [ \frac{r(r-1)}{2} + rp + \frac{p(p+1)}{2} \biggr ] \ , \label{fzeroby} \end{equation} where only the leading terms in the limit (\ref{pnlimit}) are relevant. It follows that, keeping only these leading terms, \begin{equation} \Delta f = (15/4)N^2 - (7/2)p^2 \ . \label{deltafby} \end{equation} Hence, in the large-$N$ limit, $\Delta f > 0$ only if $\lambda$ (is finite and) satisfies \begin{equation} \lambda \le \Bigl (\frac{15}{4} \Bigr )^{1/2} = 1.04 \ . \label{pnineq} \end{equation} which is a considerably stronger upper bound on $p$ than (\ref{pnaf}). (For very small $\lambda \sim const./N$, we have already noted that this realization is impossible and that $G_f$ must break.) Thus there is a finite $\lambda$ range leading to composite fermions and an unbroken $G_f$. For larger $\lambda$ values, up to the asymptotic freedom limit $\lambda = 9/2$, we expect the theory to be in the nonabelian Coulomb phase, with an interacting infrared fixed point. Near the upper end, the fixed point will be weakly interacting as determined by the first two terms in the $\beta$ function. The fixed point is then given by $\alpha_* = -2\pi \beta_1/\beta_2$, and since $\beta_{1} >0$, it exists only if $\beta_{2} <0$. In the large-$N$ limit, one finds that the two-loop fixed point exists if \begin{equation} \frac{3}{2} < \lambda < \frac{9}{2} \ , \label{lamir} \end{equation} and its value is \begin{equation} \alpha_*N = \frac{8\pi(9-2\lambda)}{13(2\lambda-3)} \ . \label{anir} \end{equation} Of course, the lower end of this range is not reliable since the fixed-point coupling, as determined by the two-loop $\beta$ function, approaches infinity. For $\lambda$ near 9/2, the infrared and ultraviolet degrees of freedom are the same, and $\Delta f$ is positive due to the (negative) perturbative correction to $f_{IR}$. Whether the nonabelian Coulomb phase persists down to the value in Eq. (\ref{pnineq}), $\lambda=1.04$, or even lower, is unknown. The inequality says simply that the confined phase with massless composite fermions and unbroken global symmetry group $G_f$ cannot exist if $\lambda > 1.04$. The full exploration of the phases of this model as a function of $\lambda$ is an interesting and unsolved strong-coupling problem. \section{Direct Product Gauge Groups} We next examine a class of models first discussed by Georgi \cite{g} with direct product gauge groups $G_k$ composed by alternating SU($N$) and SU($M$) $k$ times, so that $G_2=$ SU($N) \times$ SU($M$), $G_3=$ SU($N) \times$ SU($M) \times$ SU($N$), etc. Letting $M(N,1)$ denote $M$ copies of the representation $(N,1)$, we take the fermion content to be \begin{equation} M(N,1), \quad (\bar N, \bar M), \quad N(1,M) \quad {\rm for} \quad k=2 \ , \label{fermionsk2} \end{equation} \begin{equation} M(N,1,1), \quad (\bar N, \bar M,1), \quad (1,M,N), \quad M(1,1,\bar N) \quad {\rm for} \quad k=3 \ , \label{fermionsk3} \end{equation} \begin{equation} M(N,1,1,1), \ (\bar N,\bar M,1,1), \ (1,M,N,1), \ (1,1,\bar N,\bar M), \ N(1,1,1,M) \ \ {\rm for} \ \ k=4 \ , \label{fermionsk4} \end{equation} etc. For the $i$'th SU($N$), $\beta(g_i) = -(g_i^3/(48 \pi^2))(11N-2M) + O(g_i^5)$ and for the $j$'th SU($M$), $\beta(g_j) = -(g_j^3/(48 \pi^2))(11M-2N) + O(g_j^5)$, so that the theory is asymptotically free if \begin{equation} \frac{2}{11} < \frac{N}{M} < \frac{11}{2} \ , \label{nm411} \end{equation} which will be assumed henceforth. Then \begin{equation} f_{UV} = 2[\ell_k(N^2-1)+\ell(M^2-1)]+(7/4)(k+1)MN \ , \label{finfg} \end{equation} where $\ell_k=\ell$ if $k=2\ell$ is even and $\ell+1$ if $k=2\ell+1$ is odd. Again assuming confinement so that the massless, physical states transform as singlets under $G_k$, one can apply the 't Hooft anomaly matching conditions. Consider first even $k$; then the global flavor symmetry group is $G_{f,k \ {\rm even}}$=SU($M) \times$ SU($N) \times$ U(1), where the SU($M$) mixes the $M$ copies of $(N,1,...,1)$, the SU($N$) mixes the $N$ copies of $(1,...,M)$, and the U(1) is the one left invariant by instantons. The anomaly matching conditions for the SU($N$)$^3$, SU($M$)$^3$, SU($M$)$^2$U(1), and SU($N$)$^2$U(1) anomalies are consistent with the conclusion, also supported by large $N,M$ arguments \cite{eppz} (with $N/M$ fixed in the interval (\ref{nm411})) that the spectrum consists of massless composite fermions that transform according to the $(M,N)$ representation of $G_f$. Hence, $f_{IR, \ k \ {\rm even}} = (7/4)MN$ and \begin{equation} \Delta f_{k \ {\rm even}} = 2\ell(N^2+M^2-2)+(7/4)kMN > 0 \ . \label{delta_f_gkeven} \end{equation} Next consider odd $k$; then the global flavor symmetry group is $G_{f,k \ {\rm odd}}$=SU($M) \times$ SU($M) \times$ U(1), where the two SU($M$)'s mix the $M$ copies of $(N,1,...,1)$ among themselves and the $M$ copies of $(1,...,N)$ among themselves, respectively, and the U(1) is as before, with appropriate charges. Anomaly matching conditions and large $M,N$ arguments imply that $G_f$ is spontaneously broken to $G_{diag.}=$ SU($M)_V \times$ U(1). Hence, the massless spectrum consists of $M^2-1$ Goldstone bosons, so that $f_{IR, \ k \ {\rm odd}} = M^2-1$ and \begin{equation} \Delta f_{k =2\ell+1} = 2(\ell+1)(N^2-1) + (2\ell-1)(M^2-1) +(7/2)(\ell+1)MN > 0 \ . \label{delta_f_gkodd} \end{equation} Thus, the global symmetry is believed to remain unbroken with massless composite fermions for even $k$, and is believed to break, producing Goldstone bosons, for odd $k$. In both cases, $\Delta f > 0$. The inequality provides no new information about these models. \section{Supersymmetric Models} There are a number of chiral ${\cal N}=1$ supersymmetric gauge theories for which dual descriptions are known which are weakly coupled in the IR. It is a nontrivial test of the inequality to check that $f_{IR}$ as computed from these duals is indeed less than $f_{UV}$ . We have performed this check for several chiral supersymmetric theories and found the predictions of the inequality to be in agreement with the conjectured duals in all cases. The theories we studied include (i) all so-called s-confining theories as listed in Ref. \cite{css}; (ii) an $SO(10)$ theory with chiral superfields transforming according to a spinor and $N$ vector representations \cite{ps}, and (iii) an $SU(N)$ theory with chiral superfields comprised of a symmetric tensor $S$ and $N+4$ antifundamental representations $\bar Q$ \cite{ps2}. The theory (iii) is the ${\cal N}=1$ supersymmetric version of the model we discussed in the beginning of Section II (with no tree-level superpotential). This theory is asymptotically free for all $N$, and we obtain \begin{equation} f_{UV}=\frac{15}4\left(N^2-1+\frac12 N(N+1)+N(N+4)\right) \ . \label{fuvps} \end{equation} In \cite{ps} a dual ${\cal N}=1$ supersymmetric description of the IR dynamics of this theory was proposed. The dual theory is non-chiral and has gauge group $SO(8)$, $N+4$ copies of the vector representation of $SO(8)$, a spinor of $SO(8)$, and $(N+4)(N+5)/2 +1$ gauge singlets. For $N \ge 13$ the dual is IR free and we can calculate $f_{IR}$ by simply counting dual fields \begin{equation} f_{IR}=\frac{15}4\left(28+8(N+4)+8+\frac12(N+4)(N+5)+1\right) \ . \end{equation} The constraint $f_{IR}\le f_{UV}$ then becomes $N \ge 9$; thus the inequality is satisfied for all numbers of flavors for which we can reliably compute the quantities $f_{UV}$ and $f_{IR}$. Although we have concentrated on chiral gauge theories in this paper, we add some remarks here on a particular vectorial gauge theory for which no dual description is known. As we will see, applying the conjectured inequality yields a new constraint on the IR spectrum of this theory. The model is an ${\cal N}=1$ supersymmetric $SU(N)$ gauge theory with the same matter content as ${\cal N}=2$ supersymmetric QCD: $F$ chiral superfields $Q_i$ and $\bar Q_i$ transforming in the fundamental and antifundamental representations of $SU(N)$ and one chiral superfield $A$ in the adjoint representation. The difference with the ${\cal N} =2$ theory is that we do not include a tree level superpotential term $W=g Q A \bar Q$. Setting the superpotential coupling to zero breaks the extended supersymmetry to ${\cal N}=1$ and the resulting theory is much less well understood. Our goal is to investigate how the inequality constrains the infrared dynamics of this theory. For $F< 2N$ the theory is free in the ultraviolet, and therefore \begin{equation} f_{UV}=\frac{15}4 \left(2(N^2-1)+2FN\right) \ . \end{equation} A full description of the IR theory is not known, but a few general statements can be made. The classical scalar potential of this theory has a large number of flat directions; the theory has a moduli space of inequivalent vacua. It is well known that the classical moduli space can be parametrized in terms of the independent gauge invariant polynomials which in this case are \begin{equation} T_k={\rm Tr} A^k , \ k=2,...,N \quad {\rm and}\quad M_i= Q A^i \bar Q\ ,\ i=0,...,N-1 \ , \end{equation} in addition to several baryonic gauge invariants. Here the $T_k$ are singlets, whereas the $M_i$ have $F^2$ components. No infrared-free description to this theory has been found, and there are arguments \cite{kss} that the theory at the origin of moduli space remains interacting and is conformal in the infrared for any value of $0<F<2N$. Still, the inequality can provide useful information, as we will now discuss. An interesting picture for the infrared dynamics has been put forward in \cite{lst}, based on a study of a closely related $Sp(2N)$ theory. If one assumes that the infrared fixed point corresponds to an interacting superconformal theory, then the superconformal algebra must contain an anomaly-free $U(1)_R$ symmetry. It follows from the superconformal algebra that the scaling dimensions of all chiral composites are equal to $3/2$ times their charge under the $U(1)_R$. Unfortunately, the R-symmetry of our theory is not unique; there exists a one-parameter family of R-symmetries of which one is the $U(1)_R$ in the superconformal algebra. However one can still derive relations between the scaling dimensions of the chiral composites $M_i$ and $T_k$. The result of such an analysis is that for $F < N/2$, the scaling dimensions of some composites, as determined from the superconformal R-charges, drop below the unitarity bound. This then implies that these composites decouple from the interacting conformal theory. Similar behavior is anticipated for $N/2 \le F < 2N$. This picture for the dynamics is also supported by duality \cite{lst}. In summary, we expect that in the deep infrared this theory splits into disjoint sectors: one sector with an interacting conformal theory, and other sectors consisting only of a set of $m$ free composite fields $M_i$. We now apply the inequality to get a constraint on this scenario. To calculate $f_{IR}$ we add the contributions from each of the disjoint sectors. We cannot calculate the contribution from the conformal sector, but we know that it is positive.\footnote{ \footnotesize{It follows from the definition in Eq. (\ref{eq:firdef}) together with ${\cal F}=-p$, that $f_{IR}$ has the same sign as the pressure, which is always positive.}} This allows us to compute a lower bound on $f_{IR}$ by adding only the contributions from the $m$ free fields $M_0,..., M_{m-1}$: \begin{equation} f_{IR}\ge \frac{15}4 m F^2 \ . \end{equation} Demanding that $f_{IR}\le f_{UV}$ then gives $m\le 2(\frac{N^2}{F^2}+ \frac{N}F-\frac1{F^2})$. Since only the first $N$ of the $M_i$ are independent gauge invariants, we already have that $m\le N$, so that the inequality provides new information only for $F>2$. The constraint then simplifies to \begin{equation} m < 2\frac{N}{F}(\frac{N}{F}+1) \ . \end{equation} We see that for larger $F$ this constraint can be very significant. For example, for $F$ near the asymptotic freedom bound at $F=2N$ we find $m<3/2$, implying that only the first in the series of mesons, $M_0$, can decouple from the conformal theory and become a free field. The case of $SU(2)$ is somewhat special and leads to a tight constraint. For $SU(2)$ both $Q$ and $\bar Q$ transform in the fundamental doublet representation of $SU(2)$, and the global flavor symmetry is enhanced to $SU(2F)$. The mesons $M_{0,1}$ transform as tensors under this enlarged flavor symmetry. $M_0$ is an antisymmetric tensor, whereas $M_1$ is a symmetric tensor. The constraint then allows the following scenario: for $F\ge4$ the theory is not asymptotically free and we do not get a constraint; for $F=2,3$ only the meson $M_0$ can decouple and become free; and for $F=1$ both $M_0$ and $M_1$ may be free. \section{Summary} In summary, we have performed a quantitative comparison of the ultraviolet and infrared degrees of freedom, as measured by $\Delta f=f_{UV}-f_{IR}$, for several chiral gauge theories and for one vector-like, supersymmetric gauge theory. We have shown that in some models $\Delta f$ is automatically positive, whereas in others, such as the third model of Section II and the vector-like, supersmmetric model ( Section IV), the conjectured inequality $\Delta f \ge 0$ places interesting restrictions on the low energy structure of the theory. This research was partially supported by the DOE and NSF grants DE-FG02-92ER-4074 (TA), DE-FG02-92ER-40704 (AC), DE-AC03-76SF00515 (MS), and NSF-PHY-97-22101 (RS). T. A. acknowledges the hospitality of the ITP, Stony Brook during a visit in December, 1998, where this work was initiated.
\section{Why density matrices?} ``It is a common error always to view a mixed state as describing a system that is actually in one of a number of different possible pure states, with specified probabilities. While this `ignorance interpretation' of the mixed state can indeed be a useful practical way to describe an ensemble of completely isolated systems, it entirely misses the deep and fundamental character of mixed states: If a system has any external correlations whatever, then its quantum state cannot be pure. Pure states are a rarity, enjoyed only by completely isolated systems. The states of externally correlated individual systems are fundamentally and irreducibly mixed. This has nothing to do with `our ignorance'. It is a consequence of the existence of objective external correlation'' \cite{Mermin}. The quotation from the Mermin paper can serve as a motto for what we are going to present below. A physical system that is described by a state vector at time $t=0$ will remain in a pure state for all $t>0$ if and only if it will never interact with anything. An interaction leads to correlations, and correlations mean non-product (entangled) states. A subsystem of a bigger composite system which, as a whole, is in an entangled state, is no longer described by a state vector but by a non-pure reduced density matrix. The density matrix is an entirely quantum object whose ontological status is the same as this of the state vector it originated from. A theory that deals only with pure states deals either with the entire Universe or with objects that cannot interact and are therefore unobservable. The above remarks are especially relevant if one thinks of nonlinear generalizations of quantum mechanics. `Nonlinear quantum mechanics' is traditionally associated with nonlinear Schr\"odinger equations. Such equations have the general form \bea i|\dot\psi\rangle=H(\psi)|\psi\rangle.\label{1} \eea Assuming we have a concrete physical system which, when isolated, evolves according to (\ref{1}) we immediately face the question of how to describe this system during or after an interaction. The problem has led to a great amount of confusion as to the real status of nonlinear generalizations of quantum mechanics. Obviously, the resulting misunderstandings involve an interpretation of experiments testing linearity of quantum mechanics: A theory that does not tell us how to describe external correlations does not produce testable predictions and, hence, cannot be tested. The problem of subsystems versus (\ref{1}) was, in our oppinion, solved by Polchinski in \cite{P}. In short, his answer was the following: The equation describing a composite system is \bea i|\dot\Psi\rangle=\Big( H_{\rm s}(\rho_{\rm s}) + H_{\rm r}(\rho_{\rm r}) \Big) |\Psi\rangle\label{2} \eea where the index `s' stands for a subsystem and `r' for the `rest' or `remaining systems'. The state vector $|\Psi\rangle$ represents the pure state of the composite `subsystem plus the rest' system, and the $\rho$'s are the appropriate reduced density matrices. Let us note that, even if we assume, as we have done in (\ref{2}), that the composite system is in a pure state, the formalism inevitably leads us to density-matrix-dependent nonlinear operators $H_{\dots}(\rho_{\dots})$. To include interactions one can write \bea i|\dot\Psi\rangle=\Big( H_{\rm s}(\rho_{\rm s}) + H_{\rm r}(\rho_{\rm r}) + H_{\rm int}(\Psi) \Big) |\Psi\rangle\label{3} \eea with some interaction (linear or nonlinear) Hamiltonian operator $H_{\rm int}(\Psi)$. The main feature of (\ref{2}) and (\ref{3}) is the fact that for $H_{\rm int}(\Psi)=0$ (that is, when the interaction is over) the reduced density matrices satisfy the von Neumann-type equations \bea i\dot \rho_{\rm s} &=& [H_{\rm s}(\rho_{\rm s}),\rho_{\rm s}],\label{vNa}\\ i\dot \rho_{\rm r} &=& [H_{\rm r}(\rho_{\rm r}),\rho_{\rm r}].\label{vNb} \eea The form of these von Neumann equations illustrates the {\it locality\/} of the Polchinski formulation: Subsystems that do not interact do not `see' each other. It should be stressed that this property is typically lost when one considers different ways of extending the subsystem dynamics. An interesting discussion of the problem is given in these proceedings by L\"ucke \cite{L} who shows that there may exist local extensions different from those discussed by Polchinski. However, the nonlinear Schr\"odinger equations they apply to are linearizable by a nonlinear gauge transformation \cite{G}. The Polchinski-type extensions bear some kind of universality but simultaneously are non-unique in the following sense: There exists an infinite number of inequivalent extensions which reduce to the same equation on product states. A more detailed discussion of this and related problems can be found in \cite{P,J,MC96,MC97,MCMK98}. From what we have written so far it follows that nonlinear von Neumann equations, whether we like it or not, will always appear when quantum nonlinearity occurs. A somewhat more radical viewpoint is suggested by the non-uniqueness of the Polchinski-type extensions. To illustrate the point consider a simple spin-1/2 nonlinear `average energy' \bea H(\psi) &=& \frac{\langle\psi|\sigma_z|\psi\rangle^2}{\langle\psi|\psi\rangle}. \label{sigma} \eea where $\sigma_z$ is the Pauli matrix. This type of nonlinearity was considered in experiments involving 2-level atoms \cite{MC96}. Now, how to write $H(\rho)$ on the basis of $H(\psi)$? A natural guess is \bea H(\rho) &=& \frac{({\rm Tr\,}\sigma_z\rho)^2}{{\rm Tr\,}\rho}. \eea but \bea H(\rho) &=& \frac{{\rm Tr\,}\sigma_z\rho\sigma_z\rho}{{\rm Tr\,}\rho}. \eea and an infinite number of similar fuctions would do as well. They all reduce to (\ref{sigma}) if $\rho=|\psi\rangle\langle\psi|$. However, if we {\it start\/} with some $H(\rho)$ then the whole ambiguity disappears. This suggests that looking for a fundamental level of nonlinear quantum dynamics one should begin with von Neumann and not Schr\"odinger equations. Linear Schr\"odinger equation, as seen from a classical perspective, is simply the equation of motion of an infinite-dimensional Hamiltonian system with average energy playing a role of Hamiltonian function. For a detailed presentation of the formalism see the paper by Cirelli {\it et al.\/} published in these proceedings \cite{C}. Linear von Neumann equation is also the equation of motion of a classical Hamiltonian system with average energy in the role of the Hamiltonian function. Geometrically, density matrices form an infinite-dimensional Poisson manifold with $gl(\infty)$ Lie-Poisson bracket. This is described in the present volume by B\'ona \cite{B} and for this reason we will not spend much time on details of the Lie-Poisson formulation. \section{Nonlinear von Neumann equation} Let us denote by $\rho_a$ a matrix element of the density matrix $\rho$ taken in some basis. The Lie-Poisson version of the {\rm linear\/} von Neumann equation is \bea i\dot\rho_a=\{\rho_a,\langle H\rangle\}\label{LP} \eea where $\langle H\rangle={\rm Tr\,}H\rho$ is the Hamiltonian function. Nonlinear quantum mechanics {\it \'a la\/} B\'ona and Jordan \cite{J,B} is based on the same Lie-Poisson structure but with $\langle H\rangle$ replaced by a nonlinear Hamiltonian function $H_{\rm gen}$. The dynamics given by (\ref{LP}) or its nonlinear generalization \bea i\dot\rho_a=\{\rho_a,H_{\rm gen}\}\label{nLP} \eea defines a Hamiltonian flow on the Poisson manifold of states. Having such a flow one can consider a classical probability density $w(\rho)$ and its associated Liouville equation \bea i\dot w=\{w,H_{\rm gen}\}\label{L}. \eea The classical probability density has a clear physical meaning: It describes a classical lack of knowledge about the state of a quantum source. In any experiment one faces this type of clasical uncertainty and all experimental averages one measures in a lab are of the form \bea \langle A\rangle_{\rm exp} = \int d\rho \,w(\rho){\rm Tr\,}A\rho\label{Aexp} \eea with $\int d\rho$ meaning an integration over the parameters defining initial conditions for the quantum dynamics. The Liouville equation (\ref{L}) is {\it linear\/} in $w$ independently of whether the von Neumann equation (\ref{nLP}) is linear in $\rho_a$. Moreover, the probability density $w$ is directly accessible to the experimentalist and reflects the classical configuration of the experimental setting. For this reason it is very important that the Liouville equation is linear. Formula (\ref{Aexp}) shows that it is practical to introduce the object \bea \rho_{\rm exp} = \int d\rho \,w(\rho)\rho\label{rhoexp} \eea satisfying \bea \langle A\rangle_{\rm exp} = {\rm Tr\,}A\rho_{\rm exp} \eea and playing a role of a {\it semiclassical\/} density matrix. The `common error', mentioned by Mermin in the quotation we have started with, is the belief that all density matrices one encounters in quantum mechanics have such a semiclassical origin. The `truely quantum' density matrices one obtains by reduction of entangled states to subsystems can be written in different bases of `pure states' and all such bases are regarded as physically equivalent. Putting it differently, no particular decomposition of such a $\rho$ into a convex combination of projectors is physically special. This is one of the important impossibility principles of quantum mechanics \cite{M}. On the other hand, the decomposition defined by $w$ is not only very special but, actually, is even uniquely given by the form of the experiment. An experimentalist can arbitrarily tamper with $w$ but different convex combinations forming a concrete $\rho$ are definitely out of his reach. One can imagine also physical situations which are somehow in-between the cases we have just described. For example, consider an entangled pair of particles and an experiment where we have a shutter which is opened whenever a particle labeled `1', say, is measured and a concrete result (say, spin `up' or `down' in a `$z$' direction) is found. Each time the shutter is opened the particle labeled `2' leaves a box and, hence, the box is a source of particles in a concrete mixed state which depends on the entangled state of the pair. The mixed state, as depending on the macroscopic and clasically controlled actions undertaken by the experimentalist, is no longer `fully quantum'. The resulting mixture is of a $\rho_{\rm exp}$ type and there is no reason for the corresponding dynamics of the density matrix to be nonlinear. It seems reasonable to assume that nonlinear quantum dynamics of mixed states can occur only in cases where the very form of the `pure-state' decomposition is {\it in principle\/} out of control. Such an impossibility principle seems to eliminate all the problems analyzed by Polchinski in \cite{P}. The distinction between the von Neumann equation (\ref{nLP}) and the Liouville equation (\ref{L}) is therefore essential. One should not use the misleading term the `Liouville-von Neumann equation' which suggests that the von Neumann equation is simply a quantized version of the linear Liouville equation and, accordingly, {\it must\/} be linear as well. Historically, the linearity of the standard von Neumann equation seems to have its roots in the linearity of the Schr\"odinger equation. The two equations, \bea i|\dot\psi_k\rangle=H|\psi_k\rangle, \quad -i\langle\dot \psi_k| = \langle\psi_k|H \eea combined with $\rho=|\psi_k\rangle\langle\psi_k|$ imply \bea i\dot \rho &=& [H,\rho].\label{lvN} \eea Since the equation is linear, one can consider more general convex combinations \bea \rho=\sum_k \lambda_k |\psi_k\rangle\langle\psi_k| \eea which again satisfy (\ref{lvN}). This simple argument looks so natural that one may have not even noticed the additional assumption we have smuggled in. Indeed, we have started with (\ref{lvN}) which was valid for pure states, that is those satisfying $\rho^2=\rho$, and by linearity extended the argument to $\rho\neq \rho^2$. But a pure $\rho$ has eigenvalues 0 and 1. Therefore any function $f$ satisfying $f(0)=0$ and $f(1)=1$ will satisfy $f(\rho)=\rho$ if $\rho$ is pure! It follows that the linearity of the Schr\"odinger equation implies at most an equation of the form \bea i\dot \rho &=& [H,f(\rho)]\label{fvN} \eea with $[H,f(\rho)]=[H,\rho]$ for $\rho^2=\rho$. In the next sections we shall devote much attention to the so-called Euler-Arnold-von Neumann equation obtained if $f(x)=x^2$. Now, what is the relation between (\ref{fvN}) and (\ref{nLP})? The answer is rather surprising: (\ref{fvN}) is an example of (\ref{nLP}) with an appropriate choice of $H_{\rm gen}$ \cite{MCJN}. To see how this works assume that for $x\in [0,1]$ the function $f$ has a convergent Taylor expansion \bea f(x)=\sum_{k=0}^\infty f_k(x-a)^k \label{f} \eea and define \bea H_{\rm gen}(\rho) &=& {\rm Tr\,}\big(f(\rho)H\big) \eea (in the context of \cite{MCJN} it is important that $H_{\rm gen}(\rho)$ is 1-homogeneous; this complication is left out here). Then (\ref{nLP}) is equivalent to \bea i\dot \rho=[\hat H(\rho),\rho]\label{effvN} \eea where the effective nonlinear Hamiltonian is defined by the functional derivative \cite{MCJN} \bea \hat H(\rho) = \frac{\delta}{\delta\rho}H_{\rm gen}(\rho) = \sum_{k=1}^\infty f_k\sum_{n=0}^{k-1}(\rho- a{{\bf 1}})^{k-1-n} H (\rho- a{{\bf 1}})^{n} \eea and \bea i\dot \rho=[\hat H(\rho),\rho]=[H,f(\rho)]. \eea The form (\ref{effvN}) implies that $C_n= {\rm Tr\,} (\rho^n)$ is time-independent for any natural $n$ (such $C_n$ are Casimir functions for $\{\cdot,\cdot\}$) from which it follows that the spectrum of a solution $\rho$ of (\ref{fvN}) has to be time-invariant \cite{MCMM}. As we can see, although the dynamics of $\rho$ is nonlinear for non-pure $\rho$, we are nevertheless still quite close to linear quantum mechanics. Looking at the explicit solutions for $f(x)=x^2$ we will see that we are, in fact, {\it surprisingly\/} close to the linear dynamics but with very specific and subtle new nonlinear effects at hand. It is quite clear there are fundamental reasons for investigating nonlinear density matrix equations. But one does not have to believe in nonlinear quantum mechanics to investigate nonlinear evolutions of mixed states. It is known that nonlinear von Neumann equations arise very naturally in various mean-field approaches \cite{R} where they are solved either numerically or by approximate methods. The program we are developing aims at finding {\it exact\/} techniques of solving nonlinear von Neumann equations. \section{Lax representation of von Neumann equations and binary Darboux covariance} The class of equations we have under some control is of the form \cite{SLMC,MKMCSL} \bea i\dot\rho &=& [\hat H(\rho),\rho] = {\sum_{k=0}^n} [A^{n-k}\rho A^k,\rho] ={\sum_{k=0}^n} [A^{n-k},\rho A^k \rho] \label{n-eq} \eea where $A$ is a time-independent self-adjoint operator. For $n=1$ and $A=H$ one finds \bea \hat H(\rho)=H \rho+ \rho H \eea and the equation becomes the Euler-Arnold-von Neumann equation \bea i\dot\rho =[H \rho+ \rho H,\rho]=[H,\rho^2]. \eea For all $n$ the von Neumann equations involve Hamiltonians $\hat H(\rho)$ which are {\it linear\/} in $\rho$, and therefore the nonlinearities of the equations are quadratic. This property is important as it seems that only the quadratic nonlinearities can be treated by the binary Darboux transformation. In fact, what makes the class so interesting is the knowledge of its binary-Darboux-covariant Lax representation \bea z_\mu |\varphi\rangle &=& (\rho-\mu A)|\varphi\rangle,\label{1a}\\ i|\dot\varphi\rangle &=& \Big( \sum_{k=0}^n A^{n-k}\rho A^k-\mu A^{n+1}\Big)|\varphi\rangle, \label{1b} \eea where $z_\mu$, $\mu\in{\bf C}$. The necessary condition for the solution $|\varphi\rangle$ of (\ref{1a})-(\ref{1b}) to exist is that the relation (\ref{n-eq}) between $A$ and $\rho$ holds true. The idea of the Darboux transformation is, given a density matrix $\rho$ and a solution $|\phi\rangle$ of (\ref{1a})-(\ref{1b}), to find a new solution $|\varphi[1]\rangle$ and a new $\rho[1]$ which are again related by the pair (\ref{1a})-(\ref{1b}) with $A$ unchanged. The fact that $|\varphi[1]\rangle$ is a solution means that the nesessary condition for its existence is satisfied, and this implies that $\rho[1]$ and $A$ are again related by (\ref{n-eq}). One may think of $\rho$ and $\rho[1]$ as `ground' and `first excited' states, and the Darboux transformation is a `creation operator'. The so-called binary Darboux transformation is constructed as follows (for more details and generalizations cf. \cite{SLMC,MKMCSL,ZL,Leble,U}). Take some linear operators $V$ and $J$, a parameter $s$, and three complex numbers $\mu$, $\nu$, and $\lambda$. Now consider three linear equations \bea i\partial_s|\varphi\rangle &=& \big(V-\mu J)|\varphi\rangle, \label{ket}\\ -i\partial_s\langle\chi| &=& \langle\chi|\big(V-\nu J), \label{bra1}\\ -i\partial_s\langle\psi| &=& \langle\psi|\big(V-\lambda J). \label{bra2} \eea It is important that (\ref{bra1}) and (\ref{bra2}) involve bras and (\ref{ket}) involves a ket. Define \bea P&=&\frac{|\varphi\rangle\langle\chi|}{\langle\chi|\varphi\rangle}\label{P1}\\ V[1] &=& V + (\mu-\nu)[P,J]\label{VBDT}\\ \langle\psi[1]| &=& \langle\psi|\Big({{\bf 1}} - \frac{\nu-\mu}{\lambda-\mu}P\Big)\label{BDT} \eea (the fact that we transform here $\langle\psi|$ and not $|\varphi\rangle$ or $\langle\chi|$ is not essential; one can construct similar transformations of any of these solutions with the help of the remaining two). A straightforward calculation then shows that \bea -i\partial_s\langle\psi[1]| &=& \langle\psi[1]|\big(V[1]-\lambda J).\label{eq[1]} \eea To apply the above technique to our von Neumann equations we take three {\it pairs\/}: (\ref{1a})-(\ref{1b}), which we already have, and two more with parameters $\lambda$, $z_\lambda$, $\nu$, $z_\nu$: \bea z_\lambda \langle\psi| &=& \langle\psi|(\rho-\lambda A),\label{2a}\\ -i\langle\dot\psi| &=& \langle\psi|\Big( \sum_{k=0}^n A^{n-k}\rho A^k-\lambda A^{n+1}\Big), \label{2b}\\ z_\nu \langle\chi| &=& \langle\chi|(\rho-\nu A),\label{3a}\\ -i\langle\dot\chi| &=& \langle\chi|\Big( \sum_{k=0}^n A^{n-k}\rho A^k-\nu A^{n+1}\Big).\label{3b} \eea In what follows the $\mu$- and $\nu$-pairs (\ref{1a})--(\ref{1b}) and (\ref{3a})--(\ref{3b}) will be used to define the binary transformation of the conjugated $\lambda$-pair (\ref{2a})--(\ref{2b}). Defining $P$ as above and \bea \rho[1]=\rho+(\mu-\nu)[P,A]\label{rho[1]} \eea we find \cite{MKMCSL} \bea z_\lambda \langle\psi[1]| &=& \langle\psi[1]|(\rho[1]-\lambda A)\\ -i\langle\dot\psi[1]| &=& \langle\psi[1]|\Big( \sum_{k=0}^n A^{n-k}\rho[1] A^k-\lambda A^{n+1}\Big). \eea If all the objects necessary for the construction of $\langle\psi[1]|$ exist, then the necessary condition for its existence must be fulfilled, and this means \bea i\dot\rho[1] &=& {\sum_{k=0}^n} \big[A^{n-k}\rho[1] A^k,\rho[1]\big] ={\sum_{k=0}^n} \big[A^{n-k},\rho[1] A^k \rho[1]\big] \label{n-eq[1]}. \eea Having one solution $\rho$ we have managed to produce another solution $\rho[1]$. In this respect the Darboux transformation is a really wonderful device. However, from time to time surprises can occur. It is clear that if all the assumptions we have made are satisfied then $\rho[1]$ must be a solution. But the general theorem does not guarantee that the solution is nontrivial! In fact $\rho=A$, $\rho={{\bf 1}}$, or even $\rho=0$ are also solutions of (\ref{n-eq}), so can we guarantee that such pathological cases are excluded and $\rho[1]$ is physically interesting? The answer depends on what is meant by `interesting'. In general, each Darboux transformation has an inverse. In classical problems, such as the Korteweg-de Vries equation, it is typical to start with a trivial solution $u=0$ and $u[1]$ is already a soliton, definitely a highly nontrivial solution \cite{MS}. This means there exists a transformation that maps a soliton into 0. We will now show that this type of pathology is excluded if a binary transformation of the type we use is considered. What is not excluded, however, are the cases when, say, $\rho[1]=\rho$ or $\rho[1]=T\rho T^{-1}$, with $T$ a time-independent unitary transformation. Such a possibility leads to nontrivial practical complications. The next theorem is of fundamental importance and its proof is so elementary that we can give it here \cite{MKMCSL}. Consider the following three general Lax pairs \bea z_\mu |\varphi\rangle &=& (\rho-\mu A)|\varphi\rangle ,\label{L1a}\\ i|\dot \varphi\rangle &=& \big(V(\rho) - \mu J\big)|\varphi\rangle,\label{L1b}\\ z_\nu \langle\chi| &=& \langle\chi|(\rho-\nu A), \label{L3a}\\ -i\langle\dot \chi| &=& \langle\chi| \big(V(\rho) - \nu J\big),\label{L3b}\\ z_\lambda \langle\psi| &=& \langle\psi|(\rho-\lambda A),\label{L2a}\\ -i\langle\dot \psi| &=& \langle\psi| \big(V(\rho) - \lambda J\big).