content
stringlengths
1
15.9M
\section{Introduction} \par Particle-identification (PID) capability plays an essential role in experiments at B-factories. Especially, $\pi/$K identification in the momentum range up to 4 GeV/c over a wide angular region is crucially important for the primary physics goals to measure CP violation. In the Belle detector at KEK-B, a combination of Aerogel Cherenkov counters, TOF counters and dE/dX measurement in a central drift chamber (CDC) provides PID information for charged particles [1]. Although the present detector system covers most of the momentum region of the charged particles, the PID power for a particle with momentum above 3 GeV/c is not sufficiently satisfactory. Concerning the future upgrade of PID devices, we discuss here the effectiveness of measuring both the Time-of-Propagation (TOP) and the emission-angle of the Cherenkov photons. The time resolution of the Time-of-Flight (TOF) counter using a plastic scintillator is inherently limited by the following effects besides the transit-time spread (TTS) of a phototube: (1) a finite decay-time of photon emission, and (2) a different photon propagation-length or propagation-time in the scintillator to a phototube, depending on the photon emission angle. Thus, the conventional TOF counter measures the arrival time of the earliest photons out of many, whereby the remaining large amount of photons are not effectively used for the time measurement. While effect (1) cannot be reduced as far as the scintillation is concerned, effect (2) can be resolved by the measuring the arrival times of the earliest photons as a function of the photon emission angle, since, in the total-reflection-type scintillator bar, most photons propagate in the scintillator while keeping their original emission angles. This two-dimensional information could improve the time resolution, depending on the number of angular segmentations of the photon detector and on the number of photons collected at each segment. When Cherenkov radiation is utilized, although its number of produced photons is much smaller than in the scintillation case, effect (1) can, in principle, be disregarded. In addition, since the Cherenkov photon emission-angle is uniquely determined by the particle velocity ($\beta$) and since the propagation time of photons in a light guide of an internal-total-reflection type can be calculated as a function of the photon emission angle, a measurement of a correlation between TOP and the photon emission anlge could provide PID information by itself, as we propose below. In the DIRC concept [2, 3], Cherenkov photons produced by a quartz radiator are transported to an end of the radiator by means of internal total reflection, and its ring image is magnified and projected onto a photon-detector plane placed at the bar end. A basic study of this concept clearly proved that a photon is effectively transported in the long quartz bar, preserving its original photon direction. The Babar detector at SLAC adopted this concept for PID, and introduced a large stand-off photon detection system in order to ineffectuate the finite size of the quartz cross-section onto the ring image [3]. On the other hand, Kamae et al. [4] proposed a compact focussing type of DIRC using a small mirror instead of a large stand-off system. This approach suits particles incident normal to the quartz bar, but for inclined particles the ring image is distorted due to the finite quartz cross-section. To make this image-fusing effect insignificant, the size and geometry of the focussing mirror become unfeasible [5]. We propose here a new concept for the Cherenkov-ring image detector, the TOP detector. The two-dimensional information of the ring image is represented by the TOP of the Cherenkov photon and its horizontal angle ($\Phi$; see, Fig.1). The use of TOP information with an appropriate focussing mirror makes the image-fusing effect be disregarded and compactification can be realized while maintaining the high enough PID ability. We describe below the basic concept, characteristic features, practical configurations and simulation results on the TOP detector, and discuss some issuses for making it a realistic device. We should mention a paper by Honsheid et al. [6], which we noticed after our detailed analysis had been completed. It discusses the Cherenkov Correlated Timing (CCT) detector: CCT [6, 7] is a kind of TOF counter with a single phototube readout for detecting early arriving photons of Cherenkov radiation. In ref.[6] a similar concept as ours can be found, although no detailed study is presented and the focussing mirror approach is not introduced. \section{Conceptual Design and Expected Properties} The principal structure of the TOP counter in the ($x, y, z$) coordinate system is illustrated in Fig.1. When a charged particle passes through the radiator bar, Cherenkov photons are emitted in a conical direction defined by the emission angle ($\theta_{\rm {C}}$), where $\cos\theta_{\rm {C}}$=1/$n\beta$, $n$=refractive index. Then, photons propagate to both or either ends by means of total reflection on the internal bar surface. Photons propagating backward are reflected by a flat mirror at the end. At the forward end, photons are horizontally focused by a butterfly-shaped mirror onto a plane where position-sensitive multi-anode phototubes are equipped to measure TOP as a function of the $\Phi$-angle. \subsection{TOP and $\Phi$} The TOP is determined only by the z-component of the light velocity in the quartz bar as \begin{eqnarray} TOP &=& \left( \frac{L}{c/n(\lambda)} \right)\times \left(\frac{1}{q_{\rm {z}}}\right),\nonumber\\ &=& 4.90(ns)\times \frac{L(m)}{q_{\rm {z}}}, \end{eqnarray} where $L$ is the overall distance from the photon emission point to the detector, {\it n}$(\lambda)$ is the refractive index of the quartz at wavelength $\lambda$, $c$ is the light velocity in a vacuum, and $q_{\rm {z}}$ is the directional z-component of the photon emission. Denoting the polar and azimuthal angles of an incident charged particle as $\theta_{\rm {inc}}$ and $\phi_{\rm {inc}}$ (see, Fig.1), respectively, the directional components of the photon velocity $q_{\rm {i}}$ ({\rm i}={\it x, y, z}) are written as \begin{eqnarray} q_x &=& q_{x'}\cos\theta_{\rm {inc}}\cos\phi_{\rm {inc}} -q_{y'}\sin\phi_{\rm {inc}} +q_{z'}\sin\theta_{\rm {inc}}\cos\phi_{\rm {inc}}, \nonumber\\ q_y &=& q_{x'}\cos\theta_{\rm {inc}}\sin\phi_{\rm {inc}} +q_{y'}\cos\phi_{\rm {inc}} +q_{z'}\sin\theta_{\rm {inc}}\sin\phi_{\rm {inc}}, \\ q_z &=& -q_{x'}\sin\theta_{\rm {inc}} +q_{z'}\cos\theta_{\rm {inc}}, \nonumber \end{eqnarray} where $q_i$'s ($i$=$x^{\prime}$, $y^{\prime}$, $z^{\prime}$) are the directional components of the photon emission in the frame where the particle moves along the z$^{\prime}$-axis: \begin{eqnarray} q_{x'} &=& \sin\theta_{\rm {C}}\cos\phi_{\rm {C}}, \nonumber\\ q_{y'} &=& \sin\theta_{\rm {C}}\sin\phi_{\rm {C}}, \\ q_{z'} &=& \cos\theta_{\rm {C}}. \nonumber \end{eqnarray} The horizontal and vertical photon angles at the bar end are then given as \begin{eqnarray} \Phi &=& arctan\left(\frac{q_x}{q_z}\right),\\ \Theta &=& arctan\left(\frac{q_y}{q_z}\right). \nonumber \end{eqnarray} The $\phi_{\rm {C}}$ angle of Cherenkov photons distributes uniformly over 2$\pi$. When we fix $\phi_{\rm {C}}$, individual three-directional components ($q_i$) are uniquely related to $\theta_{\rm {C}}$. Therefore, a measurement of any two directional components or any two combinations of them among eqs.(3) and (4) provides information about $\theta_{\rm {C}}$. Accordingly, in a TOP concept, we use the two parameters ($q_z$, $\Phi$) for extracting $\theta_{\rm {C}}$, while in DIRC, ($\Theta$, $\Phi$) are used. The advantage of our approach is that the $q_z$ measurement is not seriously influenced by the quartz bar cross-section. \subsection{Accuracy of the TOP measurement} Here, we estimate various contributions to the accuracy of the TOP measurement with a detector which we are preparing for R\&D work. By modelling the detector as presented in Fig.1, the parameters of each element and the expected machining or construction accuracy are assumed as described below. The size of the radiator bar is 20 mm-thick (in $\it {y}$), 60 mm-wide (in $\it {x}$), and about 3 150 mm (in $\it {z}$), on which a butterfly-shaped forcussing mirror block and a flat reflection mirror are attached at the forward and backward ends, respectively. The focussing mirror comprises two sub-elements, each forming an arc of a circle with a radius of 250 mm. As a PID detector for the Belle experiment, we suppose these detectors to be placed at R=1 m from the beam axis to form a cylindrical configuration, and due to an asymmetric-energy collider the detector is asymmetrically configured with respect to the z-direction to cover the polar angle from 30$^{\rm {o}}$ to 140$^{\rm {o}}$ in the lab-frame. For simplicity of calculation, the photon-propagation length in the focussing mirror is set as 20 cm for all photons. The material of the radiator bar and the mirrors is synthetic optical quartz (SUPRASIL-P30) manufactured by Shin-etsu Quartz Co. and their surfaces are polished by Bikovsky-Japan Co. with a flatness of $\pm$5 $\mu$m. The quartz has a refractive index of $n$=1.47 at a wavelength ($\lambda$) of 390 nm and a transmittance of $\>$90\%/m at $\lambda$=250 nm. The reflecting surface of the mirrors comprises aluminum of a few 100 nm-thick on 200 $\rm {\AA}$-thick chromium by evaporation, which provides a reflection efficiency of 98\%. The photon detector (R5900U-00-L16) is a multi-anode (16 channel) linear array phototube with a bialkali window by Hamamatsu Photonics Co. Although this tube does not function under a magnetic field, it has a quite suitable TTS for our early R\&D work without a magnetic field. \\ Now, by taking the above structure of the TOP counter into account, we estimate various contributions to the accuracy of the PID measurement. \\ \noindent {\em (1) $\lambda$-dependence of n} \par The refractive index ($n(\lambda)$) of the quartz varies depending on $\lambda$. As a result, the Cherenkov angle ($\theta_{\rm {C}}$) and the total reflection angle vary for individual photons in addition to the propagation velocity. The effect to TOP is \begin{equation} \frac{\sigma(TOP)}{(TOP)} = \left(\frac{\sigma_n}{n}\right ) \times (1-\alpha), \end{equation} where \begin{equation} \alpha = n \left(\frac{\partial q_z /\partial n}{q_z}\right). \end{equation} Accounting for the $\lambda$-dependence of the Cherenkov photon yield and the $\lambda$-dependent quantum efficiency of the phototube, we take $n$=1.47$\pm$0.008 ($\Delta n/n$= 5$\times 10^{-3}$) for the numerical calculation described in the next subsection. \\ \noindent {\em (2) Angular resolution} \par The $\Phi$-angle resolution depends on the position resolution of the phototube and the focussing accuracy of the mirror. The designed focussing mirror has a resolution of $\sigma$=0.8 mm with a dispersion of $d\Phi/dx$=0.5$^{\rm {o}}$/mm. Since the phototube has a 0.8 mm anode width, as described below, the $\Phi$ resolution is evaluated as $\sigma_{\Phi}\approx 0.5^{\rm {o}}$. The $\Phi$-dependent behavior of this effect on TOP shows a steep increase with $\Phi$, but a very small contribution at small $\Phi$-angles, as numerically discussed in the next subsection. This limits the $\Phi$ aparture to be $\mid\Phi\mid \leq$ 45$^{\rm {o}}$. \\ \noindent {\em (3) TTS of the phototube} \par The phototube (R5900U-00-L16) has a 16 mm $\times$ 16 mm photo-sensitive area, which is segmented into 16 strips with an area of 0.8 mm $\times$ 16 mm and a 0.2 mm space between adjacent strips. The quantum efficiency ($QE$) is 25\% at 400 nm and $\sigma_{\rm {TTS}}\approx$70 ps with a pulse rise-time of 0.6 ns for a single photoelectron. \\ \noindent {\em (4) Timing accuracy of the start signal} \par We expect a start timing accuracy of 25 ps from the RF signal of the accelerator. \\ \noindent {\em (5) Effect of the quartz thickness} \par Because relativistic particle takes 66 ps to pass through the 20 mm-thick quartz, the time spread of the Cherenkov photon emission is estimated to be $\sigma_{\rm {thick}}$= 20 ps (=66 /$\sqrt{12}$). When the incident particle is inclined with respect to the normal of the bar surface, the time-spread becomes much smaller due to cancelleration of the particle flight-time and the photon propagation-time. For instance, $\sigma_{\rm {thick}}$ is 6 ps and 10 ps at $\theta_{\rm {inc}}$= 30$^{\rm {o}}$ and 60$^{\rm {o}}$, respectively.\\ \noindent {\em (6) Effect of the charged-particle tracking accuracy} \par In the Belle detector, the z position of a charged-particle track is measured by the central drift chamber (CDC) with an accuracy of about 2 mm. This means that it corresponds to a time uncertainty of 10 ps. On the other hand, the accuracy of the momentum measurement is related to the accuracy of the Cherenkov angle as \begin{equation} \sigma_{\theta_{\rm {C}}} = \left( \frac{\cot\theta_{\rm {C}}} {\gamma^2}\right) \times \left(\frac{\sigma_{\rm {p}}}{p}\right). \end{equation} Since the Cherenkov angle satisfies $\cot\theta_{\rm {C}}\sim 1$ for our particles and the CDC provides $\sigma_{\rm {p}}$/p$\approx 5\times10^{-3}$, $\sigma_{\theta_{\rm {C}}}$ can be ignored compared with the difference in the Cherenkov angle between $\pi$ and K. The former is on the order of $10^{-2}-10^{-3}$ smaller than the latter. For instance, $\Delta( \theta_{\rm {C}}(\pi) - \theta_{\rm {C}}(K) )$ = 6.5 mrad, while $\sigma_{\theta_{\rm {C}}}$ is $7\times 10^{-3}$ mrad and $7\times 10^{-2}$ mrad for 4 GeV/c $\pi$ and K, respectively. The accuracy of the incident-angle measurement contributes to the accuracy of the TOP measurement as \begin{equation} \frac{\sigma(TOP)}{(TOP)} = -\left(\frac{1}{q_z}\right)\times \left(\frac{\partial q_z}{\partial\theta_{\rm {inc}}}\right) \sigma_{\theta_{\rm {inc}}} \end{equation} and \begin{equation} \frac{\sigma(TOP)}{(TOP)} = -\left(\frac{1}{q_z}\right)\times \left(\frac{\partial q_z}{\partial\phi_{\rm {inc}}}\right) \sigma_{\phi_{\rm {inc}}}=0. \end{equation} The effect of any $\theta_{\rm {inc}}$ ambiguity on TOP is evaluated to be less than 10 ps in the case of the Belle detector [1]. It is an order-of-magnitude smaller than the contribution from item (3) ($\sigma_{\rm {TTS}}\approx$70 ps), and is thus disregarded. Furthermore, it is, on the other hand, a specific feature of our approarch that the TOP is independent of $\phi_{\rm {inc}}$, as is obvious from eq.(2). Therefore, the $\phi_{\rm {inc}}$ ambiguity does not intrinsically affect the TOP measurement. \subsection{Expected separability} From the above, the uncertainty of the TOP measurement arises predominantly from the effects of items (1), (2) and (3). In order to understand the TOP behavior and to evaluate the size of the above effects to TOP and the PID power, a calculation is performed with the following conditions. The $\lambda$ dependence of the refractive index is disregarded, but its effect is represented by $n$$=1.47 \pm 0.008$ in eq.(5); the $\Phi$ resolution is set as $\sigma_{\Phi}=0.5^{\rm {o}}$; the effects of the items (4) and (5) and the digitization bin of 25 ps are included into (3) to make an effective $\sigma_{\rm {TSS}}$=80 ps (denoted hereafter as (3)$^{\prime}$); the efficiencies of photon propagation and reflection are assumed to be 100\%. For PID in the collider geometry, in addition to the TOP measurement, we added to the TOP the particle's TOF from the colliding point to the quartz radiator placed at R=1 m. \\ \noindent {\em 2.3.1.~~$\delta$(TOP+TOF) and individual effects} {\ } \par In Fig.2, the individual uncertainties of items (1), (2) and (3)$^{\prime}$ are shown for various cases of the incident angle ($\theta_{\rm {inc}}$), in addition to the TOP+TOF and TOP differences between 4 GeV/c $\pi$ and K (denoted hereafter as $\delta$(TOP +TOF) and $\delta$(TOP), respectively). (The propagation length ($L$) in Figures (a-e) is uniquely determined for each $\theta_{\rm {inc}}$ by assuming the particle coming from the colliding point under our geometory. On the other hand, the length ($L$) in Figures (f, g) is not real, but hypothetical, only for discussion purposes.) The asymmetries seen in the figures are caused by deflection of the azimuthal angle of the incident particle due to an axial magnetic field of 1.5 T. At normal incidence (see, Fig.2(c), (d) and (g)), effect (1) is an order of magnitude smaller than $\delta$(TOP); this relation holds irrespective of $L$, because of their proportionality to $L$. Effect (2) has a strong $\Phi$ dependence. In the region of the small $\Phi$-angle the effect is negligible, while in the large-$\Phi$ region it becomes dominant in determining the resolution. Mostly from this fact, the $\Phi$-angle aperture is limitted to $\mid\Phi\mid \leq 45^{\rm {o}}$, where the TOP measurement uncertainty is comparable to or less than $\delta$(TOP+TOF) under most of the incident-particle conditions. On the other hand, because $\theta_{\rm {inc}}$ deviates from normal incidence the value of $\alpha$ decreases, and item (1) gives almost the same order of magnitude as $\delta$(TOP+TOF) at $\theta_{\rm {inc}}=30^{\rm {o}}$ (Fig.2(a)), and several to 10-times larger size at $\theta_{\rm {inc}}=140^{\rm {o}}$ (Fig.2(e)). Note that the sign of $\delta$(TOP) at small $\Phi$ reverses at around $\theta_{\rm {inc}}\sim 45^{\rm {o}}$ or $\theta_{\rm {inc}} $-90$^{\rm {o}}\sim$ 45$^{\rm {o}}$, since $\theta_{\rm {inc}}$ almost coincides with the Cherenkov angle ($\sim 45^{\rm {o}}$) and then the relative magnitude of $q$$_z$ of Cherenkov photons from $\pi$ and K reverses (see, Fig.2(a, e, f)). Thus, in the case of large $L$, as in Fig.2(f), $\delta$(TOP) is cancelled with $\delta$(TOF), which reduces the PID power compared with the small $L$ case, as in Fig.2(a). \\ \noindent {\em 2.3.2.~~$\pi$/K separability} {\ } \par In the previous discussion, the expected resolution for the TOP measurement is given for the detection of a single Cherenkov photon. However, when we use the position-sensitive photon detector, we can use every Cherenkov photon to obtain PID information. In order to estimate the PID performance, we define the $\pi$/K separability ($S$$_{\rm {o}}$) as \begin{equation} S_{\rm {o}} = \sqrt{\left \{ \sum_{\rm {i}}\left(\frac{\delta(TOP+TOF) ^i} {(\sigma_{\rm {T}})^{\rm {i}}}\right)^2\right \}\cdot \kappa}. \end{equation} In the calculation, a single Cherenkov photon is assigned at every one-degree in $\phi_{\rm {C}}$ over a range of $2\pi$, and individual $\delta$(TOP+TOF)$^i$ and the total uncertainty $(\sigma_{\rm {T}})^i$ arising from the above-mentioned items are obtained for every $i$-th photon. The summation is taken over for the photons with $\mid\Phi\mid\leq$45$^{\rm {o}}$, which are totally reflected on the internal surface of the quartz bar. Because the number of Cherenkov photons for the 20 mm-thick quartz bar is fixed as $N_{\rm {o}}^{\gamma}=160/(\sin\theta_{\rm {inc}}\times QE)$ with $QE$=25\%, $\kappa$ is a normalization factor of $N_{\rm {o}}^{\gamma}/360$. In Fig.3 the resultant $S$$_{\rm {o}}$ for $\pi$/K is presented as a function of the particle momentum ($p$) for three different $\theta_{\rm {inc}}$'s. For a normal incident particle, two cases are considered separately: (a) Cherenkov photons directly come to the photon detector (FW), or (b) go to the backward end and then return to the detector (BW). A better separation is obtained for (b) because of the longer propagation distance of photons. The incident-angle dependence of $S$$_{\rm {o}}$ is shown in Fig.4 for $p$=2, 4 and 5 GeV/c particles, where the FW and BW cases are indicated by the solid and dotted lines, respectively. The decrease in $S_{\rm {o}}$ around $\theta_{\rm {inc}}$=50$^{\rm {o}}$ is due to destructive correlation between $\theta_{\rm {inc}}$ and $\theta_{\rm {C}}$, as mentioned above. The $S_{\rm {o}}$ variation on the propagation length ($L$) is calculated by varing the quartz length for $\theta_{\rm {inc}}$= 30$^{\rm {o}}$, 60$^{\rm {o}}$ and 90$^{\rm {o}}$(FW), as shown in Fig.5. At normal incidence, $\delta$(TOP+TOF) becomes large relative to the effects of (2) and (3)$^{\prime}$ as $L$ increases, and, accordingly, $S_{\rm {o}}$ is improved. On the other hand, at around $\theta_{\rm {inc}}$=30$^{\rm {o}}$ the effect (1) becomes large with $L$, and $S_{\rm {o}}$ decreases, as can be seen in Fig.2(a) and (f). \section{Simulation Study} Based on the basic concepts considered in the preceeding section, we carried out a simulation study on the realistic environment. Photons are generated following to the Cherenkov spectrum (d$N$$(\lambda)$/d$\lambda$) convoluted by the quantum efficiency $QE$($\lambda$) of the phototube. The Cherenkov angle is determined by using the $\lambda$-dependent quartz refractivity ($n$($\lambda$)). The effect of the quartz thickness is naturally implemented in the simulation procedure. The effective TTS of the phototube is therefore set to be $\sigma_{\rm {TTS}}=$75 ps, instead of the previous value of 80 ps, including the ambiguity of only the timing reference signal; this time resolution is considered in smearing the calculated TOP+TOF values with a Gaussian of $\sigma=\sigma_{\rm {TTS}}$. The phototube is assumed to detect only the earliest photon if multi photons hit the same anode channel. The focussing resolution is also included by smearing the calculated $\Phi$-angle with a Gaussian of $\sigma_{\Phi}=0.5^{\rm {o}}$ as well. Digitizations of the arrival time and $\Phi$-angle of photons are made in 25 ps and 0.5$^{\rm {o}}$ bins, respectively. All other treatments are the same as in the previous calculation. To display the various results for $\pi$ and K tracks with the same momentum, we set parameters (as an example, $p$=4 GeV/c and $\theta_{\rm {inc}}=90^{\rm {o}}$) and require to detect only FW photons, unless the conditions are specified. Figure 6(a) shows a typical $\Phi$-angle distribution of the number of detected photons for the $\pi$ track. The number of produced photons is 137 in total, while the detected FW photons within $\mid\Phi\mid$=45$^{\rm {o}}$ are 33. The photons distribute rather uniformly with an average of less than one per bin ($\Delta\Phi$=0.5$^{\rm {o}}$). The TOP+TOF is shown in Fig.6(b), where the circles and crosses are the detected photons radiated by the $\pi$ and K track, respectively. They form two slightly aymmetric parabola-like distributions due to an inclined $\phi_{\rm {inc}}$-angle by a deflection effect due to the magnetic field. It is 4.3$^{\rm {o}}$ in this case. The dotted curves are the TOP+TOF calculated in the previous section by disregarding the $\lambda$ dependences and generating photons with $\theta_{\rm {C}}$=1/$n$$\beta$ ($n$=1.47). The upper two dotted curves in the figure correspond to the calculated TOP+TOF for the BW photons. The deviations of TOP+TOF for an incident $\pi$ and K from the calculated distribution assuming $\pi$, $\Delta$(TOP+TOF)$_{\pi/K}(\pi)$, are presented in Fig.6(c). In finding the deviation, two solutions exist at every $\Phi$-angle due to the above-mentioned deflection effect, so that we choose the smaller deviation as the correct one. The circles and closes are for photons emitted by a $\pi$ and a K, respectively. The former distributes around the null value, while the latter shifts to the positive side, as expected. The solid and dotted histograms in Fig.7 show the distribution of deviations, $\Delta$(TOP+TOF)$_{\rm {\alpha}}(\beta)$, for photons within $\mid\Phi\mid\leq$45$^{\rm {o}}$, where $\alpha$, $\beta$=$\pi$ or K; also, $\alpha$ is the particle species of the track and $\beta$ is the particle species assumed. The solid histogram presents the results under the correct assumption and the dotted histogram presents those under an incorrect assignment. Figures 7(a) and (b) show typical deviations for an event, and Figs.7(c) and (d) present the results for the accumulation of 100 tracks. We obtain the probability function ($P$$(\pi$/K)) by fitting the results of the right assumption in Figs.7(c) and (d) with Gaussian functions, as shown in the figures by the solid curves. Based on this probability function, a log-likelihood ($ln\cal L$) under the assumption of $\pi$ and K's are calculated for 400 tracks each for incoming $\pi$ and K. Figure 8(a) shows a scatter plot of the log-likelihoods thus obtained in the case of $p$=4 GeV/c and $\theta_{\rm {inc}}$= 90$^{\rm {o}}$ for the FW photons; a clear separation between $\pi$ and K can be seen. The separability ($S$) is calculated by taking the difference in the log-likelihood ($\Delta ln\cal L$) between the two assumptions (see Fig.8(b)) and using $S$=$\sqrt{2\Delta ln\cal L}$; $S$=5.7 in this case. Thus, the calculated $S$'s for different $p$'s and $\theta_{\rm {inc}}$'s are indicated in Fig.9 for both FW and BW photons. \section{Discussion and Summary} We have shown that a high PID capability is attainable by using a (TOP, $\Phi$) measurement of the Cherenkov ring image. Although R\&D is necessary, especially, on a high-quality position-sensitive phototube operable in a magnetic field and on an optimum focussing mirror, a good $\pi$/K separation is well expected at the Belle barrel detector. Some remarks are given below: \begin{itemize} \item The development of a position-sensitive phototube is most important to realize the separability so far presented. It has twofold issues. One is to make it operable under a magnetic field and we have being developed a fine-mesh-type multi-anode phototube for this purpose. The other is to remove or reduce the dead space dominated by the phototube package. The phototube (R5900U-00-L16) has a surface size of 30 mm$\times$30 mm in which the sensitive area is 16 mm$\times$16 mm, so that the effective area at our R\&D counter becomes only about 40\%. About half is in horizontal and 80\% in the vertical. The required size of the phototube is 20 mm$\times$100 mm with a bin step of 1 mm for detecting photons with $\mid\Phi\mid\leq$45$^{\rm {o}}$. \\ \item Adding the TOF information to the TOP information helps the PID, especially for low-momentum particles. However, for high-momentum particles its effect is not appreciable in our configuration. For instance, at 4 GeV/c tracks with $\theta_{\rm {inc}}$=90$^{\rm {o}}$, $\delta$(TOF) are 23 ps, which is only about 1/3 of $\sigma_{\rm {TTS}}$. However, by setting a longer flight-length (R), a sizable improvement can be realized, as listed in Table 1; in such cases the TOP detector acts as a kind of a high-resolution TOF counter by means of Cherenkov radiation, since the Cherenkov-angle difference does not produce sufficient $\delta$(TOP). \\ \item The $\pi$/K separability by the TOP measurement has an angular dependence, as can be seen in Fig.9, while DIRC gives a more-or-less flat distribution with the incident angle. Especially, TOP shows a separability decrease at $\theta_{\rm {inc}}\sim$ 50$^{\rm {o}}$, and it is in principle inevitable. On the other hand, TOP can give a larger $S$ at normal incidence. Furthermore, in this case, TOP has the merit of doubling $S$ by detecting the both FW and BW going photons those of which have an arrival time difference of about 20 ns, so that they can be easily identified by using a multi-hit TDC with a high resolution. \\ \item One possible way to improve the separability for tracks with a large incident angle is to equip an inclined surface, say, 45$^{\rm {o}}$ to the reflection mirror at the backward end. This will make the reclining Cherenkov ring image stand up to the normal direction, so that it becomes a case close to a normal incidence track and photons propagate the full bar length. Although an opposite effect occurs for a track with $\theta_{\rm {inc}}\sim$90$^{\rm {o}}$, a rather sufficient TOP difference is produced by propagating a certain long distance before the photons arrive at the reflection mirror. Optimization of the inclining angle of the reflection mirror should be examined. \\ \item Because the present design of the focussing mirror has a horizontally-extended structure, it is not suitable as a barrel detector, and therefore we need to find a practical optical structure for the compact system. Although the vertically configured mirror avoids this geometrical problem, the choise of the horizontal configuration is due to the fact that the TOP+TOF distribution for the vertical configuration has a larger differentiation to the angle $\Theta$, as shown in Fig.10, so that the TOP+TOF measurement uncertainty becomes larger for an angle resolution of $\Delta\Theta$ at the same accuracy as $\Delta\Phi$. \par The simplest and practical solution for the above is to tilt the counter around the z-axis by an angle of, for instance, 18.4$^{\rm {o}}$(=arctan(2 cm-thick/6 cm-width)) in our case, so that every mirror can avoid to interfere with each other. \par A TOP counter with a butterfly-shaped mirror, illustrated in Fig.1, has a symmterical geometry with respect to the z-axis. By dividing the butterfly mirror along the z-axis, a half butterfly-shaped counter can be constructable, although the bar width also becomes half so as to have sufficient $\Phi$-angle resolution. \\ \item For simplicity of the calculation, the probability function (P$(\pi$/K)) is obtained in the previous section from the whole $\Delta$(TOP+TOF) distributions within $\mid\Phi\mid$=45$^{\rm {o}}$, while the spread of the distribution has the $\Phi$-dependence. Therefore, if we properly take account of this $\Phi$-dependence in the probability examination, better separability will be obtained. For instance, the rms of the distribution varies from 85 ps to 140 ps within the accepted $\Phi$-angle region in the case of 4 GeV/c normal incident tracks. \\ \item We assumed the detection of only the earliest single photon at each phototube anode in the previous section. It causes the loss of some photon fraction. For instance, it is about 30\% of $\approx$ 110 detectable photons at $\theta_{\rm {inc}}$=30$^{\rm {o}}$, while it is about 10\% of $\approx$56 and $\approx$30 photons at $\theta_{\rm {inc}}$=60$^{\rm {o}}$ and 90$^{\rm {o}}$, respectively. \\ \item It may be interesting to use a scintillator instead of the quartz material, as mentioned at the beginning, although the situation is somewhat obscure without a detailed study. Because of its finite decay-time and non-unique photon emission-angle, the arrival-time distribution at individual $\Phi$-angles will spread. However, it is the usual situation at the ordinary TOF counter. Since a large number of photons are provided in this case, when an optimized configuration is found to have a sufficient number of photons at each $\Phi/\Theta$-segmentation, the TOF resolution could be improved by a factor of square-root of the number of detection segments. Angular segmentation could be arranged wider than our TOP case, so that the focussing mirror could be smaller and the number of phototubes could be less. Needless to say, the surface of the scintillator should be polished enough to allow photons to be internally reflected. \end{itemize} Although the above-mentioned issues concerning the phototube and the mirror structure remain, the proposed detector can be fully applicable as a PID device for experiments under a non-magnetic enviornment, mostly for fixed-target experiments. At fixed-target experiments, because incoming particles are incident nearly normal to the detector, the maximum separation power can be extracted. With the use of a horizontal focussing mirror together with the fact that the time ambiguity of Cherenkov radiation along the pass length is disregarded compared with the phototube TTS, the cross section of the quartz radiator does not have any effects on TOP so that we can successfully avoid the unwanted image-fusing effect. The DIRC counter [2, 3, 4] verified that the Cherenkov image is well-transported by internal total reflection, and the CCT counter [6, 7] exhibited that a time measurement for such Cherenkov photons is feasible. While the TOP detecter proposed here is based on a comprehensive compilation of the technical verification by those detecters, it can be used to exploit a new approach for measuring the TOP and $\Phi$-angle correlation by making use of the horizontal focussing mirror. As a result, it would not only provide high particle identification ability, but also a more compact and flexible detector compared with the DIRC and the gas Cherenkov imaging detecter. {\ }\\ {\ }\\ \noindent {\bf Acknowledgements} {\ } \par We greatly thank Prof. Yoshitaka Kuno for informing us about the paper on the CCT counter. We also thank Prof. Katsumi Tanimura for his consultation on optics relating to our materials and devices. This work was supported by Grant-in-Aid for Scientific Research on Priority Areas (Physics of CP violation) from the Ministry of Education, Science, and Culture of Japan. {\ }\\
\section{Introduction} The {\it Hubble Space Telescope (HST)} offers a unique opportunity for studying very young massive star formation regions in the outer galaxies. The fact that the Small Magellanic Cloud (SMC) is the most metal-poor galaxy observable with very high angular resolution makes it an ideal laboratory for investigating star formation in very distant galaxies reminiscent of those populating the early Universe. Our search for the youngest massive stars in the Magellanic Clouds started almost two decades ago on the basis of ground-based observations. This led to the discovery of a distinct and very rare class of H\,{\sc ii} regions in these galaxies, that we called high-excitation compact H\,{\sc ii} ``blobs'' (HEBs). So far only five HEBs have been found in the LMC: N\,159-5, N\,160A1, N\,160A2, N\,83B-1, and N\,11A (Heydari-Malayeri \& Testor 1982, 1983, 1985, 1986, Heydari-Malayeri et al.\,1990) and two in the SMC: N\,88A and N\,81 (Testor \& Pakull 1985, Heydari-Malayeri et al.\,1988a). These objects are expected to harbor newborn massive stars. The first part of our {\it HST} project studying the H\,{\sc ii}\, ``blobs'' was devoted to the SMC N\,81 (Heydari-Malayeri et al. \cite{hey99}, hereafter Paper I). The Wide Field Planetary Camera\,2 (WFPC2) observations allowed us to resolve N\,81 and discover a tight cluster of newborn massive stars embedded in this nebula of $\sim$\,10$''$\hspace*{-.1cm}\, across. The WFPC2 images also uncovered a striking display of violent phenomena such as stellar winds, shocks, and ionization fronts, typical of turbulent starburst regions. The SMC ``blob'' N\,88A is part of the H\,{\sc ii}\, region N\,88 (Henize \cite{hen}), or DEM\,161 (Davies et al. \cite{dav}) which lies in the Shapley Wing at $\sim$\,2$^{\circ}$\hspace*{-.1cm}.2 (2.4\,kpc) from the main body of the SMC. Other H\,{\sc ii}\, regions lying towards the Wing are from west to east N\,81, N\,84, N\,89, and N\,90. The HEB nature of N\,88A was first recognized by Testor and Pakull (\cite{tes}) who used CCD imaging at $\sim$\,2$''$\hspace*{-.1cm}\, resolution through H$\alpha$, H$\beta$, and [O\,{\sc iii}]\, filters and IDS spectroscopy (4$''$\hspace*{-.1cm}\,$\times$\,4$''$\hspace*{-.1cm}\, aperture) to study the central component N\,88A. They found a high excitation object ([O\,{\sc iii}]/H$\beta$\,=\,7.8) with an interstellar extinction $A_{V}$\,=\,1.7 mag. The chemical abundances in N\,88 had previously been determined by Dufour \& Harlow (\cite{dh}) who, using a 10$''$\hspace*{-.1cm}\,$\times$\,79$''$\hspace*{-.1cm}\, slit, found a low-metal content typical of the chemical composition of the SMC. CCD photometry and spectroscopy of 10 stars lying around N\,88A were carried out by Wilcots (\cite{wil}) using the {\sc ctio} 90\,cm telescope. However, the ground-based observations in general were unable to reveal the internal morphology and stellar content of N\,88A. The {\it HST} was used for imaging and FOS spectroscopy of N\,88A (Kurt et al. \cite{kurt}). These pre-{\sc costar} observations, in spite of the effort made in data analysis, could not clearly show the internal details of N\,88A. Garnett et al. (\cite{gar}) revisited the chemical abundances in N\,88A on the basis of {\it HST} ultraviolet FOS (0$''$\hspace*{-.1cm}.7\,$\times$\,2$''$\hspace*{-.1cm}.0\, aperture) and ground-based spectra. In this paper we present recent {\it HST} observations (GO\,6535) of N\,88A and its surroundings. In the following sections we elaborate on the extinction and emission properties of each component and suggest a plausible scenario on the star formation history of this region. \begin{figure*} \caption{A set of {\it HST} views of the SMC nebula N\,88. {\bf (a)} A ``true-color'' image of the whole WFPC2 field based on exposures taken with filters H$\alpha$\, (red), [O\,{\sc iii}]\, (green), and $U$ (blue). {\bf (b)} A False color image of the PC field in H$\alpha$. Field size $\sim$\,37$''$\hspace*{-.1cm}\,$\times$\,37$''$\hspace*{-.1cm}\, ($\sim$\,11\,$\times$\,11 pc$^{2}$). {\bf (c)} A close-up of the H$\alpha$\, image presented in Fig. 1b showing N\,88A and its neighboring N\,88B. Field size $\sim$\,12$''$\hspace*{-.1cm}\,$\times$\,12$''$\hspace*{-.1cm}\, ($\sim$\,3.5\,$\times$\,3.5\,pc$^{2}$). The star lying towards N\,88B is \#55 (see text). } \label{rgb} \end{figure*} \section{Observations and data reduction} The observations of N\,88A described in this paper were obtained with WFPC2 on board the {\it HST} on August 31, 1997 using the wide- and narrow-band filters (F300W, F467M, F410M, F547M, F469N, F487N, F502N, F656N). The observational techniques, data reduction procedures, and photometry are similar to those explained in detail in Paper I. A composite image is presented in Fig.\,\ref{rgb}. The ESO EFOSC Camera at the 3.6\,m telescope was also used on 4 June 1988 for imaging N\,88 through a narrow H$\alpha$\, filter (ESO\#507, $\lambda$\, 6565.5\,\AA, $\Delta$$\lambda$\,=\,12\,\AA) with exposure times of 1 and 5 minutes. The detector was a RCA CCD chip (\#11) with 0$''$\hspace*{-.1cm}.36 pixels and the seeing was $\sim$\,1$''$\hspace*{-.1cm}.5 ({\sc fwhm}). This H$\alpha$\, image is displayed in Fig.\,\ref{efosc}. Due to its relatively short exposure, this image displays only the brightest part of the H\,{\sc ii}\, emission. \section{Results} \subsection {Morphology} N\,88 is a relatively large concentration of ionized gas with several components (Fig.\,\ref{rgb}a). From the central region emanate a number of fine-structure filaments running southwards over 40$''$\hspace*{-.1cm}\, ($\sim$\,10 pc) which can be seen in the ``true-color'' composite image (Fig.\,\ref{rgb}a). The larger field of the H$\alpha$\, image obtained with EFOSC (Fig.\,\ref{efosc}) shows a veil of thin filaments curling southwards over more than 20 pc and brightening at some points. The main component, N\,88A, is a compact, high excitation H\,{\sc ii}\, region $\sim$\,3$''$\hspace*{-.1cm}.5\, ($\sim$\,1\,pc) in diameter surrounded by seven diffuse H\,{\sc ii}\, regions, labelled B to H in Fig.\,\ref{rgb}b. N\,88A has a complex morphology. An absorption lane crossing the nebula from north to south appears as an undulating yellow structure in Fig.\,\ref{rgb}c (see Sect. 3.2 for more details). West of this structure lies the brightest part of N\,88A, a small core of diameter $\sim$\,0$''$\hspace*{-.1cm}.3 (0.08 pc), especially apparent on the H$\alpha$\, image (white spot in Fig.\,\ref{rgb}c; see also Fig.\,\ref{core}). N\,88A is clearly ionization-bounded to the north-west since the sharp edge visible in Fig.\,\ref{rgb} indicates an ionization front in that direction. It is limited to the south-east by the weaker component B. N\,88B resembles a hollow sphere -- a shell -- centered on the bright star \#55 (see Sect. 3.4). N\,88A and N\,88B are clearly in interaction, as shown by the brightening of the shell between the two regions. Moreover, we note a high excitation narrow filament showing up in the [O\,{\sc iii}]\, emission north-east of N\,88B (Fig.\,\ref{rgb}a). The other components are situated farther away from N\,88A. N\,88\,E-F-G and H appear as more extended, diffuse, and spherical H\,{\sc ii}\, regions. Several lower excitation arc-shaped features and filaments emerging from N\,88A run outward in the north-east and south-west directions. These wind-induced structures are best seen in Fig.\,\ref{vent}, which presents an un-sharp masked image of N\,88A-B created from H$\alpha$. In this image large-scale structures have been suppressed by the technique explained in Paper I. Note also the mottled structure of the main component A, even in the direction of the absorbing lane, indicating a very inhomogeneous medium, both for gas and dust, with a typical cell size of 0$''$\hspace*{-.1cm}.4 (0.1 pc). \subsection {Nebular reddening} The Balmer H$\alpha$/H$\beta$\, intensity ratio map of N\,88A-B is presented in Fig.\,\ref{rapport}a. The most striking feature is the presence of a heavy absorption lane of $\sim$\,0$''$\hspace*{-.1cm}.7\,$\times$\,2$''$\hspace*{-.1cm}.3\, ($\sim$\,0.2\,$\times$\,0.7\,pc$^{2}$) in size, running in a north-south direction, which divides the bright N\,88A into two parts. The mean H$\alpha$/H$\beta$\, ratio in the lane is 7.10\,$\pm$\,1.42 (rms), corresponding to $A_{V}$\,=\,2.5 mag if the LMC interstellar reddening law is used (Pr\'evot et al. \cite{pre}), and reaches values as high as $\sim$\,10, or $A_{V}$\,$\sim$\,3.5 mag. The mean ratio for component A, 4.81\,$\pm$\,1.46, corresponds to $A_{V}$\,$\sim$\,1.5 mag. The extinction is also high towards component B, where H$\alpha$/H$\beta$\, keeps a relatively uniform value of 4.27\,$\pm$\,0.90 ($A_{V}$\,$\sim$\,1.1 mag). For comparison, previous lower resolution spectroscopic observations yielded $A_{V}$\,=\,1.1 mag (Dufour \& Harlow \cite{dh}, using a 10$''$\hspace*{-.1cm}\, wide slit) and $A_{V}$\,=\,1.7 mag (Testor \& Pakull \cite{tes}, 4$''$\hspace*{-.1cm}\,$\times$\,4$''$\hspace*{-.1cm}\, slit). The H$\alpha$/H$\beta$\, map was used to de-redden the H$\beta$\, flux on a pixel to pixel basis. The G component shows a sharp dividing line in the middle separating it into two distinct halves, one much fainter than the other. This feature should be due to absorbing dust. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{ms8496f2.eps}} \caption{A ground-based H$\alpha$\, image of N\,88 obtained with EFOSC attached to the ESO 3.6\,m telescope. Field size $\sim$\,1$'$\hspace*{-.1cm}.7\,$\times$\,2$'$\hspace*{-.1cm}.0\, (31\,$\times$\,37 pc$^{2}$). North is up and east is left. N\,88A and B appear as the bright ``blob'' at the top of the image. Note the absence of ionized gas and stars north-west of the ``blob''.} \label{efosc} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{ms8496f3.eps}} \caption{An un-sharp masking H$\alpha$\, image of N\,88A \& B which highlights the high spatial frequency filamentary patterns. Field size $\sim$\,11$''$\hspace*{-.1cm}\,$\times$\,7$''$\hspace*{-.1cm}.4\, (3.3\,$\times$\,2.2 pc). North is up and east is left. } \label{vent} \end{figure} \subsection{Ionized gas emission} The total H$\beta$\, flux of component A is F(H$\beta$)\,=\,3.45\,$\times$\,10$^{-12}$ erg\,\,s$^{-1}$\,cm$^{-2}$\, (accurate to $\sim$\,3\%). Correcting for the reddening (Sect. 3.2) gives a flux F$_{0}$(H$\beta$)\,=\,1.85\,$\times$\,10$^{-11}$ erg\,\,s$^{-1}$\,cm$^{-2}$. The total flux for both components A and B is F$_{0}$(H$\beta$)\,=\,1.97\,$\times$\,10$^{-11}$\,erg\,\,s$^{-1}$\,cm$^{-2}$. Thus, component B provides less than 10\% of the total H$\beta$\, energy. A Lyman continuum flux of $N_{L}$\,=\, 2.10\,$\times$\,10$^{49}$ photons\,\,s$^{-1}$\, can be estimated for component A, using a distance of 66\,kpc, if the H\,{\sc ii}\, region is ionization-bounded. A single O6V star can account for this ionizing flux (Vacca et al. \cite{vacca}, Schaerer \& de Koter \cite{sch}). Similarly, the Lyman continuum flux corresponding to component B is $N_{L}$\,$\sim$\, 3.5\,$\times$\,10$^{47}$ photons\,\,s$^{-1}$. If we take the estimated UV fluxes at face values, the exciting star of component B should be an early B type star. However, these should be considered as lower limits, since the H\,{\sc ii}\, regions are not ideally ionization-bounded. We find an rms electron density of 2700\,cm$^{-3}$\, for component A from the total H$\beta$\, flux, a $T_{e}$\,=\,14\,000\,$^{\circ}$\hspace*{-.1cm}\,K (Garnett et al. \cite{gar}), and assuming a radius of 0.5\,pc for the object. The corresponding total ionized mass of N\,88A is $\sim$\,45\,$M_{\odot}$. \begin{figure*} \resizebox{\hsize}{!}{\includegraphics{ms8496f4.eps}} \caption{Spatial variation in the extinction and ionization across N\,88A. The field size (13$''$\hspace*{-.1cm}.8\,$\times$\,13$''$\hspace*{-.1cm}.8 or $\sim$\,4\,$\times$\,4 pc) is almost identical to Fig.\,\ref{rgb}c. {\bf (a)} Balmer H$\alpha$/H$\beta$\, ratio of the inner parts of the nebulae. Darker color indicates higher H$\alpha$/H$\beta$. Note the presence of the narrow absorption lane where the extinction rises to $A_{V}$\,$\sim$\,3.5 mag. The mean value of extinction towards component A is about 1.5 mag. {\bf (b)} The [O\,{\sc iii}]/H$\beta$\, ratio. The mean value for component A is $\sim$\,7 and the ratio rises to as high as 9 at some points. Note the narrow filament north-east of component B.} \label{rapport} \end{figure*} The [O\,{\sc iii}]\lam5007/H$\beta$\, intensity map displays an extended high excitation zone towards N\,88A (Fig.\,\ref{rapport}b), where the ratio has a mean value of $\sim$\,7. The ratio peaks at some points to values as high as \,9. Taking $T_{e}$\,$\sim$\,14\,000\,$^{\circ}$\hspace*{-.1cm}\,K and $N_{e}$\,=\,2\,700\,cm$^{-3}$, a ratio [O\,{\sc iii}]/H$\beta$\,$\sim$\,7 indicates an ionic abundance O$^{++}$/H$^{+}$\,$\sim$\,9.4\,$\times$\,10$^{-5}$. Since the mean SMC oxygen abundance is O/H\,$\sim$\,10.5\,$\times$\,10$^{-5}$ (Dufour \cite{duf}), this means that $\sim$\,90\% of the oxygen atoms in N\,88A are in the form of doubly ionized O$^{++}$ ions, in agreement with the result of Garnett et al. (\cite{gar}). The [O\,{\sc iii}]/H$\beta$\, ratio for component B is comparatively much smaller, with a mean value of $\sim$\,4. The high excitation narrow filament emanating from component A is clearly visible in the [O\,{\sc iii}]/H$\beta$\, map, suggesting that the O$^{++}$ ions in the filament may be excited by shock collisions. \subsection{Stars} The {\it HST} images reveal tens of previously unknown stars towards the N\,88 complex. Many of them, especially the brightest ones, are gathered in several small groups often un-resolved in the EFOSC image (Fig.\,\ref{efosc}). The photometry obtained for the 79 brightest stars of the field using the filters wide $U$ (F300W), Str\"omgren $v$ (F410M), Str\"omgren $b$ (F467M), He\,{\sc ii}\, (F469N), and Str\"omgren $y$ (F547) is presented in Table\,\ref{phot} which also gives the coordinates (J2000) of each star. These stars are identified by their numbers in Fig.\,\ref{numbers} and Table\,\ref{phot}. The capital letters in the last column of the table identify the associated H\,{\sc ii}\, regions. The spectral types of the ionizing stars proposed by Wilcots (\cite{wil}), as well as their labels, are also listed in the table. Note that the present observations show the exciting star of N\,88D to be double (\#71 \& \#72) and the given type corresponds therefore to both of them. Table\,\ref{phot} is available in electronic form at the Centre de Donn\'ees astronomiques de Strasbourg (via anonymous ftp to cdsarc.u-strasbg.fr or via http://cdsweb.u-strasbg.fr/Abstract.html). While the exciting stars of the fainter H\,{\sc ii}\, regions are easily identified on the true-color image, a remarkable point is the absence of prominent stars towards the main component A. Nevertheless, we detect two faint stars embedded in the core of N\,88A east of the absorption lane (Fig.\,\ref{core}). These are stars \#1 and \#2 with $y=18.2$ and 18.3 mag and colors $b-y=+0.9$ and +0.6 mag respectively. It should however be underlined that these magnitudes are very uncertain since the stars lie in a very bright area where nebular subtraction is not straightforward. A third fainter star ($y$\,$\sim$\,20 mag), not visible in Fig.\,\ref{core}, is marginally detected just to the east of the bright H$\alpha$\, core of N\,88A. Its position suggests it as being a good candidate for an exciting star. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{ms8496f5.eps}} \caption{The two stars lying towards the very bright inner part of N\,88A (west of the absorption lane) are visible in this Str\"omgren $y$ image (separated by 0$''$\hspace*{-.1cm}.4, or $\sim$\,0.1 pc). The dark spot south of the stars is the brightest core of N\,88A. Field $\sim$\,4$''$\hspace*{-.1cm}\,$\times$\,2$''$\hspace*{-.1cm}\, ($\sim$\,1.2\,$\times$\,0.6 pc$^{2}$). Same orientation as all other images.} \label{core} \end{figure} Star \#55, situated towards the center of component B, has $y=16.57$ mag and is one of the brightest in the field. It has a highly elongated PSF profile which is due to multiplicity (at least three components are resolved). This may be the star ``$s1$'' detected by Wilcots (\cite{wil}) relatively close to the brightest part of N\,88. Its spectrum shows strong He\,{\sc i}\,$\lambda$\,4471\,\AA\, and He\,{\sc ii}\,$\lambda$\,4686, but weak He\,{\sc ii}\,$\lambda$\,4541\,\AA\, indicative of an O9~V star. If this spectrum belongs actually to \#55, it should correspond to the brightest component of this system. The color-magnitude diagram for the brightest stars of the sample (Fig.\,\ref{cm}) shows a main sequence with the bulk of the stars centered on Str\"omgren colors $b-y=-0.10$ and $v-b=-0.20$ mag, typical of massive OB stars (Relyea \& Kurucz \cite{rk}). These colors are equivalent to a Johnson $B-V$\,=\,--0.30 (Turner \cite{tur}), which indicates a negligible reddening (Conti et al. \cite{conti}). This result is due to the fact that the main sequence is overwhelmingly dominated by stars lying outside the N\,88 complex and means that the areas situated north-east, east, south, and south-west of N\,88 are not affected by dust. However, taking a sub-sample made up of all the exciting stars of the N\,88 H\,{\sc ii}\, regions (excluding stars \#1 \& \#2), we find the Str\"omgren colors $b-y=0.02$ and $v-b=-0.17$ which indicate $B-V$\,=\,--0.21 corresponding to a visual extinction of $A_{V}$\,=\,0.3 mag. This clearly confirms that the N\,88 complex is the most reddened part of this region of the SMC. Of particular interest are stars \#60 and \#61 situated immediately north-west of N\,88A (Fig.\,\ref{rgb}b). Assuming that these two stars are of O type, their colors suggest an extinction of $A_{V}$\,$>$\,1 mag. This result has implications for the location of the molecular cloud (Sect. 4.4). \begin{figure*} \caption{A finding chart indicating the brightest stars detected towards N\,88, based on the WFPC2 He\,{\sc ii}\, image. The field of view is the same as in Fig.\,\ref{rgb}a. North is up and east is left. The photometry of those stars is presented in Table\,\ref{phot}.} \label{numbers} \end{figure*} The brightest stars of the sample are \#39, \#12, \#19, and \#6. The first two are blue stars and the latter ones are red. The red population showing up in the color-magnitude diagram represents a collection of evolved field stars as well as young massive ones contaminated by nebulosity/dust. For instance, it is noteworthy that the very red stars \#76, and \#74 are not associated with a nebulosity, and this suggests that they should be evolved field stars. In the particular case of stars \#1 and \#2 lying inside N\,88A, in spite of their red colors, they may be young blue stars suffering from heavy extinction. Assuming that star \#1 has an O9\,V spectrum with $M_{V}$\,=\,--4.4 mag (Vacca et al. \cite{vacca}), and considering that the distance modulus of SMC is 19 mag, then an extinction of $A_{V}$\,=\,3.4 mag is necessary to make it appear as faint as $y$\,$\sim$\,18 mag. Similarly, in order for a star of spectral type O6\,V (Sect. 3.3) to have an $y$\,$\sim$\,20 mag, we need an $A_{V}$\,$\sim$\,6 mag. Thus, the main exciting star(s) of N\,88A should remain hidden in the optical due to the presence of dust by an extinction of at least 6 mag. \begin{figure*} \rotatebox{-90}{\resizebox{10.cm}{!}{\includegraphics{ms8496f7.eps}}} \caption{Color-magnitude diagram, $y$ versus $b$\,--\,$y$, of the brightest stars revealed towards the SMC compact H\,{\sc ii}\, region N\,88. Numbers refer to Fig.\,\ref{numbers} and Table\,\ref{phot}. } \label{cm} \end{figure*} \section{Discussion and concluding remarks} \subsection{Comparison with N\,81} The most striking feature of N\,88A is its lack of prominent stars, even at the WFPC2 resolution. This indicates a young age and is supported by other observational findings about N\,88A: its compactness, its high density, and its exceptionally high extinction. These facts considered together suggest that N\,88A is just hatching from its natal molecular cloud. Stars \#1 and \#2 are probably among the exciting sources of N\,88A. Other exciting stars may still be embedded in the densest part of the nebula, such as he bright spot highlighted in Fig.\,\ref{rgb}c and Fig.\,\ref{core}, and remain invisible due to the high dust content. Compared with N\,81, N\,88A is probably younger since N\,81 is more extended, less dense, and exhibits several of its exciting stars (Paper I). Although N\,81 and N\,88A are both very young, the present observations underline their difference. Apart from the exciting stars aspect, N\,88A is surrounded by several diffuse H\,{\sc ii}\, regions. In contrast, N\,81 is an isolated object. These facts point to the diversity of star forming regions belonging to the same chemical environment and also to the necessity of observing each case in detail. On the other hand, the whole N\,88 region is very reminiscent of the LMC N\,59 region studied by Armand et al. (\cite{arm}). N\,88 and N\,59 contain several individual H\,{\sc ii}\, regions, in various evolutionary states. They range from compact, bright and young components with a lot of dust hiding the exciting stars (N\,88A and N\,59A) to diffuse, spherical and evolved regions (N\,88E-F-G-H and N\,59C), and also to shell components (N\,88B and N\,59B which contains a supernova remnant). Similarly, both regions display a filamentary structure which results from the interaction of the strong stellar winds emitted by the massive stars with the surrounding medium, as well as small scale brightness variations pointing to a very inhomogeneous distribution of matter or dust inside or around these objects. \subsection{Associated neutral material} CO emission from the molecular cloud associated with N\,88A was observed by Israel et al. (\cite{isco}) using the ESO-SEST 15\,m telescope. They detected the $^{12}$CO(1--0) emission with a brightness temperature of 750 mK, a width of 2.5 km\,s$^{-1}$\, and a radial velocity of V$_{LSR}$\,=\,147.8 km\,s$^{-1}$. The molecular cloud is much brighter than the one associated with N\,81 (Paper I) and ranks among the few sources in the SMC detected in $^{13}$CO(1--0) (Israel et al. \cite{isco}). Rubio et al. (\cite{rub}) mapped the molecular cloud in $^{12}$CO(1--0) and $^{12}$CO(2--1) transitions using the SEST telescope with respective spatial resolutions of 43$''$\hspace*{-.1cm}\, ($\sim$\,13 pc) and 22$''$\hspace*{-.1cm}\, ($\sim$\,7 pc). It turns out that the cloud is in fact relatively small, $\sim$\,1$'$\hspace*{-.1cm}\, (18 pc) in size in the east-west direction and slightly smaller in north-south. More recently, Rubio et al. (private communication) have detected molecular transitions $^{12}$CO(3--2), CS(2--1), CS(3--2), HCO$^{+}$(1--0) which probably originate from the hot and dense core of the cloud. Molecular hydrogen emission also has been detected towards N\,88A (Israel \& Koornneef \cite{ik88}). The SMC is known to have an overall complex structure with several overlapping neutral hydrogen layers (McGee \& Newton \cite{McGee}). We used the recent observations by Stanimirovic et al. (\cite{stan}), with a resolution of 98$''$\hspace*{-.1cm}\, (30\,pc), to examine the H\,{\sc i}\, emission towards N\,88. The H\,{\sc i}\, spectrum profile has two main emission peaks at $\sim$\,150 and $\sim$\,175 km\,s$^{-1}$. The column density corresponding to their sum is 3.12\,$\times$\,10$^{21}$ atoms\, cm$^{-2}$, slightly smaller than that corresponding to N\,81 (Paper I). It seems that the molecular cloud is correlated with the smaller velocity H\,{\sc i}\, component. \subsection{Extinction} N\,88 was detected as a very bright IRAS source (Schwering \& Israel \cite{schwering}). The fact that near infrared photometry of N\,88A, at $J$, $H$, $K$, and $L'$ bands obtained using a 10$''$\hspace*{-.1cm}\, aperture (Israel \& Koornneef \cite{ik91}), is consistent with the IRAS spectrum (12, 25, 60, and 100\,$\mu$m) suggested that the IR emission arises mostly from the compact object in the aperture. Moreover, these authors found a quite red $K$\,--\,$L'$ color of more than 2 mag indicating the presence of hot dust. Our {\it HST} images for the first time show the heavy concentration of absorbing dust towards the inner parts of the H\,{\sc ii}\, region. More strikingly, the extinction rises to as high as $A_{V}$\,$\sim$\,3.5 mag in a narrow band towards the bright core of the nebula. This high absorption is quite unexpected for a metal-poor galaxy like the SMC. In fact N\,88A holds the record of extinction among ionized nebulosities in the SMC. The correlation between the zones of the high excitation and high extinction is an argument in favor of the physical association of the dust with hot gas. It is important to know the properties of this dust. Roche et al. (\cite{roche}) studied 8--13\,$\mu$m spectra of N\,88A and found a featureless continuum without any evidence of dust signatures attributed to silicate grains. This led them to the conclusion that the dust is likely composed of carbon grains. Further progress in this area requires appropriate IR observations using the highest spatial resolutions. \subsection{Star formation} The N\,88 nebular complex results from a small starburst which occurred recently in the Wing of the SMC. While the main stars creating N\,88A are not visible, the other members of the starburst show up in the {\it HST} images (Fig.\,\ref{rgb}). The stars exciting the diffuse H\,{\sc ii}\, regions (C to H) were formed in the outer, less dense parts of the molecular cloud, whereas the compact, very dusty N\,88A is associated with the core of this small molecular cloud (Sect. 4.2). The cloud must be to the north-east of N\,88A, as indicated by the ionization front detected in that direction (Sect. 3.1) and also by the fact that stars \#60 and \#61 situated north of the front are heavily affected by extinction (Sect. 3.4). This is further supported by ground-based higher exposure images showing a large front north-west of N\,88A beyond which no stars are visible (Fig.\,\ref{efosc}; also Testor \& Pakull \cite{tes}). The case of component B is interesting. Although it is, like N\,88A, apparently related to the core of the molecular cloud, it seems more evolved. In fact N\,88B has a significantly lower density and less dust, and reveals its exciting star (\#55). It can be considered that N\,88A has resulted from sequential star formation, that is the collapse of the shock/ionization front layer created by stars \#55. If so, we are dealing with two successive generations of stars formed in the core of the molecular cloud. The stars situated towards N\,88 are also known as HW\,81 following Hodge \& Wright (\cite{hw}) who surveyed the SMC in search of OB associations. The present observations reveal the fainter members of this association. The {\it HST} images also resolve another association in the direction of N\,88. Lying $\sim$\,50$''$\hspace*{-.1cm}\, (15 pc) south-east of N\,88A, at the lower-left corner of Fig.\,\ref{rgb}a, HW\,82 (Hodge \& Wright \cite{hw}) is composed of a dozen stars several of which are tightly packed multiple ones. HW\,82 is not associated with ionized gas, and a relatively large number of its stars are red. At present we do not know whether the red and blue stars are co-spatial members of the same cluster. Nevertheless, the facts that the ionized gas is already dispersed from there and that no significant amount of dust is detected (Sect. 3.4) suggest that HW\,82 represents an older burst of star formation in the Wing. This is confirmed by the larger field of EFOSC H$\alpha$\, image (Fig.\,\ref{efosc}) which shows no H\,{\sc ii}\, regions south of the {\it HST} field of view. Star formation must have therefore proceeded from south to north and N\,88 is the most recent site of star formation in this part of the SMC. A noteworthy aspect of the stellar population towards N\,88 is the presence of several tight clusters or multiple systems uncovered by the present observations. For example, the exciting star of N\,88B (\#55) is a multiple system of at least three components. There are also at least two stars hidden inside N\,88A, while both N\,88C and D are excited by two blue stars of comparable brightness. Note also the tight cluster in HW\,82 composed of stars \#9, \#10, \#11, \#12 , and \#13. These cases present new pieces of evidence in support of collective formation of massive stars in the SMC (see Paper I for a brief discussion). An intriguing, though unanswered question, is related to the origin of the large-scale filamentary veil visible in the EFOSC image. Our true-color image shows that filaments originating from north-east of N\,88 run towards the anonymous blue cluster in the south (stars \#16, \#17, \#21, \#22, \#23, and \#24). However, the veil significantly brightens south of that cluster and bends to the south-east. In consequence, the association of the veil with the N\,88 region is not established. It is possible that this filamentary structure is linked to the neighboring huge bubble nebula DEM\,167 (Davies et al. \cite{dav}). \begin{acknowledgements} We are grateful to an anonymous referee for his careful reading of the manuscript and comments that contributed to substantially improve the paper. VC would like to acknowledge the financial support from a Marie Curie fellowship (TMR grant ERBFMBICT960967). \end{acknowledgements}
\section{Introduction} During the last few years, Single Molecule Biophysics has become a very active field of research. Among the new explored areas, one finds the study of elastic properties of biopolymers, in particular the DNA molecule, under physical conditions close to that encountered in living organisms \cite{smith,perk,strick}. The theoretical work reported in the present paper was motivated by the results of micromanipulation experiments being performed at Ecole Normale Sup\'erieure \cite{strick} with an isolated DNA molecule, immersed in a solution of given salinity at room temperature. One extremity of the molecule is biochemically bound at multiple sites to a treated glass cover slip. The other end is similarly attached to a paramagnetic bead with a radius of a few microns which is performing a Brownian motion in the solution. An appropriate magnetic device is able to pull and rotate at the same time the magnetic bead. Because of the multiple attachments of the molecule extremities, the rotation of the bead results in a supercoiling constraint for the DNA isolated molecule. When no rotation is applied on the bead, the force versus extension curves are very well described \cite{marsig} by a simple elastic model, the so-called Worm Like Chain (WLC) model which describes a phantom chain using a single elastic constant, the bending rigidity \cite{fixman}. This model reproduces the data within experimental accuracy, in a very wide range of pulling force from $F \simeq 0.01\, pN$ to $F\simeq 10\, pN$. One has to go up to $70\, pN$ to leave the elastic regime and see a sharp increase of the molecule length which is associated with a transition to a new molecular phase, the "S-DNA" \cite{sDNA}. In contrast, in stretching experiments performed with supercoiled DNA molecules \cite{strick,strick2}, two molecular structural transitions appear within a more restricted domain of forces. An instructive way to display the supercoiled DNA stretching data is to plot the molecule extension versus the degree of supercoiling $\sigma$ (defined as the ratio of the number of bead turns to the number of double helix turns in the relaxed molecule), at various fixed values of the stretching force. Three such plots are shown on Figure \ref{fig1} for typical force values. Working with molecules having a moderate degree of supercoiling, $\sigma \le 0.1$, two changes of regime occur in the force range between $F=0.01\, pN $ and $ F=5.0\, pN $. For forces below $ 0.44\, pN $ the extension versus supercoiling curves are symmetric under the exchange $ \sigma \rightarrow - \sigma $ and have been called `hat' curves for obvious reasons. This is the domain of entropic elasticity which will be our main topic in the present paper. When $ F \ge 0.5\, pN $ a plateau is developed in the underwinding region $ \sigma <0 $. The mechanical underwinding energy is transformed into chemical energy which is used to open up the hydrogen bonds and denaturation bubbles appear along the DNA chain. Above $4 \,pN$, a plateau appears also in the overwinding region $ \sigma >0 $. This has been interpreted as a transition to a new structure of DNA called the " DNA-P"\cite{pDNA}. The double helix structure is clearly responsible for the ability of DNA to convert mechanical torsional energy into chemical energy, with a net preference for underwinding energy. Our work here will be restricted to the entropic elastic regime, and its aim is primarily to compute the extension versus supercoiling hat curves. We have used the simplest extension of the WLC model, which is still a phantom chain, with one single new elastic constant, the twist rigidity. We call the corresponding model a Rod Like Chain (RLC). We will show that the RLC model, described in terms of local Euler angle variables, is singular in the limit of a purely continuous chain. Its physical realization requires the introduction of a short length scale cutoff (which will turn out to be of the order of magnitude of the double helix pitch). Then the RLC is able to reproduce all experimental data in the restricted domain of its validity which we have discussed above. The good fit to the experimental data allows to deduce the twist rigidity to bending rigidity ratio, which was rather poorly known so far (with uncertainties up to a factor or 2). A short account of some parts of this work has been discussed in a previous letter \cite{BM1} (where the RLC model appeared under a slightly different name: the WLRC model). We shall give here a more comprehensive - and hopefully comprehensible- description of our model. Our principal aim is to show how the RLC model, once it has been properly formulated, can be solved exactly: by exact, we mean that the predictions of the model can be computed by a mixture of analytic and numerical methods to any desired precision. The first conclusive work able to fit the experimental data from a rod like chain with two elastic constants is that of Marko and Vologodskii \cite{marvol}, who used a discretized model and performed some Monte Carlo simulations. On the analytical side, Fain et al. \cite{fain} wrote the elastic energy in terms of a local writhe formula, similarly to the one which we have used, but they used it only in the zero temperature limit. In an independent work which appeared just after ours, Moroz and Nelson \cite{mornel1} used the same form of energy as we did, but with a different approach for handling the singularity of the model in the continuum limit. We shall comment extensively on the differences of these approaches, as well as on the obtained results. Developing a theory beyond the elastic regime is a much more difficult task, since such a theory should involve both the self repulsion of the chain (necessary to prevent a collapse of plectonemes), and the possibility of denaturation. Some first attempts in these directions can be found in \cite{liverpool,zhou}. Recently an explicit model, coupling the hydrogen-bond opening with the untwisting of the double helix, has been proposed \cite{cocco}. It leads to a unified description of thermal and untwisting DNA denaturation in the high stretching force limit. We now indicate the organization of the paper, with some emphasis upon the "RLC model crisis" and its solution. In sect. \ref{el_en_RLC} we introduce the RLC model and discuss its range of validity. We note first that the elastic energy used in the present paper is invariant under rotation about the molecular axis, in apparent contradiction with the double helix structure. We prove that added cylindrical asymmetric terms in the elastic energy are washed out by averaging upon the empirical length resolution $ \Delta l $, about three times the double helix pitch $p$. We also discuss the relationship to the Quantum Mechanics problem of a symmetric top, which appears naturally when one considers the configurations of the DNA chain as world lines in a quantum mechanical problem. We stress that, despite a formal analogy, there is an important difference in the formulation of the two problems. The RLC analog of the angular momentum is not quantized since, in contrast to the symmetric top, the physical states of an elastic rod are not invariant under rotations of $ 2 \pi$ about its axis. This is the origin of the "RLC crisis". In Sect. \ref{part_funct_DNA} we incorporate in the RLC partition function the supercoiling constraint coming from the rotation of the magnetic bead. First written as a boundary condition upon Euler angles used to specify the DNA configuration, the supercoiling constraint is transformed, assuming some regularity conditions on the Euler angles, into an equality between the bead rotation angle and the sum of the twist and writhe variables of the open molecular chain. These variables fluctuate independently in absence of supercoiling so that the constrained RLC partition function is given as a convolution product of the twist and writhe partition functions. The writhe partition function Fourier transform is in correspondence with the Quantum Mechanics problem of a charged particle moving in the field of a magnetic monopole with an unquantized charge. The spontaneous fluctuations of the writhe provide a spectacular signature of the RLC model pathology: the second moment of the fluctuation is predicted to be infinite in the continuum limit. This divergence comes from a singularity appearing in the RLC potential when the molecular axis is antiparallel to the stretching force. Sect. \ref{discrete_theory} explains how the angular cutoff needed in order to transform the RLC model into a sensible one can be generated by the discretization of the chain. The introduction of a length cutoff $b$, of the order $ \Delta l$, appears rather natural since an average upon $ \Delta l$ has been invoked to justify our choice of the RLC elastic energy. We proceed next with the development of the two main tools used to get an exact solution of the discretized RLC model: a direct transfer matrix approach and a Quantum Mechanics approach involving a regularized RLC Hamiltonian with a potential derived from the discretized model and now free from any singularity. In sect. \ref{quant_ana} we use the two above methods in order to compute the elongation versus supercoiling characteristics of a long DNA chain. The results are compared with experimental data in sect. \ref{ana_exp} where the small distance cutoff $ b $ and the twist elastic constant are determined. Sect. \ref{Monte_Carlo} presents a complementary approach, that of Monte Carlo simulations which provides the only method so far to introduce the constraints of self-avoidance and knots elimination. After a presentation of the results of Marko and Vologodskii \cite{marvol}, we describe our less ambitious simulations which were used to validate our analytic method. In sect. \ref{tw_wr_pl} we use the RLC model to analyse, at a force taken on the high side of our explored area, the variation of the torque and the writhe to twist ratio versus supercoiling. The two curves exhibit a rather sharp change of regime near one given value of the supercoiling angle: the torque, after a nearly linear increase, becomes almost supercoiling independent while the writhe to twist ratio, initialy confined to the $ 20 \% $ level, develops a fast linear increase. We give arguments suggesting that this quasi-transition is associated with the creation of plectoneme-like configurations, able to absorb supercoiling at constant torque. Finally the sect. \ref{fluctuations} studies, within the RLC model, the extension fluctuations of an isolated DNA molecule, subject both to stretching and supercoiling constraints. The predictions are made for two thermodynamic ensembles: one at fixed torque, the other at fixed supercoiling angle. Although the two ensembles lead, as they should, to identical results as far as average values are concerned, they yield widely different predictions for the extension fluctuations above a certain supercoiling threshold (close to the one appearing in the previous section): the fixed torque ensemble becomes much more noisy. We give a qualitative explanation for this peculiar behaviour and compare our predictions to some preliminary experimental data at fixed supercoiling. In sect. \ref{conclusion} we give a summary of the work and some perspectives on its future extensions. \section{Elastic Energy of The Rod-Like Chain} \label{el_en_RLC} A given configuration of the rod-like chain (RLC) is specified, in the continuous limit, by the local orthonormal trihedron $\{ {\hat{{\bf{e}}}}_{i}(s)\}= \{ \hat{{\bf{u}}}(s),\hat{{\bf{n}}}(s),\hat{{\bf{t}}}(s) \}$ where $s$ is the arc length along the molecule, $ \hat{{\bf{t}}} $ is the unit vector tangent to the chain, $\hat{{\bf{u}}}(s) $ is along the basis line and $\hat{{\bf{n}}}(s)=\hat{{\bf{t}}}(s)\wedge\hat{{\bf{u}}}(s)$. The evolution of the trihedron $\{ {\hat{{\bf{e}}}}_{i}(s)\}$ along the chain is obtained by applying a rotation $ {\cal R }(s) $ to a reference trihedron $\{ {\hat{{\bf{e}}}}_{i}^0(s)\}$ attached to a rectilinear relaxed molecule: ${\hat{{\bf{e}}}}_{i}(s) ={\cal R }(s)\cdot {\hat{{\bf{e}}}}_{i}^0(s)$. The rotation $ {\cal R }(s) $ is parameterized by the usual three Euler angles $\theta(s)$, $\phi(s)$ and $\psi(s)$. The reference trihedron is such that $\theta(s)=0$, $\phi(s)+\psi(s)=\omega_{0} s$, where $ \omega_{0}$ is the rotation per unit length of the base axis in the relaxed rectilinear DNA molecule. With the above definition, the set of $s$ dependent Euler angles $ \theta(s), \phi(s),\psi(s) $ describes the general deformations of the DNA molecule with respect to the relaxed rectilinear configuration. It is now convenient to introduce the angular velocity vector ${\bf{\Omega}}(s)$ associated with the rotation $ {\cal R }(s)$. The evolution of the trihedron $\{ {\hat{{\bf{e}}}}_{i}(s)\}$ along the molecular chain is given in term of ${\bf{\Omega}}(s)$ by the set of equations: \be {d\,\over ds }\, {\hat{{\bf{e}}}}_{i}(s) =\(( {\bf{\Omega}}(s) + \omega_{0} {\hat{{\bf{t}}}}(s) \))\wedge {\hat{{\bf{e}}}}_{i}(s) \ee We shall use in the following the components of ${\bf{\Omega}}(s) $ along the trihedron $\{ {\hat{{\bf{e}}}}_{i}(s) \}$: $ {\Omega}_{i}= {\bf{\Omega}}(s) \cdot {\hat{{\bf{e}}}}_{i}(s) $ with $ {\hat{{\bf{e}}}}_{1}=\hat{{\bf{u}}},\; {\hat{{\bf{e}}}}_{2}=\hat{{\bf{n}}} $ and $ {\hat{{\bf{e}}}}_{3}=\hat{{\bf{t}}}$. ( The $ {\Omega}_{i} $ are computed in terms of the Euler angles and their $ s $ derivatives in the Appendix A). The stretched RLC energy $E_{RLC}$ is written as a line integral along the chain of length $ L$, involving a sum of three contributions: \be E_{RLC}= k_{B} T\,\int_0^{L} \, ds \left( {\cal E}_{bend}(s)+{\cal E}_{twist}(s) +{\cal E}_{stretch}(s) \right) \label{EL0} \ee The above bend, twist and stretch linear energy densities are given by: \bea {\cal E}_{bend} &=& {A \over 2}\,\left( {\Omega}_{1}^2+{\Omega}_{2}^2 \right)= {A \over 2}\, \vert {d \hat{{\bf{t}}}(s) \over ds} \vert ^2 = {A \over 2} \left( {\dot{\phi}}^2 \,{\sin^2\theta}+{ \dot{\theta}}^2 \right) \nn \\ {\cal E}_{twist} &=& {C \over2}\,{\Omega}_{3}^2= { C \over 2} \,( \dot{\psi} +\dot{\phi} \, \cos\theta )^2 \nn \\ {\cal E}_{stretch} &=& - \hat{{\bf{t}}}(s)\cdot{\bf{F}} /( k_{B} T)= - {F \cos \theta \over k_{B} T} \label{EL} \eea where the dot stands for the $s$ derivative. $ {\cal E}_{bend} $ is proportional to the inverse of the square of the curvature radius and represents the resistance against bending around an axis perpendicular to the chain axis. The coefficient $ A $ is the persistence length; its typical value is $ 50 nm$. ${\cal E}_{twist}$ is proportional to the square of the projection of the angular velocity along the molecular axis and gives the energy associated with a twist constraint. The twist rigidity $C$ is known to be in the range 50 to 100 nm; its determination from the experimental measurement will be the topic of section \ref{ana_exp}. ${\cal E}_{strech}$ is the potential energy associated with the uniform stretching force $ {\bf{F}}=F.\hat{{\bf{z}}} $ which is applied at the free end of the molecule; it is written as a line integral involving $ \hat{{\bf{t}}}(s).\hat{{\bf{z}}} $. The ground state configurations of the rod like chain have been studied in \cite{fain}. These may apply to the properties of small plasmids, but in the regime which we are interested in, the thermal fluctuation effects are crucial. If no twist dependent properties are imposed or measured, one can factor out the $\psi$ integral in the partition function, or equivalently use the $ C=0$ theory, in which case one recovers the elastic energy of the Worm Like Chain model : \be {E_{WLC} \over k_{B} T}= \int_0^{L} \, ds \(( {A \over 2}\, \vert {d \hat{{\bf{t}}}(s) \over ds} \vert ^2 -\cos\theta(s) F /( k_{B} T) \)) \label{EWLC} \ee It is now appropriate to raise two questions about the above formulas. \begin{center} \subsection{ Is a cylindrical symmetric elastic rigidity tensor adequate for the DNA chain? } \end{center} The theory which we are using here is an elastic theory which cannot be valid down to atomic scales. In particular it does not take into account the microscopic charges on the DNA and their Coulomb interactions in the solvent, but it just describes the net effect by a set of elastic constants. However, even at the level of an elastic description, one may wonder whether our choice of elastic tensor, which ignores the helical structure of the DNA, is valid. Indeed the elastic energy given by equations (\ref{EL}) involves a rigidity tensor with a cylindrical symmetry around the molecular axis $ \hat{{\bf{t}}}(s) $. We would like to argue that such a description is a reasonable approximation if one deals with experimental data obtained with a finite length resolution $ \Delta \, l \simeq 10 nm$, as is the case for the experiments we shall analyse. The pitch $ p$ of the double helix contains approximately 10 basis which corresponds to a length of $ 3.4 nm$. A simple way to break the cylindrical symmetry is to introduce in the elastic energy linear density a term proportional to $ \Delta \Omega(s)= {\Omega}_{1}^2-{\Omega}_{2}^2 $. As it is shown in Appendix A, $ \Delta\Omega(s) $ can be written in terms of the Euler angles in the following way: $$ \Delta\Omega(s) = {\Omega_{\perp}}^2 \,\cos 2( \zeta(s) +\psi(s)+ \frac{2 \pi s}{p}) $$ with $ {\Omega_{\perp}}^2 ={\Omega}_{1}^2+{\Omega}_{2}^2 $ and $\zeta(s) =\arctan\((\dot{\theta}/(\sin\theta\dot{\phi} )\))$; One must note that $ \Delta\Omega(s) $ oscillates with a wavelength which is half the pitch $p$ of the double helix. An average involving a resolution length, about 6 times the oscillation wavelength, is expected to lead to a strong suppression of $ \Delta \Omega(s) $. The average $ \overline{ \Delta\Omega(s) }= \int ds_1 P(s_1 -s)\Delta\Omega(s_1)$, associated with the resolution function $ P(s)={1\over \sqrt{2 \pi } \Delta\,l}\exp( -{1\over2} s^2 /{\Delta\,l}^2) $, is computed explicitly in Appendix A in terms of simple Gauss integrals, assuming that the phase $\zeta(s) +\psi(s)$ and the amplitude ${\Omega_{\perp}}^2$ vary linearly within the interval $ ( s-\Delta\,l, s+\Delta\,l)$. Ignoring first the $\Omega_{\perp} $ variation and assuming that the supercoiling angle per unit length is $ \ll \omega_0$, we arrive in Appendix A to the result: $$ \overline{ \Delta\Omega(s) }\approx \Delta\Omega(s) \exp\(( -\frac{1}{2} (\frac{4\pi \,\Delta\,l}{p})^2 \)) $$ Taking $ p=3.4 nm $, $\Delta\,l=10 nm $ one gets: $\frac{4\pi \,\Delta\,l}{p} \approx 37$; it is clear the above expression of $\overline{ \Delta\Omega(s) }/\Delta\Omega(s)$ is zero for all practical purposes. The extra term $ \propto { \partial \, {\Omega_{\perp}}^2 \over \partial \,s}$ is also found to be exceedingly small. The same suppression factor holds for $ \overline{ {\Omega}_{1}\,{\Omega}_{2}} $ while in the case of $\overline{ {\Omega}_{1}\,{\Omega}_{3} } $ the argument of the exponential is divided by 4 but with the value quoted above for $\Delta\,l$ the suppression effect is still very important. Therefore all the cylindrical asymmetric terms in the elastic rigidity tensor are washed out by the empirical finite length resolution averaging. In the same spirit there are various effects which are not taken into account by our description. An obvious one is the heterogeneity of the sequence: a more microscopic description should include some fluctuations of the rigidities along the chain, depending on the sequence of bases. However one expects that such fluctuations will be averaged out on some long length scales, such as those involved in the experimental situation under study. This has been confirmed by some more detailed study involving some simple model of disordered sequences \cite{BDM}. One should also notice that we have not allowed the total length of the chain to vary. It is possible to add some elastic energy to bond stretching as well as some stretch-twist coupling in order to try to describe the behaviour at rather large forces and supercoiling \cite{marko,KLNO,mornel1}. But as we will see such terms are irrelevant for the elastic regime which we study here. \subsection{ What is the precise connection of the elastic rod thermal fluctuations problem with the quantum theory of a symmetric top ?} A careful inspection of the formulas (\ref{EL0}) and (\ref{EL}) suggests a close analogy of the elastic linear energy density with the classical lagrangian of the symmetric top with A and C being, up to a constant factor, the moments of inertia. For given values of the Euler angles $ \theta(s), \phi(s),\psi(s) $ at the two ends $s_0=0$ and $ s_1=L$ of the chain, the partition function of the RLC model is given by the following path integral: \begin{equation} Z({\theta}_1,{\phi}_1,{\psi}_1,s_1\vert{\theta}_0,{\phi}_0,{\psi}_0,s_0 )= \int {\cal D} \(( \theta, \phi,\psi \)) \exp\((-{E_{RLC}\over k_B T }\)) \label{Zpath} \end{equation} where $ {\cal D} \(( \theta, \phi,\psi \)) $ stands for the integration functional measure for the set of paths joining two points of the Euler angles space with $ 0 \le \theta \le \pi $ but no constraints imposed upon $ \phi $ and $ \psi $ . Let us perform in the elastic energy integral over $s$ ( Eq \ref{EL0} ) an analytic continuation towards the imaginary $s$ axis. We shall use for convenience a system of units where $ \hbar= c =k_BT=1 $ and write $ \Im{s}=t $. The elastic energy $ E_{RLC} $ is transformed into $-i \int_{t_0}^{t_1} dt {\cal L} (t)$, where the Lagrangian is that of a symmetric top in a static electric field $f$: $ {\cal L} (t)= {1\over2}\sum_1^3 {C}_i {\Omega}_{i}^2 +f \cos \theta(t) $, with inertia moments $ {C}_1={C}_2= A $, $ C_3=C$ and $ f= F/(k_BT) $. The analytically continued partition function of the RLC model is then identified with the Feynmann amplitude: \bea \langle {\theta}_1,{\phi}_1,{\psi}_1,t_1\vert{\theta}_0,{\phi}_0,{\psi}_0 ,t_0 \rangle &= & \int {\cal D}\(( \theta,\phi,\psi \)) \exp\((i \int_{t_0}^{t_1} dt {\cal L} (t) \)) \nonumber \\ &=& \langle {\theta}_1,{\phi}_1,{\psi}_1\vert\exp \(( -i(t_1-t_0){\hat{\cal H}}_{top} \)) \vert{\theta}_0,{\phi}_0,{\psi}_0 \rangle \nonumber \eea where $ {\hat{\cal H} }_{top} $ is the symmetric top Hamiltonian written as a second order differential operator acting upon Euler angles wave functions. Returning to real value of s, ones gets the RLC model partition function as a Quantum Mechanics matrix element, with the substitution $i(t_1-t_0) \to s_1-s_0$. It is now convenient to introduce the complete set of energy eigenstates of ${\hat{\cal H}}_{top}$: $ {\hat{\cal H}}_{top} \mid n \rangle =E_{n} \mid n \rangle $. One gets then: \bea Z({\theta}_1,{\phi}_1,{\psi}_1,s_1\vert{\theta}_0,{\phi}_0,{\psi}_0,s_0 ) & = & \sum_n \,\exp- L\,E_{n} \nonumber \\ & & {\Psi}_n({\theta}_1,{\phi}_1,{\psi}_1) {\Psi}_n^{\star}({\theta}_0,{\phi}_0,{\psi}_0) \label{ZQM} \eea where $ {\Psi}_{n}(\theta,\phi,\psi) =\langle \theta,\phi,\psi \vert n \rangle $ is the eigenfunction relative to the state $\vert n \rangle $. As we shall discuss later, the interest of the above formula lies in the fact that the above sum is dominated by the term of lowest energy. At this point one may get the impression that the solution of the RLC model is reduced to a relatively straightforward problem of Quantum Mechanics. Although this identification is valid in the case of the worm-like-chain model (WLC) which describes the entropic elasticity of a stretched DNA molecule with no twist constraints, it is not the case for the more general RLC model. The above analysis was somewhat formal in the sense that the Hamiltonian $ {\hat{\cal H}}_{top} $ was given as a differential operator and it is thus not completely defined, until the functional space on which it is acting is properly specified. Physical considerations will lead us to choose different functional spaces for the two problems in hands: { \it In the symmetric top problem the space is that of $2 \pi$ periodic functions of the Euler angles $ \phi$ and $ \psi $ , but in the study of RLC thermal fluctuations the space is that of general functions of these angles, without any constraint of periodicity.} As an illustration, let us discuss the simple case of a non-flexible rod with $E={C\over 2} k_{B} T\,\int_0^{L} \, ds \,{\dot{\psi}(s)}^2$. The associated quantum problem is the cylindrical rotator described by the Hamiltonian: $ {\hat{\cal H}}_{ rot}= -{1\over 2 C}\, \frac{ {\partial}^2 }{ {\partial\psi}^2 } $. The $ 2 \pi $ periodicity of the motion implies a discrete spectrum for the conjugate momentum $ p_{\psi}= -i \frac{ {\partial} }{ {\partial\psi} } $; the corresponding wave function is $\exp(i\,k \, \psi )$, where $k$ is an arbitrary integer. Imposing on the molecular chain the boundary conditions $ \psi(s_0=0)=0 \, , \psi(s_1=L)=\chi $, the partition function of the non-flexible rod is obtained by a straightforward application of formula (\ref{ZQM}): $$ Z(\chi,L)= \sum_k \exp\(( i\,k \, \chi -{ L\over 2 C}\, k^2 \)) $$ In the experiments to be discussed in the present paper $ L \gg C $ and, as a consequence, the sum over the integer $k$ is dominated by the terms $ k=0,\pm 1$, leading to the rather unphysical result: $ Z(\chi,L)=1 +2 \cos \chi \exp(- { L\over 2C}) $. On the contrary, in order to describe the twisting of an elastic rod. The spectrum of $ p_{\psi} $ has to be taken as continuous and the dicrete sum $ \sum_k ... $ must be replaced by the integral $\int_{-\infty}^{\infty} dk/( 2\pi)... $, which yields the correct result for the partition function: \be Z(\chi,L) \propto \exp-\(( { C \over 2\, L}\, {\chi}^2 \)) \label{ZT} \ee Two significant physical quantities can be easily obtained from the above partition function: First, the second moment $ < {\chi}^2 >$ relative to a situation where $\chi$ is no longer constrained but allowed to fluctuate freely. Keeping in mind that $ P( \chi) \propto Z(\chi,L) $, one gets : $< {\chi}^2 >= {L\over C}$. Second, the torque $ \Gamma $ given by the logarithmic derivative of $ Z(\chi,L) $ with respect to $ \chi $: $ \Gamma = k_B T { C \, \chi \over L}$. The above quantities look very reasonable from a physical point of view: $ < {\chi}^2 >$ scales linearly in $ L $, as expected from the fluctuations of a linear chain with a finite correlation length $C $ and the expression of $ \Gamma $ reproduces an elasticity textbook formula, if one remembers that $ k_B T C $ is the usual twist rigidity. The above considerations may look somewhat simple-minded. It turns out however that they have important physical consequences. In order to solve the RLC model, we shall have to deal with a quantum spherical top when its angular momentum along the top axis is not quantized. Such a problem is mathematically singular and as a consequence the continuous limit of the RLC model we have considered so far will not give an adequate description of supercoiled DNA. We shall have to introduce a discretized version of the RLC model, involving an elementary length scale $b$ about twice the double helix pitch $p$. This is consistent with the considerations of the previous subsection where the empirical length resolution $ \Delta\,l \approx 3 p $ was invoked in order to justify the cylindrical symmetry of the tensor of elastic rigidities. \begin{center} \section{ The Partition Function for Supercoiled DNA in the Rod-like chain Model } \label{part_funct_DNA} \end{center} \subsection{ Implementing the experimental supercoiling constraint on the partition function } The first step is to incorporate in the functional integral the supercoiling constraint which results from a rotation around the stretching force $ {\bf{F}} $ of a magnetic bead biochemically glued to the free end of the DNA chain. The rotation ${\cal R}$ in terms of the Euler angles is usually written as a product of three elementary rotations, a rotation $\theta $ about the $y$ axis sandwiched between two rotations $ \psi$ and $ \phi$ about the $z$ axis, performed in that order: ${\cal R}= R(\hat{{\bf{z}}},\phi)\,R(\hat{{\bf{y}}} ,\theta)\,R(\hat{{\bf{z}}},\psi)$, where $R(\hat{{\bf{m}}},\gamma)$ stands for a rotation $\gamma$ about the unit vector $\hat{{\bf{m}}} $. It is convenient, in the present context, to introduce a different form of ${\cal R}$\, involving a product of two rotations written in terms of a new set of variables, $\theta ,\psi $ and $ \chi=\phi+\psi $: $ {\cal R} =R(\hat{{\bf{z}}},\chi)\,R(\hat{{\bf{w}}}(\psi),\theta) $ where the unit vector $\hat{{\bf{w}}}(\psi)$ lies in the $x \; y $ plane and is given by: $ \hat{{\bf{w}}}(\psi)=R(\hat{{\bf{z}}},-\psi) \hat{{\bf{y}}} $. ( See Appendix A for a proof of the identity of the two above forms of ${\cal R}$). Let us consider, before any application of an external rotation, the DNA segment sticking out from the bead. Assume that it is short enough so that its direction is not affected by thermal fluctuations. Its orientation is specified by three Euler angles, $ { \theta}_{in},{\psi}_{in},{\chi}_{in} $. If the rotation of the magnetic bead by $ n $ turns is performed adiabatically, the final orientation of the molecular end trihedron $\{ {\hat{{\bf{e}}}}_{i}(L)\} $ is specified by the rotation: $$ {\cal R}(L)= R \(( \hat{{\bf{z}}}, 2\pi\,n +{\chi}_{in} \)) \, R \(( \hat{{\bf{w}}}({\psi}_{in}), {\theta}_{in}\)). $$ We can read off easily from the above formula: $ \chi(L)=\phi(L)+\psi(L)=2\pi\,n +{\chi}_{in} $. It is reasonable to assume that in the gluing process no large surpercoiling is involved so that the angles ${\chi}_{in}$ and $ {\psi}_{in} $ do not exceed $ 2\, \pi$. As we will see the relevant scaling variable for the partition function is: \be \eta= {\chi(L) A \over L} \ee In the experiments to be analyzed in the present paper, $ A/L \simeq {1\over 300} $ with $\eta$ of the order of unity, so that ${\chi}_{in}$ can be safely ignored in comparison to $ 2\pi\,n $. In the following we shall drop the $ L $ dependance in $ \chi(L) $ and write $ \chi= 2\pi\,n= \phi(L)+\psi(L) $. The above equation does not imply that $ \chi $ is a discrete variable: it is just for practical reasons that the experimentalists perform measurements with an integer number of turns. On the other hand, if we were dealing with a closed DNA chains instead of an open one, the supercoiling angle $ \chi(L) $, which is here an arbitrary real, would get replaced by $ 2\pi\ L_k $ where the integer $ L_k $ is the topological linking number. \centerline{\it 1. A local writhe formula } We make now an assumption which may look, at first sight, somewhat trivial but has, as we shall see, far reaching consequences: { \it The Euler angles $ \phi(s) $ and $ \psi(s) $ are regular enough functions of $s$ to allow the replacement of $\chi $ by a line integral involving the sum of Euler angles derivatives, taken along the trajectory $ \Gamma $ of the tip of the tangent vector $\hat{{\bf{t}}}(s)$.} \be \chi= 2\pi\,n= \phi(L)+\psi(L)=\int_0^L ds \ \left( \dot{\psi} +\dot{\phi}\right) \label{chif} \ee It is then convenient to introduce the total twist along the chain $ T_w $ which appears as a Gaussian variable in the partition function. \be T_w= \int_{0}^{L} ds= \int_{0}^{L} ds {\Omega}_{3} = \int_{0}^{L} ds \left( \dot{\psi} +\dot{\phi} \, \cos\theta\right) \label{tw} \ee where ${\Omega}_{3}(s) $ is the projection of the instantaneous rotation vector $ {\bf{\Omega}} $ onto the unit tangent vector $\hat{{\bf{t}}}(s)$, given in terms of Euler angles in Appendix A. We now {\it{define} } a `local writhe' contribution $ \chi_W $ by substracting the total twist $ T_w $ from the supercoiling angle $\chi $ \be \chi_W=\chi-T_w= \int_{0}^{L} ds \ \dot{\phi} (1-\cos\theta) \label{wr} \ee The above decomposition follows from our analysis of the empirical supercoiling angle and a mere line integral manipulation. This is reminiscent of the decomposition of the linking number into twist and writhe for closed chains \cite{geometry,boles}. Formula (\ref{wr}) has been first obtained by Fain { \it et al.} \cite{fain} by adapting to the case of open strings a formula first derived by Fuller \cite{fullerform} for closed strings. We have decided to include an explicit derivation here in order to be able to describe some subtleties in its use, which will require some regularization procedure. Since $\hat{{\bf{t}}}(s)$ is a unitary vector, the curve $ \Gamma $ lies upon the sphere $ S_2 $ and it is parametrized by the spherical coordinates $ ( \theta(s), \phi(s) ) $. This representation is well known to be singular at the two poles $ \theta=0,\theta=\pi $ ( the choice of the $ z $ axis is not arbritrary but dictated by the experimental conditions: it coincides with the rotation axis of the magnetic bead and gives also the stretching force direction). With the usual convention, $\theta$ is required to lie within the interval $ [0 ,\pi] $. This condition plays an essential role in the quantum-like treatment of the WLC and RLC continuous models, to be discussed later on. When the trajectory $ \Gamma $ passes through the poles the restriction imposed upon $ \theta $ implies that the function $ \phi(s) $ suffers form a discontinuity of $ \pm \pi $ and this invalidates our derivation of formula (\ref{wr}). To get around this difficulty, we have to pierce the sphere $ S_2 $ at the two poles. The two holes are defined by two horizontal circles with an arbitrary small but finite radius $ \epsilon $ with their centers lying on the $z$ axis. Note that $ S_2$ has now the topology of a cylinder so that the piercing of the sphere $ S_2 $ provides the necessary topological discrimination between the two distinct physical states $ ( \theta(s), \phi(s),\psi(s) ) $ and $( \theta(s), \phi(s)+2\pi n, \psi(s) ) $. A careful evaluation of formula (\ref{wr}), involving continuously deformed $ \Gamma $ trajectories drawn on the pierced sphere $ S_2 $ in order to avoid the poles, gives results in agreement with those given by the non local closed loop writhe formula \cite{geometry}, while a naive application of (\ref{wr}) would lead astray. \centerline{\it 2. Solenoid and plectoneme supercoil configurations} To illustrate this point, we shall evaluate $\chi_W $ for right handed solenoid and plectoneme supercoil configurations, having arbitrary orientations. Let us begin with the simple case where the solenoid superhelix is winding around the $ z $ axis. The tangent vector tip trajectory $ {\Gamma}_{sol}$ consists of $n$ turns along the horizontal circle cut upon the unit sphere $ S_2 $ by the horizontal plane $ z= \cos {\theta}_0 $ ( $ \tan {\theta}_0 = \frac{ 2 \pi R}{ P} $ where $ R $ and $P$ are respectively the radius and the pitch of the superhelix.) The writhe supercoiling angle $\chi_W ({\Gamma}_{sol} )$ is easily seen to be given by $ 2 \pi\, n\,( 1- \cos {\theta}_0 )$. The plectoneme supercoil configuration is composed of two interwound superhelices connected by a handle. The helices have opposite axis but the same helicity. The plectoneme $\hat{{\bf{t}}}$ trajectory $ {\Gamma}_{plec} $ is written as the union of three trajectories: $ {\Gamma}_{plec} = {\Gamma}_{sol}^1 \cup {\Gamma}_{han} \cup {\Gamma}_{sol}^{2}$. ${\Gamma}_{sol}^1$ is just the right handed solenoid considered above. $ {\Gamma}_{han} $ is associated with the handle connecting the two superhelices and its contribution to $\chi_W $ can be neglected in the large $ n $ limit. $ {\Gamma}_{sol}^{2} $ is the tangent vector trajectory generated by the superhelix winding down the $ z $ axis; it is given by performing upon the $ {\Gamma}_{sol}^1$ the transformation defined by the change of spherical coordinates : $ {\phi}^{\prime}=- \phi \;, \: {\theta}^{\prime}= \pi - \theta $. One readily gets $\chi_W ({\Gamma}_{plec} )$ by writing : $\chi_W ({\Gamma}_{plec} ) = \chi_W ({\Gamma}_{sol}^1 )+\chi_W ({\Gamma}_{sol}^{2} )= - 4\pi\, n\,\cos {\theta}_0$. We proceed next to the deformation of $ \hat{{\bf{t}}} $ trajectories $\chi_W ({\Gamma}_{sol} )$ and $ {\Gamma}_{plec} $ by applying a rotation about the $ y $ axis of an angle $\alpha$ which is going to vary continuously from 0 to $ \pi $. We will find that a proper use of formula (\ref{wr}) leads to rotation invariant results in the limit of small but finite angular cutoff. As a first step, let us show that $\chi_W (\Gamma )$ can be written as the circulation along $\Gamma$ of a magnetic monopole vector potential. This technical detour will also be used in the forthcoming derivation of the RLC model Hamiltonian. \centerline{\it 3. The local writhe $\chi_W $ as a magnetic monopole potential vector line integral } Our aim is to prove that the writhe supercoiled angle $ \chi_W$ can be written as the line integral $ \int d\phi \, A_{\phi}= \oint {{\bf{A}}}_{m}({\bf{r}}) \,d{\bf{r}} $ where $ A_{\phi} = \(( 1- \cos\theta \))$ will be identified as the $\phi $ spherical component of the potential vector ${{\bf{A}}}_{m}({\bf{r}})$ of a magnetic monopole of charge unity. Following the well known Dirac procedure, we write ${{\bf{A}}}_{m}({\bf{r}})$ as the potential vector of a thin solenoid of arbitrary long length, the so called Dirac String $\cal{L}$, lying along the negative half $z$-axis : \be {{\bf{A}}}_{m}({\bf{r}})= \int_{ \cal L } d {\bf{l}} \wedge {{\bf{\nabla}}}_r \frac{1}{| {\bf{r}}-{\bf{l}}|}= \frac{ \hat{{\bf{z}}} \wedge {\bf{r}}}{r (r+\hat{{\bf{z}}}\cdot{\bf{r}})} \label{monop1} \ee where $ {\bf{l}} = u \hat{{\bf{z}}} $ with $ -\infty < \, u \leq 0$. Using the above expression of ${{\bf{A}}}_{m}({\bf{r}})$, one derives the two basic equations which allow the writing of $ \chi_W $ as the circulation of the magnetic monopole potential vector, valid when $ r + \hat{{\bf{z}}}\cdot{\bf{r}} > 0 $ : \bea {{\bf{A}}}_{m}({\bf{r}})\cdot d{\bf{r}} &= & \(( 1- \cos\theta \)) d\phi \label{monop2} \\ {{\bf{B}}}_{m}({\bf{r}}) &= & {\bf{\nabla}} \wedge {{\bf{A}}}_{m}({\bf{r}})= \frac{{\bf{r}}}{r^3} \label{monop3} \eea where $ {{\bf{B}}}_{m}({\bf{r}}) $ is indeed the magnetic field produced by the magnetic charge unit. Let us now define $ S_{+}(\Gamma )$ ( resp $ S_{-}(\Gamma )$ ) as the part of the sphere $S_2$ bounded by the closed circuit $\Gamma$ which is run anticlockwise ( resp clockwise ) around the surface normal. ( Note that $ S_{+}(\Gamma ) \cup S_{-}(\Gamma )= S_2$ ). Let us call $ {\Sigma}_{\epsilon} $ the part of $S_2$ defined by the south pole hole and assume that $ S_{+}(\Gamma ) \cap {\Sigma}_{\epsilon} =\emptyset $ ( as a consequence: $ S_{-}(\Gamma ) \cap {\Sigma}_{\epsilon} = {\Sigma}_{\epsilon} $). Applying the Stokes theorem to formula ( \ref{monop3} ), we can write: \be \chi_W (\Gamma)= {\oint}_ {\Gamma } {{\bf{A}}}_{m}({\bf{r}})\cdot d{\bf{r}}= {\int \int}_{S_{+}(\Gamma )} d {\bf{S}} \cdot {{\bf{B}}}_{m}({\bf{r}})= {\cal A}( S_{+}(\Gamma )) \ee where ${\cal A}( S_{+}(\Gamma ))$ is the area of the spherical cap $ S_{+}(\Gamma )$. The closed circuit $\Gamma$ is assumed here to be run only once; if it is run $n$ times, as in solenoid and plectoneme configurations, $\cal{A}( S_{+}(\Gamma ))$ should then be multiplied by $n$ and the formulas obtained before are immediately recovered. A similar formula, valid {\it mod} $2 \pi $, has been derived previously by Fuller \cite{fullerform}, in the case of closed molecular chain. \centerline{ \it 4. On the rotation invariance of the writhe supercoiling angle $\chi_W$} We are going now to follow the variation of $ \chi_W (\Gamma) $ when the initial tangent vector $\hat{{\bf{t}}}$ trajectory ${\Gamma}_0$ is moved continuously upon the sphere by applying a rotation about the $ y $ axis by an angle $\alpha$, varying continuously from 0 to $ \pi $: $ {\Gamma}_0 \Longrightarrow \; \Gamma(\alpha)= R(\hat{{\bf{y}}} ,\alpha) {\Gamma}_0 $. We shall choose $ {\Gamma}_0 = {\Gamma}_{sol} $ but the following analysis can be easily extended to plectonemes or more general configurations involving closed $ \hat{{\bf{t}}} $ trajectories. The crossing of the north pole is harmless because $ 1-\cos\theta $ vanishes at $ \theta= 0 $ and everything goes smoothly until $ \Gamma(\alpha)$ meets the south pole hole ( $ \theta= \pi$ ) when $\alpha$ is approaching $ \pi -{\theta}_0$. When $ \alpha < \pi-{\theta}_0 $ the south pole stays outside $ S_{+}(\Gamma( \alpha ))$ and ${\chi}_W (\Gamma)$ is given by $n$ times the spherical cap area: ${\cal A}( S_{+}(\Gamma(\alpha ))= {\cal A}( S_{+}( {\Gamma}_0 )) = 2\pi ( 1- \cos {\theta_0} )$. In other words $ \chi_W (\Gamma(\alpha) ) = \chi_W ( {\Gamma}_0) $ if $ \alpha < \pi-{\theta}_0 $. In order to avoid the hole $ \Gamma(\alpha) $ must be continuously deformed for $ \alpha > \pi-{\theta}_0 $ into the circuit ${\Gamma}^{\prime}(\alpha)= \Gamma(\alpha) \cup {\Gamma}_{\epsilon} $ where the path $ {\Gamma}_{\epsilon} $ consists in $n$ turns run anticlockwise around the south pole hole of radius $\epsilon$. (We ignore the two ways narrow lane connecting the two loops which gives a vanishing contribution.) It is clear now that $ S_{+}(\Gamma(\alpha ) $ contains the south pole, so we must apply the Stokes theorem to the clockwise spherical bowl $ S_{-}(\Gamma(\alpha ))$: $$ {\chi}_W (\Gamma(\alpha))= -n {\int \int}_{S_{-}(\Gamma (\alpha ))} d {\bf{S}} \cdot {{\bf{B}}}_{m}({\bf{r}})= -n {\cal A}( S_{-}(\Gamma(\alpha )) = n ({\cal A}( S_{+}(\Gamma(\alpha ))- 4 \pi) $$ The contribution of the path ${\Gamma}_{\epsilon}$ is readily found to be $ 4 \,\pi \, n (1+ O({\epsilon}^2 ) )$ so it just cancels the $ - 4 \,\pi \, n $ term in the above result. As a consequence, the writhe supercoiling angle associated with the south pole avoiding $\hat{{\bf{t}}}$ trajectory ${\chi}_W ({\Gamma}^{\prime}(\alpha))$ coincides when $ \pi-{\theta}_0 <\alpha < \pi $ with the $ \alpha=0 $ initial result: ${\chi}_W ({\Gamma}^{\prime}(\alpha))= \chi_W ( {\Gamma}_0)= 2\pi n ( 1- \cos {\theta_0} ) $ up to corrections $ O({\epsilon}^2 ) $. {\it In conclusion, we have shown that a proper use of formula (\ref{wr}), in the simple case of solenoidal configurations, leads to rotation invariant $\chi_W$ in the limit of small $\epsilon$. } This result can be easily extended to the plectonemic configurations and even to more general ones. The crucial cancellation is taking place in the vicinity of the south pole crossing and thus does not depend on the detailed shape closed trajectory ${\Gamma}_0$ in the $\hat{{\bf{t}}}$ space. If one adds the minimal extra string section to the solenoid or plectoneme, necessary to get a closed loop in the physical space, our findings are in agreement, in the limit of large $n$, with those obtained from the non-local closed loop writhe formula \cite{geometry}, which is explicitly rotation invariant from the start. Formula (\ref{wr}) is thus a correct description of the problem provided the trajectory is obtained from the straight line $ (\theta(s)=0 ,\phi(s)=0) $ by a continuous transformation on the unit sphere pierced at $\theta=\pi$. A similar condition was obtained by Fuller \cite{fullerform} in his local Writhe formula for closed strings. An alternative derivation of our result, more similar to the approach of Fain \cite{fain}, could be to close the open string by some straight line closing at infinity and rely on Fuller's formula, but this procedure raises subtle questions about the contribution of added pieces which made us prefer the present explicit derivation. \begin{center} \subsection{ The Rod Like Chain Hamiltonian} \end{center} The partition function for a fixed value of $ \chi $ is given by the path integral in the space of Euler angles : \be Z(\chi,F) =\int { \cal D }(\theta,\phi,\psi) \, \delta \((\chi-\int_0^L ds (\dot{\phi}+\dot{\psi})\)) \exp- {E_{RLC}\over k_{B}T} \label{pathint} \ee The functional space $ \theta(s),\phi(s),\psi(s) $ should be specified in accordance with the previous considerations, in particular the tangent vector $ \hat{{\bf{t}}} $ paths must bypass the holes on the pierced sphere $ S_2 $. This can be achieved by introducing explicitly in $ E_{RLC} $ a short range repulsive potential near $ \theta= \pi $. Two extra constraints, non local in the tangent vector $\hat{{\bf{t}}}$ space should have been, in principle, implemented in the above partition function. The first one concerns self avoidance effects induced by the screened Coulomb repulsion between different molecular sections. Even the simple version of this constraint, used in supercoiled plasmids Monte Carlo simulations\cite{volog}, is exceedingly difficult to implement in the Quantum Mechanics inspired formalism of the present paper. Here Coulomb induced self avoidance wail be ignored as all quantitative analytic approaches did so far, including the celebrated WLC computation. It will appear that these effects play a limited role in the low supercoiling regime ( $\vert\sigma \vert \leq 0.02 $) to be analyzed in the present paper, a regime and experimental conditions where the WLC model is also working beautifully. The second non-local constraint not implemented in the present work concerns the exclusion of knotted chain configurations. In principle, knotting an open DNA chain is not forbidden, but such a transition is expected to be inhibited on the time scale of the actual experiment since the contour length of the molecule $L$ is only 1.7 times the circumference of the bead glued at the free end of the molecule. Using a Fourier representation of the Dirac $ \delta $ function we can write the partition function as a Fourier integral: $ Z(\chi,F)=\int dk \exp(-ik \, \chi ) \tilde Z(k) $ where the Fourier transform $ \tilde Z(k) $ is given by the functional integrals: \begin{eqnarray}\nn \tilde Z(k) & = & \int { \cal D }(\theta,\phi) \,\exp \(( i\,k\, \chi_W- {{\cal E}_{bend}+{\cal E}_{stretch} \over k_{B}T} \)) {\tilde Z}_{T}(k) \label{Ztild} \\ {\tilde Z}_{T}(k) & = & \int { \cal D }(\psi) \,\exp \(( i\, k\, T_{w} -{{\cal E}_{twist} \over k_B T}\)) \label{ZtildT} \end{eqnarray} where the above factorization follows from the identity: $\int_0^L ds (\dot{\phi}+\dot{\psi})= T_w +\chi_W$, using the elastic energy densities defined in (\ref{EL}). The Gaussian path integral upon $ \psi $ is readily performed and gives the Fourier transform of the twist partition function of equation (\ref{ZT}) $ {\tilde Z}_T(k) = \exp( -{k^2L\over 2 C})$, up to a trivial constant. The partition function Fourier transform ${\tilde Z}(k)$ can then be written as the product: $ {\tilde Z}(k)= {\tilde Z}_T(k) {\tilde Z}_W(k) $ where $ {\tilde Z}_W(k) $, which is interpreted as the writhe partition function Fourier transform, is given by a path integral upon $\theta $ and $ \phi $, with the effective energy: \be {\tilde E}_W (k) = {E_{WLC}\over k_{B}T} +i\,k\, \chi_W \label{Etild} \ee Using the results of the previous section ones sees immediately that $ i\,k\, \chi_W $ can be written as $i$ times the line integral $ \int{{\bf{A}}}_{m}({\bf{r}}) \,d{\bf{r}} $ where ${{\bf{A}}}_{m}({\bf{r}})$ is the vector potential produced by the magnetic monopole, having now a charge $k$. If one performs, as in Section II.B, an analytic continuation of the integral over $s$ in $ {\tilde E}_W (k) $ towards the imaginary axis the $ i $ factor disappears in the writhe line integral. One arrives to the action integral of a unit charge particle moving on the unit sphere under the joint action of an electric field $ f$ and a magnetic monopole of charge $k$. In order to get the corresponding Hamiltonian ${\hat H}_{RLC}(k)$ we apply a standard Quantum Mechanics rule used to implement the switching on of a magnetic field: we replace in the kinetic terms $ \frac{1}{2 A} {{\bf{p}}}^2 $, the particle momentum ${\bf{p}} $ by ${\bf{p}}-e{{\bf{A}}}_{m}({\bf{r}}) $ (here $e$=1). The term linear in ${{\bf{A}}}_{m}({\bf{r}})$ disappears by averaging over the final angle $ \phi = \phi(L)$ and we are left with the diamagnetic term: \be \frac{1}{2 A} {{\bf{A}}}_{m}^2({\bf{r}})=\frac{k^2}{2A}{1-\cos\theta \over 1+\cos\theta} \label{diam} \ee Adding (\ref{diam}) to the WLC Hamiltonian, one arrives to the dimensionless Hamiltonian ${\hat H}_{RLC}(k)$ associated with the partition function ${\tilde Z}_W(k) $: \be {\hat H}_{RLC}(k) = -\frac{1}{ 2\,\sin\theta }\frac{\partial}{\partial\,\theta}\sin\theta \, \frac{\partial}{\partial \theta }- \alpha \cos\theta+{k^2 \over 2} {1-\cos\theta \over 1+\cos\theta} \label{HRLC} \ee where we have taken, as the length unit, the persistence length $ A $ and introduced the dimensionless force parameter $\alpha= {F A \over k_{B} T}$. { \it Let us stress that the above Hamiltonian is a straightforward consequence of the expression for the elastic energy ( Eq \ref{EL0},\ref{EL}) and our expression (\ref{chif}) giving the supercoiling constraint $\chi$ in terms of Euler angles derivatives. } Moreover, as it is shown in Appendix B, ${\hat H}_{RLC}(k)$ can be derived directly, {\it without having to introduce the local writhe given by formula (\ref{wr})}. The method involves the quantization of the symmetric top Hamiltonian as a differential operator acting upon non-periodic functions of the Euler angles $ \phi $ and $\psi$. This is required to describe the thermal fluctuations of an elastic rod ( see section II.B) and $k$ appears as the continuous angular momentum along the top axis. Introducing the eigenstates and the eigenvalues of ${\hat H}_{RLC}$, $ {\hat H}_{RLC}{\Psi}_{n}(k,\theta)= {\epsilon}_n(\alpha,k^2) {\Psi}_{n}(k,\theta)$, the Fourier transformed partition function $\tilde Z$ can be written as the sum: \be \tilde Z=\sum_n { \Psi}_{n}(k,\theta_0){\Psi}_{n}(k,\theta_L) \exp\((- \frac{L}{A} \(({\epsilon}_n(\alpha,k^2)+ {k^2 A \over 2 C }\))\)) \ee (Note that ${\hat H}_{RLC}$ being a real operator the eigenfunctions can always be taken as real.) In the large $L$ limit, the sum over the eigenstates is dominated by the one with lowest energy $\epsilon_0(\alpha,k^2)$, if $ L/A \gg \Delta\epsilon $ where $\Delta\epsilon$ is the energy gap between the ground state and the nearest excited state of ${\hat H}_{RLC}$. This gives the approximate expression for the partition function $ Z ( \chi,F) $ : \be Z \simeq \int dk \ \exp\((- \frac{L}{A} \(( {\epsilon}_0(\alpha,k^2) + {k^2 A \over 2 C} \))-i\, k\,\chi \)) \label{Zchi} \ee To get the above equation, we have also neglected in $ \tilde Z(k)$ the prefactor involving the ground state wave function: ${ \Psi}_0(k,\theta_0){\Psi}_0(k,\theta_L)$. It leads to finite size corrections of order $ A/L $, which are in general dominant with respect to those associated with the excited states, which scale as $ \exp( - \Delta\epsilon L/A )$. \subsection{ The pathology of the RLC model in the continuous limit.} In order to comply with the prescription given for the path integral (\ref{pathint}) we should have added to ${\hat H}_{RLC}(k)$ a strong repulsive potential acting in the interval: $ \pi -\epsilon < \theta <\pi $, in order to avoid the $ \theta =\pi $ singular point. We shall refer to the $\epsilon \to 0$ limit as the 'continuous limit', and show that in that limit the model is pathological. Let us derive from the ground state energy $\epsilon_0(\alpha, k^2)$ of the Hamiltonian ${\hat H}_{RLC}$, some observable properties of a long RLC. In the following the variable $ z = k^2 $ is assumed to take positive real values. If instead of constraining $\chi$ one measures its thermal fluctuations, their probability distribution is just $P(\chi) \propto Z$. Let us consider the second moment as we did in subsection II.B for the case of a non-flexible rod. \bea <\, {\chi}^2 \, > & =-& \lim_{k^2\to 0}{1\over\tilde Z (k)} { \partial^2 \tilde Z (k)\over\partial\, k^2 } \\ & = & { L\over C}+ { 2\, L\over A} \lim_{k^2\to 0}{ \partial \epsilon_0(\alpha,k^2) \over \partial\, k^2 } \label{chi2} \eea We recognize in the first term of equation (\ref{chi2}) the contribution to $<\chi^2>$ from the twist fluctuations, ${ L\over C}$, obtained previously. The second piece of (\ref{chi2}) gives the contribution from $\langle \chi_W^2 \rangle$, but it turns out to be divergent. Evaluating $\epsilon_0(\alpha,k^2)$ at small $k^2$ from standard perturbation theory, we find $ \langle{\chi_W}^2 \rangle = (L/A) \langle \Phi_0\vert(1-\cos\theta)/( 1+\cos\theta)\vert \Phi_0\rangle $ where $\Phi_0(\theta)$ is the groundstate eigenfunction of the WLC Hamiltonian (which is ${\hat H}_{RLC} $ at $ k=0 $). As $\Phi_0(\theta=\pi)\ne 0$ (for any finite force), we get clearly a logarithmically divergent result. To study analytically this peculiar theory two methods have been followed, associated with different limits. {\it i)The weak force limit}: When $ \alpha =0 $ the eigenvalue problem for $ H_{RLC}$ can be solved exactly. The eigenfunctions are given in terms of Jacobi polynomials : $${\Psi}_n(k,\theta)= (1+\cos\theta)^{k} P_n^{(0,2 k)}(\cos\theta)$$ and the associated eigenvalues are: $$ {\epsilon}_0( \alpha=0,k^2)= \(( (2 n+1) k +n^2+n \))/2 $$ In the presence of a small force ($\alpha \ll 1$), the ground state energy shift is easily calculated to second order in $ \alpha $: $$ \delta_2 {\epsilon}_0(\alpha,k^2)= -{\alpha k\over 1+k} - {{\alpha}^2 \over (1 +k)^3 (3 + 2 k )} $$ Note that $ {\epsilon}_0(\alpha,k^2 ) $ is not an analytic function of $ z=k^2 $. {\it ii)The small $k $ limit}: In order to build a perturbative expansion near $k =0 $ it is convenient to factorize in the wave function the singular term near $\theta=\pi $ by writing : ${\Psi}_0(k,\theta)= (1+\cos\theta)^{k} {\chi}_0(k,\theta) $. The wave function $\chi_0(k,\theta)$ obeys a new wave equation which is amenable to a well defined perturbation expansion in $ k $. It will turn out that the values of $ k $ relevant for the evaluation of Fourier transform are of the order of $ A/L$, so that a first order perturbation is sufficient (at least when $ \alpha $ is small enough in order to guarantee that $ \Phi_0(\pi)^2 $ stays of the order of unity). We have found that the ground state energy is given to first order in $ k$ by: $ \epsilon_0(\alpha,k^2) = \epsilon_0(\alpha,0)+ k \Phi_0(\pi)^2+ O(k^2) $. It can be checked that the perturbation result of method {\it i)} is consistent with the above equation. The Fourier integral can be readily performed, leading to a Cauchy type probability distribution, which is best expressed in terms of the reduced variable $\eta=\chi A/L$: $ P(\eta ) \simeq {1\over \pi}\,{\Phi}_0(\pi)^2 /( {\Phi}_0(\pi)^4 + \eta^2 )$ and thus leading to a diverging second moment.This last point can also be seen from the fact that ${ \partial \epsilon_0(\alpha,k^2) \over\partial\, k^2 } \propto {1\over k} $ in the small $k$ limit. The relative extension of the chain in the direction of the force is given by ${\langle z \rangle / L} = (A / L) {\partial \ln Z \over \partial \alpha}$. The partition function in the present case is easily found to be : $ Z( \eta,\alpha) \propto \exp\(( -L/A\, \epsilon_0(\alpha,0) \)) \, P(\eta ) $. It leads to: $ {\langle z \rangle / L} = -{\partial \epsilon_0 \over \partial \alpha} (\alpha,0) +O(A/L) $, {\it i.e.} the same result as the WLC model. { \it The present computations show that the continuous RLC model leads to predictions which are both pathogical ( infinite second moment of the writhe spontaneous fluctuations ) and in striking contradiction with experiment ( absence of variation of the average extension with respect to the supercoiling angle $\chi$ at moderate forces ). All these peculiar features have been confirmed with explicit computations using a regularized $ H_{RLC}$ hamiltonian. We have indeed found that $< \chi_{W}^2 >$ scales like $ {L \over A} \log{1 \over \epsilon} $ and that the extension versus supercoiling curves flatten up in the small $\epsilon$ limit.} As a last remark we would like to note that if the variable $z$ is continued analytically towards negative value $ z \rightarrow -{\kappa}^2$ the groundstate energy develops an imaginary part.This a clear indication that ${\hat H}_{RLC}( i\kappa ) $ has no stable ground state. \begin{center} \section{ Discretization as a Regularization Method for the Rod Like Chain Model } \label{discrete_theory} \end{center} The introduction by hand of an angular cutoff near $ \theta =\pi$ appears as an {\it ad hoc } procedure. We are going to show in the present section that a discretization of the chain involving an elementary link of length $b$ generates an angular cutoff: $ \sin^2\theta \geq {b\over A} $. We have seen in section II.A that the RLC model with its cylindrical symmetric rigidities tensor is realistic only in presence of a finite resolution $ \Delta \,l $ in the length measurement. The existence of a length cut-off $ b \sim \Delta \,l $ appears not only natural but necessary in the present context. The non trivial fact is that the "mesoscopic " elastic properties, taking place on the length scale of the whole supercoiled molecule (about ten microns), are sensitive to the existence or not of a cutoff in the range of few nanometers. We should also stress that length cutoff effects are found to be reduced to the level of few percents when supercoiling is absent. Therefore in that case the continuous version of the WLC model remains a remarkably good description of the DNA force extension measurements. Moroz and Nelson \cite{mornel1} have devised a computation procedure where no angular cutoff is introduced from the start. They consider the situation where the torque acting upon the molecular end is kept fixed so that the relevant Hamiltonian is just ${\hat H}_{RLC}( i\kappa ) $. The pathology discussed above is expressed in the fact that this Hamiltonian has no stable ground state. In order to deal with such a situation, they use a perturbation method near the high force limit where the relative elongation is close to unity. More precisely, they construct a perturbation expansion involving negative power of $ K= \sqrt{\alpha - {1\over 4} {\kappa}^2 }$. The lowest order Hamiltonian potential is given, up to a constant, by the small angle approximation: $ {1\over 2} K^2 {\theta }^2={1\over 2} ( \alpha - {1\over 4} {\kappa}^2 ) {\theta }^2 $. Since ${\hat H}_{RLC}( i\kappa ) $ is not a self adjoint operator, the series diverges and it is expected to be, at best, an asymptotic series. The regime of validity of the prediction is difficult to assess in this approach since one is basically approximating a divergent ground state energy from the first terms of the perturbation series. As long as one sticks to a finite order, the $ \theta = \pi $ singularity does not show up. In order to make contact with experiment, Moroz and Nelson have been forced to restrict their analysis to the domain of force and supercoiling where $ K > K_c=3 $. In our approach, we have instead regularized the model explicitely, as we shall discuss. This allows to get a better control of the theory and it turns out that our results hold on a much wider range of force-supercoiling. It can even be compared in the $ F=0 $ limit to supercoiled plasmids experimental data and Monte Carlo simulations. ( See Section \ref{tw_wr_pl} ). \subsection{ Construction of the transfer matrix associated with the discretized RLC model } The continuous elastic RLC is transformed into a chain composed of $ N $ elementary links of length b. The continuous arc length $ s$ is replaced by a discrete set: $ s_n = n \,b $ with $ 0 \leq n \leq N $ and $ L= N b $. We have now to specify the discretization rules. For the bending linear energy density ${\cal E}_{bend}$, there is a rather natural prescription: it is to replace $ \vert {d \hat{{\bf{t}}}(s) \over ds} \vert ^2 $ by $ {1 \over b^2} { \vert {\hat{{\bf{t}}}}_{n} - {\hat{{\bf{t}}}}_{n-1} \vert}^2 $. We are led in this way to {\it the first discretization rule}: \be {\cal E}_{bend} \Longrightarrow { A \over b^2} \lbrack 1-\cos( {\theta}_n - {\theta}_{n-1} )+ ( 1-\cos( {\phi}_{n} - {\phi}_{n-1} )\,\sin {\theta}_{n} \, \sin {\theta}_{n-1} \rbrack \label{rule1} \ee The above discretized linear density is endowed with remarkable feature: it is $ 2\pi $-periodic with respect to the angular finite difference $ {\phi}_n - {\phi}_{n-1} $. This property plays an essential role in the derivation of the transfer matrix $ T({\theta }_n,{\theta }_{n-1})$ which allows a direct computation of the partition function ${\tilde Z}(k)$, averaged upon the azimuthal angle $ \phi(L)$. For the term generated by the supercoiling constraint: $ ik (1-\cos\theta(s) )\dot{\phi} $, we shall impose the same periodicity condition, together with a symmetry under the exchange $ {\theta }_{n+1} \Leftrightarrow {\theta}_n $, in order to insure the hermiticity of the transfer matrix. We arrive in this way to {\it the second discretization rule}: \be \(( 1- \cos \theta \)) \, \dot{\phi} \Longrightarrow { 1 \over b} \sin ( {\phi}_n - {\phi}_{n-1} ) \, \(( 1 - { \cos {\theta}_n +\cos {\theta}_{n-1} \over 2} \)) \label{rule2} \ee The discretized version ${\tilde Z}_{N}(k,{\theta}_{N},{\phi}_{N})$ of the writhe partition function can be written as the following $2 N$ dimension integral: $$ {\tilde Z}_{N}(k,{\theta}_{N},{\phi}_{N} )=\int\,\prod_{n=1}^{N} d{\phi}_{n-1}\, d(\cos{\theta}_{n-1})\, \exp -b\,\(( \Delta {\cal E}( k,{\theta}_n,{\theta}_{n-1},{\phi}_n-{\phi}_{n-1})\)) $$ where $\Delta {\cal E}( k,{\theta}_n,{\theta}_{n-1},{\phi}_n-{\phi}_{n-1}) $ is the discretized linear energy density obtained from (\ref{Etild}) by the replacement rules (\ref{rule1}) and (\ref{rule2}). It is convenient to introduce the partition function $ {\tilde Z}_{n}(k,{\theta}_{n},{\phi}_{n} )$ corresponding to the intermediate value $ s_n= b n $. One gets then the recurrence relation: \begin{eqnarray} {\tilde Z}_{n}(k,{\theta}_{n},{\phi}_{n}) &=& \int d{\phi}_{n-1}\,d\((\cos {\theta}_{n-1}\)) \, \exp -b \,\(( \Delta {\cal E}( k,{\theta}_n,{\theta}_{n-1},{\phi}_n-{\phi}_{n-1})\)) \nonumber \\ & & {\tilde Z}_{n-1}(k,{\theta}_{n-1},{\phi}_{n-1}) \label{recur1} \end{eqnarray} Since we are ultimately interested in the partition function ${\tilde Z}(k)$ averaged upon the azimuthal angle $ \phi(L)$, we shall define : \be z_n(k,{\theta}_n)= {1 \over 2 \pi M }\int_{ -\pi\, M}^{ \pi\, M} d{\phi}_{n} \, {\tilde Z}_{n}(k,{\theta}_{n},{\phi}_{n}) \label{znav} \ee where $ M $ is an integer which can be arbitrarily large. We perform the above integral upon $ {\phi}_{n} $ in the two sides of equation (\ref{recur1}). In the right hand side we change the integration variables $ {\phi}_{n} , {\phi}_{n-1} $ into the new set: $u_n= {\phi}_{n} - {\phi}_{n-1},{\phi}_{n-1} $. Using the periodicity of the integrand with respect to $ u_n $, we can make the replacement: $$ {1 \over 2 \pi M } \int_{ -\pi\, M - {\phi}_{n-1} }^{ \pi\, M- {\phi}_{n-1} } du_n ... \Longrightarrow {1 \over 2 \pi } \int_{ -\pi }^{ \pi} du_n ... $$ The integral upon $ u_n $ can then be performed exactly in terms of the Bessel function of second type $ I_0(z) $ and we arrive finally to an explicit recurrence relation involving the $ \phi$-average partition function $z_n(k,{\theta}_n)$ : \begin{eqnarray} z_n(k,{\theta}_n) &=& \int_0^{\pi} \sin {\theta}_{n-1} d{\theta}_{n-1}\, T({\theta }_n,{\theta }_{n-1} )\,z_{n-1}(k,{\theta}_{n-1})\label{recur2} \\ T({\theta }_1,{\theta }_2,k^2) & =& \exp-\lbrace \frac{A}{b}\, \left( 1 - \cos ({\theta }_1-{\theta }_2 ) + \sin {\theta }_1\, \sin {\theta }_2 \right) \nonumber \\ & & +\frac{b\,\alpha }{2\,A} \left( \cos {\theta }_1 + \cos {\theta }_2 \right) \rbrace \,I_0 (f({\theta }_1,{\theta }_2,k^2)) \label{mtrans} \end{eqnarray} where the Bessel function argument is given by : \be f({\theta }_1,{\theta }_2,k^2) = \sqrt{ -k^2 \,{ \left( 1 - \frac{\cos {\theta }_1 + \cos {\theta }_2 } {2} \right) }^2 + \frac{ A^2}{ b^2} {\left( \,\sin{\theta }_1\, \sin{\theta }_2 \right)}^2 } \label{besarg} \ee To get some basic elements of the transfer matrix formalism it is useful to rewrite the recursion relation (\ref{recur2}) within an operator formalism: \\ $ \vert z_{n} \rangle = \hat{ T }(\alpha,k^2)\vert z_{n-1} \rangle $ where $\hat{ T }(\alpha,k^2)$ is the operator associated with the transfer matrix $ T({\theta }_1,{\theta }_2,k^2)$. Let us define as $ \vert t_i(\alpha,k^2) \rangle $ the eigenstate of $\hat{ T }(\alpha,k^2)$ associated with the eigenvalue $ t_i(\alpha,k^2) $. Performing an expansion upon the above set of eigenstates, we can write $ \vert z_N \rangle $ as follows: $$ \vert z_N \rangle = {\hat{ T }(\alpha,k^2)}^N \,\vert z_{0} \rangle = \sum_{i} { t_i(\alpha,k^2)}^N \, \vert t_i(\alpha,k^2) \rangle \langle t_i(\alpha,k^2) \,\vert z_{0} \rangle $$ Let us assume that $N$ is large enough so that the above sum is dominated by the contribution from the lowest eigenvalue. The partition function $ Z(\chi) $ can then be written as: $$ Z(\chi)= Z \simeq \int dk \ \exp\((- {k^2 \over 2 C} +i\, k\,\chi \)) \, {t_0(\alpha,k^2)}^N $$ It is now convenient to introduce the { \it new definition }: \be \epsilon_0(\alpha,k^2) = - {A \over b}\, \ln t_0(\alpha,k^2) \label{epstr} \ee Note that we had already defined $ \epsilon_0(\alpha,k^2) $ as the lowest eigenvalue of $ H_{RLC} $; it is easily verified that the two definitions coincide in the limit $ b/A \ll 1$ {\it modulo } an irrelevant constant. With this new definition the equation (\ref{Zchi}) giving $ Z(\chi) $ holds true for the discretized RLC model, within the transfer matrix formalism. \subsection{ The angular cutoff induced by the discretization of the RLC model } In this section we are going to discuss what turned out to be a crucial milestone in our work: the understanding that a discretization of the DNA chain could provide the angular cutoff needed to make sense to the RLC model. The first indication came, in fact, from MonteCarlo simulations which will be discussed later on in section \ref{Monte_Carlo}. We would like to show here that, indeed, the angular cutoff comes out from the transfer matrix solution of the discretized RLC model. Moreover our analysis will lead us to the formulation of a regularized version of the continuous RLC model. Our procedure involves an analytic computation of the partial derivative with respect to $ k^2 $ of the transfer operator ground state energy $\epsilon_0(\alpha,k^2)$, as defined by equation (\ref{epstr}) : $\partial_{k^2} {\epsilon}_0= { \partial \epsilon_0(\alpha,k^2) \over \partial\, k^2 }$. This derivative is very important physically, for the following two reasons. First, we remind that $ \lim_{k^2\to 0} \partial_{k^2} {\epsilon}_0$ gives the second moment of the spontaneous "writhe" fluctuations $ \langle {\chi_w}^2 \rangle $ (see equation (\ref{chi2}) ), which was found to be infinite within the continuous RLC model. Second, for imaginary finite values of $ k=i \kappa $, $ \partial_{k^2} {\epsilon}_0$ enters in an essential way in the parametric representation of the extension versus supercoiling curves, to be derived in the next sections. The partial derivative $ \partial_{k^2} {\epsilon}_0 $ is expressed, using Eq.(\ref{epstr}), as \be { \partial \epsilon_0(\alpha,k^2) \over \partial\, k^2 } = {-A \over b\, t_0(\alpha,k^2) } \, { \partial t_0(\alpha,k^2)\over \partial\, k^2 } = {-A \over b\,t_0(\alpha,k^2) } \langle t_0 \vert {\partial\hat{ T }(\alpha,k^2)\over \partial\, k^2 } \vert t_0 \rangle \label{depsk2} \ee The partial derivative with respect to $ k^2$ of the transfer matrix $ T({\theta }_1,{\theta }_2,k^2) $ is easily obtained from equations (\ref{mtrans}) and (\ref{besarg}): \begin{eqnarray} {\partial T({\theta }_1,{\theta }_2,k^2) \over \partial\, k^2 } & = & T({\theta }_1,{\theta }_2,k^2) {\cal W}({\theta }_1,{\theta }_2) \nonumber \\ {\cal W }({\theta }_1,{\theta }_2) & = & -{1\over2} R( f({\theta }_1,{\theta }_2,k^2) ) { { \left( 1 - \frac{\cos {\theta }_1 + \cos {\theta }_2 } {2} \right) }^2 \over f({\theta }_1,{\theta }_2,k^2)} \end{eqnarray} We have defined the small $x$ cutoff function \be R(x)= {I_1(x)\over I_0( x)} \label{Rdef} \ee which behaves like $x$ when $ x \ll 1$ and goes rapidly to 1 when $ x >1 $. The next step is to investigate the variation of $ T({\theta }_1,{\theta }_2,k^2) $ as a function of $ \Delta\theta = {\theta }_2 -{\theta}_1 $ for a fixed value of $ \theta =({\theta }_1+{\theta }_2)/2 $ when $ A/b \gg 1 $. For the sake of simplicity we limit ourselves to value of $ k^2 $ such that $ \vert k^2 \vert {b^2\over A^2 }\ll 1 $ so that $ f( \theta_1 ,\theta_2 ,k^2 ) $ can be approximated by $ {A \over b} \sin\theta_1 \sin\theta_2 $. The transfer matrix is then given by the somewhat simplified expression: \be T({\theta }_1,{\theta }_2,k^2) = \exp-\lbrace \frac{A}{b}\, \left( 1 - \cos ({\theta }_1-{\theta }_2 ) \right) + \frac{b\,\alpha }{2\,A} \left( \cos {\theta }_1 + \cos {\theta }_2 \right) \rbrace B_0( \frac{A}{b} \sin {\theta }_1\, \sin {\theta }_2 ) \ee where we have introduced the auxiliary function: $ B_0(x)= \exp(-x) I_0( x) $ which is slowly varying with $x$: constant near the origin it behaves like $ 1/\sqrt{x} $ for large $x$. This formula shows that $ T({\theta }_1,{\theta }_2,k^2)$ is strongly peaked at small values of $\vert \Delta\theta \vert $, of order $\sqrt{b/A}$. This suggests the decomposition of ${\cal W}({\theta }_1,{\theta }_2)$ as a sum of two terms involving a local potential $U_w ( \theta)$ plus a small correction: \begin{eqnarray} {\cal W}({\theta }_1,{\theta }_2) & = & \(( U_w ({\theta }_1)+U_w ({\theta }_2 ) \))/2 + \Delta {\cal W}({\theta }_1,{\theta }_2) \\ U_w ( \theta) &=& {\cal W}(\theta ,\theta )= -{b\over 2 A} R ( {A \over b} \sin^2 \theta ) \(({1-\cos\theta \over 1+\cos \theta } \)) \end{eqnarray} A Taylor expansion of $ \Delta {\cal W}({\theta }_1,{\theta }_2) $ with respect to $\Delta\theta$, with $ \theta $ kept fixed, gives: $ \Delta {\cal W} \propto {\Delta\theta}^2 \(( 1 +O( {\Delta\theta}^2) \))$. It confirms that the contribution of $ \Delta {\cal W} $ relative to that of ${\cal W}$ is of the order of $ b/A $. To be more quantitative, we have computed the two averages $\langle \cal W \rangle$ and $\langle\Delta {\cal W } \rangle $ using as weight function $ T({\theta }_1,{\theta }_2) \sin{\theta }_1 \sin {\theta }_2 $. Taking $ b= .14 A $ we have obtained $ \vert \langle \Delta {\cal W} \rangle / \langle{\cal W} \rangle \vert=0.013 $. It appears then quite legitimate to set $ \Delta {\cal W} =0 $. This approximation allows us to write the partial derivative of the transfer operator $\hat{ T }$ as a symmetrized operator product: \be {\partial\hat{ T }(\alpha,k^2)\over \partial\, k^2 } = { \(( \hat{ T }(\alpha,k^2) \, {\hat{U}}_w + {\hat{U}}_w \,\hat{ T }(\alpha,k^2)\)) \over 2} \ee where $ {\hat{U}}_w $ stands for the operator associated with the local potential $ U_w ( \theta) $. Introducing this formula in eq. (\ref{depsk2}) and remembering that $ \vert t_0 \rangle $ is an eigenstate of $\hat{ T }$, the partial derivative $ \epsilon_0(\alpha,k^2) $ takes the simple form: \be { \partial \epsilon_0(\alpha,k^2)\over \partial\, k^2 } = { -A \, \langle t_0 \vert \hat{ T } \, {\hat{U}}_w+ {\hat{U}}_w \,\hat{ T } \vert t_0) \rangle \over 2 \, t_0(\alpha,k^2) }= {1 \over 2 } \langle t_0 \vert \(({1-\cos\theta \over 1+\cos \theta } \)) R ( {A \over b} \sin^2 \theta ) \vert t_0 \rangle \label{deregur} \ee The above result provides a very clear evidence that the discretization of the RLC model generates the angular cutoff around $\theta = \pi$ which we needed. Indeed, the integral involved in the above quantum average is now well defined for $ \theta= \pi$, due to the presence of the cutoff function $ R ( {A \over b} \sin^2 \theta ) $. Formula (\ref{deregur}) leads in a natural way, to a formulation of a regularized version of the continuous RLC model. The regularized operator Hamiltonian $ {\hat H}_{RLC}^{r} $ is obtained from the Hamiltonian $ {\hat H}_{RLC} $ given in equation (\ref{HRLC}) by multiplying the singular " writhe " potential by the cutoff function $ R ( {A \over b} \sin^2 \theta ) $. Standard Quantum Mechanics rules applied to $ {\hat H}_{RLC}^{r} $ give for $\partial_{k^2} {\epsilon}_0$ a formula identical to (\ref{deregur}), provided it is legitimate to neglect the small difference between $\vert t_0 \rangle $ and $\vert \Psi_0 \rangle $, respectively eigenstate of $ \hat{ T } $ and $ {\hat H}_{RLC}^{r}(k) $. We have verified that it is indeed so by comparing the results obtained from the above regularized RLC continuous model and those given by the transfer matrix method: they do coincide to the level of few $ \% $, at least in the domain of stretching force explored in the present paper. \begin{figure} \centerline{\epsfxsize=10cm \epsfysize=10cm \epsffile{dnaRLCfig1.ps} } \caption{The elongation versus the reduced supercoiling parameter $ \sigma ={ n \over L_{k0}} $ where $ L_{k0} $ stands for the number of double helix turns in absence of external constraints. The curve $ F=0.2 \, pN $, a typical "hat" curve, corresponds to the regime of entropic elasticity which is well described by the RLC model introduced in the present paper. The curve $ F=1 \, pN $ exhibits a plateau in the underwound region ($ \sigma < 0 $) which is associated with the denaturation of the DNA. In the third curve $ F=8 \, pN $ a second plateau has developed for ($ \sigma > 0 $); it has been interpreted as an induced transition to a new structure of DNA: the "P-DNA". } \label{fig1} \end{figure} \section {The RLC Model for a quantitative analysis of supercoiled DNA stretching experiments} \label{quant_ana} In this section we give the basic tools allowing to compute, within the RLC model, the various quantities measured experimentally. We shall first study the "hat" curve which is a graph giving, for a given force $ F$, the relative extension of molecule $ \langle { z(L)\over L} \rangle $ along the stretching force $ {\bf{F}} $, as a function of the supercoiling angle $ \chi= 2 \pi n $. Figure \ref{fig1} gives three examples of such "hat" curves taken within a rather large range of forces. The supercoiling angle is parameterized here by the ratio $ \sigma ={ n \over L_{k0}} $ where $ L_{k0} $ stands for the number of double helix turns in absence of external constraints. Positive values of $ \sigma$ correspond to overwinding, negative values correspond to underwinding. As it is apparent in Figure \ref{fig1}, only the curve associated with a force of $0.2 \, pN$ looks really like a "hat". That is so for all the extension versus supercoiling curves taken in the force range $ .06 \, pN \leq F \leq 0.45 \, pN $. In contrast, when $ F \ge .5 \, pN $ a plateau develops in the underwinding region. This suggests that the torsional energy is converted into chemical energy instead of entropic elasticity energy. More precisely, convincing experimental arguments have been given \cite{strick2,strick3} in favour of the following mechanism: the underwinding energy is used by the molecular system to open the hydrogen bonds linking the bases and, as a consequence, denaturation bubbles appear along the DNA molecule. When the force is further increased, say above $ 3 \, pN$, a plateau appears also in the overwinding region. This has received a very interesting interpretation \cite{pDNA} as a transition towards a new structure of the DNA: the so called " DNA-P ", which is also predicted by numerical simulations. In the present paper, we shall focus our analysis on the true " hat " curves, {\it i.e } those symmetric under the exchange $ \sigma \Leftrightarrow -\sigma $, as observed for $ F \leq 0.45 \, pN $. \subsection{ The saddle point method} The experiments suggest that the variations of the relative elongation $ \langle { z(L)\over L} \rangle $ versus the supercoiling angle $ \chi$ scale as a function of $\chi/L$. It is convenient to introduce an intensive supercoiling variable $ \eta = \chi A/L $ which will turn out to be at most of the order of few units in the domain we are going to explore. It is related to the usual variable $ \sigma $ by : \be \eta= { \chi A \over L }= { 2 \pi A \over p} \,\sigma =94.8 \,\sigma \label{etasig} \ee where we have used the values for the pitch: $ p=3.4 \, nm $ and the persistence length: $ A=51.3 \,nm $. Let us rewrite the partition function $ Z(\chi,F) $ given by equation (\ref{Zchi}) in terms of scaling variables: \be Z ( \eta,\alpha ) \simeq \int dk \ \exp - \frac{L}{A} \(( {\epsilon}_0(k^2,\alpha) + {k^2 A \over 2 C} -i \eta \,k \)) \label{Zeta} \ee The above partition function can be computed by the saddle point method in the limit $ L/A\gg 1$ with $ \eta $ kept fixed. The saddle point is imaginary, $k_c=i \kappa(\alpha)$ and is given by the equation: \be {A \over C} +2 {\partial \epsilon_0 \over \partial k^2}(\alpha,-\kappa^2)={\eta \over \kappa} \label{etakappa} \ee The saddle point contribution to the partition function $ Z ( \eta,\alpha ) $ reads as follows : \be \ln \(( Z ( \eta,\alpha ) \)) = -{ L\over A} \(( \epsilon_0(\alpha,-\kappa^2)- {\kappa^2 A\over 2 \,C }+ \eta \, \kappa \))+ O(1) \label{Zsad} \ee Let us first compute the torque $ \Gamma $ acting upon the free end of the molecule. The experiments to be analyse in this paper were not designed to measure $ \Gamma $ but there is an experimental project at ENS-Paris aiming at its direct empirical determination. \bea { \Gamma \over k_{B} \,T} & = -& {\partial \ln Z \over \partial \chi}= -{A \over L}{\partial \ln Z \over \partial \eta} \nonumber \\ & = & \kappa + { \partial \kappa \over \partial \eta }\(( \eta - {\kappa \,A\over C}- 2\, \kappa {\partial \epsilon_0 \over \partial k^2 }(\alpha,-\kappa^2) \)) = \kappa \label{torq} \eea where the term proportional to $ { \partial \kappa \over \partial \eta} $ vanishes because of the saddle point equation (\ref{etakappa}). Therefore we have found that $\kappa=\Im (\,k_c) $ is equal to the torque $\Gamma$ in units of $ k_{B} \,T$. One can introduce the thermodynamic potential ${{\cal G} \over k_B T }= -\ln \(( Z ( \eta,u ) \))- \kappa \chi $ and verify that the supercoiling angle $\chi $ given by equation (\ref{etakappa}) satisfies the thermodynamic relation: $ \chi= - {\partial \over \partial \kappa } \left( {{\cal G} \over k_B T } \right)$. We are now going to compute in the same way the relative molecule elongation $ \langle { z(L)\over L} \rangle $ : \be \frac{ \langle \, z(L) \, \rangle }{L} = \frac{k_B T}{L}\frac{\partial\ln \,Z }{ \partial F}= -\frac{\partial {\epsilon}_0(\alpha,-\kappa^2) }{ \partial \alpha} \label{extens} \ee As in the computation of the torque the term proportional to $ { \partial \kappa \over \partial \alpha } $ does not appear because it is multiplied by a factor which vanishes at the saddle point. In other words, the elongation is given by the same expression whether the experiment is performed under the conditions of fixed surpecoiling angle $ \chi $ or fixed torque $ \Gamma $. In contrast, the situation is very different for the elongation fluctuations: $ \langle \, z(L)^2 \, \rangle - {\langle \, z(L) \, \rangle} ^2 $ turns out to be much larger if measured when the torque is kept fixed instead of the supercoiling angle, as we shall discuss later. {\it From the knowledge of $\epsilon_0(\alpha, -\kappa^2)$, using jointly equations (\ref{etakappa}) and (\ref{extens}) one gets a parametric representation of the " hat " curves, the parameter being the torque in $ k_B \,T $ unit.} \begin{center} \subsection{ Solving the quantum mechanics eigenvalues problems associated with the RLC model } \end{center} In order to complete the description of the theory we now sketch the methods used to get the final theoretical ingredients: the groundstate eigenvalue $ \epsilon_0(\alpha,k^2 ) $ and its partial derivatives. We have followed two approaches. \subsubsection{ Method a): Ground state eigenvalue of the hamiltonian ${\hat{H}}_{RLC}^{r}$ associated with the regularized continuous RLC model} In the first approach, the computations are performed within the continuous RLC model regularized according to the cutoff prescription derived in section IV.B. What we have to do is to solve the ground state eigenvalue~problem: ${\hat{H}}_{RLC}^{r} \, {\Psi}_0(\theta)= {\epsilon}_0\,{ \Psi}_0(\theta)$. This ground state wave function is obtained as an appropriate solution of the ordinary differential equation associated with the eigenvalue problem: \be \frac{1}{ 2\,\sin\theta }\frac{\partial}{\partial\,\theta}(\sin\theta \ \frac{\partial}{\partial \theta}{\Psi}_{0}(\theta)) + \(( - V_{r} ( \alpha,-{\kappa }^2 )+ {\epsilon}_0 \)) {\Psi}_{0}(\theta)=0 . \label{ShRLC} \ee where the regularized potential $ V_{r} ( \alpha,k^2 ) $ is given by \be V_{r} ( \alpha,k^2 )= -\alpha \cos\theta + {k^2 \over 2} {1-\cos\theta \over 1+\cos\theta} \, { I_1( {A \over b} \sin^2 \theta )\over I_0( {A \over b} \sin^2 \theta)} \label{VRLC} \ee We search for a solution which satisfies regularity conditions both for $\theta =0 $ and $\theta=\pi$. These requirements are necessary and sufficient to guarantee that the differential operator ${\hat{H}}_{RLC}^{r} $ is a self-adjoint operator. They can be fulfilled if and only if the reduced energy $ \epsilon $ belongs to a discrete set of values. As a first step, one constructs by series expansion two solutions of (\ref{ShRLC}), ${\Psi}_{a}(\theta) $ and $ {\Psi}_{b}(\theta) $, which are respectively regular for $\theta =0$ and $ \theta = \pi $. One then proceeds to an outward numerical integration of ${\Psi}_{a}(\theta)$ and inward numerical integration of $ {\Psi}_{b}(\theta) $ up to an intermediate value of $ \theta = {\theta}_0 $. The energy eigenvalue equation is obtained as a matching condition for the two wave functions, which ensures the regularity of the eigenfunction for the whole physical domain of $\theta$. One writes, for $ \theta = {\theta}_0 $, the equality of the logarithmic derivatives of ${\Psi}_{a}(\theta)$ and ${\Psi}_{b}(\theta)$ : \begin{equation} \frac {\partial{\Psi}_{a}}{\partial \theta}/ {\Psi}_{a} - \frac {\partial{\Psi}_{b}}{\partial \theta}/ {\Psi}_{b}=0 . \label{eqvp} \end{equation} Equation (\ref{eqvp}) is then solved by a standard iteration method requiring a trial approximate eigenvalue. The construction of the parametric " hat " curve for a given value of $\alpha $ begins with a small value of $\kappa$ ( the energy $ {\epsilon}_0 (\alpha ,0) $ is easily obtained from the results of ref. \cite{bouc99} ). The trial eigenvalue needed to solve eigenvalue equation (\ref{ShRLC}) for a small value of $ \kappa $ is easily obtained from a first order perturbation calculation. Since the present method of solving the eigenvalue problem automatically yields the eigenfunction ${ \Psi}_0(\theta)$, the partial derivatives of eigenvalues are readily obtained by taking the quantum average of the corresponding partial derivatives of the potential $ V_{r} ( \alpha,k^2 ) $. Let us quote the derivative with respect to $ \alpha $ ( see equation (\ref{deregur}) of section IV.B. for the derivative with respect to $ k^2 $) \be \frac{ \langle \, z(L) \, \rangle }{L}= -\frac{\partial {\epsilon}_0(\alpha,k^2) }{ \partial \alpha} = \langle {\Psi}_0\,\vert \cos\theta \vert \, {\Psi}_0\rangle \label{cosav} \ee Once an eigenvalue is known for a value of $ k^2$, one proceeds to the neighboring value $ k^2 + \Delta k^2 $, by using $ {\epsilon}_0 + \Delta k^2 \, \frac{\partial {\epsilon}_0}{\partial\, k^2} $ as a trial eigenvalue and one proceeds to cover step by step the desired range of $ k^2=-{\kappa }^2 $. The fact that the state found with this procedure is really the ground state can be checked from the fact that the wave function has no node. \subsubsection{Method b): The transfer matrix iteration} The search for the smallest eigenvalue $ t _0(\alpha,k^2) $ of the transfer operator $ \hat{ T }(\alpha,k^2) $ defined by eq. (\ref{recur2}), (\ref{mtrans}) and (\ref{besarg}) is done by iteration of the mapping $\vert z_{n} \rangle = \hat{ T} \,\vert z_{n-1} \rangle $. Indeed, $ \vert z_{N} \rangle ={\hat{ T }}^N \,\vert z_{0} \rangle = \sum_n \vert t_n \rangle \langle t_n \vert z_0 \rangle \,t_n^N $ is reduced to the lowest eigenvalue contribution when $ N \rightarrow \,\infty $. Let us write the above mapping as a linear functional transform: \be z_{n}(k,{\theta}_n) = \int_0^{\pi} \sin {\theta}_{n-1} d{\theta}_{n-1}\, T({\theta }_n,{\theta }_{n-1} )\,z_{n-1}(k,{\theta}_{n-1}) \label{recurfon} \ee In order to perform the integral over $ {\theta }_{n-1} $ we use the following discretization procedure. We divide the variation interval $ 0 \leq {\theta }_{n-1} \leq \pi $ into $ n_s $ segments $ { (s-1)\pi \over n_s } \leq {\theta }_{n-1} \leq { s\,\pi \over n_s } $. The integral over each segment is done with the standard Gauss method involving $ n_g $ abscissas and $ n_g $ attached weights. The integral over the full $ {\theta}_{n-1}$ interval is then approximated by a discrete weighted sum over $ d =n_s \, n_d $ points: $$ \int_0^{\pi} f( {\theta }_{n-1})\sin {\theta}_{n-1} d{\theta}_{n-1}= \sum_{i=1}^ {d} w_i \,\sin{\theta}_i f( {\theta}_i ) $$ With the above integration procedure each iteration step can be reduced to a linear mapping in a $ d $ dimension Euclidean space $ {\cal {E}}_d$. To each function $ z_n(\theta )$ is attached a vector $ {{\bf{Z}}}_n $ with $d$ components $ ({{\bf{Z}}}_n)_i= \sqrt{ w_i\, \sin {\theta}_i }\, z_n( {\theta }_i ) $. The additional factor has been introduced in such a way that the Hermitian norm $ \langle z_n \vert z_n \rangle $ (defined from our discretized integration procedure) coincides with the Euclidian norm $ {{\bf{Z}}}_n\cdot {{\bf{Z}}}_n $. The linear mapping in $ {\cal{E}}_d$ is written as: $ {{\bf{Z}}}_n = {\cal T } \, {{\bf{Z}}}_{n-1} $, where the elements of symmetric $ d \times d $ matrix $ {\cal T } $ are given by: $$ {( {\cal T } )}_{i , j} = \sqrt {w_i \, \sin {\theta}_i \,w_j \, \sin {\theta}_j } \, T(\alpha,k^2,{\theta }_i,{\theta }_j) $$ It is also easily verified that within our finite sum integration procedure we have: $ \langle z_n \vert \hat{ T}( \alpha,k^2) \vert z_{n-1} \rangle = {{\bf{Z}}}_n \cdot {\cal T}{{\bf{Z}}}_{n-1} $. In order to give a simple criterion of convergence, it is convenient to require that at each iteration step the vector $ {{\bf{Z}}}_n $ stay on the $ {\cal {E}}_d$ unit sphere. This is achieved by using the following non linear mapping: \be {{\bf{Z}}}_n ={ {\cal T } \, {{\bf{Z}}}_{n-1} \over \sqrt{ {{\bf{Z}}}_{n-1} \cdot {\cal T}^2 {{\bf{Z}}}_{n-1} } } \label{receucl} \ee The criterion of convergence is set up as follows : Let us define $ \Delta {{\bf{Z}}}_n={ {\bf{Z}}}_n-{{\bf{Z}}}_{n -\Delta n }$ where $ \Delta n $ is an integer, typically of the order of one hundred. ${ {\bf{Z}}}_n $ is considered to be an appropriate fixed point ${{\bf{Z}}}_f$ if $ \sqrt { \Delta {{\bf{Z}}}_n \cdot \Delta {{\bf{Z}}}_n } \leq \delta $. In practice we have taken $ \delta =10^{-4} $; this choice , together with $ \Delta n= 200 $, leads to a very tight convergence test. Knowing ${{\bf{Z}}}_f$ it is a straightforward matter to get the groundstate energy $ {\epsilon}_0(\alpha,k^2)$ and its partial derivative with respect to $ \alpha $ and $ k^2$: \bea {\epsilon}_0(\alpha,k^2) &=& -{A\over b}\ln t_0(\alpha,k^2)= - {A\over b}\ln\(( {{\bf{Z}}}_f \cdot {\cal T } \,{{\bf{Z}}}_f \)) \nonumber \\ { \partial \epsilon_0(\alpha,k^2) \over \partial\, k^2 } & = &- {A\over b \,t_0(\alpha, k^2) }\, {{\bf{Z}}}_f \cdot { \partial {\cal T }\over \partial\, k^2 } \,{{\bf{Z}}}_f \nonumber\\ { \partial \epsilon_0 (\alpha,k^2 ) \over \partial\, \alpha } & = & - {A\over b \,t_0(\alpha, k^2) }\, {{\bf{Z}}}_f \cdot { \partial {\cal T }\over \partial\, \alpha } {{\bf{Z}}}_{f} ={{\bf{Z}}}_{f} \cdot {\cal C }os \,{{\bf{Z}}}_{f} \label{formtrans} \eea where the ${\cal C }os$ matrix elements are given by: $ {\(({\cal C }os\))}_{i,j}= {\delta}_{i,j} \cos{\theta}_i $. It is also possible to get an explicit expression of the eigenfunction $ {\Psi}_0 ( \theta) $ from the knowledge of the fixed point vector ${{\bf{Z}}}_{f}$. The eigenfunction obeys the integral equation: \be {\Psi}_0 ( \theta) = {1\over t_0}\int_0^{\pi} T(\theta, {\theta}_1) {\Psi}_0 ( {\theta}_1) \sin {\theta}_1 d \,{\theta}_1 \label{wfinteq} \ee Using our finite sum method to perform the integral over ${\theta}_1$ in the r.h.s. of the above equation, $ {\Psi}_0 ( \theta) $ is obtained immediately in terms of the components $ Z_{f,i}$ of the fixed point vector: $$ {\Psi}_0 ( \theta) ={1\over t_0} \sum_i T(\theta, {\theta}_i) \sqrt{ w_i\, \sin {\theta}_i } \, Z_{f,i} $$ In our implementation of the method we have taken $ n_g =10$ . The use of higher values of $n_g$ may lead to overfitting. In contrast we can vary more freely the number of sectors $ n_s$. The ground state energy ${\epsilon}_0(\alpha,k^2)$ and its partial derivatives vary by less than one part per million when $ n_s $ goes from 2 to 8 with $ n_g =10 $. One has to go down to $n_g=8$ and $ n_s=2$ to see a variation of about $ 10^{-5}$. We have verified, using the high precision integration subroutines provided by the Mathematica software, that the above wave function satisfies, in typical cases, the eigenvalue integral equation (\ref{wfinteq}) to better than $ 10^{-9}$ with the choice $ n_s=4$ and $ n_g=10 $. The vector $ \Delta {{\bf{Z}}}_n={ {\bf{Z}}}_n-{{\bf{Z}}}_{n -\Delta n }$ introduced to set up the convergence criterion can be used to get the energy $ {\epsilon}_1 $ and the eigenfunction $ {\Psi}_1 ( \theta) $ of the first excited state with a fairly good accuracy, say better than $ 10^{-3} $. In particular the variation of the energy gap $ \Delta \epsilon= {\epsilon}_1 - {\epsilon}_0 $ versus $ \eta $ will shed some light upon the crossover phenomenon to be discussed in section \ref{tw_wr_pl}. \begin{center} \section{ Monte Carlo Simulations with the Discretized RLC model} \label{Monte_Carlo} \end{center} The Monte Carlo procedure allows to generate configurations of the discretized RLC with a frequency proportional to their Boltzmann weight. A full simulation of the discrete model, incorporating the self avoidance effects, the check that only unknotted configurations are kept, and the estimation of the Writhe with the non local Fuller formula was developed in \cite{volog} in the case of closed DNA chains. Vologodskii and Marko \cite{marvol} were the first to adapt this formalism to the case of a single open supercoiled chain. In order to facilitate the transition from closed chain to open chain, they assume that the chain is confined between two impenetrable walls, with the two free ends sitting on different walls. From an experimental point of view, one wall is certainly welcome since in the actual experiments one molecule end is achored by biochemical links to a glass plate. The other end is biochemically glued to a magnetic bead. In the experiments studied in this paper, the bead radius is one seventh of the molecule contour length L, so that an unpenetrable wall looks somewhat inadequate to account for the geometrical obstruction of the magnetic bead, even if one neglects the very slow processes where the molecule releases the supercoiling by turning around the bead. The authors were, of course, aware of the problem and thus they limited their comparison to experiment to relative extension larger than $ 0.3 $. This wall effect is visible in the limit of zero supercoiling limit where the M.C. results differs significantly from the WLC at forces below $.1 pN$ ( in that case finite size effects may also play a role since $L/A \sim 10$ to $20$). If the two conditions $ < z >/L > 0.3 $ and $ F > 0.1 pN $ are satisfied then a reasonable agreement was achieved with experiments over a rather broad range of forces and supercoiling. The agreement is particularly satisfactory for force extension curve at $ \sigma =0.031 $ and this has allowed the authors to give an estimate of the twist rigidity $ C $, with a $ 20 \% $ error bar . We shall return to this last point in section VII. Vologodskii and Marko have also investigated the effect of the reduced ioning strength which governs the DNA effective Coulomb radius, which can be characterized by the Debye length $\lambda_D$. Passing from 20 mM to 200 mM of NaCl leads to a Debye length reduction by factor $\sqrt{10}$. The calculated changes in the force extension curves at $ \sigma= 0.01 $ and $ \sigma= 0.03 $ are barely visible for stretching forces in the range $ 0.1 pN \leq F \leq 0.4 pN $. This includes the $ (F,\sigma )$ domain to be considered in the present paper. In the actual experiments, the Debye length corresponds to 30 mM of NaCl. Therefore the Vologodskii and Marko \cite{marvol} findings suggest that the self avoidance effects associated with the finite Coulomb radius play a relative minor role in the data to be analysed in the next section. They present some evidence for the importance of knotting supression in the particular case $ F=0.2 \,pN $ and $ \sigma =0.05 $. A typical simulated knotted configuration is seen to have the ability to absorb the supercoiling more efficiently than the corresponding unknotted one: it leaves the chain with a larger extension. Our analysis of the $ F=0.2 \, pN $ hat curve of Fig \ref{fig1} will concern the range $ \vert \sigma \vert \leq 0.016 $ which lies far away from the borderline point considered by the authors. In the work presented here, we have performed a much less ambitious Monte Carlo simulation. We kept within the RLC model without self-avoidance, and the simulation was used essentially as a guide to validate our model, using the local formula for $\chi_W$ together with a discrete version of the chain: we checked that it is free of pathology and able to reproduce the experimental findings. It suggested to us that the length cutoff $ b$ provided by the elementary link of the discretized chain could generate the angular cutoff needed to regularize the continuous RLC model, as it is shown in section IV.B. Our first computations within the regularized continuous model ( the method a) of subsection V.B.1 ) were checked on few points by Monte-Carlo simulations and the agreement gave us confidence in our approach. Subsequently we gave our preference to the transfer matrix iteration method of subsection V.B.2 which uses exactly the same theoretical inputs as the Monte Carlo simulation but leads to accurate results with much less computer time. The fully discrete model involving three Euler angles per elementary link can be simulated, but it is more efficient for our purpose to make use of the fact that the $\psi$ integrals can be done analytically, and thus to work only with the two angles $\theta$ and $\phi$ for each link. Indeed, using our computation of section 3 and integrating over the momentum $k$ conjugate to the supercoiling angle $\chi$ after integrating out the $\psi$ angles, the partition function of the supercoiled DNA molecule can be written as: \be Z(\chi,F)= \int {\cal D} (\theta,\phi) \exp\((- {E_{WLC} \over k_B T} -{C \over 2 L} \(( \chi-\int ds \; \dot \phi(1-\cos \theta) \))^2 \)) \ee This is the path integral which we have discretized and simulated. The discretization procedure is exactly the one described in the previous section. We use $N$ elementary links of length $b$, and the two discretization rules (\ref{rule1},\ref{rule2}) for the energy function. The path integral measure is substituted by $\prod_{n=1}^N d \phi_n d \cos \theta_n$. The partition function is thus expressed as a $2 N$ dimensional integral, and the corresponding probability measure can be sampled by Monte Carlo. We use the standard Metropolis algorithm, where a new configuration of the chain is proposed at each step, the corresponding change of energy $\delta E$ is computed, and the change is accepted with probability $min(1,\exp(-\delta E/k_BT))$. The point on which one must be careful is the choice of moves. Clearly a choice of local moves (changing $\theta_n,\phi_n$ on one link at a time) is a very bad one with which a macroscopic change of the molecule takes very many steps, and it leads to very long thermalization times. We have checked that it is in effect useless. We need to implement more sizeable moves, but moves which do not change the energy too much. One method which we have found rather effective is the following. We have first relaxed the constraint involving $\theta_N$, which should not change the extensive properties of the chain. Then the moves consist in: \begin{enumerate} \item One picks up one link of the chain, number $n$. \item One rotates the section of the chain $j\in {n+1,...,N}$ by a global rotation. The rotation, of angle $\gamma$ around an axis $\vec n$, is picked up at random, with a uniform distribution for the choice of $\vec n$ on the unit sphere, and a uniform distribution of $\gamma$ on an interval $\lbrack -\delta,\delta \rbrack $ (a value of $\delta$ of order .5 is generally adequate). \item One moves to the next link and iterates the procedure. \end{enumerate} This is one of the types of moves which are used in the standard simulations of supercoiled DNA \cite{volog}. We have simulated mostly chains of $300$ links, each link $b$ having a length of one tenth of the persistence length. We have checked that the finite size effects can be neglected with respect to the statistical errors for this value. For a given value of the supercoiling angle $\chi$ (or rather of its intensive version $\eta$), we perform a simulation, with a number of Monte Carlo steps per link of order $10^4$. The first third of data is used for thermalization, the rest is used for measuring the distribution of elongation. We show in figure \ref{fig3} the results for $<z>/L$, and the statistical error bars. The computations were done at $C/A=1.4$, and $b/A=.10$ for values of the force equal to $.116 \, pN$, and the amount of supercoiling given by $\eta=0.0, \,0.46, \,0.92, \,1.39$. We see that they are in rather good agreement with the experiment and the transfer matrix results. Note that since we do not have to use the two impenetrable walls trick, our computation is also valid in the zero supercoiling limit. When one increases the degree of supercoiling the thermalization time becomes prohibitive, and an examination of the configurations shows that they start building up some small fluctuating plectoneme-like structures. Clearly in order to study this regime one must first include self avoidance, and then incorporate moves which are able to shift the plectonemes positions efficiently, as was done in \cite{volog,marvol}. We believe that this type of approach can be pursued further in order to get a precise comparison with experiments in the strongly supercoiled regime. It is complementary to the more restrictive, but more analytical, study at small supercoiling which we develop here. \begin{center} \section{ Analysis of the experimental data on DNA extension versus supercoiling in the small force regime. } \label{ana_exp} \end{center} In this section we are going to analyse a limited set of data obtained by the LPS-ENS group on the same single DNA molecule. They consist of three extension versus supercoiling curves with values of the stretching forces $F= 0.116\, pN , 0.197 \, pN $ and $ 0.328 \, pN $. For these data a point-by-point evaluation of the systematic uncertainties is still lacking, so only the statistical uncertainties have been considered in the determination of the ratio $ { C\over A} $. A simplified version of this analysis has been already given by us in a previous publication \cite{BM1}. We have also performed two cuts upon the data in order to exclude from the analysis regions where the validity of RLC model is questionable. The first cut allows to neglect the effect of the plane onto which the DNA is attached. The second one allows to neglect the self avoidance of the chain. First we exclude values of the relative elongation such that $ {\langle z(L) \rangle \over L}\leq 0.1 $. Experimentally, one end of the molecule is attached to a plane which thus implements a constraint $z(s) >0$, for the whole chain. We shall see later on that when the reduced supercoiling parameter $ \eta$ increases from 0 to few units the probability distribution of $ \theta $ develops a peak near $ \theta= \pi $. Specially when $ 0\leq {\langle z(L) \rangle \over L} \leq 0.1 $ , it means that the RLC model is likely to generate configurations with $ { z(s) \over L} \leq 0$. We think that, for the experimental lengths $L$ under study, the regime where $ {\langle z(L) \rangle \over L} > 0.1 $ is such that typically, the whole chain is in the half plane $z>0$. The second cut excludes the too high values of the reduced supercoiling angle by imposing the condition $ \eta \leq 1.5 $ for $ F=0.197 pN $ and $ F=0.328 pN$. In a following section we shall present evidence that when $ F= 0.328 pN $ the RLC model generates plectoneme-like configurations above a critical value $ { \eta}_c \approx1$ . Our model ignores self avoiding effects which are present under the actual experimental working conditions since the DNA Coulombic potential is only partially screened. The experimental plectonemes must have a radius \footnote{The plectoneme' radius is the common radius of the two interwound superhelices} larger than the DNA Coulombic radius. In contrast, in the RLC model plectonemes with an arbitrary small radius can be generated. For a fixed variation of the supercoiling angle, the creation of a plectoneme, having a given length, absorbs, on the average a smaller fraction of the DNA chain, in comparison with a situation where self avoiding effects are involved. Indeed we have found that when the above constraints are not fulfilled the experimental points have a tendency to fall inside the theoretical hat curves, {\it i.e } for a given $ \eta $, experiment gives a smaller $ {\langle z(L) \rangle \over L} $. \begin{figure} \centerline{\epsfxsize=12cm \epsfysize=16cm \epsffile{dnaRLCfig2.ps} } \vspace{-4 cm} \caption{Empirical determination of the cut off length b and the ratio $ C/A $ from the hat curves analysis. In the case $ F=.116 \, pN $, we have plotted, versus the cutoff $b$, the mean ratio $ \langle C/A \rangle $ obtained by averaging the $ C/A$ values predicted by the RLC model from each empirical hat curve point. The error bar represents for each $b$ the variance $ {\sigma}_r $ which leads to a measure of the RLC model ability to reproduce the hat curve data. A remarkable agreement is achieved with $ b= .14 \,A$ while it becomes very poor for $ b=.06 \,A $. This is consistent with the RLC model singularity near $ b=0$. } \label{fig2} \end{figure} In order to extract the values of $C/A$ and of the cutoff $b$ from the experimental data, we have used the following technique. Using equation (\ref{etakappa}) the ratio $ C/A $ can be written as a function of the reduced supercoiling angle $ \eta$ and the torque $ \kappa $ : \be {A \over C} = {\eta \over \kappa} -2 {\partial \epsilon_0 \over \partial k^2}(\alpha,-\kappa^2) \label{AoverC} \ee With the help of interpolation techniques, one can invert equation (\ref{extens}) in order to get $ \kappa $ as a function of $ {\langle z(L)\rangle\over L} $. In this way, each " hat " curve point of coordinates $ ( \eta ,\, {\langle z(L) \rangle \over L} )$ is associated with an empirical value of the rigidities ratio $ {A\over C} $, once a choice of the cutoff length $ b $ has been made. If the RLC model is to give a good representation of the data there must exist a value of $b $ such that the empirical values $ C/A $ obtained from all points of all hat curves cluster around the correct result for $ C/A $. \begin{figure} \centerline{\epsfxsize=12cm \epsfysize=16cm \epsffile{dnaRLCfig3.ps} } \vspace{-2cm} \caption{ The elongation versus reduced supercoiling angle, $\eta =94.8 \,\sigma $, for forces $ F=.116, \,.197, \,.328 \, pN $, from bottom to top. The smaller points are the experimental results, the bigger points on the lowest curve are from Monte Carlo simulations and the full lines are the predictions of the RLC model obtained as indicated in the text. } \label{fig3} \end{figure} For F =0.116 \, pN, we have plotted on Figure \ref{fig2} the average ratio $ \langle C/A \rangle $ and the variance $ \sigma_r=\sqrt{ \langle \, ( C/A )^2\,\rangle - (\langle \, C/A \rangle\,)^2 }$ versus the cutoff length $b$. The average and variance are computed over all the 20 experimental points of the hat curve. The point with $ \eta=0 $ is an input, used to compute the dimensionless force parameter $\alpha $. Using the scaling properties of the model, we have reduced the data analysis to a two parameters fit: $ b/A $ and $ C/A $. Since our model reduces to the WLC model in the $ \eta=0 $ limit it is legitimate to use the persistence length $ A $ obtained within the WLC model from the analysis of the force versus extension curve \cite{bouc99}. As it is apparent on Fig. \ref{fig2}, the RLC model with $ b=0.14 \, A $ leads to a remarkably good agreement with the data. It corresponds to a minimum value of $0.03 $ of the quantity $ {\sigma_r \over \langle C/A \rangle } $ which measures the ability of our model to reproduce the data. The best cutoff length $ b $ value is approximately equal to twice the double helix pitch. This is close to the length resolution $\Delta l $, which we had to invoke in order to justify the assumption of a cylindrically symmetric rigidity tensor. It is interesting also to note that the average value $ \langle C/A \rangle $ { \it varies slowly with } $b$. The variance $ \sigma_r $ increases rapidly if one goes to small values of the cut-off length ; this is consistent with the fact that the RLC model becomes singular in the limit $ b \rightarrow 0$. Similar results, somewhat less precise, have been obtained for the two other values of $ F $ considered in this section. They favour the same value of $ b/A $ and give values of $ \langle C/A \rangle $ consistent with the ones obtained for $ F= .116 \, pN$, with variances $ \sigma_r $ about three times larger. Performing a weighted average upon the whole set of $ C/A $ empirical values obtained by the three hat curves, taking $ b=0.14 \,A$, one gets the following empirical determination of the ratio of the two elastic rigidities involved in the RLC model : \be {C\over A}= 1.64 \pm 0.04 \ee This result is in agreement with the value given in our previous work \cite{BM1}: $ {C\over A}\simeq 1.68 $. Since we have used data which do not incorporate in a quantitative way the systematic uncertainties, the above number should be considered as somewhat preliminary, but as it stands, it constitutes a significative improvement upon the previous empirical estimates. We give in Figure \ref{fig3} the theoretical " hat " curves using the ratio $ {C \over A }= \langle C/A \rangle $ obtained from the data at each force and the cutoff favored value $ b= 0.14 \, A $. The agreement with the experimental data is very satisfactory and confirms the overall consistency of the procedure. It should be stressed that the rather sharp bend connecting a slow quadratic decrease to a steep nearly linear falling down, which is observed for $ \eta \simeq 1 $ on the $ F=.33 \, pN$ hat curve, is rather well reproduced by our model. Using the values of the persistence length $A$ obtained from the force versus extension curve measured on the same single molecule, one can obtain the following empirical value of the twist rigidity: \be C= 84 \pm 10\,nm \label{Cempir} \ee The statistical error on $ C $ is of the order of 2 $ nm $ but the systematic errors are expected to be non negligible and the overall uncertainty of $ 10 \, nm $ which accounts for those looks reasonable. It is here of some interest to quote a recent empirical determination of the twist rigidity obtained from an analysis which does not involve the use of the RLC model or the like. A detailed account of the procedure is to be found in reference \cite{houch98}. We shall here only quote the result: $ C=86 \pm 10 \, nm$, which is in good agreement with the number given by equation (\ref{Cempir}). Vologodskii and Marko \cite{marvol} have given previously an estimate of $C$ obtained from one force extension curve at $ \sigma =0.031 $, where the agreement with their Monte Carlo simulations is particularly striking : $ C= 75 \pm 15 \, nm $. This result agrees with ours. Unfortunately they were not able to get any value of $C$, leading to an overall good fit for the complete set of force extension curves. Moroz and Nelson \cite{mornel1,mornel2} have analyzed the "hat" curves, using a perturbation approach to the RLC model. Because of the constraint $ K= \sqrt{\alpha - {1\over 4} {\kappa}^2 } > 3 $, inherent to their method, they explore the force supercoiling angle domain given by $ F \geq .3 \, pN $ and $ \sigma \leq 0.01 $. This implies that the ranges of relative extension for a given force are very narrow : for the top curve of Fig(\ref{fig3}) ($ F= .33 \, pN $): $ 0 .72 \leq \langle z \rangle/L \leq 0.75 $, the higher the force the narrower is the interval and closer to unity is its center. It is clear from Fig \ref{fig3} that the range of relative extension analyzed in the present paper is much wider. In fact there is a rather small overlap of the experimental domains involved in the two analysis : the intersection of the two data set amounts only to 20 $ \% $ of our present data set. Moroz and Nelson have incorporated hat curves data associated with relatively high force values: $ 0.6, \,0.8, \,1.3, \,8.0 \,pN $. For such forces, the RLC model is certainly not valid when $ \sigma <0 $ and even for $ \sigma >0 $ in the case of the highest force. They \cite{mornel2} have derived from their two parameter fit $ ( A,C ) $ the twist rigidity value $ C=109 \, nm $ (in a preliminary analysis \cite{mornel1} based upon a more limited set of data they gave $ C=120 \, nm $). These authors did not give the uncertainty associated with $ C $ and the value of $ A $ coming out from their two parameter fit. Because of the limited range of relative extension values they can fit, we believe that the error bars on their results should be larger than ours. As a comparison, we have used $ C= 109 \, nm $ to compute the theoretical hat curves of Fig. \ref{fig3}. As expected, the quality of the fit becomes significantly poorer: the overall $\chi_2$ is multiplied by 5.3 when one goes from $C=86 \, nm $ to $ C= 109 \,nm $. \section{Twist, Writhe and Plectonemes.} \label{tw_wr_pl} In this section we would like to use our solution of the discretized RLC model to study, at a given force (F=.33 \, pN), the variations of the torque, twist and writhe thermal averages versus the supercoiling reduced angle $ \eta $. As we shall see, there exist two very different regimes. Below a rather sharply defined value of the supercoiling angle $ {\eta}_c $, the DNA chain behaves nearly as an elastic non-flexible rod. Above $ {\eta}_c $ the torque becomes nearly independent of supercoiling while the writhe grows linearly with it. We shall give arguments suggesting that in the high $ \eta $ regime the supercoiling constraint is satisfied by the creation of plectoneme-like configurations. \subsection{ Relation between the torque and the average twist } To begin we are going to prove that the thermal average of the rod twist $ \langle \,T_{w} \, \rangle $ is given in terms of the torque $\Gamma$ by the classical elasticity formula for a non-flexible rod: \be \langle \,T_{w} \, \rangle = { \kappa \, L \over C }= { \Gamma \, L \over \ {\cal C}} \label{twistav} \ee where $ {\cal C}= k_{B} \,T \, C $ is the usual twist rigidity. It is convenient to introduce the joint probability distribution of a DNA chain configuration $ {\cal P}( \chi_1 ,\chi_2)$ with $ T_W= \chi_1 $ and $ {\chi}_W= \chi_2 $; it can be written as the double Fourier transform: $$ {\cal P}( \chi_1 ,\chi_2)\propto Z( \chi_1, \chi_2)= \int { dk_1\,dk_2 \over 4 \pi^2} \exp(-ik_1 \, \chi_1-ik_2 \, \chi_2 ) \tilde Z(k_1,k_2) $$ where $ \tilde Z(k_1,k_2) $ is given by the functional integrals \begin{eqnarray*} \tilde Z(k_1,k_2) & = & \int { \cal D }(\theta,\phi) \,\exp \(( i\,k_2\, \chi_W- {{\cal E}_{bend}+{\cal E}_{stretch} \over k_{B}T} \)) {\tilde Z}_{T}(k_1) \\ {\tilde Z}_{T}(k_1) & = & \int { \cal D }(\psi) \,\exp \(( i\, k_1\, T_{w} -{{\cal E}_{twist} \over k_B T}\)) \end{eqnarray*} As in section III B, we perform explicitly the functional integral upon $ \psi $ and get $ {\tilde Z}_{T}(k_1) = \exp( -{k_1^2L\over 2 C}) $. This leads to the factorization property: $ \tilde Z(k_1,k_2) = {\tilde Z}_{T}(k_1) {\tilde Z}_{W}(k_2 ) $, where the writhe term ${\tilde Z}_{W}(k_2 )$ is given by the path integral: $$ {\tilde Z}_{W}(k)= \int { \cal D }(\theta,\phi) \, \exp \((i\,k\, \chi_W- {E_{WLC}\over k_{B}T} \)) $$ { \it As a physical consequence, the twist $ T_{w}$ and the writhe $ \chi_W $ fluctuate independently in the RLC model, when no supercoiling constraint is applied upon the free end of the molecule. } The thermal average $ \langle \,T_{w} \, \rangle $, in presence of the supercoiling constraint $ \chi= T_{w}+\chi_W $, is then given by: \bea \langle \,T_{w} \, \rangle &=& { 1\over Z(\chi,F)} \int d\chi_1 \, d\chi_2 \, \chi_1 \, Z_{T}( \chi_1 )\, Z_{W}( \chi_2 )\, \delta ( \chi_1+ \chi_2-\chi ) \nonumber \\ &=& {1 \over Z(\chi,F)} \int {dk \over 2 \pi} \((-i {d {\tilde Z}_T \over dk } \)) {\tilde Z}_{W}(k) \exp(-i k\, \chi ) \nonumber \eea Using the explicit form of $\tilde Z_T$, one gets the announced formula and by subtraction the writhe thermal average : \bea \langle \,T_{w} \, \rangle &= &- {L \over C} {\partial \ln Z \over \partial \chi}= {L \over C} \, { \Gamma \over k_{B} \,T } \nonumber \\ \langle \,\chi_{W} \, \rangle &=& \chi- \langle \,T_{w} \, \rangle = {2\,L\,\kappa \over A } \, { \partial \epsilon_0(\alpha,-\kappa^2) \over \partial\, k^2 } \eea \subsection{The average writhe in the zero stretching force limit: a comparison with Monte Carlo simulations of self-avoiding closed supercoiled DNA chains } The above formulas offer the opportunity to compare, on a specific point, the RLC model predictions with closed DNA chain Monte-Carlo simulations \cite{volog}, which incorporate self-avoiding effects. Using the same ratio $ C/A=1.5 $ and the same value of $ b/A=.2 $ as Vologodskii et al. \cite{volog}, we have computed $ \lim_{F=0}\, \langle \,\chi_{W} \, \rangle / \chi =< Wr/\Delta Lk > $ for $\sigma $ values taken in the range: $ 0 \le -\sigma \le 0.04 $. The comparison with the numbers taken from reference \cite{volog} is displayed in Figure \ref{figcompvog}. It is apparent that the RLC model computations, and Monte-Carlo simulations are in rather good agreement: when $ \vert \sigma \vert \leq 0.02 $ - the range explored in our previous data analysis - the results diverge by less than $ 8.5 \% $ and the difference reaches $ 10 \% $ when $ \vert \sigma \vert \rightarrow 0.04 $. They both agree rather well with the measurements performed in references \cite{boles} and \cite{adrian}. If we forget about the possible finite size effects in the Monte-Carlo simulations where $A/L \simeq 1/12$, the small deviation may tentatively be attributed to the self-avoiding effects not incorporated in our computations. \begin{figure} \vspace{-2cm} \centerline{\epsfxsize=12cm \epsfysize=16cm \epsffile{dnacompvog.ps} } \vspace{-6cm} \caption{Comparison of Monte Carlo ({\Large $\bullet$}) and the RLC model ( continuous line) results for the reduced average $ < Wr/\Delta Lk > = \lim_{F=0}\, \langle \,\chi_{W} \, \rangle / \chi$. } \label{figcompvog} \end{figure} \subsection{ A torque and writhe versus supercoiling cross-over : a possible sign for thermal excitation of plectoneme-like configurations} In Figure \ref{figplecto} we have plotted ( for the case $ F=.33 \, p N $) the torque in $ k_{B} \,T $ unit, $ \kappa $ , together with the ratio of writhe to twist $ \chi_W / {T}_{w} $ as functions of the scaled supercoiling variable $ \eta= {\chi \, A \over L} $. These theoretical curves have been obtained with the same parameters as the "hat" curves of Figure \ref{fig3}. They show a very rapid change of behaviour, a quasi-transition, for $ {\eta}_c \approx 1.0$, with the following two very different regimes: \begin{itemize} \item Below ${\eta}_c $ the twist of the DNA chain increases linearly with the supercoiling $\eta$, as in a non flexible rod with an effective twist rigidity $ C_{eff}=.82\, C$. The ratio of writhe to twist stays almost constant at the value $0.2$. \item Above ${\eta}_c $ the torque depends weakly on the supercoiling $\eta$, the twist becomes nearly constant while the writhe increases linearly with the supercoiling. \end{itemize} The behaviour in the large supercoiling regime is reminiscent of the mechanical instability leading to the formation of plectonemes which is easily observed by manipulating macroscopic elastic rods such as telephone cords. One must be careful with this analogy because the plectonemic instability corresponds to a zero temperature limit, while in the case of DNA, much of the elasticity comes from entropic effects and thermal fluctuations play a crucial role. Yet we would like to point towards the existence, in the large supercoiling regime of the RLC, of excitations which share some of the properties of plectonemes. We define $ {\cal E}_p $ as a set of undeformed plectonemes having the axis of their winding {\it helices arbitrarily oriented with respect to the force direction } ( $z-$ axis ). We then introduce the angular distribution $ P( \theta ) $ of the tangent vector $\hat{{\bf{t}}}$ about $z-$ axis, averaged along the plectoneme and upon the set $ {\cal E}_p $. Let us sketch the proof of the symmetry relation: $ P( \theta ) = P(\pi - \theta ) $. Introducing the angular distribution $ {\cal P} ( {\alpha}_p ) $ of the plectonemes axis and the tangent vector $ {\hat{{\bf{t}}} }_{plect}(s) $ relative to a vertical plectoneme, $ P( \theta )$ reads as follows : $$ P( \theta )= \int_0^{\pi} {\cal P} ( {\alpha}_p) d {\alpha}_p {1 \over L} \int_0^{L} ds \,\delta \(( \cos \theta - \hat{{\bf{z}}} \cdot R\(( \hat{{\bf{y}}},{\alpha}_p \)) {\hat{{\bf{t}}} }_{plect}(s)\)) $$ Ignoring the plectoneme handle contribution, we can rewrite $ P( \theta ) $ in terms of the vertical solenoid $\phi$-dependent tangent vector $ {\hat{{\bf{t}}} }_{sol}(\phi,{\theta}_0)= R( \hat{{\bf{z}}},\phi ) R( \hat{{\bf{y}}},{\theta}_0 )\hat{{\bf{z}}} $ : $$ P( \theta )= \int_0^{\pi} {\cal P} ( {\alpha}_p) d {\alpha}_p {1 \over 4 n_0 \pi} \int_0^{ 2 n_0 \pi } d \phi \(( \delta ( \cos \theta - \hat{{\bf{u}}} ({\alpha}_p)\cdot {\hat{{\bf{t}}} }_{sol}(\phi,{\theta}_0) ) + (\phi \rightarrow \pi-\phi, {\theta}_0 \rightarrow \pi - {\theta}_0 ) \)) $$ where we have used the idendity $ {\bf{A}}\cdot R {\bf{B}}= R^{-1}{\bf{A}} \cdot {\bf{B}} $ and defined the unit vector $ \hat{{\bf{u}}} ({\alpha}_p) = R( \hat{{\bf{y}}},-{\alpha}_p ) \hat{{\bf{z}}} $. Then we compute: $\hat{{\bf{u}}} ({\alpha}_p)\cdot {\hat{{\bf{t}}} }_{sol}(\phi,{\theta}_0)= \sin{\alpha}_p \cos \phi \sin {\theta}_0 + \cos {\alpha}_p \cos {\theta}_0 $ and by simple inspection, we arrive to the desired relation. Because of the $ \phi$ averaging, the proof holds true if one takes an arbitrary axis in the $ z= 0$ plane instead of the $ y $ axis to rotate the plectoneme. In practice the plectonemes are deformed by the thermal Brownian motion. More complex structures can also appear, like branched plectonemes. It looks however reasonable to assume that the above symmetry property is not affected, on average, by thermal fluctuations. In order to measure the degree of symmetry of the distribution of $\theta$ angle along the chain, we introduce the function $ plecto( \eta) $ defined as follows: \be plecto( \eta)= {\int_0 ^{\pi} P( \theta ,\eta) P( \pi- \theta ,\eta) d(\cos\theta) \over \int_0 ^{\pi} { P( \theta ,\eta) }^2 d(\cos\theta) } \label{plecto} \ee It is clear that $ plecto( \eta) $ reduces to unity for pure plectonemic configurations and goes to zero in the limit of rectilinear chains. Within the RLC model, the probablilty $P( \theta,\eta ) $ is given by the thermal average of the molecular axis angular distribution when one runs along the molecular chain. Exploiting the quantum mechanics analogy, it is easily proved that $ P( \theta,\eta ) $ is equal to $ {\Psi}_0(\theta)^2$, the square of the ground state wave function introduced in subsection V.B. This has been used to compute the function {\it plecto}$(\eta)$ which is plotted in figure (\ref{figplecto}). The shape of $ P( \theta ,\eta ) $ changes in a very characteristic way when ones goes from $ \eta=0 $ to $\eta \approx 4$. When $ 0 \leq \, \eta \leq {\eta}_c \approx 1$, $ P( \theta ,\eta ) $ has a rather narrow peak at $ \theta=0 $ with a nearly vanishing tail for $ \theta > {\pi \over 2} $. As a consequence, the function $ plecto(\eta )$ is practically null within this interval. A secondary peak at $ \theta= \pi $ begins to develop when $ \eta \geq {\eta}_c $ and it reaches about the same height as the primary peak at ${\eta} \approx 4 $. The building of this two bumps structure is accompanied by an almost linear increase of the function $ plecto(\eta )$ which reaches the value $ 0.9 $ near $ \eta= 4 $. This behaviour suggests that thermally deformed plectoneme-like configurations are responsible for the sharp increase of the writhe-to-twist ratio and the flattening of the curve torque versus supercoiling above the critical value ${\eta}_c \approx 1$. Finally a useful piece of information about the quasi-transition near $ \eta \geq {\eta}_c $ is the study of the variation of the energy gap between the groundstate and the first excited state $ \Delta\epsilon= {\epsilon}_1 - {\epsilon}_0$, obtained by the method given in subsection V.B.2. The corresponding curve appears in Figure \ref{figplecto} under the label "Gap". $ \Delta \epsilon $ is a decreasing function of $ \eta$, with a rather sudden fall near $ \eta \approx{\eta}_c $. This somewhat technical feature has a nice physical interpretation in terms of the correlation length associated with the $ \cos\theta $ fluctuations, {\it at fixed torque }, which is given by $ A/\Delta \epsilon$. The jump of this correlation length around $\eta_c$, is a further sign of a fast change of physical regime. \begin{figure} \centerline{\epsfxsize=14cm \epsfysize=18cm \epsffile{dnaRLCfigplecto.ps} } \vspace{-6cm} \caption{Twist, Writhe and Plectonemes in the RLC Model. We display several curves which indicate that a cross-over phenomena is taking place near $ \eta_c \approx 1$. After a linear increase, expected for a non-flexible rod, the torque becomes independent of the supercoiling above $ \eta_c $. Simultaneously, the writhe to twist ratio, which was staying constant around $ .2$, starts a rather steep linear increase. Near $ \eta_c $ the function $ plecto(\eta)$, which measures the forward backward symmetry of the $\hat{{\bf{t}}}$ angular distribution relative to $ {\bf{F}}$, takes off from 0 to the value 0.9.