\label{L2b} \eea The only assumption we make about $\rho$, $J$, $A$, and $V(\rho)$ is the covariance of (\ref{L2a})--(\ref{L2b}) under the binary Darboux transformation constructed with the help of $\langle\chi|$ and $|\varphi\rangle $. \medskip \noindent {\bf Theorem~1.} Under the above assumptions the binary Darboux transformation $\rho\mapsto\rho[1]$ is a similarity transformation, $\rho[1]=T\rho T^{-1}$, where \bea T &=& {{\bf 1}}+\frac{\mu-\nu}{\nu}P=e^{P\ln\frac{\mu}{\nu}} \eea {\it Proof\/}: By definition $\rho[1]=\rho+(\mu-\nu)[P,A]$. Eqs.~(\ref{L1a}), (\ref{L3a}) imply \bea z_\mu P &=& (\rho-\mu A)P\label{LP1a}\\ z_\nu P &=& P(\rho-\nu A)\label{LP3a} \eea from which it follows that \bea P(\rho-\mu A)P &=& (\rho-\mu A)P\label{PLP1a}\\ P(\rho-\nu A)P &=& P(\rho-\nu A).\label{PLP3a} \eea Multiplying (\ref{PLP1a}) by $\nu$, (\ref{PLP3a}) by $\mu$, and subtracting the resulting equations we get \bea [P,A]=\frac{\nu-\mu}{\mu\nu} P\rho P-\frac{1}{\mu} \rho P+\frac{1}{\nu} P\rho. \eea Inserting this expression into the definition of $\rho[1]$ we obtain \bea \rho[1]= \Big({{\bf 1}}+\frac{\mu-\nu}{\nu}P\Big) \rho \Big({{\bf 1}}+\frac{\nu-\mu}{\mu}P\Big).\label{central} \eea Q.E.D. The theorem explains why $\rho[1]=0$ etc. is excluded if $\rho$ is a density matrix: Spectra of $\rho$ and $\rho[1]$ must be identical. This property is characteristic of all equations that are compatibility conditions for binary-Darboux-covariant Lax pairs. The result will, in general, not hold if one considers a binary-Darboux-covariant zero-curvature pair of the type used for example in \cite{U}. The next important result, whose proof follows from a straightforward calculation, is the covariance under `spectrum shifting and rescaling'. \medskip \noindent {\bf Theorem~2.} Assume $[X,A]=[X,\rho]=0$, $Y\in {\bf R}$, and $\rho$ is a solution of (\ref{n-eq}). Then \bea \rho_X(t) &=& e^{-i(n+1)XA^nt}\big(\rho(t)+X\big)e^{i(n+1)XA^nt}\\ \rho_Y(t) &=& Y\rho(Yt) \eea also satisfy (\ref{n-eq}). \section{Two strategies leading to `interesting' solutions} The problem with the Darboux transformation is that in order to find a solution $\rho[1]$ we must already {\it know\/} somehow another solution $\rho$ to start with. As we have seen above, there are some obvious solutions such as 0, but by Theorem~1 they will not lead to anything nontrivial. Still, the case is not completely hopeless. \subsection{The first strategy} The strategy works for the Euler-Arnold-von Neumann equation \cite{SLMC} \bea i\dot \rho &=& [H,\rho^2].\label{EA} \eea In all the examples discussed below we assume $\nu=\bar \mu$ and $|\varphi\rangle=|\chi\rangle$ since this guarantees that hermiticity is conserved by the transformation. A pure state $\rho=\rho^2$ is simultaneously a solution of (\ref{EA}) and of the linear von Neumann equation with the same $H$. Therefore, pure state solutions of the linear equation form a nontrivial subset of solutions of the nonlinear one. Unfortunately, by Theorem~1 we will have $\rho[1]=\rho[1]^2$ and although such $\rho[1]$ cannot be claimed `trivial' they are nevertheless quite `uninteresting'. Interesting solutions are obtained if one starts with $\rho$ satisfying $[H,\rho^2-a\rho]=0$, for some $a\in {\bf R}$, but such that the operator $\Delta_a:= \rho^2-a \rho\neq 0$ is not a constant times ${\bf 1}$. In this case \cite{SLMC} \bea \rho(t)=e^{-ia Ht}\rho(0)e^{ia Ht} \eea and \bea \rho[1](t) &=& e^{-iaHt}\Big(\rho(0)+(\mu-\bar\mu)F_a(t)^{-1} e^{-\frac{i}{\mu}\Delta_a t} \big[|\varphi(0)\rangle\langle\varphi(0)|,H\big] e^{\frac{i}{\bar \mu}\Delta_a t}\Big)e^{iaHt},\nonumber\\ &=:& e^{-ia Ht}\rho_{\rm int}[1](t)e^{ia Ht}\label{rho[1]'} \eea where $\mu$ is a complex parameter of the Darboux transformation, $|\varphi(0)\rangle$ a solution of the Lax pair at $t=0$, and $$F_a(t)= \langle\varphi(0)| \exp\Big( i \frac{\mu-\bar\mu}{|\mu|^2}\Delta_a t \Big)|\varphi(0)\rangle.$$ After this first step we know were to look, but we still have to find an appropriate $\rho$ and $|\varphi\rangle$. It turns out we need three more tricks which are best illustrated by the following example (see the Appendix). Consider the Hamiltonian \bea H= \left( \begin{array}{ccc} 0 & 1 & 0\\ 1 & 0 & 0\\ 0 & 0 & \frac{1}{\sqrt{2}} \end{array} \right)\label{H} \eea and take $\mu=i$ (for real $\mu$ the binary transformation is trivial). We have to find an appropriate initial condition $\rho(0)$. The first trick we have mentioned is to begin with a solution which is neither normalized nor positive. Such a non-density-matrix solution will be denoted by $\xi $ instead of $\rho$. We can always make it positive and normalized by using the transformations of Theorem~2. We take $\xi (t)=e^{-iHt}\xi (0)e^{iHt}$ with \bea \xi (0)= \left( \begin{array}{ccc} \frac{1}{2}+ \frac{\sqrt{2}}{2} & 0 & 0\\ 0 & \frac{1}{2}- \frac{\sqrt{2}}{2} & 0\\ 0 & 0 & \frac{1}{2} \end{array} \right). \eea The first two matrix elements on the diagonal are the solutions of the equation $x^2-x=1/4$ (this is the second trick) and for this reason \bea \Delta_1= \xi (0)^2-\xi (0)=\xi (t)^2-\xi (t)= \frac{1}{4} \left( \begin{array}{ccc} 1 & 0 & 0\\ 0 & 1 & 0\\ 0 & 0 & -1 \end{array} \right) \eea which obviously commutes with $H$ ($1/4$ is not essential here; this could be basically any nonzero number). Therefore $[H,\xi (t)^2]=[H,\xi (t)]$ even though $\xi (t)^2\neq \xi (t)$ and $[H,\xi (t)]\neq 0$. It may happen that appropriate solutions of the quadratic equation $x^2-x=x_0$ do not exist. But to make the trick work it is sufficient to find solutions of $x^2-ax=x_0$ with some $a$ and then we will have $\Delta_a$ instead of $\Delta_1$. The third trick is to choose $\xi (0)$ and $|\varphi\rangle$ in such a way that the solution of the Lax pair is not an eigenstate of $\Delta_a$ (since the contributions from $F_a(t)$, $\exp(-\frac{i}{\mu}\Delta_a t)$, and $\exp(\frac{i}{\bar \mu}\Delta_a t)$ would cancel one another in (\ref{rho[1]'}), and the internal part of (\ref{rho[1]'}) would become time-independent --- that is exactly what we want to avoid). The eigenvalues of $\xi (0)-iH$ are $z_\pm=(1\pm i\sqrt{2})/2$ and $z_-$ has degeneracy 2. The two eigenvectors corresponding to $z_-$ are orthogonal: \bea |\varphi_1\rangle = \left( \begin{array}{c} 0\\ 0\\ 1 \end{array} \right), \quad |\varphi_2\rangle = \frac{1}{\sqrt{2}} \left( \begin{array}{c} e^{i\pi/4}\\ 1\\ 0 \end{array} \right). \eea Taking an arbitrary linear combination of them, say, \bea |\varphi(0)\rangle= \frac{1}{\sqrt{2}}\Big(|\varphi_1\rangle+|\varphi_2\rangle\Big) \eea we obtain an eigenvector all the three components of which are non-vanishing. This implies that the unitary transformation $T$ occuring in Theorem~1 will not be block-diagonal. The fact that the two orthogonal eigenvectors correspond to the same eigenvalue of $\xi -\mu H$ and to different eigenvalues of $\Delta_a$ allows us to construct a nonlinear dynamics involving the entire 3-dimensional space and automatically guarantees that $|\varphi\rangle$ is not an eigenstate of $\Delta_a$. We find \bea \xi [1](t) &=& e^{-iHt}\xi _{\rm int}[1](t) e^{iHt}\label{int} \eea with \bea \xi _{\rm int}[1](t) = \left( \begin{array}{ccc} \frac{1 + \sqrt{2}}{2} - \frac{\sqrt{2}}{1 + e^t} & 0 & \frac{-1 - i}{2\sqrt{2}\cosh(t/2)}\\ 0 & \frac{1 - \sqrt{2}}{2} + \frac{\sqrt{2}}{1 + e^t} & \frac{1}{2\cosh(t/2)}\\ \frac{-1 + i}{2\sqrt{2}\cosh(t/2)} & \frac{1}{2\cosh(t/2)} & \frac{1}{2} \end{array} \right). \label{int'} \eea $\xi (t)$ is not yet a density matrix since it has a negative eigenvalue $(1-\sqrt{2})/2$ and its trace is $3/2$. Now it is time to use transformations from Theorem~2. Set $X=\frac{\sqrt{2}-1}{2}{\bf 1}$ and $Y=\sqrt{2}/3$. A combination of the two transformations leads to the final solution \bea \rho_{XY}[1](t) &=& e^{-i\frac{2}{3}Ht}\rho_{\rm int}(t)e^{i\frac{2}{3}Ht} \label {finsol} \eea where \bea \rho_{\rm int}(t) = \frac{\sqrt{2}}{3} \left( \begin{array}{ccc} \sqrt{2} - \frac{\sqrt{2}}{1 + e^{\sqrt{2}t/3}} & 0 & \frac{-1 - i}{2\sqrt{2}\cosh[t/(3\sqrt{2})]}\\ 0 & \frac{\sqrt{2}}{1 + e^{\sqrt{2}t/3}} & \frac{1}{2\cosh[t/(3\sqrt{2})]}\\ \frac{-1 + i}{2\sqrt{2}\cosh[t/(3\sqrt{2})]} & \frac{1}{2\cosh[t/(3\sqrt{2})]} & \frac{\sqrt{2}}{2} \end{array} \right). \eea The density matrix $\rho_{XY}[1](t)$ has eigenvalues $2/3$, $1/3$, and 0. For $|t|\to\infty$ one gets the dynamics asymptotically {\it linear\/} but with different asymptotics for negative and positive times, and a kind of `self-scattering' (or `phase transition' as we called it in \cite{SLMC}) around $t=0$. To have some qualitative feel of what happens consider averages of spin-1 matrices \bea J_x = \left( \begin{array}{ccc} 0 & 0 & 0\\ 0 & 0 & i\\ 0 & -i & 0 \end{array} \right), \quad J_y = \left( \begin{array}{ccc} 0 & 0 & -i\\ 0 & 0 & 0\\ i & 0 & 0 \end{array} \right), \quad J_z = \left( \begin{array}{ccc} 0 & i & 0\\ -i & 0 & 0\\ 0 & 0 & 0 \end{array} \right).\nonumber \eea The figures illustrate the effect. Figures 1 and 2 show the evolution of the $x$--$y$ projection of $\langle {\bf J}\rangle$. The magnitude of the oscillation goes very quickly to 0 as $|t|$ grows. Figure~3 shows the dynamics of the $z$ component of $\langle {\bf J}\rangle$. The process of `self-scattering' maps one asymptotically linear solution into another. \begin{figure} \epsfig{figure=fig2a.eps} \caption{Projection of the average spin on the $x$--$y$ plane for times $0\leq t\leq 10$ (in dimensionless units). The amplitude of the oscillation decreases with time.} \end{figure} \begin{figure} \epsfig{figure=fig2b.eps} \caption{Seeds of destruction of the past asymptotic state: The same dynamics as in Fig.~1 but for times $-230\leq t\leq -220$. The amplitude is $10^{22}$ times smaller than at Fig.~1. The amplitude of the oscillation increases with time.} \end{figure} \begin{figure} \epsfig{figure=fig3.eps} \caption{The `self-scattering': Average $\langle J_z\rangle$ as a function of time (solid). The dynamics is asymptotically linear but with different asymptotics for $t\to +\infty$ (dashed) and $t\to -\infty$ (dotted). The self-scattering phase shift is clearly visible.} \end{figure} A discussion of more realistic systems, including a one-dimensional harmonic oscillator, can be found in \cite{SLMC}. \subsection{The second strategy} The first strategy leads to solutions involving a finite number of states and cannot be directly applied to $n>1$ equations from the Darboux covariant family (\ref{n-eq}): The non-abelian shape of, say, the $n=2$ equation \bea i\dot \rho &=& [A^2,\rho^2]+[A,\rho A\rho] \eea requires different tricks. The solutions discussed in the previous subsection started with non-stationary $\rho$'s, since those satisfying $[\rho,A]=0$ lead to the projector $P$ commuting with both $A$ and $\rho$, and the binary transformation is trivial. Still, there exists another class of stationary solutions of (\ref{n-eq}), obtained if $A\rho=-\rho A$. Now the projector $P$ will not, in general, commute with $\rho$ and $A$, and the binary transformation may be nontrivial. Assume $[P,\rho]\neq 0$. For $n=2m$ one has $\sum_{k=0}^n (-1)^k=1$ and \bea i|\dot\varphi\rangle &=& \Big( A^n\rho-\mu A^{n+1}\Big)|\varphi\rangle= z_\mu A^n|\varphi\rangle\label{15} \eea The formal solution of the latter equation is \bea |\varphi(t)\rangle = e^{-iz_\mu A^n t}|\varphi(0)\rangle \label{sol even}. \eea For odd $n$ we get $\sum_{k=0}^n (-1)^k=0$ and \bea i|\dot\varphi\rangle &=& -\mu A^{n+1}|\varphi\rangle \eea implying \bea |\varphi(t)\rangle = e^{i\mu A^{n+1} t}|\varphi(0)\rangle\label{sol odd} . \eea To have an `interesting' $\rho[1]$ we must make sure that the solution of the Lax pair is not an eigenstate of $A^{2m}$ (for $n=2m$) or $A^{2(m+1)}$ (for $n=2m+1$); otherwise we have a problem similar to this with the eigenstates of $\Delta_a$. Let us note that $\rho$ which anticommutes with $A$ does commute with $A^2$ and therefore also with the generators of the time evolution given by (\ref{sol even}) and (\ref{sol odd}), and we can look for an eigenstate of $\rho - \mu A$ at $t=0$. For simplicity we shall check the trick again on the Euler-Arnold-von Neumann equation. Let $\alpha_k$ and $\sigma_k$ be Dirac-$\alpha$ and Pauli matrices, respectively. Consider the Hamiltonian \bea H=\alpha_1\otimes {\bf 1}_{2\times 2}+{\bf 1}_{4\times 4}\otimes \sigma_1, \eea and take a non-density-matrix stationary solution \bea \xi =\alpha_2\otimes \sigma_2 + \alpha_3\otimes \sigma_3 \eea satisfying $\xi H=-H\xi $. Set $\mu=i$ and take \bea \langle\varphi| &=& (i,0,-1,0,-i,0,1,0) \eea satisfying $(\rho-iH)|\varphi\rangle=0$. Then \bea {}&{}& \xi [1](t)=\nonumber\\ &{}& \left( \begin{array}{cccccccc} {\frac{1 + {e^{4 t}}}{1 + {e^{8 t}}}} & {\frac{i {e^{4 t}}}{1 + {e^{8 t}}}} & 0 & {\frac{-1}{1 + {e^{8 t}}}} & {\frac{e^{8 t} - {e^{4 t}}}{1 + {e^{8 t}}}} & {\frac{-i {e^{4 t}}}{1 + {e^{8 t}}}} & 0 & {\frac{-e^{8 t}}{1 + {e^{8 t}}}} \\ {\frac{-i {e^{4 t}}}{1 + {e^{8 t}}}} & {\frac{{e^{4 t}}-1}{1 + {e^{8 t}}}} & {\frac{1}{1 + {e^{8 t}}}} & 0 & {\frac{i {e^{4 t}}}{1 + {e^{8 t}}}} & {\frac{-e^{8 t} - {e^{4 t}}}{1 + {e^{8 t}}}} & {\frac{{e^{8 t}}}{1 + {e^{8 t}}}} & 0 \\ 0 & {\frac{1}{1 + {e^{8 t}}}} & {\frac{-1 - {e^{4 t}}}{1 + {e^{8 t}}}} & {\frac{i {e^{4 t}}}{1 + {e^{8 t}}}} & 0 & {\frac{{e^{8 t}}}{1 + {e^{8 t}}}} & {\frac{{e^{4 t}}-e^{8 t}}{1 + {e^{8 t}}}} & {\frac{-i {e^{4 t}}}{1 + {e^{8 t}}}} \\ {\frac{-1}{1 + {e^{8 t}}}} & 0 & {\frac{-i {e^{4 t}}}{1 + {e^{8 t}}}} & {\frac{1 - {e^{4 t}}}{1 + {e^{8 t}}}} & {\frac{-e^{8 t}}{1 + {e^{8 t}}}} & 0 & {\frac{i {e^{4 t}}}{1 + {e^{8 t}}}} & {\frac{e^{8 t} + {e^{4 t}}}{1 + {e^{8 t}}}} \\ {\frac{e^{8 t} - {e^{4 t}}}{1 + {e^{8 t}}}} & {\frac{-i {e^{4 t}}}{1 + {e^{8 t}}}} & 0 & {\frac{-e^{8 t}}{1 + {e^{8 t}}}} & {\frac{1 + {e^{4 t}}}{1 + {e^{8 t}}}} & {\frac{i {e^{4 t}}}{1 + {e^{8 t}}}} & 0 & {\frac{-1}{1 + {e^{8 t}}}} \\ {\frac{i {e^{4 t}}}{1 + {e^{8 t}}}} & {\frac{-e^{8 t} - {e^{4 t}}}{1 + {e^{8 t}}}} & {\frac{{e^{8 t}}}{1 + {e^{8 t}}}} & 0 & {\frac{-i {e^{4 t}}}{1 + {e^{8 t}}}} & {\frac{{e^{4 t}}-1}{1 + {e^{8 t}}}} & {\frac{1}{1 + {e^{8 t}}}} & 0 \\ 0 & {\frac{{e^{8 t}}}{1 + {e^{8 t}}}} & {\frac{{e^{4 t}}-e^{8 t}}{1 + {e^{8 t}}}} & {\frac{-i{e^{4t}}}{1 + {e^{8t}}}} & 0 & {\frac{1}{1 + {e^{8t}}}} & {\frac{-1 - {e^{4t}}}{1 + {e^{8t}}}} & {\frac{i{e^{4t}}}{1 + {e^{8t}}}} \\ {\frac{-e^{8 t}}{1 + {e^{8t}}}} & 0 & {\frac{i{e^{4t}}}{1 + {e^{8t}}}} & {\frac{e^{8 t} + {e^{4t}}}{1 + {e^{8t}}}} & {\frac{-1}{1 + {e^{8t}}}} & 0 & {\frac{-i{e^{4t}}}{1 + {e^{8t}}}} & {\frac{1 - {e^{4t}}}{1 + {e^{8t}}}} \end{array} \right).\nonumber \eea The eigenvalues of $\xi $ and $\xi [1](t)$ are $0$ and $\pm 2$. To produce the density matrix we shift the spectrum by $\Lambda\geq 2$ and rescale to get the unit trace. This will be again a `self-scattering' solution but its explicit form will not be shown here. The second strategy has the advantage of being applicable to generically infinite-dimensional problems. A nontrivial question (requiring still more tricks) is how to produce trace-class solutions if the Hilbert space is not finite-dimensional. \section*{Acknowledgments} Our work was supported by the KBN Grant No. 2~P03B~163~15, the Flemish-Polish Project No. 007, and (MC) Alexander von Humboldt-Stiftung. \section*{Appendix} Generic solutions of the equation of motion (\ref {EA}) with $H$ given by (\ref {H}) can be obtained by a straightforward integration. Change the basis so that $H$ is of the form \bea H=-\left(\matrix{ \mu &0 &0\cr 0 &-\mu &0\cr 0 &0 &\lambda\cr }\right) \label {3by3} \eea Then the diagonal elements of $\rho$ are constants of motion. Write down the equation of motion for the remaining matrix elements. One obtains \bea i\dot\rho_{1,2}&=&2\mu [(\rho_{1,1}+\rho_{2,2})\rho_{1,2} +\rho_{1,3}\overline{\rho_{2,3}}],\cr i\dot\rho_{1,3}&=&(\mu-\lambda) [(\rho_{1,1}+\rho_{3,3})\rho_{1,3} +\rho_{1,2}\rho_{2,3}],\cr i\dot\rho_{2,3}&=&-(\mu+\lambda) [(\rho_{2,2}+\rho_{3,3})\rho_{2,3} +\overline{\rho_{1,2}}\rho_{1,3}]. \label {snled} \eea Let $W=|\rho_{1,2}|^2$. It is not difficult to show that it satisfies the equation \bea \ddot W=aW^2+bW+c \label {cleq} \eea with $a$, $b$, and $c$ constants. Equation (\ref {cleq}) is well known \cite {DA}. Its solutions are doubly-periodic functions. The result is of the form \bea W(t)=\beta^{-1}{\rm sn}^2(\alpha(t-t_0),k)+\gamma \eea with sn the Jacobi elliptic function, and with $k$, $\alpha$, $\beta$, $\gamma$, and $t_0$ constants. In the limit $k=1$ one of the periods becomes infinite. The result (\ref {finsol}) obtained by the binary Darboux transformation corresponds precisely to this $k=1$ solution. It is at the moment not clear to us what kind of a seed solution (if any) can lead, via the binary Darboux transformation, to the $k\neq 1$ class. An algebraic characterization of such additional seed solutions would be important from the perspective of more general cases, especially the infinite-dimensional ones.
\section{Introduction} It is well known that Cepheid pulsating stars constitute, through the Period-Luminosity (PL) relation, the one primary calibrator for extragalactic distance scale. Photometric data of extragalactic Cepheids are then the raw material of its calibration, and, thus, it is of greater importance to have as much data as possible. With the aim of computing new distances based on our direct calibration of the PL relation (Lanoix et al. 1999a) while taking into account the PL relation incompleteness bias (Lanoix et al. 1999b) first, and subsequently calibrating secondary distance calibrators such as Tully-Fisher relation, luminosity peak brightness of SNe Ia (Lanoix 1998) or Faber-Jackson relation for globular clusters (Di Nella-Courtois et al. 1999), we have looked for photometric data of extragalactic Cepheids. We found that Madore published in 1985 a compilation of all data available at this time. However, this compilation actually put data from 10 galaxies together, and to our knowledge, no updated version of this paper exists, whereas data from more than 30 galaxies is now available. We have therefore updated this compilation for our own work while putting quality flags to every light curve, and have now decided to put this useful tool at the astronomical community's disposal. \section{Data} We collect 3031 photometric measurements of 1061 Cepheids located in 33 galaxies (without including SMC and LMC). Table \ref{galaxy} gives the complete list and the main characteristic of those galaxies according to the LEDA database (http://www-obs.univ-lyon1.fr/leda/home$\_$leda.html). Our bibliography is as exhaustive as possible, and is complete until November 1998, while new publications are still arriving. \begin{table}[ht] \caption{List of the 33 galaxies of the database} \label{galaxy} \begin{center} {\small \begin{tabular}{llrrlcrc}\hline Name & PGC/LEDA number & RA 2000 & DEC 2000 & Morph. type & Lum. class code & Total B-magnitude & HST \\ \hline DDO 155 & PGC 44491 & 12.97775& 14.21618 & Irr & 9.000 & 14.715 & n \\ DDO 216 & PGC 71538 & 23.47614& 14.74660 & Irr & 9.000 & 12.789 & n \\ DDO 50 & PGC 23324 & 8.31831& 70.71419 & Irr & 8.279 & 11.092 & n \\ DDO 69 & PGC 28868 & 9.98995& 30.74495 & Irr & 9.000 & 12.956 & n \\ IC 10 & PGC 01305 & .34016& 59.29171 & Irr & 9.000 & 12.197 & n \\ IC 1613 & PGC 03844 & 1.08172& 2.13330 & Irr & 9.230 & 9.933 & n \\ IC 4182 & PGC 45314 & 13.09704& 37.60582 & Sm & 8.338 & 12.409 & y \\ NGC 1365 & PGC 13179 & 3.56016&-36.13807 & SBb & 1.371 & 10.350 & y \\ NGC 2090 & PGC 17819 & 5.78398&-34.25145 & Sc & 3.931 & 11.767 & y \\ NGC 224 & PGC 02557 & .71232& 41.26897 & Sb & 2.000 & 4.170 & n \\ NGC 2366 & PGC 21102 & 7.48175& 69.21442 & Irr & 8.722 & 11.430 & n \\ NGC 2403 & PGC 21396 & 7.61513& 65.59957 & SBc & 5.000 & 8.824 & n \\ NGC 2541 & PGC 23110 & 8.24451& 49.06227 & SBc & 6.696 & 12.043 & y \\ NGC 300 & PGC 03238 & .91493&-37.68250 & Scd & 5.969 & 8.785 & n \\ NGC 3031 & PGC 28630 & 9.92597& 69.06665 & Sb & 2.000 & 7.687 & y \\ NGC 3109 & PGC 29128 & 10.05185&-26.15890 & SBm & 7.924 & 10.347 & n \\ NGC 3351 & PGC 32007 & 10.73278& 11.70408 & SBb & 3.000 & 10.382 & y \\ NGC 3368 & PGC 32192 & 10.77922& 11.82098 & SBab & 3.000 & 9.916 & y \\ NGC 3621 & PGC 34554 & 11.30466&-32.81352 & SBcd & 5.849 & 10.077 & y \\ NGC 4321 & PGC 40153 & 12.38200& 15.82293 & SBbc & 1.000 & 9.992 & y \\ NGC 4414 & PGC 40692 & 12.44097& 31.22479 & Sc & 3.576 & 10.923 & y \\ NGC 4496A& PGC 41471 & 12.52771& 3.93911 & SBd & 5.600 & 12.116 & y \\ NGC 4536 & PGC 41823 & 12.57414& 2.18848 & SBbc & 2.824 & 11.012 & y \\ NGC 4725 & PGC 43451 & 12.84079& 25.50030 & SBab & 1.689 & 9.955 & y \\ NGC 5253 & PGC 48334 & 13.66551&-31.64477 & S? & / & 10.765 & y \\ NGC 5457 & PGC 50063 & 14.05356& 54.35075 & SBc & 1.000 & 8.197 & y \\ NGC 598 & PGC 05818 & 1.56414& 30.66017 & Sc & 4.000 & 6.193 & n \\ NGC 6822 & PGC 63616 & 19.74940&-14.80306 & Irr & 8.493 & 9.322 & n \\ NGC 7331 & PGC 69327 & 22.61809& 34.41949 & Sbc & 2.000 & 10.165 & y \\ NGC 925 & PGC 09332 & 2.45467& 33.57817 & SBcd & 4.000 & 10.583 & y \\ SEXTANS A& PGC 29653 & 10.18369& -4.71346 & Irr & 9.704 & 11.745 & n \\ SEXTANS B& PGC 28913 & 9.99996& 5.33256 & Irr & 8.117 & 11.834 & n \\ WLM & PGC 00143 & .03246&-15.45032 & Irr & 8.258 & 11.113 & n \\ \hline \end{tabular}} \end{center} \end{table} One can note that some old photographic data has been rejected. For instance, concerning NGC 224 (M31), we exclude data from Gaposchkin (1962) and Baade \& Swope (1963, 1964) from our base. Moreover, we also exclude measurements of that galaxy taken in very crowded fields, such as part of Welch et al. (1986) ones. However finding charts or light curves may be found in these papers. Actually, the compilation can be divided into two subclasses: ground based data and {\it Hubble Space Telescope} (HST) data. The first class appears very heterogeneous in many aspects (methods, limit magnitudes, bandpasses, time coverage, quality) since it's made of observations from many different telescopes. Moreover almost all those Cepheids were observed during a single observation campaign too, so that they cannot be compared to any other campaign. On the other hand, the spatial observations of the HST are highly homogeneous and are composed of Cepheids from 17 galaxies at present. However, concerning papers from the HST Key Project group (see Freedman et al. 1994), we have chosen to keep only data from ALLFRAME software package, while two sets of photometry may be available. We have checked every light curve (except for NGC 3368, NGC 4725 and NGC 224) in order to allocate a flag to them, and then to characterize the type of the corresponding Cepheid, or to give an idea of the reliability of its photometry. Table \ref{quality} describes the signification of the different flags we use. The ground-based measurements are divided up among 11 different bandpasses, from B (440 nm) to K (2200 nm), whereas HST observed only in V and I bands. Table \ref{band} gives the relation between notations and wavelengths. \begin{table}[ht] \caption{Bandpasses} \label{band} \begin{center} {\small \begin{tabular}{llll}\hline Band & $\lambda_{eff}$ & Band & $\lambda_{eff}$ \\ & (nm) & & (nm) \\ \hline \\ B & 440 & IV & 1050 \\ V & 550 & J & 1250 \\ r & 650 & H & 1650 \\ R & 700 & K & 2200 \\ i & 820 & g & 500 \\ I & 900 \\ \hline \end{tabular}} \end{center} \end{table} \begin{table}[ht] \caption{Description of light curves flags} \label{quality} \begin{center} {\small \begin{tabular}{ll}\hline Flag & Light curve description \\ \hline \\ N & Normal \\ S & Symetrical (but high amplitude)\\ B & Bumpy \\ B+ & Scattered or very bumpy \\ O & Overtone\\ O- & Low amplitude (but asymmetrical or with high period)\\ P & Peculiar\\ / & No curve\\ \hline \end{tabular}} \end{center} \end{table} Our base contains several types of magnitude corresponding to the calculation method that is described in table \ref{type}. In some cases the sign ``:'' may follow magnitudes or periods : it means that these magnitudes are doubtful and that these periods are just estimates or lower bounds of the real values (according to the original authors). \begin{table}[ht] \caption{Magnitudes description} \label{type} \begin{center} {\small \begin{tabular}{cl}\hline Flag & Description \\ \hline \\ mea & Intensity averaged (based on curve area) \\ max & Maximum \\ min & Minimum\\ ave & Average of minimum and maximum\\ / & Single measurement \\ \hline \end{tabular}} \end{center} \end{table} The reference codes and the corresponding authors are given in table \ref{ref}. They can allow the interested reader to look at the finding charts as well as the light curves. Finally, the Cepheid names are those chosen by the authors, except in the case of several stars in NGC 224 (Freedman \& Madore 1990, field IV) and NGC 3368 (Tanvir 1995), where we obtain data right from their figures and called them FRE -- -- and TAN -- --, respectively. \section{Table structure} Data is presented as a table of more than 5000 lines according to the structure described below for each Cepheid : \begin{itemize} \item First line: Host galaxy, Cepheid name, number of following data lines for that Cepheid \item Second line: pointer = 1, logarithm of the period in days, light curve flag \item Next lines: pointer = 2, magnitude, type, band, reference \end{itemize} This database is available as an ASCII table on request by sending an e-mail to <EMAIL> . \newpage \begin{table}[ht] \caption{References} \label{ref} \begin{center} {\small \begin{minipage}[t]{70mm} \begin{tabular}{ll}\hline Reference code & Corresponding authors \\ \hline \\ Ala83 & Mc Alary et al. 1983 \\ Ala84 & Mc Alary et al. 1984 \\ Alv95 & Alves \& Cook 1995 \\ Car90 & Carlson \& Sandage 1990 \\ Cap92 & Capaccioli et al. 1992 \\ Chr87 & Christian \& Schommer 1987 \\ Coo86 & Cook \& Aaronson 1986 \\ Fer96 & Ferrarese et al. 1996 \\ Fer98 & Ferrarese et al. 1998 \\ Fr88a & Freedman 1988 \\ Fr88b & Freedman \& Madore 1988 \\ Fre90 & Freedman \& Madore 1990 \\ Fre91 & Freedman et al. 1991 \\ Fre92 & Freedman et al. 1992 \\ Fre94 & Freedman et al. 1994 \\ Gal96 & Gallart et al. 1996 \\ Gib98 & Gibson et al. 1998 \\ Gra97 & Graham et al. 1997 \\ Hoe90 & Hoessel et al. 1990 \\ Hoe94 & Hoessel et al. 1994 \\ Hoe98 & Hoessel et al. 1998 \\ Hug98 & Hughes et al. 1998 \\ Kay67 & Kayser 1967 \\ Kel96 & Kelson et al. 1996 \\ Kin87 & Kinman et al. 1987 \\ Mad85 & Madore et al. 1985 \\ \hline \end{tabular} \end{minipage} \ \begin{minipage}[t]{70mm} \begin{tabular}{ll}\hline Reference code & Corresponding authors \\ \hline \\ Mad87 & Madore et al. 1987 \\ McA84 & Mc Alary \& Madore 1984\\ Mou87 & Mould 1987 \\ Mus98 & Musella et al. 1997\\ Pio94 & Piotto et al. 1994 \\ Phe98 & Phelps et al. 1998 \\ Raw97 & Rawson et al. 1997 \\ Sah94 & Saha et al. 1994 \\ Sah95 & Saha et al. 1995 \\ Sa85a & Sandage \& Carlson 1985a \\ Sa85b & Sandage \& Carlson 1985b \\ Sa88a & Sandage 1988 \\ Sa88b & Sandage \& Carlson 1988 \\ Sh96a & Saha et al. 1996a \\ Sh96b & Saha et al. 1996b \\ Sh96c & Saha et al. 1996c \\ Sil96 & Silbermann et al. 1996 \\ Sil98 & Silbermann et al. 1998 \\ Tam68 & Tamman \& Sandage 1968 \\ Tan95 & Tanvir et al. 1995 \\ To95a & Tolstoy et al. 1995a \\ To95b & Tolstoy et al. 1995b \\ Tur98 & Turner et al. 1998 \\ Vis89 & Visvanathan 1989 \\ Wal88 & Walker 1988 \\ Wel86 & Welch et al. 1986 \\ \hline \end{tabular} \end{minipage}} \end{center} \end{table} \begin{verbatim} Table 6: Extract of the ASCII file for the Cepheid V1 of galaxy IC1613. ---------------------------------------- IC1613 V1 9 1 .748 N 2 21.36 mea B Fr88a 2 20.79 mea V Fr88a 2 20.36 mea R Fr88a 2 20.14 mea I Fr88a 2 20.50 max B Sa88a 2 22.03 min B Sa88a 2 21.27 ave B Sa88a 2 21.39 mea B Sa88a ---------------------------------------- \end{verbatim} \refer \aba \rf{Mc Alary, C.W. et al.: 1983, ApJ. 273, 539} \rf{Mc Alary, C.W., Madore, B.F., Davis, L.E.: 1984, ApJ. 276, 487} \rf{Alves, D.R., Cook, K.H.: 1995, AJ. 110, 192} \rf{Baade, W., Swope, H.H.: 1963, AJ 68, 435} \rf{Baade, W., Swope, H.H.: 1964, AJ 70, 212} \rf{Carlson, G., Sandage, A.: 1990, ApJ. 352, 587} \rf{Capaccioli, M., Piotto, G., Bresolin, F.: 1992, AJ. 103, 1151} \rf{Christian, C.A., Schommer, R.A.: 1987, AJ. 93, 557} \rf{Cook, K.H., Aaronson, M.: 1986, ApJ. Letters, 301, L45} \rf{Di Nella-Courtois, H., Lanoix, P., Paturel, G.: 1999, MNRAS (in press)} \rf{Ferrarese, L. et al.: 1996, ApJ. 464, 568} \rf{Ferrarese, L. et al.: 1998, ApJ. 507, 655} \rf{Freedman, W.L.: 1988, ApJ. 326, 691} \rf{Freedman, W.L., Madore, B.F.: 1988, ApJ. Letters 332, L63} \rf{Freedman, W.L., Madore, B.F.: 1990, ApJ. 365, 186} \rf{Freedman, W.L., Wilson, C.D., Madore, B.F.: 1991, ApJ. 372, 455} \rf{Freedman, W.L., Madore, B.F., Hawley, S.L., Horowitz, I.K., Mould, J., Navar ette, M., Sallmen, S.: 1992, ApJ. 396, 80} \rf{Freedman, W.L. et al.: 1994, ApJ. 427, 628} \rf{Gallart, C., Aparicio, A., Vichez, J.M.: 1996, AJ. 112, 1928} \rf{Gaposchkin, S.: 1962, AJ 67, 334} \rf{Gibson B.K. et al.: astro-ph/981003} \rf{Graham, J.A. et al.: 1997, ApJ. 477, 535} \rf{Hoessel, J.G., Abbott, J., Saha, A., Mossman, A.E., Danielson, G.E.: 1990, A J. 100, 1151} \rf{Hoessel, J.G., Saha, A., Krist, J., Danielson, G.E.: 1994, AJ. 108, 645} \rf{Hoessel, J.G., Saha, A., Danielson, G.E.: 1998, AJ. 115, 573} \rf{Hughes, S.M.G. et al.: 1998, ApJ. 501, 32} \rf{Kayser, S.E.: 1967, A.J. 72, 134} \rf{Kelson, D.D. et al.: 1996, ApJ. 463, 26} \rf{Lanoix, P.: 1998, AA 331, 421} \rf{Lanoix, P., Paturel, G., Garnier, R.: 1999a, MNRAS, submitted} \rf{Lanoix, P., Paturel, G., Garnier, R.: 1999b, ApJ. 516, in press} \rf{Kinman, T.D., Mould, J.R., Wood, P.R.: 1987, AJ. 93, 833} \rf{Madore, B.F.: 1985, Proc. IAU Colloquium 82, Cambridge University press} \rf{Madore, B.F., Mc Alary, C.W., Mc Laren, R.A., Welch, D.L., Neugebauer, G., Matthews, K.: 1985, ApJ. 294, 560} \rf{Madore, B.F., Welch, D.L., Mc Alary, C.W., Mc Laren, R.A.: 1987, ApJ. 320, 26} \rf{Mc Alary, C.W., Madore, B.F.: 1984, ApJ. 282, 101} \rf{Mould, J.R.: 1987, PASP. 99, 1127} \rf{Musella, I., Piotto, G., Capaccioli, M.: 1997, AJ 114, 976} \rf{Piotto, G., Capaccioli, M., Pellegrini, C.: 1994, Astron. Astrophys. 287, 371} \rf{Phelps, R.L. et al.: 1998, ApJ. 500, 763} \rf{Rawson, D.M. et al.: 1997, ApJ. 490, 517} \rf{Saha, A., Labhardt, L., Schwengeler, H., Macchetto, F.D., Panagia, N., Sandage, A., Tammann, G.A.: 1994, ApJ. 425, 14} \rf{Saha, A., Sandage, A., Labhardt, L., Schwengeler, H., Tammann, G.A., Panagia, N., Macchetto, F.D.: 1995, ApJ. 438, 8} \rf{Saha, A., Sandage, A., Labhardt, L., Tammann, G.A., Macchetto, F.D., Panagia, N.: 1996a, ApJ. 466, 55} \rf{Saha, A., Hoessel, J.G., Krist, J., Danielson, G.E.: 1996b, AJ. 111, 197} \rf{Saha, A., Sandage, A., Labhardt, L., Tammann, G.A., Macchetto, F.D., Panagia, N.: 1996c, ApJ. SS. 107, 693} \rf{Sandage, A., Carlson, G.: 1985a, AJ. 90, 1464} \rf{Sandage, A., Carlson, G.: 1985b, AJ. 90, 1019} \rf{Sandage, A.: 1988, PASP. 100, 935} \rf{Sandage, A., Carlson, G.: 1988, AJ. 96, 1599} \rf{Silbermann, N.A. et al.: 1996, ApJ. 470, 1} \rf{Silbermann, N.A. et al.: Astro-ph 9806017} \rf{Tamman, G.A., Sandage, A.: 1968, ApJ. 151, 825} \rf{Tanvir, N.R., Shanks, T., Ferguson, H.C., Robinson, D.R.T.: 1995, Nature 337, 27} \rf{Tolstoy, E., Saha, A., Hoessel, J.G., McQuade, K.: 1995a, AJ. 110, 1604} \rf{Tolstoy, E., Saha, A., Hoessel, J.G., Danielson, G.E.: 1995b, AJ. 109, 579} \rf{Turner A. et al.: 1998, ApJ. 505, 207} \rf{Visvanathan, N.: 1989, ApJ. 346, 629} \rf{Walker, A.R.: 1998, PASP 100, 949} \rf{Welch, D.L., McAlary, C.W., Mclaren, R.A., Madore, B.F.: 1986, ApJ. 305, 583 } \rf{} \abe \addresses \rf{P. Lanoix, CRAL Observatoire de Lyon, F69230 Saint-Genis Laval, FRANCE, <EMAIL>} \rf{R. Garnier, CRAL Observatoire de Lyon, F69230 Saint-Genis Laval, FRANCE, <EMAIL>} \end{document}