( $ plecto=1$ for a plectonemic configuration). All these features suggest that the cross-over could be attributed to the creation of plectoneme-like configurations. } \label{figplecto} \end{figure} The two regimes of supercoiling displayed on Figure \ref{figplecto} are related, within the quantum mechanical formalism of sect. V.B.1, to the double well structure of the potential $ V_{r} ( \alpha,k^2,\theta )$ when $ k^2 =-{\kappa}^2$. On one hand, the potential has a relatively shallow well at $\theta =0 $, associated with the stretching potential energy. On the other hand, the regularized "writhe" potential produces a deep narrow hole at $\theta =\pi $. Let us call $ {\epsilon}_a, \Psi_a $ and $ {\epsilon}_b, \Psi_b $ the eigenvalues and eigenstates associated respectively with the semi-classical states localized in the two potential wells. As $\kappa $ is increasing the energy levels $ {\epsilon}_a$ and $ {\epsilon}_b $ are approaching each other. The level crossing is avoided near $\kappa_c $ by a quantum tunneling through the finite height wall between the two wells. This accounts for the peculiar variation of $ \Delta \epsilon $ near $\eta_c$. The mixing of the two levels $ \Psi_a $ and $ \Psi_b $ by quantum tunnelling generates the secondary bump in the probability distribution $ P(\theta )$. In the vicinity of the near crossing point, the groundstate energy ${\epsilon}_{0}$ depends almost linearly upon the tunneling amplitude $ t_{ab}$, which has typically a very rapid variation with $ \kappa^2 $. Since the ratio $ \eta / \kappa $ is a linear function of ${\partial \epsilon_0 \over \partial\kappa^2}$ ( see equation (\ref{etakappa}) ), $ \eta $ should exhibit a very sharp increase with $\kappa \ge \kappa_c \approx 1.4$ as is easily seen by turning Figure \ref{figplecto} by 90 degrees. \begin{center} \section{ Fluctuations of the Extension in a Supercoiled DNA Molecule} \label{fluctuations} \end{center} In this section we would like to give a brief analysis of preliminary measurements of the supercoiled DNA extension fluctuations, performed at a force $ F=0.33 \, pN $ by the experimental group in the ENS. We shall compare the experimental results with the $ RLC $ model predictions using the parameters $ b/A $ and $ C/A $ deduced from the extension versus supercoiling curves analysis, performed in section \ref{ana_exp}. The mean square deviations of the fluctuations of the molecule extension along the force direction is given by the second derivative of the free energy $ {\cal F}= -k_B T \log Z(F) $ with respect to F: \begin{equation} \langle \, \delta\,z ^2 \, \rangle = \langle \,z^2 \, \rangle- {\langle \,z \, \rangle }^2 = (k_B \, T)^2 { {\partial}^2 \log Z \over \partial F^2 } \end{equation} It is convenient to replace the first derivative $ { \partial \log Z \over \partial F }$ by its expression in terms of the average relative extension using equation (\ref{extens}). The extension mean square fluctuation can then be written as follows: \be \langle \, \delta\,z ^2 \, \rangle = L \(( { k_B T \over F}\)) \, \alpha \, { \partial \over \partial \alpha } \(( { \langle \,z \, \rangle \over L } \)) \label{fluctzkappa} \ee We have verified that the value of $ \langle \, \delta\,z ^2 \, \rangle $ obtained in the discretized RLC model in the limit of no supercoiling, $\chi=0$, agrees with the one given by the WLC model to better than $ 3 \, 10^{-3} $. In the situation where the free end of the molecular chain is subjected to a supercoiling constraint the fluctuations will be different depending on whether the measurement is performed at fixed torque $\kappa$ or at fixed supercoiling angle $\chi$. In the first case the result is readily obtained from equation (\ref{extens}) : \be \langle \, \delta\,z ^2 \, \rangle {\vert}_{\kappa} = - L \(( { k_B T \over F}\)) \,{ \alpha \, {\partial}^2 {\epsilon}_0(\alpha,-\kappa^2)\over \partial {\alpha }^2 } \ee In the actual experiment the extension fluctuations are measured at fixed $ \chi $ and an extra term has to be added to the above expression due to the fact that the torque is now a function of $ \alpha $ and $ \eta $: \be \langle \, \delta\,z ^2 \, \rangle {\vert}_{\eta} = L \(( { k_B T \over F}\))\, \alpha \,\(( - { {\partial}^2 {\epsilon}_0(\alpha,-\kappa^2)\over \partial {\alpha }^2 } + 2 \kappa {\partial \kappa \over \partial \alpha } { {\partial}^2 {\epsilon}_0(\alpha,-\kappa^2)\over \partial \alpha \partial k^2 } \)) \ee Using equation (\ref{etakappa}), the extra term can be transformed by writing:\\ $ 2 {\partial \epsilon_0 \over \partial k^2}(\alpha,-\kappa^2)={\eta \over \kappa} -A/C $. One gets finally the mean square extension fluctuations at fixed supercoiling angle: \be \langle \, \delta\,z ^2 \, \rangle{\vert}_{\eta} = \langle \, \delta\,z ^2 \, \rangle {\vert}_{\kappa}- L \(( { k_B T \over F }\)) \, \alpha \,{\eta \over \kappa } { \(( { \partial \kappa \over \partial \alpha } \)) }^2 \label{fluctzeta} \ee In the above formula the second term is clearly negative so we expect that $ \langle \, \delta\,z ^2 \, \rangle{\vert}_{\eta} < \langle \, \delta\,z ^2 \, \rangle {\vert}_{\kappa}$. The curves giving $ \langle \, \delta\,z ^2 \, \rangle $ versus the number of supercoils $ n = {L \eta \over 2 \pi A } \approx 50 \,\eta $ are displayed on Figure \ref{figfluct}, $ \langle \, \delta\,z ^2 \, \rangle{\vert}_{\eta} $ as a thick continuous line and $ \langle \, \delta\,z ^2 \, \rangle{\vert}_{\kappa} $ as a dashed line. It appears clearly that $ \langle \, \delta\,z ^2 \, \rangle{\vert}_{\eta} \ll \langle \, \delta\,z ^2 \, \rangle {\vert}_{\kappa} $, when $ |\eta |> {\eta}_c \approx 1 $. This implies a strong cancellation between the two terms of formula (\ref{fluctzeta}), which is then not suitable for an evaluation of $ \langle \, \delta\,z ^2 \, \rangle{\vert}_{\eta} $. To get the thick solid curve of Figure \ref{figfluct}, we have used the fact that the calculation procedure giving, at reduced force $ \alpha$, the relative extension versus the supercoiling angle, $ { \langle \,z \, \rangle \over L } (\alpha ,\eta) $, is precise enough to allow the evaluation of the partial derivative with respect to $\alpha$ by a three points finite difference formula. The difference between the two statistical ensembles can be understood qualitatively from the previous section considerations and specially by looking at the curves of Figure \ref{figplecto}. In a situation where the torque is fixed at $ \kappa = {\kappa}_c \approx 1.4 $ the supercoiling angle $ \chi $ is practically unconstrained. It can fluctuate rather freely through thermal excitation of plectonemes, which have no effect upon the torque. In contrast when $ \eta $ is fixed at a value $ \eta > {\eta}_c \approx 1 $ the ensemble is more constrained. The torque $ \kappa $ is still practically fixed at the critical value $ {\kappa}_c $. As a consequence, the writhe $ \chi_W= \chi - T_w \approx \chi - {\kappa}_c L/C $ has approximately a fixed value and then the only plectonemes which can be created are those having a very limited range of writhe. \begin{figure} \centerline{\epsfxsize=8cm \epsfysize=14cm \epsffile{dnaRLCfigfluct.ps} } \caption{Mean square fluctuations of DNA supercoiled molecule extension for $F=0.33 \, pN$. The points are the experimental data. The thick continuous line is the theory, for the experimental situation where the supercoiling turn number $ n = {L \eta \over 2 \pi A } \approx 50 \,\eta $ is fixed. The dotted line is the theoretical result for another situation where the torque would be fixed (the number of turns is then to be understood as a thermal average value). } \label{figfluct} \end{figure} As is apparent on Figure \ref{figfluct}, the experimental points agree reasonably well with the RLC model predictions, taking into account the quoted uncertainties, which are exclusively of statistical origin. ( Systematic effects are smaller). We notice the quasi-critical jump of $ { \langle \,z \, \rangle \over L } (\alpha ,\eta) $, near $ n_c \approx 50 $, which is also present in the data. \section{Summary and Conclusion} \label{conclusion} Together with new relevant contributions, this paper gives a detailed account of a work of the authors, which appeared previously under a very concise letter form \cite{BM1}. It was a rather successful attempt to describe the entropic elasticity of supercoiled single DNA molecule in terms of the thermal fluctuations of an elastic rod. As in most works based upon a similar model, the rigidity tensor was assumed to be symmetric under rotations about the molecular axis and then fully described by two elastic constants, the bending and twist rigidities. Because of the DNA helical structure, one may have expected an axially asymmetric rigidity tensor, involving extra elastic constants. We have proved in this paper that axial symmetry breaking contributions to the elastic energy are averaged out upon a " coarse graining " involving a length resolution $ \Delta l $ about three times the double helix pitch $ p $. Such a value corresponds to the actual experimental resolution. The model developed in the present paper is not expected to be realistic at length scale below twice the DNA double helix pitch. To implement in the RLC partition function the supercoiling empirical constraint $\chi$, we have followed a direct procedure, instead of adapting to open molecular DNA chains the formalism used in the study of supercoiled plasmids. Imposing a well defined regularity condition upon the Euler angles describing locally the molecular chain, we have been able to decompose the empirical supercoiling angle $\chi $ as a sum of two line integrals $ T_w $ and $\chi_W $, associated respectively with the twist and writhe contributions. The local writhe $\chi_W $ is a line integral taken along regular trajectories of the chain unit tangent vector $\hat{{\bf{t}}}$. They are drawn upon the unit sphere $ S_2 $ pierced by a hole having a small radius $\epsilon$, localized at the south pole defined by the direction of the stretching force direction. Such a prescription forbids the crossing of a spherical coordinate singularity which would invalidate our derivation of the local writhe formula. When it is applied in compliance with the above prescription, the local writhe formula allows correct evaluations of the writhe of solenoid and plectoneme configurations, having an arbitrarily space orientation. More generally the local writhe formula, though not explicitly rotation invariant, is shown to lead, in the small $\epsilon$ limit, to rotation invariant results. The partition function of the RLC having a given supercoiling angle $\chi $ is written as the convolution product of the partition function $ Z_T( T_w) $ of a twisted non flexible rod times the partition function $ Z_W( \chi_W ) $ of a worm like chain having a fixed writhe angle. The Fourier transform $ \tilde Z_W(k) $ is shown to be the analytic continuation to imaginary time of the Feynman amplitude describing the quantum evolution of unit charge particle moving upon a sphere upon the joint action of an electric dc field and a magnetic monopole having an unquantized charge $k$. The associated Hamiltonian $ {\hat H}_{RLC}(k) $ is readily obtained from standard Quantum Mechanics rules. As in the quantum magnetic monopole problem, we have found a pathology which is associated with the singular behaviour of the $ {\hat H}_{RLC}(k) $ potential term near the south pole. We have proved that an angular cutoff is generated by a discretization of the chain involving an elementary link $b$ ( this result was given without proof in our previous work \cite{BM1}). We have derived analytically from the discretized elastic energy a regularized version of RLC continuous model. The singular writhe potential is multiplied by a regulating function going smoothly to zero when $\sin^2\theta \leq {b\over A}$. We have arrived in this way to a well behaved Hamiltonian $ {\hat H}_{RLC}^{r}(k)$, which provided a precise mathematical definition of RLC model used in this paper. It leads to a partition function free of any pathology. In an independent work following ours\cite{BM1}, Moroz and Nelson proposed a high force perturbation method \cite{mornel1} where no angular cutoff is introduced from the start. To finite order, the divergent perturbation series, hopefully asymptotic, does not see the singularity at the south pole, though it is always present in the model. In order to make contact with experiment, the authors have to impose upon their expansion parameter a sharp " technical " bound. Its effect is to restrict considerably the domain of force and supercoiling where a comparison with experiment is possible. This limits the precision of their determination of $C$. Within our regularized RLC model, we have developed methods allowing the computation, to an arbitrary precision, of the relative extension versus supercoiling curves, the so called " hat " curves. The Fourier transform involved in the partition function is evaluated by the saddle point method, in the limit of a contour length $L$ much larger than the persistence length $A$. It leads to a parametric representation of the hat curves in terms of the torque acting upon the molecule's free end. The final theoretical ingredient is the ground state energy of the regulated RLC Hamiltonian. It is obtained following two methods: {\it a) } an explicit solution of the Schr\"odinger equation associated with $ {\hat H}_{RLC}^{r}(k) $, {\it b) } the iteration of the transfer matrix deduced directly from the discretized elastic energy. Though they are not strictly equivalent from a mathematical point, they lead to identical results to few $\%$. Our data analysis involves three values of the stretching forces: $F=0.116, \,0.197, \,0.328\,pN$. For each hat curve point and a fixed ${b\over A}$, the RLC model leads to an empirical value of the ratio $ {C \over A}$ as function of the measured relative extension and supercoiling angle. If the model provides an adequate description of the hat curve, the set of empirical values should cluster around the actual value $ {C \over A}$. For each cluster we have plotted the mean value $ \langle {C \over A} \rangle$ and the variance $ \sigma_r $ versus $ {b\over A}$. The best value $ {b \over A}=0.14$ corresponds to the minimum of the ratio $ \sigma_r/\langle {C \over A} \rangle $, which measures the ability of the model to fit the data. The preferred cutoff length $b$, is found to be about two times the double helix pitch and it is close to the length resolution $ \Delta l $ invoked to suppress axially asymmetric elastic energy terms. Our determination of ${C \over A}= 1.64 \pm 0.04 $ is obtained from a weighted average of $ \langle {C \over A} \rangle $ relative to the three forces involved in the fit, taking $ {b \over A}=0.14$. This number turns out to be remarkably stable under variations of $b$ within the range $ 0.08 A \leq b \leq 0.2 A$: the deviations of ${C \over A}$ from 1.64 stay below the level of $ 8 \% $. In contrast, the quality of the hat curves fit which is very satisfactory at the best $b$ value becomes poorer away from it, specially for low $b$ values. Our central value for the twist rigidity $ C=84 \pm 10 \,nm$ differs by about $ 25 \% $ from the ones given by Moroz and Nelson \cite{mornel2}. The absence of any uncertainty estimate and an analyzed data set having a small overlap with ours makes the physical significance of this apparent discrepancy difficult to assess. Although these quantities are not yet directly accessible to experiment, we have computed, as functions of the supercoiling $\chi $, the torque $ \Gamma $ and the average twist $ \langle T_w \rangle$ (or equivalently the average writhe $ \langle \chi_W \rangle= \chi- \langle T_w \rangle$). Use has been made of a remarkable property: $ \langle T_w \rangle$ is related to $ \Gamma $ by the linear elasticity formula for a non-flexible twisted rod. In the case of the highest force ($ F= 0.33 \,pN $), the two curves, torque and writhe to twist ratio versus supercoiling, exhibit a rather sharp change of regime near the same critical value $ \chi_c $ of the supercoiling angle: the torque, after a nearly linear increase, becomes almost supercoiling independent while the writhe to twist ratio, initialy confined to the $ 20 \% $ level develops a fast linear increase. This behaviour is reminiscent of the buckling instability of a twisted rubber tube associated with the creation of plectonemes, able to absorb supercoiling at constant torque. We have shown that the configurations excited in the RLC model above the critical value $ \chi_c $ share a simple global symmetry property with a set of undeformed plectonemes arbitrarily oriented with respect to the force direction $z$ axis: the angular distribution of the tangent vector $ \hat{{\bf{t}}} $ has a forward backward symmetry with respect to the $z$ axis. The existence of this rather sharp cross over has been confirmed by an analysis of the extension fluctuations versus supercoiling: the predicted fluctuations jump near $ \chi_c $ is clearly seen on the preliminary experimental data. The overall good agreement of our predictions with the analyzed experimental data seem to indicate that the self-avoiding effects, not included in our RLC model, play a limited role in the low supercoiling regime $ \vert \sigma \vert \leq 0.02 $. This was suggested by the Monte Carlo simulations of Marko and Vologodskii \cite{marvol} who use two infinite impenetrable walls to adapt the closed chain formalism to the supercoiled open chains. A further test follows from computing $ \lim_{F=0}\, \langle \,\chi_{W} \, \rangle / \chi $ for $\sigma $ values taken in the range: $ 0 \le -\sigma \le 0.04 $. This ratio compares well to the closed DNA chain writhe average $ < Wr/\Delta Lk > $ obtained by Monte Carlo simulations \cite{volog}. The two results diverge by less than $ 10\% $ when $ \vert \sigma \vert \leq 0.04 $ and both agree with the experimental data within errors \cite{boles},\cite{adrian}. It should be said that the two calculations use different formulas for the writhe: an open chain local version in the present paper, the non-local loop geometrical formula \cite{geometry} in reference \cite{volog} . In view of the wealth of experimental information, there are strong motivations to extend the validity of the present RLC model to a larger domain of the $( F,\sigma)$ plane. Further work has to be pursued in several directions. Non-local constraints in the chain tangent vector $\hat{{\bf{t}}} $ space have to be implemented. There are first the empirical geometrical constraints associated with the DNA anchoring glass plate and the finite radius of the tracking bead. The self-avoiding effects induced by the Coulomb repulsion within the DNA chain have to be studied for the range of ionic strength accessible to experiments. The only practical approach to these problems seem, for the moment, the Monte Carlo simulations technique. Another interesting perspective is to incorporate in the present model the double helix structure the DNA, in order to describe the DNA denaturation transition induced by negative supercoling above $ F=0.5 \,pN$. Some first steps in this direction have already been taken in \cite{liverpool,zhou}. Recently a model coupling the hydrogen-bond opening with the untwisting of the double helix has been proposed \cite{cocco}. It allows a unified description of DNA denaturation driven by thermal fluctuations or induced by the double helix untwisting, in the case a straight line molecular chain. It will be of interest to combine this model with the elastic RLC model in order to study the onset of the denaturation transition at moderate stretching forces, say below $ 1 pN$, where bending fluctuations can no longer be neglected. \section*{Acknowledgments} It is a great pleasure to thank all members of the experimental group at ENS, J.-F. Allemand, D. Bensimon, V. Croquette and T.R. Strick, for many valuable discussion, and for providing us some of their unpublished data. We have benefited from discussions with J.-P. Bouchaud, A. Comtet and C. Monthus on the winding of random walks. We are very grateful to D. Bensimon and M.A. Bouchiat for their careful reading of this long manuscript. The work of MM was supported in part by the National Science Foundation under grant No. PHY94-07194. \begin{appendix} \section{ Rotation Matrix Algebra} Let us denote by $ R( \hat{{\bf{n}}}, \gamma ) $ the rotation of angle $\gamma$ about the unitary vector $ \hat{{\bf{n}}} $. With this notation, the rotation $ {\cal R}(s) $, which specifies an arbitrary DNA chain configuration in terms of the three Euler angles, is given by : \be {\cal R}(s) = R\(( \hat{{\bf{z}}}, \phi(s) \)) R\(( \hat{{\bf{y}}}, \theta(s) \)) R\(( \hat{{\bf{z}}}, \psi(s) \)) \label{Re} \ee Another very useful way of writing $ {\cal R}(s) $ follows from the rotation group relation: \be \hat{R}=R_1^{-1} R( \hat{{\bf{n}}}, \gamma ) R_1= R( R_1^{-1} \hat{{\bf{n}}},\gamma ) \label{grR} \ee The proof of (\ref{grR}) follows from basic properties of a rotation matrix: first, the rotation axis is the rotation matrix eigenvector with unit eigenvalue; we verify that it is indeed the case for $ R_1^{-1} \hat{{\bf{n}}}$: $\hat{R} R_1^{-1} \hat{{\bf{n}}}= R_1^{-1} R( \hat{{\bf{n}}}, \gamma ) \hat{{\bf{n}}}= R_1^{-1} \hat{{\bf{n}}} $; second, the rotations $\hat{R}$ and R$( \hat{{\bf{n}}}, \gamma )$ have, by construction, the same eigenvalues: $ 1, \exp (i \gamma ), \exp (-i \gamma )$ and hence the same rotation angle $\gamma$. The new form of $ {\cal R}(s) $ is then obtained from the simple manipulations: \bea {\cal R}(s) &= & R\(( \hat{{\bf{z}}}, \phi \)) R\(( \hat{{\bf{y}}}, \theta \)) R\(( \hat{{\bf{z}}}, \psi \)) \nonumber \\ &=& R\(( \hat{{\bf{z}}}, \phi +\psi \)) R^{-1}\(( \hat{{\bf{z}}}, \psi \)) R\(( \hat{{\bf{y}}}, \theta \)) R\(( \hat{{\bf{z}}}, \psi \)) \nonumber \\ &=& R\(( \hat{{\bf{z}}}, \phi +\psi \)) R\(( R( \hat{{\bf{z}}}, -\psi ) \hat{{\bf{y}}}, \theta \)) \eea We are going now to discuss the two angular velocity vectors and $ {\bf{\Omega}} (s) $ and $ {\bf{\Upsilon}} (s)$ defined by relations valid for arbitrary vectors $ {\bf{X}} $ and $ {\bf{Y}} $: \bea \dot {\cal R} (s) {\cal R}^{-1}(s) {\bf{X}} &=& {\bf{\Omega}} (s) \wedge {\bf{X}} \label{Om}\\ {\cal R}^{-1}(s) \dot {\cal R} (s) {\bf{Y}} &=& {\bf{\Upsilon}}(s) \wedge {\bf{Y}} \label{Up} \eea Applying to equation (\ref{Up}) the simple identity $ {\cal R}\(({\bf{a}}\wedge{\bf{b}}\))=\(({\cal R}{\bf{a}}\))\wedge \(({\cal R}{\bf{b}}\)) $ and taking $ {\bf{Y}}= {\cal R}^{-1}(s) {\bf{X}} $, one gets immediately the relation: ${\bf{\Omega}} (s) = {\cal R} (s) {\bf{\Upsilon}}(s)$. It facilitates the evaluation of the components of $ {\bf{\Omega}} (s) $ upon the moving trihedron $\{ {\hat{{\bf{e}}}}_{i}(s)\} $ since we can write: \be {\Omega}_{i}= {\bf{\Omega}}(s) \cdot {\hat{{\bf{e}}}}_{i}(s) ={\bf{\Upsilon}}(s) \cdot {\hat{{\bf{e}}}}_{i}^0(s) \label{omi} \ee To simplify the writing in the explicit computation of $ {\bf{\Upsilon}}(s) $, we introduce the notations : $ R_1= R\(( \hat{{\bf{z}}}, \phi(s) \)) , R_2= R\(( \hat{{\bf{y}}}, \theta(s) \)) , R_3=R\(( \hat{{\bf{z}}}, \psi(s) \))$. With some elementary matrix algebra we get: $$ {\cal R}^{-1}(s) \dot {\cal R} (s) = ( R_2 R_3)^{-1} R_1^{-1} \dot R_1 R_2 R_3+ R_3^{-1} R_2^{-1} \dot R_2 R_3+ R_3 \dot R_3 $$ As an intermediary step, we compute : $$ R_a^{-1} R^{-1}(\hat{{\bf{n}}},\gamma(s) ) \dot R (\hat{{\bf{n}}},\gamma(s) ) R_a {\bf{X}}= R_a^{-1} \(( \dot \gamma(s) \hat{{\bf{n}}} \wedge R_a {\bf{X}} \))= \dot \gamma R_a^{-1} \hat{{\bf{n}}} \wedge {\bf{X}} $$ This result is nothing but the relation (\ref{grR}) applied to an infinitesimal rotation. Using the above results we arrive finally to the following expression for $ {\bf{\Upsilon}} (s) $ \be {\bf{\Upsilon}} (s)= \dot \phi R\(( \hat{{\bf{z}}}, -\psi \)) R\(( \hat{{\bf{y}}}, -\theta \)) \hat{{\bf{z}}} + \dot \theta R\(( \hat{{\bf{z}}}, -\psi \)) \hat{{\bf{y}}} + \dot \psi \hat{{\bf{z}}} \ee It is now convenient to decompose $ {\bf{\Upsilon}} (s) $ in a longitudinal ${{\bf{\Upsilon}}}_{\parallel} (s)$ and a transverse part ${{\bf{\Upsilon}}}_{\perp} (s)$ \bea {{\bf{\Upsilon}}}_{\parallel} &=&( \cos\theta \, \dot \phi + \dot \psi ) \hat{{\bf{z}}} \\ {{\bf{\Upsilon}}}_{\perp} &=& R \(( \hat{{\bf{z}}}, -\psi \)) ( -\dot \phi \, \sin\theta\hat{{\bf{x}}} + \dot \theta\, \hat{{\bf{y}}} ) \eea We get immediately the quantities appearing in the RLC elastic energy: \bea {\Omega}_3 &=& \cos\theta \, \dot \phi + \dot \psi \\ {\Omega_{\perp}}^2 &=& {{\bf{\Upsilon}}}_{\perp}^2= {\Omega}_1^2+{\Omega}_2^2 = \dot \phi^2 \, \sin^2\theta + \dot \theta^2 \eea Our next step is to compute the cylindrical symmetry breaking term $ \Delta \Omega(s)= {\Omega}_{1}^2-{\Omega}_{2}^2 $. Introducing the angle $\zeta(s) =\arctan\((\dot{\theta}/(\sin\theta\dot{\phi} )\))$, we can write: $ {{\bf{\Upsilon}}}_{\perp} = -\Omega_{\perp} \,R( \hat{{\bf{z}}},-\psi-\zeta)\,\hat{{\bf{x}}} $. A physical interpretation of $\zeta(s)$ is obtained by computing the $s$ derivative of the tangent unit vector $ \hat{{\bf{t}}}(s) $: \bea {d \hat{{\bf{t}}}(s) \over d s} &= & {\bf{\Omega}} \wedge \hat{{\bf{t}}} = {\cal R}(s) \(( {{\bf{\Upsilon}}}_{\perp} \wedge \hat{{\bf{z}}} \)) \nonumber \\ &=& -\Omega_{\perp} {\cal R}(s) \,R( \hat{{\bf{z}}},-\psi-\zeta) \((\hat{{\bf{x}}} \wedge \hat{{\bf{z}}} \)) \nonumber \\ &= & \Omega_{\perp} R\(( \hat{{\bf{z}}}, \phi \)) R\(( \hat{{\bf{y}}}, \theta \)) R\(( \hat{{\bf{z}}}, -\zeta \)) \hat{{\bf{y}}} \nonumber \eea We see that $-\zeta(s)$ plays the role of the Euler angle $\psi(s) $ vis-\`a-vis the Seret-Frenet trihedron so that $\zeta(s)$ is clearly connected with the writhe. It is now a simple matter to get the component ${\Omega}_1 $ by writing the series of equalities: \bea {\Omega}_1 &=& {\bf{\Upsilon}}(s) \cdot {\hat{{\bf{e}}}}_{1}^0(s)= -\Omega_{\perp} \(( R( \hat{{\bf{z}}},-\psi-\zeta) \,\hat{{\bf{x}}} \)) \cdot \(( R( \hat{{\bf{z}}},\omega_0 s)\,\hat{{\bf{x}}} \)) \nonumber \\ &=& -\Omega_{\perp} \hat{{\bf{x}}} \cdot \(( R( \hat{{\bf{z}}},\psi+\zeta +\omega_0 s)\,\hat{{\bf{x}}} \))= -\Omega_{\perp}\,\cos(\psi+\zeta +\omega_0 s) \nonumber \eea In a similar way we obtain ${\Omega}_1 = -\Omega_{\perp}\,\sin(\psi+\zeta +\omega_0 s)$ and we arrive finally at the following expression for $ \Delta \Omega(s)$: \be \Delta \Omega(s)= {\Omega}_{1}^2-{\Omega}_{2}^2 = {\Omega_{\perp}}^2 \cos 2 (\psi+\zeta +\omega_0 s) \ee Introducing the length resolution function $ P(s)={1\over \sqrt{2 \pi } \ell}\exp( -{1\over2} s^2 /{\ell}^2) $ ( Note the change of notation: $\ell$ stands for $\Delta\,l$ used in the main body of the paper), we proceed with the computation of the average $ \overline{ \Delta\Omega(s) }$: $$ \overline{ \Delta\Omega(s) }=\int ds_1 P(s_1-s)\Delta\Omega(s_1) = {1\over \sqrt{2 \pi } } \int du \exp \((-\frac{1}{2} u^2\)) \Delta\Omega(s + u \ell) $$ Let us first neglect the variation of $\Omega_{\perp} $ within the interval $( s-\ell,s +\ell )$ and perform a first order expansion in $\ell $ of the phase $ \psi(s+ u \ell)+\zeta (s+ u \ell) $; this is justified since its variation is expected to be of the order of $ { \ell\over A } \approx 0.2 $ ( We have proved explicitly the thermal average inequality: $ \langle \dot \psi \rangle / \eta < 1 /A $). In this way $\overline{\Delta\Omega(s) }$ is transformed into a Gauss integral: \bea \overline{ \Delta\Omega(s) } &= & \Delta \Omega(s) {1\over \sqrt{2 \pi } } \int du \exp \((-\frac{1}{2} u^2 \)) \cos\(( 2 u \ell( \dot \psi+ \dot \zeta +\omega_0 )\)) \nonumber \\ & = & \Delta \Omega(s) \exp\(( -2 \, {\ell}^2 \,(\dot \psi+ \dot \zeta +\omega_0 )^2 \)) \eea Using the estimates $ \vert \dot \psi \vert / \omega_0 \vert \sim \vert \dot \zeta \vert / \omega_0 \sim \vert \chi \vert/( L \omega_0) =\vert\sigma\vert $ and the fact that in the present paper our analysis is restricted to values of $ \vert\sigma\vert < 4 \; 10^{-2} $, we can write, introducing the pitch $p=2\pi/\omega_0$ : \be \overline{ \Delta\Omega(s) }\approx \Delta\Omega(s) \exp\(( -\frac{1}{2} (\frac{4\pi \,\ell}{p})^2 \)) \ee Taking $ p=3.4 \, nm $, $\ell=10 nm $ we find: $\frac{4\pi \,\ell}{p} \approx 37$; it means that $ \overline{ \Delta\Omega(s) }/ \Delta\Omega(s) $ is zero for all practical purposes. Using instead $ \ell=b= 7 \, nm $ would not make any difference. Let us say few words about the term $\delta \overline{ \Delta\Omega(s) }$ involving the variation of $ {\Omega_{\perp}}^2 $. A computation similar to the previous one gives the following result: $$ \delta \overline{ \Delta\Omega(s) } = { \partial \, {\Omega_{\perp}}^2 \over \partial \,s}\,\ell \, \frac{4\pi \,\ell}{p} \exp\(( -\frac{1}{2} (\frac{4\pi \,\ell}{p})^2 \)) \sin 2 (\psi+\zeta +\omega_0 s)$$ The rate of variation of $ {\Omega_{\perp}}^2 $, which is the inverse of the curvature radius square, is expected to be of the order $ 1/A $ so that $ { 1 \over {\Omega_{\perp}}^2 } {\partial \, {\Omega_{\perp}}^2 \over \partial \,s} \,\ell \sim \ell /A =0.2 $. It follows that $\delta \overline{ \Delta\Omega(s)}$ is still exceedingly small compared to $\Delta\Omega(s)$. Futhermore it is easily shown that the thermal average derivative $ { \partial \, \langle {\Omega_{\perp}}^2 \rangle \over \partial \,s} $ vanishes for $ s \gg A $. \section{ The Symmetric Top and the RLC Model} In this appendix we shall use for convenience a units system where $ \hbar= c =k_BT=1 $ and write $ \Im{s}=t $. The elastic energy $ E_{RLC} $ (eq. (2) and eq. (3) ) is transformed by an analytic continuation towards the imaginary $s$ axis into $ -i $ times the action integral: $$ {\cal A}(t_0,t_1) = \int_{t_0}^{t_1} dt \(( {1\over2}\sum_{j=1}^3 {C}_j {\Omega}_{j}^2 +f \cos \theta(t) \)) $$ where $ {C}_1={C}_2= A $, $ C_3=C$ and $ f= F/(k_BT) $. The time derivatives accounts for the relative change of sign between ${C}_i {\Omega}_{i}^2$ and the potentiel energy $ -f \cos\theta(t) $. The analytically continued partition function of the RLC model is then identified with the Feynmann path integral amplitude: $$ \langle {\theta}_1,{\phi}_1,{\psi}_1,t_1\vert{\theta}_0,{\phi}_0,{\psi}_0 ,t_0 \rangle= \int {\cal D}\(( \theta,\phi,\psi \)) \exp\((i \int_{t_0}^{t_1} dt {\cal L}_{top} (t) \)) $$ The Lagrangian $ {\cal L}_{top} (t)= {1\over2}\sum_1^3 {C}_i {\Omega}_{i}^2 +f \cos \theta(t) $ describes the motion of a spherical top with inertia moments $ {I}_i={C}_i $, under the action of a static electric field $ E_0 $. (The molecular electric moment is given by $ f/E_0 $.) In order to compute explicitly the hamiltonian $ {\cal H}_{top} $ let us write $ {\cal L}_{top} (t)$ in terms of the Euler angles and their derivatives: \be {\cal L}_{top}= {A \over 2} \left( {\dot{\phi}}^2 \,{\sin^2\theta}+{ \dot{\theta}}^2 \right) + { C \over 2} \,( \dot{\psi} +\dot{\phi} \, \cos\theta )^2 + f \cos \theta \ee One gets immediately the conjugate momentums relative to three Euler angles: $ p_{\phi} = A \, \sin^2\theta \,\dot{\phi} + \cos\theta \, p_{\psi}$, $ p_{\psi} = C \,( \dot{\psi} +\dot{\phi} \, \cos\theta )$ and $p_{\theta} = A \,\dot{\theta}$. The symmetric top hamiltonian is then readily obtained: \bea {\cal H}_{top} &=& {A \over 2} \left( {\dot{\phi}}^2 \,{\sin^2\theta}+{ \dot{\theta}}^2 \right) + { C \over 2} \,( \dot{\psi} +\dot{\phi} \, \cos\theta )^2 - f \cos \theta \nonumber \\ &=& { (p_{\phi}- \cos\theta \, p_{\psi} )^2 \over 2\,A\,\sin^2\theta }+ { {p_{\theta}}^2 \over 2\,A}+{ {p_{\psi}}^2 \over 2\,C}-f \cos \theta \eea To get the hamiltonian operator $ {\hat{\cal H}}_{top} $ we apply the standard quantization rules: $$ p_{\phi} \rightarrow {\hat{p}}_{\phi} = { \partial \over i \partial \phi } \;,\; p_{\psi} \rightarrow {\hat{p}}_{\psi}={ \partial \over i \partial \psi } \;,\; p_{\theta}^2 \rightarrow {\hat{p}}_{\theta}^2= -\frac{1}{ \sin\theta }\frac{\partial}{\partial\,\theta}\sin\theta \, \frac{\partial}{\partial \theta } $$ To get the partition function $ Z({\theta}_1,{\phi}_1,{\psi}_1,s_1\vert{\theta}_0,{\phi}_0,{\psi}_0,s_0 )$ we expand the final and initial states upon eigenfuctions of the operators $ {\hat{p}}_{\phi}$ and ${\hat{p}}_{\psi}$: $ \exp(i\,m\,\phi +i\,k \, \psi ) $ where $ k $ and $ m $ are arbitrary real numbers, in contrast with the real symmmetric top case, where they are integers. This difference, which follows from a detailed analysis of the physics involved ( see section II B. for details), is responsible for the singular features of the continuous RLC model. The partition function with the notations used in the paper to label the initial and final states reads as follows ; \bea Z(\theta(L),\phi(L),\psi(L),L\vert \theta(0),0 ) & = & \int\,dm \,dk \exp\(( i\,m\,\phi(L) +i\,k \, \psi (L) \)) \nonumber \\ & & \langle \theta(L)\vert \exp \(( -L\,\hat{\cal K} (k,m) \)) \vert \theta(0)\rangle \eea where the Hamiltonian $\hat{\cal K} (k,m) $ is given by: $$ \hat{\cal K} (k,m)= -{1 \over 2\,A \,\sin\theta } \, \frac{\partial}{\partial\,\theta}\sin\theta \, \frac{\partial}{\partial \theta }+ { (m- \cos\theta \, k )^{2} \over 2 \, A \,\sin^2 \theta }+{ k^2 \over 2\,C}-f \cos \theta $$ The experimental supercoiling constraint is implemented by averaging upon $ \phi(L) $ and $ \psi (L) $ the above partition function multiplied by the Dirac function $ \delta( \phi(L)+ \psi (L)-\chi ) $. A straightforward computation gives: $$ Z(\chi )= \int \,dk \exp(i \,k\,\chi ) \langle \theta(L)\vert \exp \(( -L\,\hat{\cal K} (k,k) \)) \vert \theta(0)\rangle $$ The hamiltonian operator ${\hat H}_{RLC}(k)$ given by equation (16) is recovered by writing: \bea {\hat H}_{RLC}(k) &=& {1\over A} \(( \hat{\cal K} (k,k)- {A\over 2\, C} k^2 \)) \nonumber \\ &=&-\frac{1}{ 2\,\sin\theta }\frac{\partial}{\partial\,\theta}\sin\theta \, \frac{\partial}{\partial \theta }- \alpha \cos\theta+{k^2 \over 2} {1-\cos\theta \over 1+\cos\theta} \eea The merit of this direct derivation is to show clearly that $ k$ is the unquantized angular momentum of the Euclidian symmetric top problem associated with the RLC model. It is the breaking of the Quantum Mechanics usual quantization rule which is at the origin of the RLC model pathology. \end{appendix}
\section{Introduction} During the last 15 years, wide-angle redshift surveys of the nearby universe have provided a basis for statistical characterization of the local large-scale structure of the universe. The CfA (Huchra, Vogeley \& Geller 1998; Huchra, Geller \& Corwin 1995; Geller \& Huchra 1989; Huchra et al. 1990; Huchra et al. 1983; Davis et al. 1982) and SSRS surveys (da Costa et al. 1994a; 1988) cover more than a third of the sky and reach to a limiting apparent magnitude $m_B \sim 15.5$. There is a rich literature analyzing these surveys to extract, for example, the galaxy luminosity function (Marzke \& da Costa 1997; Marzke, Huchra \& Geller 1994), the power spectrum for the galaxy distribution (da Costa et al. 1994b; Marzke et al. 1995; Park et al. 1994) and the velocity moments (Marzke et al. 1994). One factor limiting these analyses is the inhomogeneous nature of the databases. Here we take a step toward remedying that situation for the CfA surveys which are based on the ZC (Zwicky \mbox{ et~al.}\ 1961-1968). Measurement of redshifts for the CfA surveys began in 1978. The ZC provided an obvious and, at that time, seemingly adequate source of positions for galaxies in the northern hemisphere. The accuracy of the Zwicky catalog coordinates ($3\arcmin$ at the $3\sigma$ confidence level) was comparable with the typical pointing accuracy of the Tillinghast Reflector on Mt. Hopkins, the workhorse for the CfA surveys. Nowadays, more comprehensive scientific goals and significantly improved telescope pointing make arcsecond coordinates imperative. Fortunately, the general availability of the digitized POSS plates (DSS; Lasker et al. 1990) enables us to revise the ZC by providing $\sim 2\arcsec$ coordinates for each galaxy. Here we provide these coordinates for \mbox{19,369}\ galaxies with $m_{Zw} \leq 15.5$ that we were able to identify unambiguously. The measurement of a redshift for a nearby galaxy is now a rapid, routine process. However, at the start of the CfA surveys, the much slower pace made it seem judicious to accumulate redshifts from the literature. Given the effort required to construct the UZC, it is now clear that it would have been better to remeasure the redshifts to provide a uniform database. Approximately 30\% of the redshifts in the catalog we describe here are from the literature; the rest were measured at the CfA. Most of the CfA measurements were made with two very different spectrographs, the Z-machine (Latham 1982) and FAST (Fabricant et al. 1998). We have re-reduced all of the Z-machine and FAST data and have determined a uniformly calibrated error (Kurtz \& Mink 1998; hereafter KM98). The UZC contained here includes redshifts for 96\% of the galaxies with m$_{Zw} \leq 15.5$ along with $\sim 1,400$ additional redshifts for galaxies within multiplets in the region $20^h \geq \alpha_{1950} \leq 4^h$ and $8^h \geq \alpha_{1950} \leq 17^h$, both for $-2\pdeg5 \leq \delta_{1950} \leq 50^\circ$ (where our completeness is 98\%; hereafter we refer to this region as the CfA2 region). We define UZC multiplets as galaxies with magnitude differences $\Delta m < 0.5$ mag and positional differences $\Delta \theta <3\arcmin$. The catalog is suitable for a wide range of statistical analyses. Section 2 describes the contents of the original Zwicky-Nilson merged catalog (ZNCAT). Section 3 describes the ZCAT compilation. Section 4 outlines our procedure for matching galaxies in ZNCAT and ZCAT with the HST Guide Star catalog (GSC), as a first pass to obtain accurate coordinates, and a subsequent refinement that improved the yield of matched galaxies. Section 5 describes the CfA redshift observations and reduction procedures. Section 6 outlines the procedures we used for blunder identification and for blunder rate evaluation. Section 7 contains the catalog and comments on its possible applications, and limitations. \section{The Zwicky Catalog} Our starting point for constructing the UZC is the database of properties of galaxies in the ZC, the Zwicky-Nilson catalog (ZNCAT; Tonry \& Davis 1979; hereafter TD79). ZNCAT was created in preparation for the first CfA Redshift Survey (Davis et al. 1982); it covers the entire northern sky and contains the union of CGCG galaxies in the Zwicky catalogs (Zwicky \mbox{ et~al.}\ 1961-1968) and UGC galaxies in the Nilson catalog (Nilson 1973). ZNCAT contains position, magnitude and some morphology entries for 30,813 galaxies. We restrict ourselves to the \mbox{18,901}\ objects with m$_{Zw} \leq 15.5$, the limit where Zwicky estimated that his catalog was complete. The CfA redshift surveys include redshifts for nearly all of the galaxies to this limit over large areas. Each galaxy in ZNCAT is described by a ``Zwicky Number'', another catalog label such as NGC or IC, its B1950 coordinates, Zwicky's estimate of its Johnson $m_B$ magnitude, m$_{Zw}$, and, if known, its morphological type and its position angle on the sky and $B-$ and $R-$band diameters in arcsec. The errors in the galaxy magnitudes are 0.3 mag (Bothun \& Cornell 1990; Huchra 1976); the $3 \sigma$ error circle for the galaxies has a radius of $\sim 3\arcmin$ (Zwicky et al. 1968). \section{ZCAT} ZCAT (Huchra et al. 1992) is a compilation of redshift data for galaxies that currently contains approximately 100,000 entries. ZCAT includes positions, redshifts, magnitudes, and types from a variety of sources. We used ZCAT as an initial list of redshifts to construct a complete catalog for part of the northern sky. The coordinates of $\sim 9,600$ objects in ZCAT are identical to the corresponding ones in ZNCAT; thus, ZCAT coordinates are often accurate to only $\sim 3\arcmin$ ($3\sigma$). Revised coordinates in ZCAT frequently remain good to $\sim 1\arcmin$ ($3\sigma$) at best. There is also possible confusion with neighboring bright galaxies in $\sim 15\%$ of the entries. Such confusion cannot be eliminated without both accurate coordinates and magnitudes (see de Vaucouleurs et al. 1991; hereafter RC3). Without the latter, the only alternative was to determine accurate coordinates as we do here, and remeasure redshifts based on these coordinates. We recently began a program of redshift remeasurement; we include the current results of this program in the UZC. Magnitudes from the ZC in ZCAT are problematic. In the ZCAT compilation, the magnitudes from the original ZC are often replaced with measurements from a multitude of other sources. In cases of ``multiplets'' defined as for the UZC, ZCAT often contains eye estimates of split magnitudes. For the UZC, we restore the original Zwicky magnitudes $m_{Zw}$ because they are the only system uniformly available for all of the galaxies and because they determined the original selection of galaxies for inclusion in the CfA surveys. In the catalog, we flag these multiples; the split magnitudes are available for reference in ZCAT (Huchra \mbox{ et~al.}\ 1992). The arcsecond coordinates we provide make it easy to measure and/or include better magnitudes if more reliable sources of photometry are available (see also RC3). In using the UZC for statistical analyses, it should be noted that the flagged magnitudes of multiplets could be systematically too bright for each galaxy at the coordinates listed in the catalog. A careful reading of Zwicky et al. (1961) reveals that magnitudes in the Zwicky catalog were not compiled for multiplets in a uniform fashion. In cases where the multiplet is unresolved, the magnitude listed by Zwicky et al. (1961) is probably too bright for any single component. In resolved cases, the listed magnitude may be that for the brightest member alone. In the absence of a clear indication of the procedure of Zwicky et al. (1961), the safer alternative that we followed was to list only the original Zwicky et al. (1961) magnitude along with the brightest member. Furthermore, ours is not an exhaustive catalog of multiplets: those listed in the UZC were included because the magnitudes ($m_{Zw}$) of the components differed by less than $\sim 0.5$ mag. However, the UZC is a source for further searches for and statistical studies of multiplets. \section{Matching Catalogs to Obtain Arcsecond Positions} We compared the positions of galaxies in ZNCAT and ZCAT with an independent source of accurate coordinates, the HST Guide Star Catalog (GSC). The GSC contains coordinates for stars, and for extended objects (``non-stars''). Typical accuracies for all GSC coordinates are $\sim 1\arcsec$. However, the main focus of the GSC was on $V < 14$ mag stars. Thus, the magnitudes and identifications of galaxies are less reliable than for the stars (e.g., Lasker et al. 1990; Alonso et al. 1993, 1994). With these limitations in mind, we selected objects from the GSC which match the positions of the objects in ZNCAT. We matched GSC and ZNCAT positions with software written for this task. From a list of ZNCAT and GSC coordinates, we calculated the Cartesian distance on the sky between objects in ZNCAT and in the GSC, and searched for matches within a $6\arcmin\times6\arcmin$ box centered on each ZNCAT position; we found 19,878 matches. Out of these, 1,854 were duplicates due to overlaps of the digitized sky survey plates, for a total of 18,024 non-duplicate matches. We made $6\arcmin\times 6\arcmin$ finding charts extracted from the DSS, centered on each original ZNCAT position. Our software classified these charts as follows: \begin{enumerate} \item class Z: a GSC class 3 (a ``non-star'') object appears within a $6\arcmin\times6\arcmin$ box centered on the Zwicky position. For the 16,268 fields in this category, there is a high likelihood that each GSC object is the Zwicky galaxy. \item class N: no GSC class 3 object appears within $6\arcmin\times6\arcmin$ box centered on the Zwicky position, but one GSC class 0 object (a ``star'') is within the box. For the 3,363 fields in this category, the object nearest the Zwicky position is probably a star in the GSC. \item class B: no GSC object of any class appears within the $6\arcmin\times6\arcmin$ box centered on the Zwicky position. There are 247 fields in this category. Here, either through an error in the GSC classification, or through an error in the ZNCAT coordinates, there is no obvious candidate galaxy at the ZNCAT position. \end{enumerate} We printed DSS finding charts for all the objects in our sample. Charts of type Z and N are centered on the matching GSC coordinates. Charts of type B are centered on the ZNCAT coordinates. To each matched object, we assigned a running index increasing with apparent magnitude. Ordered in this way, objects of similar apparent brightness are sequentially indexed close together. Thus, the visual classification described below remains consistent as a function of decreasing apparent brightness. We also matched objects in ZCAT and ZNCAT by searching each $6\arcmin\times6\arcmin$ coordinate box (centered on each ZNCAT object with $m_{Zw}\leq 15.5$ mag) with corresponding objects in ZCAT (with no restriction on the magnitudes). We found 19,584 matches, a number exceeding our total count (18,024) of matched non-duplicate ZNCAT objects within our magnitude limit. The mismatch occurred because (1) there were multiple matches within our matching box and (2) there were matches with ZCAT objects fainter than our choice of magnitude limit. Because there are significant errors in both the GSC classification and in the ZNCAT coordinates, we could not be certain that our Z or N-class field classifications corresponded to the actual Zwicky galaxies. However, we had a preliminary ranking of galaxies in order of decreasing likelihood (from Z to N to B) of matching a ZNCAT galaxy. We therefore examined all of the Z, N and B fields visually, to determine a second classification of the objects. This time-consuming classification substantially improved the likelihood that a matched object in the GSC is actually a ZNCAT galaxy. The new classification consisted of numerical indices 0$-$4: \begin{enumerate} \item{} code 0: a ``hit'', where the GSC and ZNCAT objects are separated in position by $\Delta\theta\leq 180\arcsec$, are comparable in brightness as determined visually, and there is no other object within the field that could be a ZC galaxy. Our indexing scheme furnished a qualitative estimate for the range of magnitudes expected for each galaxy, which we used as a guide during the visual inspection. Note that the magnitudes m$_{Zw}$ (B band) and m$_{\rm GSC}$ (V or J bands) can differ significantly, because the expected colors for these galaxies are in the range $\sim 0.3-1.0$ mag (Frei \& Gunn 1994). \item{} code 1: a ``hit'' with $\Delta\theta\leq 180\arcsec$, again with a GSC brightness in the vicinity of the ZNCAT value. However, the field is centered on a GSC ``star'' rather than on a ``non-star'', but there is only one possible ZC galaxy nearby, within the field. Thus, index 1 merely indicated that the nearest GSC object to the ZCAT coordinates is a star rather than the appropriate galaxy; \item{} code 2: a near ``hit'' with $\Delta\theta\leq 180\arcsec$. However, there is confusion because there are at least 2 galaxies with an estimated spread in magnitudes $|\Delta m|$ within the expected range, either of which could be a ZC galaxy; \item{} code 3: also a near ``hit'' with $\Delta\theta\leq 180\arcsec$ but with $|\Delta m|$ outside the expected range, raising suspicions about the match of the GSC object to the ZC galaxy; \item{} code 4: no ``hit'' at all for $\Delta\theta\leq 180\arcsec$, i.e., there is no match for a ZC galaxy. \end{enumerate} One of us (B. Elwell) examined each chart and assigned it a code of $0-4$, according to these criteria. Our strategy was to print finding charts in order of increasing magnitude; thus, we always knew the range of brightness being examined. The visual examination yielded refined, accurate estimates for the appropriate object to match to the ZC coordinates. There were 3 distinct outcomes: \par\noindent 1) For perfect ``hits'', there is no doubt that the Zwicky galaxy is matched. The coordinates are those of the matching object in the GSC. We examined our printed DSS finding charts to confirm the match visually. These $12,154$ galaxies are in class 0 above. \par\noindent 2) For near ``hits'' in class 1, there is no doubt that we have a match for the ZC galaxy, but the GSC coordinates did not fall within $\sim 2\arcsec$ of the center of brightness of the ZC galaxy. One of us (J. Peters) confirmed the identification with software which allowed interactive examination of the DSS finding charts for the coordinates of each ZC galaxy. These $2,799$ galaxies were in class 1 above. \par\noindent 3) For confused ``hits'' in classes $2-4$, there is more than one possible match. In these cases, we have 3 sources of confusion: (a) at least 2 objects of approximately the expected magnitude lie within the 3\arcmin\ error circle; (b) the nearest match to the ZNCAT position has the wrong magnitude (invariably fainter); and (c) the 3\arcmin\ error circle is empty. We found $4,631$ galaxies in these classes, leading to a second round of visual examination using software in the WCStools package (Mink 1999). We examined the original finding charts used for the redshift measurements to associate the galaxy with its ZC position and with its measured redshift. When a match was found, the coordinates were determined interactively, with WCStools software. The success rate for this procedure was low, only $\sim 30$ \%, because of the non-uniformity of the record-keeping, and of poorly marked or missing charts. Furthermore, we found many cases ($\sim 20$\% of the charts we examined) where the redshift in ZCAT did not match the entry on the finding chart. The total number of galaxies in classes 0-4 matches the total count of 19,584. At this stage we eliminated duplicates and thus pared the total number of entries to the number of unique ZNCAT galaxies. This total number is the count of galaxies in ZNCAT with non-zero magnitudes after eliminating (1) a small number of duplicate entries; (2) 19 entries for which no clear galaxy was found within $15\arcmin$ of the Zwicky position; and (3) the matches for 063024$+$44480 (NGC2242), a known planetary nebula, and for 065012$+$16590, a known open cluster, both of which were mistakenly included in the ZC. The completeness values reported above are slightly higher if we take into account the ZC entries that we eliminated. \section{Redshifts in the UZC} The redshifts and errors listed for CfA data here take precedence over values published previously, including those measured with the Z-machine. All the CfA redshifts in the UZC were determined from observations made at Mt. Hopkins with the MMT, or with the Z-machine or FAST spectrographs on the 1.5-meter Tillinghast reflector at the Whipple Observatory. We include in the UZC 4,465 redshifts measured with FAST. Of these, 1,905 are previously unpublished redshifts in the CfA2 region, and the rest are in the remaining regions of the northern sky, as well as remeasurements. About 1,000 of the remeasurements were necessary in cases where there was ambiguity in the identification of unresolved galaxies even after looking at historical records. Our source of UZC redshifts from the literature was ZCAT; about 25\% of the redshifts in ZCAT were compiled from the literature. In the CfA2 survey region, CfA measurements make a slightly larger contribution, with 1.4\% of the redshifts from the MMT, 61\% from the Z-machine and 17\% from FAST. The Z-machine (Latham 1982) measurements were made between 1978 and 1993. For these observations, a 600 line mm$^{-1}$ grating yielded a resolution of 5\AA\ in first order over the wavelength range 4500 - 7100 \AA. The typical integration time for each object was 15-50 minutes. In 1993, the FAST spectrograph (Fabricant et al. 1998) was first mounted on the 1.5-meter Tillinghast. The FAST spectra have 6 \AA\ resolution and coverage of 3700 - 7500 \AA. Typical integration times were 3-10 minutes. For our FAST observations, when there was more than one galaxy near the ZNCAT position and within $\sim 0.5$ mag of the original Zwicky magnitude, we measured redshifts for all of the galaxies. We began the observing program in September 1996, and completed it in October 1997. All of these new redshift measurements are included in the UZC (see Table 1). We also updated the coordinates of all of these remeasured galaxies, using the DSS finding charts, with an estimated positional uncertainty of $\sim 2\arcsec$. FAST measurements (from our own and from other projects) replaced any previously measured redshifts, e.g. with the Z-machine, in the UZC. \section {Reliability of the Redshifts in the UZC} In many cases where the GSC galaxy is an unambiguous match to ZNCAT, we checked our DSS finding charts against those used for the CfA Redshift Survey. We corrected obvious typos found in this way. This procedure underscored our awareness of inadequacies in the reduction and record-keeping, and spurred us to make an objective {\it measurement} of the error rate in redshifts from the literature and from our own facilities by remeasuring redshifts for a significant number of objects chosen with a variety of criteria. We first used our current archive of repeat measurements with FAST to estimate the error rate in these new data. For 620 pairs of FAST measurements, there were no blunders (KM98). Thus, we only have an upper limit of $\sim 0.2\%$ for the error rate, which is attributable to the much better pointing of the Tillinghast 1.5-meter, much improved observing protocols, and better record-keeping and archiving procedures. We are thus confident that for the 3,446 FAST redshifts in the UZC, the blunder rate is essentially zero. We reduced the 4,366 FAST spectra of galaxies in our sample in a uniform manner, using the IRAF task {\bf rvsao} (KM98). All spectra were cross-correlated with two templates, fabtemp97, a composite absorption line spectrum (KM98), which is generally used for all FAST redshifts, and femtemp97, a synthetic emission line template (KM98). We estimated the errors for our FAST spectra as in KM98: the values are obtained by dividing a constant by $1+R$ ($R$ values are a measure of the quality of fit of a template and of S/N for the spectra, see TD79, KM98), and adding $20/\sqrt{2}$ \mbox{ km~s$^{-1}$}\ in quadrature. For fabtemp97 the constant is 350 $\mbox{ km~s$^{-1}$}$, and for femtemp97 the constant is 220 $\mbox{ km~s$^{-1}$}$. These constants were selected from Figures 10 and 12 of KM98, to optimize the contributions of statistical and systematic errors from cross-correlation with each template. We required $R>3$ for all the redshifts we report here. In cases where the difference for the two templates is less than 300 \mbox{ km~s$^{-1}$}, the redshift in Table 1 is the error-weighted combination of the two, as is its error. Of the 4,366 spectra, 30 were of insufficient quality, 1,872 matched the absorption line template fabtemp97, 692 matched the emission line template femtemp97, and 1,772 matched both templates. We also re-reduced the 10,051 Z-machine spectra of the galaxies in our sample in a uniform manner with {\bf rvsao} (KM98). All spectra were cross-correlated with the templates ztemp, a composite absorption line spectrum that has been used for all Z-machine redshifts since TD79, and femtemp97. Figure 1 shows spectrograms of ztemp and femtemp97 (compare with Figure 11 of KM98). The Z-machine spectra do not have sufficient quality to permit fully automatic reduction; the initial reduction required visual inspection of every spectrum. We compared the new redshifts, obtained by correlation with the absorption and emission line templates, with those from the original reduction; we also checked for sufficient signal to noise, requiring $R>3$. We accepted a redshift if it fell within 300 $\mbox{ km~s$^{-1}$}$ of the original reduction; of the 10,080 Z-machine spectra we accepted 9,936. We found that 4,424 matched with the absorption line template, ztemp, 2,507 matched with the emission line template femtemp97, and 3,005 matched both templates. Among the remaining 156 low-S/N spectra (FAST and Z-machine), 125 are in our 98\% complete region. These galaxies are flagged in Table 1 (see below); we plan to update their redshifts with FAST. In the cases where a single template matches the original reduction, the redshift in Table 1 is that obtained from that template, and the error is calculated with the methods outlined in KM98. The Table also indicates which template matched best, with a label of A for ztemp, E for femtemp97, and B for both. We calculated the error as for FAST but with appropriate constants, again by dividing a constant by $1+R$, and adding 35 $\mbox{ km~s$^{-1}$}$ in quadrature. For ztemp the constant is 212 $\mbox{ km~s$^{-1}$}$, and for femtemp97 the constant is 140 $\mbox{ km~s$^{-1}$}$; these constants are $3/8$ of the median width of correlation peaks where $R>4$ (KM98). In cases where both templates match the original reduction, the redshift in Table 1 is the error-weighted combination of the two, as is the error. In cases where the match with the original reduction failed, we estimated errors individually to match the system defined by successful cross-correlation. ZCAT contains another 670 low-S/N spectra that would be included in the UZC if their S/N were sufficiently high. Any galaxy that was identified unambiguously but whose redshift remains unknown is entered in Table 1 with accurate coordinates, with nothing in its redshift column. Galaxies for which spectra were obtained at CfA but where the S/N was below our acceptance criterion are flagged in Table 1 (see below). Even after the new measurements for ambiguous cases, the blunder rate in ZCAT is significant for the Z-machine data and for data from the literature. For evaluation of the reliability of redshift measurements including those obtained with the Z-machine and those compiled from the literature, we extracted a random sample of 129 galaxies from ZCAT and re-measured their redshifts with the FAST spectrograph. We also assembled several samples of FAST redshifts acquired for other projects: the Century Survey (106 galaxies: Geller et al. 1997); studies of groups of galaxies (66 galaxies: Mahdavi 1998, private communication); galaxies in voids (226 galaxies: Grogin et al. 1998); and other galaxy surveys (315 galaxies: Carter 1998, private communication). The total number of re-measured redshifts is 843. In the UZC, we have replaced the ZCAT redshift with the corresponding FAST measurement in all cases (the number of FAST redshifts discussed above includes these measurements and re-measurements). For each sample, Figure 2 shows the differences between the FAST measurements of redshifts and the corresponding original ZCAT entry, plotted against the FAST redshifts represented as $cz$. The ``blunder'' rate for the overall test sample, as well as for the individual samples, is stable at $\lesssim 3$ \% for measurements not made with FAST. We find the same rate regardless of the origin of the redshift in ZCAT (the literature, or Z-machine for example). An oddly large number of redshifts in ZCAT from the literature differ by 1,000 \mbox{ km~s$^{-1}$}\ from FAST measurements (only one of which happened to be included in Figure 2), making us suspect typographical errors or confusion of H$\alpha$ (6563\AA) with [NII] (6548\AA, 6584\AA) emission lines as major sources for the discrepancy. \section{Discussion} We have assembled a new version of the Zwicky Catalog (ZC), the Updated Zwicky Catalog (UZC), with a magnitude limit of $m_{Zw}= 15.5$ mag. The UZC is a 98\% complete redshift catalog to this magnitude limit, in the CfA2 region with $20^h \geq \alpha_{1950} \leq 4^h$ and $8^h \geq \alpha_{1950} \leq 17^h$ and $-2\pdeg5 \leq \delta_{1950} \leq 50^\circ$. The completeness of the UZC is lower at the northernmost declinations, e.g., if we restrict $\delta_{1950}\ge 50\deg$, it falls to 90\%. The main advantages of the UZC over previous compilations are (1) uniformly accurate coordinates at the $\leq 2\arcsec$ level; (2) a robust estimate of the accuracy of the CfA redshifts in the range $cz = 0-25,000\mbox{ km~s$^{-1}$}$, with a current total of $\sim 25$\% of redshifts with essentially no ``blunders''; (3) an estimate of the reliability of catalog redshifts for data from the literature and from the Z-machine: the remaining 75\% of the redshifts have a blunder rate of $\mathrel{\raise.3ex\hbox{$<$}\mkern-14mu\lower0.6ex\hbox{$\sim$}} 3$\%. A continuing problem with the UZC is that magnitudes for the galaxies in the sample are still the original Zwicky mgnitudes, which have $\sim 0.3$ mag errors. Multiples are flagged in the UZC (see Table 1), because we have not split the Zwicky magnitudes of their components. These magnitudes may be systematically too bright for each component of an unresolved UZC multiplet (\S 3). Table 1 contains a sample listing of the UZC. Column 1 (``RA (J2000) DEC'') shows the right ascension and declination (J2000) of each galaxy. Column 2 (``{\it B}'') shows the $B$-band magnitude $m_{Zw}$ from ZNCAT. Columns 3 (``c$z$'') and 4 (``$\Delta$c$z$'') show the heliocentric redshift and its error in \mbox{ km~s$^{-1}$}. Column 5 (``T'') indicates the type of the redshift for each galaxy, E (emission), A (absorption) or B (both emission and absorption). Column 6 (``U'') shows the the UZC code (0-4) assigned to each galaxy, according to the quality of the match with DSS galaxies (see \S 4). Column 7 (``$N$'') shows the number of UZC neighboring galaxies to the current entry, within 3\arcmin. Galaxies in the UZC with a number of neighbors larger than 0 are thus in UZC ``multiplet'' systems. All entries with $N>0$ are flagged with a star in Column 12, to indicate membership in a UZC ``multiplet''. Column 8 (``ZNCAT Name'') shows the ZNCAT or Zwicky label; multiplets generally share a single such label. Column 9 (``Ref. code'') indicates the origin of the redshift measurement, for data taken at the CfA: Z for Z-machine, M for the MMT and F for FAST spectra taken for our (other) project(s). The label ``X'' flags a low-S/N spectrum for Z-machine (FAST) measurements and indicates a low S/N match with a spectrum at c$z<100$ \mbox{ km~s$^{-1}$}, an indication that no redshift could be determined. Column 10 (``Ref.'') shows a ZCAT literature reference number (in turn listed in Table 2), or a blank for CfA unpublished data. Column 11 (``Other Name'') shows any other name may be associated with the current entry, such as an NGC number, as listed in ZCAT. Column 13 indicates whether the galaxy is a member of a Zwicky multiplet, as reported by NED (``P'', ``T'' and ``G'' for pairs, triples and groups, respectively). Because of the size and limited utility of a printed version, the full table is available only with the electronic version of the journal article. This table, as well as additional measurements we will carry out with FAST will also be available at URL http://tdc-www.harvard.edu. Figure 3 shows the distribution of the magnitude of the differences between UZC and Zwicky coordinates. The mode of the histogram is at $\sim 40\arcsec$, slightly better than the $1\arcmin$ ($1\sigma$) errors claimed by Zwicky. For additional confirmation of our positional accuracy, we matched the FIRST 1.4 GHz catalog (White et al. 1997) with the UZC, using a search radius of 10\arcsec. We found matches for 1,347 FIRST sources. The distribution of coordinate differences for these sources has a narrow peak at $\sim 1\arcsec$, with a width $\sigma = 1\farcs45$. Thus, for these FIRST matches, the UZC positional uncertainty is $< 1\farcs5$, which confirms the significant improvement in positional accuracy that we have achieved with the UZC. Figure 4 shows the sky distribution of UZC galaxies with measured redshifts, as listed in Table 1. The UZC covers only the northern sky; regions devoid of galaxies result from obscuration by the Milky Way. Figure 5 shows the sky distribution of galaxies without measured redshifts, also as listed in Table 1. A comparison of Figures 4 and 5 shows that there are no significant patterns in the distribution of galaxies without measured redshifts, other than, e.g., those north of $+50\arcdeg$ and south of the equator, with a known lack of coverage. Figure 6 shows cone diagrams for the distribution of galaxies in the UZC as a function of $\alpha$, $\delta$ and c$z$ in \mbox{ km~s$^{-1}$}. Each ``slice'' spans $\sim 10\arcdeg$ of declination and right ascension ranges of $8-17h$ in the northern galactic polar region and $20-4h$ in the South galactic cap. These slices cover the CfA2 region of the redshift survey. The known structure of the distribution of galaxies is readily apparent in these diagrams (Geller \& Huchra 1989). The large voids and coherent sheet-like structures appear in adjacent sets of the ``slices''. Figure 7 shows histograms of the number of galaxies $N(z)$ as a function of redshift for both polar galactic regions in the UZC. The structure known as the Great Wall (Geller \& Huchra 1989) that persists at $7,000-10,000$ \mbox{ km~s$^{-1}$}\ in each of the $8-17h$ slices in Figure 6, as well as in Figure 7. The latter figure shows significant peaks at the redshifts of the Great Wall and the Virgo cluster in the North, and at the redshift of Perseus-Pisces in the South galactic hemisphere. Such peaks confirm the presence of these large-scale structures, and demonstrate their narrowness in redshift space. These in turn are characteristic of the sheet-like structures in Figure 6. We thank E. Barton, B. Carter, N. Grogin and A. Mahdavi for use of their unpublished data. We also thank D. Fabricant for comments and J. Roll for comments and indispensable assistance with the software. We also thank the referee, Harold Corwin, for useful suggestions which improved the clarity of our presentation of the UZC. This research was supported in part by the Smithsonian Institution. \newpage \clearpage
\section{Introduction} The Virgo cluster is a fundamental milestone for the determination of $H_0$ because it is the nearest reasonably rich cluster which is tightly tied into the large-scale expansion field through excellent {\em relative\/} distances to more distant clusters, circumventing thus effects of local streaming velocities. The distance of the Virgo cluster has long been controversial mainly for three reasons. (1)~The cluster which spans $\mathrel{\hbox{\rlap{\hbox{\lower4pt\hbox{$\sim$}}}\hbox{$>$}}} 15^{\circ}$ in the sky has a considerable depth effect making even good distance determinations of {\em few\/} member galaxies vulnerable to small-number effects. This becomes particularly precarious if the selection criteria of the few members are themselves distance-dependent. (2)~Distance indicators with non-negligible intrinsic scatter lead always to too small distances if they are applied to only the brightest cluster members (Malmquist bias). This has made Virgo cluster distances useless especially from the 21cm-line width (Tully-Fisher) method until a very deep Virgo cluster catalogue became available (Binggeli, Sandage, \& Tammann 1985). The catalogue is {\em complete\/} as to normal E and spiral galaxies because it goes below the cutoff magnitude of these objects. (3)~Undue weight has been given to distance indicators in the past which had never been tested in the relevant distance range. In the case of the bright tail of the luminosity function of planetary nebulae the ineffectiveness as distance indicator is now explained by the dependence on the sample (i.e. galaxy) size (Bottinelli et~al. 1991; Tammann 1993; M{\'e}ndez et~al. 1993; Soffner et~al. 1996). The reasons why surface brightness fluctuations of E galaxies fail to provide useful absolute distances beyond $10\;$Mpc are less clear. The applicability of the method to dwarf ellipticals is presently investigated (Jerjen, Freeman, \& Binggeli 1998). With these difficulties in mind six different distance determinations of the Virgo cluster are discussed in the following. \section{The Virgo cluster distance from Cepheids} \label{sec:2} Cepheids are presently, through their period-luminosity (PL) relation, the most reliable and least controversial distance indicators. The slope and the zeropoint of the PL relation is taken from the very well-observed Cepheids in the Large Magellanic Cloud (LMC), whose distance modulus is adopted to be $(m-M) = 18.50$ (Madore \& Freedman 1991). An old PL relation calibrated by Galactic Cepheids in open clusters, and now vindicated by Hipparcos data (Sandage \& Tammann 1998), gave $(m-M)_{\rm LMC}=18.59$ (Sandage \& Tammann 1968, 1971). Hipparcos data combined with more modern Cepheid data give an even somewhat higher modulus (Feast \& Catchpole 1997). Reviews of Cepheid distances (Federspiel, Tammann, \& Sandage 1998; Gratton 1999) cluster around $18.56\pm0.05$, -- a value in perfect agreement with the purely geometrical distance determination of SN\,1987A ($18.58\pm0.05$; Gilmozzi \& Panagia 1999). In his excellent review Gratton (1999) concludes from the rich literature on RR\,Lyr stars that $(m-M)_{\rm LMC}=18.54\pm0.12$. He also discusses five distance determination methods which give lower moduli by $0.1-0.2\;$mag, but they are still at a more experimental stage. -- There is therefore emerging evidence that the adopted LMC modulus of 18.50 is too small by $\sim\!0.06$ or even $0.12\;$mag (Feast 1999) and that all Cepheid distances in the following should be increased by this amount. There has been much debate about the possibility that the PL relation of Cepheids depends on metallicity. Direct {\it observational\/} evidence for a (very) weak metallicity dependence comes from the fact that the metal-rich Galactic Cepheids give perfectly reasonable distances for the moderately metal-poor LMC Cepheids and the really metal-poor SMC Cepheids and, still more importantly, that their relative distances are wavelength-independent (Di Benedetto 1997; cf. Tammann 1997). -- Much progress has been made on the theoretical front. Saio \& Gautschy (1998) and Baraffe et~al. (1998) have evolved Cepheids through the different crossings of the instability strip and have investigated the pulsational behavior at any point. The resulting (highly metal-insensitive) PL relations in bolometric light have been transformed into PL relations at different wavelengths by means of detailed atmospheric models; the conclusion is that any metallicity dependence of the PL relations is negligible (Sandage, Bell, \& Tripicco 1998; Alibert et~al. 1999; cf. however Bono, Marconi, \& Stellingwerf 1998, who strongly depend on the treatment of stellar convection). Most extragalactic Cepheid distances are now due to {\sl HST\/ } observations. The reduction of these observations is by no means simple. The photometric zeropoint, the linearity over the field, crowding, and cosmic rays raise technical problems. The quality of the derived distances depends further on (variable) internal absorption and the number of available Cepheids in view of the finite width of the instability strip. (An attempt to beat the latter problem by using a PL-color relation is invalid because the underlying assumption of constant slope of the constant-period lines is unrealistic; cf. Saio \& Gautschy 1998). Typical errors of individual Cepheid distances from {\sl HST\/ } are therefore $\pm0.2\,$mag (10\% in distance). There are now three bona fide cluster members and two outlying members (cf. Binggeli, Popescu, \& Tammann 1993) with Cepheid distances from {\sl HST\/ } (Table~\ref{tab:Virgo_Cepheid_distances}; cf. Freedman et~al. 1998). \begin{table}[t] \footnotesize\rm \caption{The Virgo cluster members with Cepheid distances} \label{tab:Virgo_Cepheid_distances} \begin{center} \begin{minipage}{0.8\textwidth} \begin{tabular}{llll} \noalign{\smallskip} \hline \noalign{\smallskip} Galaxy & $(m-M)_{\rm Cepheids}$ & Remarks & $(m-M)_{\rm TF}$ \\ \noalign{\smallskip} \hline \noalign{\smallskip} NGC\,4321$^*$ & $31.04\pm0.21$ & highly resolved & $31.21\pm0.40$ \\ NGC\,4496A$^{*\,*}$ & $31.13\pm0.10$ & highly resolved & $30.67\pm0.40$ \\ NGC\,4536$^{*\,*}$ & $31.10\pm0.13$ & highly resolved & $30.72\pm0.40$ \\ NGC\,4571 & $30.87\pm0.15$ & extremely resolved & $31.75\pm0.40$ \\ NGC\,4639 & $32.03\pm0.23$ & poorly resolved & $32.53\pm0.40$ \\ \noalign{\smallskip} \hline \noalign{\smallskip} \end{tabular} {\footnotesize \hspace*{5pt}$^*$~From a re-analysis of the {\sl HST\/} observations Narasimha \& Mazumdar (1998) obtained $(m-M)=31.55\pm0.28$. \\ $^{*\,*}$~In the W-cloud outside the confidence boundaries of the Virgo cluster (cf. Federspiel et~al. 1998). } \end{minipage} \end{center} \end{table} The wide range of their distance moduli, corresponding to $14.9$ to $25.5\;$Mpc, reveals the important depth effect of the cluster. The first four galaxies in Table~\ref{tab:Virgo_Cepheid_distances} have been chosen from the atlas of Sandage \& Bedke (1988) because they are highly resolved and seemed easy as to their Cepheids. They are therefore {\em expected\/} to lie on the near side of the cluster. In contrast NGC\,4639 has been chosen as parent to SN\,1990N and hence independently of its distance (Saha et~al. 1997); correspondingly this distance is expected to be statistically more representative. A straight mean of the distances in Table~\ref{tab:Virgo_Cepheid_distances} is therefore likely to be an underestimate. Indeed the mean Tully-Fisher (TF) distance modulus of the five galaxies is $0\mag2$ (corresponding to $10\%$ in distance) {\em smaller\/} than the mean distance of a complete and fair sample of TF distances (Federspiel et~al. 1998). -- It should be noted that NGC\,4639 with the largest distance in Table~\ref{tab:Virgo_Cepheid_distances} has a recession velocity of only $v_0=820\nobreak\mbox{$\;$km\,s$^{-1}$}$, i.e. {\em less\/} than the mean cluster velocity of $v_0=920\nobreak\mbox{$\;$km\,s$^{-1}$}$, and that it can therefore not be assigned to the background. In fact the redshift distribution in the Virgo cluster area shows a pronounced gap behind the cluster minimizing the danger of background contamination (Binggeli et~al. 1993). B{\"o}hringer et~al. (1997) have proposed that the Cepheid distances of the spiral galaxies NGC\,4501 and NGC\,4548 would be significant because these galaxies are spatially close to the Virgo cluster center on the basis of their being stripped by the X-ray intracluster gas. In the case of NGC\,4548 there are some doubts because it is, like NGC\,4571, exceptionally well resolved (Sandage \& Bedke 1988). The resolution of NGC\,4501 is about intermediate between the two last-mentioned galaxies and the poorly resolved NGC\,4639. A first rough distance of NGC\,4501 is provided by the TF method (Section~\ref{sec:4}) which gives $(m-M)=31.5\pm0.4$ in good agreement with the preferred mean Virgo cluster distance. However, the inherent errors of the TF method if applied to individual galaxies prevent a stringent test. NGC\,4501 is therefore an interesting candidate for Cepheid observations. A preliminary Cepheid distance of the Virgo cluster is obtained by taking the Cepheid distance of the Leo group of $(m-M)=30.20\pm0.12$, based now on three galaxies with Cepheids from {\sl HST} (Saha et~al. 1999), and to step up this value by the modulus difference of $\Delta(m-M)=1.25\pm0.13$ (Tammann \& Federspiel 1997) between the Leo group and the Virgo cluster. The result is $(m-M)_{\rm Virgo}=31.45\pm0.21$, a value which is well embraced by the individual Cepheid distances in Table~\ref{tab:Virgo_Cepheid_distances}. \section{The Virgo cluster distance from Supernovae type Ia} \label{sec:3} Blue SNe\,Ia (i.e. $B_{\max}-V_{\max} \le 0.20$) at maximum light are nearly perfect standard candles. After small corrections for second parameters (decline rate and color) their scatter about the mean Hubble line amounts to only $0\mag12$ for $v > 10\,000\nobreak\mbox{$\;$km\,s$^{-1}$}$ (Parodi et~al. 1999). The equally corrected absolute peak magnitude of $M_{\rm B}^{\rm corr}=-19.44\pm0.04$, based on seven\,(!) SNe\,Ia with known Cepheid distances (Parodi et~al. 1999), is secure. The value is also in perfect agreement with present theoretical models (Branch 1998). Four blue SNe\,Ia with complete photometry are known in the Virgo cluster (Table~\ref{tab:SNeIa_Virgo}). The photometric data from the Tololo/Calan survey are compiled by Parodi et~al. (1999). The magnitudes, corrected for second parameters ($\Delta m_{15}$ and color), are calculated from equation~(30) in Parodi et~al. (1999). \begin{table}[t] \footnotesize\rm \begin{center} \caption{Blue SNe\,Ia with good photometry in the Virgo cluster} \label{tab:SNeIa_Virgo} \begin{tabular}{llrrrl} \noalign{\smallskip} \hline \noalign{\smallskip} SN & Galaxy & $m_{\rm B}$ & $m_{\rm V}$ & $\Delta m_{15}$ & $m_{\rm B}^{\rm corr}$ \\ \noalign{\smallskip} \hline \noalign{\smallskip} 1981B & NGC\,4536 & 12.04 & 11.96 & 1.10 & 11.89 \\ 1984A & NGC\,4419 & 12.45 & 12.26 & 1.20 & 12.06 \\ 1990N & NGC\,4639 & 12.76 & 12.70 & 1.03 & 12.68 \\ 1994D & NGC\,4526 & 11.86 & 11.87 & 1.27 & 11.81 \\ 1960F & NGC\,4496A & 11.60 & 11.54 & \multicolumn{1}{c}{---} & --- \\ \noalign{\smallskip} \hline \noalign{\smallskip} \multicolumn{2}{c}{mean:} & & & & $12.11\pm0.20$ \\ \noalign{\smallskip} \hline \noalign{\smallskip} \multicolumn{6}{l}{The spectroscopically unusual SN\,1991T is omitted.} \end{tabular} \end{center} \end{table} With the above calibration for $M_{\rm B}^{\rm corr}$ and the mean value of $m_{\rm B}^{\rm corr}$ in Table~\ref{tab:SNeIa_Virgo} the cluster modulus becomes $(m-M)_{\rm Virgo}=31.55\pm0.20$. The error of the result is dominated by the depth effect of the cluster. Additional SNe\,Ia in the cluster will improve the result considerably. Its present advantage over the Cepheid distance of the cluster is that the four SNe\,Ia have been discovered independently of their position within the cluster. To increase the sample one additional blue SNe\,Ia in the Virgo cluster may be added which, however, has no known decline rate $\Delta m_{15}$ (cf. Table~\ref{tab:SNeIa_Virgo}). But since the Virgo SNe\,Ia and the calibrating SNe\,Ia (except one) lie in spirals the second-parameter correction cancels in first approximation. {\em Eight\/} SNe\,Ia with Cepheid distances (including SN\,1960F without $\Delta m_{15}$) give a straight calibration of $M_{\rm B}=M_{\rm V}=-19.48\pm0.04$ (Saha et~al. 1999), while the five Virgo SNe\,Ia in Table~\ref{tab:SNeIa_Virgo} have a mean (uncorrected) apparent magnitude of $m_{\rm B}=12.14\pm0.21$, $m_{\rm V}=12.07\pm0.20$. From this follows a mean Virgo cluster modulus of $(m-M)_{\rm Virgo}=31.59\pm0.15$ which is essentially undistinguishable from the result of the four corrected SNe\,Ia. \section {The Virgo cluster distance from 21cm-line widths} \label{sec:4} The method using 21~cm line-widths, the so-called Tully-Fisher (TF) relation, has been applied many times but with variable success. Widely divergent values are in the literature which in some cases favor the short distance scale (e.g. Pierce \& Tully 1992 with $m - M = 30.9$) and in others the long scale (Kraan-Korteweg et al.~1988 with $m - M = 31.6$; Fouqu{\'e} et~al. 1990 with the same value if corrected to the modern local calibrators; Federspiel et~al. 1998). It has been shown (Federspiel et al.~1994; Sandage, Tammann, \& Federspiel 1995) that the reasons for small values of $(m - M)$ for Virgo (the short scale) using TF are two; (1) use of incorrectly small distances to the local calibrators in earlier papers by proponents of the short scale, and (2) neglect of the disastrous effect of the Teerikorpi (1987, 1990) cluster incompleteness bias. It can be shown that this bias produces errors in the modulus up to 1 mag depending on how far one has sampled into the cluster luminosity function {\em regardless how the sample is chosen\/}, if only the sample is cut by apparent magnitude. The modulus error is a strong function of the fraction of the luminosity function that remains unsampled (Kraan-Korteweg et al. 1988; Sandage et al.~1995). The calibration of the TF relation has been much improved by the advent of Cepheid distances with HST. There are now 18 Cepheid distances available for spirals suitable for the calibration. Detailed data with complete references to the extensive literature are given elsewhere (Tammann \& Federspiel 1997; Federspiel et~al. 1998) and are not repeated here. The most recent applications of the TF relation on the Virgo cluster (Schr{\"o}der 1996; Tammann \& Federspiel 1997; Federspiel et~al. 1998) use a {\em complete\/} sample of Virgo cluster spirals. Rigid criteria have been invoked in the selection of members within the cluster boundaries defined by counts, redshifts, and X-ray contours. Several subtleties, not seen in earlier studies, have been found. These include a variation of the derived modulus on the wavelength of the observations (covering UBVRI), and a correlation of the derived modulus on the degree of hydrogen depletion for the spirals. With this in mind Federspiel et~al. (1998) have derived from a complete sample of 49 sufficiently inclined spirals \begin{displaymath} (m-M)_{\rm Virgo}=31.58\pm0.24. \end{displaymath} \section{The Virgo Cluster distance from Globular Clusters} \label{sec:5} Extragalactic GCs, discovered in M\,31 by Hubble (1932), took a role in distance determinations when Racine (1968) proposed the bright end of the globular cluster luminosity function (GCLF) to be used as a ``standard candle''. First applications of this tool provided reasonable distances to M\,87 (Sandage 1968, de Vaucouleurs 1970), yet it was soon realized that the results were sensitive to the GC population size, and that the luminosity $M^{\ast}$ of the turnover point of the bell-shaped GCLF is a much more stable standard candle. This required, however, that one had to sample at least four magnitudes into the GCLF which became feasible only with the advent of CCDs. The first application of the new method to a giant E galaxy (M\,87; van den Bergh et~al. 1985) was followed by many papers such that $m^{\ast}$ magnitudes are now available for about two dozen full-size galaxies (for reviews e.g. Harris 1991; Whitmore 1997; Tammann \& Sandage 1999). The absolute magnitude $M^{\ast}$ of the peak of the globular cluster luminosity function (GCLF), approximated by a Gaussian, can be calibrated independently in the Galaxy and M\,31 through RR\,Lyr stars and Cepheids, respectively. They yield, in perfect agreement, $M^{\ast}_{\rm B} = -6.93\pm0.08$ and $M^{\ast}_{\rm V} = -7.62\pm0.08$ (Sandage \& Tammann 1995; Tammann \& Sandage 1999). Remaining differences between different authors of the luminosity calibration of RR\,Lyr stars are vanishingly small for the mean metallicity of the Galactic GCs of [Fe/H]$=-1.35$, because different adopted luminosity-metallicity relations meet for this value very closely at $M_{\rm V}({\rm RR})=0\mag54$. The calibration of $M^{\ast}$, independently confirmed by the M\,31 Cepheids, is therefore uncontroversial. Different values of $m^{\ast}_{\rm B}$ and $m^{\ast}_{\rm V}$ of bona fide members of the Virgo cluster (cf. Binggeli et~al. 1985) are compiled in Table~\ref{tab:Virgo_GC}. The values have been corrected for the small and variable Galactic absorption according to Burstein \& Heiles (1984). The $g$ magnitudes of Cohen (1988) and the $R$ magnitudes of Ajhar et~al. (1994) were transformed into $V$ magnitudes following Whitmore (1997). No (precarious) attempt was made to transfer $m^{\ast}_{\rm B}$ into $m^{\ast}_{\rm V}$. \begin{table}[t] \footnotesize\rm \begin{center} \caption{Virgo cluster members with known turnover magnitude $m^{\ast}$ of the GCLF} \label{tab:Virgo_GC} \begin{minipage}{0.9\textwidth} \begin{tabular}{lllrc} \noalign{\smallskip} \hline \noalign{\smallskip} Galaxy & \multicolumn{1}{c}{$m^{\ast}_{\rm B}$} & \multicolumn{1}{c}{$m^{\ast}_{\rm V}$} & \multicolumn{1}{c}{$m^{\ast}_{\rm B}$ - $m^{\ast}_{\rm V}$} & Source \\ \noalign{\smallskip} \hline \noalign{\smallskip} NGC\,4365 & $25.18\pm0.16(2)$ & $24.47\pm0.21(1)$ & $0.71\pm0.26$ & (1)\\ NGC\,4374 & & $24.12\pm0.30(1)$ & & (2)\\ NGC\,4406 & & $24.25\pm0.30(1)$ & & (2)\\ NGC\,4472 & $24.70\pm0.11(1)$ & $23.85\pm0.21(2)$ & $0.85\pm0.24$ & (3)\\ NGC\,4486 & $24.82\pm0.11(2)$ & $23.74\pm0.06(5)$ & $1.08\pm0.13$ & (4)\\ NGC\,4552 & & $23.70\pm0.30(1)$ & & (2)\\ NGC\,4636 & & $24.18\pm0.20(1)$ & & (5)\\ NGC\,4649 & $24.65\pm0.14(1)$ & & & (6)\\ \noalign{\smallskip} \hline \noalign{\smallskip} straight mean: & $24.84\pm0.12$ & $24.03\pm0.10$ & & \\ $(m-M)$: & $31.77\pm0.14$ & $31.65\pm0.13$ & & \\ \noalign{\smallskip} $\Rightarrow$ & \multicolumn{2}{l}{$(m-M)=31.70\pm0.10$} & (21.9$\;$Mpc)\\ \noalign{\smallskip} \hline \noalign{\smallskip} \end{tabular} {\footnotesize Sources: (1) Harris et~al. 1991; Secker \& Harris 1993; Forbes 1996 (2) Ajhar et~al. 1994 (3) Harris et~al. 1991; Ajhar et~al. 1994; Cohen 1988 (4) van den Bergh et~al. 1985; Harris et~al. 1991; Cohen 1988; McLaughlin et~al. 1994; Whitmore et~al. 1995; Elson \& Santiago 1996a,b (5) Kissler et~al. 1994 (6) Harris et~al. 1991. -- The values in parentheses in columns~2 and 3 give the number of independent determinations.} \end{minipage} \end{center} \end{table} As seen in Table~\ref{tab:Virgo_GC} the GCLF leads to \begin{displaymath} (m-M)_{\rm Virgo}=31.70\pm0.30, \end{displaymath} which is adopted in the following. For the adopted error see below. The question arises whether it is justified to apply $M^{\ast}$, calibrated in two {\em spiral\/} galaxies, to the GCs of the early-type galaxies in Table~\ref{tab:Virgo_GC}. A positive answer within the errors is provided by the Leo group. Two early-type galaxies in this group give a GCLF modulus of $(m-M)=30.08\pm0.29$ whereas the Cepheids in three spirals of the same group give $(m-M)=30.20\pm0.12$ (Tammann \& Sandage 1999). The formation of GCs not being understood, there is no theoretical reason why the value of $M^{\ast}$ should be universal. It is worrisome that the {\em width\/} of the GCLF varies significantly for different galaxies and that the two brightest galaxies in Table~\ref{tab:Virgo_GC}, NGC\,4486 (M\,87) and NGC\,4472 (M\,49), as well as NGC\,4552 (M\,89) have $m_{\rm V}^{\ast}$ $0\mag5$ {\em brighter\/} than the remaining four galaxies. Moreover, the color $m_{\rm B}^{\ast} - m_{\rm V}^{\ast}=1.08$ of NGC\,4486 is exceptionally red, and its GCLF seems to be bimodal. Finally it is alarming that the mean GCLF modulus of seven early-type Fornax cluster members is $0\mag54\pm0\mag15$ {\em smaller\/} than the secure cluster distance from three blue SNe\,Ia, the latter value being also supported by the relative distance from secondary distance indicators between the early-type members of the Virgo and Fornax clusters (Tammann \& Sandage 1999; cf. Fig.~\ref{fig:schematic_Ho_all} below). In addition the GCLF distances of some individual field galaxies are highly questionable. Account of these problems is taken by assigning a relatively large error to the GCLF distance of the Virgo cluster. \section{The Virgo cluster distance from the D\boldmath$_{n}-\sigma$ relation} \label{sec:6} It has been shown that the well known D$_{\rm n}-\sigma$ relation of early-type galaxies (Dressler et~al. 1987) applies also to the bulges of early-type spiral and S0 galaxies with surprisingly small scatter (Dressler 1987). Here D$_{\rm n}$ is an isophotally defined galaxy diameter (in arcsec), i.e. the diameter of a circle which encompasses a mean surface brightness of 19.75 B\,mag arcsec$^{-2}$, and $\sigma$ is the aperture-dependent, normalized velocity dispersion $\sigma$ (in km\,s$^{-1}$) in the bulge. The isophotal values of D$_{\rm n}$ are, of course, affected by front absorption $A_{\rm B}$. The corresponding correction amounts to $\Delta \log D=0.32 A_{\rm B}$ (Lynden-Bell et~al. 1988), which translates into a distance effect of $\Delta (m-M)=1.6 A_{\rm B}$. The absorption has paradoxically thus a stronger effect on this diameter distance than on a luminosity distance. From a sample of 26 S0--Sb Virgo cluster members Dressler (1987) has determined \begin{equation} \label{equ:Dn_sigma} \log D_{\rm n}=1.333 \log \sigma - (1.572\pm0.014) \end{equation} with a scatter of $\sigma(\log D_{\rm n})=0.06$, corresponding to $\sigma(m-M)=0\mag34$. The two deviating galaxies NGC\,4382 and NGC\,4417 were excluded. It is somewhat worrisome that equation~(\ref{equ:Dn_sigma}) is based on only 24 galaxies, i.e. one third of the total Virgo population of S0--Sb galaxies. This may invite selection bias. However, the brighter half of the sample yields the same distance modulus as the fainter one to within $0\mag07\pm0\mag14$. Moreover, the missing Virgo members are on average fainter (smaller) at any given value of $\sigma$ than the sample of 24, and they can make the constant term in equation~(\ref{equ:Dn_sigma}) only more negative, which would in any case {\em increase\/} the cluster distance. Yet for another reason the Virgo distance derived from equation~(\ref{equ:Dn_sigma}) may be somewhat low. As discussed below (Section~\ref{sec:8}) the Virgo cluster consists of two main concentrations A and B, B being, if anything, slightly more distant. The Virgo distance quoted throughout refers to the mean of {\em all\/} cluster members and therefore depends on a fair representation of A and B. In the case of equation~(\ref{equ:Dn_sigma}) 21 galaxies lie in A and only 3 in B, which is an overrepresentation of A even if one allows for the smaller size of B. The Virgo cluster distance can be derived from equation~(\ref{equ:Dn_sigma}) if the {\em linear\/} diameters D$_{\rm n}$ at given $\sigma$ are known. As calibrators have been used M\,31 and M\,81 (Dressler 1987; Sandage \& Tammann 1988) as well as the bulge of our Galaxy (Terndrup 1988); all three galaxies have been combined by Tammann (1988). The Galaxy, as a calibrator, may be of somewhat later type than the S0--Sb sample, but also the Sbc galaxy NGC\,4501 fits well equation~(\ref{equ:Dn_sigma}) (Dressler 1987). In principle the Virgo cluster distance has to be known to match the velocity dispersion $\sigma$ of the calibrators to the aperture size ($5'' \times 5''$) that was applied for Virgo. However, the remaining mismatch should introduce only negligible systematic errors (cf. Dressler 1987). The derivation of the cluster distance is repeated in Table~\ref{tab:Virgo_distance_D_sigma} with updated Cepheid distances and absorption values for M\,31 and M\,81. The Table is self-explanatory. \begin{table}[t] \footnotesize \begin{center} \caption{The local calibration of the D$_{\rm n}-\sigma$ relation of the bulges of early-type spirals and S0 galaxies} \label{tab:Virgo_distance_D_sigma} \begin{minipage}{\textwidth} \begin{tabular}{lccccrc} \noalign{\smallskip} \hline \noalign{\smallskip} Calibrator & $(m-M)^0$ & $\log$ D$_{\rm n}$ & $\sigma(\nobreak\mbox{$\;$km\,s$^{-1}$})$ & $\log$\,D$_{\rm n}$\,(eq.\,\ref{equ:Dn_sigma}) & $\Delta(m-M)$ & $(m-M)^0$ \\ \noalign{\smallskip} \hline \noalign{\smallskip} Gal. Bulge & $14.46\pm.20^{1)}$ & $4.74\pm.04^{1)}$ & $124\pm9^{1)}$ & $1.22\pm0.06$ & $17.60\pm0.30$ & $32.06\pm0.36$ \\ M\,31 & $24.44\pm.20^{2)}$ & $2.83\pm.04^{4)}$ & $150\pm5^{6)}$ & $1.33\pm0.04$ & $7.50\pm0.20$ & $31.94\pm0.28$ \\ M\,81 & $27.80\pm.20^{3)}$ & $2.15\pm.03^{5)}$ & $166\pm8^{6)}$ & $1.39\pm0.04$ & $3.80\pm0.19$ & $31.60\pm0.28$ \\ \noalign{\smallskip} \hline \noalign{\smallskip} \multicolumn{4}{l}{weighted mean:} & \multicolumn{3}{r}{$31.85\pm0.17$} \\ \noalign{\smallskip} \hline \noalign{\smallskip} \end{tabular} (1) Terndrup 1988 (2) Madore \& Freedman 1991 (3) Freedman \& Madore 1994 (4) Dressler 1987. D$_{\rm n}$ corrected for bulge absorption of $A_{\rm B}=0.33$; cf. text (5) Dressler 1987. D$_{\rm n}$ corrected for front absorption of $A_{\rm B}=0.16$ (Freedman \& Madore 1994); cf. text (6) Dressler 1987. \end{minipage} \end{center} \end{table} The method is so promising that one would hope for additional bulge data of local calibrators as well as remaining Virgo cluster members of the appropriate Hubble types. At present we adopt $(m-M)_{\rm Virgo}=31.85\pm0.17$ as the best D$_{\rm n}-\sigma$ distance from intermediate-type galaxies. The classical D$_{\rm n}-\sigma$ relation of {\em early-type\/} (E/S0) galaxies encounters the difficulty of lacking local calibrators. The only way is to use the two early-type members of the Leo group and to adopt the mean Cepheid distance of $(m-M)=30.20\pm0.12$ of the three spirals in the same group. Faber et~al. (1987) find the modulus difference between the Virgo cluster and the Leo group to be $\Delta(m-M)=0.97\pm0.29$ from the D$_{\rm n}-\sigma$ relation. This value is somewhat suspicious because it is significantly smaller than from four other relative distance indicators (cf. Tammann \& Federspiel 1997). But taken the modulus difference at face value, one obtains $(m-M)_{\rm Virgo}=31.17\pm0.31$ from E/S0 galaxies. The weighted mean distance modulus of the intermediate-type and early-type Virgo cluster members becomes then \begin{displaymath} (m-M)_{\rm Virgo}=31.70\pm0.15. \end{displaymath} \section{The Virgo cluster distance from Novae} \label{sec:7} Pritchet \& van~den Bergh (1987) found from six novae in Virgo cluster ellipticals that they are $7\mag0\pm0\mag4$ more distant than the {\em apparent\/} distance modulus of M\,31 of $(m-M)_{\rm AB}=24.58\pm0.10$ from Cepheids (Madore \& Freedman 1991) {\em and\/} Galactic novae (Capaccioli et~al 1989). Livio (1997) found from a semi-theoretical analysis of the six Virgo novae $(m-M)_{\rm Virgo}=31.35\pm0.35$. A low-weight mean of $(m-M)_{\rm Virgo}=31.46\pm0.40 $ is adopted. \section{Conclusions} \label{sec:8} The results of Sections~2\,-\,7 are compiled in Table~\ref{tab:Virgo_distance_compilation}. The individual values lead to a weighted mean distance modulus of $(m-M)_{\rm Virgo}=31.60\pm0.09$, corresponding to a distance of $r=20.9\pm0.9\;$Mpc. \begin{table}[t] \footnotesize \begin{center} \caption{Compilation of the different Virgo cluster moduli} \label{tab:Virgo_distance_compilation} \begin{tabular}{lllll} \noalign{\smallskip} \hline \noalign{\smallskip} Method & $(m\!-\!M)_{\rm Virgo}$ & Type & Calibration & Source \\ \noalign{\smallskip} \hline \noalign{\smallskip} Cepheids & $31.45 \pm 0.21$ & S & $(m\!-\!M)_{\rm LMC}=18.50$ & Tammann \& Sandage 1999 \\ Supernovae Ia & $31.55 \pm 0.20$ & S & Cepheids & Tammann \& Reindl 1999 \\ Tully-Fisher & $31.58 \pm 0.24$ & S & Cepheids & Federspiel et~al. 1998 \\ Globular Clusters & $31.70 \pm 0.30$ & E & RR\,Lyr, Cepheids & Tammann \& Sandage 1999 \\ D$_{\rm n} - \sigma$ & $31.70 \pm 0.15$ & E,\,S0,\,S & Galaxy, Cepheids & Dressler 1987 \\ Novae & $31.46 \pm 0.40$ & E & M\,31 (Cepheids) & Pritchet \& van\,den Bergh$\;$1987 \\ \noalign{\smallskip} \hline \noalign{\smallskip} Mean: & $31.60\pm0.09$ & \multicolumn{3}{l}{($\Rightarrow 20.9\pm0.9\;$Mpc)}\\ \noalign{\smallskip} \hline \noalign{\smallskip} \end{tabular} \end{center} \end{table} It is remarkable that the individual distance determinations agree to within their mean internal errors. The result gains additional weight by the fact that it is based on spiral as well as early-type cluster members. The zeropoint of the distance determination, as seen in Table~\ref{tab:Virgo_distance_compilation}, rests mainly, but not exclusively on Cepheids, and hence on the adopted LMC distance. The Virgo cluster shows clear subclustering. There are two major clumpings A (with M\,87) and B (with M\,49) (Binggeli et~al. 1985, 1993) as well as a concentration around M\,86 (Schindler, Binggeli, \& B{\"o}hringer 1999). There is rather strong evidence from the TF method that cluster B is more distant than A by $\sim\!0\mag46\pm0.18$ (Federspiel et~al. 1998). However, at present it is save to include {\em all\/} cluster members as defined by Binggeli et~al. (1993) and to quote a single mean distance of the common gravitational well. Future determinations of the Virgo cluster distance may include the brightness of the tip of the red-giant branch (TRGB). A first experiment is available (Harris et~al. 1998), and if a sufficient number of cluster members will become available to beat the cluster depth effect the method may become competitive. It has been argued many times that the recession velocity of the Virgo cluster is too small to yield the cosmic value of the Hubble constant $H_0$. Indeed the observed mean cluster velocity of $v_0=920\pm35\nobreak\mbox{$\;$km\,s$^{-1}$}$, corrected to the centroid of the Local Group (Binggeli et~al. 1993) and combined with the above cluster distance would provide too low a value of $H_0=44$ [km\,s$^{-1}\;$Mpc$^{-1}$] mainly due to the gravitational deceleration by the Virgo complex. The Virgocentric pull has decelerated the Local Group's recession velocity by $200-250\nobreak\mbox{$\;$km\,s$^{-1}$}$ (Kraan-Korteweg 1986; Tammann \& Sandage 1985; Jerjen \& Tammann 1993). $H_0$ becomes then rather 54 with some leeway as to remaining peculiar velocity effects. However the problem of the Virgo cluster velocity can be entirely circumvented by using the distances of clusters out to $10\,000\nobreak\mbox{$\;$km\,s$^{-1}$}$ {\em relative\/} to the Virgo cluster (Sandage \& Tammann 1990; Jerjen \& Tammann 1993; Giovanelli 1997). The exquisite quality of these relative distances is shown by their defining a Hubble line of slope 0.2 with very small scatter (Fig.~\ref{fig:Hubble_relative_distances}). \def3.18cm{0.665\textwidth} % \begin{figure}[t] \begin{center} \centerline{\psfig{file=13_gtammann1_1.eps,width=3.18cm}} \caption{Hubble diagram of 31 clusters with known relative distances. Asterisks are data from Jerjen \& Tammann (1993). Open circles are from Giovanelli (1997). Filled circles are the average of data from both sources. Velocities of $\ge3000\nobreak\mbox{$\;$km\,s$^{-1}$}$ are corrected for a local CMB anisotropy of $620\nobreak\mbox{$\;$km\,s$^{-1}$}$.} \label{fig:Hubble_relative_distances} \end{center} \end{figure} A linear regression through the points in Fig.~\ref{fig:Hubble_relative_distances} with forced slope of 0.2, corresponding to linear expansion, gives \begin{equation} \label{equ:log_v_relative} \log v = 0.2\,[(m-M)-(m-M)_{\rm Virgo}] + (3.070\pm0.011). \end{equation} From this follows directly \begin{equation} \label{equ:log_H_relative} \log H_0 = -0.2\,(m-M)_{\rm Virgo} + (8.070\pm0.011). \end{equation} Inserting the mean Virgo cluster modulus from Table~\ref{tab:Virgo_distance_compilation} yields \begin{displaymath} \label{equ:H0_Dn_sigma)} H_0 = 56\pm4, \end{displaymath} where the additional external error is generously estimated to be $\pm6$. The route to the Virgo cluster distance and to $H_0$ is schematically summarized in Fig.~\ref{fig:schematic_Ho_all}. \begin{figure}[t] \begin{center} {\footnotesize \sf \def3.18cm{3.18cm} \setlength{\unitlength}{1mm} \def120{120} \def151{151} \begin{picture}(120,151)(0,0) \put(0,0){\line(120,0){120}} \put(0,0){\line(0,151){151}} \put(0,151){\line(120,0){120}} \put(120,0){\line(0,151){151}} \put(110,146){\makebox(0,0)[t]{\normalsize \bf H\boldmath$_0$ $\approx$ 58}} \put(53,146){\makebox(0,0)[t]{\normalsize \bf H\boldmath$_0$ $=$ 56$\pm$4}} \put(53,138.5){\vector(0,0){3.5}} \put(40.5,134){\fbox{\parbox{2.2cm}{\centering Clusters out to \\ 10\,000\nobreak\mbox{$\;$km\,s$^{-1}$}}}} \put(36,91){\fbox{\psfig{file=13_gtammann1_1.eps,width=3.18cm}}} \put(39,86.5){Fig.~1} \put(53,121){\vector(0,0){10}} \put(53,80.3){\vector(0,0){9.5}} \put(36,74.9){\vector(-3,2){15}} \put(70.4,72){\vector(1,1){7.4}} \put(16,104){\makebox(0,0)[t]{\normalsize \bf (H\boldmath$_0$ $=$ 60$\pm$6)}} \put(16,96.5){\vector(0,0){3.5}} \put(1,90){\fbox{\parbox{2.6cm}{\centering Coma Cluster \\ (m-M)$=$35.29$\pm$0.11 \\ 114$\pm$6$\;$Mpc}}} \put(36,72){\fboxrule=1pt \fbox{\parbox{3.2cm}{\centering \vspace*{5pt} Virgo Cluster \\ (m-M)$=$31.60$\pm$0.09 \\ 20.9$\pm$0.9$\;$Mpc \vspace*{5pt}}}} \put(92,92){\makebox(0,0)[t]{\normalsize \bf (H\boldmath$_0$ $=$ 58$\pm$10)}} \put(92,84.5){\vector(0,0){3.5}} \put(78,78){\fbox{\parbox{2.6cm}{\centering Fornax Cluster \\ (m-M)$=$31.80$\pm$0.09 \\ 22.9$\pm$1.0$\;$Mpc}}} \put(58,56){\fbox{\parbox{0.8cm}{\centering Leo \\ Group}}} \put(62.5,61){\vector(-1,4){1}} \put(14,56.7){\vector(3,2){22}} \put(28,57.4){\vector(1,1){8}} \put(40,58){\vector(1,3){2.3}} \put(75.3,57.5){\vector(-2,3){5}} \def10{51} \put(10,58){\vector(0,0){27}} \put(10,58){\vector(0,0){27}} \put(10,10){\circle{40}} \put(10,10){\makebox(0,0)[c]{\parbox{1cm}{\centering \scriptsize Globular \\ clusters}}} \put(25,10){\circle{40}} \put(25,10){\makebox(0,0)[c]{\parbox{1cm}{\centering \scriptsize Novae}}} \put(40,10){\circle{40}} \put(40,10){\makebox(0,0)[c]{\parbox{1cm}{\centering \scriptsize D$_{\rm n}-\sigma$ \\ relation}}} \put(25,36.5){\vector(0,0){7.5}} \put(22,36.5){\vector(-1,1){8.3}} \put(28,36.5){\vector(1,1){8.3}} \put(78,10){\circle{40}} \put(78,10){\makebox(0,0)[c]{\parbox{1cm}{\centering \scriptsize TF \\ relation}}} \put(93,56.7){\vector(-3,2){22.3}} \put(97,58){\vector(0,0){15.0}} \put(97,10){\circle{40}} \put(97,10){\makebox(0,0)[c]{\parbox{1cm}{\centering \scriptsize SNe\,Ia}}} \def10{30} \put(15,10){\fbox{\parbox[c]{2.6cm}{\centering M\,31 \\ (m-M)$=$24.44$\pm$0.15 \\ 773$\pm$55$\;$kpc}}} \put(68,28){\fbox{\parbox{2.6cm}{\centering LMC \\ (m-M)$=$18.50$\pm$0.04 \\ 50.1$\pm$1.0$\;$kpc}}} \def10{10} \put(10,17){\vector(0,0){27}} \put(15.8,14.4){\vector(4,1){52}} \put(19.8,15.4){\vector(-4,-1){4}} \put(10,10){\circle{60}} \put(10,10){\makebox(0,0)[c]{\parbox{1cm}{\centering \scriptsize RR\,Lyr \\ stars}}} \def53{53} \put(53,64){\vector(0,0){1}} \put(53,62){\circle*{0.5}} \put(53,60){\circle*{0.5}} \put(53,58){\circle*{0.5}} \put(53,56){\circle*{0.5}} \put(53,54){\circle*{0.5}} \put(53,52){\circle*{0.5}} \put(53,50){\circle*{0.5}} \put(53,48){\circle*{0.5}} \put(53,46){\circle*{0.5}} \put(53,44){\circle*{0.5}} \put(53,42){\circle*{0.5}} \put(53,40){\circle*{0.5}} \put(53,38){\circle*{0.5}} \put(53,36){\circle*{0.5}} \put(53,34){\circle*{0.5}} \put(53,32){\circle*{0.5}} \put(53,30){\circle*{0.5}} \put(53,28){\circle*{0.5}} \put(53,26){\circle*{0.5}} \put(53,24){\circle*{0.5}} \put(53,22){\circle*{0.5}} \put(53,20){\circle*{0.5}} \put(53,18){\circle*{0.5}} \put(47,14){\vector(-3,2){16}} \put(59,14){\vector(3,2){13}} \put(60.5,15){\vector(-3,-2){1.5}} \put(54,17){\vector(2,3){19}} \put(53.4,17){\vector(1,4){8.9}} \put(57.2,16){\line(4,3){10}} \put(82.8,35.2){\vector(4,3){12}} \put(53,10){\circle{60}} \put(53,10){\makebox(0,0)[c]{\scriptsize Cepheids}} \put(98.6,41.1){\vector(-1,3){1}} \put(98.5,41.2){\circle*{0.5}} \put(99.1,39.4){\circle*{0.5}} \put(99.7,37.6){\circle*{0.5}} \put(100.3,35.8){\circle*{0.5}} \put(100.9,34.0){\circle*{0.5}} \put(101.5,32.2){\circle*{0.5}} \put(102.1,30.4){\circle*{0.5}} \put(102.7,28.6){\circle*{0.5}} \put(103.3,26.8){\circle*{0.5}} \put(103.9,25){\circle*{0.5}} \put(104.5,23.2){\circle*{0.5}} \put(105.1,21.4){\circle*{0.5}} \put(105.7,19.8){\circle*{0.5}} \put(106.3,18){\circle*{0.5}} \put(110,17){\vector(0,0){124}} \put(104,14){\vector(-3,2){13}} \put(110,10){\circle{60}} \put(110,10){\makebox(0,0)[c]{\parbox{1cm}{\centering \scriptsize Physical \\ methods}}} \end{picture} } \end{center} \caption{Schematical presentation of the distance determination of the Virgo cluster} \label{fig:schematic_Ho_all} \end{figure} The figure also shows the distance modulus of the Fornax cluster. The modulus holds strictly only for the early-type cluster members. The Cepheid distance of the large {\em spiral\/} NGC\,1365 of $(m-M)= 31.35$ (Madore et~al. 1998) proves it to lie on the near side of the cluster. Also other cluster spirals may be at relatively small distances because a compilation of about 30 determinations by various methods and authors of the relative distance between the Fornax and Virgo clusters suggests that the Fornax spirals are nearer on average than the Fornax E/S0 members by $0.35\pm0.10$ (Tammann \& Federspiel 1997). In any case the Fornax cluster distance is not helpful for the determination of $H_0$ because the observed mean cluster redshift of $\sim\!1300\nobreak\mbox{$\;$km\,s$^{-1}$}$ carries an uncertainty of $\sim\!20\%$ due to its totally unknown peculiar motion. The Coma cluster distance in Figure~\ref{fig:schematic_Ho_all} is not either very helpful for the determination of $H_0$. Its main weight hinges on distances relative to the Virgo cluster and the cluster contributes therefore only a limited amount of independent evidence. Moreover, its observed mean velocity of $6900\nobreak\mbox{$\;$km\,s$^{-1}$}$ may be affected by local streaming velocities at the level of about 10\%. The cluster lies probably outside the large local bubble which moves with $630\nobreak\mbox{$\;$km\,s$^{-1}$}$ toward the warm pole of the MWB (Smooth et~al. 1991), but it seems plausible that it has a similarly large peculiar motion by its own. The value of $H_0\approx58$, also shown in Fig.~\ref{fig:schematic_Ho_all}, is based on purely physical distance determinations, i.e. gravitationally lensed quasars, the Sunyaev-Zeldovich effect, and CMB fluctuations. The reader is referred to the abstract by G. Theureau \& G.A. Tammann in these Conference Proceedings where the original authors are quoted. \section*{Acknowledgments} G.A.T. and B.R. thank the Swiss National Science Foundation for financial support. \section*{References}
\section{Introduction} One of the oldest and most fundamental problem of physics and astrophysics is that of gravitational collapse, and, specifically, that of the ultimate fate of a sufficiently massive collapsing body. Most of the astrophysical objects that we know of, viz. galaxies, stars, White Dwarfs (WD), Neutron Stars (NS), in a broad sense, result from gravitational collapse. It is well known that the essential concept of BHs, in a primitive form, was born in the cradle of Newtonian gravitation\cite{1}. In Newtonian gravity, the mass of a collapsing gas cloud is constant even if it is radiating, and equal to its baryonic mass $M_i=M_f=M_0$. Thus as $r$ decreases, natuarally, the value of $M/r=M_0/r$ steadily increases and at a certain stage one would have $2GM/r=c^2$ when the ``escape velocity'' from the surface of the fluid becomes equal to the speed of light $c$ (henceforth $G=c=1$). Soon afterwards, a Newtonian BH would be born. And in the context of classical General Theory of Relativity (GTR), it is believed that the ultimate fate of sufficiently massive bodies is collapse to a Black Hole (BH). In contrast to Newtonian gravity, here, the gravitational mass of a radiating fluid constantly decrease and therefore unlike the Newtonian case, one can not predict with real confidence the actual value of $M_f$ when one would have $2M_f/r =1$. Given an initial gravitational mass $M_i$, the three quantities, $M_i$, $M_f$ and $M_0$ are not connected amongst themselves by means of any fundamental constants or by any basic physical principles. Thus, in a strict sense, one reqires to solve the Einstein equations for the collapsing fluid for realistic and ever evolving equation of state (EOS) and radiation transport properties. Unfortunately, the reality is that even when one does away with the EOS by assuming the fluid to be a dust whose pressure is zero everywhere including the center, there is no unique solution to the problem. Depending upon the initial conditions like density distributions and assumptions, like self-similarity, adopted, one may find either a BH or a ``naked singularity''\cite{2}. It is only when the dust is assumed to be homogeneous, a unique solution can be found\cite{3}. Homogeneous or inhomogeneous, all dust balls have a very special property: they have no internal energy and they can not radiate. Consequuenly, the graviational mass of a dust remains constant during the collapse process, which is, essentially, a Newtonian property. And if the dust collapse starts from a state of infinite dilution at $r=r_\infty =\infty$, the graviational mass of a dust must be equal to its baryonic mass $M=M_0$, another Newtonian property. Having made these remarks, we would now proceed to self-consistently analyze the pioneering work of Oppenheimer and Snyder (OS) to determine the mass of the BH whose formation it suggested. Since the OS solutions are the only (asymptotic and near) exact solutions for the GTR collapse, and are believed to explicitly show the formation of an ``event horizon'' (EH), it is extremely important to critically reexamine them. The study of any collapse problem becomes more tractable if one uses the comoving coordinates which are free from any kind of ``coordinate singularities''. By definition, for a given fluid element, the comoving coordinate $R$ is fixed. The most natural choice for $R$ is the number of baryons within a given mass shell $N(R)$ or any number proportional to it. The comoving time $\tau$ is of course the time recordrd by a clock attached to a fluid element at $R=R$. Since a dust is in perennial free fall, comoving time is synonymous with ``proper time'' and, therefore, the dust metric is \begin{equation} ds^2=d\tau^2 -e^{\bar \omega} dR^2 - e^\omega(d\theta^2 +\sin^2 \theta d\phi^2) \end{equation} It turns out that, for any spherically symmetric metric, the angular part is the same and we have \begin{equation} e^\omega= r^2 \end{equation} where $r$ is the invariant circumference radius, and this was the Eq. (27) in the OS paper. OS tried to solve the collapse equation by using the metric (2); and by skipping several intermediate equations, we shall focus attention on the solutions they oobtained. The Eq. (23-27) of their paper essentially leads to a relationship between $\tau$ and $r$: \begin{equation} \tau ={2\over 3}{R^{3/2} -r^{3/2}\over (R/R_b)^{3/2} r_0^{1/2}}; \qquad R\le R_b \end{equation} By transposing, the foregoing equation yields \begin{equation} {r\over R} =\left(1-{3\over 2} {r_0^{1/2} \tau\over R_b^{3/2}}\right)^{2/3}; \qquad R\le R_b \end{equation} Apart from the the comoving coordinate system, there is another useful coordinate system, the Schwarzschild system \begin{equation} ds^2 = e^\nu dt^2 -e^\lambda dr^2 -r^2 (d\theta^2 +\sin^2 \theta d\phi^2) \end{equation} and, in principle, it should be possible to work out the collapse (or any) problem in this coordinate system. The time $t$ appearing in the Schwarzschild cooordinate system is the proper time of a distant inertial observer. Further, by definition, the ``comoving'' coordinates can not be extended beyond the boundary of the fluid. Thus from either consideratiion, the external solutions must be expressed in terms of Schwarzschild system. OS obtained a general form of the Schwarzschild metric coefficients which involved derivatives, ${\dot r} = {\partial r/\partial \tau}$ and ${\dot t}={\partial t/\partial \tau}$ : \begin{equation} -g_{rr} = e^\lambda =(1-{\dot r}^2)^{-1} \end{equation} and \begin{equation} g_{tt} =e^\nu = {\dot t}^{-2} (1- {\dot r})^2 \end{equation} Again by skipping few intermediate steps, we note that obtained a general relationship between the coordinate time $t$ and coordinate radius $R$ ifor the region inside the collapsing body: \begin{equation} t = {2\over3} r_0 (R_b^{3/2} -r_0^{3/2} y^{3/2}) -2 r_0 y^{1/2} + r_0 \ln {y^{1/2} +1 \over y^{1/2} -1} \end{equation} where \begin{equation} y \equiv{1\over 2} \left[ (R/R_b)^2 -1\right] + {R_b r\over r_0 R} \end{equation} It is the above Eq. (8) which corresponds to Eq. (36) in the OS paper, and, for very large $t$, it attains the form \begin{equation} t = r_0 \ln {y^{1/2} +1 \over y^{1/2} -1} \end{equation} Now we trace some of the intermediate steps used by OS and not contained in their paper. By using the simple relation $\ln(m/n) = -\ln(n/m)$, one may rewrite the above equation as \begin{equation} t =- r_0 \ln {y^{1/2} -1 \over y^{1/2} +1} \end{equation} In the limit of large $t$ and $y\rightarrow 1$, the above equation becomes \begin{equation} t = r_0 \ln \left({y-1\over 4}\right) \end{equation} Or, \begin{equation} y-1 = 4 e^{-t/r_0} \end{equation} However, OS overlooked the numerical factor of ``4'' in their exercise. From Eqs. (10) and (14), we find that \begin{equation} y-1 = {1\over 2} \left[(R/R_b)^2 -3\right] + {R_b r\over r_0 R} = 4e^{-t/r_0} \end{equation} The following equation obtained from the foregoing one is used to eliminate $r$ from relevant equations \begin{equation} {R_b r\over r_0 R} = 4e^{-t/r_0} - {1\over 2} \left[(R/R_b)^2 -3\right] \end{equation} And using Eqs. (15) and (16) into (13) we obtain \begin{equation} t \sim -r_0 \ln \left\{ {1\over 8}\left[ \left({R \over R_b}\right)^2 -3\right] + {R_b \over 4 r_0} \left( 1- {3 r_0^{1/2} \tau \over 2 R_b^{3/2}}\right)^{2/3}\right\} \end{equation} However, the corresponding equation in the OS paper (Eq.[37]) contained two small errors: \begin{equation} t \sim -r_0 \ln{ \left\{ {1\over 2}\left[ \left({R \over R_b}\right)^2 - 3\right] + {R_b \over 2r_0} \left( 1- {3 r_0^{1/2} \tau \over 2 R_b^2}\right)^{2/3}\right\}} \end{equation} While the first error is a genuine one; which is due to the omission of the factor 4, the second one is a typographical error : the power of $R_b$ in the numerator of the last term of this equation should be $3/2$ and not $2$. From the foregoing equation, they concluded that, ``for a fixed value of $r$ as $t$ tends toward infinity, $\tau$ tends to a finite limit, which increases with $r$''. The fact that $\tau$ is finite at $r=r_0$ or $r=0$ is evident from Eq. (4) too (provided $r_b \neq \infty$ and $r_0 >0$). This was essentially the idea behind the occurrence of an Event Horizon (EH). By differtiating Eq. (16) with respect to $\tau$, we obtain \begin{equation} {\dot t}= {e^{t/r_0}\over 4} \left({r_0 R\over r R_b}\right)^{1/2} \end{equation} Similarly differentiating Eq. (15), and using the above result, we obtain \begin{equation} {\dot r} = {R^{3/2} r_0^{1/2}\over R_b^{3/2} r^{1/2}} \end{equation} Now using Eqs. (15), (19) and (20) in Eqs. (7-8), we obtain \begin{equation} e^{-\lambda} = 1 -(R/R_b)^2 \left\{4 e^{-t/r_0} +{1\over 2} \left[3 - (R/R_b)^2\right]\right\}^{-1} \end{equation} and \begin{equation} e^\nu = e^{\lambda- 2 t/r_0} \left\{ {e^{- t/r_0}\over 4} + {1\over 8} \left[3-(R/R_b)^2 \right]\right\} \end{equation} However, in the paper of OS, these two foregoing equations appear (Eqs. [38-39]) in a slightly erroneous form because of the omission of the nemerical factor of 4 in Eq. (13): \begin{equation} e^{-\lambda} = 1 -(R/R_b)^2 \left\{ e^{-t/r_0} +{1\over 2} \left[3 - (R/R_b)^2\right]\right\}^{-1} \end{equation} and \begin{equation} e^\nu = e^{\lambda- 2 t/r_0} \left\{ e^{- t/r_0} + {1\over 2} \left[3-(R/R_b)^2 \right]\right\} \end{equation} Note these equations were obtained by eliminating $r$ and the $t\rightarrow \infty$ limit covers both the $r\rightarrow r_0$ limit as well as the further $r\rightarrow 0$ limit. Now, if we recall that the comoving coordinates $R$ and $R_b$ are fixed, and, when there is a total collapse to a physical point at $r=0$, the {\em metric coefficients must blow up irrespective of the value of} $R$ (i.e, the the label of a given mass shell). But this {\em does not happen} for the solutions obtained by OS! OS correctly pointed out that, it is only the outer boundary ($R_b$) for which the (one of the) metric coefficients assume the desired form : $e^\lambda \rightarrow \infty$. ``For $R$ equal to $R_b$, $e^\lambda$ tends to infinity like $e^{t/r_0}$ as $t$ approaches infinity. But for any interior point, the limitting values are different, and infact, $e^\lambda = finite$ even when the collapse is supposed to be complete! OS noted ``For $R$ less than $R_b$, $e^\lambda$ tends to a finite limit as $t$ tends to infinity'' {\em without trying to find why it is so}. And on the other hand, when $e^\lambda =finite$, one sees that $e^\nu \rightarrow 0$. But, for the external boundary, for which $e^\lambda \rightarrow \infty$, it can not be predicted whether $e^\nu$ indeed approaches zero (these limits remain unchanged even when one ignores the numerical factor of 4). They admitted this problem while they wrote ``Also for $r \le r_0$, $\nu$ tends to minus infinity''. This $\nu \rightarrow -\infty$ limit would correspond to the singularity $R\rightarrow 0$. However they did not ponder why for $R=R_b$, $e^\nu $ does not behave in the desired manner. Specifically the fact that for $R <R_b$, $e^\lambda$ {\em fails to blow up at the singularity definitely hints that there is some tacit assumption, made in the OS analysis, which is not realized in Nature (GTR) or there is a basic fault in the formulation of the problem}. In our attempt for a possible resolution of this physical anomaly with regard to the unphysical aspect of OS solutions, we see from Eq. (8), that, $t\rightarrow \infty$ if either or both of the two following conditions are satisfied: \begin{equation} r_0 \rightarrow 0; \qquad t\rightarrow \infty \end{equation} and \begin{equation} y \rightarrow 1 ; \qquad t\rightarrow \infty \end{equation} OS {\em implicitly assumed} that $r_0$ is {\em finite} and then $y \rightarrow 1$ and then $y <1$ as $t\rightarrow \infty$ ``For $\lambda$ tends to a finite limit for $r \le r_0$ as $t$ approaches infinity, and for $r_b=r_0$ tends to infinity. Also for $r \le r_0$, $\nu$ tends to to minus infinity.'' But, while doing so, {\bf they completely overlooked the most important feature} of Eq. (8) (their Eq. 36), and Eq. (10) that in view of the presence of the $t \sim \ln {y^{1/2} +1\over y^{1/2} -1}$ term, in order that $t$ {\bf is definable at all}, one must have \begin{equation} y \ge 1 \end{equation} For an insight into the problem, we first focus attention on the outermost layer where $y_b = r_b/r_0$, so that the above condition becomes \begin{equation} r_b \ge r_0 \end{equation} This condition tells that $r_b$ can {\em never plunge below} $r_0$: Thus a careful analysis of the GTR homogeneous dust problem as enunciated by OS themselves actually tell that trapped surfaces can not be formed even though one is free to chase the limit $r\rightarrow r_0$. We may further rewrite this condition as \begin{equation} {r_b\over r_0} \ge 1; \qquad {2 M\over r_b} \le 1 \end{equation} This means that, if the collapse indeed proceeds upto $r_b=0$, the final gravitational mass of the configuration would be \begin{equation} M_f (r=0) = 0 \end{equation} But then, for a dust or any adiabatically evolving fluid \begin{equation} M_i = M_f = constant \end{equation} Therefore, we must have $M_i = 0$ too. This means that $r_0=0$ in this case, and then, it may be promptly verified that, irrespective of the value of $r$, Eqs. (21) and (22) lead to a unique limiting value for the metric coefficients : \begin{equation} e^{-\lambda} \approx 1 - \left(4 e^{-t/r_0} +1\right) \rightarrow 0 \end{equation} or, \begin{equation} e^\lambda \rightarrow \infty \end{equation} and \begin{equation} e^\nu \approx e^{\lambda - 2t/r_0} \rightarrow 0 \end{equation} because $t/r_0$ approaches $\infty$ much faster than $\lambda$, if $r_0=0$. Thus, technically, the final solutions of OS are correct, except for the fact {\bf they did not organically incorporate} the crucial $y \ge 0$ condition in the collapse equations (here we ignore the missing numerical factor of 4). And all we have done here is to rectify this colossal lacunae to fix the value of $r_0=0$. This conclusion that the work of OS demands $r_0=0$ could have obtained in a much more direct fashion simply from the definition of $y$ in Eq. (15). To this effect, we rewrite this equation as \begin{equation} y \equiv {1\over 2} (\alpha^2 -1) + { r\over r_0 \alpha} \end{equation} Note that during the collapse process $R/R_b\equiv \alpha$ remains fixed for a given mass shell. Suppose we are considering the collapse of an interior shell with $\alpha <1$ and $\alpha^2 -1 <0$. Then {\em if $r_0 >0$, the second term on the left hand side of the foregoing equation can ne made arbitrarily small} as $r\rightarrow 0$. Therefore $y$ {\em would become negative} as the collapse progresses if $r_0 >0$. To avoid this, we must have $r/r_0 \ge 1$. We may recall here that, long ago, Harrison et al.\cite{4} (pp. 75) mentioned that spherical gravitational collapse should come to a {\em decisive end} with $M_f=M^* =0$. And they termed this understanding as a ``Theorem'' (without offering a real proof): ``THEOREM 23: Provided that matter does not undergo collapse at the microscopic level at any stage of compression, then, regardless of all features of the equation of state - there exists for each fixed number of baryons a gravitationally collapsed configuration, in which the mass-energy $M^*$ as sensed externally is zero.'' In a somewhat more realistic way Zeldovich \& Novikov\cite{5} (see pp. 297) discussed the possibility of having an ultracompact configuration of degenerate fermions obeying an equation of state $p=e/3$, where $e$ is the proper internal energy density, with $M\rightarrow 0$. And it is also well known that the so-called ``naked singularities'' could be of zero-gravitational mass\cite{6}. We have already seen that the coordinate time required for collapse to the event horizon or beyond is $t=\infty$. And, it may be found that the proper time for collapse of the outer boundary of a dust ball to the central singularity is\cite{7} \begin{equation} \tau =\pi \left({r_\infty\over 8M}\right)^{1/2} \end{equation} where, the dust ball is assumed to be at ``rest'' at $r=r_\infty$ at $\tau =t=0$. This equation is also obtainable from Eq. (4) provided one chooses $R_b$ in such a way thay \begin{equation} R_b^{3/2} = {\sqrt 2 \pi\over 3} r_\infty^{3/2} \end{equation} Since $M=0$ for the OS problem, {\em the proper time for collapse is infinite}: $\tau =\infty$, and not finite. Physically this means that, at any finite proper or coordinate time, there is never any OS-black -hole. The collapse process goes on and on indefinitely as spacetime becomes infinitely curved near the would be singularity $r=0$. If this picture is correct, for self-consistency, Nature should not allow the existence of finite mass BHs. And at this juncture, some readers may argue that, this is not so. Irrespective of the results of the OS problem, as highlighted by us in this paper, it may be argued that GTR allows the existence of arbitray mass Schwarzschild BHs. In another related paper, we assumed the existence of a finite mass BH which is described by Kruskal- Szekeres\cite{8} metric. If this assumption of the existence of a finite mass BH is indeed allowed by GTR, one would find the world line of a material partcle to be {\em timelike}, $ds^2 >0$, everywhere except probably at the central singularity $r=0$. However, we have found that, even in the Kruskal-Szekeres metric the metric {\em appears to be null}\cite{9}, $ds^2 =0$, at $R=2M$. If we describe the region interior to the EH by Lemaitre metric\cite{10}, we have found that, in this case too, the metric becomes null at the EH. We have further shown that, in the Kruskal-Szekeres metric, the metric continues to be null for $R <2M$ too\cite{9}. And this can be explained only when we realize that the Event Horizon, itself is the end of spacetime for the free falling particle. This implies that mass of the BH is actually $M\equiv 0$, in complete agreement with the present {\em exact} analysis. This is also in complete agreement with the idea of Einstein that Schwarzschild singularity can not be realized in practice ($\tau =0$)\cite{11}.
\section{The Big Bang Model} \label{sec:bigbang} \par The discovery of the cosmic background radiation (CBR) in 1964 together with the observed Hubble expansion of the universe had established hot big bang cosmology as a viable model of the universe. The success of the theory of nucleosynthesis in reproducing the observed abundance pattern of light elements together with the proof of the black body character of the CBR then established hot big bang as the standard cosmological model. This model combined with grand unified theories (GUTs) of strong, weak and electromagnetic interactions provides an appropriate framework for discussing the very early stages of the universe evolution. A brief introduction to hot big bang follows. \subsection{Hubble Expansion} \label{subsec:hubble} \par For cosmic times $t\stackrel{_{>}}{_{\sim }} t_{P} \equiv M_{P}^{-1}\sim 10^{-44}~{\rm{sec}}$ ($M_{P}= 1.22\times 10^{19}~{\rm{GeV}}$ is the Planck scale) after the big bang, quantum fluctuations of gravity cease to exist. Gravitation can then be adequately described by classical relativity. Strong, weak and electromagnetic interactions, however, require relativistic quantum field theoretic treatment and are described by gauge theories. \par An important principle, on which the standard big bang (SBB) cosmological model \cite{wkt} is based, is that the universe is homogeneous and isotropic. The strongest evidence so far for this {\it cosmological principle} is the observed \cite{cobe} isotropy of the CBR. Under this assumption, the four dimensional spacetime in the universe is described by the Robertson-Walker metric $$ ds^{2}=-dt^{2}+ $$ \begin{equation} a^{2}(t)\left[\frac{dr^{2}}{1-kr^2} +r^{2}(d\theta^{2}+\sin^{2}\theta~d\varphi^{2}) \right]~, \label{eq:rw} \end{equation} where $r$, $\varphi$ and $\theta$ are `comoving' polar coordinates, which remain fixed for objects that have no other motion than the general expansion of the universe. The parameter $k$ is the `scalar curvature' of the 3-space and $k=0$, $k>0$ or $k<0$ correspond to flat, closed or open universe. The dimensionless parameter $a(t)$ is the `scale factor' of the universe and describes cosmological expansion. We normalize it by taking $a_{0}\equiv a(t_{0})=0$, where $t_{0}$ is the present cosmic time. \par The `instantaneous' radial physical distance is given by \begin{equation} R=a(t)\int_{0}^{r}\frac{dr}{(1-kr^{2})^{1/2}}~\cdot \label{eq:dist} \end{equation} For flat universe ($k=0$), $\bar{R}=a(t)\bar{r}$ ($\bar{r}$ is a `comoving' and $\bar{R}$ a physical vector in 3-space) and the velocity of an object is \begin{equation} \bar{V}=\frac{d\bar{R}}{dt}=\frac{\dot{a}}{a}\bar{R} +a\frac{d\bar{r}}{dt}~, \label{eq:velocity} \end{equation} where overdots denote derivation with respect to cosmic time. The second term in the right hand side (rhs) of this equation is the `peculiar velocity', $\bar{v}=a(t) \dot{\bar{r}}$, of the object, i.e., its velocity with respect to the `comoving' coordinate system. For $\bar{v}=0$, Eq.(\ref{eq:velocity}) becomes \begin{equation} \bar{V}=\frac{\dot{a}}{a}\bar{R}\equiv H(t)\bar{R}~, \label{eq:hubblelaw} \end{equation} where $H(t)\equiv \dot{a}(t)/a(t)$ is the Hubble parameter. This is the well-known Hubble law asserting that all objects run away from each other with velocities proportional to their distances and is considered as the first success of SBB cosmology. \subsection{Friedmann Equation} \label{subsec:friedmann} \par Homogeneity and isotropy of the universe imply that the energy momentum tensor takes the diagonal form $(T_{\mu}^{\nu})= {\rm{diag}}(-\rho, p, p, p)$, where $\rho$ is the energy density of the universe and $p$ the pressure. Energy momentum conservation (${T_{\mu~;\nu}^{~\nu}}=0$) then takes the form of the continuity equation \begin{equation} \frac{d\rho}{dt}=-3H(t)(\rho+p)~, \label{eq:continuity} \end{equation} where the first term in the rhs describes the dilution of the energy due to the expansion of the universe and the second term corresponds to the work done by pressure. Eq.(\ref{eq:continuity}) can be given the following more transparent form \begin{equation} d\left(\frac{4\pi}{3}a^{3}\rho\right)= -p~4\pi a^{2}da~, \label{eq:cont} \end{equation} which indicates that the energy loss of a `comoving' sphere of radius $\propto a(t)$ equals the work done by pressure on its boundary as it expands. \par For a universe described by the Robertson-Walker metric in Eq.(\ref{eq:rw}), Einstein's equations \begin{equation} R_{\mu}^{~\nu}-\frac{1}{2}~\delta_{\mu}^{~\nu}R= 8\pi G~T_{\mu}^{~\nu}~, \label{eq:einstein} \end{equation} where $R_{\mu}^{~\nu}$ and $R$ are the Ricci tensor and scalar curvature tensor and $G\equiv M_{P}^{-2}$ is the Newton's constant, lead to the Friedmann equation \begin{equation} H^{2}\equiv \left(\frac{\dot{a}(t)}{a(t)} \right)^{2}=\frac{8\pi G}{3}\rho-\frac{k}{a^{2}}~\cdot \label{eq:friedmann} \end{equation} \par Averaging $p$, we write $\rho+p=\gamma \rho$. Eq.(\ref{eq:continuity}) then becomes $\dot{\rho}= -3H\gamma\rho$, which gives $d \rho/\rho=-3 \gamma da/a$ and $\rho \propto a^{-3 \gamma}$. For a universe dominated by pressureless matter, $p=0$ and, thus, $\gamma=1$, which gives $\rho\propto a^{-3}$. This is easily interpreted as mere dilution of a fixed number of particles in a `comoving' volume due to the cosmological expansion. For a radiation dominated universe, $p=\rho/3$ and, thus, $\gamma=4/3$, which gives $\rho\propto a^{-4}$. In this case, we get an extra factor of $a(t)$ due to the red-shifting of all wave-lengths by the expansion. Substituting $\rho \propto a^{-3 \gamma}$ in Friedmann equation with $k=0$, we get $\dot{a}/a \propto a^{-3 \gamma/2}$ and, thus, $a(t)\propto t^{2/3\gamma}$. Taking into account the normalization of $a(t)$ ($a(t_{0})=1$), this gives \begin{equation} a(t)=(t/t_{0})^{2/3\gamma}~. \label{eq:expan} \end{equation} For a matter dominated universe, we get the expansion law $a(t)=(t/t_{0})^{2/3}$. `Radiation', however, expands as $a(t)=(t/t_{0})^{1/2}$. \par The universe in its early stages of evolution is radiation dominated and its energy density is \begin{equation} \rho=\frac{\pi^{2}}{30}\left(N_{b}+\frac{7}{8}N_{f} \right)T^{4}\equiv c~T^{4}~, \label{eq:boltzman} \end{equation} where $T$ is the cosmic temperature and $N_{b}$ ($N_{f}$) is the number of massless bosonic (fermionic) degrees of freedom. The combination $g_{*}=N_{b}+(7/8)N_{f}$ is called effective number of massless degrees of freedom. The entropy density is \begin{equation} s= \frac{2\pi^{2}}{45}~g_{*}~T^{3}~. \label{eq:entropy} \end{equation} Assuming adiabatic universe evolution, i.e., constant entropy in a `comoving' volume ($sa^{3}={\rm{constant}}$), we obtain the relation $aT={\rm{constant}}$. The temperature-time relation during radiation dominance is then derived from Friedmann equation (with $k=0$): \begin{equation} T^{2}=\frac{M_{P}}{2(8\pi c/3)^{1/2}t}~\cdot \label{eq:temptime} \end{equation} We see that classically the expansion starts at $t=0$ with $T=\infty$ and $a=0$. This initial singularity is, however, not physical since general relativity fails at cosmic times smaller than about the Planck time $t_{P}$. The only meaningful statement is that the universe, after a yet unknown initial stage, emerges at a cosmic time $\sim t_{P}$ with temperature $T\sim M_{P}$. \subsection{Important Cosmological Parameters} \label{subsec:parameter} \par The most important parameters describing the expanding universe are the following: \begin{list} \setlength{\rightmargin=0cm}{\leftmargin=0cm} \item[{\bf i.}] The present value of the Hubble parameter (known as Hubble constant) $H_{0}\equiv H(t_{0})=100~h~\rm{km}~\rm{sec}^{-1}~\rm{Mpc}^{-1}$ ($0.4\stackrel{_{<}}{_{\sim }}h \stackrel{_{<}}{_{\sim }}0.8$). \item[{\bf ii.}]The fraction $\Omega=\rho/\rho_{c}$, where $\rho_{c}$ is the critical density corresponding to a flat universe ($k=0$). From Friedmann equation, $\rho_{c}=3H^{2}/8\pi G$ and, thus, $\Omega=1+k/a^{2}H^{2}$. $\Omega=1$, $\Omega>1$ or $\Omega<1$ correspond to flat, closed or open universe. Assuming inflation (see below), the present value of $\Omega$ must be $\Omega_{0}=1$. However, the baryonic contribution to $\Omega$ is $\Omega_{B}\approx 0.05-0.1$ \cite{deuterium}. This indicates that most of the energy in the universe must be in nonbaryonic form. \item[{\bf iii.}]The deceleration parameter \begin{equation} q=-\frac{(\ddot{a}/\dot{a})}{(\dot{a}/a)} =\frac{\rho+3p}{2\rho_{c}}~\cdot \label{eq:decel} \end{equation} For `matter', $q=\Omega/2$ and, thus, inflation implies that the present deceleration parameter is $q_{0}=1/2$. \end{list} \subsection{Particle Horizon} \label{subsec:parhor} \par Light travels only a finite distance from the time of big bang ($t=0$) till some cosmic time $t$. From the Robertson-Walker metric in Eq.(\ref{eq:rw}), we find that the propagation of light along the radial direction is described by the equation $a(t)dr=dt$. The particle horizon, which is the `instantaneous' distance at time $t$ travelled by light since the beginning of time, is then given by \begin{equation} d_{H}(t)=a(t)\int_{0}^{t}\frac{dt^{\prime}} {a(t^{\prime})}~\cdot \label{eq:hor} \end{equation} The particle horizon is a very important notion since it coincides with the size of the universe already seen at time $t$ or, equivalently, with the distance at which causal contact has been established at $t$. Eqs.(\ref{eq:expan}) and (\ref{eq:hor}) give \begin{equation} d_{H}(t)=\frac{3\gamma}{3\gamma-2}t~, ~\gamma\neq 2/3~. \label{eq:hort} \end{equation} Also, \begin{equation} H(t)=\frac{2}{3\gamma}t^{-1}~, ~d_{H}(t)=\frac{2}{3\gamma-2}H^{-1}(t)~. \label{eq:hubblet} \end{equation} For `matter' (`radiation'), this becomes $d_{H}(t)=3t= 2H^{-1}(t)$ ($d_{H}(t)=2t=H^{-1}(t)$). The present particle horizon is $d_{H}(t_{0})=2H_{0}^{-1}\approx 6,000~h^{-1}~{\rm{Mpc}}$, the present cosmic time is $t_{0}=2H_{0}^{-1}/3\approx 6.7\times 10^{9} ~h^{-1}~{\rm{years}}$ and the present value of the critical density is $\rho_{c}=3H_{0}^{2}/8\pi G\approx 1.9\times 10^{-29}~h^{-2}~{\rm{gm/cm^{3}}}$. \subsection{Brief History of the Early Universe} \label{subsec:history} \par We will now briefly describe the early stages of the universe evolution according to GUTs \cite{ggps}. We will take a GUT based on the gauge group $G$ ($=SU(5)$, $SO(10)$, $SU(3)^{3}$, ...) with or without supersymmetry. At a superheavy scale $M_{X}\sim 10^{16}~{\rm{GeV}}$ (the GUT mass scale), $G$ breaks to the standard model gauge group $G_{S}= SU(3)_{c}\times SU(2)_{L}\times U(1)_{Y}$ by the vacuum expectation value (vev) of an appropriate higgs field $\phi$. (For simplicity, we will consider that this breaking occurs in just one step.) $G_{S}$ is, subsequently, broken to $SU(3)_{c}\times U(1)_{em}$ at the electroweak scale $M_{W}$. \par GUTs together with the SBB cosmological model (based on classical gravitation) provide a suitable framework for discussing the early history of the universe for cosmic times $\stackrel{_{>}}{_{\sim }} 10^{-44}~ {\rm{sec}}$. They predict that the universe, as it expands and cools down after the big bang, undergoes \cite{kl} a series of phase transitions during which the initial gauge symmetry is gradually reduced and several important phenomena take place. \par After the big bang, the GUT gauge group $G$ was unbroken and the universe was filled with a hot `soup' of massless particles which included not only photons, quarks, leptons and gluons but also the weak gauge boson $W^{\pm}$, $Z^{0}$, the GUT gauge bosons $X$, $Y$, ... as well as several higgs bosons. (In the supersymmetric case, all the supersymmetric partners of these particles were also present.) At cosmic time $t\sim 10^{-37}~{\rm{sec}}$ corresponding to temperature $T\sim 10^{16}~{\rm{GeV}}$, $G$ broke down to $G_{S}$ and the $X$, $Y$, ... gauge bosons together with some higgs bosons acquired superheavy masses of order $M_{X}$. The out-of-equilibrium decay of these superheavy particles can produce \cite{bau} the observed baryon asymmetry of the universe (BAU). Important ingredients for this mechanism to work are the violation of baryon number, which is inherent in GUTs, and C and CP violation. This is the second important success of the SBB model. \par During the GUT phase transition, topologically stable extended objects \cite{kibble} such as magnetic monopoles \cite{monopole}, cosmic strings \cite{string} or domain walls \cite{wall} can also be produced. Monopoles, which exist in all GUTs, can lead into cosmological problems \cite{preskill} which are, however, avoided by inflation \cite{guth,lindebook} (see Secs.\ref{subsec:monopole} and \ref{subsec:infmono}). This is a period of an exponentially fast expansion of the universe which can occur during some GUT phase transition. Strings can contribute \cite{zel} to the primordial density fluctuations necessary for structure formation \cite{structure} in the universe whereas domain walls are \cite{wall} absolutely catastrophic and GUTs predicting them should be avoided or inflation should be used to remove them from the scene. \par At $t\sim 10^{-10}~{\rm{sec}}$ or $T\sim 100~{\rm{GeV}}$, the electroweak transition takes place and $G_{S}$ breaks to $SU(3)_{c}\times U(1)_{em}$. The $W^{\pm}$, $Z^{0}$ gauge bosons together with the electroweak higgs fields acquire masses $\sim M_{W}$. Subsequently, at $t\sim 10^{-4}~{\rm{sec}}$ or $T\sim 1~{\rm{GeV}}$, color confinement sets in and the quarks get bounded forming hadrons. \par The direct involvement of particle physics essentially ends here since most of the subsequent phenomena fall into the realm of other branches. We will, however, sketch some of them since they are crucial for understanding the earlier stages of the universe evolution where their origin lies. \par At $t\approx 180~{\rm{sec}}$ ($T\approx 1~{\rm{MeV}}$), nucleosynthesis takes place, i.e., protons and neutrons form nuclei. The abundance of light elements ($D$, $^{3}He$, $^{4}He$ and $^{7}Li$) depends \cite{peebles} crucially on the number of light particles (with mass $\stackrel{_{<}}{_{\sim }} 1~{\rm{MeV}}$), i.e., the number of light neutrinos, $N_{\nu}$, and $\Omega_{B}h^{2}$. Agreement with observations \cite{deuterium} is achieved for $N_{\nu}=3$ and $\Omega_{B}h^{2}\approx 0.019$. This is the third success of SBB cosmology. Much later, at the so called `equidensity' point, $t_{\rm{eq}}\approx 3,000~ {\rm{years}}$, `matter' dominates over `radiation'. \par At cosmic time $t\approx 200,000~h^{-1} {\rm{years}}$ ($T\approx 3,000~{\rm{K}}$), we have the `decoupling' of `matter' and `radiation' and the `recombination' of atoms. After this, `radiation' evolves as an independent (not interacting) component of the universe and is detected today as CBR with temperature $T_{0}\approx 2.73~{\rm{K}}$. The existence of this radiation is the fourth important success of the theory of big bang. Finally, structure formation \cite{structure} in the universe starts at $t\approx 2\times 10^{8}~{\rm{years}}$. \section{Shortcomings of Big Bang} \label{sec:short} The SBB cosmological model has been very successful in explaining, among other things, the Hubble expansion of the universe, the existence of the CBR and the abundances of the light elements which were formed during primordial nucleosynthesis. Despite its great successes, this model had a number of long-standing shortcomings which we will now summarize: \subsection{Horizon Problem} \label{subsec:horizon} The CBR, which we receive now, was emitted at the time of `decoupling' of matter and radiation when the cosmic temperature was $T_d \approx 3,000~\rm{K}$. The decoupling time, $t_d$, can be calculated from \begin{equation} \frac {T_0}{T_d} = \frac {2.73~\rm{K}}{3,000~\rm{K}} = \frac {a (t_d)}{a(t_0)} = \left(\frac {t_d}{t_0}\right)^{2/3}\cdot \label{eq:dec} \end{equation} It turns out that $t_d \approx 200,000~h^{-1}$ years. \par The distance over which the photons of the CBR have travelled since their emission is $$ a(t_0) \int^{t_{0}} _{t_{d}} \frac {dt^\prime}{a(t^\prime)} = 3t_0 \left[1 - \left(\frac {t_d}{t_0}\right)^{2/3}\right] $$ \begin{equation} \approx 3t_0 \approx 6,000~h^{-1}~\rm{Mpc}~, \label{eq:lss} \end{equation} which essentially coincides with the present particle horizon size. A sphere around us with radius equal to this distance is called the `last scattering surface' since the CBR observed now has been emitted from it. The particle horizon size at $t_d$ was $2H^{-1} (t_d) = 3t_d \approx 0.168~h^{-1}~ \rm{Mpc}$ and expanded till the present time to become equal to $0.168~h^{-1} (a(t_0)/a(t_d))~{\rm{Mpc}} \approx 184~h^{-1}$ Mpc. The angle subtended by this `decoupling' horizon at present is $\theta_{d} \approx 184/6,000 \approx 0.03~\rm{rads} \approx 2~^o$. Thus, the sky splits into $4 \pi/(0.03)^2 \approx 14,000$ patches that never communicated causally before sending light to us. The question then arises how come the temperature of the black body radiation from all these patches is so accurately tuned as the measurements of the cosmic background explorer \cite{cobe} (COBE) require ($\delta T/T \approx 6.6 \times 10^{-6}$). \subsection{Flatness Problem} \label{subsec:flatness} The present energy density, $\rho$, of the universe has been observed to lie in the relatively narrow range $0.1 \rho_c \stackrel{_{<}} {_{\sim }}\rho \stackrel{_{<}}{_{\sim }}2 \rho_c$, where $\rho_c$ is the critical energy density corresponding to a flat universe. The lower bound has been derived from estimates of galactic masses using the virial theorem whereas the upper bound from the volume expansion rate implied by the behavior of galactic number density at large distances. Eq.(\ref{eq:friedmann}) implies that $(\rho - \rho_c)/\rho_c =3 (8 \pi G \rho_c)^{-1} (k/a^2)$ is proportional to $a$, for matter dominated universe. Consequently, in the early universe, we have $ |(\rho - \rho_c)/\rho_c|\ll 1$ and the question arises why the initial energy density of the universe was so finely tuned to be equal to its critical value. \subsection{Magnetic Monopole Problem} \label {subsec:monopole} This problem arises only if we combine the SBB model with GUTs \cite{ggps} of strong, weak and electromagnetic interactions. As already indicated, according to GUTs, the universe underwent \cite{kl} a phase transition during which the GUT gauge symmetry group, $G$, broke to $G_{S}$. This breaking was due to the fact that, at a critical temperature $T_c$, an appropriate higgs field, $\phi$, developed a nonzero vev. Assuming that this phase transition was a second order one, we have $\langle \phi\rangle(T) \approx \langle \phi \rangle(T=0)( 1 - T^2/T^2_c)^{1/2}$, $m_H (T)\approx \lambda \langle \phi\rangle (T)$, for the temperature dependent vev and mass of the higgs field respectively at $T \leq T_c$ ($\lambda$ is an appropriate higgs coupling constant). \par The GUT phase transition produces magnetic monopoles \cite{monopole} which are localized deviations from the vacuum with radius $\sim M_X^{-1}$, energy $\sim M_X/ \alpha_G$ and $\phi =0$ at their center ($\alpha_G= g^2_{G}/4\pi$ with $g_G$ being the GUT gauge coupling constant). The vev of the higgs field on a sphere, $S^2$, with radius $\gg M_{X}^{-1}$ around the monopole lies on the vacuum manifold $G/G_S$ and we, thus, obtain a mapping: $S^2\longrightarrow G/G_S$. If this mapping is homotopically nontrivial the topological stability of the monopole is guaranteed. \par Monopoles can be produced when the fluctuations of $\phi$ over $\phi=0$ between the vacua at $\pm \langle \phi\rangle(T)$ cease to be frequent. This takes place when the free energy needed for $\phi$ to fluctuate from $\langle \phi\rangle (T)$ to zero in a region of radius equal to the higgs correlation length $\xi(T) = m^{-1}_H (T)$ exceeds $T$. This condition reads $(4\pi/3) \xi^3 \Delta V \stackrel{_{>}} {_{\sim }} T$, where $\Delta V \sim \lambda^2 \langle \phi\rangle^4$ is the difference in free energy density between $\phi =0$ and $\phi= \langle \phi\rangle(T)$. The Ginzburg temperature \cite{ginzburg}, $T_G$, corresponds to the saturation of this inequality. So, at $T \stackrel{_{<}}{_{\sim }} T_G$, the fluctuations over $\phi=0$ stop and $\langle \phi\rangle$ settles on the vacuum manifold $G/G_S$. At $T_G$, the universe splits into regions of size $\xi_G \sim (\lambda^2 T_c)^{-1}$, the higgs correlation length at $T_G$, with the higgs field being more or less aligned in each region. Monopoles are produced at the corners where such regions meet (Kibble \cite{kibble} mechanism) and their number density is estimated to be $n_M \sim {\rm{p}} \xi_{G}^{-3} \sim {\rm{p}} \lambda^{6} T_{c}^{3}$, where $\rm{p} \sim \rm{1/10}$ is a geometric factor. The `relative' monopole number density then turns out to be $r_M =n_M/T^3 \sim \rm{10^{-6}}$. We can derive a lower bound on $r_M$ by employing causality. The higgs field $\phi$ cannot be correlated at distances bigger than the particle horizon size, $2t_G$, at $T_G$. This gives the causality bound \begin{equation} n_M \stackrel{_{>}}{_{\sim }}\frac {\rm{p}} {\frac{4 \pi}{3}(2t_G)^3}~, \label{eq:causal} \end{equation} which implies that $r_M\stackrel{_{>}}{_{\sim }} \rm{10^{-10}}$. \par The subsequent evolution of monopoles, after $T_G$, is governed by the equation \cite{preskill} \begin{equation} \frac {dn_M}{dt} = - D n_{M}^{2} - 3 \frac {\dot{a}}{a} n_{M}~, \label{eq:evol} \end{equation} where the first term in the rhs (with $D$ being an appropriate constant) describes the dilution of monopoles due to their annihilation with antimonopoles while the second term corresponds to their dilution by the general cosmological expansion. The monopole-antimonopole annihilation proceeds as follows. Monopoles diffuse towards antimonopoles in the plasma of charged particles, capture each other in Bohr orbits and eventually annihilate. The annihilation is effective provided the mean free path of monopoles in the plasma of charged particles does not exceed their capture distance. This happens at cosmic temperatures $T \stackrel{_{>}}{_{\sim }} \rm{10^{12}~GeV}$. The overall result is that, if the initial relative magnetic monopole density $r_{M,\rm{in}} \stackrel{_{>}}{_{\sim }}\rm{10^{-9}} ( \stackrel{_{<}}{_{\sim }}\rm{10^{-9}})$, the final one $r_{M,\rm{fin}} \sim 10^{-9} ( \sim r_{M,\rm{in}})$. This combined with the causality bound yields $r_{M,\rm{fin}} \stackrel{_{>}}{_{\sim }} \rm{10^{-10}}$. However, the requirement that monopoles do not dominate the energy density of the universe at nucleosynthesis gives \begin{equation} r_M (T \approx 1~\rm{MeV}) \stackrel{_{<}} {_{\sim }}\rm{10^{-19}}~, \label{eq:nucleo} \end{equation} and we obtain a clear discrepancy of about ten orders of magnitude. \subsection{Density Fluctuations} \label{subsec:fluct} For structure formation~\cite{structure} in the universe, we need a primordial density perturbation, $\delta \rho/ \rho$, at all length scales with a nearly flat spectrum \cite{hz}. We also need some explanation of the temperature fluctuations, $\delta T/T$, of CBR observed by COBE \cite{cobe} at angles $\theta \stackrel{_{>}}{_{\sim }} \theta_d \approx 2~^o$ which violate causality (see Sec.\ref{subsec:horizon}). \par Let us expand $\delta \rho/\rho$ in plane waves \begin{equation} \frac {\delta \rho} {\rho} (\bar{r},t) = \int d^3 k\delta_{\bar{k}}(t)e^{i\bar{k} \bar{r}}~, \label{eq:plane} \end{equation} where $\bar{r}$ is a `comoving' vector in 3-space and $\bar{k}$ is the `comoving' wave vector with $k=|\bar{k}|$ being the `comoving' wave number ($\lambda=2 \pi/k$ is the `comoving' wave length whereas the physical wave length is $ \lambda _{\rm{phys}}= a(t) \lambda$). For $\lambda_{\rm{phys}} \leq H^{-1}$, the time evolution of $\delta_{\bar{k}}$ is described by the Newtonian equation \begin{equation} \ddot{\delta}_{\bar{k}} + 2 H \dot {\delta}_{\bar{k}} + \frac {v_{s}^{2}k^2}{a^2}\delta_{\bar{k}}= 4 \pi G \rho \delta_{\bar{k}}~, \label{eq:newton} \end{equation} where the second term in the left hand side (lhs) comes from the cosmological expansion and the third is the `pressure' term ($v_s$ is the velocity of sound given by $v^{2}_{s}=dp/d\rho$, where $p$ is the mean pressure). The rhs of this equation corresponds to the gravitational attraction. \par For the moment, let us put $H$=0 (static universe). In this case, there exists a characteristic wave number $k_J$, the Jeans wave number, given by $k^{2}_{J}=4 \pi G a^{2} \rho/ v^{2}_{s}$ and having the following property. For $k \geq k_J$, pressure dominates over gravitational attraction and the density perturbations just oscillate, whereas, for $k \leq k_J$, gravitational attraction dominates and the density perturbations grow exponentially. In particular, for $p$=0 (matter domination), $v_s=0$ and all scales are Jeans unstable with \begin{equation} \delta_{\bar{k}} \propto {\rm{exp}}(t/\tau)~,~\tau= (4 \pi G \rho)^{-1/2}~. \label{eq:jeans} \end{equation} \par Now let us take $H\neq 0$. Since the cosmological expansion pulls the particles apart, we get a smaller growth: \begin{equation} \delta_{\bar{k}}\propto a(t) \propto t^{2/3}~, \label{eq:growth} \end{equation} in the matter dominated case. For a radiation dominated universe ($p \neq 0$), we get essentially no growth of the density perturbations. This means that, in order to have structure formation in the universe, which requires $\delta \rho/\rho \sim 1$, we must have \begin{equation} (\frac {\delta \rho} {\rho})_{\rm{eq}}\sim 4 \times 10^{-5} ( \Omega_0 h)^{-2}~, \label{eq:equi} \end{equation} at the `equidensity' point (where the energy densities of matter and radiation coincide), since the available growth factor for perturbations is given by $a_0/a_{\rm{eq}} \sim 2.5 \times 10^4 (\Omega_0 h)^2$. Here $\Omega_0= \rho_0/\rho_c$, where $\rho_0$ is the present energy density of the universe. The question then is where these primordial density fluctuations originate from. \section{Inflation} \label{sec:inflation} Inflation~\cite{guth,lindebook} is an idea which solves simultaneously all four cosmological puzzles and can be summarized as follows. Suppose there is a real scalar field $\phi$ (the inflaton) with (symmetric) potential energy density $V(\phi)$ which is quite `flat' near $\phi=0$ and has minima at $\phi = \pm\langle \phi\rangle$ with $V(\pm \langle \phi\rangle)=0$. At high enough $T$'s, $\phi =0$ in the universe due to the temperature corrections in $V(\phi)$. As $T$ drops, the effective potential density approaches the $T$=0 potential but a little potential barrier separating the local minimum at $\phi=0$ and the vacua at $\phi = \pm\langle \phi\rangle$ still remains. At some point, $\phi$ tunnels out to $\phi_1 \ll\langle \phi\rangle$ and a bubble with $\phi=\phi_1$ is created in the universe. The field then rolls over to the minimum of $V(\phi)$ very slowly (due to the flatness of the potential). During this slow roll over, the energy density $\rho \approx V(\phi=0) \equiv V_0$ remains essentially constant for quite some time. The Lagrangian density \begin{equation} L=\frac{1}{2} \partial_{\mu} \phi \partial^{\mu} \phi - V(\phi) \label{eq:lagrange} \end{equation} gives the energy momentum tensor \begin{equation} T_{\mu}^{~\nu} = - \partial_\mu \phi \partial^\nu \phi + \delta_{\mu}^{~\nu}\left(\frac{1}{2} \partial_\lambda \phi \partial^\lambda \phi - V (\phi)\right)~, \label{eq:energymom} \end{equation} which during the slow roll over takes the form $T_{\mu}^{~\nu} \approx - V_{0}~\delta_{\mu}^{~\nu}$. This means that $\rho \approx -p \approx V_0$, i.e., the pressure $p$ is negative and equal in magnitude with the energy density $\rho$, which is consistent with Eq.(\ref{eq:continuity}). Since, as we will see, $a(t)$ grows very fast, the `curvature' term, $k/a^2$, in Eq.(\ref{eq:friedmann}) becomes subdominant and we get \begin{equation} H^2 \equiv \left(\frac {\dot{a}}{a}\right)^2 = \frac {8 \pi G} {3} V_0~, \label{eq:inf} \end{equation} which gives $a(t) \propto e^{Ht},~H^2 =(8 \pi G/3)V_0$ = ~constant. So the bubble expands exponentially for some time and $a(t)$ grows by a factor \begin{equation} \frac {a(t_f)}{a(t_i)} ={\rm{exp}} H(t_f - t_i) \equiv {\rm{exp}} H \tau~, \label{eq:efold} \end{equation} between an initial ($t_i$) and a final ($t_f$) time. \par The inflationary scenario just described here, known as new \cite{new} inflation (with the inflaton field starting from the origin, $\phi$=0), is certainly not the only realization of the idea of inflation. Another interesting possibility is to consider the universe as it emerges at the Planck time $t_{P}$, where the fluctuations of gravity cease to exist. We can imagine a region of size $\ell_{P} \sim M_{P}^{-1}$ where the inflaton field acquires a large and almost uniform value and carries negligible kinetic energy. Under certain circumstances this region can inflate (exponentially expand) as $\phi$ rolls down towards its vacuum value. This type of inflation with the inflaton starting from large values is known as the chaotic \cite{chaotic} inflationary scenario. \par We will now show that, with an adequate number of e-foldings, $N=H \tau$, the first three cosmological puzzles are easily resolved (we leave the question of density perturbations for later). \subsection{Resolution of the Horizon Problem} \label{subsec:infhor} The particle horizon during inflation (exponential expansion) \begin{equation} d(t) = e^{Ht} \int^t_{t_{i}} \frac {d t^\prime}{e^{Ht^\prime}} \approx H^{-1}{\rm{exp}}H(t-t_i)~, \label{eq:horizon} \end{equation} for $t-t_i \gg H^{-1}$, grows as fast as $a(t)$. At the end of inflation ($t=t_f$),~ $d(t_f) \approx H^{-1}{\rm{exp}} H \tau$ and the field $\phi$ starts oscillating about the minimun of the potential at $\phi = \langle \phi\rangle$. It then decays and `reheats' \cite{reheat} the universe at a temperature $T_r \sim 10^9~{\rm{GeV}}$ \cite{gravitino}. The universe, after that, goes back to normal big bang cosmology. The horizon $d(t_{f})$ is stretched during the period of $\phi$-oscillations by some factor $\sim 10^9$ depending on details and between $T_r$ and the present era by a factor $T_r/T_0$. So it finally becomes equal to $H^{-1} e^{H \tau} 10^9 (T_r/T_0)$, which should exceed $2H_{0}^{-1}$ in order to solve the horizon problem. Taking $V_0 \approx M_{X}^{4},~M_{X} \sim 10^{16}$ GeV, we see that, with $N = H \tau\stackrel{_{>}}{_{\sim }} 55$, the horizon problem is evaded. \subsection{Resolution of the Flatness Problem} \label{subsec:infflat} The `curvature' term of the Friedmann equation, at present, is given by \begin{equation} \frac {k}{a^2} \approx \left(\frac {k}{a^2}\right)_{bi} e^{-2H \tau}~10^{-18} \left(\frac {10^{-13}~{\rm{GeV}}} {10^9~{\rm{ GeV}}}\right)^2, \label{eq:curvature} \end{equation} where the terms in the rhs correspond to the `curvature' term before inflation, and its growth factors during inflation, during $\phi$-oscillations and after `reheating' respectively. Assuming $(k/a^2)_{bi} \sim (8 \pi G/3) \rho \sim H ^2$~~ $(\rho \approx V_0)$, we get $k/a_{0}^{2} H_{0}^{2} \sim 10^{48}~e^{-2H \tau}$ which gives $(\rho_0 - \rho_c)/\rho_c \equiv \Omega_0 - 1 = k/a_{0}^{2} H_{0}^{2} \ll 1$, for $H \tau \gg 55$. In fact, strong inflation implies that the present universe is flat with a great accuracy. \subsection{Resolution of the Monopole Problem} \label{subsec:infmono} It is obvious that, with a number of e-foldings $\stackrel{_{>}}{_{\sim }} 55$, the primordial monopole density is diluted by at least 70 orders of magnitude and they become totally irrelevant. Also, since $T_r \ll m_M$ (=the monopole mass), there is no production of magnetic monopoles after `reheating'. \section{Detailed Analysis of Inflation} \label{sec:detail} The Hubble parameter is not exactly constant during inflation as we, naively, assumed so far. It actually depends on the value of $\phi$: \begin{equation} H^{2}(\phi) = \frac {8 \pi G} {3} V (\phi)~. \label{eq:hubble} \end{equation} To find the evolution equation for $\phi$ during inflation, we vary the action \begin{equation} \int \sqrt{-{\rm{det}}(g)}~ d^{4}x \left(\frac {1}{2} \partial_ {\mu} \phi \partial^{\mu} \phi - V(\phi) + M(\phi)\right)~, \label{eq:action} \end{equation} where $g$ is the metric tensor and $M(\phi)$ represents the coupling of $\phi$ to `light' matter causing its decay. We find \begin{equation} \ddot{\phi} + 3H \dot{\phi} + \Gamma_{\phi} \dot{\phi} + V^{\prime}(\phi) = 0~, \label{eq:evolution} \end{equation} where the prime denotes derivation with respect to $\phi$ and $\Gamma_{\phi}$ is the decay width \cite{width} of the inflaton. Assume, for the moment, that the decay time of $\phi $, $t_d = \Gamma_{\phi}^{-1}$, is much greater than $H^{-1}$, the expansion time for inflation. Then the term $\Gamma_{\phi} \dot{\phi}$ can be ignored and Eq.(\ref{eq:evolution}) reduces to \begin{equation} \ddot{\phi} + 3 H \dot{\phi} + V^{\prime}(\phi) = 0~. \label{eq:reduce} \end{equation} Inflation is by definition the situation where $\ddot{\phi}$ is subdominant to the `friction' term $3H \dot{\phi}$ in this equation (and the kinetic energy density is subdominant to the potential one). Eq.(\ref{eq:reduce}) then further reduces to the inflationary equation \cite{slowroll} \begin{equation} 3H \dot{\phi} = - V^{\prime} (\phi)~, \label{eq:infeq} \end{equation} which gives \begin{equation} \ddot{\phi} = - \frac {V^{\prime\prime}(\phi)\dot{\phi}} {3H(\phi)} + \frac {V^{\prime}(\phi)} {3H^{2}(\phi)} H^\prime (\phi) \dot{\phi}~. \label{eq:phidd} \end{equation} Comparing the two terms in the rhs of this equation with the `friction' term in Eq.(\ref{eq:reduce}), we get the conditions for inflation (slow roll conditions): \begin{equation} \eta \equiv \frac{M_{P}^{2}}{8 \pi} \bigg | \frac {V^{\prime\prime}(\phi)} {V(\phi)} \bigg | \leq 1~, ~\epsilon \equiv \frac {M_{P}^{2}}{16 \pi} \left(\frac {V^{\prime}(\phi)} {V(\phi)}\right)^{2} \leq 1~. \label{eq:src} \end{equation} The end of the slow roll over occurs when either of the these inequalities is saturated. If $\phi_f$ is the value of $\phi$ at the end of inflation, then $t_f \sim H^{-1}(\phi_f)$. \par The number of e-foldings during inflation can be calculated as follows: $$ N(\phi_{i}\rightarrow \phi_{f}) \equiv \ell n \left(\frac {a(t_{f})} {a(t_{i})}\right) = \int^{t_{f}} _{t_{i}} Hdt= $$ \begin{equation} \int^{\phi_{f}}_{\phi_{i}} \frac {H (\phi)}{\dot{\phi}} d \phi = - \int^{\phi_{f}}_{\phi_{i}} \frac {3 H^2 (\phi) d \phi} {V^{\prime}(\phi)}~, \label{eq:nefolds} \end{equation} where Eqs.(\ref{eq:efold}), (\ref{eq:infeq}) and the definition of $H = \dot{a}/a$ were used. For simplicity, we can shift the field $\phi$ so that the global minimum of the potential is displaced at $\phi$ = 0. Then, if $V(\phi) = \lambda \phi^{\nu}$ during inflation, we have $$ N(\phi_{i} \rightarrow \phi_{f}) = - \int^{\phi_{f}}_{\phi_{i}} \frac {3H^2(\phi)d\phi}{V^{\prime}(\phi)} = $$ \begin{equation} - 8 \pi G \int^{\phi_{f}}_{\phi_{i}} \frac {V(\phi)d\phi}{V^{\prime}(\phi)}=\frac {4 \pi G}{\nu} (\phi^{2}_{i}-\phi^{2}_{f})~. \label{eq:expefold} \end{equation} Assuming that $\phi_{i} \gg \phi_{f}$, this reduces to $N(\phi) = (4 \pi G/\nu)\phi^2$. \section{Coherent Field Oscillations} \label{sec:osci} After the end of inflation at cosmic time $t_f$, the term $\ddot{\phi}$ takes over and Eq.(\ref{eq:reduce}) reduces to $\ddot{\phi} + V^{\prime}(\phi)=0$, which means that $\phi$ starts oscillating coherently about the global minimum of the potential. In reality, due to the `friction' term, $\phi$ performs damped oscillations with a rate of energy density loss given by \begin{equation} \dot{\rho}= \frac {d}{dt}\left(\frac{1}{2} \dot{\phi}^2 + V(\phi)\right) = - 3H \dot{\phi}^2=-3H(\rho+p)~, \label{eq:damp} \end{equation} where $\rho = \dot{\phi}^2/2 + V(\phi) $ and the pressure $p=\dot{\phi}^2/2 - V(\phi)$. Averaging $p$ over one oscillation of $\phi$ and writing \cite{oscillation} $\rho+p=\gamma \rho$, we get $\rho \propto a^{-3 \gamma}$ and $a(t) \propto t^{2/3 \gamma}$ (see Sec.\ref{subsec:friedmann}). \par The number $\gamma$ for an oscillating field can be written as (assuming a symmetric potential) \begin{equation} \gamma = \frac {\int^{T}_{0} \dot{\phi}^{2} dt} {\int^{T}_{0} \rho dt} = \frac{\int^{\phi_{{\rm{max}}}}_{0} \dot{\phi}d \phi} {\int^{\phi_{{\rm{max}}}}_{0} (\rho/\dot{\phi}) d\phi}~, \label{eq:gamma} \end{equation} where $T$ and $\phi_{{\rm{max}}}$ are the period and the amplitude of the oscillation respectively. From the equation $\rho = \dot{\phi}^2/2 + V(\phi)= V_{{\rm{max}}}$, where $V_{{\rm{max}}}$ is the maximal potential energy density, we obtain $\dot{\phi}= \sqrt{2(V_{{\rm{max}}} - V(\phi))}$. Substituting this in Eq.(\ref{eq:gamma}) we get \cite{oscillation} \begin{equation} \gamma = \frac {2 \int^{\phi_{{\rm{max}}}}_{0} (1-V/V_{{\rm{max}}})^{1/2} d\phi}{\int^{\phi_{{\rm{max}}}}_{0} (1-V/V_{{\rm{max}}})^{-1/2} d \phi}~~\cdot \label{eq:gammafinal} \end{equation} For a potential of the simple form $V(\phi) = \lambda \phi^{\nu}$,~$\gamma$ is readily found to be given by $\gamma=2 \nu/(\nu +2)$. Consequently, in this case, $\rho \propto a^{-6\nu/(\nu+2)}$ and $a(t) \propto t^{(\nu+2)/3 \nu}$. For $\nu=2$, in particular, one has $\gamma$=1, ~$\rho \propto a^{-3}$,~$a(t)\propto t^{2/3}$ and the oscillating field behaves like pressureless `matter'. This is not unexpected since a coherent oscillating massive free field corresponds to a distribution of static massive particles. For $\nu$=4, however, we obtain $\gamma = 4/3$, ~$\rho \propto a^{-4}$,~$a(t) \propto t^{1/2}$ and the system resembles `radiation'. For $\nu = 6$, one has $\gamma=3/2$,~$ \rho \propto a^{-4.5}$,~$a(t) \propto t^{4/9}$ and the expansion is slower than in a radiation dominated universe (the pressure is higher than in `radiation'). \section{Decay of the Field $\phi$} \label{sec:decay} Reintroducing the `decay' term $\Gamma_{\phi} \dot{\phi}$, Eq.(\ref{eq:evolution}) can be written as \begin{equation} \dot{\rho} = \frac{d}{dt} \left(\frac{1}{2} \dot{\phi}^2 + V(\phi)\right) = - (3H + \Gamma_\phi)\dot{\phi}^2~, \label{eq:decay} \end{equation} which is solved \cite{reheat,oscillation} by \begin{equation} \rho(t)=\rho_{f} \left(\frac{a(t)}{a(t_{f})}\right)^{-3 \gamma} {\rm{exp}} [ -\gamma \Gamma_{\phi}(t-t_f)]~, \label{eq:rho} \end{equation} where $\rho_f$ is the energy density at the end of inflation at cosmic time $t_f$. The second and third factors in the rhs of this equation represent the dilution of the field energy due to the expansion of the universe and the decay of $\phi$ to light particles respectively. \par All pre-existing `radiation' (known as `old radiation') was diluted by inflation, so the only `radiation' present is the one produced by the decay of $\phi$ and is known as `new radiation'. Its energy density satisfies \cite{reheat,oscillation} the equation \begin{equation} \dot{\rho}_{r} = - 4 H \rho_{r} + \gamma \Gamma_{\phi} \rho~, \label{eq:newrad} \end{equation} where the first term in the rhs represents the dilution of radiation due to the cosmological expansion while the second one is the energy density transfer from $\phi$ to `radiation'. Taking $\rho_{r}(t_f)$=0, this equation gives \cite{reheat,oscillation} $$ \rho_{r}(t) = \rho_{f}\left(\frac {a(t)} {a(t_{f})}\right)^{-4} $$ \begin{equation} \int^{t}_{t_{f}} \left(\frac{a(t^{\prime})} {a(t_{f})}\right)^{4-3 \gamma} e^{ -\gamma \Gamma_{\phi} (t^{\prime}-t_f)} ~\gamma \Gamma_{\phi} dt^{\prime}~. \label{eq:rad} \end{equation} For $t_{f} \ll t_{d}$ and $\nu =2$, this expression is approximated by \begin{equation} \rho_{r}(t)=\rho_{f}\left(\frac {t}{t_f}\right)^{-8/3} \int^{t}_{0} \left(\frac{t^{\prime}}{t_{f}}\right)^{2/3} e^{-\Gamma_{\phi}t^{\prime}} dt^{\prime}~, \label{eq:appr} \end{equation} which, using the formula \begin{equation} \int_{0}^{u} x^{p-1} e^{-x}dx = e^{-u}~\sum^{\infty}_{k=0}~ \frac {u^{p+k}}{p(p+1)\cdot\cdot\cdot(p+k)}~~, \label{eq:formula} \end{equation} can be written as \begin{equation} \rho_{r} = \frac {3}{5}~\rho~\Gamma_{\phi}t \left[1 + \frac {3}{8}~\Gamma_{\phi}t + \frac {9}{88}~(\Gamma_{\phi}t)^2+ \cdots \right]~, \label{eq:expand} \end{equation} with $\rho = \rho_{f} (t/t_{f})^{-2}{\rm{exp}} (-\Gamma_{\phi}t)$ being the energy density of the field $\phi$ which performs damped oscillations and decays into `light' particles. \par The energy density of the `new radiation' grows relative to the energy density of the oscillating field and becomes essentially equal to it at a cosmic time $t_{d} = \Gamma_{\phi}^{-1}$ as one can deduce from Eq.(\ref{eq:expand}). After this time, the universe enters into the radiation dominated era and the normal big bang cosmology is recovered. The temperature at $t_{d}, ~T_{r}(t_{d})$, is historically called the `reheat' temperature although no supercooling and subsequent reheating of the universe actually takes place. Using the time to temperature relation in Eq.(\ref{eq:temptime}) for a radiation dominated universe we find that \begin{equation} T_{r} = \left(\frac {45}{16 \pi^{3}g_*}\right)^{1/4} (\Gamma_{\phi} M_{P})^{1/2}~, \label{eq:reheat} \end{equation} where $g_*$ is the effective number of degrees of freedom. For a potential of the type $V(\phi)=\lambda\phi^{\nu}$, the total expansion of the universe during the period of damped field oscillations is \begin{equation} \frac{a(t_{d})}{a(t_{f})} = \left( \frac{t_{d}}{t_{f}}\right)^{\frac{\nu + 2}{3\nu}}~. \label{eq:expansion} \end{equation} \section{Density Perturbations} \label{sec:density} We are ready to sketch how inflation solves the density fluctuation problem described in Sec.\ref{subsec:fluct}. As a matter of fact, inflation not only homogenizes the universe but also provides us with the primordial density fluctuations necessary for the structure formation in the universe. To understand the origin of these fluctuations, we must first introduce the notion of `event horizon'. Our `event horizon', at a cosmic time $t$, includes all points with which we will eventually communicate sending signals at $t$. The `instantaneous' (at cosmic time $t$) radius of the `event horizon' is \begin{equation} d_{e}(t) = a(t) \int ^{\infty}_{t} \frac{dt^{\prime}}{a(t^{\prime})}~\cdot \label{eq:event} \end{equation} It is obvious, from this formula, that the `event horizon' is infinite for matter or radiation dominated universe. For inflation, however, we obtain a slowly varying `event horizon' with radius $d_{e}(t) = H^{-1} < \infty$. Points, in our `event horizon' at $t$, with which we can communicate sending signals at $t$, are eventually pulled away by the `exponential' expansion and we cease to be able to communicate with them again emitting signals at later times. We say that these points (and the corresponding scales) crossed outside the `event horizon'. The situation is very similar to that of a black hole. Indeed, the exponentially expanding (de Sitter) space is like a black hole turned inside out. This means that we are inside and the black hole surrounds us from all sides. Then, exactly as in a black hole, there are quantum fluctuations of the `thermal type' governed by the `Hawking temperature' \cite{hawking,gibbons} $T_{H} = H/2\pi$. It turns out \cite{bunch,vilenkin} that the quantum fluctuations of all massless fields (the inflaton is nearly massless due to the `flatness' of the potential) are $\delta \phi = H/ 2\pi = T_{H}$. These fluctuations of $\phi$ lead to energy density fluctuations $\delta \rho = V^{\prime} (\phi) \delta \phi$. As the scale of this perturbations crosses outside the `event horizon', they become \cite{fischler} classical metric perturbations. \par The evolution of these fluctuations outside the `inflationary horizon' is quite subtle and involved due to the gauge freedom in general relativity. However, there is a simple gauge invariant quantity \cite{zeta} $\zeta \approx \delta \rho/(\rho +p)$, which remains constant outside the horizon. Thus, the density fluctuation at any present physical (`comoving') scale $\ell, ~(\delta \rho/\rho)_{\ell}$, when this scale crosses inside the post-inflationary particle horizon ($p$=0 at this instance) can be related to the value of $\zeta$ when the same scale crossed outside the inflationary `event horizon' (symbolically at $\ell \sim H^{-1}$). This latter value of $\zeta$ can be found using Eq.(\ref{eq:infeq}) and turns out to be $$ \zeta \mid_{\ell \sim H^{-1}} = \left(\frac {\delta \rho}{\dot {\phi}^2}\right) _{\ell \sim H^{-1}} = \left(\frac {V^{\prime} (\phi) H(\phi)} {2 \pi \dot{\phi}^2}\right)_{\ell \sim H^{-1}} $$ \begin{equation} =- \left(\frac {9 H^{3}(\phi)} {2 \pi V^{\prime}(\phi)}\right)_{\ell \sim H{-1}}~\cdot \label{eq:zeta} \end{equation} Taking into account an extra 2/5 factor from the fact that the universe is matter dominated when the scale $\ell$ re-enters the horizon, we obtain \begin{equation} \left(\frac {\delta \rho}{\rho}\right)_{\ell} = \frac {16 \sqrt {6 \pi}}{5}~\frac {V^{3/2}(\phi_{\ell})}{M^{3}_{P} V^{\prime} (\phi_{\ell})} ~\cdot \label{eq:deltarho} \end{equation} \par The calculation of $\phi_{\ell}$, the value of the inflaton field when the `comoving' scale $\ell$ crossed outside the `event horizon', goes as follows. A `comoving' (present physical) scale $\ell$, at $T_r$, was equal to $\ell (a(t_{d})/a(t_{0})) = \ell(T_{0}/T_{r})$. Its magnitude at the end of inflation ($t=t_{f}$) was equal to $\ell (T_{0}/T_{r})(a(t_{f})/a(t_{d})) = \ell (T_{0}/T_{r})(t_{f}/t_{d})^{(\nu +2)/3 \nu}$ $\equiv \ell_{{\rm{phys}}}(t_{f})$, where the potential $V(\phi)=\lambda\phi^{\nu}$ was assumed. The scale $\ell$, when it crossed outside the inflationary horizon, was equal to $H^{-1}(\phi_{\ell})$. We, thus, obtain \begin{equation} H^{-1}(\phi_{\ell}) e^{N(\phi_{\ell})} = \ell_{{\rm{phys}}}(t_{f})~. \label{eq:lphys} \end{equation} Solving this equation, one can calculate $\phi_{\ell}$ and, thus, $N(\phi_{\ell})\equiv N_{\ell}$, the number of e-foldings the scale $\ell$ suffered during inflation. In particular, for our present horizon scale $\ell \approx 2H_{0}^{-1} \sim10^4$~Mpc, it turns out that $N_{H_{0}}\approx 50-60$. \par Taking the potential $V(\phi)= \lambda \phi^4$, ~Eqs.(\ref{eq:expefold}), (\ref{eq:deltarho}) and (\ref{eq:lphys}) give $$ \left(\frac {\delta \rho} {\rho}\right)_{\ell} = \frac {4 \sqrt{6 \pi}} {5}\lambda^{1/2} \left(\frac{\phi_{\ell}}{M_{P}}\right)^3 = $$ \begin{equation} \frac {4 \sqrt{6 \pi}}{5}\lambda^{1/2} \left(\frac {N_{\ell}}{\pi}\right)^{3/2}~\cdot \label{eq:nl} \end{equation} The measurements of COBE~\cite{cobe}, $ (\delta \rho/ \rho)_{H_{0}} \approx 6 \times 10^{-5}$, then imply that $\lambda \approx 6 \times 10^{-14}$ for $N_{H_{0}} \approx 55$. Thus, we see that the inflaton must be a very weakly coupled field. In nonsupersymmetric GUTs, the inflaton is necessarily gauge singlet since otherwise radiative corrections will certainly make it strongly coupled. This is, undoubtedly, not a very satisfactory situation since we are forced to introduce an otherwise unmotivated extra {\it ad hoc} very weakly coupled gauge singlet. In supersymmetric GUTs, however, the inflaton could be identified \cite{nonsinglet} with a conjugate pair of gauge nonsinglet fields $\phi$, $\bar{\phi}$, already existing in the theory and causing the gauge symmetry breaking. Absence of strong radiative corrections from gauge interactions is guaranteed, in this case, by the mutual cancellation of the D terms of these fields. \par The spectrum of density fluctuations which emerge from inflation can also be analyzed. We will again take the potential $V(\phi) =\lambda \phi^{\nu}$. One then finds that $(\delta \rho/ \rho)_{\ell}$ is proportional to $\phi_{\ell}^{(\nu+2)/2}$ which, combined with the fact that $N(\phi_{\ell})$ is proportional to $\phi_{\ell}^{2}$ (see Eq.(\ref{eq:expefold})), gives \begin{equation} \left(\frac {\delta \rho}{\rho}\right)_{\ell} = \left(\frac{\delta \rho}{\rho}\right)_{H_{0}}\left( \frac{N_{\ell}}{N_{H_{0}}}\right)^{\frac{\nu+2}{4}}. \label{eq:spectrum} \end{equation} The scale $\ell$ divided by the size of our present horizon ($\approx 10^4~{\rm{Mpc}}$) should equal exp$(N_{\ell}- N_{H_{0}})$. This gives $N_{\ell}/N_{H_{0}} = 1 + \ell n(\ell/10^{4})^{1/N_{H_{0}}}$ which expanded around $\ell \approx 10^4$~Mpc and substituted in Eq.(\ref{eq:spectrum}) yields \begin{equation} \left(\frac{\delta \rho}{\rho}\right)_{\ell} = \left(\frac{\delta \rho}{\rho}\right)_{H_{0}}\left( \frac {\ell}{10^4~{\rm{Mpc}}}\right)^{\alpha_{s}}~, \label{eq:alphas} \end{equation} with $\alpha_{s}=(\nu + 2) /4 N_{H_{0}}$. For $\nu=4$, $\alpha_{s} \approx 0.03$ and, thus, the density fluctuations are essentially scale independent. \section{Density Fluctuations in `Matter'} \label{sec:matter} We will now discuss the evolution of the primordial density fluctuations after their scale enters the post-inflationary horizon. To this end, we introduce \cite{bardeen} the `conformal' time, $\eta$, so that the Robertson-Walker metric takes the form of a conformally expanding Minkowski space: \begin{equation} ds^2 = -dt^2 + a^2 (t)~d\bar{r}^2 = a^{2}(\eta)~ (-d \eta^2 + d\bar{r}^2)~, \label{eq:conf} \end{equation} where $\bar{r}$ is a `comoving' 3-vector. The Hubble parameter now takes the form $H\equiv \dot{a}(t)/a(t) =a^{\prime}(\eta)/a^{2}(\eta)$ and the Friedmann Eq.(\ref{eq:friedmann}) can be rewritten as \begin{equation} \frac{1}{a^{2}}\left(\frac{a^{\prime}}{a}\right)^{2} =\frac {8 \pi G}{3} \rho~, \label{eq:conffried} \end{equation} where primes denote derivation with respect to the `conformal' time $\eta$. The continuity Eq.(\ref{eq:continuity}) takes the form $ \rho^{\prime} = - 3 \tilde {H} (\rho+p)$ with $\tilde{H} = a^{\prime}/a$. For a matter dominated universe, $\rho \propto a^{-3}$ which gives $a=(\eta/\eta_{0})^2$ and $a^{\prime}/a =2/\eta$ ($\eta_0$ is the present value of $\eta$). \par The Newtonian Eq.(\ref{eq:newton}) can now be written in the form \begin{equation} \delta^{\prime \prime}_{\bar{k}} (\eta) + \frac {a^{\prime}}{a} \delta^{\prime}_{\bar{k}}(\eta) - 4 \pi G \rho a^{2} \delta_{\bar{k}} (\eta) =0~, \label{eq:confnewton} \end{equation} and the growing (Jeans unstable) mode $\delta_{\bar{k}} (\eta)$ is proportional to $\eta^{2}$ and can be expressed \cite{schaefer} as \begin{equation} \delta_{\bar{k}} (\eta) = \epsilon_{H} \left( \frac {k \eta}{2}\right)^{2} \hat {s} (\bar{k})~, \label{eq:growmode} \end{equation} where $\hat{s}(\bar{k})$ is a Gaussian random variable satisfying \begin{equation} <\hat{s}(\bar {k})> = 0~,~<\hat{s}(\bar {k}) \hat {s}(\bar {k}^{\prime})> = \frac {1} {k^{3}} \delta (\bar{k} - \bar {k}^{\prime})~, \label{eq:gauss} \end{equation} and $\epsilon_{H}$ is the amplitude of the perturbation when its scale crosses inside the post-inflationary horizon. The latter can be seen as follows. A `comoving' (present physical) length $\ell$ crosses inside the post-inflationary horizon when $a \ell/ 2 \pi = H^{-1}=a^2/a^{\prime}$ which gives $\ell/ 2\pi \equiv k^{-1} = a/a^{\prime}= \eta_{H}/2$ or $k \eta_{H}/2 = 1$, where $\eta_{H}$ is the `conformal' time at horizon crossing. This means that, at horizon crossing, $\delta_{\bar{k}} (\eta_{H}) = \epsilon_{H}\hat{s} (\bar{k})$. For scale invariant perturbations, the amplitude $\epsilon _{H}$ is constant. The gauge invariant perturbations of the scalar gravitational potential are given \cite{bardeen} by the Poisson's equation, \begin{equation} \Phi = - 4 \pi G \frac {a^2}{k^2}\rho \delta_{\bar{k}}(\eta)~. \label{eq:poisson} \end{equation} From the Friedmann Eq.(\ref{eq:conffried}), we then obtain \begin{equation} \Phi = - \frac {3}{2} \epsilon_{H} \hat{s} (\bar{k})~. \label{eq:scalarpot} \end{equation} \par The spectrum of the density perturbations can be characterized by the correlation function ($\bar{x}$ is a `comoving' 3-vector) \begin{equation} \xi(\bar{r}) \equiv <\tilde{\delta}^{*}(\bar{x},\eta) \tilde{\delta}(\bar{x} + \bar{r}, \eta)>~, \label{eq:corr} \end{equation} where \begin{equation} \tilde{\delta}(\bar{x},\eta) = \int d^{3}k \delta_{\bar{k}}(\eta)e^{i\bar{k}\bar{x}}~. \label{eq:fourier} \end{equation} Substituting Eq.(\ref{eq:growmode}) in Eq.(\ref{eq:corr}) and then using Eq.(\ref{eq:gauss}), we obtain \begin{equation} \xi (\bar{r}) = \int d^{3}k e^{-i\bar{k}\bar{r}} \epsilon^2_{H} \left(\frac {k \eta}{2}\right)^{4} \frac{1}{k^{3}}~, \label{eq:index} \end{equation} and the spectral function $P(k, \eta) = \epsilon^2_{H}(\eta^{4}/16)k$ is proportional to $k$ for $\epsilon_{H}$ constant. We say that, in this case, the `spectral index' $n=1$ and we have a Harrison-Zeldovich \cite{hz} flat spectrum. In the general case, $P \propto k^{n}$ with $n = 1-2 \alpha_{s}$ (see Eq.(\ref{eq:alphas})). For $V(\phi)= \lambda \phi^{4}$, we get $n \approx 0.94$. \section{Temperature Fluctuations} \label{sec:temperature} The density inhomogeneities produce temperature fluctuations in the CBR. For angles $\theta \stackrel{_{>}}{_{\sim }} 2~^o$, the dominant effect is the scalar Sachs-Wolfe \cite{sachswolfe} effect. Density perturbations on the `last scattering surface' cause scalar gravitational potential fluctuations, $\Phi$, which, in turn, produce temperature fluctuations in the CBR. The physical reason is that regions with a deep gravitational potential will cause the photons to lose energy as they climb up the well and, thus, appear cooler. For $\theta \stackrel{_{<}}{_{\sim }} 2~^o$, the dominant effects are: i) Motion of the last scattering surface causing Doppler shifts, and ii) Intrinsic fluctuations of the photon temperature, $T_{\gamma}$, which are more difficult to calculate since they depend on microphysics, the ionization history, photon streaming and other effects. \par The temperature fluctuations at an angle $\theta$ due to the scalar Sachs-Wolfe effect turn out \cite{sachswolfe} to be $(\delta T/T)_{\theta} = - \Phi_{\ell}/3$, $\ell$ being the `comoving' scale on the `last scattering surface' which subtends the angle $\theta$ [~$\ell \approx 100~h^{-1}(\theta/{\rm{degrees}}) ~{\rm{Mpc}}~]$ and $\Phi_{\ell}$ the corresponding scalar gravitational potential fluctuations. From Eq.(\ref{eq:scalarpot}), we then obtain $(\delta T/T)_{\theta} = (\epsilon_{H}/2) \hat{s} (\bar{k})$, which using Eq.(\ref{eq:growmode}) gives the relation \begin{equation} \left(\frac{\delta T}{T}\right)_{\theta} = \frac{1}{2} \delta_{\bar{k}}(\eta_{H}) = \frac {1}{2} \left(\frac{\delta \rho} {\rho}\right)_{\ell \sim 2\pi k^{-1}}\cdot \label{eq:swe} \end{equation} The COBE scale (present horizon) corresponds to $\theta \approx 60~^o$. Eqs.(\ref{eq:expefold}), (\ref{eq:deltarho}) and (\ref{eq:swe}) give \begin{equation} \left(\frac {\delta T} {T}\right)_{\ell} \propto \left(\frac{\delta \rho}{\rho}\right)_{\ell} \propto \frac{V^{3/2}(\phi_{\ell})}{M^{3}_{P} V^{\prime} (\phi_{\ell})} \propto N_{\ell}^{\frac{\nu+2}{4}}~. \label{ewq:tempflu} \end{equation} Analyzing the temperature fluctuations in spherical harmonics, the quadrupole anisotropy due to the scalar Sachs-Wolfe effect can be obtained: \begin{equation} \left(\frac{\delta T}{T}\right)_{Q-S} = \left(\frac{32 \pi} {45}\right)^{1/2} \frac{V^{3/2}(\phi_{\ell})}{M^{3}_{P} V^{\prime}(\phi_{\ell})}~\cdot \label{eq:quadrupole} \end{equation} For $V(\phi) = \lambda \phi^{\nu}$, this becomes $$ \left(\frac{\delta T}{T}\right)_{Q-S} = \left(\frac{32 \pi}{45}\right)^{1/2} \frac{\lambda^{1/2}\phi_{\ell}^{\frac{\nu+2}{2}}} {\nu M^{3}_{P}}= $$ \begin{equation} \left(\frac{32 \pi}{45}\right)^{1/2} \frac {\lambda^{1/2}}{\nu M^{3}_{P}} \left(\frac{\nu M^{2}_{P}} {4 \pi}\right)^{\frac{\nu+2}{4}} N_{\ell}^{\frac{\nu+2}{4}}~. \label{eq:anisotropy} \end{equation} Comparing this with COBE \cite{cobe} measurements, $(\delta T/T)_{Q} \approx 6.6 \times 10^{-6}$, we obtain $\lambda \approx 6 \times 10^{-14}$, for $\nu=4$, and number of e-foldings suffered by our present horizon scale during the inflationary phase $N_{\ell \sim H^{-1}_{0}} \equiv N_{Q} \approx 55$. \par There are also `tensor' \cite{tensor} fluctuations in the temperature of CBR. The quadrupole tensor anisotropy is \begin{equation} \left(\frac{\delta T}{T}\right)_{Q-T} \approx 0.77~\frac {V^{1/2}(\phi_{\ell})}{M^{2}_{P}}~\cdot \label{eq:tensor} \end{equation} The total quadrupole anisotropy is given by \begin{equation} \left(\frac{\delta T}{T}\right)_{Q} = \left[ \left(\frac{\delta T}{T}\right) ^{2}_{Q-S} + \left(\frac{\delta T}{T}\right)^{2} _{Q-T}\right]^{1/2}, \label{eq:total} \end{equation} and the ratio \begin{equation} r = \frac{\left(\delta T/T\right)^{2}_{Q-T}} {\left(\delta T/T\right)^{2}_{Q-S}} \approx 0.27 ~\left(\frac{M_{P} V^{\prime}(\phi_{\ell})} {V(\phi_{\ell})}\right)^{2}\cdot \label{eq:ratio} \end{equation} For $V(\phi) = \lambda \phi^{\nu}$, we obtain $r \approx 3.4~\nu/N_{H}\ll 1$, and the `tensor' contribution to the temperature fluctuations of the CBR is negligible. \section{Hybrid Inflation} \label{sec:hybrid} \subsection{The non Supersymmetric Version} \label{subsec:nonsusy} The most important disadvantage of the inflationary scenarios described so far is that they need extremely small coupling constants in order to reproduce the results of COBE \cite{cobe}. This difficulty was overcome some years ago by Linde \cite{hybrid} who proposed, in the context of nonsupersymmetric GUTs, an inflationary scenario known as hybrid inflation. The idea was to use two real scalar fields $\chi$ and $\sigma$ instead of one that was normally used. The field $\chi$ provides the vacuum energy which drives inflation while $\sigma$ is the slowly varying field during inflation. The main advantage of this scenario is that it can reproduce the observed temperature fluctuations of the CBR with `natural' values of the parameters in contrast to previous realizations of inflation (like the new \cite{new} or chaotic \cite{chaotic} inflationary scenarios). \par The potential utilized by Linde is \begin{equation} V ( \chi, \sigma)= \kappa^2 \left( M^2 - \frac {\chi^2}{4}\right)^2 + \frac{\lambda^2 \chi^2 \sigma^2} {4} + \frac {m^2\sigma^2}{2}~~, \label{eq:lindepot} \end{equation} where $\kappa,~\lambda$ are dimensionless positive coupling constants and $M$, $m$ are mass parameters. The vacua lie at $\langle \chi\rangle= \pm 2 M$, $\langle \sigma \rangle=0$. Putting $m$=0, for the moment, we observe that the potential possesses an exactly flat direction at $\chi=0$ with $V(\chi=0 ,\sigma)=\kappa^2 M^4$. The mass squared of the field $\chi$ along this flat direction is given by $m^2_\chi=- \kappa^2 M^2 + \lambda^2 \sigma^2/2$ and remains nonnegative for $\sigma \geq \sigma_c = \sqrt {2} \kappa M/ \lambda $. This means that, at $\chi=0$ and $\sigma \geq \sigma_c$, we obtain a valley of minima with flat bottom. Reintroducing the mass parameter $m$ in Eq.(\ref{eq:lindepot}), we observe that this valley acquires a nonzero slope. A region of the universe, where $\chi$ and $\sigma$ happen to be almost uniform with negligible kinetic energies and with values close to the bottom of the valley of minima, follows this valley in its subsequent evolution and undergoes inflation. \par The quadrupole anisotropy of CBR produced during this hybrid inflation can be estimated, using Eq.(\ref{eq:quadrupole}), to be \begin{equation} \left(\frac {\delta T}{T}\right)_{Q} \approx \left(\frac {16 \pi}{45}\right)^{1/2} \frac{\lambda \kappa^2 M^5} {M^3_Pm^2}~. \label{eq:lindetemp} \end{equation} The COBE~\cite{cobe} result, $(\delta T/T)_{Q} \approx 6.6 \times 10^{-6}$, can then be reproduced with $M \approx 2.86 \times 10^{16}$ GeV (the supersymmetric GUT vev) and $m \approx 1.3~\kappa \sqrt {\lambda}\times 10^{15}$ GeV $ \sim 10^{12}$ GeV for $\kappa, \lambda \sim 10^{-2}$. \par Inflation terminates abruptly at $\sigma=\sigma_{c}$ and is followed by a `waterfall', i.e., a sudden entrance into an oscillatory phase about a global minimum. Since the system can fall into either of the two available global minima with equal probability, topological defects are copiously produced if they are predicted by the particular particle physics model one is considering. \subsection{The Supersymmetric Version} \label{subsec:susy} The hybrid inflationary scenario is \cite{lyth} `tailor made' for application to supersymmetric GUTs except that the mass of $\sigma$, $m$, is unacceptably large for supersymmetry, where all scalar fields acquire masses of order $m_{3/2} \sim 1$ TeV (the gravitino mass) from soft supersymmetry breaking. \par To see this, consider a supersymmetric GUT with a (semi-simple) gauge group $G$ of rank $\geq 5$ with $G \to G_S$ (the standard model gauge group) at a scale $M \sim 10^{16}~$GeV. The spectrum of the theory below $M$ is assumed to coincide with the spectrum of the minimal supersymmetric standard model (MSSM) plus standard model singlets so that the successful predictions for $\alpha_{s}$, ${\rm{sin}}^{2} \theta_{W}$ are retained. The theory may also possess global symmetries. The breaking of $G$ is achieved through the superpotential \begin {equation} W = \kappa S (- M^2 + \bar{\phi} \phi), \label{eq:superpot} \end {equation} where $\bar{\phi}, \phi$ is a conjugate pair of $G_{S}$ singlet left handed superfields belonging to nontrivial representations of $G$ and reduce its rank by their vevs and $S$ is a gauge singlet left handed superfield. The coupling constant $\kappa$ and the mass parameter $M$ can be made positive by phase redefinitions. This superpotential has the most general form consistent with a $U(1)$ R-symmetry under which $W \to e^{i\theta} W, ~S \to e^{i \theta}S,~\bar {\phi} \phi \to \bar{\phi} \phi$. \par The potential derived from the superpotential $W$ in Eq.(\ref{eq:superpot}) is \begin{eqnarray*} V=\kappa^2 \mid M^2 - \bar{\phi} \phi \mid^2 +\kappa^2 \mid S \mid^2 (\mid \phi \mid^2 + \mid \bar{\phi}\mid^2) \end{eqnarray*} \begin{equation} +{\rm{ D-terms}}. \label{eq:hybpot} \end{equation} Restricting ourselves to the D flat direction $\bar{\phi}^* =\phi$ which contains the supersymmetric vacua and performing appropriate gauge and R- transformations, we can bring $S$, $\bar{\phi}$, $\phi$ on the real axis, i.e., $S \equiv \sigma/\sqrt{2}$, $\bar{\phi}=\phi \equiv \chi/2$, where $\sigma$, $\chi$ are normalized real scalar fields. The potential then takes the form in Eq.(\ref{eq:lindepot}) with $\kappa = \lambda$ and $m=0$ and, thus, Linde's potential for hybrid inflation is almost obtainable from supersymmetric GUTs but without the mass term of $\sigma$ which is, however, of crucial importance since it provides the slope of the valley of minima necessary for driving the inflaton towards the vacua. \par One way to obtain a valley of minima useful for inflation is \cite{lp} to replace the renormalizable trilinear superpotential term in Eq.(\ref{eq:superpot}) by the next order nonrenormalizable coupling. Another way, which we will adopt here, is \cite{dss} to keep the renormalizable superpotential in Eq.(\ref{eq:superpot}) and use the radiative corrections along the inflationary valley ($\bar{\phi}=\phi = 0$~, $S > S_{c} \equiv M$). In fact, the breaking of supersymmetry by the `vacuum' energy density $\kappa^{2} M^{4}$ along this valley causes a mass splitting in the supermultiplets $\bar{\phi}$, $\phi$. This results to the existence of important radiative corrections on the inflationary valley. At one-loop, and for $S$ sufficiently larger than $S_{c}$, the inflationary potential is given \cite{dss,lss} by $$ V_{{\rm{eff}}} (S) = \kappa^2 M^4 $$ \begin{equation} \left[ 1 + \frac{\kappa^2}{16\pi^2} \left( \ln \left(\frac{\kappa^{2} S^{2}}{\Lambda^2}\right) + \frac{3}{2} - \frac{S_c^4}{12S^4} + \cdots \right)\right]~, \label{eq:veff} \end{equation} where $\Lambda$ is a suitable mass renormalization scale. \par From Eqs.(\ref{eq:veff}) and (\ref{eq:quadrupole}), we find the cosmic microwave quadrupole anisotropy: \begin{equation} \left(\frac{\delta T}{T}\right)_{Q} \approx 8 \pi \left(\frac{N_{Q}}{45}\right)^{1/2} \frac{x_{Q}}{y_{Q}}\left(\frac{M}{M_{P}}\right)^2\cdot \label{eq:qa} \end{equation} Here $N_{Q}$ is the number of e-foldings suffered by our present horizon scale during inflation and $y_{Q}= x_{Q}(1-7/(12x_{Q}^2)+\cdots)$ with $x_{Q} = S_{Q}/M$, $S_{Q}$ being the value of the scalar field $S$ when the scale which evolved to the present horizon size crossed outside the de Sitter (inflationary) horizon. Also, from Eq.(\ref{eq:veff}), one finds \begin{equation} \kappa \approx \frac{8\pi^{3/2}}{\sqrt{N_{Q}}} ~ y_{Q}~\frac{M}{M_{P}}~\cdot \label{eq:kappa} \end{equation} \par Inflation ends as $S$ approaches $S_{c}$. Writing $S=xS_{c}$, $x=1$ corresponds to the phase transition from $G$ to $G_{S}$ which, as it turns out, more or less coincides with the end of the inflationary phase (this is checked by noting the amplitude of the quantities $\epsilon$ and $\eta$ in Eq.(\ref{eq:src})). Indeed, the $50-60$ e-foldings needed for the inflationary scenario can be realized even with small values of $x_{Q}$. For definiteness, we take $x_{Q}\approx 2$. From COBE \cite{cobe} one then obtains $M \approx 5.5 \times 10^{15}~$GeV and $\kappa \approx 4.5 \times 10^{-3}$ for $N_{Q} \approx 56$. Moreover, the primordial density fluctuation `spectral index' $n \simeq 0.98$. We see that the relevant part of inflation takes place at $S \sim 10^{16}$ GeV. An important consequence of this is \cite{lyth,lss,sugra} that the supergravity corrections can be brought under control so as to leave inflation intact. \par After the end of inflation the system falls towards the supersymmetric minima, oscillates about them and eventually decays `reheating' the universe. The oscillating system (inflaton) consists of the two complex scalar fields $S$ and $\theta=(\delta \bar{\phi} + \delta\phi)/\sqrt{2}$, where $\delta \bar{\phi}=\bar{\phi}-M$, $\delta \phi= \phi-M$, with mass $m_{infl}=\sqrt{2}\kappa M$. \par In conclusion, it is important to note that the superpotential $W$ in Eq.(\ref{eq:superpot}) leads to the hybrid inflationary scenario in a `natural' way. This means that a) there is no need of extremely small coupling constants, b) $W$ is the most general renormalizable superpotential which is allowed by the gauge and R- symmetries, c) supersymmetry guarantees that the radiative corrections do not invalidate inflation, but rather provide a slope along the inflationary trajectory which drives the inflaton towards the supersymmetric vacua, and d) supergravity corrections can be negligible leaving inflation intact. \acknowledgments \par This work is supported by E.U. under TMR contract No. ERBFMRX--CT96--0090.