\section{Introduction} Neutrino experiments play a key role in establishing the validity of the electroweak Standard Model. Even today with the large samples of $W$ and $Z$ boson events, precision $\nu N$ experiments are still competitive in determining the electroweak model parameters and also show that the model is valid over many orders of magnitude in $q^2$. To date, neutrino experiments have used the Llewellyn Smith\cite{smith} relationship to determine $\sin^2 \theta_W$.\cite{CCFR,others} However, the dominate systematic error, the experimental determination of the effective charm mass, has limited the precision of these measurements. Paschos and Wolfenstein\cite{paschos} have shown that if the neutral and charged current cross sections for neutrinos and anti-neutrinos could be measured separately, then the ratio \begin{eqnarray} R^- &=& \frac{\sigma(\nu_\mu N \rightarrow \nu_\mu X)-\sigma({\overline{\nu}}_\mu N \rightarrow {\overline{\nu}} X)} {\sigma(\nu_\mu N \rightarrow \mu^- X)-\sigma({\overline{\nu}}_\mu N \rightarrow \mu^+ X)} \nonumber \\ &=& \frac{R^\nu - r R^{\overline{\nu}}}{1-r} = \frac{1}{2} - \sin^2\theta_W =g_L^2-g_R^2 \label{eq:rm} \end{eqnarray} (where $R^\nu = \frac{\sigma(\nu_\mu N \rightarrow \nu_\mu X)}{\sigma(\nu_\mu N \rightarrow \mu^- X)}$, $R^{\overline{\nu}} = \frac{\sigma({\overline{\nu}}_\mu N \rightarrow \nu_\mu X)}{\sigma({\overline{\nu}}_\mu N \rightarrow \mu^+ X)}$, $r = \frac{\sigma({\overline{\nu}}_\mu N \rightarrow \mu^+ X)}{\sigma(\nu_\mu N \rightarrow \mu^- X)}$, $g_{L,R}^2 = u^2_{L,R}+d^2_{L,R}=$ the sum of the squares of the quark couplings, and the relationship to $\sin^2\theta_W$ is a tree level relationship only) eliminates the effect of the sea quarks on the determination of $\sin^2 \theta_W$ and, hence, the majority of the error from the mass of the charm quark. To make such a measurement was a goal of the NuTeV experiment at Fermilab. \section{NuTeV experiment} In order to measure $R^-$, the cross sections for $\nu N$ and ${\overline{\nu}} N$ scattering must be measured separately. For NuTeV, we designed and built a Sign Selected Quadrapole Train (SSQT) which selected the charge of the mesons that were directed into our decay volume, and, thus, selected whether neutrinos or anti-neutrinos hit the detector. The fluxes of neutrino varieties in neutrino and anti-neutrino modes are shown in Fig.~\ref{fig:lengthandflux}(a). The wrong sign contamination fraction was less than $1\times 10^{-3}$ in neutrino mode and $2\times10^{-3}$ in anti-neutrino mode. The electron neutrino contamination in the beam is a serious background in the determination of $R^-$. The majority of $\nu_e$'s come from charged kaon decays which can be kinematically related to the high energy $\nu_\mu$ spectrum that comes from the same particles. The design of the SSQT greatly reduced the contribution of the neutral kaon decays to the $\nu_e$ flux in comparison to other experiments. \begin{figure}[htb] \centerline{ \epsfxsize=2.25in \epsfbox{showy_nutev_flux.eps} \hfill \epsfxsize=2.25in \epsfbox{fancy_length.eps} } \centerline{\hspace*{1in}(a)\hspace*{1in} \hfill \hspace*{1in}(b)\hspace*{1in}} \caption{(a) {\em Preliminary} NuTeV Neutrino and Anti-Neutrino Fluxes; (b) {\em Preliminary} NuTeV Event Length Distribution. \label{fig:lengthandflux}} \end{figure} The NuTeV experiment used the refurbished CCFR iron calorimeter and toroid spectrometer that is described elsewhere.\cite{NIM} Data were collected during the '96-'97 Fermilab fixed target run. Charged current $\nu$ and ${\overline{\nu}}$ events were differentiated from neutral current events by the length in the calorimeter over which energy was deposited. Long events indicate the presence of a muon and are classified as charged current events; neutral current events only contain a hadronic shower and are therefore short. The length distribution for neutrino and anti-neutrino events is shown in Fig.~\ref{fig:lengthandflux}(b). A detailed Monte Carlo program was used to determine the contamination of neutral current events in the long distribution and of charged current events in the short distribution. The detector parameters used in this program were measured using a test beam which constantly monitored and calibrated the detector. \section{Results} Rather than measuring $R^-$ directly, we chose to measure a ``pseudo'' $R^-$ \begin{equation} {\rm pseudo}R^- = R^\nu - \alpha R^{\overline{\nu}} \label{eq:pseudoR} \end{equation} where $\alpha = 0.5136$ was chosen to minimize the effects of the charmed quark mass on our result given the differences in fluxes and cross sections between the $\nu$ and ${\overline{\nu}}$ modes. We varied the value of $\sin^2\theta_W$ in our Monte Carlo until the predicted value of pseudo$R^-$ matched the measured one. The preliminary result from the NuTeV data is \begin{eqnarray} &&\sin^2{\theta_W}^{\rm (on-shell)} = 0.2253 \pm 0.0019{\rm (stat.)} \pm 0.0010 {\rm (sys.)} \nonumber \\ &&-0.00142\times\left(\frac{M_{\rm top}^2 - (175~{\rm GeV/c}^2)^2} {(100 {\rm~GeV/c}^2)^2} \right) +0.00048 \times \log_e \left(\frac{{M_{\rm Higgs}}^2}{150 {\rm~GeV/c}^2} \right)\!.~~~ \label{eq:result} \end{eqnarray} The small residual dependence of our result on $M_{\rm top}$ and $M_{\rm Higgs}$ comes from the leading terms in the electroweak radiative corrections\cite{radiative}. The sources of the statistical and systematic errors for this result are given in Table~\ref{tab:errors}. \begin{table}[t] \caption{ Sources of uncertainties in the determination of $\sin^2\theta_W$. \label{tab:errors}} \vspace{0.2cm} \begin{center} \footnotesize \begin{tabular}{|r|c|} \hline SOURCE OF UNCERTAINTY & $\delta\sin^2\theta_W$ \\ \hline \hline { {\em Statistics}: \hfill {Data} } & {0.00188} \\ { Monte Carlo } & {0.00028} \\ \hline \hline { \bf TOTAL STATISTICS \hfill } & {0.00190} \\ \hline\hline { {\em $\nu_e/{\overline{\nu}}_e$ Flux}: \hfill $K^\pm$ ($1.1\%$) } & {0.00024} \\ { Other sources of $\nu_e$'s } & {0.00048} \\ { {\em Energy Measurement}: \hfill Calibrations ($0.5\%$)} & {0.00043} \\ { Muon Energy Deposition ($3\%$)} & {0.00004} \\ { Energy Resolution } & {0.00004} \\ { {\em Event Length}: \hfill Hadron Shower } & {0.00015} \\ { Longitudinal Vertex Determination } & {0.00015} \\ { Counter Edge Location } & {0.00010} \\ { Counter Efficiency and Noise } & {0.00016} \\ \hline { \bf TOTAL EXP. SYST. \hfill } & {0.00075} \\ \hline\hline { {\em Sea Quarks}: \hfill Strange Sea } & {0.00034} \\ { $V_{cd}$ } & {0.00004} \\ { Charm Sea } & {0.00009} \\ { Charm Mass} & {0.00009} \\ { {\em Other $\nu/{\overline{\nu}}$ Cross-Section Differences}: \hfill $\sigma^\nu/\sigma^{{\overline{\nu}}}$ } & {0.00023} \\ { Non-Isoscalar Target } & {0.00013} \\ { Radiative Corrections } & {0.00051} \\ { {\em Non-QPM Cross-Section}: \hfill Higher Twist } & {0.00013} \\ { Longitudinal Structure Function } & {0.00007} \\ \hline { \bf TOTAL PHYSICS MODEL \hfill } & {0.00070} \\ \hline \hline { \bf TOTAL UNCERTAINTY \hfill } & {0.0022} \\ \hline \end{tabular} \end{center} \end{table} Since $\sin^2{\theta_W}^{\rm (on-shell)} = 1-M_W^2/M_Z^2$, the result given in Eq.~\ref{eq:result} is equivalent to \begin{eqnarray} M_W &&= 80.26 \pm 0.10({\rm stat.}) \pm 0.05({\rm sys.}) \nonumber \\ &&+0.073\times\left(\frac{M_{\rm top}^2 - (175~{\rm GeV/c}^2)^2} {(100 {\rm~GeV/c}^2)^2} \right) -0.025 \times \log_e \left(\frac{{M_{\rm Higgs}}^2}{150 {\rm~GeV/c}^2} \right)\!.~~~ \label{eq:mw} \end{eqnarray} This result is compared with other measurements of the $W$ mass in Fig.~\ref{fig:newmwandcouplings}(a). \begin{figure}[htb] \centerline{ \hfill \epsfxsize=1.7in \epsfbox[50 40 460 540]{b_and_w_mw.eps} \hfill \epsfxsize=1.7in \epsfbox[50 40 460 540]{couplings_new.eps} \hfill } \centerline{\hfill\hspace*{.75in}(a)\hspace*{.75in}\hfill\hspace*{.75in} (b)\hspace*{.75in}\hfill} \caption{(a) Comparison of the measured $M_W$ mass from NuTeV and other experiments; (b) Comparison of the model independent couplings between NuTeV and CCFR (Point is the '98 Standard Model fit of world data by the Particle Data Group$^7$). \label{fig:newmwandcouplings}} \end{figure} It is possible to extract the NuTeV result in a model independent framework in terms of the left and right handed quark couplings. For our data \begin{eqnarray} 0.4530 - \sin^2\theta_W &=& 0.2277\pm 0.0022 \nonumber \\ &=& 0.8587 u_L^2 + 0.8828 d_L^2 -1.1657 u_R^2 -1.2288 d_R^2. \label{eq:coupling} \end{eqnarray} (Note the similarity of this with Eq.~\ref{eq:rm}.) This result is plotted on the $g^2_L$-$g^2_R$ plane in Fig.~\ref{fig:newmwandcouplings}(b) and compared with the CCFR result which measured $R^\nu = g^2_L + 0.64 g^2_R$. The results of the two experiments are consistent with the Standard Model. \section*{Acknowledgments} We gratefully acknowledge the substantial contributions in the construction and operation of the NuTeV beamlines and the refurbishment of the NuTeV detector made by the staff of the Fermilab Beams and Particle Physics Division. This work was supported by grants from the U.S. Department of Energy, from the National Science Foundation, and from the Alfred P. Sloan Foundation. \section*{References}
\section{Introduction} Observed profiles of different emission lines often show intriguing differences in width and relative velocity. In some objects, there is a rather strong correlation between line width and critical density or ionization potential (De Robertis \&\ Osterbrock 1986). These trends have led to the concept of a density gradient across the line emitting regions, with the higher critical density (or higher ionization) lines being emitted closer to the continuum source. The underlying assumption is that the velocity dispersion decreases with increasing distance from the source of ionizing photons, as expected for gas whose motion is predominantly rotational or virialized. Variations in line width complicate attempts to separate the broad (BLR) and narrow (NLR) line contributions in a source. The frequent observation of a prominent component with FWHM$\sim$ 1600 km s$^{-1}$ in the {\sc{Civ}}$\lambda$1549\/\ (hereafter CIV) profile complicates the interpretation of high ionization line (HIL) properties in particular. In an attempt to take into account the diverse widths of emission lines, some workers have proposed the existence of an ''Intermediate Line Region'' (ILR) (see e.g. Wills et al. 1993). This emission component has been routinely treated as part of the BLR emission in line profile studies either: a) because it is believed to be part of the non-NLR gas or b) because a clear inflection between the emitting components is not often seen in the CIV profile making it difficult to subtract (see e.g. Baldwin 1997, p. 92). In Section 2 we discuss three observational results: 1) an unambiguous CIV NLR component with FWHM similar to [OIII]$\lambda$5007 is seen in many Seyfert 2 galaxies, 2) a similar component is also identified in some broad line AGN (with and without an inflection) although the component is often broader with mean FWHM$\sim$ 900 km s$^{-1}$ broader than FWHM H$\beta$ ($\approx$ FWHM [OIII]$\lambda$5007) and 3) except for the broader profile width the ``narrower'' CIV component shows properties very similar to those ascribed to the classical NLR. Section 3 presents model calculations to show that the diversity of narrow line profile widths can be ascribed to a smooth gradient of physical properties in the NLR, and that a larger width, $\sim$ 1000--2000 km~s$^{-1}$, is expected for CIV without invoking a new (ILR) emitting region. Finally, in Section 4 we argue: a) that introduction of a new emitting region is unjustified and b) that BLR properties inferred from the CIV line are strongly dependent on whether or not a narrow component is subtracted. \section{An Unambiguous CIV NLR is Observed in Some AGN \label{detect}} Most of the UV spectroscopic data for Seyfert 2 sources comes from the IUE archive. While the resolution of this data is limited, it shows that the majority of Seyfert 2 galaxies exhibit a narrow line component of CIV. We assume, following unification ideas, that BLR CIV and {\sc{H}}$\beta$\/\ are obscured in Seyfert 2's and that we see only NLR emission, the same NLR in Seyfert 1 and 2 sources. Study of averaged spectra for a sample of twenty mostly ``pure'' Seyfert 2 sources (Heckmann et al. 1995) show a significant narrow CIV line. Detailed study of IUE spectra for NGC 7674 and I Zw 92 (Kraemer et al. 1994) show an NLR line intensity ratio CIV/{\sc{H}}$\beta$\/ $\approx$ 2.0. NGC 1068 and 5643 are the only Seyfert 2 sources in the HST archive with high s/n FOS spectra of the CIV region. NGC 1068 is the brightest example of a Seyfert 2 with a confirmed hidden BLR (Antonucci \& Miller 1985). Recent study of the FOS spectra for NGC 1068 (Kraemer, Ruiz \&\ Crenshaw 1998) show that it contains a strong {\sc{Civ}}$\lambda$1549\/\ NLR component (partially resolved into $\lambda\lambda$1548/1551 components) with FWHM$\approx$ 1100 $\pm$ 200 km~s$^{-1}$\ . The narrow component of H$\beta$, and {\sc [Oiii]}$\lambda$5007\, show profile FWHM $\approx$ 1000 $\pm$ 200 km~s$^{-1}$ (Caganoff et al. 1991). Clear evidence for NLR CIV (FWHM=1180 km~s$^{-1}$) is also seen in a bright Seyfert 1.5 galaxy, NGC 5548 (Goad \&\ Koratkar 1998). The NLR component in NGC 5548 remained constant while the BLR component went through large changes, reenforcing the idea that independent NLR and BLR components exist in that source. The above cited observations make it clear that NLR CIV exists in many lower luminosity AGN. It is unambiguous in Seyfert 2 sources but can be very confusing in sources that show BLR emission because a clear NLR/BLR profile inflection is only occasionally observed. It has been argued (Wills et al. 1993) that higher luminosity sources do not show an NLR component. This argument arose from profile analysis of a small sample of high luminosity radio loud sources (Wills et al. 1993). The NLR was concluded to be weak or absent in these sources because no significant profile component that matched the width of [OIII]$\lambda$5007 was found. This interpretation was widely accepted, in part, because reddening was assumed to be high in the NLR (E(B-V)$\sim$0.5). The above cited data suggests: 1) that the reddening is frequently not so high (Kraemer, Ruiz \& Crenshaw 1998 derive a nuclear reddening of only E(B-V) =0.02 for NGC1068) and/or 2) that CIV is amplified by some physical process (see section 3). The current interpretation of narrow line spectra in NGC 1068 and 5548 points toward relatively high density and ionization parameter at the inner edge of the NLR. A value of $\Gamma \sim 10^{-1.5}$ accounts for the strength of several high ionization lines (Kraemer et al. 1998). The dereddened NLR intensity ratio {\sc{Civ}}$\lambda$1549\//H$\beta$ is $\approx$ 14 for NGC 5548, and an inner NLR component with {\sc{Civ}}$\lambda$1549\//H$\beta$$\approx$ 30 appears to be necessary. Recently, Ferguson et al. (1997) interpreted the strength of the strongest coronal lines observed in the spectra of Seyfert galaxies in terms of emission from photoionized gas in a rather dense region which could also be associated with the innermost NLR. Our own line profile survey (Marziani et al. 1996: MS) of a heterogeneous sample of 52 AGN found examples of both high and low luminosity sources with a narrow CIV component comparable in width to [OIII]$\lambda$5007. However a considerable number of sources showed a narrow component of CIV (and [OIII]$\lambda$4363, when it could be resolved from H$\gamma$) that was broader than [OIII]$\lambda$5007 by as much as 1500 km s$^{-1}$. Figure 1 illustrates the distributions of broad and narrow line widths found for CIV and H$\beta$ in the MS survey. The simplest interpretation is that both CIV NLR and BLR are broader than the corresponding regions in H$\beta$. Additional support for our NLR interpretation of the narrower CIV component comes from observations of sources where the NLR is resolved. In the case of the Seyfert 2 galaxy NGC 5643, the CIV NLR feature is obviously broader, even in the extended emission line region (EELR: FWHM=1800$\pm$200 km~s$^{-1}$), than either NLR HeII$\lambda$1640 or NLR {\sc{Ciii]}}$\lambda$1909\/. PG2308+098 is an example of a higher luminosity (z=0.43) source from the MS survey that shows a clear inflection between the broad and, one of the broadest observed, narrow CIV components. We derived a FWHM $\approx$ 2630 km~s$^{-1}$\ for the narrow CIV component considerably broader than FWHM $\approx$ 700$\pm$100 km~s$^{-1}$\ measured for {\sc{H}}$\beta$\/\ and {\sc [Oiii]}$\lambda$5007\ in that source. Figure 2 shows a comparison between the CIV feature in the NGC 5643 EELR and the narrow CIV component derived for PG 2308+098. Figure 2 clearly shows that the width of the two narrow features are equal within the observational errors. Narrow component widths as broad as the one observed in PG2308+098 are clearly not unprecedented in unambiguous NLR emitting regions. \section{NLR Models: A Broader and Stronger {\sc{Civ}}$\lambda$1549\/\ Line} We calculated photoionization models of the NLR and computed the emergent spectrum using {\tt CLOUDY} (Ferland 1996). Our goal was to account for FWHM differences as large as $\rm \Delta v_r \simlt 2000$ km~s$^{-1}$ (mean difference $\Delta$V= 900 km s $^{-1}$) between {\sc{H}}$\beta$\/\ and CIV. We did not attempt sophisticated model innovations nor did we try to fit the line profiles of a particular source. We attempted instead to illustrate the implications of recent robust results involving the input to photoionization calculations for the {\sc{Civ}}$\lambda$1549\/\ and {\sc{H}}$\beta$\/\ line profiles when the NLR line emission is not spatially resolved (presently the case for the vast majority of AGN). At z $\approx$ 0.2, a 0.''25 aperture (typically employed for FOS observations) covers 1 kpc of spatial extent: thus most of the NLR should be included in a measurement. Input parameters for our models are listed in Table 1. We assumed spherical symmetry with a radial density gradient described by a power-law, $\rm n_e = n_0 r_{pc}^{-s}$ cm$^{-3}$\/\ from 1-10$^3$ pc. The density at the outer edge of the NLR was set equal to $ 10^3$ cm$^{-3}$\/\ in all cases. We assumed a power-law dependence for the covering factor, i. e. $\rm f_{c,r} \propto r^{-m}$, with total NLR covering factor 0.1. The first three models listed in Table 1 follow Netzer (1990) with a column density $\rm N_c \propto r^{-2/3s}$. The last four models assumed that $\rm N_c \approx 10^{23}$ cm$^{-2}$ = constant throughout the NLR (Column 3 lists adopted values of m). The inclusion of the effects of appreciable mechanical heating appear to be necessary for realistic modeling of the NLR. HST narrow-band images and radio maps have show a close relationship between radio ejecta and NLR emitting plasma in several nearby Seyfert galaxies (Capetti et al. 1999; Falcke et al. 1998). A crude estimate of the heating rate $\epsilon$\ can be obtained by assuming that a fraction $\rm f_P$\ $\approx 0.1$ (approximately equal to the covering factor $\rm f_c$) of the jet energy flux $\Pi$\ is converted into heating of the NLR gas. Assuming an exponential radial dependence of the heating rate we can write $\rm \epsilon = \epsilon_0 f_{c,0.1} R_{out,1 kpc}^{-3} \Pi_{45} \exp(-(r-r_0)/r_s) $ergs s$^{-1}$\/\ cm$^{-3}$\/, where we have parameterized the dependence on outer radius (R$\rm _{out}$), $\rm f_c$, and $\Pi$ (Column 4 lists the scale factor). The values $\rm r_S = 1$ and 10pc correspond, respectively, to a heating rate per unit volume at 1 pc $\log \epsilon_0 \approx -9.38$ (which is neglible) and $\approx -6.37$ (which provides significant enhancement of electron temperature, in a cocoon within 2 pc). The latter cases is also more realistic because recent hydrodynamic simulations show that most of the line luminosity is concentrated in dense clouds surrounding the jet (Steffen et al. 1997). However, recent observations (Villar-Martin et al. 1999) of high redshift radio-loud and quiet sources reveal FWHM$\simgt$ 1000 km~s$^{-1}$ emission lines in extended gas both near and far from radio jets. The jet energy flux is a poorly known quantity, and its actual relevance depends on the details of the jet/cloud interaction. Significant heating of clouds can however be expected for a wide range of AGN including radio quiet objects like Seyferts (in agreement with the analysis of Bicknell et al. 1998 for Seyfert 2 galaxies). The gas is assumed to move with velocity $\rm v(r) = (2100 r^{-1/2} + 0.1 r)$\ i.e., at the local virial velocity plus a component due to galactic rotation. A velocity dispersion $\rm \Delta v \approx 4150 r^{-0.5}_{1~pc} M^{0.5}_9$ km~s$^{-1}$\ is expected at a distance r equal to 1 pc (which we assume as the inner edge of the NLR; 1 pc $\approx$ 10$^4$ gravitational radii for a $10^9$ M$_\odot$\/\ black hole), if the gas motion is rotational or virial. We focus our profile predictions on four NLR lines: CIV, {\sc{H}}$\beta$\/, {\sc [Oiii]}$\lambda$5007, and {\sc [Fe vii]}$\lambda$6087\/. The widths reported in Table 1 (Cols. 5--8) are the weighted average of velocity over line emissivity per unit area $\rm \zeta_r$ (i.e., $\rm < v > = {\int v_r f_{c,r} \zeta_r dV}/{\int f_{c,r} \zeta_r dV}$). Observed NLR linewidth for {\sc [Oiii]}$\lambda$5007, H$\beta$, and {\sc [Fe vii]}$\lambda$6087\/\ are reproduced by the models (FWHM({\sc [Oiii]}$\lambda$5007) $\approx$ FWHM(H$\beta$) $<$ FWHM({\sc [Fe vii]}$\lambda$6087\/) $\sim$ 1000 km~s$^{-1}$), with models 5 and 7 a slightly larger with of {\sc [Oiii]}$\lambda$5007\ with respect to {\sc{H}}$\beta$\/, as it is often observed. The most relevant result is that $< v >$\ for {\sc{Civ}}$\lambda$1549\/\ appears to be significantly larger than that of any other line considered, with $< v > \sim 1000-2500$km~s$^{-1}$$ \simgt$ 1000 km~s$^{-1}$\ in all cases. {\sc{Civ}}$\lambda$1549\/\ emission is favored by (a) the larger ionization parameter in the innermost NLR, and (b) mechanical heating. A more realistic case might involve cylindrical symmetry, for example, a cylindrical NLR volume with radius 1 kpc and total height 3.4 kpc (the values appropriate for Mark 50; Falcke et al. 1998), where the axis of symmetry corresponds to the radio axis. No substantial difference from the spherically symmetric models is expected as far as $< v >$\ is concerned, if similar assumptions about gas motions are made. However, the cylindrical symmetry case may give: (a) a larger {\sc{Civ}}$\lambda$1549\/\ flux with respect to the spherically symmetric case and (b) a possibly broader {\sc{Civ}}$\lambda$1549\/\ due to the pressure exerted by the jet, which could accelerate dense gas clumps. Mechanical heating may also provide a suitable mechanism for dust destruction. More sophisticated modelling (e. g. Steffen et al. 1997; Bicknell et al. 1998) is beyond the scope of this paper. \section{Results} \subsection{The Intermediate Line Region: is it pr\ae ter necessitatem? \label{ilr}} Introduction of an additional (e.g. ILR) emitting region can be justified by the following (not mutually exclusive) circumstances: 1) spatial segregation (e.g. region associated with a jet) and/or 2) an abrupt discontinuity in physical properties (e.g. optical thickness or cloud instability at some distance). This is the physical statement of Occam's razor ({\it multiplicanda non sunt pr\ae ter necessitatem}). The above circumstances describe the justification for the concept of distinct BLR and NLR regions where suppression of collisional lines yield different spectra. However ``ILR'' emission and coronal lines ($<v>\approx 1-2\times10^3$ km~s$^{-1}$) can be produced in the inner part of the NLR which, even if associated with an inner sheet adjacent to radio plasma, does not require introduction of a new distinct and discontinuous emitting region. Part of the confusion over the identification of different line emitting components may have arisen from a difference in nomenclature. Wills et al. (1995) and Brotherton et al. (1994a,b) refer to the broader and narrower CIV components as very broad line (VBLR) and intermediate line (ILR) respectively. However they also inferred (Brotherton et al. 1994b) that the ILR properties (reverberation distance from continuum source, density, covering factor) are all similar to the standard NLR values. The ILR may be an ``inner part of the NLR'' but it is NLR and not BLR gas. Our models strengthen this view. The VBLR designation is also confusing because this name has already been applied to an even broader component sometimes seen in the HeII$\lambda$4686 feature (Ferland et al. 1990; Marziani \&\ Sulentic 1993). \subsection{BLR Implications} The model calculations support our two-component NLR-BLR interpretation of the CIV and {\sc{H}}$\beta$\/\ profiles. Of course greater broad line profile complexity is sometimes observed, with double-peaked Balmer line profiles an extreme example. The relative strength and frequency of occurence of multiple BLR components is not yet well established. However one interprets the narrower CIV feature, there is no justification for treating it as BLR gas. A number of recent studies involving the CIV BLR have analysed profile data without the benefit of subtracting any NLR component (e.g. Wills et al. 1993; Brotherton et al. 1994ab; Corbin \& Francis 1994; Corbin \& Boroson 1996). Does it matter if an NLR component is subtracted from CIV? Our results suggest that the answer is a strong ``yes''. The NLR component subtracted from CIV (in MS) represented 2-20\% of the total flux in most cases. This flux was concentrated in the narrow peak of the line and therefore affected disproportionately determinations of the half maximum level of the broad line profile. The result of nonsubtraction yields line parameters surprisingly different from those obtained by MS. Figure 3 shows a comparison of CIV centroid line shifts (relative to [OIII]5007) for sources in common between MS and Corbin \& Boroson (1996). A similar comparison also reveals that the CIV BLR profile width (FWHM) is measured, on average, 40\%\ smaller when an NLR component is not subtracted. High ionization line BLR properties inferred from CIV, arguably the least confused HIL line, are dependent on whether or not an NLR is subtracted. If an NLR component is subtracted, MS survey results suggest that: 1) radio-loud (RL) AGN show fundamentally different BLR geometry/kinematics from the radio quiet (RQ) majority, 2) in RQ--BLR CIV is blueshifted (up to 3000 km s$^{-1}$) with respect to the local rest frame while H$\beta$ show small red and blue velocity shifts (up to 600 km s$^{-1}$) about that frame and 3) In RL--BLR CIV shows zero, or small red and blue (up to 1000 km s$^{-1}$), velocity excursions relative to the rest frame while H$\beta$ shows larger predominantly red (but sometimes blue) shifts (up to 3000 km s$^{-1}$) relative to that frame. These effects must be tested with samples that cover a larger source luminosity range. \acknowledgements PM acknowledges financial support by Italian Ministry for University and Research (MURST) under grand Cofin98-02-32. \newpage \section{References} \par\noindent\hangindent 20pt Antonucci, R. and Miller, J. 1985, ApJ, 297, 621 \par\noindent\hangindent 20pt Baldwin, J. 1998, In proceedings of Emission Lines in Active Galaxies, eds. B. Peterson et al., ASP Conf. Ser., 113, 80 \par\noindent\hangindent 20pt Bicknell, G. V., Dopita, M. A., Tsvetanov, Z. I., \&\ Sutherland, R. S. 1998, ApJ, 495, 680 \par\noindent\hangindent 20pt Brotherton, M. S., Wills, B. J., Steidel, Ch., \&\ Sargent W. L. W., 1994a, ApJ, 423, 131 \par\noindent\hangindent 20pt Brotherton, M. S., Wills, B. J., Francis, P. J., \&\ Steidel, Ch. C. 1994b, ApJ, 430, 495 \par\noindent\hangindent 20pt Caganoff, S., et al. 1991, ApJ 377, 9 \par\noindent\hangindent 20pt Capetti, A., et al. 1999, MemSAIt, in press \par\noindent\hangindent 20pt Corbin, M. \&\ Francis, P. 1994, AJ, 108, 2016 \par\noindent\hangindent 20pt Corbin M., R., \&\ Boroson, T. 1996, ApJ Suppl, 107, 69 \par\noindent\hangindent 20pt De Robertis, M. M., \&\ Osterbrock, D. E. 1986, ApJ, 301, 727 \par\noindent\hangindent 20pt Falcke, H., Wilson, A. and Simpson, C. 1998, ApJ, 502, 199 \par\noindent\hangindent 20pt Ferguson, J., Korista, K. and Ferland, G. 1997, ApJ Suppl, 110, 287 \par\noindent\hangindent 20pt Ferland, G., Korista, K. and Peterson, B. 1990, ApJ, 363, L21 \par\noindent\hangindent 20pt Ferland, G. J. 1996, Hazy: a Brief Introduction to CLOUDY, Univ. Kentucky, Dept. Phys. \&\ Astron., internal report \par\noindent\hangindent 20pt Goad, M., \&\ Koratkar, A. 1998, ApJ 495, 718 \par\noindent\hangindent 20pt Heckmann, T. et al. 1995, ApJ, 452, 549 \par\noindent\hangindent 20pt Kraemer, S., et al. 1994, ApJ, 435, 171 \par\noindent\hangindent 20pt Kraemer, S., Crenshaw, D., Filippenko, A. and Peterson, B. 1998, ApJ, 499, 719 \par\noindent\hangindent 20pt Kraemer, S., Ruiz, J. and Crenshaw, M. 1998, ApJ, 508, 232 \par\noindent\hangindent 20pt Marziani, P. \&\ Sulentic, J. W. 1993, ApJ, 409, 612 \par\noindent\hangindent 20pt Marziani, P., Sulentic, J. W., Dultzin-Hacyan, D., Calvani, M. \&\ Moles, M. 1996, ApJ Suppl, 104, 37 (MS) \par\noindent\hangindent 20pt Netzer, H., 1990, in Active Galactic Nuclei, Saas Fee Advanced Course 20, R.D. Blandford, H. Netzer, L. Woltjer (Berlin:Springer), p. 57 \par\noindent\hangindent 20pt Steffen, W., Gomez, J. L., Raga, A. C., \&\ Williams, R. J. R. 1997, ApJ 491, L73 \par\noindent\hangindent 20pt Villar-Martin, M., Binette, L. and Fosbury, R. 1999, A\&A, in press \par\noindent\hangindent 20pt Wills, B. et al. 1993, ApJ, 415, 563 \par\noindent\hangindent 20pt Wills, B. J., Thompson, K. L., Han, M., Netzer, H., Wills, D., Baldwin, J. A., Ferland, G. J., Browne, I. W. A., \&\ Brotherton, M. S. 1995, ApJ 447, 139 \newpage \begin{deluxetable}{cccccccc} \tablewidth{42pc} \tablecaption{Model Computations \label{contacci}} \tablehead{ \colhead{Model} & \colhead{$\rm \log(n_0)$} & \colhead{m} & \colhead{$\rm r_s$\tablenotemark{1}} & \colhead{$< v >$} & \colhead{$< v > $} & \colhead{$< v>$} & \colhead{$< v > $ } \\ & & & &\colhead{{\sc{H}}$\beta_{\rm NC}$\/} & \colhead{{\sc [Oiii]}$\lambda$5007} & \colhead{{\sc [Fe vii]}$\lambda$6087\/} & \colhead{{\sc{Civ}}$\lambda$1549$_{\rm NC}$\/\tablenotemark{2}}\\ & \colhead{(cm$^{-3}$\/)} & & (pc) & \colhead{(km~s$^{-1}$)} & \colhead{(km~s$^{-1}$)} & \colhead{(km~s$^{-1}$)} & \colhead{(km~s$^{-1}$)}\\ \colhead{(1)} & \colhead{(2)} & \colhead{(3)} & \colhead{(4)} & \colhead{(5)} & \colhead{(6)} & \colhead{(7)} & \colhead{(8)} } \startdata 1 & 8 & $\frac{7}{18}$ & 10 & 490 & 450 & 1030 & 1370 \\ 2 & 7 & $\frac{11}{18}$ & 10 & 420 & 310 & 950 & 1250\\ 3 & 8 & $\frac{7}{18}$ & 1 & 720 & 440 & 1160 & 2680 \\ 4 & 8 & $\frac{3}{2}$ & 10 & 370 & 340 & 980 & 1380 \\ 5 & 7 & $\frac{3}{2}$ & 10 & 370 & 490 & 910 & 1210 \\ 6 & 8 & $\frac{3}{2}$ & 1 & 380 & 330& 1020 & 1450 \\ 7 & 7 & $\frac{3}{2}$ & 1 & 360 & 520 & 925 & 1950 \\ \enddata \tablenotetext{1}{Scale factor for mechanical heating, assumed to decrease exponentially from $\rm r_0$ ~~(see text).} \tablenotetext{2}{Width of single line.} \end{deluxetable} \newpage \begin{figure} \plotone{fig01.eps} \caption[1]{FWHM distributions for the broad and narrow components of CIV and H$\beta$. Data from the MS survey (Marziani et al. 1996).} \end{figure} \begin{figure} \plotone{fig02.eps} \caption[1]{CIV$\lambda$1549 profiles for: i) PG 2308+09 (upper) - smooth line shows fits to the BLR and continuum components and ii) NGC 5643 EELR (lower) - scaled and without continuum subtraction. NLR component of PG2308+09 superposed (dotted). Note the similar profile widths.} \end{figure} \begin{figure} \plotone{fig03.eps} \caption[1]{CIV BLR profile shift measurements (at FWHM level) for sources in common between the MS survey and Corbin \&\ Boroson (1996) (MC). Dashed line is an unweighted least squares fit and the thin line the locus of equal shifts. MC measures are consistent with no shift within the errors ($\pm 200$ km~s$^{-1}$).} \end{figure} \end{document}