\section{CONCLUDING REMARKS} One practical application of the quantum formalism would be to get a deeper understanding of the polarized beams. A proposal to produce polarized beams using the proposed spin-splitter devices based on the classical Stern-Gerlach kicks has been presented recently~\cite{Splitter}. Lastly it is speculated that the quantum theory of charged-particle beam optics will be able to resolve the {\em choice of the position operator} in the Dirac theory and the related question of the {\em form of the force experienced by a charged-particle in external electromagnetic fields}~\cite{Heinemann},~\cite{Barut}. This will be possible provided one can do an extremely high precision experiment to detect the small differences arising in the transfer maps from the different choices of the position operators. These differences shall be very small, i.e., proportional to powers of the de Broglie wavelength. It is the extremely small magnitude of these minute differences which makes the exercise so challenging and speculative!
\section{Introduction} Recently intensive R\&D on scintilator tile-fiber readouts is being carried out in order to satisfy the needs of calorimeters in new LHC experiments \cite{atlas}, \cite{cms}. The specific requirements for these types of detectors are: \begin{itemize} \item Operation in magnetic field up to $4Tesla$. \item Large linear dynamic range - ${10}^{5}$ for the detector preamplifier couple. \item Long lifetime of the photodetectors - $5$ to $10$ years of operation at high luminosity. \item Radiation hardness - up to $2Mrad$ integrated dose. \item Small size - because of very large number of channels in use. \item Sufficient $Signal/Noise$ ratio - to measure the signal from a minimum ionising particle ($MIP$). \item Capability of measuring the signal generated by a radioactive source as a DC current to a precision of $1\%$. \item Reasonable price. \end{itemize} Two main types of photodetectors satisfy the above requirements: hybrid photodiodes (HPD) and avalanche photodiodes (APD). The advantages of avalanche photodiodes over the other types of photodetectors are: \begin{itemize} \item Insensitivity to magnetic field. \item Linear dynamic range of ${10}^{6}$. \item Fast response ($<1ns$). \item High quantum efficiency in the range from $200nm$ to $1100nm$. \item Small size ($10\times 10\times 5{mm}^{3}$). \item Low price of APD and preamplifier. \end{itemize} Possible disadvantages of the photodiodes are low internal gain ($50-300$) compared to the PMT and HPD one and relativelly high excess noise factor.\newline The present paper is devoted to the investigation of the applicability of APD's as a readout for scintilator tile-fiber systems. The choice of the scintilator tiles design has been determined by the requirements for two of the LHC experiments under preparation - CMS and LHCb. The main goal of our research was to achieve $Signal/Noise$ ratio enough to measure the $MIP$ signal from scintilator tile fiber and good ${e}^{-}/MIP$ separation needed for the preshower detector of the LHCb experiment. For this purpose, various designs of the scintilator tile-fiber system have been developed (section \ref{section1}). The APD characteristics, electronics and calibration of the systems under investigation are presented in sections \ref{sec_apd}, \ref{sec_preamp} and \ref{section2} correspondently. The results from measurements performed with cosmic muons and muon beams at SPS accelerator are reported in section \ref{section3}. \newline It is shown that cooling of the photodiodes reduces dark current and increases gain hence allowing us to achieve by times higher $Signal/Noise$ ratio and to reach our goals. \section{Scintillator tile - fibre system} \label{section1} Two different designs of scintillator tile - fibre system have been studied (fig.\ref{fig1}). The first one is proposed for CMS HCAL \cite{hcal_tdr} while the second one is under investigation for the preshower detector in LHCb experiment \cite{lhcb_tdr}. \begin{figure}[p] \input{fig1.pic} \caption{Two designs of the scintillator tile - fibre system. (a) scintillator size $4\times 4\times 1{cm}^{3}$; (b) scintillator size $22\times 22\times 0.4{cm}^{3}$.} \label{fig1} \end{figure} Three different types of WLS fibers - green Kuraray Y11 /$0.84mm$ diameter/, green Bicron 91A /$1mm$/ and red Bicron 99-172 /$1mm$/ have been used. In order to estimate the scintilators light output, cosmic muons measurments using PMT FEU85 have been performed. The WLS fibers were directly coupled to the PMT window. The signal from PMT was read out by charge sensitive ADC LeCroy 2249W with $250fC/channel$ sensitivity using a $160 ns$ gate triggered by two scintillator counters. Calibration of PMT was done by fitting the single photoelectron distribution, produced by $15 ns$ light pulses of blue LED \cite{single_phe}. The average number of photoelectrons induced by cosmic muons are presented in Table \ref{tabl1}. \begin{table}[h] \caption{Average number of photoelectrons induced by cosmic muons.} \begin{tabular}{|c|c|} \hline Tile-fiber configuration & Light yield, ph.e. \\ \hline $22\times 22\times 0.4{cm}^{3}$ scintillator with 1 coil of Bicron 91A fiber & $4.0$ \\ \hline $22\times 22\times 0.4{cm}^{3}$ scintillator with 3 coils of Bicron 91A fiber & $7.8$ \\ \hline $4\times 4\times 1{cm}^{3}$ scintillator with 8 coils of Bicron 91A fiber & $11.1$ \\ \hline \end{tabular} \label{tabl1} \end{table} Taking into account the PMT photocathode quantum efficiency at $500nm$ emission peak of green WLS fibers ($\approx 6\%$) \cite{feu} one can estimate $\approx 180$ photons from $4\times 4\times 1{cm}^{3}$ scintillator with $8$ coils of fiber. Coupling of WLS fiber to a clear fiber with the help of optical connector leads to the reduction of the light output at the level of $20\%-35\%$. In order to improve the light collection we have matted the scintillator surface on which the WLS fibers were placed. This gaves a $20\%$ increase of the light yield. \section{Avalanche Photodiodes} \label{sec_apd} The avalanche photodiodes under investigation are manufactured by "InterQ" Ltd. using planar technology on a $p-$type high resistivity silicon with a resistivity of $2k\Omega cm$ \cite{apd1},\cite{apd2}. They are ${n}^{+}-p-\pi-{p}^{+}$ type (R'APD).The multiplication region of the diodes is produced by ion implantation and two stage diffusion of boron and phosphorus (fig.\ref{fig9}). The active surface of the detectors is $0.5 {mm}^{2}$. Some characteristics of the diodes are presented in Table \ref{tabl2}. \begin{table}[h] \caption{Avalanche photodiode characteristics at room temperature.} \begin{tabular}{|c|c|} \hline Thickness [$\mu m$] & $140$ \\ \hline Active dia. [mm] & $0.8$ \\ \hline Capacity [$pF$] at $100V$ & $0.92$ \\ \hline Breakdown voltage [$V$] & $250-380$ \\ \hline Dark current [$nA$] at $100V$ & $<1$ \\ \hline \end{tabular} \label{tabl2} \end{table} \begin{figure}[p] \input{fig2.pic} \caption{A transverse view of the avalanche photodiode. (1) high resistivity $p-Si$; (2) $p-$region; (3) ${n}^{+}-$region; (4) $n-$region; (5) ${p}^{+}-$region; (6) $p-$region; (7) dielectric cover; (8),(9) metal layers.} \label{fig9} \end{figure} \section{Electronics} \label{sec_preamp} The schematic view of the specialy designed for our investigation electronics is presented on fig.\ref{fig8}. \begin{figure}[p] \scalebox{0.95}{\includegraphics*[50,60][520,740]{./ preamp.eps}} \caption{Schematic view of the electronics.} \label{fig8} \end{figure} The chain consists of a charge-sensitive preamplifier, CR RC shaper and current feedback amplifier, which can drive a back terminated $50\Omega$ line. A JFET SST309 transistor with high forward transconductance (determined by the very short shaping time - 25 ns) about 15 mS per 10 mA drain current and a relatively low input capacity - 6 pF has been used. The noises of the shaping amplifier become significant (the noise bandwidth is 10 MHz) due to the short shaping time. For this reason, a microwave bipolar transistors with low ${R}_{BB'}$ have been used. The thermostability of the preamplifier was achivied via negative feedback. \newline Main parameters of the preamplifier are: \begin{itemize} \item Equivalent charge noise $\approx 400 {e}^{-}$ at $1pF$ detector capacity (fig.\ref{fig14}). \begin{figure}[p] \epsfig{file=./ enc_c.eps} \caption{The equivalent noise charge of the preamplifier as a function of the detector capacity.} \label{fig14} \end{figure} \item Sensitivity of the preamplifier is $160mV/{10}^{6}{e}^{-}$. (This coefficient depends on value of capacitor $C1$). \item $25ns$ shaping time. \item Full width of output signal $\approx 150ns$. \item Maximum output voltage about $3.5V$. \end{itemize} \section{Calibration of the readout} \label{section2} The calibration of the readout has been done using $15ns$ long blue light pulses emited by LED. The light has been splited up between APD and PMT readouts throught a special optical connector. The signals from both readouts have been registrated by qADC with $160ns$ gate. The measurements have been done at different light intensities, reverse bias voltages and temperatures . The summary results are presented below: \begin{itemize} \item The normalized internal gain of APD as a function of applied bias voltage is shown on fig.\ref{fig2}. The gain at $60V$ and ${22}^{o}C$ is about $6$. Cooling of the photodiodes leeds to breakbown voltage decrease and to a valuable gain increase mainly at voltages near to the breakdown. \item The APD and preamplifier noise in terms of ADC channels vs. applied bias voltage is ploted on fig.\ref{fig3}. At low temperatures the noise remains low up to particular voltage near to the breakdown and then increases drastically. On the other hand at high temperatures, noise is higher and increases smoothly. The equivalent noise charge can be estimated taking into account that the preamplifier gain is $68 {e}^{-}$ per ADC channel. The small increase of noise at low bias voltages is due to larger capacity of photodiodes which is a result of the absence of full depletion of charges in APD at these voltages. \item Fig.\ref{fig4} shows the Signal/Noise ratio as a function of the bias voltage. The cooling of APD allow us to increase drastically the Signal/Noise ratio. The maximum of this ratio can be achivied at bias voltages near the breakdown where the signal is already large while the noise is still low. Further increase of the voltage results in abrupt increase of the noise. \end{itemize} \begin{figure}[p] \epsfig{file=./ m_vt.eps} \caption{The APD gain normalized to the gain at $60V$ and ${22}^{o}C$ as a function of the applied bias voltage at different temperatures.} \label{fig2} \end{figure} \begin{figure}[p] \epsfig{file=./ noise_vt.eps} \caption{The APD and preamplifier noise RMS as a function of the applied bias voltage at different temperatures.} \label{fig3} \end{figure} \begin{figure}[p] \epsfig{file=./ sn_vt.eps} \caption{The Signal/Noise ratio as a function of the applied bias voltage at different temperatures. LED intensity $\approx 35$ photons.} \label{fig4} \end{figure} It's clear that even for low intensity ligth sources, cooled APD gives a considerable S/N ratio. The large increase of the gain of APD at low temperatures is a result from fact that the multiplication process in the APD is affected by the temperature. This happens since electrons lose energy to the phonons, whose energy density increases with the temperature, and at lower temperatures it takes shorter for the electrons to reach the energy required for impact ionization. \section{Measurments with muons} \label{section3} We have performed series of measurments of the scintilator tile fiber readout using cosmic muon flux and muon beam at SPS at CERN. The tests have been made with PMT and APD readouts in a wide temperature range. We have obtained $1.5-2$ times higher efficiency of Bicron 99-172 fiber over Bicron 91A fiber and $1.25-1.5$ times over Kuraray Y11 fiber with APD readout because of the higher photodiode quantum efficiency at longer wavelengths of incoming light. The results obtained with Bicron 99-172 fiber are presented on fig.\ref{fig6} and fig.\ref{fig7}. Due to the long right-handed tail in the muon signal distribution, in what follows the $S/N$ ratio is defined as a ratio of the signal distribution maximum $MPV$ (Most Probable Value) \cite{mpv} to the $\sigma$ of pedestal. The tail is determined by a low light yield from the scintillator and by the so called excess noise factor (F) of the photodiodes. Taking into account that F rises linearly with the bias voltage applied, we have to find the operational voltage at which the $S/N$ ratio is enough good, keeping the RMS of the signal distribution as low as possible. We have found that the compromise is reached at $U \approx 110 V$ (see fig.\ref{fig10}) for the photodiode used. We have calculated the $MIP$ detection efficiency (the ratio of the number of events in the signal distribution above $2\sigma$ from the pedestal to the number of all events in the signal distribution) for different bias voltages (fig.\ref{fig12}). Another important issue is the ability to separate clearly the signals from muons ($MIP$) and electrons. For example, for the $LHCb$ preshower detector the signal is considered as a $MIP$ when it is by five times lower than the muon $MPV$. Otherwise it is treated as an electron signal. A separation ratio (the ratio of the number of the events in signal distribution below $5 MPV$ to the number of all events in the signal distribution) as a function of the bias voltage is presented in fig.\ref{fig13}. For the bias voltage of 110 V we have satisfactory levels for the detection efficiency $\approx 85\%$ and a separation ratio of $\approx 95\%$. \begin{figure}[p] \epsfig{file=./ 3c155a-8.eps} \caption{Muon signal from $22\times 22\times 0.4{cm}^{3}$ scintillator with 2 coils of Bicron 99-172 fiber (${-8}^{o}C$, $155V$ bias voltage, $160ns$ gate).} \label{fig6} \end{figure} \begin{figure}[p] \epsfig{file=./ 7888_89.eps} \caption{Muon signal from $4\times 4\times 1{cm}^{3}$ scintillator with 10 coils of Bicron 99-172 fiber (${ 9}^{o}C$, $155V$ bias voltage, $160ns$ gate).} \label{fig7} \end{figure} \begin{figure}[p] \epsfig{file=./ fsrms_v.eps} \caption{The most probable value of the signal divided to $RMS$ of the signal as a function of the applied bias voltage. $4\times 4\times 1{cm}^{3}$ scintillator with 10 coils of Bicron 99-172 fiber (${-9}^{o}C$, $100ns$ gate).} \label{fig10} \end{figure} \begin{figure}[p] \epsfig{file=./ sn_gate.eps} \caption{The $Signal/Noise$ ratio as a function of the ADC gate. $4\times 4\times 1{cm}^{3}$ scintillator with 10 coils of Bicron 99-172 fiber (${-9}^{o}C$, $153V$ bias voltage).} \label{fig11} \end{figure} \begin{figure}[p] \epsfig{file=./ mipdet.eps} \caption{The $MIP$ detection efficiency as a function of the applied bias voltage. $4\times 4\times 1{cm}^{3}$ scintillator with 10 coils of Bicron 99-172 fiber (${-9}^{o}C$, $100ns$ gate).} \label{fig12} \end{figure} \begin{figure}[p] \epsfig{file=./ emsep.eps} \caption{The ${e}^{-}/MIP$ separation as a function of the applied bias voltage. $4\times 4\times 1{cm}^{3}$ scintillator with 10 coils of Bicron 99-172 fiber (${-9}^{o}C$, $100ns$ gate).} \label{fig13} \end{figure} We also have performed tests with different gates (fig.\ref{fig11}) in order to investigate the behavior of the readout system in various readout schemes. The best result for the $S/N$ ratio is when the ADC gate is near to full width of the signal ($160ns$). At narrower gates (for example $50ns$) $S/N$ decreases by $25\%$. At wider gates $S/N$ falls down rapidly because much more noise is integrated. \section{Conclusions} Avalanche photodiodes were tested as a scintilator tile-fiber readout for CMS HCAL and LHCb ECAL preshower. Various types of fibres were investigated and Bicron 99-172 one was determined to be the most efficient. A low noise charge sensitive preamplifier with $\approx 400 {e}^{-}$ equivalent charge noise was designed to gain signal from photodiodes. For LHCb ECAL preshower scintilator a $Signal/Noise$ ratio of $6.5$ and for CMS HCAL scintillator $Signal/Noise$ ratio of $1.3$ were achieved cooling the APDs down to ${-10}^{o}C$. A satisfactory $MIP$ detection efficiency of $\approx 85\%$ and an excellent ${e}^{-}/MIP$ separation of $\approx 95\%$ for preshower scintilator were reached. \section{Acknowledgements} We would like to express our gratitude to Dr. Y. Musienko for the useful discussions and consultations and to P.Slavchev ("Inter Q" Ltd.) for developing and manufacturing the cooling device. \listoffigures
\section{Introduction} It has been well known that series of doped manganites with the chemical formula {\it R}$_{1-x}${\it A}$_x$MnO$_3$ ({\it R }= La or Nd, and {\it A} = Ca, Sr, or Ba) near {\it x}~$\sim $ 0.3 undergo paramagnetic insulator to ferromagnetic metal transitions and show colossal magnetoresistance (CMR) phenomena near the transitions. The coexistence of metallicity and ferromagnetism in manganites has been explained by the double exchange (DE) interaction with a strong Hund coupling between {\it e}$_g$ carrier and {\it % t}$_{2g}$ spins. However, it has been argued that some additional degrees of freedom, such as lattice,\cite{millis} orbital,\cite{Ishihara} and disorder% \cite{sheng}, should be included to explain the CMR quantitatively. Optical measurements have been very useful to investigate the basic mechanisms of some insulator-metal transitions and associated electronic structure changes.\cite{okimoto,takenaka,Kaplan,Kim1} There have been numerous optical investigations on manganites, however, the experimental data and their interpretations are varied. Especially, there are several different views on the electrodynamic responses in the metallic states. One of such difficulties comes from sample preparation methods. For examples, Okimoto {\it et al}.\cite{okimoto} investigated optical responses of polished (La,Sr)MnO$_3$ single crystals and reported a small Drude weight in the metallic state. Later, Takenaka {\it et al}.\cite{takenaka} showed that cleaved (La,Sr)MnO$_3$ single crystals had optical spectra different from those of the polished crystals, and claimed that the small Drude weight in the polished samples should be originated from damaged surfaces during the polishing process. The problem due to the sample preparation method is more serious for Nd$_{0.7}$Sr$_{0.3}$MnO$_3$ thin films. Kaplan {\it et al}.% \cite{Kaplan} reported an optical response of Nd$_{0.7}$Sr$_{0.3}$MnO$_3$ thin film. They could not observe the Drude-like peak even in the metallic region at low temperature. Later, Quijada {\it et al}.\cite{Quijada} reported a Drude-like free carrier behavior for an oxygen-annealed Nd$_{0.7}$% Sr$_{0.3}$MnO$_3$ film. Therefore, careful surface treatments and exact optical measurements are inevitable to get correct electrodynamical responses and to understand the basic mechanism of the CMR. In this paper, we present optical conductivity spectra $\sigma (\omega )$ of Nd$_{0.7}$Sr$_{0.3}$MnO$_3$ (NSMO) {\it single crystal} for annealed and polished surfaces. Although {\it dc} conductivity values of the both surfaces were nearly the same, their $\sigma (\omega )$ were quite different. Using the oxygen annealing and gold the normalization techniques, surface strain and surface scattering effects could be removed. Our experimental data suggested that the low frequency spectral weight in the metallic states might be composed of a Drude part and a strong incoherent mid-infrared absorptions. The temperature dependence of the spectral weight supports that lattice effects are important in NSMO. \section{Experimental} The NSMO single crystal was prepared by floating zone methods. Details of crystal growth and characterizations were described elsewhere.\cite {Fernandez-Baca} Just before optical measurements, the sample was polished up to 0.3 $\mu m$ using diamond pastes. After the polishing, we carefully annealed the sample again in an O$_2$ atmosphere at 1000 $^oC$ for 1 hour. Figure 1 shows values of temperature $(T)$-dependent dc resistivity $\rho (T) $ of our NSMO single crystal, which were measured by the conventional four-probe method. The solid and the dashed lines represent the $\rho (T)$ curves of the polished and the annealed NSMO, respectively. The $\rho (T)$ curves show a typical insulator-metal transition around 198 K. It seems that overall features of $\rho (T)$ for the polished and the annealed NSMO are nearly the same except for small deviations around $T_C$. Although such an annealing procedure does not change $\rho (T)$ significantly, it seems to be very important for optical properties of the NSMO sample. As demonstrated in the next section, optical properties of the NSMO crystal became changed drastically after the oxygen annealing. Near normal incident reflectivity spectra, $R(\omega )$, were measured in a wide photon energy region of 5 meV $\sim $ 30 eV and at temperatures between 15 and 300 K. We used a conventional Fourier transform spectrophotometer between 5 meV and 0.8 eV. Above 0.6 eV, grating spectrometers were used. Especially above 6.0 eV, we used the synchrotron radiation from the Normal Incidence Monochromator beam line at Pohang Light Source (PLS). $T$% -dependent $R(\omega )$ below 6.5 eV were measured using a liquid He-cooled cryostat. To subtract surface scattering effects from $R(\omega )$, we used the gold normalization technique\cite{Jung97}: a gold film was evaporated onto the sample surface just after spectra were taken. Then, the reflectivity of the gold coated surface was measured and used for normalizing $R(\omega )$ of the NSMO sample. Using the Kramers-Kronig (KK) analysis with the normalized $R(\omega )$, we obtained $\sigma (\omega )$. For this analysis, $R(\omega )$ in the low frequency region were extrapolated with the Hagen-Rubens relation\cite{Kim97} for metallic states and with constants for insulating states. To obtain accurate $\sigma (\omega )$ in the far-infrared region, we also compared $% R(\omega )$ extrapolated using the Hagen-Rubens relation with those extrapolated using the Drude model, and we obtained nearly the same results in the far-infrared region. The $T$-dependent $R(\omega )$ spectra below 6.5 eV were smoothly connected with the reflectivity data obtained at room temperature from 6.0 to 30 eV, assuming that reflectivity in the high energy region should be nearly temperature independent. For the frequency region above our measurements, the reflectivity at 30 eV was extended up to 40 eV and then $\omega ^{-4}$ dependence was assumed. Using spectroscopic ellipsometry (SE), we obtained $\sigma (\omega )$ in the region of 1.5 $\sim $ 5.0 eV without relying on the KK extrapolation. The $\sigma (\omega )$ obtained by the SE method were quite consistent with those by the KK analysis. This fact suggests validity of our gold normalization techniques and the extrapolations used in the KK analysis. \section{Results and Discussion} \subsection{Effects of Oxygen Annealing} Although the oxygen annealing process does not change $\rho (T)$ very much, it affects the optical responses of NSMO significantly. Fig. 2(a) shows $% R(\omega )$ for the polished and the annealed surfaces of the NSMO single crystal. [Note that both of these spectra were corrected by the gold normalization technique, so differences are mainly due to the oxygen annealing effects.] In the insulating region above $T_C$, the differences in $R(\omega )$ are not significant. However, in the metallic region below $T_C$% , $R(\omega )$ for the polished surface are much lower in the frequency region from 2$\times $10$^{-2}$ to 1.5 eV. The dip in $R(\omega )$, so-called the ''plasma edge'', for the polished surface appears around 0.6 eV, which is much lower than the observed dip around 2.0 eV for the annealed surface. The differences in the polished and the annealed surfaces can be clearly seen in $\sigma (\omega )$, which are displayed in Fig. 2(b). After the annealing, the spectral weight in the infrared region becomes enhanced considerably. Although Drude-like peaks at the far-infrared region can be observed at 15 K for both surfaces, $\sigma (\omega )$ of the polished surface has a smaller spectral weight. Changes in the optical response due to the oxygen annealing were already noticed in NSMO thin films. Kaplan {\it et al}. reported that a NSMO film, whose resistivity peak temperature $T_P$ was located around 180 K, had a strong mid-infrared peak and no Drude-like free carrier peak.\cite{Kaplan} Later, the same group reported that $T_P$ of a NSMO film became about 230 K when it was annealed in an oxygen environment.\cite{Quijada} They showed that the strength of the mid-infrared peak became weakened considerably and that the Drude-like free carrier behavior appeared in the oxygen-annealed NSMO film. We think that the optical response changes in the NSMO film were related with the dc resistivity changes reported by Xiong {\it et al}.\cite {Xiong} They investigated the effects of oxygen annealing in NSMO films. With more annealing, their resistivity values became smaller and $T_P$ moved to a higher temperature. They suggested that the annealing effects in the NSMO films could be explained by a change of oxygen content and a variation of mixed Mn$^{3+}$/Mn$^{4+}$ ratio. Since $\rho (T)$ of our NSMO single crystal does not change by the annealing process, the optical response changes observed in Fig. 2 might not come from oxygen content changes. We think that the differences in optical spectra should be originated from damages due to the polishing process. Since the electron-phonon interactions are believed to be very strong in manganites, the stress applied during the polishing process could damage the surface region and deform optical responses significantly. \subsection{Effects of Gold Normalization} We also found that light scattering from the polished surface was quite severe in the NSMO sample. Figs. 3(a) and (b) show $R(\omega )$ and $\sigma (\omega )$ of the oxygen annealed NSMO sample at 15 K, respectively. The dashed lines are results without any correction, and the solid lines are results after the gold normalization. $R(\omega )$ in Fig. 3(a) show that the light scattering effects are not significant below 0.1 eV, but that they become very severe between mid-infrared and visible regions. The scattering loss effects become evident in $\sigma (\omega )$. Fig. 3(b) shows that the spectral weight becomes reduced in most of the frequency region. Especially, the spectral change below 0.5 eV becomes quite large. Recently, Takenaka {\it et al.}\cite{takenaka} compared optical reflectivity spectra of (La,Sr)MnO$_3$ cleavage surfaces with those of polished surfaces, and they showed that $\sigma (\omega )$ of the cleaved surface have a much more enhanced Drude weight. They claimed that the damage of the polished surface should result in a serious distortion in $\sigma (\omega )$. Since the optical spectra of their cleaved surface [i.e. Fig. 2 of Ref. 5] are very similar to our results after the oxygen annealing and the gold normalization processes, it can be suggested that the differences observed by Takenaka {\it et al.} were likely to be remedied by our experimental techniques. \subsection{Spectral Weight in the Insulating State} In Fig. 4, we show $T$-dependent $R(\omega )$ of the annealed NSMO single crystal from 5 meV to 10 eV. At 300 K, there are three sharp peaks originated from optic phonon modes in the far-infrared region. And, broad peaks above 0.1 eV come from the optical transitions between electronic levels. With decreasing $T$, the value of reflectivity approaches to 1.0 in the dc limit (i.e. $\omega \rightarrow $ 0) and the sharp phonon features become screened by free carrier. Note that there are strong $T$-dependence in $R(\omega )$ up to 5.0 eV. Similar behaviors were observed in other CMR manganites.\cite{Quijada,Jung99} Figure 5 shows $T$-dependent $\sigma (\omega )$ of the annealed NSMO single crystal, which were obtained by the KK analysis. In the temperature region above $T_C$, we can see two interesting features: an optical gap near 0.18 eV, and broad peaks around 1.2 and 4.5 eV. As $T$ decreases, a significant amount of spectral weight between 1.0 and 5.0 eV moves to a low energy region. Above 5.0 eV, the spectral weight is nearly $T$-independent. Due to the similarities in peak position and strength to those of charge transfer transitions in the other manganites,\cite{Jung99,Jung98,Liu} the peak around 4.5 eV can be assigned to a charge-transfer transition between the O 2{\it p} and the Mn {\it e}$_g$ bands. The large spectral weight change between 2.0 and 5.0 eV can be interpreted in terms of an interband transition between the Hund's rule split bands, i.e. {\it e}$_g^{\uparrow }$({\it t}$% _{2g}^{\uparrow }$) $\rightarrow $ {\it e}$_g^{\uparrow }$({\it t}$% _{2g}^{\downarrow }$) and {\it e}$_g^{\downarrow }$({\it t}$% _{2g}^{\downarrow }$) $\rightarrow $ {\it e}$_g^{\downarrow }$({\it t}$% _{2g}^{\uparrow }$), in the DE picture.\cite{Furukawa,Moritomo} [This notation indicates that the transitions occur between two {\it e}$_g$ bands with the same spin but different {\it t}$_{2g}$ spin background.] As $T$ decreases below $T_C$, the spins of the {\it t}$_{2g}$ electrons are pointing along one direction and the interband transition becomes weakened, which results in the decrease of the spectral weight. Now, let us focus on the spectral weights in the insulating state. Although the peak near 1.2 eV appears to be one broad peak, we claim that it should be interpreted in terms of a two-peak structure: one peak is located near 1.5 eV and the other is located below 1.0 eV. For most manganites with hole concentrations below 0.4, the 1.5 eV peak could not be seen clearly. However, there have been several reports which support the existence of the 1.5 eV peak.\cite{footnote1} From doping dependent conductivity studies on (La,Ca)MnO$_3$, Jung {\it et al.}\cite{Jung98} showed that such a peak should exist. Machida {\it et al.}\cite{machida} also observed similar peaks in transmission spectra of (Nd$_{0.25}$Sm$_{0.75}$)$_{0.6}$Sr$_{0.4}$MnO$_3$ and Sm$_{0.6}$Sr$_{0.4}$MnO$_3$ films. And, to explain optical spectra of Nd$% _{0.7}$Sr$_{0.3}$MnO$_3$, La$_{0.7}$Ca$_{0.3}$MnO$_3$, and La$_{0.7}$Sr$% _{0.3}$MnO$_3$ thin films, Quijada {\it et al.}\cite{Quijada} assumed the existence of the 1.5 eV peak. Moreover, if the broad peak around 1.2 eV is interpreted as one optical transition related to small polaron absorption, there will be some discrepancies between the optical data and other experimental data. For examples, recent transport measurement shows that a small polaron activation energy should be about 0.15 eV,\cite{Zhou} which is smaller than 0.3 eV.\cite {footnote2} When we accept the double-peak structure, we can present a schematic diagram for the spectral weight in the insulating state, which is shown in Fig. 6(a). As shown in this diagram, there are four main peaks below 5.0 eV: (i) a small polaron absorption peak below 1.0 eV, (ii) a peak centered around 1.5 eV, (iii) a broad peak centered around 3.5 eV, which is originated from interband transitions between the Hund's rule split bands, and (iv) a peak due to charge transfer transitions between the O 2{\it p} and the Mn {\it e}$_g$ bands. \subsection{Spectral Weight in the Metallic State} Figure 7 shows detailed $T$-dependent $\sigma (\omega )$ below 2.0 eV. The solid square, triangle, and circle represent the measured dc conductivities of 15, 120, and 180 K, respectively. Above $T_C$, we can find three optic phonon modes. With cooling, the optic phonon modes become screened by free carrier, but their signature can be observed even at 15 K. As $T$ decreases below $T_C$, the spectral features above 0.8 eV decrease, and those below ~0.8 eV increase significantly. The increasing spectral weight below $T_C$ seems to form an asymmetric band, whose peak position moves down with decreasing temperature. Even at 15 K, $\sigma (\omega )$ seem to make a downturn around 0.1 eV, and the corresponding dc conductivity is higher by about 1000 $\Omega ^{-1}$cm$^{-1}$ than $\sigma (\omega =0.1$ eV$)$. Similar spectral weight changes have been commonly observed in other manganites.\cite {Quijada,Boris,Kim2} It should be noted that reported behaviors of the broad peak structures in the low frequency regions (especially, below 0.5 eV at $T\ll T_C$) are varied quite a lot in earlier works. Okimoto {\it et al.}\cite{okimoto} measured optical responses of (La,Sr)MnO$_3$ single crystals, and claimed that the corresponding Drude peaks are very weak and sharp. Similar behaviors were observed in La$_{0.7}$Ca$_{0.3}$MnO$_3$ polycrystalline samples by Kim {\it et al.}\cite{Kim2} However, Quijada {\it et al.}\cite {Quijada} investigated $\sigma (\omega )$ of annealed Nd$_{0.7}$Sr$_{0.3}$MnO% $_3$, La$_{0.7}$Ca$_{0.3}$MnO$_3$, and La$_{0.7}$Sr$_{0.3}$MnO$_3$ films on LaAlO$_3$ substrates, and observed broad and strong $\sigma (\omega )$ features at the low frequency region. They attributed these features to Drude responses and coherent conductions. Boris {\it et al}.\cite{Boris} measured $\sigma (\omega )$ of a La$_{0.67}$Ca$_{0.33}$MnO$_3$ single crystal. At 78 K, its $\sigma (\omega )$ showed a very broad Drude-like behavior which was much stronger than those of polycrystalline samples{\it . }Takenaka {\it et al.} also measured $\sigma (\omega )$ of cleaved (La,Sr)MnO% $_3$ single crystals, and observed a Drude-like peak with a large spectral weight. The low temperature spectral responses observed by Quijada {\it et al.}\cite{Quijada} and Boris {\it et al}.\cite{Boris} are similar to our $% \sigma (\omega )$ data at 15 K, displayed in Fig. 7. It demonstrates that the low frequency responses of the manganites are very sensitive to the sample preparation conditions. One important question which we should address at this point is whether we should interpret the low-frequency broad absorption feature below 0.5 eV at 15 K as a single component Drude peak. There are three important experimental facts which we should notice. First, our experimental data showed the downturn of optical conductivity around 0.1 eV at $T\ll T_C$. The La$_{0.67}$Ca$_{0.33}$MnO$_3$ single crystal data by Boris {\it et al}. showed the same behavior. And, the thin film data by Quijada {\it et al.} showed similar trends, although such behaviors were neglected by the authors since the downturn regions were close to the low-frequency infrared cutoff of LaAlO$_3$ substrates. Second, the spectral weight changes of all the reported experiments, including the NSMO single crystal case displayed in Fig. 7, cannot be described by a decrease of one spectral component at a fixed frequency and an increase of a single Drude component near the zero frequency. Rather, they can be reasonably described by a continual evolution of an asymmetric peak whose position moves down as the sample changes from the insulating state to the metallic state. A similar trend was observed for the room temperature data of the cleaved (La,Sr)MnO$_3$ single crystals, investigated by Takenaka {\it et al. }As Sr doping increased and the sample became more metallic, a spectral weight increased in the low frequency region, which could be described by a continual movement of an asymmetric peak. Third, some optical data cannot provide reasonable values for electrodynamic quantities, when a single component Drude peak is assumed. For example, the single crystalline La$_{0.67}$Ca$_{0.33}$MnO$_3$ data at 78 K indicate mean free path {\it l }$=$ 3.9 $\sim $ 5.2 \AA , which is comparable to the lattice constant\ $a$. This value is rather too small for metallic conduction.\cite{Ioffe} On the other hand, assuming the two components response for La$_{0.7}$Ca$_{0.3}$MnO$_3$ polycrystals, Kim {\it % et al}.\cite{Kim2} reported that {\it l }$\sim $ 25 \AA\ at 15 K and that {\it l} becomes about 4.2 \AA\ near the metal-insulator transition temperature. Therefore, we think that the low frequency response should be interpreted in terms of two components, i.e. a coherent narrow Drude peak and an incoherent mid-infrared absorption band. Figure 6(b) shows a diagram of $\sigma (\omega )$ for $T$ $\ll $ $T_C$, which takes account of the two-component contributions. Other features of the diagram are following: the peak due to the charge transfer transition is nearly $T$-independent. And, as temperature decreases below $T_C$, the transition between the Hund's rule split bands becomes weakened and the spectral weight below 1.0 eV increases significantly. As displayed in Fig. 7, the incoherent mid-infrared absorption band is quite asymmetric and the Drude weight starts just below the characteristic phonon mode frequency. These optical features are quite consistent with coherent and incoherent absorptions of a large polaron, predicted by Emin.\cite{emin} From these observations, it can be argued that a crossover from small to large polaron states accompanies the insulator-metal transition in NSMO. Such a crossover from the small to the large polaron was also observed by other experiments, including Raman scattering\cite{yoon} and pulsed PDF measurements.\cite{louca} In this polaron crossover scenario, the strong oxygen annealing dependence of the optical responses in the metallic states might be explained. Just after the polishing, the surface region should be damaged quite significantly, which blocks the motion of incoherent polaron motion. However, after the annealing, such stress-induced damages could be repaired. More detailed investigations are needed to clarify this scenario. \subsection{Temperature Dependence of Low Energy Spectral Weight} To get a further understanding on the low energy spectral weight, we evaluated the effective number of carrier, $N_{eff}(\omega _C)$, below a cutoff frequency, $\omega _C$: \begin{equation} N_{eff}(\omega _C)=\frac{2m_e}{\pi e^2N}\int_0^{\omega _C}\sigma (\omega ^{\prime })d\omega ^{\prime }{\rm {,}} \end{equation} where $m_e$ is a free electron mass and $N$ represents the number of Mn atoms per unit volume. For actual estimation, the value of $\omega _C$ was chosen to be 0.8 eV, since the low frequency spectral weight below $\sim 0.8$ eV increases as $T$ decreases below $T_C$, as displayed in Fig. 7. Figure 8 shows the $T$-dependence of $N_{eff}(0.8$ eV$)$. Above $T_C$, $% N_{eff}$ is nearly independent of $T$. As $T$ decreases, $N_{eff}$ shows an increasing behavior. In the simple DE picture, the spectral weight of the Drude carriers should be proportional to $(1+M^2)/2$, where $M$ is the magnetization. However, measured magnetization data show a very sharp increase near $T_C$, so a quantitative agreement between $N_{eff}$ and $% (1+M^2)/2$ could not obtained. Recently, Kim, {\it et al.} investigated optical properties of (La,Pr)$_{0.7}$Ca$_{0.3}$MnO$_3$,\cite{Kim3} and found that all of $N_{eff}/T_C$ data for various values of Pr concentration can be scaled with $\gamma _{DE}(T)$, which represents the temperature dependent DE bandwidth due to spin ordering within the DE model predicted by Kubo and Ohata.\cite{Kubo} To explain this scaling behavior, they adopted the theoretical results\cite{Roder} based on a Hamiltonian including the DE interaction and Jahn-Teller (JT) electron-phonon coupling terms. They showed that $N_{eff}\approx t\xi \gamma _{DE}(T)$, where $\xi $ is the polaronic band narrowing term. The solid line in Fig. 8 represents the predicted behavior of $\gamma _{DE}(T)/\gamma _{DE}(0)$, which agrees quite well with the evaluated values of $N_{eff}(0.8$ eV$)$. This agreement suggested that both the DE interaction and the JT electron-phonon coupling play important roles in NSMO. \subsection{Electrodynamic Quantities} One of the important issues in barely conducting materials is to determine electrodynamic quantities, such as scattering rate $1/\tau $, carrier density $n$, and effective mass $m^{*}$. The mean free path $l${\it \ }and dc conductivity $\sigma $ can be written in terms of the electrodynamic quantities: \begin{equation} l=\tau \hbar (3\pi ^2n)^{1/3}/m^{*}{\rm {,}} \end{equation} and \begin{equation} \sigma =ne^2\tau /m^{*}. \end{equation} Assuming hole carrier conduction (i.e. $n=0.3$), {\it l} at 15 K was estimated to be about 13 \AA . [When we assumed electron carrier conduction (i.e. $n=0.7$), {\it l} should be about 7 \AA .] As shown in the inset of Fig. 7, we fitted the free carrier absorption with the simple Drude model and found that $1/\tau $ $\sim $ 120 $\pm $ 20 cm$^{-1}$ provided a reasonable fit to our low frequency data. Using this value of $1/\tau $, we found that $m^{*}/$ $m_e$ $\sim $ 21 $\pm $ 3. Values of $m^{*}$ could be estimated using other experimental techniques, such as specific heat measurements. Conduction carriers provide a $T$-linear term, whose coefficient $\gamma $ can be written as $\pi ^2k_B^2N(E_F)/3$, where $k_B$ and $N(E_F)$ are the Boltzman constant and density of states at the Fermi energy, respectively. In the free electron model, the corresponding linear coefficient $\gamma _0$ can be written as $(4\pi ^3m_ek_B^2V/3h^2)(3n/\pi )^{1/3}$, where $h$ is the Planck's constant. Therefore, $\gamma /\gamma _0$ should be the same as $m^{*}/m_e$. From specific heat measurements of Nd$_{0.67}$Sr$_{0.33}$MnO$_3$, Gordon {\it et al}.\cite{Gordon} reported the values of $\gamma (T=0)$ as 25 mJ/K$^2$mol. [Note that $\gamma _0$ was estimated to be around 0.9 mJ/K$^2$mol.] However, they claimed that the value of $\gamma /\gamma _0$, i.e. 26, was too large for the conduction carrier contribution, and that the coupling between Nd and Mn spins might enhance the value of $\gamma $. Note that our estimated value of $m^{*}/m_e$ for NSMO is larger than those of other manganites. From specific heat measurements, Hamilton {\it et al}.,% \cite{Hamilton} found that $\gamma /\gamma _0$ was around 5 for La$_{0.67}$Ba% $_{0.33}$MnO$_3$ and around 8 for La$_{0.8}$Ca$_{0.2}$MnO$_3$. And, from optical measurements, Kim {\it et al.},\cite{Kim2} reported $m^{*}/m_e$ $% \sim $ 13 for La$_{0.7}$Ca$_{0.3}$MnO$_3$. In the DE model, $T_C$ should be proportional to an effective transfer integral $t_{eff}$ for an electron hopping between Mn ions. And, in the tight binding calculation, $N(E_F)$ is inversely proportional to $t_{eff}$. Since $\gamma $ is proportional to $% N(E_F)$, $\gamma $ should also be inversely proportional to $t_{eff}$. From the fact that $T_C$ of NSMO is quite smaller than those of La$_{0.7}$Ca$% _{0.3}$MnO$_3$ and La$_{0.67}$Ba$_{0.33}$MnO$_3$, the large value of $m^{*}$ in NSMO seems to be reasonable. This also supports the validity of our analysis for the NSMO metallic state in terms of a small Drude weight and a strong incoherent peak. \section{Conclusion} We investigated the temperature dependent optical conductivity spectra of Nd$% _{0.7}$Sr$_{0.3}$MnO$_3$ single crystal. We found that the polishing process could damage the surface region of the crystal and deform the optical spectra quite significantly. Using the oxygen annealing process and the gold normalization technique, such a damage due to the polishing process seems to be recovered. Based on optical conductivity spectra, we provided schematic diagrams for spectral weight distributions in the metallic and the insulating regions. We found that the optical response in the metallic region should be interpreted in terms of a Drude part and a strong incoherent mid-infrared absorption. \section{Acknowledgements} We thank to N. J. Hur, Dr. H. C. Lee, and Dr. Y. Chung for useful discussion and helpful experiments. We acknowledge the financial support by the Ministry Education through Grant No. BSRI-98-2416, by the Korea Science and Engineering Foundation through Grant No. 971-0207-024-2, and through RCDAMP of Pusan National University. Experiments at PLS were supported in part by MOST and POSCO.
\section{Two--Loop Calculation of $ \Gamma_{1,1} $} In the Appendix we explicitly show the two--loop calculation for the model that describes the transverse phase transition in an unsymmetric random potential (cf. Section 3.1). The graphical elements of the diagrammatic perturbation expansion are shown in Fig. \ref{graphelem} directly for the quasi--static model which is here used to facilitate the calculation. The $ \tilde{\varphi} $--legs are indicated by an arrow and the $ q_{\parallel} $ of the vertex by a dash perpendicular to the propagator line. The mathematical expressions for the graphical elements are given by Equation (\ref{quasipropkorr}).\\ In this model we have to calculate the primitively divergent vertex functions $ \Gamma_{1,1} $ and $ \Gamma_{1,2} $. There are eight two--loop diagrams contributing to $ \Gamma_{1,1} $ (Fig. \ref{2loop11}) and 25 two--loop diagrams contributing to $ \Gamma_{1,2} $ each of which obeys causality that forbids closed propagator loops. Momentum conservation demands that at each vertex the sum over all wave vectors is zero. Evaluation of these diagrams requires integration over all internal wave vectors.\\ While the one--loop calculation is easy to perform by standard methods, the two--loop calculation involving mixed $ {\bf q}^{4} $--propagators requires more sophisticated tools and has not been described in literature. We introduce a technique that is based on an inverse Mellin transformation.\\ For $ \Gamma_{1,1} $ we have to compute the diagrams $ B^{(1,1)}_{1} $ to $ B^{(1,1)}_{8} $ from Fig. \ref{2loop11}. We denote the external momentum by $ {\bf q} $ and the internal ones to be integrated over by $ {\bf p} $ and $ {\bf k} $. As $ \Gamma_{1,1} $ is quadratically divergent because of CP--symmetry and dimensional reasons and the external $ \tilde{\varphi} $ already provides a factor $ q_{\parallel} $, the parts of the integrands being proportional to $ q_{\parallel} $ contain all singularities. Therefore, the integrands are first expanded with respect to the external momentum $ q_{\parallel} $ to first order. For simplicity, we now substitute $ \rho^{\frac{1}{2}} p_{\parallel} \rightarrow p_{\parallel}, \; \rho^{\frac{1}{2}} k_{\parallel} \rightarrow k_{\parallel} $ and from now on we omit the superscript ``$ \; \mathaccent28695{} $ " to characterize unrenormalized quantities.\\ In the next step all even powers of $ p_{\parallel} $ and $ k_{\parallel} $ in the numerator of the integrands are written as $ p_{\parallel}^{2} = [{\bf p}_{\perp}^{2}({\bf p}_{\perp}^{2} + \tau_{\perp}) + p_{\parallel}^{2}] - {\bf p}_{\perp}^{2}({\bf p}_{\perp}^{2} + \tau_{\perp}) $ ($ k_{\parallel}^{2} $ analogously). The first term of the right hand side cancels against factors in the denominator. For all odd powers the half of the integrand is reflected with respect to $ k_{\parallel} $. This step reduces the superficial degree of divergence of the $ {\bf p} $-- and $ {\bf k} $--integration by 2. It is allowed because the integration runs over the whole $ k_{\parallel} $--axis.\\ Then all integrals are reduced to the two types of integrals \begin{eqnarray} \label{I} I(\alpha,\beta,\gamma,\delta,\mu,\nu)& := & \\ & &\hspace{-70 pt} \int_{\bf p} \frac{{\bf p}_{\perp}^{2\alpha} } {[{\bf p}_{\perp}^{2}({\bf p}_{\perp}^{2} + \tau_{\perp}) + p_{\parallel}^{2}]^{\beta}} \int_{\bf k} \frac{({\bf p}_{\perp} -{\bf k}_{\perp})^{2 \gamma}}{[({\bf p}_{\perp} -{\bf k}_{\perp})^{4} + (p_{\parallel} - k_{\parallel})^{2}]^{\delta}} \frac{{\bf k}_{\perp}^{2\mu} }{[{\bf k}_{\perp}^{4} + k_{\parallel}^{2}]^{\nu}} \nonumber \end{eqnarray}\\[-20 pt] \begin{eqnarray*} F(A;\alpha,\beta,\gamma,\delta ,\mu,\nu)& := & \int_{\bf p} \frac{p_{\parallel}^{A} {\bf p}_{\perp}^{2\alpha} }{[{\bf p}_{\perp}^{2}({\bf p}_{\perp}^{2} + \tau_{\perp}) + p_{\parallel}^{2}]^{\beta}} \nonumber\\ & & \hspace{-110 pt}\cdot \int_{\bf k} \frac{k_{\parallel}{\bf k}_{\perp}^{2\mu} }{[{\bf k}_{\perp}^{4} + k_{\parallel}^{2}]^{\nu}} \left[\frac{({\bf p}_{\perp}-{\bf k}_{\perp})^{2 \gamma}}{[({\bf p}_{\perp}-{\bf k}_{\perp})^{4}+(p_{\parallel} -k_{\parallel})^{2}]^{\delta}} - \frac{({\bf p}_{\perp}-{\bf k}_{\perp})^{2 \gamma}}{[({\bf p}_{\perp}-{\bf k}_{\perp})^{4}+(p_{\parallel} +k_{\parallel})^{2}]^{\delta}} \right]\nonumber \end{eqnarray*} the arguments of which can be 0, 1, 2,...\\ As the integrands are already expanded with respect to $ q_{\parallel} $, the whole integration over $ {\bf p} $ and $ {\bf k} $ is at most logarithmically divergent. The integration variables are chosen such that the $ {\bf p} $--subintegration is always primitively convergent, i.e., the naive dimension $ \delta_{{\bf p}} $ of this integration (measured in powers of the external momentum scale $ \mu $) is negative. In the integrals of both types the $ {\bf k} $--integration is at most logarithmically divergent.\\ Due to these naive dimensions of the $ {\bf p} $--, $ {\bf k} $-- and the whole integration it is possible (and necessary for the following calculus) to set $ \tau_{\perp} = 0 $ in the $ {\bf k} $--integration, because the $ \tau_{\perp} \neq 0 $--parts only provide convergent contributions.\\ The sum of the two--loop diagrams for $ \Gamma_{1,1} $ expressed by the integral types $ I $ and $ F $ reads \begin{eqnarray} \label{a2} \sum_{i=1}^{8} B^{(1,1)}_{i} & = & 4\frac{g^{4}}{\rho^{2}} q_{\parallel}^{2} [3 I(1,3,0,1,1,2) - 3 I(1,3,0,1,3,3) - 2 I(3,4,0,1,1,2) \nonumber\\[4 pt] &+ & \hspace{-3mm} 2 I(3,4,0,1,3,3) - 2 \tau_{\perp} I(2,4,0,1,1,2) + 2 \tau_{\perp} I(2,4,0,1,3,3) \nonumber\\ &+ & \hspace{-3mm} \frac{1}{2} I(0,2,1,2,1,2) - \frac{3}{2} I(2,3,1,2,1,2) + I(4,4,1,2,1,2) \nonumber\\ & - &\hspace{-3mm} F(3;1,4,0,1,1,3) - \frac{1}{2} F(1;1,3,0,1,1,3) + F(1;1,4,1,2,0,1)] .\nonumber\\ \end{eqnarray} By factorizing the $ {\bf q}^{4} $--denominators in $ I $ und $ F $ into transverse and longitudinal parts we are in the position to apply the successful methods that are used for $ {\bf q}^{2} $--propagators and --correlators. This factorization is done by the Mellin transformation. We demonstrate the method for the integral type $ I $ in detail. The calculation of $ F $ goes analogously.\\ The Mellin transformation of the function $ (a + x)^{-\alpha} $ is \begin{equation} \label{mellin} \int_{0}^{\infty} \! dx x^{t-1} (a + x)^{-\alpha} = \frac{\Gamma(t)\Gamma(\alpha - t)}{\Gamma(\alpha)} a^{t - \alpha} \; , \end{equation} where the conditions $ a > 0 $ and $ 0 < Re(t) < Re(\alpha) $ must be fulfilled \cite{Erdelyi}. The corresponding inverse Mellin transformation \begin{equation} \label{invmellin} (a + x)^{-\alpha} = \int_{t_{0} - i\infty}^{t_{0} + i\infty} \frac{dt}{2\pi i} \left(\frac{\Gamma(t)\Gamma(\alpha - t)}{\Gamma(\alpha)} a^{t - \alpha}\right) x^{-t} \; , \end{equation} where the integration path parallel to the imaginary axis is restricted by $ 0 < t_{0} < Re(\alpha) $ \cite{Erdelyi}, proves to be the appropriate tool to factorize the denominators in $ I $.\\ First we only treat the $ {\bf k} $--integration of the integral type $ I $ (\ref{I}) and apply the inverse Mellin transformation to both denominators \begin{eqnarray} \label{ik} I_{{\bf k}} & := & \int_{\bf k} \frac{({\bf p}_{\perp} -{\bf k}_{\perp})^{2 \gamma}}{[({\bf p}_{\perp} -{\bf k}_{\perp})^{4} + (p_{\parallel} - k_{\parallel})^{2}]^{\delta}} \frac{{\bf k}_{\perp}^{2\mu} }{[{\bf k}_{\perp}^{4} + k_{\parallel}^{2}]^{\nu}} \\ & = & \int_{t_{0} - i\infty}^{t_{0} + i\infty} \frac{dt}{2\pi i} \int_{s_{0} - i\infty}^{s_{0} + i\infty} \frac{ds}{2\pi i} \frac{\Gamma(t) \Gamma(\delta - t) \Gamma(s) \Gamma(\nu - s)}{\Gamma(\delta)\Gamma(\nu)}\nonumber\\ & & \hspace{8 mm} \cdot \int_{\bf k} \frac{1}{[k_{\parallel}^{2}]^{s} [(p_{\parallel} - k_{\parallel})^{2}]^{t}} \frac{1}{[{\bf k}_{\perp}^{2}]^{2(\nu - s) - \mu} [({\bf p}_{\perp} - {\bf k}_{\perp})^{2}]^{2(\delta - t) - \gamma}} \; .\nonumber \end{eqnarray} In the integrand transverse and longitudinal momenta are now separated and only quadratic. The $ {\bf k} $--integration can now be performed with the help of the usual Feynman relations \begin{eqnarray} \label{feyn} \frac{1}{A^{\alpha}} & = & \frac{1}{\Gamma(\alpha)} \int_{0}^{\infty} ds s^{\alpha - 1} e^{-s A} \\ \frac{1}{\prod_{i}^{} A_{i}^{\alpha_{i}}} & = & \frac{\Gamma(\sum_{i}^{}\alpha_{i})} {\prod_{i}^{}\Gamma(\alpha_{i})} \int_{0}^{1} \prod_{i}^{} dx_{i} x_{i}^{\alpha_{i} - 1} \frac{\delta( \sum_{i}^{}x_{i} - 1)}{[\sum_{i}^{} x_{i} A_{i}]^{\sum_{i}^{}\alpha_{i}}} \; ,\nonumber \end{eqnarray} where $ \Gamma(\alpha) $ is Euler's $ \Gamma $ function. The result reads \begin{eqnarray} \label{ik2} I_{{\bf k}} & = & \frac{1}{2} \frac{{\cal O}_{d-1}}{(2\pi)^{d}} \Gamma(\frac{1}{2}) \Gamma(\frac{d - 1}{2}) \int_{t_{0} - i\infty}^{t_{0} + i\infty} \frac{dt}{2\pi i} \int_{s_{0} - i\infty}^{s_{0} + i\infty} \frac{ds}{2\pi i} \frac{\Gamma(t) \Gamma(\delta - t) \Gamma(s) \Gamma(\nu - s)}{\Gamma(\delta)\Gamma(\nu)}\nonumber\\ & & \cdot \frac{\Gamma(s + t - \frac{1}{2})\Gamma(\frac{1}{2} - s)\Gamma(\frac{1}{2} - t)}{\Gamma(s)\Gamma(t)\Gamma(1 - s - t)} \frac{\Gamma(2\nu + 2\delta - \mu - \gamma - \frac{d-1}{2} - 2(s + t))}{\Gamma(2(\nu - s)-\mu)\Gamma(2(\delta - t)-\gamma)}\nonumber\\ & & \cdot \frac{\Gamma(\frac{d-1}{2}-2(\nu - s)+\mu) \Gamma(\frac{d-1}{2}-2(\delta - t)+\gamma)}{\Gamma(d-1-2\nu-2\delta+\mu+\gamma+2(s + t))}\nonumber\\ & & \cdot |p_{\parallel}|^{1 - 2(s+t)} \; |{\bf p}_{\perp}|^{d - 1 + 4(s + t - \nu - \delta) + 2(\gamma + \mu)} \; . \end{eqnarray} The $ {\bf k} $--integral exists under the conditions \begin{eqnarray} \label{ik3} 2 (s_{0} + t_{0}) & > & 1\\ s_{0} < \frac{1}{2} \qquad\qquad t_{0} & < & \frac{1}{2}\nonumber\\ 4(\nu + \delta - (s_{0} + t_{0})) - 2(\mu + \gamma) & > & d - 1\nonumber\\ d - 1 - 4(\nu - s_{0}) + 2\mu & > & 0\nonumber\\ d - 1 - 4(\delta - t_{0}) + 2\gamma & > & 0 \nonumber \end{eqnarray} which prevent UV and IR divergences, respectively, at the $ k_{\parallel} $-- and $ {\bf k}_{\perp} $--integration. These conditions restrict the complex integration paths of the $ s $-- and $ t $--integration which are here dependent on the spatial dimension, due to the dimensional regularization.\\ The result of the $ {\bf k} $--integration is according to Equation (\ref{ik2}) proportional to powers of $ |p_{\parallel}| $ and $ |{\bf p}_{\perp}| $ with exponents that are dependent on $ s $, $ t $, and $ d $. Thus, the remaining $ {\bf p} $--integration is analogous to the one--loop problem, up to changed exponents. It is straightforwardly performed and we obtain for $ I $ \begin{eqnarray} \label{ik4} I(\alpha,\beta,\gamma,\delta,\mu,\nu) & = & C_{d} \frac{\Gamma(2(\beta + \delta + \nu) - (\alpha + \gamma + \mu) - (d + 1))}{\tau_{\perp}^{2(\beta + \delta + \nu) - (\alpha + \gamma + \mu) - (d + 1)}}\nonumber\\ & & \hspace{-35 mm} \cdot \int_{t_{0} - i\infty}^{t_{0} + i\infty} \!\!\! \frac{dt}{2\pi i} \int_{s_{0} - i\infty}^{s_{0} + i\infty} \!\!\! \frac{ds}{2\pi i} \Gamma(2\nu + 2\delta - \mu - \gamma - \frac{d-1}{2} - 2(s+t))\Gamma(s + t - \frac{1}{2})\nonumber\\ & & \hspace{-35 mm} \cdot \Gamma(\delta - t) \Gamma(\nu - s) \Gamma(\frac{1}{2} - s) \Gamma(\frac{1}{2} - t) \frac{\Gamma(d+s+t+\alpha+\gamma+\mu-\beta-2\delta-2\nu)} {\Gamma(d-1-2\nu-2\delta+\mu+\gamma+2(s+t))}\nonumber\\ & & \cdot \frac{\Gamma(\frac{d-1}{2}-2(\nu-s)+\mu) \Gamma(\frac{d-1}{2}-2(\delta-t)+\gamma)} {\Gamma(2(\nu-s)-\mu)\Gamma(2(\delta-t)-\gamma)} \; , \end{eqnarray} where $ C_{d} = \frac{1}{4} \left(\frac{{\cal O}_{d-1}}{(2\pi)^{d}}\right)^{2}\frac{\Gamma(\frac{1}{2}) \Gamma(\frac{d - 1}{2})}{\Gamma(\beta)\Gamma(\delta)\Gamma(\nu)} $ is a constant that is different for every $ I $ and depends on the dimension. The conditions for the existence of the $ {\bf p} $--integration are \begin{eqnarray} \label{ik5} 1 - (s_{0} + t_{0}) & > & 0 \\ d + s_{0} + t_{0} + \alpha + \gamma + \mu - \beta - 2\delta - 2\nu & > & 0 \nonumber\\ 2(\beta + \delta + \nu) - (\alpha + \gamma + \mu) - (d + 1) & > & 0 \; . \nonumber \end{eqnarray} After the momentum integrations there remain parameter integrations with respect to the Mellin variables $ s $ and $ t $ over parallels to the imaginary axis. The complex integration paths and the spatial dimension $ d $ are restricted by the conditions (\ref{ik3}) and (\ref{ik5}). With $ \epsilon = 9 - d $ we obtain from the first four inequalities of (\ref{ik3}) \begin{equation} \label{ik6} \frac{1}{2} < s_{0} + t_{0} < \delta + \nu - \frac{1}{2} (\gamma + \mu) - 2 + \frac{\epsilon}{4} \; , \end{equation} where $ s_{0} < \frac{1}{2},\; t_{0} < \frac{1}{2} $. All other conditions are satisfied for all $ I $ of Equation (\ref{a2}), if $ 0 < \epsilon < 2 $.\\ Since the sum $ s_{0} + t_{0} $ appears in the inequality (\ref{ik6}) we transform to the new integration variables \begin{equation} \label{ik7} z := s + t \qquad w := \frac{1}{2} (s - t) \; . \end{equation} The integration paths are now determined by \begin{eqnarray} \label{ik8} \frac{1}{2} \quad < \quad z_{0} & < & \delta + \nu - \frac{1}{2} (\gamma + \mu) - 2 + \frac{\epsilon}{4}\\ |w_{0}| & < & \frac{1}{4} \; .\nonumber \end{eqnarray} After this substitution Equation (\ref{ik4}) gives \begin{eqnarray} \label{ik9} I(\alpha,\beta,\gamma,\delta,\mu,\nu) & = & C_{d} \frac{\Gamma(2(\beta + \delta + \nu) - (\alpha + \gamma + \mu) - (d + 1))}{\tau_{\perp}^{2(\beta + \delta + \nu) - (\alpha + \gamma + \mu) - (d + 1)}} \\ & &\hspace{-38 mm} \cdot \int_{z_{0} - i\infty}^{z_{0} + i\infty} \frac{dz} {2\pi i} \int_{w_{0} - i\infty}^{w_{0} + i\infty} \frac{dw} {2\pi i} \Gamma(2\nu + 2\delta - \mu - \gamma - \frac{d-1}{2} - 2z)\Gamma(z-\frac{1}{2}) f(z,w;\epsilon) ,\nonumber \end{eqnarray} where the abbreviation \begin{eqnarray} \label{ik10} f(z,w;\epsilon) & := & \Gamma(\delta - \frac{z}{2} + w) \Gamma(\nu - \frac{z}{2} - w) \Gamma(\frac{1}{2} - \frac{z}{2} - w)\Gamma(\frac{1}{2} - \frac{z}{2} + w) \nonumber\\ & & \hspace{-1 mm}\cdot\frac{\Gamma(\frac{8-\epsilon}{2}-2(\nu-\frac{z}{2}-w)+\mu) \Gamma(\frac{8-\epsilon}{2}-2(\delta-\frac{z}{2}+w)+\gamma)} {\Gamma(2(\nu-\frac{z}{2}-w)-\mu)\Gamma(2(\delta-\frac{z}{2}+w)-\gamma)}\nonumber\\ & & \hspace{-1 mm}\cdot\frac{\Gamma(9-\epsilon +z+\alpha+\gamma+\mu-\beta-2\delta-2\nu)} {\Gamma(8-\epsilon -2\nu-2\delta+\mu+\gamma+2z)} \end{eqnarray} denotes the part of the integrand that is free of poles.\\ To extract the divergent parts of $ I $ we have to distinguish two cases.\\ Case 1: The $ {\bf k} $--integration is primitively convergent.\\ In this case the parameters satisfy \begin{equation} \label{ik11} \nu + \delta - \frac{1}{2}(\gamma + \mu) = 3 \; , \end{equation} which is true for $I(0,2,1,2,1,2), \; I(2,3,1,2,1,2) $, and $ I(4,4,1,2,1,2) $ from Equation (\ref{a2}). According to Equation (\ref{ik8}) this means for the constant $ z_{0} $ which determines the integration path \begin{equation} \label{ik12} \frac{1}{2} < z_{0} < 1 + \frac{\epsilon}{4} \; . \end{equation} For these $I $ the whole integration over $ {\bf p} $ und $ {\bf k} $ is logarithmically divergent so that their parameters fulfill the equation \begin{equation} \label{ik15} 2(\beta + \delta + \nu) - (\alpha + \gamma + \mu) - (d + 1) = \epsilon \; . \end{equation} Thus, the coefficient $ \Gamma(\epsilon) = \frac{1}{\epsilon} \Gamma(1+\epsilon) $ of the double integral over $ z $ und $ w $ (\ref{ik9}) contains an $ \epsilon $--pole, whereas the double integral itself is convergent, as the integrand is free of poles even in the limit $ \epsilon \rightarrow 0 $ and the real part of the integration path can be chosen between $\frac{1}{2} $ and 1 according to Equation (\ref{ik12}). The double integral in Equation (\ref{ik9}) can therefore be computed at $ \epsilon = 0 $, because together with the $ \epsilon $--pole as coefficient only convergent contributions are produced for $ \epsilon \neq 0 $. The convergence of the double integral is ensured by the asymptotic behaviour of the $ \Gamma $ function with complex arguments \cite{Gradshteyn} \begin{equation} \label{gammaasymp} \lim_{|y| \rightarrow \infty} |\Gamma(x + iy)| \; e^{\frac{\pi}{2}|y|} \; |y|^{\frac{1}{2} - x} \; = \; (2\pi)^{\frac{1}{2}} \; . \end{equation} Due to the accumulation of $ \Gamma $ functions we are not able to perform the integrations with respect to $ z $ and $ w $ analytically. Hence, each of the three $ I $ is computed numerically for $ \epsilon = 0 $.\\ Case 2: The $ {\bf k} $--integration is logarithmically divergent.\\ In this case the parameters are restricted to \begin{equation} \label{ik17} \nu + \delta - \frac{1}{2}(\gamma + \mu) = \frac{5}{2} \; , \end{equation} which is correct for the remaining $I $ from Equation (\ref{a2}). According to Equation (\ref{ik8}) the integration path is here restricted by \begin{equation} \label{ik18} \frac{1}{2} < z_{0} < \frac{1}{2} + \frac{\epsilon}{4} \; . \end{equation} In the limit $ \epsilon \rightarrow 0 $ the integration path is trapped by this condition between the poles of the integrand (\ref{ik9}) at $ z = \frac{1}{2} $ and $ z = \frac{1}{2} + \frac{\epsilon}{4} $. In order to extract the $ \epsilon $--poles the complex integration path is decomposed into two parts (as shown graphically in Fig. \ref{intpath}): The first part is a line parallel to the imaginary axis with an arc to the left of the singularity $ z = \frac{1}{2} $, while the second part is a circle around $ z = \frac{1}{2} $. The $ z $--integral over the circle gives the residuum at $ z = \frac{1}{2} $.\\ This way we obtain from Equation (\ref{ik9}) \begin{eqnarray} \label{ik19} I(\alpha,\beta,\gamma,\delta,\mu,\nu) & = & C_{d} \frac{\Gamma(2(\beta + \delta + \nu) - (\alpha + \gamma + \mu) - (d + 1))}{\tau_{\perp}^{2(\beta + \delta + \nu) - (\alpha + \gamma + \mu) - (d + 1)}} \\ & &\hspace{-35 mm} \cdot \left[\int_{z_{0}^{'} - i\infty}^{z_{0}^{'} + i\infty} \frac{dz} {2\pi i} \int_{w_{0} - i\infty}^{w_{0} + i\infty} \frac{dw} {2\pi i} \Gamma(1 + \frac{\epsilon}{2} - 2z)\Gamma(z-\frac{1}{2}) f(z,w;\epsilon)\right.\nonumber\\ & &\hspace{25 mm}\left. + \; \Gamma(\frac{\epsilon}{2}) \int_{w_{0} - i\infty}^{w_{0} + i\infty} \frac{dw} {2\pi i}f(z=\frac{1}{2},w;\epsilon)\right] ,\nonumber \end{eqnarray} Naturally, the integration path of the $ z $--integration is again chosen as a parallel to the imaginary axis, whose real part $ z_{0}^{'} $ now lies between $ -\frac{1}{2} $ and $ \frac{1}{2} $. After the integration path is changed, both the double integral over $ z $ and $ w $ and the single integral over $ w $ are eventually convergent. Only the coefficients of the integrals contain the $ \epsilon $--poles. For the further calculation we have again to distinguish two cases.\\ a) The whole integration is logarithmically divergent.\\ This is the case for the integrals $ I(1,3,0,1,1,2),\; I(1,3,0,1,3,3),\; I(3,4,0,1,1,2)$, and $ I(3,4,0,1,3,3) $ from Equation(\ref{a2}). Their parameters satisfy Equation (\ref{ik15}) so that in front of the entire bracket there is an $ \epsilon $--pole due to $ \Gamma(\epsilon) $.\\ For every single $ I $ the double integral is finally calculated for $ \epsilon = 0 $ numerically and provides the coefficients of simple $ \epsilon $--poles. Parts of the integrand with $ \epsilon \neq 0 $ lead to convergent contributions and are therefore omitted.\\ The coefficient $ \Gamma(\epsilon)\Gamma(\frac{\epsilon}{2}) $ of the simple integral, however, contains an $ \epsilon^{2} $--pole. Hence, the integrand $ f(z=\frac{1}{2},w;\epsilon) $ must be expanded with respect to $ \epsilon $ up to the first order \begin{eqnarray} \label{ik20} f(z=\frac{1}{2},w;\epsilon) & = & \frac{\Gamma(\frac{19}{2}-\epsilon +\alpha+\gamma+\mu-\beta-2\delta-2\nu)} {\Gamma(9-\epsilon -2\nu-2\delta+\mu+\gamma)}\Gamma(\delta - \frac{1}{4} + w)\nonumber\\ & & \hspace{-35 mm} \cdot \Gamma(\nu - \frac{1}{4} - w) \Gamma(\frac{1}{4} - w)\Gamma(\frac{1}{4} + w) \frac{\Gamma(\frac{9-\epsilon}{2}-2(\nu-w)+\mu) \Gamma(\frac{9-\epsilon}{2}-2(\delta+w)+\gamma)} {\Gamma(2(\nu-\frac{1}{4}-w)-\mu)\Gamma(2(\delta-\frac{1}{4}+w)-\gamma)}\nonumber\\ & &\hspace{-35 mm} = \frac{\Gamma(\beta - \frac{1}{2} - \epsilon)}{\Gamma(4 - \epsilon)} \Gamma(\delta - \frac{z}{2} + w) \Gamma(\nu - \frac{z}{2} - w) \Gamma(\frac{1}{2} - \frac{z}{2} - w)\Gamma(\frac{1}{2} - \frac{z}{2} + w)\nonumber\\ & & \hspace{-25 mm} \cdot \left[1 - \frac{\epsilon}{2}(\Psi(2\delta-\frac{1}{2}+2w-\gamma) + \Psi(2\nu-\frac{1}{2}-2w-\mu))+ {\cal O}(\epsilon^{2})\right] \; , \end{eqnarray} where $ \Psi(x) = \frac{\Gamma^{'}(x)}{\Gamma(x)} $ denotes Euler's $ \Psi $ function and the relations (\ref{ik15}) and (\ref{ik17}) have been used. The integral of the zeroth order in $ \epsilon $ provides the coefficients of the $ \epsilon^{2} $--poles and can even be executed analytically \cite{Gradshteyn} \begin{eqnarray} \label{ik21} \int_{w_{0} - i\infty}^{w_{0} + i\infty} \frac{dw} {2\pi i} \Gamma(\delta - \frac{1}{4} + w) \Gamma(\nu - \frac{1}{4} - w) \Gamma(\frac{1}{4} - w)\Gamma(\frac{1}{4} + w) & = & \\ & & \hspace{-40 mm} \frac{\Gamma(\delta)\Gamma(\delta+\nu-\frac{1}{2})\Gamma(\nu)\Gamma(\frac{1}{2})}{\Gamma(\delta+\nu)} \; .\nonumber \end{eqnarray} The integral of the first order in $ \epsilon $ yields the coefficients of the $ \epsilon $--poles and is numerically calculated for every single $ I $ that belongs to this case 2 a).\\ b) The whole integration is primitively convergent.\\ This statement is true for $ \tau_{\perp} I(2,4,0,1,1,2) $ and $\tau_{\perp} I(2,4,0,1,3,3) $ in Equation (\ref{a2}). Here the parameters have the property \begin{equation} \label{ik23} 2(\beta + \delta + \nu) - (\alpha + \gamma + \mu) - (d + 1) = 1 + \epsilon \; , \end{equation} i.e. in front of the brackets in Equation (\ref{ik19}) the coefficient is $ \Gamma(1 + \epsilon) $ and consequently there is no $ \epsilon $--pole.\\ Therefore the double integral needs not to be computed, as it only leads to convergent contributions. The simple integral can be evaluated at $ \epsilon = 0 $ because the coeffient contains only a simple $ \epsilon $--pole. Due to the relation (\ref{ik17}) that is also valid here, the simple integral is reduced to the one already solved in Equation (\ref{ik21}) analytically.\\ Now all integrals of type $ I $ from Equation (\ref{a2}) have been calculated. The way to solve the integrals of type $ F $ is step by step analogous to the method presented for type $ I $. $ F(3;1,4,0,1,1,3) $ and $ F(1;1,3,0,1,1,3) $ from Equation (\ref{a2}) belong to case 1, whereas $ F(1;1,4,1,2,0,1) $ belongs to case 2 a).\\ Finally we present the sum of the two--loop diagrams for $ \Gamma_{1,1} $ up to convergent parts (cf. (\ref{uvertex})) \begin{equation} \label{ik35} \sum_{i=1}^{8} B^{(1,1)}_{i} = 4\frac{g^{4}}{\rho^{2}} q_{\parallel}^{2} A_{\epsilon}^{2} \frac{\tau_{\perp}^{-\epsilon}}{\epsilon^{2}} \left(\frac{1}{18} - 0.01199 \epsilon\right) \; . \end{equation} The 25 two--loop diagrams of the primitively divergent $ \Gamma_{1,2} $ again only lead to integrals of type $ I $ and $ F $ \begin{eqnarray} \label{ik36} \sum_{i=1}^{25} B^{(1,2)}_{i} & = & -4i\frac{g^{5}}{\rho^{3}} q_{\parallel} [3 I(1,3,0,1,1,2) - I(1,3,1,2,2,2) - 3 I(3,4,0,1,1,2) \nonumber\\ &+ & \hspace{-3mm} I(3,4,1,2,2,2) - 2 I(1,3,0,1,3,3) + 2 I(3,4,0,1,3,3)\nonumber\\ & - &3 \tau_{\perp} I(2,4,0,1,1,2) + 2 \tau_{\perp} I(2,4,0,1,3,3) + \tau_{\perp} I(2,4,1,2,2,2)\nonumber\\ &- & \hspace{-3mm} \frac{1}{2} I(0,2,1,2,1,2) + I(2,3,1,2,1,2) -\frac{1}{2} I(4,4,1,2,1,2) \nonumber\\ & - &\hspace{-3mm} F(3;1,4,0,1,1,3) + F(3;0,3,1,2,1,3) + 2 F(3;1,5,1,2,0,1)] ,\nonumber\\ \end{eqnarray} which can be evaluated with the methods demonstrated for $ \Gamma_{1,1} $. This concludes the two--loop calculation. \section{Introduction} For more than a decade the long-time and critical behaviour of diffusive systems subjected to a driving force has attracted considerable interest. This is mainly caused by the richness of their highly nontrivial features which generically result from the fact that, due to the driving force, in general even the steady states of these systems are far from thermal equilibrium, especially in the case of very strong driving forces. Further, driven diffusive systems might be suitable models for fast ionic conductors which was first suggested by Katz, Lebowitz, and Spohn \cite{KLS1,KLS2}. A review of the different investigations on driven diffusive systems and their relations to other non--equilibrium systems is given by Schmittmann and Zia \cite{SZ}.\\ We are interested in the diffusive motion of uniformly driven particles with short range attractive interactions and hardcore repulsion. With such an interaction driven diffusive systems show two kinds of phase transitions. At the transverse (with respect to the driving force $ {\bf E} $) phase transition the system changes from a disordered state to an ordered one where the system is ordered in the transverse direction and remains disordered in the longitudinal (parallel to the driving force) direction. The typical configurations are strips of high-- and low--density phase, arranged parallel to the driving force ${\bf E}$ (Fig. \ref{uebergang}(a)). At the longitudinal phase transition the systems changes from a disordered state to a state where the system is ordered in the longitudinal direction and remains disordered in the transverse directions. Here the typical configurations are domains ("pancakes") of different phase with interfaces perpendicular to the driving force (Fig. \ref{uebergang}(b)). In preceding papers the long--time and critical behaviour of driven diffusion in an ordered medium \cite{JS1,JS2,LC} and the long--time behaviour of driven diffusion in a medium with quenched disorder \cite{BJ1} have been investigated by renormalized field theory. In the present paper we study the effect of quenched disorder on the two kinds of phase transitions in driven diffusive systems. Quenched disorder is important in real systems, as it models impurities and defects. We stress that this work completes the investigations of a whole model class that is graphically shown in the summary (Fig. \ref{ahnen1}). Note that the effects of quenched disorder in randomly driven diffusive systems were recently analyzed \cite{SchmBass,SchmLab}. There the driving force is itself a locally random variable with respect to its amplitude and sign, whereas here the driving force is uniform.\\ The paper is organized as follows: In Section II we set up a Langevin description of our system, directly on a mesoscopic length scale using conservation laws and symmetry arguments. The quenched disorder is modelled by random potential barriers between sites, and the symmetry properties of the random potential play a crucial role. Different realizations of the random potential are possible and are argued to lead to different kinds of noise. In Section III we study the transverse phase transition in the convenient formalism of the dynamic functional. In Section IV the longitudinal phase transition is analyzed. Section V contains the main results and a graphical overview of the entire model class. \section{Longitudinal Phase Transition} The three different realizations of the random potential I--III described in Section 2 converge to a single model in the region of the longitudinal phase transition as is sketched in the following.\\ For the three realizations of the random potential the dynamic functional still containing many irrelevant terms is given by Eq. (\ref{udynfunktional}) and by Eq. (\ref{bdynfunkt}) with $ \alpha \neq 0 $ and $ \alpha = 0 $, respectively. The longitudinal phase transition is characterized by $ \tau_{\parallel} \rightarrow 0 $ and finite $ \tau_{\perp} $. The external scale $ \mu^{2} $ here measures small $ \tau_{\parallel} $. Comparing the leading gradient terms in driving force direction ($ \sim \lambda \tilde{s} \rho \Delta_{\parallel} \kappa_{\parallel}\Delta_{\parallel} s $) and transverse direction ($ \sim \lambda \tilde{s} \tau_{\perp} \Delta_{\perp} s $) we find for length scales \begin{equation} \label{lskalierung} r_{\parallel} \sim \mu^{-1} \qquad r_{\perp} \sim \mu^{-2} \; . \end{equation} As the dynamic functional is dimensionless the coupling constants of the transverse noise scale as $ \gamma \sim \mu^{-2} $, $ \alpha \sim \mu^{-6} $, i.e., only the longitudinal noise is relevant. Hence, only a single model is required to describe the longitudinal phase transition, independent of the kind of the random potential. Further, the dimensions of fields and the other coupling constants are \begin{eqnarray} \label{ldimension} \lambda t \sim \mu^{-4}\qquad &\qquad s \sim \mu^{d - \frac{7}{2}}\qquad &\qquad \tilde{s} \sim \mu^{d + \frac{5}{2}} \nonumber\\ \kappa_{\perp} \sim \mu^{-4}\qquad & \qquad\kappa \sim \mu^{-2}\qquad & \qquad\sigma \sim \mu^{0} \nonumber\\ \qquad &\qquad g \sim \mu^{\frac{13}{2} - d}\; ,\qquad & \qquad \end{eqnarray} indicating that $ \kappa_{\perp} $ and $ \kappa $ are irrelevant and that \begin{equation} \label{lgrenzdim} d_{c} = 6.5 \end{equation} is the upper critical dimension of this model.\\ After a suitable scale change of fields and lengths the relevant dynamic functional for the longitudinal phase transition is \begin{eqnarray} \label{lJ} {\cal J}[s,\tilde{s}] & = & \int\!\! d^{d}r \left\{\int\!\! dt \left[\tilde{s} \dot{s} + \lambda \tilde{s} (- \Delta_{\perp} + \rho \Delta_{\parallel}(\rho \Delta_{\parallel} - \tau_{\parallel})) s + \frac{\displaystyle 1}{\displaystyle 2} \lambda g (\nabla_{\parallel} \tilde{s}) s^{2} \right]\right. \nonumber\\ & & \hspace{180 pt}\left. - \sigma \left[\lambda \int\!\! dt \nabla_{\parallel} \tilde{s} \right]^{2} \right\} \; . \end{eqnarray} The symmetry properties are the same as in the last model of Section 3 inclusively the invariant longitudinal scale transformation, because transverse noise exists in neither model. As a consequence $ g^{2}\sigma \rho^{-\frac{5}{2}} $ is the appropriate invariant variable of this model.\\ Due to the quenched disorder, we transform to a quasistatic Hamiltonian \begin{equation} \label{lquasihamil} {\cal H}[\varphi,\tilde{\varphi}] = \int\!\! d^{d}r \left\{\tilde{\varphi} (-\Delta_{\perp} + \rho \Delta_{\parallel}(\rho \Delta_{\parallel} - \tau_{\parallel})) \varphi + \frac{1}{2} g (\nabla_{\parallel} \tilde{\varphi}) \varphi^{2} + \sigma \tilde{\varphi} \Delta_{\parallel} \tilde{\varphi} \right\} \; . \end{equation} From this equation we directly read off the elements of perturbation theory, i.e., the vertex $ -igq_{\parallel} $ and the Gaussian propagator and correlator \begin{eqnarray} \label{lpropkorr} G_{{\bf q}} & = & \frac{1}{{\bf q}_{\perp}^{2} + \rho q_{\parallel}^{2}(\rho q_{\parallel}^{2} +\tau_{\parallel}) } \\ C_{{\bf q}} & = &\frac{2 \sigma q_{\parallel}^{2} }{[{\bf q}_{\perp}^{2} + \rho q_{\parallel}^{2}(\rho q_{\parallel}^{2} +\tau_{\parallel})]^{2}} \; .\nonumber \end{eqnarray} The vertex functions $\Gamma_{1,1} $, $ \Gamma_{2,0} $, and $ \Gamma_{1,2} $ are primitively divergent. The singular parts of $\Gamma_{1,1} $ are proportional to $ q_{\parallel}^{4} $ and $ q_{\parallel}^{2} $, the singular parts of $ \Gamma_{2,0} $ and $ \Gamma_{1,2} $ are proportional to $ q_{\parallel}^{2} $ and $ q_{\parallel} $, respectively.\\ In a one--loop calculation we obtain the primitively divergent regularized vertex functions in bare quantities up to higher orders in $ q $ and $ \epsilon = d_{c} - d $ \begin{eqnarray} \label{lvertex} \mathaccent28695{\Gamma}_{1,1}({\bf q}) & = & {\bf q}_{\perp}^{2} + \mathaccent28695{\rho} q_{\parallel}^{2}(\mathaccent28695{\rho} q_{\parallel}^{2} + \mathaccent28695{\tau}_{\parallel}) + \frac{C_{\epsilon}}{\epsilon} \frac{\mathaccent28695{g}^{2}\mathaccent28695{\sigma}}{\mathaccent28695{\rho}^{\frac{3}{2}}} q_{\parallel}^{2} \mathaccent28695{\tau}_{\parallel}^{-\epsilon}(\frac{1}{6}\mathaccent28695{\rho} q_{\parallel}^{2} + 0 \cdot \mathaccent28695{\tau}_{\parallel}) \\ \mathaccent28695{\Gamma}_{2,0}({\bf q}) & = & -2 \mathaccent28695{\sigma} q_{\parallel}^{2} -\frac{1}{6} \frac{C_{\epsilon}}{\epsilon} \frac{\mathaccent28695{g}^{2}\mathaccent28695{\sigma}^{2}}{\mathaccent28695{\rho}^{\frac{5}{2}}} q_{\parallel}^{2} \mathaccent28695{\tau}_{\parallel}^{-\epsilon}\nonumber\\ \mathaccent28695{\Gamma}_{1,2}({\bf q}) & = & i \mathaccent28695{g} q_{\parallel} - i \frac{1}{6} \frac{C_{\epsilon}}{\epsilon} \frac{\mathaccent28695{g}^{3}\mathaccent28695{\sigma}}{\mathaccent28695{\rho}^{\frac{5}{2}}} q_{\parallel} \mathaccent28695{\tau}_{\parallel}^{-\epsilon} \; , \nonumber \end{eqnarray} where \begin{equation} \label{lgeofaktor} C_{\epsilon} = \frac{1}{(2\pi)^{d}} {\cal O}_{d-1} \Gamma(1 + \epsilon) \Gamma(\frac{11 - 2\epsilon}{4}) \Gamma(\frac{5 - 2\epsilon}{4}) \end{equation} is a suitably chosen $ \epsilon $--dependent factor. These divergences are absorbed in the redefinition of the coupling constants \begin{equation} \label{lreparameter} \mathaccent28695{\rho} \; = \; Z_{\rho} \rho \qquad \mathaccent28695{\tau}_{\parallel} \; = \; Z_{\tau_{\parallel}} \tau_{\parallel} \qquad \mathaccent28695{\sigma} \; = \; Z_{\sigma} \sigma \qquad \mathaccent28695{g} \; = \; \mu^{\epsilon} Z_{u} u \; . \end{equation} In minimal subtraction we obtain the $ Z $--factors as a function of the dimensionless renormalized invariant variable $ v:= C_{\epsilon} u^{2} \sigma \rho^{-\frac{5}{2}} $ \begin{eqnarray} \label{lZ} Z_{\rho} & = & 1 - \frac{1}{12} \frac{v}{\epsilon} + {\cal O}(v^{2}) \\ Z_{\tau_{\parallel}} & = & 1 + \frac{1}{12} \frac{v}{\epsilon} + {\cal O}(v^{2}) \nonumber\\ Z_{\sigma} & = & 1 - \frac{1}{12} \frac{v}{\epsilon} + {\cal O}(v^{2}) \nonumber\\ Z_{u} & = & 1 + \frac{1}{6} \frac{v}{\epsilon} + {\cal O}(v^{2}) \; .\nonumber \end{eqnarray} Here, the renormalization group equation is of the form \begin{equation} \label{lrgg} \left [\beta_{v} \frac{\partial}{\partial v} + \rho \zeta_{\rho} \frac{\partial}{\partial \rho} +\sigma \zeta_{\sigma} \frac{\partial}{\partial \sigma} + \kappa \tau_{\parallel}\frac{\partial}{\partial \tau_{\parallel} } + \mu \frac{\partial}{\partial \mu}\right ] \Gamma_{\tilde{n},n} (\{q_{\parallel},{\bf q}_{\perp}\},\tau_{\parallel},v,\rho,\sigma,\mu) = 0 \; \end{equation} with the Wilson functions \begin{eqnarray} \label{lwilson} \beta_{v} & := & \left. \mu \frac{\partial v}{\partial \mu}\right |_{o} \; = -2v \left[\epsilon - \frac{11}{24} v + {\cal O}(v^{2})\right]\\ \zeta_{\sigma} & := & \left. \mu \frac{\partial \ln \sigma}{\partial \mu}\right |_{o} = - \frac{1}{6} v + {\cal O}(v^{2})\nonumber\\ \zeta_{\rho} & := & \left. \mu \frac{\partial \ln \rho}{\partial \mu}\right |_{o} = - \frac{1}{6} v + {\cal O}(v^{2}) \nonumber\\ \kappa & := & \left. \mu \frac{\partial \ln \tau_{\parallel}}{\partial \mu}\right |_{o} = \frac{1}{6} v + {\cal O}(v^{2}) \; . \nonumber \end{eqnarray} In addition to the flow equations for $ \bar{v}(l) $, $ \bar{\rho}(l) $, $ \bar{\sigma}(l) $, and $ \bar{\mu}(l) $ being of the same form as in the preceding model there is a flow equation for $ \bar{\tau}_{\parallel}(l) $ that reads \begin{equation} \label{lfluss} \frac{d}{dl} \ln \bar{\tau}_{\parallel}(l) = \kappa(\bar{v}(l)) \; . \end{equation} In the scaling limit $ l \ll 1 $ that corresponds to the relations \begin{equation} \label{lskalenlimes} \left |\frac{q_{\parallel}}{\mu} \right | \ll 1 \qquad \left |\frac{{\bf q}_{\perp}}{\mu^{2}} \right | \ll 1 \qquad \left |\frac{\tau_{\parallel}}{\mu^{2}} \right | \ll 1 \end{equation} $ \bar{v}(l) $ flows to an infrared stable fixed point $ v_{*} $ which is directly obtained as a zero of $ \beta_{v} $ \begin{equation} \label{lfixpunkt} v_{*} = \frac{\textstyle 24}{\textstyle 11} \epsilon + {\cal O}(\epsilon^{2}) \; . \end{equation} At the fixed point the solutions of the flow equations are given by Equation (\ref{tfluss}) and by $ \bar{\tau}_{\parallel}(l) = \tau_{\parallel} l^{\kappa_{*}} $ with $ \kappa_{*} := \kappa(v_{*}) $. Inserting the fixed point value for $ v_{*} $ into $ \zeta_{\rho} $ we find the anomaly--exponent \begin{equation} \label{leta} \eta = \frac{\textstyle 2}{\textstyle 11} \epsilon + {\cal O}(\epsilon^{2}) \; . \end{equation} The fixed point values of the other parameter functions are \begin{equation} \label{lparfix} \zeta_{\sigma *} = - \frac{4}{11} \epsilon + {\cal O}(\epsilon^{2}) \qquad\qquad \kappa_{*} = \frac{4}{11} \epsilon + {\cal O}(\epsilon^{2}) \; . \end{equation} Returning to the dynamic model and combining the solutions of the renormalization group equation at the fixed point with a dimensional analysis and the invariant scale transformation we finally arrive at the universal scaling behaviour of the vertex functions at the longitudinal phase transition \begin{eqnarray} \label{lskalkrit} \Gamma_{\tilde{n},n} (\{q_{\parallel},{\bf q}_{\perp},\omega\},\tau_{\parallel},v_{*},\lambda,\rho,\sigma,\mu) & = & \nonumber\\ & & \hspace{-80 pt}l^{-\eta (-1 + \frac{3}{2} n -\frac{1}{2} \tilde{n}) + \frac{1}{2} \zeta_{\sigma *}(\tilde{n} - n) -\frac{1}{2} \tilde{n} (2d+5) - \frac{1}{2} n(2d-7) + 2d + 3}\nonumber\\ & & \hspace{-70 pt} \cdot \Gamma_{\tilde{n},n} \left(\left\{\frac{\textstyle q_{\parallel}}{\textstyle l^{1+\eta}}, \frac{\textstyle {\bf q}_{\perp}^{}}{\textstyle l^{2}},\frac{\textstyle \omega}{\textstyle l^{4}}\right\}, \frac{\textstyle \tau_{\parallel}}{\textstyle l^{2-\kappa_{*}}},v_{*},\lambda,\rho,\sigma,\mu\right) \; . \end{eqnarray} In addition to longitudinal lengths the critical parameter $ \tau_{\parallel} $ also exhibits anomalous scaling behaviour. In analogy to the models in Section 3 we derive the scaling form of the density response function \begin{equation} \label{lchi} \chi({\bf q},t) \; = \; f\left(q_{\parallel}^{2} t^{\frac{1}{2}(1 + \eta)},q_{\perp}^{2} t \right) \; . \end{equation} While the system is normally diffusive with respect to the transverse directions, in the critical longitudinal direction typical length squares scale for long times as \begin{equation} \label{llangskal} < r_{\parallel}^{2} > \; \sim \; t^{\frac{1}{2}(1 + \eta)} \; , \end{equation} i.e. in comparison to the naively critical $ t^{\frac{1}{2}} $ (corresponding to the naive dynamical critical exponent $ z = 4 $ from model B \cite{HH}) the spread of fluctuations in the driving force direction is enhanced, due to the positive $ \eta = \frac{2}{11} \epsilon + {\cal O}(\epsilon^{2}) $.\\ A comparison with the corresponding model for the longitudinal phase transition in a system without quenched disorder \cite{JS2} shows the surprising result that the longitudinal phase transition is here continuous, evidenced by the existence of an infrared stable fixed point, whereas in the model without quenched disorder there is no infrared stable fixed point. \section{Model Building} The configurations of a driven diffusive system with homogenous driving force are fully characterized by a single conserved scalar field, i.e., the local particle density $ n({\bf r},t) $. As the particle number is conserved the order parameter for both phase transitions is the deviation of the actual density $ n({\bf r},t) $ from its uniform average $ n_{0} $: \begin{equation} \label{ord} s({\bf r},t) = n({\bf r},t) - n_{0} \; . \end{equation} This fluctuating variable satisfies a continuity equation \begin{equation} \label{konti} \dot{s} \;\; + \;\; \nabla \cdot {\bf j} \;\; = \;\; 0 \end{equation} where the current density $ {\bf j} $ consists of a deterministic and a random part. The resulting stochastic differential equation is a Langevin equation. First we model the deterministic part. Due to the anisotropy of the system caused by the driving force, $ {\bf j} $ shows different symmetries longitudinal and transverse to its direction. Henceforth the indices $ \parallel $ and $ \perp $ denote the spatial direction longitudinal to the driving force and the $ (d-1) $--dimensional subspace transverse to it, respectively. The transverse current is according to \begin{equation} \label{chempotj} {\bf j}_{\perp} = - \lambda \nabla_{\perp} \mu_{\perp} \end{equation} caused by a chemical potential $ \mu_{\perp} $ with $ \lambda $ being a kinetic coefficient. As no direction is selected in the $ (d-1) $--dimensional subspace the transverse current $ {\bf j}_{\perp} $ is isotropic and therefore a vector of the form \begin{equation} \label{jperp} {\bf j}_{\perp} = \nabla_{\perp} f(s,\Delta_{\perp},\nabla_{\parallel}) \end{equation} where the Laplacean $ \Delta_{\perp} $ as argument denotes an even number of $ \nabla_{\perp} $--operators in every term of $ f $. An expansion of $ f $ with respect to $ s $ and gradient--operators yields, up to higher order terms \begin{equation} \label{jperp2} {\bf j}_{\perp} = \nabla_{\perp}(s + s^{2} + \Delta_{\perp} s + \nabla_{\parallel} s + \Delta_{\parallel} s ) \; , \end{equation} where here and in the following equations coefficients are suppressed for the sake of simplicity.\\ The longitudinal current, however, possesses no symmetry, due to the driving force. As a scalar it can be written as \begin{equation} \label{jparallel} j_{\parallel} = g(s,\nabla_{\parallel},\Delta_{\perp}) \end{equation} where the isotropy of the transverse subspace is again taken into account by the even number of $ \nabla_{\perp} $--operators. Its expansion up to higher order terms reads \begin{equation} \label{jparallel2} j_{\parallel} = c + s + s^{2} + \nabla_{\parallel} s + \nabla_{\parallel}^{2} s + \nabla_{\parallel}^{3} s + \nabla_{\parallel} s^{2} + \Delta_{\perp} s + \nabla_{\parallel} \Delta_{\perp} s \; . \end{equation} While some terms in this equation originate in a chemical potential $ \mu_{\parallel} $, according to \begin{equation} \label{chempot} j_{\parallel} = - \lambda \nabla_{\parallel} \mu_{\parallel} \; , \end{equation} others are due to the driving force $ {\bf E} $, which at least produces the terms proportional to $c$, $s$, $s^{2}$, and $\Delta_{\perp} s$ and which, in principle, also contributes to all terms of the longitudinal current in (\ref{jparallel2}). Here, the constant c is the homogenous part of the current, and $s^{2}$ is the first nonlinear term of the longitudinal current and so the leading nonlinearity of the problem.\\ Note that a repeated coarse graining of a microscopic model typically generates anisotropic transport coefficients. Building here a model directly on a mesoscopic length scale we must take that into consideration. Thus, although containing the same terms, $ \mu_{\parallel} $ and $ \mu_{\perp} $ in general have different coefficients, due to the anisotropy. From a technical point of view anisotropic transport coefficients are required to make the model renormalizable.\\ Following microscopic models and simulations \cite{KLS1,KLS2,Marro,Leung1,Leung2} we restrict ourselves to such models which additionally hold a CP (charge and parity)--symmetry, i.e. in which the Langevin equation is invariant under the transformation \begin{equation} \label{cp} s(r_{\parallel},{\bf r}_{\perp},t) \rightarrow - s(-r_{\parallel},{\bf r}_{\perp},t) \; . \end{equation} In microscopic lattice models which also form the basis for Monte Carlo simulations of driven diffusive systems the PC--symmetry corresponds to the particle--hole symmetry in the half--filled system.\\ In models with PC--symmetry all terms from (\ref{jperp2}) and (\ref{jparallel2}) vanish whose sum of $ s $-- and $ \nabla_{\parallel} $--factors in the Langevin equation is even.\\ The stochastic part of the Langevin equation $ \zeta = - \nabla \cdot {\bf j}_{L} $ reflects a random current $ {\bf j}_{L} $ that summarizes the fast microscopic degrees of freedom (local in time in our Markovian continuum description) and the effects of the quenched disorder of the medium. After writing down the general form of the Langevin equation we show how to model the various possibilities of quenched disorder and which parts of the noise are relevant in the renormalization group sense. Note that the noise considered here is conserving, i.e., it satisfies the continuity equation. We mention that we have also analyzed driven diffusive systems with nonconserving noise that is caused by random particle sources \cite{BJ2,BJ4}.\\ Taking anisotropic transport coefficients into account we obtain the Langevin equation \begin{equation} \label{langevin} \dot{s} =\lambda [\Delta_{\perp}(\tau_{\perp} - \kappa_{\perp} \Delta_{\perp}) + \rho \Delta_{\parallel} ( \tau_{\parallel} - \kappa_{\parallel}\Delta_{\parallel}) - \kappa\Delta_{\perp}\Delta_{\parallel}] s + \frac{1}{2}\lambda g \nabla_{\parallel} s^{2} - \nabla \cdot {\bf j}_{L} \; . \end{equation} This constitutes the fundamental equation for driven diffusive systems with homogenous driving force, attractive interactions, CP--symmetry and conserving noise. It contains all terms which are relevant in the renormalization group sense for the two phase transitions and the noncritical disordered phase, but in each case it still contains irrelevant terms which must be eliminated by a dimensional analysis.\\ We proceed by investigating the influence of the quenched disorder in detail. In a microscopic driven lattice gas model the quenched disorder is modelled by random potential barriers between the sites. There are three possible realizations of their randomness which is depicted for one dimensional systems in Fig. \ref{randompot}. \begin{enumerate} \item[I] The particles are in randomly deep potential valleys that are separated by randomly high potential mountains (Fig. \ref{randompot} (a)). \item[II] The potential valleys are randomly deep, the potential mountains are equally high (Fig. \ref{randompot} (b)). \item[III] The potential mountains are randomly high, but the potential valleys are equally deep (Fig. \ref{randompot} (c)). \end{enumerate} The homogeneous driving force tilts this landscape in the longitudinal direction by an angle that depends on its strength. Thus in this direction the symmetry of the random potential in the realizations II and III is broken, whereas the symmetries in the transverse subspace are not affected. Hence in the driving force direction the random potential is unsymmetric in all realizations.\\ In the microscopic lattice gas model particles can only jump to neighbouring unoccupied sites with jump rates that depend on the external driving force, on the energetic situation of the particles due to their attractive interaction, and on the locally random height of the potential barriers between the sites. Performing a continuum limit of the microscopic model one can show that, first, the main effect of the quenched disorder results in a time--independent random current with zero mean and Gaussian fluctuations and that, second, the three different realizations of the random potential produce three different types of such a random current $ \zeta_{d}({\bf r}) $.\\ In the unsymmetric case (realization I) the correlations of the random current are given by \begin{eqnarray} \label{stromerhalt} < {\bf \zeta}_{d}({\bf r})> & = & 0 \\ < {\bf \zeta}_{d}({\bf r}) \otimes {\bf \zeta}_{d}({{\bf r'}}) > & = & 2\lambda^{2} \delta ({\bf r} - {\bf r'}) [\sigma {\bf e}_{\parallel} \otimes {\bf e}_{\parallel} + \gamma ({\bf 1} - {\bf e}_{\parallel} \otimes {\bf e}_{\parallel})] \; .\nonumber \end{eqnarray} Thus the correlations of the noise force $ \nabla \cdot \zeta_{d} $ read \begin{eqnarray} \label{erhalt} < \nabla \cdot \zeta_{d}({\bf r})> & = & 0 \\ < \nabla \cdot \zeta_{d}({\bf r}) \nabla \cdot \zeta({{\bf r'}}) > & = & -2\lambda^{2} (\gamma \Delta_{\perp} + \sigma \Delta_{\parallel}) \delta ({\bf r} - {\bf r'}) \;\; .\nonumber \end{eqnarray} By a suitable scale change of $ s $ the kinetic coefficient $ \lambda $ in equation(\ref{erhalt}) is the same as in equation(\ref{langevin}). The anisotropy of the system due to the driving force is taken into account here by the transport coefficients $ \gamma $ and $ \sigma $.\\ In realization II a particle at a given site sees equally high potential barriers in all transverse directions. As a consequence of this symmetry the random current in the transverse subspace vanishes in the lowest order and starts with a local gradient. The correlations of the random current are here given by \begin{equation} \label{brauschen} < \nabla \cdot \zeta({\bf r}) \nabla \cdot \zeta({{\bf r'}}) > \; = \; -2\lambda^{2} (-\alpha \Delta_{\perp}^{2} + \sigma \Delta_{\parallel}) \delta ({\bf r} - {\bf r'}) \; . \end{equation} As the symmetry of the random potential is broken by the driving force in the longitudinal direction the longitudinal random current behaves as in the unsymmetric case.\\ In realization III a potential barrier looks the same from both sides in the transverse subspace. This is why the transverse random current vanishes totally. Being only longitudinal the random current has the correlations \begin{equation} \label{trauschen} < \nabla \cdot \zeta({\bf r}) \nabla \cdot \zeta({{\bf r'}}) > \; = \; -2\lambda^{2} \sigma \Delta_{\parallel} \delta ({\bf r} - {\bf r'}) \; . \end{equation} Since the continuum limit of the microscopic model is not rigorous, it might be possible that even in the symmetric realizations II and III transverse noise terms with $ \alpha \neq 0 $ and $ \gamma \neq 0 $, respectively, are generated by the coarse graining procedure. This is due to the fact that the symmetry of the random potential is microscopic and concerns only a single site or bond. As we will see later each type of noise nevertheless leads to different critical behaviour governed by different stable fixed points with finite regions of attraction. We defer the discussion of the possible additional transverse noise terms until the end of the Sections 3.2 and 3.3, respectively.\\ By dimensional considerations (power counting) one can prove that each of the equations (\ref{erhalt}) -- (\ref{trauschen}) contains all relevant noise terms in the renormalization group sense for the respective type of the random potential. Also deviations from the local Gaussian nature of the random current $ \zeta $ are irrelevant for the critical and long--time properties. Even the microscopic part $ \zeta_{m} $ of the random current summarizing the fast microscopic degrees of freedom is irrelevant if $ \zeta_{d} \neq 0 $. It should be remarked that $ \zeta_{m} $ plays the role of a dangerous irrelevant field for the calculation of correlation functions with frequencies $ \omega \neq 0 $. Without $ \zeta_{m} $ correlation functions show only a "central peak" at $ \omega = 0 $. \section{Summary and Outlook} We have analyzed the transverse and longitudinal phase transition in uniformly driven diffusive systems with quenched disorder. These systems show a wide variety of possible scenarios, because the symmetry properties of the random potential are an additional distinguishing feature to which universality class a model belongs. In the region of the transverse phase transition the three different random potentials I--III (Fig. \ref{randompot}) actually define three different models and in the noncritical region they still define two different models all of which are part of different universality classes.\\ Together with earlier investigations the renormalization group studies of an entire model class are hereby completed. This model class includes the driven diffusive systems with and without quenched disorder both in the critical regions of the transverse and longitudinal phase transition and in the noncritical region. Fig. \ref{ahnen1} gives a graphical overview of the model class, where the single models are ordered chronologically from left to the right. The models in an ordered substrate \cite{JS1,JS2,LC}, i.e. without quenched disorder, had been studied prior to this work for all three regions of the phase diagram mentioned above. The models for a system with quenched disorder in the noncritical region \cite{BJ1} had also been investigated. The other models of Fig. \ref{ahnen1} have been studied in the present paper.\\ The upper critical dimension $ d_{c} $ is different from model to model and varies from 2 to 9. Below $ d_{c} $, the vertex functions being typical statistical quantities of such systems show universal anomalous scaling behaviour on large length and time scales. The deviation from pure diffusive behaviour is characterized by the anomaly--exponent $ \eta $. It indicates how strongly longitudinal lengths scale anomalously. In Fig. \ref{ahnen1} the result for $ \eta $ in the highest calculated loop order is given for every model. Note that, due to Galilean invariance, $ \eta $ is even exact in all models without quenched disorder.\\ We emphasize the following results:\\ In all models of this model class (except for the longitudinal phase transition in a system without quenched disorder) the anomaly--exponent $ \eta $ is positive implying superdiffusive spreading of density fluctuations in the driving force direction.\\ A model with quenched disorder always has a higher upper critical dimension than the corresponding model without quenched disorder. Due to the field theoretic results, in the dimension interval between these two upper critical dimensions the quenched disorder is the reason for superdiffusive spread of density fluctuations in the high temperature region and at the transverse phase transition, respectively. Notice that the anomalous diffusion of fluctuations is not directly connected with the behaviour of transport coefficients relating the mean current to the mean density, because a fluctuation--dissipation theorem (Einstein relation) does not hold in this strong nonequilibrium situation with quenched disorder. Within this model class we have not found a qualitative argument why disorder generates superdiffusion. We mention, however, a study of a one dimensional driven lattice gas \cite{KFer,E} where quenched disorder is not, as here, spatially fixed, but associated with moving particles. The authors found a superdiffusive spread of a jam behind the slowest particle and a superdiffusive spread of free spacing in front of it. Whether this situation can be transferred to our models by identifying the deepest potential valley and the highest potential mountain, respectively, with the slowest particle, should be a topic of further investigations, especially Monte Carlo simulations.\\ For the longitudinal phase transition there is no uniform statement. In the model without quenched disorder \cite{JS2} no infrared stable fixed point has been found and some analytic arguments point to a discontinuous phase transition, whereas an according two dimensional lattice gas demonstrates a continuous phase transition in Monte Carlo simulations \cite{BassZ}. In the model with quenched disorder studied here, however, we find an infrared stable fixed point and thus a continuous longitudinal phase transition.\\ Despite the partially high upper critical dimensions it appears that the anomaly--exponent (when existing) may be extrapolated quite accurately into low dimensions. Two observations lend support to this procedure: first, the two--loop correction to $ \eta $ in the critical transverse model from Section 3.1 is small and second, the coefficients of $ \eta $ are small in all models.\\ For the two--loop calculation in critical models with quenched disorder a new technique has been developed enabling us to manipulate mixed $ {\bf q}^{4} $--propagators and --correlators. This technique is based on an inverse Mellin transformation and is described for one model at the transverse phase transition in the Appendix. This method is applicable to all critical models with quenched disorder of this model class, but has only been performed for two models (Sec. 3.1 and 3.3). Furthermore we expect it to be useful for two--loop calculations in other physical problems where $ {\bf q}^{4} $--propagators are involved.\\ Monte Carlo simulations of two--dimensional driven diffusive lattice gases without quenched disorder \cite{Leung1,Leung2,KLS1,KLS2} are in excellent agreement with field theoretic predictions for the transverse phase transition and the noncritical region \cite{JS1,JS2,LC}. For the transverse phase transition in systems with quenched disorder, however, there is a simulation study \cite{LF} that is, for two reasons, hardly compatible with the models investigated here by field theory. First, the quenched disorder was there modelled by randomly blocked sites and not by random potential barriers between the sites. Second, the concentration of blocked sites is so small that only the crossover behaviour between a system with and without quenched disorder was observed. Further, we mention a recent simulation for a noncritical system with quenched disorder \cite{BarT}.\\ It is desirable to compare the field theoretic results obtained here for the various critical models with corresponding Monte Carlo simulations that are still to be done. These simulations are a nontrivial challenge, because besides the average over a huge number of realizations of the quenched disorder it is to pay attention to the fact that the periodic boundary conditions usually used let pass a particle repeatedly through the system and let it see the same quenched disorder as before. By this unwanted correlations of the randomness enter into the simulational results that make it difficult to compare them with the field theoretic results that are based on the assumption of uncorrelated disorder. Therefore Monte Carlo simulations of driven diffusive systems with open boundaries already done for systems without quenched disorder \cite{JOe2} seem to be more appropriate.\\ We finally remark that we have extended the model class investigated here in the way that we allow for random particle sources and drains in the diffusive systems. The noncritical model without quenched disorder with such a particle nonconserving randomness had already been studied \cite{HK,BJ2}. Moreover we have analyzed the influence of nonconserving noise onto the critical behaviour in systems with and without quenched disorder which is demonstrated in a paper soon to be published \cite{BJ4}. \section{Transverse Phase Transition} In the following all models are analyzed with the help of renormalized field theory. This method has successfully been applied to all driven diffusive models investigated so far.\\ The three different realizations of the random potential described in the last section actually lead to three different models for the transverse phase transition and have to be treated separately. \subsection{Unsymmetric Random Potential} The model for a driven diffusive system with frozen random unsymmetric potential is based on the equations (\ref{langevin}) and (\ref{erhalt}).\\ To set up a renormalized field theory, it is convenient to recast the model in terms of a dynamic functional \cite{J0,DeD,J1,J2,BJW,DeDP} \begin{eqnarray} \label{udynfunktional} {\cal J}[s,\tilde{s}] & \hspace{-9 pt} = &\hspace{-11 pt} \int\!\! d^{d}r \left\{\int\!\! dt [\tilde{s}\left (\dot{s} + \lambda (\Delta_{\perp}(\kappa_{\perp}\Delta_{\perp} - \tau_{\perp}) + \rho \Delta_{\parallel}(\kappa_{\parallel}\Delta_{\parallel} - \tau_{\parallel}) +\kappa \Delta_{\perp} \Delta_{\parallel}) s\right )\right.\nonumber\\ & & \hspace{30 pt}\left. + \frac{1}{2} \lambda g (\nabla_{\parallel} \tilde{s}) s^{2} ] - \gamma \left[\lambda \int\!\! dt \nabla_{\perp} \tilde{s}\right]^{2} -\sigma \left[\lambda \int\!\! dt \nabla_{\parallel} \tilde{s}\right]^{2} \right\} \; , \end{eqnarray} where $ \tilde{s}({\bf r},t) $ is a Martin--Siggia--Rose \cite{MSR} response field. Correlation and response functions can now be expressed as functional averages with weight $ \exp(-{\cal J}) $.\\ But for the description of the transverse phase transition this dynamic functional still contains irrelevant terms in the renormalization group sense. Now these terms are determined by a dimensional analysis.\\ The transverse phase transition is characterized by finite $ \tau_{\parallel} $ and $ \tau_{\perp} \rightarrow 0 $. We introduce a scale $ \mu^{2} $ for small $ \tau_{\perp} $. Then $ \mu^{-1} $ is a convenient length scale. Since $ \tau_{\perp} $ tends to 0 at the transverse phase transition, the leading term in the transverse direction is proportional to $ \tilde{s} \lambda \Delta_{\perp} \kappa_{\perp} \Delta_{\perp} s $, whereas in the longitudinal direction the leading gradient term is proportional to $ \tilde{s} \lambda \rho \Delta_{\parallel} \tau_{\parallel} s $. The comparison of the leading gradient terms demonstrates that $ \Delta_{\parallel} $ scales as $ \Delta_{\perp}^{2} $. For longitudinal and transverse length scales, this implies \begin{equation} \label{skalierung} r_{\perp} \sim \mu^{-1} \qquad r_{\parallel} \sim \mu^{-2} \; . \end{equation} Since the dynamic functional is dimensionless the dimensions of fields and coupling constants are \begin{eqnarray} \label{udimension} \lambda t \sim \mu^{-4}\qquad &\qquad s \sim \mu^{\frac{d-5}{2}}\qquad & \qquad\tilde{s} \sim \mu^{\frac{d+7}{2}} \nonumber\\ \kappa_{\parallel} \sim \mu^{-4}\qquad &\qquad \kappa \sim \mu^{-2}\qquad &\qquad \sigma \sim \mu^{-2} \nonumber\\ \gamma \sim \mu^{0}\qquad & \qquad g \sim \mu^{\frac{9-d}{2}} \; ,\qquad & \qquad \end{eqnarray} i.e. the coupling constants $ \kappa_{\parallel}$, $ \kappa $, and $ \sigma $ are irrelevant in the renormalization group sense. The irrelevancy of $ \sigma $ indicates that for the transverse phase transition in a unsymmetric random potential the noise in the driving force direction is irrelevant. From the dimension of the nonlinear coupling constant $ g $ we recognize \begin{equation} \label{ugrenzdim} d_{c} = 9 \end{equation} as upper critical dimension of this model, above which $ g $ is irrelevant und below which $ g $ is relevant. The finite $ \tau_{\parallel} $ and the transverse coupling constants are absorbed into lengths and fields by a suitable scale change. Thus, the appropriate dynamic functional to describe the transverse phase transition in a driven diffusive system with an unsymmetric random potential is given by \begin{eqnarray} \label{uJ} {\cal J}[s,\tilde{s}] & = & \int\!\! d^{d}r \left\{\int\!\! dt \left[\tilde{s} \dot{s} + \lambda \tilde{s} (\Delta_{\perp}(\Delta_{\perp} - \tau_{\perp}) - \rho \Delta_{\parallel}) s + \frac{\textstyle 1}{\textstyle 2} \lambda g (\nabla_{\parallel} \tilde{s}) s^{2} \right]\right. \nonumber\\ & & \hspace{190 pt}\left. - \left[\lambda \int\!\! dt \nabla_{\perp} \tilde{s}\right]^{2} \right\} \; . \end{eqnarray} The dynamic functional has the following symmetries. The isotropy in the transverse subspace and the CP--symmetry that reads here \begin{equation} \label{dyncp} r_{\parallel} \rightarrow -r_{\parallel} \qquad s \rightarrow -s \qquad \tilde{s} \rightarrow - \tilde{s} \end{equation} were directly integrated in the model building. After averaging over the quenched disorder, $ \cal J $ exhibits translational symmetry in space and time. The invariance of $ \cal J $ under the longitudinal scale transformation \begin{eqnarray} \label{uskalen} r_{\parallel} \rightarrow \beta r_{\parallel} & \qquad & r_{\perp} \rightarrow r_{\perp}\\ \tilde{s} \rightarrow \beta^{-\frac{1}{2}} \tilde{s} & \qquad & s \rightarrow \beta^{-\frac{1}{2}} s \nonumber\\ \rho \rightarrow \beta^{2} \rho & \qquad & g \rightarrow \beta^{\frac{3}{2}} g \nonumber \end{eqnarray} is of great importance, because the parameter combination $ g^{2}\rho^{-\frac{3}{2}} $ is found as appropriate variable of the model being invariant under this transformation. In contrast to the corresponding model without quenched disorder \cite{JS2} the dynamic functional is here not invariant under a Galilei transformation.\\ To study the critical properties of the transverse phase transition we apply standard renormalization group methods \cite{Amit,BJW,DeDP}. We use dimensional regularization in $ d = 9 - \epsilon $ followed by minimal subtraction. The one--line--irreducible vertex functions with $\tilde{n} \; \tilde{s} $--legs and $ n \; s $--legs at wavevectors $ \{{\bf q}\} $ and frequencies $ \{\omega\} $ will be denoted by $ \Gamma_{\tilde{n},n}(\{{\bf q}\},\{\omega\}) $. Taking into consideration the causality properties of the theory we only find $ \Gamma_{1,1} $ and $ \Gamma_{1,2} $ primitively divergent. The nontrivial diagrams contributing to $ \Gamma_{\tilde{n},n}(\{{\bf q}\},\{\omega\}) $ carry at least a factor $ q_{\parallel}^{\tilde{n}} $, because the interaction vertex is conserving. The primitive divergences are multiplicatively absorbed in a redefinition of the parameters \begin{equation} \label{ureparameter} \rho \rightarrow \mathaccent28695{\rho} = Z_{\rho} \rho \; , \; g \rightarrow \mathaccent28695{g} = \mu^{\frac{\epsilon}{2}} Z_{u} u \; . \end{equation} Here, in contrast to the ordered problem \cite{JS2}, the coupling constant $ g $ has to be renormalized, due to the loss of Galilean invariance in the disordered problem.\\ The elements of our perturbation expansion are the Gaussian propagator \begin{equation} \label{uprop} G_{{\bf q},\omega} \; := \; < \tilde{s}_{{\bf q},\omega} s_{-{\bf r},-\omega}>_{0} \; = \; \frac{1}{i\omega + \lambda[{\bf q}_{\perp}^{2}({\bf q}_{\perp}^{2} + \tau_{\perp}) + \rho q_{\parallel}^{2}]} \; , \end{equation} the Gaussian correlator \begin{equation} \label{ukorr} C_{{\bf q},\omega} \; := \; < s_{{\bf q},\omega} s_{-{\bf q},-\omega}>_{0} \; = \; \frac{2 \lambda^{2}{\bf q}_{\perp}^{2} \delta(\omega)}{\omega^{2} + \lambda^{2}[{\bf q}_{\perp}^{2}({\bf q}_{\perp}^{2} + \tau_{\perp}) + \rho q_{\parallel}^{2}]^{2}} \; , \end{equation} and the conserving vertex $ v({\bf q}) = -i\lambda g q_{\parallel} $. Their graphical representation is shown in the Appendix in fig.(\ref{graphelem}).\\ As the correlator of the theory $ C_{{\bf q},\omega} $ is proportional to $ \delta(\omega) $ (i.e. $ C_{{\bf q}}(t) $ is independent of time) and any loop of a diagram contains at least one correlator because of the structure of the vertex with two $ s $--legs and one $ \tilde{s} $--leg and because of causality the integration over internal frequencies cannot generate any divergences. Thus the divergent parts of the vertex functions are only in the ($ \omega = 0 $)--parts. To facilitate their calculation it is convenient to split the order parameter \begin{equation} \label{aufspaltung} s({\bf r},t) = \varphi({\bf r}) + s^{\prime}({\bf r},t) \end{equation} into a time independent $ \varphi({\bf r}) $ and a time dependent $ s^{\prime}({\bf r},t) $ that contains no ($ \omega = 0 $)--parts. Since the noise is independent of time $ s^{\prime}({\bf r},t) $ relaxes deterministically. In the limit of long times we obtain \begin{equation} \label{tlimes} s({\bf r},t \rightarrow \infty) = \varphi({\bf r}) \; . \end{equation} Together with \begin{equation} \label{quasitrans} \tilde{\varphi}({\bf r}) = \lambda \int dt \tilde{s}({\bf r},t) \end{equation} the dynamic functional $ \cal{J} $ reduces to a quasi--static (frozen) Hamiltonian \begin{equation} \label{quasihamil} {\cal H}[\varphi,\tilde{\varphi}] = \int\!\! d^{d}r \left\{\tilde{\varphi} (\Delta_{\perp}(\Delta_{\perp} - \tau_{\perp}) - \rho \Delta_{\parallel}) \varphi + \frac{1}{2} g (\nabla_{\parallel} \tilde{\varphi}) \varphi^{2} + \tilde{\varphi} \Delta_{\perp} \tilde{\varphi} \right\} \; . \end{equation} The frozen Hamiltonian generates all the zero--frequency parts of the vertex functions if causality is included in the graphical rules of perturbation theory.\\ In the quasi--static model the Gaussian propagator and correlator are only dependent on wavevectors and read \begin{eqnarray} \label{quasipropkorr} G_{{\bf q}} & = & \frac{1}{{\bf q}_{\perp}^{2}({\bf q}_{\perp}^{2} + \tau_{\perp}) + \rho q_{\parallel}^{2}} \\ C_{{\bf q}} & = &\frac{2 {\bf q}_{\perp}^{2} }{[{\bf q}_{\perp}^{2}({\bf q}_{\perp}^{2} + \tau_{\perp}) + \rho q_{\parallel}^{2}]^{2}} \; .\nonumber \end{eqnarray} In the quasi--static model we calculate the primitively divergent vertex functions in two--loop order. While the one--loop calculation is easy to perform analytically, technical difficulties have to be overcome for the two--loop diagrams with mixed $ {\bf q}^{4} $--propagators and --correlators. These two--loop diagrams have here been solved by a new technique where a special inverse Mellin transformation is used to factorize the denominators of some propagators and correlators, respectively. Thus, the $\epsilon^{2} $--poles and the simple $ \epsilon $--poles can be extracted. Their coefficients are then given by parameter integrations over paths in the complex plane. Whereas the coefficients of the $\epsilon^{2} $--poles can be analytically calculated, those of the simple $ \epsilon $--poles have to be computed numerically. The details of the two--loop calculation are shown in the Appendix.\\ In a two--loop calculation using dimensional regularization we obtain the vertex functions in an expansion in $ \epsilon = d_{c} - d $ and $ {\bf q} $ \begin{eqnarray} \label{uvertex} \mathaccent28695{\Gamma}_{1,1}({\bf q})\!\! &\!\! =\!\! & \!\!{\bf q}_{\perp}^{2}({\bf q}_{\perp}^{2} + \tau_{\perp}) + \mathaccent28695{\rho} q_{\parallel}^{2} + \frac{2}{3} \frac{A_{\epsilon}}{\epsilon} \frac{\mathaccent28695{g}^{2}}{\mathaccent28695{\rho}^{\frac{1}{2}}} q_{\parallel}^{2} \tau_{\perp}^{-\frac{\epsilon}{2}} - \frac{2}{9} \frac{A_{\epsilon}^{2}}{\epsilon^{2}} \frac{\mathaccent28695{g}^{4}}{\mathaccent28695{\rho}^{2}} q_{\parallel}^{2} \tau_{\perp}^{-\epsilon}(1 - 0.2158 \epsilon) \nonumber\\ \mathaccent28695{\Gamma}_{1,2}({\bf q})\!\! &\!\! =\!\! &\!\! i \mathaccent28695{g} q_{\parallel} - i \frac{1}{6} \frac{A_{\epsilon}}{\epsilon} \frac{\mathaccent28695{g}^{3}}{\mathaccent28695{\rho}^{\frac{3}{2}}} q_{\parallel} \tau_{\perp}^{-\frac{\epsilon}{2}} + \frac{1}{8} i \frac{A_{\epsilon}^{2}}{\epsilon^{2}} \frac{\mathaccent28695{g}^{5}}{\mathaccent28695{\rho}^{3}} q_{\parallel} \tau_{\perp}^{-\epsilon}(1 - 0.2154 \epsilon) \; , \end{eqnarray} where bare unrenormalized parameters are indicated by a superscript ``$\; \mathaccent28695{} $ " above the symbol. $ A_{\epsilon} = \frac{1}{(2\pi)^{d}} {\cal O}_{d-1} \Gamma(1 + \frac{\epsilon}{2}) \Gamma(\frac{5 - \epsilon}{2}) \Gamma(\frac{1}{2}) $ is a suitably chosen constant with $ {\cal O}_{d-1} $ being the surface of the $ (d-1) $--dimensional unit sphere and $ \Gamma(z) $ denoting Euler's $ \Gamma $--function. The $ Z $--factors defined in equation(\ref{ureparameter}) are in minimal subtraction \begin{eqnarray} \label{uZ} Z_{\rho} & = & 1 - \frac{2}{3 \epsilon} A_{\epsilon} \frac{u^{2}}{\rho^{\frac{3}{2}}} - \frac{2}{9 \epsilon^{2}}A_{\epsilon}^{2}\frac{u^{4}}{\rho^{3}}(1 + 0.2158 \epsilon) + {\cal O}(u^{6}) \\ Z_{u} & = & 1 + \frac{1}{6 \epsilon} A_{\epsilon} \frac{u^{2}}{\rho^{\frac{3}{2}}} + \frac{1}{8 \epsilon^{2}}A_{\epsilon}^{2}\frac{u^{4}}{\rho^{3}}(1 + 0.2154 \epsilon) + {\cal O}(u^{6}) \; .\nonumber \end{eqnarray} We recognize that the perturbation expansion is organized in powers of the dimensionless renormalized parameter combination $ v := A_{\epsilon} u^{2}\rho^{-\frac{3}{2}} $ that was already found as invariant variable under the longitudinal scale transformation (equation(\ref{uskalen})).\\ With the renormalizations at hand we are in a position to determine the critical behaviour of the vertex functions. We use the fact that the unrenormalized theory is independent of the momentum scale $ \mu $. This leads to the renormalization group equation \begin{equation} \label{rgg} \left [\beta_{v} \frac{\partial}{\partial v} + \rho \zeta \frac{\partial}{\partial \rho} + \mu \frac{\partial}{\partial \mu}\right ] \Gamma_{\tilde{n},n} (\{q_{\parallel},{\bf q}_{\perp}\},\tau_{\perp},v,\rho,\mu) = 0 \; . \end{equation} The Wilson parameter functions, being only dependent on $ v $, are given by \begin{eqnarray} \label{uwilson} \beta_{v} & := & \left. \mu \frac{\partial v}{\partial \mu}\right |_{o} = -v (\epsilon - \frac{4}{3} v - 0.252 v^{2} + {\cal O}(v^{3}))\\ \zeta & := & \left. \mu \frac{\partial \ln \rho}{\partial \mu}\right |_{o} = - \frac{2}{3} v - 0.0959 v^{2} + {\cal O}(v^{3}) \; ,\nonumber \end{eqnarray} where the derivatives are calculated at fixed bare parameters. As a linear partial differential equation the renormalization group equation is solvable by the method of characteristics with the result \begin{equation} \label{urggresult} \Gamma_{\tilde{n},n} (\{q_{\parallel},{\bf q}_{\perp}\},\tau_{\perp},v,\rho,\mu) = \Gamma_{\tilde{n},n} (\{q_{\parallel},{\bf q}_{\perp}\},\tau_{\perp},\bar{v}(l),\bar{\rho}(l),\bar{\mu}(l)) \; . \end{equation} The trajectories $ \bar{v}(l) $, $ \bar{\rho}(l) $, and $ \bar{\mu}(l) $ are solutions of the flow equations \begin{equation} \label{ufluss} l \frac{d}{dl} \bar{v}(l) = \beta_{v}(\bar{v}(l)) \qquad l \frac{d}{dl} \ln \bar{\rho}(l) = \zeta(\bar{v}(l)) \qquad l \frac{d}{dl} \bar{\mu}(l) = \bar{\mu}(l) \end{equation} with the flow parameter $ l $ and the initial conditions \begin{equation} \label{uanfang} \bar{v}(l=1) = v \qquad \bar{\rho}(l=1) = \rho \qquad \bar{\mu}(l=1) = \mu \; . \end{equation} The third flow equation obviously has the solution $ \bar{\mu}(l) = \mu l $.\\ With the help of characteristics the critical asymptotic region of small $ \tau_{\perp} $ and $ q $ can be mapped onto uncritical regions. In the scaling limit $ l \ll 1 $ corresponding to \begin{equation} \label{skalenlimes} \left |\frac{q_{\parallel}}{\mu^{2}} \right | \ll 1 \qquad \left |\frac{{\bf q}_{\perp}}{\mu} \right | \ll 1 \qquad \left |\frac{\tau_{\perp}}{\mu^{2}} \right | \ll 1 \end{equation} the flow of $ \bar{v}(l) $ is controlled by stable zeroes of $ \beta_{v} $. For $ \epsilon > 0 $, there is a nontrivial infrared stable fixed point \begin{equation} \label{ufixpunkt} v_{*} \; = \; \frac{\textstyle 3}{\textstyle 4} (\epsilon - 0.141 \epsilon^{2} + {\cal O}(\epsilon^{3})) \; . \end{equation} The other fixed point $ v_{*} = 0 $ being Gaussian is stable only if $ \epsilon < 0 $.\\ From the nontrivial fixed point value $ \zeta(v_{*}) $ we define the anomaly exponent $ \eta := -\frac{1}{2} \zeta(v_{*}) $ that is to two--loop order \begin{equation} \label{ueta} \eta = \frac{\textstyle 1}{\textstyle 4} \epsilon (1 - 0.302 \; \frac{\textstyle \epsilon}{\textstyle 9} + {\cal O}(\epsilon^{2})) \; . \end{equation} At the fixed point the solution of the second flow equation of equation (\ref{ufluss}) is \begin{equation} \label{rhofluss} \bar{\rho}(l) = \rho l^{-2 \eta} \; . \end{equation} The results found for the vertex functions in the quasi--static model can be directly transferred into the dynamic model because the time scale $ \lambda $ needs no renormalization.\\ The longitudinal length scale transformation according to equation (\ref{uskalen}), dimensional analysis and the renormalization group equation are now combined to derive the asymptotic critical scaling form of the vertex functions in the dynamic model at the transverse phase transition \begin{eqnarray} \label{uskalkrit} \Gamma_{\tilde{n},n} (\{q_{\parallel},{\bf q}_{\perp},\omega\},\tau_{\perp},v_{*},\lambda,\rho,\mu) & = & \nonumber\\ & & \hspace{-170 pt}l^{-\frac{1}{2} \eta (\tilde{n} + n -2) -\frac{1}{2} \tilde{n} (d+7) - \frac{1}{2} n(d-5) + d + 5} \Gamma_{\tilde{n},n} \left(\left\{\frac{\textstyle q_{\parallel}}{\textstyle l^{2+\eta}},\frac{\textstyle {\bf q}_{\perp}}{\textstyle l},\frac{\textstyle \omega}{\textstyle l^{4}}\right\}, \frac{\textstyle \tau_{\perp}}{\textstyle l^{2}},v_{*},\lambda,\rho,\mu\right) \; . \end{eqnarray} This equation implies that, below the upper critical dimension $ d_{c} = 9 $, anomalous scaling behaviour only occurs in the direction of the driving force and is completely characterized by the anomaly exponent $ \eta $ from (\ref{ueta}) that is positive. These results are in analogy to what was found in the driven diffusive systems investigated so far \cite{JS1,JS2,BJ1,BJ2}.\\ To illustrate the importance of the positive anomaly exponent $ \eta $ we especially investigate the density response function $ \chi({\bf q},t) $ which is the Fourier transform of $ \Gamma_{1,1}^{-1}({\bf q},\omega) $. Choosing the flow parameter $ l = \omega^{\frac{1}{4}} \ll 1 $, we derive the scaling form of the density response function $ \chi({\bf q},t):= <s({\bf q},t) \tilde{s}(-{\bf q},0)> $ from equation (\ref{uskalkrit}): \begin{equation} \label{usus} \chi({\bf q},t) \; = \; f\left(q_{\parallel}^{2} t^{1+\frac{1}{2} \eta},q_{\perp}^{2} t^{\frac{1}{2}}\right) \; . \end{equation} From that we conclude that for long times typical longitudinal length--squares scale with time as \begin{equation} \label{ulangskal} < r_{\parallel}^{2} > \; \sim \; t^{1+\frac{1}{2} \eta} \; . \end{equation} The positivity of $ \eta = \frac{\textstyle 1}{\textstyle 4} \epsilon (1 - 0.302 \; \frac{\textstyle \epsilon}{\textstyle 9} + {\cal O}(\epsilon^{2})) $ means that in systems with spatial dimensions $ d < d_{c} = 9 $ fluctuations spread faster than diffusively in the driving force direction. This superdiffusion was also observed in all driven diffusive systems analyzed up to now \cite{JS1,JS2,BJ1,BJ2}.\\ A comparison with the corresponding model for the transverse phase transition without quenched disorder, having an upper critical dimension $ d_{c} = 5 $, shows that the phenomenon of superdiffusion occurs here over a much greater dimensional range. As the system without quenched disorder behaves normally diffusive for $ d > 5 $ it follows that quenched disorder is the reason for superdiffusion in the dimensional interval $ 5 < d < 9 $.\\ The analogous comparison of the models with and without disorder for the noncritical region \cite{JS1,BJ1} came to an analogous result.\\ Another consequence from the scaling form of the density response function is that typical transverse length--squares scale as $ < r_{\perp}^{2} > \; \sim \; t^{\frac{1}{2}} $ for long times, i.e. subdiffusively. This behaviour corresponds to the naive dynamical exponent $ z = 4 $ in model B (in the nomenclature of Halperin and Hohenberg \cite{HH}) and has been expected, for the system here is critical with respect to the transverse directions and no renormalizations are necessary in this subspace.\\ The extrapolation of the anomaly--exponent $ \eta $ into low dimensions is difficult in view of the high upper critical dimension $ d_{c} = 9 $. Naively, we can simply set $ \epsilon = d_{c} -d = 6 $ in Equation (\ref{ueta}) to predict $ \eta $ f.e. in a 3--dimensional system, resulting in an estimate \begin{equation} \label{ueta32} \eta(d=3) \; = \; 1.20 \; . \end{equation} Although $ \epsilon $ has been considered as small quantity the two loop--correction with respect to the one--loop result is merely 20\%, even at $ \epsilon = 6 $. Due to these small corrections we expect that equation (\ref{ueta}), even though it is strictly valid only near $ d_{c} = 9 $, produces good approximations for $ \eta $ also in low dimensions. \subsection{Random Potential With Equally High Potential Mountains} This model is analyzed in analogy to the model with unsymmetric random potential of the last section. The quenched disorder according to Equation (\ref{brauschen}) resulting from the random potential with equally high potential mountains leads, in analogy to Equation (\ref{udynfunktional}), to the dynamic functional \begin{eqnarray} \label{bdynfunkt} {\cal J}[s,\tilde{s}] & \hspace{-9 pt} = &\hspace{-11 pt} \int\!\! d^{d}r \left\{\int\!\! dt [\tilde{s}\left (\dot{s} + \lambda (\Delta_{\perp}(\kappa_{\perp}\Delta_{\perp} - \tau_{\perp}) + \rho \Delta_{\parallel}(\kappa_{\parallel}\Delta_{\parallel} - \tau_{\parallel}) +\kappa \Delta_{\perp} \Delta_{\parallel}) s\right )\right.\nonumber\\ & & \hspace{30 pt}\left. + \frac{1}{2} \lambda g (\nabla_{\parallel} \tilde{s}) s^{2} ] - \alpha \left[\lambda \int\!\! dt \Delta_{\perp} \tilde{s}\right]^{2} -\sigma \left[\lambda \int\!\! dt \nabla_{\parallel} \tilde{s}\right]^{2} \right\} \; , \end{eqnarray} that still contains irrelevant terms to be eliminated.\\ Transverse and longitudinal lengths scale as before $ r_{\perp} \sim \mu^{-1},\; r_{\parallel} \sim \mu^{-2} $ . Since the dynamic functional is dimensionless the dimension of fields and coupling constants are \begin{eqnarray} \label{bdimension} \lambda t \sim \mu^{-4}\qquad &\qquad s \sim \mu^{\frac{d-3}{2}}\qquad & \qquad\tilde{s} \sim \mu^{\frac{d+5}{2}} \nonumber\\ \kappa_{\parallel} \sim \mu^{-4}\qquad &\qquad \kappa \sim \mu^{-2}\qquad & \qquad\sigma \sim \mu^{0} \nonumber\\ \alpha \sim \mu^{0}\qquad &\qquad g \sim \mu^{\frac{7-d}{2}}\qquad & \qquad\qquad\; . \end{eqnarray} Thus, the coupling constants $ \kappa_{\parallel} $ and $ \kappa $ are again irrelevant in the renormalization group sense, but $ \sigma $ is not. This signifies that in the given random potential both longitudinal and transverse noise are relevant. The dimension of the nonlinear coupling $ g $ shows that \begin{equation} \label{bgrenzdim} d_{c} = 7 \end{equation} is the upper critical dimension of this model. The finite $ \tau_{\parallel} $ and the transverse coupling constants $ \alpha $ and $ \kappa_{\perp} $ are again absorbed by a suitable scale change. Thus, the appropriate dynamic functional for the transverse phase transition in the given random potential is \begin{eqnarray} \label{bJ} {\cal J}[s,\tilde{s}] & = & \int\!\! d^{d}r \left\{\int\!\! dt \left[\tilde{s} \dot{s} + \lambda \tilde{s} (\Delta_{\perp}(\Delta_{\perp} - \tau_{\perp}) - \rho \Delta_{\parallel}) s + \frac{\textstyle 1}{\textstyle 2} \lambda g (\nabla_{\parallel} \tilde{s}) s^{2} \right]\right. \nonumber\\ & & \hspace{100 pt}\left. - \left[\lambda \int\!\! dt \Delta_{\perp} \tilde{s}\right]^{2} - \sigma \left[\lambda \int\!\! dt \nabla_{\parallel} \tilde{s} \right]^{2} \right\} \; . \end{eqnarray} This dynamic functional exhibits the same symmetries as the one in the case of the unsymmetric random potential, except the longitudinal scale transformation, where here the parameter $ \sigma $ is additionally transformed according to $ \sigma \rightarrow \beta^{2} \sigma $. In addition to the parameter combination $ g^{2} \rho^{-\frac{3}{2}} $, $ \sigma \rho^{-1} $ is also an invariant variable of this model and the perturbation expansion will be organized in powers of both variables.\\ Due to the quenched disorder, the Gaussian correlator is again time independent so that, for simplicity, we also here transform $ \cal{J} $ into a quasi--static Hamiltonian via (\ref{tlimes}) and (\ref{quasitrans}) \begin{eqnarray} \label{bquasihamil} {\cal H}[\varphi,\tilde{\varphi}] & = & \int\!\! d^{d}r \left\{\tilde{\varphi} (\Delta_{\perp}(\Delta_{\perp} - \tau_{\perp}) - \rho \Delta_{\parallel}) \varphi + \frac{1}{2} g (\nabla_{\parallel} \tilde{\varphi}) \varphi^{2} \right.\\ & & \hspace{170 pt} \left. + \tilde{\varphi} (-\Delta_{\perp}^{2} + \sigma \Delta_{\parallel}) \tilde{\varphi} \right\} \; .\nonumber \end{eqnarray} In comparison with the former model, the vertex and the Gaussian propagator remain the same (\ref{quasipropkorr}), whereas the Gaussian correlator of the quasistatic model reads \begin{equation} \label{bpropkorr} C_{{\bf q}} = \frac{2({\bf q}_{\perp}^{4} + \sigma q_{\parallel}^{2}) }{[{\bf q}_{\perp}^{2}({\bf q}_{\perp}^{2} + \tau_{\perp}) + \rho q_{\parallel}^{2}]^{2}} \; . \end{equation} By dimensional analysis taking into account the causality properties we find the vertex functions $ \Gamma_{1,1} $, $ \Gamma_{2,0} $, and $ \Gamma_{1,2} $ primitively divergent.\\ Neglecting higher orders in $ q $ and $ \epsilon = d_{c} - d $, a one--loop calculation for these vertex functions, with dimensional regularization, results in \begin{eqnarray} \label{bvertex} \mathaccent28695{\Gamma}_{1,1}({\bf q}) & = & {\bf q}_{\perp}^{2}({\bf q}_{\perp}^{2} + \tau_{\perp}) + \mathaccent28695{\rho} q_{\parallel}^{2} + \frac{B_{\epsilon}}{\epsilon} \frac{\mathaccent28695{g}^{2}}{\mathaccent28695{\rho}^{\frac{1}{2}}} q_{\parallel}^{2} \tau_{\perp}^{-\frac{\epsilon}{2}}\\ \mathaccent28695{\Gamma}_{2,0}({\bf q}) & = & -2{\bf q}_{\perp}^{4} -2 \mathaccent28695{\sigma} q_{\parallel}^{2} -\frac{1}{4} \frac{B_{\epsilon}}{\epsilon} \frac{\mathaccent28695{g}^{2}}{\mathaccent28695{\rho}^{\frac{1}{2}}} q_{\parallel}^{2} \tau_{\perp}^{-\frac{\epsilon}{2}}\left(5 + 2 \frac{\mathaccent28695{\sigma}}{\mathaccent28695{\rho}} + \left(\frac{\mathaccent28695{\sigma}}{\mathaccent28695{\rho}}\right)^{2}\right)\nonumber\\ \mathaccent28695{\Gamma}_{1,2}({\bf q}) & = & i \mathaccent28695{g} q_{\parallel} - i \frac{1}{4} \frac{B_{\epsilon}}{\epsilon} \frac{\mathaccent28695{g}^{3}}{\mathaccent28695{\rho}^{\frac{3}{2}}} q_{\parallel} \tau_{\perp}^{-\frac{\epsilon}{2}}\left(1 + \frac{\mathaccent28695{\sigma}}{\mathaccent28695{\rho}}\right) \; , \nonumber \end{eqnarray} where \begin{equation} \label{bgeofaktor} B_{\epsilon} = \frac{1}{(2\pi)^{d}} {\cal O}_{d-1} \Gamma(1 + \frac{\epsilon}{2}) \Gamma(\frac{3 - \epsilon}{2}) \Gamma(\frac{1}{2}) \end{equation} is a suitably chosen $ \epsilon $--dependent factor. Notice that although not performed a two--loop calculation could here also be done with the method described in the Appendix.\\ The primitive divergences are absorbed in the redefinition of the parameters \begin{equation} \label{breparameter} \mathaccent28695{\rho} = Z_{\rho} \rho, \; \mathaccent28695{\sigma} = Z_{\sigma} \sigma, \; \mathaccent28695{g} = \mu^{\frac{\epsilon}{2}} Z_{u} u \; . \end{equation} In comparison to the latter model the coupling constant $ \mathaccent28695{\sigma} $ has additionally to be renormalized. From Equation (\ref{bvertex}) we easily obtain the $ Z $--Factors in minimal subtraction \begin{eqnarray} \label{bZ} Z_{\rho} & = & 1 - \frac{v}{\epsilon} + {\cal O}(v^{2}) \\ Z_{\sigma} & = & 1 - \frac{1}{8} \frac{v}{\epsilon} \frac{1}{w} (5 + 2w + w^{2}) + {\cal O}(v^{2}) \nonumber\\ Z_{u} & = & 1 + \frac{1}{4} \frac{v}{\epsilon} (1 + w) + {\cal O}(v^{2}) \; ,\nonumber \end{eqnarray} expressed as functions of the dimensionless renormalized parameters $ v:= B_{\epsilon} u^{2} \rho^{-\frac{3}{2}} $ and $ w:= \sigma \rho^{-1} $ that are invariant under the longitudinal scale transformation.\\ In analogy to the former model we obtain the renormalization group equation \begin{equation} \label{brgg} \left [\beta_{v} \frac{\partial}{\partial v} + \beta_{w} \frac{\partial}{\partial w} +\rho \zeta \frac{\partial}{\partial \rho} + \mu \frac{\partial}{\partial \mu}\right ] \Gamma_{\tilde{n},n} (\{q_{\parallel},{\bf q}_{\perp}\},\tau_{\perp},v,w,\rho,\mu) = 0 \; \end{equation} with the parameter functions depending on $ v $ and $ w $ \begin{eqnarray} \label{bwilson} \beta_{v} & := & \left. \mu \frac{\partial v}{\partial \mu}\right |_{o} = -v \left[\epsilon - \frac{1}{2} v(4 + w) + {\cal O}(v^{2})\right]\\ \beta_{w} & := & \left. \mu \frac{\partial w}{\partial \mu}\right |_{o} = - \frac{1}{8} v [5 - 6w + w^{2} ] + {\cal O}(v^{2}) \nonumber\\ \zeta & := & \left. \mu \frac{\partial \ln \rho}{\partial \mu}\right |_{o} = - v + {\cal O}(v^{2}) \; . \nonumber \end{eqnarray} The associated characteristics are defined by \begin{eqnarray} \label{bfluss} l \frac{d}{dl} \bar{v}(l) = \beta_{v}(\bar{v}(l),\bar{w}(l)) & \qquad\qquad & l \frac{d}{dl} \ln \bar{\rho}(l) = \zeta(\bar{v}(l),\bar{w}(l)) \\ l \frac{d}{dl} \bar{w}(l) = \beta_{w}(\bar{v}(l),\bar{w}(l)) & \qquad\qquad & l \frac{d}{dl} \bar{\mu}(l) = \bar{\mu}(l) \nonumber \end{eqnarray} with the initial condition $ \bar{w}(l=1) = w $ and the initial conditions from Equation (\ref{uanfang}). In the scaling limit $ l \ll 1 $, $ \bar{v}(l) $ and $ \bar{w}(l) $ flow to an infrared stable fixed point $ (v_{*},w_{*}) $ given by the zeroes of $ \beta_{v} $ and $ \beta_{w} $ with a positive gradient. The zeroes of $ \beta_{w} $ in this order are \begin{equation} \label{nullw} w_{1} \; = \; 1 \qquad \qquad w_{2} \; = \; 5 \; , \end{equation} and those of $ \beta_{v} $ \begin{equation} \label{nullv} v_{1} \; = \; 0 \qquad \qquad v_{2} \; = \; \frac{2 \epsilon}{4 + w} \; . \end{equation} This yields the infrared stable fixed point \begin{equation} \label{bfixpunkt} w_{*} = 1 + {\cal O}(\epsilon) \qquad v_{*} = \frac{\textstyle 2}{\textstyle 5} \epsilon + {\cal O}(\epsilon^{2}) \; . \end{equation} The domain of attraction of this fixed point can easily be recognized as \begin{equation} \label{attrakt} v > 0 \qquad w < 5 + {\cal O}(\epsilon) \; . \end{equation} In the case $ w > 5 + {\cal O}(\epsilon) $ the critical behaviour of the system is dominated by the degenerate fixed point \begin{equation} \label{entartfix} w_{*} = \infty \qquad v_{*} = 0 \; . \end{equation} Concerning this degenerate fixed point we remark the following:\\ i) For $ w \rightarrow \infty $ the transverse noise vanishes in comparison to the longitudinal one. This can be seen easily by substituting $ \rho q_{\parallel}^{2} \rightarrow q_{\parallel}^{2} $ and extracting $ \sigma \rho^{-1} = w $ from the correlator. Then the remaining coefficient of the longitudinal part in the numerator of the correlator is $ 1 $, the coefficient of the transverse part is $ w^{-1} $. At the degenerate fixed point, this system here behaves as the model with a random potential with equally deep potential valleys and whose noise is therefore given by Equation (\ref{trauschen}). This model is analyzed in the next section, but we anticipate some results concerning the degenerate fixed point. According to Eq. (\ref{tfixpunkt}) the degenerate fixed point possesses a finite fixed point value $ (v\cdot w)_{*} = \frac{8}{3} \epsilon + {\cal O}(\epsilon^{2}) $. Although the deviations from normal diffusive behaviour only appear in order two--loop, the positive $ \eta $ from Eq. (\ref{teta}) demonstrates the system to be also superdiffusive at the degenerate fixed point.\\ ii) Both a ``normal" and a degenerate fixed point are also observed in the noncritical model with quenched disorder \cite{BJ1}. While here at the transverse phase transition the situation of two fixed points appears in the random potential with equally high mountains and the degenerate fixed point is described by the model with the random potential with equally deep valleys, in the noncritical region, however, the situation of a normal and a degenerate fixed point occurs in the model with unsymmetric random potential and the degenerate fixed point is described by both the model with equally high mountains and equally deep valleys, for these latter models are identical in the noncritical region.\\ Now we proceed to investigate the normal fixed point (\ref{bfixpunkt}). For this fixed point the anomaly--exponent reads \begin{equation} \label{beta} \eta = \frac{\textstyle 1}{\textstyle 5} \epsilon + {\cal O}(\epsilon^{2}) \; . \end{equation} At the fixed point, the second characteristic has the form \begin{equation} \label{brhofluss} \bar{\rho}(l) = \rho l^{-2 \eta} \end{equation} in analogy to Equation (\ref{rhofluss}).\\ Returning to the dynamic model and again exploiting the renormalization group equation at the fixed point, dimensional analysis, and the invariant scale transformation we obtain the universal scaling behaviour of the vertex functions in the asymptotic limit \begin{eqnarray} \label{bskalkrit} \Gamma_{\tilde{n},n} (\{q_{\parallel},{\bf q}_{\perp},\omega\},\tau_{\perp},v_{*},w_{*},\lambda,\rho,\mu) & = & \\ & & \hspace{-200 pt}l^{-\frac{1}{2} \eta (\tilde{n} + n -2) -\frac{1}{2} \tilde{n} (d+5) - \frac{1}{2} n(d-3) + d + 5} \Gamma_{\tilde{n},n} \left(\left\{\frac{\textstyle q_{\parallel}}{\textstyle l^{2+\eta}},\frac{\textstyle {\bf q}_{\perp}}{\textstyle l},\frac{\textstyle \omega}{\textstyle l^{4}}\right\}, \frac{\textstyle \tau_{\perp}}{\textstyle l^{2}},v_{*},w_{*},\lambda,\rho,\mu\right) \; .\nonumber \end{eqnarray} Thus, only longitudinal lengths scale anomalously below the upper critical dimension $ d_{c} = 7 $. From this equation we derive the scaling form of the density response function \begin{equation} \label{bsus} \chi({\bf q},t) \; = \; f\left(q_{\parallel}^{2} t^{1+\frac{1}{2} \eta},q_{\perp}^{2} t^{\frac{1}{2}}\right) \end{equation} which coincides with the form found for the preceding model. The positivity of $ \eta = \frac{1}{5} \epsilon + {\cal O}(\epsilon^{2}) $ signifies superdiffusive behaviour below $ d_{c} = 7 $. A comparison with the corresponding model without quenched disorder again shows the quenched disorder to be the reason for the enhanced spread of fluctuations in the driving force direction in a dimensional interval that is here, however, $ 5 < d < 7 $.\\ A straightforward extrapolation of the anomaly--exponent into three dimensions by setting $ \epsilon = 4 $ into the one--loop result gives the approximate value \begin{equation} \label{beta31} \eta(d=3) \; = \; \frac{4}{5} \; . \end{equation} This numerical value is very distinct from the one--loop and two--loop values for $ \eta $ in the model with unsymmetric random potential.\\ We now discuss the case that the transverse noise term with $ \gamma \neq 0 $ of Equation (\ref{udynfunktional}) is generated by coarse graining even in this symmetric random potential. Then an additional scaling variable $ \gamma/l^{\phi} $ arises in (\ref{bskalkrit}). The crossover exponent is here $ \phi = 2 $ and it simply reflects the naive dimension of the coupling constant $ \gamma $, because $ \gamma $ itself is invariant under the longitudinal scale transformation and for dimensional reasons the transverse noise term with $ \gamma \neq 0 $ needs no additional renormalization. Since $ \gamma $ is a relevant variable, for $ \gamma \neq 0 $ the system eventually flows to the fixed point of the model with unsymmetric random potential. \subsection{Random Potential With Equally Deep Potential Valleys} Having no transverse noise according to Equation (\ref{trauschen}) this model is formally obtained from the model with random potential with equally high mountains by setting the transverse noise coefficient $ \alpha = 0 $. Then the relevant dynamic functional is here \begin{eqnarray} \label{tJ} {\cal J}[s,\tilde{s}] & = & \int\!\! d^{d}r \left\{\int\!\! dt \left[\tilde{s} \dot{s} + \lambda \tilde{s} (\Delta_{\perp}(\Delta_{\perp} - \tau_{\perp}) - \rho \Delta_{\parallel}) s + \frac{\textstyle 1}{\textstyle 2} \lambda g (\nabla_{\parallel} \tilde{s}) s^{2} \right]\right. \nonumber\\ & & \hspace{180 pt}\left. - \sigma \left[\lambda \int\!\! dt \nabla_{\parallel} \tilde{s} \right]^{2} \right\} \; . \end{eqnarray} The dimensions of lengths, fields, and coupling constants as well as the upper critical dimension $ d_{c} = 7 $ are as in the latter model. The symmetry properties of this dynamic functional are the same as in both preceding models with the exception of the invariant scale transformation. The lack of the transverse noise term has the consequence that the dynamic functional is invariant under a scale transformation depending on two parameters \begin{eqnarray} \label{tskalen} r_{\parallel} \rightarrow \beta r_{\parallel} & \qquad & r_{\perp} \rightarrow r_{\perp}\\ \tilde{s} \rightarrow \alpha \tilde{s} & \qquad & s \rightarrow \alpha^{-1}\beta^{-1} s \nonumber\\ \rho \rightarrow \beta^{2} \rho & \qquad & g \rightarrow \alpha \beta^{2} g \qquad\qquad \sigma \rightarrow \alpha^{-2}\beta \sigma \; . \nonumber \end{eqnarray} Therefore, the combination of coupling constants $ g^{2} \sigma \rho^{-\frac{5}{2}} $ is the appropriate invariant variable in this model and is exactly the product of the variables being invariant each in the latter model, but not here.\\ Because of the quenched disorder we again use the transformation to a quasi--static Hamiltonian \begin{equation} \label{tquasihamil} {\cal H}[\varphi,\tilde{\varphi}] = \int\!\! d^{d}r \left\{\tilde{\varphi} (\Delta_{\perp}(\Delta_{\perp} - \tau_{\perp}) - \rho \Delta_{\parallel}) \varphi + \frac{1}{2} g (\nabla_{\parallel} \tilde{\varphi}) \varphi^{2} + \sigma \tilde{\varphi} \Delta_{\parallel} \tilde{\varphi} \right\} \; . \end{equation} Vertex and Gaussian propagator are the same as in both preceding models, whereas the Gaussian correlator reads \begin{equation} \label{tpropkorr} C_{{\bf q}} = \frac{2 \sigma q_{\parallel}^{2} }{[{\bf q}_{\perp}^{2}({\bf q}_{\perp}^{2} + \tau_{\perp}) + \rho q_{\parallel}^{2}]^{2}} \; . \end{equation} As in the previous model $ \Gamma_{1,1} $, $ \Gamma_{2,0} $, and $ \Gamma_{1,2} $ are primitively divergent. We have computed them in the lowest nonvanishing order, i.e. we have performed a one--loop calculation for $\Gamma_{2,0} $ and $ \Gamma_{1,2} $ and a two--loop calculation for $ \Gamma_{1,1} $. The two--loop calculation involves similar integrals as in the model with unsymmetric random potential and has also been performed with the help of inverse Mellin transformation (cf. Appendix).\\ In dimensional regularization we obtain the primitively divergent vertex functions, up to higher order terms in $ q $ and $ \epsilon = d_{c} - d $ \begin{eqnarray} \label{tvertex} \mathaccent28695{\Gamma}_{1,1}({\bf q}) & = & {\bf q}_{\perp}^{2}({\bf q}_{\perp}^{2} + \tau_{\perp}) + \mathaccent28695{\rho} q_{\parallel}^{2} - 4 \frac{B_{\epsilon}^{2}}{\epsilon^{2}} \frac{\mathaccent28695{g}^{4}\mathaccent28695{\sigma}^{2}}{\mathaccent28695{\rho}^{4}} q_{\parallel}^{2} \tau_{\perp}^{-\epsilon}( 0 - 0.001799 \epsilon) \\ \mathaccent28695{\Gamma}_{2,0}({\bf q}) & = & -2 \mathaccent28695{\sigma} q_{\parallel}^{2} -\frac{1}{4} \frac{B_{\epsilon}}{\epsilon} \frac{\mathaccent28695{g}^{2}\mathaccent28695{\sigma}^{2}}{\mathaccent28695{\rho}^{\frac{5}{2}}} q_{\parallel}^{2} \tau_{\perp}^{-\frac{\epsilon}{2}} \nonumber\\ \mathaccent28695{\Gamma}_{1,2}({\bf q}) & = & i \mathaccent28695{g} q_{\parallel} - i \frac{1}{4} \frac{B_{\epsilon}}{\epsilon} \frac{\mathaccent28695{g}^{3}\mathaccent28695{\sigma}}{\mathaccent28695{\rho}^{\frac{5}{2}}} q_{\parallel} \tau_{\perp}^{-\frac{\epsilon}{2}} \; , \nonumber \end{eqnarray} where $ B_{\epsilon} $ is defined by Equation (\ref{bgeofaktor}). The same redefinition of the coupling constants as in Equation (\ref{breparameter}) absorbs these divergences and yields in minimal subtraction \begin{eqnarray} \label{tZ} Z_{\rho} & = & 1 - 0.007196 \frac{v^{2}}{\epsilon} + {\cal O}(v^{3})\\ Z_{\sigma} & = & 1 - \frac{1}{8} \frac{v}{\epsilon} + {\cal O}(v^{2}) \nonumber\\ Z_{u} & = & 1 + \frac{1}{4} \frac{v}{\epsilon} + {\cal O}(v^{2}) \; ,\nonumber \end{eqnarray} where $ v:= B_{\epsilon} u^{2} \sigma \rho^{-\frac{5}{2}} $ is the dimensionless renormalized variable of the model. This variable is invariant under the longitudinal scale transformation (Eq. (\ref{tskalen})) and is the product of $ v $ and $ w $ in the latter model, but it will here merely be denoted by $ v $ for simplicity.\\ The renormalization group equation reads \begin{equation} \label{trgg} \left [\beta_{v} \frac{\partial}{\partial v} + \rho \zeta_{\rho} \frac{\partial}{\partial \rho} +\sigma \zeta_{\sigma} \frac{\partial}{\partial \sigma} + \mu \frac{\partial}{\partial \mu}\right ] \Gamma_{\tilde{n},n} (\{q_{\parallel},{\bf q}_{\perp}\},\tau_{\perp},v,\rho,\sigma,\mu) = 0 \end{equation} with the Wilson parameter functions \begin{eqnarray} \label{twilson} \beta_{v} & := & \left. \mu \frac{\partial v}{\partial \mu}\right |_{o} \; = -v \left[\epsilon - \frac{3}{8} v + {\cal O}(v^{2})\right]\\ \zeta_{\sigma} & := & \left. \mu \frac{\partial \ln \sigma}{\partial \mu}\right |_{o} = - \frac{1}{8} v + {\cal O}(v^{2})\nonumber\\ \zeta_{\rho} & := & \left. \mu \frac{\partial \ln \rho}{\partial \mu}\right |_{o} = - 0.01439 v^{2} + {\cal O}(v^{3}) \nonumber \end{eqnarray} only depending on $ v $.\\ The infrared stable fixed point, being a zero of $\beta_{v} $, is found at \begin{equation} \label{tfixpunkt} v_{*} = \frac{\textstyle 8}{\textstyle 3} \epsilon + {\cal O}(\epsilon^{2}) \; . \end{equation} Hence we obtain the anomaly--exponent \begin{equation} \label{teta} \eta = 0.0512 \epsilon^{2} + {\cal O}(\epsilon^{3}) \; . \end{equation} The characteristics of the renormalization group equation are defined by Equation (\ref{ufluss}) and an additional equation for $ \bar{\sigma}(l) $ that has the same structure as the equation for $ \bar{\rho}(l) $. At the fixed point the solutions for these flow equations read \begin{equation} \label{tfluss} \bar{\rho}(l) = \rho l^{-2 \eta} \qquad\qquad \bar{\sigma}(l) = \sigma l^{\zeta_{\sigma *}} \;\; , \end{equation} where \begin{equation} \label{tsigma} \zeta_{\sigma *} \; := \; \zeta_{\sigma}(v_{*}) = - \frac{1}{3} \epsilon + {\cal O}(\epsilon^{2}) \;\; . \end{equation} Returning to the dynamic model and combining the renormalization group equation at the fixed point, dimensional analysis, and the invariant longitudinal scale transformation we obtain the universal scaling behaviour of the vertex functions in the asymptotic limit \begin{eqnarray} \label{tskalkrit} \Gamma_{\tilde{n},n} (\{q_{\parallel},{\bf q}_{\perp},\omega\},\tau_{\perp},v_{*},\lambda,\rho,\sigma,\mu) & = & \nonumber\\ & & \hspace{-80 pt}l^{-\eta (-1 + \frac{3}{2} n -\frac{1}{2} \tilde{n}) + \frac{1}{2} \zeta_{\sigma *}(\tilde{n} - n) -\frac{1}{2} \tilde{n} (d+5) - \frac{1}{2} n(d-3) + d + 5}\nonumber\\ & & \hspace{-40 pt} \cdot \Gamma_{\tilde{n},n} \left(\left\{\frac{\textstyle q_{\parallel}}{\textstyle l^{2+\eta}},\frac{\textstyle {\bf q}_{\perp}^{}}{\textstyle l^{}},\frac{\textstyle \omega}{\textstyle l^{4}}\right\}, \frac{\textstyle \tau_{\perp}}{\textstyle l^{2}},v_{*},\lambda,\rho,\sigma,\mu\right) \; . \end{eqnarray} We again see that only longitudinal lengths scale anomalously. In contrast to both preceding models there is here an additional exponent $ \zeta_{\sigma *} $ appearing in the global $ l $--factor. As this exponent is cancelled for $ \Gamma_{1,1} $, the density response function nevertheless has the scaling form \begin{equation} \label{tsus} \chi({\bf q},t) \; = \; f\left(q_{\parallel}^{2} t^{1+\frac{1}{2} \eta},q_{\perp}^{2} t^{\frac{1}{2}}\right) \end{equation} in analogy to both preceding models. As the anomaly--exponent $ \eta = 0.0512 \epsilon^{2} + {\cal O}(\epsilon^{3}) $ is positive and the upper critical dimension $ d_{c} = 7 $ is the same as in the preceding model, all statements concerning superdiffusion and speeding--up of fluctuations by quenched disorder are here also valid.\\ The straightforward extrapolation of $ \eta $ from $ d_{c} = 7 $ to $ d = 3 $ by setting $ \epsilon = 4 $ produces the approximative value \begin{equation} \label{teta31} \eta(d=3) \; = \; 0.82 \; . \end{equation} This numerical value is close to the one--loop value for $ \eta $ in $ d = 3 $ in the model with random potential with equally high mountains, but is very far from the one--loop and two--loop values in the model with unsymmetric random potential.\\ We now investigate the possibility that even in this symmetric random potential transverse noise terms are produced by coarse graining. First, we consider the case that transverse noise proportional to $ \alpha \int d^{d}r[\lambda \int dt \Delta_{\perp} \tilde{s}]^{2} $ is generated. Then we are in the situation of the preceding model. If $ \alpha $ is small the system is in the region of attraction of the degenerate fixed point discussed in this section. This implies that the transverse noise vanishes under the renormalization flow and the results of this section remain valid.\\ Second, if transverse noise proportional to $ \gamma \int d^{d}r[\lambda \int dt \nabla_{\perp} \tilde{s}]^{2} $ is produced, an additional scaling variable $ \gamma/l^{\phi} $ arises in (\ref{tskalkrit}). Due to dimensional reasons, this relevant operator needs no additional renormalization. The invariant scale transformation according to Equation (\ref{tskalen}) implies $ \gamma \rightarrow \beta^{-1}\alpha^{-2} \gamma $. Thus, the crossover exponent is related to the exponents $ \eta $ and $ \zeta_{\sigma *} $ by $ \phi = 2 + 2\eta - \zeta_{\sigma *} $. For $ \gamma \neq 0 $ the system eventually flows to the fixed point of the model with unsymmetric random potential, because $ \gamma $ is a relevant variable.
\section{Introduction} Weak superconducting links \cite{Jo} include the tunnel structures $S-I-S$ (superconductor-insulator-superconductor) and the contacts with direct conductivity, $S-N-S$ ($N$ is the normal layer) and $S-c-S\,$ ($c$ is a geometrical constriction) . Superconducting constrictions can be modelled as the orifice with diameter $d$ in a inpenetrable sheet for electrons between two superconducting half spaces (point contact), or as a narrow channel with length $L\,$ in a contact with superconducting banks (microbridge). Aslamazov and Larkin \cite{AL} have shown on the basis of a solution of the Ginzburg-Landau (GL) equations that in the dirty limit and for small constriction's sizes $L,d\ll \xi (T)$ ( $\xi (T)$ is GL coherence length) the $S-C-S\,$ contact can be described by a Josephson model with the current-phase relation \begin{equation} I=I_{c}\sin {\varphi }, I_{c}=\pi \Delta _{0}^{2}(T)/(4eR_{N}T_{c}), \label{IAL} \end{equation} where $I_{c}$ is the Josephson critical current, $\Delta _{0}$ the absolute value of the order parameter in the bulk banks, $T_{c}$ the critical temperature and $R_{N}$ the normal-state resistance of dirty microbridge. The critical current of microbridge (\ref{IAL}) depends on the bridge length as $I_{c}\sim 1/L$. The expression (\ref{IAL}) is valid within the domain of applicability of GL approach, {\it i.e.} for temperatures $T$ close to $% T_{c} $ and $L,d\gg \xi _{0}$ ($\xi _{0}\simeq v_{F}/T_{c}$ is the coherence length at $T=0,v_{F}$ is the Fermi velocity). The present level of technology have made it possible to study the ultrasmall Josephson weak links with the dimensions up to interatomic size. For example, it can be nanosized microchannels produced by means of a scanning tunneling microscope \cite{Poza} or point contacts and microchannels obtained by using the mechanical controllable break technique \cite{Mul}, \cite{Na, Na2} . The microchannels between two superconductors can also arise spontaneously as microshorts in tunnel junctions\cite{Ya}, with the length $L$ determined by the thickness of an insulator layer. The value of the critical current $I_{c}$ of such microshorts presents the special interest in the case of tunnel structures based on high-$T_{c}$ metalloxide compounds. The case of small microconstrictions with dimensions of the order or smaller than coherence length $\xi _{0}$ , when the expression (\ref{IAL}) for the critical current $I_{c}\sim 1/L$ is not valid, requires the microscopic consideration even for $T$ near $T_{c}$. Such microscopic theory of stationary Josephson effect in microconstrictions was developed in Ref. \cite{KO} for the ballistic channel of zero length $L=0 $ in the model of the orifice of diameter $d<<\xi _{0}$. The Josephson current in this case is given by: \begin{equation} I=\frac{\pi {\Delta }_{0}(T)}{eR_{0}}\sin {\varphi /2}\tanh {\frac{{\Delta }% _{0}(T)\cos {\varphi /2}}{2T},}-\pi <\varphi <\pi , \label{IKO} \end{equation} \begin{equation} R_{0}^{-1}=\frac{1}{2}Se^{2}v_{F}N(0), \label{RSh} \end{equation} where $S=\pi d^{2}/4$ is the contact cross-section area, $N(0)=mp_{F}/(2{\pi }^{2})$ is electron density of states at the Fermi surface. At temperatures $% T_{c}-T\ll T_{c}$ expression (\ref{IKO}) coincides with the Aslamazov-Larkin result, Eq. (\ref{IAL}), in which instead of the normal resistance $R_{N}\,$ for dirty metal, the ballistic Sharvin resistance \cite{Sha}\thinspace $% R_{0}\,$(\ref{RSh}) is substituted. The purpose of this work is to present a microscopic theory of current carrying states in the ballistic microbridges having the length $L$ arbitrary in the scale of the coherence length $\xi _{0}$. We have investigated the dependence of Josephson critical current on the ratio $% L/\xi _{0}$ and analyzed the transition from the case of $I_{c}(L=0)$ (\ref {IKO}) to $I_{c}\sim 1/L$ (\ref{IAL}) with increasing of the length $L$. In Sec.2 we formulate the model of a microbridge and the microscopic equations for Green's functions with boundary conditions at the bridge edges. When the effects on the critical current of the microconstriction's length are studied, the crucial point, as always in space inhomogeneous superconducting state, consists in the self-consistent treatment of the order parameter distribution $\Delta ({\bf r})$ inside the weak link. In Sec.3 the closed integral equation for the order parameter $\Delta $ in microchannel is derived for temperatures near $T_{c}$, which in strongly inhomogeneous ($L\sim $ $\xi _{0}$) microcontact geometry replaces the differential GL equation. The critical current $I_{c}(L)$ is expressed in terms of the solution of this integral equation. The limiting cases of $L\ll \xi _{0}$ and $L\gg \xi _{0}$ are considered. We shall show that beside the characteristic scale $\xi _{0}$ the length $a_{D}\simeq v_{F}/\omega _{D}$ ($% \omega _{D}$ is Debye frequency) appears in the case of ultra small channel. The length $L\sim a_{D}$ is the length at which the frequency of the ballistic flight of electron from one bank to another becomes comparable with the frequency $\omega _{D}$ , which characterizes the retardation of the electron-phonon interaction. In conventional superconductors the value of the coherence length $\xi _{0}$ is about $10^{-4}cm$ and is much lager than $a_{D}\sim 100\AA $. But in high-$T_{c}$ metalloxide compounds we have the situation with $\xi _{0}$ comparable with $a_{D}.$ Thus, in high-$T_{c}$ compounds the critical current of the contact with dimensions $\sim a_{D}\sim \xi _{0}$ will be sensitive to the effects of strong coupling. \section{Model and basic equations} We consider the model of a contact in the form of a filament (narrow channel) that joins two superconducting half-spaces (massive banks) (Fig.1). The length $L$ and the diameter $d$ of the channel are assumed to be large as compared with the Fermi wavelength $\lambda _{F}$, so we can apply the quasi-classical approximation. In the ballistic case, we proceed from the quasi-classical Eilenberger equation for the energy-integrated Green's function \cite{Ei}: \begin{equation} {\bf v_{F}}.\frac{\partial }{\partial {\bf r}}\hat{G}+[\omega \hat{\tau}_{3}+% \hat{\Delta},\hat{G}]=0, \label{eq1} \end{equation} where \[ \hat{G}(\omega ,{\bf v_F},{\bf r})=\left( \begin{array}{cc} g_\omega & f_\omega \\ f_\omega ^{+} & -g_\omega \end{array} \right) \] is the matrix Green function, which depends on the Matsubara frequency $% \omega $, the electron velocity on the Fermi surface ${\bf v_F}$ and the spatial variable ${\bf r}$; \[ \hat{\Delta}({\bf r})=\left( \begin{array}{cc} 0 & \Delta \\ \Delta ^{\ast } & 0 \end{array} \right) \] is the superconducting pair potential; $\hat{\tau}_{i}$ ($i=1,2,3$) are Pauli matrices. Equation for matrix Green's function (\ref{eq1}) is supplemented by the normalization condition \cite{La} \begin{equation} \hat{G}^{2}=1. \label{G^2} \end{equation} The off-diagonal potential $\Delta ({\bf r)}$ must be determined from the self-consistency equation: \begin{equation} \Delta ({\bf r})=\lambda 2\pi T\sum_{\omega >0}<f> \label{delta0} \end{equation} in which $<...>$ stands for averaging over directions of ${\bf v_{F}}$ on the Fermi surface and $\lambda $ is the electron-phonon coupling constant. In the BCS model the summation over $\omega $ contains the cutoff on the frequency $\omega _{D},$ which is of the order of the Debye frequency . The equations (\ref{eq1}) and (\ref{delta0}) are supplemented by the values of Green's functions and $\Delta $ in the bulk superconductors $S_{1}$ and $% S_{2}$ far from the channel ends : \begin{equation} \hat{G}_{1,2}=\frac{\omega \hat{\tau}_{3}+\hat{\Delta}_{1,2}}{\Omega },\hat{% \Delta}_{1,2}=\Delta _{0}(\cos (\varphi /2)\hat{\tau}_{1}\pm \sin (\varphi /2)\hat{\tau}_{2}), \label{bound0} \end{equation} thus the phase $\varphi $ is the total phase difference on the contact. Also we have to determine the boundary conditions concerning the reflection of the electrons from the surface of the superconductors ${\bf r}_{S}$. For simplicity we will assume that at ${\bf r}_{S}$ electrons undergo the specular reflection. Then for quasiclassical Green's function we have the boundary condition (Ref.\cite{KO}): \begin{equation} G({\bf v_{F}},{\bf r}_{S})=G({\bf v_{F}}^{\prime },{\bf r}_{S}) \end{equation} in which ${\bf v_{F}}$ and ${\bf v_{F}}^{\prime }$ are the velocities of the incident and specular reflected electron. These velocities are related by the conditions, which conserve the component of ${\bf v}_{F}$ parallel to the reflecting surface ${\bf r}_{S}$ and changes the sign of the normal component. The solutions of the equations (\ref{eq1}) and (\ref{delta0}) allow us to calculate the current density ${\bf j}$: \begin{equation} {\bf j}(r)=-4i\pi eN(0)T\sum_{\omega >0}<{\bf v_{F}}g_{\omega }>. \label{I} \end{equation} In the case of the microconstriction shown in Fig.1, under the conditions $% d\ll \xi _{0}$ and $L\gg d$ ($d$ is the contact diameter), inside the filament we can solve the one-dimensional Eilenberger equations with $\Delta =\Delta (z)$ $.$ The banks of the bridge are equivalent here to certain boundary conditions for Green's function $\hat{G}(v_{z,}z)$ at points $z=\pm \frac{L}{2}$. Following the procedure, which was described in Refs. \cite{KO}% , one can find the Green's functions at the end points ($z=\pm \frac{L}{2}$) from the general solutions of Eq.(\ref{eq1}) in superconducting half-spaces $% S_{1}$ and $S_{2}$ with conditions (5). They are given by \begin{equation} \begin{array}{c} \hat{G}(z=\mp \frac{L}{2})=\hat{G}_{1,2}+A_{1,2}[\Delta _{0}\hat{\tau}% _{3}-(\omega \cos (\varphi /2)+i\eta \Omega \sin (\varphi /2))\hat{\tau}% _{1}\mp \\ (\omega \sin (\varphi /2)-i\eta \Omega \cos (\varphi /2))\hat{\tau}_{2}], \end{array} \label{bound} \end{equation} where $\Omega =\sqrt{\omega ^{2}+{\Delta _{0}}^{2}}$, $\eta =sign(v_{z})$. The arbitrary constants $A_{1,2}$ must be determined by matching of these boundary conditions with the solution for $\hat{G}(v_{z,}z)$ inside the channel. Taking the off-diagonal components of Eq.(\ref{eq1}), we have the following first-order differential equations for the anomalous Green's functions: \begin{equation} \begin{array}{c} v_{z}\frac{df_{\omega }}{dz}+2\omega f_{\omega }=2\Delta (z)g_{\omega }, \\ -v_{z}\frac{df_{\omega }^{+}}{dz}+2\omega f_{\omega }^{+}=2\Delta ^{\ast }(z)g_{\omega }. \end{array} \label{eq2} \end{equation} The normal Green's function $g_{\omega }$, as follows from condition (\ref {G^2}), is expressed in terms of $f_{\omega }$ and $f_{\omega }^{+}$ : \begin{equation} g_{\omega }=\sqrt{1-f_{\omega }f_{\omega }^{+}}. \label{g_om} \end{equation} From equations (\ref{I}), (\ref{delta0}),(\ref{eq2}) and \ref{g_om} one obtains the symmetry relations \begin{equation} f_{\omega }^{+}(v_{z},z)=(f_{\omega }(-v_{z},z))^{\ast },\Delta ^{\ast }(z)=\Delta (-z) \label{f+} \end{equation} and the current conservation inside the channel $dj/dz=0$ . \section{Josephson current and order parameter distribution in superconducting microchannel} In present paper we consider the case of temperatures $T$ close to the critical temperature $T_{c}$. Near the phase transition curve the order parameter $\Delta _{0}(T)$ in the banks is small. In order to find the Josephson current in the lowest order on $\Delta _{0}$ we linearize the equations (\ref{eq2} ) on $\Delta $ and obtain $f_{\omega }\sim \Delta _{0}(T)$, $g_{\omega }\simeq 1-1/2f_{\omega }f_{\omega }^{+}\sim 1-O(\Delta _{0}^{2})$, $j\sim \Delta _{0}^{2}$. The equation for $f_{\omega }$ near $% T_{c}$ takes the form \begin{equation} v_{z}\frac{df_{\omega }}{dz}+2\omega f_{\omega }=2\Delta (z), \label{eq3} \end{equation} with linearized boundary conditions (\ref{bound} ) \begin{equation} \begin{array}{c} f_{\omega }(v_{z}>0,z=-L/2)=\frac{\Delta _{0}}{\omega }e^{-i\frac{\varphi }{2% }}, \\ f_{\omega }(v_{z}<0,z=+L/2)=\frac{\Delta _{0}}{\omega }e^{+i\frac{\varphi }{2% }}. \end{array} \label{bound1} \end{equation} Its solution for arbitrary function $\Delta (z)$ is given by \begin{equation} f_{\omega }(v_{z},z)=\frac{\Delta _{0}}{\omega }e^{-i\eta \frac{\varphi }{2}% }e^{-\frac{2\omega }{v_{z}}(z+\eta L/2)}+e^{-\frac{2\omega }{v_{z}}% z}\int\limits_{-\eta L/2}^{z}dz^{\prime }\frac{2\Delta (z^{\prime })}{v_{z}}% e^{\frac{2\omega }{v_{z}}z^{\prime }}. \label{fomega} \end{equation} The Green's function $f_{\omega }^{+}(v_{z},z)$ is obtained from expression (% \ref{eq3}) with the help of relations (\ref{bound1} ). Substituting function $f_{\omega }(v_{z},z)$ (\ref{fomega}) in the self-consistency equation (\ref{delta0}), we obtain the integral equation for the space-dependent order parameter inside the contact \begin{equation} \Delta (z)=A(z)+\int\limits_{-L/2}^{L/2}dz^{\prime }\Delta (z^{\prime })K(\left| z-z^{\prime }\right| ), \label{eqdelta} \end{equation} where \begin{equation} A(z)=\lambda 2\pi T\sum\limits_{\omega >0}\frac{\Delta _{0}}{\omega }% \left\langle e^{-\frac{\omega L}{v_{z}}}\cosh (\frac{2\omega z}{v_{z}}+i% \frac{\varphi }{2})\right\rangle _{v_{z}>0}, \label{A} \end{equation} \begin{equation} K(z)=\lambda 2\pi T\sum\limits_{\omega >0}\left\langle \frac{1}{v_{z}}e^{-% \frac{2\omega }{v_{z}}z}\right\rangle _{v_{z}>0}. \label{K} \end{equation} The averaging $<...>_{v_{z}>0}$ denotes $\ <F(v_{z}=v_{F}\cos \theta )>_{v_{z}>0}=\int_{0}^{1}d(\cos \theta )F(\cos \theta ).$ In the case of strongly inhomogeneous microcontact problem the integral equation for the order parameter $\Delta $ replaces the differential Ginzburg-Landau equation . It contains the needed boundary conditions at the points of contact between the filament and the bulk superconductors. Some general properties of the solution $\Delta (z)$ of Eq.(\ref{eqdelta} ) follow from the form of the functions ( \ref{A}) and (\ref{K}). Let us write $\Delta (z)$ in the form \begin{equation} \Delta (z)=\Delta _{0}(T)(\cos \frac{\varphi }{2}+iq(z)\sin \frac{\varphi }{2% }) \label{delta} \end{equation} and substitute it in the Eq.( \ref{eqdelta}). For function $q(z)$ we obtain the equation \begin{equation} q(z)=b(z)+\int\limits_{-L/2}^{L/2}dz^{\prime }q(z^{\prime })K(\left| z-z^{\prime }\right| ), \label{q(z)} \end{equation} with $K(z)$ defined by ( \ref{K}) and the new out-integral function $b(z)$ \begin{equation} b(z)=\lambda 2\pi T\sum\limits_{\omega >0}\frac{1}{\omega }\left\langle e^{-% \frac{\omega L}{v_{z}}}\sinh (\frac{2\omega z}{v_{z}})\right\rangle _{v_{z}>0}. \label{b} \end{equation} In obtaining the Eqs(\ref{q(z)}) we have used the relation \begin{equation} \lambda 2\pi T\sum\limits_{\omega >0}^{\omega_{D}}\frac{1}{\omega }=1, \ \ {\rm for}\ \ T\rightarrow T_{c}. \label{lambda} \end{equation} It follows from (\ref{q(z)}), (\ref{K}) and (\ref{b}) that function $q(z)$ has such properties: {\it i) }function $q(z)$ is real, {\it ii) }$q(z)$ does not depend on the phase $\varphi $, {\it \ iii)} $q(-z)=-q(z),q(0)=0.$ Thus, the value of the order parameter $\Delta $ at the center of the contact always equals to $\Delta _{0}(T)\cos \frac{\varphi }{2}$ . Also, the universal phase dependence of $\Delta (z,\varphi )$, which is determined by (% \ref{delta}) and {\it i)-iii), }leads (see below) to the sinusoidal current-phase dependence $j=j_{c}\sin \varphi $. It is emphasized, that these general properties of the ballistic microchannel (within the considered case of ''rigid'' boundary conditions (\ref{bound}) and temperatures close to $T_{c}$) {\it does not depend }on{\it \ }the contact length $L$, in particular, on the ratio of $L/\xi _{0}$. Now we are going to obtain the Josephson current in the system. To calculate the total current $I=Sj$ flowing through the channel at given phase difference $\varphi ,$ we use the equation for the current density (\ref{I}) and the obtained above anomalous Green's function $f_{\omega }$ (\ref{fomega}% ). The normal Green's function $g_{\omega }$ (\ref{g_om}) in the second order on $\Delta _{0}(T)$ equals to $g_{\omega }(v_{z},z)=1-\frac{1}{2}% f_{\omega }(v_{z},z)(f_{\omega }(-v_{z},z))^{\ast }$. It is convenient to calculate the current density at the point $z=0$. By using the expression for $\Delta (z)$ (\ref{delta}), we obtain the general formula for the Josephson current $I(\varphi )$ in terms of function $q(z)$ $:$ \begin{equation} I(\varphi )=I_{c}\sin \varphi , \label{I1} \end{equation} \begin{equation} I_{c}=I_{0}\frac{16T^{2}}{v_{F}}\sum\limits_{\omega >0}\left[ \frac{1}{% \omega ^{2}}\left\langle v_{z}e^{-\frac{\omega L}{v_{z}}}\right\rangle _{v_{z}>0}+\frac{2}{\omega }\int\limits_{0}^{L/2}dzq(z)\left\langle e^{-% \frac{2\omega }{v_{z}}z}\right\rangle _{v_{z}>0}\right] . \label{Ic} \end{equation} Here $I_{0}=\pi \Delta _{0}^{2}(T)/(4eR_{0}T_{c})$ is the critical current at $L=0$. It coincides with the result of Ref. \cite{KO} for the orifice (% \ref{IKO}) at $T$ near $T_{c}$. Expression (\ref{Ic} ) jointly with equation (\ref{q(z)} ) for $q(z)$ describes the dependence of the critical current on the contact length $I_{c}(L)$. It is valid for arbitrary value of the ratio $% L/\xi _{0}$. Note, that in considered here case $T\rightarrow T_{c}$, we have the relation $\xi _{0},L\ll \xi (T).$ Let us introduce the dimensionless quantities \begin{equation} x=z/L,\ell =\frac{\pi T_{c}L}{v_{F}},\frac{\omega }{\pi T_{c}}=2n+1,J_{c}=% \frac{I_{c}}{I_{0}}. \label{dless} \end{equation} In reduced units (\ref{dless}), after taking the average $<...>_{v_{z}>0}$, the equations for $q(x)$ and $J_{c}$ take the form \begin{equation} q(x)=b(x)+\ell \int\limits_{-1/2}^{1/2}dx^{\prime }q(x^{\prime })K(\left| x-x^{\prime }\right| ), \label{q(x)1} \end{equation} \begin{equation} \begin{array}{c} J_{c}=\frac{8}{\pi ^{2}}\sum\limits_{n=0}^{N}\{\frac{\exp [-\ell (2n+1)][1-\ell (2n+1)]}{(2n+1)}-\ell ^{2}Ei[-\ell (2n+1)]+ \\ +4\ell \int\limits_{0}^{1/2}dxq(x)[\frac{\exp [-2\ell (2n+1)x]}{(2n+1)}% +2\ell xEi[-2\ell (2n+1)x]\}, \end{array} \label{Ic1} \end{equation} where \begin{equation} \begin{array}{c} b(x)=\lambda \sum\limits_{n=0}^{N}\{\frac{2\exp [-\ell (2n+1)]\sinh [2\ell (2n+1)x]}{(2n+1)}+\ell (2n+1)(1-2x)Ei[-\ell (2n+1)(1-2x)]+ \\ -\ell (2n+1)(1+2x)Ei[-\ell (2n+1)(1+2x)]\}, \end{array} \label{b1} \end{equation} \begin{equation} K(x)=-2\lambda \sum\limits_{n=0}^{N}Ei[-2\ell (2n+1)x]. \label{K1} \end{equation} Function $Ei(x)=\int_{-\infty }^{x}\frac{\exp (t)}{t}dt$ is the integral exponent. The upper limit $N$ in the sums over $n$ is related to the cutoff frequency $\omega _{D}$ in the BCS model, $N\simeq \omega _{D}/T_{c}$. The value of coupling constant $\lambda $ is related to $N$ by Eq.(\ref{lambda}% ), or in reduced units, $2\lambda \sum\limits_{n=0}^{N}\frac{1}{(2n+1)}=1.$ In the weak coupling limit of $\lambda \ll 1$, we have $N\gg 1$. In general case of the arbitrary value of the parameter $\ell $ ($\ell \simeq L/\xi _{0}$) the Eqs.(\ref{q(x)1}) are the convenient $\,$starting point for the numerical calculation of function $J_{c}(\ell )$. We consider here two limiting cases of $\ell \gg 1$ and $\ell \ll 1.$ For a long microbridge with $\ell \gg 1$ we shall seek a solution of Eq.(\ref {q(x)1}) in the form $q(x)=\alpha x$. Substituting this $q(x)$ in Eq.(\ref {q(x)1}), we find that $\alpha =2+O(1/\ell ).$ Calculating $J_{c}$ (\ref{Ic1}% ) with $q(x)=2x$, we find that the order parameter and the critical current are \begin{equation} \Delta \left( z\right) =\Delta _{0}\left( \cos \frac{\varphi }{2}+i\frac{2z}{% L}\sin \frac{\varphi }{2}\right) ,L\gg \xi _{0}, \label{delta2} \end{equation} \begin{equation} I_{c}(L)=\frac{14}{3\pi ^{2}}\zeta \left( 3\right) I_{0}\frac{\hbar v_{F}}{% T_{c}L}, L\gg \xi _{0}. \label{Ic2} \end{equation} Expressions (\ref{delta2}), (\ref{Ic2}) coincide with the solution of GL equations (with effective boundary conditions for the order parameter $% \Delta $ ) for the clean superconducting microbridge \cite{KOJ}. Thus, our microscopic approach with the boundary conditions (\ref{bound}) for Green functions (not for $\Delta )$ gives the results of the phenomenological theory at $L\gg \xi _{0}.$ For a short microbridge with $\ell \ll 1,$ in zero approximation on $\ell $ we have that $q(x)=0$ ($\Delta \left( z\right) =\Delta _{0}\cos \frac{% \varphi }{2})$, $J_{c}=1$. Or, in dimension units, $I_{c}(0)=I_{0},$ in agreement with formula (\ref{IKO}). The corrections to the zero approximation depend on the value of the product $\ell N$. For very small $% \ell \ll T_{c}/\omega _{D}$ ({\it i.e}. $L\ll a_{D}\simeq v_{F}/\omega _{D})$% , the product $\ell N$ becomes small , although the $N\gg 1$ . As a result, when $q(x,\ell )$ and $J_{c}(\ell )$ are calculated in the region $L<a_{D},$ the cutoff in the sums over $n$ must be taken into account. Apparently, when the cutoff frequency appears explicitly but not through the value of $T_{c},$% the applicability of the BCS theory becomes questionable. More rigorous consideration, based on the Eliashberg theory of superconductivity \cite{Eli}% , is needed in this case. Nevertheless, by using the BCS model with cutoff frequency we suppose qualitatively to take into account the retardation effects of electron-phonon coupling in our problem. In the domain, defined by the following inequalities : $\ell N\ll 1$, $N\gg 1$, $\ell \ll 1$ , the functions $b(x)$ (\ref{b1}) and $K(x)\,$ (\ref{K1}) have the asymptotes: \begin{equation} b(x)=4\lambda \ell N\{x\ln (\ell N)+x(C+\ln 2)+\frac{1}{4}[\ln (\frac{1+2x}{% 1-2x})+2x\ln (1-4x^{2})]\}, \label{b2} \end{equation} \begin{equation} K(\left| x\right| )=-2\lambda N[\ln (2\ell N\left| x\right| )-1]. \label{K2} \end{equation} Where $C\simeq 0.577$ is the Euler constant. As it follows from Eqs.(\ref{b2}% ), (\ref{K2}) in this case the integral term in the equation (\ref{q(x)1}) is small, and calculating the critical current in the first approximation on the small parameter $\ell N$ we can put $q(x)=b(x).$ As a result we have \begin{equation} \Delta (z)=\Delta _{0}(T)(\cos \frac{\varphi }{2}+ib(z/L)\sin \frac{\varphi }{2}),L\ll a_{D}, \label{delta3} \end{equation} with $b(x)$ defined by expression (\ref{b2}\ ), \begin{equation} I_{c}(L)=I_{0}(1-\frac{8}{\pi \lambda }\frac{T_{c}L}{v_{F}}) , L\ll a_{D}. \label{Ic3} \end{equation} In the region $\{\ell \ll 1$ and $\ell N\lesssim 1\}$ the integral term in the equation (\ref{q(x)1}) is numerically small as compared with the out-integral term $b(x)$. By using in equation (\ref{q(x)1}) the $q(x)=b(x)$ as the rough approximation, we calculate the function $J_{c}(\ell )$ shown in Fig.2. For the case $\ell \ll 1,$ and $\ell N\gg 1,$ we can put $N=\infty $ in the equation for $q(x)$ and $J_{c}(\ell ).$ The corrections to the critical current in this region of the length $L$ can be estimated as \begin{equation} I_{c}\approx I_{0}\left( 1-const\frac{L}{\xi _{0}}\ln \frac{\xi _{0}}{L}% \right) ,\quad a_{D}\ll L\ll \xi _{0.} \label{Ic4} \end{equation} The expressions (\ref{Ic2}), (\ref{Ic3}) and (\ref{Ic4}) describe the dependence of the critical current on the contact length in the limiting cases of short and long channel. With increasing length $L$ the critical current decreases. For ultrasmall $L\lesssim a_{D}$ the value of $\delta I_{c}/I_{0}\sim (1/\lambda )(L/\xi _{0})$ directly depends on the BCS coupling constant $\lambda $, and consequently it is sensitive to the effects of the strong electron-phonon coupling. \section{Conclusion} We have studied the size dependence of the Josephson critical current in ballistic superconducting microbridges. Near the critical temperature $T_{c}$, the Eilenberger equations have been solved selfconsistently. The closed integral equation for the order parameter $\Delta $ (\ref{eqdelta}) and the formula for the critical current $I_{c}$ (\ref{Ic}) are derived. Equations (\ref{eqdelta}), (\ref{Ic}) are valid for the arbitrary microbridge length $L$ in the scale of the coherence length $\xi_{0}\sim v_{F}/T_{c}$. In strongly inhomogeneous microcontact geometry they replace the differential Ginzburg-Landau equations and can be solved numerically. In the limiting cases $L\gg \xi _{0}$ and $L\ll \xi _{0}$ the analytical expressions for $% \Delta $ inside the weak link and for the $I_{c}\left( L\right) $ are obtained. In Figure 3 the dependence of $I_{c}\left( L\right) $ on $L$ is shown schematically. For long microbridge, $L\gg \xi _{0}$, the critical current $\sim 1/L$ is in the correspondence with the phenomenological consideration. The main interest presents the region $L\lesssim \xi _{0}$, where the microscopic theory is needed. We have calculated the corrections to the KO theory (\cite{KO}), which are connected with the finite value of the contact size. The expression (2) for the Josephson current was obtained in Ref.(\cite{KO}) in zero approximation on the contact size. For the $L\ll \xi _{0}$ we obtained that $\delta I_{c}/I_{0}\sim -\frac{L}{\xi _{0}}\ln \frac{\xi _{0}}{L}$ , where $I_{0}$ is the value of the critical current in KO theory. Thus, the corrections to the value $I_{0}$ are small for $L\ll \xi _{0}$, but the derivative $dI_{c}/dL$ has the singularity at $L=0$. This singularity is smeared, if we take into account the finite value of the ratio $T_{c}/\omega _{D}$. For ultra short microchannel, $L\lesssim a_{D}\sim v_{F}/\omega _{D}$ (the dashed region in the Fig.3), the length dependence of the critical current becomes $\delta I_{c}/I_{0}\sim -\frac{L}{% \lambda \xi _{0}}$ ($\lambda $ is the constant of electron-phonon coupling). In the very small microcontacts we have the unique situation, when the disturbance of the superconducting order parameter can be localized on the length $a_{D}$, making essential the effects of retardation of electron-phonon interaction. The ballistic flight of electrons through the channel is dynamical process with the characteristic frequency $\omega _{0}\sim v_{F}/L$. For $L$ smaller then $a_{D}$ this frequency becomes comparable with the Debye frequency $\omega _{D}$. Thus, the critical current $I_{c}$ for the finite contact's size is smaller then $I_{0}$. At the same time the normal state resistance $R_{N}$ of the ballistic microchannel does not depend of the length $L$ and remains equal to the Sharvin resistance $R_{0}$ (\ref{RSh}). As the result, the value of the product $I_{c}R_{N}$ is not equal $\pi \Delta _{0}^{2}/4eT_{c}$ and depends on the contact size. We have considered here the case of the quasiclassical situation, $L\gg \hbar /p_{F}$. In quantum regime, $L\sim \hbar /p_{F}$, the Sharvin resistance $R_{0}$ in formula (2) is substituted by quantized resistance of the contact, as was firstly shown by Beenakker and Houten \cite {Bee}. It follows from our consideration that for such small microcontacts with $L\lesssim a_{D}$ the rigorous calculation of the Josephson current requires to taking into account the retardation effects. \section{Acknowledgments} Authors are grateful to M. R. H. Khajehpour for useful discussions. This work has been partly supported by the Institute for Advanced Studies in Basic Sciences at Zanjan, IRAN. \newpage
\section{Slater's proof} In classical hypergeometric theory, formulae concerning summation, transformation and contiguous relations are provided in order to manipulate expressions that represent special functions. One of those is Kummer's non-terminating summation theorem: \begin{equation} \label{e:kummerth} {}_2F_1\left[{a,b \atop 1+a-b};-1\right] = \frac{ \Gamma(1+\frac{a}{2}) \: \Gamma(1+a-b) } { \Gamma(1+a) \: \Gamma(1+\frac{a}{2}-b) } \ \mbox{,} \end{equation} in which $ Re(b) < 1 $ ensures the series to be convergent. The first step in Slater's proof is to prove the following quadratic transformation, due to Kummer: \begin{equation} \label{e:kummerqua} {}_2F_1\left[{a,b \atop 1+a-b};z\right] = (1-z)^{-a} \: {}_2F_1\left[{\frac{a}{2},\frac{1}{2}+\frac{a}{2}-b \atop 1+a-b};\frac{-4\:z}{(1-z)^2}\right] \ \mbox{.} \end{equation} The proof of this transformation is as follows.\\ The right-hand side of (\ref{e:kummerqua}) can be written: \begin{eqnarray*} \displaystyle & & (1-z)^{-a} \: \sum_{k=0}^\infty \frac{(\frac{a}{2})_k \: (\frac{1}{2}+\frac{a}{2}-b)_k} {(1+a-b)_k \: k!} \: \frac{(-4)^k \: z^k}{(1-z)^{2\:k}}\\ &=& \sum_{k=0}^\infty \frac{(\frac{a}{2})_k \: (\frac{1}{2}+\frac{a}{2}-b)_k \: (-4)^k} {(1+a-b)_k \: k!} \: z^k \: (1-z)^{-a-2\:k} \ \mbox{,}\\ &=& \sum_{k=0}^\infty \frac{(\frac{a}{2})_k \: (\frac{1}{2}+\frac{a}{2}-b)_k \: (-4)^k} {(1+a-b)_k \: k!} \: z^k \: {}_1F_0\left[{a+2\:k \atop -};z\right] \quad \mbox{by the binomial theorem,}\\ &=& \sum_{k=0}^\infty \sum_{n=0}^\infty \frac{(\frac{a}{2})_k \: (\frac{1}{2}+\frac{a}{2}-b)_k \: (-4)^k} {(1+a-b)_k \: k!} \: \frac{(a+2\:k)_n}{n!} \: z^{k+n} \ \mbox{.} \end{eqnarray*} Extracting the coefficient of $z^N$ yields: \begin{eqnarray} & & \sum_{k=0}^{N} \frac{(\frac{a}{2})_k \: (\frac{1}{2}+\frac{a}{2}-b)_k \: (-4)^k \: (a+2\:k)_{N-k}} {(1+a-b)_k \: (N-k)! \: k!} \nonumber \\ \qquad & = & \frac{(a)_N}{N!} \: \sum_{k=0}^{N} \frac{(\frac{1}{2}+\frac{a}{2}-b)_k \: (a+N)_k \: (-N)_k} {(1+a-b)_k \: (\frac{1}{2}+\frac{a}{2})_k \: k!} \label{e:beforesaal} \end{eqnarray} since: \[ \frac{1}{(N-k)!} = \frac{(-1)^k \: (-N)_k}{N!} \] and \[ (a+2\:k)_{N-k} = \frac{(a)_{N+k}}{(a)_{2\:k}} = \frac{(a)_N \: (a+N)_k}{4^k \: (\frac{a}{2})_k \: (\frac{1}{2}+\frac{a}{2})_k} \ \mbox{.} \] The right-hand side of (\ref{e:beforesaal}) is summable by Saalsch{\" u}tz theorem \cite[(III.2)]{b-slater}: \[ \frac{(a)_N}{N!} \: \frac{(\frac{1}{2}+\frac{a}{2})_N \: (1-b-N)_N} {(1+a-b)_N \: (\frac{1}{2}-\frac{a}{2}-N)_N} \ \mbox{.} \] As this is the coefficient of $z^N$ in the right-hand side of (\ref{e:kummerqua}), we obtain: \begin{eqnarray*} \sum_{N=0}^{\infty} \frac{(a)_N}{N!} \: \frac{(\frac{1}{2}+\frac{a}{2})_N \: (1-b-N)_N} {(1+a-b)_N \: (\frac{1}{2}-\frac{a}{2}-N)_N} \: z^N & = & \sum_{N=0}^{\infty} \frac{(a)_N \: (b)_N}{(1+a-b)_N \: N!} \: z^N \\ & = & {}_2F_1\left[{a,b \atop 1+a-b};z\right] \ \mbox{,} \end{eqnarray*} which proves Kummer's quadratic transformation.\\ The second step is to set $z$ to $-1$ in (\ref{e:kummerqua}). Therefore \begin{equation} \label{e:kummerquaspe} {}_2F_1\left[{a,b \atop 1+a-b};-1\right] = 2^{-a} \: {}_2F_1\left[{\frac{a}{2},\frac{1}{2}+\frac{a}{2}-b \atop 1+a-b};1\right] \ \mbox{.} \end{equation} Note that these series are convergent if $ Re(b) < 1 $.\\ We can sum the series on the right-hand side by Gauss's theorem \cite[(III.3)]{b-slater}. This yields \begin{equation} {}_2F_1\left[{a,b \atop 1+a-b};-1\right] = 2^{-a} \: \frac{\Gamma(1+a-b) \: \Gamma(\frac{1}{2})} {\Gamma(1+\frac{a}{2}-b) \: \Gamma(\frac{1}{2}+\frac{a}{2})} \ \mbox{.} \end{equation} Since $ \Gamma(\frac{1}{2}+\frac{a}{2}) \: \Gamma(\frac{a}{2}+1) = 2^{a} \: \Gamma(\frac{1}{2}) \: \Gamma(a+1) $, we finally obtain Kummer's theorem: \begin{equation} {}_2F_1\left[{a,b \atop 1+a-b};-1\right] = \frac{ \Gamma(1+a-b) \: \Gamma(1+\frac{a}{2})} { \Gamma(1+\frac{a}{2}-b) \: \Gamma(1+a)} \ \mbox{.} \end{equation} \section{A new proof} In formula (\ref{e:kummerth}), let us replace $a$ by $a+2\:n$ where $n$ is a new variable that stands for a nonnegative integer. The resulting formula \begin{equation} \label{e:kummervar} {}_2F_1\left[{a+2\:n,b \atop 1+a+2\:n-b};-1\right] = \frac{ \Gamma(1+\frac{a}{2}+n) \: \Gamma(1+a+2\:n-b) } { \Gamma(1+a+2\:n) \: \Gamma(1+\frac{a}{2}+n-b) } \end{equation} is clearly equivalent to (\ref{e:kummerth}).\\ The introduction of the free parameter $n$ is the key to this new computer-assisted proof. Formula (\ref{e:kummervar}) can be written as: \begin{equation} \label{e:p1} \frac{\sum_k f(n,k)}{S(n)} = 1 \ \mbox{,} \end{equation} where \[ \displaystyle f(n,k) = \frac{(a+2\:n)_k \: (b)_k}{(1+a+2\:n-b)_k} \: \frac{(-1)^k}{k!} \] and \[ \displaystyle S(n) = \frac{\Gamma(1+\frac{a}{2}+n) \: \Gamma(1+a+2\:n-b)} {\Gamma(1+a+2\:n) \: \Gamma(1+\frac{a}{2}+n-b)} \ \mbox{.} \] (The coefficient $2$ in the substitution $a \rightarrow a+2\:n$ makes $S(n)$ hypergeometric in $n$.)\\ Formula (\ref{e:p1}) can be rewritten as: \begin{equation} \label{e:p2} \sum_k F(n,k) = 1 \qquad \mbox{where } F(n,k) = \frac{f(n,k)}{S(n)} \mbox{.} \end{equation} We shall actually prove this last formula. Using Gauthier's Maple package \texttt{HYPERG} \cite{b-user}, we apply Zeilberger's algorithm to obtain: \begin{equation} \label{e:kummerWZ} F(n,k) - F(n+1,k) = G(n,k+1) - G(n,k) \ \mbox{,} \end{equation} with $ G(n,k) = F(n,k) \: C(n,k) $ and $ \displaystyle C(n,k) = -\frac{(b-1)\:k}{(1+a+2\:n-b+k) \: (a+2\:n)} $.\\ This $C(n,k)$ is the so-called certificate of the WZ-pair $(F,G)$.\\ After summing (\ref{e:kummerWZ}) for $ k $ from $ 0 $ to $K$, the right-hand side telescopes: \begin{eqnarray} \sum_{k=0}^{K} F(n,k) - \sum_{k=0}^{K} F(n+1,k) & = & \sum_{k=0}^{K} G(n,k+1) - \sum_{k=0}^{K} G(n,k) \\ & = & G(n,K+1) - G(n,0) \\ \label{e:WZsum} & = & F(n,K+1) \: C(n,K+1) \ \mbox{.} \end{eqnarray} We now let $K$ tend to infinity. Clearly, \[ \lim_{K \rightarrow +\infty} C(n,K+1) = -\frac{b-1}{a+2\:n} \ \mbox{.} \] To determine $ \displaystyle \lim_{K\rightarrow+\infty} F(n,K+1) $, we use the well-known estimate (see e.g. \cite{b-tom}): \begin{equation} \label{e:lim} \frac{\Gamma(a+k)}{\Gamma(b+k)} \sim k^{a-b} \quad \mbox{as } k \rightarrow \infty \ \mbox{.} \end{equation} First, \begin{eqnarray*} F(n,k) & = & \frac{(a+2\:n)_k \: (b)_k}{(1+a+2\:n-b)_k} \: \frac{(-1)^k}{k!} \: \frac{1}{S(n)} \ \mbox{,} \\ & = & \frac{\Gamma(a+2\:n+k) \: \Gamma(b+k) \: (-1)^k} {\Gamma(1+a+2\:n-b+k) \: \Gamma(k+1)} \: T(n) \ \mbox{,} \end{eqnarray*} where $\displaystyle T(n)=\frac{(a+2\:n)\:\Gamma(1+\frac{a}{2}+n-b)} {\Gamma(b)\:\Gamma(1+\frac{a}{2}+n)}$ does not depend on $k$.\\ Then, \begin{eqnarray*} |F(n,k)| & \sim & k^{a+2n-(1+a+2n-b)} \: k^{b-1} \: T(n) \ \mbox{,} \\ & \sim & k^{2b-2} \: T(n) \ \mbox{.} \end{eqnarray*} We know that $ Re(b)<1 $ (necessary condition for the convergence of the series in (\ref{e:kummerth})).\\ Therefore \[ \lim_{K \rightarrow +\infty} F(n,K+1) = 0 \ \mbox{.} \] Expressing as before $F(n,k)$ in terms of Gamma functions and using the same aymptotics, it is easy to find the limit of $F(n,k)$ when $n$ tends to infinity: \begin{eqnarray*} F(n,k) & = & \frac{(a+2\:n)_k \: (b)_k}{(1+a+2\:n-b)_k} \: \frac{(-1)^k}{k!} \: \frac{\Gamma(1+a+2\:n) \: \Gamma(1+\frac{a}{2}+n-b)} {\Gamma(1+\frac{a}{2}+n) \: \Gamma(1+a+2\:n-b)} \\ & = & \frac{\Gamma(a+2\:n+k) \: \Gamma(1+a+2\:n-b)} {\Gamma(a+2\:n)\:\Gamma(1+a+2\:n-b+k)} \: \frac{(b)_k \: (-1)^k}{k!} \\ & & \times \frac{\Gamma(1+a+2\:n) \: \Gamma(1+\frac{a}{2}+n-b)} {\Gamma(1+a+2\:n-b) \: \Gamma(1+\frac{a}{2}+n)} \\ |F(n,k)| &\sim& \left| (2\:n)^{k} \: (2\:n)^{-k} \: \frac{(b)_k \: (-1)^k}{k!} \: (2\:n)^b \: n^{-b} \right| \ \mbox{.} \end{eqnarray*} So, \begin{equation} \lim_{n \rightarrow +\infty} F(n,k) = \frac{(b)_k \: (-1)^k}{k!} \: 2^b \ \mbox{.} \end{equation} By (\ref{e:lim}), we have $ 1/S(n) \sim 2^b $ when $n$ tends to infinity, so for all $n \geq n_0$, $1/S(n) < A$ where $A$ is a positive constant. Therefore, \begin{equation} \label{e:convS} |F(n,k)| = \left|\frac{f(n,k)}{S(n)}\right| \leq A \: |f(n,k)| \ \mbox{.} \end{equation} Moreover, \begin{equation} \label{e:majorF1} |f(n,k)| = \frac{ |(a+2\:n)_k| \: |(b)_k| }{ |(1+a+2\:n-b)_k| \: k! } \leq \frac{ |(b)_k| }{ k! } \ \mbox{,} \end{equation} because \[ \forall i \geq 0 \ \mbox{,} \qquad \frac{|a+2\:n+i|}{|1+a+2\:n-b+i|} < 1 \quad \mbox{when } n \mbox{ is big enough,} \] and \begin{eqnarray*} & & \frac{ |(a+2\:n)_k| }{ |(1+a+2\:n-b)_k| } \\[0.2cm] & = & \frac{ | (a+2\:n) \: (a+2\:n+1) \ldots (a+2\:n+k-1) |} { | (1+a+2\:n-b) \: (1+a+2\:n-b+1) \ldots (1+a+2\:n+b+k-1) |} \ \mbox{,} \\[0.2cm] & \leq & \frac{ |a+2\:n| \: |a+2\:n+1| \ldots |a+2\:n+k-1| } { |1+a+2\:n-b| \: |1+a+2\:n-b+1| \ldots |1+a+2\:n+b+k-1| } \ \mbox{,} \\[0.2cm] & \leq & 1 \ \mbox{.} \end{eqnarray*} We can finally establish (use (\ref{e:convS}), (\ref{e:majorF1}) and (\ref{e:lim})) that \begin{equation} \label{e:majorF2} |F(n,k)| \leq A \: \left| \frac{(b)_k}{k!} \right| \sim A \: \left| \frac{k^{b-1}}{\Gamma(b)} \right| \mbox{.} \end{equation} For $Re(b)<0$, this last expression is the general term of a convergent series, so $\sum_{k \geq 0} F(n,k) $ is dominated by a convergent series. The limit of formula (\ref{e:WZsum}) when $K$ tends to infinity gives \[ \sum_{k=0}^{\infty} F(n,k) - \sum_{k=0}^{\infty} F(n+1,k) = 0 \ \mbox{.} \] Hence, the sum $\sum_{k=0}^{\infty} F(n,k) $ is independent of $n$. Then \begin{eqnarray*} \sum_{k=0}^{\infty} F(n,k) & = & \lim_{n \rightarrow +\infty} \sum_{k=0}^{\infty} F(n,k) \ \mbox{,} \\ & = & \sum_{k=0}^{\infty} \lim_{n \rightarrow +\infty} F(n,k) \ \mbox{,} \end{eqnarray*} where the last equality is justified by dominated convergence. Moreover, \begin{eqnarray*} \sum_{k=0}^{\infty} F(n,k) & = & \sum_{k=0}^{\infty} \lim_{n \rightarrow +\infty} (F(n,k) \ \mbox{,} \\ & = & \sum_{k=0}^{\infty} 2^b \: \frac{(b)_k \: (-1)^k}{(k)!} \\ & = & 2^b \: \sum_{k=0}^{\infty} \: \frac{(b)_{k} \: (-1)^k}{k!} \ \mbox{,} \\ & = & 2^b \: (1-(-1))^b = 1 \ \mbox{,} \qquad \mbox{by binomial theorem.} \end{eqnarray*} This is precisely formula (\ref{e:p2}). So, we proved ---~with the restriction $ Re(b)<0 $~--- the validity of Kummer's theorem (specializing $n$ by $0$ in (\ref{e:kummervar})). Kummer's theorem (\ref{e:kummerth}) can be rewritten as: \[ \sum_k M(a,b,k) = \frac{ \Gamma(1+\frac{a}{2}) \: \Gamma(1+a-b) } { \Gamma(1+a) \: \Gamma(1+\frac{a}{2}-b) } \ \mbox{,} \] with $\displaystyle M(a,b,k) = \frac{(a)_k \: (b)_k}{(1+a-b)_k \: k!} \: (-1)^k $.\\ We derive by Zeilberger's algorithm, a recurrence for $M$ with respect to the parameter $b$: \[ (a-2\:b) \: M(a,b,k) + (-2\:a+2\:b) \: M(a,b+1,k) = G'(a,b,k+1) - G'(a,b,k) \] with $ G'(a,b,k) = M(a,b,k) \: \frac{(a-b+k)\:k}{b} $ and $ Re(b)<0 $. After summing both sides with respect to $k$ from $0$ to $K-1$, it gives \[ (a-2\:b) \: \sum_{k=0}^{K-1} M(a,b,k) + (-2\:a+2\:b) \: \sum_{k=0}^{K-1} M(a,b+1,k) = G'(a,b,K) - G'(a,b,0) \ \mbox{,} \] and then we let $K$ tend to infinity to obtain \[ (a-2\:b) \: \sum_{k=0}^{\infty} M(a,b,k) + (-2\:a+2\:b) \: \sum_{k=0}^{\infty} M(a,b+1,k) = 0 \ \mbox{,} \] because $ G'(a,b,0) = 0 $ and $ \displaystyle \lim_{K \rightarrow +\infty} G'(a,b,K) = 0 $. We have now for any $b$ with $Re(b)<0$ \begin{eqnarray*} \sum_{k=0}^{\infty} M(a,b+1,k) & = & \frac{\frac{a}{2}-b}{a-b} \: \sum_{k=0}^{\infty} M(a,b,k) \ \mbox{,} \\ & = & \frac{\frac{a}{2}-b}{a-b} \: \frac{ \Gamma(1+\frac{a}{2}) \: \Gamma(1+a-b) } { \Gamma(1+a) \: \Gamma(1+\frac{a}{2}-b) } \ \mbox{,} \\ & = & \frac{ \Gamma(1+\frac{a}{2}) \: \Gamma(a-b) } { \Gamma(1+a) \: \Gamma(\frac{a}{2}-b) } \ \mbox{.} \end{eqnarray*} Let us replace $b+1$ by $B$, we finally conclude that \[ {}_2F_1\left[{a,B \atop 1+a-B};-1\right] = \frac{ \Gamma(1+\frac{a}{2}) \: \Gamma(1-a-B) } { \Gamma(1+a) \: \Gamma(1+\frac{a}{2}-B) } \ \mbox{,} \] for all values of $B$ such that $Re(B)<1$. \textbf{Remark:} We can also derive this result directly from (\ref{e:majorF2}) by a convergence-acceleration argument. For this, we define a new sequence $ H(n,k) := F(n,2\:k) + F(n,2\:k+1) $, and we find that $\sum_{k \geq 0} H(n,k)$ is dominated by a convergent series. Hence, we have \begin{eqnarray*} \sum_{k=0}^{\infty} F(n,k) & = & \lim_{n \rightarrow +\infty} \sum_{k=0}^{\infty} F(n,k) \ \mbox{,} \\ & = & \lim_{n \rightarrow +\infty} \sum_{k=0}^{\infty} (F(n,2\:k)+F(n,2\:k+1)) \ \mbox{,} \\ & = & \lim_{n \rightarrow +\infty} \sum_{k=0}^{\infty} H(n,k) = \sum_{k=0}^{\infty} \lim_{n \rightarrow +\infty} H(n,k) \ \mbox{,} \\ & = & \sum_{k=0}^{\infty} \lim_{n \rightarrow +\infty} (F(n,2\:k)+F(n,2\:k+1)) \ \mbox{,} \\ & = & \sum_{k=0}^{\infty} 2^b \: \left( \frac{(b)_{2\:k} \: (-1)^{2\:k}}{(2\:k)!} + \frac{(b)_{2\:k+1} \: (-1)^{2\:k+1}}{(2\:k+1)!} \right) \\ & = & 2^b \: \sum_{k=0}^{\infty} \: \frac{(b)_{k} \: (-1)^k}{k!} \ \mbox{,} \\ & = & 1 \ \mbox{.} \end{eqnarray*} \section{Other summation theorems} The combination of Zeilberger's algorithm and asymptotic estimates suggests an approach to automatically prove identities involving hypergeometric series. It is well suited to summation theorems. As an example, we now give two apparently new proofs of classical hypergeometric identities then a list of summation theorems that the package \texttt{HYPERG} can prove automatically. \subsection{Bailey's theorem} \begin{equation} \label{e:baileyth} {}_2F_1\left[{a,1-a \atop b};\frac{1}{2}\right] = \frac{ \Gamma(\frac{b}{2}) \: \Gamma(\frac{1+b}{2}) } { \Gamma(\frac{a+b}{2}) \: \Gamma(\frac{1-a+b}{2}) } \ \mbox{,} \end{equation} Following the same scheme, let $ b \rightarrow b+2\:n $. Formula (\ref{e:baileyth}) can be written: \begin{equation} \label{e:bai1} \sum_k F(n,k) = 1 \ \mbox{,} \end{equation} where $ \displaystyle F(n,k) = \frac{f(n,k)}{S(n)} $, \\ with $ \displaystyle f(n,k) = \sum_k \frac{(a)_k \: (1-a)_k}{(b+2\:n)_k} \: \left(\frac{1}{2}\right)^k \quad \mbox{and} \quad S(n) = \frac{\Gamma(\frac{b}{2}+n) \: \Gamma(\frac{1}{2}+\frac{b}{2}+n)} {\Gamma(\frac{a+b}{2}+n) \: \Gamma(\frac{1-a+b}{2}+n)} \ \mbox{.} $\\ Zeilberger's algorithm applied to $F(n,k)$ succeeds to finding a WZ-pair: \begin{equation} \label{e:baileyWZ} F(n,k) - F(n+1,k) = G(n,k+1) - G(n,k) \ \mbox{,} \end{equation} where $ \displaystyle G(n,k) = F(n,k) \: \left(-2\:\frac{k}{b+2\:n+k}\right) $.\\ As with Kummer's theorem, we have: \begin{eqnarray} G(n,0) & = & 0 \ \mbox{,} \\ \lim_{k \rightarrow +\infty} G(n,k) & = & 0 \ \mbox{,} \end{eqnarray} and so, $ \displaystyle \sum_k F(n,k) $ is independent of $n$.\\ In this case, $ \displaystyle \lim_{n \rightarrow +\infty} F(n,k) = \delta_{k,0} $, so that: \begin{equation} \sum_{k=0}^{\infty} F(n,k) = \lim_{n \rightarrow +\infty} \sum_{k=0}^{\infty} F(n,k) = \sum_{k=0}^{\infty} \lim_{n \rightarrow +\infty} F(n,k) = \sum_{k=0}^{\infty} \delta_{k,0} = 1 \ \mbox{,} \end{equation} which proves Bailey's theorem. \subsection{Dixon's theorem} \begin{eqnarray} \label{e:dixonth} & & {}_3F_2\left[{a,b,c \atop 1+a-b,1+a-c};1\right] \nonumber \\ & = & \frac{ \Gamma(1+a-b) \: \Gamma(1+a-c) \: \Gamma(1+\frac{a}{2}) \: \Gamma(1+\frac{a}{2}-b-c) } { \Gamma(1+a) \: \Gamma(1+\frac{a}{2}-b) \: \Gamma(1+\frac{a}{2}-c) \: \Gamma(1+a-b-c) } \ \mbox{,} \end{eqnarray} where $ Re(2+a-2\:b-2\:c) > 0 $.\\ Once again, we replace $ a $ by $a+2\:n$. With the same notation (\ref{e:bai1}), we now have: \begin{eqnarray*} f(n,k) & = & \sum_k \frac{(a+2n)_k \: (b)_k \: (c)_k} {(1+a+2n-b)_k \: (1+a+2n-c)_k \: k!} \ \mbox{,} \\ \mbox{and} \quad S(n) & = & \frac{\Gamma(1+a+2n-b) \Gamma(1+a+2n-c) \Gamma(1+\frac{a}{2}+n) \Gamma(1+\frac{a}{2}+n-b-c)} {\Gamma(1+a+2n) \Gamma(1+\frac{a}{2}+n-b) \Gamma(1+\frac{a}{2}+n-c) \Gamma(1+a+2n-b-c)} \ \mbox{.} \end{eqnarray*} The WZ-pair given by Zeilberger's algorithm is \begin{equation} \label{e:dixonWZ} F(n,k) - F(n+1,k) = G(n,k+1) - G(n,k) \ \mbox{,} \end{equation} with $ \displaystyle G(n,k) = F(n,k) \: C(n,k) $, where the computed certificat $C(n,k)$ is: { \footnotesize \begin{verbatim} -(-2-8*a*n*c-a+4*b+4*c-2*n-2*n*b*c-2*a*b*k-2*a*k*c-a*b*c-8*a*n*b -4*n*k*c-4*n*b*k+a*b+2*n*b-2*a^2*c-8*n^2*c+a*c+2*n*c-2*b^2+3*a^2 +12*a*n+12*n^2+12*a^2*n-2*a^2*b+24*a*n^2+a*k^2-8*n^2*b+2*n*k^2 +2*a^3+16*n^3-6*c*b+3*k*a^2+3*k*a+6*k*n+12*k*n^2+12*k*a*n+2*b^2*c +2*b*c^2-2*c^2)*k/(2+a+2*n-2*b-2*c)/(1+a+2*n-c+k)/(1+a+2*n-b+k) /(a+2*n) \end{verbatim} } \noindent It is easy to establish the following limits: \begin{eqnarray*} G(n,0) & = & 0 \ \mbox{,} \\ \lim_{k \rightarrow +\infty} G(n,k) & = & \lim_{k \rightarrow +\infty} k^{2\:b+2\:c-2\:n-4-a} \: T(n) \\ & = & 0 \ \mbox{.} \end{eqnarray*} Again, $ \displaystyle \sum_k F(n,k) $ is independent of $n$.\\ Since $ \displaystyle \lim_{n \rightarrow +\infty} F(n,k) = \delta_{k,0} $, we conclude again that: \begin{equation} \sum_{k=0}^{\infty} F(n,k) = \lim_{n \rightarrow +\infty} \sum_{k=0}^{\infty} F(n,k) = \sum_{k=0}^{\infty} \lim_{n \rightarrow +\infty} F(n,k) = \sum_{k=0}^{\infty} \delta_{k,0} = 1 \ \mbox{.} \end{equation} \subsection{A list of theorems} These formulas can automatically be proved using the \texttt{HYPERG} package: \begin{itemize} \item Gauss's theorem \cite[(III.3)]{b-slater}: \[ {}_2F_1\left[{a,b \atop c};1\right] = \frac{ \Gamma(c) \: \Gamma(c-a-b) } { \Gamma(c-a) \: \Gamma(-b+c) } \ \mbox{.} \] \item another Dixon's theorem \cite[(III.10)]{b-slater}: \[ {}_4F_3\left[{a,1+\frac{a}{2},b,c \atop \frac{a}{2},1+a-b,1+a-c}; -1\right] = \frac{\Gamma(1+a-b) \: \Gamma(1+a-c)} {\Gamma(1+a) \: \Gamma(1+a-b-c)} \ \mbox{.} \] \item another Dixon's theorem \cite[(III.12)]{b-slater}: \begin{eqnarray*} & & {}_5F_4\left[{a,\frac{a}{2}+1,b,c,d \atop \frac{a}{2},1+a-b,1+a-c,1+a-d}; 1\right] \\ & = & \frac{\Gamma(1+a-b) \: \Gamma(1+a-c) \: \Gamma(1+a-d) \: \Gamma(1+a-b-c-d)} {\Gamma(a+1) \: \Gamma(1+a-b-c) \: \Gamma(1+a-b-d) \: \Gamma(1+a-c-d)} \ \mbox{.} \end{eqnarray*} \end{itemize}
\section{Introduction} Complete intersections in projective space have been studied extensively from many points of view. A natural generalisation is the study of complete intersections in Grassmannians. The first case that presents itself is the case of intersections with linear spaces. Indeed, there is an extensive literature on the simplest case, the Grassmannian of lines in the 3-space, where intersections are known as linear complexes and congruences of lines. L.\ Roth has studied the rationality of linear sections of Grassmannians of lines in general. If they are smooth and if the dimension of the intersection is greater than half the dimension of the Grassmannian, then they are rational. R.\ Donagi determined the cohomology and the intermediate Jacobian of some linear sections of Grassmannians of lines. \medskip In this paper we study the linear sections from the point of automorphism groups. Let $\mathbb{G}(1,N)$ be the Grassmann variety of lines in projective $N$-space, canonically embedded in $\mathbb{P}({\textstyle\bigwedge}^2\mathbb{C}^{N+1})$ and let $H^l$ be an $l$-codimensional linear subspace in this space. For general $H^l$ we determine the automorphism groups for $\mathbb{G}(1,N)\cap H$, $\mathbb{G}(1,N)\cap H^2$, $\mathbb{G}(1,4)\cap H^3$, and $\mathbb{G}(1,5)\cap H^3$. In the second case we find for example: \setcounter{section}{3} \setcounter{theorem}{4} \begin{theorem} For $N=2n-1\ge5$ the automorphism group of $\mathbb{G}(1,N)\cap H^2$ has $\mathrm{SL}(2,\mathbb{C})^n/\{1,-1\}$ as a normal subgroup and the quotient group is isomorphic to the permutation group $\mathrm{S}(3)$ for $n=3$, to $\mathbb{Z}/2\mathbb{Z} \times \mathbb{Z}/2\mathbb{Z}$ for $n=4$, and trivial otherwise. \end{theorem} \setcounter{section}{0} \setcounter{theorem}{0} \medskip We believe that apart from trivial cases these are the only general linear sections where automorphism groups of positive dimension appear. Extensive computer checks seem to confirm this. \medskip In particular we prove that the automorphism groups of $\mathbb{G}(1,2n)\cap H$, $\mathbb{G}(1,4)\cap H^2$, $\mathbb{G}(1,5)\cap H^2$, $\mathbb{G}(1,6)\cap H^2$, and $\mathbb{G}(1,4)\cap H^3$ are quasihomogenous -- those of $\mathbb{G}(1,2n-1)\cap H$ and $\mathbb{G}(1,3)\cap H^2$ are even homogenous -- whereas all others are not. \medskip As to our methods, in our proofs the rich geometry of the Grassmannian plays a decisive role. Otherwise, we mainly use well known tools like multilinear algebra, Lefschetz theorems, vanishing theorems etc. \medskip We are indebted to E.\ Opdam and A.\ Pasquale for useful remarks. The first author also thanks the Stieltjes Institut of Leiden University for financial support. \section{Preliminary} The Grassmannian $\mathbb{G}(1,N)$ of lines in $\mathbb{P}_N$ is embedded by way of the Pl\"ucker embedding into $\mathbb{P}(\bigwedge^2\mathbb{C}^{N+1})$ \[ \begin{array}{ccc} \mathbb{G}(1,N) & \longrightarrow & \mathbb{P}({\textstyle\bigwedge}^2\mathbb{C}^{N+1})\\[1ex] \mathrm{span}\,\{ v,w \} &\longmapsto & \mathbb{P}(v\wedge w). \end{array} \] We denote by $H^l$ an $l$-codimensional linear subspace of $\mathbb{P}(\bigwedge^2\mathbb{C}^{N+1})$. Roth \cite{R} examined the geometry of the general linear sections of the Grassmannians and found \begin{theorem} For a general $H^l$ with $0 \le l \le 1/2 \dim \mathbb{G}(1,N)=N-1 $ the intersection with the Grassmannians, $\mathbb{G}(1,N)\cap H^l$, is rational. \end{theorem} In this article we continue this study by describing the automorphism groups of these sections. As for the notation, given a subvariety $Y$ of a variety $X$ we define $\mathrm{Aut} (Y,X)$ to be the automorphisms of $X$ that induce automorphisms of $Y$, i.e. \[ \mathrm{Aut}(Y,X)=\{ \phi\in\mathrm{Aut}(X)\mid \phi(Y)\subseteq Y \}. \] Recall that the automorphism group of the Grassmannian itself is computed in two steps, see e.g. \cite[10.19]{H}. First one shows that all automorphisms are induced by automorphisms of $\mathbb{P}(\bigwedge^2\mathbb{C}^{N+1})$, i.e. \[ \mathrm{Aut} (\mathbb{G}(1,N))\cong \mathrm{Aut}(\mathbb{G}(1,N), \mathbb{P}({\textstyle \bigwedge}^2\mathbb{C}^{N+1})). \] Then one proves that for $N \not= 3$ the right hand side group is isomorphic to $\mathbb{P}\mathrm{GL}(N+1,\mathbb{C})$ via \[ \begin{array}{ccc} \mathbb{P}\mathrm{GL}(N+1,\mathbb{C}) & \longrightarrow & \mathrm{Aut}(\mathbb{G}(1,N),\mathbb{P}({\textstyle\bigwedge}^2\mathbb{C}^{N+1}))\\[1.5ex] \mathbb{P}(T) &\longmapsto & \left( \mathbb{P}(\sum v_i\wedge w_i)\mapsto \mathbb{P}(\sum Tv_i\wedge Tw_i)\right) . \end{array} \] For the linear sections of the Grassmannians we follow the same outline. The first step is the following theorem; the second step will be done separately for the different cases in the next sections. \begin{theorem} \label{autointersect} For a general linear subspace $H^l\subset \mathbb{P}(\bigwedge^2\mathbb{C}^{N+1})$ of codimension $l\le 2N-5$ \[ \mathrm{Aut}(\mathbb{G}(1,N)\cap H^l)=\mathrm{Aut}(\mathbb{G}(1,N),\mathbb{P}({\textstyle \bigwedge}^2\mathbb{C}^{N+1})) \cap \mathrm{Aut}(H^l,\mathbb{P}({\textstyle \bigwedge}^2\mathbb{C}^{N+1})). \] \end{theorem} \emph{Proof.} We will abbreviate $\mathbb{G}(1,N)$ by $\mathbb{G}$. The ``$\supseteq$'' inclusion is trivial. For the other one we prove first that all automorphisms of $\mathbb{G}\cap H^l$ are induced by automorphisms of $\mathbb{G}$. This follows immediately, once we show that all divisors of $\mathbb{G}\cap H^l$ are induced by divisors of $ \mathbb{P}(\bigwedge^2\mathbb{C}^{N+1})$, i.e. \[ \mathrm{Pic}(\mathbb{G}\cap H^l)=\mathrm{Pic}(\mathbb{G})=\mathbb{Z}\cdot H. \] To see this, note that by the Lefschetz hyperplane section theorem \[ \mathbb{Z}\cdot H=\mathrm{H}^2(\mathbb{G},\mathbb{Z})=\mathrm{H}^2(\mathbb{G}\cap H,\mathbb{Z})=\ldots=\mathrm{H}^2(\mathbb{G}\cap H^l,\mathbb{Z}) \] for $0\le l \le 2N-5$. From the exponential sequence \[ 0\rightarrow \mathbb{Z}_{\mathbb{G}\cap H^l}\rightarrow {\cal O}_{\mathbb{G}\cap H^l} \rightarrow {\cal O}_{\mathbb{G}\cap H^l}^* \rightarrow 0 \] we get as a part of the associated long exact sequence \[ \ldots \rightarrow \mathrm{H}^1(\mathbb{G}\cap H^l,{\cal O}) \rightarrow \mathrm{H}^1(\mathbb{G}\cap H^l,{\cal O}^*) \rightarrow \mathrm{H}^2(\mathbb{G}\cap H^l,\mathbb{Z})=\mathbb{Z}\cdot H \rightarrow 0 \] and therefore \[ \mathrm{Pic}(\mathbb{G}\cap H^l)=\mathrm{H}^1(\mathbb{G}\cap H^l,{\cal O}^*)=\mathbb{Z}\cdot H \] as soon as we know that $\mathrm{H}^1(\mathbb{G}\cap H^l,{\cal O})=0$. \medskip This is well known for $l=0$. For $l\ge 1$ we look at the restriction sequence \[ 0\rightarrow {\cal O}_{\mathbb{G}\cap H^{l-1}}(-H) \rightarrow {\cal O}_{\mathbb{G}\cap H^{l-1}} \rightarrow {\cal O}_{\mathbb{G}\cap H^l} \rightarrow 0 \] and take its associated long exact sequence \[ \begin{array}{l} \displaystyle \ldots \rightarrow \mathrm{H}^1(\mathbb{G}\cap H^{l-1},{\cal O}(-H)) \rightarrow \mathrm{H}^1(\mathbb{G}\cap H^{l-1},{\cal O}) \rightarrow \mathrm{H}^1(\mathbb{G}\cap H^l,{\cal O}) \rightarrow \\[1.5ex] \displaystyle \phantom{\ldots} \rightarrow\mathrm{H}^2(\mathbb{G}\cap H^{l-1},{\cal O}(-H)) \rightarrow \ldots \end{array} \] The right and left cohomology groups are trivial for $l\le 2N-4$ by Kodaira's vanishing theorem, so \[ 0=\mathrm{H}^1(\mathbb{G},{\cal O})=\mathrm{H}^1(\mathbb{G}\cap H,{\cal O})=\ldots= \mathrm{H}^1(\mathbb{G}\cap H^l,{\cal O}). \] \medskip Now that we know $\mathrm{Aut}(\mathbb{G}\cap H^l)\subseteq \mathrm{Aut}(\mathbb{P}(\bigwedge^2\mathbb{C}^{N+1}))$ it remains to show that these projective transformations of $\mathbb{P}(\bigwedge^2\mathbb{C}^{N+1})$ fix $H^l$. This will follow if we prove that $\mathbb{G}\cap H^l$ spans $H^l$, i.e. \[ \mathrm{h}^0(\mathbb{G}\cap H^l,{\cal O}(H))=\dim {\textstyle \bigwedge^2 } \mathbb{C}^{N+1}-l. \] This is known for $l=0$. For $l\ge 1$ we take the long exact sequence associated to the restriction sequence tensored by ${\cal O}(H)$ \[ \begin{array}{l} \displaystyle 0\rightarrow \mathrm{H}^0(\mathbb{G}\cap H^{l-1},{\cal O})=\mathbb{C} \rightarrow \mathrm{H}^0(\mathbb{G}\cap \mathrm{H}^{l-1},{\cal O}(H))\rightarrow \mathrm{H}^0(\mathbb{G}\cap \mathrm{H}^l,{\cal O}(H))\rightarrow \\[1.5ex] \displaystyle \phantom{0} \rightarrow \mathrm{H}^1(\mathbb{G}\cap \mathrm{H}^{l-1},{\cal O})=0. \end{array} \] Looking at the dimensions we get \[ \mathrm{h}^0(\mathbb{G}\cap H^l,{\cal O}(H))= \mathrm{h}^0(\mathbb{G}\cap H^{l-1},{\cal O}(H))-1, \] and the claim follows by induction. \hfill $\Box$ \bigskip It is tempting to assume that the groups $\mathrm{Aut}(\mathbb{G}(1,N),\mathbb{P}(\bigwedge^2\mathbb{C}^{N+1}))$ and $\mathrm{Aut}(H^l,\mathbb{P}(\bigwedge^2\mathbb{C}^{N+1}))$ in $\mathrm{Aut}(\mathbb{P}(\bigwedge^2\mathbb{C}^{N+1}))$ intersect transversally. Then the dimension of $\mathrm{Aut}(\mathbb{G}(1,N)\cap H^l)$ could be computed as \[ \begin{array}{c@{\;=\;}l} \dim \mathrm{Aut}(\mathbb{G}(1,N)\cap H^l) & \dim \mathrm{Aut} (\mathbb{G}(1,N))- \mathrm{codim}\, \mathrm{Aut}(H^l,\mathbb{P}(\bigwedge^2\mathbb{C}^{N+1}))\\[1ex] &(N+1)^2-1-l\left( { N+1 \choose 2}-l\right). \end{array} \] And we would find the following non-finite groups: \[\begin{array}{l@{\;=\;}l} \dim \mathrm{Aut}(\mathbb{G}(1,N)\cap H) & (N^2+3N+2)/2 \\[1ex] \dim \mathrm{Aut}(\mathbb{G}(1,N)\cap H^2)& N+4 \\[1ex] \dim \mathrm{Aut}(\mathbb{G}(1,4)\cap H^3)&3. \end{array}\] Unfortunately, the intersection is not always transversal. Our computation of the automorphism groups will show the following dimensions for $N\ge 4$: \[\begin{array}{l@{\;=\;}l} \dim \mathrm{Aut}(\mathbb{G}(1,N)\cap H) & (N^2+3N+2)/2 \\[1ex] \dim \mathrm{Aut}(\mathbb{G}(1,N)\cap H^2)& \left\{ \begin{array}{ll} N+4& \mbox{for\ }N\mbox{\ even}\\ 3(N+1)/2 & \mbox{for\ }N\mbox{\ odd} \end{array}\right. \\[2.5ex] \dim \mathrm{Aut}(\mathbb{G}(1,4)\cap H^3)&3\\[1ex] \dim \mathrm{Aut}(\mathbb{G}(1,5)\cap H^3)&1. \end{array}\] We conjecture that these are the only non-finite groups. For $N+2 \le l $ the canonical bundle $K={\cal O}(-N-1+l)$ is positive on $\mathbb{G}(1,N)\cap H^l$, and this conjecture can be proved by Serre's duality theorem and Kodaira's vanishing theorem: \[ \dim \mathrm{Aut}(\mathbb{G} \cap H^l) = h^0(\mathbb{G} \cap H^l, \Theta)= h^{2N-2-l}(\mathbb{G} \cap H^l, K \Omega^1)=0 \] A proof for the remaining cases $3\le l \le N+1$ seems difficult. By computer computations we verified the conjecture for $N\le 10$ and all $l$. \bigskip With this theorem our task of determining the automorphisms of $\mathbb{G}(1,N)\cap H^l$ has been immensely simplified. All we need to do is to find the projective transformations of $\mathrm{Aut}(\mathbb{G}(1,N),\mathbb{P}(\bigwedge^2\mathbb{C}^{N+1}))=\mathbb{P}\mathrm{GL}(N+1,\mathbb{C})$ such that their induced action on $\mathbb{P}(\bigwedge^2\mathbb{C}^{N+1})^*$ preserves $H^l$. To express this in algebraic terms we identify $(\bigwedge^2\mathbb{C}^{N+1})^*$ with $\bigwedge^2(\mathbb{C}^{N+1})^*$. If a particular basis of $\mathbb{C}^{N+1}$ is chosen, $\bigwedge^2(\mathbb{C}^{N+1})^*$ as antisymmetric forms on $\mathbb{C}^{N+1}$ can also be identified with the antisymmetric matrices of size $N+1$. In concrete terms, if $(e_0,\ldots,e_N)$ is a basis of $\mathbb{C}^{N+1}$ and $E_{ij}\in \mathrm{M}(N+1,\mathbb{C})$ the matrix, which has a $1$ in the position $(i,j)$ but is otherwise zero, then \[ \begin{array}{ccc} \left( {\textstyle \bigwedge^2}\mathbb{C}^{N+1} \right)^* & \longrightarrow & \mathrm{Antisym}(N+1,\mathbb{C})\\[2ex] \sum_{i,j} \lambda_{ij}(e_i\wedge e_j)^* &\longmapsto & {\textstyle \frac{1}{2}} \sum_{i,j} \lambda_{ij} (E_{ij}-E_{ji}). \end{array} \] In these terms a line $l=p\wedge q\in \mathbb{G}(1,N)$ is in the hyperplane $H\in \mathbb{P}(\bigwedge^2\mathbb{C}^{N+1})$ iff for a corresponding antisymmetric matrix $A\in \mathrm{Antisym}(N+1,\mathbb{C})$ with $\mathbb{P}(A)=H$ we have ${}^t\!p A q=0$. \medskip Further, the action of $\mathbb{P}\mathrm{GL}(N+1,\mathbb{C})$ on $\mathbb{P}(\bigwedge^2\mathbb{C}^{N+1})$, which was given for $\mathbb{P}(T)\in\mathbb{P}\mathrm{GL}(N+1,\mathbb{C})$ by \[ \begin{array}{ccc} \mathbb{P}\left( {\textstyle \bigwedge^2}\mathbb{C}^{N+1} \right) & \longrightarrow & \mathbb{P}\left( {\textstyle \bigwedge^2}\mathbb{C}^{N+1} \right) \\[2ex] \mathbb{P} ( \sum v_i\wedge w_i ) &\longmapsto & \mathbb{P} ( \sum Tv_i\wedge Tw_i ), \end{array} \] induces the following action on the dual space \[ \begin{array}{ccc} \mathbb{P}( \mathrm{Antisym}(N+1,\mathbb{C}) ) & \longrightarrow & \mathbb{P}( \mathrm{Antisym}(N+1,\mathbb{C}) ) \\[1ex] \mathbb{P}(A) &\longmapsto & \mathbb{P} ({}^tT^{-1} A T^{-1} ). \end{array} \] Hence an $l$-codimensional linear subspace $H^l\subseteq\mathbb{P}( \bigwedge^2\mathbb{C}^{N+1})$ which is dually given by $\mathbb{P}(\mathrm{span}\,\{ A_1,\ldots,A_l \})$ is preserved under $T$ iff every hyperplane containing $H^l$ is mapped to another hyperplane containing $H^l$, i.e. \[ \begin{array}{ll} {}^tT^{-1}\left(\sum \lambda_i A_i\right)T^{-1} \in \mathrm{span}\, \{ A_1,\ldots,A_l \} & \mbox{ for\ all\ } \lambda_i\in\mathbb{C}\\[2ex] \Longleftrightarrow {}^tT^{-1}A_iT^{-1} \in \mathrm{span}\, \{ A_1,\ldots,A_l \} & \mbox{ for\ } i=1\ldots l. \end{array} \] We conclude \medskip \begin{corollary} For $N\ge 4$, $0\le l \le 2N-5$ and a general $H^l\subset \mathbb{P}( \bigwedge^2\mathbb{C}^{N+1})$ given by $\mathbb{P}(\mathrm{span}\,\{ A_1,\ldots,A_l \}) \subset \mathbb{P}(\mathrm{Antisym}(N+1,\mathbb{C}))$ the automorphism group of $\,\mathbb{G}(1,N)\cap H^l$ is \[ \{\mathbb{P}(T)\in\mathbb{P}\mathrm{GL}(N+1,\mathbb{C}) \mid {}^t T^{-1} A_i T^{-1} \in \mathrm{span}\, \{ A_1,\dots,A_l\} \, \forall i \}. \] \end{corollary} \bigskip In the following sections we will compute the automorphism groups using this corollary. In the course of the computations we will use geometric arguments for which it is essential to know if a hyperplane $H\subset \mathbb{P}(\bigwedge^2\mathbb{C}^{N+1})$ is tangent to $\mathbb{G}(1,N)$ or not. We recall the basic facts together with their short proofs. \medskip \begin{proposition} \label{schubert} For any line $l_0\in\mathbb{G}(1,N)$ the Schubert cycle \[ \sigma:=\{ l\in \mathbb{G}(1,N)\mid l\cap l_0\not=\emptyset \} \subseteq \mathbb{G}(1,N) \] lies inside the tangent space $\mathbb{T}_{l_0}\mathbb{G}(1,N)\subseteq\mathbb{P}(\bigwedge^2\mathbb{C}^{N+1}) $ and spans it. \end{proposition} \emph{Proof.} Let $l\in\sigma$, $p\in l\cap l_0$, $q\in l_0\setminus \{p\}$ and $r\in l\setminus \{ p \}$ then \[ \begin{array}{ccc} \mathbb{C} & \longrightarrow & \mathbb{G}(1,N) \\[1ex] \lambda &\longmapsto & p\wedge (q+\lambda r) \end{array} \] is a line in $\sigma\subset\mathbb{G}(1,N) $ through $l_0$ and $l$. Therefore it is contained in the tangent space $\mathbb{T}_{l_0}\mathbb{G}(1,N)$, in particular $l\in\mathbb{T}_{l_0}\mathbb{G}(1,N)$. \smallskip We choose a basis $(e_0,\ldots,e_N)$ of $\mathbb{C}^{N+1}$ such that $l_0=\mathbb{P}(e_0\wedge e_1)$. The $2N-1$ points $\mathbb{P}(e_0\wedge e_1)$, $\mathbb{P}(e_0\wedge e_i)$, $\mathbb{P}(e_1\wedge e_i)$ for $i=2 \ldots N$ lie in $\sigma \subset \mathbb{T}_{l_0}\mathbb{G}(1,N)$ and are projectively independent, hence they span $\mathbb{T}_{l_0}\mathbb{G}(1,N)$. \hfill$\Box$ \bigskip \begin{corollary} \label{hyperplanetangency} Let $H=\mathbb{P}(A)\in\mathbb{P}(\bigwedge^2\mathbb{C}^{N+1})^*$ be a hyperplane and $l_0 \in \mathbb{G}(1,N)$ a line then \[ \mathbb{T}_{l_0}\mathbb{G}(1,N)\subseteq H \Longleftrightarrow l_0\subseteq \ker A. \] \end{corollary} \emph{Proof.} By the Proposition $\mathbb{T}_{l_0}\mathbb{G}(1,N)\subseteq H $ is equivalent to $\sigma \subseteq H$. If we use the same basis of $\mathbb{C}^{N+1}$ as in the proof of the proposition, this means that \[ \begin{array}{l} \mathbb{P}((\lambda e_0+\mu e_1)\wedge v)\in H \quad \mbox{for\ all\ } (\lambda\colon\mu)\in \mathbb{P}_1, \ v\in\mathbb{C}^{N+1}\\[1ex] \Longleftrightarrow {}^t\!(\lambda e_0+\mu e_1) A v =0 \quad \mbox{for\ all\ } (\lambda\colon\mu)\in \mathbb{P}_1, \ v\in\mathbb{C}^{N+1}\\[1ex] \Longleftrightarrow l_0 \subseteq \ker A. \end{array} \] \vspace{-4.1ex} \ \hfill$\Box$ \medskip \begin{corollary}\label{dualG} The dual variety $\mathbb{G}(1,N)^*\subset \mathbb{P}(\bigwedge^2\mathbb{C}^{N+1})^*$ of the Grassmannian variety $\mathbb{G}(1,N)$ consists of matrices of corank $\ge 2$ for $N$ odd resp. corank $\ge 3$ for $N$ even. \medskip For $N$ odd it is an irreducible hypersurface of degree $(N+1)/2$; for $N$ even it is a 3-codimensional subvariety. \end{corollary} \emph{Proof.} By the last corollary $H=\mathbb{P}(A)\in \mathbb{P}(\bigwedge^2\mathbb{C}^{N+1})^*$ is tangential to $\mathbb{G}(1,N)$ iff $\mathrm{corank}\, A\ge 2$. Recall that an antisymmetric matrix has even rank. So, for $N$ odd the matrix $A\in \mathrm{Antisym}(N+1,\mathbb{C})$ has corank $\ge 2$ iff $\det A=0$. But again since $A$ is antisymmetric, $\det A$ is the square of the irreducible Pfaffian polynomial $\mathrm{Pf} A$ \cite[5.2]{B}, which therefore defines $\mathbb{G}(1,N)^*$. \medskip For $N$ even $\mathrm{corank}\, A \ge 2$ is equivalent to $\mathrm{corank}\, A \ge 3$. We compute the dimension of $\mathbb{G}(1,N)^*$ following Mumford \cite{M} and find \[ \begin{array}{l} \begin{array}{c@{\;=\;}l} \dim\left( \begin{array}{l} \mbox{space\ of\ }A \mbox{\ with} \\ \dim \ker A =3 \end{array} \right) & \dim \mathrm{G}(3,N+1) + \dim \bigwedge^2 \mathbb{C}^{N+1}/\mathbb{C}^3\\[-0.4ex] & 3(N-2) +(N-2)(N-3)/2\\[1ex] & (N^2+N-6)/2 \end{array}\\[7ex] \Longrightarrow \mathrm{codim}\, \mathbb{G}(1,N)^*=(N+1)N/2-(N^2+N-6)/2=3. \end{array} \] \vspace{-4.3ex} \ \hfill $\Box$ \section{$\mathbb{G}(1,2n-1)\cap H$}\label{g12n-1hsection} Let the hyperplane $H\in \mathbb{P}(\bigwedge^2\mathbb{C}^{2n})$ be given by an element $A \in (\bigwedge^2\mathbb{C}^{2n})^*$, which we identify with its corresponding antisymmetric matrix. If $H$ is general, $A$ will be a matrix of full rank. This may be taken as the definition of a general $H$. We will assume from now on that $H$ is general. \medskip The line system $\mathbb{G}(1,2n-1)\cap H$ in $\mathbb{P}_{2n-1}$ does not lead to obvious special points in the $\mathbb{P}_{2n-1}$. Through every point $p\in \mathbb{P}_{2n-1}$ passes a $\mathbb{P}_{2n-3}$ of lines, namely \[ p\wedge q \in \mathbb{G}(1,2n-1) \ \mbox{with} \ q\in \ker {}^t\! pA. \] For $n\ge 3$ we can compute the automorphism group of $\mathbb{G}(1,2n-1)\cap H$ with the help of Theorem \ref{autointersect} and its Corollary. It consists of elements $\mathbb{P}(T)\in \mathbb{P} \mathrm{GL}(2n,\mathbb{C})=\mathrm{Aut}( \mathbb{G}(1,2n-1))$ such that $\mathbb{P}(T)$ as an element of $\mathbb{P}\mathrm{GL}(\bigwedge^2\mathbb{C}^{2n})$ preserves $H$, i.e. \[ {}^t T^{-1} A T^{-1} = \lambda A \ \ \mbox{\ for\ suitable\ } \lambda \in \mathbb{C}^*. \] We may choose coordinates on $\mathbb{P}_{2n-1}$ such that \[ A=\left( \begin{array}{cc} 0 & -\mathrm{E}_n \\ \mathrm{E}_n & 0 \end{array} \right). \] Then by definition \[ \mathrm{Sp}(2n,\mathbb{C})=\{ T\in \mathrm{GL}(2n,\mathbb{C}) \mid {}^t T^{-1}A T^{-1}=A \}, \] and we have an isomorphism \[ \begin{array}{c@{\;}c@{\;}c} \{ T\in \mathrm{GL}(2n,\mathbb{C}) \mid \exists\lambda_T\in \mathbb{C}^* : {}^t T^{-1} A T^{-1}=\lambda_T A\}/\mathbb{C}^* & \longrightarrow & \mathrm{Sp}(2n,\mathbb{C})/\{1,-1\} \\[1.7ex] \mathbb{C}^*\cdot T & \longmapsto & \pm \frac{1}{\sqrt{\lambda_T}} T. \end{array} \] Therefore we see \begin{proposition} The automorphism group of $\mathbb{G}(1,2n-1)\cap H$ for a general $H\subset \mathbb{P}(\bigwedge^2\mathbb{C}^{N+1})$ is $\mathrm{Sp}(2n,\mathbb{C})/\{1,-1\}$. Its action on $\mathbb{G}(1,2n-1)\cap H$ is homogeneous. \end{proposition} \emph{Proof.} The missing case of $\mathbb{G}(1,3)\cap H$ can be found in \cite[p. 278]{FH}. The transitivity of the action follows from Witt's theorem \cite[12.31]{Br}. \hfill $\Box$ \section{$\mathbb{G}(1,2n-1)\cap H^2$} A 2-codimensional linear subspace $L=H^2$ of $\mathbb{P}(\bigwedge^2\mathbb{C}^{2n})$ can be thought of as the pencil of hyperplanes containing it. So it gives a line $L^*=\mathbb{P}(\lambda A-\mu B)\subset \mathbb{P}(\bigwedge^2\mathbb{C}^{2n})^*$. We identify again $(\bigwedge^2\mathbb{C}^{2n})^*$ with the antisymmetric matrices of size $2n$. The line $L^*$ intersects the dual Grassmannian $\mathbb{G}(1,2n-1)^*$, which consists of antisymmetric matrices of rank $\le 2n-2$ and is a hypersurface of degree $n$ by Corollary \ref{dualG}, in at most $n$ points. For the moment a line $L^*$, and hence $L$, will be called general if it has $n$ points of intersection, $H_i=\mathbb{P}(\lambda_i A-\mu_i B)\in L^*$, $i=1\ldots n$, with the dual Grassmannian. These hyperplanes $H_i$ are tangent to the Grassmannian $\mathbb{G}(1,2n-1)$ at the points $l_i:= \ker (\lambda_i A-\mu_i B)\in\mathbb{G}(1,2n-1)$ by Corollary \ref{hyperplanetangency}. Therefore we get $n$ exceptional lines $l_1,\ldots,l_n$ in $\mathbb{P}_{2n-1}$. \medskip The intersection of the Grassmannian with its tangent hyperplane $H_i$ contains all lines that intersect $l_i$, because these lines are already contained in the intersection $\mathbb{G}(1,2n-1)\cap \mathbb{T}_{l_i}\mathbb{G}(1,2n-1)$ by Proposition \ref{schubert}. \medskip So, any line through a point $p\in l_i$ will be in the subspace $L\subset \mathbb{P}(\bigwedge^2\mathbb{C}^{2n})$ as soon as it is contained in any other hyperplane $H\in L^*\setminus\{H_i\}$. This gives one linear restriction to lines through $p$, so that there is at least a $\mathbb{P}_{2n-3}$ of lines through the points of the lines $l_i$. In contrast, through a general point of $\mathbb{P}_{2n-1}\setminus \bigcup l_i$ there is only a $\mathbb{P}_{2n-4}$ of lines. In fact, we have \begin{proposition}\label{linecharacter} The points of the lines $l_1,\ldots,l_n$ are characterized by the property that through each of them passes a $\mathbb{P}_{2n-3}$ of lines, i.e. \[ \left\{ p\in \mathbb{P}_{2n-1}\; \begin{array}{|l} \mbox{through\ } p \mbox{\ passes\ a\ }\mathbb{P}_{2n-3}\\ \mbox{of\ lines\ of\ } \mathbb{G}(1,2n-1)\cap L \end{array}\right\} =\bigcup l_i.\] Furthermore, the lines $l_1,\ldots,l_n$ span the whole $\mathbb{P}_{2n-1}$. \end{proposition} This may easily be seen if we write the pencil of hyperplanes $L^*$ in its normal form. \begin{proposition}[Donagi\cite{D}]\label{normalpencil} Given a pencil of hyperplanes $L^*=\mathbb{P}(\lambda A-\mu B)\subset \mathbb{P}(\bigwedge^2\mathbb{C}^{2n})^*$ such that the line $L^*$ intersects the Pfaffian hypersurface in $n$ different points. Then there is a basis of $\mathbb{C}^{2n}$ such that \[ A=\left( \begin{array}{ccc} J & & \raiseghost{-2ex}{-1ex}{{\emph{\LARGE 0}}}\\ &\ddots & \\ \raiseghost{2ex}{0ex}{{\emph{\LARGE 0}}} & &J \end{array} \right) \ \ \mbox{and} \ \ B=\left( \begin{array}{ccc} \lambda_1 J & & \raiseghost{-2ex}{-1ex}{{\emph{\LARGE 0}}}\\ &\ddots & \\ \raiseghost{2ex}{0ex}{{\emph{\LARGE 0}}} & &\lambda_n J \end{array} \right) \ \ \mbox{with} \ J=\left( \begin{array}{cc} 0 &-1 \\ 1 & 0 \end{array} \right)\!. \] The points $(\lambda_1\,\colon 1),\ldots,(\lambda_n\,\colon 1)\in \mathbb{P}_1\cong L^*$ are unique up to a projective transformation of $\mathbb{P}_1$. \end{proposition} \noindent\emph{Proof of Proposition \ref{linecharacter}.} The hyperplane $H_i=\mathbb{P}(\lambda_i A-\mu_i B)$ has, written as an antisymmetric matrix, the kernel $l_i=\mathrm{span}\,\{e_{2i-1},e_{2i}\}$ which means it is tangent to $\mathbb{G}(1,2n-1)$ at $l_i$. All lines of $\mathbb{G}(1,2n-1)\cap L$ through the point $p$ are given by $p\wedge q$ with ${}^t\!pAq={}^t\!pBq=0$. In order to have a $\mathbb{P}_{2n-2}$ of lines through $p$, the linear forms ${}^t\!pA$ and ${}^t\!pB$ must be linear dependent, i.e. there are $\lambda,\mu\in\mathbb{C}$ with \[ 0=\lambda {}^t\!pA -\mu {}^t\!pB={}^t\!p (\lambda A -\mu B). \] Therefore $p$ is in the kernel of a matrix of the pencil, but these kernels are the lines $l_1,\ldots,l_n$, so $p$ is contained in one of them. \hfill$\Box$ \bigskip Knowing the exceptional lines $l_1,\ldots,l_n$, one can immediately give some lines which are in the line system. \begin{proposition} Any line in $\mathbb{P}_{2n-1}$ which intersects two exceptional lines is an element of the line system $\mathbb{G}(1,2n-1)\cap L$. \medskip \noindent The exceptional lines themselves are not in the line system. \end{proposition} \emph{Proof.} If a line $l$ intersects $l_i$ and $l_j$, it lies -- as a point of the Grassmannian $\mathbb{G}(1,2n-1)$ -- in $H_i$ and $H_j$, hence in $L=H_i\cap H_j$. \medskip Assume that the exceptional line $l_1$ is an element of $\mathbb{G}(1,2n-1)$. By Proposition \ref{linecharacter} the lines through a point $p \in l_1$ sweep out a hyperplane. This hyperplane contains the line $l_1$ by assumption and the other exceptional lines $l_2,\ldots,l_n$ by the first part of this proposition. But this contradicts the second statement of Proposition \ref{linecharacter}. \hfill$\Box$ \bigskip \begin{remark} From Proposition \ref{normalpencil} we also see that any position of the $n$ points of the line $L^*$ is possible. In particular, we may call a line general if the position of the points is general in the sense needed below. \end{remark} \bigskip Using this geometric description we can determine the automorphisms of $\mathbb{G}(1,2n-1)\cap L$. For the moment we restrict ourselves to $n\ge 3$ in order to be able to use Theorem \ref{autointersect}. By this theorem and its Corollary we can view an automorphism of $\mathbb{G}(1,2n-1)\cap L$ as an element $\mathbb{P}(T)$ of $\mathbb{P}\mathrm{GL}(2n,\mathbb{C})$. To make the notation simpler, we will write only $T$ for $\mathbb{P}(T)$ if no confusion can result. Since the points of the exceptional lines are characterized by the property of Proposition \ref{linecharacter}, $T$ must map the union of the lines $l_i\subset \mathbb{P}_{2n-1}$ onto itself. Permutations of the lines may occur, but -- as we will presently see -- not all permutations are possible. \medskip If we view the automorphism $T$ as an element of $\mathrm{Aut} (L,\mathbb{P}({\textstyle \bigwedge\nolimits^2}\mathbb{C}^{2n}))$, it interchanges the hyperplanes containing $L$, i.e.\ it induces a projective transformation of the line $L^*\subset \mathbb{P}(\bigwedge\nolimits^2\mathbb{C}^{2n})$. Naturally, the transformation of $L^*$ must preserve the union of points of intersection of $L^*$ with the dual Grassmannian, which determine the lines $l_i$. Now, if a transformation of $\mathbb{P}_{2n-1}$ permutes the lines $l_i$, then the induced transformation of $L^*$ must permute the corresponding points of $L^*$ in the same way. \medskip Since not every permutation of four or more points on a line can be induced by a projective transformation, not all permutations are possible. In fact, if the points are in general position, we get the following subgroups of the permutation groups: \[ \begin{array}{c|c} n & \mbox{subgroup\ of\ } \mathrm{S}(n)\\[0.5ex] \hline 3 & \mathrm{S}(3)\\[0.5ex] 4 & \{ (1\,2\,3\,4),(2\,1\,4\,3),(3\,4\,1\,2),(4\,3\,2\,1)\}\cong \mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z} \\[0.5ex] \ge 5& \{\mathrm{id}\} \end{array} \] On the other hand, any permutation $\sigma\in\mathrm{S}(n)$ of the points on $L^*$ that is induced by a projective transformation $\phi$ of $L^*$ can be induced by an automorphism of $\mathbb{G}(1,2n-1)\cap L$. To see this, let us write $L^*$ in its normal form and define $T\in \mathrm{GL}(2n,\mathbb{C})$ as \[ T(e_{2i}):=e_{2\sigma(i)} \quad \mbox{and} \quad T(e_{2i-1}):=e_{2\sigma(i)-1}. \] This transformation permutes the lines in the prescribed way, and as an automorphism of $\mathbb{P}(\bigwedge^2\mathbb{C}^{2n})$ it fixes $L$ since the transformed line $L^*$ is \[ \begin{array}{l} {}^tT^{-1} \left( \lambda \left( \begin{array}{ccc} J & & \raiseghost{-2ex}{-1ex}{{\LARGE 0}}\\ &\ddots & \\ \raiseghost{2ex}{0ex}{{\LARGE 0}} & & J \end{array} \right) -\mu \left( \begin{array}{ccc} \lambda_1 J & & \raiseghost{-2ex}{-1ex}{{\LARGE 0}} \\ &\ddots & \\ \raiseghost{2ex}{0ex}{{\LARGE 0}} & &\lambda_n J \end{array} \right) \right)T^{-1} \\[6ex] = \lambda \left( \begin{array}{ccc} J & & \raiseghost{-2ex}{-1ex}{{\LARGE 0}}\\ &\ddots & \\ \raiseghost{2ex}{0ex}{{\LARGE 0}} & & J \end{array} \right) -\mu \left( \begin{array}{ccc} \lambda_{\sigma^{-1}(1)} J & & \raiseghost{-2ex}{-1ex}{{\LARGE 0}}\\ &\ddots & \\ \raiseghost{2ex}{0ex}{{\LARGE 0}} & &\lambda_{\sigma^{-1}(n)} J \end{array} \right). \end{array} \] Changing the parametrisation of the line by $\phi$ we get back the old parametrisation of the line $L^*$ by the definition of $\phi$. So this $T$ is an automorphism of $\mathbb{G}(1,2n-1)\cap L$ that induces the permutation of lines we started with. \bigskip Now we can restrict our attention to transformations that do not permute the lines since we can obtain every permutation by composing with one of the transformations from above. A transformation leaving all the lines individually fixed has the form \[ T= \left( \begin{array}{ccc} t_1 & & \raiseghost{-2ex}{-1ex}{{\LARGE 0}}\\ &\ddots & \\ \raiseghost{2ex}{0ex}{{\LARGE 0}} & &t_n \end{array} \right) \ \ \mbox{with\ }t_1,\ldots,t_n\in\mathrm{GL}(2,\mathbb{C}). \] This $T$ will fix the line system $\mathbb{G}(1,2n-1)\cap L$ in $\mathbb{P}_{2n-1}$ iff it preserves $L^*$, i.e.\ for all $\lambda,\mu\in\mathbb{C}$ there exists $\alpha,\beta\in\mathbb{C}$ such that \[ {}^tT^{-1}(\lambda A-\mu B)T^{-1}=\alpha A+\beta B. \] It is sufficient to check this for $(\lambda,\mu)=(1,0)$ and $(0,-1)$. Since \[ \begin{array}{c@{\,}c@{\,}c} {}^tT^{-1}AT^{-1} & = & \left( \begin{array}{ccc} \det t_1^{-1} J & & \raiseghost{-2ex}{-1ex}{{\LARGE 0}}\\ &\ddots & \\ \raiseghost{2ex}{0ex}{{\LARGE 0}} & &\det t_n^{-1} J \end{array} \right)\\[5.5ex] {}^tT^{-1}BT^{-1} & = & \left( \begin{array}{ccc} \lambda_1\det t_1^{-1} J & & \raiseghost{-2ex}{-1ex}{{\LARGE 0}}\\ &\ddots & \\ \raiseghost{2ex}{0ex}{{\LARGE 0}} & &\lambda_n\det t_n^{-1} J \end{array} \right) \end{array} \] this is equivalent to the question if there exist $\alpha,\beta,\gamma,\delta\in\mathbb{C}$ with \[ \begin{array}{c@{\;}c@{\;}c} (\det t_1^{-1},\ldots,\det t_n^{-1}) & = & \alpha(1,\ldots,1)+\beta(\lambda_1,\ldots,\lambda_n)\\[1ex] (\lambda_1 \det t_1^{-1},\ldots,\lambda_n \det t_n^{-1}) & = & \gamma(1,\ldots,1)+\delta(\lambda_1,\ldots,\lambda_n). \end{array} \] It follows \[ \begin{array}{l} -\gamma(1,\ldots,1)+(\alpha-\delta)(\lambda_1,\ldots,\lambda_n)+ \beta(\lambda_1^2,\ldots,\lambda_n^2)=0\\[1ex] \Longrightarrow \alpha =\delta,\ \beta=\gamma=0\\[1ex] \Longrightarrow \det t_1=\ldots=\det t_n. \end{array} \] We normalize by $\det t_1=1$, i.e.\ $t_1,\ldots,t_n\in \mathrm{SL}(2,\mathbb{C})$. Then only $T$ and $-T\in\mathrm{GL}(2n,\mathbb{C})$ give the same element in $\mathbb{P} \mathrm{GL}(2n,\mathbb{C})$. So that as a group the automorphisms of $\mathbb{G}(1,2n-1)\cap L$ that do not permute the exceptional lines are isomorphic to $\mathrm{SL}(2,\mathbb{C})^n/\{1,-1\}$. \medskip Altogether we get \begin{theorem}\label{autog12n-1h2} For $n\ge 3$ the automorphism group of the intersection of $\mathbb{G}(1,2n-1)$ with a general 2-codimensional linear subspace of $\mathbb{P}(\bigwedge^2\mathbb{C}^{2n})$ has $\mathrm{SL}(2,\mathbb{C})^n/\{1,-1\}$ as a normal subgroup and the quotient group is isomorphic to the permutation group $\mathrm{S}(3)$ for $n=3$, to $\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z}$ for $n=4$, and trivial otherwise. \medskip The automorphism group is isomorphic to the subgroup of $\mathbb{P}\mathrm{GL}(2n,\mathbb{C})$ that consists of the elements \[ P_\sigma \cdot \left( \begin{array}{ccc} t_1 & & \raiseghost{-2ex}{-1ex}{\emph{\LARGE 0}}\\ &\ddots & \\ \raiseghost{2ex}{0ex}{\emph{\LARGE 0}} & &t_n \end{array} \right) \ \ \mbox{with\ }t_1,\ldots,t_n\in\mathrm{SL}(2,\mathbb{C}) \] where $P_\sigma$ is the identity for $n\ge 5$ and otherwise defined by \[ \begin{array}{c@{\;}c@{\;}l} P_\sigma(e_{2i}) & = & e_{2\sigma(i)}\\ P_\sigma(e_{2i-1}) & = & e_{2\sigma(i)-1} \end{array} \] \[ \mbox{for} \ \sigma \in \left\{ \begin{array}{ll} \mathrm{S}(n) & \mbox{if\ }n=3\\[0.5ex] \{ (1\,2\,3\,4),(2\,1\,4\,3),(3\,4\,1\,2),(4\,3\,2\,1)\} & \mbox{if\ }n=4. \end{array}\right. \] \end{theorem} \bigskip For the sake of completeness we recall the classical case of $\mathbb{G}(1,3)\cap H^2$. \begin{remark} The automorphism group of $\mathbb{G}(1,3)\cap H^2$ is an extension of $\mathbb{Z}/2\mathbb{Z}$ by $\mathbb{P}\mathrm{GL}(2,\mathbb{C})\times\mathbb{P}\mathrm{GL}(2,\mathbb{C})$. It acts homogeneously on $\mathbb{G}(1,3)\cap H^2$. \end{remark} \emph{Proof.} The Grassmannian $\mathbb{G}(1,3)$ is a smooth quadric in $\mathbb{P}(\bigwedge^2\mathbb{C}^4)\cong\mathbb{P}_5$. Therefore $\mathbb{G}(1,3)\cap H^2$ is a smooth quadric in $\mathbb{P}_3$. Hence it is isomorphic to the Segre variety $\mathbb{P}_1\times\mathbb{P}_1$ in $\mathbb{P}_3$. The automorphism group of $\mathbb{P}_1\times\mathbb{P}_1$ is generated by $\mathbb{P}\mathrm{GL}(2,\mathbb{C})\times\mathbb{P}\mathrm{GL}(2,\mathbb{C})$ together with the automorphism that exchanges the $\mathbb{P}_1$s. All the automorphisms extend to $\mathbb{P}_3$. Obviously, the group acts transitively on $\mathbb{P}_1\times\mathbb{P}_1$. \hfill$\Box$ \bigskip For the rest of this section we consider the question if the action of the other automorphism groups is quasihomogeous on the corresponding line system, i.e.\ if there is an open orbit. \medskip This cannot be the case for $n\ge 7$ since then the dimension of the line system $\mathbb{G}(1,2n-1)\cap H^2$, $2(2n-2)-2=4n-6$, is larger than the dimension of the automorphism group, $3n$. \medskip For $n=3$ the action is quasihomogeneous. To see that one can adjust the $(\lambda_1,\lambda_2,\lambda_3)$ in the normal form of the line system to $(1,0,-1)$ by a projective transformation and compute the stabiliser of the line $(1\,\colon 0\,\colon 1\,\colon 0\,\colon 1\,\colon 0)\wedge(1\,\colon 1\,\colon 1\,\colon -2\,\colon 1\,\colon 1)$ by hand or computer and see that it is 3-dimensional. So the dimension of its orbit is $3\cdot 3-3=6$, which is just the dimension of the line system. \medskip For $n=4,5,6$ the group does not act quasihomogenously anymore. For this one computes again the dimension of the stabiliser of a general line. Since the group acts transitively on $\mathbb{P}_{2n-1}\setminus \bigcup L_i$, we may restrict our attention to lines through one of those points, e.g. $(1\,\colon 0\,\colon\, \ldots\,\colon 1\,\colon 0)$. Using a computer one sees that the stabilizer of a general line through this point has again dimension $3$. Hence the orbit has dimension $3n-3$, which is less then the dimension of the line system, $4n-6$. \section{$\mathbb{G}(1,5)\cap H^3$} Let $L=H^3\subset \mathbb{P}(\bigwedge^2\mathbb{C}^6)$ be a general 3-codimensional subspace. With our usual identification of $(\bigwedge^2\mathbb{C}^6)^*$ with the antisymmetric matrices $\mathrm{Antisym}(6,\mathbb{C})$ its dual plane $L^*=\mathbb{P}(\lambda A+\mu B +v C)\subset \mathbb{P}(\bigwedge^2\mathbb{C}^6)^*$ intersects the dual Grassmannian $\mathbb{G}(1,5)^*$, which consists of matrices of rank $\le 4$ and is a hypersurface of degree 3 by Corollary \ref{dualG}, in an irreducible cubic $C^*$. By Corollary \ref{hyperplanetangency} a point $(\lambda\,\colon \mu\,\colon \nu)\in C^*$ corresponds to the hyperplane $h_{(\lambda\,\colon \mu\,\colon \nu)}=\mathbb{P}(\lambda A+\mu B+\nu C)$ that is tangent to the Grassmannian at the point \[ l_{(\lambda:\mu:\nu)}:=\ker (\lambda A+\mu B+\nu C)\subset\mathbb{P}_5. \] In analogy to the former case we have \begin{lemma} \[ \left\{ p\in\mathbb{P}_5 \left| \begin{array}{l} \mbox{through}\ p\ \mbox{passes\ a\ }\mathbb{P}_2 \\ \mbox{of}\ \mbox{lines\ of}\ \mathbb{G}(1,5)\cap L \end{array} \right.\right\} =\bigcup_{(\lambda:\mu:\nu)\in C^*} l_{(\lambda:\mu:\nu)} \subset\mathbb{P}_5\] \end{lemma} \emph{Proof.} Since by definition the lines in $\mathbb{G}(1,5)\cap L$ that contain $p$ are $p\wedge q$ with ${}^t\!pAq={}^t\!pBq={}^t\!pCq=0$, we see that \[ \begin{array}{l} \mbox{through}\ p\ \mbox{passes\ at\ least\ a\ }\mathbb{P}_2\ \mbox{of\ lines\ of\ } \mathbb{G}(1,5)\cap L\\[0.5ex] \Longleftrightarrow {}^t\!pA, {}^t\!pB,{}^t\!pC \ \mbox{are\ linear\ dependent}\\[0.5ex] \Longleftrightarrow \exists {(\lambda\,\colon \mu\,\colon \nu)}\in\mathbb{P}_2 \ \mbox{with}\ {}^t\!p(\lambda A+\mu B+\nu C)=0\\[0.5ex] \Longleftrightarrow p\in\ker (\lambda A+\mu B+\nu C)=l_{(\lambda:\mu:\nu)}. \end{array} \] We also note that there cannot be a $\mathbb{P}_3$ of lines of $\mathbb{G}(1,5)\cap L$ through a point $p$. Because if there were one, then $\dim \mathrm{span}\,\{ {}^t\!p A,{}^t\!p B,{}^t\!pC\}=1$, i.e.\ there exist two points $(\lambda\,\colon \mu\,\colon \nu),(\lambda'\,\colon \mu'\,\colon \nu')\in\mathbb{P}_2$ with \[ {}^t\!p (\lambda A+\mu B+\nu C)={}^t\!p (\lambda' A+\mu' B+\nu' C)=0. \] It follows that all the matrices \[ (\alpha\lambda+\beta\lambda') A+(\alpha\mu+\beta\mu') B+(\alpha\nu+\beta\nu')C \ \ \ \mbox{for\ all}\ (\alpha\,\colon \beta)\in\mathbb{P}_1 \] have a non-trivial kernel. Hence the line $(\alpha\lambda+\beta\lambda'\,\colon \alpha\mu+\beta\mu'\,\colon \alpha\nu+\beta\nu')$ must lie in $L^*\cap \mathbb{G}(1,5)^*=C^*$. But this is a contradiction since the cubic $C^*$ is irreducible. \hfill$\Box$ \begin{proposition} The lines $l_{(\lambda:\mu:\nu)}\subset\mathbb{P}_5$ with $(\lambda\,\colon \mu\,\colon \nu)\in C^*$ do not intersect each other. \end{proposition} \emph{Proof.} Assume that the line $l_{(\lambda:\mu:\nu)}$ intersects the line $l_{(\lambda':\mu':\nu')}$ in the point $p$, i.e. \[ p\in\ker (\lambda A+\mu B+\nu C) \cap \ker (\lambda' A+\mu' B+\nu' C)\not= 0. \] Then \[ p\in \ker((\alpha\lambda+\beta\lambda') A+(\alpha\mu+\beta\mu') B +(\alpha\nu+\beta\nu')C)\not= 0 \ \ \ \mbox{for\ all}\ (\alpha\,\colon \beta)\in\mathbb{P}_1, \] and the line $(\alpha\lambda+\beta\lambda'\,\colon \alpha\mu+\beta\mu'\,\colon \alpha\nu+\beta\nu')$ must be contained in the irreducible cubic $C^*$, which is a contradiction. \hfill$\Box$ \bigskip Let us again derive a normal form: \begin{proposition}\label{normalcubic} For a general plane $L^*=\mathbb{P} (\lambda A+\mu B+\nu C)\subset\mathbb{P}(\bigwedge^2\mathbb{C}^6)^*$ there exists a choice of bases of $L^*$ and $\mathbb{C}^6$ such that \[ A= \left(\begin{array}{cccccc} 0 &\!\! -1 & & & &\raiseghost{-4ex}{-3.5ex}{\emph{\Huge 0}} \\ 1 & 0 & & & & \\ & & 0& \!\!-1 & & \\ & & 1& 0 & & \\ & & & &0 &0\\ \raiseghost{4ex}{1ex}{\emph{\Huge 0}} & & & &0 & 0 \end{array}\right) \ \ B= \left(\begin{array}{cccccc} 0 & 0 & & & &\raiseghost{-4ex}{-3.5ex}{\emph{\Huge 0}} \\ 0 & 0 & & & & \\ & & 0& \!\!-1 & & \\ & & 1& 0 & & \\ & & & &0 &\!\!-1\\ \raiseghost{4ex}{1ex}{\emph{\Huge 0}} & & & &1 & 0 \end{array}\right) \] \[ C= \left(\begin{array}{cc|cc|cc} 0 &0 & \!\!-\alpha &0 & \!\! -\gamma&0\\ 0 & 0 & 0&\!\!\!\!-\alpha & \!\! -\delta&\!\!\!\!-\gamma\\ \hline \alpha &0 & 0&\!\!\!\!-1 & \!\! -\beta &0\\ 0&\alpha & 1&0 &0&\!\!\!\!-\beta\\ \hline \gamma&\delta &\beta &0 & 0 &0 \\ 0&\gamma &0&\beta & 0 & 0 \end{array}\right). \] \end{proposition} \begin{remark} It is also possible to derive a more symmetric normal form where all three matrices look like $C$ only with the ${ 0\,-1 \atop 1\ \;0}$ block moved along the diagonal, but this is not more useful for our computations. \end{remark} \emph{Proof of proposition \ref{normalcubic}.} We may assume that the line $\mathbb{P}(\lambda A+\mu B)\subset L^*$ is a general line. By Proposition \ref{normalpencil} there exists a choice of coordinates (corresponding to $\lambda_1=0$, $\lambda_2=1$, $\lambda_3=\infty$) such that $A$ and $B$ are of the required form. Further, if we change the coordinates of $\mathbb{C} ^6$ by transformations of the type \[ T= \left(\begin{array}{ccc} t_1 & 0 & 0\\ 0 & t_2 & 0 \\ 0 & 0 & t_3 \end{array}\right) \ \ t_1,t_2,t_3\in \mathrm{SL}(2,\mathbb{C}), \] then $A$ and $B$ will stay the same by Theorem \ref{autog12n-1h2}. \medskip We write the matrix $C$ as \[ C=\left(\begin{array}{ccc} c_1 J &\!\! -^tC_{21} &\!\! -^tC_{31}\\[0.5ex] C_{21} & c_2J &\!\! -^t C_{32} \\[0.5ex] C_{31} & C_{32} & c_3 J \end{array}\right) \ \ \mbox{with}\ \ \begin{array}{l} J=\left(\begin{array}{cc}0&-1\\1&0\end{array}\right); \ c_1,c_2,c_3\in\mathbb{C}\\[2.5ex] C_{21},C_{31}, C_{32}\in \mathrm{M}(2,\mathbb{C}). \end{array} \] We may assume that $c_1=c_3=0$, $c_2=1$, otherwise we replace $C$ by the matrix $1/(c_2-c_1-c_3)(C-c_1 A-c_3B)$. This is possible since $c_2-c_1-c_3\not=0$, because $C$ is general. So C looks like \[ C=\left(\begin{array}{ccc} 0 &\!\! -^tC_{21} &\!\! -^tC_{31}\\[0.5ex] C_{21} & J &\!\! -^t C_{32} \\[0.5ex] C_{31} & C_{32} & 0 \end{array}\right). \] The generality of $C$ ensures that the matrices $C_{21}$ and $C_{32}$ are invertible, so \[ T=\left(\begin{array}{ccc} \frac{1}{\alpha} C_{21} & 0 &0\\[0.5ex] 0 & \mathrm{E}_2 & 0 \\[0.5ex] 0 & 0 &\frac{1}{\beta} {}^t\!C_{32} \end{array}\right) \ \ \mbox{with}\ \ \begin{array}{l} \alpha=\sqrt{\det C_{21}}\\[0.5ex] \beta=\sqrt{\det C_{32}} \end{array} \] is of the above mentioned type and transforms $C$ into \[ C':={}^tT^{-1}CT^{-1}=\left(\begin{array}{ccc} 0 &\!\! -\alpha \mathrm{E}_2 &\!\! -^t\overline{C}\\ \alpha\mathrm{E}_2 & J &\!\! -\beta \mathrm{E}_2 \\ \overline{C} &\beta \mathrm{E}_2 & 0 \end{array}\right) \ \ \mbox{with} \ \overline{C}:=\alpha\beta C_{32}^{-1}C_{31}C_{21}^{-1}. \] This matrix will be transformed under \[ T=\left(\begin{array}{ccc} t^{-1}\!\!\! & 0 &0\\ 0 & {}^t t & 0 \\ 0 & 0 & t^{-1}\!\!\! \end{array}\right) \ \ \mbox{with}\ t\in\mathrm{SL}(2,\mathbb{C}) \] into \[ {}^tT^{-1}C'T^{-1}=\left(\begin{array}{ccc} 0 & \!\!-\alpha \mathrm{E}_2 & \!\!-{}^t t ^t\overline{C}t\\[0.5ex] \alpha\mathrm{E}_2 & J & \!\!-\beta \mathrm{E}_2 \\[0.5ex] {}^t t\overline{C}t &\beta \mathrm{E}_2 & 0 \end{array}\right). \] So, all that remains to show is: Given a general matrix $\overline{C}\in\mathrm{M}(2,\mathbb{C})$ there is a matrix $t\in\mathrm{SL}(2,\mathbb{C})$ such that \[ {}^t t\overline{C}t= \left(\begin{array}{cc} \gamma & \delta\\ 0 & \gamma \end{array}\right). \] If \[ \overline{C}=\left(\begin{array}{cc} c_{11} & c_{12} \\[0.5ex] c_{21} & c_{22} \end{array}\right) \ \ \mbox{and} \ \ t= \left(\begin{array}{cc} 1 &\!\! -\frac{c_{21}}{c_{11}}\\[0.5ex] 0 & 1 \end{array}\right) \] then \[ \overline{C}'={}^t t\overline{C}t= \left(\begin{array}{cc} c_{11} & c_{12} -c_{21}\\[0.5ex] 0 & \frac{\det \overline{C}}{c_{11}} \end{array}\right) \] and an additional transformation by \[ t= \left(\begin{array}{cc} \frac{\sqrt[4]{\det \overline{C}}}{\sqrt{c_{11}}} & 0\\[0.5ex] 0 & \frac{\sqrt{c_{11}}}{\sqrt[4]{\det \overline{C}}} \end{array}\right) \] takes $\overline{C}$ into the desired form \bigskip \ \hfill $\displaystyle {}^t t\overline{C}'t= \left(\begin{array}{cc} \sqrt{\det \overline{C}} & c_{12} -c_{21} \\[0.5ex] 0 & \sqrt{\det \overline{C}} \end{array}\right) . $ \hfill$\Box$ \bigskip \begin{remark} In terms of this coordinates the cubic $C^*\subset L^*$ is given as \[ \lambda^2\mu+\mu^2\lambda+\lambda\mu\nu-(\gamma^2+\beta^2)\lambda\nu^2 -(\alpha^2+\gamma^2)\mu\nu^2+(\alpha\beta\delta -\gamma^2)\nu^3. \] One checks that the cubic is smooth for general $\alpha,\beta,\gamma,\delta$. \end{remark} \bigskip Now we start to determine the automorphism group of $\mathbb{G}(1,5)\cap L$. A given automorphism \[ \phi\in\mathrm{Aut}(\mathbb{G}(1,5)\cap L)= \mathrm{Aut}(L,\mathbb{P}({\textstyle \bigwedge\nolimits^2}\mathbb{C}^6))\cap \mathrm{Aut}(\mathbb{G}(1,5),\mathbb{P}({\textstyle \bigwedge\nolimits^2}\mathbb{C}^6)) \] induces a dual automorphism $\phi^*$ on the dual projective space $\mathbb{P}(\bigwedge\nolimits^2\mathbb{C}^6))^*$ that preserves $L^*$ and the dual Grassmannian $\mathbb{G}(1,5)^*$, i.e. \[ \phi^*\in \mathrm{Aut}(L^*,\mathbb{P}({\textstyle \bigwedge\nolimits^2}\mathbb{C}^6)^*)\cap \mathrm{Aut}(\mathbb{G}(1,5)^*,\mathbb{P}({\textstyle \bigwedge\nolimits^2}\mathbb{C}^6)^*). \] In particular, $\phi^*$ induces a projective transformation of $L^*$ preserving $C^*$. But a smooth cubic has only finitely many automorphisms that are induced by a projective linear transformation \cite[7.3]{BK}. \bigskip To find all automorphisms of $\mathbb{G}(1,5)\cap L$ that induce the identity on $L^*$, we look for the $T\in\mathbb{P}\mathrm{GL}(6,\mathbb{C})$ such that \[ {}^tT^{-1}(\lambda A+\mu B+\nu C)T^{-1} \in \mathbb{C} \cdot (\lambda A+\mu B+\nu C) \ \ \ \mbox{for\ all}\ \lambda,\mu,\nu\in\mathbb{C}. \] It suffices to check this for $(\lambda,\mu,\nu)=(1,0,0)$, $(0,1,0)$, and $(0,0,1)$. If we normalize the representation of $T$ in $\mathrm{GL}(6,\mathbb{C})$ by $\det T=1$, we know from the previous section that ${}^tT^{-1}AT^{-1}=\mathbb{C} \cdot A$ and ${}^tT^{-1}BT=\mathbb{C} \cdot B$ is equivalent to \[ T= \left(\begin{array}{ccc} t_1 & 0 & 0\\ 0 & t_2 & 0 \\ 0 & 0 & t_3 \end{array}\right) \ \ \mbox{with}\ t_1,t_2,t_3\in\mathrm{SL}(2,\mathbb{C}). \] Furthermore, we compute \[ {}^tT^{-1}CT^{-1}= \left(\begin{array}{ccc} 0 & -\alpha\;\! {}^t t_1^{-1}t_2^{-1} & \!\! -{}^t t_1^{-1}{\gamma\, 0 \choose \delta\, \gamma}t_3^{-1} \\[2ex] \alpha\;\! {}^t t_2^{-1}t_1^{-1} & 0 & -\beta\;\! {}^t t_2^{-1}t_3^{-1} \\[2ex] {}^t t_3^{-1}{\gamma\, \delta \choose 0 \, \gamma}t_1^{-1} & \beta\;\! {}^t t_3^{-1}t_2^{-1} & 0 \end{array}\right), \] so that ${}^tT^{-1}CT^{-1}=\vartheta \cdot C$ iff $t_1=\frac{1}{\vartheta} {}^t t_2^{-1}=t_3=:t$ and \[ {}^t t^{-1}\left(\begin{array}{cc} \gamma & \delta\\ 0 & \gamma \end{array}\right)t^{-1} = \vartheta \left(\begin{array}{cc} \gamma & \delta\\ 0 & \gamma \end{array}\right). \] Because of $\det t_1=\det t_2=1$, $\vartheta$ must be either 1 or -1. Setting \[ t= \left(\begin{array}{cc} a & b\\ c & d \end{array}\right) \Longrightarrow t^{-1}= \left(\begin{array}{cc} d & -b\\ -c & a \end{array}\right) \] the last condition together with $\det t=1$ requires that the following polynomials vanish: \[ \begin{array}{l} (d^2+c^2-\vartheta)\gamma- dc \delta,\ (db+ac)\gamma-(\vartheta-ad)\delta\\[0.5ex] (db+ac)\gamma-bc\delta,\ (b^2+a^2-\vartheta)\gamma-ba\delta,\ ad-bc-1 \end{array} \] The Gr\"obner basis of the ideal generated by these polynomials with respect to the lexicographical order $\gamma>\delta>a>b>c>d$ can be computed for $\vartheta=1$ as \[ \gamma a+\delta c -\gamma d,\ b+c ,\ ad+c^2-1, \] so that \[ t= \left(\begin{array}{cc} a&\!\frac{\gamma}{\delta}(a-d)\\[0.75ex] \!\frac{\gamma}{\delta}(d-a) & d \end{array}\right) \ \ \mbox{with} \ \det t=1. \] For $\vartheta=-1$ we get as the Gr\"obner basis \[ \delta, a+d,-c+b,d^2+c^2+1. \] Since in the general case $\delta\not=0$, this gives no further automorphisms. \medskip The one-dimensional subgroup of $\mathbb{P}\mathrm{GL}(2,\mathbb{C})$ consisting of elements like $t$ above acts on $\mathbb{P}_1$ with the two fixed points $(-\delta \pm \sqrt{\delta^2-4\gamma^2}:2 \gamma)$. Hence it is conjugate to the one-dimensional subgroup of $\mathbb{P}\mathrm{GL}(2,\mathbb{C})$ that acts on $\mathbb{P}_1$ with the fixed points 0 and $\infty$. Now this subgroup consists of the invertible diagonal matrices of $\mathbb{P}\mathrm{GL}(2,\mathbb{C})$, so it is isomorphic to $\mathbb{C}^*$. Therefore we have shown \begin{theorem} The component of the automorphism group of $\mathbb{G}(1,5)\cap H^3$ containing the identity is isomorphic to $\mathbb{C}^*$. The quotient of $\mathrm{Aut}(\mathbb{G}(1,5)\cap H^3)$ by this component is a subgroup of the finite group of projective automorphisms of a smooth cubic in $\mathbb{P}_2$. \end{theorem} \section{$\mathbb{G}(1,2n)\cap H$} The hyperplane $H$ is given by an element $A\in (\bigwedge^2\mathbb{C}^{2n+1})^*$ which can be thought of as an antisymmetric matrix of size $2n+1$. Since antisymmetric matrices have an even rank, the general $H$ corresponds to an $A$ of rank $2n$. The one dimensional kernel of $A$ as a point of $\mathbb{P}_{2n}$ is called the center $c$ of $H$. \medskip The center plays a special role in the geometry of the line system $\mathbb{G}(1,2n)\cap H$ in $\mathbb{P}_{2n}$. \begin{proposition}\label{centerprop} Every line through the center of the line system $\mathbb{G}(1,2n)\cap H$ is in the line system. The center is the only point with this property. \medskip Moreover, if the line $l\not\ni c$ belongs to the line system, so does every line in the plane spanned by the line $l$ and the center $c$. \end{proposition} \emph{Proof.} The line $c\wedge p$ through the center will be in the line system if ${}^t\!c A p=0$. But $c$ is the kernel of $A$, so this is true. On the other hand, if $\overline{c}$ is a point such that every line through it belongs to the line system, then ${}^t\overline{c} A p=0$ for all $p\in \mathbb{P}_{2n}$. Hence $\overline{c}$ must be in the kernel of $A$, and therefore $\overline{c}=c$. \medskip Let the line $l=p\wedge q$ be in the line system. All the lines in the plane spanned by $l$ and $c$ -- except the lines through $c$ itself -- can be written as \[ (\alpha p +\beta c)\wedge (\lambda q+\mu c) \quad \mbox{for\ }(\alpha:\beta),(\lambda:\mu)\in \mathbb{P}_1. \] These will be in the line system since \bigskip \ \hfill $\displaystyle \begin{array}{c@{\;=\;}l} (\alpha {}^t\!p +\beta {}^t\!c) A (\lambda q+\mu c) & \alpha\lambda {}^t\!p A q+\alpha\mu {}^t\!p A c +\beta\lambda {}^t\!c A q+\beta\mu {}^t\!c A c\\[0.5ex] & \alpha\lambda {}^t\!p A q =0. \end{array} $ \hfill$\Box$ \bigskip Let us for a moment look at the projection $\mathbb{P}(\mathbb{C}^{2n+1}/c)$ of $\mathbb{P}_{2n}$ from the center $c$. This projection maps all lines in a plane through the center -- except the lines through $c$ itself -- to only one line. Hence we get a codimension one line system inside $\mathbb{P}_{2n-1}$. In fact, it is of the form $\mathbb{G}(1,2n-1)\cap \overline{H}$, which is most easily seen in coordinates. We choose a basis $(e_0,\ldots,e_{2n})$ of $\mathbb{C}^{2n+1}$ such that the hyperplane $H$ is given by the matrix \[ A=\left( \begin{array}{c|c} \begin{array}{cc} {\displaystyle 0} &{\displaystyle -\mathrm{E}_n}\\[0.5ex] {\displaystyle\mathrm{E}_n} &{\displaystyle 0} \end{array} & \begin{array}{c} \!\! 0 \\[-0.8ex]\!\! \vdots \\[-0.3ex] \!\!0 \end{array} \\ \hline 0\, \cdots\, 0 &\!\! 0 \end{array} \right)\in \mathrm{Antisym}(2n+1,\mathbb{C}). \] The center of $H$ is $c=\mathbb{P}(e_{2n})$. So the projected line system is $\mathbb{G}(1,2n-1)\cap \overline{H}$, where $\overline{H}$ is given by the matrix $A$ with the last row and column deleted. \bigskip This description helps to determine the automorphism group of $\mathbb{G}(1,2n) \cap H$. \medskip First of all, any of the automorphism must -- as a transformation $T\in\mathbb{P}\mathrm{GL} (2n+1,\mathbb{C})$ -- preserve the center, i.e. $Tc=c$. Therefore it induces a transformation $\overline{T}$ of the projected space $\mathbb{P}(\mathbb{C}^{2n+1}/c)$. This induced transformation $\overline{T}$ has to preserve the projected line system $\mathbb{G}(1,2n-1)\cap \overline{H}$. Since this case has been treated in Section \ref{g12n-1hsection}, we know that if we normalize $\overline{T}$ by $\det \overline{T}=1$, then $\overline{T}\in \mathrm{Sp}(2n,\mathbb{C})$. Therefore $T$ must have been of the form \[ T= \left( \begin{array}{c|c} {\displaystyle \overline{T}} & \begin{array}{c} \!\!0 \\[-0.8ex]\!\! \vdots \\[-0.3ex]\!\! 0 \end{array} \\ \hline a_0 \, \cdots\, a_{2n-1} & \!\! b \end{array} \right) \quad \mbox{with}\ \begin{array}{l} \overline{T}\in\mathrm{Sp}(2n,\mathbb{C})\\ a_i\in \mathbb{C} \\ b\in \mathbb{C}^*. \end{array} \] One immediately checks that ${}^tT^{-1}AT=A$, so that the automorphism group as a subset of $\mathbb{P}\mathrm{GL}(2n+1,\mathbb{C})$ consists of all elements of the above type. Since we normalized $\overline{T}$, we have up to multiplication by $-1$ an unique representative in the class of $\mathbb{P}\mathrm{GL}(2n+1,\mathbb{C})$. \medskip A small computation shows that \[ N:=\left\{ \left( \begin{array}{c|c} {\displaystyle \mathrm{E}_{2n}} & \begin{array}{c} \!\!0 \\[-0.8ex]\!\! \vdots \\[-0.3ex]\!\! 0 \end{array} \\ \hline a_0 \, \cdots\, a_{2n-1} & \!\! 1 \end{array} \right) \mid a_i\in\mathbb{C} \right\} \subset \mathrm{Aut}(\mathbb{G}(1,2n)\cap H) \] is a normal subgroup which is isomorphic to $(\mathbb{C}^{2n},+)$. \medskip Collecting everything together we have \begin{proposition} The automorphism group of $\mathbb{G}(1,2n)\cap H$ for a general hyperplane $H\subset \mathbb{P}(\bigwedge^2\mathbb{C}^{2n+1})$ is an extention of $\mathrm{Sp}(2n,\mathbb{C})\times\mathbb{C}^*/\{1,-1\}$ by $(\mathbb{C}^{2n},+)$ and is isomorphic to the group \[ \left\{ \left( \begin{array}{c|c} {\displaystyle T} & \begin{array}{c} \!\!0 \\[-0.8ex]\!\! \vdots \\[-0.3ex]\!\! 0 \end{array} \\ \hline a_0 \, \cdots\, a_{2n-1} & \!\! b \end{array} \right) \begin{array}{|l} T\in\mathrm{Sp}(2n,\mathbb{C})\\ a_i\in \mathbb{C} \\ b\in \mathbb{C}^* \end{array} \right\}/\{1,-1\}. \] \end{proposition} \bigskip The action of the automorphism group on the line system is described by the following \begin{proposition} The action of the automorphism group of $\mathbb{G}(1,2n)\cap H$ on the lines of $\mathbb{G}(1,2n)\cap H$ has two orbits: \begin{enumerate} \item the lines containing the center $c$ \item the lines that do not. \end{enumerate} \end{proposition} \emph{Proof.} Since all the automorphisms preserve the center any orbit will be contained in these two sets. \medskip First we show that the lines containing $c$ form one orbit. For two lines $c\wedge p$ and $c\wedge q$, we may assume $p,q\in\mathbb{P}(\mathbb{C}^{2n}\!\times 0)$. Take a $\overline{T}\in \mathrm{Sp}(2n,\mathbb{C})$ that maps $p$ to $q$. The trivial extention of $\overline{T}$ to $T\in\mathrm{SL}(2n+1,\mathbb{C})$ will take $c\wedge p$ to $c\wedge q$. \medskip The other lines will form the second orbit since any line not containing the center can be pushed into the hyperplane $\mathbb{P}(\mathbb{C}^{2n}\!\times 0)$ by a transformation with an element of the normal subgroup $N$. There one can use the transitive action of the $\mathrm{Aut}(\mathbb{G}(1,2n-1)\cap H)$ subgroup to show that all these lines can be mapped onto each other. \hfill$\Box$ \section{$\mathbb{G}(1,2n)\cap H^2$} Let $L=H^2$ be a 2-codimensional linear subspace of $\mathbb{P}(\bigwedge^2\mathbb{C}^{2n+1})$. We want to study the linear line system $\mathbb{G}(1,2n)\cap L$. To $L$ corresponds the line $L^*=\mathbb{P}(\lambda A-\mu B)\subset \mathbb{P}( \bigwedge^2\mathbb{C}^{2n+1} )^*$ of the hyperplanes $H_{(\lambda:\mu)}=\mathbb{P}(\lambda A-\mu B)$ containing $L$. We identify as always $(\bigwedge^2\mathbb{C}^{2n+1})^*$ with the antisymmetric matrices $\mathrm{Antisym}(2n+1,\mathbb{C})$. The locus of antisymmetric matrices of corank~3 in $\mathrm{Antisym}(2n+1,\mathbb{C})$ is 3-codimensional by Corollary \ref{dualG}. Therefore a line $L^*$ may be called general if it does not intersect it. Hence for the general line $L^*$ the antisymmetric matrices $\lambda A-\mu B$ corresponding to the hyperplanes $H_{(\lambda:\mu)}$ have all corank~1. So each of the hyperplane sections $\mathbb{G}(1,2n)\cap H_{(\lambda:\mu)}$ has a unique center $c_{(\lambda:\mu)}\in \mathbb{P}_{2n}$ by Proposition \ref{centerprop}. These centers play an important role in the geometry of $\mathbb{G}(1,2n)\cap L$. \begin{proposition} The centers $c_{(\lambda:\mu)}$ are those points of $\mathbb{P}_{2n}$ through which there passes a $\mathbb{P}_{2n-2}$ of lines of the line system $\mathbb{G}(1,2n)\cap L$. Through all the other points of $\mathbb{P}_{2n}$ passes only a $\mathbb{P}_{2n-3}$ of lines. \end{proposition} \emph{Proof.} The lines of the line system through a point $p\in \mathbb{P}_{2n}$ are $p\wedge q$ with ${}^t\!p Aq={}^t\!pBq=0$. So we need to show that ${}^t\!p A$ and ${}^t\!pB$ are linear dependent iff $p$ is a center of a hyperplane $H_{(\lambda:\mu)}$. Now ${}^t\!p A$ and ${}^t\!pB$ are linear dependent precisely if there exists a $(\lambda\,\colon \mu)\in\mathbb{P}_1$ with $0=\lambda {}^t\!pA-\mu{}^t\!pB={}^t\!p (\lambda A-\mu B)$, i.e.\ $p$ is the kernel of $\lambda A-\mu B$, which is by definition the center of $H_{(\lambda:\mu)}$. \hfill $\Box$ \bigskip \begin{remark}\label{twocenteroncurve} Any line that contains two centers is a member of the line system $\mathbb{G}(1,N)\cap L$. \end{remark} \emph{Proof.} If the line contains the centers $c_{(\alpha:\beta)}$ and $c_{(\lambda:\mu)}$, it is contained in the hyperplanes $H_{(\alpha:\beta)}$ and $H_{(\lambda:\mu)}$ by Proposition \ref{centerprop} and therefore in their intersection $L=H_{(\alpha:\beta)}\cap H_{(\lambda:\mu)}$. \hfill$\Box$ \bigskip Next we want to know more about the curve $c_{(\lambda:\mu)}$. \begin{proposition} Let $A$, $B$ be two antisymmetric matrices of size $2n+1$ such that every non-zero linear combination of them has corank 1. Then the map \[ \begin{array}{c@{\,}ccc} c: & \mathbb{P}_1 & \longrightarrow & \mathbb{P}_{2n}\\[0.3ex] & (\lambda\,\colon \mu) &\longmapsto & \ker (\lambda A-\mu B) \end{array} \] is a parametrisation of a rational normal curve of degree $n$. \end{proposition} \emph{Proof.} (compare \cite[X,4.3]{SR} for $n=2$.) First we show that the map is injective. If it is not, there are two points of $\mathbb{P}_1$ with the same image. We may assume that this is the case for $(1\,\colon 0)$ and $(0\,\colon 1)$, i.e.\ $A$ and $B$ have the same kernel, say $e_0$. Writing $A$ and $B$ in a basis with $e_0$ as first element, we have \[ A= \left(\!\!\!\begin{array}{cc} 0 &\!\!\!\!\!\!\!\; \cdots\, 0 \\[-0.8ex] \begin{array}{c} \vdots \\[-0.2ex] 0 \end{array} & \!\!\!\!\!\!\!\; {\displaystyle \widetilde{A}} \end{array} \right) \ \mbox{and}\ \ B= \left(\!\!\!\begin{array}{cc} 0 &\!\!\!\!\!\!\!\; \cdots\, 0 \\[-0.8ex] \begin{array}{c} \vdots \\[-0.2ex] 0 \end{array} & \!\!\!\!\!\!\!\; {\displaystyle \widetilde{B}} \end{array} \right) \ \ \mbox{with}\ \widetilde{A},\widetilde{B}\in\mathrm{Antisym}(2n,\mathbb{C}). \] Since $\det(\lambda \widetilde{A}-\mu \widetilde{B})$ is a homogeneous polynomial of degree $2n$, there exist a $(\lambda'\,\colon \mu')\in\mathbb{P}_1$ with $\det(\lambda' \widetilde{A}-\mu' \widetilde{B})=0$. But then $\lambda' A-\mu' B$ has corank at least two, which contradicts our assumption. \medskip Secondly, we proof that the map is of maximal rank everywhere. If it is not, we may assume that it is not maximal at $(1\,\colon 0)$. Restricting to the chart $\lambda=1$, this means $c'(0)=0$. Now from \[ \begin{array}{l} (A-\mu B)c(\mu)=0\\[0.8ex] \Longrightarrow Ac'(\mu) - B c(\mu)-\mu B c'(\mu)=0 \\[0.8ex] \Longrightarrow Ac'(0) - B c(0)=0 \end{array} \] $Bc(0)=0$ follows. Therefore $A$ and $B$ have the same kernel $c(0)$, and we are back in the above chain of arguments. \medskip Finally, we have to show that the embedding $c$ is of degree $n$. For this we give an explicit form of the map. Recall \cite[5.2]{B} that the determinant of an antisymmetric matrix $C=(c_{ij})$ of size $2n$ is the square of the irreducible Pfaffian polynomial $\mathrm{Pf}\, C$, \[ \mathrm{Pf}\, C:=\sum_\sigma \mathrm{sgn}\,(\sigma) c_{\sigma(1) \sigma(2)}\ldots c_{\sigma(2n-1) \sigma(2n)}, \] where $\sigma$ runs through all permutations $\mathrm{S}(2n)$ with $\sigma(2i-1)<\sigma(2i)$ for $i=1\ldots n$ and $\sigma(2i)<\sigma(2i+2)$ for $i=1\ldots n-1$. \medskip Let $c_i(\lambda\,\colon \mu)$ denote $(-1)^i$-times the Pfaffian of the matrix $\lambda A-\mu B$ with the $i$-th row and column deleted. Then the $c_i(\lambda\,\colon \mu)$ are irreducible polynomials of degree $n$ and by a straightforward but messy computation one can check that $(c_0(\lambda\,\colon \mu)\,\colon \ldots\,\colon c_{2n}(\lambda\,\colon \mu))$ is the kernel of $\lambda A-\mu B$. Therefore $c=(c_0\,\colon \ldots\,\colon c_{2n})$, which shows that $c$ is a degree $n$ embedding of $\mathbb{P}_1$. \hfill$\Box$ \bigskip After we have determined the special points in $\mathbb{P}_{2n}$ of the line system $\mathbb{G}(1,2n)\cap L$, we are nearly ready to compute its automorphism group. It remains to give a normal form for the line $L^*\subset \mathbb{P}(\bigwedge^2\mathbb{C}^{2n+1})$ to make computations easier. This normal form was found by Donagi \cite[2.2]{D}. But he did not give a proof for it since his main interest was lines in $\mathbb{P}_{2n-1}$ and not in $\mathbb{P}_{2n}$. So we give the proof here. \begin{proposition} Let $L^*$ be a line in $\mathbb{P}(\bigwedge^2\mathbb{C}^{2n+1})^*$ such that the antisymmetric matrices corresponding to the points of $L^*$ have all corank 1. Then there exists a basis $(e_0,\ldots,e_{2n})$ of $\mathbb{C}^{2n+1}$ such that the line can be taken as $L^*=\mathbb{P}(\lambda A-\mu B)$ with the matrices \[ A= \left(\begin{array}{c|c} \begin{array}{cc} {\displaystyle 0} &\!\! {\displaystyle-\mathrm{E}_n}\\[0.5ex] {\displaystyle\mathrm{E}_n} &{\displaystyle 0} \end{array} & \begin{array}{c} \!\! 0 \\[-0.8ex]\!\! \vdots \\[-0.3ex]\!\! 0 \end{array} \\ \hline 0\, \cdots\, 0 & \!\! 0 \end{array}\right) \ \ \mbox{and} \ \ \ B= \left( \begin{array}{c|c|c} {\displaystyle 0}\! & \begin{array}{c}\!\!\! 0\!\!\!\! \\[-0.8ex] \!\!\!\vdots\!\!\!\! \end{array} & \! {\displaystyle -\mathrm{E}_n} \\ \cline{1-1}\cline{3-3} \multicolumn{3}{c}{\hspace{0.4ex} 0\; \cdots \; 0 \;\cdots\; 0} \\[-0.35ex]\cline{1-1}\cline{3-3} {\displaystyle \raiseghost{4ex}{0ex}{${\displaystyle \mathrm{E}_n}$}\phantom{-\mathrm{E}_n}}\! & \begin{array}{c} \!\!\!\vdots\!\!\!\! \\[-0.3ex]\!\!\!0\!\!\!\! \end{array} &\! {\displaystyle 0} \end{array}\right). \] \end{proposition} \emph{Proof.} Let $A$ and $B$ any two matrices of $L^*$ in an arbitrary basis. We will adjust the basis in three steps to achieve the required form for $A$ and $B$. \medskip \noindent\emph{$1^{st}$ Step:} We know that the map \[ \begin{array}{c@{\,}ccc} c: & \mathbb{P}_1 & \longrightarrow & \mathbb{P}_{2n}\\[0.5ex] & (\lambda\,\colon \mu) &\longmapsto & \ker (\lambda A-\mu B) \end{array} \] is a parametrisation of a rational normal curve of degree $n$. Modulo projective transformations of $\mathbb{P}_1$ and $\mathbb{P}_{2n}$ such parametrisations are all the same. So we can pick a basis of $\mathbb{P}_1$ and $n+1$ linear independent vectors $e_n,\ldots,e_{2n}$ of $\mathbb{C}^{2n+1}$ such that \[ \begin{array}{c@{\,}ccc} c: & \mathbb{P}_1 & \longrightarrow & \mathbb{P}_{2n}\\[0.5ex] & (\lambda\,\colon \mu) &\longmapsto & \mathbb{P} \left(\sum\limits_{i=0}^n \lambda^i\mu^{n-i} e_{n+i} \right). \end{array} \] Extending $(e_n,\ldots,e_{2n})$ to a basis $(e_0,\ldots,e_{2n})$ of $\mathbb{C}^{2n+1}$ and denoting by $a_0,\ldots,a_{2n}$ resp. $b_0,\ldots,b_{2n}$ the columns of $A$ resp. $B$, the fact that $c(\lambda\,\colon \mu)$ is the kernel of $\lambda A-\mu B$ for all $(\lambda\,\colon \mu)\in\mathbb{P}_1$ has the following consequences for $A$ and $B$: \bigskip $ \hspace{9ex}( \lambda A-\mu B) \left(\sum\limits_{i=0}^n \lambda^i\mu^{n-i} e_{n+i} \right)=0$\vspace{1ex} $\hspace{9ex} \Longrightarrow \sum\limits_{i=0}^n \lambda^{i+1}\mu^{n-i} a_{n+i} -\sum\limits_{i=0}^n \lambda^i\mu^{n+1-i} b_{n+i} =0 $\vspace{1ex} $\hspace{9ex}\Longrightarrow -\mu^{n+1}b_n +\sum\limits_{i=0}^{n-1} \lambda^{i+1}\mu^{n-i}(a_{n+1}-b_{n+i+1}) +\lambda^{n+1} a_{2n} =0 $\vspace{1ex} $\hspace{9ex}\Longrightarrow b_n=0,\ a_{2n}=0,\ a_{n+i}=b_{n+i+1} \ \ \mbox{for\ }i= 0\ldots n-1. \hfill (*)$ \bigskip We claim that this implies: \[ a_{n+i,n+j}=0 \ \ \ \ \mbox{for\ }i,j=0\ldots n. \] Indeed, for $1\le i \le n$, $0\le j\le n-1$ we have using $(*)$ \[ a_{n+i,n+j}=b_{n+i,n+j+1}=-b_{n+j+1,n+i}=-a_{n+j+1,n+i-1}=a_{n+i-1,n+j+1}. \] This shows that the $a_{n+i,n+j}$ are all the same for $i+j=const$, in particular $a_{n+i,n+j}=a_{n+j,n+i}$. On the other hand, by the antisymmetricity of $A$ we have $a_{n+i,n+j}=-a_{n+j,n+i}$, and the claim follows. \medskip Using $(*)$ again we know that $A$ and $B$ look in our basis like \[ A= \left(\begin{array}{c|c} \begin{array}{cc} {\displaystyle \widetilde{A}} &\!\! {\displaystyle-{}^t\! M}\\[0.5ex] {\displaystyle M} &{\displaystyle 0} \end{array} & \begin{array}{c} \!\! 0 \\[-0.8ex]\!\! \vdots \\[-0.3ex]\!\! 0 \end{array} \\ \hline 0\, \cdots\, 0 & \!\! 0 \end{array}\right) \ \ \mbox{and} \ \ B= \left( \begin{array}{c|c|c} {\displaystyle \widetilde{B}}\! & \begin{array}{c}\!\!\! 0\!\!\!\! \\[-0.8ex] \!\!\!\vdots\!\!\!\! \end{array} & \! {\displaystyle -{}^t\! M} \\ \cline{1-1}\cline{3-3} \multicolumn{3}{c}{\hspace{0.4ex} 0\; \cdots \; 0 \;\cdots\; 0} \\[-0.35ex]\cline{1-1}\cline{3-3} {\displaystyle \raiseghost{4ex}{0ex}{${\displaystyle M}$}\phantom{-{}^t\! M}}\! & \begin{array}{c} \!\!\!\vdots\!\!\!\! \\[-0.3ex]\!\!\!0\!\!\!\! \end{array} &\! {\displaystyle 0} \end{array}\right) \] with $\widetilde{A},\widetilde{B}\in \mathrm{Antisym}(n,\mathbb{C})$ and $M\in\mathrm{GL}(n,\mathbb{C})$. \medskip \noindent\emph{$2^{nd}$ Step:} Here we will improve the choice of $(e_0,\ldots,e_{n-1})$ to achieve $\widetilde{A}=0$ and $M=\mathrm{E}_n$. We claim: \medskip Let an antisymmetric matrix $A\in\mathrm{Antisym}(2n+1,\mathbb{C})$ of rank $2n$ and linear independent vectors $e_n,\ldots, e_{2n}\in\mathbb{C}^{2n+1}$ with ${}^t\!e_iAe_j=0$ for $n\le i,j\le 2n$ and $Ae_{2n}=0$ be given. Then $(e_n,\ldots,e_{2n})$ can be extended to a basis $(e_0,\ldots,e_{2n})$ of $\mathbb{C}^{2n+1}$ such that in this basis $A$ is given as \[ A=\left(\begin{array}{c|c} \begin{array}{cc} {\displaystyle 0} & \!\!{\displaystyle-\mathrm{E}_n}\\[0.5ex] {\displaystyle\mathrm{E}_n} &{\displaystyle 0} \end{array} & \begin{array}{c} \!\! 0 \\[-0.8ex]\!\! \vdots \\[-0.3ex]\!\! 0 \end{array} \\ \hline 0\, \cdots\, 0 & \!\! 0 \end{array}\right) .\] The proof is by induction. The statement is trivial for $n=0$. Assuming the claim for $n-1$, we prove it for $n$. Let $W:=\bigcap_{i=1}^n\ker {}^t\!e_{n+i}A$, then there exists an $e_0\in W$ with ${}^t\!e_n A e_0=1$. If not, we would have $W=W\cap \ker {}^t\!e_nA$ and with $e_{2n}\in \ker A$ \[ \dim\, \mathrm{span}\,\{{}^t\!e_nA,\ldots,{}^t\!e_{2n}A\}= \dim\, \mathrm{span}\,\{{}^t\!e_{n+1} A,\ldots,{}^t\!e_{2n-1}A\}\le n-1, \] which contradicts $\mathrm{rank}\, A=2n$. \medskip Set $V:=\ker {}^t\!e_0A \cap \ker{}^t\!e_nA$, then $\dim V =2(n-1)+1$ and $e_{n+1},\ldots,e_{2n}\in V$. Therefore the induction hypothesis can be applied to $\left.A\right|_V$. Together with ${}^t\!e_0Ae_n=1$ and ${}^t\!e_0Av={}^t\!e_nAv=0$ for $v\in V$ this implies the stated form of the matrix. \medskip So up to now $A$ and $B$ look like \[ A= \left(\begin{array}{c|c} \begin{array}{cc} {\displaystyle 0} & \!\!{\displaystyle-\mathrm{E}_n}\\[0.5ex] {\displaystyle\mathrm{E}_n} &{\displaystyle 0} \end{array} & \begin{array}{c} \!\! 0 \\[-0.8ex]\!\! \vdots \\[-0.3ex]\!\! 0 \end{array} \\ \hline 0\, \cdots\, 0 & \!\! 0 \end{array}\right) \ \ \ B= \left( \begin{array}{c|c|c} {\displaystyle \widetilde{B}}\! & \begin{array}{c}\!\!\! 0\!\!\!\! \\[-0.8ex] \!\!\!\vdots\!\!\!\! \end{array} & \! {\displaystyle -\mathrm{E}_n} \\ \cline{1-1}\cline{3-3} \multicolumn{3}{c}{\hspace{0.4ex} 0\; \cdots \; 0 \;\cdots\; 0} \\[-0.35ex]\cline{1-1}\cline{3-3} {\displaystyle \raiseghost{4ex}{0ex}{${\displaystyle \mathrm{E}_n}$}\phantom{-\mathrm{E}_n}}\! & \begin{array}{c} \!\!\!\vdots\!\!\!\! \\[-0.3ex]\!\!\!0\!\!\!\! \end{array} & {\displaystyle 0} \end{array}\right). \] \noindent\emph{$3^{rd}$ Step:} We adjust the vectors $(e_0,\ldots,e_{n-1})$ so that $\widetilde{B}=0$ and $A$ stays the same. \medskip We note that a transformation of $\mathbb{C}^{2n+1}$ by \[ T= \left(\begin{array}{c|c} \begin{array}{cc} {\displaystyle \mathrm{E}_n} & {\displaystyle 0}\\[0.5ex] {\displaystyle t} &{\displaystyle \mathrm{E}_n} \end{array} & \begin{array}{c} \!\! 0 \\[-0.8ex]\!\! \vdots \\[-0.3ex]\!\! 0 \end{array} \\ \hline 0\, \cdots\, 0 & \!\! 1 \end{array}\right)^{\!\!-1} \ \ \mbox{with}\ t\in \mathrm{Sym}(n,\mathbb{C}) \] does not change $A$ since \[ {}^tT^{-1}AT^{-1}={}^tT^{-1}(AT^{-1})= \left(\begin{array}{c|c} \begin{array}{cc} {\displaystyle \mathrm{E}_n} & {\displaystyle t}\\[0.5ex] {\displaystyle 0} &{\displaystyle \mathrm{E}_n} \end{array} & \begin{array}{c} \!\! 0 \\[-0.8ex]\!\! \vdots \\[-0.3ex]\!\! 0 \end{array} \\ \hline 0\, \cdots\, 0 & \!\! 1 \end{array}\right) \! \left(\begin{array}{c|c} \begin{array}{cc} {\displaystyle \!\!-t} &\!\! {\displaystyle-\mathrm{E}_n}\\[0.5ex] {\displaystyle\mathrm{E}_n} &{\displaystyle 0} \end{array} & \begin{array}{c} \!\! 0 \\[-0.8ex]\!\! \vdots \\[-0.3ex]\!\! 0 \end{array} \\ \hline 0\, \cdots\, 0 & \!\! 0 \end{array}\right) =A. \] If we denote by $\overline{t}$ resp. $\left|t\right.\in \mathrm{M}(n\times n,\mathbb{C})$ the matrix that we obtain by deleting the first row resp. column of $t$ and adding a row of zeroes below resp. a column of zeroes on the right side, we can write down the transformation of $B$ as follows: \[\begin{array}{l} {}^tT^{-1}BT^{-1}={}^tT^{-1}(BT^{-1})= \left(\begin{array}{c|c} \begin{array}{cc} {\displaystyle \mathrm{E}_n} & {\displaystyle t}\\[0.5ex] {\displaystyle 0} &{\displaystyle \mathrm{E}_n} \end{array} & \begin{array}{c} \!\! 0 \\[-0.8ex]\!\! \vdots \\[-0.3ex]\!\! 0 \end{array} \\ \hline 0\, \cdots\, 0 & \!\! 1 \end{array}\right) \left( \begin{array}{c|c|c} {\displaystyle\! \widetilde{B}-\overline{t}}\! & \begin{array}{c}\!\!\! 0\!\!\!\! \\[-0.8ex] \!\!\!\vdots\!\!\!\! \end{array} & {\displaystyle -\mathrm{E}_n} \\ \cline{1-1}\cline{3-3} \multicolumn{3}{c}{\hspace{0.65ex} 0\; \cdots \; 0 \;\cdots\; 0} \\[-0.35ex]\cline{1-1}\cline{3-3} {\displaystyle \raiseghost{4ex}{0ex}{${\displaystyle \mathrm{E}_n}$}\phantom{-\mathrm{E}_n}}\! & \begin{array}{c} \!\!\!\vdots\!\!\!\! \\[-0.3ex]\!\!\!0\!\!\!\! \end{array} & {\displaystyle 0} \end{array}\right) \\[4.8ex] \phantom{{}^tT^{-1}BT^{-1}}= \left( \begin{array}{c|c|c} {\displaystyle\! \widetilde{B}-\overline{t}+\left|t\right.}\! & \begin{array}{c}\!\!\! 0\!\!\!\! \\[-0.8ex] \!\!\!\vdots\!\!\!\! \end{array} & {\displaystyle\ \ \ -\mathrm{E}_n \ \ \ } \\ \cline{1-1}\cline{3-3} \multicolumn{3}{c}{\hspace{0.5ex} 0 \;\ \ \cdots\ \ \; 0 \; \ \ \cdots\ \ \; 0} \\[-0.35ex]\cline{1-1}\cline{3-3} \mathrm{E}_n & \begin{array}{c} \!\!\!\vdots\!\!\!\! \\[-0.3ex]\!\!\!0\!\!\!\! \end{array} & {\displaystyle 0} \end{array}\right) . \end{array} \] So to finish this step, we need to show that every antisymmetric matrix $\widetilde{B}=(b_{ij})\in\mathrm{Antisym}(n,\mathbb{C})$ can be written as $\overline{t}-\left|t\right.$ for a symmetric matrix $t\in\mathrm{Sym}(n,\mathbb{C})$. The entries of $\overline{t}-\left|t\right.$ are \[ (t_{i+1,j}-t_{i,j+1})_{i j} \] where $t_{n+1, i}:=t_{i, n+1}:=0$ for all $i=1\ldots n$. Obviously, $\overline{t}-\left|t\right. $ is antisymmetric. We set $t_{1 i}:=t_{i1}:=0$ for $i=1\ldots n$ and define recursively for $j$ from $n$ down to $2$ \[ t_{i+1, j}:=t_{i, j+1}+b_{i j} \ \ \ \mbox{for}\ i=1\ldots j-1. \] Then by the symmetry of $t$ the whole matrix $t$ is defined and $\overline{t}-\left|t\right.=B$ \hfill$\Box$ \bigskip For the linear system in normal form the rational curve of centers has the parametrisation \[ \begin{array}{l} c(\lambda\,\colon \mu) = \ker (\lambda A-\mu B) = \ker\left( \begin{array}{c@{\,}c@{\,}c|c@{\,}c@{\,}c@{\,}c} & & & -\lambda &\mu & & \\[-0.5ex] &\raiseghost{-1ex}{0ex}{\huge 0}& & &\ddots &\ddots & \\[-0.5ex] & & & & &-\lambda &\mu \\ \hline \lambda & & & & & & \\[-0.5ex] -\mu &\ddots & & & & & \\[-0.5ex] &\ddots &\lambda & &\raiseghost{2ex}{1ex}{\Huge 0} & & \\[-0.5ex] & &-\mu & & & & \end{array} \right) \\[11ex] \phantom{c(\lambda\,\colon \mu)} = (0\,\colon \ldots\,\colon 0\,\colon \mu^n\,\colon \mu^{n-1}\lambda\,\colon \ldots\,\colon \lambda^n). \end{array} \] The $\mathbb{P}_{2n-2}$ of lines through a center $c_{(\lambda:\mu)}$ is given by \[ c_{(\lambda:\mu)} \wedge q \ \ \ \ \mbox{where\ }q\in\mathbb{P}_{2n}\ \mbox{with} \ {}^t\!c_{(\lambda:\mu)} A q= {}^t\!c_{(\lambda:\mu)} B q=0, \] i.e.\ $q$ must be an element of the hyperplane $h_{(\lambda:\mu)}\in\mathbb{P}_{2n}^*$ \[ \begin{array}{c@{\;=\;}l} h_{(\lambda:\mu)} & \ker {}^t\!c_{(\lambda:\mu)} A \cap {}^t\!c_{(\lambda:\mu)} B\\[1.8ex] & \ker\left( \begin{array}{ccccccc} \mu^n & \mu^{n-1}\lambda &\cdots & \mu \lambda^{n-1} & 0 & \cdots & 0\\ \mu^{n-1}\lambda &\mu^{n-2}\lambda^2 &\cdots & \lambda^n & 0 & \cdots & 0 \end{array}\right)\\[3.5ex] & \ker\left( \begin{array}{ccccccc} \mu^{n-1} & \mu^{n-2}\lambda &\cdots & \lambda^{n-1} & 0 & \cdots & 0\\ \end{array}\right). \end{array} \] So the hyperplanes $h_{(\lambda:\mu)}$, which are traced out by the $\mathbb{P}_{2n-2}$ of lines through the centers, give rise to a rational normal curve of degree $n-1$ in the space of hyperplanes containing the center curve. That the hyperplanes $h_{(\lambda:\mu)}$ contain the center curve could already be seen from the Remark \ref{twocenteroncurve}, by which $h_{(\lambda:\mu)}$ must contain any line connecting $c_{(\lambda:\mu)}$ with any other point of the center curve. \bigskip Now we are ready to study the automorphism group of $\mathbb{G}(1,2n)\cap L$. \medskip Any automorphism $T\in\mathrm{Aut}(\mathbb{G}(1,2n)\cap L)\subseteq\mathbb{P}\mathrm{GL}(2n+1,\mathbb{C})$ has to map the center curve onto itself and also the projective space $P\cong\mathbb{P}_n$ spanned by the center curve onto itself. It is known \cite[10.12]{H} that the group of automorphisms of $\mathbb{P}_n$ fixing a rational normal curve of degree $n$ is isomorphic to $\mathbb{P}\mathrm{GL}(2,\mathbb{C})$. If the rational normal curve is given by \[ \begin{array}{c@{\,}ccc} c: & \mathbb{P}_1 & \longrightarrow & \mathbb{P}_{n}\\ & (\lambda\,\colon \mu) &\longmapsto & (\mu^n\,\colon \mu^{n-1}\lambda \,\colon \ldots\,\colon \lambda^n), \end{array} \] this isomorphism $\mathbb{P}\mathrm{GL}(2,\mathbb{C})\cong\mathrm{Aut}(c,\mathbb{P}_n)$ maps \[ t=\left(\begin{array}{cc} a&b\\c&d\end{array} \right)\in\mathbb{P}\mathrm{GL}(2,\mathbb{C}) \] to $t_{n+1}\in\mathbb{P}\mathrm{GL}(n+1,\mathbb{C})$ where $t_{n+1}$ is the unique matrix such that \[ t_{n+1} \left(\begin{array}{c} \mu^n\\ \mu^{n-1}\lambda \\ \vdots \\ \lambda^n \end{array}\right) = \left(\begin{array}{c} (d\mu+c\lambda)^n \\ (d\mu+c\lambda)^{n-1}(b\mu+a\lambda) \\ \vdots \\ (b\mu+a\lambda)^n \end{array}\right); \] for example \[ t_2= \left(\begin{array}{cc} d & c\\ b & a \end{array} \right) \qquad t_3= \left(\begin{array}{ccc} d^2&2cd & c^2\\ bd & ad+bc &ac\\b^2 & 2ab & a^2 \end{array} \right). \] Applying this to the center curve restricts the form of the transformation $T$ to \[ T= \left(\begin{array}{c|c} * & 0\\ \hline {\displaystyle *} & \rule[-1.5ex]{0ex}{4.5ex} {\displaystyle t_{n+1}} \end{array} \right). \] We know further that if $T$ maps $c_{(\lambda:\mu)}$ to $c_{(a\lambda+b\mu :c\lambda+d\mu)}$, then it must map the hyperplane $h_{(\lambda:\mu)}$ to $h_{(a\lambda+b\mu :c\lambda+d\mu)}$. Therefore it induces also an automorphism on the rational curve $h$ of degree $n-1$ in the dual projective space $(\mathbb{P}_{2n}/P)^*$ of hyperplanes containing $P$. Hence $T$ must be of the form \[ T= \left(\begin{array}{c|c} \alpha\, \rule[-2.5ex]{0ex}{5.5ex} {}^t\hspace{-0.1ex} t_n^{-1} & 0\\ \hline * & \rule[-3ex]{0ex}{6.5ex} {\displaystyle \ t_{n+1}\ } \end{array} \right) \ \ \mbox{with\ }\alpha\in\mathbb{C}^*. \] We make the following \emph{claim}: \[ T= \left(\begin{array}{c|c} \rule[-1.5ex]{0ex}{4ex}{}^t t_n^{-1} & 0\\ \hline 0 & \rule[-1.5ex]{0ex}{4.5ex} {\displaystyle t_{n+1}} \end{array} \right)\in \mathbb{P}\mathrm{GL}(2n+1,\mathbb{C}) \] is an automorphism of the linear system $\mathbb{G}(1,2n)\cap L$. \medskip \noindent\emph{Proof.} We need to check that for every $t\in\mathbb{P}\mathrm{GL}(2,\mathbb{C})$ \[ {}^tT^{-1}(\lambda A-\mu B) T^{-1}\in \mathrm{span}\,\{ A,B\} \] for all $\lambda,\mu \in\mathbb{C}$. Since $\mathbb{P}\mathrm{GL}(2,\mathbb{C})$ is a group, this is equivalent to the statement that for every $t\in\mathbb{P}\mathrm{GL}(2,\mathbb{C})$ \[ {}^tT(\lambda A-\mu B) T\in \mathrm{span}\,\{ A,B\} \] for all $\lambda,\mu \in\mathbb{C}$. Because of the linearity it is enough to do this for $(\lambda,\mu)=(1,0)$ and $(0,-1)$. Denoting by $\overline{t_{n+1}}$ resp. $\underline{t_{n+1}}$ the matrix $t_{n+1}$ with the first resp. last row deleted, we compute: \[ \begin{array}{l} {}^tT(AT)=\! \left(\!\begin{array}{c|c} \rule[-1.5ex]{0ex}{4ex} t_n^{-1} & 0\\ \hline 0 & \rule[-1.5ex]{0ex}{4.5ex} {\displaystyle{}^t t_{n+1}} \end{array}\! \right) \!\!% \left(\!\begin{array}{c|c} \rule[-1.5ex]{0ex}{5ex} 0 & -\underline{t_{n+1}}\\ \hline \!\begin{array}{c}{}^tt_n^{-1} \\[-1ex] {\scriptstyle 0 \cdots 0} \end{array} \! & \rule[-1.5ex]{0ex}{5.5ex} {\displaystyle 0} \end{array}\!\right) \!=\! \left(\!\begin{array}{c|c} \rule[-2.5ex]{0ex}{7ex} \mbox{\Large 0} & -t_n^{-1}\underline{t_{n+1}} \!\\ \hline {}^t\!(t_n^{-1} \underline{t_{n+1}})\! & \rule[-2.5ex]{0ex}{7.5ex} \mbox{\LARGE 0} \end{array}\!\right) \\[8ex] {}^tT(BT)=\! \left(\!\begin{array}{c|c} \rule[-1.5ex]{0ex}{4ex} t_n^{-1} & 0\\ \hline 0 & \rule[-1.5ex]{0ex}{4.5ex} {\displaystyle{}^t t_{n+1}} \end{array}\! \right) \!\!% \left(\!\begin{array}{c|c} \rule[-1.5ex]{0ex}{5ex} 0 & -\overline{t_{n+1}}\\ \hline \!\begin{array}{c} {\scriptstyle 0 \cdots 0} \\[-0.5ex]{}^tt_n^{-1} \end{array} \! & \rule[-1.5ex]{0ex}{5.5ex} {\displaystyle 0} \end{array}\!\right) \!=\! \left(\!\begin{array}{c|c} \rule[-2.5ex]{0ex}{7ex} \mbox{\Large 0} & -t_n^{-1}\overline{t_{n+1}}\!\\ \hline {}^t\!(t_n^{-1} \overline{t_{n+1}})\! & \rule[-2.5ex]{0ex}{7.5ex} \mbox{\LARGE 0} \end{array}\!\right)\!. \end{array} \] So, if we show \[ \begin{array}{l} \underline{t_{n+1}}=d\left( t_n {}_0^0 \right)+ c \left( {}_0^0 t_n \right) \Longrightarrow t_n^{-1} \underline{t_{n+1}}=d\left( \mathrm{E}_n {}_0^0 \right)+ c \left( {}_0^0 \mathrm{E}_n \right) \\[3ex] \overline{t_{n+1}}=b\left( t_n {}_0^0 \right)+ a \left( {}_0^0 t_n \right) \Longrightarrow t_n^{-1} \overline{t_{n+1}}=b\left( \mathrm{E}_n {}_0^0 \right)+ a \left( {}_0^0 \mathrm{E}_n \right), \end{array} \] where ${}_0^0$ stands for adding a column of zeroes, then \[ \begin{array}{l} {}^tT A T=dA+cB\\[1ex] {}^tT BT=bA+aB. \end{array} \] To show the equality for $\underline{t_{n+1}}$ note that on the one hand $\underline{t_{n+1}}$ is the unique matrix with \[ \underline{t_{n+1}} \left(\begin{array}{c} \mu^n\\ \mu^{n-1}\lambda \\ \vdots \\ \lambda^n \end{array}\right) = \left(\begin{array}{c} (d\mu+c\lambda)^n \\ (d\mu+c\lambda)^{n-1}(b\mu+a\lambda) \\ \vdots \\ (d\mu+c\lambda)(b\mu+a\lambda)^{n-1} \end{array}\right) \] and on the other hand \[ \begin{array}{l} \left(\begin{array}{c} (d\mu+c\lambda)^n \\ (d\mu+c\lambda)^{n-1}(b\mu+a\lambda) \\ \vdots \\ (d\mu+c\lambda)(b\mu+a\lambda)^{n-1} \end{array}\right) =(d\mu+c\lambda) \left(\begin{array}{c} (d\mu+c\lambda)^{n-1} \\ (d\mu+c\lambda)^{n-2}(b\mu+a\lambda) \\ \vdots \\ (b\mu+a\lambda)^{n-1} \end{array}\right) \\[7ex] = (d\mu+c\lambda)t_n \left(\begin{array}{c} \mu^{n-1}\\ \mu^{n-2}\lambda \\ \vdots \\ \lambda^{n-1} \end{array}\right) = d t_n \left(\begin{array}{c} \mu^n\\ \mu^{n-1}\lambda \\ \vdots \\ \mu\lambda^{n-1} \end{array}\right) + c t_n \left(\begin{array}{c} \mu^{n-1}\lambda\\ \mu^{n-2}\lambda^2 \\ \vdots \\ \lambda^n \end{array}\right)\\[7ex] = \left( d\left(t_n {}_0^0\right)+ c\left( {}_0^0 t_n\right) \right) \left(\begin{array}{c} \mu^n\\ \mu^{n-1}\lambda \\ \vdots \\ \lambda^n \end{array}\right). \end{array} \] Of course, the proof for $\overline{t_{n+1}}=b\left( t_n {}_0^0 \right)+ a \left( {}_0^0 t_n \right)$ is analogous. \hfill$\Box$ \bigskip Given any automorphism of the line system $\mathbb{G}(1,2n)\cap L$ we can compose it with one of the above automorphisms such that the composition fixes the center curve pointwise. So, we can focus our attention to automorphisms of the last type. \begin{lemma}\label{centerfix} All automorphisms of $\mathbb{G}(1,2n) \cap L$ that fix the center curve pointwise are of the form \[ T=\left(\begin{array}{cc} \alpha \mathrm{E}_n & 0\\ S & {\displaystyle \mathrm{E}_{n+1}} \end{array}\right) \ \mbox{with\ }\alpha\in\mathbb{C}^*,\ S\in\mathrm{M}( (n+1)\times n,\mathbb{C}), \] where the matrix $S\in\mathrm{M}( (n+1)\times n,\mathbb{C})$ has the same entries along the minor diagonals, i.e.\ $s_{ij}=s_{kl}$ for $i+j=k+l$. \medskip As a group these matrices are isomorphic to the semi direct product $\mathbb{C}^{2n}\ltimes \mathbb{C}^*$, $\ (s,\alpha)\cdot(s',\alpha')=(\alpha's+s',\alpha\alpha')$. \end{lemma} \emph{Proof.} We need only to check the property of $S$ and the group structure. $T$ is an automorphism iff \[ {}^tT^{-1}AT^{-1},{}^tT^{-1}BT^{-1}\in \mathrm{span}\,\{A,B\}. \] The inverse of $T$ is \[ T^{-1}=\left(\begin{array}{cc} \frac{1}{\alpha} \mathrm{E}_n & 0\\[1ex] \!- \frac{1}{\alpha} S & {\displaystyle \mathrm{E}_{n+1}} \end{array}\right). \] Now if $\overline{S}$ resp. $\underline{S}\in\mathrm{M}(n\times n,\mathbb{C})$ denote the matrix $S$ with the first resp. last row deleted, then \[\begin{array}{l} {}^tT^{-1}(AT^{-1})= \!\!\left(\begin{array}{cc} \!\frac{1}{\alpha} \mathrm{E}_n\! &\!\!-\frac{1}{\alpha} {}^t\!S \!\\[1ex] \! 0 & \!\! {\displaystyle\mathrm{E}_{n+1}} \! \end{array}\right) \!\!% \left(\begin{array}{c|c} \begin{array}{cc} \!\!\! \frac{1}{\alpha}{\displaystyle \underline{S}} &\!\!\! {\displaystyle-\mathrm{E}_n}\!\!\\[1ex] \!\!\! \frac{1}{\alpha} {\displaystyle \mathrm{E}_n} &{\displaystyle 0} \end{array}\!\!\!\! & \begin{array}{c} \!\! 0 \\[-0.8ex]\!\! \vdots \\[-0.3ex]\!\! 0 \end{array} \\ \hline 0\, \cdots\, 0 & \!\! 0 \end{array}\right) \\[6.5ex] \phantom{{}^tT^{-1}(AT^{-1}) } = \left(\begin{array}{c|c} \begin{array}{cc} \!\!\! \frac{1}{\alpha^2}{\displaystyle(\underline{S}-{}^t\!\underline{S})} &\!\! -\frac{1}{\alpha} {\displaystyle\mathrm{E}_n}\\[1ex] \!\!\frac{1}{\alpha}{\displaystyle\mathrm{E}_n} &{\displaystyle 0} \end{array}\!\!\!\! & \begin{array}{c} \!\! 0 \\[-0.8ex]\!\! \vdots \\[-0.3ex]\!\! 0 \end{array} \\ \hline 0\, \cdots\, 0 & \!\! 0 \end{array}\right) \\[6.5ex] {}^tT^{-1}(BT^{-1})= \left(\begin{array}{cc} \!\frac{1}{\alpha} \mathrm{E}_n\! & \!\! -\frac{1}{\alpha} {}^t\!S\!\\[1ex] \!0 &\!\! {\displaystyle \mathrm{E}_{n+1}}\!\! \end{array}\right) \!\!% \left( \begin{array}{c|c|c} \frac{1}{\alpha}{\displaystyle\overline{S}}\! & \begin{array}{c}\!\!\! 0\!\!\!\! \\[-0.8ex] \!\!\!\vdots\!\!\!\! \end{array} & \! {\displaystyle -\mathrm{E}_n} \\ \cline{1-1}\cline{3-3} \multicolumn{3}{c}{\hspace{0.0ex} 0\; \cdots \; 0 \;\cdots\; 0} \\[-0.35ex]\cline{1-1}\cline{3-3} \frac{1}{\alpha}{\displaystyle \mathrm{E}_n}\! & \begin{array}{c} \!\!\!\vdots\!\!\!\! \\[-0.3ex]\!\!\!0\!\!\!\! \end{array} &\! {\displaystyle 0} \end{array}\right) \\[8.5ex] \phantom{{}^tT^{-1}(BT^{-1}) } = \left( \begin{array}{c|c|c} \!\!\frac{1}{\alpha^2}{\displaystyle (\overline{S}-{}^t\overline{S}) }\!\!\! & \begin{array}{c}\!\!\! 0\!\!\!\! \\[-0.5ex] \!\!\!\vdots\!\!\!\! \end{array} & \! \ \; -\frac{1}{\alpha} {\displaystyle \mathrm{E}_n}\ \; \\ \cline{1-1}\cline{3-3} \multicolumn{3}{c}{\hspace{0.5ex} 0\ \ \; \cdots \ \ \; 0 \ \ \; \cdots\ \ \; 0} \\[-0.35ex]\cline{1-1}\cline{3-3} \frac{1}{\alpha}{\displaystyle \mathrm{E}_n}\! & \begin{array}{c} \!\!\!\vdots\!\!\!\! \\[-0.0ex]\!\!\!0\!\!\!\! \end{array} &\! {\displaystyle 0} \end{array}\right)\!\!. \end{array} \] Therefore $T$ is an automorphism iff $\underline{S}={}^t\!\underline{S}$ and $\overline{S}={}^t\overline{S}$. In other words \[\left. \begin{array}{c@{\;=\;}c} s_{ij} & s_{ji}\\ s_{i+1, j} & s_{j+1, i} \end{array} \right.\quad \mbox{for\ }1\le i,j\le n, \] so \[ s_{ij} = s_{ji} = s_{(j-1)+1, i} = s_{i+1, j-1} \] for $j>1$ and $i<n$, hence $s_{ij}=s_{kl}$ for $i+j=k+l$. \medskip The statement about the group action follows from \bigskip \ \hfill $\displaystyle \left(\begin{array}{cc} \alpha \mathrm{E}_n &\!\! 0\\ S &\!\! \mathrm{E}_{n+1} \end{array}\right) \left(\begin{array}{cc} \alpha' \mathrm{E}_n &\!\! 0\\ S' &\!\! \mathrm{E}_{n+1} \end{array}\right) = \left(\begin{array}{cc} \!\alpha\alpha' \mathrm{E}_n &\!\! 0\\ \!\alpha'S+S' & \!\!\mathrm{E}_{n+1} \end{array}\right). $ \hfill$\Box$ \bigskip Collecting the results we have \begin{theorem}\label{autog12nh2} The automorphism group of $\mathbb{G}(1,2n)\cap L$ is an extention of $\mathbb{P}\mathrm{GL}(2,\mathbb{C})$ by the semi direct product $\mathbb{C}^{2n}\ltimes\mathbb{C}^*$. \medskip It is isomorphic to the matrix subgroup of $\,\mathbb{P}\mathrm{GL}(2n+1,\mathbb{C})$ given by \[ \left(\begin{array}{cc} \alpha \mathrm{E}_n &\! 0\\ S & \!\mathrm{E}_{n+1} \end{array}\right) \left(\begin{array}{cc} {}^tt_n^{-1} &\! 0\\ 0 & \!t_{n+1} \end{array}\right) \] where $\alpha\in\mathbb{C}^*$, $S\in\mathrm{M}((n+1)\times n,\mathbb{C})$ with $s_{i j}=s_{kl}$ for $i+j=k+l$ and $t_n\in\mathrm{Aut}(h,\mathbb{P}_{n-1})$ resp. $t_{n+1}\in\mathrm{Aut}(c,\mathbb{P}_n)$ are the transformations that are induced by the $\mathbb{P}\mathrm{GL}(2,\mathbb{C})$ action on the rational normal curve $h\subset\mathbb{P}_{n-1}$ resp. $c\subset\mathbb{P}_n$. \end{theorem} \emph{Proof.} It remains to show that the automorphism fixing the center curve pointwise form a normal subgroup, but that can be easily computed.\hfill$\Box$ \bigskip \begin{remark}\label{centerdetermine} An automorphism of $\mathbb{G}(1,2n)\cap L$ is determined by its action on the lines intersecting the center curve. \end{remark} \medskip In contrast to that, the line system, i.e.\ the position of the line $L^*\subset \mathbb{P}(\bigwedge^2\mathbb{C}^{2n+1})^*$, is not determined by these lines, as a simple dimension count shows. Giving these lines is equivalent to giving the two rational curves $c\subset\mathbb{P}_{2n}$ and $h\subset \mathbb{P}_{2n}/P\cong\mathbb{P}_{n-1}$ and a correspondence between them, so that we have the following dimension count \[ (2(2n+1)-4)+(2n-4) +3 < \dim \mathbb{G}(1,\mathbb{P}({\textstyle \bigwedge}^2\mathbb{C}^{2n+1}))= 2\left({ 2n+1 \choose 2}-2\right). \] \emph{Proof of the remark.} We need to show that only the identity fixes these lines one by one. First a transformation $T$ that fixes the lines must fix the center curve, hence by the Lemma \ref{centerfix} it is of the form \[ T= \left(\begin{array}{cc} \alpha \mathrm{E}_n &\! 0\\ S & \!\mathrm{E}_{n+1} \end{array}\right). \] We compute the induced action $\widetilde{T}$ of $T$ on $\{ l\in\mathbb{G}(1,2n)\cap L\mid c_{(0:1)} \in l\}$, the $\mathbb{P}_{2n-2}$ of lines through $c_{(0:1)}=e_n$. A line $l\in\mathbb{G}(1,2n)$ through $c_{(0:1)}$ will be in the line system $\mathbb{G}(1,2n)\cap L$ iff it lies in the hyperplane $h_{(0:1)}=\ker(1\,\colon 0\,\colon \ldots\,\colon 0)$. Therefore the $\mathbb{P}_{2n-2}$ of lines through $e_n$ is given by \[ e_n\wedge x \ \quad \mbox{with\ }x\in\mathbb{P}(\mathrm{span}\,\{e_1,\ldots,e_{n-1},e_{n+1},\ldots,e_{2n}\}). \] Using $(e_1\wedge e_n,\ldots,e_{n-1}\wedge e_n,e_{n+1}\wedge e_n,\ldots,e_{2n}\wedge e_n)$ as a basis, the induced action $\widetilde{T}$ is \[ \widetilde{T}= \left(\begin{array}{cc} \alpha \mathrm{E}_{n-1} &\! 0\\[0.5ex] |\overline{S} & \!\mathrm{E}_n \end{array}\right). \] Here $|\overline{S}$ denotes the matrix $S$ with the first row and column deleted. In order to have $\widetilde{T}=\mathrm{E}_{2n-1}$, we must have $\alpha=1$ and $|\overline{S}=0$. \medskip The same computation for the lines through $c_{(1:0)}=e_{2n}$ yields $\alpha=1$ and $\underline{S}|=0$ from which $S=0$ and the remark follow. \hfill $\Box$ \bigskip For the rest of the section we analyze the action of the automorphism group on the line system $\mathbb{G}(1,2n)\cap L$. We start with $\mathbb{G}(1,4)\cap L$. \begin{proposition} The action of $\mathrm{Aut}( \mathbb{G}(1,4)\cap L)$ on the lines has four orbits: \begin{enumerate} \item tangents of the center conic \item secants of the center conic \item lines through the center conic that do not lie in the plane of the center conic \item lines that do not intersect the plane of the center curve. \end{enumerate} \end{proposition} \emph{Proof.} Since any automorphism maps the center conic onto itself, it is clear by the geometric description that all the mentioned lines lie in different orbits. \medskip Any line in the plane $P$ of the center conic intersects the conic twice, so by the Remark \ref{twocenteroncurve} it is a member of the line system. Since the automorphism group acts like $\mathrm{Aut}(c,P)\cong\mathbb{P}\mathrm{GL}(2,\mathbb{C})$ on the plane $P$, the first two orbits are obvious. \medskip To see that the lines of 3) form one orbit, we have to exhibit an automorphism that given two lines of 3) maps one onto the other. Since the $\mathbb{P}\mathrm{GL}(2,\mathbb{C})$ part of the automorphism group acts transitively on the center conic, we may assume that both lines pass through the same point of the center conic, say $e_2=c_{(0:1)}$. Now the induced action $\widetilde{T}$ of an automorphism $T$ fixing the center conic pointwise on the $\mathbb{P}_2$ of lines through $e_2$ was computed in the proof of the Remark \ref{centerdetermine} as \[ \widetilde{T}= \left(\begin{array}{ccc} \alpha & 0 & 0 \\ f & 1 & 0 \\ g & 0 & 1 \end{array} \right) \ \ \mbox{with\ }\alpha\in\mathbb{C}^*;\ f,g\in \mathbb{C}. \] These transformations act transitively on $\mathbb{P}_2\setminus\mathbb{P}(\mathrm{span}\,\{\widetilde{e_1},\widetilde{e_2}\})$, where the line $\mathbb{P}(\mathrm{span}\,\{\widetilde{e_1},\widetilde{e_2}\})$ corresponds to the lines through $e_2$ that lie in the plane of the center conic. \medskip The lines of 4) are all the remaining lines since there are no lines that intersect the plane $P$ of the center conic but not the conic $c$ itself. This is clear because the $\mathbb{P}_1$ of lines through a point $p\in P\setminus c$ is formed by the lines through $p$ in the plane $P$, so there can be no other line. \medskip Finally, we have to show that the lines of 4) form one orbit. By a small computation one checks that $\mathrm{Aut}(\mathbb{G}(1,4)\cap L)\subset \mathbb{P}\mathrm{GL}(5,\mathbb{C})$ acts transitively on $\mathbb{P}_4\setminus P$. So, it suffices to show that the line $e_0\wedge e_1$ can be mapped to any other line through $e_0$ by an automorphism. Any of these lines can be written as \[ e_0\wedge (e_1+\beta e_4) \quad \ \mbox{with\ }\beta \in \mathbb{C}, \] and the automorphism \[ T= \left( \begin{array}{cc|ccc} 1 & 0 & 0 &0 &0 \\ 0 & 1 &0 & 0 &0 \\ \hline 0 & 0 & 1 & 0 &0 \\ 0 & 0 &0 & 1 &0 \\ 0 & \beta &0 & 0 &1 \end{array}\right) \] will take $e_0\wedge e_1$ to it. \hfill $\Box$ \begin{proposition} The automorphism group acts quasihomogeneously on $\mathbb{G}(1,6)\cap H^2$. \end{proposition} \emph{Proof.} For this it is enough to show that the stabiliser of the line $l=e_0\wedge e_2$ is a 2-dimensional subgroup since then \[ \begin{array}{c@{\,=\,}l} \dim \mathrm{Orbit}(l) & \dim \mathrm{Aut} (\mathbb{G}(1,6)\cap H^2) -\dim \mathrm{Stab}(l) =10-2=8\\ & \dim \mathbb{G}(1,6)\cap H^2. \end{array} \] If we normalize the $t\in\mathbb{P}\mathrm{GL}(2,\mathbb{C})$ by $\det t=1$, every $T\in \mathrm{Aut} (\mathbb{G}(1,6)\cap H^2)$ can be written by the Theorem \ref{autog12nh2} as \[ \begin{array}{l} T= \left( \begin{array}{ccc|cccc} \alpha & &0 & & & & \\ & \alpha & & & \raiseghost{3.5ex}{-1ex}{\Huge 0} & & \\ 0 & &\alpha & & & & \\ \hline e & f & g & 1& & \raiseghost{3ex}{-2ex}{\LARGE 0} & \\ f & g & h & & 1& & \\ g & h & i & & & 1& \\ h & i & j & \raiseghost{3.5ex}{0.5ex}{\LARGE 0} & & & 1 \end{array} \right) \left(\begin{array}{cc} {}^tt_3^{-1} &\! 0\\ 0 & \!t_4 \end{array}\right)\\[11ex] \ \ \mbox{with}\ {}^tt_3^{-1}= \left( \begin{array}{ccc} a^2 & -ab &b^2 \\ -2ac& ab+cd &-2bd \\ c^2 & -cd &d^2 \end{array} \right). \end{array} \] To compute the stabilizer we start by looking only at the first three entries of \[ \begin{array}{l} Te_0=(\alpha a^2 , -2\alpha ac, \alpha c^2,\ldots)\\[1ex] Te_2=(\alpha b^2 , -2\alpha bd, \alpha d^2,\ldots). \end{array} \] Since we must have $Te_0,Te_2\in l$, $ac=bd=0$ follows. By $\det t=ad-bc=1$ we have the two possibilities $b=c=0$, $d=a^{-1}$ and $a=d=0$, $c=-b^{-1}$. We examine only the first case, the second being similar. Now we have \[ \begin{array}{l} Te_0=a^2(\alpha , 0, 0 ,e,f,g,h)\\[1ex] Te_2=a^2(0,0,\alpha,g,h,i,j). \end{array} \] From $Te_0,Te_2\in l$ we conclude $e=f=g=h=0$ resp. $g=h=i=j=0$. Therefore, including the case $(a=d=0,\ c=-b^{-1})$, the stabilizer of $l$ is \bigskip \hfill \mathrm{Stab}(l)=\left\{ \left( \begin{array}{c@{\,}c@{\,}c@{\,}c@{\,}c@{\,}c@{\,}c} \alpha a^2\! & & & & & & \\ &0 & & & &\raiseghost{-1ex}{-1ex}{\Huge 0} & \\ & & \alpha a^{-2}\! & & & & \\ & & &a^{-3} & & & \\ & & & &\!\!a^{-1}\, & & \\ &\raiseghost{-1ex}{0ex}{\Huge 0} & & & &\!\!\!\!\!a & \\ & & & & & &a^3 \end{array} \right), \left( \begin{array}{c@{\,}c@{\,}c@{\,}c@{\,}c@{\,}c@{\,}c} \!\!0 & &\!\alpha b^2 & & & & \\ &0 & & & &\raiseghost{0ex}{-0.5ex}{\LARGE 0} & \\ \!\!\alpha b^{-2} & &0 & & & & \\ & & & & & &\!\!\!\!\!-b^{-3} \\ & & & & &\;b^{-1}\! & \\ &\raiseghost{0ex}{1ex}{\LARGE 0} & & &\!\!\!-b & & \\ & & &\!b^3 & & &\raiseghost{0ex}{1ex}{\Large 0} \end{array} \right)\right\}.$ \hfill $\Box$ \bigskip \begin{proposition} For $n\ge4$ the action of the automorphism group on $\mathbb{G}(1,2n)\cap H^2$ is not quasihomogeneous. \end{proposition} \emph{Proof.} We project the $\mathbb{P}_{2n}$ from the space $P$ of the center curve onto $\mathbb{P}_{2n}/P\cong \mathbb{P}_{n-1}$. This projects the lines of $\mathbb{G}(1,2n)\cap H^2$ not intersecting $P$ surjectively onto the lines $\mathbb{G}(1,\mathbb{P}_{2n}/P)$ of $\mathbb{P}_{2n}/P$. The automorphisms of $\mathbb{G}(1,2n)\cap H^2$ induce automorphisms of $\mathbb{P}_{2n}/P$. As matrices these are the upper left $n\times n$ matrices of the matrices of Theorem \ref{autog12nh2}, i.e.\ they are of the form ${}^tt_n^{-1}$. So, as a group this induced automorphism group is isomorphic to $\mathrm{Aut}(h^*,\mathbb{P}_{2n}/P)\cong \mathbb{P}\mathrm{GL}(2,\mathbb{C})$. If $\mathrm{Aut}(\mathbb{G}(1,2n)\cap H^2)$ acts quasihomogeneously, then $\mathrm{Aut}(h^*,\mathbb{P}_{2n}/P)$ would have to act quasihomogeneously on $\mathbb{G}(1,\mathbb{P}_{2n}/P)\cong\mathbb{G}(1,n-1)$, but this contradicts \bigskip \ \hfill $\displaystyle \dim \mathbb{P}\mathrm{GL}(2,\mathbb{C})=3<\dim \mathbb{G}(1,n-1)=2n-4.$ \hfill $\Box$ \section{$\mathbb{G}(1,4) \cap H^3$} Let $L=H^3$ be a general 3-codimensional subspace of $\mathbb{P}(\bigwedge\nolimits^2 \mathbb{C}^5) \cong \mathbb{P}_9$. To $L$ corresponds the plane $L^*=\mathbb{P}(\lambda A+\mu B+\nu C) \subseteq \mathbb{P}(\bigwedge\nolimits^2\mathbb{C}^5)^*$ of hyperplanes containing $L$. Since the locus of antisymmetric matrices of corank 3 is 3-codimensional in $\mathbb{P}(\bigwedge\nolimits^2 \mathbb{C}^5)^*$ by Corollary \ref{dualG}, $L^*$ does not contain any. Therefore to each of the hyperplanes $H_{(\lambda:\mu:\nu)}\subset L$ corresponds a unique center $c_{(\lambda:\mu:\nu)} \in \mathbb{P}_4$. In complete analogy to the last case we get \begin{lemma} \label{centerofg14} The centers $c_{(\lambda:\mu:\nu)}$ are those points of $\mathbb{P}_4$ through which there passes a $\mathbb{P}_1$ of lines of the line system $\mathbb{G}(1,4) \cap L$. Through all the other points passes a unique line. \end{lemma} \begin{proposition}\label{centervero} The map of centers \[ \begin{array}{rccc} c: & L^* \cong \mathbb{P}_2 & \longrightarrow & \mathbb{P}_4\\[0.5ex] & (\lambda:\mu:\nu) & \longmapsto & c_{(\lambda:\mu:\nu)}=\ker (\lambda A+\mu B+\nu C) \end{array} \] is an embedding of $\mathbb{P}_2$ in $\mathbb{P}_4$ of degree 2, i.e. its image is a smooth projected Veronese surface. \end{proposition} \begin{remark} Any line that contains three centers is in the line system. \end{remark} \emph{Proof.} Let $c(p_0)$, $c(p_1)$ and $c(p_2)$ with $p_0,p_1,p_2 \in L^*$ be the three centers on the line $l$. By the definition of the centers we have $l\in H_{p_i}$. Since $c$ maps lines in $L^*$ onto conics in $\mathbb{P}_4$, the three points $p_0,p_1,p_2$ do not lie on a line, hence they span $L^*$. So $ l\in H_{p_0} \cap H_{p_1} \cap H_{p_2} =L$. \hfill $\Box$ \bigskip From the statements we get a complete picture of the lines of $\mathbb{G}(1,4) \cap H^3$ in $\mathbb{P}_4$. We define the trisecant variety $\mathrm{Tri}(X)$ of a variety $X\subseteq \mathbb{P}_N$ by: \[ \mathrm{Tri}(X):=\overline{\{ l\in\mathbb{G}(1,N) \mid \#(X \cap l)\ge 3 \}} \] Then we have \begin{corollary}[Castelnuovo] $\mathbb{G}(1,4) \cap L$ is the trisecant variety of the smooth projected Veronese surface $\mathrm{Im}\, c \subset \mathbb{P}_4$. \end{corollary} \emph{Proof.}(see \cite{C} or \cite[X, 4.4]{SR}) By the remark above the trisecant variety is contained in the irreducible variety $\mathbb{G}(1,4) \cap L$. So it is enough to show that both varieties have the same dimension. The Lemma \ref{centerofg14} together with the Proposition \ref{centervero} shows that there is an unique line of $\mathbb{G}(1,4) \cap L$ through a general point of $\mathbb{P}_4$ and that $\mathbb{G}(1,4) \cap L$ is the closure of such lines. The same statement for the trisecant variety is classical \cite[VII,3.2]{SR}. Hence both varieties have dimension three. \hfill $\Box$ \bigskip The general trisecant intersects the projected Veronese surface $\mathrm{Im}\, c$ in three different points. Their inverse image under $c$ are triples of points in $\mathbb{P}_2$ that have to fulfill some conditions since there is only a 3-dimensional family of these triples. To see what these conditions are, we recall some facts about the Veronese surface \cite{H}. \medskip The Veronese surface $V$ is the image of the embedding \[ \begin{array}{rccc} \upsilon :& \mathbb{P}_2=\mathbb{P}(\mathbb{C}^3) & \longrightarrow & \mathbb{P}(\mathrm{Sym}^2 \mathbb{C}^3)\\[0.5ex] & \mathbb{P}(v) & \longmapsto & \mathbb{P}( v\cdot v)\, . \end{array} \] Its secant variety consists of the points of $\mathbb{P}(\mathrm{Sym}^2 \mathbb{C}^3)$ that are the product of two vectors of $\mathbb{C}^3$, \[ \sec(V)=\left\{ \mathbb{P}(v \cdot w) \mid v,w\in \mathbb{C}^3\setminus \{0\}\right\}.\] The projected Veronese surface will be smooth -- like in our case -- iff the center of projection $P$ is not in the secant variety. \bigskip An intersection of the Veronese surface $V$ with a hyperplane $H\in \mathbb{P}(\mathrm{Sym}^2 \mathbb{C}^3)^*$ gives the conic $\upsilon^{-1}(V\cap H)\subset \mathbb{P}_2$ which is described by the equation $H$ if we identify $\mathbb{P}(\mathrm{Sym}^2 \mathbb{C}^3)^*$ with the polynomials of degree 2 modulo $\mathbb{C}^*$. The conics that we get as hyperplane sections of the projected Veronese surface are precisely the conics that we get as hyperplane sections of the Veronese surface by hyperplanes that contain the projection center $P$. So these conics fulfill one linear condition given by $P$. We view $P\in \mathbb{P}(\mathrm{Sym}^2 \mathbb{C}^3)$ as a conic $C^*_P$ in $\mathbb{P}_2^*$. Since $P$ does not lie in the secant variety of the Veronese surface, $P$ is not the product of two elements of $\mathbb{C}^3$, hence $C^*_P$ is not the union of two lines. Therefore it is smooth. We denote the dual conic of $C^*_P$ by $C_P\subset \mathbb{P}_2$. \medskip Now, three different points of the projected Veronese surface $\mathrm{Im}\, c\subset \mathbb{P}_4$ lie on a line, the trisecant, iff any hyperplane that contains two of them contains all three. Under the inverse of the embedding $c$ that means the following on the $\mathbb{P}_2$: \medskip Three different points of $\mathbb{P}_2$ are the inverse image $c^{-1}(l)$ of a trisecant $l$ of the projected Veronese surface $\mathrm{Im}\, c$ iff all conics that fulfill the linear condition given by $P$ (or equivalently by $C_P$) and pass through two of the points pass through all three of them. \medskip The propositions in the appendix tell us that these triples of points are the vertices of the non-degenerated polar triangles of the conic $C_P$. By a continuity argument the trisecants that are tangent to $\mathrm{Im}\, c$ at one point and intersect it in another correspond to the degenerated polar triangles, and the trisecants that intersect $\mathrm{Im}\, c$ in only one point ``with multiplicity three'' correspond to a triple point on the conic $C_P$. We also see that there are no 4-secants. Since if there is one, there would be four points in $\mathbb{P}_2$ such that any three of them build a different polar triangle. But this is impossible because a polar triangle is already determined by two of its vertices. \bigskip With this geometric description it is easy to compute the automorphism group of $\mathbb{G}(1,4)\cap H^3$. Any automorphism as a projective linear transformation of $\mathbb{P}_4$ maps by definition the trisecants of the projected Veronese surface $\mathrm{Im}\, c$ onto themselves. Further, it must fix the projected Veronese surface $\mathrm{Im}\, c$, since $\mathrm{Im}\, c$ is the union of the centers. In fact, the automorphism is already determined by its action on $\mathrm{Im}\, c$ since this action on $\mathrm{Im}\, c$ determines the images on the trisecants. Under the inverse of the embedding $c$, this automorphism of $\mathrm{Im}\, c$ preserving the trisecants corresponds to an automorphism of $\mathbb{P}_2$ preserving the polar triangles of the conic $C_P \subset \mathbb{P}_2$. Such an automorphism of $\mathbb{P}_2$ maps the degenerated polar triangles onto themselves. In particular, it maps the tangents to the conic $C_P$ onto themselves. Therefore it has to fix the conic $C_P$. \medskip So we have seen how an automorphism of $\mathbb{G}(1,4)\cap L$ induces an unique automorphism of $\mathbb{P}_2$ fixing $C_P$, hence an automorphism of $C_P$ since $\mathrm{Aut}(C_P,\mathbb{P}_2)\cong \mathrm{Aut}(C_P)\cong \mathbb{P} \mathrm{GL}(2,\mathbb{C})$. \medskip On the other hand, any projective linear transformation of $\mathbb{P}_2$ that fixes the conic $C_P$ preserves the polar triangles of $C_P$. Therefore it induces via the embedding $c$ an automorphism of the projected Veronese surface $\mathrm{Im}\, c$ that preserves triples of points that lie on a line. So it defines an automorphism of the trisecants of $\mathrm{Im}\, c$, which is the same as an automorphism of $\mathbb{G}(1,4)\cap L$. \medskip We summarize: \begin{theorem} The automorphism group of $\mathbb{G}(1,4)\cap H^3$ is isomorphic to $\mathbb{P} \mathrm{GL} (2,\mathbb{C})$. \end{theorem} The description of the orbits of this automorphism group follows immediately. \begin{proposition} The action of $\mathrm{Aut} ( \mathbb{G}(1,4)\cap H^3 )$ on the linear system $\mathbb{G}(1,4)\cap H^3$ has three orbits: \begin{enumerate} \item trisecants of the projected Veronese surface that intersect it in three points \item trisecants that are tangent to the projected Veronese surface at one point and intersect it in another \item trisecants that intersect the projected Veronese surface in only one point ``with multiplicity three''. \end{enumerate} \end{proposition} \emph{Proof.} By what was said above, this is equivalent to the classical statement that the action of group $\mathrm{Aut} (C_P,\mathbb{P}_2)$ on the polar triangles has three orbits: the non-degenerated triangles, the degenerated ones and the triple points on $C_P$.\hfill$\Box$ \section{Appendix: Polar Triangles} Here we prove the needed propositions about polar triangles. The whole appendix may be seen as a modern exposition of \cite[348]{SF}. First we recall the basic definitions. \medskip Let $C_A$ be a smooth conic in $\mathbb{P}_2$, which is given by the quadratic equation ${}^t\!x A x=0$, where $A\in \mathrm{GL}(3,\mathbb{C})$ is a symmetric, invertible matrix. Then $C_A$ induces a polarity $P$ by \[\begin{array}{rccc} P:&\mathbb{P}_2 & \longrightarrow & \mathbb{P}_2^* \\[0.5ex] & \mathbb{P}(x) & \longmapsto & \mathbb{P}({}^t\!x A)\, . \end{array} \] For a point $p\in \mathbb{P}_2$ the line $P(p)$ is called the polar of $p$ and $p$ the pole of $P(p)$. \medskip A polar triangle is given by three points, at least two of which are different, such that the polar of each point contains the other two points, i.e. $(p,q,r)$ is a polar triangle if ${}^t\!p A q={}^t\!q A r={}^t\!r A p=0$. The sides of the triangle are the polars of the points. In the non-degenerated case when all three points are different, the three points cannot lie on a line and therefore span the whole $\mathbb{P}_2$. In the degenerated case, $(p,p,q)$, $p$ lies on the conic and $q$ on the tangent to the conic at the point $p$. The sides are the polar of $q$ and twice the tangent. \bigskip \parbox[b]{5.8cm}{ \mbox{ \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 <0.750000cm,0.7500000cm> \unitlength=0.750000cm \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) \setshadesymbol ({\thinlinefont .}) \setlinear \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \ellipticalarc axes ratio 2.697:2.697 360 degrees from 5.554 21.907 center at 2.857 21.907 \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 3.810 22.860 4.763 25.718 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 3.810 22.860 6.668 21.907 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 4.763 25.718 6.668 21.907 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 6.668 21.907 7.144 20.955 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 4.286 26.670 4.763 25.718 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 4.763 25.718 5.048 26.670 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 6.668 21.907 7.525 21.622 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 3.810 22.860 2.953 23.146 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 3.810 22.860 3.524 22.003 / \linethickness=0pt \putrectangle corners at 0.144 26.695 and 7.550 19.196 \endpicture} }\\ Non-degenerated polar triangle} \hfill \parbox[b]{6.38cm}{ \mbox{ \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 <1.06066cm,1.06066cm> \unitlength=1.06066cm \linethickness=1pt \setplotsymbol ({\makebox(0,0)[l]{\tencirc\symbol{'160}}}) \setshadesymbol ({\thinlinefont .}) \setlinear \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \ellipticalarc axes ratio 1.905:1.905 360 degrees from 4.763 22.860 center at 2.857 22.860 \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \put{\makebox(0,0)[l]{\circle*{ 0.1}}} at 5.400 24.765 \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \putrule from 0.953 24.765 to 6.0 24.765 \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 2.857 24.765 5.133 21.738 / \linethickness= 0.500pt \setplotsymbol ({\thinlinefont .}) \plot 2.857 24.765 2.381 25.400 / \linethickness=0pt \putrectangle corners at 0.927 25.425 and 6.693 20.940 \endpicture} }\\ Degenerated polar triangle} { \begin{proposition} Let $C_A=\{x\in \mathbb{P}_2 \mid {}^t\!x A x=0\}$ be a smooth conic and $C_A^*=\{x\in \mathbb{P}_2^* \mid {}^t\!x A^{-1} x=0\}$ its dual conic. Further, let $C_B=\{x\in \mathbb{P}_2 \mid {}^t\!x B x=0\}$ be a conic such that \bigskip \ \hfill $\displaystyle \sum_{i,j=0}^2 a^{ij} b_{ij}=0$\hfill $(*)$ \bigskip \noindent where $A^{-1}=(a^{ij}),B=(b_{ij})\in \mathrm{Sym}(3,\mathbb{C})$. Finally, let $(p,q,r)$ be a polar triangle of $C_A$ then: \medskip If two of the three points $p,q,r$ lie on the conic $C_B$, then also the third. \end{proposition} In the case of a degenerate polar triangle, $(p,p,q)$, the condition that $C_B$ contains $p$ twice means that $C_B$ contains $p$ and $C_B$ is either singular in $p$ or its tangent in $b$ is the polar of $q$. \medskip \noindent\emph{Proof.} One can show that all the properties in the statement of the proposition are independent of the choice of coordinates, so we may pick nice ones. We have to distinguish between the two cases of the polar triangle being degenerated or not. We treat the case of the non-degenerated polar triangle first. \medskip By a suitable choice of coordinates we may assume that $p=(1\,\colon 0\,\colon 0)$, $q=(0\,\colon 1\,\colon 0)$ and $r=(0\,\colon 0\,\colon 1)$. Then the assumption that $(p,q,r)$ is a polar triangle of $C_A$ translates into \[ {}^t\!p A q={}^t\!q A r={}^t\!r A p=0 \Longleftrightarrow a_{0 1}=a_{1 2}=a_{0 2}=0 . \] By a scaling of the coordinates we can achieve that $ a_{0 0}=a_{1 1}=a_{2 2}=1$, so that $C_A=\{x\in \mathbb{P}_2 \mid x_0^2+x_1^2+x_2^2=0\}$. Then the condition $(*)$ reads $ b_{0 0}+b_{1 1}+b_{2 2}=0$. If the two points $p$ and $q$ are on the conic $C_B$, we have ${}^t\!p B p =b_{0 0}=0$ and ${}^t\!q B q =b_{1 1}=0$. By $(*)$ we see $0=b_{2 2}= {}^t\!r B r$, i.e.\ the third point $r$ lies also on the conic $C_B$. \medskip Now we treat the case of the degenerated polar triangle $(p,p,q)$. We choose coordinates such that $C_A=\{ x\in \mathbb{P}_2 \mid x_0^2+x_1^2+x_2^2=0\}$ and $p=(1\,\colon i\,\colon 0)$. The point $q\neq p$ must lie on the tangent to $C_A$. So it has coordinates $q=(\lambda\,\colon i\lambda\,\colon 1)$, and its polar is spanned by $p$ and $(1\,\colon 0\,\colon -\lambda)$. Now using the assumptions \bigskip \hspace{3cm}$\displaystyle b_{0 0}+b_{1 1}+b_{2 2}=0$ \hfill $(*)$ \medskip \hspace{3cm}$\displaystyle p\in C_B \Longleftrightarrow {}^t\!p B p=0 \Longleftrightarrow b_{0 0}+2 i b_{0 1} -b_{1 1}=0, $ \hfill $(**)$ \bigskip \noindent we have to show \[ q\in B \Longleftrightarrow C_B \mbox{\ singular\ in\ }p \ \ \mbox{\ or\ } \ \ \mathbb{T}_pC_B=\; \mbox{polar\ of\ }q. \] We rewrite this as \[ {}^t\!q B q= {\vphantom{ \left({\textstyle \begin{array}{c}\lambda\\ i\lambda \\1 \end{array}}\right)} }^t\!\!\!\! \left({\textstyle \begin{array}{c}\lambda\\ i\lambda \\1 \end{array}}\right) \! B \! \left({\textstyle \begin{array}{c}\lambda\\ i\lambda \\1 \end{array}}\right) =0 \Longleftrightarrow {}^t\! p B \! \left({\textstyle \begin{array}{c}1 \\ 0 \\ \!\! -\lambda \end{array}}\right)= {\vphantom{ \left(\begin{array}{c}1\\ i \\1 \end{array}\right)} }^t\!\!\!\! \left(\begin{array}{c}1\\ i \\0 \end{array}\right) \! B\! \left(\begin{array}{c}1 \\ 0 \\\!\!-\lambda \end{array}\right)=0. \] But this is true since $-2$ times the left hand side plus $(\lambda^2+1)$ times $(**)$ plus $(*)$ gives the right hand side. \hfill $\Box$ \bigskip Now we prove the converse of the last proposition. \begin{proposition} Given a smooth conic $C_A=\{x\in \mathbb{P}_2 \mid {}^t\!x A x=0\}$ \[{\cal B}:= \left\{ C_B=\{x\in \mathbb{P}_2 \mid {}^t\!x B x=0\} \left| \sum_{i,j=0}^2 a^{ij} b_{ij}=0\right.\right\} \] is a four dimensional family of conics. Let $p,q,r\in \mathbb{P}_2$ be three points, at least two of which are different, with the property that if two of them lie on a conic $C_B\in {\cal B}$ then also the third. \medskip Then $(p,q,r)$ is a polar triangle of $C_A$. \end{proposition} \emph{Proof.} We will show that if $(p,q,r)$ is not a polar triangle then there exits a $C_A\in {\cal B}$ for which this property is violated. We have to treat several cases. \medskip First let the three points be all different, then they cannot lie on a line. Because if they would, the conics in the at least two dimensional family \[ {\cal B}_{p,q}:=\{ C_B \in {\cal B}\mid p,q\in C_B\} \] of conics of ${\cal B}$ passing through $p$ and $q$ must split off the line through the three points. If we pick coordinates such that this line is given by $\{x_2=0\}$, then ${\cal B}_{p,q}$ must be \[ {\cal B}_{p,q} =\left\{ \mathrm{V}( x_2(\lambda_0 x_0+\lambda_1 x_1+\lambda_2 x_2)) \mid (\lambda_0\,\colon \lambda_1\,\colon \lambda_2)\in \mathbb{P}_2 \right\}.\] This means that the conics $C_B$ with the matrices \[ B=\left({\textstyle \begin{array}{ccc} 0 & 0 & b_{0 2}\\ 0 & 0 & b_{1 2}\\ b_{0 2} & b_{1 2} & b_{2 2} \end{array}}\right) \quad \mbox{with} \ b_{0 2}, b_{1 2}, b_{2 2}\in \mathbb{C} \] are all in ${\cal B}$. Hence the matrix $A^{-1}$ must be of the type \[ A^{-1}=\left({\textstyle \begin{array}{ccc} a^{00} & a^{01} & 0 \\ a^{01} & a^{11} & 0 \\ 0 & 0 & 0 \end{array}}\right) \quad \mbox{with} \ a^{0 0}, a^{0 1}, a^{1 1}\in \mathbb{C}, \] but this contradicts the invertibility of $A^{-1}$. \medskip Now since $p,q,r$ span the $\mathbb{P}_2$ we may pick coordinates such that $p=(1\,\colon 0\,\colon 0)$, $q=(0\,\colon 1\,\colon 0)$ and $r=(0\,\colon 0\,\colon 1)$. That $(p,q,r)$ is not a polar triangle of $C_A$ means that ${}^t\!p A q=a_{01}\not= 0$, ${}^t\!q A r=a_{12}\not= 0$ or ${}^t\!r A p=a_{02}\not= 0$. Assuming $\det A=1$ we conclude that not all of the $a^{0 1}=a_{0 2}a_{1 2}-a_{0 1}a_{2 2}$, $a^{0 2}=a_{0 1}a_{1 2}-a_{0 2}a_{1 1}$ and $a^{1 2}=a_{0 1}a_{0 2}-a_{0 0}a_{1 2}$ can be zero. If for example $a^{0 2}\not= 0$, then \[ B=\left({\textstyle \begin{array}{ccc} 0 & 0 & - a^{1 1}\\ 0 & 2 a^{0 2} & 0\\ - a^{1 1} & 0 & 0 \end{array}}\right)\] gives a conic $C_B\in {\cal B}$ that contains the points $p$ and $q$, but not $r$. \medskip Now let us look at the case where two of the points $p,q,r$ are the same. The points $(p,p,q)$ will not form a polar triangle if ${}^t\!p A p\not= 0$ or ${}^t\!p A q \not= 0$. \medskip For the case ${}^t\!p A p\not= 0$ we pick coordinates such that $A$ is the identity matrix and $p=(1\,\colon 0\,\colon 0)$. Let \[ B=\left\{ \begin{array}{ll} \left({\textstyle \begin{array}{ccc} 0 & \!\!\!-q_2 & q_1\\ \!\!\!-q_2 & 0 & 0\\ q_1 & 0 & 0 \end{array}}\right) & \mbox{for\ } q_0\not=0 \mbox{\ or\ }q_1^2+q_2^2\not=0\\[4.5ex] \left({\textstyle \begin{array}{ccc} 0 & 0 & 1\\ 0 & 1 & \!\!\pm i\\ 1 & \!\!\pm i & \!-1 \end{array}}\right) & \mbox{for\ } q=(0\,\colon 1\,\colon \pm i), \end{array} \right. \] then $C_B$ is a conic of ${\cal B}$ that contains $p$ and $q$, but is smooth in $p$, and its tangent in $p$ is not the polar of $q$, so it does not contain $p$ twice. \medskip Finally, if ${}^t\! p A p=0$ and ${}^t\! p A q \not=0$, we pick coordinates such that $A$ is the identity matrix and $p=(1\,\colon i\,\colon 0)$. Let \[ B=\left\{ \begin{array}{ll} \left({\textstyle \begin{array}{ccc} -2i & 2 & \!\!\! iq_0-2q_1\\ 2 & 2i & -iq_1\\ \!\! iq_0-2q_1 &\!\! -iq_1 & 0 \end{array}}\right) & \mbox{for\ } q=(q_0\,\colon q_1\,\colon 1) \\[4.5ex] \left({\textstyle \begin{array}{ccc} 0 & 0 & 1\\ 0 & 0 & 0\\ 1 & 0 & 0 \end{array}}\right) & \mbox{for\ } q_2=0, \end{array} \right. \] then we are in the same situation as above. \hfill $\Box$
\section{Introduction} Naturalness is the only problem of the SM that requires new physics at energies $\Lambda$ accessible to accelerators, $\Lambda\circa{<} 4\pi M_Z$. Supersymmetry (SUSY) is one possible solution to the problem: in this case the scale of new physics is given by the mass of the supersymmetric particles (`sparticles'), that are expected to be just around the $Z$-boson: $m_{\rm SUSY}=(\frac{1}{3}\div 3)M_Z$. However experiments now say that most of the sparticles must be heavier than the $Z$ boson. It is thus interesting to studying if the existing `conventional' supersymmetric models (supergravity and gauge mediation) prefer a chargino mass $M_\chi$ around $\sim 30\,{\rm GeV}$ or $\sim 300\,{\rm GeV}$. We would like to discuss the following questions: \cQ{ \begin{quotation} Experiments are saying that naturalness is a problem for supersymmetry? If yes, can we found a solution and learn something about the sparticle spectrum? \end{quotation}} {$\bullet$ Experiments are saying that naturalness is a problem for supersymmetry? $\bullet$ If yes, can we found a solution and learn something about the sparticle spectrum?} \section{Naturalness and fine-tuning} Usually questions of this kind are discussed using fine tuning (`FT')\cite{FT} as a quantitative measure of naturalness. In a supersymmetric model the $Z$ boson mass squared, $M_Z^2(\wp)$, can be computed as function of the parameters $\wp$ ($\mu$-term, gaugino masses, etc) of the model, by minimizing the potential. The FT (for example with respect to the $\mu$ term) is defined as \cQ{$${\rm FT}(\mu)\equiv \frac{\mu^2}{M_Z^2}\frac{\partial M_Z^2}{\partial \mu^2}$$} {${\rm FT}(\mu)\equiv (\mu^2/M_Z^2)(\partial M_Z^2/\partial \mu^2)$} and tells how sensitive is the $Z$-mass with respect to variation of the parameters. The result of a FT analysis in `minimal' supergravity (`mSuGra', sometimes also called `constrained' suGra) MSSM is the following\cite{FTLEP}: the most recent bounds require a FT greater than about 6. Is this an answer to our questions? We believe not, due to the following reasons \cQ{\begin{enumerate}}{} \cQ{\item}{(1.)} Replacing $M_Z^2\to M_Z$ in the definition of the FT would reduce the minimal FT from 6 to 3. \cQ{\item}{(2.)} Replacing $\mu^2\to \mu$ would increase the minimal FT from 6 to 12. \cQ{\item}{(3.)} Should we worry if FT $> 5$, or 10, or 20, or\ldots? \cQ{\item}{(4.)} The FT can be large in some cases where nothing of unnatural happens\cite{FTcritica}. \cQ{\end{enumerate}}{} So, the normalization of the FT is arbitrary and setting an upper bound on the FT is a subjective choice. Summarizing, a FT analysis only gives the following not very interesting statement: \cQ{$$\hbox{ If we change the parameters of the mSuGra model by $1\%$, the $Z$ mass changes by more than 3\%.}$$} {``If we change the parameters of the mSuGra model by 1\% the $Z$ mass changes by more than 3\%''} that does not seem to indicate a problem for supersymmetry. \section{Naturalness and the allowed $\%$ of parameter space} Since the FT does not give a useful answer, we approach the naturalness problem in a more direct way\cite{nat}t. We will illustrate the question with various plots, all generated by the following procedure: we consider the model that we want to study and \begin{table}[t] \caption{How naturally 3 different models generate the EW scale, using 2 different ways of quantifying naturalness.\label{tab:exp}} \vspace{0.4cm} \begin{center} \begin{tabular}{c|ccc} &\special{color cmyk 1. 1. 0.3 0} ElectroWeak scale & Fine-tuning & acceptable \% of \special{color cmyk 0 0 0 1.}\\ &\special{color cmyk 1. 1. 0.3 0} $M_Z^2$ generated as & parameter & parameter space \special{color cmyk 0 0 0 1.}\\ \hline \special{color cmyk 0 1. 1. 0.5} Standard Model\special{color cmyk 0 0 0 1.} & $\sum \frac{\alpha}{4\pi}({\cal O}} \def\SU{{\rm SU}(\Lambda^2)+M^2)$&$\sim 10^{30}$&$\sim 10^{-30}$\\ \special{color cmyk 0 1. 1. 0.5} Techni-colour\special{color cmyk 0 0 0 1.} & $\Lambda^2 e^{-4\pi b/\alpha}$ & $\sim 50$~ & $\sim 1$\\ \special{color cmyk 0 1. 1. 0.5} MSSM (mSuGra)\special{color cmyk 0 0 0 1.} & $9 M_{1/2}^2-2\mu^2$ & $>6 $& $\sim 5\%$\\ \end{tabular} \end{center} \end{table} \begin{enumerate} \item We choose random values of ${\cal O}} \def\SU{{\rm SU}(1)$ for all the dimensionless ratios of soft parameters (like $\mu/M_{1/2}$); \item We do not impose how heavy are the sparticles; the overall mass scale is instead fixed by requiring that the the minimum of the one-loop effective potential $V(h)$ be at its physical value $M_Z=g_2\langle h\rangle=91.18\,{\rm GeV}$; \item Now everything is fixed and everything can be computed. Since we want to study naturalness we check if the obtained spectrum of sparticles is sufficiently heavy for satisfying all accelerator bounds; \item We repeat the previous steps many times and we compute $\%\equiv \frac{\hbox{number of acceptable spectra}}{\hbox{total number of spectra}}$. \end{enumerate} This Monte Carlo-like procedure computes how frequently numerical accidents can make the $Z$ boson sufficiently lighter than the unobserved supersymmetric particles. The main result is shown in table~1, where we compare how natural are different models according to the FT and to our $\%$. Both metods agree that the SM is perfectly unnatural in presence of high mass scales $M,\Lambda\circa{>}10^{16}\,{\rm GeV}$. Fine-tuning says that techni-colour models generate the $Z$ scale in an unnatural way\cite{FTcritica}, while this is not true and our $\%$ is of order one (concrete models would have various problems). {\em The FT does not give a clear answer about the naturalness of the mSuGra MSSM --- our procedure says that roughly $95\%$ of its parameter space (with properly broken EW symmetry) is now excluded.} Before analizing more models in more detail, we comment about the solidity of the procedure that we follow. It can be viewed as an efficient way of making plots with density of points {\em proportional} to 1/FT: thus our $\%$ does not depend on the largely subjective normalization of the FT. The $\%$ is normalized to be 100\% in absence of experimental bounds. So a small $\%$ has a clear meaning. Only {\em one aspect of the definition of the $\%$ is subjective\/}: in step~1 we must specify what we mean with `order-one'. In more general terms, we sample the parameter space according to some arbitrary distribution of probability $p(\wp)$. The results (like the allowed percentage of parameter space) depend on the choice of $p(\wp)$. Actually this is a standard situation in inferential statistics: for example an experimentalist that wants to report a lower bound on a neutrino mass $m_\nu$ needs to assume an arbitrary distribution of probability $p(m_\nu)$ (this assumption is explicit in the Bayesian approach\cite{Bayes}, where $p$ is called `prior distribution'; the situation is not better in the frequentistc approach, where different `confidence intervals' are obtained using different techniques). In conclusion the procedure that we will use \cQ{\begin{itemize}}{} \cQ{\item}{~~$\bullet$~~} is more intuitive (and simpler to compute) than FT; \cQ{\item}{~~$\bullet$~~} does not tell that natural situations are unnatural; \cQ{\item}{~~$\bullet$~~} does not tell that unnatural situations are natural (differently from our $\%$, the FT only looks at how the $Z$ mass is obtained: but other unnatural situations can be present, as we will exemplificate); \cQ{\item}{~~$\bullet$~~} is not more subjective than --- say --- a 90\% C.L. bound on a neutrino mass. \cQ{\end{itemize}}{ } In short, our procedure is a sort of Bayesian analysis with an unconventional use of the prior distribution: the only subjective aspect of our analysis is the usual one, that no one knows how to to avoid. So the real question is if this procedure is safe enough (i.e.\ is there a decent choice of $p(\wp)$ that transforms the 5\% in table~1 into 50\%?): we will discuss this question in the next section when we analyze concrete models. \section{The supersymmetric naturalness problem}\label{masses} In this section we discuss how serious is the supersymmetric naturalness problem in the various different motivated scenarios of SUSY breaking. We will not consider models with extra fields at low energy beyond the minimal ones present in the MSSM and we will assume that $R$ parity is conserved. To be conservative {\em we will consider as excluded only those spectra that violate the experimental bounds coming from direct searches\/} at accelerators. \subsection{A simple example}\label{example} We now try to illustrate the previous discussions with a simple and characteristic example. We consider the `most minimal' gauge mediation scenario with very heavy messengers (one $5\oplus\bar{5}$ multiplet of the unified gauge group SU(5) with mass $M_{\rm GM}=10^{15}\,{\rm GeV}$). We also assume that the unknown mechanism that generates the $\mu$-term does not give additional contributions to the other parameters of the higgs potential. Since the $B$ term vanishes at the messenger scale, we obtain moderately large values of $\tan\beta\sim20$. This model is a good example because it is not completely irrelevant and \begin{itemize} \item its spectrum is not very different from a typical supergravity spectrum and is the most natural possibility within `conventional' models. It has the lowest FT (we fix the unprecisely measured parameter $\lambda_t(M_{\rm GUT})$ at a small value (0.5) that alleviates the naturalness problem). \item it is simple because it has only two free parameters: the overall scale of gauge mediated soft terms and the $\mu$ term. The condition of correct electroweak breaking fixes the overall mass scale, and only one dimensionless parameter remains free. We choose it to be $\wp\equiv \mu/M_2$ (renormalized at low energy), where $M_2$ is the mass term of the $\SU(2)_L$ gaugino. \end{itemize} In figure~1a we plot, as a function of $\wp$, the lightest chargino mass in GeV. The dark gray regions where $\mu$ is too small or too large are always excluded from our analysis because the electroweak gauge symmetry cannot be properly broken. Values of $\wp$ shaded in \Orange light gray\special{color cmyk 0 0 0 1.}{} are excluded because some supersymmetric particle is too light (in this example the chargino gives the strongest bound). The chargino is heavier than its LEP2 bound only in small ranges of the parameter space at $\mu/M_2\approx 2.3$ (left unshaded in fig.~\ref{fig:example}) close to the region where EW symmetry breaking is not possible because $\mu$ is too large. {\em The fact that the allowed regions are very small and atypical is a naturalness problem for supersymmetry}. About $90\%$ of the acceptable values of $\wp$ are now excluded by the chargino mass bound. This percentage would not be significantly different if we had used $\mu^2$ in place of $\mu$ as a parameter; it could be significantly higher only choosing a rather ad-hoc parametrization. Indirect bounds (that we neglect) are not completely negligible: about half of the allowed parameter space of this example is excluded because the $b\to s\gamma$ decay is too fast. \begin{figure}[t] \caption[SP]{ Fig.~1a: the naturalness problem in a simple supersymmetric model: values of $\mu/M_2$ marked in dark gray are unphysical, while \Orange light gray\special{color cmyk 0 0 0 1.}{} regions have too light sparticles. Only the small white vertical band is now experimentally acceptable. Fig.~1b: Scatter plot of chargino mass vs higgs mass done with sampling density proportional to the naturalness probability in minimal supergravity. Only the \Green light gray\special{color cmyk 0 0 0 1.}{} sampling points are allowed. The shading covers excluded regions. Fig.~1c: naturalness distribution of chargino mass: allowed spectra give the $\sim 5\%$ tail in \GreenDark dark gray\special{color cmyk 0 0 0 1.}{}. \label{fig:example}} \begin{center} \begin{picture}(17,5) \putps(-0.5,-1)(-0.5,-1){figMoriond}{figMoriond.ps}\special{color cmyk 0 1. 1. 0.5} \put(2.5,4.5){fig.~1a} \put(8.3,4.5){fig.~1b} \put(14.1,4.5){fig.~1c}\special{color cmyk 0 0 0 1.} \end{picture} \cQ{\vspace{5mm}}{} \end{center}\end{figure} \subsection{Minimal supergravity} ``Minimal supergravity'' assumes common values $m_0$, $M_{1/2}$ and $A_0$ at the unification scale $M_{\rm GUT}\approx2\cdot10^{16}\,{\rm GeV}$ for the sfermion masses, the three gaugino masses, and the $A$-terms. The parameters $\mu_0$ and $B_0$ are free. \cQ{Since we adopt standard notations, we do not list the exact definitions of the various supersymmetric parameters.}{} As explained in the previous section we randomly fix the dimensionless ratios of the soft terms and compute the overall supersymmetric mass scale ``$m_{\rm SUSY}$'' from the minimization condition of the MSSM potential. More precisely we scan the parameters within the following ranges $$m_0,|\mu_0|,|M_{1/2}|,|B_0|,|A_0|=(0\div 1)m_{\rm SUSY},\qquad (a\div b)\equiv \hbox{a random number between $a$ and $b$}.$$ The alternative scanning procedure \cQ{$$m_0=(\frac{1}{9}\div 3) m_{\rm SUSY},\qquad |\mu_0|,|M_{1/2}|=(\frac{1}{3}\div 3) m_{\rm SUSY},\qquad A_0,B_0=(-3\div 3) m_0$$} {$m_0=(\frac{1}{9}\div 3) m_{\rm SUSY}$, $|\mu_0|,|M_{1/2}|=(\frac{1}{3}\div 3) m_{\rm SUSY}$, $A_0,B_0=(-3\div 3) m_0$} (with the samplings of $m_0, M_{1/2}$ and $\mu_0$ done with flat density in logarithmic scale) would give the same final results: {\em \cQ{in the MSSM with soft terms mediated by `minimal supergravity'}{} the present experimental bounds exclude $95\%$ of our parameter space.\/} As in the previous example, reasonable different definitions of `order one' (i.e.~of the prior Bayes distribution) do not alleviate the naturalness problem: we could make the naturalness problem apparently more dramatic by restricting the dimensionless ratios to a narrow region that does not include some significant part of the experimentally allowed region, or by extending the range to include larger values that produce a larger spread in the spectrum so that it is more difficult to satisfy all the experimental bounds. Allowing one or few soft terms to be much smaller or much larger than the others does not help. Complex soft terms only give problems with electric dipoles. \smallskip \cQ{We now exhibit the results of this analysis in a series of figures.}{} In fig.~1b we show a scatter plot with sampling density proportional to the naturalness probability. The sampling points corresponding to excluded (still allowed) spectra are drawn in black, (in \Green light gray\special{color cmyk 0 0 0 1.}{}). Fig.~1b shows the correlation between the masses of the lightest chargino and of the lightest higgs, $m_\chi$ and $m_h$. We see that the most natural value for the chargino mass is around $30\,{\rm GeV}$ (rather than $300\,{\rm GeV}$): however numerical accidents can sometimes make the chargino so light that it can be produced at $B$-factories, or make it so heavy that it cannot be produced at LEP2. Fig.~1b also shows that the experimental bounds on $m_\chi$ and $m_h$ are the most important ones in minimal supergravity (and that the bound on $m_\chi$ is more important than the one on $m_h$). The LEP2 bounds together with the assumption of gaugino mass universality at the unification scale give stronger constraints on the mass of coloured sparticles than the direct bounds from Tevatron experiments. In fig.~1c we show the same kind of results using a different format: we plot the density of the points in fig.~1b as function of $m_\chi$. We obtain a bell centered around $30\,{\rm GeV}$. Allowed spectra are confined in fig.~1c to the small ($\sim 5\%$) upper tail where the chargino is anomaously heavy. \subsection{`Unified' supergravity} If we only impose unification relations between soft terms, the naturalness problem remains unchanged: an unnatural cancellation remains necessary even if there are more parameters. From the point of view of phenomenology, maybe the most interesting new possibility is that a mainly right handed stop can `accidentally' become significantly lighter than the other squarks. A very light stop can mediate large loop effects (in $b\to s\gamma$, $\Delta m_B$, $\varepsilon_K$, \ldots) and allows baryogenesis at the weak scale\cite{EWbaryogenesis}. However this possibility is severely limited by naturalness considerations\cite{lightStop,nat}: it requires not only that one numerical accident makes the $Z$ boson sufficiently lighter than the unobserved chargino, but also that a second independent numerical accident gives a stop lighter than the other squarks. This unnatural combination occurs very rarely ($p\sim10^{-3}$) in our analysis. \subsection{Gauge mediation} For our purposes `gauge mediation' models\cite{GaugeSoft} can be characterized saying that they predict the soft terms $M_i$, $m_R$, $A$ of sparticles in terms of their gauge charges $c_R^i$ in the following way: $$ M_i(M_{\rm GM})=\frac{\alpha_i(M_{\rm GM})}{4\pi} M_0,\qquad m_R^2(M_{\rm GM}) = \frac{c^i_R }{\sqrt{n}}M_i^2(M_{\rm GM}) ,\qquad A(M_{\rm GM})=0 $$ ($R$ runs over the sfermions and $i$ over the gauginos). These predictions depend on three unknown parameters: the overall scale of soft terms $M_0$, the messenger mass $M_{\rm GM}$ and the `number of messengers' $n$. A computation shows that: \cQ{\begin{itemize}}{} \cQ{\item}{~~$\bullet$~~} If the messengers are heavy, $M_{\rm GM}\circa{>}10^{12}\,{\rm GeV}$ the gauge mediated sparticle spectrum, and the naturalness problem, are not much different than in supergravity. \cQ{\item}{~~$\bullet$~~} If $M_{\rm GM}\sim 10^{(6\div 10)}\,{\rm GeV}$ gauge mediation models predict a right handed selectron mass significantly smaller than the higgs mass term which sets the scale of electroweak symmetry breaking. This makes the naturalness problem more acute\cite{GMFT}. \cQ{\item}{~~$\bullet$~~} If the messengers are very light ($M_{\rm GM}\circa{<}100\,{\rm TeV}$) and the lightest neutralino $N$ decays in a detectable way ($N\to \gamma\,\hbox{gravitino}$) within the detector, experimentalists can put a {\em very\/} strong bound on its mass that makes the model very unnatural\cite{nat}. \cQ{\end{itemize}}{} \cQ{Extremely light messengers ($M_{\rm GM}\approx 10\,{\rm TeV}$) could give rise to a more natural sparticle spectrum, but unknown NLO corrections, that depend on unknown couplings between messengers, become relevant in this limiting case.}{} \section{The credibility of supersymmetric models after LEP2} If we employ the standard notation \cQ{$$}{$} p(A|B)\equiv \hbox{(probability of $A$ assuming that $B$ is true)} \cQ{$$}{$} we can compactely rewrite our main result as \cQ{\begin{equation}}{$\special{color cmyk 1. 1. 0.3 0}} p(\hbox{``LEP''}|\hbox{``SUSY''})=5\%\label{eq:5} \cQ{\end{equation}}{$\special{color cmyk 0 0 0 1.}} where \hbox{``LEP''}{} and \hbox{``SUSY''}{} are abbreviations for \begin{eqnarray*} \hbox{``LEP''} &=&\hbox{``no supersymmetric signal found at LEP and Tevatron experiments''}\\ \hbox{``SUSY''} &=&\hbox{``MSSM with minimal supergravity, with comparable soft terms''} \end{eqnarray*} Our result means that if \hbox{``SUSY''}{} is true, it is strange that sparticles have not been discovered at LEP\footnote{We repeat that {\bf \hbox{``SUSY''}{} does not mean supersymmetry\/} --- it is just a particular representative model often used for discussing capabilities of accelerators and for other purporposes (we are also assuming that soft terms are comparable, and that there is no special correlation between them). However a similar naturalness problem is present in all `conventional' supersymmetric models.}. We do {\bf not} claim that $p(\hbox{``SUSY''}|\hbox{``LEP''})=5\%$, i.e.\ that \hbox{``LEP''}{} exclude \hbox{``SUSY''}{} at $95\%$ C.L. Nothing says that the inverse probability $p(\hbox{``SUSY''}|\hbox{``LEP''})$ is equal to the direct one $p(\hbox{``LEP''}|\hbox{``SUSY''})$. However we can say something on the more interesting inverse probability because the Bayes theorem\cite{Bayes} says how they are related: $$\frac{p(\hbox{``SUSY''}|\hbox{``LEP''})}{p(\hbox{``SUSY''})}=\frac{5\%}{1-95\%p(\hbox{``SUSY''})}$$ In order to apply the theorem we have introduced an additional quantity $p(\hbox{``SUSY''})=$ degree of belief that \hbox{``SUSY''}{} (rather than some other unspecified mechanism) is the solution to the SM naturalness problem before knowing the experimental results. The theorem says how much one must reduce its confidence in \hbox{``SUSY''}{} after knowing the LEP2 results. Of course it gives obvious results. For example who very strongly believes in $\hbox{``SUSY''}$ would insert $p(\hbox{``SUSY''})=1$ and get $p(\hbox{``SUSY''}|\hbox{``LEP''})=1$: experiments are explained by saying that we live in the small allowed corner of the parameter space of \hbox{``SUSY''}{} with heavy sparticles. Who believes strongly in supersymmetry could instead suspect that the true model has not yet been found. \section{Conclusions}\label{fine} In conclusion, the negative results of the recent searches for supersymmetric particles pose a naturalness problem to all `conventional' supersymmetric models. What could be the reason of these negative experimental results? \cQ{\begin{itemize}}{} \cQ{\item}{~~$\bullet$~~} Maybe SUSY has escaped detection beacuse a numerical accident has made the sparticles too heavy for LEP2. This happens with $\sim5\%$ probability. Once one accepts the presence of a numerical accident, it is no longer extremely unlikely that the coloured sparticles have mass of few TeV (beyond the LHC reach) due to an accidental cancellation stronger than the `minimal one'~\cite{nat}; \cQ{\item}{~~$\bullet$~~} Maybe SUSY has not been discovered because it does not exists at the weak scale; \cQ{\item}{~~$\bullet$~~} Maybe SUSY exists at the weak scale, there is no unluck accident, but some ingredient is missing in the `conventional' models. \cQ{\end{itemize}}{ } Some recent papers have studied few different possible ways of alleviating the naturalness problem: \cQ{\begin{enumerate}}{} \cQ{\item}{(1.)} A correlation between the $\mu$ term and the gaugino mass\cite{corr}. However this correlation reduces the FT with respect to $\mu$, but not the FT with respect to $\lambda_{\rm top}$ or to $\alpha_{\rm strong}$. The problem is that the required value of $\mu/M_{1/2}$ depends on the values of $\lambda_{\rm top},\alpha_{\rm strong},\tan\beta\ldots$ (i.e.\ even if the high-energy theory predicts a $Z$ boson ligher than sparticles, the prediction is destabilized by large RGE corrections). \cQ{\item}{(2.)} Models with SUSY breaking mediated at lower energy could predict a light $Z$ boson, not destabilized by quantum corrections. However concrete models of this kind (gauge mediation\cite{GaugeSoft} and higher dimensional Scherk-Schwarz\cite{SS}\footnote{The predictions of these models depend on the form of the ultraviolet cutoff; even for the cutoff used in\cite{SS} the dependence of the higgs potential on the arbitrary matching scale (beteween the full theory and the low energy 4-dimensional theory) is so strong to require a NLO computation, which result would depend on the unknown details of the high energy theory.} models) turn out to be not more natural than supergravity. \cQ{\item}{(3.)} A sparticle spectrum more degenerate than the `conventional' one. Since the Tevatron direct bound on the gluino mass is weaker than the indirect bound obtained from LEP2 assuming gaugino mass universality, it is possible to reduce the mass of coloured particles (and consequently their large RGE corrections to the $Z$ mass, that sharpen the naturalness problem) if gaugino mass universality is strongly broken\cite{FTred,bau,nat}. The phenomenology is more interesting than in `conventional' models: the gluino is lighter (could be discovered at Tevatron) and the minimal cross section for detecting a neutralino dark matter is $\sim 50$ times higher. However even this case is severely constrained by experimental bounds: often at least one sparticle happens to be lighter than the $Z$ boson and contradics one of the many experimental bounds. \cQ{\end{enumerate}}{ } In conclusion it is certainly possible to debate if there is a supersymmetric naturalness problem, sometimes with considerations that also make the SM natural, thus killing the motivation for supersymmetry. Once that the problem is accepted it is however a solid problem: as usual it cannot be eluded with mild assumptions about the high energy theory. \special{color cmyk 0 0 0 1.} \frenchspacing \section*{References}
\section{Introduction} Over the last two decades various attempts have been aiming at a qualitative understanding and modeling of two basic properties of QCD: quark confinement and chiral symmetry breaking. Two well-known, complementary schemes are the instanton liquid model \cite{ILM} and the dual superconductor picture of the QCD vacuum \cite{DSC1,DSC2}. While the first model explains chiral symmetry breaking and solves the $U_A(1)$ problem, the second provides the language to discuss and to quantitatively describe the confinement mechanism within an effective field theory. Nowadays, it has become of central importance in the lattice community to quantify the properties of the lattice vacuum guided by the instanton liquid picture \cite{NEGELE98}. The evidence for a particular instanton structure is still controversial, in pure Yang-Mills theory and in full QCD. In the effective dual Abelian Higgs theory \cite{ADHM}, the vacuum is viewed as a dual superconductor, where condensation of color magnetic monopoles \cite{SmitVdS} (described by the Higgs phase of that model) leads to confinement of color charges by flux tubes through a dual Meissner effect. In order to establish this correspondence step by step, it has first been demonstrated that the Abelian components of the gauge fields are effective degrees of freedom (among the original non-Abelian gauge fields) at large distances, leading to the concept of Abelian dominance \cite{SuzYot}. The role of monopoles and their different percolation properties for confinement and deconfinement were empirically substantiated by a large number of lattice simulations over the last years. It was shown that in the confinement phase monopoles percolate through the $4D$ volume \cite{BORN} and are responsible for the dominant contribution to the string tension (monopole dominance) \cite{ShiSuz,BALI}. More recently, these monopoles have been shown to be predominantly selfdual (in confinement) \cite{SchiBo} and to encode the non-perturbative structure of the gauge field at large distances as seen by quarks, explaining the chiral condensate \cite{MIYAMURA} as well as the topological density \cite{SasaMiya}. In order to establish the link to the effective dual theory, the action that describes the first quantized theory of monopoles has been derived from the configurations of magnetic currents \cite{SuzAct}. At the same time, inherited from the original gauge fields, these ``Abelian monopoles" are characterized by a self-action (which is essentially non-Abelian!) and a complicated interaction with the remaining degrees of freedom. In the last couple of years also a deeper connection between monopoles and instantons has been pointed out, both on the lattice \cite{THUR95,THUR96,SUGA97} and in the continuum \cite{CHER95,REIN97,WIPF,LENZ,BROW97}. Following 't Hooft \cite{DSC1}, monopoles should be searched for as pointlike singularities of some gauge transformation dictated by a local, gauge covariant composite field. The choice of this field implies the choice of a particular gauge to analyze the non-Abelian gauge field. The best known examples are the maximally Abelian gauge (MAG) \cite{KRON87} and the Polyakov gauge (PG). It has become practice to identify the monopoles as particle-like singularities on the lattice after Abelian projection from one of the gauges chosen, identifying their trajectories in the resulting compact Abelian lattice field. This is known as the DeGrand-Toussaint (DGT) construction \cite{degrandT}. An attractive alternative, related to the Laplacian gauge (LG), has been revived recently \cite{VINK,SIJS}. This method offers the possibility to identify monopoles on the lattice without the problems that the conventional methods have and bears a close relationship to the 't Hooft-Polyakov monopoles. One can define them without really performing a gauge transformation. The LG can, however, also be the starting point for an Abelian projection followed by a DGT monopole construction. In this paper we intend to compare these three gauges used to define DGT monopoles with respect to their physical significance at the deconfinement transition. Furthermore we want to give a justification of the DGT construction of monopoles in the case of the LG, in view of their gauge independent localization similar to that of 't Hooft-Polyakov monopoles. An important ingredient of our discussion is a smoothing method for non-Abelian gauge fields used to remove UV fluctuations that are without relevance for the long distance physics. From our previous experience \cite{FEUE} with this method we know that this procedure is suitable to eliminate UV lattice artifacts from the monopole-related observables, too, without destroying the confining structure as usual cooling procedures finally do. Originally, methods like cooling or smoothing have been developed in order to study the instanton structure which otherwise is possible only by fermionic (spectral flow) methods \cite{fermionic}. A particular feature of cooling or smoothing, that should interest us here, is that it maps gauge fields on full gauge field configurations which still can then be put into various gauges for a closer study of one or the other qualitative picture. Concerning the monopole degrees of freedom, it is advantageous that smoothed configurations carry monopole currents with short loops removed. The renormalization group (RG) based smoothing method which uses (classically) perfect actions has been strongly recommended in the last two years \cite{degrand,FEUE,PRD98} because it is not afflicted with some problems of the simple or improved cooling \cite{coolimp} methods. UV fluctuations near to the lattice discretization scale are removed from the gauge field by one blocking step followed by a smooth interpolation of the field on the original lattice. By this procedure, performed locally below a well-defined (the blocked) scale, the RG smoothed configurations remain confining (if they are derived from MC runs in the confining phase). The string tension is reproduced almost completely. At the lattice sizes (and values of $\beta$) that were to our disposal, the two RG levels involved in the procedure are not totally decoupled from the confinement scale. Fluctuations are removed by the blocking--inverse--blocking scheme, which contribute a few percent to the string tensions at the distances where we can study it. In this paper we put special emphasis on the possibility of a gauge invariant definition of monopole trajectories and test it against the conventional DGT method usually applied to configurations being Abelian projected from various gauges. We were particularly interested in temperatures below and above the deconfinement critical point. This gave us the opportunity to compare the behavior of the respective monopole degrees of freedom at the transition and examine their dynamical relevance. The paper is organized as follows. In Sec. II we briefly discuss the RG smoothing technique and recall some properties of action and topological charge densities after smoothing. In Sec. III we describe the different gauge fixing procedures being used and report certain technical details how they work for smoothed configurations. In Sec. IV various aspects of the monopole degrees of freedom are presented, pinning down the main difference between the different gauges by means of the different physical behavior of the corresponding monopoles across the phase transition. In Sec. V we present some incidental results of our study concerning Abelian dominance, comparing the Abelian string tensions obtained by Abelian projection from various gauges with the fully non-Abelian non-perturbative potential extracted by smoothing. In Sec. VI we summarize our conclusions. \section{RG Smoothing and Physical Densities \label{localact}} To recognize semiclassical structure in ensembles of gauge field configurations generated by lattice simulations, cooling methods have been used, that minimize the action. However, even improved versions of cooling \cite{coolimp} gradually reduce the string tension. These versions are designed to stabilize instantons above a certain threshold for their scale size $~\rho~$ (typically $~\rho > 2a~$ with the lattice spacing $a$). However, they let close instanton-antiinstanton pairs annihilate. Up to now, most lattice studies are performed using the Wilson action, even if they focus on the topological structure of the vacuum. This action is known to be afflicted with so-called dislocations, structures of a size near one lattice spacing. For dislocations certain lattice definitions of the topological charge $Q$ give a signal of unit charge, although their total action $S$ violates the inequality $S \geq |Q| 8\pi^2 / g^2$ and the entropic bound $S > \bar{S}=48\pi^2/11~N_c^2$. For discretized instantons, the Wilson action decreases with size $\rho$ of an instanton, such that isolated instantons shrink under cooling before they decay as a dislocation. Thus, cooling with Wilson action is unsuitable to unambiguously determine the instanton size. Other methods like APE smearing let instantons even grow. To avoid these ambiguities we have proposed to use the renormalization group motivated method of ``constrained smoothing'' \cite{FEUE} which is based on the concept of perfect actions \cite{HASE}. For discretized instantons, these actions guarantee a size independent action for instantons, even with radii near to the lattice spacing. They allow a theoretically consistent ``inverse blocking'' operation which, for instance, reconstructs an instanton solution on a finer lattice with the same value of action. Inverse blocking is a method to find a smooth interpolating field on a fine lattice by constrained minimization of the perfect action, provided the configuration is given on a coarse lattice. This way to interpolate allows to define an unambiguous topological charge \cite{HASE}. For non-classical configurations like generic Monte Carlo configurations, one is interested to study smoothed configurations which are made free from UV fluctuations below a well-defined scale. Constrained smoothing in the way we use it, consists in first blocking the fields $\{U\}$, sampled on a fine lattice with lattice spacing $a$ using the perfect action, to a coarse lattice configuration $\{V\}$ with lattice spacing $2~a$ by a standard blockspin transformation. Then inverse blocking is used to find a smoothed field $\{U^{\mathrm{sm}}\}$ replacing $\{U\}$ on the fine lattice. We refer to this method as ``renormalization group'' (RG) smoothing. This procedure can be cyclically repeated, choosing different blocked lattices for the blocking step in subsequent cycles. This has been practized in Ref. \cite{boulder_cycling}. Experience shows that this RG cycling method makes configurations locally more and more classical. Unfortunately, there is no well-defined smoothing scale, but in contrast to cooling this method preserves long range features like confinement rather well. An important advantage of the simpler RG smoothing method outlined above is that it does not drive configurations into classical fields as unconstrained minimization of the action (and RG cycling) would do. It saves the long-range structure of the Monte Carlo configuration in $\{V\}$ and allows to study semiclassical objects in the smoothed background deformed by classical and quantum interaction. It is the upper blocking scale which roughly defines the border line between ``long'' and ``short range''. In this work we used a simplified fixed-point action \cite{degrand,PRD98} for Monte Carlo sampling and for RG smoothing. Before describing in detail the various gauge conditions we should explain which {\it gauge invariant} quantities we shall use to characterize the smoothed configurations. Partly they are suggested by the perfect action $S_{FP}$ itself which is parametrized in terms of only two types of Wilson loops, plaquettes $U_{C_{1}}=U_{x,\mu,\nu}$ (type $C_{1}$) and tilted $3$-dimensional $6$-link loops (type $C_{2}$) of the form \begin{equation}\label{sixlinks} U_{C_{2}}=U_{x,\mu,\nu,\lambda}= U_{x,\mu} U_{x+\hat{\mu},\nu} U_{x+\hat{\mu}+\hat{\nu},\lambda} U_{x+\hat{\nu}+\hat{\lambda},\mu}^{\dag} U_{x+\hat{\lambda},\nu}^{\dag} U_{x,\lambda}^{\dag} \, \, , \end{equation} and contains several powers of the linear action terms corresponding to each loop of both types that can be drawn on the lattice \begin{equation}\label{eq:fp_action} S_{FP}(U)=\sum_{type~ i} \sum_{C_{i}} \sum_{j=1}^{4} w(i,j) (1 - {1 \over 2}~\mbox{Tr}~U_{C_{i}} )^j \, \, . \end{equation} The parameters of this action are reproduced in Table \ref{tab:weights}. We define the action density $s_{site}(x)$ per lattice point (a local action) as follows \begin{equation} \label{eq:local_action} s_{site}(x) = \sum_{j=1}^{4} \left( \sum_{C_{1}(x)} \frac{w(1,j)}{4} (1 - {1 \over 2}~\mbox{Tr}~U_{C_{1}(x)})^j + \sum_{C_{2}(x)} \frac{w(2,j)}{6} (1 - {1 \over 2}~\mbox{Tr}~U_{C_{2}(x)})^j \right) \,. \end{equation} Here, $~C_{j}(x)$ ($j=1,2$) means loops of type $j$ running through the lattice site $x$. Summing $s_{site}(x)$ with respect to $x$ yields the total action of the configuration. The topological charge density definition we are using is ({\it i}) the simplest ``field theoretic'' one constructed out of plaquettes around a site $x$ \begin{equation}\label{eq:q_naive} q(x) = - \frac1{2^9 \pi^2} \sum_{\mu,\nu,\sigma,\rho=-4}^{+4} \epsilon_{\mu\nu\sigma\rho} {\mathrm tr} \left(U_{x,\mu,\nu} U_{x,\sigma,\rho}\right) \end{equation} and ({\it ii}) the charge contributed by hypercubes according to L\"uscher's definition of topological charge \cite{luescher}. We did not attempt to improve the field theoretic definition. Both topological densities used are known to behave regularly for smooth configurations \cite{letter96,FEUE}. In order to give a characterization of Abelian monopole currents in terms of locally defined {\it gauge independent} observables, we consider also the $3$-cubes of the original lattice dual to an elementary piece of monopole world line (carried by a link of the dual lattice). This leads us to a natural definition of a local action on the monopole world line, $s_{3-cube}(c)$ instead of $s_{site}(x)$. We include into $s_{3-cube}(c)$ the contribution of all plaquettes forming the $6$ faces of the $3$-cube $c$ plus the contribution of all $6$-link loops which wind around its surface. In short, it is the part of the total action which lives on that cube. \section{Monopole Identification Methods} \subsection{Maximum Abelian gauge and Laplacian gauge} The most popular gauge is the maximally Abelian gauge (MAG) \cite{KRON87}. Abelian projection of configurations put into MAG exhibits Abelian dominance. This gauge is often used to study the dynamics of monopoles on the lattice. It is enforced by an iterative minimization procedure, which can get stuck in local minima that are gauge copies of each other (so-called {\it technical} Gribov copies). The Laplacian gauge (LG) is not afflicted with this problem. The basic idea has been suggested \cite{Seixas} more than a decade ago but has apparently not been used until very recently \cite{SIJS}. Originally, in the context of the adjoint Higgs (Georgi-Glashow) model, it has been proposed to introduce, besides the quantized scalar matter field, a purely auxiliary adjoint slave field. Such a Higgs field can be used to define a gauge transformation to the ``unitary'' gauge with respect to the auxiliary field except at its zeroes. Thus, the corresponding monopoles are located at the zeroes of the auxiliary Higgs field and are representing the singularities of the gauge rotation. They can be identified as t'Hooft-Polyakov monopoles. For each given (generic, non-classical) non-Abelian gauge field the auxiliary field is determined as the eigenvector related to the lowest eigenvalue of the adjoint covariant Laplacian. This is why the resulting gauge is named Laplacian gauge (LG). The monopole identification via the Higgs zeroes is obviously gauge invariant (see also Ref. \cite{HOLLANDS}). It is this prescription we want to apply to pure Yang-Mills theory on the lattice and to compare with other ways for the Abelian projection. Strictly speaking, in order to describe the location of the monopole world lines (Higgs zeros), we would need algorithms interpolating the Higgs field between the lattice sites. For our present purposes it will be sufficient to detect clusters of lattice points with small Higgs modulus. Technically, MAG and LG are closely related to each other. For MAG, the gauge functional to be minimized can be written as follows \begin{eqnarray} \label{functional} F(\Omega) & = &\sum_{x,\mu} (1-\frac{1}{2} \, tr \, (\sigma_3 U^{(\Omega)}_{x,\mu} \sigma_3 U^{(\Omega)\dag}_{x,\mu} ) ) \nonumber \\ &=& \sum_{x,\mu,a} (X_x^a - \sum_{b} R^{a,b}_{x,\mu} X_{x+\hat \mu}^b )^2 \rightarrow \int_V (D_{\mu} X)^2 \, , \end{eqnarray} with the gauge transformation $\Omega_x$ acting on $\{U\}$ \begin{displaymath} U^{(\Omega)}_{x,\mu} = \Omega_x U_{x,\mu} \Omega^{\dag}_{x+\hat \mu} \end{displaymath} encoded in an {\it auxiliary} adjoint Higgs field \begin{displaymath} \Phi_x = \Omega^{\dag}_x \sigma_3 \Omega_x = \sum_a X^a_x \sigma_a \end{displaymath} subject to local constraints $\sum_a (X^a_x)^2 = 1$ and with adjoint links \begin{displaymath} R^{a,b}_{x,\mu} = \frac{1}{2} \, tr \, (\sigma_a U_{x,\mu} \sigma_b U_{x,\mu}^{\dag}) \, . \end{displaymath} To change from MAG to LG, the local constraints $ \sum_{a} (X^a_x)^2 = 1 $ are relaxed and replaced by a global normalization: $ \sum_{x,a} (X^a_x)^2 = V $, such that Eq. (\ref{functional}) can be further written: \begin{equation} \int_V (D_{\mu} X)^2 \rightarrow \sum_{x,a} \sum_{y,b} X^a_x\{-\Box_{x,y}^{a,b}(R) \} X^b_y \, . \end{equation} Thus, the minimization of the gauge functional is reduced to a search for the lowest eigenmode of the covariant lattice Laplacian. The LG of a given lattice configuration is unambiguously defined, except for the case that degenerate lowest eigenmodes exist. The corresponding LG configurations would then define {\it true} Gribov copies of each other. For both MAG and LG, the gauge transformation is finally accomplished by finding $\Omega_x$ that diagonalizes the field $\Phi_x$. Here a technical observation ought to be made. As usual, the MAG is found iteratively by minimizing the functional (\ref{functional}). The fact that we analyze smoothed configurations does not mean that the MAG is computationally less demanding to find. Still, as in the case for unsmoothed configurations, we need $O(1000)$ iterations to fulfill our stopping criterion. We require that the next gauge transformation should deviate from unity uniformly by less than $10^{-7}$. By abelianicity $R$ we denote the average Abelian fraction per link after completion of the minimization. We show the $\beta$ dependence of this quantity in Fig. \ref{fig:abelianicity} in comparison with other gauges which do not explicitly attempt to get a large abelianicity. There is a monotonous increase with $\beta$ across the phase transition, but no ``deconfinement signal'' related to it. For the search of the lowest eigenvalue for the LG and its corresponding eigenvector, we apply the conjugate gradient algorithm to minimize the following Ritz functional \begin{equation} F=\frac{\langle X ,( - D^2 ) X \rangle}{\langle X , X \rangle} \, , \end{equation} where $D^2$ is the adjoint lattice Laplacian. We choose 5 random start vectors $X_x=X^a_x \sigma_a$ and follow the relaxation of the estimated eigenvalue over 3000 iterations. The vector with the lowest eigenvalue is then taken further representing the given gauge field configuration. The average over an ensemble of 100 gauge fields (per $\beta$ value) of the respective lowest eigenvalue is shown in Fig. \ref{fig:gaugefixing_lap} as a function of $\beta$ (or temperature). This average is $\beta$ dependent in both phases with a steep drop across the transition. The corresponding abelianicity in Fig. \ref{fig:abelianicity} is always smaller than the abelianicity for MAG, but approaches it with increasing $\beta$ (in the deconfined phase). Among the different gauges, the biggest rise of abelianicity accompanying the transition is found in the LG case. \subsection{Polyakov gauge} This gauge condition puts special emphasis on the (gauge group valued) Polyakov loop $P(x)$, the gauge transporter along the shortest loop starting from and arriving at a given lattice point which is closed by periodic thermal boundary conditions. The Polyakov gauge is enforced by simultaneous diagonalization of $P(x)$ at each point. In Fig. \ref{fig:abelianicity} we also show the abelianicity achieved in the PG. Except near the transition (where it has a shallow maximum) the abelianicity is practically $\beta$ (temperature) independent. This gauge has the smallest abelianicity among the discussed gauges, but it might be surprising that it comes near to $95$ \%. \subsection{Monopole identification} After the gauge of choice has been fixed one extracts the Abelian degrees of freedom by Abelian projection. The Abelian link angles representing a $U(1)$ field with residual gauge symmetry can then be used for the identification of monopoles. This is done, like in pure compact $U(1)$ gauge field theory, by computing the $U(1)$ gauge invariant (magnetic) fluxes through the surfaces of elementary 3-dimensional cubes or -- what is equivalent -- by searching for the ends of Dirac strings. Monopoles identified in this manner are generally referred to as DeGrand-Toussaint (DGT) monopoles \cite{degrandT}. We are going to localize DGT monopoles in our smoothed configurations put into all three gauges and to compare their possible significance. As stressed before, for the LG there exists an independent, gauge invariant way to localize monopoles in the sense of 't Hooft that can be used to test the reliability of the DGT method. The Higgs field introduced in the LG provides an alternative, physically more satisfactory possibility of monopole identification. In the continuum, lines of $\rho_x=|\Phi_x|=0$, where $\rho_x$ is defined as \begin{equation} X^a_x = \rho_x \hat X^a_x \,\,\ , \,\,\, \rho_x = \sqrt{\sum_{a=1}^3 (X^a_x)^2 } \, , \end{equation} directly define lines of gauge fixing singularities (mo\-no\-po\-les). For the 't Hooft-Polyakov monopole in the adjoint Higgs model, regions with $\rho=0$ of the {\it physical} Higgs field are identified with the centers of such monopoles. Note that this way of monopole identification in the pure Yang-Mills theory, too, does not require to actually perform the gauge transformation and an Abelian projection! It is gauge invariant by definition. Let us demonstrate the localization of the singularities of the LG transformation and how this is related to the DGT monopoles. We have plotted in Fig. \ref{fig:Xlokal} all lattice points with the local modulus squared $X^2(x)=\sum_a (X_x^a)^2$ being less than a threshold value of $10^{-5}$, for a generic configuration of the confined phase ($t$ fixed, top) and for one of the deconfined phase ($z$ fixed, bottom). The eigenvectors are always globally (over $12^3 \times 4$ lattice points) normalized to one, $\sum_x \sum_a (X_x^a)^2 = 1$. Although there are no true singularities (exact zero modulus of $X$) located on lattice sites it can be seen that small values $X^2(x)$ (less than one order of magnitude below the average) mark connected regions of space on the lattice where one may expect to find the monopole world lines (see also Ref. \cite{SIJS}). The dark lines correspond to DGT monopoles that have been extracted after Abelian projection has been applied to the LG. These monopoles are shown for a confinement configuration in Fig. \ref{fig:schicht} by arrows, where non-vertical and non-horizontal directions encode world lines leaving (entering) each $(x,y)$-miniplot along the $z$ or $t$ direction. They are located almost completely inside the shaded regions having $X^2(x)<10^{-5}$. Due to the smoothness property of the gauge transformations dictated by the local field $X_x^a \sigma_a$ it has been conjectured in Ref. \cite{SIJS} and demonstrated for normal Monte Carlo configurations that monopole currents are more abundant in the LG compared with MAG, as far as DGT monopoles are concerned. For RG smoothed gauge field configurations, we confirm this conjecture to be true as well. A small local modulus squared $X^2(x)=\sum_a (X_x^a)^2$ expresses the obstruction to unwind the gauge field (to the unitary gauge) near the site $x$, while the local contribution to the Ritz functional $\sum_a X^a_x ( - D^2 X)^a_x$ expresses the disordering of the gauge field. Here we concentrate on the Higgs modulus squared because this is directly related to the monopole localization. In Fig. \ref{fig:binX} the upper plot shows the average value of $X^2$ provided, site by site, the local action $s_{site}$ falls into one of the bins on the abscissa. This plot refers to $\beta=1.4$. It lets us expect that only points with $s_{site}>4 a^{-4}$ have a chance to meet the condition $X^2<10^{-5}$ mentioned before with non-vanishing probabiltity. In Fig. \ref{fig:binX} the lower plot (also for $\beta=1.4$) shows the average value of $X^2$ similarly depending on the local topological density $q$ falling into one of the bins. According to this, only points with $|q|>0.01$ have a chance to meet the $X^2$ condition mentioned before with some probability. In Fig. \ref{fig:probX} we show the probability distribution for encountering a certain $X^2$ for the two cases, having no DGT monopole in a $3D$ box (top) and having one or more monopoles present (bottom), both in the confinement and in the deconfinement phase. To the (normalized) histograms, values of $X^2$ found at the eight corners of each (occupied or empty) $3D$ cube have contributed. The histograms are averages over all configurations. It is obvious that in the case of a DGT monopole the $X^2$ distribution on the corners is strongly concentrated near zero. Roughly $50$ \% of the neighboring lattice sites fall below the threshold mentioned above. Finally, we should add one remark on the problem of potential Gribov copies. We have checked that the DGT monopoles, detected on RG smoothed configurations after transforming to MAG and subsequent Abelian projection, do basically not change their positions if random gauge transformations are applied before relaxation to MAG. From several RG smoothed configurations we have produced 10 different randomly gauged copies and have analyzed the relative shifts of individual monopole trajectories that emerge after MAG, Abelian projection and DGT construction. We found that more than $60$ \% of the dual links occupied by monopole trajectories will stay unchanged in position. About $30$ \% of the links will be shifted by one lattice spacing, shifts of two and more units were very rare. This gives reason to believe that RG smoothed fields are smooth enough such that the MAG gauge and Abelian projection will identify singularities without ambiguity. Notice that this was not the case without smoothing \cite{BORN,BALI}. Concerning the LG method, the eigenvalue estimators slightly differ numerically according to the 5 different random starting vectors that we considered, the regions characterized by a Higgs modulus below the threshold do practically not differ between the relaxed vectors. We interprete our observations mentioned before as an {\it a posteriori} justification for the use of the DGT method in the case of LG, too. In the following we will exclusively use the DGT prescription to extract LG monopoles. \section{Physical results on Abelian monopoles} Our configurations are generated on a $12^3 \times 4$ lattice at various $\beta$-values with the perfect action tested for consistency in Ref. \cite{PRD98}. In that paper we have found a critical $\beta_c=1.545(10)$ of the deconfinement transition for $N_{\tau}=4$. According to the second order of the transition, we have determined it from the intersection of the Binder cumulant of Polyakov lines (unblocked and blocked ones) for different spatial volumes. The present values of $\beta=1.4$, $1.5$, $1.6$, $1.7$ and $1.8$ enclose the critical point. The smoothing is performed as described in Ref. \cite{PRD98}. We have extracted the monopole degrees of freedom in the respective gauge by projection onto the diagonal part of the $SU(2)$ gauge field. For this Abelian gauge field, we have then detected the DGT monopoles. \subsection{Density of monopoles in various gauges} In Fig. \ref{fig:asymmetrie} (top) we show the temperature dependence of the average number of monopole currents $\langle M_s+M_t \rangle$ (total length of all monopole trajectories for a lattice of size $12^3 \times 4$), for the different gauges. If monopoles are important for confinement, one would expect to see a decrease accompanying the phase transition. In fact, this is only the case for the LG and, less drastically, for the MAG. For all temperatures the monopole number in LG is bigger than that in MAG as expected in Ref. \cite{SIJS}. The monopole number in PG is almost independent of the temperature! It is interesting to ask for the anisotropy ratio $\frac{\langle M_s \rangle}{3\langle M_t \rangle} $ as a function of the temperature, shown in Fig. \ref{fig:asymmetrie} (bottom). At zero temperature this ratio should be equal to one for isotropic gauges like MAG and LG. But it is expected to differ from one at higher temperatures. In our first study of RG smoothing \cite{FEUE} for MAG, we observed a drop towards the deconfined phase much more pronounced than for unsmoothed, Monte Carlo configuration. This is due to the strong isotropic noise which is present in the magnetic currents before smoothing. We find the ratio smaller than one at our different finite temperatures, and it drops towards and across the deconfining transition. In the case of MAG, beyond the deconfinement temperature the anisotropy ratio is approaching zero. For LG the ratio is closer to one in the confinement phase and drops more abruptly crossing the transition. It decreases slowly for $T > T_c$. We have observed that also the LG monopole network decays (similar to the MAG monopoles \cite{PRD98}), into a few temporally closed world lines which are, however, not perfectly static as in the case of MAG. In view of this, both the MAG and LG data suggest the conclusion that in the deconfinement phase the corresponding Abelian monopoles become preferably static objects. What becomes static in the case of the LG are the extended ``channels'' of small Higgs modulus (see Fig. \ref{fig:Xlokal}). For MAG, this observation has been first discussed in Ref. \cite{BORN}. For the PG the anisotropy ratio of the corresponding monopoles does not depend on temperature at all. Moreover it is far from isotropy, with $\frac{\langle M_s \rangle}{3\langle M_t \rangle}\approx0.05$ (almost no spacelike magnetic currents) even in the confinement phase. Thus, the magnetic currents observed in the PG are hardly related to the confinement phenomenon. For clarity we show in Fig. \ref{fig:separate} the average density of timelike magnetic currents per $3D$ hyperplane orthogonal to its direction, as a function of the temperature in units of $\Lambda_L^3$. The upper data points represent the LG monopoles, the lower the MAG monopoles. In both of the gauges the physical density of the (almost static) timelike monopoles starts rising as a function of the temperature for $T > 1.2~T_{\mathrm{c}}$. It has been speculated that the deconfinement phase is characterized by a bosonic gas of 't Hooft-Polyakov monopoles with a mass linearly rising with the temperature \cite{LAURSEN,BORN}. \subsection{Local properties of monopoles in various gauges} By definition, DGT monopole currents are localized along the links of the dual lattice which are related to $3D$ cubes of the original lattice. They test the properties of the non-Abelian gauge field most locally. We have defined in Sec. II local actions and topological charge densities. Now we want to discuss the different gauges from the point of view of how their respective monopole trajectories probe the vacuum, i.e., whether these monopoles are equipped with particular, gauge independent properties. We expect this picture to become clearer when, as in our case, UV fluctuations are removed from the gauge field configurations. First let us have a look how the average monopole occupation number on the dual links near to a site depends on the local action and charge, in analogy to Fig. \ref{fig:binX}. Fig. \ref{fig:actbinmon} shows this for the three gauges. In this respect, no drastic difference between the gauges is seen. Again the occupation numbers of LG monopoles are bigger than in the case of MAG, those for PG somewhat lower. No distinction is made here with respect to the direction (spatial vs. temporal) of the magnetic currents. In Fig. \ref{fig:probact140} we show the distribution of $s_{3-cube}$, separately for cubes free of monopoles (top) and for cubes distinguished by monopoles detected in the LG (bottom), for the lowest and the highest $\beta$-value. This local action contains all loops contributing to $S$ localized on the $3D$ cube under consideration. We see that the distributions for the two cases are different. Mean and variance in the case of presence of monopoles are almost a factor of two larger, indicating that on average significantly more action is picked up per unit of length along an actual monopole trajectory than would be picked up along a random walk in the given gauge field background. The situation is very similar for the MAG case, the difference between the action distributions for monopole-free and monopole-carrying cubes is somewhat more pronounced. Actually, these distributions are insensitive to the {\it direction} of the monopole current only in the confinement phase. In the deconfined phase there are important differences. For LG, monopoles occupying a cube dual to a spacelike link (which are still frequent in the deconfined phase, in contrast to MAG!) are accompanied by a probability distribution of local action which is similar to that of an empty cube, and for timelike ones the action is not as high as in the MAG case. This explains the higher multiplicity of LG monopoles and the lack of suppression of spacelike ``detours'' for world lines closing in timelike direction. For cubes along the trajectories of PG monopoles the action distribution is practically the same as for the empty cubes. This holds in both phases. Looking back to section III.C, we conclude that among the gauge invariant quantities $X^2$ is the observable, whose distribution reflects the absence or presence of monopoles most clearly. This holds independently of the direction of the monopole current in both phases (Fig. \ref{fig:probX}). \subsection{Excess action and charge} In a more concise way than by the distributions of local action just discussed we can characterize the different types of monopoles and the change with $\beta$ by an average excess action of monopoles defined as \begin{equation} S_{\mathrm{ex}}= \frac{< S_{\mathrm{monopole}}-S_{\mathrm{no~monopole}} > } {<S_{\mathrm{no~monopole}} >} \,\, , \end{equation} where $S_{\mathrm{monopole}}$ is again the action localized on a three-dimensional cube mentioned before. Replacing the action in the above expression by the modulus of the topological charge density according to the L\"uscher method we can define the charge excess $q_{\mathrm{ex}}$ compared with the noise of $|q(x)|$ anywhere else in the vacuum. For details of the definition of the local operators see Ref. \cite{PRD98}. Fig. \ref{fig:excess} shows that already just below $T_{\mathrm{c}}$ the excess action for MAG and LG monopoles is clearly above one (indicating an excess of action of more than a factor of two compared to the bulk average) and rises across the transition. Around $T_{\mathrm{c}}$, the excess charge is even a few times as big for MAG and LG monopoles. These results are more pronounced compared with a $T=0$ study \cite{BAKK98} (using Wilson action, without cooling or smoothing) and emphasize the dynamical importance of these two types of monopoles at the deconfining transition. \section{Does Abelian dominance of the string tension hold in all gauges?} The Abelian dominance of the string tension \cite{SuzYot} in the case of MAG was a strong argument for choosing this gauge to look deeper for the role of monopoles. Finally, this has given rise to the construction of infrared effective actions for QCD \cite{SuzAct}. Originally, of course, the concept of Abelian dominance was referring to generic Monte Carlo configurations. As a by-product of our study on monopoles in various gauges, we present here results concerning this question for smoothed configurations. RG smoothing preserves the non-Abelian string tension within $7$ \% and removes the Coulombic part of the heavy quark force at small distance. In Fig. \ref{fig:LLpot} we display the static quark-antiquark potential as obtained from Polyakov line correlators, for the smoothed $SU(2)$ fields and, in the case of LG and MAG, after Abelian projection. The corresponding values for the string tensions are found in Table \ref{tension}. The Abelian string tension of the MAG is about $5$ \% less than that calculated for the non-Abelian $SU(2)$ field. The Abelian string tension of LG is again somewhat smaller than for the MAG. In qualitative respect, we still find Abelian dominance for smoothed lattice fields, independent of the gauge chosen for projection. Notice, however, that in the PG (where the temporal links $U_{({\bf x},x_4),4}$ become diagonal and $x_4$ independent) the non-Abelian string tension extracted from Polyakov line correlators is trivially identical with the Abelian one. \section{Conclusion} We have collected additional evidence that the RG smoothing technique (with an approximate classically perfect action) is a powerful tool, here used to investigate semiclassical aspects of the monopole structure of the Yang-Mills vacuum. We have presented further results concerning the recently proposed, physically better motivated and gauge invariant method to identify monopoles on the lattice, the so-called Laplacian gauge (LG). We contrasted the corresponding monopoles with other types of monopole trajectories which usually are obtained in a gauge dependent way. What parallels exist and how different the monopole content can be if different gauges are used to localize them has been examplified by considering three gauges (MAG, LG and PG). In a first step we showed that the gauge invariant way to localize monopoles is consistent with the DeGrand-Toussaint (DGT) method being applied to Abelian projected configurations after LG fixing. This correspondence becomes clearer if smoothed Monte Carlo configurations are analyzed. This justifies to take over the DGT method to study the monopole content of configurations also in the LG, at least after RG smoothing. Accepting this technique, we also found strong correlations between positions of MAG and LG monopoles. While MAG and LG monopoles (as extracted by the DGT method) behave similar at the deconfining phase transition concerning density and anisotropy, monopoles identified in the Polyakov gauge apparently lack a dynamical relevance for confinement and at the deconfinement transition. In a second step we then analyzed trajectories of monopoles identified in various gauges and found that monopoles (in all three gauges studied) appear preferably in regions which are characterized by enhanced action and topological charge density. The reverse is not universally true: only monopoles detected in LG and MAG are characterized by significantly different local distributions of action. We have shown earlier \cite{PRD98} that strong gauge fields are locally (anti)selfdual. This is further, circumstantial evidence that the monopoles are in fact dyons in the confinement phase. As a by-product of our study we found that almost the complete string tension can be recovered from the Abelian projected field corresponding to various Abelian gauges, including LG and PG. For the latter this is trivially true, as long as we measure the string tension by means of the Polyakov line correlator. We have shown that for smoothed gauge field background monopole trajectories carry an excess action of about twice the bulk average of local action. Similarly, monopoles also carry an even bigger excess topological charge. In the confinement phase this observation applies to monopoles independent of which gauge has been chosen for identifying them, but PG monopoles are very different in the deconfinement phase. We conclude that the DGT monopoles related to MAG and LG behave similar physically and can be semiclassically interpreted as physical objects which carry considerable action and topological charge in the vicinity of the deconfinement transition. This work was supported in part by FWF under project P11456. One of us (M. M.-P.) acknowledges support by the European TMR network {\it Phase Transitions in Hot Matter} under contract FMRX-CT97-0122. E.-M. I. wishes to thank T. Suzuki and M. I. Polikarpov for discussions on dual descriptions of confinement.