\section{The geometric set-up} Let $\Sigma_n\subset S^{[n]}\times S$ be the universal family of subschemes parameterized by $S^{[n]}$, and let $I_n\subset {\cal O}_{S^{[n]}\times S}$ and ${\cal O}_n$ denote its ideal sheaf and structure sheaf, respectively. \beeq{dispA} \begin{array}{ccccc} \Sigma_n&\subset&S^{[n]}\times S&\stackrel{q}{\longrightarrow}&S\\ &&\scriptstyle{p}\Big\downarrow\phantom{\scriptstyle{p}}\\ &&S^{[n]}. \end{array} \end{eqnarray} The induction step involves the incidence variety $S^{[n,n+1]}$ of all pairs $(\xi,\xi')\in S^{[n]}\times S^{[n+1]}$ satisfying $\xi\subset \xi'$. If $\xi'$ is obtained by extending $\xi$ at the closed point $x\in S$, there is an exact sequence \beeq{version1} 0\longrightarrow{I_{\xi'}}\longrightarrow{I_\xi}\stackrel{\lambda}{\longrightarrow}k(x)\lra0. \end{eqnarray} Observe that $\lambda$ defines a point in the fibre of the morphism $$\sigma=(\phi,\rho):\IP(I_n)\longrightarrow S^{[n]}\times S$$ over $(\xi,x)$. (Here for any coherent sheaf $F$ we denote $\IP(F):=Proj(Sym^*(F))$, and ${\cal O}_F(1)$ is the tautological quotient line bundle.) Conversely, any point in the fibre of $\sigma$ defines an extension $\xi'$. In fact, let ${\cal L}:={\cal O}_{I_n}(1)$, let $\Gamma\subset \IP(I_n)\times S$ be the graph of $\rho$, and let $j:=({\rm id},\rho)$. Then the kernel of the composite epimorphism $$\beta:(\phi\times{\rm id}_S)^*I_n\longrightarrow (\phi\times{\rm id}_S)^*I_n|_{\Gamma} \cong j_*\sigma^*I_n\longrightarrow j_*{\cal L}$$ is an $\IP(I_n)$-flat family of ideal sheaves and induces a classifying morphism $$\psi:\IP(I_n)\longrightarrow S^{[n+1]},$$ such that $Ker(\beta)=\psi_S^*(I_{n+1})$. This leads to a scheme theoretic isomorphism $S^{[n,n+1]}\cong\IP(I_n)$ and the basic diagram \beeq{dispB} \begin{array}{ccccc} S&\stackrel{\rho}{\longleftarrow}&\IP(I_n)&\stackrel{\psi}{\longrightarrow}&S^{[n+1]}\\[1ex] &&\phantom{\scriptstyle{\phi}}\Big\downarrow\scriptstyle{\phi}\\[2ex] &&S^{[n]} \end{array} \end{eqnarray} together with two short exact sequences of universal families \beeq{basicseq}\ses{\psi_S^*I_{n+1}}{\phi_S^*I_n}{j_*{\cal L}} \end{eqnarray} \beeq{basicseq2}\ses{j_*{\cal L}}{\psi_S^*{\cal O}_{n+1}}{\phi_S^*{\cal O}_n}. \end{eqnarray} Note that $j_*{\cal L}=p^*{\cal L} \otimes{\cal O}_\Sigma=p^*{\cal L}\otimes \rho_S^*{\cal O}_\Delta$. Here and throughout the paper we will use the short form $f_S:=f\times {\rm id}_S$ in order to simplify the notations. Since $S$ is smooth of dimension $2$ and $I_n$ is $S^{[n]}$-flat and fibrewise torsion free, there is a resolution of length 1 $$0\longrightarrow{A}\longrightarrow{B}\longrightarrow{I_n}\longrightarrow 0$$ by locally free ${\cal O}_{S^{[n]}\times S}$-sheaves $A$ and $B$ of rank $a$ and $a+1$, respectively. Then $\IP(I_n)$ embeds into $\IP(B)$ as the zero locus of the homomorphism $$\pi^*A\longrightarrow\pi^*B\longrightarrow{\cal O}_B(1)$$ (where $\pi:\IP(B)\to S^{[n]}\times S$ is the projection) such that ${\cal L}={\cal O}_B(1)|_{\IP(I_n)}$. The incidence variety $S^{[n,n+1]}$ is irreducible \cite{EllStrIrr, CheahSmooth}. (In fact, it is also smooth, though we will not use this. This result has been independently proven by Cheah \cite{CheahSmooth}, Ellingsrud (unpublished) and Tikhomirov \cite{TikhoSmooth}.) As $S^{[n,n+1]}$ has the correct dimension, it is a locally complete intersection and its fundamental class is given by $$[\IP(I_n)]=c_a({\cal L}\otimes\pi^*A\makebox[0mm]{}^{{\scriptstyle\vee}})\in A_{2n+2}(\IP(B)).$$ \begin{lemma}--- Let $\ell=c_1({\cal L})$. Then for any class $u\in A_*(S^{[n]}\times S)$ one has \beeq{pushdownofell} \sigma_*(\ell^i)=(-1)^ic_i({\cal O}_n)=(-1)^ic_i(-I_n). \end{eqnarray} \end{lemma} \begin{proof}{} Let $\varepsilon:=c_1({\cal O}_B(1))$. Then \begin{eqnarray*} \sigma_*(\ell^i)&=&\pi_*(\varepsilon^i\cdot[\IP(I_n)])=\pi_*(\varepsilon^i c_a(\pi^*A\makebox[0mm]{}^{{\scriptstyle\vee}}\otimes{\cal O}_B(1))\\ &=&\pi_*(\sum_{k=0}^a\varepsilon^{i+a-k}\pi^*c_k(A\makebox[0mm]{}^{{\scriptstyle\vee}}))\\ &=&\sum_{k=0}^as_{i-k}(B\makebox[0mm]{}^{{\scriptstyle\vee}})c_k(A\makebox[0mm]{}^{{\scriptstyle\vee}})\\ &=&c_i(A\makebox[0mm]{}^{{\scriptstyle\vee}}-B\makebox[0mm]{}^{{\scriptstyle\vee}})=(-1)^ic_i({\cal O}_n). \end{eqnarray*} \end{proof} \begin{proposition}\label{hilbertchowvanishing}--- Let $f_n:S^{[n]}\longrightarrow S^{(n)}$ denote the Hilbert-Chow morphism from the Hilbert scheme to the symmetric product. Then $$(f_n)_*{\cal O}_{S^{[n]}}={\cal O}_{S^{(n)}}\quad\mbox{and}\quad R^i(f_n)_*{\cal O}_{S^{[n]}}=0\quad\mbox{for all }i>0.$$ \end{proposition} \begin{proof}{} As $f_n$ is a birational proper morphism of normal varieties, it follows that $(f_n)_*{\cal O}_{S^{[n]}}={\cal O}_{S^{(n)}}$. $S^{(n)}$ has rational singularities, as the quotient of a smooth variety by a finite group (see \cite{Bu}, Proposition 4.1). Therefore its resolution $f_n:S^{[n]}\to S^{(n)}$ satifies $R^i(f_n)_*{\cal O}_{S^{[n]}}=0$. \end{proof} \section{The tangent bundle and tautological bundles} For a smooth projective variety $X$ let $K(X)$ denote the Grothendieck group generated by locally free sheaves, or equivalently, arbitrary coherent sheaves. We will denote a sheaf and its class by the same symbol. $K(X)$ is endowed with a ring structure which for classes of locally free sheaves $F$ and $G$ is given by $F\cdot G:=F\otimes G$, and a ring involution ${}\makebox[0mm]{}^{{\scriptstyle\vee}}$, which extends $F\mapsto \kh om(F,{\cal O}_X)$ for a locally free sheaf. Thus for arbitrary coherent sheaves $F$, $G$ one has $F\makebox[0mm]{}^{{\scriptstyle\vee}}=\sum_i(-1)^i\ke xt^i(F,{\cal O}_X)$ and $F\makebox[0mm]{}^{{\scriptstyle\vee}}\cdot G=\sum_i(-1)^i\ke xt^i(F,G)$. For $f:X\to Y$ a morphism of smooth projective varieties there is a push-forward $f_!:K(X)\to K(Y); \ F\mapsto \sum_i (-1)^iR^if_*(F)$, and a pullback $f^!$ which for a locally free sheaf $G$ on $Y$ is given by $f^!G=f^*G$. In the paragraphs below, we will use the following well-known fact, which holds in greater generality: let $G$ be an $S$-flat family of sheaves on $Y\times S$, and let $p:Y\times S\to Y$ denote the projection. If $f:Y'\to Y$ is a morphism of schemes, then $f^!(p_!G)=p_!(f_S^!G)$. Moreover, if the support of $G$ is finite over $S$, there is an isomorphism of sheaves $f^*p_*G\cong p_*f_S^*G$. In this section we describe recursion relations for the classes of the tangent sheaf $T_n$ of $S^{[n]}$ and certain tautological sheaves $F^{[n]}$ on $S^{[n]}$ which are defined as follows: For any locally free sheaf $F$ on $S$ let $F^{[n]}:=p_*({\cal O}_n\otimes q^*F)$ (with $p$ and $q$ as in diagram (\ref{dispA})). This extends to a group homomorphism ${}^{[n]}:K(S)\to K(S^{[n]})$. We will also write $p$ and $q$ for the projections of $S^{[n,n+1]}\times S$ to $S^{[n,n+1]}$ and $S$. \begin{lemma}--- The following relation holds in $K(S^{[n,n+1]})$: \beeq{tautorecursion} \psi^!F^{[n+1]}=\phi^!F^{[n]}+{\cal L}\cdot\rho^!F. \end{eqnarray} \end{lemma} \begin{proof}{} Let $F$ be a locally free sheaf on $S$. Apply the functor $p_*(\,.\,\otimes q^*F)$ to the exact sequence (\ref{basicseq2}) and observe that $$p_*(\phi_S^*({\cal O}_n)\otimes q^*F)=p_*(\phi_S^*({\cal O}_n\otimes q^*F)) =\phi^*(p_*({\cal O}_n\otimes q^*F))=\phi^*F^{[n]},$$ i.e. $p_!(\phi_S^!({\cal O}_n)\cdot q^!F)=\phi^!F^{[n]}$ and similarly $p_!(\psi_S^!({\cal O}_{n+1})\cdot q^!F)=\psi^!F^{[n+1]}$. Using $p\circ j={\rm id}_{S^{[n,n+1]}}$ we get $$p_*(j_*({\cal L})\otimes q^*F))=(p\circ j)_*({\cal L}\otimes (q\circ j)^*F)= {\cal L}\otimes \rho^*F,$$ i.e. $p_!(j_!({\cal L})\cdot q^!F)) ={\cal L}\cdot \rho^!F$. \end{proof} Next, we turn to the tangent sheaf: \begin{proposition}\label{Tprop}--- The class of the tangent sheaf in $K(S^{[n]})$ is given by the relation \beeq{} T_n=\chi({\cal O}_S)\cdot 1- p_!(I_n\makebox[0mm]{}^{{\scriptstyle\vee}}\cdot I_n). \end{eqnarray} \end{proposition} \begin{proof}{} We have the following isomorphisms $$T_n\cong{\rm Hom}_p(I_n,{\cal O}_n)\cong{\rm Ext}^1_p({\cal O}_n,{\cal O}_n)\cong p_*\ke xt^1_{{\cal O}_{S^{[n]}\times S}}({\cal O}_n,{\cal O}_n).$$ Here ${\rm Ext}_p$ are the higher derived functors of the composite functor $p_*\circ\kh om$. For the first isomorphism see e.g.\ \cite{Lehn}. The last equality is a consequence from the spectral sequence $R^ip_*\ke xt^j \Rightarrow{\rm Ext}_p^{i+j}$ and the observation that the sheaves $\ke xt^*({\cal O}_n,{\cal O}_n)$ are supported on the universal family $\Sigma_n$, which is finite over $S^{[n]}$ so that all higher direct images vanish. Moreover, $$\ke xt^0({\cal O}_n,{\cal O}_n)\cong\ke xt^0({\cal O}_{S^{[n]}\times S},{\cal O}_n)\cong{\cal O}_n$$ and, by $\ke xt^0({\cal O}_n,{\cal O}_{S^{[n]}\times S})=0$, $\ke xt^1({\cal O}_n,{\cal O}_{S^{[n]}\times S})=0$ also $$\ke xt^2({\cal O}_n,{\cal O}_n)\cong\ke xt^2({\cal O}_n,{\cal O}_{S^{[n]}\times S})={\cal O}_n\makebox[0mm]{}^{{\scriptstyle\vee}}.$$ Hence we get \begin{eqnarray*} T_n&=&p_*\ke xt^1({\cal O}_n,{\cal O}_n)=p_!{\cal O}_n-p_!({\cal O}_n\makebox[0mm]{}^{{\scriptstyle\vee}}\cdot{\cal O}_n)+p_!{\cal O}_n\makebox[0mm]{}^{{\scriptstyle\vee}}\\ &=&p_!(1)-p_!((1-{\cal O}_n)\makebox[0mm]{}^{{\scriptstyle\vee}}(1-{\cal O}_n))=\chi({\cal O}_S)\cdot1-p_!(I_n\makebox[0mm]{}^{{\scriptstyle\vee}}\cdot I_n) \end{eqnarray*} as an identity in $K(S^{[n]})$. \end{proof} \begin{proposition}--- The following relation holds in $K(S^{[n,n+1]})$: \beeq{tangentrecursion} \psi^!T_{n+1}=\phi^!T_n+{\cal L}\cdot \sigma^! I_n\makebox[0mm]{}^{{\scriptstyle\vee}} +{\cal L}\makebox[0mm]{}^{{\scriptstyle\vee}}\cdot\sigma^!I_n\cdot\rho^!\omega_S\makebox[0mm]{}^{{\scriptstyle\vee}} -\rho^!(1-T_S+\omega_S\makebox[0mm]{}^{{\scriptstyle\vee}}). \end{eqnarray} \end{proposition} \begin{proof}{} We have \begin{eqnarray*} \psi^!T_{n+1}&=&\psi^!(\chi({\cal O}_S)\cdot 1) -\psi^!p_!(I_{n+1}\makebox[0mm]{}^{{\scriptstyle\vee}}\cdot I_{n+1})\\ &=&\phi^!(\chi({\cal O}_S)\cdot 1)-p_!(\psi_S^!I_{n+1}\makebox[0mm]{}^{{\scriptstyle\vee}}\cdot\psi_S^!I_{n+1}) \end{eqnarray*} Now use (\ref{basicseq}) to replace $\psi_S^!I_{n+1}$ by $\phi_S^!I_n-p^!{\cal L}\cdot\rho_S^!{\cal O}_\Delta$, where $\Delta\subset S\times S$ denotes the diagonal. This yields \begin{eqnarray*} \psi^!T_{n+1}&=&\phi^!T_n-\rho^!p_!({\cal O}_\Delta\makebox[0mm]{}^{{\scriptstyle\vee}}\cdot{\cal O}_\Delta)\\ &&+p_!(\rho_S^!{\cal O}_\Delta\cdot\phi_S^!I_n\makebox[0mm]{}^{{\scriptstyle\vee}})\cdot{\cal L} +p_!(\rho_S^!{\cal O}_\Delta\makebox[0mm]{}^{{\scriptstyle\vee}}\cdot\phi_S^!I_n). \end{eqnarray*} The two last summands can be simplified as follows: $$p_!(\rho_S^!{\cal O}_{\Delta}\cdot\phi_S^!I_n\makebox[0mm]{}^{{\scriptstyle\vee}})= p_!(j_!(j^!\phi_S^!I_n\makebox[0mm]{}^{{\scriptstyle\vee}}))=p_!j_!\sigma^!I_n\makebox[0mm]{}^{{\scriptstyle\vee}}=\sigma^!I_n\makebox[0mm]{}^{{\scriptstyle\vee}}$$ and similarly $p_!(\rho_S^!{\cal O}_\Delta\makebox[0mm]{}^{{\scriptstyle\vee}}\cdot\phi_S^!I_n)= \sigma^!I_n\cdot\rho^!\omega_S\makebox[0mm]{}^{{\scriptstyle\vee}}$, since $\rho_S^!{\cal O}_\Delta\makebox[0mm]{}^{{\scriptstyle\vee}}=\rho_S^!\Delta_!\omega_S\makebox[0mm]{}^{{\scriptstyle\vee}}=j_!\rho^!\omega\makebox[0mm]{}^{{\scriptstyle\vee}}$. Finally, $p_!({\cal O}_\Delta\makebox[0mm]{}^{{\scriptstyle\vee}}\cdot{\cal O}_\Delta)={\cal O}_S-T_S+\omega_S\makebox[0mm]{}^{{\scriptstyle\vee}}$, which follows e.g.\ by Proposition \ref{Tprop}. \end{proof} \section{The induction step} We want to relate integrals on $S^{[n+1]}\times S^m$ to integrals on $S^{[n]}\times S^{m+1}$. Let $Z=S^{[n,n+1]}\times S^{m}$. The maps $\phi$ and $\psi$ from diagram (\ref{dispB}) generalize to morphisms $\Psi=\psi\times{\rm id}_{S^m}:Z\to S^{[n+1]} \times S^m$ and $\Phi=\sigma\times{\rm id}_{S^m}:Z\to S^{[n]}\times S^{m+1}$. For any $I\subset \{0,1,\ldots,m\}$ let $pr_I$ denote the projection from $S^{[n+1]}\times S\times\ldots\times S$ to the product of the factors indexed by $I$. \begin{proposition}\label{technoprop}--- Let $f$ be a polynomial in the Chern classes of the following sheaves on $S^{[n+1]}\times S^m$: $$pr_0^*T_{n+1},\,\,\, pr_{0i}^*I_{n+1},\,\,\, pr_{ij}^*{\cal O}_\Delta,\,\,\, pr_i^*T_S\,\,\mbox{ for any }1\leq i,j\leq m.$$ Then there is a polynomial $\tilde f$, depending only on $f$, in the Chern classes of the analogously defined sheaves on $S^{[n]}\times S^{m+1}$ such that $$\int_{S^{[n+1]}\times S^{m}}f=\int_{S^{[n]}\times S^{m+1}}\tilde f.$$ \end{proposition} \begin{proof}{} The morphism $\Psi$ is generically finite of degree $n+1$, so that $$\int_{S^{[n+1]}\times S^{m}}f=\frac{1}{n+1}\int_Z\Psi^*f.$$ Because of an index shift resulting from the insertion of the additional factor $S$ between $S^{[n]}$ and $S^m$ we have $$\Psi^!pr_i^*T_S=\Phi^!pr_{i+1}^*T_S,\,\Psi^!pr_{ij}^*{\cal O}_\Delta=\Phi^! pr_{i+1,j+1}^*{\cal O}_\Delta.$$ Using (\ref{basicseq}) we get $$\Psi^!pr_{0i}^*I_{n+1}=\Phi^!pr_{0,i+1}^*I_n+pr_Z^!{\cal L}\cdot pr_{1,i+1}^*{\cal O}_\Delta.$$ And finally by (\ref{tangentrecursion}) \begin{eqnarray} \Psi^!pr_{0}^*T_{n+1}&=&\Phi^!pr_0^*T_n +pr_{S^{[n,n+1]}}^!{\cal L}\cdot pr_{01}^*I_n\makebox[0mm]{}^{{\scriptstyle\vee}}\\ &&+pr_{S^{[n,n+1]}}^!{\cal L}\makebox[0mm]{}^{{\scriptstyle\vee}}\cdot pr_{01}^!I_n\cdot pr_1^!\omega_S\makebox[0mm]{}^{{\scriptstyle\vee}}-pr_1^!({\cal O}_S-T_S+\omega_S\makebox[0mm]{}^{{\scriptstyle\vee}}). \end{eqnarray} It follows that there are polynomials $f_\nu$ depending only on $f$, in the Chern classes of the sheaves $$pr_0^*T_{n},\, pr_{0i}^*I_{n},\, pr_{ij}^*{\cal O}_\Delta,\, pr_i^*T_S$$ such that $$\int_{S^{[n+1]}\times S^{[m]}}f=\frac{1}{n+1}\int_Z\Psi^*f= \int_Z\big(\sum_{\nu\geq0}\Phi^*f_\nu\cdot pr_Z^*(-c_1({\cal L}))^\nu\big).$$ As we are only trying to prove a general structure result we make no attempt to derive from the above recursion relations for the classes in the $K$-groups more explicit formulae for the dependence of $f_\nu$ on $f$. Now, according to (\ref{pushdownofell}), the last integral equals: $$\int_{S^{[n]}\times S^{m+1}}\Phi_*\big(\sum_{\nu\geq0} \Phi^*f_\nu\cdot pr_Z^*(-c_1({\cal L}))^\nu\big) =\int_{S^{[n]}\times S^{m+1}}\sum_{\nu\geq0}f_\nu\cdot c_\nu(-pr_{01}^*I_n).$$ The integrand in this expression is the polynomial $\tilde f$. \end{proof} \begin{proof}{ of Proposition \ref{propB}} Suppose we are given a polynomial $P$ in the Chern classes of $T_n$. Applying the proposition repeatedly, we may write $$\int_{S^{[n]}}P=\int_{S^n} \tilde P$$ for some polynomial $\tilde P$, which depends only on $P$, in the Chern classes of sheaves on $S^n$ of the form $pr_i^*T_S$ and $pr_{ij}^*{\cal O}_\Delta$. Any such expression $\int_{S^n}\tilde P$ can be universally reduced to a polynomial expression of integrals of polynomials in the Chern classes of $T_S$ (to see this for the Chern classes of $pr_{ij}^*{\cal O}_\Delta$ one applies Riemann-Roch without denominators \cite{Jouanolou}). This finishes the proof. \end{proof} \section{A generalization} We can generalize the methods used above to prove Proposition \ref{propB} to cover integrals of polynomial expressions in the Chern classes of tautological sheaves. Let $F_1,\ldots,F_\ell\in K(S)$. We require that the rank $r_i$ of $F_i$ ($i=1,\ldots,\ell$) is the same on all connected components of $S$. \begin{theorem}\label{GeneralTheorem}--- Let $P$ be a polynomial in the Chern classes of the tangent bundle $T_n$ of $S^{[n]}$ and the Chern classes of $F_1^{[n]}$,\ldots,$F_\ell^{[n]}$. Then there is a universal polynomial $\widetilde P$, depending only on P, in the Chern classes of $T_S$, the $r_1,\ldots,r_k$ and the Chern classes of $F_1$,\ldots,$F_\ell$ such that $$\int_{S^{[n]}}P =\int_S \widetilde P .$$ \end{theorem} \begin{proof}{} The proof goes along similar lines as that of Proposition \ref{propB}. The result immediately follows from a modified version of Proposition \ref{technoprop}: We now allow $f$ to be a polynomial in the $r_k$ and the Chern classes of the $$pr_0^*F_k^{[n+1]},pr_0^*T_{n+1},pr_{0i}^*I_{n+1}, pr_{ij}^*{\cal O}_\Delta, pr_i^*F_k,pr_i^*T_S$$ on $S^{[n+1]}\times S^{m}$, and get $\widetilde f$ to be a polynomial in the $r_k$ and the Chern classes of the analogously defined bundles on $S^{[n]}\times S^{m+1}$. To prove this use the recursion relation (\ref{tautorecursion}) and the formula $$c_i({\cal L}\otimes \rho^*F_k)=\sum_{a+b=i} \binom{r_k-b}{a} c_1({\cal L})^a\rho^*(c_b(F_k)),$$ to obtain $$\int_{S^{[n+1]}\times S^{m}}f= \int_Z\big(\sum_{\nu\geq0}\Phi^*f_\nu\cdot pr_Z^*(-c_1({\cal L}))^\nu\big),$$ where the $f_\nu$ are universal polynomials in the $r_k$ and the Chern classes of $$pr_0^*F_k^{[n]},pr_0^*T_{n},pr_{0i}^*I_{n}, pr_{ij}^*{\cal O}_\Delta, pr_i^*F_k,pr_i^*T_S.$$ Then we again use (\ref{pushdownofell}). \end{proof} We can use Theorem \ref{GeneralTheorem} to make predictions on the algebraic structure of certain formulae: let $\Psi:K\longrightarrow H^\times$ be a multiplicative function, i.e.\ for any complex manifold $X$ we are given a group homomorphism from the additive group $K(X)$ into the multiplicative group of units in $H(X;\IQ)$, which is functorial with respect to pull-backs and is polynomial in the Chern classes of its argument. The total Chern class, the total Segre class or the Chern character of the determinant are of this type. Also let $\phi(x)\in\IQ[[x]]$ be a formal power series and, for a complex manifold $X$ of dimension $n$ put $\Phi(X):=\phi(x_1)\cdot\ldots\cdot\phi(x_n)\in H^*(X,\IQ)$ with $x_1,\ldots,x_n$ the Chern roots of $T_X$. Let $S$ be a smooth projective surface and let $x\in K(S)$. For any functions $\Phi$ and $\Psi$ as above we define a power series in $\IQ[[z]]$ as follows: $$H_{\Psi,\Phi}(S,x):=\sum_{n=0}^\infty\int_{S^{[n]}} \Psi(x^{[n]})\Phi(S^{[n]})\,z^n.$$ \begin{theorem}\label{uni}--- For each integer $r$ there are universal power series $A_i\in\IQ[[z]]$, $i=1,\ldots,5$, depending only on $\Psi$, $\Phi$ and $r$, such that for each $x\in K(S)$ of rank $r$ (on every component of $S$) one has $$H_{\Psi,\Phi}(S,x)=\exp(c_1^2(x)A_1+c_2(x)A_2+c_1(x)c_1(S)A_3+c_1^2(S)A_4+ c_2(S)A_5).$$ \end{theorem} For simplicity we have suppressed the integrals $\int_S$ in the statement of the theorem and interpret the expressions $c_1(x)c_1(S)$ etc.\ as intersection numbers. \begin{proof}{} Let ${\cal K}_r:=\{(S,x)|S\mbox{ an algebraic surface}, x\in K(S), {\rm rank}(x)=r\}$, and let $\gamma:{\cal K}_r\to \IQ^5$ be the map $(S,x)\mapsto(c_1^2(x),c_2(x),c_1(x)c_1(S),c_1^2(S),c_2(S))$. The images of the five elements $(\IC\IP_2, r\cdot 1)$, $(\IC\IP_2,{\cal O}(1)+(r-1)\cdot 1$, $(\IC\IP_2,{\cal O}(2)+(r-1)\cdot 1)$, $(\IC\IP_2,2{\cal O}(1)+(r-2)\cdot 1)$, and $(\IC\IP_1^2,r\cdot 1)$ under $\gamma$ are the linearly independent vectors $(0,0,0,9,3)$, $(1,0,3,9,3)$, $(4,0,6,9,3)$, $(4,1,6,9,3)$, and $(0,0,0,8,4)$. Now, if $S=S_1\sqcup S_2$ we may decompose $(S,x)\in {\cal K}_r$ as $(S_1,x_1)+(S_2,x_2)$, where $x_i=x|_{S_i}$, and get $\gamma(S,x)=\gamma(S_1,x_1)+\gamma(S_2,x_2)$. Moreover, there is a decomposition of the class of the tautological sheaf analogous to the decomposition (\ref{disjointunion}) of the Hilbert scheme: $$x^{[n]}|_{S_1^{[n_1]}\times S_2^{[n_2]}} =pr_1^*(x_1^{[n_1]})\cdot pr_2^*(x_2^{[n_2]}).$$ >From the multiplicative behaviour of $\Psi$ and $\Phi$ we deduce $$\int_{S^{[n]}}\Psi(x^{[n]})\Phi(S^{[n]})=\sum_{n_1+n_2=n} \int_{S_1^{[n_1]}}\Psi(x_1^{[n_1]})\Phi(S_1^{[n_1]}) \int_{S_2^{[n_2]}}\Psi(x_2^{[n_2]})\Phi(S_2^{[n_2]})$$ and get \beeq{multiplicprop}H_{\Psi,\Phi}(S,x) =H_{\Psi,\Phi}(S_1,x_1)H_{\Psi,\Phi}(S_2,x_2). \end{eqnarray} By Theorem \ref{GeneralTheorem}, the function $H_{\Psi,\Phi}:{\cal K}_r\to\IQ[[z]]$ factors through $\gamma$ and a map $h:\IQ^5\longrightarrow\IQ[[z]]$. As the image of $\gamma$ is Zariski dense in $\IQ^5$ we conclude from (\ref{multiplicprop}), that $$\log h(y_1+y_2)=\log h(y_1)+\log h(y_2),\mbox{ for all }y_1,y_2\in\IQ^5,$$ i.e.\ $\log h$ is a linear function. This proves the theorem. \end{proof} \section{Riemann-Roch numbers} In this section we want to compute the Riemann-Roch numbers $\chi(M)$ for line bundles and vector bundles on $S^{[n]}$. Let $L_n:=f^*g_*(\otimes_{i=1}^n pr_i^*L)^{{\mathfrak S}_{n}}$, for $L$ a line bundle on $S$, where $f:S^{[n]}\to S^{(n)}$ and $g:S^n\to S^{(n)}$ are the natural morphisms, and $pr_i:S^n\to S$ is the $i$-th projection. This gives a monomorphism $-_n:Pic(S)\to {\rm Pic}(S^{[n]}), L\mapsto L_n$. It is well known that for $n\ge 2$ $$Pic(S^{[n]})=(Pic(S))_n\oplus {\mathbb Z} E, \qquad E:=\det({\cal O}_S^{[n]}),$$ ($E={\cal O}_S$ in case $n\le 1$). Moreover, $c_1(E)=-\frac{1}{2}D$, where $D$ is the exceptional divisor of $S^{[n]}\to S^{(n)}$. Note that for $F\in K(S)$ $$\det(F^{[n]})=\det(F)_n\otimes E^{{\rm rk}(F)}.$$ \begin{lemma}\label{chivonLn}--- The Euler characteristics of $L_n\otimes E$, $L_n$ and $L$ are related by the formulae $$\chi(L_{n})=\binom{\chi(L)+n-1}{n}\quad\mbox{ and } \quad\chi(L_n\otimes E)=\binom{\chi(L)}{n}.$$ \end{lemma} \begin{proof}{} Consider the cartesian diagram \beeq{dispC} \begin{array}{ccccc} \widehat S^n_*&\stackrel{\widehat f}{\longrightarrow}&S^n_*&\\[1ex] \scriptstyle{\widehat g}\Big\downarrow \phantom{\scriptstyle{\phi}}&&\phantom{\scriptstyle{\phi}}\Big\downarrow \scriptstyle{g}\\[2ex] S^{[n]}_*&\stackrel{ f}{\longrightarrow}&S^{(n)}_*, \end{array} \end{eqnarray} where $S^{(n)}_*$ is the open subscheme of zero cycles of length $n$ whose support consists of at least $(n-1)$ points and $S^{[n]}_*$ and $S^n_*$ are the preimages of $S^{(n)}_*$ under $f$ and $g$. It is easy to see that $\widehat S^n_*$ is the blow-up of $S^n_*$ along the (disjoint) diagonals $$\Delta_{i,j}=\big\{(x_1,\ldots,x_n)\in S^n_*\bigm| x_i= x_j\big\}.$$ Let $\widehat\Delta_{i,j}$ denote the corresponding exceptional divisors, and $\widehat\Delta:=\sum_{i<j} \widehat\Delta_{i,j}$ their sum. The family $\Gamma\subset \widehat S^n_*\times S$ corresponding to the classifying morphism $\widehat g$ is the union $\bigcup_{i=1}^n\Gamma_i$ of the graphs $\Gamma_i$ of the $i$-th projection $\widehat S^n_*\to S^n_*\to S$. Projecting the short exact sequence $$0\longrightarrow{\cal O}_{\Gamma}\longrightarrow\bigoplus_{i=1}^n{\cal O}_{\Gamma_i} \longrightarrow\bigoplus_{i<j}{\cal O}_{\Gamma_i\cap\Gamma_j}\lra0$$ to the base $\widehat S^n_*$ of these families we obtain a short exact sequence $$0\longrightarrow p_*{\cal O}_{\Gamma}\longrightarrow{\cal O}_{\widehat S^n_*}^{\oplus n}\longrightarrow \bigoplus_{i<j}{\cal O}_{\hat\Delta_{ij}}={\cal O}_{\hat\Delta}\longrightarrow 0$$ of ${\mathfrak S}_n$-linearized sheaves. ${\mathfrak S}_n$ acts on the sheaf in the middle by permutation of the factors. If we take the determinant of the first homomorphism in this sequence, we obtain a short exact sequence $$0\longrightarrow\det(p_*{\cal O}_{\Gamma})\longrightarrow{\cal O}_{\widehat S^n_*}^{\varepsilon}\longrightarrow {\cal O}_{\widehat \Delta}\longrightarrow 0$$ of ${\mathfrak S}_n$-linearized sheaves with equivariant homomorphisms, where the upper index $\varepsilon$ indicates that the ${\mathfrak S}_n$-linearization of ${\cal O}_{\widehat S^n_*}$ is the standard one twisted by the alternating character $\varepsilon:{\mathfrak S}_n\to{\mathbb Z}/2$. (Recall that any two $G$-linearizations of a line bundle on a $G$-scheme differ by a character of the group $G$). Thus we can identify $\widehat g^*E=\det(p_*{\cal O}_{\Gamma})$ with the ${\mathfrak S}_n$-linearized subsheaf ${\cal O}_{\widehat S^n_*}(-\widehat\Delta)^{\varepsilon}\subset{\cal O}_{\widehat S^n_*}^{\varepsilon}$, endowed with the alternating linearization. Now $\widehat g^*L_n=\widehat f^*L^{\boxtimes n}$ and $\widehat g^*(L_n\otimes E)=\widehat f^*L^{\boxtimes n} \otimes {\cal O}(-\widehat\Delta))$. As $S^{[n]}$ is smooth and $S^{[n]}\setminus S^{[n]}_*$ has codimension $2$, this gives \begin{eqnarray*} H^0(S^{[n]},L_n)&=&H^0(S^{[n]}_*,L_n)=H^0(\widehat S^n_*,\widehat f^*L^{\boxtimes n})^{{\mathfrak S}_n}\\ &=&H^0(S^n_*,L^{\boxtimes n})^{{\mathfrak S}_n}=H^0(S^n,L^{\boxtimes n})^{{\mathfrak S}_n}\\ &=&(H^0(S,L)^{\otimes n})^{{\mathfrak S}_n}\cong S^nH^0(S,L), \end{eqnarray*} and, similarly, \begin{eqnarray*} H^0(S^{[n]},L_n\otimes E)&\cong& H^0(S^{[n]}_*,L_n\otimes E))\\ &=&H^0(\widehat S^{n}_*, \widehat f^*L^{\boxtimes n} \otimes{\cal O}(-\widehat\Delta))^{{\mathfrak S}_n}\\ &=&H^0(S^n_*,L^\boxtimes)^{{\mathfrak S}_n,\varepsilon} =(H^0(S,L)^{\otimes n})^{{\mathfrak S}_n,\varepsilon}\\ &\cong&\Lambda^n(H^0(S,L)). \end{eqnarray*} Here the notation $V^{{\mathfrak S}_n,\varepsilon}$ means $\{v\in V|\pi(v)=\varepsilon(\pi)\cdot v,\forall\,\pi\in{\mathfrak S}_n\}$. Taking dimensions, we get \beeq{formulaforh0} h^0(L_n)=\binom{h^0(L)+n-1}{n}\quad\mbox{and} \quad h^0(L_n\otimes E)=\binom{h^0(L)}{n}. \end{eqnarray} Now let $H$ be an ample line bundle on $S$. By Grothendieck's construction of the Hilbert scheme it follows that $H_n\otimes E\cong\det(H^{[n]})$ is very ample for sufficiently ample $H$ on $S$. Applying this to $L\otimes H^k$ for sufficiently large $k$, We conclude that $$\chi(L_n\otimes H^k_n\otimes E) =h^0(L_n\otimes H_n^k\otimes E)=\binom{h^0(L\otimes H^k)}{n} =\binom{\chi(L\otimes H^k)}{n}. $$ Evaluating this equation of polynomials in $k$ at $k=0$ one finds $\chi(L\otimes E)=\binom{\chi(L)}{n}$ for all line bundles $L$. Let $L$ be an arbitrary line bundle on $S$ and let $H$ be ample $S$. Then $H^{\boxtimes n}$ is an ${\mathfrak S}_n$-linearized ample line bundle on $S^{n}$ and descends to an ample line bundle $H^{(n)}$ on $S^{(n)}$. For sufficiently large $k$, the line bundles $L\otimes H^k$ on $S$ and $(L\otimes H^k\otimes \omega_S\makebox[0mm]{}^{{\scriptstyle\vee}})^{(n)}$ on $S^{[n]}$ will be very ample. In particular, $f^*(L\otimes H^k\otimes\omega_S\makebox[0mm]{}^{{\scriptstyle\vee}})^{(n)}\cong L_n\otimes H_n^k\otimes\omega_{S^{[n]}}\makebox[0mm]{}^{{\scriptstyle\vee}}$ is globally generated and big. It follows from the Grauert-Riemenschneider vanishing theorem \cite{GrauertRiemenschneider} that $H^i(S^{[n]},L_n\otimes H_n^k)=0$ for all $i>0$. This gives $$\chi(L_n\otimes H_n^k)=h^0(L_n\otimes H_n^k) =\binom{h^0(L\otimes H^k)+n-1}{n}= \binom{\chi(L\otimes H^k)+n-1}{n}$$ for all sufficiently large $k$. As both sides are polynomials in $k$, we may take $k=0$. \end{proof} Consider the following power series in $\IQ[a,y][[z]]$: $$f_{y,a}:=\sum_{n\ge 0}\binom{y-a(n-1)}{n}z^n, \quad g_{y,a}:=\sum_{n\ge 0}\frac{y}{y-an}\binom{y-an}{n}z^n.$$ $f_{y,a}$ and $g_{y,a}$ play a role in combinatorics: Take $k$ points on a line with a distance of $1$ from one to the next. Then $\binom{k-a(n-1)}{n}$ is the number of ways to choose $n$ of them with a distance $>a$ among each other, and $\frac{k}{k-an}\binom{k-an}{n}$ is the same number if the $k$ points lie on a circle \cite{Rio}. In the proof of the following lemma we make use of the B\"urmann-Lagrange expansion formula in the following form: if $L(z)$ is a power series, and $\alpha(z)$ is a power series with $\alpha(0)=\alpha'(0)=0$, then $$\sum_{n\geq 0}\frac{1}{n!}\frac{d^n}{dz^n}(L(z)\alpha(z)^n) =\frac{L(\zeta)}{1-\alpha'(\zeta)}\Big|_{\zeta=z-\alpha(\zeta)}.$$ For a proof in an analytic context see \cite[p 128-135]{WhitWat}. \begin{lemma}\label{powseries}--- The power series $f_{y,a}$ and $g_{y,a}$ are related as follows: $$g'_{y,a}=yf_{y-2a-1,a},\qquad g_{y,a}=g_{1,a}^y,\qquad f_{y,a}=g_{1,a}^y\cdot f_{0,a}.$$ \end{lemma} \begin{proof}{} Using new formal variables $u,x$ and $v$ that are related by $z=u^a$, $u=x+x^{a+1}$, and $v=x^a$, so that $z=v(1+v)^a$, and, in the last line of the computation, the B\"urmann-Lagrange formula, we have \begin{eqnarray*} f_{y,a}(z)&=&\sum_{n\geq 0}\binom{y-a(n-1)}{n}z^n\\ &=&\sum_{n\geq 0}\binom{(a+1)n-(y+a+1)}{n}(-z)^n\\ &=&\sum_{n\geq 0}u^{y+a+1}\frac{1}{n!}\frac{d^n}{du^n} \frac{(-u^{a+1})^n}{u^{y+a+1}}\\ &=&\left(\frac{u}{x}\right)^{y+a+1}\frac{1}{1+(a+1)x^a}= (1+v)^y\cdot\frac{(1+v)^{a+1}}{1+(a+1)v} \end{eqnarray*} This gives $f_{0,a}=(1+v)^{a+1}/(1+(a+1)v)$ and $f_{y,a}(z) =f_{0,a}(z)\cdot (1+v(z))^{y}$. Now $dz=(1+(a+1)v)(1+v)^{a-1}dv$ and hence $$\frac{d}{dz} (1+v(z))=\frac{1}{(1+v)^{a-1}(1+(a+1)v)}=\frac{f_{0,a}}{(1+v)^{2a}}.$$ Differentiating $g_{y,a}$ we find \begin{eqnarray*} \frac{d}{dz}g_{y,a}(z)&=&\sum_{n\geq 1}\frac{yn}{y-an}\binom{y-an}{n}z^{n-1}\\ &=&\sum_{n\geq 1}y\binom{y-an-1}{n-1}z^{n-1}\\ &=&yf_{y-2a-1,a}=y(1+v)^{y-1-2a}f_{0,a}\\ &=&y(1+v)^{y-1}\cdot\frac{f_{0,a}}{(1+v)^{2a}}=\frac{d}{dz}(1+v)^y. \end{eqnarray*} As both $g_{y,a}$ and $(1+v)^y$ are power series in $z$ with constant term 1, we have $g_{y,a}=(1+v)^y$. Collecting the proven relations we conclude that $g_{1,a}=1+v$, $g_{y,a}=g_{1,a}^y$ and $f_{y,a}=g_{1,a}^yf_{0,a}$. \end{proof} \begin{theorem}\label{chithm}--- Let $K$ denote the canonical divisor of $S$. For each $r\in {\mathbb Z}$, there exist universal power series $A_r,B_r\in \IQ[[z]]$, such that for all $L\in Pic(S)$ $$\sum_{n\ge 0} \chi(L_n\otimes E^r))z^n=g_{1,r^2-1}(z)^{\chi(L)}\cdot f_{0,r^2-1}(z)^{\frac{\chi({\cal O}_S)}{2}}\cdot A_r(z)^{KL-\frac{K^2}{2}} \cdot B_r(z)^{K^2}.$$ Moreover, $A$ and $B$ satisfy the symmetry relations $A_{-r}=1/A_r$ and $B_{-r}=B_r$ for arbitrary $r$, and $A_r=B_r=1$ for $r=-1,0,1$. In particular, $$\chi(L_n\otimes E^{\pm 1})= \binom{\chi(L)}{n}.$$ If $S$ is a K3-surface, then $$\chi(L_n\otimes E^r)=\binom{\chi(L)-(r^2-1)(n-1)}{n},$$ and if $S$ is an abelian surface, then $$\chi(L_n\otimes E^r)=\frac{\chi(L)}{n}\binom{\chi(L)-(r^2-1)n-1}{n-1}.$$ \end{theorem} \begin{proof}{} Let $F=L+r\cdot 1\in K(S)$. Then $\det(F^{[n]})= L_n\otimes E^r$, $c_1(F)=c_1(L)$, $c_2(F)=0$. Therefore, by Theorem \ref{uni}, \begin{eqnarray*} \sum_{n\geq 0}\chi(L_n\otimes E^r)z^n &=&Z_1(z)^{K^2}Z_2(z)^{KL}Z_3(z)^{c_2(S)}Z_4(z)^{L^2}\\ &=&A_r(z)^{K L-\frac{K^2}{2}}B_r(z)^{K^2} F_r(z)^{\frac{\chi({\cal O}_S)}{2}}G_r(z)^{\chi(L)}, \end{eqnarray*} for suitable power series $Z_i,A_r,B_r,G_r,F_r\in \IQ[[z]]$. For the second equality we have used the identities $\chi(L)=\frac{L(L-K)}{2}+\chi({\cal O}_S)$ and $\chi({\cal O}_S)= \frac{c_1(S)^2+c_2(S)}{12}$. It is well-known that $\omega_{S^{[n]}}=(\omega_S)_n$. We get by Serre duality $$\chi(L_n\otimes E^{-r}))=\chi(\omega_{S^{[n]}}\otimes L_n\makebox[0mm]{}^{{\scriptstyle\vee}}\otimes E^r) =\chi((\omega_S\otimes L\makebox[0mm]{}^{{\scriptstyle\vee}})_n\otimes E^r)).$$ Using $\chi(L)=\chi(\omega_S\otimes L\makebox[0mm]{}^{{\scriptstyle\vee}})$ and $K(K-L)-\frac{K^2}{2}= -(KL-\frac{K^2}{2})$, this gives $B_{-r}=B_r$, $F_{-r}=F_r$, $G_{-r}=G_r$ and $A_{-r}=1/A_{r}$. To determine $F_r$ and $G_r$ explicitly, let $S$ be a K3-surface. Then by \cite{Bea} the Hilbert scheme $X:=S^{[n]}$ is an irreducible symplectic complex manifold. There exists a natural quadratic form $q$ on $H^2(X,{\mathbb Z})$ (see \cite{Bea},\cite{Fu},\cite{Huybas}), which on $H^{1,1}(X)$ is defined as follows: let $\omega\in H^2(X,{\mathbb Z})$ be the everywhere non-degenerate holomorphic $2$-form, normalized by $\int_X \omega^n\bar\omega^n =n!$. Then for $\alpha\in H^{1,1}(X)$: $$q(\alpha)=\frac{1}{(n-1)!}\int_X(\omega\bar\omega)^{n-1}\alpha^2.$$ Moreover, there exists a universal polynomial $h\in \IQ[z]$ such that $\chi(M)=h(q(c_1(M)))$ for all $M$ in $Pic(X)$. For $L\in Pic(S)$ Beauville \cite[Lemma 9.1, 9.2]{Bea} showed that $q(c_1(L_n))=L^2$, $q(c_1(E))=-2(n-1)$, and $E$ and $L_n$ are orthogonal with respect to $q$; thus $q(c_1(L_n\otimes E^r))=L^2-2r^2(n-1)$. The polynomial $h$ is determined by the formula of Lemma \ref{chivonLn} $$\chi(L_n)=\binom{\chi(L)+n-1}{n}=\binom{L^2/2+n+1}{n}=\binom{q(L)/2+n+1}{n} $$ applied to sufficiently many $L$ on $S$ with distinct values of $L^2$: \begin{eqnarray*}\chi(L_n\otimes E^r)&=&h(q(c_1(L_n\otimes E^r))) \\[1ex] &=&\binom{q(c_1(L_n\otimes E^r))/2+n+1}{n}\\[1ex] &=&\binom{\chi(L)-(r^2-1)(n-1)}{n}. \end{eqnarray*} As $\chi({\cal O}_S)=2$, we get $F_r\cdot G_r^{\chi(L)}=f_{\chi(L),r^2-1}$. It follows from Lemma \ref{powseries} that we can identify $F_r=f_{0,r^2-1}$ and $G_r= g_{1,{r^2-1}}$. Finally, the computations of Lemma \ref{chivonLn} show that the equation $\chi(L\otimes E^r)=\binom{\chi(L)-(r^2-1)(n-1)}{n}$ holds for all surfaces, if we restrict to small ranks $r=-1,0,1$. This implies that $A_r=B_r=1$ for $r=-1,0,1$. \end{proof} \begin{remark}\label{maple}--- Using Bott's residue formula the coefficients of $A_r$ and $B_r$ can be determined. We computed $A_r$ and $B_r$ up to order $8$ in $z$. Up to order $5$ in $z$ we get \begin{eqnarray*} \log(A_r)&\equiv&\left(\frac{1}{6}r-\frac{1}{6}{r}^{3}\right)z^2 +\left(\frac{1}{5}r-\frac{5}{8}r^3+\frac {17}{40}r^5\right)z^3\\ && +\left(\frac{29}{140}r-\frac{209}{180}r^3+\frac{88}{45}r^5 -\frac{631}{630}r^7\right)z^4\\ && +\left(\frac{13}{63}r-\frac{31259}{18144}r^3+\frac{16979}{3456}r^5 -\frac{69619}{12096}r^7+\frac{171215}{72576}r^9\right)z^5 \end{eqnarray*} \begin{eqnarray*} B_r&\equiv&1+\left (\frac{1}{24}{r}^{2}-\frac{1}{24}{r}^{4}\right ){z}^{2} +\left ({\frac {29}{360}}{r}^{2}-{\frac {31}{144}}{r}^{4}+{\frac {97}{720}}{r}^{6} \right){z}^{3}\\ &&+\left(\frac{139}{1260}r^2-\frac{3053}{5760}r^4+ \frac{2273}{2880}r^6-\frac{14899}{40320}r^8\right)z^4\\ &&+\left(\frac{187}{1400}r^2-\frac{6257}{6480}r^4 +\frac{421267}{172800}r^6-\frac{311701}{120960}r^8 +\frac{503377}{518400}r^{10}\right ){z}^{5} \end{eqnarray*} The method for showing this result is as follows. By Theorem \ref{chithm} it is enough to compute $\chi((kH)_n+rE)$ for $S=\IC\IP_2$ and $H$ the hyperplane bundle. By the Riemann-Roch theorem we have to compute $$\int_{\IC\IP_2^{[n]}}td_{\IC\IP_2^{[n]}} \cdot \exp((kH)_n+rE).$$ For this we use Bott's residue formula. The maximal torus $\Gamma$ of $SL(2,\IC)$ acts on $\IC\IP_2^{[n]}$ with finitely many fix points $Z$, which together with the structure of the tangent space as $\Gamma$-module are explicitly described in \cite{GE-SAS}. In \cite{GE-SAS2} also the structure of the fibres $(nH)^{[n]}$ as $\Gamma$ module is given. This information is enough to apply Bott's residue formula to compute the above intersection number following the strategy outlined in \cite{GE-SAS2}. The computations are carried out with a suitable Maple program. \end{remark} \begin{remark}\label{maple2} Theorem \ref{MainTheorem} allows us to use the Bott residue formula to compute the Chern numbers of $S^{[n]}$ for all $n$ and any surface $S$: We computed them for $n\le 7$. As an illustration we give the numbers for $S$ a $K3$ surface (in this case the odd Chern classes vanish) and $n\le 4$. For a partition $(n_1,\ldots,n_r)$ of $2n$ we write $(n_1,\ldots,n_r)$ for $c_{n_1}(S^{[n]})\cdot\ldots\cdot c_{n_r}(S^{[n]})$. We obtain $(4) = 324$, $(2^2) = 828$, $(6) = 3200$, $(4,2) = 14720$ $(2^3)= 36800$, $(8) = 25650$, $(6,2) = 182340$, $(4^2) = 332730$, $(4,2^2) = 813240$, $(2^4)= 1992240$. The strategy is similar to that of Remark \ref{maple}: by Theorem \ref{MainTheorem} it is enough to compute the Chern numbers in case $S=\IP_2$ and $S=\IP_1\times \IP_1$. In both cases we have an action of a 2-dimensional torus $\Gamma$ on $S^{[n]}$ with finitely many fix points. We proceed as above, again using a suitable Maple program. It is remarkable that for any surface $S$ and all $n\le 7$ all the Chern numbers of $S^{[n]}$ are polynomials in $c_1(S)^2$ and $c_2(S)$ with nonnegative coefficients (this was observed by G. Thompson). G. H\"ohn has used these numbers to check the conjecture of \cite{DMVV} about the elliptic genus of the Hilbert scheme of points $S^{[n]}$ on a $K3$ surface for $n\le 6$. \end{remark} Finally, we compute the holomorphic Euler characteristics $\chi(F^{[n]})$ of the tautological bundles $F^{[n]}$ on $S^{[n]}$. Related results are shown, using different methods in the recent preprint \cite{D}. \begin{proposition}\label{chifn}--- i) Let $F$ be a vector bundle on $S$. Then $$ \sum_{n,i} h^i(S^{[n]},F^{[n]})z^it^{n-1}=\Big(\sum_j h^j(S,F)z^j\Big) \frac{(1+zt)^{h^1({\cal O}_S)}}{(1-t)^{h^0({\cal O}_S)}(1-z^2t)^{h^2({\cal O}_S)}}. $$ In particular, if $S$ is connected, then $h^0(S^{[n]},F^{[n]})=h^0(S,F)$, and if furthermore $h^i(S,{\cal O}_S)=0$ for $i>0$, then $h^i(S^{[n]},F^{[n]})=h^i(S,F)$ for all $i$. ii) For all $F\in K(S)$ one has $$\chi(F^{[n]})=\chi(F)\binom{\chi({\cal O}_S)+n-2}{n-1}.$$ \end{proposition} \begin{proof}{} Part ii) follows from i) by putting $z=-1$. In order to prove part i) consider the cartesian diagram for $n\geq 1${} $$\begin{array}{ccccc} \Sigma_n&\stackrel{\overline f}{\longrightarrow}&Z_n&\\[1ex] \scriptstyle{p}\Big\downarrow \phantom{\scriptstyle{p}}&&\phantom{\scriptstyle{\overline p}}\Big\downarrow \scriptstyle{\overline p}\\[2ex] S^{[n]}&\stackrel{ f}{\longrightarrow}&S^{(n)}, \end{array}$$ where $f$ is the Hilbert-Chow morphism and $$Z_n:=\big\{(\eta,x)\in S^{(n)}\times S\bigm| x\in supp(\eta)\big\}$$ is the image of the universal family $\Sigma_n\subset S^{[n]}\times S$ in $S^{(n)}\times S$. We denote by $q:\Sigma_n\to S$, $\overline q:Z_n\to S$ the restriction of the projection. It is easy to see that there is an isomorphism $Z_n\cong S^{(n-1)}\times S$ which identifies $\overline q$ with the projection to the second factor. The maps $p$ and $\overline p$ are finite. By proposition \ref{hilbertchowvanishing} the higher direct images $R^if_*{\cal O}_{S^{[n]}}$ vanish. An easy spectral sequence argument shows that then the higher direct image sheaves $R^i\overline f_*{\cal O}_{\Sigma_n}$ must vanish as well. Moreover, $\overline f$ is a birational morphism of integral varities and $Z_n$ is normal. This shows that $\overline f_*{\cal O}_{\Sigma_n}={\cal O}_{Z_n}$. In particular, for any locally free sheaf $G$ on $Z_n$ one has $H^i(\Sigma_{n},\overline f^*G)=H^i(Z_n,G)$. By definition, $F^{[n]}=p_*q^*F$. As $p$ is finite, we get \begin{eqnarray*} H^i(S^{[n]},F^{[n]})&\cong& H^i(\Sigma_n,q^* F)=H^i(\Sigma_n,\overline f^*\overline q^*F)\\ &=&H^i(Z_n,\overline q^*F)=H^i(S^{(n-1)}\times S,{\cal O}_{S^{(n-1)}}\boxtimes F) \end{eqnarray*} for all $i$. By the K\"unneth formula this gives $$\sum_i h^i(S^{[n]},F^{[n]})z^i=\Big(\sum_ih^i(S,F)z^i\Big)\cdot \Big(\sum_ih^i(S^{(n-1)},{\cal O}_{S^{(n-1)}})z^i\Big).$$ By proposition \ref{hilbertchowvanishing}, $h^i(S^{(n-1)},{\cal O}_{S^{[n-1]}}) =h^i(S^{[n-1]},{\cal O}_{S^{[n-1]}})$. Hence $$\sum_{n}\sum_{i}h^i(S^{[n]},F^{[n]})z^it^{n-1} =\Big(\sum_ih^i(S,F)z^i\Big)\cdot\Big(\sum_{\nu} \sum_ih^i(S^{[\nu]},{\cal O}_{S^{[\nu]}})z^it^\nu\Big) $$ The second factor on the right hand side was computed in \cite[Proposition 3.3]{Goe} and equals $$\frac{(1+zt)^{h^1(S,{\cal O}_S)}}{(1-t)^{h^0(S,{\cal O}_S)}(1-z^2t)^{h^2(S,{\cal O}_S)}}.$$ \end{proof}
\section{Introduction} The idea that scale invariance is at the origin of critical phenomena associated with equilibrium second order phase transitions has proven to be very fruitful. The analysis of scale transformations in equilibrium statistical systems, now known as renormalization group (RG), has indeed allowed for the explicit calculation of critical exponents and, moreover, has led to the introduction of new fundamental concepts such as scaling and universality. The extension of the RG approach to non-equilibrium phenomena, where scale invariance is widely observed, and the identification of new universality classes, is of great importance from both theoretical and practical points of view. Technically, the RG ideas can be implemented in different ways. The most standard one for systems at equilibrium is to consider their stationary probability distribution written in terms of continuum coarse-grained fields, and study them perturbatively around their corresponding upper critical dimension. The most systematic way to extend the previous methods to non-equilibrium systems, where in general the stationary probability distribution is not known, is to cast them into a continuum dynamical equation \cite{HH}, or equivalently into a generating functional or action \cite{ZJ}. This last one can, in principle, be treated using the same perturbative techniques developed to deal with equilibrium systems. However, there are some cases where perturbative methods around a mean field solution are not suitable. In these cases the $\epsilon$-expansion fails to give information on the relevant physics. This turns out to be governed by a strong coupling, perturbatively inaccessible fixed point. The prototypical example of this class of systems is the well known Kardar-Parisi-Zhang (KPZ) equation for surface growth \cite{KPZ}, where the properties of rough surfaces have not been so far explained satisfactorily in generic spatial dimension. This is a problem of great theoretical importance since the KPZ describes not only the properties of rough surfaces\cite{HZ,Krug,Laszlo}, but is also related to the Burgers equation of turbulence \cite{Burger}, to directed polymers in random media \cite{DiP}, and to systems with multiplicative noise \cite{MN}. In particular, one of the most debated issues in this context is the existence of an upper critical dimension, above which the system is well described by the nontrivial infinite dimensional limit\cite{dinfi}. Although the usual approach fails for KPZ, the presence of generic scale invariance suggests that also for this system the basic idea of the RG approach should be applicable in some form. Real space approaches have proven useful wherever standard perturbative techniques fail \cite{VL,FST,Binney}. This, for example, is the case of fractal growth, and in particular for Diffusion-Limited-Aggregation\cite{FST}. However, the attempts to apply standard real space techniques to the KPZ problem (and to surface growth in general) fail because of a fundamental technical difficulty: The anisotropy of the scaling properties of the system. That is, in order to cover with blocks (in the Kadanoff sense \cite{Binney}) a surface, isotropic blocks cannot be used: Lengths in different directions must scale in different ways, and the relative shape of blocks has to depend upon the scale via an exponent that is unknown. This makes conceptually non-trivial the application of real space RG procedures to surface growth processes. In this paper we investigate the scale invariant properties of generic interface growth processes through the introduction of a real-space method. To achieve this goal we introduce some new ingredients permitting us to overcome the aforementioned problem. In particular, we introduce the idea that the statistical properties of growing surfaces on large scales can be described in terms of an effective scale invariant dynamics for renormalized blocks. Such dynamics is the fixed point of a RG transformation relating the parameters of the dynamics at different coarse-graining levels. The study of the RG flow, of the fixed points and of their stability gives the universality classes and their associated exponents. As a first application of the method, we study the KPZ growth dynamics and obtain accurate estimates for the roughness exponent (when compared with numerical results) in spatial dimensions from $d=1$ to $d=9$. Furthermore, an analytical approximation allows us to exclude the existence of a finite upper critical dimension for KPZ dynamics and suggests that the roughness exponent decays as $1/d$ for large dimensions, shedding light on a currently much debated issue. In order to show the generality of the new real space scheme and test its accuracy, we also apply it to the well known linear theory, the Edwards-Wilkinson (EW) equation. We reproduce the expected behavior in different dimensions, confirming the general applicability of the method. The paper is organized as follows. In section II we present the general RG method, the main concepts, the basic equations, and discuss all the approximations involved. In section III we review some results associated with KPZ growth and apply the new RG method to such problem. We present some simple analytical approximations, explicit results for spatial dimensions up to $d=9$, and discuss the large dimensional limit in detail. In section IV we report results on the analysis of the Edwards-Wilkinson equation. In section V a critical discussion of the method and of the results is reported. Partial accounts of the work presented here have already been published recently, with a slightly different notation\cite{Castellano98a,Castellano98b}. \section{Real Space RG for Surface Growth} In order to present the RG method let us consider a generic surface growth model where the height is a single-valued function $h(\vec x,t)$, with $\vec x$ the position in a $d$-dimensional substrate and $t$ denoting time. The possibility of having overhangs will not be considered here, as they are known to be irrelevant for the asymptotic behavior of KPZ-like growth \cite{overhangs}. The generic growth model under consideration can be either described at the microscopic level by a stochastic equation or by a discrete dynamical rule. In the first case $h$ and $\vec x$ are continuous variables, while in the latter they are discrete. The roughness of a system, when considered on a substrate of linear size $L$, is defined by \begin{equation} W^2(L,t)= {1 \over L^d} \sum_{\vec x} \left [h(\vec x,t)- {\bar h}(t) \right]^2, \end{equation} where \begin{equation} {\bar h}(t) = {1 \over L^d} \sum_{\vec x} h(\vec x,t). \end{equation} If we start the growth process from a flat configuration, for short times the roughness grows as \begin{equation} W(L,t) \sim t^\beta \end{equation} until it reaches a stationary state characterized by \begin{equation} W(L) \sim L^\alpha. \end{equation} The crossover between the two behaviors occurs at a characteristic time $t_s$, that scales with $L$ as $L^z$. This is the time scale over which correlations decay in the stationary state. The exponents $\alpha$, $\beta$ and $z$ are to a large extent universal for many different growth processes, and are related by the trivial scaling relation $\beta=\alpha/z$. We now introduce the real space renormalization group (RSRG) procedure aimed at the study the stationary state and in particular at the determination of the roughness exponent $\alpha$. The following subsections are structured as follows. In A) we introduce the geometric elements or blocks (equivalent to the Kadanoff blocks in standard RSRG methods) suitable to deal with anisotropic situations. In B) we discuss the effective dynamics of the previously defined blocks at a generic scale. In C) we introduce the RG equation and explain how the roughness exponent is determined. Finally in D) we analyze critically the approximations involved in general in the method. \subsection{Geometric description} The first nontrivial problem in the development of a RSRG approach is to find a sensible description of the geometry of the growing surface at a generic scale, i. e. how to build the analog of a block-spin transformation\cite{Binney}. Given the anisotropy of the system, the shape of the blocks must depend on the scale. Therefore, subdividing a cell in subcells is not a feasible task and the explicit construction of the block-spin transformation is not possible. Hence we develop an alternative strategy. To obtain a description at a generic scale $k$ of the growing surface, we consider a partitioning of the $(d+1)$-dimensional space in cells of lateral size $L_k = L_0 b^k$ and vertical size $h_k$. Here $b$ is a constant and $k$ labels the scale (Fig.~\ref{Fig01}). A cell is declared to be empty or filled according to a majority rule. In this way we pass from the microscopic description $h(\vec x,t)$ to a coarse-grained one at scale $k$, fully defined by the number $h(i,k,t)$ of filled blocks in the column $i$. Heights at scale $k$ are measured in units of $h_k$. The only characteristic vertical length at scale $k$ is that fixed by typical intrinsic fluctuations of the surface of a lateral size $L_k$. This suggests to take \begin{equation} h_k = \sqrt{c} W(L_k) \sim L_k^\alpha, \label{h_k} \end{equation} where $\sqrt{c}$ is a proportionality constant that will be discussed later. This equation expresses the requirement of scale invariance in the geometric description. Any other choice would result either in a redundant description (if $h_k/W(L_k)\to 0$ as $k\to\infty$) where too many (infinite) blocks would be needed to describe fluctuations in the same column, or in a too coarse description (if $h_k/W(L_k)\to \infty$ as $k\to\infty$). By imposing Eq. (\ref{h_k}), we always have a meaningful covering of the surface upon scale changes. Observe that since in general $\alpha \neq 1$, the shape (i.e. the ratio of vertical to horizontal length) of the blocks changes with the scale $k$. Contrarily to the usual RG approach, the definition of the block-spin transformation depends explicitly on the roughness exponent $\alpha$, the calculation of which is the final goal of the method. The constant $c$ in Eq.~(\ref{h_k}) fixes the unit of measure of our blocks. Its optimal value can be determined as follows. The distribution of microscopic height fluctuations within a block can be mapped into an {\em effective} distribution with the same average $\bar h$ and standard deviation. For simplicity we take it to be bimodal \begin{eqnarray} P[h(x)] = p &\delta & \{h(x)-[{\bar h}+(1-p) h_k]\} \nonumber \\ + (1-p) &\delta & \{h(x)-[{\bar h}-p h_k]\}. \label{pofh} \end{eqnarray} This distribution results from mapping all points with microscopic height larger than $\bar h$ to $\bar h+ (1-p) h_k$ and those smaller than $\bar h$ to $\bar h -p h_k$. The parameter $p$ describes the degree of asymmetry of the distribution: The fluctuations inside a block can then be calculated, using Eq.~(\ref{pofh}), as \begin{equation} W^2(L_k)= p (1-p) h^2_k, \label{prop} \end{equation} which implies that the constant $c$ is given by \begin{equation} c = {1 \over p(1-p)}. \end{equation} For a symmetric distribution $p=1/2$ and therefore $c=4$. In general the height distribution is not symmetric, i. e. there is some nonvanishing skewness and one must consider $c \ne 4$. \subsection{Dynamic description} The second step in the construction of the RG procedure is the definition of the effective dynamics at a generic scale $k$, i.e. the determination of the growth rules for the blocks defined in the previous subsection. The effective dynamics will depend on a set of scale dependent parameters. The changing of scale induces a flow in the parameter space whose fixed points correspond to the scale invariant dynamics. Analogously to what happens in the usual application of the RG approach to equilibrium systems, it may happen that mechanisms not appearing in the microscopic rule are generated upon coarse-graining. In the language of equilibrium systems this means that operators not included in the bare Hamiltonian can be generated iteratively. Conversely, microscopic ingredients can prove to be irrelevant and be progressively eliminated when going to coarser scales. Therefore, exactly as in the equilibrium case the choice of the parametrization of the effective dynamics is not trivial: Principles as the preservation of symmetries and conservation laws must be the guidelines. In general, the effective dynamics will be defined in terms of the transition rates for the addition of occupied blocks at a generic coarse-grained scale, that is, \begin{equation} r[h(i,k) \to h'(i,k)] = r(x_k^1,x_k^2,\ldots,x_k^n). \end{equation} The number of parameters $x_k^i$ is in principle arbitrary, although in the applications presented below it will be limited to one \cite{bi}. It is clear that the more complete the parametrization the better the final description of the statistical scale invariant state. We will discuss this problem in detail in subsection D. \subsection{The RG Equations} So far we have defined the geometrical and dynamical aspects of the coarse-graining procedure. These give us the necessary ingredients to introduce the RG transformation. The explicit derivation of it is based on the following property of the roughness $W$. Let us consider a $d$-dimensional system of linear size $L$ and partition it in $(L/b)^d$ blocks of size $b^d$ (labeled by the index $j$). It is straightforward to verify that the total roughness can be decomposed as \begin{eqnarray} W^2(L) &= & {1 \over (L/b)^d} \sum_{j=1}^{(L/b)^d} \left \{ {1\over b^d} \sum_{i \in j} \left[h(i)-{\bar h}(j) \right]^2 \right \} \nonumber \\ &+& {1\over (L/b)^d} \sum_{j=1}^{(L/b)^d} \left[{\bar h}(j)-{\bar h} \right]^2, \end{eqnarray} where ${\bar h}(j)$ is the average height within block $j$. The interpretation of this formula is simple: The first term on the right hand side is the averaged value of the roughness within blocks of size $b^d$, while the second term is the fluctuation of the average value of $h$ among blocks. In our coarse-graining procedure this property is read as follows: If one takes $L=L_{k+1}=b L_k$ the first term on the right hand side is $W^2(L_k)$, the total roughness within a block of size $L_k$; the second is the roughness of the configuration in which blocks of size $L_k$ are considered as flat objects. This second contribution is obviously proportional to the square of the height of a block $h_k^2$. Hence, employing Eq. (\ref{h_k}), \begin{eqnarray} W^2(L_{k+1}) &= & W^2(L_k)+ \omega^2(b,k) h_k^2 \nonumber \\ & = & \left[ 1 + c\omega^2(b,k)\right] W^2(L_k) \nonumber \\ & = & F_b(k) W^2(L_k), \label{main} \end{eqnarray} where $\omega^2(b,k)$ is the roughness in the stationary state of a system of $b^d$ sites of unit height that evolves according to the dynamical rules specified by $(x_k^1,x_k^2,\ldots,x_k^n)$, and \begin{equation} F_b(k) \equiv \left[ 1 + c\omega^2(b,k)\right]. \label{F} \end{equation} Note that the dependence on the scale $k$ is only through the parameters $\{x_k\}$. {\it Eq. (\ref{main}) is the equation that relates the width at scales $k$ and $k+1$.} In order to proceed further, we must evaluate the function $F_b(k)$, or equivalently $\omega^2(b,k)$. To do so, we identify all the possible surface configurations of a system of composed of $b^d$ sites, and write down a master equation for their associated probabilities $\rho_i$ \begin{equation} \partial_t \rho_i = \sum_j \rho_j P_{j \to i} - \rho_i \sum_j P_{i \to j}. \end{equation} $P_{i \to j}$ is the rate for the transition between configuration $i$ and $j$ and depends on the set of parameters $\{x_k\}$. Imposing the stationarity condition $\partial_t \rho_i=0$ and the normalization $\sum_i \rho_i=1$ the master equation can be solved. If we call $W^2_i$ the roughness of configuration $i$, then we can write \begin{equation} \omega^2(b,k) = \sum_i \rho_i(k) W^2_i. \end{equation} Depending on the particular structure of the master equation the explicit solution of the previous equation may be difficult or impossible. In such cases it may be more useful to determine $\omega^2(b,k)$ numerically by performing (relatively small) Monte Carlo simulations. We will describe examples of both analytical and numerical computations of $\omega^2(b,k)$. Let us suppose now that $\omega^2(b,k)$ has been determined. Eq.~(\ref{main}) gives an explicit relation between the roughness at two different scales. Observe that so far the scale invariance idea has not been implemented. We have just studied how the width changes upon changing the level of description. The last task to be performed is the determination of the RG transformation relating the parameters of the dynamics at scale $k$ with those at scale $k+1$. This is done by means of a self-consistency requirement for the description of the same system at two different levels of detail, i.e. the total width of a system should be independent on the size of the blocks we use to describe it. To make this idea more precise, let us consider the case of a dynamics parametrized by only one parameter $x_k$. Let us take a system of size $L=L_{k+2}$. By applying Eq.~(\ref{main}) we have \begin{equation} W^2(L_{k+2}) = F_b(x_{k+1}) W^2(L_{k+1}). \end{equation} This procedure can be iterated again on each of the resulting systems of size $L_{k+1}$, obtaining \begin{equation} W^2(L_{k+2}) = F_b(x_{k+1}) F_b(x_k) W^2(L_k). \end{equation} The same quantity can alternatively be computed by considering directly the whole system as composed by $b^{2d}$ systems of size $L_k$. Applying again Eq.~(\ref{main}) \begin{equation} W^2(L_{k+2}) = F_{b^2}(x_k) W^2(L_k). \end{equation} Imposing the consistency of the two procedures one has an implicit RG transformation for $x_k$ \begin{equation} F_b(x_{k+1}) = {F_{b^2}(x_k) \over F_b(x_k)}, \label{RGeq} \end{equation} or explicitly \begin{equation} x_{k+1} = R(x_k) \equiv F_b^{-1} \left[{F_{b^2}(x_k) \over F_b(x_k)} \right]. \end{equation} This equation provides the evolution of the parameter under a change of scale. If a fixed point $x^*$ such that \begin{equation} x^*= R(x^*) \label{FP} \end{equation} exists, then the parameter $x^*$ characterizes the scale invariant dynamics of the system. The knowledge of it directly allows the determination of the exponent $\alpha$. Since $W^2(L_{k+1})/W^2(L_k)$ is equal to $ b^{2 \alpha}$ we have \begin{eqnarray} \alpha & = & \lim_{k \to \infty} {1 \over 2} \log_b \left[ {W^2(L_{k+1}) \over W^2(L_k)} \right ] \nonumber \\ & = & \lim_{k \to \infty} {1 \over 2} \log_b F_b(k)={1 \over 2} \log_b F_b(x^*). \label{alpha} \end{eqnarray} To analyze the stability of the fixed point we linearize the RG transformation around it \begin{equation} x_{k+1} - x^* = R(x_k) - x^* \simeq R'(x^*) (x_k-x^*). \end{equation} Hence if $|R'(x^*)|<1$ the scale invariant dynamics specified by $x^*$ is an attractive fixed point under changes of scale. Extension of the previous formalism to the case of $n$ parameters of the dynamics is straightforward. The $n$ RG transformations are obtained by imposing the consistency of the description of the same system when divided in $2^d$ and $4^d$ blocks, in $4^d$ and $16^d$ blocks, and so on \cite{bi}. \subsection{Approximations} Let us discuss now the approximations involved in the method. There are two steps where approximations come into play: The first is the choice of the parametrization of the scale invariant dynamics. The second is the computation of $\omega^2(b,k)$. With respect to the first problem, it is reasonable to expect that under coarse-graining the microscopic dynamics will flow towards a scale invariant dynamics depending in principle on an infinite number of parameters. This proliferation is analogous to what happens in RSRG approaches to equilibrium systems. The restriction to a finite (and small) number of parameters involves unavoidably an approximation, due to the projection of the RG flow onto the sub-space spanned by these parameters. However, a very important difference with respect to equilibrium critical phenomena is that here the scale invariant dynamics is ``self-organized critical'', that is, there are no relevant operators. Only irrelevant fields, with negative scaling dimensions, need to be parametrized. The system is by definition {\em on the critical manifold} and, by iteration of the RSRG transformation, it converges to the stable fixed point, without any fine tuning of parameters. The projection onto a low-dimensional parameter space yields a projected RG flow which will share these same properties. The fixed point in this sub-space, being the projection of the actual fixed point in the high-dimensional space, will have the same qualitative properties. Even the simplest parametrization capturing the correct symmetries of the dynamics can provide a quite accurate determination of the properties of the system in this case. On the contrary, when relevant fields are present, as in second order phase transitions, truncation effects are quite dramatic. The reason is that relevant fields have, in general, a non-vanishing component on any discrete (lattice) operator. Any approximation due to truncation is amplified by the RG iteration thus driving the flow {\em out of} the critical manifold along the relevant directions. The determination of the fixed point becomes then very difficult\cite{Baillie}. The second source of approximation is the computation of $\omega^2(b,k)$. As stated above this quantity is the stationary roughness of a system composed of $b^d$ substrate sites evolving according to the dynamical rules specified by the parameters $(x_k^1,\ldots,x_k^n)$. This is a perfectly well defined quantity that may in principle be computed to any degree of accuracy by solving the master equation. However very often the structure of the master equation is too complicated to allow for a full solution. One then has to devise suitable simplifications to make the analytical computation feasible. This involves approximations that affect the final result. We will see an example of this way of proceeding and discuss how the effect of the approximation can be controlled. Alternatively, when $b$ and $d$ are not too large, one can resort to the numerical evaluation of $\omega^2(b,k)$. In practice this boils down to performing Monte Carlo simulations of small systems evolving with different values of the parameters $\{x_k\}$. It is important to stress that the MC procedure involves no approximation, except for the fluctuations associated with statistical sampling. We will describe below an example of this alternative way of computing $\omega^2(b,k)$. A delicate issue is also the choice of the boundary conditions. In the conceptual framework described above, $\omega^2(b,k)$ is the roughness of a section of size $b$ of an infinitely extended surface. This would suggest the use of open boundary conditions. On the other hand, when integrating out degrees of freedom relative to height fluctuations inside the cell, one should not consider the fluctuation of the average slope. This slope effect is eliminated if one uses closed (i.e. periodic) boundary conditions. Even though the choice of the appropriate boundary conditions is not trivial, we will see, in the KPZ case, that the use of periodic or open boundary conditions has little effect on the value of the exponent. Furthermore, one expects that both truncation errors and those induced by neglecting fluctuations of boundary conditions vanish as the parameter $b$ grows. Arguments in support of this conclusion are reported in the Appendix I. \section{RG for KPZ dynamics} \subsection{The problem of KPZ growth} The Kardar-Parisi-Zhang equation is the minimal continuum equation capturing the physics of rough surfaces. After appearing in 1986 \cite{KPZ}, an overwhelming number of studies has been devoted to elucidate its properties\cite{HZ,Laszlo}. It reads \cite{KPZ} \begin{equation} {\partial h(x,t) \over \partial t} = \nu \nabla^2 h + {\lambda \over 2} (\nabla h)^2 + \eta(x,t). \end{equation} where $h(x,t)$ is a height variable at time $t$ and position $x$ in a $d$-dimensional substrate of linear size $L$. $\nu$ and $\lambda$ are constants and $\eta$ is a gaussian white noise. As a consequence of a tilting (Galilean) invariance \cite{Burger,HZ,Krug} $\alpha + z=2$, and since in general $z=\alpha/\beta$, there is only one independent exponent, say $\alpha$. The difference between the KPZ equation and the linear equation (Edwards-Wilkinson), describing surfaces growing under the effect of random deposition and surface tension, is the presence of a nonlinear term proportional to $\lambda$. This nonlinear term is generated by microscopical processes giving rise to lateral growth, i.e. the fact that growth velocity is normal to the local surface orientation. Exact results\cite{HZ,Laszlo} indicate that in $d=1$ there is only a rough phase for KPZ with $\alpha=1/2$. Instead, standard field theoretical methods predict the presence of a {\it roughening transition } above $d=2$ \cite{Terry}; i.e., there are two RG attractive fixed points and an unstable fixed point separating them. More specifically, there is a gaussian fixed point with $\alpha=0$ describing a flat phase (characterized by a vanishing renormalized nonlinear coupling) and a nontrivial one describing the rough phase (in which the renormalized nonlinear coupling diverges in perturbation theory). Perturbative methods fail to give any prediction for the exponents in the rough phase. For $d > 2$, an $\epsilon$-expansion ($d=2+ \epsilon$) around the gaussian solution can be performed and the exponents at the roughening transition evaluated to all orders in perturbation theory \cite{Nuclear,MN}. These results seem to indicate the presence of an anomaly in $d=4$ for the roughening transition. This has been interpreted as an indication that $d_c=4$ is the upper critical dimension for the rough phase, i.e., for the strong coupling fixed point \cite{MCDC,Lassig,Bhatta}. Above this dimension the exponents should take the values known for $d=\infty$ \cite{dinfi}. Applications of non-perturbative methods such as functional renormalization group\cite{hh} and Flory-type arguments \cite{Flory} also suggested that $d_c=4$, in agreement with a $1/d$-expansion\cite{1/d} around the $d=\infty$ limit. The mode-coupling approximation led to contradictory results, suggesting the existence of a finite $d_c$ \cite{MCDC} or $d_c=\infty$ \cite{Yuhai}. Arguments for a finite $d_c$ based on directed\cite{DP} or invasion\cite{Cie} percolation have also been proposed. On the other hand, numerical results seem to indicate that the exponent $\alpha$ decays continuously with the system dimensionality up to $d=7$, excluding therefore $d=4$ as upper critical dimension \cite{debate}. Finally, some doubts have been cast on the validity of the continuum approach to study rough surfaces \cite{discre}. Summing up, the issue of the behavior of the KPZ dynamics for $d \ge 2$ is a highly debated one, and it is extremely desirable to have alternative approaches shedding light into the problem. In what follows we present the application of our new RG scheme to KPZ growth. \subsection{Simplest RG scheme} \subsubsection{Parametrization of the dynamics} The modelization of the dynamics at a generic scale should keep the number of parameters to a minimum and catch all the relevant physical mechanisms of the process. The main feature of the KPZ dynamics is lateral growth. Therefore we take as the only parameter defining the dynamics at a generic scale $k$, the ratio $x_k$ between lateral and vertical growth (i.e. random deposition). More formally, the growth rate for the addition of an occupied block on column $i$ is \begin{eqnarray} r_i & \equiv & r\left[h(i) \to h(i)+1 \right] \\ &= & 1 + x_k \sum_{j n.n. i} \max\left[0,h(j)-h(i)\right]. \label{r_i} \end{eqnarray} Eq.~(\ref{r_i}) states that the rate for lateral growth is proportional to the difference in height between neighboring columns (Fig.~\ref{Fig02}). Overhangs, known to be irrelevant on large scales \cite{overhangs}, are not allowed. This dynamics can be seen as a generalization of the Eden growth model. Few observations are in order. We call the parameter $x$ appearing in Eq.~(\ref{r_i}) ``lateral growth'' parameter, but this is an abuse of language: $x_k$ cannot be identified with the parameter $\lambda$ of the KPZ equation. Instead the term that multiplies $x_k$ in Eq.~(\ref{r_i}) is a combination of the discretized Laplacian, of the discretized square gradient and of other discrete operators. The explicit dependence of $x$ on $\nu$ and $\lambda$ cannot be disentangled. Other parametrizations are clearly possible and they will be discussed below. Eq.~(\ref{r_i}) has the nice feature that it contains as limiting cases both the random deposition process ($x_k=0$) and the infinitely strong ``lateral growth'' ($x_k=\infty$) leading to flat surfaces. Most importantly, it is easy to see that $x^*=\infty$ is, by construction, a fixed point of the RSRG with $\alpha=0$. This feature makes it possible the determination of the upper critical dimension above which the stable solution leads to $\alpha=0$. In this situation we expect $x^*=\infty$ to be an attractive fixed point. Below the critical dimension, on the other hand, the fixed point $x^*=\infty$ must be unstable and an intermediate fixed point with finite $\alpha$ must appear. The RSRG accommodating for a fixed point at $x^*=\infty$, naturally allows to address the issue of the upper critical dimensionality. \subsubsection{d=1} We restrict ourselves for the moment to the one-dimensional case and illustrate in detail the application of the RG approach, i. e. the computation of $\omega^2(b,k)$, the determination of the scale invariant dynamics and of the exponent $\alpha$. It is very instructive to consider first the dynamics Eq.~(\ref{r_i}) supplemented by the condition that the height difference between adjacent columns is restricted to values such that ($|\Delta h|\leq \Delta h_{max}$), with $\Delta h_{max}=1$. This greatly reduces the number of possible surface configurations allowing for a full analytical treatment. For the system of size $b=2$, assuming periodic boundary conditions, there are only 2 nonequivalent configurations, while for the system of size 4 there are 6 of them (Fig.~\ref{Fig03}). Using the definition Eq.~(\ref{r_i}) of the growth rates for the addition of a block, one has simply, for $b=2$, \begin{eqnarray} P_{1 \to 1} &=& 0 \nonumber \\ P_{1 \to 2} &=& 2 \nonumber \\ P_{2 \to 1} &=& 1+2 x_k \nonumber \\ P_{2 \to 2} &=& 0. \end{eqnarray} In configuration 1 only vertical growth is possible (in two sites) leading always to configuration 2. Only one site can instead grow in configuration 2 and the rate for this is the sum of the rate of one vertical and two lateral contributions. Hence the master equation reads \begin{eqnarray} {\partial \rho_1 \over \partial t} &=& (1+2 x_k) \rho_2 - 2 \rho_1 \\ {\partial \rho_2 \over \partial t} &=& 2 \rho_1 -(1+2 x_k) \rho_2. \end{eqnarray} Imposing the stationarity condition $\partial_t \rho_1=\partial_t \rho_2=0$ and the normalization $\rho_1+\rho_2=1$ one has \begin{equation} \rho_2={2 \over 3+2 x_k} \end{equation} and then considering that the width associated with configurations 1 and 2 is, respectively, $0$ and $1/4$ \begin{equation} \omega^2(2,x_k)={2 \over 4(3+2x_k)}. \end{equation} For the system of size $b^2=4$ one finds in an analogous way \begin{equation} \omega^2(4,x_k)={51+86 x_k +40 x_k^2 \over 4 (47+106 x_k+68 x_k^2 + 8 x_k^3)}. \end{equation} Plugging these two expressions into the RG equation~(\ref{RGeq}) with $c=4$ \footnote{In $d=1$ the distribution is known to be symmetric.