\section{Introduction} Experimental studies of the time-scale of fission of hot nuclei have recently been carried out using the emission rates of neutrons \cite{Hinde}, $\gamma$-rays \cite{Hofman}, and charged particles \cite{Lestone} as "clocks" for the fission process. These experiments have shown that the fission process is strongly hindered relative to expectations based on the statistical model description of the process. The observed effects extend well beyond any uncertainties in the model parameters. It therefore appears that a dynamical description of the fission process at these energies is more appropriate \cite{Froebrich} and that the experimental data are able to shed light on dissipation effects in the shape degree of freedom. However, these experiments are not very sensitive to whether the emission occurs mainly before or after the traversal of the saddle point as the system proceeds toward scission. Various dissipation models are, however, strongly dependent on the deformation and shape symmetry of the system. As an alternative to these methods we therefore measure the evaporation probability for hot nuclei formed in heavy-ion fusion reactions, which is sensitive only to the dissipation strength inside the fission barrier. As the hot system cools down by the emission of neutrons and charged particles there is a finite chance to undergo fission after each evaporation step. If the fission branch is suppressed due to dissipation there is therefore a strongly enhanced probability for survival which manifests itself as an evaporation residue cross section which is larger than expected from statistical model predictions. This effect depends, however, only on the dissipation strength inside the saddle point and may therefore provide the desired separation between pre-saddle and post-saddle dissipation. In this paper, we report on recent measurements of evaporation residue cross sections for the $^{32}$S+$^{184}$W system over a wide range of beam energies using the Argonne Fragment Mass Analyzer (FMA). In sect. II we describe the experimental procedure followed by a discussion of the measurements of absolute evaporation residue cross section in sect. III. The results are compared to statistical model calculations and other relevant data in sect. IV followed by the conclusion, sect. V. \section{Experimental arrangement} The measurements were carried out using $^{32}$S-beams from the ATLAS superconducting linac at Argonne National Laboratory. The cross sections for evaporation residues produced in the $^{32}$S+$^{184}$W reaction were measured at beam energies of 165, 174, 185, 195, 205, 215, 225, 236, 246, and 257 MeV. Targets of isopically separated $^{184}$W with thickness 200~$\mu$g/cm$^2$ on a 100~$\mu$g/cm$^2$ carbon backing were used. The Argonne Fragment Mass Analyzer \cite{Davids} was used for identification of evaporation residues. A schematic illustration of the setup is shown in Fig. 1. In these experiments a sliding seal target chamber was used, which allows for measurements at angles away from 0$^\circ$. This is required in order to obtain the angular distributions for integration of the total evaporation residue cross section. Elastically scattered S-ions were registered in a Si detector placed at 30$^\circ$ relative to the beam axis with a solid angle of $\Omega_{mon}$ = 0.249 msr. These data were used for normalization purposes. A~ 40 $\mu$g/cm$^2$ carbon foil was placed 10 cm downstream from the target to reset the charge state of reaction products, which may be abnormally highly charged as a result of Auger electron emission following the $\gamma$ decay of short-lived isomers. A square entrance aperture for the FMA covering $\theta,\phi =4.5^\circ\times4.5^\circ$ ($\Omega_{FMA} = 6.24$ msr) was used. Reaction products transmitted through the FMA were dispersed in M/q (mass/charge) at the focal plane, where the spatial distribution was measured by a thin x-y position sensitive avalanche detector. When the FMA was placed at 0$^\circ$ some settings of the electrostatic and magnetic fields of the instrument allows beam particles scattered off the anode of the first electrostatic dipole, ED1, to be transported to the focal plane (presumably after a subsequent forward scattering in the vacuum chamber of the magnetic dipole MD1). When measuring small cross sections, as in the present study, it is therefore mandatory to achieve a clean separation between evaporation residues and beam particles. This was achieved by measuring their flight-time over the 40~cm distance to a double-sided Si strip detector (DSSD) placed behind the focal plane. This detector has a total active area of 5$\times$5 cm$^2$ and is divided into 16 strips on both the front and rear surface arranged orthogonally to each other. The information on the particle mass obtained from the time-of-flight and energy measurement provided by the Si-detector gave a clean discrimination against the scattered beam as illustrated in Fig. 2. The efficiency for transporting evaporation residues from the focal plane to the Si-detector was determined from the spatial distribution over the face of the DSSD detector as shown in Fig. 3 for these beam energies. By Gaussian extrapolation of the distribution beyond the edge of the detector it is estimated that this efficiency is around $\epsilon_{PPAC-Si}$ = 87\%. The transport efficiency of the FMA as a function of the mass, energy amd charge-state of the ion has been determined in a separate experiment \cite{Back}. \section{Cross sections} The evaporation residue cross section for the $^{32}$S+$^{184}$W reaction was measured for beam energies in the range $E_{beam}$=165-257 MeV. Evaporation residues were identified by time-of-flight and energy measurement using the focal plane PPAC detector and the Si-strip detector placed ca.~40~cm behind the focal plane. The charge state distributions, which were measured at three beam energies, are shown in Fig. 4. The dashed curves represent the formula of Shima {\it et al.} \cite {Shima}, whereas a somewhat better fit to the data is given by the Gaussian fit (solid curves) with a fixed standard deviation of $\sigma$=3. The arrows indicate the charge state setting of the FMA used for the cross section measurement. The derivation of the evaporation residue cross section at intermediate beam energies is based on an interpolation between these measured charge state distributions. Since the FMA disperses in $M/q$ at the focal plane there will be cases of ambiguities in the mass identification, since overlaps between lighter mass products from one charge state, $q$, will invariably overlap with heavier products from the neighboring charge state, $q+1$, when compound nuclei with high excitation energy are studied, see Fig. 5. We are not able to resolve this ambiguity with the present setup, and have therefore obtained the cross sections by integrating all counts that fall between the positions for $M/(q-\frac{1}{2})$ and $M/(q+\frac{1}{2})$ along the focal plane. Since the FMA is set up for the most abundant charge state, $q$ and mass, $M$, we expect that the loss of residues with charge state, $q$, and masses that fall outside this window is compensated by the acceptance of residues with charge states, $q+1$ and $q-1$ that fall inside this window. \subsection{Detection efficiency} The transport efficiency as a function of recoil energy and mass relative to the setting for the FMA has been measured for monoenergetic particles by observing the recoils from elastic scattering of $^{32}$S + $^{197}$Au, $^{208}$Pb, $^{232}$Th \cite{Back}. To correctly estimate the transport efficiency for evaporation residues, which have an extended energy distribution, it is necessary to fold the energy distribution with the measured acceptance curve. The energy distribution was not measured directly in the present experiment, but the yield of residues as a function of the energy setting of the FMA was measured as shown in Fig. 6 (top panel). In principle, since the energy acceptance of the FMA is known, it should be possible to convert this measurement into an energy distribution with some accuracy. We have, however, used a slightly different method which incorporated both this measurement and the measurement of the angular distributions. Assuming that both the angular distribution of evaporation residues and their energy distribution at 0$^\circ$ arise from isotropic multiparticle emission from the hot compound nucleus, these two entities are related by the kinematics of the particle decay cascade. We assume that the recoil energy distribution is isotropic in the center-of-mass system and that it has a Maxwellian form, namely \begin{equation} \frac{dP}{dE_{cm}} = \frac{2}{\sqrt{\pi}}\frac{\sqrt{E_{cm}}}{a^{3/2}}\; \exp\left(-E_{cm}/a\right), \end{equation} where $E_{cm}$ is the recoil energy in the center-of-mass system and $a = \frac{2}{3}\langle E_{cm}\rangle$ is two thirds of its average value. The energy distribution in the laboratory system at $\theta = 0^\circ$ is then \begin{equation} \frac{dP}{dE_{lab}}|_{_{0^\circ}} = \frac{1}{2} \left( \frac{1}{\pi a} \right)^{3/2} \sqrt{E_{lab}}\; \exp\left[ -\left(\sqrt{E_{lab}}-\sqrt{E_{CN}}\right)^2/a\right]. \end{equation} Here, $E_{CN}$ is the laboratory energy of the compound nucleus prior to the particle evaporation cascade. A small correction to $E_{CN}$ arising from the mass loss due to particle evaporation has been ignored in eqs. 1-3. Similarly we find the angular distribution \begin{equation} \frac{dP}{d\Omega_{lab}} = \frac{1}{2} \left( \frac{1}{\pi a}\right)^{3/2} \int_0^{\infty} \sqrt{E_{lab}}\; \exp\left[-\left(E_{lab}+E_{CN}-2\sqrt{E_{lab}E_{CN}}\cos\theta\right)/a\right]\;dE_{lab}. \end{equation} We find that a value of $a$ = 0.5 MeV gives a good representation of both the transmission as a function of the energy setting of the FMA, $E_{FMA}$ and the measured angular distribution, see Fig. 6. For the angular distribution we have also taken the effects of multiple scattering in the target and backing material as well as the charge state reset foil into account. This increases the width of the angular distribution somewhat and results in good agreement with the data as shown by the solid curve in Fig. 6b. This value of $a$ = 0.5 MeV corresponds to a transport efficiency of the FMA for evaporation residues of $\overline{\epsilon_{FMA}} \approx 0.60$, see Table I. \subsection{Angular distributions} The angular distributions of evaporation residues were measured at three beam energies utilizing the sliding-seal target chamber for the FMA. Differential cross sections, $d\sigma/d\Omega$, as a function of the mean angle, $<\theta_{lab}>$, relative to the beam axis are shown in the left side panel of Fig. 7. The right side panel shows the cross sections converted to $d\sigma/d\theta$, which is relevant for the angular integration of the total evaporation residue cross section. The angle integrated cross sections are thus derived from a fit to the data expressed in terms of $d\sigma/d\theta$ using the function $2\pi\sin\theta dP/d\Omega_{lab}$. The curves shown in the left side panel of Fig. 7 are computed by removing the $2\pi\sin\theta$ term. We observe that these latter curves underrepresent the differential cross section at small angles indicating that the angular distribution really has two components. However, we do not feel that the data are of sufficient quality to allow for a reliable separation of two components and by observing the fits to the $d\sigma/d\theta$ data it is clear that only a very small error could arise from this simplification. The data shown in Fig. 7 are corrected for the efficiency for transporting evaporation residues through the FMA. We estimate this transport efficiency, $\epsilon_{FMA}$, by folding the energy distribution of the evaporation residues with the energy acceptance of the FMA, which was measured by Back {\it et al} \cite{Back} for the entrance aperture used in this experiment. The mean energy of the compound system, $E'_{CN}$ ( corrected for the energy losses in the target material, backing and the reset foil ), is determined from the reaction kinematics and listed in Table I. The parameter, $a = \frac{2}{3}\langle E_{cm}\rangle$ was determined to have a value of about $a \approx 0.5$ for the $E_{beam} = $ 246 MeV point by simultaneously fitting the angular distribution of evaporation residues and a scan of the energy setting of the FMA, see Fig. 6. The value of $a$ was for the other beam energies scaled according to $a \propto \sqrt{E^*}/22.4$, which was found to reproduce also the angular distributions measured at $E_{beam}$ = 174 and 205 MeV, see Fig. 7. \subsection{Total evaporation residue cross sections} The total evaporation residue cross section, $\sigma_{ER}$, is obtained from the measurement of the differential cross section at $\theta = 5^\circ$, which was performed at all beam energies. The ratio, $f(E_{beam}) = \sigma_{ER}/\frac{d\sigma_{ER}}{d\theta}(5^\circ)$, of the angle integrated cross section, $\sigma_{ER}$, to the measured differential cross section, $\frac{d\sigma_{ER}}{d\theta}(5^\circ)$, is obtained by smooth interpolation between the values of $f(E_{beam})$ = 0.089, 0.088, and 0.086 rad obtained from the angular distribution measurements at $E_{beam}$ = 174, 205, and 246 MeV, respectively. The total evaporation residue cross sections are then given by \begin{eqnarray} \sigma_{ER}& =& f(E_{beam}) \frac{d\sigma(5^\circ)}{d\theta} \\ &=& f(E_{beam})\frac{N_{ER}(5^\circ)}{N_{mon}}\; \frac{\Omega_{mon}}{\Omega_{FMA}}\; 2\pi \sin(5^\circ)\frac{d\sigma_{Ruth}}{d\Omega}(30^\circ)\; \frac{1}{\epsilon_{FMA} \epsilon_{PPAC-Si} P(q)}\nonumber \end{eqnarray} where, $N_{ER}(5^\circ), N_{mon}$ are the number of evaporation residue counts observed at the FMA focal plane Si-detector, and the number of elastically scattered $^{32}$S ions registered in the monitor detector, respectively. The differential Rutherford cross section in the laboratory system is denoted $d\sigma_{Ruth}/d\Omega$, and $P(q)$ is the fraction of evaporation residues in the charge state, $q$, for which the FMA was tuned. The charge state fraction, $P(q)$, was obtained by interpolation of the central charge state, $q_0$ resulting from the fits to the measured distributions with a Gaussian with a standard deviation of $\sigma$ = 3 charge state units. The resulting evaporation residue cross sections for the $^{32}$S+$^{184}$W reaction are shown as filled circles in Fig. 8 and listed in Table I. The measurements are assigned a systematic error of 20\%, mainly due to the procedure for estimating the transport efficiency through the FMA. Fission-like cross sections and a derived estimate of the complete fusion cross sections for the $^{32}$S+$^{182}$W reaction are shown as open circles \cite{Glagola,Keller} and open squares \cite{Keller}, respectively, along with theoretical calculations using a modified Extra Push model \cite{Toke}. \section{Comparison with statistical model calculations} In Fig. 8, the evaporation residue data are compared with a statistical model calculation obtained with the code CASCADE (long dashed curve labeled $\gamma$=0) using Sierk fission barriers \cite{Sierk} scaled by a factor of 0.9 to approximately account for the cross section at low beam energies, and using level density parameters of $a_n = a_f = A/8.8$ MeV$^{-1}$. We observe that the measured cross section increases with beam energy, whereas the statistical model predicts a decreasing cross section because of an increased probability for fission during the longer evaporation cascades. For comparison we have also performed CASCADE calculations using level density parameters of $a_n = A/8.68$ MeV$^{-1}$ and $a_f = A/8.49$ MeV$^{-1}$ as suggested by T\={o}ke and Swiatecki \cite{Toke81} (dotted curve), and $a_n = A/11.26$ MeV$^{-1}$ and $a_f = A/11.15$ MeV$^{-1}$ by Ignatyuk {\it et al} \cite{Ignatyuk} (dotted-dashed curve). Using these values results in an even sharper decrease of the predicted evaporation residue cross section with beam energy as shown in Fig. 8. This is a consequence of the fact that the fission decay rate increases more rapidly with excitation energy when values of $a_f > a_n$ are used. Although it is expected that $a_f > a_n$ on rather firm theoretical grounds we have, however, used the standard values of $a_f = a_n = a/8.8$ in order to be able to compare to other works, where this value was used in the analysis. We hypothesize that the observed increase of the measured evaporation residue cross section with excitation energy, which is at variance with the statistical model calculations, can be attributed to an increased hindrance of the fission motion with excitation energy. Fission hindrance at high excitation has previously been shown to explain observations of enhanced emission of pre-scission neutrons \cite{Hinde}, charged particles \cite{Lestone}, and $\gamma$-rays \cite{Hofman}, as well as recent observation of an enhanced survival probability of excited target recoils from deep inelastic scattering reactions \cite{Hofman98}. The inclusion of friction in the fission motion results in a modification of the normal Bohr - Wheeler expression \cite{Bohr-Wheeler} for the fission decay width, $\Gamma_f^{BW}$ as pointed out by Kramers \cite{Kramers}, {\it i.e. } \begin{equation} \Gamma_f^{Kramers} = \Gamma_f^{BW} (\sqrt{1+\gamma^2} - \gamma)[1-\exp(-t/\tau_f)] \end{equation} where $\gamma = \beta/2\omega_0$ is a reduced nuclear friction coefficient, and $\tau_f$ is a charactistic time for the building of the fission flux over the saddle point. $\beta$ denotes the reduced dissipation constant and $\omega_0$ describes the potential curvature at the fission saddle point. The modification to the Bohr-Wheeler expression for the fission width thus consists of an overall reduction given by the so-called Kramers factor, $\sqrt{1+\gamma^2} - \gamma$, as well as a time dependent in-growth of the fission rate given by the factor $1-\exp(-t/\tau_f)$ \cite{Grange}. These modifications to the fission decay width has been incorporated into the CASCADE statistical model code in an approximate way \cite{Butsch}, which has, however, been shown \cite{Back93} to be very accurate over the applied range of parameters. Because the evaporation residue cross section is such a small fraction of the complete fusion cross section we find that it is very sensitive to the nuclear viscosity of the system inside the barrier. The thin solid curve in Fig. 8 represent a statistical model calculation where the effects of viscosity are included using a linear normalized dissipation coefficient of $\gamma$=5, corresponding to a strongly overdamped motion in the fission degree of freedom. This is approximately the dissipation strength expected from the one-body dissipation mechanism \cite{Blocki}. We see that this leads to an increase of about a factor 10-20 in the evaporation residue cross section relative to the pure statistical model estimate (long dashed curve), but the overall shape of the excitation function is virtually unchanged. Within this framework it therefore appears that the viscosity (or dissipation) increases rather rapidly over this range of beam energies i.e. from 200 to 260 MeV, which corresponds to an excitation energy range of $E_{exc}$=85-136 MeV. Similar effects have been observed in studies of pre-scission $\gamma$-rays \cite{Hofman} albeit in that case it appears to take place over an even smaller excitation energy interval. In order to deduce the temperature dependence of the dissipation strength in the fission degree of freedom $\gamma(T)$ that reproduces the observed increase of the evaporation residue cross section, we have performed a series of calculations at each beam energy, varying the value of $\gamma$ to reproduce the measured cross section. This procedure leads to the thick solid cross section curve going through the data points in Fig. 8; the corresponding values of $\gamma(T)$ are plotted as solid triangles in Fig. 9. Note that there is some inconsistency in this approach because the value of the dissipation strength is ${\it not}$ allowed to vary as the system cools down during the particle evaporation cascade. Rather, the dissipation strength is kept constant throughout the cascade with the value needed to fit the measured evaporation residue cross section for this particular beam energy. Although this has been recognized as a shortcoming of these calculations, we have employed this procedure to be able to compare to other published data analyzed in the same way. \section{Discussion} The dissipation strength in the fission process has recently been measured by several methods, and it is of interest to compare these different results. In Fig. 9 we show the normalized dissipation strength parameter, $\gamma$, obtained from the analysis of 1) the survival probability of Th-like nuclei excited in deep-inelastic scattering reaction of 400 MeV $^{40}$Ar+$^{232}$Th \cite{Hofman98} (solid squares), 2) the evaporation residue cross section and pre-scission $\gamma$-ray emission from the $^{16}$O+$^{208}$Pb \cite{Brinkmann,Hofman96} (solid diamonds), 3) the present data (solid triangles), and 4) the fission cross section for $^3$He+$^{208}$Pb reaction \cite{Rubehn,Back98} (open circles). We observe that the dissipation strength required to reproduce the different data falls into two groups, namely one which increases rather sharply above an excitation energy of $E_{exc} \sim$ 40 MeV, and another group that increases slowly only above $E_{exc} \geq$ 80 MeV. It is interesting to note that this behaviour may be related to the shell structure of the compound system. The two systems that have a closed (or nearly closed) neutron shell at N=126 show only moderate fission dissipation strength up to high excitation energy, whereas the mid-shell systems with N = 134 , 142 display a strong increase in $\gamma$ above $E_{exc} \sim$ 40 MeV. Recently, there has been much theoretical interest in the study of the dynamics of the fission process, both in terms of the description of experimental observables on the basis of phenomelogical assumptions of the dissipation strength \cite{Froebrich,Dhara} as well as more fundamental theories for the dissipation mechanism itself \cite{Hofmann,Kolomietz,Magierski,Mukhopadhyay}. Although the overall dissipation strength found to reproduce the present data is in fair agreement with estimated based on the simple one-body dissipation model, namely $\gamma \approx 5-6$ the rather striking increase with excitation energy (or temperature) is unexplained within this mechanism, which has no temperature dependence. It is interesting to note that the linear response theory approach \cite{Hofmann} appears to predict the increase in dissipation strength although the present development level of this theory is not directly applicable for comparison with the experimental data. \section{Conclusion} Measurements of evaporation residue cross sections for heavy fissile systems are shown to provide rather direct evidence for the fission hindrance (or retardation) which is caused by strong nuclear dissipation in the fission degree of freedom for hot nuclei. The data obtained for the $^{32}$S+$^{184}$W system show an increasing evaporation cross section with excitation energy, whereas a decrease is expected on the basis of statistical model considerations and calculations. The data indicate an increase in the linear normalized dissipation coefficient $\gamma$ from $\gamma$=0 at $E_{exc}$=85 MeV to $\gamma$=5 at $E_{exc}$=135 MeV. Although hints of such an increase have been obtained within the framework of linear response theory, no direct comparison can be made with the experimental data. Further study, both experimental and theoretical, of this phenomenon is warrented, This work was supported by the U. S. Department of Energy, Nuclear Physics Division, under contact No. W-31-109-ENG-38. \vspace{0.375in}
\subsection{A.~~Low-Mass Systems} It is now clear that stellar jets, with speeds of 100-300\,\hbox{km~s$^{-1}$}\ and densities of order $10^3\,\hbox{${\rm cm}^{-3}$}$, are responsible for accelerating much of the molecular gas in many of the youngest low-mass outflow systems (e.g. Richer et al.\ 1992; Padman and Richer 1994; Bachiller et al.\ 1995). However, there is also good evidence for momentum being deposited into the flows by wind components with wider opening angles, and that this component perhaps becomes relatively more powerful as the flows age (e.g. Bence et al.\ 1998). $$\psfig{file=fig1.ps,angle=270,width=5.2in}$$ \caption{Figure~1.\hskip\capskipamount The HH211 molecular jet mapped with the Plateau de Bure interferometer. The main panel shows the high-velocity CO jet in greyscale superposed on the lower velocity outflowing gas, which forms a cavity around the jet. The thick contours show the 1.3mm continuum emission. The panels top left and lower right show the fast and slow CO emission overlaid on the shocked H$_2$ line emission. Data taken from McCaughrean and Zinnecker 1996, and Gueth and Guilloteau 1999.} Some of the first clear evidence for a jet-dominated outflow origin was identified in the flows from L1448C (Bachiller et al.\ 1990; Guilloteau et al.\ 1992; Dutrey et al.\ 1997) and NGC2024-FIR5 (Richer et al.\ 1989, 1992). In L1448, a spectacular SiO jet is seen emanating from the driving source, aligned with the large-scale CO flow and unresolved across its width. In the NGC2024-FIR5 flow, the fastest gas is seen to form an elongated jet-like feature on the axis of nested shells of lower velocity gas. The collimation ratio $q$, defined as the width of the flow to the distance to the driving source, is as high as 30 for the high velocity (30\,\hbox{km~s$^{-1}$}) gas in this source, but only 4 or so for the lower velocity (5\,\hbox{km~s$^{-1}$}) envelope. Many other flows are now known which show this structure, with high-velocity elongated components tracing jet-like activity lying inside cavities of lower velocity gas; examples include L1157 (Gueth et al.\ 1996), HH111 (Cernicharo and Reipurth 1996; Nagar et al.\ 1997) and HH211 (Gueth and Guilloteau 1999). All of these flows appear to be driven by very young, low-mass objects, based on their low luminosities, and non-detection at even infrared wavelengths. The CO flow from HH211 (Fig.\ 1) is perhaps the most striking image to date of this phenomenon, showing an unresolved CO jet with high velocity gas ($v>10\,$\hbox{km~s$^{-1}$}\ with no correction for inclination) lying within an ovoid cavity of slower moving gas ($<10\,$\hbox{km~s$^{-1}$}). HH211 is probably also one of the youngest low-mass outflows known, having a {\it dynamical age} $\hbox{$\tau_{\rm d}$} = 0.07{\rm pc} / 10\,\hbox{km~s$^{-1}$} = 7000$ years; if the source is inclined close to the plane of the sky, as is expected given the clear separation of red and blue outflow lobes and the relatively modest projected flow speeds, the true dynamical age could well be a factor of 5 or so lower. It appears that HH211 is at the very start of its main accretion phase, and this perhaps explains the relative simplicity of the outflow structure. Note that the full opening angle at the base of the flow is only 22\hbox{$^{\circ}$}. It is natural to identify these so-called ``molecular jets'' as the deeply-embedded counterparts of the Herbig-Haro (HH) jets seen in less obscured systems and as the neutral counterparts of the ionized jets seen in the radio (see chapters by Eisl\"offel et al., and by Hartigan et al., this volume). Important work by Raga (1991) and Hartigan et al.\ (1994) demonstrated that HH jets were denser and hence more powerful than initial estimates suggested (e.g. Mundt et al.\ 1987), so that the total momentum flux in HH jets integrated over the lifetime of a typical source is sufficient to drive the observed CO flows (e.g. Richer et al.\ 1992; Mitchell et al.\ 1994). This unified picture of jet-driven outflows is entirely consistent with the observed CO structures discussed above. (Masson and Chernin 1993; Cabrit et al. 1997; Smith et al. 1997; Gueth and Guilloteau 1999). However, we caution that the actual composition of the driving jet, whether primarily atomic or molecular, is unknown. It is still unclear whether the CO jets seen in sources such as HH211 and NGC2024-FIR5 arises (1) from the body of a jet where molecules have formed in the gas phase (Glassgold et al.\ 1989; Smith et al.\ 1997); (2) from molecules formed in the post-shock region of shocks in the jet; or (3) from ambient molecular gas turbulently entrained along the jet's edge and at internal working surfaces (Raga et al.\ 1993; Taylor and Raga 1995). $$\psfig{file=fig2.ps,angle=0,height=120mm}$$ \caption{Figure~2.\hskip\capskipamount CO 2-1 image (in greyscale) of the RNO~43 molecular outflow (Bence at al.\ 1996); the driving FIR source is at the origin, marked with a cross. The solid white patches show the extent of the \hbox{H$\alpha$}\ emission. Note the large flow extent and correlation of the \hbox{H$\alpha$}\ emission with the CO hotspots. The panel to the right shows a detail from the northernmost \hbox{H$\alpha$}\ emission patch, showing the bow-shock shaped \hbox{H$\alpha$}\ structure in greyscale, and CO 2-1 contours overlaid. } RNO43-FIR is another low-luminosity outflow source, and is probably a flow in middle-age (Bence et al.\ 1996; Cabrit et al.\ 1988): the driving source, with $\hbox{$L_{*}$}=6$ \hbox{L$_{\odot}$} , is a heavily embedded Class~0 protostar invisible even at 2\hbox{$\mu{\rm m}$}, but the flow extends over a total size of 3\,pc. The outflow axis is close to the plane of the sky, so the image of CO integrated intensity shown in Fig.\ 2 shows both blue and red-shifted sides of the flow. The dynamical age is $\hbox{$\tau_{\rm d}$} \sim 1.5\,{\rm pc} / 10\,\hbox{km~s$^{-1}$} = 1.5\times10^5$ years which is an upper limit due to the small but unknown inclination angle. The CO flow is much more complex than HH211 and this most likely reflects the clumpy nature of the molecular gas through which the driving jet has passed; nonetheless, a very approximate S-shape symmetry can be seen about the driving source, both in the spatial and velocity data, suggesting that the jet direction has changed over time (Bence et al.\ 1996). A precessing jet model, with a full cone angle of 27\hbox{$^{\circ}$}\ and a precession timescale of order $3\times10^4$ years, provides a reasonable description of the overall flow properties (Bence et al.\ 1996); such precession could be driven by a binary companion in an inclined orbit. The evidence that this flow is jet-driven even at a distance of 1.5\,pc from the driving source is demonstrated by the \hbox{H$\alpha$}\ images of the object, which show strong emission coincident with the brightest CO points (see Fig.\ 2). In addition, at the northernmost CO feature in the flow, bright \hbox{H$\alpha$}\ coincident with a bow-shock shaped CO feature suggests that the whole of the RNO43 outflow can be explained by the gradual sweeping up of a clumpy molecular cloud by a powerful stellar jet whose ejection axis varies slowly with time. It is also interesting to note that parsec-scale flows such as RNO43 are the natural counterparts to the parsec-scale Herbig-Haro flows (Reipurth et al.\ 1997) seen in wide-field optical imaging: if there is molecular gas in the jet's path, then flows such as RNO43 result, whereas jets in essentially empty space such as HH34 (Devine et al.\ 1997) show only optical emission. This points strongly to the jets being primarily atomic in composition. However, in the HH111 outflow system, CO ``bullets'' associated with the optical jet and having similar velocities to the optical gas have been detected far beyond the molecular cloud boundary (Cernicharo and Reipurth 1996). This suggests that in some cases the jets may have a molecular component. Not all low-mass flows are jet-dominated: many of the older flows show CO emission dominated by low-velocity cavities with little evidence for elongated high-velocity CO jet features. Examples include L43 (Bence et al.\ 1998), L1551 (Moriarty-Schieven et al.\ 1988) and B5 (Velusamy and Langer 1998). These flows are typically $10^{5-6}\,$years old, and associated with nebulosities visible in the optical or at 2\,\hbox{$\mu{\rm m}$}, suggesting they are Class~I or II protostars, older than the Class~0 objects responsible for the HH211, RNO43 and NGC2024-FIR5 outflows. The lack of high-velocity CO and obvious jet-like features in these flows, in contrast to their younger counterparts such as HH211, and the presence of much wider cavities at the base of the flows, suggests that the jet power has declined over time, and that a wider opening angle wind is now primarily responsible for driving the outflow. However, even in these older sources there is usually some evidence for weak jet activity: L1551 has an optical jet and extended Herbig-Haro emission apparently on one of its cavity walls (Davis et al.\ 1995; Fridlund and Liseau 1998), as well as fast CO emission suggestive of a jet-origin (Bachiller et al.\ 1994), and B5 has optical jets stretching 2.2 pc away from the driving source. However, the L43 flow shows no signs of shocks, jets or very fast CO, and this may be a true ``coasting'' flow, which is no longer being accelerated by a stellar wind or jet. The recent interferometric images of the B5 outflow (Velusamy and Langer 1998) reveal a beautiful example of wide, hollow cavities at the base of the outflow, much like the reflection nebulosities seen in the near-infrared in many of these systems; in order to reproduce the clearly separated red- and blue-shifted emission lobes, the CO must be flowing along these cavity walls, presumably being accelerated by a poorly-collimated radial wind from the star. The entire outflow driven by B5 has a dynamical age of $10^6\,$years, and a collimation factor of about 5. The cavity opening angle at the protostar is extremely large, in the range $90-125\hbox{$^{\circ}$}$, and Velusamy and Langer (1998) suggest that if this angle further broadens with time it may ultimately cut off the accretion flow. This idea is consistent with most of the available data: the young flows such as HH211, RNO43 and L1157 have opening angles less than 30\hbox{$^{\circ}$} or so, while the older flows such as L1551, L43 and B5 are significantly greater than 90\hbox{$^{\circ}$}. With maps at good resolution of a larger sample of outflows, it will be possible to test the hypothesis that flow opening angle is a measure of the source age. \subsection{B.~~High-Mass Systems} Our understanding of massive flows is beginning to change because we are starting to find a few isolated systems that can be studied in depth. However, we are still observationally biased toward older flows that are easier to identify and study. This bias is likely to diminish in the future as more massive young flows are identified (e.g. Cesaroni et al.\ 1997; Molinari et al.\ 1998; Zhang et al.\ 1998). Most luminous YSOs have relatively wide-opening angle outflows as defined by their CO morphology. Although the statistics are poor because few massive flows have been studied with sufficient resolution to adequately determine the morphology, collimation factors {\it q} for 7 well-mapped flows produced by YSOs with $\hbox{$L_{\rm bol}$} > 10^3\,\hbox{L$_{\odot}$}$ range from 1 to 1.8 (NGC 7538 IRS1: Kameya et al.\ 1989; HH~80-81: Yamashita et al.\ 1989; NGC 7538 IRS9: Mitchell et al.\ 1991; GL 490: Mitchell et al.\ 1995; Ori~A : Chernin and Wright 1996; W75N: Davis et al.\ 1998; G192.16: Shepherd et al.\ 1998). The dynamical time scales for these outflows range from 750 years to $\sim 2 \times 10^5$ years and there is no obvious dependence of flow collimation on age. In comparison, collimation factors in low-mass outflows range from $\sim 1$ to as high as 10 with a typical value being $\sim$~2 or 3 (Fukui et al.\ 1993 and references therein). It appears that more luminous YSOs do not in general produce very well-collimated CO outflows and this result is independent of outflow age, unlike outflows produced by low-luminosity YSOs. This may be due to the fact that outflows from luminous YSOs tend to break free of their molecular cloud core at a very early stage in the outflow process. Hence, massive molecular flows are frequently the truncated base of a much larger outflow that extends well beyond the cloud boundaries (e.g. HH~80--81: Yamashita et al.\ 1989; DR21: Russell et al.\ 1992; G192.16: Devine et al.\ 1999). The outflow from Orion~A is perhaps the best known and best studied high-mass outflow system. The driving source is believed to be an O star, and the estimated age of the flow is $\sim 750$ years which makes this one of the youngest known outflows. The flow differs significantly from those produced by low-luminosity sources and represents a spectacular example of the outflow phenomenon. The CO outflow is poorly collimated at all velocities and spatial resolutions and the H$_2$ emission is dispersed into a broad fan shape that is unlike any other known outflow. Its morphology is highly suggestive of an almost isotropic, explosive origin. McCaughrean and Mac~Low (1997) model the H$_2$ bullets as a fragmented stellar wind bubble using the fragmentation model of Stone et al.\ (1995) and suggest that the bullets are caused by several young sources within the BN-KL cluster. Chernin and Wright (1996) argue that the flow is driven by a single massive YSO, source I. The estimated opening angle, corrected for inclination, is approximately 60\hbox{$^{\circ}$}\ for the blue lobe and 120\hbox{$^{\circ}$}\ for the red lobe. $$\psfig{file=fig3.ps,angle=270,width=4.8in}$$ \caption{Figure~3.\hskip\capskipamount The G192.16 outflow mapped with the OVRO interferometer. Contours of redshifted CO (thick lines) and blueshifted CO (thin lines) delineate the bipolar flow emanating from a dense core traced by the central dust continuum emission represented in greyscale. } The outflow from G192.16 is perhaps more typical of outflows from B stars. Figure 3 shows an interferometric CO J=1-0 image of the outflow together with a closeup of the 3~mm continuum emission showing a flattened distribution of hot dust (Shepherd et al.\ 1998; Shepherd and Kurtz 1999). The 3~pc long molecular flow represents the truncated base of a much larger flow identified by \hbox{H$\alpha$}\ and [SII] emission that extends almost 5~pc from the YSO (Devine et al.\ 1998). The CO flow is $\sim 10^5$ years old and appears to be driven by an early B star. Despite the very different masses and energies involved, the CO morphology looks very similar to L1551-IRS5: the extent of the G192.16 CO outflow is $2.6 \times 0.8$~ pc, approximately twice that of L1551, but the mass in the outflow ($\approx95\,\hbox{M$_{\odot}$}$) is much greater. The opening angle at the base of the flow is $\sim 90\hbox{$^{\circ}$}$. The present lack of well-collimated CO outflows with $q >$ 3 from YSOs with $\hbox{$L_{\rm bol}$} > 10^3\hbox{L$_{\odot}$}$ does not mean that jets and well-collimated structures are not present in these massive sources. For example, the central source in HH~80-81 ($\hbox{$L_{\rm bol}$}\sim 2 \times 10^4$ L$_{\odot}$) powers the largest known Herbig-Haro jet with a total projected length of 5.3 pc, assuming a distance of 1.7 kpc (Mart{\'\i} et al.\ 1993, 1995 and references therein). However, the CO flow appears poorly collimated with $q\sim 1$ when mapped at moderate resolution (Yamashita et al.\ 1989). Also, the biconical thermal radio jet from Cepheus A HW2 ($\hbox{$L_{\rm bol}$} \sim 10^4$ L$_{\odot}$) appears to be responsible for at least part of the complicated molecular flow seen in CO and shock-enhanced species such as H$_2$, SiO, and SO (e.g., Doyon and Nadeau 1988, Mart{\'\i}n-Pintado et al. 1992, Hughes 1993, Torrelles et al. 1993, Rodr{\'\i}guez et al. 1994, Rodr{\'\i}guez 1995, Garay et al. 1996, Hartigan et al. 1996, Narayanan and Walker 1996). The HH~80-81 and Cepheus A HW2 systems demonstrate that high-luminosity YSOs can produce well-collimated jets like those found in association with less luminous stars, even though the CO flow may appear chaotic or poorly collimated. Other examples of possible jets in massive outflows include IRAS~20126 and W75N IRS1. IRAS~20126 ($\hbox{$L_{\rm bol}$} \sim 1.3 \times 10^4\,\hbox{L$_{\odot}$}$) appears to drive a compact jet seen in SiO and H$_2$ (Cesaroni et al.\ 1997). W75N IRS1 ($\hbox{$L_{\rm bol}$} \sim 1.4 \times 10^5\,\hbox{L$_{\odot}$}$) shows H$_2$ 2.12$\mu$m\ shock-excited emission at the end and sides of the CO lobe, with a morphology and emission characteristics highly suggestive of a jet bowshock (Davis et al.\ 1998). However, this ``bowshock'' is 0.3\,pc wide, i.e. 30 times larger than the H$_2$ 2.12$\mu$m\ bowshock at the end of the HH211 flow (cf. Fig.~1). It is not fully clear whether such wide bows are created by protostellar jets, or by a low collimation wind component. Scaling up from current hydrodynamical jet simulations (e.g. Suttner et al. 1997), one would need a jet radius of $\sim$ 0.03\,pc at 1.3\,pc from the star, hence a jet opening angle $\sim$ 2.6$^{\circ}$. \subsection{C.~~Outflow Velocity Structure} From the above discussion, we conclude that jet activity, bow shocks, and CO cavities are common to outflows from low and high-mass systems. There is some evidence that high-mass systems are less well collimated than low-mass ones, but this may be due to selection effects --- young, high-mass systems with small opening angles may simply be missing from the small sample currently known (although the Orion outflow does appear to be very young). This conclusion suggests it is sensible to consider the possibility that a common driving mechanism is responsible for all outflows. Several authors have noted that molecular flows seem to be characterized by a power-law dependence of flow mass $\hbox{$M_{\rm CO}$} (v)$ as a function of velocity. The power-law exponent $\gamma$ (where $\hbox{$M_{\rm CO}$} (v) \propto v^{\gamma}$) is typically $\sim -1.8$ for most low-mass outflows, although the slope often steepens at velocities greater than 10~{\hbox{km~s$^{-1}$}} from $v_{LSR}$ (e.g., Masson and Chernin 1992; Rodr{\'\i}guez et al.\ 1982; Stahler 1994; Chandler et al.\ 1996; Lada and Fich 1996; Gibb, personal communication). $$\hbox{\psfig{file=fig4a.ps,height=61mm} \hskip1pc\psfig{file=fig4b.ps,height=61mm}}$$ \caption{Figure~4.\hskip\capskipamount The slope $\gamma$ of the mass spectrum $M(v)$ plotted as a function of (a) source bolometric luminosity and (b) flow dynamical age \hbox{$\tau_{\rm d}$}. Triangles represent $\gamma$ for gas with projected speeds less than 10~{\hbox{km~s$^{-1}$}} relative to the source, and squares are for gas moving at more than 10~{\hbox{km~s$^{-1}$}}. The dashed, horizontal line separates the low and high-velocity $\gamma$'s for clarity. Sources with $\hbox{$L_{\rm bol}$} < 10^3\,\hbox{L$_{\odot}$}$ are plotted with open symbols, those with $\hbox{$L_{\rm bol}$} > 10^3\,\hbox{L$_{\odot}$}$ with filled symbols. Representative error bars are displayed in the lower left corner of each plot. The solid line in Fig.\ 4b is a linear least squares fit (slope $-1.7 \pm 0.6$) to high-velocity $\gamma$'s in luminous sources versus log($t_{dyn}$). The sources plotted here are VLA1623, IRAS03282, L1448-C, L1551-IRS5, NGC2071-IRS1, and Ori~A~IRC2 (Cabrit and Bertout 1992); TMC-1 and TMC-1A (Chandler et al.\ 1996); L379-IRS1-S (Kelly and MacDonald 1996); NGC2264G (Lada and Fich 1996); G5.89-0.39 (Acord et al. 1997). CephE (Smith et al.\ 1997) HH251-254, NGC7538-IRS9, W75N-IRS1 and NGC7538-IRS1 (Davis et al.\ 1998); G192.16 (Shepherd et al.\ 1998); and HH26IR, LBS17-H, G35.2-0.74, HH25MMS (Gibb, personal communication).} Figure 4 plots (a) $\gamma$ versus $\hbox{$L_{\rm bol}$}$ and (b) $\gamma$ versus the dynamical time scale $t_{dyn}$ for a new compilation of 22 sources with luminosities ranging from 0.58 L$_\odot$ to $3 \times 10^5$ L$_\odot$. Triangles represent slopes derived from gas moving less than 10~{\hbox{km~s$^{-1}$}} relative to $v_{LSR}$ while squares represent slopes derived from gas moving more than 10~{\hbox{km~s$^{-1}$}} relative to $v_{LSR}$. Sources with $\hbox{$L_{\rm bol}$} < 10^3\,\hbox{L$_{\odot}$}$ are plotted with open symbols while those with $\hbox{$L_{\rm bol}$} > 10^3\,\hbox{L$_{\odot}$}$ are plotted with filled symbols. Both well-collimated and poorly-collimated outflows are represented in the sample. The most striking result from Fig.\ 4a is that $\gamma$'s for low-velocity gas are similar in sources of all luminosities. This suggests that a common gas acceleration mechanism may operate over nearly six decades in \hbox{$L_{\rm bol}$}. In addition, there is a clear separation between $\gamma$'s in high and low-velocity gas which supports the interpretation that there are often two distinct outflow velocity components, perhaps corresponding to a recently accelerated component and a slower, coasting component. Hydrodynamic simulations of jet-driven outflows from low-lu\-m\-in\-os\-ity YSOs predict such a change of slope at high velocity. They also predict that $\gamma$ should steepen over time possibly due to the collection of a reservoir of low-velocity gas (Smith et al.\ 1997). Figure 4b reveals marginal evidence that the mass spectrum in flows from luminous YSOs does becomes steeper with time, both in the low and high velocity ranges. The solid line in Fig.\ 4b is a linear least squares fit (slope $-1.7 \pm 0.6$) to high-velocity $\gamma$'s in luminous sources versus log($t_{dyn}$). There is no indication of time evolution of the mass spectrum slopes in outflows from low-luminosity sources. The decrease of $\gamma$ with time in more luminous sources may be due to a difference in the driving mechanism, or it may simply be more prominent because the mass outflow rate is several orders of magnitude greater than in outflows from low-luminosity YSOs (thus allowing more precise determination of $\gamma$) and because their flow ages cover a broader range, from 750 to $2\times 10^5$ yrs. \mainsection{{I}{I}{I}.~~TESTS OF PROTOSTELLAR WIND AND ACCRETION MODELS} \backup \subsection{A.~~Flow Energetics} It is well known that outflow energetics correlate reasonably well with \hbox{$L_{\rm bol}$}\ over the entire observed luminosity range. In Fig.\ 5a, we show a recent compilation of the mean momentum deposition rate \hbox{$F_{\rm CO}$}\ as a function of bolometric luminosity of the driving star. It must be remembered that \hbox{$F_{\rm CO}$}\ is the time-averaged force required to drive the CO outflow: $\hbox{$F_{\rm CO}$}=\hbox{$M_{\rm CO}$}\,\hbox{$v_{\rm CO}$}/\hbox{$\tau_{\rm d}$}$, where we assume $\hbox{$\tau_{\rm d}$}$ is a good approximation to the flow age. If the CO-emitting material is accelerated by a separate stellar wind (or jet), the wind momentum flux \hbox{$F_{w}$}\ may be quite different from \hbox{$F_{\rm CO}$}, depending on the nature of the wind-cloud interaction. There are clearly large uncertainties in the measured flow properties, primarily due to difficulties in estimating the inclination angle of the outflow, the optical depth and excitation of the CO, and the correctness of the assumption that the dynamical timescale is a good estimate of the flow age (Cabrit and Bertout 1992; Padman et al.\ 1997). However, as seen in Fig.\ 5a, the correlations are roughly consistent with a single power law (here of slope $\sim$ 0.7) across the full range of source luminosities. It has been suggested that this correlation argues for a common entrainment and/or driving mechanism for molecular flows of all masses; but this is by no means a compelling argument, given that we would surely expect most physically reasonable outflow mechanisms to generate more powerful winds if the source mass and luminosity are increased. Regardless of the details of how the stellar wind or jet entrains material, if the shock cooling times are short compared to flow dynamical timescales (as we expect for wind speeds less than 300\,\hbox{km~s$^{-1}$}: Dyson [1984], or if there is efficient mixing at the wind/molecular gas interface: Shu et al.\ [1991]), then the wind and molecular flow momenta will be equal. This is often called the momentum-conserving limit. Then, molecular outflows represent a good opportunity to test proposed protostellar ejection mechanisms. In particular we show in this section that they can be used to estimate the ratio $f$ of ejection to accretion rates that would be necessary if the wind is accretion-powered. We first recall that optically thin, line-driven radiative winds (such as those present in main-sequence O and B stars) are insufficient to drive the flows if the flows are momentum driven. The solid line in Fig.\ 5a shows the maximum momentum flux available in stellar photons in the single-scattering limit, $\hbox{$L_{\rm bol}$}/c$. It falls short of the observed amount in molecular flows by one to three orders of magnitude. If flow lifetimes have been underestimated, the discrepancy is reduced, but not by a sufficient amount. In principle, higher momentum flux rates could be reached with multiple scattering. Wolf-Rayet stars ($L_{\rm bol} \sim$ a few $10^5$ \hbox{L$_{\odot}$}) have a wind force reaching $20-50L_{\rm bol}/c$, similar to the momentum rate in molecular flows from sources of comparable luminosity (cf. Fig. 5a). Such high values are attributed to multiple scattering of each photon by many lines closely spaced in frequency (e.g. Gayley et al.\ 1995 and refs. therein). However, excitation conditions in protostellar winds are very different from those in hot, ionized Wolf-Rayet winds. The dashed line in Fig.\ 5a shows the typical momentum flux in the {\it ionized} component of protostellar winds, inferred from recombination lines or radio continuum data (Panagia 1991): the values lie a factor of 10 below \hbox{$F_{\rm CO}$}, implying that the driving winds must be 90\% neutral. If dust grains instead provide the dominant opacity source in protostellar winds, comparison with winds from cool giants and supergiants might be more appropriate. Their wind velocities are $\simeq$ 5-30 km s$^{-1}$ and mass-loss rates do not exceed 10$^{-4}$ \hbox{M$_{\odot}$}\ \hbox{yr$^{-1}$}, hence $F_w \le 0.5-3L_{\rm bol}/c$ for a typical $L_{\rm bol} \sim 5\times 10^4$ \hbox{L$_{\odot}$}. Calculations by Netzer and Elitzur (1993) show that 10 times larger mass-loss rates could in principle be achieved in oxygen-rich stars, where silicates dominate the opacity curve. The wind force could then become comparable to the molecular outflow momentum rate for $L_{\rm bol} \sim 5\times 10^4\,\hbox{L$_{\odot}$}$. Hence it is just possible that luminous stars ($\hbox{$L_{*}$}>5\times10^4\,\hbox{L$_{\odot}$}$) could drive their flows by radiative acceleration if dust opacity plays a significant role; further work is needed to investigate if such models are viable. In lower luminosity sources, however, the opacity required to lift material above escape speeds largely exceeds typical values for circumstellar dust. Therefore, molecular outflow sources with $\hbox{$L_{*}$}<5\times10^4\hbox{L$_{\odot}$}$ must possess an efficient non-radiative outward momentum source. It thus appears most likely that both low and high-mass systems possess an efficient non-radiative wind-generation mechanism in their embedded protostellar phase. If energetic bipolar winds are the chief means of angular momentum loss during the main accretion phase for stars of all mass, this is not surprising. However, the details of the wind ejection mechanism could differ; in particular massive accreting stars are likely to have thinner convective layers and probably rotate faster, so that the magnetic field and accretion geometry close to the star may be very different from that in low-mass stars. There exist quite a number of efficient accretion-powered wind mechanisms from the stellar surface, disk, or disk-mag\-ne\-to\-sphere bou\-nd\-ary. The most efficient models use a strong magnetic field in the star or disk to drive the wind and to carry off angular momentum from the accreting gas (see e.g. reviews by K\"onigl and Pudritz, and Shu et al., this volume). This wind is further collimated by magnetic or hydrodynamic processes (e.g. Mellema et al.\ 1997), generating a high-Mach number wind and/or jet with a speed of order 200-800\,\hbox{km~s$^{-1}$}. A fraction $f$ $<1$ of the accretion flow \hbox{$\dot{M}_{a}$}\ is ejected in the wind: $\hbox{$\dot{M}_{w}$} = f\,\hbox{$\dot{M}_{a}$}$. A different scenario recently explored by Fiege and Henriksen (1996a, 1996b) is that molecular flows are not predominantly swept-up by an underlying wind, but represent infalling gas that has been deflected into polar streams by magnetic forces. Only a small fraction of the infalling gas actually reaches the star to produce accretion luminosity. In that case, $f > 1$. This model has been invoked in particular to explain the large flow masses $\hbox{$M_{\rm CO}$} \sim 10M_*$ observed in flows from luminous sources. In the following we use molecular flow observations to set constraints on these two classes of proposed models. \subsection{B.~~The Ejection/Accretion Ratio in Protostellar Winds} Two simple, independent methods have been used in the literature to estimate $f$, assuming that the driving mechanism is steady over the source lifetime. First, in low-mass protostars where the luminosity is accretion-dominated, it is possible to use the observed correlation of flow force with \hbox{$L_{\rm bol}$}\ (Fig.\ 5a) to derive the ejection fraction $f$. The source bolometric luminosity is $\hbox{$L_{\rm bol}$}=G\hbox{$M_{*}$}\hbox{$\dot{M}_{a}$}/\hbox{$R_{*}$}=\hbox{$\dot{M}_{a}$}\,\hbox{$v_{K}$}^2$, where \hbox{$v_{K}$}\ is the Keplerian speed at the stellar surface. The flow and wind force (assuming momentum conservation) is $\hbox{$F_{\rm CO}$} = f\,\hbox{$\dot{M}_{a}$}\,\hbox{$v_{w}$}$. Hence $\hbox{$F_{\rm CO}$}/\hbox{$L_{\rm bol}$} = f\,\hbox{$v_{w}$}/\hbox{$v_{K}$}^2$. A value $f \sim 0.1$ is inferred in both Class 0 and Class I low-luminosity objects (Bontemps et al.\ 1996). Alternatively, if one estimates \hbox{$M_{*}$}\ from \hbox{$L_{\rm bol}$}\ via the ZAMS relationship (which is probably only valid for sources with \hbox{$L_{\rm bol}$} $> 10^3$ \hbox{L$_{\odot}$}), one can use the {\it accumulated} flow momentum to infer $f$, using: $\hbox{$P_{\rm CO}$} = \hbox{$v_{w}$} \, \hbox{$M_{w}$} = f \, \hbox{$M_{*}$}\, \hbox{$v_{w}$}$. Values of $f$ ranging from 0.1-1 are inferred for a wind speed of 150 \hbox{km~s$^{-1}$}\ (Masson and Chernin 1994; Shepherd et al.\ 1996). However, we point out that wind speeds are unlikely to remain constant over the whole \hbox{$L_{\rm bol}$}\ range. There is evidence for higher wind velocities in luminous flow sources: for example, proper motions up to 1400\,\hbox{km~s$^{-1}$}\ are seen in HH80-81 (Mart\'{\i} et al.\ 1995) and Z~CMa shows optical jet emission with speeds up to 650\,\hbox{km~s$^{-1}$}\ (Poetzel et al. 1989). These are significantly higher than the 100-200\,\hbox{km~s$^{-1}$}\ typically seen in low-mass sources. To re-examine this issue in a homogeneous way over the whole luminosity range, we plot in Fig.\ 5b the values of $f\hbox{$v_{w}$}/\hbox{$v_{K}$}$ obtained using a new combination of the above two methods (Cabrit and Shepherd, in preparation). The solid circles are derived assuming the luminosity is accretion dominated, while the open circles (for sources more luminous than $10^3\,\hbox{L$_{\odot}$}$) assume the ZAMS relationship given above. Typically, \hbox{$v_{K}$}\ ranges from 100\,\hbox{km~s$^{-1}$}\ in low-luminosity sources to 800\,\hbox{km~s$^{-1}$}\ in high-luminosity sources. A rather constant value $f \hbox{$v_{w}$}/\hbox{$v_{K}$} \sim 0.3$ seems to hold over the whole range of \hbox{$L_{\rm bol}$}. This value is in line with both popular MHD ejection models. In the X-wind model (Shu et al.\ 1994, and this volume), the wind is launched close to the stellar surface ($\hbox{$v_{w}$} \sim \hbox{$v_{K}$}$) and a large fraction of the accreting gas is ejected ($f \sim 0.3$). In the self-similar disk-wind models (Ferreira 1997; K\"onigl and Pudritz, this volume), less material is ejected ($f \sim 0.03$) but the long magnetic lever arm accelerates it to many times the Keplerian speed ($\hbox{$v_{w}$} \sim 10\hbox{$v_{K}$}$). Thus we conclude that the {\it energetics} of the flows over the entire luminosity range are broadly consistent with a unified MHD ejection model for all flow luminosities, but we reiterate that given the very different physics involved around high and low-mass protostars, the details of such a model for high-mass systems are still unclear. $$\psfig{file=fig5.ps,angle=270,height=99mm}$$ \caption{Figure~5.\hskip\capskipamount (a) The average momentum flux in the CO flows as a function of source luminosity. The solid line shows the force available in stellar photons (assuming single scattering), and the dashed line the force available from the ionized wind components (Panagia 1991). (b) The factor $f\hbox{$v_{w}$}/\hbox{$v_{K}$}$ derived for the same set of sources is plotted as solid circles (small open circles assume that luminous sources are on the ZAMS). Large open circles show the required $\langle U(\theta)\rangle$ factors for the same objects if the circulation models are applicable (see text for details). } We stress again that this plot assumes perfect momentum conservation in the wind/flow interaction. There could be strong deviations from this key assumption. First, we should keep in mind the possibility that massive flows enter the energy-driven regime: for high wind speeds, the gas cooling behind the shock will be slow, and the snowplow or momentum-conserving flow will turn into an energy driven one. In this case, the shell momentum can exceed the momentum in the wind itself by a factor of order $\hbox{$v_{w}$}/\hbox{$v_{\rm CO}$}$ (Cabrit and Bertout 1992; Dyson 1994) so reducing markedly the momentum requirement of the driving wind. Of course, there are objections to energy driven flows, in particular their inability to reproduce the bipolar velocity fields of many flows (Masson and Chernin 1992); but given the apparently poorer collimation of high-mass outflows this issue should perhaps be re-examined. Second, if the flow is entrained in a jet bow-shock, the efficiency of momentum transfer will depend on the ratio of ambient to jet density: it will only be close to 1 if the jet is less dense than the ambient medium. Highly overdense jets will pierce through the cloud without depositing much of their momentum (Chernin et al.\ 1994), and in that case the ratio $f\hbox{$v_{w}$}/\hbox{$v_{K}$}$ plotted in Fig.\ 5b would have to be increased. We conclude that for current protostellar jet models to apply, we have the additional condition that most of the jet momentum must be in a component that is not significantly denser than the ambient molecular cloud on scales of 0.1-1pc. The above analysis also provides important constraints on accretion rates in protostellar objects of various masses: if $f\hbox{$v_{w}$}/\hbox{$v_{K}$} \sim 0.3$, then the values of \hbox{$F_{\rm CO}$}\ in Fig.\ 5a show that $\hbox{$\dot{M}_{a}$} = 3\hbox{$F_{\rm CO}$}/\hbox{$v_{K}$}$ must range from a few $10^{-6}$ \hbox{M$_{\odot}$} \hbox{yr$^{-1}$} at \hbox{$L_{\rm bol}$} $\sim$ 1\hbox{L$_{\odot}$}\ to a few $10^{-3}$ \hbox{M$_{\odot}$} \hbox{yr$^{-1}$} at \hbox{$L_{\rm bol}$} $\sim 10^5$ \hbox{L$_{\odot}$}. Then, protostellar sources would not be characterized by a single infall rate across the whole stellar mass range, and massive stars would form with much higher infall rates than low-mass stars (Cabrit and Shepherd, in preparation). The wealth of data available on low-mass systems also allows one to break the samples down by estimated age, and so look for {\it evolution} of outflow properties with stellar age. Bontemps et al.\ (1996) made an important study of low-mass systems in Taurus and Ophiuchus, and found evidence for a secular decline in outflow power with age, while $f$ remained constant. Intriguingly, the best correlation of outflow power was with circumstellar mass (as measured by the millimeter continuum flux) rather than with source bolometric luminosity; Saraceno et al (1996) also presented a similar correlation between millimeter continuum flux and outflow kinetic luminosity in systems with $L<10^3\,\hbox{L$_{\odot}$}$. These results strongly suggest that the accretion rate and the outflow strength both decline in proportion to the disk and envelope mass. \subsection{C.~~Deflected Infall Models} Although the above analysis shows that $f$ need not be necessarily higher in high-mass outflows, the very large masses involved, combined with the inefficiency of entrainment and momentum transfer by dense jets (especially once they escape their parent clouds), has led to some discussion of whether the sweeping up of ambient molecular gas by an accretion driven wind is a viable mechanism for these objects (e.g. Churchwell 1997a). A recent class of outflow models, termed circulation models, can naturally generate outflow masses much greater than the stellar mass. In these, most of the infalling circumstellar material is diverted magnetically at large radii into a slow-moving outflow along the polar direction, while infall proceeds along the equatorial plane (Fiege and Henriksen 1996a, 1996b). The main attraction of these models is that they generate large outflow masses for even small stellar masses and can generally explain the observed opening angles and velocity structure seen in high-mass systems. In particular, self-similar models predict that the velocity and density laws should take the form $V(r,\theta) = U(\theta) \sqrt{GM_*/r}$ and $\rho(r,\theta) = \mu(\theta) M_* r_o^{-3} (r/r_o)^{2\alpha-0.5}$, where $0.25 \ge \alpha > -0.5$, $r_o$ is an unspecified radial scale, and $U(\theta)$ and $\mu(\theta)$ are dimensionless functions. It is then straightforward to show that the force and mass-flux in the outflow are related by $\hbox{$F_{\rm CO}$}/\dot{M}_{CO} = \sqrt{G\hbox{$M_{*}$}/R_{CO}}\langle U(\theta)\rangle$ where $\langle U(\theta) \rangle$ = $\int{U(\theta)^2 \mu(\theta) d\omega} / \int{U(\theta) \mu(\theta) d\omega}$ is the density-weighted average velocity over the outflow solid angle. The inferred $\langle U(\theta)\rangle$ (using the same \hbox{$M_{*}$}\ as for our estimates of $f \, \hbox{$v_{w}$} / \hbox{$v_{K}$}$) is plotted in the top part of Fig.\ 5b as open circles. It is clear that high values (between 1000 and 10) are required. Generalized circulation models that include Poynting flux driving yield values of $\langle U(\theta) \rangle$ between 5 and 200 (Lery, Henriksen, Fiege, in preparation). In model cases where radiation transport is important in setting up the flow, a power-law slope of 0.8 is predicted between $F_{CO}$ and \hbox{$L_{\rm bol}$}, which is close to the observed slope $\sim 0.7$ (see also Henriksen 1994). Observations seem to indicate a systematic decline of $\langle U(\theta) \rangle$ with \hbox{$L_{\rm bol}$}, which would also have to be explained. There are several concerns about these circulation models. First, there are many {\it unipolar} CO outflows known, such as NGC2024-FIR5 (Richer et al.\ 1992), and the almost-unipolar HH46-47 system (Chernin and Masson 1991). These are naturally explained by swept-up wind models if the protostar is forming on the edge of a cloud or close to an HII region interface: the jet or wind propagating into the cloud will then sweep up a large CO flow, while in the opposite direction little evidence for a CO lobe will be seen. In circulation models, anisotropic solutions may also occur, but it is unclear why the weaker lobe would necessarily be on the side where the large-scale cloud density is low Second, in some objects such as B5 (Velusamy and Langer 1998), there is an apparent lack of molecular material in the equatorial plane which can feed a circulation flow. Third, as discussed in section II, in some high-mass systems such as HH80-81 there is direct evidence for fast jets and bow-shock entrainment of molecular gas. Consequently, it seems more probable given the evidence presented that even high-mass outflows can be generated by the sweeping up of ambient cloud material by an accretion-driven stellar wind or jet. The details of the MHD driving mechanism in these cases, and of the momentum transfer between the wind and the jet, remain open issues. \mainsection{{I}{V}.~~SHOCK CHEMISTRY AND ENERGETICS} The interaction between a supersonic protostellar wind and surrounding quiescent material is expected to drive strong shock fronts. Shocks can be of type C (continuous) or J (jump) depending upon the shock velocity, the magnetic field, and the ionization fraction of the pre-shock gas (Draine and McKee 1993; Hollenbach 1997). In the last few years, spectacular gains in sensitivity in the mm and IR domains have allowed us for the first time to witness the chemical and thermal effects of these shocks in molecular outflows. These observations yield direct estimates of the flow age, energetics, and entrainment conditions which represent an important new step toward a complete description of the outflow phenomenon. \subsection{A.~~Chemical Processing of ISM in Molecular Flows} {\it Theoretical expectations:} By compressing and heating the gas, shock waves trigger new chemical processes which lead to a specific ``shock chemistry'' (see chapter by Langer et al., this volume). The most active molecular chemistry is expected to occur in C-shocks, as they increase the temperature to moderate values of about 2000 K in a thick layer where molecules can survive, and reactions that overcome energy barriers can proceed. In particular, the very reactive OH radical can be formed by O+H$_2$ $\to$ OH + H (which has an energy barrier of 3160 K), and will contribute to the formation of H$_2$O by further reaction with H$_2$: OH + H$_2$ $\to$ H$_2$O + H (energy barrier: 1660 K). In dissociative J-shocks, molecules are destroyed in the hot (T $\sim$ 10$^5$ K) thin post-shock layer and only reform over longer timescales, in a plateau of gas at $\sim$ 400 K. Since some of these chemical processes are fast, and the cooling times are short, the chemical composition of the shocked regions is expected to be strongly time-dependent. Shocks also process dust grains. In the most violent J shocks, destruction of grain cores and thermal sputtering inject refractory elements (such as Si and Fe) into the gas phase (e.g. Flower et al.\ 1996). In slower C-type shocks, non-thermal sputtering will inject refractory and volatile species mainly from the grain mantles into the gas phase (e.g. Flower and Pineau des For\^ets 1994). The entrance of this fresh material, together with the high abundance of OH, will produce oxides such as SO and SiO (see Bachiller 1996, van Dishoeck and Blake 1998, and references there in). As the shocked gas cools, the dominant reactions will be again those of the usual low temperature chemistry, and depletion onto dust grain surfaces will reduce the abundances of some of the newly formed molecules (e.g. H$_2$O, see Bergin et al.\ 1998). However, the chemical composition of both the gas and the solid phases will remain altered with respect to pre-shock ones. {\it Observations:} The chemical effects of shocks have been observed in a number of outflows from low-mass Class 0 objects, which are particularly energetic and contain shocked regions well separated spatially from the quiescent protostellar envelope. Recent examples include IRAS16293 (Blake et al.\ 1994; van Dishoeck et al.\ 1995), NGC1333 IRAS4 (Blake et al.\ 1995), and NGC1333 IRAS2 (Langer et al.\ 1996; Blake 1997; Bachiller et al.\ 1998). A comprehensive study of many different species has been recently carried out on L1157 (Bachiller and P\'erez Guti\'errez 1997). The abundances of many molecules (e.g. CH$_3$OH, H$_2$CO, HCO$^+$, NH$_3$, HCN, HNC, CN, CS, SO, SO$_2$) are observed to be enhanced by factors ranging from a few to a few hundred. The extreme case is SiO which is enhanced by a factor of $\sim 10^6$. There are significant differences in spatial distribution among the different species: some molecules such as HCO$^+$ and CN peak close to the central source, while SO and SO$_2$ have a maximum in the more distant shocks, with OCS having the most distant peak. Other molecules such as SiO, CS, CH$_3$OH, and H$_2$CO show an intermediate behavior. Such differences cannot be attributed solely to excitation conditions: an important gradient in chemical composition is observed along the outflow. It is very likely that this strong gradient is related to the time dependence of shock-chemistry. As an example, consider the chemistry of SO, SO$_2$, and OCS, which has been recently modeled by Charnley (1997). It is believed that sulfur is released from grains in the form of H$_2$S, and that it is then oxidized to SO and SO$_2$ in a few 10$^3$ yr. The formation of OCS needs a few 10$^4$ yr. This is in general agreement with observations, since the SO/H$_2$S and SO$_2$/H$_2$S ratios do increase with distance from the source (i.e. with time), and OCS emission is only observed in the most distant position (i.e the oldest shock). Hence, chemical studies are of high potential to constrain the age and time evolution of molecular outflows. \subsection{B.~~Shock Cooling and Energetics} Millimeter observations of sh\-ock-en\-hanced mo\-le\-cules trace chemically processed gas that has already cooled down to 60-100 K, as indicated e.g. by multi-line NH$_3$ studies (Bachiller et al.\ 1993, Tafalla and Bachiller 1995). Emission from hotter post-shock gas, on the other hand, is important to obtain information on the instantaneous energy input rate and pre-shock conditions in outflows. {\it Hot (T $\ge$ 1000 K) shocked gas:} Hollenbach (1985) suggested that the [O~I]\,63$\mu$m\ line should offer a useful, extinction-insensitive measure of {\it dissociative} J-shocks in molecular flows. Because [O~I]\,63$\mu$m\ is the main coolant below $\sim$ 5000 K, its intensity is roughly proportional to the mass flux {\it into the J-shock}, $\dot{M}_{JS}$, through the relation $L_{[O I]}/L_\odot = 10^{4} \times \dot{M}_{JS} / M_\odot {\rm yr}^{-1}$, as long as the line remains optically thin (i.e. $n_o V_{JS} < 10^7$ \hbox{km~s$^{-1}$}\ cm$^{-3}$, where $n_o$ is the pre-shock density and $V_{JS}$ is the J-shock velocity). First detections of [O~I]\,63$\mu$m\ in outflows were obtained with the Kuiper Airborne Observatory (KAO) toward 3 HH objects and 5 highly collimated Class 0 outflows (Cohen et al. 1988; Ceccarelli et al. 1997). Since the advent of the Infrared Space Observatory (ISO), the Long Wavelength Spectrometer (LWS) has revealed [O~I]\,63$\mu$m\ emission in at least 10 more HH objects and molecular outflows (Liseau et al. 1997; Saraceno et al. 1998). These authors find a surprisingly good correlation between {\it current} values of $\dot{M}_{JS}$ derived from [O~I]\,63$\mu$m\ and {\it time-averaged} $\dot{M}_{ave}$ values derived from mm observations of the outflow assuming ram pressure equilibrium at the shock (i.e. \hbox{$F_{\rm CO}$} = $\dot{M}_{ave}\times V_{JS}$). Both values agree for $V_{JS} \sim$ 100 \hbox{km~s$^{-1}$}. The dispersion in this correlation, roughly a factor of 3, is of the same order as the uncertainties in \hbox{$F_{\rm CO}$}\ caused by opacity and projection effects (Cabrit and Bertout 1992). Hence, CO-derived momentum rates in outflows do not appear to suffer from large systematic errors. With the development of large format near-IR arrays in the early 1990's, it has also become possible to map molecular outflows in the 2.12$\mu$m\ $v = 1-0$ S(1) line of \hbox{${\rm H}_2$}, a tracer of hot ($\sim$ 2000 K) shocked molecular gas. In both low-luminosity and high-luminosity outflows, the \hbox{${\rm H}_2$}\ emission delineates single or multiple bow-shaped features associated with the leading edge of the CO emission (Davis and Eisl\"offel 1995; Davis et al.\ 1998), as illustrated in Fig.\ 1 for HH211; in a few cases, \hbox{${\rm H}_2$}\ emission also traces collimated jets and cavity walls (e.g Bally et al.\ 1993; Eisl\"offel et al.\ 1994). Observed surface brightnesses and rotational temperatures $\sim$ 1500-2500 K indicate moderate velocity shocks: either J-shocks of speed 10-25 \hbox{km~s$^{-1}$}\ or C-shocks with $V_s \sim 30$ \hbox{km~s$^{-1}$}\ and low filling-factor (Smith 1994; Gredel 1994). In particular, the morphology, line profile shapes, intensity, and proper motions of \hbox{${\rm H}_2$}\ 2.12$\mu$m bows are well reproduced by hydrodynamical simulations of jets propagating into the surrounding cloud, where \hbox{${\rm H}_2$}\ 2.12$\mu$m emission arises mostly in the non-dissociative wings of the bowshock (Raga et al.\ 1995; Micono et al.\ 1998; Suttner et al.\ 1998). If these bowshocks are also where most of the slow molecular outflow is being accelerated, and if \hbox{${\rm H}_2$}\ emission dominates the cooling (as expected e.g. in 2000K molecular gas at densities of $10^5-10^8$ cm$^{-3}$), then $L$(\hbox{${\rm H}_2$})$/\hbox{$L_{\rm CO}$}$ should be of order unity (see e.g. Hollenbach 1997). In the 5 flows studied by Davis and Eisl\"offel (1995), the observed ratio $L(\hbox{${\rm H}_2$})/\hbox{$L_{\rm CO}$}$ has a median value of $\sim$0.4, but it covers a very broad range from 0.001 to 30. The discrepancies could be caused by uncertainties in 2$\mu$m\ extinction (corrections typically amount to 10-100 for $A_V = 20-50$ mag), by the use of unreliable \hbox{$L_{\rm CO}$}\ estimates, or --- in the case of very low ratios --- by a strong decrease in outflow power over time (see W75N; Davis et al. 1998). Hence the \hbox{${\rm H}_2$}\ 2.12$\mu$m\ line alone is not a sufficient diagnostic of the outflow entrainment process. {\it Warm (T $\sim$ 300-1000 K) molecular gas}: The ISO mission has led to the detection of a new component of warm post-shock gas at T $\sim$ 300-1000 K in several outflows. Fig.\ 6 shows a map of the L1157 outflow in the $v = 0-0$ S(5) pure rotational line of \hbox{${\rm H}_2$}\ at 6.9$\mu$m\ obtained with ISOCAM (Cabrit et al.\ 1998). A series of bright emission spots are seen along the outflow axis. They coincide spatially with hot shocked gas emitting in the \hbox{${\rm H}_2$}\ 2.12$\mu$m\ line (Eisl\"offel and Davis 1995), and with the various peaks of shock-en\-han\-ced molecules identified by Bachiller and P\'erez Guti\'errez (1997; see Sect. IV.A). However, they trace an intermediate temperature regime of $\sim$ 800 K, considerably lower than the 2000 K observed in ro-vibrational \hbox{${\rm H}_2$}\ lines. Warm \hbox{${\rm H}_2$}\ at 700-800K was also found with the ISO Short Wavelength Spectrometer (SWS) in two outflows from very luminous sources, Cep~A and DR~21 (Wright et al.\ 1996; Smith et al.\ 1998). Finally, warm CO at $T \sim 330-1600 K$ was detected in high-J lines ($J_{up}$ = 14 to 28) with ISO-LWS toward 5 outflows of various luminosities, while H$_2$O and OH lines were detected in two cases (Nisini et al.\ 1996, 1997; Ceccarelli et al.\ 1998). The observed emission fluxes and temperatures in \hbox{${\rm H}_2$}\ and CO are well explained by non-dissociative J-shocks with $V_s \sim$ 10 \hbox{km~s$^{-1}$}\ or slow C-shocks with $V_s \sim$ 10-25 \hbox{km~s$^{-1}$}, and $n_o \sim 10^4-3\times 10^5$ cm$^{-3}$\ (e.g. Wright et al. 1996; Nisini et al. 1997; Cabrit et al. 1998). One important constraint is the rather low [H$_2$O]/[\hbox{${\rm H}_2$}] abundance ratio $\sim 1-2\times 10^{-5}$ observed in HH54 and IRAS16293 (Liseau et al.\ 1996; Ceccarelli et al. 1998); steady-state C-shocks would predict complete conversion of O into water. The relatively high [OH]/[H$_2$O] ratio $\sim$ 1/4-1/10 in these two flows suggests that the shock age is too short for conversion to be complete, and points to the need for time-dependent C-shock models for proper interpretation of the data (e.g. Chi\`eze et al.\ 1998). Mid and far-infrared emission from this warm molecular gas component appears more tightly correlated with \hbox{$L_{\rm CO}$}\ than the 2$\mu$m\ \hbox{${\rm H}_2$}\ lines. In 4 out of the 5 outflows studied by Nisini et al.\ (1997), the FIR CO luminosity represents 10-30\% of the flow kinetic luminosity, in good agreement with C-shock calculations in the inferred density and velocity range (Kaufman and Neufeld 1996). The only large discrepancy is observed in IC1396N, an object contaminated by PDR emission (Molinari et al.\ 1997). In L1157, the \hbox{${\rm H}_2$}\ luminosity of warm gas is also around 10\% of \hbox{$L_{\rm CO}$}\ (Cabrit et al., 1999). Hence, these slow shocks seem sufficient to drive the whole outflows. Detailed comparisons between \hbox{${\rm H}_2$}\ and FIR-CO lines in the same objects are now under way to further narrow the range of possible shock models and perhaps allow us to discriminate between wide-angle wind and jet scenarios for the entrainment of outflows. $$\psfig{file=fig6.ps,height=140mm}$$ \caption{Figure~6.\hskip\capskipamount The L1157 outflow mapped in the 6.9$\mu$m\ pure rotational line of \hbox{${\rm H}_2$}\ (adapted from Cabrit et al.\ 1998) with CO(2-1) contours superimposed (from Bachiller and P\'erez Guti\'errez 1997).} \mainsection{{V}~~CONCLUSIONS} We have shown that the current data on the structure and energetics of molecular outflows suggest broad similarities across the entire luminosity range, from 1 to $10^5\,\hbox{L$_{\odot}$}$. If the flows are swept up by a stellar wind or jet, we find that $\hbox{$\dot{M}_{w}$}\hbox{$v_{w}$}/\hbox{$\dot{M}_{a}$}\hbox{$v_{K}$}$ has a value of about 0.3 for all flows, perhaps suggesting that flows have a common drive mechanism. However, it remains unclear if the MHD disk and X-wind models which have been used to explain low-mass outflows are appropriate in the very different physical regime of high-mass YSOs. While observational data continue to improve these constraints, there remains an urgent need for a larger sample of molecular outflows, particularly from high-mass stars, to be fully mapped at high resolution; at the moment it is very possible that our estimates of the properties of high-mass systems are biased by strong selection effects. Single dish data, especially from the new focal plane arrays, plus interferometric images at millimeter wavelengths will continue to accumulate. However, only when the large millimeter interferometer (MMA/LSA) is operational will it be possible to acquire high-resolution data quickly enough to study large samples of outflows in detail. The nature of the shocks which drive outflows is slowly becoming clearer. We now have diagnostics of all the temperature components in the outflows, from the 2000K gas seen in the 2\,\hbox{$\mu{\rm m}$}\ \hbox{${\rm H}_2$}\ lines, to the several hundred Kelvin component recently detected by ISO, through to the cool massive component seen in the millimeter waveband where most of the momentum is eventually deposited. The relationship between these components is providing valuable tests of the outflow mechanism, although a fuller understanding will require observations at higher angular resolution than ISO provided. The SOFIA and especially the FIRST missions will provide valuable data in this area. \vskip 0.25in {\bf Acknowledgments:~~} DSS would like to thank Andy Gibb and Chris Davis for providing unpublished outflow data for a number of sources and A. Gibb, A. Sargent, and L. Testi for useful discussions. JSR acknowledges support from the Royal Society. \vfill\eject\null \vskip .5in \centerline{\bf REFERENCES} \vskip .25in \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Acord, J. M., Walmsley, C. M., and Churchwell, E. 1997. The Extraordinary Outflow toward G5.89-0.39. In {\refit Astrophy. J.\/}, 475:693--704.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Bachiller}, R., {Codella}, C., {Colomer}, F., {Liechti}, S. and {Walmsley}, C. M. 1998. Methanol in protostellar outflows. Single-dish and interfero\-metric maps of NGC 1333-IRAS~2. {\refit Astrophys.\ J.\/} 335:266--276.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Bachiller}, R., {Tafalla}, M. and {Cernicharo}, J. 1994. Successive ejection events in the L1551 molecular outflow. {\refit Astrophys.\ J.\ Lett.\/} 425:93--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Bachiller, R., Cernicharo, J., Mart\'{\i}n-Pintado, J., Tafalla, M., and Lazareff, B. 1990. High-velocity molecular bullets in a fast bipolar outflow near L1448/IRS3. {\refit Astron.\ Astrophys.\/} 231:174--86.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Bachiller, R., Martin-Pintado, J. and Fuente, A. 1993. High-Velocity Hot Ammonia in Bipolar Outflows. {\refit Astrophys.\ J.\ Lett.\/} 417:45--48.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Bachiller}, R. and {P\'erez Guti\'errez}, M. 1997. Shock Chemistry in the Young Bipolar Outflow L1157. {\refit Astrophys.\ J.\ Lett.\/} 487:93-97.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Bachiller, R., Guilloteau, S., Dutrey, A., Planesas, P., and Mart\'{\i}n-Pintado, J. 1995. The jet-driven molecular outflow in L 1448. CO and continuum synthesis images. {\refit Astron.\ Astrophys.\/} 299:857--.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Bachiller, R. 1996. Bipolar Molecular Outflows from Young Stars and Protostars. {\refit Ann.\ Rev.\ Astron.\ Astrophys.\/} 34:111--154.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Bally, J., {Devine}, D., {Hereld}, M., and {Rauscher}, B. J. 1993. Molecular Hydrogen in the IRAS 03282+3035 Stellar Jet. {\refit Astrophys.\ J.\ Lett.\/} 418:75--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Bence}, S. J., {Padman}, R., {Isaak}, K. G. {Wiedner}, M. C., and {Wright}, G. S. 1998. L~43: the late stages of a molecular outflow. {\refit Mon.\ Not.\ Roy.\ Astron.\ Soc.\/} 299:965-- } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Bence}, S. J., {Richer}, J. S. and {Padman}, R. 1996. RNO~43: a jet-driven super-outflow. {\refit Mon.\ Not.\ Roy.\ Astron.\ Soc.\/} 279: 866--883. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Bergin}, E. A., {Neufeld}, D. A. and {Melnick}, G. J. 1998. The Postshock Chemical Lifetimes of Outflow Tracers and a Possible New Mechanism to Produce Water Ice Mantles. {\refit Astrophys.\ J.\/} 499:777--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Blake}, G. A., {van Dishoeck}, E. F., {Jansen}, D. J., {Groesbeck}, T. D. and {Mundy}, L. G. 1994. Molecular abundances and low-mass star formation. 1: Si- and S-bearing species toward IRAS 16293-2422. {\refit Astrophys.\ J.\/} 428:680--692.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Blake}, G. A., {Sandell}, G., {Van Dishoeck}, E. F., {Groesbeck}, T. D., {Mundy}, L. G. and {Aspin}, C. 1995. A molecular line study of NGC~1333/IRAS~4. {\refit Astrophys.\ J.\/} 441:689--701.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Blake G.A. 1997. High angular resolution observations of the gas phase composition of young stellar objects. In {\refit IAU Symposium No. 178\/}, ed. E. van Dishoeck, pp. 31--.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Bontemps, S., Andr\'e, P., Terebey, S. and Cabrit, S. 1996. Evolution of outflow activity around low-mass young stellar objects. {\refit Astron.\ Astrophys.\/} 311:858--872.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Cabrit, S. and Bertout, C. 1992. CO line formation in bipolar flows. 3.~The energetics of molecular flows and ionized winds. {\refit Astron.\ Astrophys.\/} 261:274--284.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Cabrit S., Raga, A. and Gueth, F. 1997. Models of bipolar molecular outflows. In {\refit IAU Symposium No. 182\/}, eds.\ B. Reipurth and C. Bertout (Kluwer Academic Publishers), pp.\ 163--180.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Cabrit}, S., {Goldsmith}, P. F. and {Snell}, R. L. 1988. Identification of RNO 43 and B335 as two highly collimated bipolar flows oriented nearly in the plane of the sky. {\refit Astrophys.\ J.\/} 334:196--208. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Cabrit}, S., {Couturier}, Andre, P., {Boulade}, O., {Cesarsky}, C. J. and {Lagage}, P. O., {Sauvage}, M., {Bontemps}, S., {Nordh}, L. and {Olofsson}, G., {Boulanger}, F., {Sibille}, F., and {Siebenmorgen}, R. 1998. Mid-Infrared Emission Maps of Bipolar Outflows with ISOCAM: an in-depth study of the L1157 outflow. In {\refit ASP Conf. Ser. 132: Star Formation with the Infrared Space Observatory}, eds. Yun, J. and Liseau, R. (ASP), pp.\ 326--329. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Ceccarelli}, C., {Haas}, M. R., {Hollenbach}, D. J. and {Rudolph}, A. L. 1997. OI 63 Micron-determined Mass-Loss Rates in Young Stellar Objects. {\refit Astrophys.\ J.\/} 476:771--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Cernicharo, J. and Reipurth, B. 1996. Herbig-Haro Jets, CO Flows, and CO Bullets: The Case of HH 111. {\refit Astrophys.\ J.\ Lett.\/} 460:57--61. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Cesaroni, R., Felli, M., Testi, L., Walmsley, C. M., and Olmi, L. 1997. The disk-outflow system around the high-mass (proto)star IRAS 20126$+$4104. {\refit Astron.\ Astrophys.\/} 325:725--744. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Chandler, C. J., Terebey, S., Barsony, M., Moore, T. J. T. and Gautier, T. N. 1996. Compact outflows associated with TMC-1 and TMC-1A. {\refit Astrophys. J.\/} 471:308--320.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Chandler, C. J., and Depree, C. G. 1995. Vibrational Ground-State SiO J=1-0 Emission in Orion IRc2 Imaged with the VLA. {\refit Astrophys.\ J.\ Lett.\/} 455:L67--L71. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Chang}, C. A. and {Martin}, P. G. 1991. Partially dissociative jump shocks in molecular hydrogen. {\refit Astrophys.\ J.\/} 378:202--213.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Charnley}, S. B. 1997. Sulfuretted Molecules in Hot Cores. {\refit Astrophys. J.\/} 481:396--.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Chernin, L. M. and Wright, M. C. H. 1996. High-Resolution CO observations of the molecular outflow in the Orion IRc2 region. {\refit Astrophys. J.\/} 467:676--683.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Chernin, L., {Masson}, C., {Gouveia Dal Pino}, E. M. and {Benz}, W. 1994. Momentum transfer by astrophysical jets. {\refit Astrophys. J.\/} 426:204--214. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Chernin}, L. M. and {Masson}, C. R. 1991. A nearly unipolar CO outflow from the HH 46 - 47 system. {\refit Astrophys. J.\ Lett.\/} 382:93-96. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Chieze}, J-P., {Pineau Des Forets}, G. and {Flower}, D. R. 1998. Temporal evolution of MHD shocks in the interstellar medium. {\refit Mon.\ Not.\ Roy.\ Astron.\ Soc.\/} 295:672--682.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Churchwell, E. 1997a. Origin of the Mass in Massive Star Outflows. {\refit Astrophys. J.\ Lett.\/} 479:59--61.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Churchwell, E. 1997b. Massive star formation: observational constraints and the origin of the mass in massive outflows. In {\refit IAU Symposium No. 182\/}, eds.\ B. Reipurth and C. Bertout (Kluwer Academic Publishers), pp.\ 525--536.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Cohen}, M., {Hollenbach}, D. J., {Haas}, M. R. and {Erickson}, E. F.", 1988. Observations of the 63 micron forbidden O I line in Herbig-Haro objects. {\refit Astrophys.\ J.\/} 329:863--873 } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Davis, C. J., Moriarty-Schieven, G., Eisl{\"o}ffel, J., Hoare, M. G. and Ray T. P. 1998. Observations of shocked H$_2$ and entrained CO in outflows from luminous young stars. {\refit Astron. J.\/} 115:1118-1134.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Davis, C. J., Mundt, R., Ray, T. P., and Eisl{\"o}ffel, J. 1995. Near-Infrared and optical imaging of the L1551-IRS5 region -- The importance of poorly collimated outflows from young stars. {\refit Astron. J.\/}110:766--775.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Davis}, C.J. and {Eisl\"offel}, J. 1995. Near-infrared imaging in H\_2\_ of molecular (CO) outflows from young stars. {\refit Astron.\ Astrophys.\/} 300:851--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Davis, C. J., and Smith, M. D. 1995. Near-IR Imaging and Spectroscopy of DR21: A Case for Supersonic Turbulence. {\refit Astron. and Astrophys. \/} 310:961--969. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Devine, D., Bally, J., Reipurth, B., Shepherd, D. S., and Watson, A. M. 1999. A Giant Herbig-Haro Flow from a Massive Young Star in G192.16-3.82. {\refit Astron. J.\/} in press.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Devine}, D., {Bally}, J., {Reipurth}, B. and {Heathcote}, S. 1997. Kinematics and Evolution of the Giant HH34 Complex. {\refit Astron. J.\/} 114:2095--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Doyon, R. and Nadeau, D. 1988. The Molecular Hydrogen Emission from the Cepheus A Star-Formation Region. {\refit Astrophys.\ J.\/} 334:883--890. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Draine, B. T. and {McKee}, C. F. 1993. Theory of interstellar shocks. {\refit Ann.\ Rev.\ Astron.\ Astrophys.\/} 31:373--432. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Dutrey}, A., {Guilloteau}, S. and {Bachiller}, R. 1997. Successive SiO shocks along the L1448 jet axis. {\refit Astron.\ Astrophys.\/} 325:758--768.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Dyson, J. E. 1984. The interpretation of flows in molecular clouds. {\refit Astrophys.\ Space Sci.\/} 106: 181--197. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Eisl\"offel}, J., {Davis}, C. J., {Ray}, T. P. and {Mundt}, R. 1994. Near-infrared observations of the HH 46/47 system. {\refit Astrophys.\ J.\ Lett.\/} 422:91-93. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Ferreira, J. 1997. Magnetically-driven jets from Keplerian accretion discs. {\refit Astron.\ Astrophys.\/} 319:340--359} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Fiege, J. D. and Henriksen, R. N. 1996a. A global model of protostellar bipolar outflow - I. {\refit Mon.\ Not.\ Roy.\ Astron.\ Soc.\/} 281:1038} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Fiege, J. D. and Henriksen, R. N. 1996b. A global model of protostellar bipolar outflow - II. {\refit Mon.\ Not.\ Roy.\ Astron.\ Soc.\/} 281:1055} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Flower, D.R., Pineau des For\^ets, G., 1994. Grain-mantle erosion in magnetohydrodynamic shocks. {\refit Mon.\ Not.\ Roy.\ Astron.\ Soc.\/} 268:724} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Flower, D.R., Pineau des For\^ets, G., Field, D., May, P.W., 1996. The structure of MHD shocks in molecular outflows: grain sputtering and SiO formation. {\refit Mon.\ Not.\ Roy.\ Astron.\ Soc.\/} 280:447} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Fridlund}, C. V. M. and {Liseau}, R. 1998. Two Jets from the Protostellar System L1551~IRS5. {\refit Astrophys.\ J.\ Lett.\/} 599:75--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Fukui, Y., Iwata, T., Mizuno, A., Bally, J., and Lane, A. P. 1993. Molecular outflows. In {\refit Protostars \& Planets {I}{I}{I}\/}, eds.\ E. H. Levy and J. I. Lunine (Tucson: Univ.\ of Arizona Press), pp.\ 603--639. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Giannakopoulou, J., Mitchell, G. F., Hasegawa, T. I., Matthews, H. E., and Maillard, J-P. 1997. The Star-Forming Core of Monoceros R2. {\refit Astrophys. J.\/} 487:346--364. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Garay, G., Ramirez, S. Rogr{\'\i}guez, L. F., Curiel, S., Torrelles, J. M. 1996. The Nature of the Radio Sources within the Cepheus A Star-Forming Region. {\refit Astrophys. J.\/} 459:193--208. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Garden, R. P., Hayashi, M., Gatley, I., Hasegawa, T. and Kaifu, N. 1991. A spectroscopic Study of the DR~21 Outflow Source. III the CO Line Emission. {\refit Astrophys.\ J.\/} 374:540--554. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Garden, R. P., and Carlstrom, J. E. 1992. High-Velocity HCO$^+$ Emission Associated with the DR~21 Molecular Outflow. {\refit Astrophys.\ J.\/} 392:602--615. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Gayley, K.G., Owocki, S.P., Cranmer, S.R. 1995. Momentum deposition in Wolf-Rayet winds: nonisotropic diffusion with effective gray opacity {\refit Astrophys.\ J.\/} 442:296--310.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Glassgold, A. E., {Mamon}, G. A. and {Huggins}, P. J. 1989. Molecule formation in fast neutral winds from protostars. {\refit Astrophys.\ J.\ Lett.\/} 336:29--31. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Gredel, R. 1994. Near-infrared spectroscopy and imaging of Herbig-Haro objects. {\refit Astron.\ Astrophys.\/} 292:580--592.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Greenhill, L. J., Gwinn, C. R., Schwartz, C., Moran, J. M., and Diamond, P. J. 1998. Coexisting Conical Bipolar and Equatorial Outflows from a High-Mass Protostar. {\refit Nature\/} 396:650--653. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Gueth}, F., {Guilloteau}, S. and {Bachiller}, R. 1996. A precessing jet in the L1157 molecular outflow. {\refit Astron.\ Astrophys.\/} 307:891--897.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Gueth}, F., and Guilloteau, S., 1998. {\refit Astron.\ Astrophys.\/} in press.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Guilloteau, S., Bachiller, R., Fuente, A. and Lucas, R. 1992. First observations of young bipolar outflows with the IRAM interferometer - 2 arcsec resolution SiO images of the molecular jet in L 1448. {\refit Astron.\ Astrophys.\ Lett.\/} 265:49--52.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Hartigan, P., Morse, J. and Raymond, J. 1994. Mass Loss Rates, Ionization Fractions, Shock Velocities and Magnetic Fields of Stellar Jets. {\refit Astrophys.\ J.\/} 436: 125--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Hartigan, P., Carpenter, J. M., Dougados, C., and Skrutskie, M. F. 1996. Jet Bow Shocks and Clumpy Shells of H$_2$ Emission in the Young Stellar Outflow Cepheus A. {\refit Astron.\ J.\/} 111:1278--1285. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Harvey, P. M., Lester, D. F., Colom{\'e}, C., Smith, B., Monin, J-L., and Vaughlin, I. 1994. G5.89-0.39: A Compact HII Region with a Very Dense Circumstellar Dust Torus. In {\refit Astrophy. J.\/}, 433:187--198. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Henriksen, R.N. 1994. Theory of bipolar outflows. In {\refit The Cold Universe\/}, eds.\ Th. Montmerle, Ch.J. Lada, I.F. Mirabel, J. Tran Thanh Van (Editions Fronti\`eres: Gif-sur-Yvette), pp.\ 241--254.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Hollenbach, D. 1997. In {\refit IAU Symposium No. 182\/}, eds.\ B. Reipurth and C. Bertout (Kluwer Academic Publishers), pp.\ 181--198.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Hollenbach, D. 1985. Mass loss rates from protostars and OI(63 micron) shock luminosities. {\refit Icarus}, 61:36--39.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Hughes, S. M. G. 1993. Is Cepheus A East a Herbig-Haro Object? {\refit Astron. J.\/} 105:331--338. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Hunter, T. R., Phillips, T. G. and Menten, K. M. 1997. Active Star Formation toward the Ultracompact HII Regions G45.12+0.13 and G45.07+0.13. {\refit Astron. J.\/} 478:283--294. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Kameya, O., Hasegawa, T. I., Hirano, N., Takakubo, K. 1989. {\refit Astrophys. J.\/} 339:222-230.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Kaufman}, M. J. and {Neufeld}, D.A. 1996. Far-Infrared Water Emission from Magnetohydrodynamic Shock Waves. {\refit Astrophys. J.\/} 456:611--.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Kelly, M. L. and Macdonald, G. H. 1996. Two new young stellar objects with bipolar outflows in L379. {\refit Mon.\ Not.\ Roy.\ Astron.\ Soc.\/} 282, 401-412.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Lada, C. J. and Fich, M. 1996. The structure and energetics of a highly collimated bipolar outflow: NGC 2264G. {\refit Astrophys. J.\/} 459:638--652.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Langer}, W. D., {Castets}, A. and {Lefloch}, B. 1996. The IRAS 2 and IRAS 4 Outflows and Star Formation in NGC 1333. {\refit Astrophys.\ J.\ Lett.\/} 471:111--115.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Liseau}, R., {Giannini}, T., {Nisini}, B., {Saraceno}, P. , {Spinoglio}, L., {Larsson}, B., {Lorenzetti}, D. and {Tommasi}, E. 1997. Far-IR Spectrophotometry of HH Flows with the ISO Long-Wavelength Spectrometer. In {\refit IAU Symposium No. 182\/}, eds.\ B. Reipurth and C. Bertout (Kluwer Academic Publishers), pp.\ 111--120. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Mart{\'i}, J., Rodr{\'\i}guez, L. F., and Reipurth, B. 1993. HH~80-81: A highly collimated Herbig-Haro complex powered by a massive young star. {\refit Astrophys. J.\/} 416:208--217.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Mart{\'i}, J., Rodr{\'\i}guez, L. F., and Reipurth, B. 1995. Large proper motions and ejection of new condensations in the HH 80-81 thermal radio jet. {\refit Astrophys. J.\/} 449:184-187.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Martin-Pintado, J., Bachiller, R., Fuente, A. 1992. SiO Emission as a Tracer of Shocked Gas in Molecular Outflows. {\refit Astron.\ Astrophys.\/} 254:315--326.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Masson, C. R. and Chernin, L. M. 1992. Properties of swept-up molecular outflows. {\refit Astrophys. J.\/} 387:L47--L50.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Masson, C. R. and Chernin, L. M. 1993. Properties of jet-driven molecular outflows. {\refit Astrophys. J.\/} 414:230--241.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Masson, C. R. and Chernin, L. M. 1994. Observational constraints on outflow models. In {\it Clouds, Cores, and Low-Mass Stars}, eds. D. Clemens and R. Barvainis, A.S.P.Conference Series 65:350} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{McCaughrean, M.J., Rayner, J.T., Zinnecker, H., 1994. Discovery of a molecular hydrogen jet near IC 348. {\refit Ap.\ J.\/} 436:L189--L192} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{McCaughrean, M. and Mac Low, M-M. 1997. The OMC-1 molecular hydrogen outflow as a fragmented stellar wind bubble. {\refit Astron.\ J.\/} 113:391--400.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Mellema}, G. and {Frank}, A. 1997. Outflow collimation in young stellar objects. {\refit Mon.\ Not.\ Roy.\ Astron.\ Soc.\/} 292:795--807. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Menten, K. M., and Reid, M. J. 1995. What is Powering the Orion Kleinmann-Low Nebula? {\refit Astrophys.\ J.\ Lett.\/} 445:L157--L160. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Micono, M., Davis, C.J., Ray, T.P., Eisl\"offel, J., Shetrone, M.D. 1998. Proper motions and variability of the H$_2$ emission in the HH 46/47 system. {\refit Ap.\ J.\/} 494:L227--L230.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Mitchell, G. F., Maillard, J. P., Hasegawa, T. I. 1991. {\refit Astrophys. J.\/} 371:342--356.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Mitchell, G. F., Lee, S. W., Maillard, J. P., Matthews, H. E., Hasegawa, T. I., and Harris, A. I. 1995. A multitransitional CO study of GL 490. {\refit Astrophys. J.\/} 438:794--812. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Mitchell}, G. F., {Hasegawa}, T. I., {Dent}, W. R. F., {Matthews}, H. E. 1994. A molecular outflow driven by an optical jet. {\refit Astrophys.\ J.\ Lett.\/} 436:177--180. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Molinari}, S., {Testi}, L., {Brand}, J., {Cesaroni}, R.and {Pallo}, F. 1998. IRAS~23385+6053: A Prototype Massive Class 0 Object. {\refit Astrophys. J.\ Lett.\/} 505:39--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Molinari}, S., {Saraceno}, P., {Nisini}, B., {Giannini}, T., {Ceccarelli}, C., {White}, G. J., {Caux}, E. and {Palla}, F. 1998. Shocks and PDRs in an intermediate mass star forming globule: the case of IC1396N. In {\refit ASP Conf. Ser. 132: Star Formation with the Infrared Space Observatory}, eds. Yun, J. and Liseau, R. (ASP), pp.\ 390--.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Moriarty-Schieven, G. H. and Snell, R. L. 1988. High-resolution images of the L1551 molecular outflows. II. Structure and kinematics. {\refit Astrophys. J.\/} 332:364--378. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Mundt, R., Brugel, E. W. and {B\"uhrke}, T. 1987. Jets from Young Stars: {CCD} Imaging, Long-slit Spectroscopy And Interpretation Of Existing Data. {\refit Astrophys. J.\/} 319:275--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Nagar}, N. M., {Vogel}, S. N., {Stone}, J. M. and {Ostriker}, E. C. 1997. Kinematics of the Molecular Sheath of the HH 111 Optical Jet. {\refit Astrophys.\ J.\ Lett.\/} 482:195--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Narayanan, G., and Walker, C. K. 1996. Evidence for Multiple Outbursts from the Cepheus A Molecular Outflow. {\refit Astrophys.\ J.\/} 466:844--865. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Netzer}, N., Elitzur, M. 1993. The dynamics of stellar outflows dominated by interaction of dust and radiation. {\refit Astrophys.\ J.} 410:701--713.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Neufeld}, David A. and {Kaufman}, Michael J. 1993. Radiative Cooling of Warm Molecular Gas. {\refit Astrophys.\ J.\/} 418:263--.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Nisini}, B., {Giannini}, T., {Molinari}, S., {Saraceno}, P., {Caux}, E., {Ceccarelli}, C., {Liseau}, R., {Lorenzetti}, D., {Tommasi}, E. and {White}, G. J. 1998. High-J CO lines from YSOs driving molecular outflows. In {\refit ASP Conf. Ser. 132: Star Formation with the Infrared Space Observatory}, eds. Yun, J. and Liseau, R. (ASP), pp.\ 256--.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Nisini}, B., {Lorenzetti}, D., {Cohen}, M., {Ceccarelli}, C., {Giannini}, T., {Liseau}, R., {Molinari}, S., {Radicchi}, A., {Saraceno}, P., {Spinoglio}, L., {Tommasi}, E., {Clegg}, P.E., {Ade}, P.A.R., {Armand}, C., {Barlow}, M.J., {Burgdorf}, M., {Caux}, E., {Cerulli}, P., {Church}, S.E., {Di Giorgio}, A., {Fischer}, J., {Furniss}, I., {Glencross}, W.M., {Griffin}, M.J., {Gry}, C., {King}, K.J., {Lim}, T., {Naylor}, D.A., {Texier}, D., {Orfei}, R., {Nguyen-Q-Rieu}, {Sidher}, S., {Smith}, H.A., {Swinyard}, B.M., {Trams}, N., {Unger}, S.J. and {White}, G.J. 1996. LWS-spectroscopy of Herbig Haro objects and molecular outflows in the Cha II dark cloud. {\refit Astron.\ Astrophys.\ Lett.\/} 315:321-324.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Padman, R., Bence, S., and Richer, J. 1997. Observational properties of molecular outflows. In {\refit IAU Symposium No. 182\/}, eds.\ B. Reipurth and C. Bertout (Kluwer Academic Publishers), pp.\ 123--140. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Padman}, R. and {Richer}, J. S. 1994. Interactions between molecular outflows and optical jets. {\refit Astrophys.\ Space Sci.\/} 216:129--134.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Panagia}, N. 1991. Ionized winds from young stellar objects. In {\refit The Physics of Star Formation and Early Stellar Evolution\/}, eds. C.J. Lada and N. Kylafis (Kluwer Academic Publishers), pp.\ 565--593.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Poetzel}, R., {Mundt}, R. and {Ray}, T. P. 1989. Z CMa - A large-scale high velocity bipolar outflow traced by Herbig-Haro objects and a jet. {\refit Astron.\ Astrophys.\ Lett.\/} 224:13--16} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Raga, A. 1991. A New Analysis of the Momentum and Mass-loss Rates of Stellar Jets. {\refit Astron.\ J.\/} 101:1472-- } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Raga, A. C., Cant\'o, J., Calvet, N., Rodr\'{\i}guez, L. F. and Torrelles, J.M. 1993. A Unified Stellar Jet/Molecular Outflow Model. {\refit Astron.\ Astrophys.\/} 276:539--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Raga}, A.C., {Taylor}, S.D., {Cabrit}, S. and {Biro}, S. 1995. A simulation of an HH jet in a molecular environment. {\refit Astron.\ Astrophys.\/} 296:833--.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Reipurth}, B., {Bally}, J. and {Devine}, D. 1997. Giant Herbig-Haro Flows. {\refit Astron. J.\/} 114:2708--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Richer}, J.S., {Hills}, R.E. and {Padman}, R. 1992. A fast CO jet in Orion B. {\refit Mon.\ Not.\ Roy.\ Astron.\ Soc.\/} 254:525--538. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Richer}, J.S., {Hills}, R.E., {Padman}, R. and {Russell}, A.P.G. 1989. High-resolution molecular line observations of the core and outflow in Orion B. {\refit Mon.\ Not.\ Roy.\ Astron.\ Soc.\/} 241: 231-246. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Rodr{\'\i}guez, L. F., Carral, P., Moran, J. M., and Ho, P. T. P. 1982. Anisotropic mass outflow in regions of star formation. {\refit Astrophys. J.\/} 260:635--646.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Rodr{\'\i}guez, L. F., Garay, G., Curiel, S., Ram{\'\i}rez, S., Torrelles, J. M., G{\'o}mez, Y., and Vel{\'a}zquez, A. 1994. Cepheus A HW2: A powerful Thermal Radio Jet. {\refit Astrophys.\ J.\ Lett.\/} 430:L65--L68. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Rodr{\'\i}guez, L. F. 1995. Subarcsecond Observations of Radio Continuum from Jets and Disks. In {\refit Revista Mexicana de Astronom{\'\i}a y Astrof{\'\i}sica, Serie de Conferencias, Volumen 1\/}, eds.\ P. Pismis and S. Torres-Peimbert, pp.\ 1--10. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Russell, A. P. G., Bally, J., Padman, R., and Hills, R. E. 1992. Atomic and Molecular Outflow in DR~21. {\refit Astrophys. J.\/} 387:219--228.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Saraceno}, P., {Andr\'e}, P., {Ceccarelli}, C., {Griffin}, M., and {Molinari}, S. 1996. An evolutionary diagram for young stellar objects. {\refit Astron.\ Astrophys.\/} 309:827--839. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Saraceno}, P., {Nisini}, B., {Benedettini}, M., {Ceccarelli}, C., {Di Giorgio}, A. M., {Giannini}, T., {Molinari}, S., {Spinogl}, L., {Clegg}, P. E., {Correia}, J. C., {Griffin}, M. J., {Leeks}, S. J., {White}, G. J., {Caux}, E., {Lorenzetti}, D., {Tommasi}, E., {Liseau}, R. and {Smith}, H. A. 1998. LWS Observations of Pre Main Sequence Objects. In {\refit ASP Conf. Ser. 132: Star Formation with the Infrared Space Observatory}, eds. Yun, J. and Liseau, R. (ASP), pp.\ 233--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Schulz, A., Henkel, C., Beckmann, U., Kasemann, C., Schneider, G. Nyman, L. A., Persson, G., Gunnarsson, G. G., and Delgado, G. 1995. A High-Resolution CO J=4-3 Map of Orion-KL. {\refit Astron. and Astrophys.\/} 295:183--193. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Shepherd, D. S. and Churchwell, E. 1996. Bipolar molecular outflows in massive star-formation regions. {\refit Astrophys. J.\/} 472:225--239.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Shepherd, D. S., Watson, A. M., Sargent, A. I. and Churchwell, E. 1998. Outflows and luminous YSOs: A new perspective on the G192.16 massive bipolar outflow. {\refit Astrophys. J.\/} accepted.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Shepherd, D. S., Kurtz, S. E. 1998. {\refit In preparation.\/}} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Shu}, F. H. and {Shang}, H. 1997. Protostellar X-rays, Jets, and Bipolar Outflows. In {\refit IAU Symposium No. 182\/}, eds.\ B. Reipurth and C. Bertout (Kluwer Academic Publishers), pp.\ 225--239.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Shu}, F.H., {Ruden}, S.P., {Lada}, C.J., and {Lizano}, A. Star formation and the nature of bipolar outflows. 1991. {\refit Astrophys. J. L.\/} 370:31-34.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Shu}, F., {Najita}, J., {Ostriker}, E.,{Wilkin}, F., {Ruden}, S. and {Lizano}, S. 1994. Magnetocentrifugally driven flows from young stars and disks. 1: A generalized model. {\refit Astrophys. J.\ Lett.\/} 429:781-796. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Smith}, M.D. 1994. Jump shocks in molecular clouds - speed limits and excitation levels. {\refit Mon.\ Not.\ Roy.\ Astron.\ Soc.\/} 266:238--.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Smith, M. D., Suttner, G. and Yorke, H. W. 1997. Numerical hydrodynamic simulation of jet-driven bipolar outflows. {\refit Astron.\ Astrophys.\/} 323:223--230. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Smith}, M. D., {Eisl\"offel}, J. and {Davis}, C. J. 1998. ISO observations of molecular hydrogen in the DR21 bipolar outflow. {\refit Mon.\ Not.\ Roy.\ Astron.\ Soc.\/} 297:687--691.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Suttner}, G., {Smith}, M.D., {Yorke}, H.W. and {Zinnecker}, H. 1997. Multi-dimensional numerical simulations of molecular jets. {\refit Astron.\ Astrophys.\/} 318:595--607.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Stahler, S. W. 1994. The kinematics of molecular outflows. {\refit Astrophys. J.\/} 422:616--620.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Stone}, J. M., {Xu}, J. and {Mundy}, L.G. 1995. Formation of bullets by hydrodynamical instabilities in stellar outflow. {\refit Nature\/} 377:315--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Tafalla}, M. and {Bachiller}, R. 1995. Ammonia emission from bow shocks in the L1157 outflow. {\refit Astrophys. J.\ Lett.\/} 443:37-40.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Taylor}, S. D. and {Raga}, A. C. 1995. Molecular mixing layers in stellar outflows. {\refit Astron.\ Astrophys.\/} 296:823--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Torrelles, J. M., Verdes-Montenegro, L., Ho, P. T. P. Rodr{\'\i}guez, L. F., and Jorge, C. 1993. >From Bipolar to Quadrupolar -- The Collimation Processes of the Cepheus A Outflow. {\refit Astrophys.\ J.\/} 410:202--217. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{van Dishoeck, E. F., {Blake}, G. A., {Jansen}, D. J. and {Groesbeck}, T. D.", 1995. Molecular Abundances and Low-Mass Star Formation. II. Organic and Deuterated Species toward IRAS 16293-2422. {\refit Astrophys.\ J.\/} 447:760--.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{van Dishoeck E.F. and Blake G.A. 1998. {\refit Ann.\ Rev.\ Astron.\ Astrophys.\/} in press.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Velusamy, T. and Langer, W. D. 1998. Outflow-Infall interactions as a mechanism for terminating accretion in protostars. {\refit Nature\/} 392:685-687. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{V\"olker, R., Smith, M.D., Suttner, G., Yorke, H.W. 1999. Numerical hydrodynamical simulations of molecular outflows driven by hammer jets. {\refit Astron.\ Astrophys.\ }, in press} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Wilner, D. J., Welch, W. J., Forster, J. R. 1992. The G5.89-0.39 UC HII Region: Kinematics and Millimeter Aperture Synthesis Maps. {\refit Am.\ Astron.\ Soc.\ Meeting, 181\/}, number 22.06. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Wright, M. C. H., Plambeck, R. L., and Wilner, D. J. 1996. A Multiline Aperture Synthesis Study of Orion-KL. {\refit Astrophys.\ J.\/} 469:216--237. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Wolf, G. A., Lada, C. J., and Bally, J. 1990. The Giant Molecular Outflow in Mon R2. {\refit Astron. J.\/}, 100:1892--1902. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Wood, D. O. S. 1993. Bipolar Molecular Outflow from a Massive Star: High Resolution Observations of G5.89-0.39. In {\refit ASP conf. Series\/}, 35:108--110. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Wright}, C.M., {Drapatz}, S., {Timmermann}, R., {Van Der Werf}, P.P., {Katterloher}, R. and {De Graauw}, T. 1996. Molecular hydrogen observations of Cepheus A West. {\refit Astron.\ Astrophys.\ Lett.\/} 315:301--L304.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Wu}, Y., {Huang}, M. and {He}, J. 1996. A catalogue of high velocity molecular outflows. {\refit Astron.\ Astrophys.\ Suppl.\/} 115:283--. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Yamashita, T., Suzuki, Kaifu, N., Tamura, M., Mountain, C. M. and Moore, T. J. T. 1989. A new CO bipolar flow and dense disk system associated with the infrared reflection nebula GGD 27 IRS. {\refit Astrophys. J.\/} 373:560--566.} \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{{Zhang}, Q., {Hunter}, T. R. and {Sridharan}, T. K. 1998. A Rotating Disk around a High-Mass Young Star. {\refit Astrophys.\ J.\ Lett.\/} 505:151-154. } \noindent \parshape2 0truein 4.25truein 0.25truein 4truein{Zijlstra, A. A., Pottasch, S. R., Engels, D., Roelfsema, P. R., Hekkert, P. L., and Umana, G. 1990. Mapping the Outflow of OH5.89-0.39. {\refit Mon.\ Not.\ Roy.\ Astron.\ Soc.\/} 246:217--236. } \end
\section{Introduction} Directed percolation (DP) has emerged as one of the generic absorbing state type dynamic processes. It describes epidemic processes, e.g., forest fires and various types of surface catalysis processes \cite{KinzelBook,Schlogl,Grassberger79,Grassberger78,Ziff}. Such processes include a so-called absorbing state, typically the vacuum, from which it can not escape. The relevant tunable parameter is the propagation probability $p$. The system undergoes a phase transition from the absorbing phase at small $p$, where the stationary state is the absorbing state, into an active stationary phase at large $p$, where the system refuses to die. The scaling properties at DP dynamic phase transitions are known for almost two decades, and it's now realized that DP critical behaviour is the generic universality class for dynamic absorbing state type processes \cite{KinzelBook}. At DP type critical points the equilibration time $\xi_\parallel$ diverges. It scales as $\xi_\parallel\sim \xi_\perp^z$ compared to the spatial correlation length $\xi_\perp$, with dynamic exponent $z = 1.581$\cite{Jensen}. For example, starting from a single seed, the survival probability obeys the scaling form \begin{equation} P_s(\epsilon, t) = b^{-x_s} P_s( b^{1/\nu_\perp} \epsilon, b^{-z} t) \label{surv1} \end{equation} with $\epsilon=p_c-p$ the distance from the critical point. This leads to \begin{equation} P_s\sim \epsilon^\beta \exp(\frac{-t}{\xi_\parallel}), \label{surv2} \end{equation} with exponent $\beta=x_s \nu_\perp$. The exponential factor reflects that deep inside the absorbing phase $P_s$ decays exponentially in time. The equilibration time diverges at the DP critical point as $\xi_\parallel\sim \epsilon^{-\nu_\parallel}$ with $z=\nu_\parallel/\nu_\perp$. At $p_c$ the survival probability decays as a powerlaw, $P_s(t)\sim t^{-\delta}$ with $\delta = x_s/z =\beta/\nu_\parallel$. A recent direction of research in this topic concerns the scaling properties near boundaries \cite{Essam,Frojdh,Lauritsen98,Lauritsen}. Those studies address absorbing and reflective walls. The scaling properties are modified by surface type critical exponents. In particular, the survival probability for a seed near the boundary obeys the same scaling form as above, but with a new interface critical exponent $x$, and therefore a modified value for $\beta$. In this study we discuss the scaling properties near active boundaries. Consider a stationary active vertical wall in the system. All sites in the wall are alive. The critical exponent $\beta$ is not an issue, because the system remains active near the wall for all $p$. However, in the absorbing phase the cloud of active sites near the wall has a specific stationary state width, which is expected to diverge as $W \sim \xi_\perp\sim \epsilon^{-\nu_\perp}$. Widths like this diverge with bulk exponents. Assume that this wall is slanted, with an arbitrary angle $\theta \neq 90^\circ$ with respect to the horizontal direction (see Fig.\ref{curtain}). In the space-time interpretation of the configurations, the wall moves with a constant velocity. It acts as a slanted active curtain rod. A curtain of active sites hangs down from it as illustrated in Fig.\ref{curtain}. For $p<p_c$ the curtain has a finite width $l_\perp$ and length $l_\parallel = l_\perp \tan(\theta)$. In this study we address how the stationary state width of this slanted curtain scales near the DP critical point. Naively this seems a simple question. One would expect that the curtain width diverges with the same exponent as the equilibration time scale, $W\sim \epsilon^{-\nu_\parallel}$, i.e., with the same exponent as the length of a curtain hanging down from an horizontal curtain rod ($\theta=0$). The latter is equivalent to asking for the survival probability in the set-up without any walls where all sites are active in the initial state. This expectation is based on the anisotropic scaling properties. Consider a system with a rod at angle $\theta\neq 0$. The horizontal and vertical bulk lengths diverge with different exponents, as $\xi_\parallel\sim \xi_\perp^z$. Therefore, a system at $p_c-p=\epsilon$ and wall angle $\theta$ is equivalent by renormalization to a system with a smaller wall angle $\theta^\prime$ at $\epsilon^\prime= b^{-1/\nu_\perp}\epsilon$ with $\tan(\theta^\prime)\simeq b^{z-1} \tan(\theta)$. The scaling properties of $W$ should not depend on the angle $\theta$, since the rod renormalizes towards the horizontal position. We should expect the same scaling behaviour as at $\theta=0$. However, a recent numerical study \cite{Park} seems to contradict this. Kwon {\it et. al.}\cite{Park} studied a model with two absorbing states. It undergoes a dynamic phase transition which belongs to Directed Ising (DI) universality class when the two absorbing states are symmetric, and belongs to the Directed Percolation (DP) universality class when a symmetry breaking field is introduced. They studied the interface dynamics of the active domain between two asymmetric absorbing states. As one absorbing state dominates over the other, the interface is driven into the unpreferred absorbing region with a constant velocity. Therefore they expected the width of the active domain to scale like the horizontal width of the active curtain in the above setup for ordinary DP models. A simple power-law fit of their data suggests that the active domain width scales as $W\sim \epsilon^{-x}$ with $x\simeq 2.00(5)$, which does not agree with the DP exponent $\nu_\parallel\simeq 1.734$. In this paper we address the same issue more directly. We insert a slanted active wall into the most basic model for DP, the one studied originally by Kinzel\cite{Kinzel,Domany}, see section 2. We find a similar anomalous value for the width exponent. $W\sim \epsilon^{-x}$ scales as $x\simeq 1.95(5)$. In section 3 we develop a qualitative scaling theory. It predicts that the curtain width scales with the conventional exponent $\nu_\parallel$ but with an additional logarithmic factor as $W\simeq A \epsilon^{-\nu_\parallel} \ln(\frac{\epsilon_0}{\epsilon})$. In section 4 we show that the numerical Monte Carlo data fits this form well. In section 5 we illustrates how DP type processes with slanted walls can be studied in the Master equation formalism. Our finite size scaling (FSS) results, using exact numerical enumeration of the eigenvalue spectrum, show that at $p_c$ the width of the slanted curtain diverges as $W\sim L^z$ with system size. This confirms the absence of a new independent exponent. The logarithmic factor arises only in the $\epsilon$ dependence. \section{ Numerical Results for the Curtain Width.} Consider the square space-time lattice shown in Figure \ref{lattice}. All bonds run under 45 degrees. The black (open) circles represent the active (inactive) sites. Time evolves from top to bottom in half units $t \to t+\frac{1}{2}$. Bonds between nearest neighbor sites at $t$ and $t \to t+\frac{1}{2}$ are being created with probability $p$ but only if the upper site is active. Each bond activates the lower site. Kinzel studied this model in detail with Master equation type FSS in the early eighties \cite{Kinzel}. The critical exponents and the location of the DP transition are known quite accurately. For example, the latest series expansion results put the DP phase transition at $p_c\approx 0.6447$\cite{Jensen}. We modify the boundary conditions in this model to accommodate an active wall. The lattice is semi-infinite, bound to the left by the wall, which runs away under $\theta=45^\circ$ as shown in Fig.~\ref{lattice}. $45^\circ$ is its natural angle for the curtain rod for this specific lattice. We can restrict ourselves to this angle because the scaling properties of the curtain width should not depend on the angle according to the anisotropic scaling argument outlined above. Moreover, the angle is a continuous parameter in the model by Kwon {\em et. al.}\cite{Park} and their results show no angle dependence. We perform Monte Carlo simulations with as initial configuration an active wall in an inactive bulk. The horizontal curtain width is defined as the distance of the last active site from the rod in each time slice. For $p<p_c$, the width grows initially approximately linear in time, until it saturates at the stationary state value which varies with $\epsilon=p_c-p$. Figure \ref{width} shows the active width versus $\epsilon$ on a logarithmic scale. The line is quite linear over the two decades shown. The slope is clearly distinct from the expected value $\nu_\parallel\approx 1.734$ and close to the value found by Kwon {\em et. al.}\cite{Park}. In Fig.\ref{width-exp} we perform a more careful FSS analysis to the same data. We fit the numerical data from two nearby points, $\epsilon_2=\sqrt2 \epsilon_1$, to the form $W\simeq a\epsilon^{-x}$ and plot $x$ as a function of $\epsilon$, the exponent $x$ appears to be around $1.95$. This fit is remarkably stable, and shows virtually no power law type corrections to scaling. Taken out of context it is strongly suggestive of a new independent critical exponent. The other curves in Fig.\ref{width-exp} relate to the FSS analysis assuming an additional logarithmic factor as discussed in the next two sections. \section{Independent Cluster Approximation} Figure \ref{curtain} shows a typical curtain configuration in a Monte Carlo simulation at a $p$ just below the percolation threshold $p_c$. The most striking features are the needles in the curtain. Isolated clusters are expected to be needle like. The correlation length in the time direction diverges faster than in the spatial direction, as $\xi_\parallel\sim \xi_\perp^z$. Therefore, active clusters (when grown from a single seed) become needle shaped near the percolation threshold. Figure \ref{curtain} gives the impression that close to $p_c$, the curtain consists of a set of weakly interacting needle shaped clusters when viewed from length scales larger than $\xi_\perp$. In this section we pursue the implications of the assumption that such needles are completely uncorrelated. In that approximation the probability that the curtain extends over a horizontal distance $l$ is given by the probability that a needle longer than $\tau = l \tan (\theta)$ hangs down from the curtain rod vertically above that site. Let $P$ be that probability. It must have the same form as the survival probability from a single seed, Eq.(\ref{surv2}), The actual value of the exponent $\beta$ turns out to be irrelevant in this section, but it must be identical to the single seed value, according to a time reversal symmetry argument \cite{CCC}. The spatial coordinate needs to be coarse grained, because the needles can only be uncorrelated beyond the horizontal correlation length $\xi_\perp\sim \epsilon^{-\nu_\perp}$. Define $n= x/ \xi_\perp$ as the coarse grained discrete spatial coordinate and recall that $t= x \tan(\theta)$ is the corresponding vertical distance from the curtain rod to the same point. The probability for the curtain to have width $n$ factorizes in the independent needle approximation as \begin{equation} P_w(n)=P (n) \prod_{n^\prime>n} \bigg[ 1- P ( n^\prime ) \bigg]. \label{Pw} \end{equation} This equation can be rewritten into a derivative form \begin{equation} \frac{P_w(n+1)-P_w(n)}{P_w(n+1)} = \frac{P (n+1)-P(n)[1-P(n+1)]}{P (n+1)} \label{Pwd} \end{equation} The maximum of the distribution obeys the relation \begin{equation} P_w(\tilde n-1)=P_w(\tilde n) \end{equation} and can be written as \begin{equation} \frac{1}{P(\tilde n)}-\frac{1}{P(\tilde n-1)}=1. \label{Pmax} \end{equation} Assume that $P$ has the same asymptotic form as the single seed survival probability, in Eq.(\ref{surv2}), and that the maximum of the distribution occurs in this range of $n$. The transformation to the coarse-grained $n= x/ \xi_\perp\sim x \epsilon^{\nu_\perp}$ variable changes the critical exponent inside the exponential factor \begin{equation} P\simeq B \epsilon^\beta e^{- b n \epsilon^\Delta} \label{Psurv} \end{equation} with $\Delta = \nu_\parallel-\nu_\perp$, and $b \sim \tan(\theta)$. Inserting this form into Eq.(\ref{Pmax}) leads to \begin{equation} 1- e^{-b \epsilon^\Delta} = B \epsilon^\beta e^{-b \tilde n \epsilon^\Delta} \end{equation} and, after expanding the exponential on the left hand side, to \begin{equation} b \tilde n |\epsilon|^\Delta \simeq \ln(\frac{B}{b}) + (\beta-\Delta) \ln(\epsilon) \end{equation} In original units this reads \begin{equation} \tilde W \simeq A \epsilon^{-\nu_\parallel}\ln(\frac{\epsilon_0}{\epsilon}) \label{W-scaling} \end{equation} The characteristic probability depends on the wall angle as $\epsilon_0\sim 1/\tan(\theta)$. The most probable width $\tilde W$ scales with the expected exponent $\nu_\parallel$ but contains an additional logarithmic factor. Asymptotically the most probable and the average widths coincide. Eq.(\ref{Pwd}) can be approximated in the continuum limit as \begin{equation} \frac{1}{P_w} \frac{d P_w}{d n} = 1-\frac{P(n)}{P(n+1)} + P(n) \end{equation} Close to $p_c$ and for large $n$, where $P$ obeys Eq.(\ref{Psurv}), we can integrate this \begin{eqnarray} P_w (n) & \sim & \exp[(1-e^{b\epsilon^\Delta}) n -\frac{B}{b}\epsilon^{\beta-\Delta} e^{-bn \epsilon^\Delta}] \\ \nonumber & \sim & e^{b n \epsilon^\Delta}\exp[ -\frac{B}{b}\epsilon^{\beta-\Delta} e^{-bn \epsilon^\Delta}] \end{eqnarray} This distribution decays exponentially on both sides of the most probable value and becomes sharp at the critical point, $\epsilon\to 0$. We checked explicitly that the most probable and average coincide in this limit, and scale asymptotically with the same logarithmic factor, as in Eq.(\ref{W-scaling}). \section{Logarithmic Corrections to Scaling Analysis. } The logarithmic factor in the independent needle approximation formula for the curtain width \begin{equation} W(\epsilon) \simeq A \epsilon^{-\nu_\parallel} \ln(\frac{\epsilon_0}{\epsilon}) \label{W-log} \end{equation} does not change the asymptotic exponent. It is still equal to $\nu_{\parallel}$. However the finite size scaling (FSS) approach to this value is very singular. A conventional FSS analysis involves the construction of approximants for the critical exponent $x$ by fitting the values of $W$ at to nearby $\epsilon$ to a pure power law form, $W\sim \epsilon^{-x}$. This is equivalent to defining $x(\epsilon)$ as a derivative and yields for the above logarithmic form \begin{equation} x = -\frac{\epsilon}{W} \frac{d W}{d\epsilon} =\nu_\parallel + \frac{1}{\ln(\frac {\epsilon_0}{\epsilon})} \end{equation} This function approaches $\nu_\parallel$ in a singular manner. In the interval $0.01<\epsilon/\epsilon_0<0.3$, $x$ seems to converge convincingly with a linear correction to scaling term to an effective exponent which is about $0.2$ too large. One would have to go to extremely small $\epsilon$'s to see the true convergence. The power law fit in Fig.\ref{width-exp} shows signs of this. The two other curves in Fig.\ref{width-exp} show the FSS estimates for the exponent $\nu_\parallel$ according to the form Eq.(\ref{W-log}) with $\epsilon_0=1$ or $\epsilon_0=0.5$. $\epsilon_0$ is unknown, but likely of order one. Both curves converge towards the conventional value $\nu_\parallel=1.734$. This is strong evidence for the presence of the logarithmic factor. \section{finite size scaling at the percolation threshold} The logarithmic factor originates from the screening of independent needles. It should not play a role in the FSS at the percolation threshold itself, because there $\xi_\perp$ diverges, and the independent needle concept becomes meaningless. So the curtain width must scale as $W\sim L^z$ at $p_c$, if it is really true that no independent new exponent is involved. To confirm this we present in this section numerical data from master equation type finite size scaling using exact enumeration. We also performed Monte Carlo simulations but prefer to present our master equation data since this method requires a technical novelty. A moving wall is inconvenient in simulations. The lattice is finite by necessity and the moving wall requires a much bigger lattice than the one actually used by the process. This is a handicap in particular for master equation calculations where one evaluates the rate at which the stationary state is being reached by letting time go to infinity at each lattice size $L$. Those systems sizes are typically small, $L\leq 20$ in our case, because phase space scales exponentially with $L$. Compared to MC simulations the master equation method trades system size for numerical accuracy, and the ability to perform a detailed corrections to scaling analysis. The accuracy of the two methods is typically comparable, except for specific issues, like the logarithmic factor in the previous sections, which require intrinsic large lattice sizes. The solution to the moving wall problem is to distinguish between the time and spatial directions of the dynamic process, $\hat e_\perp$ and $\hat e_\parallel$, and the ones used in the master equation. There is no need for them to coincide. We choose a set-up where the master equation's time and space directions are redirected in the following manner. Lines of constant time are parallel to $\hat e_\parallel- \hat e_\perp$, such that the moving wall coincides with the $t=0$ line. Lines of constant position are parallel to the x-axis, which in the dynamic process represented lines of constant time. The following skewed dynamic rule implements this pace-time rotation. Consider a square space time lattice (Fig.\ref{lattice} rotated over 45 degrees). Each site in the master equation time slice $t$ is updated sequentially from right to left. The probability for site $x$ at time $\tau$ to be active depends on whether site $x-1$ was active at the previous time $t-1$ and/or at this moment in time, $t$. This set-up requires screwed boundary conditions. The forest fire runs under an angle. In this new interpretation the active wall represents a fully active initial configuration. The energy gap in the spectrum of the time evolution operator (transfer matrix), is related to the curtain width in the following manner. Let $|I\rangle$ be the initial state of the master equation, $|0\rangle$ the absorbing state, and $\hat T$ be the transfer matrix. The stochastic nature of the transfer matrix implies that the disordered state $|D\rangle$ is a left eigenvector with eigenvalue $\lambda_0=1$. Define $\hat a_x$ as the projection operator which returns one (zero) when site $x$ active (inactive). The curtain width is associated with the probability distribution for site $x$ to be active at time $t$ but after that never again. This takes the form \begin{equation} P(x,t) = \lim_{t_F\to \infty} \langle D |~ [(1-\hat a_x)\hat T]^{t_F-t}~ \hat a_x \hat T^t | I \rangle \end{equation} The operator $(1-\hat a_x)\hat T$ has $\lambda_0=1$ as largest eigenvalue since \begin{equation} (1-\hat a_x)\hat T] |0\rangle = |0\rangle \end{equation} and because attaching a projection operator to $\hat T$ can not result in an eigenvalue larger than the largest in $\hat T$. Let $\langle L_x|$ be the corresponding left eigenvector (which can be evaluated numerically). Inserting this leads to \begin{eqnarray} P(x,t) & = & \langle L_x| \hat a_x \hat T^t | I \rangle \\ \nonumber & \simeq & \langle L_x| \hat a_x|\lambda_1 \rangle \lambda_1^t \langle \lambda_1 | I \rangle \\ \nonumber & \sim & \exp[-t/\xi_t] \end{eqnarray} with $\xi_t= \log(\lambda_1)$ and $\lambda_1$ the next largest eigenvalue of $\hat T$. This illustrates that the curtain width scales in the same manner as the the characteristic time $\xi_t$ needed to reach the stationary state, when the latter is measured in this space-time twisted coordinate system. Fig.\ref{trans} shows the FSS estimates for the dynamic exponent $z$ according to $\xi_t\sim L^z$ and $W\sim L^z$. Both converge clearly to the DP dynamic exponent $z=1.58$. This confirms that no new independent curtain width exponent is presents. \section{Final Remarks} The analysis presented in this paper explains the anomalous scaling of the width of the slanted curtain boundary in DP type processes. The needles screen each other, and that leads an extra logarithmic factor according to the independent needle approximation. Our numerical data confirm the validity of this assumption. The same mechanism must apply to other dynamic processes, like directed Ising type absorbing state dynamics, and also to other quantities. Consider the following example. Directed percolation describes epidemic growth processes without immunization, where the probability to be sick at time $t+1$ requires that you yourself or at least one of your neighbours is already sick at time $t$. Consider an initial condition that everybody is sick at time $t=0$. A stationary local observer will conclude that below the percolation threshold the life time of the epidemic scales as $t\sim \epsilon^{-\nu_\parallel}$. A moving observer concludes it diverges faster, as $t\sim \epsilon^{-\nu_\parallel} \log(\frac{\epsilon_0 }{\epsilon})$. This research is supported by NSF grant DMR-9700430, by the Korea Research Foundation (98-015-D00090), and by an Inha University research grant (1998).
\section{Introduction} Recent observations of core collapse supernovae provide increasing evidence that the core collapse process is intrinsically asymmetric: 1) The spectra of these supernovae are significantly polarized indicating asymmetric envelopes (M\'endez et al. 1988; H\"oflich 1991, 1995; Jeffrey 1991; Trammel et al. 1993, Tran et al. 1997). The degree of polarization tend to vary inversely with the mass of the hydrogen envelope, being maximum for Type Ib/c events with no hydrogen (Wang et al. 1996; Wang, Wheeler \& H\"oflich 1999; Wheeler, H\"oflich \& Wang 1999). 2) After the explosion, neutron stars are observed with high velocities, up to 1000 km~s$^{-1}$ (Strom et al. 1995). 3) Observations of SN~1987A showed that radioactive material was brought to the hydrogen rich layers of the ejecta very quickly during the explosion (Lucy 1988; Sunyaev et al. 1987, Tueller et al. 1991). 4) The remnant of the Cas A supernova shows rapidly moving oxygen-rich matter outside the nominal boundary of the remnant (Fesen \& Gunderson, 1996) and evidence for two oppositely directed jets of high-velocity material (Fesen 1999; Reed, Hester, \& Winkler 1999). 5) High velocity ``bullets" of matter have been observed in the Vela supernova remnant (Taylor et al. 1993) Understanding the mechanism of producing supernovae explosions by core collapse is a physics problem that has challenged researchers for decades (Hoyle \& Fowler 1960; Colgate \& White 1966). The current most sophisticated calculations based on the neutrino energy deposition mechanism are multidimensional and involve the convection of the newly formed neutron star. These, however, but have failed to produce robust explosions (Herant et al. 1994; Burrows, Hayes \& Fryxell 1995; Janka \& M\"uller 1996; Mezzacappa et al 1997; Lichtenstadt, Khokhlov \& Wheeler 1999). Even when successful, these models do not explain why SN 1998bw produced the strongest radio source ever associated with a supernova, probably requiring a relativistic blast wave (Kulkarni et al. 1998), or account for a probable link between SN 1998bw and the $\gamma$-ray burst\ GRB980425 observed in the same general location in the same general time frame (Galama et al. 1998). The discovery of pulsars led to early considerations of the role of rotating magnetized neutron stars in the explosion mechanism (LeBlanc and Wilson 1970; Ostriker \& Gunn 1971; Bisnovatyi-Kogan 1971). LeBlanc and Wilson studied the magneto-rotational core collapse of a $7 M_{\odot}$ star. They numerically solved the two-dimensional MHD equations coupled to the equation for neutrino transport. Their simulations showed the formation of two oppositely directed, high-density, supersonic jets of material emanating from the collapsed core. They estimated that at the surface located $\sim 4 \times 10^8$~cm from the center, the jet carried away $\sim 10^{32}$~g with $\sim 1-2 \times 10^{51}$~ergs in $\sim 1$~s. The magnetic field generated in this calculation was $\sim 10^{15}$ Gauss. Evidence now exists for strongly magnetized neutron stars, ``magnetars" (Duncan \& Thompson 1992; Kouveliotou et al. 1998). The LeBlanc-Wilson mechanism is extremely asymmetric and contains jets. Their calculations only followed the jet to a distance of $\sim 10^8$ cm, whereas a stellar core has a radius of $10^{10}$~cm or more. The issues that arise are: how can this asymmetry propagate to much larger distances inside the star? Can these jets induce asymmetry at distances comparable to the stellar radius, or even push through the entire star and exit? In this paper, we model the explosion of a core collapse supernova assuming that the LeBlanc-Wilson mechanism has operated in the center. We take a $15 M_{\odot}$ main-sequence star evolved to the point of the explosion (Straniero, Chieffi \& Limongi 1999) and assume that the star has lost all of its hydrogen envelope before the explosion. The resulting $4.1$\ifmmode M_\odot\else M$_\odot$\fi\ model of a helium star corresponds to the explosion of a Type Ib or Ic supernova. The simulations show that the jets cause a very asymmetric explosion of the star. Most of the observations of asymmetries listed above can be explained by this process. \section{Numerical Simulations} Figure 1 presents a schematic of the setup of the computation. The computational domain is a cube of size $L = 1.5 \times 10^{11} $~cm with a spherical helium star of radius $R_{\rm star} = 1.88 \times 10^{10}$~cm and mass $M_{\rm star}\simeq~4.1$\ifmmode M_\odot\else M$_\odot$\fi\ placed in the center. The distribution of physical parameters inside the star is shown in Figure 2. The innermost part with mass $M_{\rm core} \simeq 1.6 M_{\odot}$ and radius $R_{\rm core} = 3.82 \times 10^8$~cm, consisting of Fe and Si, is assumed to have collapsed on a timescale much faster than the outer, lower-density material. It is removed and replaced by a point gravitational source with mass $ M_{\rm core}$ representing the newly formed neutron star. The remaining mass, from $\simeq 1.6$ to $\simeq 4.1 M_{\odot}$, consists of an O-Ne-Mg inner layer surrounded by the C-O and He-envelopes. This structure is mapped onto the computational domain from $R_{\rm core}$ to $R_{\rm star}$. At $R_{\rm core}$ and the outer boundary of the computational domain, we impose an outflow boundary condition assuming zero pressure, velocity, and density gradients. At two polar locations where the jets are initiated at $R_{core}$, we impose an inflow with velocity $v_j$, density $\rho_j$ and pressure $P_j$. The jet parameters are chosen to represent the results of LeBlanc \& Wilson (1970). At $R_{\rm core}$, the jet density and pressure are the same as those of the background material, $\rho_j = 6.5 \times 10^5$~g~cm$^{-3}$ and $P_j = 1.0 \times 10^{23}$~ergs~cm$^{-3}$, respectively. The radii of the cylindrical jets entering the computational domain are approximately $r_j = 1.2 \times 10^8$~cm. For the first 0.5~s, the jet velocity at $R_{\rm core}$ is kept constant at $v_j = 3.22 \times 10^9~cm~s^{-1}$. This results in a mass flux rate of $\sim 9.5\times10^{31}$ \gm-s with an energy deposition rate $d E / dt = 5\times 10^{50} $~ergs/s for each jet. After 0.5~s, the velocity of the jets at $R_{\rm core}$ was gradually decreased to zero at approximately 2~s. The total energy deposited by the jets is $E_j \simeq 9\times 10^{50}$~ergs and the total mass ejected is $M_j \simeq 2\times 10^{32}$~grams or $\simeq 0.1$\ifmmode M_\odot\else M$_\odot$\fi. These parameters are consistent within, but somewhat less than, those of the LeBlanc-Wilson model. The amount of material ejected is less than that which falls through the inner boundary during the jet operation, $\simeq 4\times10^{32}$~g. This amounts to an implicit assumption that $\sim 1/2$ of the matter accreted is channeled back out into the jets. More accurate jet parameters can only be determined by self-consistently modeling the formation of the jets in the vicinity of a neutron star. The stellar material was described by the time-dependent, compressible, Euler equations for inviscid flow with an ideal gas equation of state $P=E(\gamma-1)$ with constant $\gamma=5/3$. The Euler equations were integrated using an explicit, second-order accurate, Godunov type, adaptive-mesh-refinement, massively parallel, Fully-Threaded Tree (FTT) program, ALLA (Khokhlov 1998, Khokhlov \& Chtchelkanova 1999). Euler fluxes were evaluated by solving a Riemann problem at cell interfaces. FTT discretization of the computational domain allowed the mesh to be refined or coarsened at the level of individual cells. Physical scales involved in the simulation range from the size of the computational domain ($1.5 \times 10^{11} $~cm) to the jet diameter ($ \sim 10^8$~cm) and span at least three orders of magnitude. We used a cartesian, nonuniformly refined FTT mesh with fine cells $\Delta_{\rm min}\simeq 3.7\times 10^7$~cm near $R_{\rm core}$ to resolve the jets, and with cell size increasing towards the outer boundary of the computational domain where the cell size was $\Delta_{\rm max} = 2.3\times 10^9$~cm. This mesh was fixed from initial time $0$ to $ 6$~s of physical time. After that, the inner parts were coarsened near the center by a factor of four, and the central hole was eliminated. At this moment, the jets have exited the star and the details of the flow near $R_{\rm core}$ do not affect the essential features of the explosion. In this first, demonstration calculation, we did not use the time-adaptive mesh refinement capability of ALLA. It will be used to follow shocks and mixing processes with higher resolution in future simulations. We computed the entire configuration including both jets and assuming no symmetries. The total number of computational cells used in the simulation was $\sim 2\times 10^6$, whereas a uniform resolution $\Delta_{\rm min}$ would have required $\sim 7\times 10^{10}$ cells. \section{Results and Discussion} Figure 3 shows the propagation of the jet inside the star. As the jets move outwards, they remain collimated and do not develop much internal structure. A bow shock forms at the head of the jet and spreads in all directions, roughly cylindrically around each jet. The sound crossing time $\tau(r) = r/ a_s(r)$ is shown as a function of stellar radius $r$ in Figure 2, where $a_s(r)$ is the sound speed at a given radius for the initial stellar model. It might be expected that if energy were released at the center of a star on a timescale much shorter that $\tau(r)$, the effect of energy deposition at $r$ would resemble that of a strong point explosion. In particular, the jet characteristic time $\tau_j\sim 1$~s is much shorter than the sound crossing time of the star, $\tau(R_{\rm star}) \sim 10^3$~s (Figure 2). Nonetheless, these jets stay collimated enough to reach the surface as strong jets. It is known that supersonic jets stay collimated for a long distance. For example, Norman et al. (1983) simulated supersonic jets with densities $\rho_j$ both less than and greater than a uniform background, $\rho_b$. Jets with $\rho_j/\rho_b \ge 1$ developed a bow shock and little internal structure. Our jets resemble those with $\rho_j/\rho_b\ge 1$. The stellar matter is shocked by the bow shock, and then flows out and acts as a high-pressure confining medium by forming a cocoon around the jet. The sound crossing time of the dense O-Ne-Mg mantle, $\tau(R\sim 10^9~{\rm cm}) \simeq 10$~s, is only ten times longer than $\tau_j$, and the jets are capable of penetrating this dense inner part of the star in $\sim 2$~s. By the time the jets penetrate into the less dense C-O and He layers, the inflow of material into the jets has been turned off. By this time, however, the jets have become long bullets of high-density material moving through the background low-density material almost ballistically. The higher pressures in these jets cause them to spread laterally. This spreading is limited by a secondary shock that forms around each jet between the jet and the material already shocked by the bow shock. The radius of the jets, $\sim 3\times10^9$ cm as they emerge from the star, is larger than the initial radius, $\sim 10^8$ cm, but it is still significantly less than the radius of the star. After about 5.9 s, the bow shock reaches the edge of the star and breaks through. Figure 4 shows the subsequent evolution of the star after the breakthrough. By $\simeq 20$~s, most of the material in the jets has left the star propagates into the interstellar medium ballistically. We estimate the total mass in these two jets as $M_j \approx 0.05 M_\odot$ and the total kinetic energy $E_j \approx 2.5 \times 10^{50}$~ergs. The average velocity of the jet is about 25,000~$km~s^{-1}$. The laterally expanding bow shocks generated by the jets (Figure 3) move towards the equator where they collide with each other. The collision of the shocks first produces a regular reflection that then becomes a Mach reflection. The Mach stem moves outwards along the equatorial plane. The result is that the material in the equatorial plane is compressed and accelerated more than material in other directions (excluding the jet material). At $t\simeq 29$~s, the Mach stem reaches the outer edge of the star, and the star begins to settle into the free expansion regime. The computation was terminated at $\simeq 35$~s, before free expansion was attained. The stellar ejecta at this time is highly asymmetric. The density contour of $50~{\rm g~cm^{-3}}$, which is the average density of the ejecta at this time, forms an oblate configuration with the equator-to-polar velocity ratio $\simeq 2/1$. Complex shock and rarefaction interactions inside the expanding envelope will continue to change the distribution of the parameters inside the ejecta. Nonetheless, we expect that the resulting configuration will resemble an oblate ellipsoid with a very high degree of asymmetry, axis ratios $\ge 2$. \section{Conclusions} We have numerically studied the explosion of a supernova caused by supersonic jets generated in the center of the supernova as a result of the core collapse into a neutron star. We simulated the process of the jet propagation through the star, jet breakthrough, and the ejection of the supernova envelope by the lateral shocks generated during jet propagation. The end result of the interaction is a highly nonspherical supernova explosion with two high-velocity jets of material moving in polar directions ahead of an oblate, highly distorted ejecta containing most of the supernova material. Below we argue that such a model explains many of the observations that are difficult or impossible to explain by the neutrino deposition explosion mechanisms. We have assumed that the jets were generated by a magneto-rotational mechanism during core collapse and neutron star formation (LeBlanc \& Wilson 1970). That collimated jets could be a common phenomenon in core collapse supernovae and be associated with $\gamma$-ray\ bursts was raised recently by Wang \& Wheeler (1998). A different mechanism of jet generation involving neutrino radiation during collapse of a very massive star into a black hole has been recently discussed by MacFayden \& Woosley (1999) in the context of a ``failed'' supernova, also to explain $\gamma$-ray\ bursts. Low density relativistic jets may also be produced by the intense radiation of the newly born pulsar, as discussed by Blackman \& Yi (1998) and Yi et al. (1999). We found in our preliminary simulations (not presented here) that lower density and higher velocity jets than the one considered in this paper may produce similar hydrodynamical effects. The asymmetric explosion generated in this calculation provides ejection velocities that are comparable to those observed in supernovae. For this particular calculation, an energy of $2.5\times 10^{50}$ ergs is invested in the jet and the star of $\simeq 2.5$\ifmmode M_\odot\else M$_\odot$\fi\ is ejected with kinetic energy of $6.5\times 10^{50}$ ergs and average velocity $3,000-4,000~km~s^{-1}$. Increasing the jet opening angle, jet duration, or jet velocity would result in a more powerful explosion. The density and velocity profiles of the main ejecta (excluding jets) are oblate with equator to polar ratios greater than 2/1. This structure will produce significant polarization, of order 1\% or more as observed in bare-core supernovae (H\"oflich, Wheeler \& Wang 1999). The two polar jets move outward from the star with a speed $\sim 25,000~km~s^{-1}$, much greater than the ejecta itself. They may be detected in supernova remnants and might account for the evidence of jets in in Cas A (Fesen \& Gunderson 1996; Reed, Hester \& Winkler 1999). The composition of the jets must reflect the composition of the innermost parts of the star, and should contain heavy and intermediate-mass elements. During the explosion, the jets would bring heavy and intermediate mass elements into the outer layers. This will influence the spectral and polarization properties of a supernova. Here we considered a bare helium core, but if the core were inside a hydrogen envelope, the explosion would remain very inhomogeneous. Radioactive elements could potentially be carried into the hydrogen envelope. This could explain the early appearance of X-rays, as in SN1987A. It is plausible that a sufficiently powerful jet could even penetrate a hydrogen envelope. We assumed that the jets are identical which is not the general case. Any momentum imbalance might impart a kick to the neutron star. From momentum conservation, we estimate the required difference between the inflow velocities of the jets, $\Delta v_j$, be of the order of $$ {\Delta v_j\over v_j} \simeq {M_{NS} \over M_j} \, {v_{NS}\over v_j} \simeq 1.0 \, \left( v_{NS}\over 1,000{\rm km/s}\right)\, \left( 30,000{\rm km/s}\over v_j \right)~, $$ where $V_{NS}$ is the kick velocity, and we have taken the neutron mass $M_{NS}= 1.5$\ifmmode M_\odot\else M$_\odot$\fi, and the jet mass $M_j= 10^{32}~g$. Although the required jet asymmetry, ${\Delta v_j\over v_j}=1$, to produce a 1,000 km/s kick may seem extreme, the parameters of jets selected for this calculations are mild. If the duration of the jets is increased by a factor of two, asymmetry of only 0.5 would be required. When the jets break through the stellar photosphere, a small amount of mass will be accelerated through the density gradient to very high velocities. Our resolution was not enough and the code does not have a relativistic Riemann solver incorporated to make quantitative predictions; however, a small fraction of the material at the stellar surface was observed to move with a velocity of up to $\sim$ 90,000 km~s$^{-1}$. This may, in principle, lead to the $\gamma$-ray burst\ and the radio outburst similar to those associated with SN~1998bw/GRB980425. The jet-induced explosion of a supernova computed in this paper is entirely due to the action of the jet on the surrounding star. The mechanism that determines the energy of such an explosion must be related to the shut-off of the accretion onto the neutron star by the lateral shocks that accelerate the material outwards. The explosion thus does not depend on neutrino transport or re-acceleration of the stalled shock. This work opens many issues that require further investigation. A study must be made of different input parameters, including properties of the jets and of the initial star, and the jet engine mechanisms must be studied as well. These studies are currently underway. \acknowledgments Computations were performed on Origin 2000 at the Naval Research Laboratory. The authors are grateful to Rob Duncan and Insu Yi for helpful discussions, and to Almadena Chtchelkanova for the development of massively parallel software used in the simulations. This research was supported in part by NSF Grant 95-28110, NASA Grant NAG 5-2888, NASA Grant LSTA-98-022 and a grant from the Texas Advanced Research Program. The Laboratory for Computational Physics and Fluid Dynamics at the Naval Research Laboratory thanks NASA Astrophysics Theory Program for support.
\section{Introduction} The importance of the notion of uniformity of point sets in the problem of numerical integration has been stressed in many publications (see e.g. \cite{nieder1,prepubs,jhk}). In general, a {\em discrepancy} is a measure of non-uniformity of point sets in an integration region. For a set of $N$ random points it is a function of $N$ random variables. For the {\em Lego} discrepancy, the integration region is considered to be divided into a number of, say $M$ bins, so that it becomes a function of $M$ random variables, namely the number of points in each of the bins. Furthermore, the Lego discrepancy is defined such that it is exactly the $\chi^2$-statistic of the binned data points. In \cite{hkh1}, we have given criteria for the asymptotic probability distribution of various quadratic discrepancies to become Gaussian when a certain free parameter becomes infinitely large. This parameter often is the dimension $s$ of the integration region. In the case of the Lego discrepancy, it is the number of bins $M$. In \cite{leeb}, it is shown that for the {\em Fourier diaphony} a Gaussian limit is obtained when both $N$ and $s$ go to infinity such that $c^s/N\ra0$, where $c$ is some constant larger than $1$. This theorem clearly gives more information about the behavior of the probability distribution than the statements of $\cite{hkh1}$, for it relates $s$ and $N$, whereas in \cite{hkh1} the limit of $N\rightarrow\infty$ is assumed before considering the behavior with respect to $s$ or $M$. In \cite{hak1} and \cite{hak3}, we introduced techniques to calculate the $1/N$ expansion of the moment generating function of the probability distribution of quadratic discrepancies, and applied it to the $L_2^*$-discrepancy, the Fourier diaphony (both for $s=1$), and the Lego discrepancy. Such an expansion can be used to calculate corrections to the asymptotic distribution for $N\rightarrow\infty$, but, presented as in \cite{hak3}, it cannot give information about limits if $M$ or $s$ become infinite also. In this paper we use the mentioned technique to calculate limits for the Lego discrepancy if $M$ as well as $N$ become infinite. First, we will show that the natural expansion parameter in the calculation of the moment generating function is $M/N$, and calculate a few terms. We will see, however, that a strict limit of $M\rightarrow\infty$ does not exist, and, in fact, this is well known because the $\chi^2$-distribution, which gives the lowest order term in this expansion, does not exist if the number of degrees of freedom becomes infinite. We overcome this problem by going over to the standardized variable, which is obtained from the discrepancy by shifting and rescaling it such that it has zero expectation and unit variance. In fact, it is this variable for which the results in \cite{hkh1} and \cite{leeb} were obtained. In this paper, we derive similar results for the Lego discrepancy, depending on the behavior of the sizes of the bins if $M$ goes to infinity. We will see that various asymptotic probability distributions occur if $M,N\rightarrow\infty$ such that $M^\alpha/N\rightarrow\textsl{constant}$ with $\alpha\geq0$. If, for example, the bins become asymptotically equal and $\alpha>{\textstyle\frac{1}{2}}$, then the probability distribution becomes Gaussian. Notice that this includes limits with $\alpha<1$, which is in stark contrast with the rule of thumb that, in order to trust the $\chi^2$-distribution, each bin has to contain at least a few, say five (see e.g. \cite{knuth}), data points. Our result states that, for large $M$ and $N$, the majority of bins is allowed to remain empty! \section{The $\bs{\chi^2}$-statistic as a discrepancy} The $\chi^2$-statistic for $N$ data points distributed over $M$ bins with expected number of points $w_nN$ for bin $n=1,\ldots,M$ is given by \begin{equation} \chi^2 \;=\; \sum_{n=1}^M\frac{(\mathcal{N}_n-w_nN)^2}{w_nN} \;\;, \end{equation} where $\mathcal{N}_n$ is the number of points in bin $n$. If the data points are distributed truly random, i.e., if they are independent and if the probability for a point to fall in bin $n$ is equal to $1/w_n$, then it is distributed along a multinomial distribution: \begin{equation} \Prop{\chi^2\leq t} \;=\; \sum_{\{\mathcal{N}_m\}}\theta(t-\chi^2)\, N!\prod_{m=1}^M\frac{w^{\mathcal{N}_m}}{\mathcal{N}_m!} \;\;, \end{equation} where the sum is over all configurations $\{\mathcal{N}_m\}$ with $\sum_{m=1}^M\mathcal{N}_m=N$. In the limit of an infinite number of data points, this becomes the $\chi^2$-distribution, which has moment generating function \begin{equation} \frac{1}{(1-2z)^{\frac{M-1}{2}}} \;=\; G_0(z) \;\;. \end{equation} A first step in the interpretation of this statistic as a quadratic discrepancy is the insight that it is quadratic in the variables $\mathcal{N}_n$ and can be characterized by a matrix with elements \begin{equation} \mathcal{B}_{n,m} \;=\; \frac{\delta_{n,m}}{w_n} - 1 \;\;, \label{LegEq002} \end{equation} so that $\chi^2=\frac{1}{N}\sum_{n,m}^M\mathcal{N}_n\mathcal{B}_{n,m}\mathcal{N}_m$. Furthermore, we assume that the bins are disjunct subsets of an integration region, and that the data points $x_k$, $1\leq k\leq N$ are in this integration region. The measure of subset $n$ is then equal to $w_n$. The union of the subsets we call ${\bf K}$; it has measure \begin{equation} \sum_{n=1}^Mw_n \;=\; 1 \;\;. \end{equation} The number of points $\mathcal{N}_n$ is equal to $\sum_{k=1}^N\vartheta_n(x_k)$, where $\vartheta_n$ is the characteristic function of bin $n$, so that \begin{equation} \chi^2 \;=\; \frac{1}{N}\sum_{k,l=1}^N\sum_{n,m=1}^M\vartheta_n(x_k)\mathcal{B}_{n,m}\vartheta_m(x_l) \;=\; \frac{1}{N}\sum_{k,l=1}^N\mathcal{B}(x_k,x_l) \;=\; D_N \;\;. \end{equation} The interpretation of a quadratic discrepancy in the definition following \cite{hak1,hak3} is possible, because the two-point function $\mathcal{B}$ as obtained above integrates to zero with respect to each of its variables. From now on, we will call $D_N$ the Lego discrepancy. \section{Sequences and notation} In the following, we will investigate limits in which the number of bins $M$ goes to infinity. Note that for each value of $M$, we have to decide on the values of the measures $w_n$. They clearly have to scale with $M$, because their sum has to be equal to one. There are, of course, many possible ways for the measures to scale, i.e., many double-sequences $\{w_n^{(M)},1\leq n\leq M,M>0\}$ of positive numbers with \begin{equation} \sum_{n=1}^Mw_n^{(M)} \;=\; 1 \quad\forall\,M>0 \quad\quad\textrm{and}\quad\quad \lim_{M\rightarrow\infty}\sum_{n=1}^Mw_n^{(M)} \;=\; 1 \;\;. \end{equation} We, however, want to restrict ourselves to discrepancies in which the relative sizes of the bins stay of the same order, i.e., sequences for which \begin{equation} \inf_{n,M}Mw_n^{(M)}\in(0,\infty) \qquad\textrm{and}\qquad \sup_{n,M}Mw_n^{(M)}\in(0,\infty) \;\;. \label{LegEq001} \end{equation} It will appear to be appropriate to specify the sequences under consideration by another criterion, which is for example satisfied by the sequences mentioned above. It can be formulated in terms of the objects \begin{equation} M_p \;=\; \sum_{n=1}^{M}\frac{1}{\left[w_n^{(M)}\right]^{p-1}} \;\;,\quad p\geq1 \;\;, \label{LegEq022} \end{equation} and is given by the demand that \begin{equation} \lim_{M\rightarrow\infty} \frac{M_p}{M^p} \;=\; h_p \in[1,\infty) \quad\forall\,p\geq1 \;\;. \label{LegEq018} \end{equation} Within the set of sequences we consider, there are those with for which the bins become asymptotically equal, i.e., sequences with \begin{equation} w_n^{(M)} = \frac{1 + \varepsilon_n^{(M)}}{M} \quad\textrm{with}\quad \varepsilon_n^{(M)}>-1,\;\;1\leq n\leq M \quad\textrm{and}\quad \lim_{M\rightarrow\infty}\max_{1\leq n\leq M}|\varepsilon_n^{(M)}| = 0 \;\;. \end{equation} They belong to the set of sequences with $h_p=1$ $\forall\,p\geq1$, which will allow for special asymptotic probability distributions. In the following analysis, we will consider functions of $M$ and their behavior if $M\rightarrow\infty$. To specify relative behaviors, we will use the symbols ``$\sim$'', ``$\asymp$'' and ``$\prec$''. The first one is used as follows: \begin{equation} f_1(M) \sim f_2(M) \quad\Longleftrightarrow\quad \lim_{M\rightarrow\infty} \frac{f_1(M)}{f_2(M)} \;=\; 1 \;\;. \end{equation} If a limit as above is not necessarily equal to one and not equal to zero, then we use the second symbol: \begin{equation} f_1(M) \asymp f_2(M) \quad\Longleftrightarrow\quad f_1(M) \sim cf_2(M) \;\;,\quad c\in(0,\infty) \;\;. \end{equation} We only use this symbol for those cases in which $c\neq0$. For the cases in which $c=0$ we use the third symbol: \begin{equation} f_1(M) \prec f_2(M) \quad\Longleftrightarrow\quad \lim_{M\rightarrow\infty} \frac{f_1(M)}{f_2(M)} \;=\; 0 \;\;. \end{equation} We will also use the ${\mathcal O}$-symbol, and do this in the usual sense. We can immediately use the symbols to specify the behavior of $M_p$ with $M$, for the criterion of \eqn{LegEq018} tells us that \begin{equation} M_p \asymp M^p \;\;, \label{LegEq019} \end{equation} and that \begin{equation} M_p \sim M^p \quad\textrm{if}\quad h_p=1 \;\;. \end{equation} In our formulation, also the number of data points $N$ runs with $M$. We will, however, never denote the dependence of $N$ on $M$ explicitly and assume that it is clear from now on. Also the upper index at the measures $w_n$ we will omit from now on. \section{Feynman rules} In \cite{hak1} and \cite{hak3} we have shown that the moment generating function $G:z\mapsto \Exp{e^{zD_N}}$ of the probability distribution of the Lego discrepancy can be written as an integral over $M$ ordinary degrees of freedom (the {\em bosons}) and $2N$ Grasmannian degrees of freedom (the {\em fermions}). The integral is not unique and contains a {\em gauge freedom}. Furthermore, it introduces a natural expansion parameter \begin{equation} g=\sqrt{\frac{2z}{N}} \;\;, \end{equation} to calculate the generating function perturbatively. In one particular gauge, the {\em Landau} gauge, the terms are equal to the contribution of {\em Feynman diagrams}, that can be calculated according to the following rules: \begin{align} &\textrm{boson propagator:}\hspace{6pt}\quad n\,\diagram{dcP2}{50}{-1}\,m \;=\; \mathcal{B}_{n,m} \;\;; \tag*{rule 1}\\ &\textrm{fermion propagator:}\quad i\,\diagram{dcP1}{50}{2}\,j\hspace{5pt} \;=\; \delta_{i,j}\;\;;\tag*{rule 2}\\ &\textrm{vertices:}\hspace{60pt}\quad \diagram{dcV1}{65}{34}\hspace{4pt} \;=\; -g^p\times\textrm{convolution} \;\;,\quad p\geq 2 \label{CorEq007}\;\;. \tag*{rule 3} \end{align} To calculate a term in the expansion series, the contribution of all {\em vacuum} diagrams, i.e. all diagrams without external legs, carrying the same power of $g$ has to be taken into account. In the Landau gauge, only {\em connected} diagrams have to be calculated, for \begin{equation} \log G(z) \;=\; \textsl{the sum of the connected vacuum diagrams.} \tag*{rule 4} \end{equation} In the vertices, boson propagators are convoluted as $\sum_{m=1}^Mw_m\mathcal{B}_{m,n_1}\mathcal{B}_{m,n_2}\cdots\mathcal{B}_{m,n_p}$, fermion propagators as $\sum_{j=1}^{\raisebox{-1pt}{${\scriptstyle N}$}}\delta_{i_1,j}\delta_{j,i_2}$, and then these convolutions are multiplied. As a result of this, the bosonic part of each diagram decouples completely from the fermionic part. The contribution of the fermionic part can easily be determined, for \begin{equation} \textsl{every fermion loop only gives a factor $-N$.}\tag*{rule 5} \end{equation} \subsection{Rules for the bosonic parts} Because of the rather simple expression (\ref{LegEq002}) for the bosonic propagator, we are able to deduce from the basic Feynman rules some effective rules for the bosonic parts of the Feynman diagrams. Remember that the bosonic parts decouples completely from the fermionic parts. The following rules apply after having counted the number of fermion loops and the powers of $g$ coming from the vertices, and after having calculated the symmetry factor of the original diagram. When we mention the {\em contribution} of a diagram in this section, we refer to the contribution apart from the powers of $g$ and symmetry factors. This contribution will be represented by the same drawing as the diagram itself. The first rule is a consequence of the fact that \begin{equation} \sum_{n=1}^Mw_m\mathcal{B}_{n_1,m}\mathcal{B}_{m,n_2} \;=\; \mathcal{B}_{n_1,n_2} \end{equation} and states that all vertices with only two legs that do not form a single loop can be removed. The second rule is a consequence of the fact that for any $M\times M$-matrix $f$ \begin{equation} \sum_{n,m=1}^Mw_nw_mf_{n,m}\mathcal{B}_{n,m} \;=\; \sum_{n=1}^Mw_nf_{n,n} - \sum_{n,m=1}^Mw_nw_mf_{n,m} \;\;, \end{equation} and states that the contribution of a diagram is the same as that of the diagram in which a boson line is contracted and the two vertices, connected to that line, are fused together to form one vertex, minus the contribution of the diagram in which the line is simply removed and the vertices replaced by vertices with one boson leg less. This rule can be depicted as follows \begin{equation} \diagram{lcH3}{75}{17} \;=\; \diagram{lcH2}{52}{17} - \diagram{lcH1}{75}{17} \;\;.\tag*{rule 6} \end{equation} By repeated application of these rules, we see that the contribution of a connected bosonic diagram is equal to the contribution of a sum of products of so called {\em daisy} diagrams~\footnote{For example $\diagram{lcBr1}{20}{8}=\diagram{lcBr2}{20}{8}-\diagram{lcBr3}{20}{8} =\diagram{lcBr4}{34}{6}-2\,\diagram{lcBr3}{20}{8} =\diagram{lcBr5}{26}{10}-\diagram{lcBr6}{30}{4} - 2\left(\diagram{lcBr6}{30}{4}-\diagram{lcBr7}{15}{4}\right)$}, which are of the type \begin{equation} \diagram{lcDia2}{70}{24} \;\;. \label{LegEq003} \end{equation} They are characterized by the fact that all lines begin and end on the same vertex and form single loops. The contribution of such a diagram is given by \begin{equation} d_p(M) \;=\; \sum_{n=1}^Mw_n\mathcal{B}_{n,n}^p \;=\; \sum_{q=0}^p\binom{p}{q}(-1)^{p-q}M_q \;=\; M_p[1+{\mathcal O}(M^{-1})]\;\;, \label{LegEq021} \end{equation} where the last equation follows from \eqn{LegEq019}. The maximal number of leaves, a product in the sum of daisy diagrams contains, is equal to the number of loops $L_B$ in the original diagram, so that \begin{equation} \textsl{the contribution of a diagram with $L_B$ boson loops is $M_{L_B}[1+{\mathcal O}(M^{-1})]$.} \tag*{rule 7} \end{equation} The leading order contribution of a diagram with $L_B$ boson loops is thus of the order of $M^{L_B}$. \subsection{Extra rule if $\bs{h_p=1}$} If $h_p=1$ $\forall\,p\geq1$, then all kind of cancellations between diagrams occur, because in those cases $M_p\sim M^p$ $\forall\,p\geq1$. As a result of this, the contribution of a daisy diagram is $d_p(M)\sim M^p$, and we can deduce the following rule: the contribution of a diagram that falls apart in disjunct pieces if a vertex is cut, is equal to the product of the contributions of those disjunct pieces times one plus vanishing corrections. Diagrammatically, the rule looks like \begin{equation} \diagram{lcV1}{60}{13} \;\sim\; \diagram{lcVA}{30}{13}\times\diagram{lcVB}{30}{13} \;\;, \label{LegEq016} \end{equation} In \cite{hak3} we called discrepancies for which \eqn{LegEq016} is exact one-vertex decomposable, and have shown that for those discrepancies only the {\em one-vertex irreducible} diagrams contribute, i.e., diagrams that do not fall apart in pieces containing bosonic parts if a vertex is cut. The previous rule tells us that, if $h_p=1$ $\forall\,p\geq1$, then \begin{equation} \log G(z) \;\sim\; \textsl{sum of all connected one-vertex irreducible diagrams}. \tag*{rule 8} \end{equation} The connected one-vertex irreducible diagrams we call {\em relevant} and the others {\em irrelevant}. \section{Loop analysis} We want to determine the contribution of the diagrams in this section, and in order to do that, we need to introduce some notation: \begin{align} L_B \;&=\; \textrm{the number of boson loops} \;\;;\\ L_F \;&=\; \textrm{the number of fermion loops} \;\;;\\ L \;&=\; \textrm{the total number of loops} \;\;;\\ I_B \;&=\; \textrm{the number of bosonic lines} \;\;;\\ I_F \;&=\; \textrm{the number of fermionic lines} \;\;;\\ v \;&=\; \textrm{the number of vertices} \;\;;\\ L_M \;&=\; L-L_B-L_F \;=\; \textrm{number of mixed loops}\;\;. \end{align} These quantities are in principle functions of the diagrams, but we will never denote this dependence explicitly, for it will always be clear which diagram we are referring to when we use the quantities. With the foregoing, we deduce that the contribution $C_A$ of a connected diagram $A$ with no external legs satisfies \begin{equation} C_A \;\asymp\; M^{L_B}N^{L_F}g^{2I_B} \;\;. \end{equation} The Feynman rules and basic graph theory tell us that, for connected diagrams with no external legs, $v=L_F$ and $L=I_B+I_F-v+1$, so that \begin{equation} I_B \;=\; L-1 \;=\; L_B+L_F+L_M-1 \;\;. \end{equation} If we furthermore use that $g=\sqrt{2z/N}\,$, we find that the contribution is given by \begin{equation} C_A \;\asymp\; \frac{M^{L_B}}{N^{L_M}N^{L_B-1}}\,(2z)^{I_B} \;\;. \label{LegEq005} \end{equation} Notice that this expression does not depend on $L_F$. Furthermore, it is clear that, for large $M$ and $N$, the largest contribution comes from diagrams with $L_M=0$. Moreover, we see that we must have $N={\mathcal O}(M)$, for else the contribution of higher-order diagrams will grow with the number of boson loops, and the perturbation series becomes completely senseless. If, however, $N\asymp M$, then the contribution of each diagram with $L_M=0$ is more important than the contribution of each of the diagrams with $L_M>0$. Finally, we also see that the contribution of the ${\mathcal O}(M^{-1})$-corrections of a diagram (\eqn{LegEq021}) is always negligible compared to the leading contribution of each diagram with $L_M=0$. These observations lead to the conclusion that, if $N$ and $M$ become large with $N\asymp M$, then the leading contribution to $\log G(z)$ comes from the diagrams with $L_M=0$, and that there are no corrections to these contributions. If we assume that $M/N$ is small, then the importance of these diagrams decreases with the number of boson loops $L_B$ as $(M/N)^{L_B}$. \subsection{The loop expansion of $\bs{\log G(z)}$} Now we calculate the first few terms in the loop expansion of $\log G(z)$. We start with the diagrams with one loop (remember that it is an expansion in boson loops and that we only have to calculate {\em connected} diagrams for $\log G(z)$). The sum of all $1$-loop diagrams with $L_M=0$ is given by \begin{align} \frac{1}{2}\,\diagram{dcZ1}{40}{8} \;+\; \frac{1}{4}\,\diagram{dcZ2}{60}{8} \;+\; \frac{1}{6}\,\diagram{dcZ3}{50}{16} \;+\; \cdots \;&=\; (M-1)\sum_{p=1}^{\infty}\frac{1}{2p}\,(Ng^2)^p \notag\\ \;&=\; -\frac{M-1}{2}\,\log(1-2z) \;\;, \label{LegEq007} \end{align} and this is exactly equal to $\log G_0(z)$ as we know it for the Lego-discrepancy (cf.\cite{hak1}). The fractions before the diagrams are the symmetry factors. From now on, we will always write them down explicitly. To calculate the higher loop diagrams, we introduce the following effective vertex: \begin{equation} \diagram{dcV2}{70}{30} \;\equiv\; \diagram{dcV3}{70}{30} \;\;, \label{LegEq006} \end{equation} and the following partly re-summed propagator: \begin{align} n\,\diagram{dcP6}{50}{-1}\,m \;&=\; n\,\diagram{dcP2}{50}{-1}\,m \;+\; n\,\diagram{dcP3}{50}{0}\,m \;+\; n\,\diagram{dcP4}{75}{0}\,m \notag\\ &\hspace{183pt} \;+\; n\,\diagram{dcP5}{100}{0}\,m \;+\; \cdots \notag\\ \;&=\; \sum_{p=0}^{\infty}(Ng^2)^p\times n\,\diagram{dcP2}{50}{-1}\,m \;=\; \frac{1}{1-2z}\times n\,\diagram{dcP2}{50}{-1}\,m \;\;. \end{align} It is the propagator $\mathcal{G}_{n,m}^{(z)}$ from \cite{hak3}. The contribution of the $2$-loop diagrams with $L_M=0$ is given by \begin{align} & \frac{1}{8}\,\diagram{dcA11}{40}{5} \;+\; \frac{1}{8}\,\diagram{dcA22}{60}{7} \;+\; \frac{1}{8}\,\diagram{dcA13}{58}{5} \;+\; \frac{1}{12}\,\diagram{dcA12}{30}{11} \notag\\ \;&=\; \left[ - \frac{1}{8}\,\frac{Ng^2M^2}{(1-2z)^2} + \frac{1}{8}\,\frac{Ng^2M_2}{(1-2z)^2} + \frac{1}{8}\,\frac{(Ng^3)^2(M_2-M^2)}{(1-2z)^3} + \frac{1}{12}\,\frac{(Ng^3)^2M_2}{(1-2z)^3}\right] [1+{\mathcal O}(M^{-1})] \notag\\ \;&=\; \frac{1}{N} \left[\frac{1}{8}(M_2-M^2)\eta(z)^2 + \left(\frac{5M_2}{24}-\frac{M^2}{8}\right)\eta(z)^3\right] [1+{\mathcal O}(M^{-1})] \;\;, \end{align} where we define \begin{equation} \eta(z) \;=\; \frac{2z}{1-2z} \;\;. \end{equation} Notice that the first three diagrams vanish if $h_p=1$ $\forall\,p\geq1$. The contribution of the $3$-loop diagrams with $L_M=0$ is given by \begin{align} & \frac{1}{48}\,\diagram{dcB12}{30}{11} \;+\;\frac{1}{24}\,\diagram{dcB111}{30}{12} \;+\;\frac{1}{8}\,\diagram{dcB18}{30}{12} \;+\;\frac{1}{16}\,\diagram{dcB115}{40}{13} \;+\;\frac{1}{48}\,\diagram{dcB11}{35}{11} \;+\;\frac{1}{12}\,\diagram{dcB14}{50}{9} \notag\\ &\;+\;\frac{1}{8}\,\diagram{dcB16}{75}{5} \;+\;\frac{1}{16}\,\diagram{dcB13}{60}{6} \;+\;\frac{1}{8}\,\diagram{dcB17}{50}{11} \;+\;\frac{1}{16}\,\diagram{dcB19}{60}{16}\notag\\ &\;+\;\frac{1}{8}\,\diagram{dcB212}{95}{7} \;+\;\frac{1}{16}\,\diagram{dcB114}{90}{5} \;+\;\frac{1}{8}\,\diagram{dcB113}{65}{11} \;+\;\frac{1}{12}\,\diagram{dcB110}{65}{9} \notag\\ &\;+\;\frac{1}{8}\,\diagram{dcB21}{80}{7} \;+\;\frac{1}{16}\,\diagram{dcB22}{55}{17} \;+\;\frac{1}{8}\,\diagram{dcB26}{65}{11} \;+\;\frac{1}{12}\,\diagram{dcB27}{70}{10} \notag\\ &\;+\;\frac{1}{48}\,\diagram{dcB112}{60}{20} \;+\;\frac{1}{16}\,\diagram{dcB213}{60}{27} \;+\;\frac{1}{8}\,\diagram{dcB214}{60}{14} \;+\;\frac{1}{16}\,\diagram{dcB15}{45}{19} \;+\;\frac{1}{24}\,\diagram{dcB32}{54}{18}\notag\\ &\;+\;\frac{1}{16}\,\diagram{dcB34}{90}{7} \;\;. \end{align} If $h_p=1$ $\forall\,p\geq1$, then only the first four diagrams are relevant, and their contribution $C$ satisfies \begin{equation} C \;\sim\; \frac{M^3}{N^2}\left[ \frac{1}{48}\eta(z)^4 \;+\; \frac{1}{8}\eta(z)^5 \;+\; \frac{5}{48}\eta(z)^6\right] \;\;. \end{equation} \section{Various limits} In the previous calculations, $M/N$ was the expansion parameter and the expansion of the generating function only makes sense if it is considered to be small. In fact, a limit in which $M\rightarrow\infty$ does not even exist, because the zeroth order term is proportional to $M$. In order to analyze limits in which $M$ as well as $N$ go to infinity, we can go over to the standardized variable $(D_N-E)/\sqrt{V}\,$ of the discrepancy, where \begin{align} E \;&=\; \Exp{D_N} \;=\; M-1 \\ V \;&=\; \Var{D_N} \;=\; 2(M-1) + \frac{M_2-M^2-2(M-1)}{N} \;\;. \end{align} In terms of the standardized variable, the generating function is given by \begin{equation} \hat{G}(\xi) \;=\; \Exp{e^{\xi\,\frac{D_N-E}{\sqrt{V}}}} \;=\; \exp\left(-\frac{E\xi}{\sqrt{V}}\right)\, G\left(\frac{\xi}{\sqrt{V}}\right) \;\;. \label{LegEq008} \end{equation} Instead of the variable $z$, the variable $\xi=z\sqrt{V}$ is considered to be of ${\mathcal O}(1)$ in this perspective and the contribution of a diagram changes from (\ref{LegEq005}) to \begin{equation} C_A \;\asymp\; \frac{M^{L_B}}{N^{L_M}N^{L_B-1}V^{{\scriptstyle\frac{1}{2}}(L_B+L_F+L_M-1)}}\, (2\xi)^{I_B} \;\;. \end{equation} In the following we will investigate limits of $M\rightarrow\infty$ with, at first instance, the criterion of \eqn{LegEq018} as only restriction. The fact that the variance $V$ shows up explicitly in the contribution of the diagrams, forces us to specify the behavior of $M_2$ more precisely. We will take \begin{equation} M_2-M^2 \asymp M^\gamma \;\;,\quad 0\leq\gamma\leq2 \;\;. \label{LegEq023} \end{equation} Notice that $h_2=1$ if $\gamma<2$ and that $h_2$ does not exist if $\gamma>2$. Furthermore, we cannot read off the natural expansion parameter from the contribution of the diagrams anymore, and have to specify the behavior of $N$. We will only consider limits in which \begin{equation} N \asymp M^\alpha \;\;,\quad \alpha>0 \;\;. \label{LegEq024} \end{equation} Although they are a small subset of possible limits, those that can be specified by a pair $(\alpha,\gamma)$ show an interesting picture. We will derive the results in the next section, but present them now in the following phase diagram: \begin{center} \begin{picture}(160,160)(5,5) \LongArrow(0,0)(150,0) \LongArrow(0,150)(150,150) \Line(0,0)(0,149.5) \Line(0,105)(3,105) \Line(31.5,105)(34.5,105) \DashLine(35,0)(35,105){3} \DashLine(35,105)(0,150){3} \Text(0,-10)[]{$0$} \Text(35.5,-10)[]{$\frac{1}{2}$} \Text(-10,1)[]{$0$} \Text(-10,106)[]{$\frac{3}{2}$} \Text(-10,150)[]{$2$} \LongArrow(-20,70)(-20,80) \Text(-28,75)[]{$\gamma$} \Text(140,-10)[]{$\alpha$} \Text(18,70)[]{$\bs{T}$} \Text(100,100)[]{$\bs{U}$} \LongArrow(75,30)(40,40) \Text(80,30)[]{$\bs{\ell}$} \end{picture} \end{center} \vspace{20pt} It shows the region $\bs{S}=\{(\alpha,\gamma)\in{\bf R}^2\,|\,\alpha\in[0,\infty),\,\gamma\in[0,2]\}$ of the real $(\alpha,\gamma)$-plane. In this region, there is a {\em critical line} $\bs{\ell}$, given by \begin{equation} \bs{\ell}=\{(f_{\bs{\ell}}(t),t)\in\bs{S}\,|\,t\in[0,2]\} \quad\textrm{with}\quad f_{\bs{\ell}}(t) = \begin{cases} {\textstyle\frac{1}{2}} &\textrm{if $0\leq t\leq\frac{3}{2}$}\;\;,\\ 2-t &\textrm{if $\frac{3}{2}\leq t\leq 2$}\;\;. \end{cases} \end{equation} It separates $\bs{S}$ into two regions $\bs{T}$ and $\bs{U}$, neither of which contains $\bs{\ell}$. Our results are the following. Firstly, \begin{equation} \textsl{in the region $\bs{T}$, the limit of $M\rightarrow\infty$ is not defined.} \label{LegRes01} \end{equation} In this region, the standardized variable is not appropriate, and we see that there are to many diagrams that grow indefinitely with $M$. Secondly, \begin{equation} \textsl{in the region $\bs{U}$, the limit of $M\rightarrow\infty$ gives a Gaussian distribution.} \label{LegRes02} \end{equation} Because we used the standardized variable, this distribution has necessarily zero expectation and unit variance. Finally, \begin{equation} \textsl{on the line $\bs{\ell}$, various limits exist, depending on the behaviour of $M_p$, $p>2$.} \label{LegRes03} \end{equation} One of these limits we were able to calculate explicitly. It appears if $M_p-M^p\prec M^{p-\frac{1}{2}}$ $\forall p\geq1$, which is, for example, satisfied in the case of equal binning. In this limit, the generating function is given by \begin{equation} \log\hat{G}(\xi) \;=\; \frac{1}{\lambda^2}\left(e^{\lambda\xi}-1-\lambda\xi\right) \;\;,\quad \lambda=\lim_{M\rightarrow\infty}\frac{\sqrt{2M}}{N} \;\;. \label{LegEq014} \end{equation} In Appendix A, we show that the probability distribution $\hat{H}$ belonging to this generating function, which is the inverse Laplace transform, is given by \begin{equation} \hat{H}(\tau) \;=\; \sum_{n\in{\bf N}} \delta\left(\tau-\left[n\lambda-\frac{1}{\lambda}\right]\right) \frac{1}{n!}\left(\frac{1}{\lambda^2}\right)^n \exp\left(-\frac{1}{\lambda^2}\right) \;\;. \label{LegEq020} \end{equation} It consists of an infinite number of Dirac delta-distributions, weighed with a Poisson distribution. The delta-distributions reveal the fact that, for finite $N$ and $M$, the Lego discrepancy, and also the $\chi^2$-statistic, can only take a finite number of values, so that the probability density {\em should} consist of a sum of delta-distributions. In the usual limit of $N\rightarrow\infty$, the discrete nature of the random variable disappears, and the $\chi^2$-distribution is obtained. In our limit, however, the discrete nature does not yet disappear. A continuous distribution is obtained if $\lambda\ra0$, which corresponds with going over from $\alpha={\textstyle\frac{1}{2}}$ to $\alpha>{\textstyle\frac{1}{2}}$. Then $\hat{G}(\xi)\rightarrow\exp({\textstyle\frac{1}{2}}\xi^2)$. \section{Derivation of the various limits} We will treat the cases $\gamma=2$ and $\gamma<2$ independently. \subsection{$\bs{\gamma=2}$} We distinguish the three cases $0<\alpha<1$, $\alpha=1$ and $\alpha>1$. If $\alpha>1$, then $V\asymp M$, and the contribution $C_A$ of a diagram $A$ satisfies $C_A\asymp M^\beta$, with \begin{equation} \beta \;=\; ({\textstyle\frac{1}{2}} -\alpha)L_B - {\textstyle\frac{1}{2}} L_F + (\alpha + {\textstyle\frac{1}{2}})(1-L_M) \;\;. \end{equation} A short analysis shows that only diagrams with $(L_B,L_F,L_M)=(1,1,0)$ or $(L_B,L_F,L_M)=(1,2,0)$ give a non-vanishing contribution, and those diagrams are \begin{align} \frac{1}{2}\,\diagram{dcZ1}{40}{8} \;&=\; \frac{1}{2}\,\frac{N(M-1)2\xi}{N\sqrt{V}} \;=\; \frac{E\xi}{\sqrt{V}}\label{LegEq009}\\ \frac{1}{4}\,\diagram{dcZ2}{60}{8} \;&=\; \frac{1}{4}\frac{N^2(M-1)4\xi^2}{N^2V} \;=\; \frac{\xi^2}{2} + {\mathcal O}(M^{-1}) \label{LegEq010} \;\;. \end{align} The first diagram gives a contribution that is linear in $\xi$ and cancels with the exponent in \eqn{LegEq008}. This has to happen for every value of $\alpha$, and as we will see, this diagram will occur always. Notice that the diagrams above are the first two diagrams in the series on the l.h.s of \eqn{LegEq007}. The logarithm of the generating function becomes quadratic, so that the probability distribution becomes Gaussian. If $\alpha=1$, then again $V\asymp M$, so that $\beta=-{\textstyle\frac{1}{2}}(L_B+L_F)+\sfrac{3}{2}(1-L_M)$, and we have to add the diagrams with $(L_B,L_F,L_M)=(2,1,0)$: \begin{equation} \frac{1}{8}\,\diagram{lcF1B2b}{40}{14} \;+\; \frac{1}{8}\,\diagram{lcF1B2}{60}{10} \;=\; \frac{1}{8}\,\frac{N(M_2-M^2)4\xi^2}{N^2V} \;=\; \frac{(M_2-M^2)\xi^2}{2NV} \;\;. \label{LegEq011} \end{equation} If $0<\alpha<1$, then $V\asymp M^{2-\alpha}$ and $\beta=-\sfrac{\alpha}{2}L_B-(1-\sfrac{\alpha}{2})L_F-(\sfrac{\alpha}{2}+1)L_M+\sfrac{\alpha}{2}+1$, so that, besides the diagram of \eqn{LegEq009}, only the diagrams of \eqn{LegEq011} give a non-vanishing contribution, and this contribution is equal to $\xi^2/2$. \subsection{$\bs{0\leq\gamma<2}$} We can distinguish the two cases $\gamma-\alpha\leq1$ and $\gamma-\alpha>1$. \subsubsection{$\bs{\gamma-\alpha\leq1}$} In this case, $V\asymp M$, and the contribution $C_A$ of a diagram $A$ satisfies $C_A\asymp M^\beta$ with \begin{equation} \beta \;=\; ({\textstyle\frac{1}{2}} -\alpha)L_B - {\textstyle\frac{1}{2}} L_F + (\alpha + {\textstyle\frac{1}{2}})(1-L_M) \;\;. \end{equation} If $\alpha<{\textstyle\frac{1}{2}}$, then $\beta$ increases with the number of boson loops $L_B$, and we are not able to calculate the limit of $M\rightarrow\infty$. If $\alpha>{\textstyle\frac{1}{2}}$, then the only diagrams that have a non-vanishing contribution are those with $(L_B,L_F,L_M)=(1,1,0)$, $(1,2,0)$ or $(2,1,0)$. These are exactly the diagrams of \eqn{LegEq009}, \eqn{LegEq010} and \eqn{LegEq011}. Notice, however, that the diagrams of \eqn{LegEq011} cancel if $\gamma-\alpha<0$: then they are {\em irrelevant}. The resulting asymptotic distribution is Gaussian again. If $\alpha={\textstyle\frac{1}{2}}$, then $L_B$ disappears from the equation for $\beta$, and we obtain a non-Gaussian asymptotic distribution. The diagrams that contribute are those with $(L_F,L_M)=(1,0)$ or $(2,0)$. There is, however, only one {\em relevant} diagram with $(L_F,L_M)=(1,0)$, namely the diagram of \eqn{LegEq009} that gives the linear term. We have to be careful here, because the other diagrams with $(L_F,L_M)=(1,0)$ still might be non-vanishing. A short analysis shows that they are given by the sum of all ways to put daisy diagrams to one fermion loop, and that their contribution is given by \begin{equation} C_1(M) \;=\; N\log\left(1+\sum_{p=1}^\infty\frac{({\textstyle\frac{1}{2}} g^2)^pd_p(M)}{p!}\right) \;\;. \end{equation} We know that, if $h_p=1$, then $d_p(M)=M^p[1+\varepsilon_p(M)]$ with $\lim_{M\rightarrow\infty}\varepsilon_p(M)=0$, so that \begin{equation} C_1(M) \;=\; {\textstyle\frac{1}{2}} NMg^2 + N\log\left(1 + e^{-{\scriptstyle\frac{1}{2}} Mg^2} \sum_{p=1}^\infty\frac{({\textstyle\frac{1}{2}} Mg^2)^p\varepsilon_p(M)}{p!}\right) \;\;. \label{LegEq004} \end{equation} The first term gives the leading contribution; the contribution of the relevant diagram, which consists of a boson loop and a fermion loop attached to one vertex. The second term is irrelevant with respect to the first, but can still be non-vanishing, depending on the behavior of $\varepsilon_p(M)$. Remember that $\alpha={\textstyle\frac{1}{2}}$ and $V\asymp M$, so that $Mg^2=2\xi M/(N\sqrt{V}\,)\rightarrow\textsl{constant}$, and we can see that the contribution is only vanishing if \begin{equation} \lim_{M\rightarrow\infty}N\varepsilon_p(M) \;=\;0 \quad\forall\,p\geq1 \quad\Longleftrightarrow\quad M_p-M^p\prec M^{p-\frac{1}{2}} \quad\forall\,p\geq1 \;\;. \end{equation} For $p=1$ this relation is satisfied because $\varepsilon_1(M)=0$. For $p=2$ this relation is also satisfied if $\gamma<\frac{3}{2}$. If the relation is also satisfied for the other values of $p$, then the only diagrams that contribute to the generating function are the relevant diagrams with $(L_F,L_M)=(2,0)$: \begin{equation} \diagram{lcBB2}{40}{15} \;+\; \diagram{lcBB3}{40}{15} \;+\; \diagram{lcBB4}{40}{15} \;+\; \cdots \;\;, \label{LegEq012} \end{equation} where we used the effective vertex (\ref{LegEq006}) again. The contribution of a diagram of this type with $p$ boson lines is given by \begin{equation} \frac{1}{2\,p!}\left(\frac{2\xi}{N\sqrt{V}}\right)^p N^2M_p[1+{\mathcal O}(M^{-1})] \;\sim\;\frac{N^2}{2M}\,\frac{1}{p!} \left(\frac{2M\xi}{N\sqrt{V}}\right)^p. \label{LegEq013} \end{equation} The factor $1/2\,p!$ is the symmetry factor of this type of diagram. If we sum the contribution of these diagrams and use that $V\sim2M$, we obtain \begin{equation} \log\hat{G}(\xi) \;\sim\; \frac{1}{\lambda^2}\left(e^{\lambda\xi}-1-\lambda\xi\right) \;\;,\quad \lambda=\lim_{M\rightarrow\infty}\frac{\sqrt{2M}}{N} \;\;. \end{equation} \subsubsection{$\bs{\gamma-\alpha>1}$} In this case, $V\asymp M^{\gamma-\alpha}$ and the contribution $C_A$ of a diagram $A$ satisfies $C_A\asymp M^\beta$ with \begin{equation} \beta \;=\; (1-\sfrac{\gamma+\alpha}{2})L_B - \sfrac{\gamma-\alpha}{2}\,L_F + \sfrac{\gamma+\alpha}{2}(1-L_M) \;\;. \end{equation} If $\gamma+\alpha<2$, then $\beta$ increases with the number of boson loops $L_B$, and we are not able to calculate the limit of $M\rightarrow\infty$. If $\gamma+\alpha>2$, then the only diagrams that have a non-vanishing contribution are those with $(L_B,L_F,L_M)=(1,1,0)$, $(1,2,0)$ or $(2,1,0)$. These are exactly the diagrams of \eqn{LegEq009}, \eqn{LegEq010} and \eqn{LegEq011}. Notice, however, that the diagrams of \eqn{LegEq011} cancel if $\gamma-\alpha<0$: then they are {\em irrelevant}. The resulting asymptotic distribution is Gaussian. If $\gamma+\alpha=2$, then $\beta=(\alpha-1)L_F+1-L_M$. Because $\gamma-\alpha>1$, we have $\alpha<{\textstyle\frac{1}{2}}$, and non-vanishing diagrams have $(L_F,L_M)=(1,0)$. Their contribution is given by the r.h.s. of \eqn{LegEq004}, the first term of which gives the term linear in $\xi$. The second term is non-vanishing, because $Mg^2\asymp M^{1-(\gamma+\alpha)/2}\rightarrow\textsl{constant}$ and $N\varepsilon_2(M)\asymp M^{\alpha+\gamma-2}\rightarrow\textsl{constant}$. \section{Conclusions} We have shown that the Lego discrepancy with $M$ bins is equivalent to a $\chi^2$-statistic with $M$ bins. We have presented a procedure to calculate the moment generating function of the probability distribution of the discrepancy perturbatively if $M$ and $N$, the number of uniformly and randomly distributed data points, become large. The natural expansion parameter we have identified to be $M/N$, and we have calculated the first few terms in the series explicitly. In order to calculate limits in which $N,M\rightarrow\infty$, we have introduced the objects of \eqn{LegEq022} and restricted the behavior of the size of the bins such that they satisfy \eqn{LegEq018}. Furthermore, we have gone over to the standardized variable of the discrepancy. For this variable, we have derived a phase diagram, representing the limits specified by \eqn{LegEq023} and \eqn{LegEq024}. We have formulated the results in (\ref{LegRes01}), (\ref{LegRes02}) and (\ref{LegRes03}). On of these results is that there are non-trivial limits if $N,M\rightarrow\infty$ such that $M^\alpha/N\rightarrow\textsl{constant}$ with $\alpha<1$. This result is in stark contrast with the rule of thumb that, in order to trust the $\chi^2$-distribution, each bin has to contain at least a few data points. \section*{Appendix A} We want to calculate the integral \begin{equation} \hat{H}(\tau) \;=\; \frac{1}{2\pi i}\int\limits_{-i\infty}^{i\infty}e^{f_{\tau}(z)}\,dz \quad,\quad f_{\tau}(z) =\frac{1}{\lambda^2}\left(e^{\lambda z}-1-\lambda z\right) - z\tau \;\;. \end{equation} We will make use of the fact that \begin{equation} f_{\tau}\left(z+\frac{2\pi in}{\lambda}\right) \;=\; f_{\tau}(z) - 2\pi in\frac{1+\lambda\tau}{\lambda^2} \end{equation} for all $n\in{\bf Z}$, so that \begin{equation} \hat{H}(\tau) \;=\; \frac{1}{2\pi i}\sum_{n\in{\bf Z}}\, \int\limits^{\frac{(2n+1)\pi i}{\lambda}} _{\frac{(2n-1)\pi i}{\lambda}} e^{f_{\tau}(z)}\,dz \;=\; \frac{1}{2\pi i}\sum_{n\in{\bf Z}} e^{-2\pi in\frac{1+\lambda\tau}{\lambda^2}} \int\limits^{\frac{\pi i}{\lambda}}_{\frac{-\pi i}{\lambda}} e^{f_{\tau}(z)}\,dz \;\;. \label{CorEq019} \end{equation} Notice that the integral is independent of $n$, so that the sum can be interpreted as a sum of Dirac delta-distributions: \begin{equation} \sum_{n\in{\bf Z}}e^{-2\pi in\frac{1+\lambda\tau}{\lambda^2}} \;=\; \sum_{n\in{\bf Z}} \delta\left(\frac{1+\lambda\tau}{\lambda^2} - n\right) \;=\; \sum_{n\in{\bf Z}} \lambda\delta\left(\tau-\left[n\lambda -\frac{1}{\lambda}\right]\right) \;\;. \end{equation} These delta-distributions restrict the values that $\tau$ can take. If we use these restrictions and do the appropriate variable substitutions, the remaining integral in (\ref{CorEq019}) can be reduced to \begin{equation} \int\limits^{\frac{\pi i}{\lambda}}_{\frac{-\pi i}{\lambda}} e^{f_{\tau}(z)}\,dz \;=\; \frac{e^{-\frac{1}{\lambda^2}}}{\lambda} \int\limits^{\pi i}_{-\pi i}\exp\left(\frac{e^\varphi}{\lambda^2} -n\varphi\right)\,d\varphi \;=\; \frac{e^{-\frac{1}{\lambda^2}}}{\lambda} \oint \frac{e^{\frac{1}{\lambda^2}w}}{w^{n+1}}\,dw \;\;, \end{equation} where $n\in{\bf Z}$ and the contour is closed around $w=0$. According to Cauchy's theorem, the final integral is only non-zero if $n\in{\bf N}$, and in that case its value is $2\pi i\frac{1}{n!}(\frac{1}{\lambda^2})^n$. The combination of these results gives \eqn{LegEq020}.
\section{Introduction} Nuclear magnetic resonance (NMR), which is one of most powerful tools for investigating various physical properties, has a number of peculiarities for magnetically ordered materials. Last time, a number of new classes of magnets have been studied by this method, e.g., heavy-fermion compounds \cite {hf}, ferromagnetic films and monolayers \cite{films}, low-dimensional systems including copper-oxide perovskites \cite{Mehr}, etc. Thus the problem of theoretical description of various NMR characteristics of magnets is topical again. This problem was already a subject of great interest since 50-60s when the interaction of nuclear magnetic moments with spin waves in localized-spin Heisenberg model was studied \cite{Mor56,Turov}. However, this model is inadequate to describe the most interesting systems mentioned above where the role of conduction electrons is essential in magnetic properties. Usually the data on the longitudinal nuclear magnetic relaxation rate $1/T_1$ (this NMR characteristic is probably most convenient to compare experimental results with theoretical predictions) are discussed within itinerant-electron models such as Hubbard model or phenomenological spin-fluctuation theories. Ueda and Moriya \cite{UM,Mor} calculated the dependences $1/T_1(T)$ for weak itinerant magnets, especial attention being paid to the paramagnetic region, and obtained strong temperature effects. Later this approach was extended to the two-dimensional case and extensively developed in connection with high-$T_c$ superconductors and related compounds \cite{Mor1,Pines}. On the other hand, in a number of systems (e.g., in most rare-earth compounds which are also a subject of NMR investigations, see, e.g., Refs.\cite{re}) the $s-d(f)$ exchange model with well-separated localized and itinerant magnetic subsystems is more adequate. Magnetic properties in such a situation differ essentially from those in the paramagnon regime (see, e.g., discussion in Refs.\cite{IKJP90,afm}). At the same time, the contributions to nuclear magnetic relaxation rate owing to electron-magnon interaction are not investigated in detail. In the present work we obtain the dependences of $1/T_1(T)$ and the linewidth $1/T_2(T)$ in the spin-wave region for three- and two-dimensional (% $3D$ and $2D$) metallic magnets with well-defined local magnetic moments. In Sect.2 we discuss the general formalism and physical picture of hyperfine interactions. In Sects.3 and 4 we calculate various contributions to the relaxation rates in metallic ferro- and antiferromagnets. In Sect.5 we analyze the isotropic $2D$ case where at finite temperatures the long-range order is absent, but the correlation length is very large. \section{Hyperfine interactions} We start from the standard Hamiltonian of the hyperfine interaction \cite {abragam} \begin{equation} H_{hf}={\bf hI,\,}h_\alpha =A_{\alpha \beta }S_\beta \end{equation} $\widehat{A}$ being the hyperfine interaction matrix, which contains the Fermi (contact) and dipole-dipole contributions, \begin{equation} A_{\alpha \beta }=A^F\delta _{\alpha \beta }+A_{\alpha \beta }^{dip}. \end{equation} The Fermi hyperfine interaction is proportional to the electron density at the nucleus and therefore only $s$-states participate in it, the contribution of core $s$-states (which are polarized due to local magnetic moments) being much larger than of conduction electrons. It is just the consequence of considerably smaller localization area (and therefore higher density on nuclei) for the core states. The dipole contribution to $H_{hf}$ can be represented as \cite{abragam} \begin{equation} H_{hf}^{dip}=\frac a2\{[\frac 16(I^{+}S^{-}+I^{-}S^{+})-\frac 13% I^zS^z]F^{(0)}+I^{+}S^{+}F^{(2)}+2(I^zS^{+}+I^{+}S^z)F^{(1)}\}+h.c\text{.} \label{hdip} \end{equation} where \begin{eqnarray} F^{(0)} &=&\langle (1-3\cos ^2\theta )/r^3\rangle ,F^{(1)}=\langle \sin \theta \cos \theta \exp (-i\phi )/r^3\rangle , \nonumber \\ F^{(2)} &=&\langle \sin ^2\theta \exp (-2i\phi )/r^3\rangle ,a=-\frac 32% \gamma _e\gamma _n \end{eqnarray} Here $\langle ...\rangle $ is the average over the electron subsystem states, $\gamma _e$ and $\gamma _n$ are gyromagnetic ratios for electron and nuclear moments, respectively. In the case of the {\it local} cubic symmetry we have $F^{(a)}=0.$ It is obvious that magnetic $f$- or $d$-electrons dominate also in dipole interactions because of large spin polarization. Hence the direct interaction of nuclear spins with that of conduction electrons can be neglected in magnets with well-defined local magnetic moments. Nevertheless, conduction electrons do effect nuclear relaxation via their influence on the local-moment system; besides that, as we shall see below, such contributions possess large exchange enhancement factors. The investigation of these effects is one of the main aims of this work. A general way to consider all these contributions is using the Green's function method which leads to the following expression for the longitudinal nuclear magnetic relaxation rate \cite{Mor63} \begin{eqnarray} \frac 1{T_1} &=&-\frac T{2\pi }\text{Im}\sum_{{\bf q}}\langle \langle h_{% {\bf q}}^{+}|h_{-{\bf q}}^{-}\rangle \rangle _{\omega _n}/\omega _n, \label{tt1} \\ \frac 1{T_2} &=&\frac 1{2T_1}-\frac T{2\pi }\lim_{\omega \rightarrow 0}\text{% Im}\sum_{{\bf q}}\langle \langle h_{{\bf q}}^z|h_{-{\bf q}}^z\rangle \rangle _\omega /\omega \label{tt2} \end{eqnarray} ($\omega _n=\langle h^z\rangle \ll T$ is the NMR frequency). As follows from (\ref{hdip}), \begin{eqnarray} h^{-} &=&(A^F+\frac 13aF^{(0)})S^{-}+aF^{(2)}S^{+}+2aF^{(1)}S^z, \\ h^z &=&(A^F-\frac 23aF^{(0)})S^z+a(F^{(1)}S^{+}+aF^{(1)*}S^{-}) \end{eqnarray} Then we derive \begin{eqnarray} \frac 1{T_1} &=&\frac T2\{[(A^F+\frac 13aF^{(0)})^2+a^2|F^{(2)}|^2]K^{+-} \nonumber \\ &&\ \ \ \ \ \ \ +2a(A^F+\frac 13aF^{(0)})\text{Re}% F^{(2)}K^{++}+4a^2|F^{(1)}|^2K^{zz}\} \label{FK} \end{eqnarray} \begin{equation} \frac 1{T_2}=\frac 1{2T_1}+\frac T2\{(A^F-\frac 23% aF^{(0)})^2K^{zz}+a^2[2|F^{(1)}|^2K^{+-}+(F^{(1)})^2K^{++}+(F^{(1)*})^2K^{--}]\} \label{FK1} \end{equation} where the quantities $K^{\alpha \beta }$ are defined by \begin{equation} K^{\alpha \beta }=-\frac 1\pi \lim_{\omega \rightarrow 0}\text{Im}\sum_{{\bf % q}}\langle \langle S_{{\bf q}}^\alpha |S_{-{\bf q}}^\beta \rangle \rangle _\omega /\omega \end{equation} As we shall see below in Sect.5, $\omega _n\neq 0$ which enters (\ref{FK}) (but not the second term of (\ref{FK1})) may become important in the case of very small magnetic anisotropy. Formula (\ref{tt1}) has a rather general character. On the other hand, the problem of calculating the NMR linewidth is much more complicated. The Moriya formula (\ref{tt2}) is in fact applicable only in the case where the line has the Lorentz form (i.e., charactersitic frequency of hyperfine field fluctuations is large in comparison with their amplitude) \cite{abragam}. In insulating crystals the latter condition is usually violated, the lineform being close to Gaussian with the width determined by dipole interactions of nuclear spins. At the same time, in metals the Korringa relaxation described by the formula (\ref{tt2}) usually dominates, so that we will use this. A peculiar case is provided by conducting systems which are on the borderline of ferro- or antiferromagnetic instability (i.e. with large correlation length $\xi $), e.g., copper-oxide superconductors \cite{thelen}. Under this condition, the anisotropic Ruderman-Kittel interaction between nuclear spins turns out to be greatly ehanced and dominates over the dipole interaction. The lineform turns out to be Gaussian with the width being estimated as \begin{equation} \left( \frac 1{T_2}\right) ^2\propto A^2\sum_{{\bf q}}\chi ^2({\bf q},\omega =0) \label{gauss} \end{equation} where $\chi ({\bf q},\omega )$ is the dynamical spin susceptibility of electron system. We will use this result in Sect.5 when discussing 2D systems which do possess large correlation length. \section{Ferromagnetic metals} We proceed with the $s-d(f)$ exchange model Hamiltonian \begin{equation} \ H=\sum_{{\bf k}\sigma }t_{{\bf k}}c_{{\bf k}\sigma }^{\dagger }c_{{\bf k}% \sigma }-I\sum_{i\alpha \beta }{\bf S}_i\mbox {\boldmath $\sigma $}_{\alpha \beta }c_{i\alpha }^{\dagger }c_{i\beta }+\sum_{{\bf q}}J_{{\bf q}}{\bf S}_{% {\bf -q}}{\bf S}_{{\bf q}}+H_a \label{H} \end{equation} where $t_{{\bf k}}$ is the band energy, ${\bf S}_i$ and ${\bf S}_{{\bf q}}$ are spin-density operators and their Fourier transforms, $\sigma $ are the Pauli matrices, $H_a$ is the anisotropy Hamiltonian which results in occurrence of the gap $\omega _0$ in the spin-wave spectrum. For convenience we include explicitly in the Hamiltonian the Heisenberg exchange interaction with the parameters $J_{{\bf q}},$ although really this may be, e.g., the Ruderman-Kittel-Kasuya-Yosida (RKKY) interaction. It should be noted that similar results may be reproduced for the localized-moment Hubbard magnets (cf.\cite{IKJP90,Ent}). First we consider the ferromagnetic (FM) case. Then $K^{++}=0$ and the relaxation rates (\ref{FK}), (\ref{FK1}) are the sums of transverse $% (\propto K^{+-})$ and longitudinal $(\propto K^{zz})$ terms. Passing to the magnon representation we obtain \begin{equation} \langle \langle S_{{\bf q}}^{+}|S_{-{\bf q}}^{-}\rangle \rangle _\omega =2S/[\omega -\omega _{{\bf q}}+i\gamma _{{\bf q}}(\omega )] \label{fgf} \end{equation} where $\omega _{{\bf q}}=2S(J_{{\bf q}}-J_0)+\omega _0$ is the magnon frequency, $\gamma _{{\bf q}}(\omega )\propto \omega $ is the magnon damping. Then we have \begin{equation} K^{+-}=2S\sum_{{\bf q}}\frac{\gamma _{{\bf q}}(\omega _n)}{\pi \omega _n\omega _{{\bf q}}^2} \label{gfm} \end{equation} (cf. Refs.\cite{IKJP90,UFN}). The damping in the denominator of (\ref{gfm}) can be neglected for both localized-moment and itinerant-electron magnets (in the latter case the expression (\ref{fgf}) corresponds to the RPA structure, see Ref.\cite{IKJP90}) due to smallness of $\omega _n.$ On the contrary, temperature dependences of magnetization, resistivity etc. in weak itinerant magnets are just determined by the damping in the denominator, i.e. by paramagnon excitations rather than by spin waves \cite{Mor}. The damping owing to the one-magnon decay processes is given by the well-known expression \begin{eqnarray} \gamma _{{\bf q}}^{(1)}(\omega ) &=&-2\pi I^2S\sum_{{\bf k}}(n_{{\bf % k\uparrow }}-n_{{\bf k-q\downarrow }}) \label{G1F} \\ \ \times \delta (\omega +t_{{\bf k\uparrow }}-t_{{\bf k-q\downarrow }}) &\simeq &2\pi I^2S\omega \lambda _{{\bf q}} \nonumber \end{eqnarray} where $t_{{\bf k}\sigma }=$ $t_{{\bf k}}-\sigma IS,\,t_{{\bf k}}$ is referred to the Fermi level, $n_{{\bf k}\sigma }=n(t_{{\bf k}\sigma })$ is the Fermi function, \begin{equation} \lambda _{{\bf q}}=\sum_{{\bf k}}\delta (t_{{\bf k\uparrow }})\delta (t_{% {\bf k-q\downarrow }}). \end{equation} The linearity of spin fluctuation damping in $\omega $ is the characteristic property of metals. According to (\ref{FK}) this leads to $T$-linear contributions to $1/T_1$ which is the Korringa law \cite{Kor}. Note that the simplest expression for the Korringa relaxation \begin{equation} 1/T_1\simeq 1/T_2\simeq A^2\rho _{\uparrow }\rho _{\downarrow }T, \label{kor} \end{equation} where $A$ is an effective hyperfine interaction constant, $\rho _\sigma $ are the partial densities of electron states at the Fermi level, is practically never applicable for magnetic metals (e.g., exchange enhancement factors can change even the order of magnitude of $1/T_1$ \cite{Mor,UFN}). Accurate expression for the ``Korringa'' contribution in the case under consideration can be derived by the substitution (\ref{gfm}) and (\ref{G1F}) into (\ref{FK}). The damping (\ref{G1F}) has the threshold value of $q,$ which is determined by the spin splitting $\Delta =2|I|S$, $q^{*}=\Delta /v_F$ ($v_F$ is the electron velocity at the Fermi level). The quantity $q^{*}$ determines a characteristic temperature and energy scale \begin{equation} \omega ^{*}=\omega (q^{*})={\cal D}(q^{*})^2\sim (\Delta /v_F)^2T_C \label{T*} \end{equation} with ${\cal D}$ the spin-wave stiffness. Besides $3D$ magnets, consideration of the $2D$ case is of interest (this may be relevant, e.g., for layered magnets and ferromagnetic films; for more details see Sect.5). We have \begin{equation} \lambda _{{\bf q}}=\theta (q-q^{*})\times \left\{ \begin{tabular}{ll} $(qv_F)^{-1},$ & $D=3$ \\ $\frac 1\pi (q^2v_F^2-\Delta ^2)^{-1/2},$ & $D=2$% \end{tabular} \right. \end{equation} After integration for the parabolic electron spectrum ($q^{*}$ plays the role of the lower cutoff), the one-magnon damping contribution to (\ref{gfm}% ) takes the form \begin{equation} \delta ^{(1)}K^{+-}=\frac{\rho _{\uparrow }\rho _{\downarrow }}{{\cal D}^2m^2% }\times \left\{ \begin{tabular}{ll} $1/4,$ & $D=3$ \\ $1/(\pi q^{*}),$ & $D=2$% \end{tabular} \right. \label{t1f} \end{equation} with \begin{equation} \rho _\sigma =\frac{m\Omega _0}{2\pi }\times \left\{ \begin{tabular}{ll} $k_{F\sigma }/\pi ,$ & $D=3$ \\ $1,$ & $D=2$% \end{tabular} \right. \end{equation} $m$ the electron effective mass, $\Omega _0$ the lattice cell volume (area). Thus in the $3D$ case the factor of $I^2$ is canceled, and the factor of $% I^{-1}$ occurs in the $2D$ case and we obtain a strongly enhanced $T$-linear Korringa-type term (remember that ${\cal D}\sim J\sim I^2\rho $ for the RKKY interaction). This means that the contribution of conduction electrons to $T$% -linear relaxation rate via their interaction with localized spins is indeed much more important than the ``direct'' contribution (\ref{kor}): perturbation theory in the $s-d$ exchange coupling parameter $I$ turns out to be singular. Earlier such contributions (for the $3D$ case) were calculated by Weger \cite{Weg} and Moriya \cite{Mor64} for iron-group metals. However, Moriya has concluded that for these materials they are not important in comparison with orbital current contributions. In the case under consideration (where magnetic subsystem is well separated from the conductivity electrons) the situation is different and the spin-wave contribution in $1/T_1$ is normally the most important. The one-magnon decay contribution (\ref{t1f}) is absent for so-called half-metallic ferromagnets, e.g., some Heusler alloys, where electron states with one spin projection only are presented at the Fermi surface \cite {HFM,UFN}. In such a situation we have to consider two-magnon scattering processes. In this connection, it is worthwhile to note an important difference between relaxation processes via phonons and via magnons. The main difference is due to the gap in magnon spectrum. Usually $\omega _0>\omega _n$ and therefore one-magnon processes contribute to the relaxation rate due to magnon damping only (cf. discussion of the phonon-induced relaxation processes in Ref.\cite{abragam}). However, the mechanisms of magnon damping in magnetic dielectrics (magnon-magnon interactions) are different from those in magnetic metals and degenerate semiconductors \cite{Aus,afm}. The damping in a conducting ferromagnet owing to electron-magnon (two-magnon) scattering processes is calculated in Refs.\cite{Aus,IKJP90} and has the form \begin{eqnarray} \frac{\gamma _{{\bf q}}^{(2)}(\omega )}\omega &=&\pi I^2\sum_{{\bf kp}\sigma }\left( \frac{t_{{\bf k+q}}-t_{{\bf k}}}{t_{{\bf k+q}}-t_{{\bf k}}+2\sigma IS% }\right) ^2(\omega _{{\bf p}}-\omega )\frac{\partial n_{{\bf k}\sigma }}{% \partial t_{{\bf k}}}\frac{\partial N_{{\bf p}}}{\partial \omega _{{\bf p}}} \nonumber \\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \times \delta (t_{{\bf k}}-t_{{\bf k-p+q}}) \label{Gam2gen} \end{eqnarray} where $N_{{\bf p}}=N(\omega _{{\bf p}})$ is the Bose function. Substituting this into (\ref{gfm}) and performing integration we obtain for $D=3$% \begin{equation} \delta ^{(2)}K^{+-}=\frac{\Omega _0T^{1/2}}{128\pi ^2Sm^2{\cal D}^{7/2}}% \sum_\sigma \rho _\sigma ^2\times \left\{ \begin{tabular}{ll} $3\pi ^{1/2}\zeta (\frac 32)T,$ & $T\ll \omega ^{*}$ \\ $8M_3\omega ^{*},$ & $T\gg \omega ^{*}$% \end{tabular} \right. \label{f2} \end{equation} where $\zeta (z)$ is the Riemann function, \begin{equation} M_3=\int_0^\infty dx\left[ \frac 1{x^2}-\frac{x^2\exp x^2}{(\exp x^2-1)^2}% \right] \simeq 0.65 \end{equation} The contribution (\ref{f2}) should play the dominant role in the half-metallic ferromagnets \cite{UFN}. Besides that, this contribution may modify considerably the temperature dependence of $1/T_1$ in ``usual'' ferromagnets, a crossover from $T^{5/2}$ to $T^{3/2}$ dependence of the correction taking place. For $D=2$ at $T,\omega ^{*}\gg \omega _0$ small magnon momenta of order of $% \left( \omega _0/{\cal D}\right) ^{1/2}$ make the main contribution to (\ref {gfm}). To calculate the integral one can use the high-temperature expression for $N_{{\bf p}}=T/\omega _{{\bf p}}$. As a result, one gets \begin{equation} \delta ^{(2)}K^{+-}=\frac{\Omega _0^3k_FM_2}{8\pi ^4S{\cal D}^{5/2}\omega _0^{1/2}}T \label{f22} \end{equation} with \begin{eqnarray} M_2 &=&\int_0^\infty \frac{dx}{1+x^2}\int\limits_0^{\pi /2}\frac{d\varphi \sin ^2\varphi }{\left( \sin ^2\varphi +x^2\right) ^{3/2}} \nonumber \\ \ &=&\int_0^\infty dy\left[ 1+\frac{y^2}{\sqrt{1+y^2}}\ln \frac{\sqrt{1+y^2}% -1}y\right] \simeq 1.23 \end{eqnarray} Thus in the $2D$ FM case, in contrast with $3D$ one, the relaxation rate $% 1/T_1$ is strongly dependent on the anisotropy gap. Consider now the second term in the tranverse relaxation rate $1/T_2(T)$ (% \ref{FK1}), which is normally determined by $K^{zz},$ and the longitudinal contribution to relaxation rate $1/T_1$ in (\ref{FK}), which is due to dipole-dipole interactions with the characteristic constant $\widetilde{A}% \sim a|F^{(1)}|$. The simplest calculation from the longitudinal Green's function for the localized-spin subsystem gives \begin{eqnarray} \langle \langle S_{{\bf q}}^z|S_{-{\bf q}}^z\rangle \rangle _\omega &=&\sum_{% {\bf p}}\frac{N_{{\bf p}}-N_{{\bf p-q}}}{\omega -\omega _{{\bf p-q}}+\omega _{{\bf p}}}, \\ K^{zz} &=&\sum_{{\bf qp}}\left( -\frac{\partial N_{{\bf p}}}{\partial \omega _{{\bf p}}}\right) \delta (\omega _{{\bf q}}-\omega _{{\bf p}}). \label{kzz} \end{eqnarray} The quantity (\ref{kzz}) has been considered in Refs.\cite{Rob,Turov} as a contribution to the NMR line width $1/T_2.$ The integration in the $3D$ case gives the logarithmic singularity \begin{equation} K^{zz}=\frac{\Omega _0^2}{16\pi ^4{\cal D}^3}T\ln \frac T{\omega _0} \label{zz3d} \end{equation} For $D=2$ this singular term is inversely proportional to the magnetic anisotropy parameter and very large: \begin{equation} K^{zz}=\left( \frac{\Omega _0}{4\pi {\cal D}}\right) ^2N(\omega _0)\simeq \left( \frac{\Omega _0}{4\pi {\cal D}}\right) ^2\frac T{\omega _0},T\gg \omega _0 \label{zz2d} \end{equation} For small enough $\omega _0$ and $\widetilde{A}\sim A$ this contribution can dominate over the ``Korringa'' contribution (\ref{t1f}) in $1/T_1$ at $% T>\omega _0/|I\rho |$. The leading contribution to $K^{zz}$ from the $s-d$ interaction is determined by \begin{eqnarray} \delta \langle \langle S_{{\bf q}}^z|S_{-{\bf q}}^z\rangle \rangle _\omega &=&2I^2S\sum_{{\bf kp}\sigma }\frac 1{(\sigma \omega +\omega _{{\bf p-q}% }-\omega _{{\bf q}})^2} \nonumber \\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \times \frac{n_{{\bf % k\downarrow }}(1-n_{{\bf k+p-q\uparrow }})+N_{{\bf p}}(n_{{\bf k\downarrow }% }-n_{{\bf k+p-q\uparrow }})}{t_{{\bf k\downarrow }}-t_{{\bf k+p-q\uparrow }% }+\sigma \omega -\omega _{{\bf p}}} \end{eqnarray} However, it is not singular in $\omega _0$ and practically never important. \section{Antiferromagnetic metals} Now we consider the spiral antiferromagnetic (AFM) structure along the $x$% -axis with the wavevector {\bf Q \ } \[ \langle S_i^z\rangle =S\cos {\bf QR}_i,\,\langle \,S_i^y\rangle =S\sin {\bf % QR}_i,\,\langle S_i^x\rangle =0 \] We introduce the local coordinate system \begin{eqnarray*} S_i^z &=&\hat S_i^z\cos {\bf QR}_i-\hat S_i^y\sin {\bf QR}_i, \\ \,S_i^y &=&\hat S_i^y\cos {\bf QR}_i+\hat S_i^z\sin {\bf QR}_i,\,S_i^x=\hat S% _i^x \end{eqnarray*} Further we pass from spin operators ${\bf \hat S}_i$ to the spin deviation operators $b_i^{\dagger },b_i$ and, by the canonical transformation $b_{{\bf % q}}^{\dagger }=u_{{\bf q}}\beta _{{\bf q}}^{\dagger }-v_{{\bf q}}\beta _{-% {\bf q}},$ to the magnon operators. Hereafter we consider for simplicity two-sublattice AFM ordering (2{\bf Q }is equal to a reciprocal lattice vector, so that $\cos ^2{\bf QR}_i=1,\,\sin ^2{\bf QR}_i=0$). Calculating the Green's functions to second order in $I$ (to second order in the formal quasiclassical parameter $1/2S,$ cf. Refs.\cite{afm,kondo}) we derive \begin{eqnarray} \langle \langle b_{{\bf q}}|b_{{\bf q}}^{\dagger }\rangle \rangle _\omega &=&% \frac{\omega +C_{{\bf q-}\omega }}{(\omega -C_{{\bf q}\omega })(\omega +C_{% {\bf q-}\omega })+D_{{\bf q}\omega }^2} \label{FG} \\ \langle \langle b_{-{\bf q}}^{\dagger }|b_{{\bf q}}^{\dagger }\rangle \rangle _\omega &=&\frac{D_{{\bf q}\omega }}{(\omega -C_{{\bf q}\omega })(\omega +C_{{\bf q-}\omega })+D_{{\bf q}\omega }^2} \label{FGa} \end{eqnarray} with \begin{eqnarray} C_{{\bf q}\omega } &=&S(J_{{\bf Q+q},\omega }^{tot}+J_{{\bf q}\omega }^{tot}-2J_{{\bf Q}0}^{tot})+\sum_{{\bf p}}[C_{{\bf p}}\Phi _{{\bf pq}\omega } \nonumber \\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ -(C_{{\bf p}}-D_{{\bf p}})\Phi _{% {\bf p}00}+\phi _{{\bf pq}\omega }^{+}+\phi _{{\bf pq}\omega }^{-}]+g_{{\bf q% }} \label{CD} \\ D_{{\bf q}\omega } &=&D_{{\bf q-}\omega }=S(J_{{\bf q}\omega }^{tot}-J_{{\bf % Q+q},\omega }^{tot})+\sum_{{\bf p}}D_{{\bf p}}\Phi _{{\bf pq}\omega }+h_{% {\bf q}} \nonumber \end{eqnarray} The $s-d$ exchange contributions of the first order in $1/2S$ correspond to the RKKY approximation \begin{equation} J_{{\bf q}\omega }^{tot}=J_{{\bf q}}+I^2\sum_{{\bf k}}\frac{n_{{\bf k}}-n_{% {\bf k-q}}}{\omega +t_{{\bf k}}-t_{{\bf k-q}}} \label{RKKY} \end{equation} ($n_{{\bf k}}=n(t_{{\bf k}})$ is the Fermi function), the second summand in (% \ref{RKKY}) being the $\omega $-dependent RKKY indirect exchange interaction. The function $\Phi $, which determines the second-order corrections, is given by \begin{eqnarray} \Phi _{{\bf pq}\omega } &=&(\phi _{{\bf pq}\omega }^{+}-\phi _{{\bf pq}% \omega }^{-})/\omega _{{\bf p}}, \label{Fi} \\ \phi _{{\bf pq}\omega }^{\pm } &=&I^2\sum_{{\bf k}}\frac{n_{{\bf k}}(1-n_{% {\bf k+p-q}})+N(\pm \omega _{{\bf p}})(n_{{\bf k}}-n_{{\bf k+p-q}})}{\omega +t_{{\bf k}}-t_{{\bf k+p-q}}\mp \omega _{{\bf p}}} \nonumber \end{eqnarray} where \[ \omega _{{\bf p}}=(C_{{\bf p}}^2-D_{{\bf p}}^2)^{1/2}=[4S^2(J_{{\bf p}}-J_{% {\bf Q}})(J_{{\bf Q+p}}-J_{{\bf Q}})+\omega _0^2]^{1/2} \] is the magnon frequency to zeroth order in $I\,$and $1/2S$. The $\omega $% -independent corrections $g_{{\bf q}},h_{{\bf q}}$ that describe the ``direct'' magnon-magnon interaction are written down in Refs.\cite {afm,kondo}. Now we consider the effects of electron-magnon interaction. The intrasubband one-magnon damping (which is absent in the FM case) is finite at arbitrarily small $q$ \cite{IKFMM}. Similar to the FM case, the contributions of intersubband transitions (which correspond to small $|{\bf q-Q}|$) are cut at the characteristic temperature and energy scale \begin{equation} \omega ^{*}=\omega (q^{*})=cq^{*}\sim (\Delta /v_F)T_N, \label{T**} \end{equation} We have \begin{equation} K^{+-}=-\frac{2S}\pi \lim_{\omega \rightarrow 0}\text{Im}\sum_{{\bf q}% }\omega ^{-1}C_{{\bf q}\omega }/\omega _{{\bf q}}^2, \end{equation} and the term with $K^{++}$ in (\ref{FK}) vanishes due to the property $D_{% {\bf q}\omega }=-D_{{\bf q+Q}\omega }.$ The one-magnon contribution owing to the imaginary part of (\ref{RKKY}) in the 3D case takes after integration the form \begin{equation} \delta ^{(1)}K^{+-}=\frac{S^2\Omega _0}{\pi ^2c^2}\left( P_0\ln \frac{\omega _{\max }}{\omega _0}+P_{{\bf Q}}\ln \frac{\omega _{\max }}{\omega _0^{*}}% \right) . \label{K1} \end{equation} Here $c$ is the magnon velocity defined by $\omega _{{\bf p}}^2=\omega _{% {\bf p+Q}}^2=\omega _0^2+c^2p^2$, \begin{equation} P_{{\bf p}}=I^2\lim_{{\bf q}\rightarrow 0}|{\bf q-p}|\sum_{{\bf k}}\delta (t_{{\bf k}})\delta (t_{{\bf k-q+p}}), \end{equation} (the quantity $P_0$ depends, generally speaking, on the direction of the vector ${\bf q,}$ see Refs.\cite{Mor,Pi,Kal}), the second logarithm in the brackets of (\ref{K1}) contains the cutoff \[ \omega _0^{*}=\sqrt{\omega _0^2+(\omega ^{*})^2} \] The ``enhancement'' factor in (\ref{K1}) is smaller than in the FM case because of the linear dispersion law of magnons, but this contribution still dominates over the ``usual'' Korringa term (\ref{kor}). Besides that, a large logarithmic factor occurs (in the isotropic case, this is cut at $% \omega _n$ only). Note that a similar logarithmic singularity in $1/T_1$ takes place for $3D$ itinerant-electron antiferromagnets \cite{UM}. It is interesting that the intersubband contribution does not lead here to enhancing the singularity, unlike the situation for the magnon damping, magnetic and transport properties\cite{IK95,afm}. Under the ``nesting'' conditions ($t_{{\bf k+Q}}\simeq -t_{{\bf k}}$ in a large part of the Fermi surface) the singularity is not enhanced as well. The singularity becomes stronger in the $2D$ case where integration gives \begin{equation} \delta ^{(1)}K^{+-}=\frac{S^2\Omega _0}{\pi c\omega _0}\left( \frac \pi 2% P_0+P_{{\bf Q}}\arctan \frac{\omega _0}{\omega ^{*}}\right) \label{a12} \end{equation} This fact may be important when treating experimental data on layered AFM metals. The contribution owing to electron-magnon scattering processes is determined by the imaginary part of the function (\ref{Fi}). After a little manipulation we obtain \begin{equation} \delta ^{(2)}K^{+-}\simeq 2SL\sum_{{\bf p}\rightarrow 0,{\bf q}}\frac 1{% q\omega _{{\bf q+p}}^2}\left( -\frac{\partial N_{{\bf p}}}{\partial \omega _{% {\bf p}}}\right) [P_0+P_{{\bf Q}}\widetilde{\phi }(q)] \end{equation} where $L=2S(J_0-J_{{\bf Q}}),\widetilde{\phi }(q<q^{*})=0,\,\widetilde{\phi }% (q\gg q^{*})=1.$ The integration in the $3D$ case yields \begin{equation} \delta ^{(2)}K^{+-}=\frac{SL\Omega _0^2}{8\pi ^4c^4}\left[ P_0f(T,\omega _0)+P_{{\bf Q}}f(T,\omega _0^{*})\right] \end{equation} where \begin{eqnarray} f(T,\omega _0) &=&\int_{\omega _0}^\infty d\omega \omega \left( -\frac{% \partial N(\omega )}{\partial \omega }\right) \ln \frac{\omega _{\max }}% \omega \nonumber \\ \ &\simeq &T\ln \frac T{\omega _0}\left( \ln \frac{\omega _{\max }}{\omega _0% }-\frac 12\ln \frac T{\omega _0}\right) ,T\gg \omega _0 \end{eqnarray} Thus we have $1/T_1\propto T^2\ln T.$ In the $2D$ case we derive \begin{equation} \delta ^{(2)}K^{+-}\simeq T\frac{SL\Omega _0^2}{4\pi ^4c^4}\left( P_0\ln ^2% \frac T{\omega _0}+P_{{\bf Q}}\ln ^2\frac T{\omega _0^{*}}\right) , \label{a22} \end{equation} so that the singularity is not enhanced in comparison with the $3D$ case. The contributions owing to longitudinal fluctuations will be estimated for the localized subsystem only. We obtain \begin{equation} K^{zz}\simeq \sum_{{\bf pq}}\frac{L^2}{2\omega _{{\bf p}}^2}\left( -\frac{% \partial N_{{\bf p}}}{\partial \omega _{{\bf p}}}\right) \delta (\omega _{% {\bf q}}-\omega _{{\bf p}}) \label{kzza} \end{equation} The corresponding contribution to $1/T_2$ was considered in Ref.\cite{Mor56}% . The term in the longitudinal relaxation rate determined by (\ref{kzza}) is estimated as \begin{equation} \delta ^{(z)}(1/T_1)\propto \widetilde{A}^2\times \left\{ \begin{tabular}{ll} $T^3/J^4,$ & $D=3$ \\ $T^2/J^3,$ & $D=2$% \end{tabular} \right. \end{equation} Provided that the dipole-dipole contributions in (\ref{FK}) are considerable ($\widetilde{A}\sim A$), this term can dominate over the ``Korringa'' term (% \ref{K1}) of order of $A^2I^2\rho ^2T\ln |J/\omega _0|/J^2$ at $T/|J|>|I\rho |\ln ^{1/2}|J/\omega _0|$ only. Note that this two-magnon contribution is similar to the two-phonon (Raman) contribution in the spin-lattice relaxation. The existence of the gap $\omega _0$ is not important here (at least if it is sufficiently small), but the matrix elements of interaction of nuclear spins with magnons are singular, unlike those for acoustic phonons ($|M_{{\bf q}\rightarrow 0}|^2\sim 1/q$ instead of $q$ ). Therefore we have a $T^3$ law instead of $T^7$ one for the phonon scattering \cite {abragam}. \section{Isotropic 2D case and NMR in layered and frustrated magnets} Now we investigate the case of layered magnets, in particular, the isotropic $2D$ limit. A detailed treatment of the spin correlation functions and corresponding spin-fluctuation contributions to $1/T_1$ in the isotropic $2D$ Heisenberg antiferromagnets with $\omega _n\rightarrow 0$ was performed in Ref.\cite{t13}. Here we calculate also corrections owing to electron-magnon interaction. The magnetic ordering temperature is determined by magnetic anisotropy or interlayer coupling, \begin{equation} T_M\sim |J|S^2/\ln (|J|S^2/\max \{\omega _0,|J^{\prime }|\}) \label{tm} \end{equation} (for more details see, e.g., Refs.\cite{SSWT}). Despite the absence of the long-range ordering (LRO) at finite temperatures, spin-wave description holds even in the pure $2D$ isotropic case in the broad temperature region up to $T\sim |J|S$ (i.e., $T_M\rightarrow |J|S^2$) owing to strong short-range order (SRO). In a more general case of finite $\omega _0$ and $% J^{\prime }$, this description holds at $T\gg T_M.$ A possibilty to describe LRO without introducing anomalous averages (like sublattice magnetization) in terms of singularities of the spin correlation function was demonstrated in Ref.\cite{IKZ}. Such an approach enables one to obtain an unified description of ordered and disordered phases. In the pure $2D$ case the magnetization (or sublattice magnetization for the AFM case) $\overline{S}$ is replaced in both magnetic and electronic properties by the square root of the Ornstein-Cernike peak (see, e.g., \cite{IKJP91}). The gap in the effective spin-wave spectrum appears at finite temperatures, which is determined by the inverse correlation length. The correlation length in the situation under consideration is estimated as \cite{Arovas} \begin{equation} \xi \propto \exp \left( \pi |J|S^2/T\right) \text{\ } \label{korr} \end{equation} As shows the two-loop scaling theory\cite{Chakraverty}, the preexponentional factor is temperature independent; quantum effects can renormalize the exchange parameter $J$ \cite{SSWT}. To describe formally NMR in the absence of LRO ($\langle {\bf S}_i\rangle =0$% ) we follow to Ref.\cite{IKZ} and consider the autocorrelation function of the nuclear spin ${\bf I}$ \cite{Mori}. Performing calculations with the simplest Hamiltonian $H_{hf}=A{\bf IS}_i$ to second order in $A$ we derive \[ (I^{+},I^{-})^\omega =\frac 23I(I+1)[-i\omega +\Sigma (\omega )] \] with the memory function \begin{eqnarray} \Sigma (\omega ) &=&A^2\int_0^\infty dt\exp (i\omega t)\sum_{{\bf q}}\langle S_{-{\bf q}}^z(t)S_{{\bf q}}^z+\frac 12S_{-{\bf q}}^{-}(t)S_{{\bf q}% }^{+}\rangle , \\ \langle S_{-{\bf q}}^\alpha (t)S_{{\bf q}}^\beta \rangle &=&\int_{-\infty }^\infty d\omega \exp (i\omega t){\cal J}_{{\bf q}}^{\alpha \beta }(\omega ),% {\cal J}_{{\bf q}}^{\alpha \beta }(\omega )=-\frac 1\pi N(\omega )\text{Im}% \langle \langle S_{{\bf q}}^\beta |S_{-{\bf q}}^\alpha \rangle \rangle _\omega . \end{eqnarray} As discussed in Ref.\cite{IKJP91}, the spectral density ${\cal J}_{{\bf q}% }^{\alpha \beta }(\omega )$ contains an almost singular contribution \begin{equation} \delta {\cal J}_{{\bf q}}^{\alpha \beta }(\omega )\propto \Delta _{{\bf q-Q}% }\Delta _\omega \end{equation} where $\Delta _{{\bf q}}$ and $\Delta _\omega $ are delta-like functions smeared at the scales $q\sim \xi ^{-1}$ and $\omega \sim \omega _\xi $ with the characteristic spin-fluctuation energy $\omega _\xi \sim {\cal D}\xi ^{-2}$ (FM), $\omega _\xi \sim c\xi ^{-1}$ (AFM). To obtain the singular term in $\Sigma (\omega )$ with the correct factor of $S^2$ we can introduce a very small magnetic anisotropy (which does not violate time-reversal symmetry), so that the whole singular contribution passes to $K_{{\bf q}% }^{zz}(\omega )$ (cf. Ref.\cite{IKZ}). Then the term $iA^2S^2/\omega $ occurs at $\omega \gg $ $\omega _\xi $, and we obtain the expression \begin{equation} (I^{+},I^{-})^\omega =\frac i3I(I+1)\left( \frac 1{\omega -AS}+\frac 1{% \omega +AS}\right) \end{equation} which describes precession of the nuclear spin with {\it both} frequencies $% \pm AS$. We see that the resonance picture holds at $\omega _n\gg $ $\omega _\xi $ only. In the opposite case the NMR line is smeared, but we can calculate the quantity $1/T_1$ according to (\ref{tt1}). Provided that $\omega _0\ll $ $\omega _\xi $ the quantity $\omega _\xi $ plays a role of the gap in the magnon spectrum. Therefore at $\omega _n\ll \omega _\xi $ the cutoffs in the singular contributions to $1/T_1$ described by (\ref{a12}), (\ref{a22}), (\ref{zz2d}) are determined by very small inverse correlation length, so that they have very large values and possess unusual temperature behavior. In the $2D$ FM case the expression (\ref{f22}) is also applicable, but with another expression for $\omega _0,\omega _0\rightarrow {\cal D}\xi ^{-2}$ which is exponentially small. We have \begin{equation} \delta ^{(2)}(1/T_1)\propto I^2A^2T^2/{\cal D}^{5/2}\omega _\xi ^{1/2}\propto I^2\omega _n^2\xi T^2/{\cal D}^3. \label{d21} \end{equation} In the isotropic $2D$ AFM case we obtain from (\ref{a12}) \begin{equation} \delta ^{(1)}(1/T_1)\propto I^2A^2T/c\omega _\xi \propto I^2\omega _n^2\xi T/c^2. \label{d22} \end{equation} As follows from (\ref{a22}), (\ref{korr}), \begin{equation} \delta ^{(2)}(1/T_1)\propto I^2A^2(T^2/c^4)\ln ^2\xi \simeq \text{const}(T). \label{d23} \end{equation} Possibility to observe such dependences experimentally is of great interest. Unfortunately, most experimental data for layered compunds deal with copper-oxide systems which are on the bordeline of AFM instability. The latter results in a specific temperature dependence of the spin susceptibility and strong deviations from the Korringa law. This makes separation of one- and two-magnon contributions impossible. At $\omega _n\gg $ $\omega _\xi $ the cutoff in the singular contributions to $1/T_1$ described by (\ref{a12}), (\ref{a22}), (\ref{zz2d}) is the NMR frequency $\omega _n.$ However, such a cutoff is absent for the corresponding terms in $1/T_2$ owing to the interaction $\widetilde{A}\sim a|F^{(1)}|$. Provided that $\widetilde{A}\sim A$ we reproduce for these terms the dependences (\ref{d21})-(\ref{d23}). However, as discussed in Sect.2, for large $\xi $ the lineform turns out to be Gaussian owing to strong Ruderman-Kittel interaction between nuclear spins, the Moriya formula (\ref{tt2}) is inapplicable, and one can estimate the linewidth from (\ref{gauss}). In the 2D case we obtain $1/T_2\propto \xi .$ A more detailed discussion of this situation and an application to copper-oxide systems is given in Ref.\cite{thelen}. We see that NMR investigations can be in principle used to obtain the temperature dependence of the correlation length. When crossing the magnetic ordering point in layered systems the NMR picture should not change radically% $.$ Formally, as follows from (\ref{tm}), (\ref{korr}) at $T\sim T_M$ we have $\ln \xi \simeq \ln |JS^2/\max \{\omega _0,|J^{\prime }|\}|,$ so that the cutoffs are joined smoothly. A similar situation can take place for other systems with suppressed LRO and strong SRO, e.g., for frustrated 3D magnetic systems \cite{Kat} (where $\xi $ is also large, the lineform is Gaussian, and we obtain from (\ref{gauss}) $% 1/T_2\propto \xi ^{1/2}$). This may explain why the problem of detecting long-range magnetic ordering is frustrated systems with small ordered moments with the use of the NMR method so difficult. Indeed, the NMR data for heavy-fermion systems are doubtful and contradict to results of other experiments \cite{hf}. \section{Conclusions} In the present paper we have investigated in detail various mechanism of nuclear magnetic relaxation in metallic ferro- and antiferromagnets in the spin-wave temperature region. In the most cases the main contribution to $% 1/T_1$ is of Korringa type, but its physical origin is more complicated than in paramagnetic metals. Formally, it results from the interaction of nuclear magnetic moments with the {\it localized }electronic subsystem with taking into account the ``Stoner'' (Landau) damping of spin waves via conduction electrons. This contribution is greatly enhanced in comparison with the standard Korringa term by inverse powers of exchange interaction ($s-d(f)$ parameter), especially in ferromagnets. In $3D$ antiferromagnets such a contribution contains the logarithmic singularity which is cut at the gap in the magnon spectrum (magnetic anisotropy energy) $\omega _0.$ Thus we can conclude that the ``Korringa'' relaxation rate in magnetic metals should be much larger than in paramagnetic ones where the relaxation is determined by direct interaction with conduction electrons (such a term is also present in the magnetically ordered state, but is much smaller than the contribution discussed). In the $2D$ AFM case we have $1/T_1\propto \omega _0^{-1}$. In the isotropic limit the singularity in $1/T_1$ is cut at very small inverse correlation length, so that the one-magnon contribution becomes very large and possesses unusual temperature behavior. Besides that, we have calculated contributions from more complicated magnon damping processes (electron-magnon scattering). For antiferromagnets and $2D$ ferromagnets they contain singular logarithmic or (in the $2D$ FM case) power-law factors which are also cut at $\omega _0$. These contributions may result in considerable deviations of the temperature dependence of $1/T_1$ from the linear Korringa law. In the $3D$ FM case this contribution is also noticeable and probably can be separated when fitting experimental data. For half-metallic ferromagnets, where the ``Stoner'' damping is absent, this scattering should be the main nuclear relaxation mechanism (see the discussion of experimental data in the review \cite{UFN}). Provided that the ``longitudinal'' matrix elements of dipole-dipole hyperfine interactions in (\ref{FK}) are not too small, the two-magnon (``Raman'') relaxation processes may be also important in $1/T_1$, especially for $2D$ ferromagnets. The research described was supported in part by Grant No.99-02-16279 from the Russian Basic Research Foundation.
\section{Introduction} Searches for rare signals often fail to detect a signal of sufficient statistical significance and thus face the need to set an upper limit on the signal rate. Unfortunately, there is no standard prescription for setting such limits, and a number of techniques have been employed in the past to meet this challenge. This problem is particularly important for the particle physics community. Upper limits on rare and forbidden decays can provide valuable constraints on physics beyond the Standard Model of Electroweak Interactions. This paper is stimulated by the observation that often authors do not pay enough attention to the choice of a procedure for upper limit estimation. Furthermore, many experimentalists do not provide a sufficient description of the procedure used in their analysis technique, assuming that the reader is smart enough to figure out the details. An example given in Section~\ref{sec:class} shows that, in some situations, the choice of a procedure can change the value of an upper limit {\em by an order of magnitude}, which suggests that this problem should not be taken lightly. In Section~\ref{sec:survey} the most popular approaches to upper limit estimation are surveyed and criticism is provided, where applicable. In Section~\ref{sec:ss} definitions of the statistical significance of a signal are discussed and a new technique, which estimates upper limits using these definitions, is proposed. Unlike commonly used procedures, this approach does not just count the number of events in the signal region, but takes into account statistical fluctuations of the background as well. In Section~\ref{sec:toy_model} various techniques are compared using a toy model. In Section~\ref{sec:cleo} an example of a recent search for the lepton number violating decay $\tau\to\mu\gamma$~\cite{mugamma}, performed by the CLEO collaboration, is given. The emphasis is made on estimation of upper limits, though the discussion can be easily generalized to include construction of confidence intervals whose lower bound is not constrained to zero. There is no discussion of systematic effects in this paper. Throughout the paper the number of observed events, the expected background rate and, if applicable, the coordinates of observed events are assumed to be measured with perfect accuracy. \section{Commonly Used Techniques for Upper Limit Estimation} \label{sec:survey} \subsection{Bayesian Approach} \label{sec:bayes} In a Bayesian approach one has to assume a {\em prior} probability density function (pdf) of an unknown parameter and then perform an experiment to update the prior distribution. A prior pdf reflects knowledge that is available to an experimentalist {\em before} an experiment is performed. The updated prior is called the {\em posterior} pdf and is used to draw inference on the unknown parameter. This updating is done with the use of Bayes' Rule. For the moment let us ignore the issue of background, i.e., let us assume that the background rate is measured very accurately and thus can be treated as a known constant. Then the only unknown parameter is the signal rate $s$. Bayes' Rule gives: \begin{equation} \label{eq:poster} \pi(s|n) = \frac{ f(n|s) \pi(s) }{ \int_{0}^{\infty} f(n|s) \pi(s) ds }\ , \end{equation} where $n$ represents the number of observed events, $f(n|s)$ is the conditional probability to observe $n$ events, given the signal rate $s$, $\pi(s)$ is the prior pdf, and $\pi(s|n)$ is the conditional posterior pdf. Now, for any given confidence level, one can compute a Bayesian confidence interval for the signal rate $s$. For upper limit estimation the natural choice of the Bayesian confidence interval is of the form $(0,s_0)$. Here, $s_0$ denotes an upper limit and can be found from the equation: \begin{equation} \label{eq:bayes_cl} 1-\alpha = \int_0^{s_0} \pi(s|n) ds\ . \end{equation} The confidence level is denoted by $(1-\alpha)$ following the conventional statistical notation. A nice feature of the Bayesian approach is that the zero value of an upper limit $s_0$ always corresponds to the zero value for the confidence level $(1-\alpha)$. As you will see below, this is not necessarily true for the classical approach. The most important step here is to define a prior distribution of the parameter. Naturally, this is the step which brings most of controversy (and sometimes confusion) into Bayesian methods. Many statistical textbooks treat a prior pdf as a purely subjective assumption which is based on experimenter's belief. At the same time there are authors~\cite{jeffr1,jeffr2,jaynes,prosp1} who advocate the objectivity of prior assumptions. In particular, Jaynes~\cite{jaynes} stated the ``basic desideratum'' of the objective approach as follows: {\em in two problems where we have the same prior information, we should assign the same prior probabilities}. These authors have offered a number of mathematical procedures that can be used to convert prior knowledge into an exact formula for the prior distribution. Another important question is whether one should assume an {\em informative} prior, i.e., a prior which incorporates results of previous experiments, or a {\em non-informative} prior, i.e., a prior which claims total ignorance. Naively, it would seem unnatural to use non-informative priors since the power of the Bayesian approach comes from the fact that we update our prior knowledge rather than start every analysis from scratch. In particle physics the major objection against informative priors is based on the following argument: if we assume a prior which incorporates results of earlier experiments, then our experiment will not be independent of those and thus we will not be able to combine our results with the results of previous experiments just by taking a weighted average. So far, particle physicists have been largely ignoring informative priors. Thus, the discussion below covers only those Bayesian methods that assume a non-informative prior pdf for the positive parameter of a Poisson distribution. In the absence of background the conditional pdf $f(n|s)$ is given by: \begin{equation} \label{eq:poiss} f(n|s) = e^{-s} \frac{s^n}{n!}\ . \end{equation} Bayes and Laplace~\cite{bayes,laplace}, who pioneered statistical research employing Bayesian methods, stated that the non-informative prior for any parameter must be flat. This conclusion was not based on any strict mathematical argument, it was merely a product of their intuition. Modern advocates~\cite{hel1,hel2,hel3,hel4} of this approach do not offer any mathematical explanation either, they just consider a flat prior pdf as the most natural choice one can make. As natural as this assumption may seem, there are several objections. The most obvious argument is that if one can assume a flat distribution of an unknown parameter, then one can also assume a flat distribution for any function of this parameter, and these two assumptions are clearly not identical. If an experiment is measuring the mean of the Poisson statistic~(\ref{eq:poiss}), then one can argue that the most natural candidate for the unknown parameter is the signal rate $s$, and we have no physical reasons to consider any functions of this parameter except the parameter itself. In other experiments the situation may be not so simple. For example, if an experiment is measuring a neutrino mass, then typically the measured quantity is not the neutrino mass itself, but the neutrino mass squared. In this situation it's not clear whether one should choose mass or mass squared as a candidate for the flat distribution. Jeffreys~\cite{jeffr1} resolved this problem by introducing an {\em invariate} prior pdf $1/\theta$ which, he stated, was a valid non-informative prior for all problems where the unknown parameter $\theta$ could vary from $0$ to $+\infty$. His choice was mostly motivated by the fact that $d\theta/\theta \propto d\theta^n/\theta^n$, i.e., the pdf of $\theta$ stays invariant under any power transformation. All non-power functions of $\theta$ were rejected by Jeffreys as non-physical under the assumption that the parameter $\theta$ is a dimensional quantity. This argument was put on a more rigorous mathematical basis by Jaynes~\cite{jaynes} who stipulated that a prior pdf has to stay invariant under any symmetry transformation\footnote{Strictly speaking, Jaynes' argument is not applicable to the Poisson distribution~(\ref{eq:poiss}) with a non-dimensional mean rate $s$. Originally Jaynes' non-informative prior was derived for the Poisson pdf $f(n|s) = e^{-st}(st)^n/n!$, where $n$ is the number of counts, $t$ is the counting time, and $s$ is the signal rate per unit time. However, the requirement of dimensionality seems to be somewhat arbitrary. For example, in the above formula $t$ can be the amount of statistics accumulated in the experiment and $s$ can be the signal rate normalized to this amount of statistics. A prior pdf has to stay invariant under the transformation $s'=qs$, $t'=t/q$, i.e., Jaynes' argument is applied.} that does not change the physics of an experiment. His conclusion was similar to that of Jeffreys, namely that the non-informative prior for a Poisson statistic~(\ref{eq:poiss}) has to be proportional to $1/s$. An alternative approach was developed by Box and Tiao~\cite{box} who introduced the notion of a {\em data-translated} likelihood. In their approach a prior pdf is non-informative if the location, but not the shape, of the corresponding posterior likelihood\footnote{As usually, a likelihood function is defined by swapping argument and parameter in the expression of the corresponding pdf.} is determined by the unknown parameter. Thus, the location of the posterior likelihood is completely defined by data, while its shape had been determined before the data were seen (hence the name ``data-translated'' likelihood). It may not be possible to construct a data-translated likelihood for every distribution. In this case one can use Taylor expansion to construct an approximate data-translated likelihood. In particular, a non-informative prior which produces an approximate data-translated likelihood for the Poisson statistic~(\ref{eq:poiss}) is given by $1/\sqrt{s}$. In the presence of background, the Poisson pdf~(\ref{eq:poiss}) has to be modified to account for the non-zero background rate $b$: \begin{equation} \label{eq:poiss_b} f(n|s) = e^{-(s+b)} \frac{(s+b)^n}{n!}\ . \end{equation} If the background rate is accurately measured, it can be treated as a known constant. In this case an argument similar to that of Ref.~\cite{jaynes} gives $1/(s+b)$ and an argument similar to that of Ref.~\cite{box} gives $1/\sqrt{s+b}$ for the non-informative prior. In general, for the prior pdf \begin{equation} \label{eq:prior} \pi(s) \propto \frac{1}{(s+b)^m}\ ;\ \ \ \ 0\leq m\leq 1\ ; \end{equation} the posterior distribution is given by \begin{equation} \label{eq:poster_1sm} \pi(s|n) = \frac{ e^{-(s+b)} (s+b)^{n-m} }{ \Gamma(n-m+1,b) }\ , \end{equation} where \begin{equation} \Gamma(p,\mu) = \int_{\mu}^{\infty} s^{p-1} e^{-s} ds ; \ \ p>0 ;\ \ \mu>0 ; \end{equation} is an incomplete gamma-function. Substituting the posterior pdf~(\ref{eq:poster_1sm}) into Eqn.~(\ref{eq:bayes_cl}), we obtain: \begin{equation} \label{eq:bayes_ul} 1-\alpha = 1 - \frac{ \Gamma(n-m+1,s_0+b) }{ \Gamma(n-m+1,b) }\ . \end{equation} A gamma-function is not well-defined if its first argument is equal to zero or a negative integer. Thus, at $n=0$ (no events observed) and $m=1$ (the $1/(s+b)$ prior) Eqn.~(\ref{eq:bayes_ul}) fails to find an upper limit. Under the flat prior $m=0$, Eqn.~(\ref{eq:bayes_ul}) turns into the formula \begin{equation} \label{eq:pdg96} 1-\alpha = 1 - \frac{ \sum_{k=0}^{n} e^{-(s_0+b)} \frac{(s_0+b)^k}{k!} } { \sum_{k=0}^{n} e^{-b} \frac{b^k}{k!} }\ , \end{equation} which was adopted by Particle Data Group~\cite{pdg96}. Here and below, $0^0$ is by definition equal to one. If the background rate is measured with large uncertainty, then the situation is more complicated. The usual approach is to modify the Poisson statistic~(\ref{eq:poiss_b}) as: \begin{equation} f(n|s) = \int_0^{+\infty} e^{-(s+b)} \frac{(s+b)^n}{n!} f(b) db\ , \end{equation} where $f(b)$ is the measured or predicted pdf of the background rate $b$. The argument of the previous paragraph, leading to the $1/(s+b)$ and $1/\sqrt{s+b}$ expressions for the non-informative prior, is not necessarily valid in this case. A more consistent Bayesian approach would be to derive a joint non-informative prior $\pi(s,b)$, thus treating both the signal and background rates as unknown parameters. An example of such derivation is given in Ref.~\cite{prosp1}. A discussion of intricacies, that might arise from inclusion of background uncertainty into the upper limit calculation, is beyond the material covered in this paper. \subsection{Classical Approach} \label{sec:class} The classical (or frequentist) approach is traditionally interpreted in the following way: if a $(1-\alpha)$ classical confidence set is constructed for the unknown parameter, then the probability for this confidence set to cover the true value of the parameter equals $(1-\alpha)$, that is, if an infinite number of identical independent experiments is performed and for each of these a $(1-\alpha)$ confidence set is constructed, then $100(1-\alpha)\%$ of these confidence sets will and $100\alpha\%$ of these confidence sets will not contain the true value of the parameter. Rules for construction of classical intervals were outlined in a famous work~\cite{neyman} by Neyman. Here I treat the terms ``classical'' and ``frequentist'' as equivalent. The implication is that the classical approach is inevitably connected to the concept of an identical independent experiment. Sometimes this concept is misunderstood, and are curious attempts to treat certain problems in the ``classical'' vein while, in fact, this treatment has nothing to do with the frequentist approach. An example of such misunderstanding is shown in this Section. A confidence set for the unknown signal rate $s$ is typically constructed~\cite{garwood} as a confidence interval $s_1\leq s\leq s_2$ satisfying: \begin{equation} \label{eq:neyman} 1 - \alpha = \sum_{k=n}^{\infty} f(k|s_1) + \sum_{k=0}^{n} f(k|s_2)\ . \end{equation} For upper limit estimation it is natural to consider intervals $0\leq s\leq s_0$, where $s_0$ is the value of an upper limit. Eqn.~(\ref{eq:neyman}) is thus reduced to: \begin{equation} \label{eq:class_ul} 1 - \alpha = 1 - \sum_{k=0}^{n} e^{-(s_0+b)} \frac{(s_0+b)^k}{k!}\ , \end{equation} where $f(k|s_0)$ was replaced by its definition~(\ref{eq:poiss_b}). Eqn.~(\ref{eq:class_ul}) looks similar to the Bayesian formula~(\ref{eq:pdg96}), except that the denominator is now absent. The denominator in Eqn.~(\ref{eq:pdg96}) represents the probability of observing $n$ or less background events in the signal region, and thus, it is always less than 1 except at $b=0$. Therefore, for the non-zero background rate, i.e., $b\neq 0$, the classical approach~(\ref{eq:class_ul}) always provides a smaller value of an upper limit than the Bayesian approach~(\ref{eq:pdg96}) with a flat prior. The difference becomes significant when the denominator in Eqn.~(\ref{eq:pdg96}) is small, i.e., when the number $n$ of events observed in the signal region is small as compared to the expected background rate $b$. For example, for the observed number of events in the signal region $n=3$ and the expected background rate $b=6.5$ the classical approach~(\ref{eq:class_ul}) gives an upper limit $s_0=0.18$ at 90\% confidence level, while the Bayesian approach~(\ref{eq:pdg96}) with a flat prior gives $s_0=3.39$, i.e., a difference of more than an order of magnitude. Another feature of Eqn.~(\ref{eq:class_ul}) is that now a zero upper limit: $s_0=0$, does not give you a zero confidence level. When the expected background rate $b$ is large, one can achieve a situation when $1-\alpha < 1-\sum_{k=0}^{n} e^{-b} b^k/k!$, i.e., formula~(\ref{eq:class_ul}) gives an unphysical negative value of an upper limit. Advocates of the Bayesian method consider this as an undesirable feature of the classical approach. But the failure to set an upper limit is not necessarily a bad feature. It just implies that the observed outcome of an experiment is highly improbable, and one should question the experimental technique that was used to measure the signal and the background. Zech~\cite{zech} made an attempt to derive the Bayesian formula~(\ref{eq:pdg96}) using the classical approach. To arrive at Eqn.~(\ref{eq:pdg96}), Ref.~\cite{zech} postulates that the number of background events cannot exceed the number $n$ of observed events and therefore one has to renormalize the probability~(\ref{eq:class_ul}) according to this constraint. This approach misunderstands the concept of frequentist coverage. Eqn.~(\ref{eq:class_ul}) answers the question: what is the probability of observing $n$ or more events in an {\em identical independent} experiment? This implies that the number $n$ of events cannot be fixed at any particular value. If you fix $n$, then you confine yourself to this particular drawing; the concept of an identical independent experiment is therefore inapplicable. In this case a binomial distribution should be used instead of a Poisson pdf. Confidence intervals are not uniquely defined by Eqn.~(\ref{eq:neyman}) unless one imposes specific criteria on their construction. In the situation discussed above, the uniqueness was introduced by requiring that the confidence set must be an interval $(0,s_0)$. One can choose another requirement and obtain a different confidence set. In general, if the problem is reduced to one variable and one unknown parameter, then confidence intervals are obtained via construction of confidence belts~\cite{eadie}. The idea behind this technique is the following. For every assumed value of the signal rate $s$, one finds an acceptance interval $n_1(s)\leq n\leq n_2(s)$ which satisfies: \begin{equation} \label{eq:c_belt} 1 - \alpha = \sum_{n=n_1}^{n_2} f(n|s)\ . \end{equation} Due to the discrete nature of the Poisson distribution, it is usually impossible to find values of $n_1$ and $n_2$ that satisfy Eqn.~(\ref{eq:c_belt}). The standard solution is to stay on the conservative side and to search for an acceptance interval that gives at least the required coverage: \begin{equation} \label{eq:c_belt_2} 1 - \alpha \leq \sum_{n=n_1}^{n_2} f(n|s)\ . \end{equation} The obtained functions $n_1(s)$ and $n_2(s)$ define two curves on the $s$-vs-$n$ plane. Then, for every value of $n$, one obtains an interval $s_1(n)\leq s\leq s_2(n)$ which lies between these two curves. The obtained interval is a $(1-\alpha)$ confidence interval for the specific value of $n$. The order in which values of $n$ are added to the acceptance region $(n_1,n_2)$ for a specific value of $s$ is called an {\em ordering principle}. Thus, every ordering principle corresponds to a specific set of confidence intervals $(s_1,s_2)$ indexed by the observation variable $n$. The well-known examples are an old paper~\cite{crow} by Crow and Gardner and a recent paper~\cite{feldman} by Feldman and Cousins. Ref.~\cite{crow} minimizes the length of the acceptance interval $(n_1,n_2)$ defined by Eqn.~(\ref{eq:c_belt_2}). Ref.~\cite{feldman} employs an ordering principle based on likelihood ratios. The latter was adopted by the last release of the Particle Data Group Review~\cite{pdg98}. In both approaches an experimentalist does not have to decide whether she/he wants to quote an upper limit or a confidence interval: she/he simply applies the chosen procedure to construct a confidence interval $(s_1,s_2)$. If the lower bound turns out to be strictly equal to zero: $s_1=0$, then an upper limit is quoted, and if the lower bound is positive: $s_1>0$, then a confidence interval is quoted. This versatile procedure, however, has one subtle problem. The fact that the constructed 90\% confidence interval has a non-zero lower bound does not guarantee that a signal of high statistical significance is observed. There is nothing surprising about that. A measured signal rate is usually quoted if the statistical significance of the signal exceeds 3 which corresponds to 99.87\% of the area under a Gaussian peak. At the same time, typical confidence levels used to quote upper limits are 90\% and 95\%. For example, at $b=1$ and $n=3$ the 90\% confidence interval obtained by the procedure of Ref.~\cite{feldman} equals $(0.10,6.42)$. At the same time, the statistical significance $\sigma$ calculated as $1/\sqrt{2\pi} \int_{\sigma}^{\infty} e^{-{x^2}/{2}} dx = \sum_{k=n}^{\infty} e^{-b} b^k/k!$ is only 1.4. Thus, the non-zero lower bound has no clear interpretation: it is a vague indication that a clear signal might be observed in a future experiment. On the contrary, if one chooses to construct a one-sided interval and thus quote an upper limit, this clearly expresses the fact that a signal of sufficient statistical significance was not observed and therefore the signal rate is believed to be consistent with zero. \subsection{Unbinned Extended Maximum Likelihood Fit} \label{sec:lkh} Unbinned extended maximum likelihood fits~\cite{lyons,barlow} have become popular in the past few years. This is an excellent analysis tool for rare signal searches since, unlike a standard maximum likelihood, an extended maximum likelihood correctly incorporates the Poisson error on the number of observed events. Thus, when one observes contributions from two processes (e.g., signal and background), one should use an extended unbinned likelihood: \begin{equation} \label{eq:lkh} {L}(s,b) = \frac{e^{-(s+b)}}{N!} \prod_{i=1}^{N} (s{S}_i + b{B}_i) \end{equation} instead of a standard likelihood: $$ {L}(f_s) = \prod_{i=1}^{N} (f_s{S}_i + (1-f_s){B}_i)\ . $$ Here ${S}_i$ and ${B}_i$ represent the signal and background spatial pdf's, respectively. $N$ is the total number of events observed in the signal region and in the vicinity, and $s$ and $b$ are the signal and background rates, respectively. It has also become a common practice to extract an upper limit value by integrating a likelihood function: \begin{equation} \label{eq:int_s} 1 - \alpha = \frac{ \int_{0}^{s_0} {L}(s)ds }{ \int_{0}^{\infty} {L}(s)ds }\ , \end{equation} where $s_0$, as usually, denotes the value of an upper limit. The likelihood ${L}(s)$ is typically obtained by integrating the two-dimensional likelihood ${L}(s,b)$ over the parameter space $b$: \begin{equation} \label{eq:lkh_int} {L}(s) = \int_{0}^{\infty} {L}(s,b) db\ . \end{equation} This technique is applied, for instance, to estimate upper limits in rare $B$ decay studies~\cite{rare_b1,rare_b2}. It is important to realize that the integration of the likelihood implicitly uses the Bayesian approach. As it was pointed out by Cousins~\cite{cousins}, {\em in particle physics, the prior is almost always taken uniform (where non-zero), although this assumption goes unemphasized by those who merely report that they ``integrated the likelihood function''}. To understand this comment, one should recall the definition of a likelihood function: \begin{equation} \label{eq:lkh_def} {L}(s,b) = f(\vec{x}|s,b)\ , \end{equation} where $\vec{x}=\left\{N,x_1,x_2,...,x_N\right\}$ is a set of observables, the quantity $x_i$ ($i=1,...,N$) is the coordinate of the $i$th event, the symbol $f$ denotes the corresponding pdf, and a vertical line $|$ always implies conditional distribution. Applying Bayes' Theorem, one obtains: $$ {L}(s,b) = \frac{ f(s,b|\vec{x}) f(\vec{x}) }{ f(s,b) }\ . $$ If signal and background are assumed independent\footnote{The assumption of their independence seems natural. However, Prosper~\cite{prosp1} derived a non-informative joint prior $f(s,b)$, where the signal and background contributions are non-factorizable. In fact, this assumption is not as obvious as it may seem.}, their joint pdf must factorize: $f(s,b)=f(s)f(b)$. Multiplying both sides of the above equation by $f(b)$ and integrating over $b$, one obtains: $$ \int_{0}^{\infty} {L}(s,b) f(b) db = \frac{ f(s|\vec{x}) f(\vec{x}) }{ f(s) }\ . $$ The quantity on the right side of the equation can be recognized as ${L}(s)$. Therefore, \begin{equation} \label{eq:lkh_b} {L}(s) = \int_{0}^{\infty} {L}(s,b) f(b) db\ . \end{equation} In the small signal limit ${S}_i\ll{B}_i$, the likelihood~(\ref{eq:lkh}) is expressed as: $$ {L}(s,b) = \frac{e^{-(s+b)}}{N!} b^N \prod_{i=1}^{N} {B}_i\ . $$ Due to the fact that the contributions from the signal and background rates $s$ and $b$ decouple, the background pdf $f(b)$ in Eqn.~(\ref{eq:lkh_b}) cannot change the ratio $\int_{0}^{s_0} {L}(s)ds / \int_{0}^{\infty} {L}(s)ds$. But when the approximation ${S}_i\ll{B}_i$ is no longer valid, this is not so. Thus, Eqn.~(\ref{eq:lkh_int}) treats the pdf $f(b)$ as prior knowledge and assumes that it is flat in the interval $(0,+\infty)$. Furthermore, the posterior distribution $f(s|\vec{x})$ can be represented as: $$ f(s|\vec{x}) = \frac{ {L}(s) f(s) }{ f(\vec{x}) }\ , $$ which is similar to the Bayesian formula~(\ref{eq:poster}). Thus, Eqn.~(\ref{eq:int_s}) implicitly assumes a flat prior $f(s)$ and integrates $f(s|\vec{x}) \propto {L}(s)$ to extract an upper limit. When the experiment is dominated by background, i.e., ${S}_i\ll{B}_i$, the likelihood~(\ref{eq:lkh_int}) is proportional to $e^{-s}$ and the extracted value of an upper limit is equal to 2.3 at 90\% confidence level. The approximation ${S}_i\ll{B}_i$ is equivalent to the situation when no events are observed in the signal region ($n=0$), but the expected number of background events is non-zero ($b\neq 0$). Under these conditions, the Bayesian method~(\ref{eq:pdg96}) with a flat prior gives the $e^{-s}$ behavior as well. Neither the classical approach, nor the Bayesian methods with alternative priors give the $e^{-s}$ behavior for $n=0$ and $b\neq 0$; all these techniques give smaller upper limit values. Thus, the integration of the likelihood and the Bayesian method with a flat prior give the most conservative estimates in background-dominated analyses. One can easily avoid the implicit use of the Bayesian method and set an upper limit using the frequentist definition. It is straightforward to implement this approach by running a Monte Carlo job. First, the observed data $\vec{x}_{obs}$ is fitted to the likelihood function~(\ref{eq:lkh}) to extract estimates $s_{obs}$ and $b_{obs}$ of the true signal and background rates. Then, for every assumed value of an upper limit $s_0$, a Monte Carlo sample, consisting of a large number of experiments, is generated. For each experiment the number of signal and background events are generated assuming Poisson distributions: $$ f(s) = e^{-s_0} \frac{ s_0^s }{s!};\ \ \ f(b) = e^{-b_{obs}} \frac{ b_{obs}^b }{b!};\ \ \ N=s+b; $$ and the coordinates of events are generated, based on the spatial pdf's ${S}_i$ and ${B}_i$. Then the outcome of every experiment is fitted to the same likelihood function~(\ref{eq:lkh}) and the distribution of measured signal rates $f(s_{meas})$ is plotted. The confidence level corresponding to this value of $s_0$ is estimated as: \begin{equation} \label{eq:ul_mc} 1 - \alpha = \frac{ \int_{s_{obs}}^{\infty} f(s_{meas}) ds_{meas} } { \int_0^{\infty} f(s_{meas}) ds_{meas} }\ . \end{equation} Thus, a value of $s_0$ is chosen that gives the required confidence level $(1-\alpha)$. This agrees with the frequentist approach when an experimentalist draws inference about an unknown parameter on the basis of observed data only, without making any subjective assumption. \section{Calculation of Upper Limits, Based on Statistical Significance of a Signal} \label{sec:ss_gen} The Bayesian and classical approaches described in Sections~\ref{sec:bayes} and \ref{sec:class} are based on the conditional probability $f(n|s)$, i.e., the probability to observe $n$ events in the signal region, given the signal rate $s$. However, the number of events observed in the signal region by itself is not so important, since the usual goal of an experiment is not to count events in the signal region, but to observe a signal of high statistical significance. These approaches do not take into account the fact that the number of background events fluctuates too. For example, the classical method~(\ref{eq:class_ul}) estimates the confidence level as a probability to observe a larger number of events in the signal region than the number observed in the experiment, given the assumed value $s$ of the signal rate. But, at the same time, the number of background events in the sideband can fluctuate up to high values or down to zero. The former implies that no signal is observed, and the latter implies that a very clean signal is observed. The methods based on the pdf $f(n|s)$ take into account both these possibilities and thus do not distinguish between signals of high and low statistical significance. Therefore, one should replace the conditional pdf $f(n|s)$ with the conditional pdf $f(\sigma|s)$, where $\sigma$ represents the statistical significance of a signal. Before we proceed to derivation of the related mathematical formalism, we need to discuss various definitions of statistical significance. \subsection{Definition of Statistical Significance} \label{sec:comment} If $n$ events are observed in the signal region and $b$ is the estimated background rate, then the statistical significance $\sigma$ of a signal is defined by \begin{equation} \label{eq:ss_prob} \frac{1}{\sqrt{2\pi}} \int_{\sigma}^{\infty} e^{-{x^2}/{2}} dx = \sum_{k=n}^{\infty} e^{-b} \frac{b^k}{k!}\ , \end{equation} that is, it represents the probability of observing $n$ or a larger number of background events in an identical independent experiment. The background rate $b$ can be estimated in a number of ways. If the background rate is estimated {\em independently} of the data seen, e.g., from a Monte Carlo analysis or from another data sample that is known to contain no signal events, then it is common to assume that the estimated background rate is proportional to the rate observed in the independent experiment: \begin{equation} \label{eq:b_ind} b = \zeta_{ind} n_{ind}\ , \end{equation} where $n_{ind}$ is the number of events observed in the independent experiment, $\zeta_{ind}$ is the corresponding scale factor and the subscript $ind$ implies that the background rate is estimated independently of the data seen. If the background rate is estimated {\em from data}, then the situation is more complicated. In this case, the background rate is usually estimated from sideband, i.e., a region which is located near the signal region and which contains no signal events: \begin{equation} \label{eq:b_stand} b = \zeta_{sb} n_{sb}\ . \end{equation} Here $n_{sb}$ is the number of events observed in the sideband, and $\zeta_{sb}$ is the sideband-to-signal scale factor. Under the assumption of flat background, the scale factor is given by: \begin{equation} \zeta_{sb} = \frac{A_{sig}}{A_{sb}}\ , \end{equation} where $A_{sig}$ is the area of the signal region, and $A_{sb}$ is the area of the sideband. The definition~(\ref{eq:b_stand}) of the background rate is, in my opinion, incorrect. It works only if the sideband is much larger than the signal region, and thus the estimate of the expected background rate $b$ is accurate enough. However, in a situation when the areas of signal and sideband regions are comparable, Eqn.~(\ref{eq:b_stand}), combined with Eqn.~(\ref{eq:ss_prob}), overestimates the significance of a signal for large numbers of observed events $n>b$, increasing the probability of a ``discovery''. This is caused by the fact that the formulas~(\ref{eq:ss_prob}) and (\ref{eq:b_stand}) answer the wrong question. The correct question is: what is the probability of observing $n$ or a larger number of events in the signal region {\it if the true signal rate is zero}? To answer this question, one has to assume that {\it all} events, that were observed in the experiment, came from background, and therefore one has to redefine: \begin{equation} \label{eq:b_my} b = \zeta N\ , \end{equation} where $N$ is the number of events observed in the entire region, which includes the signal region, sideband and, perhaps, an intermediate region between the signal and sideband regions, and $\zeta$ is the corresponding scale factor. Under the assumption of flat background, the scale factor is given by: \begin{equation} \zeta = \frac{A_{sig}}{A}\ , \end{equation} where $A$ is the area of the entire region. In the same spirit, statistical significance, as defined via likelihood, is given by: \begin{equation} \sigma = \sqrt{ -2\, ln\, {L}(0)/{L}_{max} }\ , \end{equation} i.e., it gives a number of standard deviations from the observed signal rate, which maximizes ${L}(s)$, to the zero signal rate, under the assumption that $-2\, ln\, {L}(0)/{L}_{max}$ is distributed as $\chi_1^2$. Formula~(\ref{eq:b_my}) should be used only to estimate statistical significance of a signal; for upper limit calculation one is not allowed to assume that all observed events come from background and one has to apply the standard definition(\ref{eq:b_stand}) of an expected background rate. A numerical discrepancy between the definitions~(\ref{eq:b_stand}) and (\ref{eq:b_my}) can be illustrated on the following hypothetical example. Let us assume that the background spatial pdf is flat in the vicinity of the signal region, that the area of the sideband is equal to that of the signal region, and that one event is observed in the signal region and no events are observed in the sideband. Then an experimentalist, who uses the definition~(\ref{eq:b_stand}) of the statistical significance, would claim that she/he observes a very clean signal ($\sigma=+\infty$), while an experimentalist, who uses the definition~(\ref{eq:b_my}), will estimate the statistical significance of the observed signal as $\sigma=0.27$. The latter reflects the fact that, due to the specific choice of the sideband and signal regions, the observed statistic (one event) is insufficient for positive identification of the signal. This example is purely hypothetical. Of course, in the situation when only one event is observed and the area of the sideband is comparable to that of the signal region, most of experimentalists would choose to quote an upper limit instead of a measurement. A more realistic situation occurs when the sideband is somewhat larger than the signal region, there are some events both in the signal region and in the sideband and the calculated statistical significance is close to three. Then one has to make a binary decision: if the statistical significance is larger than three, then a measurement is quoted, otherwise an upper limit is quoted. In this situation the choice of a specific procedure becomes fairly important, and the example above shows that the numerical discrepancy between the two approaches can be significant. The conclusion that I would like to reach in this Section is that there are two situations that should be treated differently. The first situation occurs when the background rate is estimated independently of the data seen. In this case one has to adopt the definition~(\ref{eq:b_ind}) of the background rate and therefore define statistical significance as: \begin{equation} \label{eq:ss_ind} \frac{1}{\sqrt{2\pi}} \int_{\sigma}^{\infty} e^{-{x^2}/{2}} dx = \sum_{k=n}^{\infty} e^{-\zeta_{ind}n_{ind}} \frac{(\zeta_{ind}n_{ind})^k}{k!}\ . \end{equation} The second situation takes place when the background rate is estimated from the same data sample which is used to draw inference about the unknown signal rate. As shown above, in this case one has to adopt the definition~(\ref{eq:b_my}). Statistical significance is then defined as: \begin{equation} \label{eq:ss_my} \frac{1}{\sqrt{2\pi}} \int_{\sigma}^{\infty} e^{-{x^2}/{2}} dx = \sum_{k=n}^{\infty} e^{-\zeta N} \frac{(\zeta N)^k}{k!}\ . \end{equation} These two formulas will be used to estimate upper limits, based on the statistical significance of a signal. \subsection{Upper Limit Calculation} \label{sec:ss} As shown in Section~\ref{sec:comment}, the statistical significance $\sigma$ of a signal is defined either by Eqn.~(\ref{eq:ss_ind}) or by Eqn.~(\ref{eq:ss_my}). I will take Eqn.~(\ref{eq:ss_my}) as an example and proceed to derive a {\em cumulative density function} (cdf) $P(\sigma\leq\sigma')$. If Eqn.~(\ref{eq:ss_ind}) is chosen, the derivation goes through similar steps; thus, only the final result will be quoted. When the background and signal rates are estimated from the same data sample, one can rewrite Eqn.~(\ref{eq:ss_my}) as: \begin{equation} \label{eq:ss_use} \frac{1}{\sqrt{2\pi}} \int_{\sigma}^{\infty} e^{-{x^2}/{2}} dx = \sum_{k=n}^{\infty} e^{-\zeta(n+n_{out})} \frac{\left[\zeta(n+n_{out})\right]^k}{k!}\ , \end{equation} where $n$ is the number of events observed in the signal region, $n_{out}$ is the number of events observed in the outer region, which covers the sideband and, perhaps, an intermediate region between the signal region and the sideband, and $\zeta$ is the corresponding scale factor. Under the assumption of flat background, the scale factor is given by: \begin{equation} \zeta = \frac{A_{sig}}{A_{sig}+A_{out}}\ , \end{equation} where $A_{sig}$ and $A_{out}$ are the areas of the signal and outer regions respectively. The random variables $n$ and $n_{out}$ are drawn from independent Poisson distributions: \begin{equation} \label{eq:dists} n \sim \mbox{Poisson}(s+b);\ \ \ n_{out} \sim \mbox{Poisson}(\lambda_{out}); \end{equation} where $\lambda_{out}$ is the mean of the Poisson distribution that controls the number of events in the outer region $A_{out}$. The best unbiased estimator of $\lambda_{out}$ is the number of events actually observed outside the signal region: \begin{equation} \hat{\lambda}_{out} = n_{out,obs}\ , \end{equation} and thus the equality $\lambda_{out} = n_{out,obs}$ is implied in the further discussion. In Eqn.~(\ref{eq:dists}) the expected number of background events $b$ in the signal region is estimated in the traditional way: $b=\zeta_{sb}n_{sb,obs}$, and its value is not used to determine the statistical significance of the signal. The cdf $P(\sigma\leq\sigma')$ cannot be expressed in a convenient analytical form. However, the problem can be simplified by introducing a new variable: \begin{equation} \label{eq:p} p = \sum_{k=0}^{n-1} e^{-\zeta(n+n_{out})} \frac{\left[\zeta(n+n_{out})\right]^k}{k!}\ . \end{equation} Hence, $$\frac{1}{\sqrt{2\pi}} \int_{\sigma}^{\infty} e^{-{x^2}/{2}} dx = 1 - p\ ;\ \ \ 0\leq p\leq 1\ .$$ The variable $p$ is a monotone function of $\sigma$. Therefore, its cdf can be obtained by a one-to-one transformation: \begin{equation} \label{eq:p_and_sigma} P( p(\sigma)\leq p(\sigma') ) = P( \sigma\leq\sigma' )\ , \end{equation} and one can use the variable $p$ instead of $\sigma$ to set an upper limit. The cdf of $p$ is given by: \begin{equation} \label{eq:1-p} 1 - P(p\leq p'|s) = \sum_{n=1}^{\infty} \sum_{n_{out}=0}^{n'_{out}-1} e^{-(s+b)} \frac{(s+b)^n}{n!} e^{-\lambda_{out}} \frac{\lambda_{out}^{n_{out}}}{n_{out}!}\ , \end{equation} where $n'_{out}=n'_{out}(n)$ is the smallest non-negative integer which satisfies the inequality: \begin{equation} \label{eq:n_out} \sum_{k=0}^{n-1} e^{-\zeta(n+n'_{out})} \frac{\left[\zeta(n+n'_{out})\right]^k}{k!} \leq p' \end{equation} for the given value of $n$. The infinite sum over $n$ in Eqn.~(\ref{eq:1-p}) converges quickly and can be easily calculated numerically. In a simulation, described in Section~\ref{sec:toy_model}, the summation over $n$ from 1 to 1000 was enough to achieve an accuracy of $10^{-6}$ or better for $(1-P(p\leq p'|s))$. Now one can choose the approach one would like to use. All the techniques described in Sections~\ref{sec:bayes} and \ref{sec:class} are applicable, but now instead of a conditional pdf $f(n|s)$ one would have to use the conditional cdf $P(p\leq p'|s)$. The easiest thing to try is to employ the classical approach and define the confidence interval as $(0,s_0)$. In this case, through the reasoning described in Section~\ref{sec:class}, one arrives at the formula similar to Eqn.~(\ref{eq:class_ul}): \begin{equation} \label{eq:ul_p} 1 - \alpha = 1 - P(p\leq p_{obs}|s_0)\ , \end{equation} where \begin{equation} \label{eq:p_obs} p_{obs} = \sum_{k=0}^{n_{obs}-1} e^{-\zeta(n_{obs}+n_{out,obs})} \frac{\left[\zeta(n_{obs}+n_{out,obs})\right]^k}{k!} \end{equation} is the observed value of the variable $p$. When no events are observed in the signal region ($n_{obs}=0$), one obtains $p_{obs}=0$, therefore $n'_{out}=+\infty$ for any $n\geq 1$, and Eqn.~(\ref{eq:ul_p}) is reduced to: $$1 - \alpha = 1 - e^{-(s_0+b)}\ , $$ which coincides with the classical expression~(\ref{eq:class_ul}). This reflects the fact that in this case background fluctuations are irrelevant as the statistical significance $\sigma$ is always equal to $-\infty$, no matter how many background events we observe in the sideband. When no events are observed outside of the signal region ($\lambda_{out}=n_{out,obs}=0$ and therefore $b=0$), one obtains $n'_{out}\geq 1$ for any $n\geq 1$ and $p_{obs}>0$, therefore the sum over $n_{out}$ in Eqn.~(\ref{eq:1-p}) is always equal to 1, and the formula~(\ref{eq:ul_p}) is reduced to: $$1 - \alpha = 1 - e^{-s_0}\ ,$$ which again is identical to the classical expression~(\ref{eq:class_ul}). This reflects the fact that no background fluctuations are expected in this case, and hence, the statistical significance $\sigma$ is a function of the signal rate $s$ only. If the background rate is estimated independently of the data seen, then one should start from the definition~(\ref{eq:ss_ind}) of statistical significance and repeat the same logical steps to arrive at an equation similar to Eqn.~(\ref{eq:1-p}). The random variables $n$ and $n_{ind}$ are drawn from independent Poisson distributions: \begin{equation} n \sim \mbox{Poisson}(s+b);\ \ \ n_{ind} \sim \mbox{Poisson}(\lambda_{ind}); \end{equation} where $\lambda_{ind}$ is the mean of the Poisson distribution that controls the number of events in an independent sample which is used to estimate the background rate. The best unbiased estimator of $\lambda_{ind}$ is the number of events actually observed in the independent data sample: \begin{equation} \hat{\lambda}_{ind} = n_{ind,obs}\ , \end{equation} and thus the equality $\lambda_{ind} = n_{ind,obs}$ is implied. Now the background rate used in the definition of statistical significance and the background rate used to calculate an upper limit are estimated similarly. Without losing generality, one can stipulate that: \begin{equation} \lambda_{ind} = b/\zeta_{ind}\ . \end{equation} The cdf of $p$ is now given by: \begin{equation} 1 - P(p\leq p'|s) = \sum_{n=1}^{\infty} \sum_{n_{ind}=0}^{n'_{ind}-1} e^{-(s+b)} \frac{(s+b)^n}{n!} e^{-b/\zeta_{ind}} \frac{(b/\zeta_{ind})^{n_{ind}}}{n_{ind}!}\ , \end{equation} where $n'_{ind}=n'_{ind}(n)$ is the smallest non-negative integer which satisfies the inequality \begin{equation} \sum_{k=0}^{n-1} e^{-\zeta_{ind}n'_{ind}} \frac{(\zeta_{ind}n'_{ind})^k}{k!} \leq p' \end{equation} for the given value of $n$. \section{Comparison of Various Approaches Using a Toy Model} \label{sec:toy_model} The performance of all approaches discussed in this paper is compared using the following toy model. The total observation region is defined as an interval $(-10,10)$. The signal spatial pdf is taken to be a Gaussian with zero mean and unit variance, and the background spatial pdf is taken to be flat. Under these conditions, the signal region is defined as $(-2.5,2.5)$, and the sidebands are defined as $(-10,-5)$ and $(5,10)$. The value of the expected background rate is taken to be $b=0, 1, ..., 5$ consecutively, which corresponds to the number of events observed in the sidebands $n_{sb}=0, 2, ..., 10$. The number of events in the intermediate region $(-5,-2.5)$ and $(2.5,5)$ is set equal to the expected number $b$ of background events in the signal region, since the area of the signal region is equal to that of the intermediate region. Positions of events in the sideband and intermediate regions are generated under the assumption of a uniform spatial pdf ${B}_i$. For every value of the expected background rate $b$, upper limits obtained with various approaches are estimated for the number $n$ of events in the signal region varying in integer steps from 0 to 6. The results are plotted in Fig.~\ref{fig:all}. Inside the signal region all events are positioned precisely at zero. It is assumed that the background rate in this experiment is estimated from sidebands; thus, to implement the classical method of Section~\ref{sec:ss}, formulas~(\ref{eq:ul_p}) and (\ref{eq:p_obs}) are used. To estimate a 90\% CL upper limit by the Monte Carlo technique of Section~\ref{sec:lkh}, I find the value $s_0$ of an upper limit by using the method of binary division. For every assumed value of $s_0$, I generate a Monte Carlo sample consisting of 50,000 experiments and use the formula~(\ref{eq:ul_mc}) to estimate the corresponding confidence level. This procedure is repeated until the value of $s_0$ that gives a 90\% confidence level is obtained. The required accuracy for the computation of $s_0$ is taken $10^{-2}$. For comparison, lower and upper bounds of 90\% confidence intervals obtained by the procedure~\cite{feldman} are shown with crosses. One cross for a specific value of $n$ corresponds to the situation when the lower bound is strictly zero. \sixplex{htbp} {8b0_freq.eps}{8b1_freq.eps} {8b2_freq.eps}{8b3_freq.eps} {8b4_freq.eps}{8b5_freq.eps} {all}{Upper limits as functions of the number $n$ of observed events in the signal region under the assumptions described in the text. The expected number of background events in the signal region takes values of 1) $b=0$; 2) $b=1$; 3) $b=2$; 4) $b=3$; 5) $b=4$; 6) $b=5$.} As shown in Fig.~\ref{fig:all}, at $n=0$ the Bayesian method~(\ref{eq:pdg96}) with a flat prior pdf and the integration of the likelihood, described in Section~\ref{sec:lkh}, always give 2.30, while all the other approaches give somewhat smaller values. For this particular model, the likelihood integration technique of Section~\ref{sec:lkh} turns out to be the most conservative approach, except for $b=0$. The Bayesian method with a flat prior pdf always gives larger upper limit values than the Bayesian methods with the $1/\sqrt{s+b}$ and $1/(s+b)$ priors and both classical approaches discussed in Sections~\ref{sec:class} and \ref{sec:ss}. The results produced by the Bayesian method with the $1/(s+b)$ prior can be obtained by shifting the corresponding results obtained with a flat prior one step to the right, e.g., $\{n=0,m=0\}$ and $\{n=1,m=1\}$ obviously produce the same result when substituted into the gamma-function $\Gamma(n-m+1,s_0+b)$. The classical approaches of Sections~\ref{sec:class} and \ref{sec:ss} fail to set an upper limit at $n\ll b$, i.e., when background dominates over signal. The classical procedure, based on the statistical significance of a signal, always gives a smaller upper limit value, compared to the standard classical approach of Section~\ref{sec:class}, except at $n=0$. In the specific situation, when the expected background is zero, the Bayesian method with a flat prior pdf, both classical approaches and both likelihood techniques of Section~\ref{sec:lkh} give identical results. \section{An Example of a CLEO Analysis: $\tau\to\mu\gamma$} \label{sec:cleo} In 1996 the CLEO Collaboration searched~\cite{mugamma} for the neutrinoless decay $\tau\to\mu\gamma$ and set an upper limit, which is, so far, the most stringent limit on the $\tau\to\mu\gamma$ branching fraction. In this analysis 3 events were observed in the signal region and the expected number of background events was estimated to be 5.5. The signal Monte Carlo and data energy-vs-mass distributions are shown in Fig.~\ref{fig:mugamma}. The Bayesian approach with a flat prior pdf was used and the upper limit value was estimated as 3.6 at 90\% confidence level. This value was divided by the integrated luminosity and efficiency factor and thus an upper limit of $3.0\times 10^{-6}$ at 90\% confidence level was obtained for the $\tau\to\mu\gamma$ branching fraction. In Table~1 are shown the values of upper limits for this analysis calculated with the alternative techniques. \simplex{htbp}{mug_evsmass.ps}{mugamma} {$(E_{\mu\gamma}-E_{beam})$ vs $(m_{\mu\gamma}-m_{\tau})$ distribution. Solid squares represent the data, open circles represent the signal Monte Carlo distribution.} \begin{table}[bthp] \caption{}{\bf Upper limits at 90\% CL for the $\tau\to\mu\gamma$ analysis of Ref.~\cite{mugamma}.} \begin{tabular}{|c|c|}\hline {\bf Method} & {\bf Upper limit at 90\% CL} \\ \hline Bayesian with flat prior & {\bf 3.57} \\ \hline Bayesian $1/\sqrt{s+b}$ & 3.30 \\ \hline Bayesian $1/(s+b)$ & 3.06 \\ \hline classical & 1.18 \\ \hline classical, based on statistical significance & 1.03 \\ \hline integration of likelihood & 2.30 \\ \hline Monte Carlo likelihood technique & 1.37 \\ \hline Feldman \& Cousins~\cite{feldman} & $\sim 2.5$ \\ \hline \end{tabular} \end{table} The value given by the procedure~\cite{feldman} of Feldman and Cousins is a rough estimate only, since the combination of input parameters $\{n=3,b=5.5\}$ is not shown in their tables. To implement the maximum likelihood approach of Section~\ref{sec:lkh}, the signal Monte Carlo two-dimensional distribution on the energy-vs-mass plane was fitted to a bivariate Gaussian plus a non-Gaussian tail produced by initial and final state radiation and other effects: \begin{eqnarray} {S}_i (m,E) & = & \frac{1}{A+B} \left\{ \frac{A}{2\pi\sigma_m\sigma_E\sqrt{1-\rho^2}} \right. \times \nonumber \\ & & \left. \times exp \left\{ -\frac{1}{2(1-\rho^2)} \left[ \left( \frac{m-m_0}{\sigma_m} \right)^2 - 2\rho \left(\frac{m-m_0}{\sigma_m}\right)\left(\frac{E-E_0}{\sigma_E}\right) + \left( \frac{E-E_0}{\sigma_E} \right)^2 \right] \right\} \right. + \nonumber \\ & & \left. + B\epsilon(m,E) \right\}\ ;\\ & & \nonumber \\ \epsilon(m,E) & = & \left\{ \begin{array}{cl} \frac{1}{\sqrt{2\pi}\sigma_m} exp\left[ -\frac{1}{2} \left( \frac{m-m_0}{\sigma_m} \right)^2 \right] \frac{1}{\sigma_E\Gamma(\alpha)\beta^\alpha} \left( \frac{E_0-E}{\sigma_E} \right)^{\alpha-1} exp\left[ -\frac{E_0-E}{\beta\sigma_E} \right] & \mbox{if $E<E_0$} \\ 0 & \mbox{otherwise} \end{array} \right. \nonumber \end{eqnarray} where $A$, $B$, $\sigma_m$, $\sigma_E$, $\rho$, $m_0(\approx m_\tau)$, $E_0(\approx E_{beam})$, $\alpha$ and $\beta$ are the fit parameters. The background spatial pdf ${B}_i$ was obtained by fitting data events observed in the vicinity of the signal region to a linear function: \begin{equation} {B}_i (m,E) = \frac{1}{m_2-m_1} \frac{1}{ (a_0-a_1E')(E_2-E_1) + 0.5a_1(E_2^2-E_1^2) } \left[ a_0 + a_1(E-E') \right]\ , \end{equation} where $a_0$, $a_1$ and $E'$ are the fit parameters, and $(m_1,m_2)$ and $(E_1,E_2)$ are the limits defining the fit region. The three events observed in the signal region are located far from the peak of the signal Monte Carlo distribution. Thus, the maximum likelihood fit treats these events as background, and the extracted signal rate is consistent with zero. This explains why the likelihood integration technique gives the value of 2.30. \section{Conclusion} There is no such thing as the ``best'' procedure for upper limit estimation. An experimentalist is free to choose any procedure she/he likes, based on her/his belief and experience. The only requirement is that the chosen procedure must have a strict mathematical foundation. The Bayesian method with a flat prior and the likelihood integration technique of Section~\ref{sec:lkh} seem to have been the two most popular choices in the past few years. Typically, these two approaches give the most conservative values of upper limits. The purpose of this note is to show that there are other approaches, equally justified by mathematical formalism, which produce less conservative estimates. \section{Acknowledgements} The author thanks Prof.~W.T.~Ford, Prof.~R.~Stroynowski, Prof.~H.L.~Gray and Dr.~I.~Volobouev for useful comments and suggestions.
\section{Introduction} Three-dimensional Quantum Electrodynamics ($QED_3$), with an even number of fermion flavours, apart from serving as a toy model for studying chiral-symmetry breaking and confinement, also constitutes a physically interesting theory {\it per se}, in view of its possible applications in modelling novel (high-temperature) superconductors~\cite{dor}-\cite{csm}. Chiral symmetry breaking or, equivalently, dynamical mass generation for fermions, in even-flavour $QED_3$ has still many unresolved issues. One of those is the existence of a (dimensionless) critical coupling, above which dynamical mass generation for the fermions occurs~\cite{app}. In the context of large-$N$ treatment, which at present constitutes the only well-studied approach, the r\^ole of the dimensionless coupling~\footnote{In three dimensions the coupling $e^2$ has dimensions of mass. One can still define, however, dimensionless couplings by dividing with a dynamically generated scale, which in the large $N$-treatment arises by demanding that~\cite{app} $e^2 N =8 \alpha$, with the scale $\alpha$ kept fixed as $N \rightarrow \infty$. The dimensionless coupling is then defined as the ratio $e^2/8\alpha = 1/N$.} is played by the inverse of the fermion flavour number $N$. The issue of the existence of a critical coupling in $QED_3$ is a delicate one~\cite{penn}. Many of the original approximations~\cite{app} leading to its existence have been questioned, in particular the fact that wave-function renormalization effects have not been properly accounted for. Recently, however, the incorporation of such effects, still within the large -$N$ context, appears to corroborate~\cite{amm}-\cite{maris} the qualitative picture advocated in \cite{app}. In addition, the latter is also supported by lattice simulations~\cite{kogut}. Nonetheless, the situation is far from being conclusive. The fact that the critical $N$, below which dynamical mass generation occurs, is found to be of order $3$ (in a four-component notation for the fermions) provides motivation for searches beyond the large-$N$ treatment. Moreover, up to now the gauge coupling in the infrared has been treated as an arbitrary parameter, whose size has not been restricted by an additional dynamical constraint. At present we are lacking a self-consistent treatment of the dynamical Schwinger-Dyson (SD) equations involving the vertex function on an equal footing with the self-energy and gap functions. In all the approaches so far, at least within the large $N$ treatment that we are aware of, one invokes a specific Ansatz for the vertex, by the sole requirement of satisfying some truncated form of the Ward-Takahashi identity stemming from gauge invariance~\cite{kondo}-\cite{am}. The lack of dynamical information for the coupling poses problems; for instance, its size in the infrared is treated as an arbitrary parameter, being assumed to merely exceed a critical value, if one wishes to trigger chiral symmetry breaking. Another point, related to the above, which is already familiar from studies in the case of four-dimensional non-Abelian gauge theories, is whether chiral-symmetry breaking is associated with confinement of charges~\cite{conf}. This issue acquires physical importance in view of the condensed-matter applications. In particular, it may shed more light in the dynamics of spin-charge separation, by analogy with the physics of strong interactions~\cite{laughlin}. In this work we shall not deal with issues of confinement, which exists in $QED_3$ despite its abelian nature. Instead, we shall attempt a novel approach to chiral symmetry breaking, independently of a large $N$ treatment, by studying the {\it coupled} fermion and photon self-energies and vertex SD equations in the context of a method first introduced for the case of four-dimensional $QCD$~\cite{pc}. The novel ingredient is that we concentrate on the semi-amputated vertex, defined in section 2, which is the correct gauge-invariant quantity to determine a physically meaningful ``running'' coupling (`effective charge'). The fact that $QED_3$ is superrenormalizable in the ultraviolet does not preclude the possibility of defining such a quantity, having non-trivial structure in the infrared. As we show in sections 2 and 3, the existence of a non-trivial infrared fixed point is a property {\it only} of the infrared-regularized theory, as conjectured in \cite{am}. In other words, in the absence of fermion (or photon) masses, the infrared singularities of the vertex equation will force the effective charge to vanish at zero momentum transfer, thereby excluding the possibility of an interesting infrared behaviour. From a condensed-matter application point of view this would correspond to what is usually called the Landau-Fermi liquid theory (trivial infrared fixed point)~\cite{shankar,am}. In the presence of masses, and in particular fermion masses, we show in section 3 that there is a non-trivial infrared fixed point structure, stemming from the fact that the effective charge obtained as a self-consistent solution of the {\it non-linear} vertex SD equation, is driven to a finite positive value, which can be large enough to trigger dynamical generation of a fermion mass. This implies that the phenomenon of chiral symmetry breaking is intimately associated with deviations from the trivial infrared fixed-point structure. We should stress that, as a result of the non-linearity of the vertex equation, there are delicate constraints between the fermion mass and the effective charge, which are responsible for the appearance of regions of the latter for which dynamical generations occurs. At present, these restrictions appear as a consequence of mathematical self-consistency of the truncated equations. It is not clear to us, whether the upper bounds on the effective charge, imposed by the present cubic approximations for the vertex corrections, will survive the inclusion of higher orders. In contrast, we believe that the lower bounds will survive such a treatment, thereby indicating the existence of a critical coupling above which dynamical mass generation will occur. This is physically appealing, given that one would not expect a weakly-coupled theory to be capable of breaking dynamically chiral symmetry. The layout of this article is as follows: in section 2 we set up the SD equations that we wish to study, and discuss in detail the approximations employed. We demonstrate that, under certain assumptions to be justified retrospectively in section 4, the equation for the semi-amputated vertex decouples from the rest, and hence can be solved separately. Moreover, we establish the absence of a non-trivial infrared fixed point rigorously (within the cubic approximation for the vertex), by casting the SD equation for the effective charge in a form of a non-linear differential equation, known as Emden-Fowler equation~\cite{emden,kamke}. In section 3 we study the equation for the vertex in the presence of a fermion mass, acting as an infrared regulator. We derive the appropriate non-linear differential equation describing the infrared behaviour of the running coupling, and solve it to demonstrate the existence of a non-trivial infrared fixed point. The so-derived running charge is a monotonically decreasing function of the momentum, tending asymptotically to a constant positive value in the ultraviolet. In section 4, we solve the equations for the photon and fermion self-energies, which in our approximations decouple from each other and depend only on the semi-amputated vertex function. We then verify the self-consistency of the approach. In section 5 we examine the self-consistency of the dynamical generation of the fermion mass, by solving the appropriate SD equation upon substituting the solution for the semi-amputated vertex found in previous sections. The self-consistency of the approach restricts the allowed regions of the effective charge, implying the existence of a lower bound (critical coupling) but also of an upper one. In section 6, we examine an alternative type of infrared cut-off, namely that of a (bare) covariant photon mass term. This case also exhibits a non-trivial infrared fixed-point structure but, in contrast to the monotonic decrease of the effective charge in the case of fermion masses, here the coupling initially increases in the infrared, then displays a local maximum, and eventually decreases, tending asymptotically to a constant value in the ultraviolet. Some possible applications of this behaviour, inspired by condensed-matter physics, are briefly discussed. Finally, in section 7 we present our conclusions and outlook. \medskip \setcounter{equation}{0} \section{The SD equation for the semi-amputated vertex} In this section we will first set up the SD equations for the photon and electron self-energies, and the photon-electron vertex; then we will define the semi-amputated vertex and derive its corresponding SD equation. As we will explain, the latter governs the behaviour of the effective coupling in the infra-red. The derivation of the SD equations for the photon propagator $\Delta_{\mu\nu}$, the electron propagator $S_{F}$, and the photon-electron vertex $\Gamma_{\mu}$ proceeds following standard methods \cite{cjt,cm} (see figures \ref{fig1a},\ref{fig1b},\ref{fig1c}). The full photon propagator $\Delta_{\mu\nu}$, its inverse $\Delta_{\mu\nu}^{-1}$, and the full vacuum polarisation $\Pi_{\mu\nu}$ in Euclidean space are related by \begin{eqnarray} \Delta_{\mu\nu}^{-1}(q) &=& \Delta^{-1}_{0\mu\nu}(q) + \Pi_{\mu\nu}(q) \nonumber\\ {\Delta}^{-1}_{0\mu\nu}(q) &=& q^2 \delta_{\mu\nu} - (1-\frac{1}{\xi})q_{\mu}q_{\nu}\nonumber\\ \Pi_{\mu\nu}(q) &=& (q^2~\delta_{\mu\nu} - q_{\mu}q_{\nu})\Pi(q) \nonumber\\ \Delta_{\mu\nu}(q) &=& (\delta_{\mu\nu} - q_{\mu}q_{\nu}/q^2)[q^2 + q^2~\Pi(q)]^{-1} + \xi q_{\mu}q_{\nu}/q^4 \label{vacpol} \end{eqnarray} where $\xi$ is the gauge fixing parameter (in covariant gauges). The corresponding SD equation reads \begin{equation} \Delta_{\mu\nu}^{-1}(q)= \Delta^{-1}_{0\mu\nu}(q) + e^2 \int \frac{d^3k}{(2\pi)^3} {\rm Tr}[\Gamma_{\mu} S_{F}\Gamma_{\nu}S_{F}] ~+\dots \label{SD2} \end{equation} The SD equation for the electron propagator $S_{F}$ is given by \begin{equation} S_{F}^{-1}(p)= -i\nd{p} - e^2 \int \frac{d^3k}{(2\pi)^3} \Gamma_{\mu} S_{F}\Gamma_{\nu}\Delta^{\mu\nu} ~+ \dots \label{SD1} \end{equation} Finally, the SD equation for the photon-electron vertex $\Gamma_{\mu}$ has the form \begin{eqnarray} \Gamma_{\mu}(p_1,p_2,p_3) &=& \gamma_{\mu} -e^2 \int \frac{d^3k}{(2\pi)^3} \Gamma_{\alpha} S_{F}(k-p_1)\Gamma_{\mu}S_{F}(k) \Gamma_{\beta}\Delta^{\alpha\beta}(k+p_2) \nonumber\\ && +\dots \label{SD3} \end{eqnarray} with $p_1+p_2+p_3=0$. The ellipses on the right-hand side of (\ref{SD2})-(\ref{SD3}), denote the infinite set of terms containing the two-particle irreducible four-point function \cite{cjt,cm}. Although we are not working in the context of a large $N$ analysis, we note that the above truncation is compatible with working to leading order in resummed $1/N$ expansion. \begin{figure}[hbt] \begin{center} \begin{picture}(300,50) \Text(5,30)[]{$\Big($} \Photon(10,30)(35,30){3}{2.5} \Vertex(40,30){5} \Photon(45,30)(70,30){3}{2.5} \Text(80,31)[]{$\Big)^{-\!1}$} \Text(95,30)[]{$=$} \Text(110,30)[]{$\Big($} \Photon(115,30)(175,30){3}{5.5} \Text(185,31)[]{$\Big)^{-\!1}$} \Text(200,30)[]{$+$} \Photon(215,30)(240,30){3}{2.5} \Photon(270,30)(295,30){3}{2.5} \GOval(255,43)(6,8)(0){0} \GOval(255,17)(6,8)(0){0} \CCirc(255,30){14}{White}{White} \CArc(255,30)(15,0,360) \GCirc(240,30){4}{0.2} \GCirc(270,30){4}{0.2} \end{picture} \caption{Schematic form of the SD equation for the gauge field propagator in resummed perturbation theory. The blobs on the right-hand-side indicate the full (non-perturbative) vertex corrections.} \label{fig1a} \end{center} \end{figure} \begin{figure}[hbt] \begin{center} \begin{picture}(300,50)(0,0) \Text(5,30)[]{$\Big($} \Line(10,30)(70,30) \Text(80,31)[]{$\Big)^{-\! 1}$} \GOval(40,30)(6,8)(0){0} \CBox(32,23)(48,29){White}{White} \Text(95,30)[]{$=$} \Text(110,30)[]{$\Big($} \Line(115,30)(175,30) \Text(185,31)[]{$\Big)^{-\! 1}$} \Text(200,30)[]{$-$} \Line(215,30)(295,30) \GOval(255,30)(6,8)(0){0} \CBox(247,23)(263,29){White}{White} \PhotonArc(255,30)(25,9,79){-3}{2.5} \PhotonArc(255.4,30)(25,101,171){-3}{2.5} \Vertex(255,55){5} \GCirc(230,30){4}{0.2} \GCirc(280,30){4}{0.2} \end{picture} \caption{Schematic form of the SD equation for the full fermion propagator. Blobs indicate full non-perturbative quantities. Notice the full vertex appears on both ends of the internal photon line. \label{fig1b}} \end{center} \end{figure} \begin{figure}[hbt] \begin{center} \begin{picture}(330,65)(0,0) \ArrowLine(0,30)(30,30) \GCirc(34,30){4}{0.2} \ArrowLine(38,30)(68,30) \Photon(34,64)(34,34){3}{3} \Text(83,30)[]{$=$} \ArrowLine(98,30)(128,30) \GCirc(129,30){1}{0.2} \ArrowLine(130,30)(160,30) \Photon(129,64)(129,31){3}{3} \Text(175,30)[]{$+$} \ArrowLine(190,30)(205,30) \GCirc(209,30){4}{0.2} \Line(213,30)(277,30) \GCirc(245,30){4}{0.2} \GCirc(281,30){4}{0.2} \ArrowLine(285,30)(300,30) \GOval(227,30)(6,8)(0){0} \GOval(263,30)(6,8)(0){0} \CBox(218,23)(236,29){White}{White} \CBox(254,23)(272,29){White}{White} \Photon(245,64)(245,34){3}{3} \PhotonArc(245,64.7)(50,228,266){3}{3} \PhotonArc(245,64.7)(50,274,312){-3}{3} \Vertex(245,14.7){5} \Text(315,30)[l]{$+\quad \cdots$} \end{picture} \caption{The SD equation for the vertex $\Gamma_{\mu}$ \label{fig1c}} \end{center} \end{figure} We next define the scalar quantities $A$, $B$ and ${\cal G}$ as follows: \begin{equation} S_{F}(k)= \frac{1}{A(k)~\nd{k}} \label{pertfermion} \end{equation} \begin{equation} \Delta _{\mu\nu}(k) \equiv \frac{\delta_{\mu\nu}}{B(k)k^2} \label{photon} \end{equation} and \begin{equation} \Gamma _\mu (p_1,p_2,p_3)={\cal G}(p_1,p_2,p_3) \gamma _\mu ~. \label{vertexgen} \end{equation} The quantity $B$ is related to $\Pi$ defined in (\ref{vacpol}) by $B(q)= 1 + \Pi(q)$. The definition in (\ref{photon}) implies that the longitudinal pieces of the photon propagator will be discarded in what follows. Of course, there are no rigorous field-theoretic arguments justifying their omission or inclusion. The correct treatment of such terms necessitates a formalism which would allow for the self-consistent truncation of the SD series in a manifestly gauge-invariant way; unfortunately, no such formalism exists to date. The standard lore when writing down SD equations is to use in (\ref{SD1}) the form of $\Delta_{\mu\nu}$ given in (\ref{vacpol}), setting $\xi=0$ (Landau gauge). While in $QED_4$ this choice renders the vertex corrections unimportant in the ultraviolet, it appears to be less compelling in the context of the superrenormalizable $QED_3$. In addition, it is known that, while the conventionally defined fermion self-energy and photon-fermion vertex depend explicitly on the gauge-fixing parameter $\xi$, it is possible to construct --at least at one-loop -- a $\xi$-independent fermion self-energy and vertex, by resorting to the diagrammatic rearrangement of the $S$-matrix known as the pinch technique \cite{PT}. It turns out that the $\xi$-independent fermion self-energy and vertex so constructed {\it coincide} with their conventional counterparts, if we choose for the latter the special value $\xi=1$ (Feynmann-t' Hooft gauge)~\cite{PT2}. Furthermore, as has been formally shown in \cite{Stagnant}, {\it all} longitudinal pieces appearing in (\ref{vacpol}) vanish from {\it physical} observables, such as $S$-matrix elements, to all orders in perturbation theory. Thus, one is led to a generalized form of the Feynman-t' Hooft gauge, known as the ``stagnant gauge'', where only the $\delta_{\mu\nu}$ part of the vacuum polarization contributes, to {\it all} orders in perturbation theory. This gauge will be adopted throughout the present article. Following \cite{pc} and \cite{cm} we define the semi-amputated vertex ${\hat G}$ as \begin{equation} {\hat G}(p_1,p_2,p_3) \equiv Z(p_1,p_2,p_3) {\cal G}(p_1,p_2,p_3) \end{equation} \label{amp} with \begin{equation} Z(p_1,p_2,p_3) = B^{-1/2}(p_1)A^{-1/2}(p_2) A^{-1/2}(p_3) \label{zed} \end{equation} This definition proves very useful in reducing the complexity of the set of equations (\ref{SD1})-(\ref{SD3}), under certain approximations to be discussed in detail in the next sections. In addition, the quantity \begin{equation} g_R(p_1,p_2,p_3) \equiv e {\hat G}(p_1,p_2,p_3) \label{runcoupl} \end{equation} provides a natural generalization of the concept of the running or effective charge in the context of superrenormalizable gauge theories \cite{jmcres}, such as $QED_3$. This running of the coupling should be understood as a Wilsonian rather than Gell-Mann-Low type, in the sense that it is not associated with ultraviolet infinities; instead, it expresses a non-trivial infrared structure of the theory~\cite{am}. Note that in four-dimensional $QED$ ($QED_4$) the effective charge $e^2_{\rm eff}$ is defined in terms of the photon vacuum polarisation as: \begin{equation} e^2_{\rm eff}(q^{2}) = e^2 [1 + \Pi(q^{2})]^{-1}, \label{pertphot} \end{equation} and is a gauge-, scale-, and scheme-independent quantity \cite{JW} to all orders in perturbation theory. $e^2_{\rm eff}$ depends explicitly on $q^2$ {\it and} the masses of the fermions inside the vacuum polarisation loop. In the limit where the fermion masses can be neglected, $e^2_{\rm eff}(q^{2})$ coincides with the running coupling obtained by the $\beta$ function of $QED_4$, i.e. the solution of the usual renormalization-group differential equation. An advantage of the definition given in (\ref{runcoupl}) is that it captures the running coupling even in the case of scalar theories \cite{cm}, where, due to the absence of Ward-Takahashi identities, the role of the gauge boson self-energies is not as prominent as in gauge theories. As we shall see in section 5, the interpretation of $g_R$ defined in (\ref{runcoupl}) as a running coupling is also justified by the form of the SD for the fermion mass gap. The equation for the semi-amputated vertex $\hat{G}$ may be obtained from (\ref{SD3}) by multiplying both sides by the factor $Z(p_1,p_2,p_3)$, i.e. \begin{equation} {\hat G}(p_1,p_2,p_3)\gamma_{\mu} = Z(p_1,p_2,p_3)\gamma_{\mu} -e^2 \int \frac{d^3k}{(2\pi)^3} {\hat G}^3 \gamma^{\alpha}\frac{1}{\nd{k}-\nd{p_1}}\gamma_{\mu} \frac{1}{\nd{k}} \gamma_{\alpha}\frac{1}{(k+p_2)^2} \label{SDamp} \end{equation} where \begin{equation} {\hat G}^3 \equiv {\hat G}(p_3,k+p_2,p_1-k){\hat G}(p_1,-k,k-p_1) {\hat G}(p_2,k,-k-p_2) \end{equation} In what follows we shall restrict ourselves to the case where the photon momentum is vanishingly small, and thus one is left with a single momentum scale $p$. One can then define a renormalization-group $\beta$ function from this ``running'' coupling $G(p)$ by setting \begin{equation} \beta \equiv p \frac{d}{d p}{\hat G}(p) \label{beta} \end{equation} In order to further simplify the SD equation for $G(p)$ we make the additional approximation that ${\hat G}^3 = {\hat G}^3(k)$, i.e. a cubic power of a single ${\hat G}(k)$ depending only on the integration variable $k$. This approximation will be justified by the self-consistency of the solutions. Carrying out the gamma-matrix algebra using the formulae $ \gamma_{\mu}\gamma_{\mu} = -d $, and $\gamma_{\mu}\gamma_{\rho}\gamma_{\mu} = (d-2)\gamma_{\rho}$ valid for $4\times 4$ gamma matrices in $d(=3)$-dimensional Euclidean space, one obtains in a straightforward manner: \begin{equation} {\hat G}(p) = Z(p) + \frac{1}{3}e^2 \int \frac{d^3k}{(2\pi)^3} {\hat G}^3 (k) \frac{1}{k^2}\frac{1}{(k-p)^2} \label{asymptotic} \end{equation} Several remarks are now in order. First, one observes that $Z(p) \rightarrow 1$ for $p \rightarrow \infty$, where perturbation theory is valid. This is inferred from the fact that in such a case, as can be readily verified, the functions $A(p), B(p) \rightarrow 1 + {\cal O}(e^2/p)$. In addition, in the ultraviolet region, $p \rightarrow \infty$, gauge invariance requires ${\cal G}(p) \sim A(p)$. Second, from (\ref{asymptotic}) one observes that, if ${\hat G}$ stays positive, which is expected for any physical theory, then, as a result of the positivity of the integrand, ${\hat G(p)} \ge Z(p)$ for any $p$. Thus, one has the following basic properties of ${\hat G}(p)$, which stem directly from the integral equation (\ref{asymptotic}): \begin{equation} {\hat G}(p) \ge Z(p),~ {\rm for~all}~ p, \quad ; \quad {\hat G}(p) \sim B^{-1/2}(p) \rightarrow 1, \quad p \rightarrow \infty \label{ascond} \end{equation} Notice that ${\hat G}(p)$ in the ultraviolet is thereby given by $\left(1 + \Pi (p)\right)^{-1/2}$ as in $QED_4$. This is the perturbative result, which, as we shall see later, is modified non-trivially in the infrared, in a way consistent with gauge invariance. Moreover, as we shall see later, the self-consistency of the approximations employed will require ${\hat G}(0) >> \sqrt{3/2}$. Our next assumption is that in the infrared regime, $k/\alpha << 1$, which is of interest to us here, the effects of the inhomogeneous term $Z(p)$ can be ignored. This assumption will be justified later on, when we consider a non-trivial self-consistency check of the solutions. We now remark that by ignoring the effects of $Z(p)$ one can decouple the equation for the (gauge-invariant) {\it amputated} vertex from the equations for $A(p)$,$B(p)$. As we shall discuss in subsequent sections, these latter equations also decouple from each other, depending only on the vertex function ${\hat G}(p)$. Because of this, we commence our analysis from the SD equation for the vertex ${\hat G}(p)$, which we solve upon ignoring the effects of the inhomogeneous $Z(p)$ term. Thus we arrive at the homogeneous equation \begin{equation} {\hat G}(p) = \frac{1}{3}e^2 \int \frac{d^3k}{(2\pi)^3} {\hat G}^3 (k) \frac{1}{k^2}\frac{1}{(k-p)^2} \label{ampapp} \end{equation} This integral equation involves only one unknown function, namely ${\hat G}$, which must be self-consistently determined. Note that this equation is invariant under the rescaling ${\hat G} \rightarrow {\hat G}/e$. This indicates a straightforward extension of the analysis to a large-$N$ treatment, given that $N$ can be absorbed in a redefinition of $e^2$. It is easy to see that, written in the form (\ref{ampapp}), the equation does not admit {\it physically acceptable} solutions, i.e. solutions with ${\hat G} \ge 0$ and {\it finite}~\footnote{Solutions that blow up in any point of the integration region are discarded.}. Indeed, setting $p =0$ one obtains after the (trivial) angular integration~\footnote{These arguments remain unaffected even in the presence of an inhomogeneous term $Z(p)$, such that $Z(0)$ is non-negative and finite. The non-negative nature of $Z(p)$ (\ref{zed}) stems from that of $A(p)$, which is guaranteed from general renormalization-group arguments~\cite{books}.}: \begin{equation} {\hat G}(0) = \frac{e^2}{12\pi^2} \int _0^\infty \frac{dk}{k^2} {\hat G}^3(k) \label{limito} \end{equation} Finiteness of ${\hat G}(0)$ requires that the integrand of the right hand side of (\ref{limito}) converges at $y \rightarrow 0~{\rm and}~ \infty$. The ultaviolet limit does not present a problem, because the kernel vanishes like $y^{-2}$, which is consistent with the superrenormalizability of the theory as well as the fact that the amputated vertex tends to 1. In the infrared limit $y \rightarrow 0$, however, the kernel blows up. For the integral to remain finite at that point, as required by the finiteness assumption for ${\hat G}(0)$, $G^3(y)$ must approach zero as $y^\alpha,~\alpha > 1/3$, thereby implying that ${\hat G}(0) =0$. However for that to happen the integrand in (\ref{limito}) must change sign, which would in turn imply that ${\hat G}(y)$ itself must change sign somewhere in $y$. According to our assumption above this is not a physically acceptable situation. To show rigorously that there are no {\it physically} acceptable solutions leading to a non-trivial infrared structure we next convert the integral equation into a non-linear differential equation of the type known in the mathematical literature as {\it Emden-Fowler} equation~\cite{emden,kamke}. To this end, we perform the angular integration in (\ref{ampapp}), to arrive at the equation: \begin{equation} {\hat G}(p) = \frac{2}{3\pi^2} \frac{\alpha}{p} \int _0^\infty \frac{dk}{k} {\hat G}(k)^3 {\rm ln}|\frac{k+p}{k-p}| \label{v2} \end{equation} where we have set $e^2 \equiv 8\alpha$ to make contact with the usual large-$N$ definition~\cite{app}. For us, however, the number of fermion flavours is not assumed to be necessarily large. In fact, for brevity we set $N=1$ (in a four-component notation for the fermions) throughout this work. Next we introduce the dimensionless variables $x \equiv p/\alpha$ and $y\equiv k/\alpha$. Since we are interested in the infrared behaviour of the model we consider the limit $x <<1 $, for which one obtains by expanding the logarithms in the integrand: \begin{eqnarray} &~& {\hat G}(x) =\frac{2}{3\pi^2 x } \int _0^\infty \frac{dy}{y} {\hat G}^3 (y) {\rm ln}|\frac{y+x}{y-x}| \simeq \nonumber \\ &~& \frac{4}{3\pi^2 x^2}\int _0^x dy {\hat G}(y)^3 + \frac{4}{3\pi^2 } \int _x^\infty \frac{dy}{y^2} {\hat G}^3 (y) \label{v4} \end{eqnarray} Differentiating appropriately with respect to $x$, we arrive at the following differential equation for small $x$: \begin{equation} x^3 \frac{d^2 {\hat G}}{d x^2} + 3 x^2 \frac{d {\hat G}}{d x} + \frac{8}{3\pi^2}{\hat G}^3(x) =0 \qquad x << 1 \label{fcaleq} \end{equation} It is convenient to rescale ${\hat G}$ by setting: \begin{equation} G \equiv \sqrt{\frac{8}{3\pi^2}} {\hat G} \label{definighat} \end{equation} Then, eq. (\ref{fcaleq}) becomes: \begin{equation} x^3 \frac{d^2 G}{d x^2} + 3 x^2 \frac{d G}{d x} + G^3(x) =0 , \qquad x << 1 \label{fcaleq2} \end{equation} Upon the change of variables $\xi = \frac{1}{2 x^2}$, $G=2^{3/4}~\eta(\xi)$, the equation becomes of Emden-Fowler type~\cite{emden,kamke}: \begin{equation} \frac{d^2}{d \xi^2} \eta (\xi) + \xi^{-3/2} \eta^3 (\xi) =0, \qquad \xi \rightarrow +\infty \label{emden2} \end{equation} As discussed in the mathematical literature~\cite{emden}, the {\it only non-trivial real solutions} of (\ref{emden2}), as $\xi \rightarrow +\infty$, are {\it oscillatory} about zero of simple sinusoidal form, which however oscillate infinitely rapidly as $x \rightarrow 0$. The amplitude of the above solutions behaves like $x^{-1/2}$, for $x << 1$. We interprete this behaviour as indicating an {\it instability} of the massless-fermion ground state. It is interesting to notice that the only power-law solution of (\ref{fcaleq2}) for $x << 1$, is {\it purely imaginary}, i.e. \begin{equation} G(x) = i\frac{\sqrt{5}}{2}x^{1/2}, ~\qquad x << 1 \label{fp} \end{equation} This solution would imply a `trivial infrared fixed point structure' given that its associated $\beta$ function vanishes at $x=0$. However, the fact that (\ref{fp}) is {\it purely imaginary } would again suggest instability. The above analysis constitutes a rigorous proof that, within the context of {\it proper} (i.e. finite and with finite-derivatives) solutions, and modulo the approximations discussed, no non-trivial infrared-fixed point is possible in $QED_3$ in the absence of an infrared cut-off. This was conjectured in ref. \cite{am}, but here we have given an analytic proof. This motivates one to look for the existence of a possible non-trivial infrared fixed-point structure in the presence of fermion and/or photon masses. In the next section we shall discuss the case when the fermions develop a mass $m$. As we shall show, the existence of a non-trivial infrared fixed point is guaranteed due to the form of the resulting equations. \setcounter{equation}{0} \section{Equation for the Vertex in the case of non-zero fermion mass} As a first kind of infrared cut-off in the integral equation (\ref{ampapp}) we shall consider the case of a fermion mass gap $m(p)=\frac{\Sigma (p)}{A(p)}$, where $\Sigma (p)$ is the fermion self energy. In that case the fermion propagator $S_F$ becomes: \begin{equation} S_F (k) = \frac{i}{A(k) \left( \nd{k} + m_f(k)\right)} \label{fermass} \end{equation} For our purposes below we assume that $m_f(p) \simeq m_f(0) \equiv m_f \ne 0$. In that case the integral equation (\ref{ampapp}) becomes: \begin{equation} {\hat G}(p) = \frac{1}{3}e^2 \int \frac{d^3k}{(2\pi)^3} {\hat G}^3 (k) \frac{1}{(k^2+ m_f^2 )(k-p)^2} + \frac{2m_f^2}{3} \int \frac{d^3k}{(2\pi)^3}{\hat G}^3 (k) \frac{1}{(k^2 + m_f^2)^2 (k-p)^2} \label{ineq2} \end{equation} Notice that the effects of the fermion mass are not given simply by just adding a mass squared term in the fermion denominators, but they result in additional structures in the integral equation. Performing the angular integrations one arrives at: \begin{equation} {\hat G}(x) = \frac{2}{3\pi^2 x}\int dy f(y) {\rm ln}\left|\frac{y+x}{y-x}\right|{\hat G}^3(y) \label{equntitl} \end{equation} where $x \equiv p/\alpha$, $m \equiv m_f/\alpha$ are dimensionless, and \begin{equation} f(y) \equiv y \frac{y^2 + 3m^2}{(y^2 + m^2)^2} \ge 0 \label{fdef} \end{equation} Differentiation with respect to $x$ yields: \begin{equation} x\frac{d}{d x}{\hat G}(x) =-\frac{2}{3\pi^2 x}\int _0^\infty dy f(y) \left( {\rm ln}\left|\frac{y+x}{y-x}\right|+\frac{2xy}{x^2-y^2}\right) {\hat G}^3(y) \label{kernel} \end{equation} One observes that formally as $x\rightarrow 0$ the right-hand-side vanishes, provided that ${\hat G}$ is finite. This indicates the existence of a fixed point. As we shall show below this is confirmed analytically by converting the integral equation into a non-linear differential equation. An additional feature which one would have hoped to study already at the level of the integral equation (\ref{kernel}) is the monotonicity of ${\hat G}$. Unfortunately the kernel in (\ref{kernel}) is not manifestly positive to allow for such an analytic proof at the level of the integral equation for generic values of $x$, and one has to resort to numerical treatments, which fall beyond the scope of this article. However one can already infer from (\ref{kernel}) that, for high momenta $x >> 1$, a monotonically decreasing ${\hat G}$ is consistent with the expectation that in this regime ${\hat G}$ is essentially given by its perturbative expression which asymptotes to $1$. The analysis is omitted because it is straightforward. For low momenta, on the other hand, the behaviour of ${\hat G}(x)$ will also be shown to be monotonically decreasing, starting from a non trivial fixed point. This will be achieved by converting the integral equation into a differential one. Unfortunately, at present, we cannot analytically derive the monotonicity for intermediate momenta. To derive the differential equation from (\ref{equntitl}) we follow a similar analysis to the one leading to (\ref{v4}). First, one expands the logarithms for small $x << 1$, thereby writing the equation as: \begin{eqnarray} &~& {\hat G}(x) \simeq \frac{4}{3\pi^2 x^2}\int _0^x dy {\hat G}(y)^3 \frac{y^2}{y^2 + m^2} + \frac{4}{3\pi^2 } \int _x^\infty \frac{dy}{y^2 + m^2} {\hat G}^3 (y) + \nonumber \\ &~& \frac{8m^2}{3\pi^2 x^2}\int _0^x dy {\hat G}(y)^3 \frac{y^2}{(y^2 + m^2)^2} + \frac{8m^2}{3\pi^2 } \int _x^\infty \frac{dy}{(y^2 + m^2)^2} {\hat G}^3 (y) \label{eq2} \end{eqnarray} Differentiating with respect to $x$ one arrives at: \begin{equation} x(x^2 + m^2)^2 \frac{d^2}{d x^2}{\hat G}(x) + 3(x^2 + m^2)^2\frac{d}{dx}{\hat G}(x) + \frac{8}{3\pi^2}(x^2 + 3m^2){\hat G}^3(x) =0 \label{emdenmass} \end{equation} The above equation can be solved numerically, to which we will come later on. However, in the infrared region $x << m$ the equation accepts an analytic treatment, as we discuss below. In this region the equation (\ref{emdenmass}) is approximated by: \begin{equation} x \frac{d^2}{dx^2}{\hat G}(x) + 3 \frac{d}{d x}{\hat G}(x) + \frac{8}{\pi^2m^2}{\hat G}^3(x)=0 \label{papmav} \end{equation} It is immediate to see that a special power-law solution is given by (for positive $G(x)$) by: \begin{equation} {\hat G}(x) =m \pi \frac{\sqrt{3}}{4\sqrt{2}}x^{-1/2} \label{ss} \end{equation} Notice the infrared divergence of this type of solutions {\it even} in the presence of a (bare) fermion mass. The associated renormalization-group $\beta$ function (\ref{beta}) for this case reads: \begin{equation} \beta (x) = -\frac{1}{2}{\hat G} \sim x^{-1/2} \rightarrow +\infty,~{\rm as}~x \rightarrow 0 \label{beta2} \end{equation} indicating the absence of an infrared fixed point. The associated operator appears to be {\it relevant} (negative scaling dimension), which implies the possibility that the theory be driven to a non-trivial fixed point. However, in the infrared regime $x << 1$, one can find a different type of solution~\cite{emden}: \begin{equation} {\hat G} = m \pi \frac{\sqrt{3}}{2\sqrt{2}}\frac{c}{1 + c^2 x}, \qquad x \rightarrow 0 \label{es} \end{equation} where $c$ is a constant of integration to be fixed by the boundary condition at $x=0$ implied by the integral equation, to be discussed later on. For physical solutions $c$ is assumed positive. This type of solutions has a renormalization-group $\beta$-function (\ref{beta}) of the form: \begin{equation} \beta = -{\hat G}(x) + \frac{2\sqrt{2}}{\sqrt{3}\pi m c}{\hat G}^2(x) \sim -\frac{x}{\left(1 + c^2 x\right)^2} \rightarrow 0, \qquad x \rightarrow 0 \label{beta3} \end{equation} from which we observe the existence of a non-trivial (non-perturbative) infrared fixed point at ${\hat G}^* = \frac{\pi m \sqrt{3}c}{2\sqrt{2}} >0$. Such a fixed point is the result of the dynamical generation of a parity-invariant, chiral-symmetry breaking fermion mass~\cite{app}, indicating the connection of the phenomenon of chiral symmetry breaking in $QED_3$ to a non-trivial infrared fixed point structure, in agreement with the expectations of ref. \cite{am}. The non-trivial fixed-point solution (\ref{es}) {\it is not} compatible with the integral equation (\ref{ineq2}) for {\it any} value of the fermion mass $m$. Indeed one can derive a {\it boundary condition} for ${\hat G}(0)$ from (\ref{ineq2}), which reads: \begin{equation} {\hat G}(0) = \frac{4}{3\pi^2}\int _0^\infty dy \frac{1}{y^2 + m^2} {\hat G}^3(y) + \frac{8m^2}{3\pi^2}\int _0^\infty dy \frac{1}{(y^2 + m^2)^2}{\hat G}^3(y) \label{bc2} \end{equation} In contrast to the massless case (\ref{limito}), ${\hat G}(0)$ is now a finite constant, $\pi m c \sqrt{3/8}$, as seen from (\ref{es}), and this allows for a compatibility of the solution (\ref{es}) with (\ref{bc2}), {\it provided} that $m {\hat G}(0)$ satisfies certain conditions to be specified below. To this end, we split the $y$-integration in (\ref{bc2}) into two intervals: (i) $y \in [0, m)$, where the form (\ref{es}) is valid to a good approximation, and (ii) $y \in [m, \infty)$, where we approximate ${\hat G}(x)$ by its perturbative asymptotic value ${\hat G} \simeq 1$. By this latter approximation we overestimate the actual value of the integration, given that ${\hat G}$ is actually slightly smaller than unity for finite $p$, approaching it only asymptotically (c.f. (\ref{ascond})). However, this is sufficient for our qualitative purposes of demonstrating the existence of constraints on the fermion mass implied by the boundary condition (\ref{bc2}). With these in mind, the boundary condition (\ref{bc2}) reduces to: \begin{equation} 1 = \frac{mc^2}{2}\int _0^1 dy \frac{1}{y^2 + 1} \frac{1}{(1 + mc^2 y)^3} \left( 1 + \frac{2}{y^2 + 1}\right) + \frac{4}{3\pi^2 m^2c}\left(1 - \frac{1}{\pi}\right)\sqrt{\frac{2}{3}} \label{gap} \end{equation} To obtain the condition imposed on $m$ by the boundary condition (\ref{bc2}) it suffices to observe that the first term on the right-hand-side is a function of $mc^2$ alone, and that, after the (elementary) $y$ integration, the resulting function of $mc^2$ asymptotes rapidly to the value $3/4$ (see figure \ref{fig2}). This implies the following inequality: \begin{equation} m {\hat G}(0) < \frac{8}{3\pi}\left(1 - \frac{1}{\pi}\right) \label{cond} \end{equation} As already mentioned this bound is overestimated, given that in the actual situation the function ${\hat G}(y)$ is not exactly $1$ immediately after the region $y \ge m$. \begin{figure}[htb] \epsfxsize=3in \bigskip \centerline{\epsffile{integr.eps}} \caption{\it\baselineskip=12pt Plot of the function $f(z)=\frac{z}{2}\int _0^1 dy \frac{1}{y^2 + 1} \frac{1}{(1 + z y)^3} \left( 1 + \frac{2}{y^2 + 1}\right)$, where $z=mc^2$. The function asymptotes rapidly to $3/4$.} \bigskip \label{fig2}\end{figure} We next remark that $m$ should actually be determined self-consistently from a solution of the pertinent gap equation. This will be done in section 5. However, at the moment, and for completeness, we shall assume that $m$ is determined by its approximate form derived within the context of a large-$N$ treatment~\cite{app,dor}. Compatibility of the dynamical solution with the constraint (\ref{cond}) will then lead to further restrictions on the range of the allowed masses $m$. As we shall see in section 5, there is good agreement between the allowed fermion-mass ranges, obtained within the context of a large-$N$ treatment, and those obtained from a self-consistent solution of the mass gap equation within our approach. In the context of a large $N$ treatment, and to leading order in $1/N$ resummation, the following solution for the dynamically-generated $m$ is found~\cite{app,dor}, \begin{equation} m \sim {\cal O}(1) {\rm exp}\left(-\frac{2\pi}{\sqrt{\frac{g^2}{g_c^2}-1}}\right) \label{appelquist} \end{equation} where $g_c^2 =\frac{\pi^2}{32}$ is the critical coupling, above which dynamical mass generation occurs~\cite{app}. Compatibility of the solution (\ref{appelquist}) with the constraint (\ref{cond}) implies the existence of an {\it upper bound} on fermion masses, $m < m_{max}$, where $m_{max}$ is defined through the intersection of the curves (\ref{cond}) and (\ref{appelquist}) in the $\left(m,g\right)$ plane (see fig. \ref{fig3}). This yields $m_{max} \simeq 0.3$. \begin{figure}[htb] \epsfxsize=3in \bigskip \centerline{\epsffile{appel.eps}} \caption{\it\baselineskip=12pt Fermion mass versus the infrared-value of the coupling ${\hat G}(0)$. The solid curve represents the condition derived from the integral equation for the vertex, whereas the dashed line represents the solution obtained from the standard gap equation in the large-N treatment.} \bigskip \label{fig3}\end{figure} On the other hand, for large momenta, we know that ${\hat G} \rightarrow 1$. Physically one expects {\it a monotonic decrease} of ${\hat G}(x)$ over the {\it entire} range of $x \in [0, \infty)$. This would occur in our case if and only if ${\hat G}(0) > 1$, which, in the context of the large-$N$ result of \cite{app,dor}, implies a {\it minimum} bound for the fermion masses $m > m_{min}\simeq 0.03$. Actually, as we shall argue in the next section, ${\hat G}(0)$ should be comfortably larger than $\sqrt{3/2}$ for self-consistency of our approximations. Hence, we see that the monotonicity of the running coupling can be achieved in the context of a large-$N$ treatment, if the mass $m$ lies in the following regime: \begin{equation} 0.03 \mathrel{\rlap {\raise.5ex\hbox{$ < $} m \mathrel{\rlap {\raise.5ex\hbox{$ < $} 0.3 \label{zone} \end{equation} or equivalently if the coupling at the infrared point ${\hat G}(0)$ is restricted in the regime (see fig. \ref{fig2}): \begin{equation} 1 < {\hat G}(0) < 2.5,\qquad \label{regionc} \end{equation} At this point it is useful to turn to a numerical study of the equation (\ref{emdenmass}), supplemented with the boundary conditions imposed by the solutions of the form (\ref{es}), specifically: \begin{equation} {\hat G}(0)=m\frac{\pi}{2}\sqrt{\frac{3}{2}}c \qquad; \qquad {\hat G}'(0)=-\frac{8 {\hat G}^3(0)}{3\pi^2 m^2} \label{bc3} \end{equation} where the constant $c$ should be chosen in such a way that the bound (\ref{cond}) be satisfied. The numerical solution of (\ref{emdenmass}) is given in figure \ref{fig4}, for a typical case, where $m=0.1$, and ${\hat G}(0)=2$. As we observe, the figure clearly demonstrates {\it the monotonic decrease} of ${\hat G}$ in the small $x$ interval, where equation (\ref{emdenmass}) is valid. We also note that the solution asymptotes quickly to a constant positive value, smaller than $1$ in contrast to (\ref{ascond}), which would require ${\hat G}(x) \sim 1$, for $x >> 1$. This, however, presents no contradiction, because equation (\ref{emdenmass}) was based on ignoring the inhomogeneous term $Z(x)$. As one goes to a region of larger $x$, this assumption is no longer valid. In that case the inhomogeneous term $Z(x)$ becomes important, and equation (\ref{emdenmass}) receives additive (positive) contributions, resulting in ${\hat G}(p) \sim 1$, for large $p$, as required by (\ref{ascond}). \begin{figure}[htb] \epsfxsize=3in \bigskip \centerline{\epsffile{asympt.eps}} \caption{\it\baselineskip=12pt Numerical solution of the equation (\ref{emdenmass}) versus $p/\alpha$, for a typical set of values $m \sim 0.1$ and ${\hat G}(0)=2$; the solution decreases monotonically and asymptotes quickly to a positive constant value.} \bigskip \label{fig4}\end{figure} \setcounter{equation}{0} \section{Equations for Photon and Fermion self-energies} As a consistency check of our assumption about ignoring the effects of the inhomogeneous term $Z(p)$ as $p \rightarrow 0$, next to ${\hat G}$, we now turn our attention to the equations that determine the fermion and photon self-energies $A(p)$ and $B(p)$ respectively. As shown below, upon neglecting $Z(p)$ in the infrared, these equations decouple from each other and both functions can be determined solely from knowledge of ${\hat G}(x)$. It should also be stressed that in the massless case the system does not admit a self-consistent solution. In contrast, the presence a fermion mass term changes the situation drastically by yielding self-consistent solutions for $A(p),B(p)$ which are such that $Z(p) \rightarrow 3\sqrt{6}/5$, but at a rate slower than the one with which ${\hat G}$ approaches ${\hat G}(0)$ as $x \rightarrow 0$. Thus the approximation of ignoring the effects of $Z(x)$ in the region $x << 1$ is qualitatively correct. This is a highly non-trivial check of our approach, and justifies fully the approximations used above. To this end, we begin from the integral equation for $A(x)$, which reads: \begin{equation} A(p)~\nd{p} = \nd{p} - e^2\int \frac{d^3k}{(2\pi)^3}{\cal G}^2(k) \gamma _\mu \frac{i}{A(k)\nd{k}}\gamma _\nu \frac{\delta _{\mu\nu}}{B(k-p)(k-p)^2} \label{fermiself} \end{equation} which in terms of the semi-amputated vertex ${\hat G}$ becomes: \begin{equation} A(p)~\nd{p} = \nd{p} - e^2A(p)\int \frac{d^3k}{(2\pi)^3}{\hat G}^2(k) \frac{\nd{k}}{k^2(k-p)^2} \label{fermiself2} \end{equation} To arrive at the above equation we have carried out the $\gamma$-matrix algebra and we have used the approximation that ${\hat G}(k,p) ={\hat G}(k)$. This approximation is equivalent to the one made for ${\hat G}$ in (\ref{SDamp}). It will be justified below, by the self-consistency of the solution. In terms of the dimensionless $x,m$ parameters introduced previously, the equation (\ref{fermiself2}) becomes: \begin{equation} x^4=A^{-1}(x)x^4 -\frac{2}{\pi^2}\int _0^x dy \frac{y^4 {\hat G}^2 (y)}{y^2 + m^2} - \frac{2 x^4}{\pi^2}\int _x^1 dy \frac{{\hat G}^2 (y)}{y^2 + m^2} \label{eqwf} \end{equation} Differentiating twice with respect to $x$ one obtains the following differential equation: \begin{equation} \frac{d}{dx}\left[x^{-3} \frac{d}{dx}\left(x^4 A^{-1}(x)\right)\right] + \frac{8}{\pi^2}\frac{{\hat G}^2 (x)}{x^2 + m^2} = 0 \label{eqdiffwf} \end{equation} Restricting our attention to the infrared region $x \rightarrow 0$, one may ignore $x^2 $ in front of $m^2$, and use the form (\ref{es}) for ${\hat G}(x)$. After elementary integrations one then finds for $A(x)^{-1}$: \begin{eqnarray} &~&A^{-1}(x) =\frac{3}{2}\left[-\frac{2}{c^7}x^{-3} + \frac{1}{c^5}x^{-2} - \frac{2}{3c^3}x^{-1} + \frac{2}{c^9}x^{-4}{\rm ln}\left(1 + c^2 x \right)\right] + \frac{1}{4}c' \simeq \nonumber \\ &~&\frac{1}{4}(c'-\frac{3}{c}) + \frac{3}{5}cx -\frac{1}{2}c^3x^2 + {\cal O}(x^5), \qquad x \rightarrow 0 \label{awf} \end{eqnarray} where $c>0$ is the integration constant of the solution (\ref{es}), which obeys the restriction (\ref{cond}). We choose the constant of integration $c'=3/c$, so that \begin{equation} A(x)^{-1} \simeq \frac{3}{5}c x - \frac{1}{2}c^3 x^2 + \dots \label{awff} \end{equation} as $x \rightarrow 0$. As we shall show later, this choice is compatible with the boundary behaviour $Z(x) \rightarrow {\rm const}$, as $x \rightarrow 0$, advocated for the inhomogeneous term. Next we turn our attention to the equation for the photon function $B(p)$, in the presence of a fermion mass. Following similar steps to the ones leading to (\ref{fermiself2}), the pertinent integral equation reads: \begin{equation} B(p)\left(p^2 \delta _{\mu\nu} - p_\mu p_\nu \right)= p^2 \delta _{\mu\nu} - p_\mu p_\nu +B(p)e^2 \int \frac{d^3k}{(2\pi)^3} {\hat G}^2(k){\rm Tr}\left[\gamma _\mu \frac{\nd{k}}{k^2} \gamma _\nu \frac{\nd{k}-\nd{p}}{(k-p)^2}\right] \label{photonself2} \end{equation} To trasnform the above equation into a scalar equation for $B(p)$ it is necessary to take the trace on both sides. In doing so one assumes that the integral term on the right-hand-side is also transverse. This is expected from gauge invariance. In our truncated scheme it can be demonstrated that this is indeed the case, provided one makes our working hypothesis that ${\hat G}(k,p)={\hat G}(k)$. Specifically if one contracts both sides of (\ref{photonself2}) by $p^\mu$, the right-hand-side vanishes (Ward identity for the photon) only upon making the above assumption for the semi-amputated vertex, and shifting appropriately the integration variable. After standard manipulations, similar to the ones above, one arrives at: \begin{equation} \frac{d}{dx}\left[x^{-1}\frac{d}{dx}\left(x^{-1}\frac{d}{dx}\left(x^4 B^{-1}\right)\right)\right] + \frac{32}{\pi^2}\frac{1}{x^2 + m^2}{\hat G}^2 (x) = 0, \qquad x \rightarrow 0 \label{photonwf} \end{equation} Making the same approximations in the infrared $x << 1$ as in the case of $A(x)$, and substituting the Emden's solution (\ref{es}) for ${\hat G}$ on the right-hand-side of (\ref{photonwf}), we obtain after some elementary integrations, upon setting their constants but one ($c_1$) to zero: \begin{equation} x^4 B^{-1}(x)=\frac{6}{c^2}x^2 -\frac{6}{c^6}x + \frac{3}{c^4}x^2 + \frac{6}{c^8}{\rm ln}\left(1 + c^2 x\right) - \frac{6x^2}{c^4} {\rm ln}\left(1 + c^2 x\right) + \frac{1}{8}c_1 x^4 \label{betacapital} \end{equation} which for $x << 1$ yields: \begin{equation} B^{-1}(x) \simeq \frac{6}{c^2}x^{-2} \left[1 - \frac{2}{3} x + \left(\frac{c_1}{8}+ \frac{3}{2}\right)\frac{c^2}{6} x^2 + \dots \right] \label{photonwff} \end{equation} This form of $B(x)$ gives rise to the following form for the photon propagator in the infrared $k \rightarrow 0$, \begin{equation} \Delta _{\mu\nu} \sim \frac{\delta _{\mu\nu}}{k^4} , \qquad k \rightarrow 0 \label{photinfr} \end{equation} Notice, as a consistency check, that the photon continues to be massless in the chiral-symmetry broken phase. The corresponding static effective potential is given by the appropriate Fourier transform of the $00$ (temporal) component of the photon propagator for $k_0=0$. In the case (\ref{photinfr}) this yields formally an effective potential scaling like $R^2$ for large distances $R$, suggestive of confining behaviour~\footnote{However, this is a formal result, given that the corresponding momentum integrals are infrared divergent and hence need proper regularization, which falls beyond our scope here. For standard treatments see discussions in \cite{conf}.}. The fact that this behaviour is found in the chiral-symmetry broken case is consistent with general (four-dimensional) arguments~\cite{conf,pc} that confinement is a sufficient but not necessary condition for chiral-symmetry breaking. An important feature of the expressions (\ref{awff}) and (\ref{photonwf}) is that, although they are derived only in case where a fermion mass $m \ne 0$, however they do not explicitly depend on the magnitude of the mass. From (\ref{awff}) and (\ref{photonwff}) one obtains for $Z(x)$ in the infrared region $x \rightarrow 0$: \begin{equation} Z(x) =B^{-1/2}(x)A^{-1}(x) \simeq \frac{3\sqrt{6}}{5}-\frac{2\sqrt{6}}{5}x \left( 1 + \frac{5}{4}c^2\right) + {\cal O}(x^3)\dots, \qquad x \rightarrow 0 \label{vanishingZ} \end{equation} where we have chosen the constant of integration $c_1$ such that there are no ${\cal O}(x^2)$ terms. In order to check under which conditions the omission of the $x$-dependent parts of $Z$ next to those of ${\hat G}(x)$ is justified, we must compare the corresponding linear terms in $x$ of both quantities. Using that \begin{equation} {\hat G}(x) \sim \sqrt{\frac{3}{2}}\left(m\frac{\pi}{2}c - m\frac{\pi}{2}c^3x + \dots\right), \qquad x \rightarrow 0 \label{eszero} \end{equation} this requires: \begin{equation} \sqrt{8/3} {\hat G}^3(0) - 2 {\hat G}^2 (0) - (3\pi^2/5)m^2 >> 0 \label{cond2} \end{equation} This inequality furnishes a condition on ${\hat G}(0)$ under the assumption $m << 1$, which, as we shall see in section 5, turns out to be correct. The condition is \begin{equation} {\hat G}(0) >> \sqrt{\frac{3}{2}} \simeq 1.22 \label{limit} \end{equation} In the above formulas the symbol $>>$ means actually at least an order of magnitude. As we shall see below, such a regime for ${\hat G}(0)$ arises self-consistently. Note that the condition (\ref{limit}) is in agreement with (\ref{ascond}) given that $Z(0)=3\sqrt{6}/5 \simeq 1.47$ (c.f. Eq. (\ref{vanishingZ})) is only slighty larger than $\sqrt{3/2}$. In view of the existence of a non-zero constant value of $Z(x)$ in the infrared region, our analysis on the restrictions (\ref{cond}),(\ref{regionc}) has to be repeated, given that these restrictions stem from the integral equation. From (\ref{cond2}), and the value $Z(0)=3\sqrt{6}/5$ it becomes clear that the boundary condition (\ref{gap}) is now modified to: \begin{equation} 1 = \frac{12}{5\pi mc} + \frac{mc^2}{2}\int _0^1 dy \frac{1}{y^2 + 1} \frac{1}{(1 + mc^2 y)^3} \left( 1 + \frac{2}{y^2 + 1}\right) + \frac{4}{3\pi^2 m^2c}\left(1 - \frac{1}{\pi}\right)\sqrt{\frac{2}{3}} \label{gap2} \end{equation} The first term continues to asymptote to $3/4$, but now one obtains the following restrictions on the coupling: \begin{equation} {\hat G}(0) < \frac{12\sqrt{6}}{5} + \frac{8}{3}\left(1 - \frac{1}{\pi}\right)\frac{1}{\pi m} \label{newcond} \end{equation} From the above relation it becomes clear that, if ${\hat G}(0) < 12\sqrt{6}/5$ there is no restriction on $m$. However, for this regime of the couplings the condition (\ref{limit}), necessary for safely ignoring the effects of the inhomogeneous term $Z(p)$ in the infrared next to ${\hat G}(p)$, is only marginally satisfied. On the other hand, if one allows ${\hat G}(0)$ to exceed this value, which is physically acceptable, then the following upper bound on $m$ as a function of ${\hat G}(0)$ is obtained: \begin{equation} m < \frac{\frac{8}{3\pi}\left(1 - \frac{1}{\pi}\right)} {{\hat G}(0) - \frac{12\sqrt{6}}{5}}, \qquad {\hat G}(0) > \frac{12\sqrt{6}}{5} \simeq 5.88 \label{fincond} \end{equation} which replaces (\ref{cond}). One may repeat the comparison with the large $N$ limit results~\cite{app}, as in the previous section, to determine the new region of allowed values of the mass. The analysis is given in figure (\ref{app2}). \begin{figure}[htb] \epsfxsize=3in \bigskip \centerline{\epsffile{appel2.eps}} \caption{\it\baselineskip=12pt Fermion mass versus the infrared-value of the coupling ${\hat G}(0)$ for the case where the effects of the inhomogeneous term (almost constant) $Z(x)$ have been taken into account in the infrared. The solid curve represents the condition derived from the integral equation for the vertex, whereas the dashed line represents the solution obtained from the standard gap equation in the large-N treatment.} \bigskip \label{app2}\end{figure} From fig. \ref{app2} we then obtain the following allowed range of the coupling ${\hat G}(0)$: \begin{equation} 5.88 < {\hat G}(0) < 7 \label{regionc2} \end{equation} to be compared with (\ref{regionc}) associated with the monotonicity requirement for ${\hat G}(p)$. This analysis shows that the new allowed region for the fermion masses, in case (\ref{fincond}) is: \begin{equation} 0.55 < m < 0.6 \label{zone2} \end{equation} where the lowest bound is the value of $m$ corresponding to ${\hat G}=5.88$ from the large-$N$-analysis formula (\ref{appelquist}). The region (\ref{zone2}) is to be compared with (\ref{zone}). As we shall see in section 5, the dynamically generated mass in our scenario, which does not resort to a large-$N$ treatment, lowers significantly the upper bound in (\ref{zone2}). It is important to emphasize that, as a consequence of (\ref{awf}),(\ref{photonwff}), the non-amputated vertex ${\cal G}(p)$, defined in (\ref{vertexgen}), approaches a non-zero {\it finite} constant value in the infrared $p=0$. Notice that this constant behaviour is compatible with the Ward identity in the infrared limit $p=0$, thereby implying that the information obtained about a non-trivial infrared fixed-point structure is gauge invariant, as it should be. This completes our analysis of including the effects of an almost constant $Z(x)\simeq Z(0)$ in the infrared. The above considerations justify retrospectively the omission of $Z$ in deriving the differential equation (\ref{emdenmass}). Indeed, a constant shift of ${\hat G}$ inside the derivative operators in that equation, which would account for a constant $Z$, does not affect its form. These results, therefore, constitute a highly non-trivial check of our approach. \setcounter{equation}{0} \section{Dynamical Derivation of the Fermion Mass Gap} In the previous section we have assumed the presence of a finite fermion mass, which we have treated effectively as an arbitrary parameter of the model. In this section we turn to the full problem, and study the dynamical generation of this mass, by deriving it self-consistently from the corresponding SD mass-gap equation. The equation for the gap $\Sigma (p)$ is derived from the graphs of fig. \ref{fig1b}, which yields: \begin{equation} A(p)\nd{p} + \Sigma (p) = \nd{p} + A(p)e^2 \int \frac{d^3k}{(2\pi)^3} {\hat G}(k)^2 \gamma _\mu \frac{1}{\nd{k} + M(k)}\gamma ^\mu \frac{1}{(k-p)^2} \label{massgap} \end{equation} where $M(k) \equiv \Sigma (k)/A(k)$ is the mass function, and we have pulled out factors of $A(p)$ appropriately so as to be able to define an amputated vertex function ${\hat G}(k)$. The consistency of the approach will be demonstrated below by the existence of solutions. Taking the trace in the above equation, and performing the angular integration, one arrives easily at: \begin{eqnarray} &~&M(p) = 3e^2 \int \frac{d^3k}{(2\pi)^3}{\hat G}^2(k) \frac{M(k)}{k^2 + M^2(k)} \frac{1}{(k-p)^2} =\nonumber \\ &~&\frac{6\alpha}{\pi^2 p}\int dk \frac{k}{k^2 + M^2(k)}{\hat G}(k)^2 M(k){\rm ln}\left|\frac{k + p}{k - p}\right| \label{tracem} \end{eqnarray} From the middle equation (\ref{tracem}) it becomes clear that the quantity $g_R (k) \equiv e {\hat G}(k)$, defined in (\ref{runcoupl}), plays indeed the r\^ole of a ``running coupling''. The situation is analogous, but not identical, to that of \cite{am}, where a ``running coupling'' has also been obtained from the gap equation in the context of a large $N$ analysis. However, in that case, one assumed an Ansatz for the vertex, satisfying a truncated form of the Ward identities. Instead, in our approach there was no necessity for a vertex Ansatz, since the non-perturbative vertex function was determined self-consistently from the SD equations. As a consequence, the running coupling (\ref{runcoupl}) is constructed out of the amputated vertex function alone; the latter is a manifestly gauge-independent quantity, at least perturbatively. Passing into dimensionless variables, in units of $\alpha=e^2/8$, ${\tilde M}\equiv M(k)/\alpha$, $x\equiv p/\alpha$, $y \equiv k/\alpha$, and working in the regime of low momenta $x << 1$ as usual, one obtains after some straightforward algebra, involving a truncated expansion of the logarithmic kernel: \begin{equation} {\tilde M}(x) =\frac{12}{\pi^2} \left[ \frac{1}{x^2}\int _0^x dy \frac{y^2 {\tilde M}(y)}{y^2 + {\tilde M}^2(y)}{\hat G}^2(y) + \int _x^\infty dy \frac{{\tilde M}(y)}{y^2 + {\tilde M}^2(y)}{\hat G}^2(y)\right] \label{tildemeq} \end{equation} Differentiating twice with respect to $x$ one arrives at: \begin{equation} x \frac{d^2}{d x^2} {\tilde M}(x) + 3 \frac{d}{dx}{\tilde M}(x) + \frac{24}{\pi^2}\frac{{\tilde M}(x)}{x^2 + {\tilde M}^2(x)}{\hat G}(x)^2 = 0 \label{gapeq} \end{equation} Given that dynamical mass generation is expected to be an infrared phenomenon, we now restrict ourselves to the region~\footnote{Note that this is the opposite limit than the one usually considered in the framework of large-$N$ analysis of dynamical mass generation, where one arrives at a gap equation by making the assumption $\alpha >> p >> m$ for the momenta $p$ of the pertinent excitations.}, \begin{equation} x^2 << {\tilde M}^2 << 1 \label{condmx} \end{equation} In this region we neglect $x^2$ next to ${\tilde M}^2$ in (\ref{gapeq}) and use the solution (\ref{es}) for ${\hat G}(x) \simeq {\tilde M}\sqrt{\frac{3}{8}}\pi c$ as $x \rightarrow 0$. The result is: \begin{equation} x \frac{d^2}{d x^2} {\tilde M}(x) + 3 \frac{d}{dx}{\tilde M}(x) + 9c^2{\tilde M}(x)= 0, \qquad x \rightarrow 0 \label{bessel} \end{equation} Changing variables $x \rightarrow \xi = x^{-1}$, the equation reduces to a Bessel equation~\cite{kamke} \begin{equation} \xi^2 \frac{d^2}{d \xi^2} {\tilde M}(\xi) - \xi \frac{d}{d\xi}{\tilde M}(\xi) + 9c^2 \xi^{-1} {\tilde M}(\xi)= 0, \qquad \xi \rightarrow \infty \label{bessel2} \end{equation} with (formally) general solution: \begin{equation} {\tilde M}(\xi) = C_1 \xi J_{-2}(-6c \xi^{-1/2}) + C_2 \xi Y_{-2}(-6c \xi^{-1/2}) \label{solutionbess} \end{equation} where $C_i,i=1,2$ are arbitrary constants, $J_{-n} (x) = (-1)^n J_{n}(x)$, $n = 1,2,3, \dots$, is a Bessel function of the first kind, and $Y_{n}(x)$ is a generalized Bessel function of the second kind~\cite{kamke}. The latter is only defined for positive $n$ and this imposes the choice $C_2=0$ in (\ref{solutionbess}). Expressing the solution in terms of $x$, $x\rightarrow 0$, one has the following power series expression for the dynamical mass: \begin{equation} {\tilde M}(x) = C_1 x^{-1} \sum _{n=0}^{\infty} (-1)^n \left(3c\right)^{2n + 2} x^{1 + n} \frac{1}{ n! \Gamma (3 + n)} \simeq \frac{9}{2} C_1 c^2 + {\cal O}(x), \qquad x \rightarrow 0 \label{fermionmass} \end{equation} From this one obtains the following relation between ${\hat G}(0)$ and ${\tilde M}(0)\equiv m_f /\alpha $: \begin{equation} {\tilde M}(0) \equiv m_f/\alpha = \left(\frac{12}{\pi^2}\right)^{1/3}C_1^{1/3}{\hat G}^{2/3}(0) \label{appelgm} \end{equation} This result is to be compared the result (\ref{appelquist}) within the context of a large-$N$ analysis. In particular, at first sight it seems that the relation (\ref{appelgm}) does not have a critical coupling, above which dynamical mass generation occurs. However, because the result (\ref{appelgm}) has been derived in the context of the solution (\ref{es}), one should bear in mind the restrictions characterizing that situation, in particular (\ref{newcond}). This implies an appropriate restriction for $C_1$~\footnote{The restriction (\ref{newcond}) is to be viewed as a boundary condition. We remind the reader that the requirement of finiteness of ${\tilde M}$ had already fixed the other constant $C_2$ to zero.}, \begin{equation} {\hat G}^{5/3}(0) - \frac{12\sqrt{6}}{5}{\hat G}^{2/3}(0) - \frac{8}{3 \left(12 \pi C_1\right)^{1/3}}\left(1 - \frac{1}{\pi}\right) < 0 , \qquad {\hat G}(0) > \frac{12\sqrt{6}}{5} \label{c1} \end{equation} This restriction implies a critical coupling, ${\hat G}_c = 12\sqrt{6}/5 \simeq 5.88$ but it is derived in a way independent of any large-$N$ analysis. The way to understand (\ref{c1}) is the following: one should first fix a range of ${\hat G}(0)$, with ${\hat G}(0) > 5.88$, and then use a $C_1$ that will be such that, within this range of the couplings, Eq. (\ref{c1}) is satisfied for masses ${\tilde m} << 1$. As can be readily seen, the bound for $C_1$ obtained from the requirement that $m <<1$ is far less restrictive than the one associated with (\ref{c1}), provided ${\hat G}(0)$ is not too close to the critical ${\hat G}_c$, where the mass $m$ vanishes. For instance, for ${\hat G}(0) ={\cal O}(8)$, the upper bound on $C_1 $ from (\ref{c1}) is of order ${\cal O}(10^{-4})$, while for ${\hat G}(0) = 6$ the upper bound is $C_1 < 4$. Notice that the bound is very sensitive to small changes in ${\hat G}(0)$. A typical situation is depicted in fig. \ref{gmpm} for two values of $C_1 =10^{-5}, 10^{-2}$. We observe that the case $C_1=10^{-2}$ yields an upper bound in the mass which is of order $0.8$ and hence should be discarded on the basis that it is not small enough. On the other hand, the value $C_1=10^{-5}$ yields an acceptable upper bound $ m \sim 0.1$. In that case, from fig. \ref{gmpm}, we observe that the allowed region of $m$ is \begin{equation} 0.08 \mathrel{\rlap {\raise.5ex\hbox{$ < $} m \mathrel{\rlap {\raise.5ex\hbox{$ < $} 0.12 \label{zone4} \end{equation} to be compared with (\ref{zone2}), derived using the result for the mass in the context of a large-$N$ analysis~\cite{app}. The corresponding regime of the couplings ${\hat G}(0)$ is: \begin{equation} 5.88 < {\hat G}(0) < 11 \label{regionc4} \end{equation} \begin{figure}[htb] \epsfxsize=3in \bigskip \centerline{\epsffile{appel4.eps}} \caption{\it\baselineskip=12pt Fermion mass versus the infrared-value of the coupling ${\hat G}(0)$ using (\ref{appelgm}) (dashed curves), for two values of $C_1=10^{-5}$ (lower dashed curve) and $C_1=10^{-2}$ (upper dashed curve). The continuous curve is (\ref{fincond}), viewed as a boundary condition. The value $C_1 =10^{-2}$ should be excluded on grounds of yielding too high values of the mass ${\tilde m}$.} \bigskip \label{gmpm}\end{figure} Before closing this section we would like to discuss possible applications of the above behaviour; specifically, the restriction (\ref{regionc4}) of the allowed values of ${\hat G}(0)$ may have interesting applications in case the above model turns out to describe the physics of high-temperature superconductors. We remind the reader that in such models dynamical mass generation coincides with superconductivity~\cite{dor}-\cite{fm}. In that case, the fermion fields (`electrons') of $QED_3$ represent `holons', i.e. electrically-charged excitations of fermionic statistics, which are constituents of the physical electron. The latter is believed to exhibit an effective spin-charge separation~\cite{anderson} in the complicated ground-state of high-temperature superconductors. The `photon' of $QED_3$ then represents an effective Heisenberg spin-spin antiferromagnetic interaction, responsible for binding the holons in Cooper-like pairs. In some models~\cite{dor,fm} the effective (gauge-invariant) coupling ${\hat G}(0)$ may be expressed in terms of the parameters of the microscopic condensed-matter lattice systems, whose long-wavelength limit is equivalent to the above $QED_3$ model, as: \begin{equation} {\hat G}(0)^2 \sim \frac{J}{e^2} (1-\eta) \label{dopheis} \end{equation} where $\eta$ expresses the concentration of impurities in the system (doping), and $J$ denotes the Heisenberg (antiferromagnetic) exchange energy. Hence, on account of (\ref{regionc4}), Eq. (\ref{dopheis}) implies that $ 6 \mathrel{\rlap {\raise.5ex\hbox{$ < $} (J/e^2)(1 - \eta) \mathrel{\rlap {\raise.5ex\hbox{$ < $} 11$ for superconductivity to occur. In phenomenologically acceptable models~\cite{dor} $e^2/J \sim 0.1$, which implies an upper bound on $\eta \sim 0.4$. However, the reader should bear in mind that the above-described limiting values are rather indicative at present, given that a complete quantitative understanding of the underlying dynamics of high-temperature superconductivity from an effective gauge-theory point of view is still lacking. \setcounter{equation}{0} \section{Equation for the Vertex in the case of non-zero photon mass} In this section we study the case where a (small) covariant photon mass $\delta $ is added to the photon propagator (\ref{photon}): \begin{equation} \Delta _{\mu\nu}(k) \sim \frac{\delta_{\mu\nu}}{k^2 + \delta ^2 } \label{photon2} \end{equation} For the purposes of our analysis in this work, such a mass term will simply be added by hand, without further discussions about its origin. However, we point out for completeness that a small photon mass may be the result of non-perturbative configurations (instantons) in compact $U(1)$ three-dimensional electrodynamics~\cite{polyakov}. Moreover, a term acting like a photon mass term arises in real-time finite-temperature considerations~\cite{aitchison}; in the latter case, of course, one loses manifest Lorentz covariance. It is straightforward to see that, in the presence of a photon mass, small compared to the scale $\alpha=e^2/8$, the resulting integral equation for the amputated vertex reads: \begin{equation} {\hat G}(p) = \frac{1}{24\pi^2}\frac{e^2}{p} \int \frac{dk}{k}{\hat G}^3(k) {\rm ln}\left|\frac{(k + p)^2 +\delta ^2}{(k-p)^2 + \delta^2}\right| \label{vertphot} \end{equation} We are interested in the limit $p << \delta $, which would allow us to study the effects of a photon mass on the infrared regime of the theory. Expanding the logarithm in (\ref{vertphot}) appropriately, we obtain: \begin{equation} {\hat G}(x) =\frac{4}{3\pi^2}\frac{1}{{\hat \delta} ^2} \int _0^x dy {\hat G}^3(y) + \frac{4}{3\pi^2}\int _x^\infty \frac{dy}{y^2 + {\hat \delta}^2} {\hat G}^3(y) \label{eqf} \end{equation} where ${\hat \delta} \equiv \delta/\alpha$, $x=p/\alpha$. Differentiating once with respect to $x$ one obtains: \begin{equation} \frac{d}{dx}{\hat G}(x) =\frac{4}{3\pi^2{\hat \delta ^2}}{\hat G}^3 (x) \frac{x^2}{x^2 + {\hat \delta}^2} \label{deriv} \end{equation} which can be integrated to yield \begin{equation} {\hat G}(x) =\frac{1}{\sqrt{c + \frac{8}{3\pi^2{\hat \delta}^2} \left({\hat \delta }{\rm arctg}(\frac{x}{{\hat \delta}}) - x\right)}} \label{rcdelta} \end{equation} where $c$ is an integration constant to be fixed by the boundary condition imposed by the integral equation (see below). The running coupling $G(x)$ tends to ${\hat G}(0) = \frac{1}{\sqrt{c}}$ as $x \rightarrow 0$. There is a non-trivial fixed point at $x=0$, given that the renormalization-group $\beta$-function (\ref{beta}) vanishes: \begin{equation} x\frac{d}{dx}{\hat G}(x) \rightarrow \frac{4}{3\pi^2}{\hat G}^3 (x) \frac{x^3}{x^2 + {\hat \delta}^2} =0, \qquad {\rm as}~x \rightarrow 0 \label{fp2} \end{equation} Notice that the function ${\hat G}(x)$ increases monotonically with increasing $x$, for small $x$. We next note that by choosing $c$ appropriately one can satisfy the condition (\ref{limit}) ${\hat G}(0) >> \sqrt{3/2}$, required for self-consistency of the approximation of ignoring the effects of the inhomogeneous term $Z(p)$ in the infrared. In such a case one expects that at larger $x$ the increase of the coupling stops at a certain (small) $x=x_0$, and then the coupling starts decreasing to reach asymptotically the perturbative result ${\hat G} \rightarrow 1$ (see fig. \ref{fig5}). The existence of a local maximum is characteristic of the effects of the photon mass on the infrared behaviour, to be contrasted with the monotonically decreasing situation in the case of a non-zero fermion mass (see fig. \ref{fig4}). We expect that in the general case, when both photon and fermion masses are present, the running coupling will exhibit a local maximum at low $x$ when the photon mass is larger than the fermion mass, whilst this behaviour will be replaced by a monotonic decrease, in the case when the fermion mass is considerably larger than the photon one. This situation should be compared with the corresponding one in four-dimensional QCD in the presence of non-vanishing gluon and quark masses~\cite{pc}. \begin{figure}[htb] \epsfxsize=3in \bigskip \centerline{\epsffile{bump.eps}} \caption{\it\baselineskip=12pt Running coupling (versus momentum) in the case of a small photon mass. Note the bump at low momenta, which is absent in the case of non-zero fermionic masses. The dashed part of the curve is conjectural at present, and can only be derived numerically. Continuous curves are the result of analytic studies.} \bigskip \label{fig5}\end{figure} Before closing we discuss briefly a physical situation which could be qualitatively similar to the case of a covariant infrared cut-off in the form of a non-zero photon mass, discussed in this section. This situation has been argued in \cite{am} to simulate finite-temperature effects, given that in such a case the photon propagator acquires a longitudinal plasmon mass term: \begin{equation} \Delta_{\mu\nu}(p_0=0, P \rightarrow 0, T) = \frac{\delta _{\mu 0}\delta _{0\nu}}{P^2 + M_{00}^2 } + {\rm longitudinal~parts} \label{photfinitet} \end{equation} where, for simplicity, we restricted ourselves to the instantaneous approximation~\cite{dor}, $p_0=0$, and the plasmon mass term \begin{equation} M_{00}(T) =\sqrt{\frac{2\alpha {\rm ln}2 T}{\pi}} \label{moo} \end{equation} We observe from the form (\ref{photfinitet}) that the presence of the mass term $M_{00}$ behaves for low $T$ somewhat analogously (but not identical to) a small photon cut-off mass $\delta$. Prompted by this observation we would like now to make some speculations regarding the low temperature effects of the $\delta $ cut-off term in (\ref{rcdelta}). We hasten to emphasize that a discussion could at best only qualitatively correct, given that a quantitative understanding would require a finite-temperature extension of the above analysis. Due to the loss of Lorentz covariance in the respective propagators the analysis is far more complicated than the zero-temperature one. However, when the temperature $T$ is sufficiently low, its effects on the pole structure of the photon propagator may be simulated by a covariant photon-mass cut-off, at least qualitatively~\cite{am,aitchison}. For this reason, we are motivated to examine the effects of ${\hat \delta } <<1$ on the low-temperature scaling of the resistivity~\cite{csm} $\rho$; the latter is defined as the inverse of the conductivity $\sigma _f$, which expresses the response of the system to a change in an externally-applied electromagnetic potential $A_\mu$. Ohm's law, in a gauge with $A_0=0$, then, gives~\cite{ioffe,am,csm}: \begin{equation} \sigma _f \equiv \rho ^{-1} = \frac{1}{\left(P^2 + p_0^2\right) \left(1 + \Pi (P,p_0,T) \right)}\left|_{P=0}\right. \label{condt} \end{equation} where $\Pi (P,p_0,T)$ denotes the dimensionless photon polarization used in this work. Following the analysis of \cite{aitchison} we may assume for our discussion below that~\footnote{The analysis of \cite{aitchison} is based on a real-time finite temperature approach. In such an approach the resistivity is defined in the infrared limit $p_0 \rightarrow 0, P=0$ in the presence of a plasmon mass term $M_{00}$ (\ref{moo}) in the photon propagator; as noted in \cite{aitchison}, the way in which the two mass scales ($p_0,P$) approach $0$ suffers from ambiguities, related to physical Landau-damping processes; it is not our purpose here to resolve such issues.} : \begin{equation} p_0^2/\alpha \rightarrow M_{00}^2/\alpha \sim {\hat \delta} ^2 \sim \frac{2{\rm ln}2}{\pi} T \label{deltaT} \end{equation} From the discussion following (\ref{runcoupl}) one immediately sees that the resistivity, defined for $P=0,x^2={\hat \delta}^2$, would be given in terms of the running coupling ${\hat G}(p)$ by: \begin{equation} \rho = {\hat \delta} ^2/{\hat G}^2(0,T) \label{rho} \end{equation} Using (\ref{rcdelta}) in the limit $x^2={\hat \delta} ^2$, one has: \begin{equation} \rho \sim \frac{2c T}{\pi}{\rm ln}2 + \frac{8}{3\pi^2}\left(\frac{\pi}{2} - 1\right) \sqrt{\frac{2T}{\pi}{\rm ln}2} \label{tbeh} \end{equation} The function $\rho$ is plotted in figure (\ref{fig6}), versus a single-power scaling behaviour of the form: $\rho \sim T^{1-\chi}$, with $\chi$ small; the latter is characteristic of certain theoretical condensed-matter models~\cite{byczuk}, and has been used in order to bound possible deviations from the experimentally-observed linear-$T$ behaviour of the resistivity of high-temperature superconductors at optimal doping~\cite{csm}. We observe that, upon appropriate choice of the integration constant $c$, the temperature dependence of (\ref{tbeh}) for a low-temperature regime, accessible to experiment, is hardly distinguishable from the single power scaling behaviour. \begin{figure}[htb] \epsfxsize=3in \bigskip \centerline{\epsffile{finiteT.eps}} \caption{\it\baselineskip=12pt Plot of the resistivity (\ref{tbeh}) versus temperature $T$ (in units of $\alpha$), for a low-temperature region (continuous curve), and comparison with the experimentally observed single-scaling (basically linear) behaviour $T^{1-\chi}, \chi=0.1$ (dashed curve). The agreement is very good for a regime of temperatures accessible to experiment.} \bigskip \label{fig6}\end{figure} \section{Conclusions and Outlook} In this work we have presented a study of chiral-symmetry breaking in $QED_3$, based on a system of coupled non-linear SD equations. The novel ingredient is the introduction of a semi-amputated vertex, whose dynamics is governed by a suitably truncated SD equation of cubic order. This allows for a self-consistent determination of the infrared value of the effective charge, in the presence of infrared cutoffs provided by either the fermion mass or a covariant photon mass. The theory is characterized by a (gauge-invariant) non-trivial infrared fixed-point, suggestive of non-Fermi liquid behaviour~\cite{shankar,am}. The non-linear vertex equation furnishes highly non-trivial constraints among the infrared value of the effective coupling and the mass of the fermions. When these constraints are combined with the fermion mass-gap equation, they select specific regions of the coupling space for which dynamical mass generation occurs. It will be interesting to study how these constraints are affected if one goes beyond the one-loop dressed approximation considered here. From the physical point of view, we remark that dynamical fermion mass generation in $QED_3$ is associated with superconductivity~\cite{dor,fm}. In the relevant statistical models, whose low-energy effective theories are of $QED_3$ type, the couplings depend on the doping concentration. Therefore the restrictions in the parameter space we have found above might impose restrictions in the allowed models. An interesting feature of our approach is the fact that the infrared structure of the photon propagator, derived as a consistent solution of a SD equation, implied formally a confining effective potential. The fact that this behaviour occurs in the chiral-symmetry broken phase of the theory, appears to be in agreement with generic expectations that confinement is a sufficient condition for chiral symmetry breaking. An obvious next step in this programme is the attempt to solve the coupled system of SD equations numerically, without the approximation of ignoring the effects of the inhomogeneous term $Z(p)$ in the infrared. We remind the reader that in the present work we have restricted the allowed values of the coupling such that this approximation was self-consistent. This, however, does not imply that solutions with $Z(p)$ not meeting these requirements are impossible; to establish or exclude their existence one should solve the whole system of equations, which at present can only be done numerically. Such a numerical treatment will have the additional advantage of not resorting to the approximations made to the kernel of the integral equations in order to convert them into differential ones. Furthermore, an analysis at finite temperatures will reveal whether our conjecture in section 6, on the scaling behaviour of the resistivity, is correct. As we have discussed there, this issue acquires great importance in view of very accurate experiments in high-temperature superconductors indicating a basically linear scaling with temperatures of the resistivity in the normal phase (no fermion mass gap) of the system. Finally, an extension of these ideas to the non-Abelian case appears as a challenge for the future, especially in view of recent claims~\cite{fm,wen} that non-Abelian gauge symmetries might describe the dynamics of spin-charge separation in realistic antiferromagnetic condensed-matter systems~\cite{anderson}. \section*{Acknowledgements} The authors would like to thank I.J.R. Aitchison for useful discussions. The work of N.E.M. is partially supported by a United Kingdom P.P.A.R.C. Advanced Research Fellowship, and that of JP is funded by a Marie Curie Fellowship (TMR-ERBFMBICT 972024).