} one finds that the explicit form of the RG transformation is \begin{equation} R(x)={293+804 x+636 x^2 +160 x^3 + 32 x^4 \over 2(59+148 x + 156 x^2 + 64 x^3)} \end{equation} and that there exists only one finite fixed point for \begin{equation} x^* \simeq 2.08779 \ldots \end{equation} Such a fixed point is attractive since \begin{equation} R'(x^*)\simeq -0.03548\ldots \end{equation} Hence no matter how small or large the microscopic value of $x$ is, upon coarse-graining the dynamics flows towards an attractive scale invariant dynamics, characterized by a ratio $x^*$ of the lateral to vertical growth rates. The roughness associated with this scale invariant dynamics is \begin{equation} \alpha= {1 \over 2} \log_2 F_{b=2} (x^*) \simeq 0.177352\ldots \end{equation} that must be compared with the known exact value $\alpha=1/2$. The apparent poor performance of the method is due to the assumption that $\Delta h_{max}=1$, which allows for full analytical treatment, but is clearly wrong. The point is that even if at the microscopic level the dynamics is of restricted type, the effective dynamics at generic scale defined by the renormalization procedure will proliferate in a non restricted one. Allowing larger steps ($\Delta h_{max}>1$) increases the number of superficial configurations and makes the analytical determination of the function $\omega^2(b,x_k)$ impossible. Still this task can be performed numerically via simulation of systems of such a small size. Fig.~\ref{Fig04} reports the results obtained by considering increasing values of $\Delta h_{max}$. The value of $x^*(\Delta h_{max})$ converges already for $\Delta h_{max}=8$ to $x^* \simeq 0.726$, corresponding to a value $\alpha=0.507\ldots$ in excellent agreement with the exact value. Further increases of $\Delta h_{max}$ do not change the results, indicating that in the scale invariant dynamics the probability of steps larger than 8 is negligible. \subsubsection{$d>1$} The computation via Monte Carlo method of $\omega^2(2,x_k)$ and $\omega^2(4,x_k)$ can be performed with very little computational effort also in higher dimension. We considered $d=1,\ldots,9$ with less than a week of CPU time of a workstation. The results are reported in Table I and summarized graphically in Fig.~\ref{Fig05}. We find a finite attractive fixed point for all dimensions, with an exponent $\alpha$ in remarkably good agreement with the best numerical results available\cite{debate}. This is the first theoretical approach providing estimates for the roughness exponent that match in all dimensions with numerics. No anomalies are found for $d=4$ where other approaches find an upper critical dimension. The extrapolation to $d \to \infty$ suggests that the fixed point is always stable and that $\alpha$ decreases with $d$ but remains always nonvanishing. The fixed point parameter $x^*$ grows exponentially with the dimension. These results are confirmed by an analytical expansion of the method in high dimensions, that is presented in subsection D. \subsection{Robustness of the results} \subsubsection{$b>2$} In order to analyze the stability of the results upon increasing the value of $b$ it is convenient to introduce the function \begin{equation} \alpha_\ell(x) = {1 \over 2} \log_\ell F_\ell(x). \label{alphal} \end{equation} With this definition one can express the fixed point condition Eq.~(\ref{FP}) as \begin{equation} \alpha_b(x^*) = \alpha_{b^2}(x^*), \label{FPl} \end{equation} and see that the fixed point is stable if \begin{equation} |R'(x^*)| = \left| 2 {\alpha'_{b^2}(x^*) \over {\alpha'_b(x^*)}}-1 \right| < 1, \label{Ralpha} \end{equation} i. e. \begin{equation} 0<{\alpha'_{b^2}(x^*) \over \alpha'_b(x^*)} < 1. \end{equation} Such a formula can be extended also to the case where the size of the larger system considered is not $b^2$ but a generic $b'>b$. We study the stability of the results for growing $b$ by computing $\alpha_b(x)$ with $b_i=2,4,8,16,\ldots$ and imposing the consistency between two successive $b_i$. The value of $b$ indicated in the plots is the smaller one; for instance, $b=4$ label the results obtained imposing the consistency between $b=4$ and $b'=8$. In Fig.~\ref{Fig06} we report the plot of the curves $\alpha_b(x)$ in $d=1$ for $b=2,4,8,16,32$. Remarkably they all meet practically at the same point indicating that $x^*$ and $\alpha$ virtually do not change by increasing the number of cells. Fig.~\ref{Fig07} reports the values of the exponent $\alpha$ in $d=1$ (empty circles). Observe that fluctuations are extremely small. The value of the fixed point parameter $x^*$ is reported in Fig.~\ref{Fig08} (empty circles). Again it remains practically unchanged when $b$ grows. In higher dimensions the results are less stable. The values of $\alpha$ and of $x^*$ for $d=2, 3$ and 4 are reported in Fig.~\ref{Fig07} and~\ref{Fig08}, respectively (empty symbols). A clear trend is present for $d=2$: the exponent initially decreases as $b$ is increased, then reaches a minimum and starts growing. This behavior is reflected in the value of $x^*$ that first grows and then decreases. The decreasing part of the pattern is present in the analogous plots for $d=3$ and $d=4$. For large dimensions however, it is increasingly more time consuming to perform the computation for large systems. In particular for $d=4$, the largest system that could be simulated is $b=16$ and for such a system size the trend is still decreasing. Therefore it is not possible to decide from a numerical point of view whether for larger values of $b$ $\alpha$ would converge to zero or to a finite value. These data do not provide any conclusive indication on whether $d=4$ is the upper critical dimension for KPZ growth. However, as it will be shown below, such a conclusion is ruled out by the results with other parametrizations and by the analytical large-$d$ expansion of the method. The reason for the difference in the stability of the results for large number of cells in $d=1$ and $d \geq 2$ is probably related to crossover phenomena. In the RG flow there are two competing fixed points; this reflects the existence of two universality classes, strong coupling KPZ and EW. In $d=1$ this fixed points are associated with the same roughness exponent $\alpha=1/2$ and to similar values of $x^*$. Therefore, any crossover phenomenon between the two fixed points has little effect in our formalism. In $d \geq 2$ instead, the two scale invariant dynamics are associated with different exponents and also very different values of the parameter $x^*$, which is finite for KPZ and infinite for EW. We interpret the initial decrease in the value of $\alpha$ in the KPZ case as the effect of a crossover caused by the presence of the EW fixed point. It is not clear to us, however, why the fixed point found for $b=2$ is so close to the results of the numerical simulations. \subsubsection{Open boundary conditions} The calculation of $\omega^2(b,k)$ can also be performed with open boundary conditions, that is assuming that the height of the columns outside the system which are in contact with the boundary is the same of their neighbors inside the system. This means that no ``lateral growth'' event can be caused in the system by the environment around it. The results are also presented in Fig.~\ref{Fig07} and~\ref{Fig08} (filled symbols). Interestingly, in this case the accuracy of the method for $b=2$ is not as good as for periodic boundary conditions, but the error remains below 10\%, indicating a low sensitivity to the boundary conditions even for small number of cells. For larger number of cells the difference goes quickly to zero. For higher dimensions the general dependence of $x^*$ and $\alpha$ on $b$ remains unchanged: In $d=2$ $\alpha$ is initially high, then decreases and finally increases again. The variations with $b$ are however less strong than when periodic boundary conditions are considered. For $d=4$ it is more clear than in the case with periodic boundary conditions that $\alpha$ does not converge to zero for large $b$. \subsubsection{Other parametrizations of the dynamics} As stated above the parametrization~(\ref{r_i}) of the KPZ scale invariant dynamics is by no means unique. Actually, given the problems of slow and nonmonotonic convergence towards the asymptotic values, it turns out clearly that the parametrization~(\ref{r_i}) is quite far from being optimal and better parametrizations would help. In order to keep things as simple as possible we started considering transition rates of the form \begin{equation} r_i = 1 + x_k \sum_{j n.n. i} \left\{\max\left[0,h(j)-h(i)\right]\right\}^{\gamma} \label{gamma} \end{equation} with $\gamma$ constant; for $\gamma=1$ it coincides with~(\ref{r_i}). By comparing the values of $\alpha$ obtained with several $\gamma$ an interesting pattern can be spotted (Fig.~\ref{Fig09}). While for small number of cells $b$ the estimate gets worse with increasing $\gamma$, the opposite is true for large $b$. For large values of $\gamma$ the estimate for $\alpha$ converges quite rapidly. For $\gamma=9$ and $\gamma=20$ we find on the largest systems $\alpha=0.399$, suggestive of a convergence towards 0.4. For $d=4$ the sizes that can be simulated are too small to allow the determination of the asymptotic value of $\alpha$. However, it is clearly seen that $\alpha$ does not go to zero as $b$ is increased. The same type of behavior is found by using an exponential parametrization of the dynamics \begin{equation} r_i = 1 + x_k \sum_{j n.n. i} \exp\left\{\gamma \cdot \max\left[0,h(j)-h(i)\right]\right\}-1. \label{exp} \end{equation} In $d=2$ for large $\gamma$ the estimate of $\alpha$ on the largest system is 0.399, exactly as with Eq.~(\ref{gamma}). In $d=4$ again we cannot precisely determine where $\alpha$ is converging to. Again the data strongly suggest that this limit is finite. The study of these two alternative parametrizations of the dynamics consistently indicates a value of $\alpha=0.399$ in $d=2$ and a finite $\alpha>0$ in $d=4$ suggesting strongly that 4 is not the upper critical dimension of the KPZ. One could in principle imagine a parametrization of the effective dynamics, more in the spirit of the KPZ equation, of the type \begin{equation} r_1 = 1 + \nu_k |\nabla^2 h(i)| + \lambda_k [\nabla h(i)]^2 \label{r_i2} \end{equation} However, there is no reason for believing that such a parametrization would be better for the KPZ rough phase; additional operators are very likely to be present in the scale invariant dynamics. Moreover the dynamics described by Eq.~(\ref{r_i2}) is plagued by numerical instabilities, as pointed out by Bray and Newman\cite{Bray97}. \subsection{The $d\to\infty$ limit and the upper critical dimension.} The results presented so far show that $\alpha>0$ even for large number of cells in $d=4$, thus indicating that 4 is not the upper critical dimension for KPZ growth processes. By using the RG procedure it is actually possible to go beyond this numerical conclusion: {\it The existence of any finite upper critical dimension can be ruled out.} This result is obtained when the function $\omega^2(b,x_k)$ is computed analytically in the large-$d$ limit. The basic fact allowing this calculation is that when $d \gg 1$ one expects $\alpha \ll 1$, which suggests that surface fluctuations are small \begin{equation} \omega(b,x_k) \sim b^\alpha \simeq 1+\alpha \ln b + O(\alpha^2). \end{equation} For small $b$ one may reasonably account for the fluctuations of the interface by considering only two possible values of $h(i)$, $h_0$ (``low sites'') and $h_0+1$ (``high sites'') (Fig.~\ref{Fig11}). Starting from a flat surface ($h(i)=0, \forall i$), one considers growth events occurring according to the rates Eq.~(\ref{r_i}), with the restriction that no block can be deposited on top of an already grown one. Only when the whole layer at height 1 is grown one allows growth to level 2 and so on. This approximation allows the analytical evaluation of $\omega^2(b,k)$, the identification of the fixed points and the study of their stability. We will check a posteriori the consistence of the results with the assumption, and see that the existence of a finite upper critical dimension can be excluded. Let us now present the details of the calculation. Within the ``two layers'' approximation it is convenient to group together all configurations with the same number of high sites: we will call ``state'' $n$ the set of all surface configurations with $n$ sites at height $h_0+1$ and the remaining $b^d-n$ sites at height $h_0$. The state $n=0$, corresponding to a flat surface, is equivalent to the state with $n=b^d$. This classification is useful because the only transitions permitted from state $n$ are those to state $n+1$. The master equation for the probability $\rho_n$ of being in state $n$ (i.e. of having any of the configurations with $n$ high sites) is then greatly simplified \begin{equation} \partial_t \rho_n = \rho_{n-1} r(n-1 \to n) - \rho_n r(n\to n+1). \label{ME} \end{equation} $r(n \to n+1)$ is the average of all the rates~(\ref{r_i}) for the growth processes that transform one configuration with $n$ high sites in one with $n+1$ of them. We can write this quantity in the form \begin{equation} r(n \to n+1) = (b^d-n) + x_k \Omega_n. \end{equation} The first term on the right hand side is simply the total rate of vertical growth [1 in Eq.~(\ref{r_i})] for configurations with $n$ high sites. Observe that it is obviously equal to the number ($b^d-n$) of sites where vertical growth is allowed. $x_k \Omega_n$ is the rate for lateral growth: $\Omega_n$ is the average number of lateral walls in configurations with $n$ high sites. Its precise computation is not easy, since it would require the knowledge of the stationary probability for each configuration belonging to state $n$. However, when $x_k=0$ the computation is trivial since growth occurs only via uncorrelated deposition and high and low sites are randomly distributed. The number of low sites, where growth is allowed, is $b^d-n$; each of them has $2d$ neighbors which are occupied with probability $n/b^d$. Hence the average number of lateral walls is \begin{equation} \Omega_n = (b^d-n) 2d {n \over b^d} \label{Omega_n} \end{equation} The form of $\Omega_n$ for $x\neq 0$ is in general more complicated, but a numerical computation for large dimensions, namely for $d=7$, shows that Eq.~(\ref{Omega_n}) is a good approximation for all values of $x_k$ (Fig.~\ref{Fig12}). We assume the validity of Eq.~(\ref{Omega_n}) for all values of $x_k$. This leads to \begin{equation} r(n \to n+1) = (b^d-n) \left[1 + 2d {n \over b^d} x_k \right]. \end{equation} The stationary solution of Eq.~(\ref{ME}) is \begin{equation} \rho_n = \rho_{0} \frac{r(0 \rightarrow 1)} {r(n \rightarrow n+ 1)}, ~~~~n=1,\ldots,b^d-1. \label{ro} \end{equation} By imposing the normalization condition $\sum_{n=0}^{b^d-1} \rho_n =1$ and approximating the sum by an integral one obtains \begin{equation} \rho_0 = \left\{ 1 + \frac{b^d}{2 d x_k} \left[ 2 d \ln b + \ln \left(\frac{1+2dx_k}{b^d+2 d x_k} \right) \right] \right\}^{-1}. \label{r0} \end{equation} Equations (\ref{ro}) and (\ref{r0}) provide a complete description of the stationary probability density. Given that the roughness of all configurations with $n$ high sites is $n/b^d (1-n/b^d)$, the total roughness of the surface can be computed as \begin{equation} \omega^2(b,x_k) = \sum_{n=0}^{b^d-1} \rho_n \left(1-{n \over b^d}\right) {n \over b^d}. \end{equation} Using the fact that $b^d \gg 1$ and assuming $d x_k \gg 1$, we obtain \begin{equation} \omega^2(b,x_k) = \rho_0 {b^d \over 2 d x_k}. \label{omega2} \end{equation} Inserting Eq.~(\ref{omega2}) with $b=2$ and $b=4$ in the fixed point equation~(\ref{FP}) yields, to leading order in $d$, \begin{equation} x^* = 2^{d+1} \ln 2. \label{x*d} \end{equation} The assumption $d x_k \gg 1$ is therefore self-consistent for sufficiently large $d$. Notice that an exponential dependence of $x^*$ on $d$ was already found in the numerical implementation of the method (Fig.~\ref{Fig08}). Using Eq.~(\ref{alpha}) one obtains the value of the roughness exponent \begin{equation} \alpha \simeq {1 \over 3(\ln 2)^2} {1 \over d}. \label{alphad} \end{equation} Finally, by computing the derivative of the RG transformation at the fixed point \begin{equation} R'(x^*) = -1 + {1 \over 2 \ln 2} {1 \over d} + O(1/d^2), \end{equation} we see that the fixed point is attractive for all finite dimensions. In conclusion we find that for large-$d$ the RG has a fixed point $x^*$ corresponding to an exponent $\alpha \sim 1/d$ and therefore strictly positive in all finite dimensions. On the contrary the existence of a finite upper critical dimension would have implied, for $d>d_c$, either the absence of a finite fixed point or its instability. At this point we must use the analytical result to check the consistency of the two layers assumption. Such assumption is correct provided the rate of its violation is negligible for all values of $n$. Processes violating the assumption are those in which an event of vertical growth occurs on top of an high site. Their rate in state $n$ is $r_{up}=n$, that must be compared to the total rate of processes not violating the restriction $r(n \to n+1)$ computed for $x=x^*$. By imposing $r_{up}(n) \ll r(n \to n+1)$ we get \begin{equation} n \ll (b^d-n) \left(1+{2dn \over b^d} x^*\right). \end{equation} Let us consider $n=b^d-1$ which is the situation that maximizes $r_{up}$ and minimizes $r(n \to n+1)$. Then \begin{equation} b^d-1 \ll 1+{2d (b^d-1) \over b^d} x^*. \end{equation} Since $b^d \gg 1$, this means \begin{equation} b^d \ll 2dx^* \sim 2^{d+2} d. \end{equation} Hence the two layers assumption is correct for $b=2$ but fails for $b=4$. Therefore the value of $\omega^2(4,x_k)$ is systematically underestimated by Eq.~(\ref{omega2}) since fluctuations involving more than two layers are neglected despite being likely. The consequence of this on our results is understood by considering Eq.~(\ref{RGeq}). In such a formula we estimate correctly the left hand side, while the right hand side is underestimated (Fig.~\ref{Fig13}). Since the fixed point parameter $x^*$ and the exponent $\alpha$ are given by the intersection of the curves it is clear that we get an upper bound for $x^*$ and a lower bound for $\alpha$. This is confirmed by the comparison of our estimate of $\alpha$, Eq. (\ref{alphad}), and $x^*$ with the numerical results of Ala-Nissila {\em et al.} and the value of $x^*$ computed numerically for $d=1,\ldots,9$ (Fig.~\ref{Fig14}). These results have been obtained for the smallest values of $b$, namely $b=2$. In the previous sections we showed that for low dimensions the results for small $b$ are in good agreement with numerics, but for larger $b$ there are deviations. A very reasonable question therefore concerns the robustness of the large-$d$ results when $b$ grows. As we have shown above the two layers approximation breaks down for $b>2$. In order to extend the above calculation to larger $b$ one should replace the two layers approximation with some less restrictive but still doable calculational scheme. We have not been able to fulfill this task and hence we cannot directly show whether a fixed point exists for finite $x^*$ when $b \to \infty$. However, the two layers assumption is valid for any $b$ in the neighborhood of the fixed point $x^*=\infty$ that gives a flat surface $\alpha=0$. This fixed point exists in any dimension, and its stability can be safely analyzed using the previous two layer assumption as follows. Let us introduce $\epsilon=1/x$. The derivative of the RG transformation at the fixed point $\epsilon^*=0$ is (see Eq.~(\ref{Ralpha})) \begin{equation} R'(\epsilon=0)= {2 \alpha'_b(\epsilon=0) \over \alpha'_{b^2}(\epsilon=0)}-1 \end{equation} where now prime indicates derivative with respect to $\epsilon$. To first order in $\epsilon$ we have (see Appendix II) \begin{equation} \alpha_{b^2}(\epsilon) = {1 \over 4 \ln b} \ln \left[ 1 + c \mu \epsilon b^{2d+2} \right]. \label{eq54} \end{equation} Then \begin{equation} \alpha'_{b^2}(\epsilon=0) = {1 \over 4 \ln b} c \mu b^{2d+2}. \end{equation} Analogously \begin{equation} \alpha'_b(\epsilon=0) = {1 \over 2 \ln b} c \mu b^{d+1}. \end{equation} Hence \begin{equation} R'(\epsilon=0) = b^{d+1}-1 \gg 1 \end{equation} and the fixed point corresponding to $\alpha=0$ is unstable. As a consequence, a finite fixed point with $\alpha>0$ must exist and be stable when $b \to \infty$ for any large and finite $d$. This supports the conclusion that there is no finite upper critical dimension. \section{RG for the Edwards-Wilkinson dynamics} So far we have applied the new RG method to a KPZ-like dynamics. Now we intend to show that it is more general and can be applied to growth mechanisms belonging to universality classes other than KPZ. In particular we study in this section its application to the exactly solvable, Edwards-Wilkinson (EW) equation for which the roughness exponent is known in any dimension. In particular $\alpha=1/2$ in $d=1$, while $\alpha=0$ for $d \geq 2$ with logarithmic corrections at $d=2$. The parametrization Eq.~(\ref{r_i}) of the dynamics describes a growth model where only deposition events can take place and the symmetry between up and down in the $h$ direction is clearly broken. Such a dynamics is inherently out of equilibrium and therefore cannot accommodate the scale invariant dynamics of the Edwards-Wilkinson growth process, which is an equilibrium one, with growth rules symmetric along the growth direction. We now introduce a generalized dynamics which admits the KPZ and EW dynamics as particular limiting cases. Let us consider the quantities \begin{eqnarray} K_d(i) & = & \sum_{j nn i} \max[0,h(j)-h(i)] \nonumber \\ K_u(i) & = & \sum_{j nn i} \max[0,h(i)-h(j)]. \end{eqnarray} In the KPZ case described so far we have allowed only deposition of particles and written \begin{equation} r_i = 1+x_k K_u(i). \end{equation} We now allow also for evaporation of particles. That is, we consider the transition rate for site $i$ as \begin{equation} r_i = 1+x_k |\epsilon K_u(i)-(1-\epsilon) K_d(i)| \end{equation} and with probability \begin{equation} P_b=1/r_i \end{equation} a random deposition/evaporation event takes place ($h_i \to h_i +1$ with probability $\epsilon$ and $h_i \to h_i-1$ with probability $1-\epsilon$), while with probability \begin{equation} P_l=x_k |\epsilon K_u(i)-(1-\epsilon) K_d(i)|/r_i \end{equation} we have a ``lateral'' event \begin{equation} h_i \to h_i +{\epsilon K_u(i)-(1-\epsilon) K_d(i) \over |\epsilon K_u(i)-(1-\epsilon) K_d(i)|}. \end{equation} For $\epsilon=1$ only deposition is allowed and we have the transition rates for the KPZ dynamics. For $\epsilon=1/2$, we have up-down (deposition-evaporation) symmetry and the rates are \begin{equation} r_i = 1+x_k |\nabla^2 h(i)| \end{equation} where $\nabla^2 h(i)=[K_u(i)-K_d(i)]/2$ is the discretized Laplacian evaluated at site $i$. Therefore we expect this case to correspond to EW dynamics. Let us verify that for $\epsilon=1/2$ the average interface velocity does not depend on the interface configuration (which is a basic property of EW dynamics) \begin{equation} v = {1 \over L^d} \sum_i v_i = {1 \over L^d} \sum_i r_i \Delta h_i. \end{equation} Since \begin{eqnarray} \Delta h_i & = & P_b \cdot 0 + P_l \cdot {K_u(i)-K_d(i)\over |K_u(i)-K_d(i)|} \nonumber \\ & = & {1 \over r_i} \left(x |K_u(i)-K_d(i)| { K_u(i)-K_d(i)\over |K_u(i)-K_d(i)|} \right) \nonumber \\ & = & {1 \over r_i} \left(x [K_u(i)-K_d(i)] \right) \end{eqnarray} and as $\sum_i K_u(i) = \sum_i K_d(i)$, we have that \begin{equation} v = {x \over L^d} \sum_i [K_u(i)-K_d(i)] = 0. \end{equation} With this generic dynamics we can perform the RG procedure exactly as in the KPZ case. The evaluation of the function $\omega^2(b,x)$ is carried out again using small Monte Carlo simulations with periodic boundary conditions. The results are reported in Fig.~\ref{Fig15}. For $d=1$ the value of $\alpha$ for $b=2$ is $\simeq 0.4$ below the exact value $1/2$, but for $b>2$ the correct value is rapidly approached. The situation is completely different in $d=2$. In such case for $b=2$ the exponent $\alpha$ is around 0.25, but when $b$ is increased, the fixed point is shifted monotonically towards $\infty$. In $d=3$ the behavior of the fixed point for finite $x^*$ is similar. From these plots we can conclude that the behavior of the EW dynamics is very different in $d=1$ and $d>1$. For $d=1$ there is a stable fixed point with $\alpha=1/2$. For $d>1$ there is no stable fixed point with $\alpha \neq 0$. Hence the RG method is able to capture the difference between KPZ and EW dynamics and correctly describes both the rough and the flat phases of the Edwards-Wilkinson growth. We speculate that the reason why one needs to consider $b>2$ is probably related to the fact that on a system of size $b=2$ the discretized Laplacian and square gradient take the same form. \section{Discussion} In the previous sections we have introduced a general method for studying surface growth models by means of a real space renormalization group procedure. The anisotropy of the scale invariant properties of surfaces makes the very definition of the RSRG highly nontrivial, since the direct integration of degrees of freedom at small scale cannot be performed. For this reason we had to devise an alternative route: the main ingredient is that the integration of degrees of freedom is performed implicitly by imposing the self-consistency between two descriptions of the same object at different levels of coarse-graining. The application of such an approach to the KPZ dynamics yields several results that can be summarized as follows. \begin{itemize} \item The scale invariant dynamics is identified and parametrized as a function of the ``lateral growth'' parameter $x$. This parameter turns out to have a non-trivial attractive fixed point under RG transformation for all dimensions. \item The KPZ roughness exponent $\alpha$ estimated for small $b$ is in very good agreement with large scale simulations of discrete models. \item For larger values of $b$ the estimate of $\alpha$ is stable in $d=1$ while it changes noticeably for $d \geq 2$, presumably owing to crossover effects. When $b \to \infty$ it converges towards the correct result. \item The results are robust with respect to changes in the parametrization of the dynamics and in the boundary conditions. \item No evidence is found of the existence of an upper critical dimension for KPZ. Moreover, we show very strong evidence that no such an upper critical dimension exists. \item By changing the nature of the parametrization of the dynamics at generic scale, the method is able to describe the EW dynamics and capture the existence for it of an upper critical dimension above which only a trivial (flat) phase exists. \end{itemize} Regarding the general nature of the approach it is worth remarking that the key point in the method is the identification of the scale invariant dynamics. In some sense the procedure can be seen as a kind of finite size scaling approach allowing for the evaluation of scaling exponents via the extrapolation of small size MC simulations. However the crucial point is that the MC data do not directly determine the exponent; they rather allow the identification of the scale invariant parameters of the dynamics which in turn determines the exponent. With respect to the estimates of the roughness exponent for small $b$, it is remarkable that the accuracy (in the sense of the discrepancy with known numerical results) seems to be the same in all dimensions $d \geq 2$. This is not the typical situation in ordinary critical phenomena, where usually RSRG methods fail in high dimensions. There are at least two reasons why usual RSRG schemes are inaccurate in high dimensions: a) The necessity of defining an explicit geometrical mapping between degrees of freedom at two scales (spanning rule, majority rule, bond-moving, etc.). b) The presence of relevant fields and the need to compute exponents from the derivatives of the RG transformation at the fixed point. Due to the success of field theory in high dimensions (close to $d_c$) in usual critical phenomena, RSRG methods have been mostly devised to work in low dimensions where the predictions of the $\epsilon$-expansion become less reliable. RSRG methods based on an explicit geometric mapping (block-spin transformation) are quite accurate (and sometimes even exact) close to the lower critical dimension. Problems related with this geometric transformations become worse and worse as the dimension increases. In our perspective, the only limitation has to do with the quality of the parametrization of the RG transformation. For example, the parametrization of the RSRG transformation for ferromagnetic systems based on the Migdal transformation of Ising spins gives inaccurate results in high $d$. However, if one uses the parametrization of the $\phi^4$ theory, one recovers $d_c=4$ within the RSRG Migdal approach \cite{LM} even for ferromagnetic systems. In any case, notice that our RSRG method does not need an explicit geometrical definition of the RG transformation. Therefore it bypasses the problems related to a). In some sense, this is similar to the phenomenological RG method \cite{DV} where the RG transformation is defined implicitly through finite size scaling arguments. Remarkably, phenomenological RG calculations are quite accurate. With respect to point b), as discussed above the absence of relevant fields makes truncation errors much less important than in ordinary applications of the RG. Furthermore we note that in the KPZ problem one has to compute exponents depending only on the RG transformation at the fixed point, i. e. the critical parameter. This is profoundly different from what happens in Ising-like problems, where some exponents (as, for example, the correlation length exponent $\nu$) depend on the derivatives of the RG transformation around it: As a consequence $\nu$-type exponents are rather difficult to estimate since, even if the location of the fixed point is determined accurately, the computation of the derivatives is much less precise. No exponents of such type exist in the KPZ case. This is, in our opinion, a further reason for the great accuracy of the new method with respect to the usual RSRG. As a final point, it is worth discussing the current limitations of the method. It is clear from the results presented, that a most important role in the method is played by the choice of the parametrization of the scale invariant dynamics. This is particularly true since one deals with a monoparametric description of the growth process: If one could easily introduce several parameters and study their flow under the RG, the stability of the results when details are changed would be greatly improved. Within the present framework, the inclusion of additional parameters is however not straightforward. The problem is that additional RG equations are provided only by the use of equation~(\ref{main}) with $b$, $b^2$, $b^3$ and so on. This requires the computation of $\omega^2(b,k)$ on systems whose size becomes quickly prohibitively large. A remarkable improvement of the method would therefore be the identification of additional RG transformations independent from Eq.~(\ref{main}). Despite this difficulty, we believe that the theoretical framework presented here constitutes an important new element in the field of surface growth and deserves further investigation, in particular with respect to possible applications to other open problems, and generalization to deal with time-dependent properties. In summary, in this paper we have presented a real space renormalization group method developed to deal with surface growth processes. The new method overcomes the difficulties inherent to standard real space renormalization group analysis of anisotropic situations. It is based on the definition of anisotropic blocks of generic scale and of a parametrized effective dynamics for the evolution of such blocks. Imposing the surface width to be the same when using different scales (different block sizes) we write a renormalization group equation. Its associated fixed points define the scale invariant effective dynamics and permit to determine the roughness exponent $\alpha$. We have employed the new method to study the Kardar-Parisi-Zhang and the Edwards-Wilkinson universality classes. In particular, for KPZ we compute the $\alpha$ exponent in dimensions from $d=1$ to $d=9$. The results are in very good agreement with the best numerical estimates in all dimensions. Moreover, we present analytical calculation excluding the possibility of KPZ having a finite upper critical dimension. On the other hand, well known results for the EW universality class are obtained, confirming the generality of the method. \section{Acknowledgments} We acknowledge interesting discussions with A. Gabrielli, A. Maritan, G. Parisi, A. Stella, C. Tebaldi, G. Bianconi, and A. Vespignani. This work has been partially supported by the European network contract FMRXCT980183, and by a M. Curie fellowship, ERBFMBICT960925, to M. A. M.