\part{Figure Caption} Fig.1 : Typical evolution of the intensity profiles $\mid q_1\mid ^2$ and $% \mid q_2\mid ^2$ of the two-soliton solution (4) (a) showing the suppression in switching between the two modes of s1 soliton and (b) showing the suppression in switching of s1 soliton and enhancement in s2 soliton (while undergoing a large phase shift) for the parameter values given in the text. \end{document}
\section{#1}\setcounter{equation}{0}} \renewcommand{\theequation}{\thesection.\arabic{equation}} \begin{document} {}~ \hfill\vbox{\hbox{hep-th/9904207}\hbox{MRI-PHY/P990411} }\break \vskip 3.5cm \centerline{\large \bf Non-BPS States and Branes in String Theory} \medskip \vspace*{6.0ex} \centerline{\large \rm Ashoke Sen \footnote{E-mail: <EMAIL>, <EMAIL>}} \vspace*{1.5ex} \centerline{\large \it Mehta Research Institute of Mathematics} \centerline{\large \it and Mathematical Physics} \centerline{\large \it Chhatnag Road, Jhoosi, Allahabad 211019, INDIA} \vspace*{4.5ex} \centerline {\bf Abstract} We review the recent developments in our understanding of non-BPS states and branes in string theory. The topics include 1) construction of unstable non-BPS D-branes in type IIA and type IIB string theories, 2) construction of stable non-BPS D-branes on various orbifolds and orientifolds of type II string theories, 3) description of BPS and non-BPS D-branes as tachyonic soliton solutions on brane-antibrane pair of higher dimension, and 4) study of the spectrum of non-BPS states and branes on a system of coincident D-brane $-$ orientifold plane system. Some other related results are also discussed briefly. \vfill \eject \tableofcontents \baselineskip=18pt \sectiono{Introduction} \label{ss1} In this article I shall review the recent progress in our understanding of stable non-BPS branes and states in string theory. These lectures will be based mainly on refs.\cite{9803194,9805019,9805170,9806155,9808141,9809111,9810188,9812031,9812135,9901014}. We shall work in the convention $\hbar=1$, $c=1$, and $\alpha'=1$ (string tension=$(2\pi)^{-1}$) unless mentioned otherwise. Let us begin with some motivation for studying non-BPS branes. There are several reasons: \begin{enumerate} \item Stable non-BPS states and branes are very much part of the spectrum of string theory, and our understanding of string theory remains incomplete without a knowledge of these states. \item Stable non-BPS states are the simplest objects whose masses are not protected by supersymmetry, and yet are calculable in different limits of the string coupling. Hence studying the spectrum of these states in these different limits might provide new insight into what happens at finite string coupling. \item A system of coincident non-BPS D-branes typically has, as its world-volume theory, a non-supersymmetric gauge theory. Thus they may be useful in getting results about non-supersymmetric field theories from string theory, in the same way that a configuration of supersymmetric branes can be used to study non-perturbative aspects of supersymmetric gauge theories. \item Non-BPS branes may be relevant for constructing string compactification with broken supersymmetry. \end{enumerate} The plan of this article is as follows. In section \ref{ss2} we shall discuss the construction of unstable non-BPS D-branes in type IIA and type IIB string theories. Whereas type IIA (IIB) string theory admits stable BPS branes of even (odd) dimensions, we shall see that they also admit unstable non-BPS branes of odd (even) dimensions. In section \ref{ss3} we shall show how on certain orientifolds / orbifolds of type II string theories these non-BPS branes may give rise to stable non-BPS states and branes. The main point here will be to note that under this orbifolding / orientifolding operation the tachyonic mode responsible for the instability of the non-BPS brane gets projected out. The resulting brane is free from tachyonic mode and hence is stable. In section \ref{ss4} we shall discuss the interpretation of the non-BPS branes in type IIA and IIB string theories as tachyonic kink solution on a BPS D-brane - anti-D-brane pair of one higher dimension in the same theory. We shall also show how the BPS D-brane (anti-D-brane) can be regarded as a tachyonic kink (anti-kink) solution on a non-BPS D-brane of one higher dimension. This gives a set of descent relations between BPS and non-BPS D-branes of type II string theories, and form the basis of identifying the D-brane charge with elements of K-theory\cite{9810188,9812135,9812226,9901042,9902102,9902116,9902160,9904153}. Since the actual proof of these relations is technically somewhat complicated, we postpone the details to the appendix. In section \ref{ss5} we discuss the spectrum of stable non-BPS states and branes on a coincident D-brane and orientifold plane system. The masses (tensions) of these states (branes) can be calculated in the strong coupling limit using various duality symmetries of string theory. Although for each system we use a completely different method for finding the spectrum, the final spectrum of non-BPS states and branes on a D-$p$-brane $-$ orientifold $p$-plane system exhibits an unusual regularity as a function of $p$. Whether this signifies any deep aspect of string theory remains to be seen. Since the only similarity between these systems is in their weak coupling perturbation expansion, we suspect that the strong coupling result may be governed by large order behaviour of this perturbation expansion. Finally in section \ref{ss6} we discuss some related developments in this subject. This includes a discussion of some other non-BPS branes in type I string theory, the relationship between K-theory and D-brane charges, and the application of boundary state formalism in the study of non-BPS D-branes. We end with a discussion of some open questions. \sectiono{Unstable Non-BPS D-branes in Type II String Theories} \label{ss2} \subsection{BPS D-branes in type II string theories} \begin{figure}[|ht] \begin{center} \leavevmode \epsfbox{fig1.eps} \end{center} \caption{Open strings ending on a BPS D-brane.} \label{f1} \end{figure} Let us begin by reviewing what we know about BPS D-branes in type IIA/IIB string theories\cite{DBRANE}. The defining property of the D-brane is that fundamental strings can end on a D-brane as shown in Fig.\ref{f1}, although type II string theories in the bulk only contains closed string states without any end. The open strings with ends on the D-brane can be interpreted as the dynamical modes of the D-brane. In order to compute the spectrum of these open string states with ends on the D-brane, we impose Dirichlet boundary condition on the open string coordinates along directions transverse to the D-brane, and Neumann boundary condition along directions parallel to the brane world-volume (including time). A D-brane with $p$ tangential spatial directions is called a D-$p$-brane. Let us now list some of the properties of D-branes in type II string theories which will be useful to us later. \begin{itemize} \item Type IIA (IIB) string theory admits BPS D-$2p$-brane (D-$(2p+1)$-brane) which are invariant under half of the space-time supersymmetry transformations of the theory. \item A D-$p$-brane is charged under a $(p+1)$-form gauge field arising in the Ramond-Ramond (RR) sector of the theory. \item These BPS D-branes are oriented. D-branes of opposite orientation carry opposite RR charge and will be called anti-D-branes ($\bar {\rm D}$-branes). \end{itemize} \begin{figure}[|ht] \begin{center} \leavevmode \epsfbox{fig2.eps} \end{center} \caption{Open strings living on a coincident D-brane anti-D-brane pair. Although for clarity we have displayed the brane and the anti-brane to be spatially separated, we shall analyse the case where they coincide.} \label{f2} \end{figure} Next we shall review properties of coincident D-brane $-$ ${\bar {\rm D}}$-brane pair shown in Fig.\ref{f2}. They are as follows: \begin{itemize} \item Spectrum of open strings living on the world-volume contains four different sectors. These four sectors can be labelled by $2\times 2$ Chan Paton (CP) factors: \begin{eqnarray}\displaystyle \label{es2.1} && (a): \quad \pmatrix{0 & 0 \cr 0 & 1}, \qquad (b):\quad \pmatrix{ 1 & 0\cr 0 & 0} \nonumber \\ && (c):\quad \pmatrix{0 & 0 \cr 1 & 0}, \qquad (d):\quad \pmatrix{ 0 & 1\cr 0 & 0} \, . \end{eqnarray} \item GSO projection: Physical states in sectors (a) and (b) should have $(-1)^F=1$ whereas those in sectors (c) and (d) should have $(-1)^F=-1$. Here $F$ denotes the {\it world-sheet fermion number} carried by the state. We use the convention that the $(-1)^F$ eigenvalue of the Neveu-Schwarz (NS) sector ground state is $-1$. The GSO projection rule follows from the observation that the closed string exchange interaction between a D-brane and a ${\bar {\rm D}}$-brane and that between a pair of D-branes have the same sign for NSNS sector closed string exchange and opposite sign for RR sector closed string exchange. In the open string channel this corresponds to replacing the GSO projection operator ${1+(-1)^F\over 2}$ for DD strings by ${1-(-1)^F\over 2}$ for D${\bar {\rm D}}$ strings. \item Since the NS sector ground state has $(-1)^F=-1$, it survives the GSO projection in sectors (c) and (d) and gives tachyonic excitations with\cite{9403040,9511194,9604091,9604156,9612215} \begin{equation} \label{es2.2} m^2 = -(1/2)\, . \end{equation} Since the tachyon comes from two different sectors it is a complex scalar field. \item Although individually the D-brane as well as the ${\bar {\rm D}}$-brane is invariant under half of the space-time supersymmetry transformations, the combined system breaks all supersymmetries. \end{itemize} We shall now study the action of $(-1)^{F_L}$ on the coincident D-brane $-$ ${\bar {\rm D}}$-brane system, where $F_L$ denotes the contribution to the {\it space-time fermion number} from the left-moving sector of the string world-sheet. $(-1)^{F_L}$ is known to be an exact symmetry of type IIA and type IIB string theories. Acting on the closed string Hilbert space, it changes the sign of all the states on the left-moving Ramond sector, but does not change anything else. Thus it has trivial action on the world-sheet fields. {}From this definition it follows that the space-time fields originating in the Ramond-Ramond (RR) sector of the world-sheet change sign under $(-1)^{F_L}$. Since D-branes are charged under RR field, it follows that $(-1)^{F_L}$ must take a D-brane to a ${\bar {\rm D}}$-brane. Thus a single D-brane or a single ${\bar {\rm D}}$-brane is not invariant under $(-1)^{F_L}$, but a coincident D-brane $-$ ${\bar {\rm D}}$-brane system is invariant under $(-1)^{F_L}$. Hence it makes sense to study the action of $(-1)^{F_L}$ on the open strings living on this system, which is what we shall do now. We begin with the observation that since $(-1)^{F_L}$ has no action on the world-sheet fields, we only need to study its action on the CP factors.\footnote{We shall focus our attention on the NS-sector states, but a similar analysis can be done separately for the R-sector states.} Since $(-1)^{F_L}$ exchanges D-brane with ${\bar {\rm D}}$-brane, it acts on the CP matrix $\Lambda$ as \begin{equation} \label{es2.3} \Lambda \to \sigma_1 \Lambda (\sigma_1)^{-1}\, , \end{equation} where \begin{equation} \label{es2.4} \sigma_1 = \pmatrix{0 & 1\cr 1 & 0}\, . \end{equation} This shows that states with CP factors $I$ and $\sigma_1$ are even under $(-1)^{F_L}$, whereas those with CP factors $\sigma_3$ and $i\sigma_2$ are odd. (We could replace $\sigma_1$ by $\sigma_2$ in \refb{es2.3}, but this just amounts to a change in convention.) \subsection{Non-BPS D-branes in type II string theories} We are now ready to define a non-BPS D-$2p$-brane of type IIB string theory\cite{9812031}. This is done by following the steps listed below. \begin{itemize} \item We start with a D-$2p$ $-$ ${\bar {\rm D}}$-$2p$-brane pair in type IIA string theory and take the orbifold of this configuration by $(-1)^{F_L}$. \item In the bulk, modding out IIA by $(-1)^{F_L}$ gives IIB. \item Acting on the open strings living on the D-${\bar {\rm D}}$-brane world-volume, $(-1)^{F_L}$ projection keeps states with CP factors $I$ and $\sigma_1$ and throws out states with CP factors $\sigma_3$ and $i\sigma_2$. \end{itemize} This defines a {\it non-BPS D-$2p$-brane} of type IIB string theory. In order to see that it describes a single object rather than a pair of objects, we simply note that before the projection the degree of freedom of separating the two branes reside in the sector with CP factor $\sigma_3$. Since states in the CP sector $\sigma_3$ are projected out, we lose the degree of freedom of separating the brane antibrane pair away from each other. Similarly, starting from a D-$(2p+1)$-brane ${\bar {\rm D}}$-$(2p+1)$-brane pair of IIB, and modding it out by $(-1)^{F_L}$, we can define a non-BPS $(2p+1)$-brane of IIA. Thus type IIB string theory contains BPS D-branes of odd dimension and non-BPS D-branes of even dimension, whereas type IIA string theory contains BPS D-branes of even dimension and non-BPS D-branes of odd dimension. Let us now list some of the properties of the non-BPS D-$2p$-brane of type IIB string theory. (Similar results also hold for the non-BPS D-$(2p+1)$-brane of type IIA string theory.) These properties follow from their definition, and properties of coincident brane-antibrane pair reviewed earlier. \begin{itemize} \item Excitations on its world-volume are open strings with Dirichlet boundary condition on the $(9-2p)$ transverse directions, and Neumann boundary condition on $2p+1$ tangential directions (including time). \item These open strings carry Chan Paton factors $I$ or $\sigma_1$. \item Physical states with CP factor $I$ has $(-1)^F=1$ and physical states with CP factor $\sigma_1$ has $(-1)^F=-1$. (Note again that $F$ denotes {\it world-sheet fermion number}.) \item The NS sector ground state carrying CP factor $\sigma_1$ has $(-1)^F=-1$ and hence is physical. Thus there is a tachyonic mode with \begin{equation} \label{es2.5} m^2 = -{1\over 2} \, . \end{equation} \item Since tachyon comes from only one sector, it is a real scalar field. \item The tension of the non-BPS D-$2p$-brane of type IIB string theory is given by: \begin{equation} \label{es2.6} (2\pi)^{-2p} (\sqrt 2/g)\, , \end{equation} where $g$ denotes the coupling constant of the string theory. This property can be derived by taking into account the effect of modding out by $(-1)^{F_L}$, and the fact that the original brane-antibrane system before $(-1)^{F_L}$ modding had a tension equal to \begin{equation} \label{es2.7} (2\pi)^{-2p} (2/g)\, . \end{equation} Similarly, the tension of a non-BPS D-$(2p+1)$-brane of type IIA string theory is given by \begin{equation} \label{eiia} (2\pi)^{-(2p+1)}(\sqrt 2/g)\, . \end{equation} \end{itemize} One can also derive the spectrum of open strings with one end on the non-BPS brane and other end on a BPS brane, but we shall not discuss it here. \subsection{BPS D-branes from non-BPS D-branes} Let us now consider the effect of modding out a non-BPS D-$2p$-brane of IIB by $(-1)^{F_L}$, where $(-1)^{F_L}$ now denotes the corresponding symmetry of the type IIB string theory\cite{9812031}. In the bulk, modding out type IIB string theory by $(-1)^{F_L}$ gives us back a type IIA string theory. The question we shall be interested in is: what happens to the D-$2p$-brane after this modding? This question makes sense as the non-BPS D-brane does not carry any RR charge and hence is invariant under the action of $(-1)^{F_L}$. In order to answer this question we need to study the action of $(-1)^{F_L}$ on the open string states living on the D-$2p$-brane. As before $(-1)^{F_L}$ does not act on the world-sheet fields, but acts only on the CP factors. Thus we need to find the action of $(-1)^{F_L}$ on CP factors. This is done with the help of the following observations:\footnote{For definiteness we shall focus our attention on the NS sector states, but a similar analysis can also be carried out for R sector states.} \begin{figure}[|ht] \begin{center} \leavevmode \epsfbox{fig3.eps} \end{center} \caption{The disk amplitude for two point function of the graviton and translation mode of the D-brane.} \label{f3} \end{figure} \begin{itemize} \item There is a non-zero two point function of the graviton $g_{m\mu}$ from the closed string sector and the translation modes $X^m$ of the D-$2p$-brane originating in the CP sector $I$ of the form: \begin{equation} \label{egfmet} \eta^{\mu\nu} g_{m\mu} \partial_\nu X^m \, , \end{equation} where $m$ denotes a direction transverse to the brane and $\mu,\nu$ denote directions tangential to the brane. This coupling follows from expanding the Dirac-Born-Infeld action on the brane world-volume around the configuration of a flat brane in a flat space-time background. This can also be seen by computing a disk amplitude with a graviton vertex operator inserted at the center of the disk, and the tachyon vertex operator inserted at the boundary of the disk, as shown in Fig.\ref{f3}.\footnote{This does not mean that that a physical on-shell scalar particle on the brane has a finite transition probability into a graviton state in the bulk. This is disallowed due to various kinematic reasons. However, the existence of the coupling \refb{egfmet} can still be deduced by evaluation the disk amplitude in a region of unphysical (complex) external momenta; as is done {\it e.g.} in deducing the Yang-Mill's three gauge boson vertex from three string amplitude\cite{SCHPHYSREP}.} Since the graviton is even under $(-1)^{F_L}$, this shows that states with CP factor $I$ must also be even under $(-1)^{F_L}$. \begin{figure}[|ht] \begin{center} \leavevmode \epsfbox{fig3a.eps} \end{center} \caption{The disk amplitude for the two point function of the tachyon and the RR sector $2p$-form gauge field. The dotted line denotes the $(-1)^{F_L}$ cut extending from the RR vertex operator to the disk boundary.} \label{f3a} \end{figure} \item There is a non-zero two point function of the RR-sector $2p$-form gauge field $A^{(2p)}$ from the closed string sector and the tachyonic mode $T$ of the D-$2p$-brane originating in the CP sector $\sigma_1$ of the form: \begin{equation} \label{etachrr} \int A^{(2p)}\wedge dT\, . \end{equation} This can be seen by computing the disk amplitude with a RR-sector gauge field vertex operator inserted at the center of the disk, and the tachyon vertex operator inserted at the boundary, as shown in Fig.\ref{f3a}.\footnote{Again, as before, the actual transition between a massless RR sector state and the tachyon is absent due to kinematic reasons.} The fact that this amplitude is non-zero may seem surprising, as the tachyon vertex operator carries a CP factor $\sigma_1$, and there seems to be no other CP factor inserted at the boundary of the disk. However, since the RR sector states in type IIB string theory appear in the twisted sector when we regard type IIB string theory as type IIA string theory modded out by $(-1)^{F_L}$, there is a cut extending from the RR sector vertex operator at the center all the way to the boundary of the disk. At the point where the cut hits the boundary we need to insert an extra factor of $\sigma_1$, since $(-1)^{F_L}$ action on the CP factors correspond to conjugation by $\sigma_1$. This gives a total of two factors of $\sigma_1$ on the disk boundary and makes the amplitude non-vanishing. {}From this it follows that since RR-sector fields are odd under $(-1)^{F_L}$, states with CP factor $\sigma_1$ are also odd under $(-1)^{F_L}$. \end{itemize} The net result of this analysis is that states with CP factor $I$ are $(-1)^{F_L}$ even and states with CP factor $\sigma_1$ are $(-1)^{F_L}$ odd. Thus under modding out by $(-1)^{F_L}$, only states with CP factor $I$ survive the projection. As we have already seen before, GSO projection requires these states to be even under $(-1)^F$. Thus the spectrum is identical to that of open strings living on a {\it BPS D-$2p$-brane of IIA}, and we conclude that the non-BPS D-$2p$-brane of type IIB string theory, modded out by $(-1)^{F_L}$, gives a BPS D-$2p$-brane of type IIA string theory.\footnote{Note that at this stage we cannot determine whether the resulting brane is a D-brane or a ${\bar {\rm D}}$-brane, as both carry the same spectrum of open string. This is a reflection of the fact that the orbifolding procedure has a two-fold ambiguity, so that we could end up either with a D-brane or a ${\bar {\rm D}}$-brane by following these steps.} \begin{figure}[|ht] \begin{center} \leavevmode \epsfbox{fig4.eps} \end{center} \caption{Relationship between BPS and non-BPS D-branes in type II string theories. The horizontal arrow represents the effect of modding out the theory by $(-1)^{F_L}$.} \label{f4} \end{figure} The results of this section have been summarized in Fig.\ref{f4}. There is also a similar relation with IIA $\leftrightarrow$ IIB and $(2p)\rightarrow(2p+1)$. \sectiono{Stable Non-BPS D-branes on Type II Orbifolds and Orientifolds} \label{ss3} Although we have constructed non-BPS D-branes in type IIA/IIB string theory in the last section, they are all unstable due to the presence of the tachyonic mode. As we shall discuss in section \ref{ss4}, if the tachyon condenses to its minimum, then the configuration is indistinguishible from the vacuum\cite{9805170}. Thus it is natural to ask: what is the use of such a D-brane? In this section we shall show that although they are unstable in type IIA/IIB string theory, we may get {\it stable} non-BPS D-branes in certain orbifolds/orientifolds of IIA/IIB {\it if the tachyonic mode is projected out under this operation.} We shall illustrate this through two examples. \subsection{Type I D-particle} \label{stypei} Let us consider the following construction: \begin{itemize} \item Start with the non-BPS D0-brane (D-particle) of type IIB as defined in the last section. \item Mod out the configuration by the world-sheet parity transformation $\Omega$. \end{itemize} The result can be described as a non-BPS D-particle of type I string theory, since in the bulk type IIB string theory modded out by $\Omega$ gives a type I string theory. The crucial question is: is this D-particle stable? Or equivalently we may ask: is the tachyonic mode on the type IIB D-particle odd under $\Omega$? The answer to this question follows from eq.\refb{etachrr} for $p=0$, {\it i.e.} that the two point function of the tachyonic mode on the D-particle world-volume and the RR sector scalar field $\phi$ of type IIB string theory is non-vanishing. Since the field $\phi$ is known to be odd under $\Omega$, we conclude that the tachyonic mode of the D-particle is also odd under $\Omega$. Thus it is projected out in type I string theory. In other words, the type I D-particle is stable\cite{9809111}! The spectrum of open strings on type I D-particle also includes open strings with one end on the D-particle and the other end on any one of the 32 nine branes which are present in type I string theory. The Ramond sector states from this sector can be shown to give rise to 32 massless fermionic zero modes living on the D0-brane. Quantization of these zero modes gives rise to a ground state which transforms in the spinor representation of the type I gauge group SO(32). This also gives an additional explanation of the stability of the D-particle. Since all perturbative states of type I string theory are in the scalar conjugacy class of SO(32), and since a spinor state cannot decay into states in the scalar conjugacy class, the D-particle is prevented from decaying into perturbative string states due to charge conservation. If we consider two or more coincident D-particles in type I string theory, then there are also possible tachyonic modes coming from open strings with two ends on two different D-particles. It turns out that the $\Omega$ projection does not remove all the tachyonic modes from these sectors, and two or more coincident D-particles describe an unstable system. This is consistent with the observation that two particles in the spinor representation of SO(32) can combine and annihilate into perturbative string states, as there is no conservation law preventing this process. The existence of the type I D-particle is also relevant for testing the conjectured duality between type I and heterotic string theory\cite{9503124,9506160,9506194,9510169}. SO(32) heterotic string theory contains states in the perturbative spectrum which transform in the spinor representation of SO(32). These states are massive, and non-BPS. But the lightest state belonging to the spinor representation of SO(32) is stable at all values of the coupling, as they cannot decay into anything else. Thus these states must also exist in the strong coupling limit of the SO(32) heterotic string theory, which is nothing but the weakly coupled type I string theory. The type I D-particles provide explicit realization of these states. It is instructive to compare the mass formulae for these SO(32) spinor states at the two extreme ranges of the coupling constant. We shall use the variables of the heterotic string theory to express this mass formula at the two ends. For small heterotic coupling $g_H$, the perturbative mass formula in the heterotic string theory holds: \begin{equation} \label{es3.1} \sqrt{T_H}\, (a_0+a_1\, g_H^2+a_2\, g_H^4+\ldots)\, , \end{equation} where $T_H$ is the heterotic string tension, and $a_i$ are numerical coefficients. $a_0$ is computed at tree level of heterotic string theory, whereas $a_m$ is computed at $m$-loop order. For large heterotic coupling we can use the description of this state as type I D-particle to compute its mass. As we saw earlier, this has mass of order $\sqrt{T_I}/g_I$, where $T_I$ and $g_I$ are the string tension and coupling constant respectively of the type I string theory. Using standard relationship between the heterotic and type I variables\cite{9503124} \begin{equation} \label{es3.2} T_I = T_H g_H^{-1}, \qquad g_I=g_H^{-1}\, , \end{equation} we see that for large $g_H$ the mass of this state is proportional to: \begin{equation} \label{es3.3} \sqrt{T_H} (g_H)^{1/2}\, . \end{equation} It will be interesting to see if the perturbation expansion \refb{es3.1} contains any information about the large $g_H$ behaviour given in \refb{es3.3}. \subsection{D-branes wrapped on non-supersymmetric cycles of K3 orbifold} In this section we shall discuss another example where the tachyonic mode of a non-BPS D-brane is projected out under an orbifolding operation\cite{9812031}. We proceed as follows: \begin{itemize} \item Start with a non-BPS D-string of type IIA string theory wrapped on a circle along $x^9$ of radius $R_9$ and placed at $x^i=0$ for $1\le i\le 8$. \item Compactify three other directions $x^6,x^7,x^8$. \item Mod out the theory by a $Z_2$ transformation ${\cal I}_4$ which changes the sign of $x^6,\ldots x^9$: \begin{equation} \label{es3.4} {\cal I}_4: \quad (x^6,x^7,x^8,x^9)\to (-x^6,-x^7,-x^8,-x^9)\, . \end{equation} \end{itemize} In the bulk this gives type IIA string theory on an orbifold K3. We shall now analyze the fate of the tachyon field on the D-string in this orbifold theory. Since the D-string lies along $x^9$, the tachyon field on its world-steet is a function of $x^9$ and time $t$. Again by considering a two point function between the tachyon and an RR sector gauge field, one can show that under the $Z_2$ transformation ${\cal I}_4$, \begin{equation} \label{es3.5} T(x^9,t)\to - T(-x^9,t)\, . \end{equation} If we expand $T(x^9,t)$ in its Fourier mode as: \begin{equation} \label{es3.6} T(x^9,t) = \sum_n T_n(t) e^{inx^9/R_9}\, , \end{equation} then under ${\cal I}_4$: \begin{equation} \label{es3.7} T_n(t) \to - T_{-n}(t)\, . \end{equation} Thus \begin{itemize} \item $T_0$ is projected out. \item For $n\ne 0$ the combination $T_n - T_{-n}$ survives the projection under ${\cal I}_4$. \end{itemize} Since the effective mass$^2$ of $T_n-T_{-n}$ is given by \begin{equation} \label{es3.8} m_n^2 = (n^2/R_9^2)-(1/2)\, , \end{equation} we see that there is no tachyon in the spectrum for \begin{equation} \label{es3.9} R_9\le \sqrt 2\, . \end{equation} There are also possible tachyonic modes from open string states stretched between the original D-string and its image under translation along $x^6$, $x^7$ or $x^8$. Demanding that there are no tachyonic modes from these sectors we also get\footnote{These relations can be found from eq.\refb{es3.9} by a T-duality transformation.} \begin{equation} \label{es3.10} R_8\ge{1\over \sqrt 2}, \quad R_7\ge {1\over \sqrt 2}, \quad R_6\ge {1\over \sqrt 2}\, . \end{equation} \begin{figure}[|ht] \begin{center} \leavevmode \epsfbox{fig8.eps} \end{center} \caption{The D0-${\bar {\rm D}}$0 brane configuration obtained by marginal deformation of the non-BPS D-string wrapped along $x^9$.} \label{f8} \end{figure} The net result of this analysis is that we have a stable non-BPS state in type IIA string theory on $T^4/{\cal I}_4$ in the range of parameters described in \refb{es3.9}, \refb{es3.10}. The next question would be: what is the interpretation of this state? There are many ways to answer the question; we shall explain it by studying the physics at the critical radius $R_9=\sqrt 2$. At this radius $T_{\pm 1}$ are massless modes. In fact one can show that the potential for $(T_1-T_{-1})$ vanishes identically.\footnote{This and various other issues related to this discussion will be discussed in some detail in section \ref{ss4}.} Thus $(T_1-T_{-1})$ denotes an exactly marginal deformation of the boundary conformal field theory (CFT) describing the D-brane. We can study this deformation using CFT techniques. We shall only quote the result here (see section \ref{ss4}, the appendix and ref.\cite{9812031} for some of the details). It turns out that this marginal deformation takes the non-BPS D-string of IIA to a D$0$-${\bar {\rm D}}$0-brane pair situated at the two fixed points $x^9=0$ and $x^9=\pi R_9$ respectively as shown in Fig.\ref{f8}. So far what we have described could have been done even before modding out the theory by ${\cal I}_4$. Let us now study the result of modding out this configuration by ${\cal I}_4$. It was shown in ref.\cite{9603167} that after modding out by ${\cal I}_4$ a $D0$-brane at $x^9=0$ can be interpreted as a D2-brane of type IIA string theory, wrapped on the supersymmetric 2-cycle\cite{9507158} associated with the fixed point of ${\cal I}_4$ at $x^9=0$.\footnote{Although the cycle has zero area, the wrapped D-brane has a finite mass due to the presence of the anti-symmetric tensor field flux through the two cycle\cite{9507012}.} A similar interpretation can be given for the ${\bar {\rm D}}$0-brane at $x^9=\pi R_9$. Thus in the orbifold theory the marginal perturbation by $(T_1-T_{-1})$ at $R_9={\sqrt 2}$ takes the original non-BPS state to a pair of D2-branes, wrapped on the 2-cycles associated with the fixed points at $x^9=0$ and $x^9=\pi R_9$ respectively. This suggests that the original configuration is a D-2-brane of IIA wrapped simultaneously on both these 2-cycles. This represents a D2-brane wrapped on a non-supersymmetric 2-cycle. Before the projection, the mass of the wrapped non-BPS D-string is given by $(\sqrt 2 R_9/g)$, whereas the sum of the masses of the D0-${\bar {\rm D}}$0 pair is given by $(2/g)$. Modding out by ${\cal I}_4$ reduces the mass of each state to half its original value. By comparing the masses of the various (wrapped) branes we arrive at the following picture: \begin{itemize} \item At the critical radius the D-2-brane wrapped on the non-supersymmetric cycle is degenerate with the pair of D-2-branes wrapped on the supersymmetric cycles. \item Below the critical radius the D-2-brane wrapped on the non-supersymmetric cycle is lighter than the pair of D-2-branes wrapped on the two supersymmetric cycles. Hence this wrapped brane is stable. \item Above the critical radius the D-2-brane wrapped on the non-supersymmetric cycle is heavier than the pair of D-2-branes wrapped on the two supersymmetric cycles. As a result this wrapped brane is unstable against decay into a pair of supersymmetric brane configurations. \end{itemize} This construction can be generalized to describe a $(2p+2)$-brane ($(2p+1)$-brane) of IIA (IIB) wrapped on a non-supersymmetric cycle of K3. Using this procedure one can also construct examples of D-branes wrapped on non-BPS 2- and 3-cycles of Calabi-Yau manifolds. These generalizations have been discussed in ref.\cite{9812031}. Before concluding this discussion we note that the world-volume theory of $N$ coincident branes of this type gives rise to a non-supersymmetric U(N) gauge theory. This might be useful in solving non-supersymmetric field theories via branes. \sectiono{D-branes as Tachyonic Kink Solutions} \label{ss4} \subsection{Non-BPS D-brane as tachyonic kink on the brane-antibrane pair} \begin{figure}[|ht] \begin{center} \leavevmode \epsfbox{fig5.eps} \end{center} \caption{The tachyon potential on D-brane $-$ ${\bar {\rm D}}$-brane pair.} \label{f5} \end{figure} In this section we shall give an alternative construction of the non-BPS D-branes discussed in section \ref{ss2}\cite{9805019,9808141}. Let us start with a coincident pair of D-$2p$ $-$ ${\bar {\rm D}}$-$2p$ branes ($p\ge 1$) of type IIA string theory. As discussed in section \ref{ss2}, there is a complex tachyon field $T$ living on the world-volume of this system. This reflects the fact that $T=0$ is the maximum of the tachyon potential $V(T)$ obtained after integrating out all other massive modes on the world-volume. There is a U(1)$\times$U(1) gauge field living on the world-volume of the brane-antibrane system, and the tachyon picks up a phase under each of these U(1) gauge transformations. As a result, $V(T)$ is a function only of $|T|$, and the minimum of the potential occurs at $T=T_0 e^{i\theta}$ for some fixed $T_0$ but arbitrary $\theta$, as shown in Fig.\ref{f5}. As we shall argue shortly, at the minimum, the sum of the tension of the D-brane ${\bar {\rm D}}$-brane pair and the (negative) potential energy of the tachyon is {\it exactly zero\cite{9805170} i.e.}, \begin{equation} \label{es4.1} 2 T_D + V(T_0) = 0\, , \end{equation} where $T_D$ is the D-brane tension. This shows that the tachyonic ground state $T=T_0$ is indistinguishible from the vacuum, since it carries neither any charge nor any energy density. \begin{figure}[|ht] \begin{center} \leavevmode \epsfbox{fig6.eps} \end{center} \caption{Tachyonic kink solution on the brane-antibrane pair.} \label{f6} \end{figure} But now, instead of considering tachyonic ground state, let us consider a tachyonic kink solution. For this, consider the minimum energy configuration with the following properties: \begin{itemize} \item $Im(T)=0$. \item $Re(T)$ independent of time and $(2p-1)$ of the $2p$ spatial coordinates. \item $Re(T)$ depends on the remaining spatial coordinate $x$ such that \begin{eqnarray}\displaystyle \label{es4.2} T(x) \to T_0 \quad &\hbox{as}& \quad x\to \infty, \nonumber \\ T(x)\to -T_0 \quad &\hbox{as}& \quad x\to -\infty\, . \end{eqnarray} \end{itemize} This has been shown in Fig.\ref{f6}. {}From this it is clear that as $x\to \pm\infty$ the solution goes to vacuum configuration. Thus the energy density is concentrated around a $(2p-1)$ dimensional subspace, and the solution describes a $(2p-1)$-dimensional brane. We now claim that {\it this $(2p-1)$-brane associated with the tachyonic kink solution on the brane-antibrane pair is identical to the non-BPS D-$(2p-1)$-brane of IIA.} Note that $V(T)$ cannot be explicitly calculated. Thus one might ask how one could show the equivalence between the non-BPS D-brane described in section \ref{ss2} and the tachyonic kink on the brane $-$ antibrane pair described here. This will be discussed in some detail in the appendix; but here we shall describe the outline of the proof. \begin{itemize} \item There is a marginal deformation involving bulk and boundary operators which interpolates between the $T=0$ configuration and the kink solution. \item One can study the fate of the CFT describing the brane-antibrane pair under this marginal deformation. \item The end result turns out to be a CFT which is identical to the CFT describing the non-BPS D-brane. \end{itemize} One can also give an intuitive understanding of why a tachyonic kink should behave like a D-brane. For this note that far away from the kink (large $|x|$) the configuration represents the vacuum, and hence strings cannot end there. On the other hand, on the subspace $x=0$, the tachyon field vanishes, and hence we expect the configuration to behave in a way that a D-brane $-$ ${\bar {\rm D}}$-brane pair would have behaved in the absence of tachyon vev, {\it i.e.} open strings should be able to end there. Thus the tachyonic kink should at least qualitatively behave as a D-brane located at $x=0$. Note that the manifold ${\cal M}$ describing the minimum of the tachyon potential is a circle $S^1$. In order to get a topologically stable kink solution, we need $\pi_0({\cal M}) \ne 0$. But $\pi_0(S^1)=0$ since $S^1$ is connected. Thus the kink is not topologically stable. Indeed it has tachyonic mode correponding to the freedom of changing $T$ at $x\to\infty$ to $T_0 e^{i\theta}$. As $\theta\to\pi$ we get back the vacuum configuration, since $T\to -T_0$ as $x\to\pm\infty$ in this case. This however is completely consistent with the identification of this kink solution with the non-BPS D-$(2p-1)$-brane of type IIA string theory, since, as we have seen earlier, the latter also has a tachyonic mode living on it. The tachyonic mode on the kink solution can be identified as the tachyonic mode on the non-BPS D-$(2p-1)$-brane of IIA discovered in section \ref{ss2}. Before we move on to the next subject, let us give an argument in favour of eq.\refb{es4.1}. For this, note that if \refb{es4.1} had not been true, then the tachyonic kink solution described here will not have a finite energy per unit $(2p-1)$-volume, since the energy density, integrated along the transverse direction (denoted by $x$ in eq.\refb{es4.2}) would give infinite answer. On the other hand from the analysis of section \ref{ss2} we certainly know that a non-BPS D-$(2p-1)$ brane of type IIA string theory has finite tension. Thus once we establish the equivalence of the tachyonic kink solution and the non-BPS D-brane (as will be discussed in the appendix), it automatically establishes eq.\refb{es4.1}. \subsection{The BPS D-brane as the tachyonic kink on the non-BPS D-brane} \begin{figure}[|ht] \begin{center} \leavevmode \epsfbox{fig5a.eps} \end{center} \caption{The tachyon potential on the non-BPS D-brane} \label{f5a} \end{figure} We can now continue one step further. Let us start with a non-BPS D-$(2p-1)$-brane of IIA. As was discussed in section \ref{ss2}, it has a real tachyon $\widetilde T$. By studying the disk amplitude it can be easily seen that there is a $Z_2$ symmetry on the world-volume of this non-BPS D-brane under which $\widetilde T$ (and all other modes originating in the CP sector $\sigma_1$) changes sign. Let $\pm \widetilde T_0$ be the minimum of the tachyon potential $\widetilde V(\widetilde T)$ obtained after integration out the other massive modes, as shown in Fig.\ref{f5a}. Again one can argue that: \begin{equation} \label{es42.1} \widetilde V(\widetilde T_0) + \widetilde T_D =0\, , \end{equation} where $\widetilde T_D$ is the tension of the non-BPS D-brane. We now consider a kink solution on this D-$(2p-1)$-brane world-volume such that: \begin{figure}[|ht] \begin{center} \leavevmode \epsfbox{fig6a.eps} \end{center} \caption{Tachyonic kink solution on a non-BPS D-brane.} \label{f6a} \end{figure} \begin{itemize} \item $\widetilde T$ is independent of time as well as $(2p-2)$ of the spatial coordinates. \item It depends on the remaining world-volume coordinate $y$ such that: \begin{eqnarray}\displaystyle \label{es42.3} \widetilde T(y)\to \widetilde T_0 \quad &\hbox{as}& \quad y\to \infty, \nonumber \\ \widetilde T(y)\to -\widetilde T_0 \quad &\hbox{as}& \quad y\to -\infty\, . \end{eqnarray} \end{itemize} This configuration has been shown in Fig.\ref{f6a}. By the same argument as in the previous subsection, this describes a $(2p-2)$ dimensional brane. We shall show in the appendix that this can be identified as the BPS D-$(2p-2)$ brane of type IIA string theory. The analysis is again based on finding a series of marginal deformations involving bulk and boundary operators which connect the $\widetilde T=0$ configuration of the non-BPS D-$(2p-1)$ brane to a solution representing a kink-antikink pair, and using conformal field theory techniques to show that this marginal deformation actually interpolates between the non-BPS D-$(2p-1)$-brane and a BPS D-$(2p-2)$-brane $-$ ${\bar {\rm D}}$-$(2p-2)$-brane pair. Note that now the manifold $\widetilde{\cal M}$ describing the minimum of the tachyon potential consists of a pair of points $\pm \widetilde T_0$. Thus $\pi_0(\widetilde{\cal M}) \ne 0$, and hence the kink is stable as is expected of a BPS D-brane. An argument similar to the one in the previous subsection can be used to give an intuitive explanation of why the kink should behave as a D-brane near $y=0$ but as vaccuum for large $|y|$. We can also explain the origin of the RR charge of the kink from the coupling \refb{etachrr} and the fact that $\partial_y T$ is non-zero at $y=0$. Since $\partial_y T$ has opposite sign for the anti-kink, this also shows that the anti-kink must represent the BPS ${\bar {\rm D}}$-$(2p-2)$ brane. \begin{figure}[ht] \begin{center} \leavevmode \epsfbox{fig7.eps} \end{center} \caption{Descent relations among BPS and non-BPS D-branes in type II string theories. The horizontal arrows denote the effect of modding out by $(-1)^{F_L}$, and the vertical arrows denote the ffect of considering tachyonic kink solution.} \label{f7} \end{figure} The results of this section, combined with the results of section \ref{ss2} leads to the set of `descent relations' between BPS and non-BPS D-branes shown in Fig.\ref{f7}. By combining the two main results of this section, we can also represent a BPS D-$p$-brane as a soliton (vortex) solution on the D-$(p+2)$-brane $-$ ${\bar {\rm D}}$-$(p+2)$-brane pair in the same theory\cite{9808141,9810188}. This construction is relevant for relating allowed D-brane charges to elements of the K-group of space-time\cite{9810188}. \sectiono{Stable Non-BPS Branes on the D-brane $-$ Orientifold Plane System} \label{ss5} \subsection{Summary of the results} \label{s1} We have already introduced the notion of a D-$p$-brane in type II string theory. We now introduce the concept of an orientifold $p$-plane (O-$p$-plane)\cite{ORIENT,9601038}. For this we consider type II string theory on $R^{p+1}\times (R^{9-p}/{\cal I}_{9-p}\cdot\Omega\cdot g)$, where ${\cal I}_{9-p}$ reverses the sign of all the coordinates on $R^{9-p}$, $\Omega$ is the world-sheet parity transformation (L $\leftrightarrow$ R) and $g$ is identity for $(9-p)=4m$ or $(4m+1)$ and $g=(-1)^{F_L}$ for $(9-p)=(4m+2)$ or $(4m+3)$. One can show that ${\cal I}_{9-p}\cdot\Omega\cdot g$ is a symmetry transformation of order 2 in type IIA string theory if $p$ is even, and in type IIB string theory if $p$ is odd. The origin of $R^{9-p}$ will be called an orientifold $p$-plane. Thus type IIA (IIB) string theory contains orientifold $p$-planes of even (odd) dimensions. \begin{figure}[|ht] \begin{center} \leavevmode \epsfbox{fig13.eps} \end{center} \caption{Transverse section of the coincident D-$p$-brane $-$ O-$p$-plane system. Although for clarity we have shown the D-brane and the O-plane as separated in space, we shall analyze the case where they are on top of each other.} \label{f13} \end{figure} Our focus of attention in this section will be a system of parallel D-$p$-brane $-$ O-$p$-plane system. This corresponds to starting with a D-brane and its image under ${\cal I}_{9-p}\cdot\Omega\cdot g$, and then modding out the theory by ${\cal I}_{9-p}\cdot\Omega\cdot g$, as shown in Fig.\ref{f13}. The world volume theory of a Dirichlet $p$-brane (D-$p$-brane) on top of an orientifold $p$-plane (O-$p$-plane) has as its low energy limit an $N=4$ supersymmetric SO(2) gauge theory.\footnote{There is some ambiguity in how we choose the action of this $Z_2$ transformation on the CP factors; and due to this ambiguity we can get different kinds of orientifold planes\cite{9601038,9712028}. Throughout this paper we shall only consider orientifold planes of SO-type $-$ also known as the O$^+$ planes\cite{9712028} $-$ carrying negative RR charge compared to that of a D-brane.} The spectrum of stable states in this theory contains a massive non-BPS state carrying unit charge under this SO(2) gauge field. These arise from open strings stretched between the D-brane and its image.\footnote{Before the orientifold projection the ground state in this sector is massless and corresponds to the charged vector bosons and their superpartners, but the orientifold projection removes this state from the spectrum.} In the weak coupling limit these states have mass of the order of the string scale $m_S$ with corrections expressible as a perturbation series in the string coupling $g_S$: \begin{equation} \label{eperturb} m_S(K_0+K_1 g_S + K_2 g_S^2 +\cdots)\, . \end{equation} Here $K_0,K_1,K_2,\ldots$ are numerical constants, with $K_m$ computed from a diagram with $m$ open string loops. Since the lowest mass state carrying SO(2) electric charge must be stable at all values of the string coupling, it makes sense to ask what would be the masses of these states in the strong coupling limit. This is one of the questions we address in this section. The answers were obtained in refs.\cite{9803194,9805019,9806155,9808141} and have been summarised in table 1. \begin{center} \begin{tabular}{|c|c|c|} \hline $p$ & mass & $C_p$ \\ \hline 6 & $C_6 m_S g_S$ & known \\ \hline 5 & $C_5 m_S (g_S)^{1\over 2}$ & known \\ \hline 4 & $C_4 m_S (g_S)^{1\over 3}$ & unknown \\ \hline 3 & unknown & $-$ \\ \hline \end{tabular} \medskip Table 1: Masses of electrically charged states on the D-$p$-brane O-$p$-plane system in the strong coupling limit. \end{center} In this table, the first column denotes the value of $p$, the second column denotes the mass of the lightest stable electrically charged state on the D-$p$-brane $-$ O-$p$-plane system, $C_p$ denotes a numerical constant, and the last column denotes whether the numerical constant $C_p$ is known or unknown at present. We have restricted $p$ in the range $3\le p\le 6$ due to the following reason. For $p\ge 7$, the dilaton does not go to a constant value asymptotically\cite{9605150}, and as a result the string coupling $g_S$ is not a well defined quantity. On the other hand, for $p\le 2$, the self-energy of an electrically charged particle blows up due to the long range Coulomb field associated with the particle, and hence the mass of such a state is not a well defined quantity. We shall review the arguments leading to these results in subsection \ref{s2}. As we can see from this table, we still do not know the mass of the electrically charged particle on the D-3-brane $-$ O-3-plane system in the strong coupling limit. Although it may be somewhat premature to look for a pattern among three data points, we note that there seems to be some regularity in the dependence of this mass on $g_S$ for $4\le p\le 6$, namely it seems to go as \begin{equation} \label{e1} m_S (g_S)^{1\over 7-p}\, . \end{equation} Considering that for different values of $p$ these results are derived using very different techniques, one might wonder if there is a deeper lesson about strongly coupled string theory in this spectrum. Since the only feature that is common between different values of $p$ is the structure of weak coupling perturbation theory, it is tempting to speculate that the regularity of the strong coupling spectrum is a reflection of the regularity of the weak coupling perturbation theory as a function of $p$. In that case we can expect that the information about the strong couping result is somehow contained in the weak coupling perturbation theory, $-$ in particular in its large order behaviour. Besides stable non-BPS states which are electrically charged under the SO(2), the brane world-volume theory also contains branes which are magnetically charged under the SO(2). On the D-$p$-brane O-$p$-plane system these are $(p-3)$ branes, and come from a D-$(p-2)$-brane, stretched between the brane and its image. Such configurations are allowed according to the rules of refs.\cite{9512059,9512062}. Naively, when the D-$p$-brane and its image coincide these stretched branes will have vanishing tension. But quantum corrections must give non-vanishing contribution to the tension, reflecting the fact that these are non-BPS branes.\footnote{Otherwise we should expect a singularity in the moduli space of this system for coincident D-brane $-$ orientifold plane system. This is known not to be present.} Unfortunately calculating tensions of these non-BPS branes in the weak coupling limit remains an open problem.\footnote{As we shall see later, this problem is related to finding the last row of table 1.} However as we shall see in subsection \ref{s2}, for every value of $p$ between 3 and 6, one can calculate the tensions of these non-BPS branes in the strong coupling limit. The answer has been summarized in table 2. \begin{center} \begin{tabular}{|c|c|c|} \hline $p$ & (tension)$^{1\over (p-2)}$ & $\widetilde C_p$ \\ \hline 6 & $\widetilde C_6 m_S$ & known \\ \hline 5 & $\widetilde C_5 m_S g_S^{-{1\over 6}}$ & known \\ \hline 4 & $\widetilde C_4 m_S g_S^{-{1\over 3}}$ & unknown \\ \hline 3 & $\widetilde C_3 m_S g_S^{-{1\over 2}}$ & known \\ \hline \end{tabular} \medskip Table 2: Tensions of magnetically charged $(p-3)$-branes on the D-$p$-brane O-$p$-plane system in the strong coupling limit. \end{center} The first column in this table describes the value of $p$ as before. The second column represents the $(p-2)$-th root of the tension of the magnetically charged $(p-3)$-brane. This root is taken in order to make it into a quantity of mass dimension 1. $m_S$ and $g_S$ denote, as before, the square root of the fundamental string tension and the string coupling constant respectively, and $\widetilde C_p$ denote numerical constants. The last column shows that at present the coefficients $\widetilde C_p$ are known for $p=3$, 5 and 6, but not for $p=4$. We again observe that there is a regularity in this spectrum. In particular the $(p-2)$-th root of the tension of the $(p-3)$-brane on the D-$p$-brane $-$ O-$p$-plane system goes as: \begin{equation} \label{e2} m_S (g_S)^{{p\over 6}-1}\, . \end{equation} Again it is natural to suspect that this reflects some deeper aspect of string theory which is not understood at present. \subsection{Strong coupling description of electrically charged states and magnetically charged branes} \label{s2} In this subsection we shall review the analysis leading to tables 1 and 2. We shall discuss each value of $p$ separately, since the strong coupling description of the D-$p$-plane $-$ O-$p$-plane system is different for each value of $p$. \noindent \underline{$p=6$} In this case the system under study is a D6-brane on top of an O6-plane in type IIA string theory. The strong coupling description of this system is known to be M-theory on $R^{6,1}\times {\cal N}$, where $R^{6,1}$ is along the world-volume of the D6$-$O6 system, and ${\cal N}$ is the double cover of the Atiyah-Hitchin space\cite{AH} with a rescaled metric\cite{9607163,9803194}. Asymptotically, ${\cal N}$ locally looks like $R^3\times S^1$. The Planck mass $m_p$ of the M-theory, and the radius $R$ of this $S^1$ are related to $m_S$ and $g_S$ of type IIA string theory via the relations: \begin{equation} \label{e3} m_p=m_S (g_S)^{-{1\over 3}}, \qquad R = m_S^{-1} g_S\, . \end{equation} The metric on ${\cal N}$ is given by \begin{equation} \label{eathi} ds^2 = {R^2\over 4} ds_{AH}^2\, , \end{equation} where $ds_{AH}^2$ is the standard Atiyah-Hitchin metric\cite{AH}. The SO(2) gauge field $A$ on the brane world-volume is related to the three form gauge field $C_{\mu\nu\rho}$ of M-theory as \begin{equation} \label{ex1} C = \omega \wedge A + \cdots \end{equation} where $\omega$ is the unique normalizable harmonic two form on ${\cal N}$\cite{GIBRUB,MANSCH,9402032}, and $\cdots$ denotes terms involving other normalizable and non-normalizable differential forms on ${\cal N}$. The topology as well as the metric on ${\cal N}$ is completely known. In particular ${\cal N}$ contains a non-contractible two cycle of minimal area $-$ called the bolt $-$ which has the property that the integral of the two form $\omega$ over the bolt is non-vanishing. {}From the relation \refb{ex1} and the fact that a membrane is electrically charged under $C$, it follows that a membrane wrapped on the bolt will be electrically charged under $A$. In other words, the electrically charged stable non-BPS state on the world-volume of the D6$-$O6 system is described by the M-theory membrane wrapped on the bolt of ${\cal N}$\cite{9803194}. The area of the bolt is equal to $\pi^3 R^2$. On the other hand, the membrane tension is proportional to $m_p^3$. Thus the mass of the state is given by: \begin{equation} \label{e4} C_6 m_p^3 R^2= C_6 m_S g_S\, , \end{equation} where $C_6$ is a known constant. Following the same logic, the magnetically charged three brane on the D6-O6 world-volume can be identified as the M-theory five-brane wrapped on the bolt of ${\cal N}$. The tension of this 3-brane can be calculated by multiplying the five-brane tension ($m_p^6$) with the area of the bolt. This is given by \begin{equation} \label{ex2} (\widetilde C_6)^4 m_p^6 R^2 = (\widetilde C_6)^4 (m_S)^4\, , \end{equation} where $\widetilde C_6$ is a known numerical constant. Eqs.\refb{e4} and \refb{ex2} reproduce the first rows of tables 1 and 2 respectively. \noindent \underline{$p=5$} The system under study is a D5-brane on top of an O5-plane in type IIB string theory. In the strong coupling limit, this theory is S-dual to the weakly coupled type IIB string theory on $R^{5,1}\times (R^4/(-1)^{F_L}\cdot{\cal I}_4)$ where $R^{5,1}$ is along the D5-O5 world volume, ${\cal I}_4$ changes the sign of the coordinates of $R^4$ $-$ the directions transverse to the D5-O5 world-volume, $-$ and $(-1)^{F_L}$ changes the sign of all the Ramond sector states on the left-moving sector of the string world-sheet\cite{9604070}. This can be argued by noting that under S-duality of type IIB string theory $\Omega$ gets transformed to $(-1)^{F_L}$ and a D5-brane is transformed to an NS 5-brane. Thus naively one would think that the dual system should correspond to the orbifold described above together with an NS 5-brane. But upon examining the spectrum of massless states originating in the twisted sector of the orbifold theory one finds that they are already in one to one correspondence with the massless degrees of freedom living on the D5-O5 system. Thus there is no need to add another NS five brane; in fact adding it will double the number of massless degrees of freedom, and will describe the dual of a system of two D5-branes on top of an O5-plane. The relationship between the string scale $\widetilde m_S$ and the coupling constant $\widetilde g_S$ of this dual theory, and those of the original theory is given by: \begin{equation} \label{e5} \widetilde m_S = m_S (g_S)^{-{1\over 2}}, \qquad \widetilde g_S = (g_S)^{-1}\, . \end{equation} The SO(2) gauge field on the D5-brane $-$ O5-plane world volume corresponds to massless vector fields originating in the twisted sector of this orbifold theory. The state carrying electric charge under the SO(2) gauge field corresponds to, in this orbifold theory, the non-BPS D0-brane of IIB placed on the orbifold plane\cite{9805019,9806155,9808141}. This has mass \begin{equation} \label{e6} C_5 \widetilde m_S (\widetilde g_S)^{-1} = C_5 m_S (g_S)^{1\over 2}\, , \end{equation} where $C_5$ is a known constant. Similarly the two brane carrying magnetic charge under this SO(2) gauge field corresponds to a non-BPS D2-brane of type IIB string theory, placed inside the orbifold fixed plane. Its tension is given by: \begin{equation} \label{e7} (\widetilde C_5)^3 \widetilde m_S^3 (\widetilde g_S)^{-1} = (\widetilde C_5)^3 (m_S)^3 (g_S)^{-{1\over 2}}\, , \end{equation} where $\widetilde C_5$ is another known constant. Eqs.\refb{e6} and \refb{e7} reproduce the second rows of tables 1 and 2 respectively. \noindent \underline{$p=4$} The configuration under study is a D4-brane on top of an O4-plane in type IIA string theory. The strong coupling limit of this theory is best described as M-theory on $R^{4,1}\times S^1\times (R^5/{\cal I}_5\cdot\sigma)$, together with a five-brane (and its image under ${\cal I}_5\cdot\sigma$) placed at the origin of $R^5$ with its world-volume extending along $R^{4,1}\times S^1$\cite{9803194}. Here $R^{4,1}$ is along the world-volume of the original D4-O4 system, $S^1$ is a circle of radius $R$ given in eq.\refb{e3}, ${\cal I}_5$ reverses the sign of the coordinates of $R^5$ transverse to the brane world-volume, and $\sigma$ denotes the transformation which changes the sign of the three form gauge field of M-theory. This can be seen by noting that under the type IIA - (M-theory on $S^1$) duality, $\Omega$ of type IIA is mapped to $\sigma$ of M-theory, and the four brane of type IIA is mapped to a five brane of M-theory wrapped on $S^1$. The Planck mass $m_p$ of M-theory is given in terms on $m_S$ and $g_S$ as in eq.\refb{e3}. \begin{figure}[|ht] \newpage \begin{center} \leavevmode \epsfbox{fig14.eps} \end{center} \caption{Membrane stretched between the M5-brane and its image under ${\cal I}_5\cdot\sigma$. We shall consider the case where the 5-brane (and its image) coincide with the orbifold plane.} \label{f14} \end{figure} The five-brane world-volume carries a self-dual anti-symmetric tensor field $B_{MN}$. The component $B_{1\mu}$, where $x^1$ denotes the coordinate along $S^1$ and $\mu$ is the coordinate along $R^{4,1}$, is the SO(2) gauge field $A_\mu$ on the D4-O4 system. As displayed in Fig.\ref{f14}, the world-volume of the five brane placed at the origin of $R^5$ also contains a non-BPS string from the membrane stretched between the five-brane and its image under ${\cal I}_5\cdot\sigma$\cite{9803194}.\footnote{Classically this string should have zero tension when the five-brane approaches its image, but if this had been true also quantum mechanically then the moduli space would have a singularity when the five brane coincides with the orbifold plane. Using the duality between M-theory on $T^5/Z_2$ and type IIB on K3\cite{9512196,9512219}, one can see that there is no singularity in this region of the moduli space.} Although we cannot explicitly compute the tension of this string, by dimensional analysis we see that this tension must be proportional to $m_p^2$, since this is the only scale in the problem. Since the non-BPS string carries $B_{MN}$ charge, this string wrapped on $S^1$ will be electrically charged under $B_{1\mu}=A_\mu$. The mass of this state is given by: \begin{equation} \label{e8} C_4 m_p^2 R = C_4 m_S (g_S)^{1\over 3}\, , \end{equation} where $C_4$ is an unknown numerical constant. On the other hand, the non-BPS string with world-volume along $R^{4,1}$ will be magnetically charged under the gauge field $B_{1\mu}=A_\mu$, and its tension will be given by: \begin{equation} \label{e9} (\widetilde C_4)^2 m_p^2 = (\widetilde C_4)^2 (m_S)^2 (g_S)^{-{2\over 3}}\, , \end{equation} where $\widetilde C_4$ is a numerical constant related to $C_4$. Eqs.\refb{e8} and \refb{e9} reproduce the third rows of tables 1 and 2 respectively. \noindent \underline{$p=3$} The system under study is the D3-brane on top of an O3-plane in type IIB string theory. The strong coupling limit of this system is dual to a weakly coupled type IIB string theory in the same background, with the parameters of the dual theory related to those in the original theory by eq.\refb{e5}. The electrically charged state in the original theory is mapped to the magnetically charged state in the dual theory. Unfortunately at present we do not know anything about this state, as was discussed earlier in subsection \ref{s1}. On the other hand, the magnetically charged state in the original theory is mapped to the electrically charged state in the dual theory. This is a perturbative open string state, and has mass proportional to $\widetilde m_S$ for small $\widetilde g_S$. Thus the mass of the magnetically charged state in the original theory in the strong coupling limit is give by: \begin{equation} \label{e10} \widetilde C_3 \widetilde m_S = \widetilde C_3 m_S (g_S)^{-{1\over 2}}\, , \end{equation} where $\widetilde C_3$ is a known numerical constant. This reproduces the last row of table 2. \subsection{Electrostatic self-energy of the electrically charged non-BPS particle on the D3-brane $-$ O3-plane system} \label{s3} In this subsection we shall give a lower bound on the electrostatic self-energy of the electrically charged non-BPS particle on the D3-brane $-$ O3-plane system in the strong coupling limit. To do this we go to the dual weakly coupled description where this particle corresponds to a magnetically charged particle on the D3-brane $-$ O3-plane world-volume. Although we do not know at present how to explicitly construct this state, it is clear that sufficiently far away from the center, the magnetic field around the state will look like the magnetic field of a point monopole. Let $r_c$ be the distance beyond which this happens. If we normalize the gauge field on the D3-brane so that the action has the form: \begin{equation} \label{ez1} {1\over \widetilde g_S}\int d^4 x F_{\mu\nu} F^{\mu\nu}\, , \end{equation} where $\widetilde g_S$ as usual is the string coupling constant in this dual string theory, then the magnetic field for $r>>r_c$ is of order $(1/r^2)$, and hence its contribution to the total energy of the system from the region $r\ge r_c$ is of order \begin{equation} \label{ez2} {1\over \widetilde g_S}\int_{r\ge r_c} {d^3 r\over r^4} \sim (\widetilde g_S)^{-1}(r_c)^{-1}\, . \end{equation} In order to give a lower bound to this expression we need an upper bound on $r_c$. This is obtained by noting that $r_c$ cannot be larger than the string scale $(\widetilde m_S)^{-1}$ in this dual string theory, since for small $\widetilde g_S$ we expect the lightest massive states in this theory to have mass of order $\widetilde m_S$. Thus beyond the distance $(\widetilde m_S)^{-1}$, the magnetic field of the monopole should approach that of a point monopole. This gives the following lower bound to the magnetostatic energy: \begin{equation} \label{ez3} \widetilde m_S (\widetilde g_S)^{-1} \sim m_S (g_S)^{1\over 2}\, . \end{equation} This exceeds the expected answer $m_S (g_S)^{1\over 4}$ from eq.\refb{e1}. This suggests that eq.\refb{e1} is applicable, if at all, only to the `intrinsic mass' of the non-BPS particle (if it could be defined at all), and cannot account for the contribution from the Coulomb energy. Presumably the issue will be clarified once we have an explicit construction of this non-BPS state. It is the same problem which appears in a more severe form in the case of $p=2$. Here the electrostatic self-energy is infinite, and completely masks the `intrinsic mass' of the particle. \sectiono{Some Related Developments} \label{ss6} In this section we shall briefly discuss some other related developments in this field. In particular, we shall discuss \begin{enumerate} \item construction of other non-BPS states in type I string theory\cite{9810188}, \item relationship between D-brane charge and K-theory\cite{9810188,9812135}, and \item application of boundary state formalism to the study of non-BPS states\cite{9806155}. \end{enumerate} There are several other related developments\cite{9901159,9902158,9903129,9812003,9807138,9804160,9808073,9902181} which will not be discussed here. At the end we shall also briefly discuss some open problems. \subsection{Other non-BPS branes in type I string theory} In the same way that we constructed a D0-brane in type I string theory, one can construct a D8-brane in this theory. The idea is to start with the non-BPS D8-brane of type IIB string theory, and mod it out by the world-sheet parity transformation $\Omega$. The result is a non-BPS D8-brane of type I string theory. The tachyonic mode of the open string with both ends on the 8-brane is projected out as in the case of the D0-brane. However in type I string theory there are also space filling D9-branes, and it turns out that open strings with one end on the D8-brane and the other end on a D9-brane has tachyonic modes which are not projected out\cite{9903123}. Thus these branes are not stable. One can also construct non-BPS D-instantons in the type I string theory as follows\cite{9810188}. We can start from a D-instanton anti-D-instanton pair of type IIB string theory, and mod out the theory by the world-sheet parity transformation $\Omega$. The result is a non-BPS D-instanton of type I string theory. One can show that the tachyonic mode is projected out under this operation; so that the D-instanton is a stable configuration of type I string theory. A similar construction can be done by starting with a D7-brane $-$ ${\bar {\rm D}}$7-brane pair of type IIB string theory, and modding out the configuration by $\Omega$. Again the tachyonic mode originating in open strings with both ends on the D7-brane is projected out. But in this case there is a tachyonic mode in the open strings with one end on the D7-brane and the other end on the D9-brane. Thus the D7-brane is not a stable configuration in type I string theory\cite{9903123}. \subsection{K-theory} Another related development in this field has been the discovery of the relationship between elements of the K-group of space-time manifold and D-brane charges on the same manifold\cite{9810188}. This is related to the idea of representing a D-brane as a tachyonic soliton on a D-brane - anti-D-brane pair of higher dimensions. The simplest example is that of type IIB string theory, so we shall only discuss this case. In this case, following the discussions of section \ref{ss4} we see that a BPS D-$(2p+1)$-brane can be regarded as a soliton solution on a D-$(2p+3)$ $-$ ${\bar {\rm D}}$-$(2p+3)$-brane pair. Each of the D-$(2p+3)$ branes on the other hand can be regarded as a soliton solution on a D-$(2p+5)$-brane ${\bar {\rm D}}$-$(2p+5)$-brane pair. Following this argument we see that each stable D-brane in type IIB string theory can be regarded as some kind of soliton solution on a sufficient number ($N$) of 9-brane $-$ anti-9-brane pair. It was shown by Witten\cite{9810188} that the solution representing a system of D-branes (possibly with some gauge field configurations on them) can be completely classified by specifying the $U(N)$ gauge bundles $E$ and $F$ on the 9-brane and the anti-9-brane which characterize the gauge field configurations corresponding to this soliton. Furthermore, if we add equal number of extra 9-branes and anti-9-branes to the system with identical gauge bundles $H$ on them, then the tachyon associated with the open strings stretched between the 9-brane and the anti-9-brane is a section of a trivial bundle, and hence can condense to the minimum ($T_0$) of the potential everywhere on the 9-brane $-$ anti-9-brane world-volume. Since this configuration is identical to the vacuum, we conclude that adding such extra pairs of 9-brane and anti-9-brane has no effect on the topological class of the soliton. Thus the D-brane charges are classified by specifying a pair of U(N) vector bundles $(E,F)$ subject to the equivalence relation \begin{equation} \label{ek1} (E,F)\equiv (E+H,F+H)\, , \end{equation} for any U(M) vector bundle $H$. This is precisely the definition of the K-group of the space-time manifold. This is the basic idea of using K-theory to classify D-brane charges. Similar analysis can be carried out for type I and type IIA string theories as well. In type I theory the starting point is the representation of all D-branes as solitons on D9-${\bar {\rm D}}$9-brane system\cite{9810188}, whereas in type IIA string theory the starting point is the representation of a D-brane as a soliton on a system of non-BPS D9-branes\cite{9812135}. \subsection{Boundary state approach to non-BPS branes} The boundary state approach\cite{POLCAI,CLNY,ONOISH,ISHIBASHI,BOUNDARY} to the study of non-BPS D-branes was pioneered by Bergman and Gaberdiel\cite{9806155}. Corresponding to any D-brane in string theory, we can associate a boundary state $|B\rangle$ in the closed string sector whose inner product with a closed string state describes the amplitude for a closed string emission from the D-brane. Furthermore, if $|B\rangle$ and $|B'\rangle$ denote the boundary states associated with a pair of (not necessarily identical) D-branes, then $\langle B|B'\rangle$ describes the one loop partition function of an open string stretched from the first D-brane to the second D-brane. The boundary state $|B_p\rangle$ of a BPS D-$p$-brane in type IIA or type IIB string theory can be written as a sum of two terms: \begin{equation} \label{ebou1} |B_p\rangle = {1\over\sqrt 2}(|NSNS_p\rangle + |RR_p\rangle)\, , \end{equation} where $|NSNS_p\rangle$ and $|RR_p\rangle$ denote the contribution to the boundary state from the NSNS and RR sector closed strings respectively. The inner product $\langle B_p|B_q\rangle$ can be expressed as \begin{equation} \label{ebou2} \langle B_p|B_q\rangle = Tr_{p-q}{1 + (-1)^F\over 2} \, , \end{equation} where $Tr_{p-q}$ denotes trace over the open string states stretched from the D-$p$-brane to the D-$q$-brane. In this equation the contribution proportional to 1 comes from the NSNS component of the boundary state, whereas the contribution proportional to $(-1)^F$ comes from the RR component of the boundary state. In this notation the boundary state $|\widetilde B_p\rangle$ describing the non-BPS D-$p$-brane of type IIB or type IIA string theory is given by\cite{9806155,9809111} \begin{equation} \label{ebou3} |\widetilde B_p\rangle = |NSNS_p\rangle\, . \end{equation} Note that the contribution from the RR sector is absent, reflecting the fact that the non-BPS D-brane does not carry any RR charge. Also the NSNS sector contribution to \refb{ebou3} has an extra factor of $\sqrt 2$ compared to that in \refb{ebou1}. This reflects the fact that the non-BPS D-brane has an extra multiplicative factor of $\sqrt 2$ in its tension. {}From eqs.\refb{ebou1}-\refb{ebou3} it follows that \begin{equation} \label{ebou4} \langle \widetilde B_p|\widetilde B_p \rangle = Tr_{p-p}(1)\, . \end{equation} This shows that the partition function of open string states living on the non-BPS D-brane has no GSO projection. One can also analyse the fate of stable non-BPS D-branes in various orbifolds and orientifolds of type II string theories using the boundary state approach. Let us consider, for example, the case of type I D-particle. In this case the boundary state is described by \begin{equation} \label{ebou5} {1\over \sqrt 2} (|\widetilde B_0\rangle + 32 |B_9\rangle + |C\rangle)\, , \end{equation} where $|\widetilde B_0\rangle$ is the boundary state of the non-BPS D0-brane (the $(1/ \sqrt 2)$ factor is due to the $\Omega$ projection), $32|B_9\rangle$ denotes the boundary state corresponding to the 32 BPS D9-branes in the vacuum, and $|C\rangle$ is the crosscap state\cite{POLCAI,CLNY,ONOISH,ISHIBASHI} reflecting the effect of $\Omega$ projection. The terms involving $|\widetilde B_0\rangle$ in the inner product of this boundary state with itself is given by \begin{equation} \label{ebou6} {1\over 2}(\langle \widetilde B_0|\widetilde B_0\rangle + \langle \widetilde B_0|C\rangle + \langle C|\widetilde B_0\rangle +32 \langle \widetilde B_0|\widetilde B_9\rangle + 32 \langle \widetilde B_9 | \widetilde B_0\rangle)\, . \end{equation} The sum of the first three terms gives \begin{equation} \label{ebou7} Tr_{0-0}{1+\Omega\over 2}\, , \end{equation} where $Tr_{0-0}$ denotes trace over open strings with both ends on the D0-brane. On the other hand the last two terms give \begin{equation} \label{ebou8} 32 Tr_{0-9}(1)\, , \end{equation} where $Tr_{0-9}$ denotes the trace over open string states stretched from the D0-brane to the D9-brane. There is no $\Omega$ projection in this term, since $\Omega$ relates these open strings to open strings stretched from the D9-brane to the D0-brane. Thus the effect of $\Omega$ projection is to simply include either the $0-9$ or the $9-0$ sector, but not both. Since $|\widetilde B_0\rangle$, $|C\rangle$ and $|B_9\rangle$ are all explicitly known, we can evaluate each term in \refb{ebou6} explicitly. Comparing these with \refb{ebou7} we can explicitly derive the $\Omega$ projection rules for the open strings with both ends on the non-BPS D0-brane, and check that these rules agree with the ones derived following the arguments in subsection \ref{stypei}. In particular, one can verify that the tachyonic mode on this D-particle is projected out under $\Omega$. \subsection{Open questions and speculations} We shall conclude this article by reviewing some of the open questions and with some speculations. \begin{enumerate} \item The various arguments given in favour of the idea that the tachyonic ground state on the brane anti-brane pair is indistinguishible from the vacuum are all indirect, and involves first compactifying one or more directions tangential to the brane world-volume, followed by switching on the tachyon vev and then taking the radius back to infinity. A direct proof of this on a non-compact brane-antibrane pair, presumably based on the construction of an explicit classical solution in the open string field theory on the brane-antibrane pair describing the tachyonic ground state, is still lacking. Similarly, one should be able to construct an explicit classical solution in this open string field theory representing the tachyonic kink solution and show that this solution describes a non-BPS D-brane. \item One of the difficulties in understanding the phenomenon of tachyon condensation on the brane-antibrane pair has been in understanding what happens to the various U(1) gauge fields living on the original system. The tachyon is charged under one combination of the two U(1) gauge fields, and hence breaks this gauge symmetry. However the other linear combination, which we shall denote by ${\cal A}_\mu$, does not get broken since the tachyon, as well as all other perturbative open string states living on the brane-antibrane world-volume, are neutral under this gauge field. It has been suggested in ref.\cite{9901159} that the other U(1) is in the confining phase. The suggested mechanism for this confinement is the condensation of the tachyonic $(p-3)$-branes obtained from D-$(p-2)$-branes stretched between the original D-$p$-brane and the anti-D-$p$-brane. Thus for example for $p=3$, it involves condensation of the tachyonic mode of the D-string stretched between the D3-brane and the ${\bar {\rm D}}$3-brane. It was shown in \cite{9901159} that this tachyon is magnetically charged under ${\cal A}_\mu$, and hence condensation of this tachyon will imply that the corresponding U(1) gauge theory is in the confining phase. Whereas the general idea is quite appealing, this mechanism is highly non-perturbative from the point of view of the world-volume theory of the D3-brane $-$ ${\bar {\rm D}}$3-brane pair. On the other hand the indirect arguments reviewed in this article showing that the tachyonic ground state is identical to the vacuum configuration are based on open string tree level analysis. Thus there must be a way to see the phenomenon of confinement of the U(1) gauge field ${\cal A}_\mu$ at open string tree level. Presumably once we understand how to describe tachyon condensation using classical open string field theory, this issue will be automatically resolved. \item Another open problem, which has already been discussed earlier, is the construction of magnetically charged non-BPS D-$(p-3)$-brane on the D-$p$-brane $-$ O-$p$-plane system. It is clear that these stable branes must exist in the spectrum, so one should be able to find them in the weakly coupled string theory. \item It would be interesting to investigate the relationship between weak coupling perturbation expansion for the mass of a non-BPS state and its strong coupling limit. This might give us new insight into string theory at finite coupling. \item One of the main lessons from our analysis (and of refs.\cite{9510227,9704006,9708075}) is that the existence of tachyons in the spectrum of open string theory does not necessarily signify a sickness of the theory, but often simply indicates the existence of a ground state with energy (density) lower than that of the starting configuration. It would be interesting to investigate if closed string tachyons have a similar interpretation. \item {}From our discussion in this article it is clear that all D-branes in type IIA (IIB) string theory can be regarded as classical solutions in the open string field theory living on a system of non-BPS D9-branes (D9-brane $-$ ${\bar {\rm D}}$9-brane pair). It would be interesting to see if this can also be done for other known objects in string theory, namely the fundamental string and the NS 5-branes.\footnote{K-theory does not contain these states, but K-theory uses only a small subset of available open string fields, namely the tachyon and the gauge bosons.} Actually fundamental strings appear as bound state poles in the S-matrix computed from Witten's open string field theory\cite{WITTBFT,WITTSFT,CUBIC}. On the other hand a formal construction was presented in \cite{CUBIC} showing that any string background represented by a two dimensional conformal field theory (of which the NS five-brane is an example) can be represented as a classical solution in the purely cubic open string field theory. If these results can be made more concrete, then one could take open string field theory on the non-BPS D9-brane (D9-${\bar {\rm D}}$9 brane pairs) as the fundamental formulation of type II string theories and their orbifolds/orientifolds, since all states in string theory could be constructed from this field theory. \end{enumerate} \noindent{\bf Acknowledgement}: I would like to thank O.~Bergman, S.~Elitzur, M.~Gaberdiel, P.~Horava, N.~Manton, B.~Pioline, E. Rabinovici, A.~Recknagel, V.~Schomerus and E. Witten for useful correspondence at various stages of this work. I would also like to thank the organisers of the APCTP winter school for organising an excellent school, and the Tata Institute at Mumbai, IAS at the Hebrew University and CTP at MIT for their hospitality during the preparation of this manuscript.
\section{Introduction} \label{sec:intro} \subsection{Background} The principles of thermodynamics were established in the last century as the universal laws characterizing the thermal and mechanical behavior of macroscopic systems. The fact that we cannot control all the details of energy transfer leads to the concept of {\it heat} as a form of energy flow, and the {\it Carnot cycle} has played a crucial role in the course of investigation leading to the introduction of entropy as a state variable in addition to energy\cite{Fermi}. On the other hand, Brownian motion and the stochastic dynamics of {meso}scopic systems in general have also been studied for many years, and projection methods have allowed for the derivation of Langevin dynamics from microscopic Hamiltonian mechanics. In a properly defined Langevin equation, the influence of the unpredictable microscopic dynamics, which essentially represent the heat, is taken into account by Markovian random forces obeying the fluctuation-dissipation relationship. In this manner, such an equation describes the canonical equilibrium distribution of the variables in question\cite{LandauSTD}. Very recently, the concept of the heat on mesoscopic scales has been unambiguously defined in terms of Langevin dynamics\cite{ks1}. We refer to the formalism providing this definition as {\it stochastic energetics}.% The essential point of the thinking behind this formalism is that the heat transfered to the system is nothing but the microscopic work done by both frictional forces {\it and} the random force in the Langevin equation. The theoretical framework resulting from this realization widens the scope of application of Langevin dynamics to the extent that it can be used to describe not merely equilibrium states of system in contact with heat baths, but also general thermodynamic {\it processes} connecting different equilibrium states. As a result, we can derive the first and the second laws of thermodynamics\cite{ks1,ks-ss} from stochastic energetics. This formalism, together with projection methods, bridge a longstanding gap between microscopic Hamiltonian mechanics and macroscopic thermodynamics. In this paper we apply the method of stochastic energetics to the investigation of the Carnot cycle in the context of small systems. To make this paper self-contained, we briefly summarize the framework of stochastic energetics in \S~\ref{sec:steng}. Stochastic energetics has also been applied to the study of thermodynamic processes under non-equilibrium conditions, such as processes including two heat baths\cite{ks1} and processes in the presence of steadily driving forces\cite{ks2}. In particular, Feynman's ratchet model\cite{feynman} has been analyzed. Regarding this model, the doubt has been cast by Parrondo\cite{parrondo}, and later by Sekimoto\cite{ks1} independently, about the attainability of reversible energy conversion with the `Carnot efficiency' $1-T_{\rm L}/T_{\rm H}$, where $T_{\rm L}$ and $T_{\rm H}$ are the temperatures of cool and hot heat baths. Analysis using stochastic energetics has shown explicitly that the efficiency of Feynman's ratchet is much less than that of the Carnot efficiency mentioned above\cite{ks1}. \subsection{Problems} \label{subsec:problems} With the descriptive power of stochastic energetics in hand, we wish to reconsider the Carnot cycle. We consider the Carnot cycle as an {\it object} of analysis within the theoretical framework of stochastic energetics. Note that since we can derive the laws of thermodynamics directly using the stochastic energetics based on the Langevin description, the role of the Carnot cycle in our study is {\it not} a source of theoretical results from which one derives the laws of thermodynamics as was its historical role. (Of course both the Langevin description and thermodynamics have a microscopic basis in mechanics.) We now describe our viewpoint in more detail. Usually the Carnot heat engine is considered in an ideally macroscopic context,working in the thermodynamic limit. There, the small relative fluctuations of the variables, typically on the order of the inverse square root of the system size, are neglected. Also, the cost involved in the operations of attaching/detaching the system under study to/from heat baths is neglected, since this is not an the extensive quantity. It is important to notice that the second law of thermodynamics, which is consistent with such a macroscopic Carnot engine, can exclude {\it only marginally} the existence of perpetual machine of the second kind whose cycles yield positive work in an isothermal environment. Thus we may gain a deeper insight into the nature of statistical thermodynamics and mechanics if we can formulate a method to take account of the finiteness of the the system under study as well as the cost involved in operations of changing its interaction with heat baths, in particular considering reversibility and the second law of thermodynamics. The approach of the present work is to construct the simplest model of the Carnot heat engine with a finite number (actually only three) of degrees of freedom, including the apparatus connecting/disconnecting it with heat baths, and to determine the effect of the finiteness of the system and the change resulting from operations of the type mentioned above. As the system of study (or the `working material') we choose a single harmonic oscillator. We show that there is an inevitable source of dissipation due to the {\it intrinsically} irreversible nature of the operations of connecting and disconnecting it with heat baths, and that, with the exception of such loss, our model can attain the Carnot maximal efficiency defined as a properly defined average over infinitely many cycles, each of which is performed infinitely slowly. At the same time this study reveals several basic problems which should be further scrutinized in the future: one regards the smooth connection between the adiabatic process and the isothermal processes, and the other regards the irreversibility of adiabatic processes. In the last section we discuss these problems as well as the problem involving energy conversion with no help from external operations. In the remaining part of this section we give a qualitative description of the aspects of the Carnot cycle that we study in detail in the later sections.\\ \noindent {\it 1.} {\sf The operations of connection to and disconnection from the heat bath.}\\ We ask first how we can describe {\it mechanically} the connection and disconnection of the system with the heat baths. In an idealized picture, this basically consists of the switching on and off of the interaction between the system and each heat bath. In \S~\ref{sec:model} we describe explicitly a model that realizes these operations. We represent the influence of the heat baths by a frictional force and the random force of a Langevin equation, and we control the strength of the coupling between the system and the heat baths by controlling the values of the corresponding interaction potentials. We call these interaction potentials `couplers'. (In an actual mechanical system, the control of such a coupler could be exercised by a system of clutches.) One could also imagine such control exercised through the change of the friction constants that appear in the Langevin equation. In consideration of the absence of a definition of the required work to change these friction constants, however, this idea is not pursued in the present paper. .\\ \noindent {\it 2.}{\sf The reversible and irreversible work of operating the couplers}\\ The operation of the couplers can, {\it in principle}, never be carried out quasi-statically, but, at the same time, that accompanying irreversible work can be made arbitrarily small. The former part of this assertion is based on the following argument: When the interaction between the system and a heat bath is strong, the energy transfer between them occurs with a short relaxation time. However, if we gradually weaken this interaction, this relaxation time increases more and more until it diverges when the system is completely detached from the heat bath. As long as the time-scale of the operation (i.e. the switching-off) is finite, this operation can never remain `slow' in comparison to the diverging relaxation time. Thus the switching-off process is by no means quasi-static, or quasi-equilibrium. (This is analogous to the non-adiabaticity encountered in chemical reactions; the Born-Oppenheimer approximation is inevitably invalid when the distance between nuclei is neither sufficiently large nor sufficiently small.) Inevitable irreversibility can also exist in the process of strengthening the interaction between the system and the heat bath. Such extreme strengthening of the interaction leads to the freezing of some degree(s) of freedom involved in the interaction, and the mean first-passage time associated with these degrees of freedom may become larger than the timescale of the operation (i.e. the strengthening). In such a situation also, the operation can never be carried out quasi-statically. Unlike the switching-off process, however, the indefinite strengthening of the interaction is not necessarily a part of the Carnot cycle. Despite this fact, we consider the latter process in the sections that follow, because this allows us to minimize the calculations needed to reach a general conclusion. We should, however, note that the inevitably irreversibile nature of the operations described above does not necessarily imply an associated large amount of irreversible work. In \S~\ref{sec:clutch} we analyze the work involved in operating the coupler and show that the amount of irreversible work resulting from these inevitably irreversible operations can be made arbitrarily small in the limit that the timescale of the operation becomes large. We also show that the reversible part of the work associated with these operations remains finite in this limit, but that it cancels out within a cycle.\\ \noindent {\it 3.}{\sf The condition for reversible contact between the system and a heat bath}\\ Temporarily putting aside the concept of irreversibility in the sense described in {\sf 2} above, we can scrutinize the remaining part of the cycle and ask if and how the Carnot maximum efficiency can be attained. With regard to a macroscopic Carnot cycle, according to textbook descriptions, in order to realize a reversible cycle, `the temperature of the system should be the same as that of the heat bath with which the system is to make contact after an adiabatic process.' Strictly speaking, however, the energy, rather than the temperature, takes a definite value in a thermally isolated system, and the above statement needs to be refined in terms of the language of probability. We argue in \S~\ref{sec:match} that reversible contact requires the probability distribution of the energy of the system just before interaction with a heat bath to be identical to the canonical distribution at the temperature of the heat bath. This condition can be satisfied if the system consists of harmonic oscillators, as the model described below. Generally, however, this is not the case, and in the general situation an irreversible process takes place when system contacts the heat bath, even though there occurs no net irreversible energy transfer between the two (\S~\ref{subsec:1}). In \S~\ref{sec:efficiencymax} we summarize the necessary conditions for Carnot cycle to realize the maximal efficiency $1-T_H/T_L$ without assuming the thermodynamic limit. We show at the same time that this actually is the case for the model described in \S~\ref{sec:model}.\\ \noindent {\it 4.} {\sf Statistics of the efficiency of a finite number of cycles} \\ The efficiency of the energy conversion of an {\it individual} cycle is statistically distributed, because the energy possessed by the system is different each time the system is disconnected from a heat bath. This fact reflects the indeterminate nature of the details of the microscopic states of the system and of the heat bath upon disconnection.% As a consequence, if we define the cumulative `bonus' work as the difference between the cumulative work obtained over $n$ cycles and what we would expect from the Carnot maximal efficiency, this bonus work takes the form of a discrete random walk as a function of $n$. We show in \S~\ref{sec:efficiency} and the Appendix that the so-called null-recurrence property of a one-dimensional random walk insures that, although, if we actually carry out a sequence of these cycles, the cumulative bonus work we obtain will with probability 1 first become positive after a finite number of repetitions $n^*$, the statistical average of $n^*$ is infinite. \noindent % {\it 5.} {\sf Irreversibility of adiabatic processes} \\ The Carnot cycle includes an adiabatic process that is purely mechanical. We are interested in determining what work can be obtained through the cycle including {\it non}-quasi-static adiabatic processes. If the efficiency in this case is increased in comparison to the quasi-static case, the existence of a perpetual machine of the second kind is inspired, because our Carnot cycle can attain the maximal (reversible limit) efficiency under certain conditions specified in \S~\ref{sec:efficiencymax}. In \S~\ref{subsec:adiab} we show that, in relation to the impossibility of a perpetual machine, there emerges the concept of the irreversibility of purely mechanical processes (with no assumption of the thermodynamic limit or mixing properties necessary) under the influence of `macroscopic' operations by an external agent. Here, designation of an operation as `macroscopic' implies that (i) we are ignorant of the initial phase point on a given energy surface, and (ii) we are interested only in the statistical average over such an initial ensemble at a given energy. \section{Brief summary of stochastic energetics} \label{sec:steng} We consider a Langevin equation that represents a system in contact with a heat bath at temperature $T$, \begin{eqnarray} \frac{dx}{dt}&=& \frac{p}{m}\cr \frac{dp}{dt}&=&-\gamma \frac{p}{m}-\frac{\partial U(x,a)}{\partial x}+\xi(t). \label{eq:Langevin2} \end{eqnarray} Here we denote by $x$ and $p$ the dynamical variable of the system and its conjugate momentum, while $m$ represents the mass, $\gamma$ the friction constant, and $U$ the potential energy for $x$. We assume that $U$ may depend on, in addition to $x$, the variable (or variables) $a$, which is controlled by an external agent (or agents). The function $\xi(t)$ represents, as usual, Gaussian white noise obeying the relations (hereafter we adopt units in which $k_{\rm B}=1$) \begin{equation} \langle\xi(t)\rangle=0, \qquad\langle\xi(t)\xi(t^\prime)\rangle=2\gamma T\delta(t-t^\prime). \end{equation} The second relation (Einstein's relation) insures a canonical distribution of $x$ and $p$ at temperature $T$ if the parameter $a$ is held fixed for an infinitely long time. Multiplication of each term in the second equation in (\ref{eq:Langevin2}) by the displacement $dx$ yields the equation\cite{gardiner} \begin{equation} \frac{d}{dt}\left(\frac{p^2}{2m}\right)dt =\left(-\gamma\frac{p}{m}+\xi(t)\right)dx -\frac{\partial U(x,a)}{\partial x}dx, \label{eq:balance2} \end{equation} where we have used the first equation of Eq.(\ref{eq:Langevin2}) and also the identity $\frac{dp}{dt}dx=\frac{dp}{dt}\frac{p}{m}dt$. We note that $-\left(-\gamma \frac{dx}{dt}+ \xi(t)\right)$ is the {\it reaction} force exerted by the system against the heat bath, since the frictional force $-\gamma\frac{dx}{dt}$ and the random force $\xi(t)$ are both due to % the heat bath. We identify the work done by the reaction force as the {\it heat} transferred from the system to the heat bath, which we denote by $(-d{ Q})$\cite{ks1}: \begin{equation} -d{ Q} \equiv -\left(-\gamma\frac{dx}{dt}+\xi(t)\right)dx. \end{equation} (The minus sign in front of ${ d Q}$ is included to conform to the convention of thermodynamics textbooks.) The key point of introducing the concept of heat is that, although the heat bath is idealized and not affected by the system's dynamics, the heat bath can still be subject to a reaction force exerted by the system. Adding the total differential $dU$ to both sides of Eq.(\ref{eq:balance2}), we obtain the general expression for the {\it energy balance} as \begin{equation} d\left(\frac{p^2}{2m}+U\right) =\frac{\partial U}{\partial a}da+d{ Q}. \label{eq:first} \end{equation} Now, because the l.h.s. is the total increase of the energy, and $d{ Q}$ is the energy input to the system as a heat, the first term on the r.h.s. of Eq.(\ref{eq:first}) must be identified as the {\it work} done by the external system, $dW$, on the system through the change of the variable $a$, \begin{equation} dW \equiv \frac{\partial U}{\partial a}da. \label{eq:work} \end{equation} We conclude that the law of energy balance expressed as \begin{equation} dE=dW+d{ Q}, \qquad E\equiv \frac{p^2}{2m}+U \label{eq:balance} \end{equation} is satisfied for any {\it single} realization of the stochastic process described by Eq.(\ref{eq:Langevin2}). For a quasi-static process, in which $|\frac{da}{dt}|$ is arbitrarily small, the work is reversible and is equal to the change in the Helmholtz free energy $F(T,a)$ with probability 1. That is, in an ensemble of infinitely many realizations of such a process, the probability distribution of the work becomes a point distribution concentrated at the value of $F(T,a)$. \begin{equation} dW =dF(T,a), \qquad (\mbox{for a quasi-static process with $T$ fixed}), \label{eq:second0} \end{equation} with \begin{equation} F(T,a)\equiv -T \log \left[ \int\!\!\!\int e^{-\frac{E}{T}}{\rm d}x{\rm d}p\right] . \label{eq:helmholtz} \end{equation} The derivation of Eq.(\ref{eq:second0}) is as follows. We first note that, for $a$ to change by any small but finite amount $da$, it takes a time $|da|/|\frac{da}{dt}|$, which is indefinitely large in the quasi-static limit. During this time interval the state point $(x,p)$ comes arbitrarily close to almost all possible values, and its empirical distribution becomes asymptotically equal to the canonical distribution, $P_{\rm eq}(x;T,a)=\exp\left[\frac{F(T,a)-E}{T}\right].$ (The exception here is the case in which the interval $[a,a+da]$ includes a point at which the equilibration time diverges. See {\sf 2} of \S~\ref{subsec:problems} and \S~\ref{sec:clutch} below.) We can then evaluate $\frac{\partial U}{\partial a}da$ using its average with respect to $P_{\rm eq}$ in the quasi-static limit. Using the identity, \begin{equation} \int dx \frac{\partial U(x,a)}{\partial a}P_{\rm eq}(x;T,a)= \frac{\partial F(T,a)}{\partial a} \quad\mbox{($T$ fixed)}, \end{equation} we reach the result Eq.(\ref{eq:second0}). In fact Eq.(\ref{eq:second0}) is a stronger statement than the usual second law of thermodynamics for extensive systems, Note that for a thermodynamic system constituted by an ensemble of a large number of independent stochastic systems obeying Eq.(\ref{eq:Langevin2}), the first law of thermodynamics is obtained from Eq.(\ref{eq:balance}), \begin{equation} \left\langle dE\right\rangle =\langle dW\rangle +\langle d{ Q}\rangle, \label{eq:firstav} \end{equation} and the second law of thermodynamics for quasi-static processes is obtained from Eq.(\ref{eq:second0})\cite{ks-ss,chris}, \begin{equation} \langle dW\rangle =dF(T,a), \qquad (\mbox{for a quasi-static process with $T$ fixed}). \label{eq:second} \end{equation} These relations are concerned with only the ensemble averages denoted by $\langle \cdot \rangle$. It has also been shown\cite{ks-ss} that, for a finite rate of change of $a(t)$, the Clausius inequality \begin{equation} \langle dW\rangle \ge dF(T,a) \label{eq:second2} \end{equation} holds, and an explicit formula for the {\it irreversible} work, $\langle dW\rangle -dF(T,a)$, has been obtained up to the second order in $da/dt$. \section{Model} \label{sec:model} Figure \ref{fig:notion} schematizes the idea of our model. We employ a single harmonic oscillator with mass $m$ and spring constant $k\, (>0)$ as the system under study, which we call simply the ``system''. We denote by $x$ and $p$ the position and momentum of the system. correspond to compressing [decompressing] the ideal gas. Below, we consider $k$ to be a quantity that can be controlled, as the volume of a gas system is controlled in macroscopic Carnot cycles. In order to allow independent and variable interaction with each heat bath, we represent each such interaction in the form of a mechanical force, which subsumes the corresponding frictional and Gaussian random forces. Such mechanical forces should be related in some way to the degrees of freedom that directly interact with the heat baths, which we denote by $y_{\rm H}$ and $y_{\rm L}$. For simplicity, we do this by writing the mechanical forces as interaction forces, $-\frac{\partial \phi_{\rm H}}{\partial x}$ and $-\frac{\partial \phi_{\rm L}}{\partial x}$. As interaction potentials, we choose functions $\phi_{\rm H}(x-y_{\rm H},\chi_{\rm H})$ and $\phi_{\rm L}(x-y_{\rm L},\chi_{\rm L})$, where $\chi_{\rm H}$ and $\chi_{\rm L}$ are the control parameters. We call $\phi_{\rm H}$ and $\phi_{\rm L}$ the `couplers', because their values directly indicate the strength of the coupling between the system and the respective heat baths. We use the expressions like `control the coupler(s)' in reference to changes made in the values of these control parameters. We assume that the functions $\phi_\alpha$ ($\alpha$=H, L) are $2\pi$-periodic functions of $x-y_\alpha$. (See, for details, \S~\ref{sec:clutch} and Fig.~\ref{fig:clutch}.) \begin{figure}[b] \postscriptbox{7.80cm}{3.66cm}{notion.eps} \caption{Schematic view of a Carnot heat engine. The spring and the shaded linear `gear' represent the harmonic oscillator as the ``system.'' The left end of the spring (the black box) is fixed. Heat baths of temperatures $T_{\rm H}$ and $T_{\rm L}$ (the square shaded boxes) exert forces on the vanes (the star-shaped symbols inside the heat baths) whose angles of rotation are denoted by $y_{\rm H}$ and $y_{\rm L}$, respectively. These vanes are tightly connected to the circular gears. These circular gears can interact with the system in a manner that depends on the control parameters $\chi_{\rm H}$ and $\chi_{\rm L}$ of the couplers.} \label{fig:notion} \end{figure} \noindent The degrees of freedom $y_{\rm H}$ and $y_{\rm L}$ are subject to the frictional forces $-\gamma_{\rm H} \frac{dy_{\rm H}}{dt}$ and $-\gamma_{\rm L} \frac{dy_{\rm L}}{dt}$ and the random forces $\xi_{\rm H}(t)$ and $\xi_{\rm L}(t)$ exerted by the heat baths at temperatures $T_{\rm H}$ and $T_{\rm L}$, respectively, as well as the interaction forces from the system, $-\partial \phi_{\rm H}/\partial y_{\rm H}$ and $-\partial \phi_{\rm L}/\partial y_{\rm L}$. Here, $\gamma_{\rm H}$ and $\gamma_{\rm L}$ are the friction constants, and $\xi_{\rm H}(t)$ and $\xi_{\rm L}(t)$ are the white Gaussian random forces satisfying $\langle \xi_{\rm{H}}(t)\xi_{\rm{H}}(t^\prime)\rangle = 2 \gamma_{\rm{H}} T_{\rm{H}} \delta(t-t^\prime)$, $\langle \xi_{\rm{L}}(t)\xi_ {\rm{L}}(t^\prime)\rangle = 2 \gamma_{\rm{L}} T_{\rm{L}} \delta(t-t^\prime)$, a nd $\langle \xi_{\rm{H}}(t)\xi_{\rm{L}}(t^\prime)\rangle = 0$. The equations of motion for $x$, $p$, $y_{\rm H}$, and $y_{\rm L}$ are given as follows: \begin{mathletters} \label{eq_of_mot}% \begin{equation} \frac{dx}{dt}=\frac{p}{m}, \label{equationx} \end{equation} \begin{equation} \frac{dp}{dt}= -kx-\frac{\partial \phi_{\rm H}}{\partial x} -\frac{\partial \phi_{\rm L}}{\partial x}, \label{equationp} \end{equation} \begin{equation} \gamma_{\rm H}\frac{dy_{\rm H}}{dt} =-\frac{\partial \phi_{\rm H}}{\partial y_{\rm H}}+\xi_{\rm H}(t), \label{equationyH} \end{equation} \begin{equation} \gamma_{\rm L}\frac{dy_{\rm L}}{dt} =-\frac{\partial \phi_{\rm L}}{\partial y_{\rm L}}+\xi_{\rm L}(t). \label{equationyL} \end{equation} \end{mathletters} % We consider the gears of the heat bath and the system to be `tightly connected' (i.e. completely engaged) for $\chi_\alpha =1$, that is, the interaction $\phi_\alpha(x-y_{\alpha},1)$ is so strong that the difference $x-y_{\alpha}$ is fixed except for a small thermal fluctuation around its mean value, while these gears are `disconnected' (i.e. completely disengaged) for $\chi_{\alpha}=0$, that is, $\phi_\alpha(x-y_{\alpha},0)\equiv 0.$ We have neglected the inertia effect related to $y_{\rm H}$ and $y_{\rm L}$, as they would play only secondary role for our analysis. The protocol by which we control the parameters is represented in the space of $(k, \chi_{\rm H}, \chi_{\rm L})$ as shown in Fig. \ref{fig:protocol}. \begin{figure}[ht] \hspace*{2cm} \postscriptbox{5.23cm}{6.66cm}{scheme.eps} \caption{The cycle undergone by the control parameters. The various legs of this cycle correspond to the following processes of the system: isothermal processes (${\rm B}_{\rm H}\to {\rm C}_{\rm H}$ and ${\rm D}_{\rm L}\to {\rm A}_{\rm L}$), adiabatic processes (${\rm A}_{\rm 0}\to {\rm B}_{\rm 0}$ and ${\rm C}_{\rm 0}\to {\rm D}_{\rm 0}$), and the remaining processes where only one of the two control parameters of the couplers is changed (i.e. $\chi_{\rm H}\neq 0$ exclusively or $\chi_{\rm L}\neq 0$) while $k$ is kept constant. } \label{fig:protocol} \end{figure} \noindent In the figure, the paths along the axis $(\chi_{\rm H},\chi_{\rm L})=(0,0)$ correspond to the adiabatic processes, while the vertical paths with $(\chi_{\rm H},\chi_{\rm L})=(1,0)$ and $(\chi_{\rm H},\chi_{\rm L})=(0,1)$ correspond to the isothermal processes. The values of $k$ corresponding to these four horizontal paths are the parameters. \section{Reversible and irreversible work of operating the couplers} \label{sec:clutch} As we discussed in \ref{sec:intro}, the operation of the couplers can never be made quasi-static, because the time-scale of this operation inevitably becomes shorter than the equilibration time for the system when the coupling between the system and a heat bath becomes either absent or extremely tight. In addition to the work due to these irreversible processes, there is also reversible work associated with operation of coupler. Figure~\ref{fig:clutch} illustrates the generic features of the potential $\phi_\alpha(z,\chi_\alpha)$ for three different values of $\chi_\alpha$. Here $\Phi_0$, $\Phi_1$, and $\Phi_\infty$ represent the height of each potential profile. We assume that the height of $\phi_\alpha(z,\chi_\alpha)$ is a monotonically increasing function of $\chi_\alpha$ and satisfies $\max_z \phi(z,0)=0$, $\max_z \phi(z,\chi_{\alpha 0})=\Phi_0 \,(\ll T_\alpha)$, $\max_z \phi(z,\chi_{\alpha 1})=\Phi_1 \,(\gg T_\alpha)$, and $\max_z \phi(z,1)=\Phi_\infty \,(>\Phi_1)$ with $0 < \chi_{\alpha 0}< \chi_{\alpha 1}< 1$. \begin{figure}[ht] \hspace*{0.5cm} \postscriptbox{6.73cm}{4.66cm}{clutch.eps} \caption{The profiles of the interaction potential $\phi_\alpha$ are given as functions of $z\equiv x-y_\alpha$ for the three typical values of the maximum of $\phi_\alpha$, $\Phi_0$, $\Phi_1$ and $\Phi_\infty$, where $\Phi_0 \ll T_\alpha \ll \Phi_1 < \Phi_\infty$ (see the text). } \label{fig:clutch} \end{figure} There are two situations in which the time-scale of measurement and/or operation cannot exceed the equilibration time of the system. One is when the height of $\phi_\alpha$ is very small, and the other is when the height of $\phi_\alpha$ is very large. Let us assume that the regime $0<\chi_\alpha <\chi_{\alpha 0}$ corresponds to the former case; that is, for $\max_z \phi_\alpha (z,\chi_\alpha) \le \Phi_0$, the interaction $\phi_\alpha$ is so weak that the equilibration time of the system with the heat bath ($T=T_\alpha$) is beyond the timescale of measurement and/or operation. We call this the {\it loose regime}. Then, we assume that the regime $\chi_{\alpha 1}<\chi_\alpha <1$ corresponds to the latter case; that is, for $\max_z \phi_\alpha (z,\chi_\alpha) \ge \Phi_1$, the interaction $\phi_\alpha$ is so strong that the equilibration time characterized by the over-barrier transition of $z$ (see Fig~\ref{fig:clutch}) is again beyond the timescale of measurement and/or operation. In particular, we assume that for $\chi_\alpha =1$ there occur essentially no thermal activation events over the barrier $\Phi_\infty$. We call this the {\it tight regime}. In the remaining regime, $\chi_{\alpha 0}\leq \chi_\alpha \leq \chi_{\alpha 1}$ we assume that the operation can be carried out in a manner that arbitrarily closely approximates the quasi-static limit. Note that what Kramers calls the `small viscosity' and `large viscosity' cases in Ref.\cite{kramers} correspond, respectively, to the limits of the loose and tight regimes. We now evaluate the work % \begin{equation} W_\alpha(\chi_{\alpha 1}\!\to \!\chi_{\alpha 2})\equiv \!\!\! \int_{\chi_\alpha=\chi_{\alpha 1}}^{\chi_{\alpha 2}} \!\!\!\!\!\! \frac{\partial \phi_{\alpha}(x(t)\!- \! y_\alpha(t),\chi_\alpha(t)) }{\partial \chi_{\alpha}} d\chi_{\alpha}(t), \label{eq:clutchwork} \end{equation} for the loose and tight regimes using Eq.(\ref{eq:work}). We will describe the case of $\alpha=$H for concreteness. (The case of $\alpha=$L can be treated similarly.) In the loose regime (the processes near B$_0$ and C$_0$ in Fig.~\ref{fig:protocol}), $W_{\rm H}(0\to \chi_{{\rm H}0})$ is of order $\Phi_0$ and is small $(\ll T_{\rm H})$, although this work may be mostly irreversible. Since $\Phi_0$ is associated with the lower limit of quasi-static operation, the timescale of the operation is required to be large enough to satisfy the condition $\Phi_0 \ll T_{\rm H}$. In the tight regime (the processes near $B_{\rm H}$ and $C_{\rm H}$ in Fig.~\ref{fig:protocol}), the situation is more subtle. The contribution to $W_{\rm H}(\chi_{{\rm H}1}\to 1)$ consists of {\it (i)} the contribution produced when $z$ moves around the valley regions of $\phi_{\rm H}(z,\chi_{\rm H})$ with $\phi_{\rm H}\lesssim \Phi_1$ and {\it (ii)} the contribution produced when $z$ visits, by {\it rare} thermal excitation, the barrier regions of $\phi_{\rm H}$. To simplify the analysis we exclude the former contribution by assuming that $\partial \phi_{\rm H}(z,\chi_{\rm H})/\partial \chi_{\rm H}=0$ for $z$ in the valley regions of $\phi_{\rm H}$ (Fig.~\ref{fig:clutch}) in the tight regime. (This assumption is only technical; one can reach the conclusion of this paragraph without it.) The evaluation of the contribution {\it (ii)} above is carried out as follows. The probability to find $z$ in the barrier region is $\sim e^{-{\Phi_1}/{T}}$, and for such values of $z$, the change of $\chi_{\rm H}$ from $\chi_{{\rm H}1}$ to 1 results in an amount of work $\sim (\Phi_\infty -\Phi_1)$. Thus we have $W_{\rm H}(\chi_{{\rm H}1}\to 1)$ $\sim e^{-{\Phi_1}/{T}}(\Phi_\infty -\Phi_1) \sim e^{-{\Phi_1}/{T}}\Phi_\infty $. Because the timescale of operation is sufficiently large to allow large values of $\Phi_1$. For this reason, the conditions $e^{-{\Phi_1}/{T}}\Phi_\infty \ll T$ and $\Phi_\infty \gg T$ can be satisfied simultaneously. In conclusion, the irreversible part of the work associated with both the loose regime and the tight regime can be made as small as we wish by making the timescale of the operation sufficiently long in these regimes. The same conclusion holds for the case of $\alpha={\rm L}$. The quasi-static work associated with the change of $\chi_{\alpha}$ within the region $\chi_{\alpha 0}\le \chi_{\alpha}\le \chi_{\alpha 1}$ can be evaluated using Eq.(\ref{eq:second0}). Below we show that such quasi-static work cancels exactly when summed over the consecutive operations of connection and disconnection with a heat bath. Again, considering the case of $\alpha={\rm H}$, we denote by $F(T_{\rm H},k,\chi_{\rm H},0(=\chi_{\rm L}))$ the Helmholtz free energy of the composite system of the harmonic oscillator and the couplers, $\{p, x, y_{\rm H},y_{\rm L}\}$: \begin{eqnarray} & &e^{- F(T_{\rm H},k,\chi_{\rm H},0)/{T_{\rm H}}} \equiv 2\pi \int_{-\infty}^\infty dp \int_{-\infty}^\infty dx \int_{0}^{2\pi} dy_{\rm H}\cr & & \exp\left\{ -\frac{1}{T_{\rm H}} \left [ \frac{p^2}{2m}+\frac{kx^2}{2}+\phi_{\rm H}(x-y_{\rm H},\chi_{\rm H}) \right ]\right\}. \label{helmholtz} \end{eqnarray} Note that here $\chi_{\rm L}=0$ and the factor $2\pi$ in front of the integration on the r.h.s. comes from the phase integration over $y_{\rm L}$. Performing the integration over $y_{\rm H}$ first, we have \begin{equation} F(T_{\rm H},k,\chi_{\rm H},0)= -\frac{T_{\rm H}}{2}\log \frac{ {(2\pi)}^4 {(T_{\rm H})}^2 m}{k} +\tilde{F}(T_{\rm H},\chi_{\rm H},0), \label{eq:clutchF} \end{equation} with $\tilde{F}$ defined by \begin{equation} e^{-\tilde{F}(T_{\rm H},\chi_{\rm H},0)/T_{\rm H}}= \int_{0}^{2\pi} dz e^{-{\phi_{\rm H}(z,\chi_{\rm H})}/{T_{\rm H}}}. \label{eq:etildeF} \end{equation} The first term on the r.h.s of Eq.(\ref{eq:clutchF}) is independent of $\chi_{\rm H}$, while $\tilde{F}$ is independent of $k$. Using the notation of Eq.(\ref{eq:clutchwork}), we find from Eq.(\ref{eq:clutchF}) that $W_{\rm H}(\chi_{{\rm H}0}\to \chi_{{\rm H}1})=$ $\tilde{F}(T_{\rm H},\chi_{{\rm H}1},0)- \tilde{F}(T_{\rm H},\chi_{{\rm H}0},0)$ along B$_0 \to$B$_{\rm H}$ in Fig.~\ref{fig:protocol}, and $W_{\rm H}(\chi_{{\rm H}1}\to\chi_{{\rm H}0})=$ $\tilde{F}(T_{\rm H},\chi_{{\rm H}0},0)- \tilde{F}(T_{\rm H},\chi_{{\rm H}1},0)$ along C$_{\rm H} \to$C$_0$. These two cancel exactly: \begin{equation} W_{\rm H}(\chi_{{\rm H}0}\to \chi_{{\rm H}1})+ W_{\rm H}(\chi_{{\rm H}1}\to \chi_{{\rm H}0}) =0. \label{eq:workadBC} \end{equation} The actual time interval required to change $\chi_\alpha$ between $\chi_{\alpha 0}$ and $\chi_{\alpha 1}$ is finite, say, $t_{\rm 01}$. The irreversible work due to this finiteness has been shown to be ${\cal O}({t_{\rm 01}}^{-1})$ quite generally\cite{ks-ss}. Thus the irreversible work associated with the process in which $\chi_\alpha$ is changed between $\chi_{\alpha 0}$ and $\chi_{\alpha 1}$ can be made as small as we wish by making the timescale of operation sufficiently long. For later use, we now also estimate the heat exchanged upon the operation of the couplers. As we have shown above, the amount of work in the loose and tight regimes can be made arbitrarily small. Also, changes in the parameters $\chi_\alpha$ lead to only small changes of the internal energy of the composite system. These facts together with the energy balance principle (see Eq.(\ref{eq:balance})) lead to the conclusion that the heat exchanged in these two regimes can be made as small as we wish. Next, the heat exchanged during the quasi-equilibrium processes with $\chi_{\alpha 0}<\chi<\chi_{\alpha 1}$ is assessed as follows. From Eq.(\ref{eq:clutchF}) the ensemble average of the internal energy of the composite system is given by $T_{\rm H}+(1-T_{\rm H}\frac{\partial}{\partial T_{\rm H}}) \tilde{F}(T_{\rm H},\chi_{\rm H},0)$. If we define by $\langle Q({\rm B}_0\to {\rm B}_{\rm H})\rangle$ the average heat influx to the composite system during the quasi-equilibrium operation along B$_0\to$B$_{\rm H},$ the condition of energy balance, Eq.(\ref{eq:firstav}), yields $\langle Q({\rm B}_0\to {\rm B}_{\rm H})\rangle= -T_{\rm H}\frac{\partial}{\partial T_{\rm H}} [\tilde{F}(T_{\rm H},\chi_{{\rm H}1},0) -\tilde{F}(T_{\rm H},\chi_{{\rm H}0},0)].$ This heat cancels exactly the average heat input $\langle Q({\rm C}_{\rm H}\to {\rm C}_0)\rangle$ similarly defined along C$_{\rm H}\to$C$_0$: \begin{equation} \langle Q({\rm B}_0\to {\rm B}_{\rm H})\rangle +\langle Q({\rm C}_{\rm H}\to {\rm C}_0)\rangle =0. \label{eq:heatadBC} \end{equation} In the same manner, we can show the cancellation of both the work and the heat during the quasi-static part of the operation of coupler along D$_0 \to$D$_{\rm L}$ and A$_{\rm L} \to$A$_0$. \section{Matching the `temperature' of the system and a heat bath} \label{sec:match} Here we study the meaning of the idea of matching the `temperature' of the small system with that of a heat bath. For comparison, we note that this meaning is unambiguous for an isolated macroscopic thermodynamic system, for which the energy and the temperature are simultaneously well-defined quantities, and in order to realize a reversible Carnot cycle, the temperature of the system should be the same as that of the heat bath with which it makes contact. Contrastingly, for isolated small systems, the energy takes a definite value, while the temperature is not generally well-defined. We will show in this section that if the small system is a harmonic oscillator, the concept of the temperature is still useful, and reversibility can be obtained as in macroscopic systems. Discussion regarding the general case is given in \S~\ref{subsec:1}. Suppose that a coupler is operated quasi-statically up to the edge of a loose regime ($\chi_{\rm L}=\chi_{{\rm L}0}$ along ${\rm A}_{\rm L}\to {\rm A}_0$ or $\chi_{\rm H}=\chi_{{\rm H}0}$ along ${\rm C}_{\rm H}\to {\rm C}_0$ in Fig.~\ref{fig:protocol}). The energy of the oscillator in this situation fluctuates as a function of time and though the temporal fluctuation of the energy is very slow, it still obeys the canonical distribution at the temperature of the heat bath ($T_{\rm L}$ for A$_0$ and $T_{\rm H}$ for C$_0$) up to a small error of ${\cal O}(\Phi_0)\, (\ll T_\alpha)$. By the definition of the loose regime, further weakening the connection results in a situation in which there is no appreciable exchange of energy between the system and the heat bath (see \S~\ref{sec:intro} and \ref{sec:clutch}). Then, the complete disconnection leaves an isolated system whose energy is distributed according to the canonical distribution corresponding to the temperature of the heat bath. This is in fact true even if the small system is not a harmonic oscillator. For a harmonic oscillator, however, both the energy distribution in the canonical ensemble and the transformation of the system's energy through the quasi-static adiabatic process have special features. The energy distribution in the canonical ensemble at temperature $T$, $P_{\rm can}(E;T)$ is independent of the parameter $k$: \begin{equation} P_{\rm can}(E;T)=\frac{1}{T}e^{-\frac{E}{T}}. \label{eq:canonical} \end{equation} (See Appendix~\ref{sec:derivation} for a derivation.) Then, at A$_0$ and C$_0$, the energy of the oscillator is distributed according to $P_{\rm can}(E;T_{\rm L})$ and $P_{\rm can}(E;T_{\rm H})$, respectively. If the oscillator has a specific (initial) energy $E$ and undergoes the quasi-static adiabatic process represented as A$_0\to$B$_{\rm 0}$ or C$_0\to$D$_{\rm 0}$ in Fig.~\ref{fig:protocol}, its energy, $E(k)$, changes so that the value $J(E,k)$ given by (\ref{eq:harmonicJ}) remains constant \cite{landau:cm}: \begin{equation} J(E(k),k)= \frac{E(k)}{2\pi}\sqrt{\frac{m}{k}}=\mbox{constant}. \label{invariant} \end{equation} The relation (\ref{invariant}) determines the change of the energy distribution when the system undergoes a quasi-static adiabatic process through the change of $k$: For a change from $k$ to $k'$, the altered distribution $P'(E')$ must obey $P_{\rm can}(E,T)dE=P'(E')dE'$ with $E'/({2\pi\sqrt{\frac{k'}{m}}})=E/({2\pi\sqrt{\frac{k}{m}}})$. Thus we obtain the energy distribution of a canonical ensemble {\it{after}} the change in $k$, $P^\prime (E')=P_{\rm can}(E',T')$ with $T'$ being defined by $T/\sqrt{k}=T'/\sqrt{k'}$. However, we note that this simple situation is due to the special nature of the harmonic oscillator system. The general case is discussed in \S~\ref{subsec:1}. With these facts in mind, we can now characterize the condition for a quasi-equilibrium transition between adiabatic processes and subsequent isothermal processes: \begin{description} \item[(1)] The energy distribution of the system before the connection with the heat bath must be that of a canonical ensemble. \item[(2)] The temperature characterizing this canonical ensemble must be the same as that of the heat bath in question. \end{description} The reason is that, under these conditions, the system upon connection behaves statistically {\it as if} it had been in contact with the heat bath for a long time. Note that the connection should be begun through a sufficiently weak interaction $\sim \Phi_0$ (see \S~\ref{sec:clutch}). In our case with the protocol described by Fig.~\ref{fig:protocol}, we require that $T'=T_{\rm H}$ at B$_0$ and $T'=T_{\rm L}$ at D$_0$ hold, respectively. Denoting the value of $k$ at that between $A_0$ and $A_{\rm{L}}$ as $k_A$, that between $B_0$ and $B_{\rm{H}}$ as $k_B$, etc., the condition for a quasi-equilibrium connection to the heat bath is given explicitly as follows: \begin{equation} \frac{T_{\rm L}}{\sqrt{k_{\rm A}}}=\frac{T_{\rm H}}{\sqrt{k_{\rm B}}}, \quad \frac{T_{\rm H}}{\sqrt{k_{\rm C}}}=\frac{T_{\rm L}}{\sqrt{k_{\rm D}}}. \label{matching1} \end{equation} \section{Efficiency} \label{sec:efficiencygen} \subsection{How is the Carnot limit approached?} \label{sec:efficiencymax} We now evaluate the maximal overall efficiency. We must take into account (i) the operation of the couplers, (ii) the isothermal processes, and (iii) the adiabatic processes. (i) We assume that by making the time in the loose and tight regimes sufficiently long, the intrinsically irreversible work and heat flow can be made as small as we wish. The remaining part of the operation of couplers is assumed to be made under quasi-equilibrium conditions. The accompanying work and heat flow cancel exactly when summed over an infinite number of consecutive connection and disconnection with the heat bath (see Eqs.(\ref{eq:workadBC}) and (\ref{eq:heatadBC})). (ii) For the isothermal parts of the cycle, B$_{\rm H}\to$C$_{\rm H}$ and D$_{\rm L}\to$A$_{\rm L}$, we assume a quasi-static change of $k$. The accompanying work is then given using the general formula Eq.(\ref{eq:second0}). For the part B$_{\rm H}\to$C$_{\rm H}$, $F(T,a)$ in Eq.(\ref{eq:second0}) is replaced by $F(T_{\rm H},k,\chi_{\rm H},0)$ of Eq.(\ref{eq:clutchF}), and the work done by the system is $ \frac{T_{\rm H}}{2} \log\left(\frac{k_{\rm B}}{k_{\rm C}}\right),$ which we denote by $-W({\rm B}_{\rm H}\to {\rm C}_{\rm H})$. Similarly, for the part D$_{\rm L}\to$A$_{\rm L}$ the work is $-W({\rm D}_{\rm L}\to {\rm A}_{\rm L}) = \frac{T_{\rm L}}{2} \log\left(\frac{k_{\rm D}}{k_{\rm A}}\right).$ Then, the relation (\ref{matching1}), we have \begin{equation} -W({\rm B}_{\rm H}\to {\rm C}_{\rm H})- W({\rm D}_{\rm L}\to {\rm A}_{\rm L}))=\frac{T_{\rm H}-T_{\rm L}}{2} \log\left(\frac{k_{\rm B}}{k_{\rm C}}\right). \label{sumBCDA} \end{equation} (iii) For the adiabatic part of the cycle, A$_{\rm 0}\to$B$_{\rm 0}$ and C$_{\rm 0}\to$D$_{\rm 0}$, we also assume a quasi-static change of $k$. The energy of the system then obeys the law (\ref{invariant}) and the amounts of work done by the system in the adiabatic processes $-W({\rm A}_{\rm 0}\to {\rm B}_{\rm 0})$ and $-W({\rm C}_{\rm 0}\to {\rm D}_{\rm 0})$ are given by $-W({\rm A}_{\rm 0}\to {\rm B}_{\rm 0})=% E(k_{\rm A})(1-\sqrt{k_{\rm B}/k_{\rm A}})$ and $-W({\rm C}_{\rm 0}\to {\rm D}_{\rm 0})=% E(k_{\rm C})(1-\sqrt{k_{\rm D}/k_{\rm C}}).$ As the energies $E_{\rm A}$ and $E_{\rm C}$ obey the distribution Eq.(\ref{eq:canonical}) with $T=T_{\rm L}$ and $T=T_{\rm H}$, respectively, their statistical averages are $\langle E_{\rm A}\rangle =T_{\rm L}$ and $\langle E_{\rm C}\rangle =T_{\rm H}$ [up to a small error of ${\cal O}(\Phi_0)$]. Using (\ref{matching1}), we then have \begin{eqnarray} && -\langle W({\rm A}_{\rm 0}\to {\rm B}_{\rm 0})\rangle - \langle W({\rm C}_{\rm 0}\to {\rm D}_{\rm 0})\rangle \cr &&= T_{\rm L}\left[1-\sqrt{\frac{k_{\rm B}}{k_{\rm A}}}\right]+ T_{\rm H}\left[1-\sqrt{\frac{k_{\rm D}}{k_{\rm C}}}\right]=0. \label{sumCD} \end{eqnarray} While we obtain this simple result in the present case, it is important to note that the cancellation of the contributions from the adiabatic processes on the average is not a generic feature of the Carnot cycle (consider, for example, non-ideal gases). The heat influx from the high temperature heat bath is evaluated as follows. While the work during the isothermal quasi-equilibrium processes is a non-fluctuating quantity (see Eq.(\ref{eq:second0})), both the energy influx from the heat bath and the system's energy fluctuate subject to the constraint of energy balance described by Eq.(\ref{eq:balance}). From Eqs.(\ref{eq:clutchF}) and (\ref{eq:etildeF}) we can show that the average internal energy of the composite system with degrees of freedom $\{p, x, y_{\rm H},y_{\rm L}\}$ is independent of the parameter $k$. Therefore, the statistical average of the heat influx during the isothermal process ${\rm B}_{\rm H}\to{\rm C}_{\rm H}$, which we denote by $\langle Q({\rm B}_{\rm H}\to{\rm C}_{\rm H})\rangle$, satisfies \begin{equation} 0= \langle Q({\rm B}_{\rm H}\to{\rm C}_{\rm H})\rangle +W({\rm B}_{\rm H}\to {\rm C}_{\rm H}). \label{eq:energy0} \end{equation} Thus we have \begin{equation} \langle Q({\rm B}_{\rm H}\to{\rm C}_{\rm H})\rangle =\frac{T_{\rm H}}{2} \log\left(\frac{k_{\rm B}}{k_{\rm C}}\right). \label{eq:input} \end{equation} As we have seen in \S~\ref{sec:clutch}, there is no net heat flow due to the quasi-static part of the operation of coupler (see Eq.(\ref{eq:heatadBC})), while the heat transfer associated with the loose and tight regimes can be made as small as we wish. Collecting the above results, the maximal overall efficiency $\eta_{\rm max}$ of the cycles is reduced to the following formula: \begin{equation} \eta_{\rm max}=-\frac{W({\rm B}_{\rm H}\to {\rm C}_{\rm H})- W({\rm D}_{\rm L}\to {\rm A}_{\rm L}) }{\langle Q({\rm B}_{\rm H}\to{\rm C}_{\rm H})\rangle}. \label{eq:eta} \end{equation} Then, using (\ref{sumBCDA}) and (\ref{eq:input}) we have \begin{equation} \eta_{\rm max}= 1-\frac{T_{\rm L}}{T_{\rm H}}. \label{eq:etamax} \end{equation} We would like to stress that the attainment of this efficiency (whose expression is familiar from textbook treatments) is not due to the quasi-static operation of the whole system. In the situation we consider, we have seen that some parts of the cycle can never be carried out quasi-statically, due to the intrinsically irreversible operation of the couplers (\S~\ref{sec:clutch}), as well as the intrinsic irreversibility resulting from the non-canonical energy distribution of the systems caused by the adiabatic processes(\S~\ref{sec:match}). \subsection{Statistics over a finite number of cycles} \label{sec:efficiency} The maximal efficiency Eq.(\ref{eq:etamax}) obtained above represents exclusively the ratio of the total work done by the system to the total energy influx from the high temperature heat bath through infinite number of cycles. Here we consider the efficiency for a single cycle, $\eta_{\rm loc}$, which can be written as \begin{equation} \eta_{\rm loc}=\frac{\frac{T_{\rm H}-T_{\rm L}}{2} \log\left(\frac{k_{\rm B}}{k_{\rm C}}\right)-\delta_W}{ \frac{T_{\rm H}}{2} \log\left(\frac{k_{\rm B}}{k_{\rm C}}\right)+\delta_Q}, \label{eq:etaloc} \end{equation} with \[-\delta_W= - W({\rm A}_{\rm 0}\to {\rm B}_{\rm 0}) - W({\rm C}_{\rm 0}\to {\rm D}_{\rm 0}),\] and \begin{eqnarray} \delta_Q &=& \left\{Q({\rm B}_0\to {\rm B}_{\rm H}) + Q({\rm C}_{\rm H}\to {\rm C}_0)\right\} \cr &&+\left\{ Q({\rm B}_{\rm H}\to{\rm C}_{\rm H})- \frac{T_{\rm H}}{2} \log\left(\frac{k_{\rm B}}{k_{\rm C}}\right) \right\}. \end{eqnarray} We first note several properties of $\eta_{\rm loc}.$ {\it 1.}~The deviations $\delta_W$ and $\delta_Q$ do not vanish, even in the quasi-equilibrium limit, since the system continues to exchange energy with a heat bath until the moment that it is disconnected from the heat bath. {\it 2.} If we choose the initial point of an individual cycle to be somewhere between D$_{\rm L}$ and A$_{\rm L}$, then the values of $\eta_{\rm loc}-\langle \eta_{\rm loc}\rangle$ for different cycles are statistically independent. In fact, the statistical deviations of $W({\rm A}_{\rm 0}\to {\rm B}_{\rm 0})$ and $W({\rm C}_{\rm 0}\to {\rm D}_{\rm 0})$ are mutually uncorrelated because of the intervening isothermal Markov processes ${\rm B}_{\rm H}\to{\rm C}_{\rm H}$ and ${\rm D}_{\rm L}\to{\rm A}_{\rm L}$, while $W({\rm A}_{\rm 0}\to {\rm B}_{\rm 0})$ and $Q({\rm B}_{\rm 0}\to{\rm B}_{\rm H})$ are statistically correlated through the shared point B$_0$, as are $Q({\rm C}_{\rm H}\to{\rm C}_{\rm 0})$ and $W({\rm C}_{\rm 0}\to {\rm D}_{\rm 0})$ through the point C$_0$. {\it 3.} One can define the `excess output', $-\stackrel{\circ}{W}$, by \begin{eqnarray} -\stackrel{\circ}{W}&\equiv& (\eta_{\rm loc}-\eta_{\rm max})\left[{\frac{T_{\rm H}}{2} \log\left(\frac{k_{\rm B}}{k_{\rm C}}\right)+\delta_Q}\right] \cr &=& -\delta_W-\eta_{\rm max}\,\delta_Q. \label{eq:bonus} \end{eqnarray} A positive value of $-\stackrel{\circ}{W}$ implies that we happened to get more work than that expected from the Carnot maximal efficiency (i.e., $-\delta_W>\eta_{\rm max}\,\delta_Q$). Such a situation can result through fluctuations, and it is not in contradiction with the second law of thermodynamics. We may then, however, ask how many cycles on average we must carry out before we first obtain a cumulative excess output, $-\stackrel{\circ}{w}_n\equiv \sum_{i=1}^n (-\stackrel{\circ}{W}_i)$, where $-\stackrel{\circ}{W}_i$ is the excess output of the $i$-th cycle and $n$ is the total number of consecutive cycles. The point of this question can be understood in terms of the following apparent paradox: Suppose one monitors $-\stackrel{\circ}{w}_n$ as a function of $n$ and stops when it becomes positive for the first time. If one could repeat such a procedure of monitor-and-stop indefinitely many times, one could construct a perpetual machine of the second kind. This, of course, would be in contradiction with the second law of thermodynamics. A pitfall of this false argument is that, although for a given sequence of cycles, the condition $-\stackrel{\circ}{w}_n>0$ will be satisfied at some finite $n$ with probability 1, the {\it average} over separate sequences of the smallest value of $n$ with positive $-\stackrel{\circ}{w}_n$ is divergent. This fact is closely related to the fact that the one-dimensional random walk is `null-recurrent' (see Appendix~\ref{sec:nullrecurrent}). \section{Discussion} \label{sec:conclusion} \subsection{Irreversibility resulting from contact with a heat bath} \label{subsec:1} In \S~\ref{sec:match} we have used the fact that for the harmonic oscillator system, as a result of the quasi-static adiabatic process, the energy changes in such a manner that the energy distribution remains in canonical form, with simply a change of the temperature, $T\to T'$. However, the harmonic oscillator represents a special system, and this is not generically the case with any Hamilton. More generally, the energy distribution $P_{\rm can}(E,T)$ is distorted into some non-canonical form, $P'(E)$, as a result of the quasi-static adiabatic process. When an ensemble of systems following the distribution $P'(E)$ are brought into (weak) contact with a heat bath of arbitrary temperature $T'$, the energy distribution {\it irreversibly} relaxes to the canonical form $P_{\rm can}(E,T')$. This is the case even if $T'$ is chosen so that $\int EP'(E)dE =\int E P_{\rm can}(E,T')dE$, i.e. even in the case that no net heat is transferred on the average from the bath to the system. The relevance of this irreversible relaxation to the energetics of small systems requires further scrutiny. This will be discussed in more detail in a separate paper\cite{SSHT}. \subsection{Irreversible adiabatic process} \label{subsec:adiab} The adiabatic process in the Carnot cycle is a mechanical process. The ergodic invariant theorem\cite{kasuga} tells us that under quasi-static and adiabatic change of a system parameter, say $a$, by a finite amount, the phase volume enclosed by the energy surface defined by the system's energy at each moment, $J(E,a)$ (see (\ref{eq:defJ}) for definition), remains constant. The theorem does not assume the thermodynamic limit nor the presence or absence of chaotic trajectories. Contrastingly, for a non-quasi-static process $J(E,a)$ can either increase or decrease, depending on both the nature of the change of $a$ and the initial conditions of the system. In the context of the present paper, however, it is most meaningful to confine ourselves to only `macroscopic' external operations, excluding `demonic' ones which depend on detailed information of the system. More precisely, we focus on an unprejudiced choice of the initial conditions among those with a given energy, and also focus only on statistical averages of the energy, rather than consider particular results obtained from particular initial conditions. Sato\cite{ksato} has recently studied a harmonic oscillator under a time-dependent force, % $ m\ddot{x}=-x +a(t)$, as the simplest non-trivial example of a system with only macroscopic external operations. Here, the change of $a(t)$ amounts to a horizontal displacement of the potential. He showed analytically, as one can easily confirm, that for an arbitrary function $a(t)$, the energy of the oscillator, $\frac{m}{2}\dot{x}^2+\frac{1}{2}(x-a)^2$, is strictly non-decreasing {\it if} the average is taken with respect to the initial condition over all initial conditions represented by states with a given energy, i.e., over a micro-canonical ensemble. Only in the limit of a quasi-static process $(|\frac{da}{dt}|\to 0)$ is the energy unchanged This example demonstrates the irreversibility of a mechanical system of non-macroscopic size {with}\, properly defined macroscopic operation. A recent numerical work\cite{komatsu2} investigating a harmonic oscillator with time-dependent spring constant, $ m\ddot{x}=-k(t)x$, reveals the same phenomena when $k(t)$ is constrained to return to its initial value. In the present context of the analysis of the efficiency of Carnot cycle (\S~\ref{sec:efficiencygen}), these findings are both natural and important, since, if the case were different, one could find functional form for $k(t)$ for which the adiabatic work on the system would be less than what we expect for a quasi-static process, and the whole cycle could be used to construct a perpetual machine of the second kind. (Note that in the analysis of \S~\ref{sec:efficiencygen} we have excluded {only marginally} the existence of such a perpetual machine if the loss due to the other sources is made arbitrarily small.) It is desirable to obtain a general proof (or counter-evidence) of the irreversibility resulting from non-quasi-static processes. A related analysis using path probabilities has been performed for chaotic systems\cite{sasa00}. However, we suspect that the essential mechanism of the irreversibility related to the characterization of macroscopic operators, which are ignorant of the initial microscopic state of the system, can be elucidated without resort to chaotic statistics. It is also a future problem to scrutinize the case in which the topology of the energy contour surface in the phase space (${\cal H}_a=E$) changes at some parameter value, say $a_c$. In such cases, quasi-static processes cannot be extended across $a_c$\cite{8figure}. \subsection{Control of processes by the system itself} \label{subsec:3} As in the case of the ordinary macroscopic Carnot cycle, we have introduced control parameters, $\{k,\chi_{\rm H},\chi_{\rm L}\}$. We assumed that the values of these parameters are changed by some external agent whose dynamics are external to the equation of motion of the system. In fact, however, there are many `self-controlled' energy transducers which contain their own control systems. In such cases, the identification of the control system is more or less a matter of interpretation. In the macroscopic world, DC electric motors and steam engines are examples, while motor proteins, such as myosins, kinesins and dyneins, etc., are microscopic examples. Theoretically, the so-called Feynman ratchet and pawl system\cite{feynman} has been proposed as a microscopic energy transducer working by itself between hot and cool heat baths. In this model, the role of the control system is played either by the pawl or by the ratchet, depending on which of these two is in direct contact with the cool heat bath. The stochastic energetics of this model have been analyzed\cite{ks1,magnasco98}. B\"utticker's model\cite{buttiker} is another self-controlled microscopic transducer. In this model a massive particle moves while in contact with a heat bath of position-dependent temperature, $T(x)=T_{\rm H}$ or $T_{\rm L}$. In this model, the inertia of the particle serves to switch the particle's environment from $T=T_{\rm H}$ to $T=T_{\rm L}$, or vice versa. The stochastic energetics of this model have also been analyzed\cite{mm-ss,hondou-ks}. Some people have claimed that the Carnot limiting efficiency $\eta_{\rm max}= 1-\frac{T_{\rm L}}{T_{\rm H}}$ can be attained in Feynman's ratchet and pawl system (see \cite{feynman} and \cite{sakaguchi}) and in B\"utticker's model (see \cite{astumian}). With the exception of the original work by Feynman\cite{feynman}, where no implementation details are given, these studies introduced into their analyses some `gate' mechanism. The study of the energetics of such systems, including the action of these gates, has not yet been made. \acknowledgements Discussions with K. Sato, S. Sasa, T. Komatsu and T. Chawan-ya on the adiabatic processes of harmonic oscillator are gratefully acknowledged. The helpful comments by G. Paquette on the manuscript are also acknowledged. This work is supported in part by the Inamori Foundation (T.H.), and a Grant in Aid by the Ministry of Education, Culture and Science (Priority Area, Nos.11156216 and 12030215) (K.S.).
\section{Introduction} Gravitational structure formation is intrinsically non-linear. As matter is pulled toward a density maximum by gravity to form a condensate and pulled away by gravity from a density minimum to form a void, both gravitational forces are increased by the resulting redistribution of mass so that the mass flow rates are accelerated. This is a classical positive feedback process. Linear theories are notoriously misleading when applied to non-linear processes. Fluid mechanics provides a familiar example, where steady laminar solutions for the momentum equations with non-linear terms neglected are vastly different from the random turbulent motions actually observed when the Reynolds number (the ratio of inertial to viscous forces) is large. Jeans \cite{jns02,jns29} applied a linear perturbation stability analysis to the problem of gravitational structure formation in a motionless gas with uniform properties. Neglecting non-linear terms of the momentum and density conservation equations reduces the problem to one of acoustics. Sound wave-crests provide density maxima that can nucleate gravitational condensation only if a wave-length $\lambda$ propagation time $\tau_{\lambda} = \lambda/V_S$ is greater than the ``free fall'' gravitational time scale $\tau_G = (\rho G)^{-1/2}$, where $V_S$ is the speed of sound in the gas, $\rho$ is the gas density and $G$ is Newton's gravitational constant $6.7\times10^{-11}$ $ \rm m^3 \> kg^{-1} \> s^{-2}$. Thus, acoustic density perturbations with sizes $L \ge L_J = V_S/(\rho G)^{1/2}$ can grow, but those smaller cannot. This Jeans criterion for gravitational instability has been universally adopted in astrophysics and cosmology, with modifications to take into account effects of general relativity and the rapid expansion rate of the early universe when the first structures formed \cite {pbl93,kol94}. However, the Jeans criterion is unreliable. Its widespread application is responsible for the present dark matter paradox. Jeans asserted that acoustic density maxima with $\lambda \ge L_J$ remain acoustic with velocity $V_S$ after condensation begins. This is one of many misconceptions in \cite {jns29}, including the claim that cores of galaxies consist of hot gas and not stars, and are sources of matter flowing from other Universes. Actually, acoustic density nuclei must rapidly stop moving with sound speed $V_S$ as they accumulate zero momentum from the motionless ambient fluid. Gibson \cite {gib96} shows that non-acoustic density nuclei are absolutely unstable to gravitational structure formation in a motionless fluid. Most density maxima and minima are produced by turbulent mixing of density. They move approximately with the fluid velocity and have a characteristic length scale $L_B = (D/\gamma)^{1/2}$ termed the Batchelor scale, where D is the molecular diffusivity of the density and $\gamma$ is the rate-of-strain of the turbulence \cite {gib68}. Gravitational condensation on non-acoustic density maxima and void formation from non-acoustic density minima are limited either by the dominant fluid force or by molecular diffusivity to length scales larger than the largest ``Schwarz'' length scale $L_{SX}$. Schwarz scales are derived in the following Section by comparing dispersive forces with the force of gravity, and the diffusion velocity with the gravitational velocity. The subscript ``X'' may be $T,V,M,D$ referring to either turbulent, viscous, magnetic or diffusive limitations to gravitational condensation. \section{Theory} A density nucleus with scale $L \ll L_J$ and mass excess $M'$ in a large body of motionless, homogeneous gas is gravitationally unstable and will form a condensate, just as the same sized density nucleus with mass deficit $-M'$ is gravitationally unstable and will form a void. Both cases contradict Jeans' theory. If the gas is not motionless, gravitational forces $F_G \approx \rho ^2 G L^4$ on scale L must be larger than turbulence forces $F_T \approx \varepsilon ^{2/3} L^{8/3}$, viscous forces $F_V \approx \rho \nu \gamma L^2$, or magnetic forces $F_M \approx (H^2 /\rho) L^2$, where $\varepsilon$ is the viscous dissipation rate, $\nu$ is the kinematic viscosity, $H$ is the magnetic field, and $H^2 /\rho$ is the magnetic pressure. Equality occurs at the turbulent Schwarz scale $L_{ST} = \varepsilon ^{1/2} / (\rho G)^{3/4}$, the viscous Schwarz scale $L_{SV} = (\nu \gamma / \rho G)^{1/2}$, and the magnetic Schwarz scale $L_{SM} = [(H^2 /\rho)/\rho G]^{1/2} = [(\chi_H /\varepsilon ^{1/3} )/\rho ^2 G]^{2/3}$, where $\chi_H$ is the diffusive dissipation rate of the magnetic field variance. From turbulent mixing theory, density fluctuations on scale $L$ diffuse with velocity $D/L$ \cite {gib68}. Equating this to gravitational velocities $L/\tau_G$ gives the Schwarz diffusive scale $L_{SD} = (D^2/\rho G)^{1/4}$. Condensation of non-baryonic fluids on galactic scales is impossible because $L_{SD} \gg L_{galaxy}$ \cite {gib96}. \section{Application of the theory} By the new theory \cite {gib96}, gravitational decelerations to form proto-superclusters and proto-galaxies begin in the plasma epoch at 13,000 years after the big bang, limited by viscous forces at scale $L_{SV}$. The initial conditions of the primordial hydrogen and helium gas are constrained by the COBE satellite observation that $\delta T/T \approx 10^{-5}$. Turbulence levels were thus extremely weak, with maximum rates-of-strain only slightly larger than that of the expanding universe $\gamma = 1/t = 10^{-13} s^{-1}$. The viscosity $\nu$ and diffusivity $D$ of the mixture were about $10^{13}$ $\rm m^2 s^{-1}$, giving the condensation mass $M_{SV} = {L_{SX} {}_{max}}^3 \rho = {L_{SV}}^3 \rho = 10^{23-24} \rm kg$. The entire universe of primordial gas condensed to form small planetary mass objects (Mercury to Mars) termed ``primordial fog particles'', or PFPs. \section{Observations} Schild \cite {sch96} concludes that his lensing galaxy consists mostly of small planetary objects ``likely to be the missing mass'', based on continuous month-period fluctuations between quasar image Q0957+651A,B light curves with a 1.1 year time delay. Three observatories confirm the same time delay and the same fluctuating (microlensed) signals \cite {gs98}. MACHO and EROS collaborations exclude Galactic planetary masses as the halo dark matter assuming homogeneous spatial distributions of the objects \cite {alc97}. However, large clumping corrections to star microlensing mass exclusion diagrams are required for $10^{-7} M_{\odot}$ objects because such small objects develop extreme spatial intermittency during their 15 billion years of nonlinear gravitational accretion. Such corrections could resolve the apparent observational inconsistency between quasar-microlensing \cite {sch96} and star-microlensing \cite {alc97} studies of small planetary mass objects as galaxy dark matter. \section{Conclusions} The Jeans criterion for gravitational instability is unreliable, and is replaced by a new criterion that density fluctuations in a gas will experience gravitational condensation and void formation on length scales larger than the maximum Schwarz scale; that is, $L \ge L_{SX} {}_{max} $. The baryonic ``missing mass'' or ``dark matter'' consists of those PFPs that have not aggregated to form stars. Because non-baryonic dark matter has $L_{SD}$ scales larger than galactic scales it is a negligible part of the galactic dark matter, but dominates supercluster dark matter as super-halos \cite {gib96}.