\section{Introduction} Electromagnetic emission generated by relativistic electrons moving in a magnetic field is referred to as {\it synchrotron} radiation if the electrons are ultra-relativistic, and as {\it gyrosynchrotron} if the electrons are mildly relativistic. In the present paper we deal only with gyrosynchrotron radiation, using the formalism developed by Ramaty \shortcite{R}, except that we also include the longitudinal component of the radiation which is important at low multiples of the cyclotron frequency. We also take into account the correction pointed out by Trulsen and Fejer \shortcite{Trulsen}. This formalism has been tested by other workers (e.g. Holman \& Benka 1992) and was found to be accurate. Gyrosynchrotron absorption can become negative and this amplification process is referred to as the Electron-Cyclotron-Maser (ECM). \par The observation of millisecond microwave spikes from the Sun (reviewed by Benz 1986) has been interpreted as gyrosynchrotron maser emission since the late 1970`s (Wu \& Lee 1979; Holman, Eichler \& Kundu 1980), and it is now the most widely accepted explanation for the spike emission \cite{BenzBook}. Spikes have been observed in the frequency range 0.3--8 GHz, exhibit a duration of a few milliseconds to a hundred milliseconds, a narrow bandwidth (0.5--2 per cent or a few MHz), and until recently have not been resolved in angle \cite{Benzetal}. By assuming that the homogeneity scale of the magnetic field is of order $10^7$ cm, and that this is the size of the source, it is possible to deduce a characteristic brightness temperature of $10^{15-18}$ K . \par In order that ECM appears the electron number distribution must either increase towards higher energies or an anisotropy in pitch angle must exist, as in a loss cone distribution. A loss cone forms naturally where the magnetic field increases, as electrons with small pitch angles continue and electrons with large pitch angles mirror. At the footpoints of flare loops the density becomes high enough that electrons continuing into the chromosphere lose their energy and do not reflect back upwards, and thus a loss-cone distribution appears (Aschwanden \& Benz 1988; Kuncic \& Robinson 1993). We therefore expect maser emission from the region of the footpoints. \par The study of the ECM mechanism is of interest because of the radio emission, which serves as a probe into conditions in the flare, and also because of the energetic particle precipitation which produces X-rays, and for local heating of coronal material. The ECM reduces the trapping time and increases the flux of precipitating particles by causing a reduction in the pitch angle of electrons within a short distance \cite{BenzBook}. Electrons which otherwise would be mirrored back and forth are thus precipitated into the chromosphere after one reflection. The maser thus transfers energy from the fast particle distribution to the ambient plama, and serves as a heating mechanism. \par Several papers have explored the possibilities of ECM. Melrose and Dulk \shortcite{D+M} have used a semi-relativistic approximation to derive approximate formulae for the frequency of the largest growth rate, and for the growth rate. Winglee \shortcite{Winglee} approximated the effects of the ambient plasma temperature on the maser emission. Aschwanden (Aschwanden 1990a,b) followed the diffusion of electrons into the loss-cone as a result of the maser emission, and therefore the self-consistent closure of the loss-cone as kinetic energy is transformed to radiation through the maser. Aschwanden \shortcite{Aschwanden b} concluded that the millisecond spikes are produced in the region of $\nu_{\rm p}/\nu_{\rm B}<0.25$. Kuncic and Robinson \shortcite{Kuncic} performed ray-tracing in a model loop with dipole field, and concluded that maser emission can escape from lower levels. Fleishman and Yastrebov \shortcite{Fleishman1} present results which show simultaneous amplification in two or more frequencies. \par An important difference in our work is that all previous studies concentrated on the {\it growth rate} of the amplified wave, and defined the {\it dominant mode} as the particular wave -- frequency $\nu$ and direction to the magnetic field $\theta$ -- with the largest growth rate. However, the important criterion for long-term emission is the {\it saturation length}, which is the distance beyond which the maser ceases to amplify, and not the growth rate \cite{BenzBook}. We compute the absorption coefficient and, using our model, the saturation length, and determine which wave will be observed, and at what intensity. \par Another difference is that most previous studies used soft energy distributions for the fast electrons, such as the Dory, Guest, Harris distribution \cite{Winglee2} or a hot thermal component distribution (Aschwanden 1990a,b). However, observations of Microwaves (MW) and X-rays from solar flares indicate a power law energy spectrum with energy up to 10 MeV and even higher, and with a power index typically $\delta=3$, and in the range $\delta=2-6$ (Ramaty 1969; Zirin 1989; Fleishman \& Yastrebov 1994b for other references). We therefore chose to use a power law distribution function for our fast non-thermal electrons. We also use an idealized one sided loss-cone distribution \cite{D+M} while other studies use a one sided $sin^N(\phi)$ distribution (Aschwanden 1990a,b) or a two-sided gaussian (Fleishman \& Yastrebov 1994a,b) pitch angle distribution. \par The ratio $\nu_{\rm p}/\nu_{\rm B}$, where $\nu_{\rm p}=\sqrt{n_e/m/\pi}\ e$ is the plasma frequency and $\nu_{\rm B}=eB/(m\gamma c)/(2\pi)$ the cyclotron frequency, is an important parameter for ECM emission. It is used to distinguish regions where different magneto-ionic modes dominate. From the previous studies we conclude that the dominant mode depends on the distribution function of the fast electrons. In some studies the fundamental eXtraOrdinary (XO) mode is dominant for $\nu_{\rm p}/ \nu_{\rm B} < 0.3$ \cite{B+A2}, while in others the fundamental Ordinary Mode (OM) is dominant for $\nu_{\rm p}/ \nu_{\rm B} < 0.5$ \cite{Winglee2}. In some the fundamental OM is dominant for $0.3 < \nu_{\rm p}/\nu_{\rm B}<1$ \cite{B+A2}, and in others the second harmonic XO is dominant for $0.5 < \nu_{\rm p}/\nu_{\rm B}<1.5$ \cite{Winglee2}. The unobservable Z-Mode (ZM) is dominant for $\nu_{\rm p}/\nu_{\rm B} > 1.5$ \cite{Winglee2}, or even for $\nu_{\rm p}/\nu_{\rm B} > 0.5$ \cite{Fleishman2}. In our work the ZM is the most highly amplified mode (i.e. with the largest in absolute value negative absorption) for the range $0.5 < \nu_{\rm p}/\nu_{\rm B} < 1.1$ and for $1.8 < \nu_{\rm p}/\nu_{\rm B}$, similar to what was found by others who have used a power law \cite{Fleishman2}. \par The high amplification of the ZM would seem to indicate that this unobservable mode is the dominant mode, and would therefore preclude the maser mechanism as the producer of MW spikes, except in the narrow range $1.1 < \nu_{\rm p}/\nu_{\rm B}<1.8$. We show that for the range $\nu_{\rm p}/ \nu_{\rm B} < 2$ it is possible for the OM or the XO mode to be amplified enough that they become observable. Our absorption-coefficient approach allows a simple procedure to estimate the saturation length, which is difficult to derive from growth rate calculations, and is the important parameter in deciding which mode dominates \cite{BenzBook}. Our derived saturation lengths are consistent with the assumed homogeneity of the magnetic field and density. We therefore conclude that the Electron-Cyclotron-Maser mechanism can produce microwave spikes. \par In section 2 we give the theory and equations for gyrosynchrotron emission and wave propagation through cold-plasma. In section 3 we show that the gyrosynchrotron is the important process in absorption as well as emission, for the frequencies and plasma parameters we assume. Section 4 describes our estimate of the energy available to the loss-cone, and a derivation of the saturation length is given. Section 5 describes our calculations and the parameters we used. In section 6 we discuss the observability of the maser emission under different conditions, in section 7 we give a comparison with previous studies, and in section 8 we summarize our results. \section{Gyrosynchrotron Absorption} Solar flares are known to be a complex phenomena. The solar flare plasma is composed of both an ambient thermal plasma, and a population of fast particles. It is customary to simplify and assume that any waves which are excited are a phenomena of the ambient plasma alone. The properties of the ambient plasma therefore determine which waves are excited and how the waves propagate. However, the fast particles can absorb waves present in the ambient plasma, and also emit in these waves` frequencies. We used the cold plasma approximation to describe the ambient plasma, and assumed the waves which result from magnetoionic theory \cite{MelroseB}. For a given electron distribution $f({\bmath p})$ and wave $(\nu,\theta)$ it is then possible to compute the emission and absorption coefficients, which describe the interaction of the wave with the distribution. \par The cold plasma parameters which determine the propagation are the refraction index and the components of the polarization vector \cite{MelroseB}: \begin{eqnarray} \label{polvec} &&\hat\epsilon_\sigma = { {L_{\sigma}({\bmath k}){\hat k} + T_\sigma({\bmath k}) {\bmath t} + i{\bmath c}} \over \sqrt{1+L_\sigma^2+T_\sigma^2} } \end{eqnarray} where $\bmath k$ is the wave vector,$\hat k$ a unit vector in the $\bmath k$ direciton, $\bmath c$ a unit vector perpendicular to $\bmath k$ and to the magnetic field ${\bmath B}$, and ${\bmath t}$ is a unit vector perpendicular to ${\bf k}$, in the plane of ${\bf k}$ and ${\bf B}$. Also $T_\sigma$ and $L_\sigma$ are functions depending on $\bmath k$, and $\sigma$ is `+` (OM) or `-` (XO mode). The transverse component of the polarization is given by $T_\sigma$ and the longitudinal component is described by $L_\sigma$. We assumed ${\bmath B}$ to be in the $\hat z$ direction, and ${\bmath k}$ to be in the $(x,z)$ plane with angle $\theta$ to the magnetic field. With these definitions ${\bmath c}=\hat y$ , $\hat k=(\sin(\theta),0,\cos(\theta))$, and ${\bmath t}=(\cos(\theta),0,-\sin(\theta))$. \par The cold plasma modes are the OM, which exists for \begin{equation} \nu>\nu_{\rm p} \end{equation} the XO which exists for \begin{equation} \nu>\nu_{\rm x}=\nu_{\rm B}/2 + \sqrt{\nu_{\rm p}^2+\nu_{\rm B}^2/4} \end{equation} and the low frequency branch of the XO, called here the ZM, which exists for \begin{eqnarray} \nonumber \nu_{\rm x}-\nu_{\rm B} & < &\nu < \sqrt{0.5} \times \\ &&\nonumber \sqrt{ (\nu_{\rm p}^2+\nu_{\rm B}^2)+ [(\nu_{\rm p}^2+\nu_{\rm B}^2)^2-4\nu_{\rm p}^2\nu_{\rm B}^2\cos^2(\theta)]^{0.5} } \end{eqnarray} and is bound from above by the upper-hybrid frequency \begin{equation} \nonumber \nu_{\rm uh}=\sqrt{\nu_{\rm B}^2+\nu_{\rm p}^2} \end{equation} The Whistler mode exists for \begin{equation} \nonumber \nu< {\sqrt{ { (\nu_{\rm p}^2+\nu_{\rm B}^2)- [(\nu_{\rm p}^2+\nu_{\rm B}^2)^2-4\nu_{\rm p}^2\nu_{\rm B}^2\cos^2(\theta)]^{0.5}}\over{2} } } \end{equation} and is a low frequency mode which we ignore \cite{MelroseB}. \par The equations for the cold-plasma parameters are (Ramaty 1969; Melrose 1989) \begin{eqnarray} \label{Tpm} T_\pm &=& [2\nu (\nu_{\rm p}^2-\nu^2)\cos(\theta)] \times \\ &&\nonumber \big[ -\nu^2\nu_{\rm B} \sin^2(\theta) \pm \\ &&\nonumber \sqrt{ \nu^4\nu_{\rm B}^2\sin^4(\theta)+ 4 \nu^2(\nu_{\rm p}^2-\nu^2)^2\cos^2(\theta) }\ \big] ^{-1} \end{eqnarray} \begin{equation} \label{Lpm} L_\pm={{\nu_{\rm p}^2\nu_{\rm B}\sin(\theta)}\over{\nu^3-\nu\nu_{\rm p}^2}} {{\nu T_\pm}\over{\nu T_\pm-\nu_{\rm B}\cos(\theta)}} \end{equation} And the refraction index $n_{\rm ref}$, also referred to as $n_\pm$, is given by \begin{eqnarray} n_\pm^2 &=& 1 + [2\nu_{\rm p}^2(\nu_{\rm p}^2-\nu^2)] \times \\ &&\nonumber \big[ -2\nu^2(\nu_{\rm p}^2-\nu^2)-\nu^2\nu_{\rm B}^2\sin^2(\theta) \pm \\ &&\nonumber \sqrt{\nu^4\nu_{\rm B}^4\sin^4(\theta)+ 4\nu^2\nu_{\rm B}^2(\nu_{\rm p}^2-\nu^2)^2\cos^2(\theta)}\ \big] ^{-1} \end{eqnarray} With the symbols being $\nu_{\rm p}$ the plasma frequency, $\nu_{\rm B}$ the cyclotron frequency, and $\nu$ the emitted frequency. \par The cold plasma approximation is valid only if the two following criteria are are true \cite{MelroseB} \begin{equation} \label{criteria1} \left({{\nu- s\nu_{\rm B}}\over{n_{\rm ref}\nu\cos(\theta)}}\right)^2 >> {{k_{\rm B} T}\over{m_e c^2}} = 0.001683 {T\over{10^7\ {\rm K}}} \end{equation} and \begin{equation} \label{criteria2} \left({{\nu_{\rm B}}\over{n_{\rm ref}\nu\sin(\theta)}}\right)^2 >> {{k_{\rm B} T}\over{m_e c^2}}= 0.001683 {\rm T\over{10^7\ K}} \end{equation} Where we have used $k_{||}=n_{\rm ref}\omega\cos(\theta)/c$ and $v_{\rm thermal}=\sqrt{k_{\rm B} T /m_e}$ to get the above from Melrose`s equations. The $s$ multiplier in the first criterion is an integer such that $s\nu_{\rm B}$ is the closest cyclotron harmonic to $\nu$. \par Our rule for when the cold plasma approximation ceases to be valid is when the left side of the criteria is less then ten times larger than the right side. The cold plasma approximation fails near the resonances of the cold plasma modes at the cyclotron harmonics, and for the low-frequency Z-mode when the frequency of the waves is near the upper-hybrid frequency. \par Most authors calculate the growth rate, but our derivation follows Ramaty`s \shortcite{R} in calculating the absorption coefficient which is (Ramaty 1969; Aschwanden 1990a): \begin{eqnarray} \label{gyroe} k_\pm \left( {\nu ,\theta } \right)& = & {1 \over B}4\pi ^2e {{2\pi } \over {\left| \cos \left( \theta \right) \right|}} {{\nu_{\rm B}} \over \nu }{1 \over {n_\pm ^2}} \int_1^\infty {\rm d}\gamma {{u\left( \gamma \right)} \over \beta }\times \\ &&\nonumber \sum\limits_{\rm s=s_{\rm 1}}^{\rm s_{\rm 2}} {{g\left( {\phi_{\rm s}} \right)} \over {1+T_\pm ^2}} \Bigl[ \beta \sin \left( {\phi_{\rm s}} \right)J_{\rm s}^\prime\left( {x_{\rm s}} \right)+ {{L_\pm}\over{n_\pm}}J_{\rm s}\left(x_{\rm s} \right)+ \\ &&\nonumber T_\pm \left({{\cot \left( \theta \right)} \over {n_\pm }}- {{\beta \cos \left( {\phi_{\rm s}} \right)} \over {\sin \left( \theta \right)}} \right)J_{\rm s}\left( {x_{\rm s}} \right) \Bigr]^2\times\\ &&\nonumber \Bigl[ {{-\beta \gamma ^2} \over {u\left( \gamma \right)}} {{\rm d} \over {{\rm d}\gamma }} {{u\left( \gamma \right)} \over {\beta \gamma ^2}}+ \\ &&\nonumber {{n_\pm \beta \cos \left( \theta \right)-\cos \left( {\phi_{\rm s}} \right)} \over {\gamma \beta ^2\sin \left( {\phi_{\rm s}} \right)}} {1 \over {g\left( {\phi_{\rm s}} \right)}} {{{\rm d}g\left( \phi \right)} \over {{\rm d}\phi }} \Bigr] \end{eqnarray} Where we corrected Ramaty's \shortcite{R} formula by deleting a factor $1+\partial \ln(n_{\rm ref}) / \partial \ln(\nu)$, and changed the minus sign multiplying the $J_{\rm s}^\prime$ term because our $T_\pm$ have the opposite sign to Ramaty`s $a_{\theta\pm}$. We also added the longitudinal contribution which is important for frequencies close to the cyclotron frequency. $J_{\rm s}$ are Bessel functions, $J_{\rm s}^\prime$ their derivatives, with the argument $x_{\rm s}=(\nu/\nu_{\rm B}) n_\pm\gamma\beta\sin(\theta)\sin(\phi_{\rm s})$. The limits of the summation are $s1,2={{\gamma \nu}\over{\nu_{\rm B}}} [1 \pm n_{\rm ref} \beta \cos(\theta)]$, and $\pm$ designates the mode ('+' for OM and '-' for XO). The angle $\phi_{\rm s}$ comes from the resonance condition (Ramaty 1969; Melrose 1989) \begin{equation} \label{resonance} {{s\nu_{\rm B}}\over \gamma}=\nu [1-n_{\rm ref}\beta\cos(\theta)\cos(\phi)] \end{equation} and is derived by \begin{equation} \cos(\phi_{\rm s})={ {1- { {s\nu_{\rm B}}\over{\gamma \nu}}}\over{n_\pm \cos(\theta) \beta} } \end{equation} The distribution function of the electrons was assumed to be separable into the form \begin{equation} \label{dist} f({\bmath p}){\rm d}^3p=2\pi u(\gamma)g(\phi){\rm d}\gamma {\rm d}(\cos(\phi)) \end{equation} Where $\phi$, the pitch angle, is the angle of the velocity to the magnetic field, and $\gamma=(1-\beta^2)^{-0.5}$ the Lorentz factor. \par It is important to note that the equation for the absorption coefficient has the same properties as the equation for the growth rate since the growth rate is \begin{equation} \label{growth} \Gamma_\pm=-k_\pm\cdot v_{\rm g} \end{equation} where $v_{\rm g}$ is the group velocity given by \cite{MelroseB} \begin{equation} \label{vg} {\bmath v}_{\rm g} = {{\partial\omega}\over{\partial k}}= {c\over {{\partial(\omega n_{\rm ref})}\over{\partial\omega}}} \left(\hat k-{{L_\sigma T_\sigma}\over{1+T_\sigma^2}} {\bmath t}\right) \end{equation} As in all previous studies, we neglected the angle between the group velocity and the wave vector. The group velocity in our study is therefore assumed to be in direction $\hat k$, and of magnitude \begin{equation} \nonumber v_{\rm g} ={c\over {{\partial(\omega n_{\rm ref})}\over{\partial\omega}}} \end{equation} \par In this work we assumed the power-law electrons have a power-law dependence on kinetic energy \begin{equation} \label{powerlaw} u(\gamma)=n_{\rm hot}\cdot(\delta-1)\cdot(\gamma_{\rm min}-1)^{\delta-1}\cdot (\gamma-1)^{-\delta} \end{equation} where $n_{\rm hot}$ is the number density of the fast electrons, and the power law starts from a low-energy cut-off $\gamma_{\rm min}$ and ends at a high-energy cut-off $\gamma_{\rm max}$. The distribution is normalized to $n_{\rm hot}$ when we neglect $(\gamma_{\rm min}-1)/(\gamma_{\rm max}-1)$. For the ambient plasma we used a thermal distribution \cite{Lang} \begin{equation} \label{thermal} u(\gamma){\rm d}\gamma=n_{\rm cold} {2 \over \sqrt{\pi}} \left({{mc^2}\over{kT}}\right)^{3/2} \sqrt{\gamma-1}\times {\rm e}^{-{{mc^2}\over{kT}}(\gamma-1)} {\rm d}\gamma \end{equation} where $n_{\rm cold}$ is the number density of the cold, ambient plasma, electrons. \par For the pitch angle dependence in the fast distribution we used an ideal loss-cone \cite{D+M} \begin{equation} \label{losscone} g(\phi)=\cases{ Const., &$\cos(\phi) < \cos(\alpha)$ \cr Const.*{ {cos(\alpha-\Delta\alpha)-cos(\phi)} \over {cos(\alpha-\Delta\alpha) - \cos(\alpha)} }, &$\cos(\alpha) < \cos(\phi) < $ \cr &$\cos(\alpha-\Delta\alpha)$ \cr 0, & $\cos(\alpha-\Delta\alpha) < \cos(\phi)$ \cr} \end{equation} This describes an isotropic distribution for pitch angles above some $\alpha$, no electrons below another angle $\alpha - \Delta \alpha$, and a linear in $cos(\phi)$ decrease within the $\Delta \alpha$ interval. The normalizing constant in $g(\phi)$ depends on the loss cone parameters. For an idealized loss cone as above it is $Constant =1 / \pi /(2+\cos(\alpha)+\cos(\alpha-\Delta\alpha))$. With this definition the integral is normalized so that $2\pi \int_{-1}^{1} g(\phi) {\rm d}(\cos\phi)= 1$. \par For the thermal plasma the distribution is isotropic and therefore $g(\phi)=1/4/\pi$. \par \section{Other Processes} There are other processes in a thermal plasma which have to be taken into account. The free-free absorption coefficient for $T > 3\cdot 10^5$ K is \cite{Lang} \begin{eqnarray} \label{freefree} k_{\rm ff}= 0.01 {{n_e^2}\over {\nu^2 T^{3/2}}} \left(24.57+ \ln{{T}\over {\nu}} \right) \end{eqnarray} where $n_e$ is the density of the thermal electrons in ${\rm cm^{-3}}$, $\nu$ the frequency of the absorbed radiation in Hz, and $k_{\rm ff}$ the absorption coefficient in ${\rm cm^{-1}}$ . \par Our assumed number density is of order $n_e \le 10^{11}$, our standard temperatures are of order $T \approx 10^7$, and the emission frequency for our standard values is of order $\nu \approx 1 {\rm GHz}$. The free-free absorption for these values is $k_{\rm ff} \approx 5\cdot 10^{-8}$. For a standard number of fast electrons of order $n_{\rm hot} \approx 10^6\ {\rm cm^{-3}}$ the free-free absorption normalized to one electron is $k_{\rm ff}/n_{\rm hot} < 5\cdot 10^{-14}$, which is negligible compared to the gyrosynchrotron absorption of the dominant modes. \par The deflection time for a particle in a thermal plasma, with a particle velocity much higher than the thermal velocity but still non-relativistic, is \cite{BenzBook} \begin{equation} t_{\rm d} = 3\cdot 10^{-20} {{v^3}\over{n_e}} \end{equation} where $v$ is the particle velocity, and $t_{\rm d}$ is the time for a deflection of $90$ degrees. \par Taking an electron with $\gamma=1.02$, which is the low energy cutoff of our fast electron distribution, the deflection time is $t_{\rm d} \approx 62$ msec. The time for a deflection of only a few degrees is much shorter, but the timescale of saturation is a few milliseconds, and therefore the deflection should not be an important factor \cite{Aschwanden b}. Faster particles have even longer deflection times, and are even less affected. \par \section{Conditions for Maser and Saturation Length} Inspection of equation \ref{gyroe} shows that the absorption for a distribution which is isotropic in pitch angle is negative only if the slope of $u(\gamma)/\gamma^2/\beta$ is positive. This is the equivalent of an increase of the fast particles with energy, a distribution which is not likely to occur. The other possibility for amplification is when the distribution is not isotropic in pitch angle. In that case if the derivative of $g(\phi)$ is large enough, and there are solutions such that the pitch angle dependent term in equation \ref{gyroe} is negative, then the absorption will be negative. One such distribution is the loss-cone distribution, which we use. The loss-cone distribution arises naturally in a solar flare where the magnetic field and the density increase with decreasing height from the corona to the chromosphere, and we expect a loss cone to appear near the foot-points of a flare. \par A numerical computation of equation \ref{gyroe} with a loss-cone (equation \ref{losscone}) and a power-law (equation \ref{powerlaw}) results in negative absorption coefficients for several frequencies and angles to the magnetic field. Waves of these frequencies, and propagating with these angles, are amplified, and this is the Electron-Cyclotron-Maser mechanism. \par The major factor determining which mode is the most important, or dominant, is $k\times L$, or $\Gamma\times L/v_{\rm g}$, where $L$ is the length of amplification \cite{BenzBook}. If $L$ is determined by homogeneity length-scales in the loop then the mode with largest absolute value negative absorption is dominant. However, if the maser stops operating after a distance smaller than the homogeneity scales of the loop, the amplification length is intrinsic to the ECM itself. Our goal was to derive the absorption coefficients, and the length of amplification $L$. \par The intrinsic length of the maser was estimated with a model for the evolution of the fast electron distribution. We assumed the electrons are accelerated for a short time at the loop top, so that a pulse of fast electrons with a duration of a few milliseconds to a few hundred milliseconds is produced. The fast electrons reach an abrupt increase in the magnetic field, and the electrons with large pitch angles are reflected back. We assumed that before reaching the mirror the fast electron distribution is isotropic in pitch angle, and that all the electrons below a critical pitch angle pass the field increase. These assumptions result in the ideal loss cone distribution described by equation \ref{losscone} with a loss-cone opening angle of $\alpha$. Our evolution model therefore began with an ideal loss-cone distribution over some volume, and at time $t=0$ the maser is ``turned on``. We then assumed that the anisotropic loss-cone distribution relaxes into a new distribution via the maser, and for simplicity we assumed the new distribution is also an ideal loss cone. The new distribution has a smaller opening angle $\alpha\prime$, and a power law in energy of the same power index as the original distribution. If the dominant process which causes the relaxation is the maser, the difference in energy between the initial loss-cone distribution and the final distribution is the energy available to the maser. \par A useful measure of the amplification process is the {\it saturation length}, $l_{\rm sat}$. We define the saturation length as the distance a wave can be amplified before the loss-cone distribution is relaxed so that the loss cone disappears, or the particular wave is no longer amplified. Complete relaxation for a volume element occurs when the energy in the element is reduced from the anisotropic distribution level to the isotropic distribution level. Our method for calculating the saturation length assumes that the greatest possible amplification is when the entire energy difference between initial and final loss-cone configurations is turned into maser radiation. We estimate this energy difference by assuming that the relaxation process leaves the number of fast electrons and the total parallel momentum unchanged. The length of amplification can not be larger than the smallest saturation length, and therefore if $l_{\rm sat}$ is smaller than the loop length-scales it determines the dominant mode. The mode with the smallest saturation length is therefore the dominant mode. \par We followed the gradual closing of the loss cone in detail for some test cases, and our results justify an all-or-nothing approximation. In this approximation we assume that the distribution remains unchanged until the energy radiated by the maser is equivalent to some per cent of the kinetic energy, and then the relaxation to the lower energy configuration is instantaneous. When the closing of the loss-cone is followed it is found that the loss-cone opening angle is usually changed only by a few degrees before the absorption coefficient for the mode which had the largest amplification becomes positive. However, this happens after a distance which is only larger by $20$ to a $100$ per cent of the $l_{\rm sat}$ calculated with the all-or-nothing approximation, when we assume the energy turned into the maser is $10$ per cent of the kinetic energy. The energy difference with the gradual closing method is usually 7--10 per cent of the total kinetic energy. Our all-or-nothing method is simple and gives a transparent relation between the absorption and the saturation length. \par Our calculations are predicated upon the assumptions that the magnetic field strength $B$, the direction to the observer $\theta$, and the densities $n_{\rm cold}, n_{\rm hot}$ are constant. Figures \ref{OMT=5}, \ref{XOT=5}, and \ref{ZMT=5} show contours of negative absorption coefficients around the area of maximum negative absorption. From these figures it is evident that a change in direction of the magnetic field by even $1\degr$ removes a propagating wave from the peak of negative absorption. Likewise a change of 1 per cent in the magnetic field makes the normalized frequency $\nu/\nu_{\rm B}$ of a wave change by about 1 per cent, and also removes the wave from the peak of amplification. Estimates of the scale lengths for the magnetic field are usually such that changes of $1$ per cent occur over distances of about $L=10^7$ cm \cite{B+A2}. Therefore, for saturation lengths much larger than $10^7$ cm, we expect the amplification ceases before the particular wave reaches its saturation length, and the loss-cone either persists through the loop, or perhaps relaxes by some other means. \subsection{Energy in the Loss-Cone} \label{E in loss-cone} The total momentum parallel to the magnetic field for a one-sided loss cone distribution with constant density in angle up to $\cos(\phi)=\cos(\alpha)$ and no electrons for higher $\cos(\phi)$, and with power law in kinetic energy is \begin{eqnarray} \label{P_par} &&P_{||} = \\ &&\nonumber \int_{\gamma_{\rm min}}^{\gamma_{\rm max}}\int_{\phi=0}^{\phi=\pi/2} {\rm d}\gamma {\rm d}(\cos(\phi)) m\gamma\beta c \cos(\phi) 2\pi u(\gamma)g(\phi)=\\ &&\nonumber = { {mc} \over {2\pi(1+\cos(\alpha))} } { {\cos^2(\alpha)} \over 2 } W_{\gamma} \end{eqnarray} Where the dependence on energy is included in the integral \begin{eqnarray} W_{\gamma} = n_{\rm hot} (\delta - 1) (\gamma_{\rm min} -1 )^{\delta-1} \int{\sqrt{\gamma^2-1} (\gamma - 1)^{-\delta} {\rm d}\gamma} \end{eqnarray} and $\gamma_{\rm min}$ is the lowest energy, $\gamma_{\rm max}$ the highest energy, of the fast electrons. \par The dependence on the opening angle of the loss-cone $\alpha$ is included in the first part of the formula for $P_{||}$, where the integral over $\phi$ was done on the positive side of the distribution, $\phi \leq 90$, and where we assume the transition zone from isotropic distribution to the empty region is negligible. \par The energy dependent integral $W_{\gamma}$ is especially simple for the case of a power law in energy with $\delta=3$ and is \begin{eqnarray} \label{delta=3} W_{\gamma}= {{2 n_{\rm hot}} \over 3} ( 1+ \gamma_{\rm min})^{3/2} (\gamma_{\rm min}-1)^{1/2} \end{eqnarray} where we have neglected, as is usually done, terms of order $ (\gamma_{\rm min}-1) / (\gamma_{\rm max}-1)$. We use $\gamma_{\rm min}=1.02$ and $\gamma_{\rm max}\geq 3$ in our calculations. \par We also need the equation for the kinetic energy density in a power law distribution which is \begin{equation} \label{Ek} E_{\rm k} = { {\delta -1} \over {\delta-2} } mc^2 \ n_{\rm hot} \ (\gamma_{\rm min} -1) \left[{\rm erg\ cm^{-3}}\right] \end{equation} where $n_{\rm hot}$ is the number density of fast electrons. \par We demand conservation of parallel momentum when the distribution changes from a loss-cone with opening angle $\alpha$ and $\gamma_{\rm min}$ to a distribution with $\gamma_{\rm min} \prime$ and $\alpha \prime$, so that $P_{||}(\cos(\alpha),\gamma_{\rm min}) = P_{||}(\cos(\alpha\prime),\gamma_{\rm min} \prime)$. From the parallel momentum conservation condition, and given $\cos(\alpha),\gamma_{\rm min},\cos(\alpha\prime)$, we calculate $\gamma_{\rm min} \prime$. From $\gamma_{\rm min} \prime$ and equation \ref{Ek} we calculate the amount of energy lost in the transition from a loss cone with $\alpha$ to a loss cone with opening angle $\alpha\prime$. The transition to a lower cut off energy is physically reasonable since we expect electrons with low $\gamma$ to emit in the low frequencies more than electrons with high $\gamma$, and when these electrons lose energy they move to smaller $\gamma$-es. \par As an example we took an initial distribution with $\cos(\alpha)=0.8, \gamma_{\rm min}=1.02$, which is similar to our standard loss-cone. From equations \ref{P_par} to \ref{delta=3} the isotropic distribution ($\cos(\alpha\prime)=1$) which has the same total parallel momentum is a distribution with $\gamma_{\rm min} \prime = 1.01026$. Therefore, about fifty per cent of the energy was lost in the transition from a loss-cone with $\cos(\alpha)$=0.8 and $(\gamma_{\rm min}-1)=0.02$, to an isotropic distribution with $\cos(\alpha \prime)=1$ and $(\gamma_{\rm min} \prime-1)=0.01026$. An estimate of fifty per cent is two orders of magnitude larger than that found by Mackinnon, Vlahos \&Vilmer \shortcite{Mackinnon} who estimated the energy as 1 per cent of total kinetic energy, and one order of magnitude larger than that given by White, Melrose \& Dulk \shortcite{WMD}. However, this estimate is about the same as quoted by White et.al. \shortcite{WMD} as resulting from analytic arguments. \par We also expect that a particular wave is no longer amplified once the loss cone characteristics have changed even slightly. Therefore we should have calculated the energy loss when the loss cone angle changed from $\cos(\alpha)=0.8$ to, say, $\cos(\alpha\prime)=0.82$. The energy loss for this case turns out to be only seven (7) per cent. Since our purpose was merely to get an estimate for the energy available to the maser, we can take as a reasonable estimate ten per cent of the kinetic energy. The exact number has very little effect on the saturation length, but the brightness temperature is directly proportional to the energy (see equation \ref{TB2}). \par \subsection{Saturation Length} \label{saturation-length} To calculate the saturation length from the energy estimated to be available to the maser we assume that the maser is dominated by one mode $(\nu,\theta)$. We assume that the entire energy is concentrated in this one frequency and direction, with a typical angular width according to our numerical computations of $\Delta\Omega = 0.1\ \pi$, and a typical frequency width of $\Delta\nu=0.015\nu_{\rm B}$. \par To decide which is the most important mode we used the typical frequency and angular widths, calculated the saturation length for every mode as if that mode was the dominant one, and then compared saturation lengths. Usually the mode with the largest absolute value absorption coefficient also has the smallest saturation length, and is therefore the dominant mode. \par A justification of this approach is seen for example in Fig. \ref{OMT=5}. The figure shows contours of equal absorption coefficient for the case $\nu_{\rm p}/\nu_{\rm B}=1, T=5\cdot 10^6$, and the OM. The absorption coefficients are negative, and therefore the figure shows equal amplification contours. The maximum amplification in Fig. \ref{OMT=5} is in the region $0.795<\cos(\theta)<0.8$ and $1.04<\nu/\nu_{\rm B}<1.045$. If the magnetic field strength decreases by one per cent the frequency of the wave moves from approximately $1.04 \nu_{\rm B}$ to approximately $1.05 \nu_{\rm B}$, and the amplification is decreased by approximately half. Likewise, if the inclination of the field changes by one degree, the amplification is also reduced by approximately half. It is therefore justified to use narrow frequency and angular widths. \par The simplest method of calculating the saturation length is the all-or-nothing method. The amplification of any mode is calculated from the length over which there is negative absorption, which is assumed to be arbitrary at first. On the other hand the intensity produced from a given volume must be limited by the total energy in the volume. The length for which the loss cone persists is calculated self-consistently by equating the energy in the wave, and the energy in the volume. \par The total power of the maser per unit area is \begin{equation} P = I(\nu,\cos(\theta)) \Delta\Omega \Delta\nu \end{equation} The energy increases over a volume equal to the path of amplification times unit area, and therefore the total energy available to the maser is \begin{equation} 0.1\times E_{\rm k}\times l_{\rm sat}\ \ [{\rm erg\ cm^{-2}}] \end{equation} where we used section \ref{E in loss-cone} and our estimate of 10 per cent of the total kinetic energy as the energy available to the maser, and $l_{\rm sat}$ is the length of amplification -- which is unknown as yet. \par Assuming that the maser lasts for a given time $\Delta t$, and that the intensity of the emission is constant for that time, the equation which has to be solved is \begin{eqnarray} P \Delta t =0.1\times E_{\rm k}\times l_{\rm sat} \end{eqnarray} By using the radiation transfer equation for negative absorption \begin{equation} \label{radtran} I = {j\over k} (1-{\rm e}^{-k x}) \approx {j\over {|k|}} {\rm e}^{|k|x} \end{equation} where we have neglected $1$ compared to the exponent, and assumed that $j,k$ are constant everywhere, we have \begin{eqnarray} \label{lsat2} &&{j\over{|k|}} {\rm e}^{n_{\rm hot} |k| l_{\rm sat}} \Delta\Omega \Delta\nu \Delta t = \\ &&\nonumber 0.1\times 2\times 8.2\cdot 10^{-7}\times n_{\rm hot}\times (\gamma_{\rm min} -1) \times l_{\rm sat} \end{eqnarray} where we used equation \ref{Ek} with $\delta=3$, and also introduced normalized (to $1$ particle) absorption and emission coefficients. \par If we assume the maser lasts for a given time $\Delta t$ the equation can be solved numerically for any $\Delta t$. However, if we assume $\Delta t=l_{\rm sat}/v_{\rm g}$, with $v_{\rm g}$ given by equation \ref{vg}, then the solution is straightforward, and the result of the simple method is \begin{equation} \label{satlen} l_{\rm sat} = \ln\left({{0.1\cdot E_{\rm k}\cdot v_{\rm g}\cdot |k|}\over {\Delta\Omega\Delta\nu\cdot j}}\right) /(n_{\rm hot} |k|) \end{equation} A more complicated method of deriving the saturation length is the differential approach. We assume a loss-cone distribution over some volume of space. We further assume that the maser is ``turned on`` at time $t=0$. The intensity at time $t$ arriving into a volume element is given by equation \ref{radtran} with $x=v_{\rm g} t$, since a wave starting at time $t=0$ has advanced this distance by time $t$. A volume element adds to the intensity a fractional intensity \begin{equation} \nonumber {\rm d}I=(j-kI){\rm d}x \end{equation} and again neglecting $1$ compared to the exponent, the intensity increase for a volume element is \begin{equation} {\rm {dI}\over{dx}}\approx |k|I\approx j {\rm e}^{|k|v_{\rm g}t} \end{equation} The power in ${\rm erg\ s^{-1}}$ drawn from a unit volume at time $t$ after the maser was turned on, assuming that $k$ and $j$ remain independent of $t$ and the same everywhere, is therefore: \begin{equation} \Delta P(t) \approx \int {\rm d}\Omega \int {\rm d}\nu j {\rm e}^{|k|v_{\rm g} t} \end{equation} The energy drawn from from a unit volume from time $t_1$ to time $t_2$ is \begin{equation} \Delta E=\int_{t_1}^{t_2}\Delta P(t) {\rm d}t \approx \int {\rm d}\Omega \int {\rm d}\nu {{j}\over{|k|v_{\rm g}}}({\rm e}^{|k|v_{\rm g} t_2}-{\rm e}^{|k|v_{\rm g} t_1}) \end{equation} Since energy is being drawn from the volume element the distribution function changes, and here we use section \ref{E in loss-cone}. We assume that the distribution changes so that it remains a power-law with index $\delta$ in energy, and a loss-cone. The new distribution function, with less energy, has a new and smaller $\gamma_{\rm min}\prime$ and a new and larger $\cos(\alpha\prime)$. From our calculated energy loss for some interval $t_2-t_1$ we write $E_{\rm k}^\prime=E_{\rm k} - \Delta E$, and derive $\gamma^\prime$. From the condition that the parallel momentum is constant $P\prime_{||}=P_{||}$, and using $\gamma^\prime$, we derive $\alpha^\prime$. We can then calculate the new absorption and emission coefficient, the new energy loss, and so on, and repeat the process iteratively. If we take $t_1=0$ and $t_2=\tau=l_{\rm sat}/v_{\rm g}$, and substitute for the integrals the constants $\Delta\Omega$ and $\Delta\nu$, the equation for the saturation length we derive from the differential approach is identical to the equation of our simple method (equation \ref{satlen}). \par We followed the gradual closing of the loss-cone iteratively for a few test cases. We took into account the change of the absorption and emission coefficients for waves arriving in later times, which passed through regions where the distribution has already changed. The results of this method are that about 3--7 per cent of the kinetic energy is turned into maser radiation, mostly in the dominant mode, that this happens after a time $<2\cdot \tau$, and that the intensity of the maser is very near to the intensity after amplification with the original maximum absorption over length $l_{\rm sat}=\tau v_{\rm g}$. Thus the natural growth time of the maser to maximum is $\leq 2\ \tau$ and the length over which maser amplification is possible is $\leq 2v_{\rm g}\tau$, and our simple method is a very good approximation. \par Since the results of following the closing of the loss-cone and just assuming a fixed energy available to the maser are so close, we used for the rest of our calculations a fixed 10 per cent energy availability. The results given in this work were calculated with this assumption. \section{Calculations} Our calculations were carried out within the frame of the cold-plasma and the magneto-ionic approximations \cite{MelroseB}. We therefore disregard any negative absorption for frequencies which do not obey either of the two cold-plasma criteria \ref{criteria1} and \ref{criteria2}, where ``much larger`` in these criteria is taken to be at least $10$ times larger. \par We calculated the absorption coefficient with equation \ref{gyroe}, and the emission coefficient with the equation appropriate for it \cite{R}. We performed the calculation using a power-law distribution in equation \ref{gyroe} to get the absorption due to the fast electrons, and also using a thermal distribution in the equation to get the absorption due to the ambient plasma, and the absorptions were then added to derive the total absorption. The total emissivity was derived in the same manner. The calculations were carried out without any approximations to the Bessel functions, and the full sum over the $s$ values was performed. The integral over the pitch angle $\phi$ was performed using the fully relativistic resonance condition \ref{resonance}, and the distribution functions $f({\bmath p})$ were normalized to the number density of the fast electrons for the power law, and the ambient plasma for the thermal distribution. The parameters used for the calculation were \begin{itemize} \item $\delta=3$ ; \item $\gamma_{\rm min}=1.02$ ; \item $\gamma_{\rm max}=3$; \item $\cos(\alpha)=0.81$ ; \item $\cos(\alpha-\Delta\alpha)=0.83$ ; \item $B=360$ gauss ; \item $n_{\rm cold}/n_{\rm hot}=10^4$, unless otherwise indicated ; \item The ambient density is taken to be $n_{\rm cold}=1.24\cdot 10^{10} (\nu_{\rm p}/\nu_{\rm B})^2\ {\rm cm^{-3}}$, which gives $\nu_{\rm p}=1$ GHz for $\nu_{\rm p}/\nu_{\rm B}=1$. \end{itemize} We used $\gamma_{\rm max}=3$ because several tests we performed showed that electrons with higher kinetic energy do not contribute to emission in the frequencies of the maser. \par For every given pair of $(\nu_{\rm p}/\nu_{\rm B},T)$ we calculated the absorption and emission coefficients on a grid of $\cos(\theta)=0.02-0.92$, with a spacing of $\Delta\cos(\theta)=0.02$. For every $\cos(\theta)$ we calculated the coefficients for the frequency range from $\nu=\nu_{\rm B}$ to $\nu=3\nu_{\rm B}$, or higher if required, and with a changing resolution. The initial scanning resolution was changed from $\Delta\nu/\nu_{\rm B}=10^{-4}$ for the small $\cos(\theta)$ to $\Delta\nu/\nu_{\rm B}=0.01$ for the larger $\cos(\theta)$. When negative absorption coefficients were found the resolution was increased ten fold, and the calculation for the range where the absorption is negative was performed with the higher resolution. In this way we are confident that no negative absorption in the $(\cos(\theta),\nu)$ range was missed. Sometimes calculation was performed with better resolution in $\Delta\cos(\theta)$ or $\Delta\nu/\nu_{\rm B}$ if the absorption coefficients changed too much with the coarser resolutions. We performed calculations for the entire range $\nu_{\rm p}/\nu_{\rm B}=0.1-3$, but present only examples. \par Negative absorption appeared for the OM and XO near the harmonics of the cyclotron frequency, and for the ZM also near the upper hybrid frequency. Therefore it was usually not necessary to search the whole range $\nu_{\rm B}-2\nu_{\rm B}$ and the whole range $2\nu_{\rm B}-3\nu_{\rm B}$ to find the negative coefficients. \par Calculation of the saturation length was performed as described in the section {\it Saturation Length}, with the assumption of constant width in angle and frequency. The widths enter the saturation length only logarithmically, so the result is only weakly dependent on them. Typical numbers are $\Delta\Omega=0.1\pi$ and $\Delta\nu=0.015\nu_{\rm B}$. \par The tables summarizing our results are arranged according to $\nu_{\rm p}/\nu_{\rm B}$. Every row gives the characteristics of the most important wave for the given mode and temperature, and for the $\nu_{\rm p}/\nu_{\rm B}$ of the table. The waves are either those with the largest (in absolute value) absorption coefficient, or those with largest growth rate, for all $\theta$ and $\nu$. The first column is the temperature used to calculate the absorption coefficient for the ambient plasma, in units of $10^6$ K. The temperature affects only the thermal absorption and not the absorption calculated from the power law electrons. The second column is the frequency in units of the cyclotron frequency $\nu_{\rm B}$. The third column is the emission angle to the magnetic field, or the angle of the wave vector ${\bmath k}$ (assumed to be in the $(x,z)$ plane) to the z-axis ( $\bmath B$ was assumed to be in the $z$ direction). The fourth column gives the absorption coefficient itself, normalized to a number density of the fast particles of $1$, and given for a magnetic field of $360$ gauss which gives $\nu_{\rm B}=1$ GHz. The absorption coefficient is therefore $n_{\rm hot}$ times the $k$ which is given in the table. The fifth column is the saturation length in units of $10^6$ cm, computed with our formula \ref{satlen}, and the assumption that $n_{\rm hot}=1.24\cdot 10^6\ (\nu_{\rm p}/\nu_{\rm B})^2\ {\rm cm^{-3}}$. The sixth column is the group velocity, the seventh is the growth rate, and the last column gives the mode for which these parameters were computed. \par The contour figures show only negative absorption contours, multiplied by a factor of $-10^{12}$ to make the numbers easier to display. These are examples showing only the region of $(\cos(\theta),\nu/\nu_{\rm B})$ plane around the largest in absolute magnitude negative absorption coefficient for some mode. \section{Observability of the Maser} A previous study which used a power law in energy for the fast electrons found that the ZM dominates for all $\nu_{\rm p}/\nu_{\rm B} > 0.5$ \cite{Fleishman2}, and therefore concluded that the emission must come from regions with smaller $\nu_{\rm p}$. Another study also concluded that the maser must come from a region of low density $\nu_{\rm p}/\nu_{\rm B} < 0.25$, because for higher densities the dominant modes reach saturation much faster than the observed durations of MW spikes \cite{Aschwanden b}. \par We found that for $\nu_{\rm p}/\nu_{\rm B}< 1.1$, and $\nu_{\rm p}/\nu_{\rm B} > 1.8$ the ZM has the largest in absolute value negative absorption (see Fig. \ref{absorption}), but has the largest growth rate only for $\nu_{\rm p}/\nu_{\rm B}< 0.6$ (Fig. \ref{growth-rate}). If the pulse of fast electrons has a much longer duration than the time scale of maser saturation the important factor is not the growth rate, but the saturation length. At a time after the maser saturates, the mode with the smallest saturation length has closed the loss-cone beyond its own saturation length, and therefore the smallest saturation length becomes the amplification length for all modes. Modes with longer saturation lengths can reach only a fraction of the full amplification possible to them, being amplified by a factor $exp(|k|\cdot L)$, with $L$ the saturation length of the mode with the smallest saturation length, and not ${\rm exp}(|k|\cdot l_{\rm sat})$, with $l_{\rm sat}$ the saturation length of the mode which is $L<l_{\rm sat}$. Therefore what is required is to derive the saturation length for all modes, and find the mode with the smallest saturation length. Since the saturation length is inversely proportional to the absorption coefficient, the mode with smallest saturation length is usually the mode with the largest (in absolute value) negative absorption coefficient. \par These findings present a problem in that we expect a loss-cone distribution to appear near the foot-points of flare loops where the magnetic field and the density increase sharply. The cyclotron frequency-plasma frequency ratio for the footpoints is probably in the range $\nu_{\rm p}/\nu_{\rm B} \approx 0.5-3$, if we take as typical $n_{\rm cold} \approx 10^{10}-10^{11}\ {\rm cm^{-3}}$ for the density of ambient plasma \cite{B+A}, and $B \approx $ a few times $100$ gauss as typical for the magnetic field. These values are in the range where our calculations, and previous work, predict the unobservable ZM is the dominant mode. The high amplification of the ZM would therefore seem to preclude the ECM mechanism as the producer of observable MW spikes, except in the range $1.1 < \nu_{\rm p}/\nu_{\rm B} < 1.8$. \par We have, however, to consider the ambient thermal plasma at the foot points, which we assume to be at a temperature of order $10^7$ K. The thermal effect was not taken into account in most studies, and those which did -- assumed much lower temperatures for the ambient plasma than we do. We use the range of temperatures inferred from SXR observations, since we assume the model of a hot dense plasma, heated by the fast electrons, flowing upwards to fill the loop, and emission where the fast electrons intersect the evaporation front \cite{B+A}. \par The temperature has two effects on our results. The first is to increase the positive absorption near the harmonics of the cyclotron frequency. If the frequency where the fast electrons create negative absorption is close to a harmonic, the thermal plasma absorption decreases the amplification, or even turns the negative coefficient positive. For the XO and OM this may eliminate the amplification at a low harmonic, but these modes appear in higher harmonics and usually have negative absorption for the higher harmonics as well. The ZM is limited to frequencies less than the upper-hybrid frequency, and therefore it is possible that the introduction of high temperatures will eliminate the ZM negative absorption while leaving higher frequency OM and XO negative absorptions. The second, and more important, consideration high temperatures introduce is the cold plasma conditions. For a temperature of $10^6$ K these criteria are valid for frequencies very close to the harmonics, and for very large refraction indices. However, the high temperature we assumed forced us to reject certain negative absorption results as not valid within the framework of our assumptions. We assumed that if there is any amplification for these cases it is orders of magnitude smaller than the cold plasma result, and therefore negligible. This is supported by the known effect of thermal corrections, which is to decrease the refraction indice and the absorption \cite{MelroseB}. \par Even if the thermal positive contribution leaves a negative ZM absorption which is valid for the cold plasma approach, and is the largest in absolute value of all negative absorptions, the OM and XO may be observable. We assumed in the previous section that the largest in absolute value negative absorption coefficient dominates, and that all the energy goes into amplifying this mode. With this assumption we derived the saturation length and the amplification for this dominant mode. This is similar to the assumption in previous studies, which give the dominant mode in some parameter region as the mode with the highest growth rate (White et.al. 1986; Aschwanden 1990a), and it is justified because of the exponential nature of the intensity. There are, however, two scenarios where modes with smaller absorption coefficients can be observed. The first scenario, presented in section \ref{non-dominant}, depends on the high brightness temperature of the maser. The second scenario , presented in section \ref{slow-mode}, depends on the slowness of the Z-mode. If the ZM is slow enough, another mode may be able to reach its full amplification before the ZM saturates, and then, for a short time until the ZM saturates, this mode can also be observed. \par Finally, there is a region where the ZM does not have negative absorption. For the region $1.5\leq \nu_{\rm p}/\nu_{\rm B} \leq 1.6$ the lowest frequency for which the ZM exists is higher than $\nu_{\rm B}$ and the highest frequency for which the ZM exists is lower than $2\nu_{\rm B}$. Negative absorption usually appears near the harmonics of the cyclotron frequency, and for the ZM frequencies with this region of $\nu_{\rm p}/\nu_{\rm B}$ there are no negative absorption coefficients. \par All the above considerations are summarized in Fig. \ref{absorption} and in Fig. \ref{growth-rate} giving the maximum absorption coefficients and growth rates for the modes. The coefficients are the largest in all frequencies and angles for the given $\nu_{\rm p}/\nu_{\rm B}$, wave mode, and with $T=5\cdot 10^6$ K. The ZM does not appear for $1.6<\nu_{\rm p}/\nu_{\rm B}$ because where negative absorption reappears at $\nu_{\rm p}/\nu_{\rm B} > 1.6$ the frequencies fail the cold plasma criteria. It is important to note that the largest growth rate does not necessarily occur at the same angle and frequency as the largest in absolute value negative absorption coefficient, as can be seen in Tables \ref{ffp=0.8} to \ref{ffp=1.8}. \par Whatever mode is amplified, the ECM is difficult to observe because of the narrow bandwidth, and especially because of the small angular width of the emission. Fig. \ref{OMT=5} shows contours of negative absorption coefficients around the area where the OM has the largest negative absorption (largest in absolute value). To observe the emission it is necessary to be within a cone of less than $1$ degree around the angle of maximum amplification, and within $1$ per cent of the frequency of maximum amplification. The absorption $1$ per cent away from the maximum drops by half, which for an exponential growth with a maximum power of the exponent of about $30$ means a drop of $7$ orders of magnitude in amplification. In Fig. \ref{XOT=5} and Fig. \ref{ZMT=5} the same narrow bandwidth and small angular width are seen for the other modes, and the phenomena occurs for all other $\nu_{\rm p}/\nu_{\rm B}$ values. This general result means that observation of ECM should be rare, even if ECM is very common and occurs at all times. \par \subsection{Thermal Effect} In most cases the thermal plasma tends to suppress maser emission, as was noted in previous studies (Aschwanden 1990b; Kuncic \& Robinson 1993). For example for $\nu_{\rm p}/\nu_{\rm B}=0.8$ (Table \ref{ffp=0.8}) and $T=5\cdot 10^6$ K, the most important OM was found at frequency $\nu/\nu_{\rm B}=1.03$. At a higher temperature of $10^7$ K the most important OM moved to a higher frequency of $\nu/\nu_{\rm B}=1.079$. The fast electrons still gave the same negative absorption at $\nu/\nu_{\rm B}=1.03$, only for $T=10^7$ the thermal absorption was large enough to turn the absorption at this frequency positive. On the other hand, the ZM at $\nu/\nu_{\rm B}=1.2745$ was not affected by the thermal absorption, because of its distance from the cyclotron frequency. The change there was the result of the application of the cold plasma criteria, which for $T=10^7$ this frequency violates, and therefore the largest acceptable absorption was at a lower frequency, further from the resonance and with a smaller refraction index. All these changes were not enough to change the dominant mode, which remained the ZM, because the OM absorption was reduced by a factor of $8$, while the ZM absorption was reduced only by a factor of $4$. The total effect of the higher temperature was therefore to reduce the relative importance of the OM emission. \par The other ZM given in Table \ref{ffp=0.8} is the ZM wave with the largest growth rate. This is a wave at a much lower frequency, and a slightly different angle than the wave with largest (negative) allowable absorption coefficient. This mode also changes with the increase in temperature from frequency $\nu/\nu_{\rm B}=1.0351$ for $T=5\cdot 10^6$ to frequency $\nu/\nu_{\rm B}=1.087$ for $T=10^7$. In this case the absorption was only reduced by a factor of $2$, and the growth rate by about the same. This is an example where the mode with the largest growth rate is not the dominant mode. \par A counter example is demonstrated in Table \ref{ffp=1.8}. The first row of Table \ref{ffp=1.8} gives a negative absorption coefficient for the frequency $\nu/\nu_{\rm B}=2.04885$, which is very close to the ZM resonance for this case. The refraction index for this frequency and for the angle $\cos(\theta)=0.06$ is $n_{\rm ref}=4.42$, which was large enough for this solution to barely violate the 1st plasma criterion (eq. \ref{criteria1}), giving a value only $9.5$ times larger than $0.001683/2$. This is not allowed, which is why the row is marked as ZMN in the table. Therefore, for $\nu_{\rm p}/\nu_{\rm B}=1.8$ the temperature leads us to discard the ZM waves as unimportant, and leaves us with the OM as the dominant mode. \par When we made the calculations in Table \ref{ffp=1.8} with a temperature of $T=10^7$ K, all the ZM negative coefficients disappeared entirely, overcome by the thermal absorption, and therefore there was no need to even consider the cold plasma criteria. \par It is important to note a problem with Table \ref{ffp=1.8} which is that the negative absorption coefficients are very small. The saturation lengths were therefore very large, and may violate the assumption that the maser is very small compared to the dimensions of the loop. For example for $T=10^7$ K the saturation length for the OM was found to be $l_{\rm sat}=1.8\cdot 10^8$ cm, which is large enough for our assumption that the magnetic field, the density, and the temperature are constant to be incorrect. In such a case the maser does not reach its saturation intensity, and the loss cone may relax by some other means. \par \subsection{Non-Dominant Mode} \label{non-dominant} Since the intensity $I$ is difficult to measure we would like to estimate the brightness temperature \begin{equation} \label{TB} T_{\rm B}={{c^2}\over{2\nu^2}} {I\over{k_{\rm B}}} = 3.26\cdot 10^{18} {I\over{(\nu / {\rm GHz})^2}} \end{equation} We can express the intensity in terms of the saturation length, use equation \ref{satlen}, and write the intensity (equation~\ref{radtran}) as \begin{equation} \nonumber I={j\over{|k|}}\left({{0.1 E_{\rm k} v_{\rm g} |k|}\over{\Delta\Omega\Delta\nu\ j}}\right) \end{equation} Using this, the brightness temperature for the dominant mode can now be written as \begin{equation} \label{TB2} T_{\rm B}={{3.26\cdot 10^{18}}\over{(\nu / {\rm GHz})^2}} {{0.1 E_{\rm k} v_{\rm g}}\over{\Delta\Omega\Delta\nu}} \end{equation} Therefore for the dominant mode the brightness temperature depends only on the mode (through $v_{\rm g}$) and on the energy assumed to be available to the maser, and not on the saturation length $l_{\rm sat}$. \par We estimated the importance of any wave with negative absorption by calculating a saturation length assuming this particular wave is the only amplified one. We then compared the various non-dominant saturation lengths to the saturation length of the dominant mode, where by definition the dominant mode is the mode with the shortest saturation length. The amplification length of a non-dominant mode is limited to the loss-cone length determined by the dominant mode $l_{\rm dom}$. The intensity of the non-dominant mode can then be calculated by using the non-dominant absorption and emission coefficients with the dominant mode`s saturation length. The intensity to be used in equation \ref{TB} is for this case \begin{equation} I_{\rm non}= {{j_{\rm non}}\over{|k_{\rm non}|}} {\rm e}^{|k_{\rm non}|\cdot l_{\rm dom}} \end{equation} or, alternatively \begin{equation} I_{\rm non}={{j_{\rm non}}\over{|k_{\rm non}|}}\left({{0.1 E_{\rm k} v_{\rm g} |k_{\rm dom}|}\over {\Delta\Omega\Delta\nu\ j_{\rm dom}}}\right)^{k_{\rm non}/k_{\rm dom}} \end{equation} Where $k_{\rm dom},j_{\rm dom}$ are for the dominant mode, which determines the saturation length, and $k_{\rm non},j_{\rm non}$ are for some other mode, at a different frequency and another emission angle, but which also has a negative absorption, and is therefore amplified. This formulation is true no matter which is the dominant mode, also for the ZM. \par If $|k_{\rm non}|\cdot l_{\rm dom}$ is large, the resulting intensity is large and observable. For example for $\nu_{\rm p}/\nu_{\rm B}=1$ (Table \ref{ffp=1}) and $T=5\cdot 10^6 $ K the XO saturation length was the shortest of all modes, though very close to the ZM saturation length. However, the OM absorption coefficient was also very close to the XO absorption coefficient, and obviously the non-dominant situation is relevant. This is also one of the rare cases where the OM was slower than the ZM, so the growth rate of the OM was much smaller than the growth rate of the ZM, and certainly smaller than the growth rate of the XO. In spite of that, our model predicts a strong emission in the OM for this table's $\nu_{\rm p}/\nu_{\rm B}=1$. For this case $j/k$ for the OM is about $1/16$ of the $j/k$ ratio for the XO. Therefore even when the power of the exponent is the same, the total intensity in the OM is only $1/16$ of the intensity in the XO. \par If the non-dominant mode has a much smaller absorption, for example as in $\nu_{\rm p}/\nu_{\rm B}=1$ (Table \ref{ffp=1}) and $T=10^7$ K, where $k_{\rm non}\approx k_{\rm dom}/3$, then the non-dominant intensity is much smaller. For example if we assume $k_{\rm dom}\cdot l_{\rm dom} \approx 25-30$, as it often is, and the same $j/k$ for the dominant and non-dominant modes, the factor between the intensities is ${\rm e}^{17}\approx 10^7$. This factor reduces a brightness temperature of $10^{18}$ K to a brightness temperature of only $10^{11}$ K, which is lost against the background of the entire loop, and is in any case much lower than is observed for microwave spikes. \par The way to decide when a non-dominant mode could be observed is to compare absorption coefficients. In Fig. \ref{growth-rate} the XO growth rate for $\nu_{\rm p}/\nu_{\rm B}=1$ (of the example above) appears to be much larger than the OM growth rate. However, in Fig. \ref{absorption} we see that the absorption coefficients are equal. The example of the previous paragraph shows that the OM is observable whenever such a situation occurs. \subsection{Slow Mode Saturation} \label{slow-mode} Another possibility for observation of a non-dominant mode is for the case of the dominant mode being slow. The Z-mode is usually a slow wave, especially near its resonance frequency, where its group velocity can be even as low as $c/500$. It is therefore possible that when a certain set of conditions is met another mode is amplified to its maximum. This can happen if the Z-mode is slow enough that the other mode reachs its saturation length before the ZM saturates. In that case until the ZM saturates and reduces the enegry in the distribution, the other mode is emitted with the full amplification of its own saturation length. \par The growth rate equation \ref{growth} already includes the group velocity, and therefore a comparison of growth rates suffices to show if this slow-mode scenario is possible. It is also possible to take the group velocity into account from the perspective of the spatial absorption coefficient, though we need a few assumptions. We assumed a typical emitting electron energy of $\gamma=1.1$, and that the important electrons are on the edge of the loss cone with pitch angle $\cos(\phi)=0.8$. These parameters gave a velocity parallel to the magnetic field of $c/3$ for the most important electrons, while the OM or XO wave was usually faster than $c/3$. \par For an example we used Table \ref{ffp=1.4} for temperature of $10^7$ K. The OM group velocity is $0.783\cdot c$, and the ZM group velocity is $0.1253\cdot c$. An OM wave of speed $0.783\ c$ beginning near the mirror point at time $t_1$ after the first electrons mirrored overtakes these first electrons (of parallel velocity $c/3$) at time $t_2=1.74 \cdot t_1$, and at a point $0.58\ c\cdot t_1$ from the mirror point. In our model (section \ref{saturation-length}), when this distance is the saturation length, $0.58\cdot c\cdot t_1=l_{\rm sat}^{OM}$, the intensity of the radiation is so great that the end-most volume element with a loss-cone distribution is drained of its free energy, and the loss-cone distribution disappears from that end-most volume. A Z-mode wave emitted at time $t=0$, and having a group velocity of $0.1253\ c$, has propagated by this time only to a distance of $0.1253\cdot c\cdot t_2=0.376\cdot l_{\rm sat}^{OM}$. If the ZM with this group velocity has an absorption coefficient such that its saturation length is for example $0.4$ of the OM saturation length, the Z-mode had not saturated by the time the OM emission has reached its peak and is beginning to close the loss-cone. However, at some later time the ZM will have had time to saturate near the mirror-point, and has closed the loss-cone close to the mirror. An OM/XO wave starting from the mirror point at that time does not see a loss-cone, and is not amplified. For the case in Table \ref{ffp=1.4} the ZM has a saturation length only $0.27$ of the OM saturation length, and therefore the ZM has saturated by the time the OM wave reaches its saturation length. This however does not mean that the OM is not amplified as explained below. \par A better calculation than the above takes into account that the OM wave advances faster than the electrons and much faster than the ZM wave. The requirement then is that the electrons the OM wave passes have not themselves gone previously through a saturated Z-mode wave, and therefore have retained the loss-cone distribution. The result of that requirement is that for a mode with $l_{\rm sat}$ to be amplified, the Z-mode saturation length $l_{\rm z}$ must be large enough so that the inequality below is correct : \begin{equation} {{l_{\rm z}}\over{l_{\rm sat}}} \geq {{v_{\rm z}}\over{v_e}} {{v_{\rm g}-v_e}\over{v_{\rm g}-v_{\rm z}}} \end{equation} With $v_e$ the velocity of the electrons, $v_{\rm g}$ the group velocity of the OM or XO, $v_{\rm z}$ the group velocity of the ZM, $l_{\rm z}$ the saturation length of the ZM, $l_{\rm sat}$ the saturation length of the OM. The same equation holds for any two modes, one with group velocity $v_{\rm z}$ slower than the electrons, and the other with group velocity $v_{\rm g}$ faster than the electrons. \par For the case of Table \ref{ffp=1.4} the left side of the equation is $0.27$ while the right side is $0.257$, and therefore the OM can reach its maximum possible amplification before the ZM closes the loss-cone and prevents it. \par A simpler estimate for the relative importance of the modes is to compute $\tau=l_{\rm sat}/v_{\rm g}$ for each of the modes. This estimate is very close to calculating the growth rate, since dividing the saturation length by $v_{\rm g}$ is equivalent to replacing the absorption $k$ by the growth rate $\Gamma=-k\cdot v_{\rm g}$ in equation \ref{satlen}. This criterion tends to be smaller for a larger growth rate, and the smallest value is the most important. For our example (Table \ref{ffp=1.4}, $T=10^7$) the 7th column of the table shows that the growth rate of the OM is larger than the growth rate of the ZM. Therefore, even though the saturation length of the OM is much larger than the saturation length of the ZM, and after the ZM saturates the OM disappears, the growth rate criterion, and our calculations above, show that the OM can be observed for a short duration, until the ZM saturates . \par Another example is for $\nu_{\rm p}/\nu_{\rm B}=0.8$ (Table \ref{ffp=0.8}), $T=5\cdot 10^6$ and the ZM. In the Table we give both the ZM wave with largest (absolute value) negative absorption, and the ZM wave with the largest growth rate. The ZM wave with the largest absorption is very slow and has a very small growth rate, and therefore the important wave is at first the one with the largest growth-rate. For $\nu_{\rm p}/\nu_{\rm B}=0.8$ the largest growth rate in the ZM is still smaller than the growth rate of the OM wave, and therefore we conclude that the OM reaches its saturation length before the ZM closes the loss-cone, and we expect to see an OM wave for a short time. It turns out that the difference between the time to saturation for the OM and the time to saturation for the largest growth rate ZM is very small, about $0.1\ {\rm msec}$. The time we expect to see the OM at its full amplification is therefore very small. \par In this case it was usefull to check the non-dominant-mode approach of the previous section, and we found that the OM is observed until the small saturation length ZM saturates. When we used the OM absorption coefficient with the saturation length of the largest growth rate ZM we got a brightness temperatures of order $T_{\rm B}=10^{15}$, which is observable. \subsection{Time Dependence} The duration of the pulse of fast electrons is obviously unrelated to the loss-cone time scales. If the pulse is shorter than the loss-cone time-scale, less than $t_2=3\cdot l_{\rm sat}/c$ for the example of the previous section, the maser intensity in the OM is not strong enough to close the loss-cone. The loss-cone distribution therefore moves from the mirror point at the typical loss-cone velocity, and the maser continues to emit, though at a weak intensity. Using the terminology of the example from the previous section, and assuming a pulse duration of $t_1$, the maximum amplification length is $c\cdot t_1/3$. In the example we had $l_{\rm sat}=0.58\cdot c\cdot t_1$, and therefore the maximum amplification length is about $l_{\rm sat}/2$. Since the intensity is exponential with the amplification length, this translates into a much weaker emission. Typically the power of the exponent for the dominant mode is 20--30, and with $l_{\rm sat}/2$ the intensity is more than a hundred thousand times weaker. For such a case it is unlikely that the maser emission can be observed at all. \par The other extreme is that the pulse of fast electrons persists for a time much longer than typical maser saturation times. For such a case, the loss-cone closes at distance $l_{\rm sat}$ from the mirror point, but new fast electrons move into place constantly. At first glance it would seem that the maser persists for approximately the duration of the pulse, however, if there is a ZM with a smaller saturation length than the observable modes this does not happen. After the Z-mode reaches its saturation length it quenches the fast electron loss-cone distribution, and the observable modes can not be amplified enough to be seen. In this picture the ZM waves saturate and maintain an intensity just strong enough to drain every new volume element of its energy when it overtakes the ZM waves. An observable OM or XO wave starting from the mirror at a time $t > t_1$ of the previous section, so that the ZM has already reached saturation $t > l_{\rm sat}^{ZM}/v_{\rm g}^{ZM}$, does not see a loss-cone, and therefore is not amplified. For our example above, the intensity of an observable mode at first increases, reaching a maximum at time $t_2$ after the first electrons are mirrored. At this time the OM wave which left the mirror at $t_1$ reaches $l_{\rm sat}$. This high intensity can last only until OM waves leaving the mirror encounter saturated ZM waves. For the example from Table \ref{ffp=1.4} the ZM reaches its saturation length near the mirror at time $1.25\cdot t_1$. An OM wave leaving the mirror point at $1.25\cdot t_1$ reaches $l_{\rm sat}$ at $t_2+0.25 t_1$, and the maser then starts to decrease in intensity. The maser is therefore at peak intensity only for a time $0.25 t_1=1.7$ msec. Of course, if the shortest saturation length is for an observable mode, the maser does persist for the length of the pulse of accelerated particles. \par Our conclusion is therefore that if the smallest $l_{\rm sat}$ is for the ZM, the maser persists for a time less then but of order $[l_{\rm sat}/v_{\rm g}]_{ZM}$, regardless of the duration of the fast electron acceleration pulse. This is the case for Table \ref{ffp=0.8} at all temperatures. Of course, after the pulse ends, the waves disappear, and when a new pulse reaches the mirror point, the process can start again. \par An example of a different kind is for $\nu_{\rm p}/\nu_{\rm B}=1$ (Table \ref{ffp=1}) and $T=5\cdot 10^6$ K. For this case the saturation length of the XO was shorter than the saturation length of the ZM, and the saturation length of the OM was not much longer. Therefore we expect the XO radiation to last, and the OM radiation to be amplified as well. For the other tables, at $T=5\cdot 10^6$ the smallest saturation length was always for an observable mode, and therefore at this temperature the loss cone should always produce long lasting maser emission. \par The differential approach described in section \ref{saturation-length} also gives a model for the time-profile of the maser. The emission should appear as a rising exponent, and then decrease faster than an exponent, because the elements whose energy was exhausted begin to positively absorb radiation from the still amplifying elements. This picture is at odds with the characteristics observed, which were a rise as ${\rm e}^(t^2)$ and a decrease as an exponent \cite{G+B}. Perhaps the gaussian rise is a feature of the acceleration mechanism, for example if $n_{\rm hot}$ also increases with time. \par \section{Comparison with Previous Work} We can compare our results with only one previous case \cite{Aschwanden b} where saturation lengths were estimated, and given for example for $\nu_{\rm p}/\nu_{\rm B}=0.5$ as $l_{\rm sat}^{ZM}=10$ km and $l_{\rm sat}^{OM}=30$ km.\par We incorporated in our program Aschwanden`s thermal distribution in energy and $\sin^6(3\times\phi)$ loss-cone. Aschwanden used the standard parameters $T_{\rm hot}=10^8$ K, $T_{\rm cold}=10^6$ K, $n_{\rm cold}=1.25\cdot 10^8\ {\rm cm^{-3}}$, and $n_{\rm hot}/n_{\rm cold}=0.01$. With the above and from the standard $\nu_{\rm p}/\nu_{\rm B}=0.1$ a magnetic field of $B=360$ gauss was the standard field. If the number of particles was kept constant the magnetic field for $\nu_{\rm p}/\nu_{\rm B}=0.5$ as in the example above should have been reduced to $B=72$ gauss. \par Our calculations with Aschwanden`s distribution functions gave the normalized results, with $B=100$ gauss and divided by $n_{\rm hot}$, for the OM: $k=-1.1\cdot 10^{-11}$, $j=1.69\cdot 10^{-24}$, $c/v_{\rm g}=1.144$, $\nu/\nu_{\rm B}=1.019$, $\cos(\theta)=0.24$ and for the ZM: $k=-3.6\cdot 10^{-11}$, $j=4\cdot 10^{-23}$, $c/v_{\rm g}=7.3$, $\nu/\nu_{\rm B}=1.039$, $\cos(\theta)=0.2$. Where these numbers are for the respective largest growth rate in every mode, since that was Aschwanden`s criterion for the dominant wave. \par Using our simple method to calculate the saturation length, and the scaling laws $k\propto n_{\rm hot}/B$ and $j \propto n_{\rm hot}\times B$, we got $l_{\rm sat}^{ZM}=4.89$ km and $l_{\rm sat}^{OM}=17.9$ km. These saturation lengths are about a factor $2$ smaller than Aschwanden`s. We also have to take into account Aschwanden`s assumption of an initial energy density equivalent to $T_{\rm B}=10^{14}$ K. In our model the waves amplify the spontaneous emission, and the amplification is according to the radiative transfer equation \ref{radtran}. A typical value is $j/k \approx 10^{-13}$ and if this is the intensity $I$ the equation for the brightness temperature (equation \ref{TB}) gives the equivalent temperature as only $T_{\rm B}=10^7$ K. The brightness temperature reaches $T_{\rm B}=10^{14}$ K only after the wave has been amplified by a factor of $10^7 \approx {\rm e}^{16}$, and the comparison of the saturation lengths should be only from this stage. \par The total saturation length gives a power in the exponent of about $30$, therefore if we had started with a wave at $T_{\rm B}=10^{14}$ K the saturation length would have been only half of what we calculate. The conclusion is that computing the saturation length with our method resulted in saturation lengths of a factor of 4 smaller than those calculated by Aschwanden`s method. This was the result of neglecting the change in the loss-cone during the amplification of the wave, and assuming that the absorption coefficient remains constant until the wave reaches its maximum energy. The differential iterative method shows that the actual saturation length is up to twice what the simple method gives, and therefore we get a factor of about $2$ back. Considering the simplicity of the equations we use, a factor of two is extremely close. \section{Summary} We have developed a model for estimating the intensity and duration of emission through the Electron-Cyclotron-Maser (ECM) mechanism. \par Observations led us to use a power law for the fast electrons, with a minimum energy of $10$ keV and maximum energy of $1$ MeV. This distribution is harder than the ones studied in the past (Winglee \& Dulk 1986; Aschwanden 1990a). \par From recent results \cite{B+A} we estimate the important ratio $\nu_{\rm p}/\nu_{\rm B}$ in the foot-points of flares, which is the region of interest for this study, to be in the range 0.5--3. For most of this range a loss cone in a power law distribution of electrons results in the ZM being the dominant magneto-ionic mode. The ZM is unobservable, because when a ZM wave reaches a region with a smaller plasma frequency than the plasma frequency where the wave originated, it is absorbed. Thus it seems at first glance that the maser mechanism can not produce observable emission. \par We use an absorption coefficient approach, differing from previous studies which used the growth rate. Using this approach we calculate the {\it saturation length} for the competing magneto-ionic modes, and show that in many cases the observable OM and XO are amplified to observable intensities. \par The conditions which enable the OM and XO modes to be amplified in the presence of a ZM with much larger (in absolute magnitude) negative absorption are : first, the inclusion of an ambient thermal plasma with temperature of order $T=10^7$ K ; second, the low group velocity of the ZM which enables a competing mode to reach its saturation before the ZM saturates ; third, the high amplification of the maser, which makes even non-dominant modes observable. \par Our conclusion is that in the range $\nu_{\rm p}/\nu_{\rm B}=0.5-1.8$ the maser mechanism can produce observable emission in the form of intense spikes with a narrow bandwidth -- about one per cent of the frequency -- and narrow angular width -- about $0.1 \pi$ solid angle. \par The examples given in the tables indicate that long duration emission in the OM and XO appears at angles of 30--60 degrees to the magnetic field, and at 3--6 per cent of the cyclotron frequency above the nearest cyclotron harmonic. This is usually the second harmonic for the OM, and always the second harmonic for the XO. We expect the long duration emission to be determined by the duration of the pulse of accelerated electrons, and the brightness temperature to be of order $T_{\rm B}=10^{16}$ K and above. \par Short duration emission appears when the OM or XO are emitted only until the ZM saturates. For these cases the saturation lengths of the OM and XO are longer than the ZM saturation length, but the ZM group velocity is smaller than the group velocity of the OM or XO. The short duration emission lasts in our model from fractions of a millisecond to a millisecond, for example for $\nu_{\rm p}/\nu_{\rm B}=0.8$ (Table \ref{ffp=0.8}) and a temperature of $5\cdot 10^6$ K, where the OM has the largest growth rate. \par From the results given in Tables \ref{ffp=1} to \ref{ffp=1.8} for a temperature of $5\cdot 10^6$ K, we expect emission for the parameters of the tables for the duration of the accelerated electrons` pulse, and at frequency $\nu \approx 2.06\nu_{\rm B}$. The exception is the case of $\nu_{\rm p}/\nu_{\rm B}=1$ (Table \ref{ffp=1}) where the XO emission for $T=5\cdot 10^6$ K is at $\nu/\nu_{\rm B} \approx 2.06$, and there is also OM emission of observable intensity at $\nu/\nu_{\rm B} \approx 1.03$. We also expect long duration emission for $\nu_{\rm p}/\nu_{\rm B}=1.2$ (Table \ref{ffp=1.2}) and $T=10^7$ K. For this case $l_{\rm sat}^{XO} \approx l_{\rm sat}^{ZM}$, and therefore the XO will be amplified to observable intensities. The XO frequency is $\nu = 2.129\nu_{\rm B}$. For the cases $\nu_{\rm p}/\nu_{\rm B}=1.6$ and $\nu_{\rm p}/\nu_{\rm B}=1.8$ (Table \ref{ffp=1.6} and Table \ref{ffp=1.8}) and $T=10^7$ and $T=2\cdot 10^7$, the OM is the dominant mode, but the saturation lengths are much too large. We conclude that the loss-cone relaxes by some other means, and not through the ECM mechanism. \par In summary, this study shows that the ECM produces millisecond radio spikes such as observed from the Sun. The ECM emission reproduces the frequency of the observed spikes, the bandwidth of the spikes, the source size as recently observed \cite{Benzetal}, the brightness temperature, and the duration of the spikes. \par The ECM does not produce emission at two frequencies simultaneously, and therefore can not reproduce the observed bands \cite{Krucker}. However, if a loss cone distribution forms independently in different regions, and the magnetic field is not the same in these regions, then ECM in several frequencies and angles can be produced. For example in Table \ref{ffp=1} for temperature $T=5\cdot 10^6$ the XO emission is at angle $69$ degrees. In Table \ref{ffp=1.2} for $T=5\cdot 10^6$ the XO emission is at angle $64$ degrees. If two regions, one with $\nu_{\rm p}/\nu_{\rm B}=1$ and one with $\nu_{\rm p}/\nu_{\rm B}=1.2$, emit ECM, and if the magnetic field changes direction by a few degrees between the regions, it may become possible to observe both. Since the magnetic field and the density both increase towards the foot points, the ratio between the emitted frequencies in the two regions can be larger than $1.2$. \par Our model also does not reproduce the reported gaussian rise time \cite{G+B}. However, if the number of fast electrons also increases with time,the rise time is a feature of the acceleration mechanism, since the emission in ECM is directly proportional to the number density of the fast electrons. \par It is possible to conclude from our results that the observed frequency of the spikes should be higher at the beginning of a flare, since the hot and dense chromospheric plasma is then still close to the chromosphere, where a large magnetic field causes a large cyclotron frequency. As the flare advances, and the evaporation front propagates upwards, temperatures below increase, the unobservable ZM dominates for the larger temperatures, and the magnetic field in the region from which ECM emission is observed is smaller. \par \clearpage