\section{Introduction} SU(2) Yang-Mills-Higgs (YMH) theory, with the Higgs field in the adjoint representation, possesses globally regular particle-like solutions, such as the 't Hooft-Polyakov magnetic monopole \cite{thooft} and the Julia-Zee dyon, which carries both magnetic and electric charge \cite{jul}. In SU(2) Einstein-Yang-Mills-Higgs (EYMH) theory, for small gravitational constant the gravitating fundamental monopole \cite{old,ewein,bfm} and the gravitating fundamental dyon \cite{bhk} emerge smoothly from the corresponding flat space solutions. The mass of the gravitating fundamental monopole \cite{ewein,bfm} and dyon \cite{bhk} solutions decreases with increasing gravitational constant, corresponding to increasing coupling constant $\alpha$, which is proportional to the ratio of vector meson mass and Planck mass. The regular solutions cease to exist beyond some maximal value of $\alpha$, depending on the electric charge $Q$ \cite{ewein,bfm,bhk,ewein3}. Besides these globally regular solutions, in SU(2) EYMH theory magnetic \cite{ewein,bfm,bfm2,aichel} and dyonic \cite{bhk} non-abelian black hole solutions exist. They emerge from the globally regular solutions, when a finite regular event horizon $x_h$ is imposed. Consequently, they have been characterized as ``black holes within magnetic monopoles'' \cite{ewein} or ``black holes within dyons'' \cite{bhk}. Distinct from the corresponding embedded Reissner-Nordstr\o m (RN) black holes \cite{rn}, these non-abelian black hole solutions represent counterexamples to the ``no-hair'' conjecture, because they carry non-trivial non-Coulomb-like fields outside their horizon. In contrast, pure SU(2) EYM theory possesses neither regular dyon solutions \cite{gal1}, nor dyonic black holes other than embedded RN solutions \cite{gal2,popp}. The magnetic non-abelian black holes exist only in a limited domain of the $\alpha-x_h$ plane \cite{ewein,bfm,bfm2}. The domain of existence consists of two regimes, where for fixed coupling constant $\alpha$ and varying horizon radius $x_h$, the solutions approach limiting solutions in two distinct ways. At the transition point between these two regimes, given by the critical value $\hat \alpha$, the maximal horizon radius is attained \cite{bfm,bfm2}. Limiting our investigations to vanishing Higgs self-coupling, we here determine the domain of existence of the dyonic non-abelian black holes, finding also two regimes for the coupling constant $\alpha$. Besides the fundamental gravitating monopole and dyon solutions there are radially excited solutions \cite{bfm,bfm2,bhk}, where the gauge field function possesses nodes. Both the radially excited monopole and dyon solutions are related to the globally regular Einstein-Yang-Mills (EYM) solutions, found by Bartnik and McKinnon \cite{bm}, and, like these solutions, have no flat space counterparts. Besides the radially excited globally regular dyon solutions, we here consider the radially excited dyonic non-abelian black hole solutions and determine their domain of existence. In section II we derive the EYMH equations of motion. We present the equations both in Schwarzschild-like coordinates, employed in the numerical calculations \cite{bhk}, and in isotropic coordinates, to extend the analytical calculations \cite{bfm2} to the dyonic case. In section III we briefly review the embedded RN solutions. We discuss the globally regular dyon solutions in section IV. In addition to vanishing Higgs self-coupling, we consider the fundamental solutions also for finite Higgs self-coupling, in order to demonstrate that, as in flat space \cite{bkt}, a maximally possible charge $Q_{\rm max}$ arises. In section V we discuss the dyonic non-abelian black hole solutions. In particular, we derive the critical value $\hat \alpha$ for the dyonic solutions analytically and confirm it numerically. For vanishing Higgs self-coupling, we determine the domain of existence for the fundamental and the first radially excited dyonic non-abelian black hole solutions, and we discuss the pattern how the black hole solutions approach limiting solutions. We conclude in section VI. In Appendix A, finally we show the local existence of the non-abelian regular dyon and dyonic black hole solutions at the origin and at the horizon, respectively. \section{Einstein-Yang-Mills-Higgs Equations of Motion} We consider the SU(2) EYMH action \begin{equation} S=S_G+S_M=\int L_G \sqrt{-\rm g} d^4x + \int L_M \sqrt{-\rm g} d^4x \ \label{action} \end{equation} with \begin{equation} L_G=\frac{1}{16\pi G}{\cal R} \ , \end{equation} and \begin{equation} L_M = - \frac{1}{4} F_{\mu\nu}^a F^{a\mu\nu} - \frac{1}{2} D_\mu \phi^a D^\mu \phi^a - \frac{\beta^2 g^2}{4} (\phi^a \phi^a - v^2)^2 \ , \end{equation} where \begin{equation} F_{\mu\nu}^a = \partial_\mu A_\nu^a - \partial_\nu A_\mu^a + g \epsilon^{abc} A_\mu^b A_\nu^c \ , \end{equation} \begin{equation} D_\mu \phi^a = \partial_\mu \phi^a + g \epsilon^{abc} A_\mu^b \phi^c \ , \end{equation} $g$ is the gauge coupling constant, $\beta$ is the dimensionless Higgs coupling constant, proportional to the ratio of Higgs boson mass and vector boson mass, and $v$ is the Higgs field vacuum expectation value. Variation of the action eq.~(\ref{action}) with respect to the metric $g_{\mu\nu}$, the gauge field $A_\mu^a$ and the Higgs field $\phi^a$ leads to the Einstein equations and the matter field equations. \subsection{Ansatz} For static spherically symmetric globally regular and black hole solutions the metric can be parametrized as \cite{bfm2,bck} \begin{equation} ds^2=g_{\mu\nu}dx^\mu dx^\nu= - e^{2 \nu (R)} dt^2 + e^{2 \lambda (R)} dR^2 + r^2(R) (d\theta^2 + \sin^2\theta d\phi^2) \ . \label{genmet} \end{equation} For the gauge and Higgs field we employ the spherically symmetric ansatz \cite{bhk} \begin{equation} \vec A_t = \vec e_r J(r) v \ , \end{equation} \begin{equation} \vec A_r=0 \ , \ \ \ \vec A_\theta = -\vec e_\phi \frac{1- K(r)}{g} \ , \ \ \ \vec A_\phi = \vec e_\theta \frac{1- K(r)}{g} \sin \theta \ , \end{equation} and \begin{equation} \vec \phi = \vec e_r H(r) v \ , \end{equation} with the standard unit vectors $\vec e_r$, $\vec e_\theta$ and $\vec e_\phi$. For $A_t^a=0$, the solutions carry only magnetic charge \cite{ewein,bfm,aichel}. With these ans\"atze we obtain the action \begin{eqnarray} {S} = \frac{1}{G} \int dt dR e^{\nu+\lambda} && \Biggl( {1\over 2} \left[1+e^{-2\lambda} \left((r')^2 +\nu'(r^2)'\right)\right] \Biggr. \nonumber\\ &&+4 \pi G \left[ e^{-2(\lambda+\nu)} \frac{r^2}{2} (J')^2 v^2 + e^{-2\nu} K^2J^2v^2 \right. \nonumber\\ && \Biggl. \left. - e^{-2\lambda} \left( \frac{(K')^2}{g^2} + {r^2\over 2} (H')^2v^2) \right) - V_2 \right] \Biggr) \ , \label{ac1} \end{eqnarray} where $\nu$, $\lambda$, $r$, $K$, $J$ and $H$ are functions of $R$, the prime indicates the derivative with respect to $R$, and \begin{equation} V_2 = {(1-K^2)^2\over{2g^2r^2}} + K^2 H^2 v^2 + {\beta^2 g^2 \over 4} r^2 (H^2-1)^2 v^4 \ . \label{v2} \end{equation} \subsection{Schwarzschild-like coordinates} We first employ Schwarz\-schild-like coordinates, corresponding to the gauge $r(R)=R$. Renaming $R=r$, the spherically symmetric metric reads \begin{equation} ds^2= -A^2N dt^2 + N^{-1} dr^2 + r^2 (d\theta^2 + \sin^2\theta d\phi^2) \ , \end{equation} with the metric functions \begin{equation} A(r)= e^{\lambda + \nu} \ , \label{a} \end{equation} \begin{equation} N(r)=1-\frac{2m(r)}{r} = e^{-2 \lambda} \ , \label{n} \end{equation} and the mass function $m(r)$. We now introduce the dimensionless coordinate $x$ and the dimensionless mass function $\mu$, \begin{equation} x = g v r \ , \ \ \ \mu=g v m \ , \label{xm} \end{equation} as well as the dimensionless coupling constant $\alpha$ \begin{equation} \alpha^2 = 4 \pi G v^2 \ . \end{equation} The $tt$ and $rr$ components of the Einstein equations then yield the equations for the metric functions, \begin{eqnarray} \mu'&=&\alpha^2 \Biggl( \frac{x^2 J'^2}{2 A^2} + \frac{J^2 K^2}{A^2 N} + N K'^2 + \frac{1}{2} N x^2 H'^2 \nonumber\\ & &\phantom{ \alpha^2 \Biggl( } + \frac{(K^2-1)^2}{2 x^2} + H^2 K^2 + \frac{\beta^2}{4} x^2 (H^2-1)^2 \Biggr) \ , \label{eqmu} \end{eqnarray} and \begin{eqnarray} A'&=&\alpha^2 x \Biggl( \frac{2 J^2 K^2}{ A^2 N^2 x^2} + \frac{2 K'^2}{x^2} + H'^2 \Biggr) A \ , \label{eqa} \end{eqnarray} where the prime now indicates the derivative with respect to $x$. For the matter functions we obtain the equations \begin{eqnarray} (A N K')' = A K \left( \frac{K^2-1}{x^2} + H^2 - \frac{J^2}{A^2 N} \right) \ , \label{eqk} \end{eqnarray} \begin{equation} \left( \frac{x^2 J'}{A} \right)' = \frac{2 J K^2}{AN} \ , \label{eqj} \end{equation} and \begin{eqnarray} ( x^2 A N H')' = A H \left( 2 K^2 + \beta^2 x^2 (H^2-1) \right) \ . \label{eqh} \end{eqnarray} Introducing the new function \begin{equation} P = \frac{J}{A} \label{p} \end{equation} and using eq.~(\ref{eqa}), we see that the metric function $A$ can be eliminated from the set of equations (\ref{eqmu}), (\ref{eqk})-(\ref{eqh}). These can then be solved separately, with the metric function $A$ being determined subsequently by (\ref{eqa}). In the following we refer to the above set of functions, variables and parameters, also employed in \cite{bhk}, as notation I. \subsection{Isotropic coordinates} To make contact with the analytical results of \cite{bfm,bfm2} and to extend them to the case of gravitating dyons and dyonic black holes, we now present the equations of motion in the notation of \cite{bfm2}, in the following referred to as notation II. In the Schwarzschild gauge the following identifications hold \begin{equation} \begin{array}{ccccccccc} \cite{bhk} : & x & A & N & K & H & J & \alpha & \beta^2\\ \cite{bfm2} : & \alpha r & e^{\lambda+\nu} & e^{-2\lambda} & W & \alpha H & J & \alpha & \frac{\beta^2}{2} \end{array} \end{equation} Along with \cite{bfm2}, we now employ the gauge choice $e^{\lambda}={r(R)\over R}$, corresponding to isotropic coordinates, and introduce the coordinate \begin{equation} \tau = \ln R \ , \label{tau} \end{equation} and the dot ( $\dot {} $ ) now indicates the derivative with respect to $\tau$. To obtain a system of first order equations we further employ for the first derivatives the new variables \cite{bfm2} \begin{equation} N={\dot r\over r}\ , \ \ \kappa=\dot \nu + N \ , \ \ U={\dot W\over r}\ , \ \ V=\dot H \ , \end{equation} and we introduce \begin{equation} B = e^{-\nu} J\ , \ \ C=\dot B \ . \label{b} \end{equation} (Note, that the symbol $N$ here has a totally different meaning from eq.~(\ref{n}).) Then the following system of first order autonomous equations is obtained \begin{eqnarray} \label{autoi} \dot r &=& Nr\\ \dot N &=& N(\kappa-N)-(2U^2+V^2) - 2\alpha^2B^2W^2\\ \dot \kappa &=& 1-\kappa^2+2U^2-{\beta^2r^2\over 2} (H^2-\alpha^2)^2-2W^2H^2 + 2\alpha^2 W^2B^2\\ \dot H &=& V\\ \dot V &=& {\beta^2r^2\over 2} (H^2-\alpha^2)H+2W^2H- \kappa V\\ \dot W &=& rU\\ \dot U &=& U(N-\kappa)+ {W(W^2-1)\over r} + Wr (H^2-\alpha^2B^2)\\ \dot B &=& C\\ \dot C &=& -\kappa C+B(2W^2-4U^2-V^2+\kappa^2-1 \nonumber \\ &\ & +2W^2H^2+{\beta^2r^2\over 2} (H^2-\alpha^2)^2) - 4\alpha^2B^3W^2 \label{autof} \end{eqnarray} supplemented by the constraint \begin{equation} \label{contr} 2\kappa N = 1+N^2+2U^2+V^2-2V_2 + 2\alpha^2W^2B^2-\alpha^2(C+B(\kappa-N))^2 \end{equation} compatible with the equations. \section{Embedded Reissner-Nordstr\o m Solutions} As noted long ago \cite{rn}, the Einstein-Yang-Mills-Higgs system admits embedded RN solutions. In notation I, RN solutions with mass $\mu_\infty$, unit magnetic charge and arbitrary electric charge $Q$ are given by \begin{equation} \mu(x) = \mu_\infty - \frac{\alpha^2 (1 + Q^2)}{2x} , \ \ \ A(x)=1 \ , \end{equation} \begin{equation} K(x)=0 \ , \ \ \ J(x) = J_\infty - \frac{Q}{x} \ , \ \ \ H(x)=1 \ , \label{Q} \end{equation} where $J_\infty$ is a priori arbitrary, but here chosen as $J_\infty = Q/x_h$ (see eq.~(\ref{Jhor})). At the regular horizon $x_h$ the metric function \begin{equation} N(x) = {x^2 - 2x \mu_{\infty} + \alpha^2(1 + Q^2) \over x^2} \end{equation} vanishes, $N(x_h)=0$, yielding \begin{equation} x_h = \mu_{\infty} + \sqrt{\mu_{\infty}^2 - \alpha^2 (1+Q^2)} \end{equation} or equivalently \begin{equation} \mu_{\infty} = {x_h^2 + \alpha^2 (1+Q^2) \over 2 x_h} \ . \end{equation} For fixed values of $\alpha$ and $Q$, RN solutions exist for $x_h \geq \alpha \sqrt{1+Q^2}$, independent of $\beta$. In the extremal case \begin{equation} x_h = \mu_{\infty} = \alpha \sqrt{1+Q^2} \ , \label{exh} \end{equation} extremal RN solutions are obtained. Extremal RN solutions are characterized by $Q$ and $\alpha$. In particular, for extremal RN solutions the function $B(x)$ defined in (\ref{b}) becomes constant \begin{equation} B = {J \over A \sqrt{N}} = {Q \over \alpha \sqrt{1+Q^2}} \ . \label{brn} \end{equation} \section{Dyon Solutions} Let us first consider the globally regular particle-like solutions of the SU(2) EYMH system. In the Prasad-Sommerfield limit, $\beta=0$, the dyon solutions in flat space are known analytically \cite{jul}, whereas for finite $\beta$ they are obtained only numerically \cite{bkt}. In the presence of gravity, the corresponding gravitating monopole and dyon solutions are obtained only numerically as well. The gravitating dyon solutions have many features in common with the gravitating monopole solutions \cite{bhk}. After presenting the boundary conditions for the asymptotically flat globally regular solutions, we here briefly discuss the fundamental dyon solutions both in the Prasad-Sommerfield limit and for finite $\beta$. Then we turn to the excited dyon solutions. \subsection{Boundary conditions} Requiring asymptotically flat solutions implies in notation I that the metric functions $A$ and $\mu$ both approach a constant at infinity. We here adopt \begin{equation} A(\infty)=1 \ , \end{equation} and $\mu(\infty)=\mu_\infty$ represents the dimensionless mass of the solutions. The matter functions also approach constants asymptotically, \begin{equation} K(\infty)=0 \ , \ \ \ J(\infty)= J_\infty \ , \ \ \ H(\infty) = 1 \ , \end{equation} where for magnetic monopole solutions $J_\infty = 0$. The asymptotic fall-off of the function $J(x)$ determines the dimensionless electric charge $Q$ (see eq.~(\ref{Q})). Regularity of the solutions at the origin requires \begin{equation} \mu(0)=0 \ , \end{equation} and \cite{jul} \begin{equation} K(0)=1 \ , \ \ \ J(0) = 0 \ , \ \ \ H(0)=0 \ . \label{bc0} \end{equation} The local existence of a family of analytic solutions obeying these conditions is shown in Appendix A.1. \subsection{Fundamental dyons} \boldmath \indent {\bf $\beta=0$} \unboldmath Like the gravitating monopole solutions, the gravitating dyon solutions exist up to a maximal value of the coupling constant $\alpha$. Beyond this value no dyon solutions exist. The fundamental dyon branch does not end at the maximal value $\alpha_{\rm max}$, but bends backwards and extends up to the critical value $\alpha_c$. Since $\alpha^2=4\pi G v^2$, variation of the coupling constant along the fundamental dyon branch can be considered as first increasing $G$ with $v$ fixed up to the maximal value $\alpha_{\rm max}$, and then decreasing $v$ with $G$ fixed up to the critical value $\alpha_c$. Completely analogously to the fundamental monopole branch \cite{bfm}, the fundamental dyon branch reaches a limiting solution at the critical value $\alpha_c$, where it bifurcates with the branch of extremal RN solutions of unit magnetic charge and electric charge $Q$ \cite{bhk}. The critical value $\alpha_c(Q)$ depends only slightly on the charge $Q$ \cite{bhk}. For $\alpha \rightarrow \alpha_c$ the minimum of the metric function $N(x)$ of the dyon solutions decreases monotonically, approaching zero at $x_c = \alpha_c \sqrt{1+Q^2}$. For $x \ge x_c$, the metric function $N(x)$ of the limiting solution corresponds to the metric function of the extremal RN black hole with horizon $x_h = x_c = \alpha_c \sqrt{1+Q^2}$, unit magnetic charge and electric charge $Q$. Likewise, for $x \ge x_c$, the other functions of the limiting solution correspond to those of this extremal RN black hole \cite{bhk}. Consequently, also the mass of the limiting solution coincides with the mass of this extremal RN black hole. On the interval $0 \le x \le x_c$ the functions $N(x)$, $K(x)$ and $H(x)$ of the limiting solution vary smoothly from their respective boundary values at $x=0$ to those of the extremal RN solution at $x_c$, whereas the function $J(x)$ is identically zero. The metric function $A(x)$ is identically zero, as well, on the interval $0 \le x < x_c$, but discontinuous at $x_c$. Consequently, in the limiting spacetime the inner and outer parts are not analytically connected \cite{ewein,bfm}. \boldmath \indent {\bf $\beta>0$} \unboldmath Let us briefly consider the dyon solutions for finite Higgs self-coupling, $\beta > 0$, because we here encounter a phenomenon, also present for the black hole solutions at vanishing Higgs self-coupling, $\beta=0$. In flat space dyon solutions of arbitrarily large charge exist only for $\beta = 0$. In contrast, for finite values of $\beta$ dyon solutions exist only up to a maximal value of the charge, $Q_{\rm max}(\beta)$ \cite{bkt}. As the maximal value of the charge is approached, the function $J(x)$ approaches asymptotically the limiting value \begin{equation} \lim_{Q \rightarrow Q_{\rm max}} J_\infty = 1 \ . \label{Jlim} \end{equation} At this point the solutions change character and become oscillating instead of asymptotically decaying \cite{bkt}. As illustrated in Figs.~1-2, this phenomenon persists in curved space. In Fig.~1 we show the mass of the flat space dyon (solid line) and of the gravitating dyon (dashed line, corresponding to $\alpha = 0.5$) as a function of the charge $Q$ for $\beta^2=0$ and $\beta^2=0.5$. In Fig.~2 the corresponding asymptotic values of the function $J(x)$ are presented. We note, that for finite $\beta$ also gravitating dyon solutions exist only up to a maximal value of the charge, where $J_\infty \rightarrow 1$. \subsection{Radially excited dyons} Besides the branch of fundamental dyon solutions there are branches of radially excited dyon solutions, where the gauge field function $K(x)$ of the $n$-th radially excited dyon solution has $n$ nodes. These solutions have no flat space counterparts, and the variation of $\alpha$ along a branch of excited solutions must be interpreted as a variation of $v$ with $G$ fixed. In the limit $\alpha \rightarrow 0$ the Higgs field vacuum expectation value therefore vanishes, while $G$ remains finite. Because of the particular choice of dimensionless variables (\ref{xm}), in this limit the solutions shrink to zero size and their mass $\mu$ diverges. As for the radially excited monopole solutions \cite{bfm}, the coordinate transformation $\tilde x = x/\alpha$ leads to finite limiting solutions in the limit $\alpha \rightarrow 0$, which are the Bartnik-McKinnon solutions \cite{bhk}. In Fig.~3 we show the normalized mass $\mu_\infty/\alpha$ as a function of $\alpha$ for the first excited dyon branch with electric charge $Q=1$ and $\beta=0$. For comparison also the first excited monopole branch is shown. Like the radially excited monopole solutions, the radially excited dyon solutions exist only below a critical value of the coupling constant $\alpha$. For the radially excited monopole solutions this critical value is $\alpha_c = \sqrt{3}/2$ \cite{bfm,bfm2}. For the radially excited dyon solutions the critical value is larger and will be discussed in the context of the dyonic black hole solutions. \section{Black Hole Solutions} We now turn to the dyonic black hole solutions of the SU(2) EYMH system, choosing $\beta=0$. Again, the SU(2) EYMH dyonic black holes have many features in common with the magnetic non-abelian black holes \cite{bhk}. In particular, dyonic non-abelian black hole solutions exist in a limited domain of the $\alpha$-$x_h$ plane, which depends on the charge $Q$, and, in the limit $x_h \rightarrow 0$, the corresponding globally regular solutions are obtained. But a new phenomenon occurs for small values of $\alpha$, where the non-abelian solutions no longer bifurcate with non-extremal RN solutions. In the following, we briefly present the boundary conditions. We generalize the considerations of \cite{bfm2} to dyonic solutions and derive analytically the critical value of the coupling constant, $\hat \alpha$. We then present our numerical calculations for the fundamental dyonic black hole solutions and their radial excitations. \subsection{Boundary conditions} Imposing the condition of asymptotic flatness, the black hole solutions satisfy the same boundary conditions at infinity as the regular solutions. The existence of a regular event horizon at $x_h$ leads in notation I to the conditions for the metric functions \begin{equation} \mu(x_h)= \frac{x_h}{2} \ , \end{equation} and $A(x_h) < \infty $, and for the matter functions \begin{eqnarray} {N' K' }|_{x_h} = \left. K \left( \frac{K^2-1}{x^2} + H^2 \right) \right|_{x_h} \ , \end{eqnarray} \begin{equation} J |_{x_h} =0 \ , \label{Jhor} \end{equation} and \begin{eqnarray} {x^2 N' H' }|_{x_h} = \left. H \left( 2 K^2 + \beta^2 x^2 (H^2-1) \right) \right|_{x_h} \ . \end{eqnarray} The local existence of a family of solutions obeying these conditions at the horizon is demonstrated in Appendix A.2. \boldmath \subsection{Critical value of $\alpha$} \unboldmath For magnetic black holes, there are two regimes of the coupling constant $\alpha$, corresponding to two distinct patterns of reaching a limiting solution as a function of the horizon radius $x_h$. For larger values of $\alpha$, the solutions bifurcate with an extremal RN black hole solution, when the horizon radius appoaches its critical value $x_h^{\rm cr}$. Here the solutions tend towards the corresponding extremal RN solution with horizon radius $x_h^{\rm RN} = \alpha$ only on the interval $x \ge x_h^{\rm RN}$, whereas they tend to a generic non-abelian limiting solution on the interval $x_h^{\rm cr} \le x \le x_h^{\rm RN}$ \cite{bhk}. In contrast for smaller values of $\alpha$ the solutions bifurcate with a non-extremal RN solution on the full interval $x \ge x_h^{\rm cr}$ \cite{bfm,bfm2}. The transition point between these two regimes corresponds to the critical value $\hat \alpha$, where the non-abelian black hole solutions tend towards the extremal RN solution with horizon radius $x_h^{\rm RN} = \hat \alpha$ on the full interval $x \ge x_h^{\rm cr} = x_h^{\rm RN}$, i.e.~$x_h^{\rm cr}$ and $x_h^{\rm RN}$ coincide. For the magnetic non-abelian black holes this critical value is given by $\hat \alpha = \sqrt{3}/2$ \cite{bfm,bfm2}. Analogously, for dyonic non-abelian black holes there are also two regimes of the coupling constant $\alpha$, with $\hat \alpha$ marking the transition point between these two regimes. We now derive this critical value $\hat \alpha$ for the dyonic non-abelian black hole solutions, following closely the reasoning of \cite{bfm2}. Let us then consider the fixed points of the system (\ref{autoi})-(\ref{autof}). Employing notation II, the relevant fixed point is given by the configuration \begin{equation} \label{fixp} H_h = \alpha \ , \ \ W_h = 0 \ , \ \ \kappa_h=1 \ , \ \ r_h = {1\over \sqrt{1- \alpha^2 B_h^2}} \ , \end{equation} where $N_h = V_h = U_h = C_h = 0$, while $B_h$ is such that $0 \leq B_h \leq 1/ \alpha$. Linearization of the equations about this fixed point leads to the following eigenvalues of the matrix defining the linear part (adopting the order $N$, $W$, $U$, $H$, $V$, $P$, $Q$, $\kappa$) \begin{eqnarray} \label{eig0} \gamma_0 &=& 1\\ \gamma_{1,2} &=& -{1\over 2} \pm i \sqrt{ {3\over 4}- \alpha^2 {{1 -B^2_h} \over {1 - \alpha^2 B_h^2}} } \label{eig12}\\ \gamma_{3,4} &=& -{1\over 2} \pm \sqrt{{1\over 4} + \alpha^2\beta^2r_h^2}\\ \gamma_5 &=& 0\\ \gamma_6 &=& -1\\ \gamma_7 &=& -2 \label{eig7} \end{eqnarray} In particular, we see that the eigenvalues $\gamma_{1,2}$, which are related to the function $W$ and determine the critical value $\hat \alpha$ for the magnetic black holes \cite{bfm2}, now depend on $B_h$. The zero mode $\gamma_5$ is related to the occurrence of this free parameter. For $Q=0$, the critical value of $\alpha$ corresponds to the transition of the eigenvalue $\gamma_{1,2}$ (eq.~(\ref{eig12}) for $Q=0$) from complex to real value \cite{bfm2}. Repeating this reasoning for dyonic black holes, we conclude, that the corresponding critical value of the coupling constant for the dyonic non-abelian black holes is given by \begin{equation} \hat \alpha^2 = {3\over{4-B^2_h}} \ . \label{acrit} \end{equation} Here the relation of the critical value $\hat \alpha$ to the charge $Q$ seems only indirect, since the value of the function $B(x)$ at the horizon enters in eq.~(\ref{acrit}). However, for the dyonic non-abelian black holes $\hat \alpha$ also represents the special point, where the non-abelian black hole solutions tend towards the corresponding extremal RN solution with horizon radius $x_h^{\rm RN} = \hat \alpha \sqrt{1+Q^2}$ on the full interval $x \ge x_h^{\rm cr} = x_h^{\rm RN}$. This implies for the function $B(x)$ according to eq.~(\ref{brn}) \begin{equation} \label{valp} \lim_{\alpha \rightarrow \hat \alpha} B(x) = {Q \over \hat \alpha \sqrt{1+Q^2} } \ , \ \ \forall x > x_h^{\rm RN} \ . \end{equation} Consequently, we may insert for $B_h$ in eq.~(\ref{acrit}) the constant RN value, eq.~(\ref{brn}), and obtain for the critical coupling constant the expression \begin{equation} \hat \alpha(Q) = \sqrt{{3 + 4 Q^2 \over 4(1+Q^2)}} \ , \label{ahat} \end{equation} which now depends on the electric charge $Q$. \subsection{Fundamental dyonic black holes} Let us now define some quantities helpful in the analysis of the numerical results. Denoting the set of functions of the extremal RN solution with horizon radius $x_h^{\rm RN} = \alpha \sqrt{1+Q^2}$ by $(N_{\rm RN},K_{\rm RN},H_{\rm RN},P_{\rm RN})$, and denoting the set of functions of the non-abelian black hole solution with horizon radius $x_h$ (with $x_h < x_h^{\rm RN}$) by $(N,K,H,P)$, we define \begin{equation} D N = {\rm max}_{x \in [x_h^{\rm RN}, \infty]} \vert N_{\rm RN}(x) - N(x) \vert \ , \label{dn} \end{equation} and analogously we define the quantities $D K$, $D P$ and $D H$. Clearly the quantities $DN$, $DK$, $DP$ and $DH$ must vanish, when the non-abelian black hole solutions bifurcate with the corresponding extremal RN solution. Therefore these quantities help to determine the critical values with good accuracy in the regime $\alpha \ge \hat \alpha$. \boldmath {\rm $Q=0$} \unboldmath Our numerical evaluation of the domain of existence of the magnetic non-abelian black holes in the $\alpha$-$x_h$ plane is in agreement with the results of \cite{bfm,bfm2}, and in particular we confirm $\hat \alpha = \sqrt{3}/2$. Considering the quantities $DN$, $DK$ and $DH$ (eq.~(\ref{dn})) as functions of the horizon radius $x_h$, we observe that for $\alpha \ge \hat \alpha$ the three functions approach zero at a common value $x_h^{\rm cr}(\alpha) \le \hat \alpha$. As $\alpha - \hat \alpha$ changes sign, in particular the quantity $DK$ changes drastically, becoming a convex function when the critical horizon radius $x_h^{\rm cr}$ is approached. \boldmath {\rm $Q \ne 0$} \unboldmath We have numerically determined the domain of existence of the fundamental dyonic non-abelian black hole solutions in the $\alpha$-$x_h$ plane, which is shown in Fig.~4 for $Q=1$. The straight line indicates the extremal RN solutions, dividing the domain of existence into a lower region with non-abelian black holes only and an upper region with both non-abelian and non-extremal abelian black holes, completely analogously to the case of magnetic black holes. To start the detailed discussion of the diagram, we choose a small fixed value of the horizon radius $x_h$ and vary $\alpha$. Then the dyonic black hole solutions exist only up to some maximal value $\alpha_{\rm max}(x_h)$. >From $\alpha_{\rm max}(x_h)$ a second branch extends backwards until a critical value $\alpha_c(x_h)$ is reached. With increasing $x_h$ the second branch becomes increasingly smaller, and finally disappears. The critical values $\alpha_{\rm max}(x_h)$ and $\alpha_c(x_h)$ are both shown in Fig.~4. The presence of the two branches is illustrated in Fig.~5 for $Q=1$ and $x_h = 0.8$, where the quantities $DN$, $DK$, $DP$ and $DH$ are shown as functions of the parameter $\alpha$. Here the second branch of solutions only exists in the small interval $1.1986 \le \alpha \le 1.2002$. For fixed $x_h$ and varying $\alpha$, the functions of the black holes solutions evolve completely analogously to those of the regular dyon solutions. Consequently, the limiting solution reached at the critical value $\alpha_c(x_h)$ consists of two parts, the outer part corresponding to the exterior of the extremal RN black hole with horizon $x_h^{\rm RN} = \alpha_c \sqrt{1+Q^2}$, unit magnetic charge and electric charge $Q$ and the inner part representing a genuine non-abelian solution. Similarly for fixed $\alpha > \hat \alpha$ and varying $x_h$, the limiting solution reached consists of two parts, the outer part corresponding to the exterior of the extremal RN black hole with horizon $x_h^{\rm RN} = \alpha \sqrt{1+Q^2}$, unit magnetic charge and electric charge $Q$, as demonstrated in Figs.~6-7 for the black hole solutions with horizon $x_h=1$, 1.1 and 1.153, charge $Q=1$ and $\alpha=1$ for the functions $N$ and $P$ and the functions $K$ and $H$, respectively. According to eq.~(\ref{ahat}), for $Q=1$ the critical value $\hat \alpha$ is given by $\hat \alpha = \sqrt{7/8} \approx 0.935$. Here the transition between the two regimes with different bifurcation patterns is supposed to occur. Indeed, our numerical calculations confirm this critical value, as demonstrated in Figs.~8-9, where the four quantities $DN$, $DK$, $DP$ and $DH$ are shown as functions of the horizon radius $x_h$ for $\alpha=0.94$ and $\alpha=0.93$, respectively. For $\alpha=0.94$, which is slightly above $\hat \alpha$, the curves in Fig.~8 clearly approach zero at $x_h^{\rm cr} \approx 1.235$. This value is well below the horizon value of the corresponding extremal RN solution, $x_h^{\rm RN}= \alpha \sqrt{1+Q^2} \approx 1.329$. Moreover, in this limit $x_h \rightarrow x_h^{\rm cr}$, the corresponding values of $K(x_h)$ and $H(x_h)$ do not tend to their RN values of zero and one, respectively, but they converge to generic numbers in the interval $[0,1]$. In contrast, for $\alpha=0.93$, which is slightly below $\hat \alpha$, the function $DK$ becomes a convex function when the critical horizon radius $x_h^{\rm cr} \approx 0.93 \sqrt{2} \approx 1.315$ is approached. Furthermore, the quantities $K(x_h)$ and $H(x_h)$ tend to their RN values zero and one, respectively, for $x_h \rightarrow x_h^{\rm cr}$. Thus for $\alpha = 0.93$ the limiting solution represents a RN solution on the full interval $x \ge x_h^{\rm cr}$. Clearly, the transition point $\hat \alpha$ occurs inbetween these two values. Approaching $\hat \alpha$ further from both sides, the value $\hat \alpha = \sqrt{7/8} \approx 0.935$ is confirmed. The gap in Fig.~4 on the rhs of $\hat \alpha$ represents a tiny region, numerically not accessible with sufficient accuracy \cite{foot}. We have confirmed the validity of eq.~(\ref{ahat}) also for other values of the charge $Q$. In the regime $\alpha < \hat \alpha$, we observe that the maximal value of the horizon radius $x_h$ decreases with decreasing $\alpha$, completely analogously to the magnetic case. Continuing our detailed discussion of the approach to the limiting solution in this regime, however, we observe differences to the magnetic case for smaller values of $\alpha$ and larger values of $Q$. For fixed $\alpha$ close to $\hat \alpha$ and increasing $x_h$, the dyonic black hole solutions bifurcate with a RN solution when the maximal value of the horizon is reached. The limiting solution corresponds to the exterior of the RN solution on the full interval $x \ge x_h^{\rm cr}$. For smaller values of $Q$ and somewhat smaller fixed $\alpha$, with increasing $x_h$ the solutions do not yet bifurcate with a non-extremal RN solution, when the maximal value of the horizon is reached. Instead a second branch of solutions emerges, extending backwards to lower values of the horizon radius, until it bifurcates with a non-extremal RN solution at the critical value of the horizon $x_h^{\rm cr}$ \cite{bfm,bfm2,bhk}. Moving along both branches, the values of $K(x_h)$ and $H(x_h)$ change monotonically, reaching their respective RN values of zero and one as the critical horizon is reached. For the magnetic black holes as well as for the dyonic black holes with smaller values of $Q$ this latter bifurcation pattern persists with decreasing $\alpha$ \cite{bfm,bfm2}. In contrast, for the dyonic non-abelian black holes with larger values of $Q$ new phenomena occur. We exhibit these new phenomena in Fig.~10, where the critical region of the $\alpha$-$x_h$ plane shown in Fig.~4 for $Q=1$ is enlarged and more details are given. First, we observe the occurrence of bifurcations, which lead to the small triangular region $ABC$, where three solutions exist. Thus below $\hat \alpha$ we find four distinct regions in the $\alpha$-$x_h$ plane. With decreasing $\alpha$ and increasing $x_h$ there are i) for $0.605 < \alpha$ first one, then zero solutions, ii) for $0.575 < \alpha < 0.605$ first one, then three, then one, then zero solutions, iii) for $0.535 < \alpha < 0.575$ first one, then three, then two, then zero solutions and iv) for $\alpha < 0.535$ first one, then two, then zero solutions. Second, with decreasing $\alpha$, the value $J_\infty$ of the limiting solution increases. When it reaches its maximal value of one, the function $K(x)$ ceases to decay exponentially. The limiting solution then no longer represents a non-extremal RN solution, instead an oscillating solution is reached as the limiting solution. The corresponding critical point, where the transition from the limiting non-extremal RN solutions to the oscillating solutions occurs, is labelled $D$ in Fig.~10. Along the curve $BD$ the maximal value of the horizon radius is reached when the limiting value $J_\infty=1$ is attained, whereas along the curves $AB$ and $A'A$ an intermediate extremal value of the horizon radius is reached when the limiting value $J_\infty=1$ is attained. Figs.~11-13 illustrate both phenomena, representing the values $K(x_h)$, $H(x_h)$ and $J_\infty$ as functions of the horizon radius for $\alpha=0.7$, 0.55 and 0.2, respectively. As seen in Fig.~11, for $\alpha=0.7$ only one branch of solutions exists. Here, with increasing $x_h$ $K(x_h)$ decreases monotonically to zero, $H(x_h)$ increases monotonically to one, while $J_\infty$ increases monotonically to a limiting value smaller that one. Decreasing $\alpha$, the critical point $C$ of Fig.~10 is reached, below which three branches of solutions exist. Moving along the three branches, $K(x_h)$ decreases monotonically, $H(x_h)$ increases monotonically and $J_\infty$ increases monotonically until it reaches one, as seen in Fig.~12 for $\alpha = 0.55$. For still smaller values of $\alpha$, at the critical value $A$, the third branch disappears, leaving two branches which end when $J_\infty$ reaches one. This is seen in Fig.~13 for $\alpha = 0.2$. The maximal horizon radius in this case is approximately 0.917. For small values of $\alpha$ (i.e.~below the critical point $A$) the lower curve in Fig.~4 represents the critical value of the horizon, where the asymptotically exponentially decaying solutions cease to exist. Considering the $Q$-dependence of the new phenomena, we observe that the bifurcations leading to three branches arise only for $Q \ge 0.8$, and the limiting value $J_\infty = 1$ is reached only for $Q \ge 0.75$. The latter is seen in Fig.~14, where the critical value $D$ is shown, which marks the transition from the limiting non-extremal RN solutions to the oscillating solutions. Also shown in Fig.~14 is the critical value $\hat \alpha$. We note, that with increasing $Q$ the two curves get increasingly closer. \subsection{Excited dyonic black holes} Besides the fundamental magnetic black hole solutions there exist radially excited black hole solutions, for which the gauge function $K(x)$ possesses $n$ nodes \cite{bfm,bfm2}. These radially excited solutions exist only for $\alpha \le \hat \alpha = \sqrt{3}/2$ \cite{bfm,bfm2}. Our numerical analysis of the radially excited dyonic black hole solutions strongly indicates that, at least for $n=1$, the excited dyonic black hole solutions penetrate into a small domain of the $\alpha > \hat \alpha$ regime, as can already be anticipated from Fig.~3, where the first excited regular dyon solution is shown for charge $Q=1$. Indeed, we observe that the maximal value of $\alpha$, where the radially excited black hole solutions cease to exist, increases more strongly with $Q$ than the critical value $\hat \alpha$. Let us now consider the domain of existence of the black hole solutions in the $\alpha-x_h$ plane in more detail. In the regime $\alpha > \hat \alpha$, for fixed $x_h$ and increasing $\alpha$ the solutions again reach a limiting solution at the critical value $\alpha_c(x_h)$, which consists of two parts, the outer part corresponding to the exterior of the extremal RN black hole with horizon $x_h^{\rm RN} = \alpha_c \sqrt{1+Q^2}$, unit magnetic charge and electric charge $Q$. For $\alpha \rightarrow \alpha_c(x_h)$ the function $N(x)$ develops a progressively decreasing minimum, $N_{\rm min}$, reaching zero at $x_h^{\rm RN}= \alpha_c(x_h) \sqrt{1+Q^2}$. This is illustrated in Fig.~15, where the value of $N_{\rm min}$ is shown as a function of $\alpha$ for the horizon radii $x_h=0.01$, 0.4 and 0.8 and charge $Q=1$. With increasing $x_h$ the value of $\alpha_c(x_h)$ decreases, though it remains greater than $\hat \alpha$ even for $x_h=0.8$, as inspection of the four quantities $DN$, $DK$, $DP$ and $DH$ reveals. In Fig.~16 we show the functions $N(x)$, $K(x)$, $H(x)$ and $P(x)$ of the first radially excited black hole solution for charge $Q=1$, horizon radius $x_h=0.4$ and $\alpha=0.938 > \hat \alpha$. The minimum of the function $N(x)$, which here occurs close to $x_h^{\rm RN}= \alpha_c(x_h) \sqrt{1+Q^2}$, remains well separated from the horizon $x_h$ for $\alpha \rightarrow \alpha_c(x_h)$. Considering finally the first excited black hole solutions for small fixed $\alpha$ and increasing $x_h$, we observe that as for the fundamental black hole solutions, the quantity $J_\infty$ here plays a major role. With increasing $x_h$, $J_\infty$ approaches its maximal value one, where the asymptotically exponentially decaying solutions cease to exist. For $\alpha = 0.2$, for instance, this happens at $x_h \approx 0.28$. \section{Conclusions} In analogy to gravitating monopoles \cite{ewein,bfm,ewein3} also gravitating dyons \cite{bhk} exist. When a regular horizon $x_h$ is imposed, from these solutions ``black holes within monopoles'' \cite{ewein,bfm} and ``black holes within dyons'' \cite{bhk} emerge. Besides the fundamental regular and black hole solutions also radially excited solutions exist \cite{bfm,bfm2,bhk}, related to the corresponding EYM solutions \cite{bm,su2}. The static spherically symmetric ``black holes within monopoles'' and ``black holes within dyons'' provide counterexamples to the ``no-hair conjecture''. The domain of existence of the fundamental dyonic non-abelian black hole solutions in the $\alpha-x_h$ plane is similar to the one of the fundamental magnetic non-abelian black hole solutions. There are two regimes of the coupling constant $\alpha$. For $\alpha < \hat \alpha$, the maximal horizon radius increases with increasing $\alpha$, whereas for $\alpha > \hat \alpha$ it decreases. The transition point between these two regimes, $\hat \alpha$, depends on the charge $Q$ of the solutions, $\hat \alpha = \sqrt{(3 + 4 Q^2)/(4(1+Q^2))}$. For fixed $\alpha$ and varying horizon radius, a limiting solution is approached. For $\alpha > \hat \alpha$ the limiting solution consists of two distinct parts. The outer part corresponds to the exterior of an extremal RN black hole solution with horizon radius $x_h^{\rm RN} = \alpha \sqrt{1+Q^2}$, unit magnetic charge and electric charge $Q$, whereas on the interval $x_h \le x < x_h^{\rm RN}$ the limiting solution has the features of generic non-abelian solutions. At the transition point $\alpha = \hat \alpha$ the limiting solution corresponds precisely to the exterior of the extremal RN black hole solution with horizon radius $x_h^{\rm RN} = \hat \alpha \sqrt{1+Q^2}$, on the full interval $x \ge x_h$. In contrast, for $\alpha < \hat \alpha$ the limiting solution represents the exterior of a non-extremal RN solution for the larger values of $\alpha$, while for the smaller values of $\alpha$ the dyonic non-abelian black hole solutions encounter a critical point, where the solutions no longer are exponentially decaying but become oscillating. The domain of existence of the radially excited dyonic non-abelian black hole solutions in the $\alpha-x_h$ plane is similar to the one of the fundamental solutions. Again, there are two regimes of the coupling constant $\alpha$ with the transition point $\hat \alpha$. This is in contrast to the radially excited magnetic non-abelian black hole solutions, which exist only up to $\hat \alpha$ \cite{bfm,bfm2}. In a large part of their domain of existence, the magnetic non-abelian black hole solutions are classically stable \cite{ewein,bfm,bfm2,holl}. In contrast, dyonic non-abelian black hole solutions should be classically unstable. As for the classical dyon solutions in flat space, their mass should be lowered continuously, by lowering the electric charge, as long as there is no charge quantization \cite{jul}.
\section{Introduction} We recall the results of \cite{SS} restricting ourselves to the ``absolute'' case for the sake of simplicity. \subsection{The index theory for elliptic pairs}\label{section:ell} Let $X$ be a complex manifold of dimension $\dim_{{\Bbb C}}X = d$. Say that a ${\cal D}_X$-module ${\cal M}$ is {\em good} is for every relatively compact open subset $U$ of $X$ the restriction ${\cal M}\vert_U$ admits a finite filtration $G_\bullet{\cal M}\vert_U$ by ${\cal D}_U$-submodules, such that $Gr^G_\bullet{\cal M}\vert_U$ is ${\cal D}_U$-coherent with good filtration. An {\em elliptic pair} $({\cal M}^\bullet,{\cal F}^\bullet)$ on $X$ consists of \begin{itemize} \item a complex ${\cal M}^\bullet$ of ${\cal D}_X$-modules with bounded good cohomology, \item a complex ${\cal F}^\bullet$ of ${\Bbb C}$-vector spaces with bounded, ${\Bbb R}$-constructible cohomology, \end{itemize} which satisfy \[ \operatorname{char}({\cal M}^\bullet)\cap\operatorname{SS}({\cal F}^\bullet)\subseteq T^*_XX\ . \] In the case when $\operatorname{Supp}{\cal M}^\bullet\cap\operatorname{Supp}{\cal F}^\bullet$ (where the support of a complex of sheaves is understood to be the union of the supports of the cohomologies) is compact, P.~Schapira and J.-P.~Schneiders proved that \[ \dim H^\bullet(X;{\cal F}^\bullet\otimes{\cal M}^\bullet\otimes^{{\Bbb L}}_{{\cal D}_X}{\cal O}_X) <\infty \ . \] Thus, the Euler characteristic \[ \chi(X;({\cal M}^\bullet,{\cal F}^\bullet))\overset{def}{=}\sum_{i} (-1)^i \dim H^i(X;{\cal F}^\bullet\otimes{\cal M}^\bullet\otimes^{{\Bbb L}}_{{\cal D}_X}{\cal O}_X) \] is defined. The index theorem for elliptic pairs (\cite{SS}, Theorem 5.1) says that \[ \chi(X;({\cal M}^\bullet,{\cal F}^\bullet))= \int_{T^*X}\mu\operatorname{eu}({\cal M}^\bullet)\smile\mu\operatorname{eu}(F^\bullet) \] where $\mu\operatorname{eu}({\cal M}^\bullet)\in H^{2d}_{\operatorname{char}({\cal M}^\bullet)}(T^*X;{\Bbb C})$ and $\mu\operatorname{eu}(F^\bullet)\in H^{2d}_{\operatorname{SS}(F^\bullet)}(T^*X;{\Bbb C})$ are defined in \cite{SS}. The class $\mu\operatorname{eu}(F^\bullet)$ is the characteristic cycle of the constructible complex $F^\bullet$ as defined by Kashiwara (see \cite{KS} for more details). For example, if $Y\subset X$ is a closed real analytic submanifold, one has $\mu\operatorname{eu}({\Bbb C}_Y) = [T^*_YX]$. With regard to $\mu\operatorname{eu}({\cal M}^\bullet)$ P.~Schapira and J.-P.~Schneiders conjectured that it is related to a certain characteristic class of the symbol $\sigma({\cal M}^\bullet)$ of ${\cal M}^\bullet$ (\cite{SS}, Conjecture 8.5, see Conjecture \ref{conj:ch}). \subsection{The Riemann-Roch type formula} Let $\pi : T^*X\to X$ denote the canonical projection. In what follows we will assume that the complex ${\cal M}^\bullet$ admits a filtration $F_\bullet{\cal M}^\bullet$ by ${\cal O}_X$-submodules which is compatible with the action of ${\cal D}_X$ and the filtration $F_\bullet{\cal D}_X$ by order and such that the {\em symbol complex} of ${\cal M}$ defined by \[ \sigma({\cal M}^\bullet) = \pi^{-1}Gr^F_\bullet{\cal M}^\bullet\otimes_{\pi^{-1} Gr^F_\bullet{\cal D}_X}{\cal O}_{T^*X} \] has bounded ${\cal O}_{T^*X}$-coherent cohomology. Note that, by definition, $\operatorname{char}({\cal M}^\bullet)\overset{def}{=}\operatorname{Supp}\sigma({\cal M}^\bullet)$. For $\Lambda$ a closed subvariety of $T^*X$ let $K^0_\Lambda(T^*X)$ denote the Grothendieck group of perfect complexes of ${\cal O}_{T^*X}$-modules supported on $\Lambda$ (i.e. acyclic on the complement of $\Lambda$ in $T^*X$). Let $K^0(T^*X)\overset{def}{=}K^0_{T^*X}(T^*X)$. For $\Lambda$ containing $\operatorname{char}({\cal M})$ let $\sigma_\Lambda({\cal M}^\bullet)$ denote the class of $\sigma({\cal M}^\bullet)$ in $K^0_\Lambda(T^*X)$. \begin{remark} Both the characteristic variety and the class of the symbol in the Grothendieck group are independent of the choice of the good filtration. Local existence of good filtration in coherent ${\cal D}_X$-modules is sufficient to define $\sigma_\Lambda({\cal M}^\bullet)$ for ${\cal M}^\bullet$ with bounded good cohomology. \end{remark} The Chern character $ch : K^0(T^*X)\to\bigoplus_i H^{2i}(T^*X;{\Bbb C})$ admits a natural extension to the Chern character with supports (Theorem \ref{thm-intro:char-cl}) \begin{equation}\label{map:ch-with-supp} ch_\Lambda : K^0_\Lambda(T^*X)\to\bigoplus_i H^{2i}_\Lambda(T^*X;{\Bbb C}) \end{equation} which is functorial with respect to change of support. In \cite{SS}, P.~Schapira and J.-P.~Schneiders make the following conjecture. \begin{conj}\label{conj:ch} Suppose that ${\cal M}^\bullet$ is a complex of ${\cal D}_X$-modules with bounded good cohomology and $\Lambda$ is a closed conic subvariety of $T^*X$ containing $\operatorname{char}({\cal M}^\bullet)$. Then, \[ \mu\operatorname{eu}({\cal M}^\bullet) = \left[ ch_\Lambda(\sigma({\cal M}^\bullet))\smile \pi^*Td(TX)\right]^{2d}\ . \] \end{conj} (For $\alpha$ an element of a graded object we denote by $[\alpha]^p$ the homogeneous component of $\alpha$ of degree $p$.) We will refer to formulas such as the one in \ref{conj:ch} as Riemann-Roch formulas and refer to the left hand side as ``non-commutative'' and the right hand side as ``commutative''. What follows is an informal discussion of the Riemann-Roch formula and the key ideas and observations which enter into the statment and the proof of the main result of this paper (Theorem \ref{thm:mainE}) from which Conjecture \ref{conj:ch} follows. \subsubsection{} Although Conjecture \ref{conj:ch} is stated in terms of ${\cal D}$-modules, it is, in fact of micro-local (i.e. local on $T^*X$) nature and is naturally formulated in terms of modules over the ring ${\cal E}_X$ of micro-differential operators. Recall that ${\cal E}_X$ is a sheaf of algebras on $T^*X$ equipped with the canonical filtration $F_\bullet{\cal E}_X$ by order, the symbol map $\sigma : Gr^F_\bullet{\cal E}_X @>>> {\cal O}_{T^*X}$ and the canonical faithfully flat map $\pi^{-1}{\cal D}_X @>>> {\cal E}_X$ of filtered $\pi^{-1}{\cal O}_X$-algebras. The characteristic variety of a (coherent) ${\cal D}_X$-module ${\cal M}$ is the support of the ${\cal E}_X$-module $\pi^{-1}{\cal M}\otimes_{\pi^{-1}{\cal D}_X}{\cal E}_X$. Let $K^0_\Lambda({\cal D}_X)$ (respectively $K^0_\Lambda({\cal E}_X)$) denote the Grothendieck group of the category of perfect complexes of ${\cal D}_X$-modules whose characteristic variety is contained in $\Lambda$ (respectively perfect complexes of ${\cal E}_X$-modules supported on $\Lambda$). Extension of scalars gives rise to the map $K^0_\Lambda({\cal D}_X) @>>> K^0_\Lambda({\cal E}_X)$. The microlocal Euler class which appears on the non-commutative side of the Riemann-Roch formula is a map $K^0_\Lambda({\cal D}_X) @>>> H^{2d}_\Lambda (T^*X;{\Bbb C})$, but, in fact, by it's very definition, is the composition of extension of scalars with the map $\mu\operatorname{eu} : K^0_\Lambda({\cal E}_X) @>>> H^{2d}_\Lambda (T^*X;{\Bbb C})$. Similarly, the commuative side of the Riemann-Roch formula depends not on the filtered ${\cal D}_X$-module but on the filtered ${\cal E}_X$-module obtained by extension of scalars. \subsubsection{} The presence of the filtration can be accounted for in ring theoretic terms by the traditional device of the Rees construction. The Rees ring of the filtered ring $({\cal E}_X,F_\bullet)$ is the (graded) ${\Bbb C}[t]$-algebra ${\cal R}{\cal E}_X = \bigoplus_p F_p{\cal E}_X t^p\subset{\cal E}_X[t]$. A filtered ${\cal E}_X$-module is the same as a (graded) $t$-torsion-free module over ${\cal R}{\cal E}_X$: ${\cal R}$ is an exact functor defined by the formula ${\cal R}({\cal M},F_\bullet)=\bigoplus_p F_p{\cal M} t^p$. The algebra ${\cal R}{\cal E}_X$ comes equipped with the symbol map $\sigma : {\cal R}{\cal E}_X @>>> {\cal O}_{T^*X}$ which annihilates $t$ and the canonical isomoprhism ${\cal R}{\cal E}_X[t^{-1}]\overset{\sim}{=}{\cal E}_X[t^{-1},t]$. The symbol complex is given by the formula $\sigma({\cal M}^\bullet) = {\cal R}{\cal M}^\bullet\otimes_{{\cal R}{\cal E}}{\cal O}_{T^*X}$ and the complex is recovered via the canonical isomorphism ${\cal R}{\cal M}^\bullet [t^{-1}]\overset{\sim}{=}{\cal M}^\bullet [t^{-1},t]$. The natural input for the Riemann-Roch formula is conveniently summarized as a perfect complex of ${\cal R}{\cal E}_X$-modules. The use of the Rees construction is not just a neat device to keep track of the filtration. The appearance of ${\cal E}_X$ and $Gr^F_\bullet{\cal E}_X$ as, respectively, the generic and the special fiber in a one parameter deformation allows application of deformation theoretic metods which play a crucial role in the proof of the Riemann-Roch formula. \subsubsection{} It becomes necessary, particularly with the appearance of algebras such as ${\cal R}{\cal E}_X$ to use Hochschild homology, rather than appeal directly to the duality theory for the definition of the micro-local Euler class. The map $\mu\operatorname{eu} : K^0_\Lambda({\cal E}_X) @>>> H^{2d}_\Lambda (T^*X;{\Bbb C})$ factors naturally into the composition \[ K^0_\Lambda({\cal E}_X) @>{\operatorname{Eu}^{{\cal E}_X}_\Lambda}>> H^0_\Lambda(T^*X;{\cal E}_X\otimes^{{\Bbb L}}_{{\cal E}_X\otimes{\cal E}_X^{op}}{\cal E}_X) @>{\mu_{{\cal E}}}>> H^{2d}_\Lambda (T^*X;{\Bbb C}) \] where the first map is the Euler class with values in Hochschild homology and is defined in the generality of perfect complexes of modules over sheaves of algebras, and the second map is induced by the morphism \[ \mu_{{\cal E}} : {\cal E}_X\otimes^{{\Bbb L}}_{{\cal E}_X\otimes{\cal E}_X^{op}}{\cal E}_X @>>> {\Bbb C}_{T^*X}[2d] \] in the derived category of sheaves on $T^*X$ called {\em the (canonical) trace density map}. The existence of the canonical trace density map is a non-trivial fact which has to do with special properties of the algebra of micro-differential operators. The above mentioned factorization of the microlocal Euler class essentially reduces the Riemann-Roch formula to the problem of expressing the canonical trace density map in ``commutative'' terms, i.e., loosely speaking, factoring the composition \[ {\cal R}{\cal E}_X\otimes^{{\Bbb L}}_{{\cal R}{\cal E}_X\otimes{\cal R}{\cal E}_X^{op}}{\cal R}{\cal E}_X @>>> {\cal E}_X\otimes^{{\Bbb L}}_{{\cal E}_X\otimes{\cal E}_X^{op}}{\cal E}_X[t^{-1},t] @>{\mu_{{\cal E}}}>> {\Bbb C}_{T^*X}[t^{-1},t][2d] \] through the symbol map. (This reduction is not so surprising, considering the fact that the relevant portion of the Hochschild homology of micro-differential operators is spanned by the Euler classes.) Before addressing the latter, simpler, problem we make some more observations with regard to the Riemann-Roch formula intended to motivate a gearchange that follows. \subsubsection{} The added complication in the Riemann-Roch formula is the fact that commutative side is comprised of a homogeneous component of a manifestly non-homogeneous expression. As much as one would like to remove the brackets in the formula it is not clear how to extend the micro-local Euler class to a manifestly non-homogeneous characteristic class without the reformulation of the problem in terms of Hochschild homology. Hochschild homology is the natural recipient of the Lefshetz trace map. The Euler class is defined as the Lefshetz trace of the identity and posesses special symmetries not shared by the traces of other endomorphisms. As a result, the Euler class extends to the Chern character \[ ch^{{\cal E}_X}_\Lambda : K^0_\Lambda({\cal E}_X) @>>> H^0_\Lambda(T^*X;CC^{per}_\bullet({\cal E}_X)) \ . \] On the non-commutative side of the Riemann-Roch formula the transition from Hochschild homology to periodic cyclic homology amounts to switching from the integer grading to the even/odd grading, i.e. allowing not necessarily homogeneous classes which is a welcome change. Due to the same special homological properties of micro-differential operators the canonical trace density map extends in a unique fashion to the morphism (in the derived category of sheaves on $T^*X$) \[ \tilde\mu_{{\cal E}} : CC^{per}_\bullet({\cal E}_X) @>>> {\Bbb C}_{T^*X}[u^{-1},u]] \] where $u$ denotes a formal variable of degree $2$ (so that, for a complex $A^\bullet$ multiplication by $u$ establishes the isomorphism $A^\bullet @>>> uA^\bullet [2]$). \subsubsection{} The preceeding discussion was intended as a justification of the following reformulation of the Riemann-Roch formula in terms of the (non-commutative) Chern character: for a perfect complex ${\cal M}^\bullet$ of ${\cal R}{\cal E}_X$-modules \[ \tilde\mu_{{\cal E}}(ch^{{\cal E}_X}_\Lambda(\iota{\cal M}^\bullet)) = ch_\Lambda(\sigma{\cal M}^\bullet)\smile\pi^*Td(TX)\ , \] where $\iota{\cal M}^\bullet = {\cal M}^\bullet [t^{-1}]$, $\sigma{\cal M}^\bullet = {\cal M}^\bullet\otimes^{{\Bbb L}}_{{\cal R}{\cal E}_X}{\cal O}_{T^*X}$, and the equality takes place in $H^{2d}_\Lambda(T^*X;{\Bbb C}_{T^*X}[t^{-1},t][u^{-1},u]])\overset{\sim}{=} H^{even}_\Lambda(T^*X;{\Bbb C}_{T^*X}[t^{-1},t])$. \subsubsection{} The Chern character which appears on the commutative side of the Riemann-Roch formula may be formulated in terms of periodic cyclic homology of ${\cal O}_{T^*X}$ as the composition \[ K^0_\Lambda({\cal O}_{T^*X}) @>{ch^{{\cal O}_{T^*X}}_\Lambda}>> H^0_\Lambda(T^*X;CC^{per}({\cal O}_{T^*X}) @>{\tilde\mu_{{\cal O}}}>> H^0_\Lambda(T^*X;{\Bbb C}_{T^*X}[u^{-1},u]]) \] where the second map is induced by the Hochschild-Kostant-Rosenberg-Connes map $\tilde\mu_{{\cal O}} : CC^{per}({\cal O}_{T^*X}) @>>> \Omega^\bullet_{T^*X}[u^{-1},u]]$ and the de Rham isomorphism. With these notations the Riemann-Roch formula reads \[ \tilde\mu_{{\cal E}}(ch^{{\cal E}_X}_\Lambda(\iota{\cal M}^\bullet)) = \tilde\mu_{{\cal O}}(ch^{{\cal O}_{T^*X}}_\Lambda(\sigma{\cal M}^\bullet)) \smile\pi^*Td(TX)\ . \] Furthermore, the naturality of the Chern character implies that $ch^{{\cal E}_X}_\Lambda(\iota{\cal M}^\bullet) = \iota(ch^{{\cal E}_X}_\Lambda({\cal M}^\bullet))$ (respectively $ch^{{\cal O}_{T^*X}}_\Lambda(\sigma{\cal M}^\bullet) = \sigma(ch^{{\cal O}_{T^*X}}_\Lambda({\cal M}^\bullet))$) where $\iota : CC^{per}_\bullet({\cal R}{\cal E}_X) @>>> CC^{per}_\bullet({\cal E}_X)[t^{-1},t]$ (respectively $\sigma : CC^{per}_\bullet({\cal R}{\cal E}_X) @>>> CC^{per}_\bullet({\cal O}_{T^*X})$) is induced by $\iota : {\cal R}{\cal E}_X @>>> {\cal E}_X[t^{-1},t]$ (respectively $\sigma : {\cal R}{\cal E}_X @>>> {\cal O}_{T^*X}$). The Riemann-Roch formula now takes the shape \[ \tilde\mu_{{\cal E}}(\iota(ch^{{\cal E}_X}_\Lambda({\cal M}^\bullet))) = \mu_{{\cal O}}(\sigma(ch^{{\cal O}_{T^*X}}_\Lambda({\cal M}^\bullet)))\smile\pi^*Td(TX) \] and, in the present formulation is valid for any cycle in $CC^{per}_\bullet ({\cal R}{\cal E}_X)$ (because the image of the canonical trace density map is, in essence, spanned by Chern characters). The above formula is equivalent to the commutativity of the diagram \[ \begin{CD} CC^{per}_\bullet({\cal R}{\cal E}_X) @>{\sigma}>> CC^{per}_\bullet({\cal O}_{T^*X}) \\ @V{\iota}VV @VV{\tilde\mu_{{\cal O}}\smile\pi^*Td(TX)}V \\ CC^{per}_\bullet({\cal E}_X)[t^{-1},t] @>{\tilde\mu_{{\cal E}}}>> {\Bbb C}_{T^*X}[t^{-1},t][u^{-1},u]] \end{CD} \] in the derived category of sheaves on $T^*X$. \subsubsection{} The non-triviality of the issue of defining the canonical trace density map has to do with the fact that the canonical morphism ${\Bbb C}_{T^*X}[2d] @>>> {\cal E}_X\otimes^{{\Bbb L}}_{{\cal E}_X\otimes{\cal E}_X^{op}}{\cal E}_X$ furnished by the duality theory is not an isomorphism (in the derived category) and that can be traced to the fact that the canonical map $Gr^F_\bullet{\cal E}_X @>>> {\cal O}_{T^*X}$ is not an isomorphism. The trace density map is a retraction of the morphism ${\Bbb C}_{T^*X}[2d] @>>> {\cal E}_X\otimes^{{\Bbb L}}_{{\cal E}_X\otimes{\cal E}_X^{op}}{\cal E}_X$ and is arranged for by embedding ${\cal E}_X$ into a larger algebra. The choice more natural in the context of micro-local analysis is the algebra of micro-local operators ${\cal E}_X^{{\Bbb R}}$ which bears very little resemblance to ${\cal O}_{T^*X}$. The same may be achieved by embedding ${\cal R}{\cal E}_X$ into a formal one-parameter deformation (with $t$ being the parameter) ${\Bbb A}^t_{T^*X}$ of ${\cal O}_{T^*X}$ which is uniquely determined by the fact that it contains ${\cal R}{\cal E}_X$. The Riemann-Roch formula in its diagramatic reformulation, follows from the analogous statement about ${\Bbb A}^t_{T^*X}$: \[ \begin{CD} CC^{per}_\bullet({\Bbb A}^t_{T^*X}) @>{\sigma}>{t\mapsto 0}> CC^{per}_\bullet({\cal O}_{T^*X}) \\ @V{\iota}VV @VV{\tilde\mu_{{\cal O}}\smile\pi^*Td(TX)}V \\ CC^{per}_\bullet({\Bbb A}^t_{T^*X})[t^{-1}] @>{\tilde\mu^{{\Bbb A}}}>> {\Bbb C}_{T^*X}[t^{-1},t]][u^{-1},u]] \end{CD} \] commutes in the derived category of sheaves on $T^*X$. The latter fact is a particular case of a general Riemann-Roch type theorem conserning symplectic deformations of structure sheaves and whose proof relies on the techniques of formal geometry and invariant theory and constitutes the sequel to the present paper. \subsection{Acknowledgements} The authors would like to thank J.-L.Brylinski and one another for inspiring discussions. \section{Characteristic classes and trace maps}\label{section:outline} In this section we introduce the ingredients which go into the statement of the main result of this paper (Theorem \ref{thm:main}) and show how Conjecture \ref{conj:ch} follows from it. Proofs are postponed until later sections as indicated. \subsection{Characteristic classes of perfect complexes} The following theorem (Theorem \ref{thm-intro:char-cl}) summarizes the relevant aspects of the requisite characteristic classes for perfect complexes. The details of the construction will appear in a separate publication. Let $X$ be a topological space, $Z$ a closed subset of $X$. Let ${\cal A}$ denote a sheaf of algebras on $X$ such that there is a global section $1\in\Gamma(X;{\cal A})$ which restricts to $1_{{\cal A}_x}$. We denote by $K^i_Z({\cal A})$ the $i$-th $K$-group of the category of perfect complexes of ${\cal A}$-modules which are acyclic on the complement of $Z$ and refer the reader to Section \ref{section:CC} for other notations. \begin{thm}\label{thm-intro:char-cl} There exists the Chern character $ch^{{\cal A}}_{Z,i} : K^i_Z({\cal A})\to H^{-i}_Z(X;CC^-_\bullet({\cal A}))$ and the Euler class $\operatorname{Eu}^{{\cal A}}_{Z,i} : K^i_Z({\cal A})\to H^{-i}_Z(X;C_\bullet({\cal A}))$ natural in $X$, $Z$, and ${\cal A}$ such that \begin{itemize} \item the composition $K^i_Z({\cal A}) @>{ch^{{\cal A}}_{Z,i}}>> H^{-i}_Z(X;CC^-_\bullet({\cal A})) @>>> H^{-i}_Z(X;C_\bullet({\cal A}))$ coincides with $\operatorname{Eu}^{{\cal A}}_{Z,i}$; \item for a perfect complex ${\cal F}^\bullet$ of ${\cal A}$-modules supported on $Z$ the class $\operatorname{Eu}^{{\cal A}}_{Z,0}({\cal F}^\bullet)\in H^0_Z(X;C_\bullet({\cal A}))$ coincides with the compostion \begin{multline*} k @>{1\mapsto\operatorname{id}}>> \operatorname{\bold R}\underline{\operatorname{Hom}}^\bullet_{{\cal A}}({\cal F}^\bullet,{\cal F}^\bullet) @<{\overset{\sim}{=}}<< (\operatorname{\bold R}\underline{\operatorname{Hom}}^\bullet_{{\cal A}}({\cal F}^\bullet,{\cal A})\otimes_k{\cal F}^\bullet) \otimes^{{\Bbb L}}_{{\cal A}\otimes_k{\cal A}^{op}}{\cal A} \\ @>{ev\otimes\operatorname{id}}>> {\cal A}\otimes^{{\Bbb L}}_{{\cal A}\otimes_k{\cal A}^{op}}{\cal A}\ . \end{multline*} \end{itemize} \end{thm} In what follows we will only consider the components of the Euler class and the Chern character defined on $K^0$ and write $\operatorname{Eu}^{{\cal A}}_Z$ (respectively $ch^{{\cal A}}_Z$) instead of $\operatorname{Eu}^{{\cal A}}_{Z,0}$ (respectively $ch^{{\cal A}}_{Z,0}$). Nor shall we make notational distinction between $ch^{{\cal A}}_Z$ and its image under the natural map $H^\bullet_Z(X;CC^-_\bullet({\cal A}))\to H^\bullet_Z(X;CC^{per}_\bullet({\cal A}))$. \subsection{The microlocal Euler class} For the reader's convenience we give the definition of the microlocal Euler class. Suppose that $\Lambda$ is a closed subvariety of $T^*X$ and ${\cal N}^\bullet$ is a perfect complex of ${\cal E}^{{\Bbb R}}_X$-modules acyclic on the complement of $\Lambda$. Consider the morphism in the derived category (the Lefshetz trace map) given by the composition \begin{eqnarray*} \operatorname{\bold R}\underline{\operatorname{Hom}}_{{\cal E}^{{\Bbb R}}_X}({\cal N}^\bullet, {\cal N}^\bullet) & \overset{\sim}{=} & {\cal N}^\bullet\underline{\boxtimes}\underline{D}{\cal N}^\bullet \otimes^{{\Bbb L}}_{{\cal E}^{{\Bbb R}}_{X\times X}}{\cal C}^{{\Bbb R}}_{\Delta\vert X\times X} \\ & @<{\overset{\sim}{=}}<< & \operatorname{\bold R}\Gamma_\Lambda\left( {\cal N}^\bullet\underline{\boxtimes}\underline{D}{\cal N}^\bullet \otimes^{{\Bbb L}}_{{\cal E}^{{\Bbb R}}_{X\times X}}{\cal C}^{{\Bbb R}}_{\Delta\vert X\times X} \right) \\ & @>>> & \operatorname{\bold R}\Gamma_\Lambda\left({\cal C}^{(2d){\Bbb R}}_{\Delta\vert X\times X} \otimes^{{\Bbb L}}_{{\cal E}^{{\Bbb R}}_{X\times X}}{\cal C}^{{\Bbb R}}_{\Delta\vert X\times X} \right) \\ & \overset{\sim}{=} & \operatorname{\bold R}\Gamma_\Lambda\left({\Bbb C}_{T^*X}[2d]\right) \end{eqnarray*} where \[ \underline{D}{\cal N}^\bullet\overset{def}{=} \operatorname{\bold R}\underline{\operatorname{Hom}}_{{\cal E}^{{\Bbb R}}_X}({\cal N}^\bullet,{\cal E}^{{\Bbb R}}_X\otimes_{\pi^{-1}{\cal O}_X} \pi^{-1}\Omega^d_X[d])\ , \] $\pi : T^*X @>>> X$ is the canonical projection, and the last isomorphism is as in ... . The microlocal Euler class $\mu\operatorname{eu}({\cal N}^\bullet)$ is defined as the Lefshetz trace of the identity, i.e. as the composite \[ \mu\operatorname{eu}({\cal N}^\bullet) : {\Bbb C}_{T^*X} @>{1\mapsto\operatorname{id}}>> \operatorname{\bold R}\underline{\operatorname{Hom}}_{{\cal E}^{{\Bbb R}}_X}({\cal N}^\bullet, {\cal N}^\bullet) @>>> \operatorname{\bold R}\Gamma_\Lambda\left({\Bbb C}_{T^*X}[2d]\right)\ . \] For a perfect complex ${\cal N}^\bullet$ of ${\cal E}_X$-modules the microlocal Euler class is defined by \[ \mu\operatorname{eu}({\cal N}^\bullet)\overset{def}{=}\mu\operatorname{eu}({\cal N}^\bullet \otimes_{{\cal E}_X}{\cal E}^{{\Bbb R}}_X)\ . \] Similarly, the microlocal Euler class of a complex ${\cal N}^\bullet$ of ${\cal D}_X$-modules with $\operatorname{char}{\cal N}^\bullet\subseteq\Lambda$ is defined as $\mu\operatorname{eu}({\cal N}^\bullet\otimes_{\pi^{-1}{\cal D}_X}{\cal E}^{{\Bbb R}}_X)$. The microlocal Euler class gives rise to the map \[ \mu\operatorname{eu} : K^0_\Lambda({\cal E}_X) @>>> H^{2d}_\Lambda(T^*X;{\Bbb C})\ . \] \subsection{The canoncial trace density map} Let $\Delta\subset X\times X$ denote the diagonal. Let $a: T^*X @>>> T^*X$ denote the antipodal map ($(x,\xi)\mapsto (x,-\xi)$). There is a natural isomorphism of sheaves of algebras $a^{-1}{\cal E}^{({\Bbb R})}_X \overset{\sim}{=}\left({\cal E}^{({\Bbb R})}_X\right)^{op}$. Put $\left({\cal E}^{({\Bbb R})}_{X\times X}\right)^a = (\operatorname{id}\times a)^{-1} {\cal E}^{({\Bbb R})}_{X\times X}$. Then, there is a natural map ${\cal E}^{({\Bbb R})}_X\boxtimes\left({\cal E}^{({\Bbb R})}_X\right)^{op} @>>> \left({\cal E}^{({\Bbb R})}_{X\times X}\right)^a$, and the canonical structure of an $\left({\cal E}^{({\Bbb R})}_{X\times X}\right)^a$-module on ${\cal E}^{({\Bbb R})}_X$ which extends the canoncial structure of an ${\cal E}^{({\Bbb R})}_X\boxtimes \left({\cal E}^{({\Bbb R})}_X\right)^{op}$-module. By a theorem of Kashiwara the canonical map \[ {\Bbb C}_{T^*X} @>>> \operatorname{\bold R}\underline{\operatorname{Hom}}_{{\cal E}_{X\times X}^{{\Bbb R}}}( {\cal C}_{\Delta\vert X\times X}^{{\Bbb R}}, {\cal C}_{\Delta\vert X\times X}^{{\Bbb R}}) \] is an isomorphism in the derived category. There are canonical isomorphisms \begin{eqnarray*} {\Bbb C}_{T^*X}[2d] & \overset{\sim}{=} & \operatorname{\bold R}\underline{\operatorname{Hom}}_{{\cal E}_{X\times X}^{{\Bbb R}}}( {\cal C}_{\Delta\vert X\times X}^{{\Bbb R}}, {\cal C}_{\Delta\vert X\times X}^{{\Bbb R}})[2d] \\ & \overset{\sim}{=} & \operatorname{\bold R}\underline{\operatorname{Hom}}_{{\cal E}_{X\times X}^{{\Bbb R}}} ({\cal C}_{\Delta\vert X\times X}^{{\Bbb R}},{\cal E}_{X\times X}^{{\Bbb R}})[2d] \otimes^{{\Bbb L}}_{{\cal E}_{X\times X}^{{\Bbb R}}} {\cal C}_{\Delta\vert X\times X}^{{\Bbb R}} \\ & \overset{\sim}{=} & {\cal C}^{(2d){\Bbb R}}_{\Delta\vert X\times X}\otimes^{{\Bbb L}}_{{\cal E}_{X\times X}^{{\Bbb R}}} {\cal C}_{\Delta\vert X\times X}^{{\Bbb R}} \\ & \overset{\sim}{=} & {\cal E}^{{\Bbb R}}_X\otimes^{{\Bbb L}}_{\left({\cal E}^{{\Bbb R}}_{X\times X}\right)^a} {\cal E}^{{\Bbb R}}_X\ . \end{eqnarray*} By abuse of notation we will denote by $\mu_{{\cal E}}$ and refer to as {\em the canonical trace density map} the composition \[ {\cal E}_X\otimes^{{\Bbb L}}_{\left({\cal E}_{X\times X}\right)^a}{\cal E}_X @>>> {\cal E}^{{\Bbb R}}_X\otimes^{{\Bbb L}}_{\left({\cal E}^{{\Bbb R}}_{X\times X}\right)^a} {\cal E}^{{\Bbb R}}_X @>{\overset{\sim}{=}}>> {\Bbb C}_{T^*X}[2d] \] and the composition of the latter with the canonical map \[ {\cal E}_X\otimes^{{\Bbb L}}_{{\cal E}_X\boxtimes{\cal E}_X^{op}} {\cal E}_X @>>> {\cal E}_X\otimes^{{\Bbb L}}_{\left({\cal E}_{X\times X}\right)^a}{\cal E}_X\ . \] \subsection{Comparison of the Euler classes} The Euler class of Theorem \ref{thm-intro:char-cl} combined with the canonical trace density map gives rise to the map \begin{equation}\label{map:tr-Eu} K^0_\Lambda({\cal E}_X) @>{\operatorname{Eu}^{{\cal E}_X}}>> H^0_\Lambda(T^*X;{\cal E}_X\otimes^{{\Bbb L}}_{{\cal E}_X\boxtimes{\cal E}_X^{op}}{\cal E}_X) @>{\mu_{{\cal E}}}>> H^{2d}_\Lambda(T^*X; {\Bbb C})\ . \end{equation} \begin{prop}\label{prop:mueu-is-Eu} The composition \eqref{map:tr-Eu} coincides with the micro-local Euler class. In other words, if ${\cal N}^\bullet$ is a perfect complex of ${\cal E}_X$-modules and $\Lambda$ is a closed subvariety of $T^*X$ containing $\operatorname{Supp}{\cal N}^\bullet$, then $\mu\operatorname{eu}({\cal N}^\bullet) = \mu_{{\cal E}}(\operatorname{Eu}^{{\cal E}_X}_\Lambda({\cal N}^\bullet))$. \end{prop} \begin{pf} The Lefshetz trace map \[ \operatorname{\bold R}\underline{\operatorname{Hom}}_{{\cal E}_X}({\cal N}^\bullet, {\cal N}^\bullet) @>>> {\cal C}^{(2d)}_{\Delta\vert X\times X}\otimes^{{\Bbb L}}_{{\cal E}_{X\times X}} {\cal C}_{\Delta\vert X\times X} \] is compatible with the Lefshetz trace map \[ \operatorname{\bold R}\underline{\operatorname{Hom}}_{{\cal E}_X}({\cal N}^\bullet, {\cal N}^\bullet) @>>> {\cal E}_X\otimes^{{\Bbb L}}_{{\cal E}_X\otimes{\cal E}_X^{op}}{\cal E}_X @>>> {\cal E}_X\otimes^{{\Bbb L}}_{\left({\cal E}_{X\times X}\right)^a}{\cal E}_X \] and the isomorphisms \[ {\cal C}^{(2d){\Bbb R}}_{\Delta\vert X\times X}\otimes^{{\Bbb L}}_{{\cal E}^{{\Bbb R}}_{X\times X}} {\cal C}^{{\Bbb R}}_{\Delta\vert X\times X}\overset{\sim}{=} {\cal E}^{{\Bbb R}}_X\otimes^{{\Bbb L}}_{\left({\cal E}^{{\Bbb R}}_{X\times X}\right)^a}{\cal E}^{{\Bbb R}}_X \overset{\sim}{=} {\Bbb C}_{T^*X}[2d]\ . \] \end{pf} \subsection{Complexes of Hochschild chains} With the view onto the passage to cyclic homology we describe the canonical trace density map in terms of the standard complexes of Hochschild chains. The object ${\cal E}_X\otimes^{{\Bbb L}}_{{\cal E}_X\boxtimes{\cal E}_X^{op}}{\cal E}_X$, regarded as an object of the derived category of sheaves on $T^*X$, is represented by the standard complex of Hochschild chains $C_\bullet({\cal E}_X)$ (see \ref{subsection:C-CC}). Put \begin{eqnarray*} \left({\cal E}^{({\Bbb R})}_X\right)^{\frak{e}} &= & \delta_{T^*X}^{-1}(\operatorname{id}\times a)^{-1} {\cal E}^{({\Bbb R})}_{X\times X} \\ \widehat{C}_p({\cal E}^{({\Bbb R})}_X) & = & \delta_{T^*X}^{-1} {\cal E}^{({\Bbb R})}_{X^{\times p+1}}\ , \end{eqnarray*} The Hochschild differential extends to the map $b: \widehat{C}_p({\cal E}^{({\Bbb R})}_X) @>>> \widehat{C}_{p-1}({\cal E}^{({\Bbb R})}_X)$. The complex $\widehat{C}_\bullet({\cal E}^{({\Bbb R})}_X)$ represents ${\cal E}^{({\Bbb R})}_X\otimes^{{\Bbb L}}_{\left({\cal E}^{({\Bbb R})}_X\right)^{\frak{e}}} {\cal E}^{({\Bbb R})}_X$ in the derived category of sheaves on $T^*X$. There is a canoncial map $C_\bullet({\cal E}^{({\Bbb R})}_X) @>>> \widehat C_\bullet({\cal E}^{({\Bbb R})}_X)$. Let $\Phi^{{\cal E}}$ denote the global section of $H^{-2d}\widehat{C}_\bullet({\cal E}^{{\Bbb R}}_X)$ which corresponds to the global section $1$ under the isomorphism induced by $\mu_{{\cal E}}$. \begin{lemma} Let $x_1,\ldots,x_d$ be a local coordinate system on $X$ and let $\partial_i =\displaystyle\frac{\partial}{\partial x_i}$. Then, the locally defined Hochschild chain $Alt(1\otimes x_1\otimes\cdots\otimes x_d \otimes\partial_1\otimes\cdots\otimes\partial_d)\in C_{2d}({\cal E}_X)$ is a cycle which represents (the restriction to the domain of the coordinate system of) $\Phi^{{\cal E}}\in\Gamma (T^*X;H^{-2d}C_\bullet({\cal E}^{{\Bbb R}}_X))$. \end{lemma} \subsection{Cyclic homology of microdifferential operators}\label{subsec:CCE} Our immediate goal is to recast the calculation of the micro-local Euler class in terms of the negative (and, ultimately, the periodic) cyclic homology. The passage to cyclic homology is motivated by two developments. On the one hand, the Euler class (which takes values in Hochschild homology) factors canonically through the Chern character which takes values in negative cyclic homology; thus, the Chern character should be a (significant) refinement of the Euler class. On the other hand, the periodic cyclic homology is invariant under formal deformations of algebras and this feature will, ultimately, allow us to evaluate the ``non-commutative'' Chern character explicitly in ``commutative'' terms. The cyclic differential $B$ extends to the map $B: \widehat{C}_p({\cal E}^{({\Bbb R})}_X) @>>> \widehat{C}_{p+1}({\cal E}^{({\Bbb R})}_X)$ and the cyclic complexes $\widehat{CC}^-_\bullet({\cal E}^{({\Bbb R})})$ and $\widehat{CC}^{per}_\bullet({\cal E}^{({\Bbb R})}_X)$ are defined as in \ref{subsection:C-CC} starting from $\widehat C_\bullet({\cal E}^{({\Bbb R})}_X)$. The analogs of \eqref{inclusions} and \eqref{ses:CC-C} hold and there are canonical maps ${CC}^-_\bullet({\cal E}^{({\Bbb R})}) @>>> \widehat{CC}^-_\bullet({\cal E}^{({\Bbb R})})$ and ${CC}^{per}_\bullet({\cal E}^{({\Bbb R})}_X) @>>> \widehat{CC}^{per}_\bullet({\cal E}^{({\Bbb R})}_X)$. The complexes $(\widehat{CC}^-_\bullet({\cal E}^{({\Bbb R})}), b+uB)$ and $(\widehat C_\bullet({\cal E}^{({\Bbb R})}_X)[[u]], b)$ have canonical filtrations by powers of $u$ which will be denoted by $F_\bullet$. Note that the associated graded complexes $Gr^F_\bullet\widehat{CC}^-_\bullet({\cal E}^{({\Bbb R})})$ and $Gr^F_\bullet\widehat C_\bullet({\cal E}^{({\Bbb R})}_X)[[u]]$ are identical. \begin{lemma}\label{lemma:CC-E} There exists an isomorphism of filtered complexes \[ (\widehat{CC}^{-}_\bullet({\cal E}^{{\Bbb R}}_X),b+uB, F_\bullet)\overset{\sim}{=} (\widehat C_\bullet({\cal E}^{{\Bbb R}}_X)[[u]],b,F_\bullet) \] which induces the identity map on the associated graded complexes and any two such are (filtered) homotopic. \end{lemma} \begin{pf} Consider the spectral sequence of the filtered complex \\ $(\operatorname{Hom}^\bullet(\widehat{CC}^{-}_\bullet({\cal E}^{{\Bbb R}}_X), \widehat C_\bullet({\cal E}^{{\Bbb R}}_X)[[u]]),F_\bullet)$ and use the isomorphism (in the derived category) $\widehat C_\bullet({\cal E}^{{\Bbb R}}_X)\overset{\sim}{=}{\Bbb C}_{T^*X}[2d]$. \end{pf} \begin{cor}\label{cor:CC-E} There exists an isomorphism of filtered complexes \[ (\widehat{CC}^{per}_\bullet({\cal E}^{{\Bbb R}}_X),b+uB,F_\bullet) \overset{\sim}{=} (\widehat C_\bullet({\cal E}^{{\Bbb R}}_X)[u^{-1},u]],b,F_\bullet) \] which induces the identity map on the associated graded objects and any two such are homotopic. \end{cor} \subsection{The trace density on the periodic cyclic complex} \label{subsec:trace-density} The isomorphism (in the derived category) $\widehat{C}_\bullet({\cal E}^{{\Bbb R}}_X) \overset{\sim}{=}{\Bbb C}_{T^*X}[2d]$ is represented, using the standard truncation functors, by the following diagram of quasi-isomorphisms of complexes: \begin{multline}\label{map:muE-trunk} \widehat{C}_\bullet({\cal E}^{{\Bbb R}}_X) @<<< \tau^{\leq -2d}\widehat{C}_\bullet({\cal E}^{{\Bbb R}}_X) @>>> \\ @>>> \tau^{\geq -2d}\tau^{\leq -2d}\widehat{C}_\bullet({\cal E}^{{\Bbb R}}_X) \overset{\sim}{=} {\Bbb C}_{T^*X}[2d]\ . \end{multline} The diagram \eqref{map:muE-trunk} gives rise to the diagram of complexes of ${\Bbb C}[u^{-1},u]]$-modules \begin{multline*} \widehat{C}_\bullet({\cal E}^{{\Bbb R}}_X)[u^{-1},u]] @<<< \left(\tau^{\leq -2d}\widehat{C}_\bullet({\cal E}^{{\Bbb R}}_X)\right)[u^{-1},u]] @>>> \\ @>>> \left(\tau^{\geq -2d}\tau^{\leq -2d}\widehat{C}_\bullet({\cal E}^{{\Bbb R}}_X)\right) [u^{-1},u]] \overset{\sim}{=} {\Bbb C}_{T^*X}[2d][u^{-1},u]] \end{multline*} which represents a morphism \begin{equation}\label{map:muE-trunk-u} \widehat{C}_\bullet({\cal E}^{{\Bbb R}}_X)[u^{-1},u]] @>>> {\Bbb C}_{T^*X}[2d][u^{-1},u]] \end{equation} in the derived category. We will denote by $\tilde\mu_{{\cal E}}$ and refer to as the canonical trace density map the composition \begin{multline*} {CC}^{per}_\bullet({\cal E}_X) @>>> {CC}^{per}_\bullet({\cal E}^{{\Bbb R}}_X) @>>> \\ \widehat{CC}^{per}_\bullet({\cal E}^{{\Bbb R}}_X) @>>> (\widehat C_\bullet({\cal E}^{{\Bbb R}}_X)[u^{-1},u]] @>>> {\Bbb C}_{T^*X}[2d][u^{-1},u]] \end{multline*} where the last two maps are furnished, respectively, by Corollary \ref{cor:CC-E} and \eqref{map:muE-trunk-u}. Note that, by construction, the map $\tilde\mu_{{\cal E}}$ induces the map $\mu_{{\cal E}}$ on the associated (to the filtrations by powers of $u$) graded objects. \begin{lemma}\label{cor:Eu-comp-ch} For a perfect complex ${\cal N}^\bullet$ of ${\cal E}_X$-modules and $\Lambda$ a closed subvariety of $T^*X$ containing $\operatorname{Supp}{\cal N}^\bullet$, $\mu_{{\cal E}}(\operatorname{Eu}^{{\cal E}_X}_\Lambda({\cal N}^\bullet)) = \left[\tilde\mu_{{\cal E}} (ch^{{\cal E}_X}_\Lambda({\cal N}^\bullet))\right]^{2d}$. \end{lemma} \begin{pf} Consider the commutative diagram \[ \begin{CD} H^0_\Lambda(T^*X;CC^-_\bullet({\cal E}_X)) @>{\tilde\mu_{{\cal E}}}>> H^0_\Lambda(T^*X;{\Bbb C}_{T^*X}[2d][[u]]) \\ @VVV @VV{u\mapsto 0}V \\ H^0_\Lambda(T^*X;C_\bullet({\cal E}_X)) @>{\mu_{{\cal E}}}>> H^{2d}_\Lambda(T^*X;{\Bbb C}) \end{CD} \] \end{pf} \section{The Schapira--Schneiders conjecture} In this section we arrive at the statement of the main result of this paper which implies the Schapira--Schneiders conjecture and explain how it follows from the the analogous theorem for formal symplectic deformations. The proof of the latter is presented in the sequel to this paper. \subsection{The Rees construction} The sheaf of algebras ${\cal E}_X$ carries a natural filtration $F_\bullet{\cal E}_X$ by order. The Rees ring ${\cal R}{\cal E}_X$ is the (graded) algebra flat over ${\Bbb C}[\hbar]$ defined by \[ {\cal R}{\cal E}_X\overset{def}{=}\bigoplus_p F_p{\cal E}_X\cdot t^p\ . \] The Rees module ${\cal R}{\cal N}$ (which is a graded module over ${\cal R}{\cal E}_X$) associated to a filtered module $({\cal N},F_\bullet)$ is defined similarly. Thus defined, the Rees construction extends to an exact functor ${\cal R}$ form the (exact) category of filtered ${\cal E}_X$-modules to the (Abelian) category of graded ${\cal R}{\cal E}_X$-modules. The functor ${\cal R}$ is an embedding with the essential image consisting of the subcategory of $t$-torsion free modules. The induced functor between the respective derived categories is an equivalence. The filtered complex $({\cal N}^\bullet,F_\bullet)$ is {\em good} if and only if the complex of ${\cal R}{\cal E}_X$-modules ${\cal R}{\cal N}^\bullet$ is perfect. The Rees construction restricts to an equivalence between the derived categories of good filtered complexes of ${\cal E}_X$-modules and perfect complexes of graded ${\cal R}{\cal E}_X$-modules. In particular it induces an isomorphism of respective Grothendieck groups. Setting $t = 0$ one obtains the map \[ \sigma : {\cal R}{\cal E}_X @>>> Gr^F_\bullet{\cal E}_X @>>> {\cal O}_{T^*X}\ . \] Note that \[ \sigma({\cal N})\overset{def}{=}Gr^F_\bullet{\cal N}\otimes_{Gr^F_\bullet{\cal E}_X} {\cal O}_{T^*X} = {\cal R}{\cal N}\otimes_{{\cal R}{\cal E}_X}{\cal O}_{T^*X}\ . \] Consider a filtered complex $({\cal N}^\bullet,F_\bullet)$ with ${\cal R}{\cal N}^\bullet$ perfect over ${\cal R}{\cal E}_X$ with support contained in $\Lambda\subset T^*X$. Then, \[ ch^{{\cal O}_{T^*X}}_\Lambda(\sigma({\cal N}^\bullet)) = \sigma\left(ch^{{\cal R}{\cal E}_X}_\Lambda({\cal R}{\cal N}^\bullet)\right)\ . \] The canonical homomorphism of algebras \[ \iota :{\cal R}{\cal E}_X @>>> {\cal R}{\cal E}_X[t^{-1}] @>{\overset{\sim}{=}}>> {\cal E}_X[t^{-1},t] \] induces the natural isomorphism ${\cal R}{\cal N}^\bullet[t^{-1}]\overset{\sim}{=}{\cal N}^\bullet \otimes_{{\cal E}_X}{\cal E}_X[t^{-1},t]$. Therefore, \[ ch^{{\cal E}_X}_\Lambda({\cal N}^\bullet) = \iota\left(ch^{{\cal R}{\cal E}_X}_\Lambda({\cal R}{\cal N}^\bullet)\right) \ . \] In particular the right hand side is independent of $t$. \subsection{Cyclic homology of the structure sheaf} We recall, briefly, the well known calculation of the cyclic homology of the structure sheaf ${\cal O}_M$ of a complex manifold $M$ of Hochschild-Kostant-Rosenberg-Connes. Let $\widehat C_p({\cal O}_M)$ denote the completion of ${\cal O}_{M^{\times p+1}}$ along the diagonal. The Hochschild differential extends to the map $b: \widehat C_p({\cal O}_M) @>>> \widehat C_{p-1}({\cal O}_M)$ and the resulting complex $\widehat C_\bullet({\cal O}_M)$ represents ${\cal O}_M\otimes^{{\Bbb L}}_{{\cal O}_{M\times M}}{\cal O}_M$ in the derived category. There is a canonical map $C_\bullet({\cal O}_M) @>>> \widehat C_\bullet({\cal O}_M)$. The assignment $f_0\otimes\dots\otimes f_p\mapsto\displaystyle\frac1{p!} f_0df_1\wedge\dots df_p$ extends to a of complexes \[ \mu_{{\cal O}} : \widehat C_\bullet({\cal O}_M) @>>> \bigoplus_p\Omega^p_M[p] \] which, according to a theorem of Hochschild-Kostant-Rosenberg, is a quasi-isomorphism. The cyclic differential $B$ extends to the map $B : \widehat C_p({\cal O}_M) @>>> \widehat C_{p+1}({\cal O}_M)$ so that the square \[ \begin{CD} \widehat C_p({\cal O}_M) @>>> \Omega^p_M \\ @V{B}VV @VV{d}V \\ \widehat C_{p+1}({\cal O}_M) @>>> \Omega^{p+1}_M \end{CD} \] commutes. The cyclic complex $\widehat{CC}^{per}_\bullet({\cal O}_X)$ is defined as in \ref{subsection:C-CC}. Thus, the Hochschild-Kostant-Rosenberg map induces the map of complexes \[ \tilde\mu_{{\cal O}} : \widehat{CC}^{per}_\bullet({\cal O}_X) @>>> \Omega^\bullet_M[u^{-1},u]] \] which, according to A.~Connes, is a quasi-isomorphism. The natural inclusion ${\Bbb C}_M[u^{-1},u]] @>>> CC^{per}({\Bbb C}_M)$ is easily seen to be a quasi-isomorphism and the composition \begin{multline*} {\Bbb C}_M[u^{-1},u]] @>>> CC^{per}({\Bbb C}_M) @>>> \\ CC^{per}_\bullet({\cal O}_X) @>>> \widehat{CC}^{per}_\bullet({\cal O}_X) @>>> \Omega^\bullet_M[u^{-1},u]] \end{multline*} coincides with the canoncal map ${\Bbb C}_M @>>> \Omega^\bullet_M$. By abuse of notation we will denote by $\tilde\mu_{{\cal O}}$ the morphism in the derived category represented by \begin{equation}\label{map:muO} \tilde\mu_{{\cal O}} : CC^{per}_\bullet({\cal O}_X) @>>> \Omega^\bullet_M[u^{-1},u]] @<{\overset{\sim}{=}}<< {\Bbb C}_M[u^{-1},u]] \end{equation} which, by the discussion above, is a one-sided inverse to the map induced by the inclusion ${\Bbb C}_M @>>> {\cal O}_M$. The Chern character with supports (as in \eqref{map:ch-with-supp}) for perfect complexes of ${\cal O}_X$-modules supported on a closed subvariety $Z$ of a complex manifold $M$ is defined as the composition \[ K^0_Z({\cal O}_X) @>{ch^{{\cal O}_X}_Z}>> H^0_Z(X;CC^{per}_\bullet({\cal O}_X)) @>{\tilde\mu_{{\cal O}}}>> \bigoplus_{p}H_Z^{2p}(X;{\Bbb C})\ . \] \subsection{The Riemann-Roch formula} In view of Lemma \ref{cor:Eu-comp-ch} and Proposition \ref{prop:mueu-is-Eu}, to prove Conjecture \ref{conj:ch} it is sufficient to show that \begin{multline}\label{formula:RR-for-E} \tilde\mu_{{\cal E}}\left(\iota\left(ch^{{\cal R}{\cal E}_X}_\Lambda({\cal R}{\cal N}^\bullet)\right) \right) = \tilde\mu_{{\cal O}}\left(\sigma\left(ch^{{\cal R}{\cal E}_X}_\Lambda({\cal R}{\cal N}^\bullet)\right) \right)\smile\pi^*Td(TX)\ . \end{multline} The above reformulation of Conjecture \ref{conj:ch} in terms of the Chern character with values in periodic cyclic homology is crucial since, as it turns out, the formula \eqref{formula:RR-for-E} holds with $ch^{{\cal R}{\cal E}_X}_\Lambda({\cal R}{\cal N}^\bullet)$ replaced by an arbitrary cycle in the periodic cyclic complex. The latter fact constitutes the main result of the paper and is formulated in the following theorem. \begin{thm}\label{thm:mainE} The diagram in the derived category of sheaves on $T^*X$ \[ \begin{CD} CC^{per}_\bullet({\cal R}{\cal E}_X) @>{\sigma}>> CC^{per}_\bullet({\cal O}_{T^*X}) \\ @V{\iota}VV @VV{\tilde\mu_{{\cal O}}\smile \pi^*Td(TX)}V \\ CC^{per}_\bullet({\cal E}_X)[t^{-1},t] @>{\tilde\mu_{{\cal E}}}>> {\Bbb C}_{T^*X}[t^{-1},t] [u^{-1},u]] \end{CD} \] is commutative. \end{thm} \begin{pf} The statement follows from Theorem \ref{thm:main} applied to the case $M=T^*X$ and ${\Bbb A}^t_{T^*X}$ furnished by Proposition \ref{prop:DQcotan} (with $\theta = \pi^*c_1(TX)/2$), Corollary \ref{cor:EtoA}, and the equality $\widehat A(TM)\smile e^\theta = \pi^*Td(TX)$. \end{pf} \section{Deformation quantization} \subsection{Weyl quantization} Consider the vector space ${\Bbb C}^{2d}$ with coordinates $x_1,\ldots,x_d,\xi_1, \ldots,\xi_d$ as a symplectic manifold equipped with the standard symplectic form $dx_1\wedge d\xi_1\wedge\dots\wedge dx_d\wedge d\xi_d$. The sheaf of algebras ${\Bbb A}^t_{{\Bbb C}^{2d}}$ is defined as follows. For $U$ an open subset of ${\Bbb C}^{2d}$ the ${\Bbb C}[[t ]]$-module underlying ${\Bbb A}^t_{{\Bbb C}^{2d}}(U)$ is ${\cal O}_{{\Bbb C}^{2d}}(U)[[t]]$ and the product on ${\Bbb A}^t_{{\Bbb C}^{2d}}(U)$ is given by the standard Moyal--Weyl product \[ (f\ast g)(\underline x,\underline\xi) = \\ exp\left(\frac{t}{2}\sum_{i=1}^d\left( \frac{\partial\ }{\partial\xi_i}\frac{\partial\ }{\partial y_i}- \frac{\partial~}{\partial\eta_i}\frac{\partial~}{\partial x_i}\right)\right) f(\underline x,\underline\xi)g(\underline y,\underline\eta) \vert_{\overset{\underline x=\underline y}{\underline\xi =\underline\eta}} \] where $\underline x = (x_1,\ldots,x_d),\ \underline\xi = (\xi_1,\ldots,\xi_d), \ \underline y = (y_1,\ldots,y_d),\ \underline\eta = (\eta_1,\ldots,\eta_d)$. \subsection{Symplectic deformation quantization} Let $(M,\omega)$ denote a symplectic manifold of dimension $\dim_{\Bbb C} M = 2d$. For purposes of this paper a {\em (symplectic) deformation quantization} of $M$ is a formal one parameter deformation of the structure sheaf ${\cal O}_M$, i.e. a sheaf of algebras ${\Bbb A}^t_M$ flat over ${\Bbb C} [[t]]$ equipped with an isomorphism of algebras ${\Bbb A}^t_M\otimes_{{\Bbb C} [[t]]}{\Bbb C} @>>> {\cal O}_M$ and having the following properties: \begin{itemize} \item it is locally isomorphic to the standard deformation ${\Bbb A}^t_{{\Bbb C}^{2d}}$ of ${\Bbb C}^{2d}$ (see below); \item the Poisson bracket on ${\cal O}_M$ defined by $\displaystyle{ \lbrace f, g\rbrace = \frac{1}{t}[\tilde f,\tilde g ] + t\cdot {\Bbb A}^t_M}$ where $f$ and $g$ are two local sections of ${\cal O}_M$ and $\tilde f$, $\tilde g$ are their respective lifts ${\Bbb A}^t_M$, coincides with the one induced by the symplectic structure. \end{itemize} To a symplectic deformation quantization ${\Bbb A}^t_M$ one associates a characteristic class $\theta\in H^2(M;\displaystyle\frac{1}{t}{\Bbb C}[[t]])$ (see \cite{NT3} and \ref{subsection:Fedosoff} below) with the property that the coefficient of $\displaystyle\frac{1}{t}$ is the class of the symplectic form. In what follows we will use the canonical maps $\sigma :{\Bbb A}^t_M\to{\cal O}_M$ and $\iota:{\Bbb A}^t_M\to{\Bbb A}^t_M[t^{-1}]$. \subsection{Review of Fedosov connections}\label{subsection:Fedosoff} We recall briefly the definition of the characteristic class $\theta$ in terms of the Fedosov connections. Let $W = {\Bbb C}[[\hat{z}_1,\dots,\hat{z}_d,\hat{\xi}_1,\dots,\hat{\xi}_d,t]]$ equipped with the Weyl product denoted $\ast$ while the ``commutative'' product will not be indicated by any special notations. Thus, equipped with the Weyl product, $W$ is a deformation quantization of the formal completion of ${\Bbb C}^{2d}$ at the origin. The induced symplectic structure is the standard one. Let $\frak g$ denote the Lie algebra of continuous derivations of $W$. Let $\widetilde{\frak g}$ denote the Lie algebra $\displaystyle\frac1{t}W$ with the bracket $[\displaystyle\frac1{t}f,\displaystyle\frac1{t}g] = \displaystyle\frac1{t}(f\ast g-g\ast f)$. The exact sequence \[ 0 @>>> \frac1{t}{\Bbb C}[[t]] @>>> \widetilde{\frak g} @>{\frac1{t}ad}>> {\frak g} @>>> 0 \] is a central extension of Lie algebras. Let $\deg\hat{z}_i =\deg\hat{z}_i = 1$, $\deg t = 2$, let $\widetilde{\frak g}_k$ denote the subspace of homogeneous elements of degree $k$, and let ${\frak g}_k$ denote the image of $\widetilde{\frak g}_k$. Then $\sum_k\widetilde{\frak g}_k$ (respectively $\sum_k{\frak g}_k$) is a ${\Bbb Z}$-graded dense subalgebra of $\widetilde{\frak g}$ (respectively $\frak g$). The Lie algebra ${\frak sp}(2d,{\Bbb C})$ is naturally embedded in both $\widetilde{\frak g}_0$ and ${\frak g}_0$ as the span of the monomials $\displaystyle\frac1{t}\hat{z}_i\hat{z}_j$, $\displaystyle\frac1{t}\hat{z}_i\hat{\xi}_j$, and $\displaystyle\frac1{t}\hat{\xi}_i\hat{\xi}_j$. The (adjoint) action of ${\frak sp}(2d,{\Bbb C})$ on $\widetilde{\frak g}$ and ${\frak g}$ integrates to an action of the group $Sp(2d,{\Bbb C})$. Suppose that $M$ is a complex manifold of dimension $\dim_{{\Bbb C}}M = 2d$, $\omega\in\Gamma(M;\Omega^2_M)$ a symplectic form on $M$. Let ${\cal U} = \{U_\alpha\}_{\alpha\in I}$ be a cover of $M$ which trivializes $TM$. Set $U_{\alpha,\beta} = U_\alpha\cap U_\beta$ and let $g_{\alpha,\beta}: U_{\alpha,\beta} @>>> Sp(2d,{\Bbb C})$ denote the transition functions for $TM$. Let ${\Bbb W}_M$ denote the sheaf of ${\cal O}_M$ algebras which corresponds to the representation $W$ of $Sp(2d,{\Bbb C})$ (``the sheaf of holomorphic section of the associated bundle''). A Fedosov connection is a flat $\frak g$-connection $\nabla^\Phi$ on ${\Bbb W}_M\otimes_{{\cal O}_M}{\cal C}^\infty_M$ of a particular kind which we proceed to describe. Locally (i.e. on a $U_\alpha\in{\cal U}$) a Fedosov connection is given by $\nabla^\Phi = d + A_\alpha$, where $A_\alpha = \sum_{k\geq -1}A^{(k)}_\alpha$ with $A^{(k)}_\alpha\in {\cal A}^1_M(U_\alpha)\otimes{\frak g}_k$. Moreover, in Darboux coordinates $z_1,\dots,z_d,\xi_1\dots\xi_d$, $A^{(-1)}_\alpha = \displaystyle\frac1{t}\sum_i dz_i\otimes\hat{\xi}_i - d\xi_i\otimes\hat{z}_i$, and $A^{(0)}_\alpha\in{\cal A}^{1,0}_M(U_\alpha)\otimes{\frak sp}(2d,{\Bbb C})$ so that $\nabla^{(0)} = d+A^{(0)}_\alpha$ is a torsion free connection. A Fedosov connection $\nabla^\Phi$ admits a lifting to a $\widetilde{\frak g}$-connection $\widetilde\nabla^\Phi$ (which is not flat) with curvature $c(\widetilde\nabla^\Phi)\in\Gamma(M;{\cal A}^{2,cl}_M\widehat\otimes \displaystyle\frac1{t}{\Bbb C}[[t]])$ where ${\cal A}^{2,cl}_M$ is the sheaf of closed 2-forms. {\em The subsheaf of horizontal sections $\left({\Bbb W}_M\otimes_{{\cal O}_M}{\cal C}^\infty_M\right)^{\nabla^\Phi}$ is a deformation quantization of $(M,\omega)$ with characteristic class $\theta$ the cohomology class of $c(\widetilde\nabla^\Phi)$ which is independent of the choice of a lifting.} It is shown in \cite{NT3} that every deformation quantization arises in from a Fedosov connection and that the characteristic classes of isomorphic deformations coincide. \subsection{Homological properties of symplectic deformations} The sheaf of algebras ${\Bbb A}^t_M\widehat\boxtimes_{{\Bbb C}[[t]]} \left({\Bbb A}^t_M\right)^{op}$ is contained in a unique symplectic deformation quantization ${\Bbb A}^t_{M^{\times 2}}$ of the symplectic manifold $(M,\omega)\times(M,-\omega)$. Let \[ \left({\Bbb A}^t_M\right)^{\frak{e}} \overset{def}{=}\delta^{-1}{\Bbb A}^t_{M^{\times 2}}\ . \] Then there is a natural map of sheaves of algebras ${\Bbb A}^t_M\widehat\otimes_{{\Bbb C}[[t]]} \left({\Bbb A}^t_M\right)^{op} @>>> \left({\Bbb A}^t_M\right)^{\frak{e}}$. Let \[ \widehat{C}_p({\Bbb A}^t_M) \overset{def}{=} \underset{n}{\varprojlim} \left({\Bbb A}^t_M\right)^{\widehat\otimes_{{\Bbb C}[[t]]}p+1}/{\cal I}_{(p+1)}^n\ , \] where (the two-sided ideal) ${\cal I}_{(p+1)}$ is the kernel of the composition of the multiplication map with the symbol map $\left({\Bbb A}^t_M\right)^{\widehat\otimes_{{\Bbb C}[[t]]}p+1} @>>> {\Bbb A}^t_M @>>> {\cal O}_M$. The Hochschild differential \eqref{diffl:b} extends to this setting and gives rise to the complex $\widehat{C}_\bullet ({\Bbb A}^t_M)$ and the natural map of complexes $C_\bullet({\Bbb A}^t_M) @>>> \widehat{C}_\bullet({\Bbb A}^t_M)$. The complex $\widehat{C}_\bullet({\Bbb A}^t_M)$ represents ${\Bbb A}^t_M\otimes^{{\Bbb L}}_{\left({\Bbb A}^t_M\right)^{\frak{e}}} {\Bbb A}^t_M$ in the derived category. The cyclic complexes $\widehat{CC}^{-}_\bullet({\Bbb A}^t_M)$ and $\widehat{CC}^{per}_\bullet({\Bbb A}^t_M)$ are defined as in \ref{subsection:C-CC} starting from $\widehat{C}_p({\Bbb A}^t_M)$ The analogs of \eqref{inclusions} and \eqref{ses:CC-C} hold and there are natural maps of complexes $CC^{-}_\bullet({\Bbb A}^t_M) @>>> \widehat{CC}^{-}_\bullet({\Bbb A}^t_M)$ and $CC^{per}_\bullet({\Bbb A}^t_M) @>>> \widehat{CC}^{per}_\bullet({\Bbb A}^t_M)$. \begin{prop}\label{prop:CCA} Suppose that ${\Bbb A}^t_M$ is a symplectic deformation quantization of $M$. \begin{enumerate} \item Let $x_1,\ldots,x_d,\xi_1,\ldots,\xi_d$ be a local Darboux coordinate system on $M$. Then the local section $\Phi^{{\Bbb A}}$ of $H^{-2d}\widehat{C}_\bullet({\Bbb A}^t_M)[t^{-1}]$ represented by $Alt(1\otimes x_1\otimes\cdots\otimes x_d\otimes \displaystyle\frac{\xi_1}{t}\otimes\cdots\otimes\displaystyle\frac{\xi_d} {t})$ is independent of the choice of the Darboux coordinates. \item The section $\Phi^{{\Bbb A}}$ generates $H^{-2d}\widehat{C}_\bullet({\Bbb A}^t_M)[t^{-1}]$ as a ${\Bbb C}_M[t^{-1},t]]$-module and $H^{p}\widehat{C}_\bullet({\Bbb A}^t_M)[t^{-1}]=0$ for $p\neq -2d$, thus and there is a unique isomorphism \begin{equation}\label{map:muA} \mu_{{\Bbb A}}:\widehat{C}_\bullet({\Bbb A}^t_M)[t^{-1}] @>{\overset{\sim}{=}}>> {\Bbb C}_M[t^{-1},t]][2d] \end{equation} in the derived category of sheaves which maps $\Phi^{{\Bbb A}}$ to $1$. \item There exists an isomorphism of filtered complexes \[ (\widehat{CC}^{-}_\bullet({\Bbb A}^t_M)[t^{-1}],b+uB,F_\bullet) \overset{\sim}{=} (\widehat{C}_\bullet({\Bbb A}^t_M)[t^{-1}][[u]],b,F_\bullet) \] which induces the identity map on the associated graded objects and any two such are homotopic. \item There exists an isomorphism of filtered complexes \[ (\widehat{CC}^{per}_\bullet({\Bbb A}^t_M)[t^{-1}],b+uB,F_\bullet) \overset{\sim}{=} (\widehat{C}_\bullet({\Bbb A}^t_M)[t^{-1}][u^{-1},u]],b,F_\bullet) \] which induces the identity map on the associated graded objects and any two such are homotopic. \end{enumerate} \end{prop} \begin{pf} Since the statement is local it is sufficient to consider the stalk at the origin of the standard deformation of ${\Bbb C}^{2d}$. Note that the completion of ${\Bbb A}^t_{{{\Bbb C}^{2d}},0}$ with respect to the powers of the two-sided ideal which is the kernel of the composition ${\Bbb A}^t_{{{\Bbb C}^{2d}},0} @>{\sigma}>> {\cal O}_{{{\Bbb C}^{2d}},0} @>>> {\Bbb C}_0$ coincides with the Weyl algebra $W=W({\Bbb C}^{2d})$. The induced map $\widehat{C}_\bullet({\Bbb A}^t_{{\Bbb C}^{2d}})_0 @>>> \widehat{C}_\bullet(W)$ is a quasiisomorphism and the calculation of Hochschild homology follows from the standard results for the Weyl algebra. To prove the last two statements one argues as in Lemma \ref{lemma:CC-E} and Corollary \ref{cor:CC-E}. \end{pf} By abuse of notation we will denote by $\mu_{{\Bbb A}}$ the composition \[ C_\bullet({\Bbb A}^t_M) @>>> \widehat{C}_\bullet({\Bbb A}^t_M)[t^{-1}] @>{\text{\eqref{map:muA}}}>> {\Bbb C}_M[t^{-1},t]][2d] \] as well as \eqref{map:muA}. Proceeding as in \ref{subsec:trace-density} and using Proposition \ref{prop:CCA} one defines the morphism \[ \tilde\mu_{{\Bbb A}} : CC^{per}_\bullet({\Bbb A}^t_M)[t^{-1}] @>>> {\Bbb C}_M[t^{-1},t]][2d][u^{-1},u]]\ . \] \subsection{Deformation quantization of a cotangent bundle} For applications of our results to microlocal analysis we will require the particular deformation quantization of a cotangent bundle furnished by the following proposition. \begin{prop}\label{prop:DQcotan} There exists a deformation quantization ${\Bbb A}^t_{T^*X}$ of $T^*X$ and faithfully flat map ${\cal R}{\cal E}_X @>>> {\Bbb A}^t_{T^*X}$ of algebras over ${\Bbb C} [t ]$. The characteristic class $\theta$ of the deformation ${\Bbb A}^t_{T^*X}$ is equal to $\displaystyle\frac12\pi^*c_1(TX)$ (note that the symplectic form is exact in this case). \end{prop} \begin{pf} Suppose that $x_1,\ldots,x_d$ is a local coordinate system on $X$ and let $\xi_i$ denote the symbol of $\displaystyle\frac{\partial}{\partial x_i}$ viewed as a function on $T^*X$. Then $x_1,\ldots,x_d,\xi_1,\ldots,\xi_d$ form a Darboux coordinate system on $T^*X$. The map alluded to in Proposition \ref{prop:DQcotan} is given in local coordinates by $x_i\mapsto x_i$, $t\displaystyle\frac{\partial}{\partial x_i}\mapsto\xi_i$. The only non-trivial part is the calculation of the characteristic class $\theta$. This will be achieved by explicitly manufacturing a Fedosov connection $\nabla^\Phi$ on the sheaf of Weyl algebras ${\Bbb W}_{T^*X}$ whose sheaf of horizontal sections is ${\Bbb A}^t_{T^*X}$ and computing the Weyl curvature. Let ${\cal U} = \{ U_\alpha \}_{\alpha\in I}$ denote a cover of X which trivializes $TX$. Set $U_{\alpha,\beta} = U_\alpha\cap U_\beta$ and let $g_{\alpha,\beta} : U_{\alpha,\beta} @>>> GL(d,{\Bbb C})$ denote the transition functions for $TX$. Let ${\frak w}_d$ denote the Lie algebra of continuous derivations of the (topological) ${\Bbb C}$-algebra ${\Bbb C}[[\hat{z}_1,\dots,\hat{z}_d]]$. Let ${\frak w}_{d,k} = \{\sum_i P_i\displaystyle\frac{\partial\ }{\partial\hat{z}_i}\}$ where $P_i$ are homogeneous polynomials (in $\hat{z}_1,\dots,\hat{z}_d$) of degree $\deg P_i = k+1$. Then $\left[{\frak w}_{d,k},{\frak w}_{d,l}\right]\subseteq {\frak w}_{d,k+l}$ and $\sum_k{\frak w}_{d,k}$ is a dence subalgebra of ${\frak w}_d$. The Lie algebra ${\frak gl}(d,{\Bbb C})$ is identified with ${\frak w}_{d,0}$ by the map $\left(a_{i,j}\right)\mapsto\sum_{i,j}a_{i,j}\hat{z}_i\displaystyle \frac{\partial\ }{\partial\hat{z}_j}$. The (adjoint) action of ${\frak gl}(d,{\Bbb C})$ on ${\frak w}_d$ integrates to the action of $GL(d,{\Bbb C})$. The action of ${\frak w}_{d,0}$ on ${\frak w}_{d,-1}$ is identified with the standard representation of ${\frak gl}(d,{\Bbb C})$ on ${\Bbb C}^d$. We begin with a torsion free ${\frak gl}(d,{\Bbb C})$-connection $\nabla^{(0)}$ on $TX$ given on $U_\alpha$ by $d + A^{(0)}_\alpha$, where $A^{(0)}_\alpha\in{\cal A}^{1,0}_X\otimes {\frak gl}(d,{\Bbb C})$. The connection $\nabla^{(0)}$ extends to a flat ${\frak w}_d$-connection (called a Kazhdan connection) $\nabla^K$ which is, locally, of the form $d + A_\alpha$, where $A_\alpha\in{\cal A}^1_X\widehat\otimes{\frak w}_d$, $A_\alpha = \sum_k A^{(k)}$ with $A^{(-1)}_\alpha=-\sum_i dz_i\otimes\displaystyle\frac{\partial\ }{\partial\hat{z}_i}$ and $A^{(0)}_\alpha$ as above. Let $\widehat S = {\Bbb C}[[\hat{z}_1,\dots,\hat{z}_d]]$; this is a $({\frak w}_d,GL(d,{\Bbb C}))$-module. Let $\widehat{\cal S}$ denote the corresponding sheaf of ${\cal O}_X$-algebras (``holomorphic sections of the associated bundle''). The Kazhdan connection $\nabla^K$ is a connection on $\widehat{\cal S}\otimes_{{\cal O}_X}{\cal C}^\infty_X$, and the sheaf of horizontal sections $\left(\widehat{\cal S}\otimes_{{\cal O}_X}{\cal C}^\infty_X\right)^{\nabla^K}$ is a sheaf of algebras (since ${\frak w}_d$ acts by derivations) isomorphic to ${\cal O}_X$. Let $\widehat R = {\Bbb C}[[\hat{z}_1,\dots,\hat{z}_d,\hat{\xi}_1,\dots,\hat{\xi}_d,t]]$. We consider $\widehat R$ as an algebra with the Weyl product denoted $\ast$. The algebra $\widehat R$ is an algebra over $\widehat S$ in the obvious way and is isomorphic to the completion of the Rees algebra of the algebra of differential operators on $\widehat S$ by the assignment $t\displaystyle\frac{\partial\ }{\partial\hat{z}_i}\mapsto\hat{\xi}_i$. Let $Der(\widehat R)$ denote the Lie algebra of continuous, ${\Bbb C}[[t]]$-linear derivations of $\widehat R$. The exact sequence \[ 0 @>>> \frac1{t}{\Bbb C}[[t]] @>>> \frac1{t}\widehat R @>{\frac1{t}ad}>> Der(\widehat R) @>>> 0 \] is a central extension of Lie algebras. Just as in the case of the Weyl algebra in \ref{subsection:Fedosoff}, the Lie algebra ${\frak sp}(2d,{\Bbb C})$ is contained in $\frac1{t}\widehat R$ and $ Der(\widehat R)$ is the standard way. The inclusion of ${\frak gl}(d,{\Bbb C})$ into ${\frak sp}(2d,{\Bbb C})$ is given in terms of the standard embedding by $(a_{i,j})\mapsto\sum_{i,j}a_{i,j}\hat{z}_i\hat{\xi}_j$. The action of ${\frak gl}(d,{\Bbb C})$ integrates to an action of $GL(d,{\Bbb C})$. The identification with the differential operators (above) restricts to the map of Lie algebras $i : {\frak w}_d @>>>\displaystyle\frac1{t}\widehat R$ whose restriction to ${\frak gl}(d,{\Bbb C})$ is \begin{equation}\label{map:gl-to-R} \left(a_{i,j}\right)\mapsto \sum_{i,j}a_{i,j}\hat{z}_i\ast\displaystyle\frac{\hat{\xi}_j}{t} = \sum_{i,j}a_{i,j}\hat{z}_i\displaystyle\frac{\hat{\xi}_j}{t} - \frac12 tr\left(a_{i,j}\right)\ . \end{equation} Note that the first summand in the last expression is the standard embedding of ${\frak gl}(d,{\Bbb C})$. Since the second summand in the last expression is central the standard embedding and $i$ coincide in $Der(\widehat R)$. Let $\widehat{\cal R}$ denote the sheaf of ${\cal O}_X$-algebras on $X$ which corresponds to $\widehat R$. The Kazhdan connection $\nabla^K$ gives rise to a connection, still denoted $\nabla^K$, on $\widehat{\cal R}\otimes_{{\cal O}_X}{\cal C}^\infty_X$ via the composition $\displaystyle\frac1{t}ad\circ i$. The sheaf of horizontal sections $\left(\widehat{\cal R}\otimes_{{\cal O}_X}{\cal C}^\infty_X\right)^{\nabla^K}$ is isomorphic to the completion of the Rees algebra ${\cal R}{\cal D}_X$. There is a canonical isomorphism of algeras $\widehat{\cal R}\overset{\sim}{=}{\Bbb W}_{T^*X}\otimes_{{\cal O}_{T^*X}}{\cal O}_X$. Since the sheaves of algebras ${\Bbb W}_{T^*X}\otimes_{{\cal O}_{T^*X}}{\cal C}^\infty_{T^*X}$ and $\pi^*\widehat{\cal R}\otimes_{{\cal O}_{T^*X}}{\cal C}^\infty_{T^*X}$ are (non-canonically) isomorphic, it suffices to construct a ``Fedosov'' $Der(\widehat R)$-connection on the latter and this is what we will do explicitly, at least in the formal neighborhood of the zero section of $T^*X$. Let $z_1,\dots,z_d$ denote local coordinates on $X$, $\xi_i$ the symbol of $\displaystyle\frac{\partial\ }{\partial z_i}$, so that $z_1,\dots,z_d,\xi_1\dots\xi_d$ is a local Darboux coordinate system on $T^*X$. The automorphism $\Psi$ of sheaf of algebras $\pi^*{\cal R}$ given in local coordinates by the formula $\Psi=\exp(ad(-\sum_i\xi_i\displaystyle\frac{\hat{z}_i}{t}))$ is, in fact globally well defined on the formal neighborhood of the zero section of $T^*X$. The $\Psi$-conjugate of the pull-back $\pi^*\nabla^K$ of the Kazhdan connection is the desired Fedosov connection: $\nabla^\Phi = \left(\nabla^K\right)^\Psi$. Suppose $\tilde{\pi^*\nabla^K}$ is a lifting of $\pi^*\nabla^K$ to a $\displaystyle\frac1{t}\widehat R$-connection. Then $\left(\tilde{\pi^*\nabla^K}\right)^\Psi$ is a lifting of $\nabla^\Phi$, and $c(\tilde{\pi^*\nabla^K}) = c\left(\left(\tilde{\pi^*\nabla^K}\right)^\Psi\right)$ since the curvature is central and $\Psi$ is an ``inner'' automorphism. In order to find an explicit lifting $\tilde{\pi^*\nabla^K}$, consider the $\displaystyle\frac1{t}\widehat R$-valued forms $i(A_\alpha)$. Since the (${\frak w}_d$-valued) connection forms $A_\alpha$ satisfy \[ A_\alpha = dg^t_{\alpha,\beta}(g^t_{\alpha,\beta})^{-1} + g^t_{\alpha,\beta}A_\beta(g^t_{\alpha,\beta})^{-1}\ , \] the forms $iA_\alpha$ satisfy \begin{eqnarray*} i(A_\alpha) & = & i(dg^t_{\alpha,\beta}(g^t_{\alpha,\beta})^{-1}) + i(g^t_{\alpha,\beta}A_\beta(g^t_{\alpha,\beta})^{-1}) \\ & & = dg^t_{\alpha,\beta}(g^t_{\alpha,\beta})^{-1} - \frac12 tr(dg^t_{\alpha,\beta}(g^t_{\alpha,\beta})^{-1}) + g^t_{\alpha,\beta}i(A_\beta)(g^t_{\alpha,\beta})^{-1}\ , \end{eqnarray*} the last equality obtained using \eqref{map:gl-to-R}. Since $A^{(0)}_\alpha$ are (${\frak gl}(d,{\Bbb C})$-valued) connection forms and trace remains invariant under conjugation the identity \[ \frac12 tr(A^{(0)}_\alpha) = \frac12 tr(dg^t_{\alpha,\beta}(g^t_{\alpha,\beta})^{-1}) + \frac12 tr(A^{(0)}_\beta) \] holds. Adding the latter identity to the preceding one one easily concludes that the collection of $\displaystyle\frac1{t}\widehat R$-valued forms $\widetilde A_\alpha = i(A_\alpha) + \displaystyle\frac12 tr(A^{(0)}_\alpha)$ determine a connection $\tilde\nabla^K$ which lifts the Kazhdan connection. Since the Kazhdan connection is flat and $i$ is a morphism of Lie algebras, it follows that $c(\tilde\nabla) = \displaystyle\frac12 tr((\nabla^{(0)})^2)$ which represents $\displaystyle\frac12 c_1(TX)$ in de Rham cohomology. Hence, the cohomology class of the curvature of the connection $\tilde\pi^*\nabla^K$ is equal to $\displaystyle\frac12 \pi^*c_1(TX)$ and so is the characteristic class of the deformation. \end{pf} In what follows ${\Bbb A}^t_{T^*X}$ will always refer to the deformation furnished by Proposition \ref{prop:DQcotan}. \begin{lemma} The diagram \[ \begin{CD} C_\bullet({\cal R}{\cal E}_X)[t^{-1}] @>{\overset{\sim}{=}}>> C_\bullet({\cal E}_X)[t^{-1},t]\\ @VVV @VV{\mu_{{\cal E}}}V \\ C_\bullet({\Bbb A}^t_{T^*X})[t^{-1}] @>{\mu^t_{{\Bbb A}}}>> {\Bbb C}_{T^*X}[t^{-1},t]][2d] \end{CD} \] is commutative. \end{lemma} \begin{pf} The statement is equivalent to the commutativity of the diagram of respective cohomology sheaves in degree $-2d$. The latter fact follows from the local coordinate representation of $\Phi^{{\cal E}}$, $\Phi^{{\Bbb A}}$, the map ${\cal R}{\cal E}_X @>>> {\Bbb A}^t_{T^*X}$, and the definitions of $\mu_{{\cal E}}$ and $\mu^t_{{\Bbb A}}$. \end{pf} \begin{cor}\label{cor:EtoA} The diagram \[ \begin{CD} CC_\bullet({\cal R}{\cal E}_X)[t^{-1}] @>{\overset{\sim}{=}}>> CC_\bullet({\cal E}_X)[t^{-1},t] \\ @VVV @VV{\tilde\mu_{{\cal E}}}V \\ CC_\bullet({\Bbb A}^t_{T^*X})[t^{-1}] @>{\tilde\mu^t_{{\Bbb A}}}>> {\Bbb C}_{T^*X}[t^{-1},t]][2d][u^{-1},u]] \end{CD} \] is commutative. \end{cor} \subsection{The Riemann-Roch theorem for periodic cyclic cocycles} We are finally ready to state the Riemann-Roch type theorem for symplectic deformations. Consider a complex manifold $M$ and a symplectic deformation quantization ${\Bbb A}^t_M$ of $M$ with the characteristic class $\theta$. The deformation quantization ${\Bbb A}^t_M$ induces the symplectic structure on $M$. In particular the symplectic structure determines a reduction of the structure group of $TM$ to $Sp(2d,{\Bbb C})$. The characteristic classes of $Sp(2d,{\Bbb C})$-bundles are given by the invariant functions on the Lie algebra ${\frak sp}(2d,{\Bbb C})$. In particular the class $\widehat A(TM)$ corresponds to the function \[ {\frak sp}(2d,{\Bbb C})\ni X\mapsto \det\left(\frac{\operatorname{ad}(X)} {\exp(\operatorname{ad}(\frac{X}2)) - \exp(\operatorname{ad}(-\frac{X}2))}\right)\ . \] \begin{thm}\label{thm:main} The diagram \[ \begin{CD} CC^{per}_\bullet({\Bbb A}^t_M) @>{\sigma}>> CC^{per}_\bullet({\cal O}_M) \\ @V{\iota}VV @VV {\tilde\mu_{{\cal O}}\smile \widehat A(TM)\smile e^\theta}V \\ CC^{per}_\bullet({\Bbb A}^t_M)[t^{-1}] @>{\tilde\mu^t_{{\Bbb A}}}>> {\Bbb C}_M[t^{-1},t]][u^{-1},u]] \end{CD} \] is commutative. \end{thm} Theorem \ref{thm:main} may be deduced from \cite{NT1} and \cite{NT2}. A complete proof will appear in the sequel to the present paper. \section{Appendix: Review of Hochschild and cyclic homology} \label{section:CC} In this section we review the basic definitions of the Hochschild and the (negative, periodic) cyclic complex of an algebra, of a sheaf of algebras on a space as well as of the ``topological'' versions of the above. In particular we establish notational conventions with regard to Hochschild and cyclic complexes which are used in the body of the paper. In addition we review the results concerning the Hochschild and cyclic homologies of certain examples of (sheaves of) algebras which appear in this paper for the reader's convenience. \subsection{Hohschild and cyclic complexes of algebras} \label{subsection:C-CC} Let $k$ denote a commutative algebra over a field of characteristic zero and let $A$ be a flat $k$-algebra with $1_A\cdot k$ contained in the center, not necessarily commutative. Let $\overline A = A/k$, and let $C_p(A)\overset{def}{=}A\otimes_k\overline A^{\otimes_k p}$ and let \begin{eqnarray}\label{diffl:b} b : C_p(A) & @>>> & C_{p-1}(A) \\ \nonumber a_0\otimes\cdots\otimes a_p & \mapsto & (-1)^p a_pa_0\otimes\cdots\otimes a_{p-1} + \\ & & \sum_{i=0}^{p-1}(-1)^ia_0\otimes\cdots\otimes a_ia_{i+1}\otimes\cdots\otimes a_p\ .\nonumber \end{eqnarray} Then $b^2 = 0$ and the complex $(C_\bullet(A), b)$, called {\em the standard Hochschild complex of $A$} represents $A\otimes^{{\Bbb L}}_{A\otimes_kA^{op}}A$ in the derived category of $k$-modules. The map \begin{eqnarray}\label{diffl:B} B : C_p(A) & @>>> & C_{p+1}(A) \\ \nonumber a_0\otimes\cdots\otimes a_p & \mapsto & \sum_{i=0}^p (-1)^{pi} 1\otimes a_i\otimes\cdots\otimes a_p\otimes a_0\otimes\cdots\otimes a_{i-1} \end{eqnarray} satisfies $B^2 = 0$ and $[B,b] = 0$ and therefore defines a map of complexes \[ B: C_\bullet(A) @>>> C_\bullet(A)[-1]\ . \] For $i,j,p\in{\Bbb Z}$ let \begin{eqnarray*} CC^-_p(A) & = & \prod_{\overset{i\geq 0}{i+j = p \mod 2}} C_{i+j}(A) \\ CC^{per}_p(A) & = & \prod_{i+j = p \mod 2} C_{i+j}(A)\ . \end{eqnarray*} The complex $(CC^-_\bullet(A),B+b)$ (respectively $(CC^{per}_\bullet(A),B+b)$) is called the {\em negative} (respectively {\em periodic}) {\em cyclic complex of $A$}. There are inclusions of complexes \begin{equation}\label{inclusions} CC^-_\bullet(A)[-2]\hookrightarrow CC^-_\bullet(A)\hookrightarrow CC^{per}_\bullet(A) \end{equation} and the short exact sequence \begin{equation}\label{ses:CC-C} 0 @>>> CC^-_\bullet(A)[-2] @>>> CC^-_\bullet(A) @>>> C_\bullet(A) @>>> 0\ . \end{equation} In what follows we will use the notation of Getzler and Jones (\cite{GJ}). Let $u$ denote a variable of degree $-2$ (with respect to the homological grading. Then the negative and periodic cyclic complexes are described by the following formulas: \begin{eqnarray} \label{ses:gejo} CC^-_\bullet(A) & = & (C_{\bullet}(A)[[u]], b + uB) \\ \label{ses:gejoper} CC^{per}_\bullet(A) & = & (C_{\bullet}(A)[[u, u^{-1}], b + uB)\ . \end{eqnarray} \subsection{Hochschild and cyclic complexes of sheaves of algebras} Suppose that $X$ is a topological space and ${\cal A}$ is a flat sheaf of $k$-algebras on $X$ such that there is a global section $1\in\Gamma(X;{\cal A})$ which restricts to $1_{{\cal A}_x}$ and $1_{{\cal A}_x}\cdot k$ is contained in the center of ${\cal A}_x$ for every point $x\in X$. Let $C_\bullet({\cal A})$ (respectively $CC^-_\bullet({\cal A}),\ CC^{per}_\bullet({\cal A})$) denote the complex of sheaves of $k$-modules associated to the presheaf with value $C_\bullet({\cal A}(U))$ (respectively $CC^-_\bullet({\cal A}(U)),\ CC^{per}_\bullet({\cal A}(U))$) on an open subset $U$ of $X$. Then $C_\bullet({\cal A})$ represents ${\cal A}\otimes^{{\Bbb L}}_{{\cal A}\otimes_k{\cal A}^{op}}{\cal A}$ in the derived category of sheaves of $k$-modules on $X$.
\section{Introduction} Attempts to conciliate General Relativity with Quantum Theory yielded to propose theories of gravitation including torsion fields, so that the natural arena is the space--time $U_4$ that is a generalization of Riemann manifold $V_4$. The advantage to pass from $V_4$ to $U_4$ is due to the fact that the spin of a particle turns out to be related to the torsion just as its mass is responsable of the curvature. From this point of view, such a generalization tries to include the spin fields of matter into the same geometrical scheme of General Relativity. One of the attempts in this direction is the Einstein--Cartan--Sciama--Kibble (ECSK) theory \cite{hehl}. However the torsion seems to play an important role in any fundamental theory. For instance: a torsion field appears in (super)string theory if we consider string fundamental modes; we need, at least, a scalar mode and two tensor modes: a symmetric and antisymmetric one. The latter, in the low energy limit for string effective action, gives the effects of a torsion field \cite{GSW}; any attempts of unification between gravity and electromagnetism require the inclusion torsion in four and in higher--dimensional theories as Kaluza--Klein ones \cite{GER}; theories of gravity formulated in terms of twistors need the inclusion of torsion \cite{HOW}; in the supergravity theory torsion, curvature and matter fields are treated under the same standard \cite{LOS}; in cosmology torsion could have had a relevant role into dynamics of the early universe because it gives a repulsive contribution to the energy--momentum tensor so that cosmological models become singularity--free \cite{SIV}, and if the universe undergoes one or several phase transitions, torsion could give rise to topological defects ({\it e.g. } torsion walls \cite{FIG}) which today can result as intrinsic angular momenta for cosmic structures as galaxies. Some macroscopic observable effects of torsion in the framework of cosmology has been studied in Ref. \cite{ZZI} where it is shown that the presence of torsion into effective energy--momentum tensor alters the spectrum of cosmological perturbations giving characteristic lengths for large scale structures. As a final remark, we have to note that spacetime torsion, being related to the intrinsic spin degrees of freedom of matter \cite{hehl}, cannot be transformed away, so that we have to expect its remnants at any epoch of cosmological evolution. All these arguments do not allow to neglect torsion in any comprehensive theory of gravity which takes into account non-gravitational counterpart of fundamental interactions. The purpose of this paper is to show that, in presence of torsion, the helicity of fermion particles is not conserved. This effects could be important for testing some astrophysical consequences of torsion \cite{HAM} because of smallness of coupling constant with respect to the other fundamental interactions. Our starting point is to consider the Dirac equation in the space--time $U_4$. Due to the torsion, it acquires an additional coupling term of the form $(1/4)S_{\alpha\beta\sigma}\gamma^{\alpha}\gamma^{\beta}\gamma^{\sigma}$, where $S_{\alpha\beta\sigma}$ is related to the antisymmetric part of the affine connection, $\Gamma^{\sigma}_{[\alpha, \beta]}=S_{\alpha\beta}^{\sigma}$. This term is, as we will see, responsable of the spin flipping of fermions. It is worthwhile to note that helicity flips are induced also by gravitational fields, as consequence of coupling between spin and curvature \cite{PAP}. The paper is organized as follows. In Section 2 we will shortly review the basic concepts leading to the Dirac equation in presence of torsion fied. In Section 3 we show that the helicity operator of a fermionc particle is not conserved. The probability that the flip helicity occurs is calculated in Section 4. Conclusions are discussed in Section 5. \section{The Dirac Hamiltonian} The Dirac equation in curved space--time is written in terms of the vierbeins formalism \cite{BIR}. One introduces the vierbein fileds $e_{\mu}^a(x)$ where the Latin indices refer to the locally inertial frame and Greek indices to a generic non--inertial frame. The non--holonomic index $a$ labels the vierbein, while the holonomic index $\mu$ labels the components of a given vierbein. The connection in non--holonomic coordinates is given by \cite{HAM} \begin{equation}\label{eq2.1} \Gamma_{abc}=-\Omega_{abc}+\Omega_{bca}-\Omega_{cab}+S_{abc}\,{,} \end{equation} where $\Omega_{\alpha\beta}^c\equiv e^c_{[\alpha,\beta]}$, $\Omega^c_{ab}= e_a^{\alpha}\,e_b^{\beta}\,c_{\sigma}^c\, \Omega_{\alpha\beta}^{\sigma}$, and $S_{abc}$ is the anti-symmetric part of the affine connection. The covariant derivative is defined as \begin{equation}\label{eq2.2} D_{\mu}\equiv \partial_{\mu}-\frac{1}{4}\Gamma_{\mu ab}\gamma^a\gamma^b\,{,} \end{equation} and the Dirac equation is given by \begin{equation}\label{eq2.3} \gamma^a D_a\psi+i\frac{mc}{\hbar}\psi=0\,{.} \end{equation} In the spirit to study only the effects due to the torsion, we will neglect gravitational effects to the spin flip (they have been analyzed in details in Ref. \cite{PAP}). It means to neglect the $\Omega_{abc}$ terms in eq. (\ref{eq2.1}) so that the Dirac equations assumes the form \begin{equation}\label{eq2.4} \gamma^{\alpha}\psi_{,\alpha}+i\frac{mc}{\hbar}\psi=\frac{1}{4} S_{\alpha\beta\sigma}\gamma^{\alpha}\gamma^{\beta}\gamma^{\sigma}\psi \end{equation} >From it one derives the Hamiltonian \begin{equation}\label{eq2.5} H=c\vec{\alpha}\cdot\vec{p}+mc^2\beta+\frac{i}{4}S_{\alpha\beta\sigma} \gamma^0\gamma^{\alpha}\gamma^{\beta}\gamma^{\sigma}=H_0+H^{\prime}\,{,} \end{equation} where $H^{\prime}$ is a perturbation of the unperturbed Hamiltonian $H_0=c\vec{\alpha}\cdot\vec{p}+mc^2\beta$. \section{Helicity flip of fermions} In this section we will prove that the helicity of a fermion is not conserved in a space $U_4$. This follows by calculating the time variation of the helicity operator in the Heisenberg representation and showing that it does not vanish. The helicity operator is defined as \cite{ITZ} \begin{equation}\label{eq3.1} h=\vec{\Sigma}\cdot\vec{\hat{p}}\,{,} \end{equation} where the spin matrix $\vec{\Sigma}$ and the versor $\vec{\hat{p}}$ are \begin{equation}\label{eq3.2} \Sigma^i= \left( \begin{array}{cc} \sigma^i & 0 \\ 0 & \sigma^i \end{array} \right) \,{,} \quad \hat{p}^i=\frac{p^i}{|\vec{p}|}\,{.} \end{equation} $\sigma^i, i=1,2,3$ are the Pauli matrices and $p^{\mu}=(p^0, \vec{p})$ is the momentum. In the Heisenberg representation the dynamical evolution of the helicity operator is given by \begin{equation}\label{eq3.3} i\hbar\dot{h}=[h,H]\,{,} \end{equation} where $H$ is the Hamiltonian of the system under consideration. For the Hamiltonian (\ref{eq2.5}) one gets \begin{equation}\label{eq3.4} i\dot{h}=\frac{cp^k}{4|\vec{p}|}\varepsilon_{ijk}S_{\alpha\beta\sigma} \gamma^0\left[ g^{i\sigma}\gamma^{\alpha}\gamma^{\beta}\gamma^j+ 2g^{i\alpha}g^{j\beta}\gamma^{\sigma}\right] \,{.} \end{equation} Eq. (\ref{eq3.4}) implies that $\dot{h}\neq 0$ so that the helicity of fermion particle is not conserved. \section{Probability of spin flipping} In this Section, we will calculate the probability of the helicity flip induced by the torsion term in Eq. (\ref{eq2.5}). We consider the totally anti-symmetric dual or a null vector, $S^{\sigma}= (|\vec{S}|, \vec{S})$. We also used the approximation $g_{\mu\nu}\sim \eta_{\mu\nu}$. Then the Hamiltonian (\ref{eq2.5}) can be recast in the form \begin{equation}\label{eq4.1a} H^{\prime}=-\frac{3\hbar c|\vec{S}|}{2}\gamma^5+i\frac{3}{2} \vec{S}\cdot\left( \begin{array}{cc} \vec{\sigma} & 0 \\ 0 & \vec{\sigma} \end{array} \right) \,{.} \end{equation} where $\gamma^5$ is defined as \cite{ITZ} \begin{equation}\label{gamma5} \gamma^5=i\gamma^0\gamma^1\gamma^2\gamma^3= \left( \begin{array}{cc} 0 & 1 \\ 1 & 0 \end{array} \right) \,{.} \end{equation} The state of a fermion particle is described by spinor $$\psi (x)=\left( \begin{array}{c} \psi_R \\ \psi_L \end{array} \right)\,{,}$$ so that it can be rewritten as a superposition of states $|\psi_R>$ and $|\psi_L>$. For istance, at $t=0$ one has \begin{equation}\label{eq4.1} |\psi(0)>=a_0|\psi_R>+b_0|\psi_L>\,{,} \end{equation} where $a_0, b_0$ are constants, $|\psi_R>$ and $|\psi_L>$ are eigenkets of energy, i.e. $H_0|\psi_{R/L}>=E|\psi_{R/L}>$. We choose the independent kets $$ |\psi_R>\equiv \left( \begin{array}{c} 1 \\ 0 \end{array} \right)\,{,} \quad |\psi_L>\equiv \left( \begin{array}{c} 0 \\ 1 \end{array} \right)\,{.} $$ The time evolution of the state (\ref{eq4.1}) \begin{equation}\label{eq4.2} |\psi(t)>=a(t)|\psi_R>+b(t)|\psi_L>\,{,} \end{equation} is given by recasting Dirac's equation as a Schr\"odinger like one \begin{equation}\label{eq4.3} i\hbar\frac{\partial}{\partial t}|\psi (t)>=(H_0+H^{\prime})|\psi(t)>\,{.} \end{equation} Inserting Eq. (\ref{eq4.2}) into Eq. (\ref{eq4.3}), at first order of perturbative calculation, one gets \begin{equation}\label{eq4.4} i\hbar\frac{\partial}{\partial t} \left( \begin{array}{c} a \\ b \end{array} \right) =M \left( \begin{array}{c} a \\ b \end{array} \right)\,{.} \end{equation} where $M$ is the matrix \begin{equation}\label{eq4.5} M=\left( \begin{array}{cc} <\psi_R|H_0+H^{\prime}|\psi_R> & <\psi_R|H^{\prime}|\psi_L> \\ <\psi_L|H^{\prime}|\psi_R> & <\psi_L|H_0+H^{\prime}|\psi_L> \end{array} \right) \,{.} \end{equation} Explicit calculation of the matrix elements yields \begin{equation}\label{eq4.6} M=\left( \begin{array}{cc} E +i3\hbar c|\vec{S}|^2 & \hbar c|\vec{S}| \\ \hbar c|\vec{S}| & E+i3\hbar c|\vec{S}|^2 \end{array} \right) \,{.} \end{equation} By diagonalizing the matrix (\ref{eq4.6}), one derives the eigenvalues \begin{equation}\label{eq4.7} \lambda_{\pm}=E+i3|\vec{S}|^2\pm \frac{3\hbar c}{2}|\vec{S}|\,{,} \end{equation} and the corresponding normalized eigenkets \begin{equation}\label{eq4.8} |\lambda_{\pm}>=\frac{1}{\sqrt{2}}[|\psi_R>\pm\eta_{\pm}|\psi_L>]\,{.} \end{equation} In Eq. (\ref{eq4.8}), $\eta_{\pm}$ are the phase factors that we choose to be equal to one. It implies that at $t=0$, $|\psi (0)>=|\psi_R>$, i.e. $a_0=1, b_0=0$ in the Eq. (\ref{eq4.1}). Then, the evolution of the state $|\psi (t)>$ can be written as \begin{eqnarray}\label{eq4.9} |\psi (t)> & = & \frac{1}{\sqrt{2}}[e^{-i\lambda_+ t/\hbar}|\lambda_+> +e^{-i\lambda_- t/\hbar}|\lambda_- >]= \\ & = & e^{-i(E/\hbar+3c|\vec{S}|/4) t}\, e^{- 3c|\vec{S}|^2 t}\, [\cos\frac{c\vert S\vert}{2}t\, |\psi_R> + \sin\frac{c\vert S\vert }{2}t\,|\psi_L>]\,{.} \nonumber \end{eqnarray} Eq. (\ref{eq4.9}) describes the state of a fermion at time $t$ if it starts as $|\psi_R>$. The probability to find it in state $|\psi_R>$ at time $t$ is $P_R(t)\sim\cos^2(3c|\vec{S}|/4)t$, while the probability that the spin flip occurs is $P_L(t)\sim\sin^2(3c|\vec{S}/4)t$. The frequency of spin flipping is $\omega=3c|\vec{S}|/4$, from which follows the characteristic length $L=8\pi/3|\vec{S}|$. Due to the dissipative term, the state decrease exponentially. This fact has important consequences in the very early universe. \section{Conclusions} In this paper, we calculate the probability that a background torsion source induces a spin flip on fermion particles moving in it. The torsion field is described by a null vector. We are dealing with high energy fermion particles, so that helicity can be identified with spin; this method can work both for fermion massive and massless particles. This phenomenology can occur in a regime where the effects of torsion become of the same order of magnitude or bigger than those due to energy momentum tensor at extremely high densities and at sufficiently high polarization of fermion particles. Such a scenario could realize at early cosmological epoch where particle density becomes similar to the critical cosmological density; for example, this happens if electrons are taken into account and $kT\simeq 10^{11}$GeV \cite{BAU}. It means that, at this epoch, the probability $P_L(t)$ has to be different from zero. In this sense, torsion and spin density can assume relevant roles in the today observed astrophysical structures, resulting, for example, as intrinsic macroscopic angular momenta \cite{FIG}. \vspace{1. cm}
\section*{Introduction} There is mounting experimental evidence that novel charge- and spin-ordered phases are generic to underdoped cuprates\cite{Cuprates}. The doped holes in Nd-doped La$_{1-x}$Sr$_x$CuO$_4$ segregate into periodic stripes that separate antiferromagnetically ordered, hole-poor domains. Lattice distortions pin these stripes, and this pinning is most effective for planar hole concentrations near $p$=1/8 where the stripe modulation wavelength is commensurate with the lattice. In the absence of pinning, stripe modulations are presumed to be fluctuating and/or disordered, and this is the emerging picture for La$_{1-x}$Sr$_x$CuO$_4$ (La-214) and YBa$_2$Cu$_3$O$_{6+x}$ (Y-123)\cite{DynamicStripes}. We have recently demonstrated, through measurements of thermal conductivity ($\kappa$)\cite{CohnHg}, that both Y-123 and HgBa$_2$Ca$_{m-1}$Cu$_m$O$_{2m+2+\delta}$ [Hg-12($m$-1)$m$, $m$=1, 2, 3] exhibit doping anomalies near $p$=1/8 that can be attributed to the presence of localized charge and associated lattice distortions. Here we discuss the oxygen doping behavior of the anomalies in Hg cuprates and present new results on vacancy-free Ca-doped YBa$_2$Cu$_4$O$_8$ (Y-124) that suggest elastic properties may reflect stripe dynamics and stripe fragments may pin near oxygen-vacancy clusters. \begin{figure}[t!] \centerline{\epsfig{file=hts99_f1.psc,height=7.52in,width=5.8125in}}% \vskip -3.4in \caption{(a) Thermal resistivity at $T$=100K relative to that at $p_{opt}=0.16$ for Y-123 polycrystals and the ab-plane of single crystals \protect\cite{Popoviciu} [$W_{opt}=0.21$ mK/W (0.08mK/W) for the polycrystal (crystals)]. The hole concentration per planar Cu atom, $p$, was determined from thermopower measurements and the empirical relations with bond valence sums 59 established by Tallon {\it et al.}\protect\cite{PfromTEP}. Also plotted ($\times$'s) is the relative intensity of anomalous $^{63}$Cu NQR signals\protect\cite{HamScal}. The dashed curve is a guide to the eye. (b) The normalized slope change in $\kappa(T)$ at $T_c$ {\it vs} doping for each of the Y-123 specimens in (a), for the a-axis of single-crystal Y-124\protect\cite{Cohn124xtals}, and for polycrystal Y$_{1-x}$Ca$_x$-124\protect\cite{CohnYCa124}. The Y$_{1-x}$Ca$_x$-124 data are multiplied by 2.3 so that they coincide with the Y-124 crystal data at x=0. Solid (open) symbols are referred to the left (right) ordinate. Also shown ($\times$'s) are the normalized electronic specific heat jump\protect\cite{Loram}, $\Delta\gamma/\gamma$, and (solid curve) the $\mu$SR depolarization rate\protect\cite{muSRY123} (in $\mu$s$^{-1}$), divided by 1.4 and 2.7, respectively, and referred to the right ordinate. (c) Thermal resistivity for Hg-1201 and Hg-1212 polycrystals relative to that of Hg-1223\protect\cite{CohnHg}. Dashed curves are guides. (d) The normalized slope change in $\kappa(T)$ at $T_c$ {\it vs} doping for Hg cuprates. The solid line is $1.71-250(p-0.157)^2$. Dashed curves are guides.} \label{fig1} \end{figure} The $p$=1/8 features (Fig.~\ref{fig1}) are evident in the doping behavior of the normal-state thermal resistivity, $W=1/\kappa$, and the normalized change in temperature derivative of $\kappa$ that occurs at $T_c$, $\Gamma\equiv -d(\kappa^s/\kappa^n)/dt|_{t=1}$ [$t=T/T_c$ and $\kappa^s (\kappa^n)$ is the thermal conductivity in the superconducting (normal) state]. For Y-123, there is a compelling correlation of $W(p)$ and $\Gamma(p)$ with the doping behavior of anomalous $^{63}$Cu NQR spectral weight\cite{HamScal}, attributed to localized holes, and the electronic specific heat jump\cite{Loram}, $\Delta\gamma/\gamma$, respectively [crosses in Fig.'s 1 (a) and (b)]. Thus $W$ probes lattice distortions associated with localized holes, and $\Gamma$ the change in low-energy spectral weight induced by superconductivity. Since both $\Gamma$ and $\Delta\gamma/\gamma$ provide bulk measures of the superfluid volume, it is significant that the muon spin rotation ($\mu$SR) depolarization rate [$\sigma_0$ in Fig. 1 (b)], proportional to the superfluid density, exhibits no anomalous behavior near 1/8 doping. The $\mu$SR signal originates in regions of the specimen where there is a flux lattice. The apparent discrepancy between $\sigma_0$ and $\Gamma$ or $\Delta\gamma/\gamma$ is resolved if the material is inhomogeneous, composed of non-superconducting clusters embedded in a superconducting network. The suppression of $\Gamma$ and $\Delta\gamma/\gamma$ below the scaled $\sigma_0$ curve in Fig. 1 (a) are then measures of the non-superconducting volume fraction. Taken together, the $W$ and $\Gamma$ data imply that the non-superconducting regions are comprised of localized holes and associated lattice distortions, i.e. polarons. It is plausible that these hole-localized regions are stripe domains akin to those inferred from neutron scattering\cite{Cuprates}. That they produce lattice thermal resistance implies they are static on the timescale of the average phonon lifetime, estimated as a few ps in the ab-plane of Y-123\cite{CohnHg}. Given that $T_c$ is not substantially suppressed near $p$=1/8 implies that these regions do not percolate, and thus the picture is one of small clusters of localized holes and their associated lattice distortions, perhaps no larger than the stripe unit cell ($2a\times8a$ where $a$ is the lattice constant\cite{Cuprates}), separated by a distance comparable to the phonon mean free path (about 100${\rm\AA}\sim 25a$ at 100K). For the Hg materials the $p$=1/8 enhancement of $W$ and suppression of $\Gamma$ is most prominent in single-layer Hg-1201, less so in double-layer Hg-1212, and absent or negligible in three-layer Hg-1223. This trend follows that of the oxygen vacancy concentration: a single HgO$_{\delta}$ layer per unit cell contributes charge to $m$ planes in Hg-12($m$-1)$m$ so that the oxygen vacancy concentration, $1-\delta$, increases with decreasing $m$ \cite{Occupancy}. The absence of suppression in $\Gamma$ near 1/8 doping for Hg-1223 [Fig.~1~(d)] suggests that this material has sufficiently few localized-hole domains that their effects in $W$ and $\Gamma$ are unobservable. Thus we employ the Hg-1223 $W(p)$ data as a reference and plot the differences for the other two compounds in Fig. 1 (c). Comparing Fig.'s~1~(c) and (d) we see that for both Hg-1201 and Hg-1212 $W$ is enhanced and $\Gamma$ is suppressed relative to values for Hg-1223 in common ranges of $p$, with maximal differences near $p$=1/8. \begin{figure}[t!] \vskip -.5in \centerline{\epsfig{file=hts99_f2.psc,height=7.2in}} \vskip -3in \caption{(a) Thermal conductivity of Hg-1223 polycrystal versus doping. The dotted and dashed curves are proposed electronic (right ordinate) and lattice thermal conductivities, respectively. The solid line is a guide to the eye. (b) sound velocity data for single-crystal La-214 from Ref.~\protect\cite{LSCOelastic}.} \label{fig2} \end{figure} \section*{Stripes Alter the phonon dispersion?} That 1/8 doping anomalies are observed in both Y-123 and Hg cuprates suggests they are a generic feature of underdoped CuO$_2$ planes. In this context the doping behavior of $\kappa$ for Hg-1223 [Fig. 2 (a)] is of interest since it presumably approximates the phase behavior in the absence of localized holes. We expect the electronic contribution, $\kappa_e$, to increase smoothly with increasing $p$. The upturn in $\kappa$ above optimal doping ($p$=0.16) is presumably attributable to this rising $\kappa_e$. The most reliable estimate of $\kappa_e$ in the cuprates comes from thermal Hall conductivity measurements\cite{Krishana} on Y-123 which imply $\kappa_e/\kappa\simeq 0.1$ for in-plane heat conduction near optimal doping. The data in Fig. 1 (a), (b) suggest that this ratio is roughly the same in polycrystals, thus motivating the dotted and dashed curves in Fig. 2 (a) as an educated guess for the electronic and lattice terms, respectively, in Hg-1223. The lattice conductivity, $\kappa_L=\kappa-\kappa_e$, predominates in the underdoped regime and is peaked near $p$=1/8. There is experimental support for the proposal that this peak in $\kappa$ is associated with doping-dependent changes in the phonon dispersion which are generic to underdoped cuprates. Figure 2~(b) shows the behavior of the normal-state, transverse shear elastic constant (proportional to the square of the sound velocity) for single-crystal La-214\cite{LSCOelastic}, the only material for which doping-dependent measurements for all the main symmetry directions have been reported to our knowledge. A substantial hardening of the lattice in the underdoped regime, with a maximum near $p=x\simeq$1/8, was observed for all symmetries, indicating that the changes with doping are systemic. It is possible that the phase behavior of the elastic constants for La-214 and $\kappa_L$ for Hg-1223 reflect a renormalization of the lattice dispersion due to changes in stripe dynamics with doping. At $p$=1/8 the stripes are maximally commensurate with the lattice, and their fluctuations should be minimal. Enhanced fluctuations are to be expected at higher doping, due to the destabilizing role of repulsive interactions in stripes that neutron scattering results (Yamada {\it et al.}\cite{DynamicStripes}) suggest, are charge compressed. For $p<$1/8, the cluster spin-glass state observed by $\mu$SR studies\cite{Niedermayer} is characterized by increasing magnetic disorder with decreasing $p$ in the range $0.08 < p <0.12$, possibly associated with increasing disorder\cite{Emery} in the stripe period. It is plausible that disorder in the stripe system at both higher and lower doping about $p$=1/8 induces a softening of the lattice that is reflected in Fig.~2. Theoretical investigations of elastic coupling to the stripe system would certainly be of interest. \begin{figure}[b!] \centerline{\epsfig{file=hts99_f3.psc,height=6.47in,width=5in}} \vskip -3in \caption{(a) The enhancement in thermal resistivity and suppression in $\Gamma$ for the Hg compounds at $p$=1/8, relative to values for Hg-1223, plotted versus the oscillator strength ratio of Raman-scattering defect modes (Ref.~\protect\cite{HgRaman}).} \label{fig3} \end{figure} \section*{Oxygen Vacancies and 1/8 Anomalies} Returning now to the enhancement of $W$ and suppression of $\Gamma$ near $p$=1/8, our measurements suggest these phenomena reflect stripe pinning (or reduced stripe fluctuations) associated with oxygen vacancy clusters. Raman studies of defect modes in the Hg materials\cite{HgRaman} provide a measure of the density of vacancy clusters. The 590 cm$^{-1}$ Raman mode in Hg cuprates is attributed to c-axis vibrations of apical oxygen in the presence of oxygen vacancies on each of the four nearest-neighbor sites in the HgO$_{\delta}$ layers. The amount by which the thermal resistivity and $\Gamma$ values of Hg-1201 and Hg-1212 differ from those of Hg-1223 at $p$=1/8 both correlate well with the integrated oscillator strength of this vibrational mode normalized by that of the 570 cm$^{-1}$ mode (Fig.~\ref{fig3}), the latter attributed to apical vibrations in the presence of fewer than four vacancies, and common in the spectra\cite{HgRaman} of all three Hg materials. We infer that oxygen-vacancy clusters, of size four or more in the Hg cuprates, can pin a stripe fragment. Supporting the role of oxygen vacancies in the observed 1/8 features for Y-123 are our recent measurements on oxygen-stoichiometric Y$_{1-x}$Ca$_x$-124\cite{CohnYCa124}. The $\Gamma$ values for $x$=0, 0.10, 0.15 are plotted in Fig.~1~(b). There is no evidence for suppression of $\Gamma$ near 1/8 doping, and the data follow the scaled $\mu$SR curve quite well. Within the context of the interpretation we have outlined, we conclude that Ca substitution for Y (at the level of $\leq$ 15\%) is not as effective in stripe pinning as are oxygen vacancy clusters. It remains to be determined by what mechanism these clusters induce pinning. \section*{acknowledgements} This work was supported by NSF Grant No. DMR-9631236.
\section{Introduction} Coherent states provide a versatile tool in quantum mechanics both from a conceptual and a technical point of view. Originally discovered \cite{schroedinger26} in the search for `particle solutions' (stable, nondispersing wave packets) of Schr\"odinger's equation, they turned out to be useful \cite{perelomov86}, for example, in semiclassical descriptions of quantum systems, quantum optics, quantum statistical mechanics, quantum field theory$\ldots$ To a large extent, this wide range of applications is due to the intimate link of coherent states to group theory. Having once realized this fundamental connection, the way is open to generalize the initial version of coherent states which is related to the Heisenberg-Weyl group of phase-space translations of a quantum particle. The purpose of the present note is twofold. First, it will be pointed out that coherent states enable one to provide a simple answer to the question: {\em How to express the state of a quantum mechanical system by measurable quantities?} Due to recent experimental progress in handling individual quantum systems \cite{smithey+93}-\cite{kurtsiefer+97}, this long-standing question \cite{pauli33} has turned into a rapidly expanding field known as {\em state reconstruction} \cite{leonhardt97}. Second, a {\em discrete phase-space formulation} of the quantum mechanics will be shown to emerge naturally from the coherent-state approach to state reconstruction for a spin. The new `expectation value representation' deals with {\em positive} probabilities in phase space only, contrary to the Wigner functions or `quasi-probabilities' which may take negative values, too. The resulting symbolic calculus provides a {\em generalized} Stratonovich-Weyl correspondence between operators and symbols as follows from a comparison to the Moyal representation for a spin. There is a large variety of other methods to determine the state of a spin in terms of expectations. Many references can be found in Refs.\ \cite{weigert92}-\cite{weigert99/1}, and \cite{brif+99}, for example. \section{State reconstruction based on coherent states} Let us briefly review the quantum mechanical description of a spin which is based on a vector operator $\widehat {\bf S} \equiv \hbar \widehat{\bf s}$. Its components act irreducibly in a $(2s+1)$-dimensional Hilbert space ${\cal H}_s$, and they satisfy the commutation relations of the algebra $su(2)$: $[ {\hat s}_x, {\hat s}_y ]= i{\hat s}_z, \ldots$ The standard basis of the space ${\cal H}_s$ is given by the eigenvectors of the $z$ component of the spin, ${\widehat S}_z = \widehat {\bf S} \cdot {\bf n}_z$, denoted by $\ket{\mu, {\bf n}_z},$ $-s \leq \mu \leq s$. Observables are represented by hermitian operators, ${\widehat {A}}^\dagger = {\widehat {A}}$, all of which are linear combinations of polynomials in the operators ${\hat s}_x$, ${\hat s}_y$ and ${\hat s}_z$ of degree $2s$ at most. The ensemble of all hermitean operators acting on ${\cal H}_s$ can be considered as a vector space ${\cal A}_s$ of dimension $N_s = (2s+1)^2$. A {\em coherent spin state} $\ket{ {\bf n}}$ is associated to each point ${\bf n} = ( \sin \vartheta \cos \varphi, \sin \vartheta \sin \varphi,$ $\cos \vartheta) \in {\cal S}^2$ of the surface of the unit sphere \cite{arecchi+72}. It coincides with the eigenstate of the operator $ \hat {\bf s} \cdot {\bf n} $ along the direction ${\bf n} $ and with eigenvalue $s$: \be \ket{{\bf n}} \equiv \exp [ -i \, \vartheta \, {\bf m}(\varphi) \cdot {\hat {\bf s}} \, ] \, \ket{s,{\bf n}_z} \, , \label{defineaxes} \ee where ${\bf m}(\varphi) = (- \sin \varphi,\cos\varphi,0)$. Thus, the coherent state $\ket{{\bf n} }$ is obtained from rotating the state $\ket{s,{\bf n}_z}$ about the axis ${\bf m}(\varphi)$ in the $xy$ plane by an angle $\vartheta$. The ensemble of all coherent states provides an overcomplete basis of the Hilbert space ${\cal H}_s$: \be \frac{(2s+1)}{4\pi} \int_{{\cal S}^2} d {\bf n} \, \ket{{\bf n}} \bra{{\bf n}} = {\widehat E} \, , \label{complete} \ee where ${\widehat E}$ is the unit operator on ${\cal H}_s$. It is convenient to combine $(\vartheta,\varphi)$ into a single complex variable, $z= \tan(\vartheta/2)\exp [ i \varphi ]$. This provides a stereographic projection of the surface of the sphere to the complex plane. In terms of $z$, a coherent state has the expansion \cite{amiet+91} \be \ket{ z } = \frac{1}{(1+|z|^2)^s}\sum_{k= 0}^{2s} \left( \begin{array}{c} 2s \\ k \end{array} \right)^{1/2} z^{k}\ket{s-k, {\bf n}_z} \, . \label{expandcs} \ee In order to show that the density matrix ${\hat \rho}$ of a spin $s$ is determined unambiguously by appropriate measurements with a Stern-Gerlach apparatus one preceeds as follows. Distribute $N_s = (2s+1)^2$ axes ${\bf n}_{\mu\nu}, -s \leq \mu,\nu \leq s$, over $(2s+1)$ cones about the $z$ axis with different opening angles such that the set of the $(2s+1)$ directions on each cone is invariant under a rotation about $z$ by an angle $2\pi/(2s+1)$. As shown in \cite{amiet+99/2}, an unnormalized statistical operator ${\hat \rho}$ is then fixed by measuring the $N_s$ relative frequencies \be p_s({\bf n}_{\mu\nu}) = \bra{ {\bf n}_{\mu\nu}} \hat \rho \ket{ {\bf n}_{\mu\nu}} \, , \qquad -s \leq \mu,\nu \leq s \, , \label{relfreq} \ee that is, by the expectation values of the statistical operator $\hat \rho$ in the coherent states $\ket{{\bf n}_{\mu\nu}}$. Here is the idea of the proof: take the expectation value of $\hat \rho$ in the coherent states $\ket{ {\bf n}_{\mu\nu}}$ as given in Eq. (\ref{expandcs}). You obtain $N_s$ linear relations between probabilities $p_s({\bf n}_{\mu\nu})$ and the matrix elements of the density matrix with respect to the basis $\ket{s-k, {\bf n}_z}$. This set of equations can be inverted by standard techniques if the directions ${\bf n}_{\mu\nu}$ are chosen as described above. For a spin $s$, the projection operators \be {\widehat Q}_{{\mu\nu}} = \ket{{\bf n}_{{\mu\nu}}} \bra{{\bf n}_{{\mu\nu}}} \, , \qquad -s \leq \mu,\nu \leq s \, , \label{mixedrho} \ee constitute thus a {\em quorum} ${\cal Q}$. In general, a quorum is defined as a collection of (hermitean) operators having the property that their expectation values are sufficient to reconstruct the quantum state of the system at hand. The operators ${\widehat Q}_{{\mu\nu}}$ do even define an {\em optimal} quorum since exactly $(2s+1)^2$ numbers have to be determined experimentally which equals the number of free real parameters of the (unnormalized) hermitean density matrix $\hat \rho$. If the state to be reconstructed is known to be a {\em pure} one, only $4s$ real parameters need to be specified in order to identify it. The set of operators ${\widehat Q}_{{\mu\nu}} $ still provides a quorum but the data required for reconstruction now is highly redundant. In Ref.\ \cite{amiet+99/2}, another reconstruction method based on coherent states has been worked out which takes into account the reduced number of parameters. \section{A symbolic calculus for spin systems} It will be argued now that Eqs.\ (\ref{relfreq}) and (\ref{mixedrho}) provide the basis for a {\em symbolic calculus} in the spirit of the Wigner formalism \cite{wigner32,amiet+91} of quantum mechanics. For simplicity, the labels $\mu$ and $\nu$ are replaced by a single index $n= (\mu,\nu)$, say, with $1 \leq n \leq N_s$. The set of all unnormalized hermitean density matrices for a spin $s$ is equal to the set of all hermitean operators acting on the Hilbert space ${\cal H}_s$. Therefore, the expectation values \begin{equation} {\sf A}_{n} = \mbox{ Tr } \left[ {\widehat {A}} \, {\widehat Q}_{n} \right] \, , \qquad 1 \leq n \leq N_s \, , \label{symbol1} \end{equation} of a hermitean operator $\widehat{A}$ in the coherent states $\ket{{\bf n}_n}$ suffice to characterize unambiguously the operator. In other words, the is a one-to-one relation between the operator $\widehat{A}$ and its {\em symbol\ }, the collection of $N_s$ real numbers ${\sf A}_{n}$. Considering the trace as a scalar product, Eq. (\ref{symbol1}) can be interpreted as the projection of the operator $\widehat{A}$ on $N_s$ linearly independent elements ${\widehat Q}_{n}$ of the vector space ${\cal A}_s$. The operators ${\widehat Q}_{n}$, however, do {\em not} constitute an orthogonal basis: $\mbox{ Tr } [ {\widehat Q}_{n}{\widehat Q}_{n'} ] = | \langle {\bf n}_n \ket{{\bf n}_{n'}} |^2 \neq 0$. Nevertheless, the inversion of Eq. (\ref{symbol1}) is achieved easily by means of the operators ${\widehat {\sf Q}}^{n}$, a second basis of the space ${\cal A}_s$, defined as the dual of the quorum (\ref{mixedrho}): \begin{equation} \frac{1}{2s+1} \mbox{ Tr }\left[ {\widehat Q}_{n} {\widehat {\sf Q}}^{n'} \right] = \delta_{n}^{n'} \, , \qquad 1 \leq n,n' \leq N_s \, . \label{orthogonalityquorum} \end{equation} This implies an expansion for hermitean operators, \begin{equation} {\widehat {A}} = \frac{1}{2s+1} \sum_{n=1}^{N_s} {\sf A}_{n} {\widehat {\sf Q}}^{n} \, , \label{expandA} \end{equation} which is indeed the inversion of (\ref{symbol1}): ${\widehat {A}}$ is given explicitly as a function of the measurable numbers ${\sf A}_{n}$ and a specified set of operators. Due to the symmetry between the quorum ${\cal Q}$ and its dual ${\cal Q}^D$, there is a second way to expand hermitean operators, \begin{equation} {\widehat {A}} = \frac{1}{2s+1} \sum_{n=1}^{N_s} A^{n} {\widehat Q}_{n} \, , \qquad A^{n} = \mbox{ Tr } \left[ {\widehat {A}} \, {\widehat {\sf Q}}^{n} \right] \, , \label{expandquorum} \end{equation} with coefficients $A^n \neq {\sf A}_{n}$, providing a {\em second} symbol of ${\widehat {A}}$. Upon expanding the operator ${\widehat Q}_n$ in terms of the dual basis, \be {\widehat Q}_n = \frac{1}{2s+1} \sum_{n'=1}^{N_s} {\sf G}_{nn'} {\widehat {\sf Q}}^{n'} \, , \qquad {\sf G}_{nn'} = \mbox{ Tr } \left[ {\widehat Q}_{n} {\widehat Q}_{n'} \right] \, , \label{trfqdualq} \ee one discovers the existence of a real symmetric matrix ${\sf G}$ which can be shown to be {\em positive definite}. It can be used as a {\em metric} for raising and lowering indeces of both symbols and operators. In terms of symbols, tracing the product ${\widehat {A}} \, {\widehat {B}}$ of two hermitean operators gives \be \mbox{ Tr } \left[ {\widehat {A}}\, {\widehat {B}} \right] = \frac{1}{(2s+1)} \sum_{n=1}^{N_s} {\widehat {\sf A}}^{n} {\widehat B}_{n} = \frac{1}{(2s+1)} \sum_{n=1}^{N_s} {\widehat A}_{n} {\widehat {\sf B}}^{n} \, , \label{symbprod} \ee using the expansions (\ref{expandA}), (\ref{expandquorum}), and (\ref{orthogonalityquorum}). Invoking the metric ${\sf G}$, one can also write \be \mbox{ Tr } \left[ {\widehat {A}}\, {\widehat {B}} \right] = \frac{1}{(2s+1)^2} \sum_{n,n'=1}^{N_s} {\sf G}_{nn'} {\widehat {\sf A}}^{n} {\widehat {\sf B}}^{n'} = \frac{1}{(2s+1)^2} \sum_{n,n'=1}^{N_s} {\sf G}^{nn'} {\widehat A}_{n} {\widehat B}_{n'} \, . \label{symbprod2} \ee Let us now consider the properties of the statistical operator $\hat \rho$ when expanded in the basis ${\widehat {\sf Q}}^{n}$ dual to the original quorum, \begin{equation} {\hat {\rho}} = \frac{1}{2s+1} \sum_{n=1}^{N_s} {\sf P}_{n} {\widehat {\sf Q}}^{n} \, , \label{expandrhodual} \end{equation} where the coefficients ${\sf P}_{n} = \mbox{ Tr } [ \hat \rho\, {\widehat Q}_{n} ] \equiv \bra{{\bf n}_{n}} \hat \rho \ket{{\bf n}_{n}}$ satisfy \be 0 \leq {\sf P}_{n} \leq 1 \, , \qquad 1 \leq n \leq N_s \, . \label{positive!} \ee Since the density matrix $\hat \rho$ is a positive operator, the ${\sf P}_{n}$ are {\em non-negative} throughout, and each of the $N_s$ numbers ${\sf P}_{n}$ has a value less or equal to one due to the normalization condition $\mbox{ Tr }[\, \hat \rho \, ] = 1 $. The positivity of the numbers ${\sf P}_{n}$ is characteristic for the basis $ {\widehat {\sf Q}}^{n} $, and it is {\em not} valid for the expansion coefficients of $\hat \rho$ with respect to the basis ${\widehat Q}_{n}$. The interpretation of the coefficients $ {\sf P}_{n}$---to measure the value $s$ along the axis ${\bf n}_{n}$---is clearly compatible with (\ref{positive!}). It is important to note that, although each of the $ {\sf P}_{n}$ {\em is} a probability, they do {\em not} sum up to unity: \be 0 < \sum_{n=1}^{N_s} {\sf P}_{n} < (2s+1)^2 \, . \label{largersum} \ee This is due to the fact that they all refer to {\em different orientations} of the Stern-Gerlach apparatus, being thus associated with the measurement of {\em incompatible} observables, \be \left[ {\widehat Q}_{n}, {\widehat Q}_{n'} \right] \neq 0 \, , \qquad 1 \leq n,n' \leq N_s \, , \label{commutator} \ee since the scalar product $\bra{{\bf n}_{n}} {\bf n}_{n'} \rangle$ of two coherent states is different from zero. The sum in (\ref{largersum}) cannot take the value $(2s+1)^2 $ since this would require a common eigenstate of all the operators ${\widehat Q}_{n}$ which does not exist due to (\ref{commutator}). By an appropriate choice of the directions ${\bf n}_{n}$ (all in the neighborhood of one single direction ${\bf n}_0$, say), the sum can be arbitrarily close to $(2s+1)^2 $ for states `peaked' about ${\bf n}_0$. Similarly, the sum of all ${\sf P}_{n}$ cannot take on the value zero since this would require a vanishing density matrix which is impossible. If, however, considered as a sum of {\em expectation values}, there is no need for the numbers ${\sf P}_{n}$ to sum up to unity. Nevertheless, they are not completely independent when arising from a statistical operator: its normalization implies that \be \mbox{ Tr } \left[ \, \hat \rho \, \right] = \frac{1}{2s+1} \sum_{n=1}^{N_s} \mbox{ Tr} \left[ {\widehat {\sf Q}}^{n} \right] {\sf P}_{n} = 1 \, , \label{restriction} \end{equation} turning one of the probabilities into a function of the $(2s+1)^2-1 = 4s(s+1)$ others, leaving us with the correct number of free real parameters needed to specify a density matrix. In Ref.\ \cite{weigert99/2}, the time evolution of the density matrix, generated by a Hamiltonian $\widehat H$, is expressed in terms of linear differential equations of first-order which couple the probabilities ${\sf P}_{n}$. The resulting `expectation-value representation' for a spin is equivalent to any other representation. \section{Comparison to the Moyal representation} In the following, the symbolic representation introduced above is compared to the Moyal representation associated with the group $SU(2)$ as given by V\'arilly and Gracia-Bond\'\i a \cite{varilly+89}. For a spin $s$, they define a `Stratonovich-Weyl' correspondence as a rule which maps each operator ${\widehat A}$ on the $(2s+1)$-dimensional Hilbert space ${\cal H}_s$ to a function $W_{\! \! A} ({\bf n})$ on the phase space of the classical spin, ${\cal S}^2$. This prescription must satisfy various properties which are conveniently expressed in terms of a family of operator kernels ${\widehat \Delta ({\bf n})}$. The kernels establish a correspondence between an operator ${\widehat A}$ and its symbol via \be {W_{\! \! A} ({\bf n})} = \mbox{ Tr } \left[ {\widehat A} \, {\widehat \Delta ({\bf n})} \right] \, . \label{definesymbol} \ee Here are the requirements to be satisfied by the kernels: \begin{enumerate} \item[($\alpha$):]{Each kernel ${\widehat \Delta} ({\bf n})$ is a hermitean operator, \be {\widehat \Delta}^\dagger ({\bf n}) = {\widehat \Delta} ({\bf n}) \, , \qquad {\bf n} \in {\cal S}^2 \, , \label{herm} \ee implying that the symbol ${W_{\! \! A} ({\bf n})}$ in (\ref{definesymbol}) of a hermitean operator ${\widehat A}$ is {\em real}.} \item[($\beta$):]{The set of all kernels is (over-) {\em complete}: \be \frac{(2s+1)}{4\pi} \int_{{\cal S}^2} d{\bf n} \, {\widehat \Delta ({\bf n})} = {\widehat E} \, , \label{compl} \ee} providing thus an explicit resolution of the identity. \item[($\gamma$):]{Two kernels with labels ${\bf n}$ and ${\bf n}'$ satisfy an {\em orthogonality} relation: \be \frac{(2s+1)}{4\pi} \mbox{ Tr } \left[ \, {\widehat \Delta} ({\bf n}) \, {\widehat \Delta} ({\bf n}') \, \right] = \delta ({\bf n} - {\bf n}') \, , \qquad {\bf n}, {\bf n}' \in {\cal S}^2 \, \label{ortho} \ee This property, also called {\em traciality}, allows one to invert (\ref{definesymbol}): \be {\widehat A} = \frac{(2s+1)}{4\pi} \int_{{\cal S}^2} d{\bf n} \, W_{\! \! A} ({\bf n}) \, {\widehat \Delta} ({\bf n}) \, , \label{invertdefinesymbol} \ee that is, to express the operator ${\widehat A}$ in terms of its symbol. Note that (\ref{definesymbol}) and (\ref{invertdefinesymbol}) are effected by means of the {\em same} operator kernel.} \item[($\delta$):]{The kernels transform {\em covariantly} with respect to the group $SU(2)$: \be {\widehat U}_{\sf R}\, {\widehat \Delta ({\bf n})} \, {\widehat U}_{\sf R}^\dagger = {\widehat \Delta ( {\sf R} \, {\bf n})} \label{covar} \ee where the unitary ${\widehat U}_{\sf R}$ in ${\cal H}_s$ represents a rotation ${\sf R}$ .} \end{enumerate} These conditions have been used to derive the explicit form of the kernels ${\widehat \Delta ({\bf n})}$ for a spin. For a quantum {\em particle} in one dimension, the family of kernels generating the Moyal representation is given by ${\widehat \Delta} (\alpha) = {\widehat T} (\alpha) \, {\widehat P}\, {\widehat T}^\dagger (\alpha)$, where ${\widehat T} (\alpha)$, $\alpha \in {\sf C}$, represents a phase-space translation on the particle Hilbert space, and ${\widehat P}$ is the parity operator. Upon replacing the group of rotations by the translations in phase space, the operator ${\widehat \Delta} (\alpha)$ satisfies properties structurally equal to ($\alpha$)-($\delta$), giving thus rise to the standard phase-space representation of quantum mechanics for a particle. In the following, the properties of the expectation-value representation will be compared to those of the Stratonovich-Weyl correspondence. A fundamental difference is obvious from the outset: the representation based on the quorum $\cal Q$ and its dual ${\cal Q}^D$ gives rise to {\em two} intimately connected types of symbols, $A^n$ and ${\sf A}_n$. This `doubling' is due to fact that,individually, the collections $\{ {\widehat Q}_n \}$ and $\{ {\widehat {\sf Q}}^n\} $ do not form two orthogonal bases but only together they define a {\em bi-orthogonal} basis of the space ${\widehat {\cal A}}_s$. The comparison will be simplified by slightly modifying the notation. Denote the symbols of ${\widehat A}$ in analogy to (\ref{definesymbol}) by \be V_{\!\!A}^{n} \equiv A^{n} = \mbox{ Tr } \left[ {\widehat {A}} \, {\widehat {\sf Q}}^{n} \right] \quad \mbox{ and } \quad {\sf V}_{\!\!A,n} \equiv {\sf A}_{n} = \mbox{ Tr } \left[ {\widehat {A}} \, {\widehat Q}_{n} \right] \, . \label{definesymbol2} \ee The operator kernels ${\widehat Q}_n$ and ${\widehat {\sf Q}}^n$ have the following properties: \begin{enumerate} \item[($\alpha'$):]{Each of the kernels ${\widehat Q}_n$ and ${\widehat {\sf Q}}^n$ is a hermitean operator, \be {\widehat Q}_n^\dagger = {\widehat Q}_n \quad \mbox{ and } \quad \left( {\widehat {\sf Q}}^n \right)^\dagger = {\widehat {\sf Q}}^n \, , \qquad n=1,\ldots , N_s \, , \label{herm2} \ee implying that the symbols $V_{\! \! A,n}$ and $V_{\!\!A}^{n}$ of a hermitean operator ${\widehat A}$ are {\em real}.} \item[($\beta'$):]{Both sets of kernels are {\em complete}: \be \frac{1}{2s+1} \sum_{n=1}^{N_s} \mbox{ Tr} \left[ \, {\widehat Q}_n \, \right] \, {\widehat {\sf Q}}^n = \frac{1}{2s+1} \sum_{n=1}^{N_s} \mbox{ Tr } \left[ \, {\widehat {\sf Q}}^n \, \right] \, {\widehat Q}_n = {\widehat E} \, , \label{compl2} \ee providing thus two explicit resolutions of the identity. The analogy to ($\beta$) becomes more obvious if one notes that the kernel in (\ref{compl}) is multiplied by the symbol of the unity, $W_E ({\bf n}) \equiv \mbox{ Tr } [ {\widehat E} {\widehat \Delta} ({\bf n}) ] = 1$, and the coefficients in (\ref{compl2}) are the symbols of the unity, too: ${\sf V}_{E,n} \equiv \mbox{ Tr } [ {\widehat E} \, {\widehat Q}_n ]$ and $V_E^n \equiv \mbox{ Tr } [ {\widehat E} \, {\widehat {\sf Q}}^n ]$.} \item[($\gamma'$):]{Two kernels with labels $n$ and $n'$ satisfy the {\em bi-orthogonality} relation (\ref{orthogonalityquorum}) in the space ${\cal H}_s$: \be \frac{1}{2s+1} \mbox{ Tr }\left[ {\widehat Q}_{n} {\widehat {\sf Q}}^{n'} \right] = \delta_{n}^{n'} \, , \qquad 1 \leq n,n' \leq N_s \, . \label{ortho2} \ee This property allows one to invert (\ref{definesymbol2}): \be {\widehat A} = \frac{1}{2s+1} \sum_{n=1}^{N_s} {\sf V}_{\!\!A,n} {\widehat {\sf Q}}^n = \frac{1}{2s+1} \sum_{n=1}^{N_s} V_{\!\!A}^{n} \, {\widehat Q}_n \, , \label{invertdefinesymbol2} \ee that is, to express the operator ${\widehat A}$ in terms of its symbols. Contrary to (\ref{definesymbol}) and (\ref{invertdefinesymbol}) the transformations (\ref{definesymbol2}) and (\ref{invertdefinesymbol2}) are not effected by means of the same but the {\em dual} operator kernel.} \item[($\delta'$):]{The kernels ${\widehat Q}_{n}$ transform {\em covariantly} with respect to the group $SU(2)$: \be {\widehat U}_{\sf R} \, {\widehat Q}_{n} \, {\widehat U}_{\sf R}^\dagger = {\widehat U}_{\sf R} \, \ket{ {\bf n}_n} \bra{{\bf n}_n} \, {\widehat U}_{\sf R}^\dagger = \ket{{\sf R}\, {\bf n}_{n}} \bra{{\sf R}\, {\bf n}_{n}} = {\widehat Q}_{{\sf R},n}\, , \label{covar2} \ee since a rotation $\sf R$ maps a coherent state with label ${\bf n}$ to another coherent state with label $\sf R \, {\bf n}$. Hence, the operators ${\widehat Q}_{n}$ transform covariantly, and a similar relation holds for the dual family.} \end{enumerate} Conceptually, the quorum ${\cal Q}$ and its dual ${\cal Q}^D$ do thus provide a generalization of the Stratonovich-Weyl correspondence which is based on a bi-orthogonal basis of the space of operators ${\cal A}_s$. Therefore, this formulation differs also from the Wigner formalism for Hilbert spaces of finite systems introduced in Ref.\ \cite{leonhardt95}. \section{Outlook} It has been shown that state reconstruction for a spin based on coherent states gives rise to a generalized Stratonovich-Wey correspondence mapping operators on phase-space functions. In a natural manner, {\em two} symbols (each a collections of $N_s$ numbers associated with specific points of phase space) with qualitatively different properties emerge. One of the symbols allows one to represent a density matrix in terms {\em positive} numbers which correspond to probabilities associated with incompatible measurements. \medskip \noindent {\bf Acknowledgments} Financial support by the {\em Schweizerische Nationalfonds} is gratefully acknowledged.
\section{Introduction} In Veilleux, Sanders, \& Kim (1997b; hereafter VSK), a near-infrared search for obscured broad-line regions (BLRs) and [Si~VI] emission feature was carried out in a set of twenty-five ULIGs selected from the {\em IRAS} 1-Jy survey of 118 ULIGs (Kim \& Sanders 1998). Prior to selecting their set of 25 objects, VSK excluded the $\sim$ 10\% of the 118 galaxies that already were known optically to show direct signs of quasar activity, i.e. optically classified as Seyfert~1. Broad ($\Delta V_{\rm FWHM} \ga$ 2,000~km~s$^{-1}$) infrared recombination lines were detected in 5 (possibly 8) of the 10 optically classified Seyfert 2 galaxies in the sample. The high-ionization [Si~VI] feature ($\chi = 164$~eV), a clear signature of an AGN, was detected in 3 or 4 of these Seyfert 2 galaxies. Interestingly, all 6 of the `warm' ($f_{25}/f_{60} > 0.2$) optically classified Seyfert 2 galaxies in their sample showed either obscured BLRs or [Si~VI] emission at near-infrared wavelengths, as well as large Pa$\alpha$-to-infrared luminosity ratios, strongly suggesting that the screen of dust obscuring the cores of these ULIGs is optically thin at 2~\micron. In constrast, no obvious signs of an obscured BLR or strong [Si~VI] emission were detected in any of the 15 optically classified LINERs and H~II galaxies in sample of VSK. The obscured BLRs detected in the Seyfert 2 galaxies were found to have dereddened broad-line luminosities which are similar to those of optically selected quasars of comparable bolometric luminosity. This result was used by VSK to argue that most of the bolometric luminosity in these broad-line objects is powered by the same mechanism as that in optical quasars, namely mass accretion onto a supermassive black hole. The results of VSK rely on very small numbers of objects (e.g., there were only 10 optically classified Seyfert 2 galaxies in their sample). Their conclusions are therefore affected by large statistical uncertainties. For this reason and because of the success of this initial survey, another set of 39 objects from the 1-Jy sample was observed using the same techniques as VSK (cf.~Table 1). F11058--1131, a luminous infrared galaxy (LIG) with log(L$_{\rm IR}$/L$_{\odot}$) = 11.3, was also observed to test the reported detection of broad polarized emission lines in this object (Young et al. 1993). The present paper describes the results of this new study and combines them with the earlier data of VSK to produce a sample of 64 ULIGs and 1 LIG which comprises 23 optically classified Seyfert 2s, 25 LINERs, 16 H~II galaxies, and one object with undetermined optical spectral type. Section 2 of this paper describes the techniques used to acquire and analyze the new data. The results of the new survey are presented in Section 3 and combined with the results of VSK. In Section 4, the combined sample is used to test the conclusions of VSK. The conclusions from this analysis are summarized in Section 5 and combined with our recent optical spectroscopic results. This allows us to make statistically meaningful statements about the fraction of ULIGs which harbor a quasar, how this fraction varies with infrared luminosity and optical spectral type, and whether the quasars detected at optical and near-infrared wavelengths are energetically significant. A separate discussion of the emission-line properties of each object in the new sample, with particular emphasis on the presence or absence of broad components in the profiles of the infrared hydrogen recombination lines, is presented in an Appendix. We adopt H$_{\rm o}$ = 75 km s$^{-1}$ Mpc$^{-1}$ and q$_{\rm o}$ = 0 throughout this paper. \section{Observations and Data Analysis} To allow direct comparisons with the results of VSK, all of the new data were acquired, reduced and analyzed using the same procedures as VSK (see also Goodrich et al. 1994 and Veilleux, Goodrich, \& Hill 1997a). All of these data were obtained with CGS4 on UKIRT (Mountain et al. 1990). A summary of the observations is presented in Table 2. The 256 $\times$ 256 InSb array was used in conjunction with the 300 mm focal length camera and the 75 l/mm grating. The slit width and pixel scale were 1.$\arcsec$2. All of the spectra were obtained under photometric conditions. However, as described in VSK, the absolute fluxes derived from these spectra are sensitive to slit losses, seeing, centering errors, and guiding effects. The line fluxes listed in Table 3 have been corrected for slit losses (based on seeing measurements taken throughout the night) but not for the other effects. This source of errors is estimated to be less than about 50\%. A one-pixel (1.$\arcsec$2) extraction window was used to transform the long-slit data into one-dimensional spectra. \section{Results} The reduced spectra are presented in Figure 1 and discussed individually in the Appendix. Following the same format as in VSK, Table 3 lists the fluxes, equivalent widths, and line widths of the emission lines detected in these spectra. Also listed in this table are the intensities of H$\alpha$ and H$\beta$, the line widths of [O~III]~$\lambda$0.5007 and the color excess, $E(B-V)$, determined from the optical emission-line flux ratios. Unless otherwise noted, these measurements were taken from Kim et al. (1995), Kim, Veilleux, \& Sanders (1998), Veilleux et al. (1995), or Veilleux, Kim, \& Sanders (1999; hereafter VKS). The last two columns of Table 3 present the color excesses determined from the infrared emission-line flux ratios and based on the intrinsic flux ratios and extinction coefficients of Veilleux et al. (1997a; their Table 2). As found by VSK, the spectra of ULIGs are characterized by strong Pa$\alpha$ emission and weaker lines from H$_2$. Weak emission from H$_2$ 1--0 S(5)~$\lambda$1.835, 1--0 S(2)~$\lambda$2.033, and 2--1 S(3)~$\lambda$2.073, and from Br$\gamma$~$\lambda$2.166, Br$\delta$~$\lambda$1.945, Br$\epsilon$~$\lambda$1.817, and He~I~$\lambda$2.058 is also visible in some galaxies (cf. Table 3). Figure 2 shows the distributions of the intensities of H$_2$ 1--0 S(3)~$\lambda$1.958 (possibly blended with [Si~VI]~$\lambda$1.962; see below) and H$_2$ 1--0 S(1)~$\lambda$2.122 relative to narrow Pa$\alpha$ as a function of optical spectral types. Ratios from the combined set of objects in VSK and the present sample are presented. This figure indicates that the H$_2$ emission in optical H~II region-like galaxies is generally weaker than in optically classified LINERs or Seyfert 2s. The average H$_2$ 1--0 S(3)~$\lambda$1.958/Pa$\alpha$ ratios are 0.13, 0.23, and 0.32 for the H~II galaxies (14 objects), LINERs (24) and Seyfert 2 galaxies (17), respectively. The average H$_2$ 1--0 S(1)~$\lambda$2.122/Pa$\alpha$ ratios are 0.11, 0.20, and 0.20 for the H~II galaxies (8 objects), LINERs (15) and Seyfert 2 galaxies (10), respectively. Kolmogorov-Smirnov (K-S) tests confirm that the difference between the H~II galaxies and the other types of objects is significant. The H$_2$~$\lambda$1.958 and $\lambda$2.122 luminosities in the combined sample of galaxies range from 5 $\times$ 10$^6$~$L_\odot$ to 3 $\times$ 10$^8$~$L_\odot$. These luminosities translate into hot H$_2$ masses of order 10,000--500,000~$M_\odot$ if the hot H$_2$ molecules are thermalized at T = 2,000 K (Scoville et al. 1982). There are only two strong cases for broad ($\Delta V_{\rm FWHM} \ga$ 2,000~km~s$^{-1}$) line emission in the new sample of ULIGs: F05189--2524 and F13305--1739. We also detect very broad ($\Delta V_{\rm FWHM} \sim$ 7,200~km~s$^{-1}$) Pa$\alpha$ emission in the LIG F11058--1131, therefore confirming the discovery of a hidden BLR in this object from optical spectropolarimetry (Young et al. 1993). In addition, many of our objects present blue asymmetric wings at Pa$\alpha$ which can be attributed to faint broad-line emission with $\Delta V_{\rm FWHM} \approx$ 2,000--4,000~km s$^{-1}$. However, in most cases (with the possible exceptions of F04103--2838, F13443+0802~SW, F14394+5332, and F16156+0146; see A4, A28, A33, and A35, respectively), this excess emission on the blue-shifted side of Pa$\alpha$ is probably associated with faint H$_2$ emission from the 7--5 O(3)~$\lambda$1.8721 and 6--4 O(5)~$\lambda$1.8665 transitions. As discussed in Black \& van~Dishoeck (1987) and VSK, each of these lines is expected to be about 1/3--1/4 the strength of the 1--0 S(3) transition if H$_2$ is excited through fluorescence (Black \& van~Dishoeck 1987). Combining the current sample with that of VSK, we therefore find seven ULIGs with obvious obscured BLRs. All seven of these objects are optically classified as Seyfert~2 galaxies, and all but one object (F23499+2423) are `warm' galaxies with $f_{25}/f_{60} >$ 0.2 (the LIG F11058--1131 also is a Seyfert 2 galaxy with warm colors). As in VSK, the resolution of our spectra is insufficient to deblend the H$_2$ line at 1.958~\micron~from any possible [Si~VI]~$\lambda$1.962 emission. Once again, we have to rely on the fact that fluorescent excitation of molecular hydrogen at low density predicts an H$_2$ 1--0 S(3)/1--0 S(1) flux ratio less than unity (typically $\sim$~0.7; Black \& van Dishoeck 1987), while collisional excitation predicts a ratio that ranges from 0.5 to 1.4 (Shull \& Hollenbach 1978; Black \& van Dishoeck 1987; Sternberg \& Dalgarno 1989). We conservatively assume that [Si~VI]~$\lambda$1.962 is present in galaxies where we measure what appears to be H$_2$ 1--0 S(3)/1--0 S(1) flux ratios larger than 1.4. As pointed out by VSK, this lower limit is confirmed in active galaxies with measured H$_2$~$\lambda$1.958 and $\lambda$2.122 line fluxes (e.g., Marconi et al. 1994). Only one object in the new sample clearly falls in this category: the Seyfert 2 galaxy F13305--1739, one of the two ULIGs in the new sample with broad Pa$\alpha$. Two other Seyfert 2 galaxies, Mrk 273 and F13454--2956, may also present [Si~VI] emission, but the evidence is less convincing. The H$_2$ 1--0 S(3) line intensity of the LINERs F04103--2838 and F08572+3915NW appears significantly larger than that of H$_2$ 1--0 S(1) but the H$_2$ $\lambda$1.958 feature is partially blended with Br$\delta$ in both objects, making the intensity of this H$_2$ line uncertain. If it is assumed that the H$_2$ 1--0 S(3)/1--0 S(1) flux ratio in ULIGs is of order unity---as generally seems to be the case in optically selected Seyfert~2s (Marconi et al. 1994)---the [Si~VI]~$\lambda$1.962 luminosities of F13305--1739, Mrk 273, and F13454--2956 are: 15, 0.8, and 2.0 $\times$ 10$^7$~$L_\odot$ (these luminosities have been corrected for reddening using the color excesses derived from the Pa$\alpha$/H$\alpha$ ratios listed in Table 3). F13305--1739 is therefore more luminous than F12072--0444, the most powerful [Si~VI] emitter previously known (cf. VSK and Ward et al. 1991). \section{Discussion} Our new data confirm the tendency reported by VSK for the warmer Seyfert 2 galaxies in the 1-Jy sample to reveal obscured BLRs at near-infrared wavelengths. All five objects in the combined sample of ULIGs with $f_{25}/f_{60} > 0.33$ (Pks~13451+1232, F13305--1739, Mrk~463E, F20460+1925, and F23060+0505) present obvious broad Paschen emission lines (note that the LIG F11058--1131 would also fall in this category). F05189--2524, another ULIG with unambiguous broad Pa$\alpha$, is only marginally cooler than these objects ($f_{25}/f_{60}$ = 0.25). The only other object in our sample with convincing broad Pa$\alpha$, F23499+2423, has $f_{25}/f_{60}$ = 0.12. This BLR detection rate among Seyfert 2 ULIGs appears to be higher than that found by Veilleux et al. (1997a) among optically-selected Seyfert 2s. However, it is important to mention that the results on the (low-redshift) optical Seyfert 2s relied on the detection of a broad-line component to the weak Pa$\beta$ line rather than Pa$\alpha$. Consequently, one cannot directly compare the results from these two studies. Our combined dataset allows us to make stronger statements on the frequency of occurrence of nuclear activity among Seyfert 2 galaxies. Of the 22 ULIGs in the combined sample with optical Seyfert~2 spectra, 10 of them show either obvious broad Pa$\alpha$ or Pa$\beta$ emission, or strong evidence for [Si~VI]~$\lambda$1.962 emission. They are F05189--2524, F12072--0444, F13305--1739, Mrk~273, Pks 1345+12, F13454--2956, Mrk~463E, F20460+1925, F23060+0505, and F23499+2423 (recall that F11058--1131 is a LIG not a ULIG). Six additional Seyfert 2s from the combined sample may fall in this category (F08559+1053, F13443+0802~SW, F14394+5332, F16156+0146, F17179+5444, and F23233+2817), but data of higher spectral resolution are needed to confirm the presence of [Si~VI] emission in these objects or to determine the origin of the broad asymmetric wings to their Pa$\alpha$ profiles. Near-infrared signs of nuclear activity are therefore detected in 50 -- 70\% of the Seyfert 2 galaxies of the combined sample. These percentages are somewhat lower than those found by VSK. The slight difference is probably due to the larger fraction of `cool' ULIGs in the combined sample. The combined sample constitutes a more representative subset of the whole (non-Seyfert 1) 1-Jy sample in terms of {\em IRAS} colors than the original sample of VSK. The results derived from the combined surveys should therefore be used when generalizing to the entire 1-Jy sample Our near-infrared results suggest that at least half of the optically classified Seyfert 2 galaxies in the 1-Jy sample are genuine AGN. This is a lower limit to the actual number since our near-infrared method of AGN detection requires either the presence of a BLR or sufficient ionizing radiation with energies above 164 eV to produce detectable [Si~VI] $\lambda$1.962 emission. Some of the known optically selected Seyfert~2 galaxies do not present this feature (cf. Marconi et al. 1994). Note that the BLR detection rate among Seyfert~2 galaxies appears to be lower than the 100\% (5 out of 5 or 6 out of 6 if the LIG F11058--1131 is also considered) detection rate among objects with $f_{25}/f_{60}~>$ 1/3. This stringent {\em IRAS} 25-to-60 $\mu$m color criterion is therefore a better indicator of BLR activity in ULIGs than the Seyfert 2 characteristics at optical wavelengths. The high BLR detection rate among `warm' ULIGs suggests that the screen of dust in these objects is often optically thin at 2~\micron. Following VSK, we have plotted in Figure 3 the reddening-corrected Pa$\alpha$ luminosities of our sample galaxies as a function of their infrared luminosities. The reddening correction was carried out using the narrow-line extinctions listed in Table 3 and the lower limits on the broad-line extinctions derived in the Appendix. The solid line represents the relation found by Goldader et al. (1997a, 1997b) among lower luminosity infrared galaxies assuming a Pa$\alpha$/Br$\gamma$ ratio of 12 (case B recombination). A significant fraction of the data points fall below the solid line, therefore suggesting a deficit in Pa$\alpha$ emission in many ULIGs. This result is different from that of VSK, but is qualitatively similar to that of Goldader et al. (1995), who suggested that this deficit is an indication that dust obscuration is still important at 2~$\mu$m in some of these objects. This apparent discrepancy with VSK is once again due to the larger percentage of `cool' ULIGs in the combined sample. Figure 4 illustrates this point. The reddening-corrected Pa$\alpha$-to-IR luminosity ratios of our sample galaxies are plotted as a function of their {\em IRAS} $f_{25}/f_{60}$ ratio. As first pointed out by VSK, there is a clear tendency for the `warm' ULIGs to present larger Pa$\alpha$-to-IR luminosity ratios than `cool' ULIGs. Note that this tendency is only visible when the Pa$\alpha$ fluxes include the contribution from the broad-line component. Our new results therefore confirm that the {\em IRAS} $f_{25}/f_{60}$ ratio is the primary factor determining the Pa$\alpha$-to-IR luminosity ratio and the near-infrared detectability of obscured BLRs in ULIGs. `Warm' ULIGs present optical and infrared properties which are intermediate between those of `cool' ULIGs and optically selected quasars. But are the putative `buried quasars' detected in these objects powerful enough to provide the bulk of the energy output? This question was first addressed by VSK using a sample of five objects with obscured BLRs. VKS expanded this analysis by including ten luminous and ultraluminous infrared galaxies with optically detected BLRs (Seyfert 1s). Here, we further extend the analysis to include the data from the three buried quasars/AGN detected in the new sample (F05189--2524, F13305--1739, and the LIG F11058--131). The results from this new analysis are presented in Figure 5, where the dereddened emission-line luminosities of the optical and obscured BLRs in infrared galaxies and in optically selected quasars are plotted as a function of bolometric luminosity. A discussion of the methods and assumptions which were used to create this figure is presented in VKS and summarized in the caption to Figure 5. The typical uncertainties on the data points of Figure 5 are of order $\pm$ 30\%. The solid line in the figure is the best log-linear fit through the data of the optical quasars. To first order, this line therefore represents the relation found for pure broad-line AGNs\footnote{Note, however, that even quasars may not be ``pure AGN'' since part of the far-infrared/submm emission may be produced by a dusty starburst (Rowan-Robinson 1995). Here, we assume that all of the far-infrared/submm emission is powered by the central AGN (Sanders et al. 1989). The far-infrared/submm emission generally contributes about 20\% of the total bolometric luminosity of optical quasars.}. Starburst-dominated ULIGs are expected to fall below this line. Figure 5 shows that F05189--2524 falls very close to the correlation derived for optical quasars, while F11058--1131 and F13305--1739 lie less than 0.5 dex below the correlation. Our new results thus provide further support to the idea first suggested by VSK and VKS that most ($\sim$ 80\%) of the ULIGs with optical or near-infrared BLRs in the 1-Jy sample are powered predominantly by the quasar rather than by a powerful starburst. In other words, {\em the detection of an optical or near-infrared BLR in a ULIG (about 20\% of the total 1-Jy sample) appears to be an excellent sign that the AGN is the dominant energy source in that ULIG.} The good agreement in Figure 5 between optical quasars/Seyfert 1s and Seyfert 2s with obscured BLRs is not expected within the standard unification model of Seyfert galaxies. In this model, Seyfert 1s and 2s are fundamentally the same kind of objects and the viewing angle is the only parameter of importance in determining their appearance (cf., e.g., Antonucci 1993). Consequently, Seyfert 1s and 2s are expected to look the same when isotropic properties are compared. One such property is the far-infrared luminosity (since the far-infrared emission is usually assumed to be optically thin thermal emission; cf., e.g., Mulchaey et al. 1994 for a discussion of this issue). Another isotropic property is the luminosity of the BLR {\em after} correction for the extinction. The unification model therefore predicts that quasars/Seyfert 1s and Seyfert 2s should follow the same relation between the far-infrared luminosity and the extinction-corrected broad-line H$\beta$ luminosity. The agreement should break down when the bolometric luminosity is considered because the near-infrared, big-blue bump, and X-ray emission that contributes significantly to the bolometric luminosity of quasars/Seyfert 1s (but not to that of Seyfert 2s) is supposedly radiated anisotropically. This is not what we observe in Figure 5. Given that L$_{\rm IR}$ $\approx$ L$_{\rm BOL}$ in ULIGs and L$_{\rm IR}$ $<<$ L$_{\rm BOL}$ in quasars, Figure 5 implies that the BLRs of obscured Seyfert 1 ULIGs are {\em underluminous} for a given far-infrared luminosity compared to the quasars. Within the context of the unification model, one could then interpret Figure 5 as evidence that the ULIGs produce a lot more far-infrared emission than ``genuine'' AGN with the same broad-line luminosity, thus that the AGN does {\em not} dominate their energetics. Alternatively, the standard unification model may not apply to ULIGs. VKS have already pointed out that the standard unification model has difficulties explaining the larger percentage of Seyfert 1s relative to Seyfert 2s among ULIGs of higher infrared luminosities. It may be that the obscuring mass distribution in ULIGs varies with the luminosity of the energy source and with the optical spectral type. For instance, the trends with infrared luminosities can be explained if the covering factor decreases with increasing infrared luminosities (VKS). On the other hand, Figure 5 may indicate that Seyfert 2 ULIGs have much larger dust covering factors (and therefore reradiate a larger fraction of their bolometric luminosity in the far infrared) than optical quasars and Seyfert 1 ULIGs. Perhaps the obscuring material in Seyfert 2s has somehow had less time to settle into a disk than the material in optical quasars/Seyfert 1s. A recent comparison of our optical/near-infrared results with those obtained with {\em ISO} (Lutz, Veilleux, \& Genzel 1999) lends support to the scenario in which ULIGs with optical or obscured BLRs are powered predominantly by an AGN. Lutz et al. (1999) conclude that quasars constitute the dominant energy source in nearly all (10 out of the 11) optically classified Seyfert ULIGs in common between the {\em ISO} and optical samples. Strong AGN activity, once triggered, appears to quickly break the obscuring screen at least in certain directions, thus becoming detectable over a wide wavelength range. The situation in LINERs and H~II galaxies is more ambiguous. These objects represent about 70\% of the entire 1-Jy sample (VKS). None of the 41 optically classified LINERs and H~II galaxies in our near-infrared sample show any obvious signs of an obscured BLR or strong [Si~VI] emission\footnote{Broad Pa$\alpha$ may be present in the LINER F04103--2838 (cf. A4), but spectra of higher spectral resolution are needed to confirm this result. [Si~VI] $\lambda$1.962 emission may also be present in this object and the LINER F08572+3915 NW, but contamination by Br$\delta$ makes the intensity of this feature uncertain. }. This apparent lack of an energetic AGN in these objects is consistent with the recent {\it ISO} results (Genzel et al. 1998; Lutz et 1999). As discussed by VSK, two possible scenarios can explain the near-infrared results on the LINERs and H~II galaxies: (1) the cores of these ULIGs do not contain any AGN, (2) the optical depth due to dust is large enough to hide the AGN even at 2 \micron. The broad range of {\em IRAS} colors and Pa$\alpha$-to-IR luminosity ratios among LINERs and H~II galaxies (cf. Fig. 4) suggests that these objects are affected by dust screens of various optical depths. Dust obscuration is therefore unlikely to be the {\em only} reason why these objects present no obvious signs of AGN activity. \section{Conclusions} A sensitive near-infrared search for obscured broad-line regions was carried out in a set of thirty-nine ULIGs selected from the {\em IRAS} 1-Jy survey of 118 ULIGs (Kim \& Sanders 1998). The results of this survey were combined with those obtained by VSK to produce a near-infrared spectroscopic database on sixty-four ULIGs. Prior to selecting these objects, the $\sim$ 10\% of the 118 galaxies that already were known optically to show direct signs of quasar activity, i.e. optically classified as Seyfert~1, were excluded. The sixty-four objects in the combined sample constitute a representative subset of the whole (non-Seyfert 1) 1-Jy sample, in terms of redshift, infrared luminosity, and {\em IRAS} colors. The results from the present survey can therefore be generalized with a high degree of confidence to the entire 1-Jy sample. These results can be summarized as follows: \begin{itemize} \item[1.] Broad infrared recombination lines are detected in seven objects (all optically classified Seyfert 2s): F05198--2525, F13305--1739, Pks~1345+12 (= F13451+1232), Mrk~463E (= F13536+1836), F20460+1925, F23060+0505, and F23499+2423. Excess broad emission is also detected at the base of the Pa$\alpha$ profile in seven objects (all but one are optically classified Seyfert 2s): F04103--2838 (the only LINER), F08559+1053, F13443+0802~SW, F14394+5332, F16156+0146, F17179+5444, and F23233+2817. But new data are required to determine whether this emission is from H$_2$~$\lambda\lambda$1.8665,~1.8721 rather than from a genuine BLR. In addition, we confirm the spectropolarimetric discovery of a hidden BLR in the optically classified Seyfert 2 LIG F11058--1131. \item[2.] The [Si~VI] feature appears to be present in three additional objects (all optically classified Seyfert 2 galaxies), F12072--0444, Mrk~273 (= F13428+5608), and F13454--2956, and three or quite possibly four of the objects mentioned above: F13305--1739, Pks~1345+12, and F23233+2817, and perhaps F17179+5444. If these findings are confirmed by high-resolution spectroscopy, several of these objects would be among the most luminous [Si~VI] emitters presently known ($\ga$ 6 $\times$ 10$^7$~$L_\odot$). \item[3.] From results 1 and 2 above, we find that all nine `warm' ($f_{25}/f_{60} > 0.2$) optically classified Seyfert 2 galaxies in our sample of ULIGs show either obscured BLRs or [Si~VI] emission at near-infrared wavelengths. None of these objects presents deficient Pa$\alpha$-to-infrared luminosity ratios. These results support those derived from the smaller sample of VSK, and suggest that the screen of dust obscuring the cores of most `warm' Seyfert 2 ULIGs is optically thin at 2~\micron. \item[4.] No obvious signs of an obscured BLR or strong [Si~VI] emission are detected in any of the 41 optically classified LINERs and H~II galaxies in our sample. This result is consistent with recent mid-infrared {\it ISO} observations of ULIGs. The LINERs and H~II galaxies in our sample span a wide range of {\em IRAS} colors and Pa$\alpha$-to-IR luminosity ratios. The apparent lack of AGN activity in these objects is unlikely to be due solely to dust obscuration. \end{itemize} Recent results from an optical spectroscopic survey of the entire 1-Jy sample of 118 ULIGs (VKS) indicate that about 10\% of these objects are optically classified as Seyfert 1s, while about 20\% are optically classified as Seyfert 2s. Above $L_{\rm ir} = 10^{12.3}\ L_\odot$, the Seyfert 1s and 2s represent $\sim$ 26\% and 23\% of the ULIG population, respectively. Our near-infrared study of a representative subset of 64 of these ULIGs indicates that {\em at least} 50 -- 70\% (10 or possibly 16 out of 22) of the Seyfert 2s show signs of AGN activity at rest wavelenths shortward of $\sim$ 2 $\mu$m (BLR or [Si~VI] emission). The optical and near-infrared data taken together, therefore suggest that the total fraction of objects in the 1-Jy sample with signs of a bonafide AGN is {\em at least} $\sim$ 20 -- 25\%. This fraction reaches 35 -- 50\% for objects with $L_{\rm ir} > 10^{12.3}\ L_\odot$. These percentages are lower limits because the near-infrared method often fails to detect AGN activity ([Si~VI] emission) in known optically selected Seyfert 2 galaxies (Marconi et al. 1994). Comparisons of the dereddened emission-line luminosities of the optical or obscured BLRs detected in the ULIGs of the 1-Jy sample with those of optical quasars suggest that the AGN/quasar is the main source of energy in $\sim$ 80\% of all ULIGs with optical or near-infrared BLR. {\em At least} 15 -- 25\% of all ULIGs in the 1-Jy sample are therefore powered primarily by the AGN/quasar rather than by a starburst. This fraction is closer to 30 -- 50\% among ULIGs with $L_{\rm ir} > 10^{12.3}\ L_\odot$. ULIGs with powerful AGN/quasar but with no or highly obscured BLRs would increase these percentages. These results are consistent with those derived from {\em ISO} data. \acknowledgments S. V. and D. B. S. thank the organizers of the 1998 Ringberg meeting where some of the issues in this paper were discussed, and the referee, Tim Heckman, for his comments on the standard unification model of Seyfert galaxies which helped improve this paper. This research was supported in part by JPL contract no. 961566 to the University of Hawaii (D. B. S.). S. V. gratefully acknowledges the financial support of NASA through LTSA grant NAG 56547 and Hubble fellowship HF-1039.01-92A awarded by the Space Telescope Science Institute which is operated by the AURA, Inc. for NASA under contract No. NAS5--26555. This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with NASA. \clearpage \centerline{APPENDIX} \vskip 0.2in \centerline{Notes on Individual Objects} \vskip 0.1in In this Appendix the emission-line properties of each object are discussed, with particular emphasis on the presence or absence of broad components in the profiles of the infrared hydrogen recombination lines. In the cases where no positive detection of broad emission is reported, very conservative upper limits to the broad-line fluxes are given. These broad-line fluxes were determined by fitting the maximum broad Gaussian consistent with the data. \vskip 0.2in \centerline{A1. IRAS~F00188--0856} An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.6 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 7,000~km~s$^{-1}$, is derived in this optically classified LINER. \vskip 0.2in \centerline{A2. IRAS~F03250+1606} An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 1 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 6,400~km~s$^{-1}$, is derived in this optically classified LINER. \vskip 0.2in \centerline{A3. IRAS~F04074--2801} A wing is seen extending blueward from the base of Pa$\alpha$ in this optically classified LINER. A conservative fit to this wing emission gives $F$(Pa$\alpha_{\rm bl}$) $\approx$ 0.5 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$ with $\Delta V_{\rm FWHM} \approx$ 7,800~km~s$^{-1}$. However, contamination from H$_2$~$\lambda\lambda$1.8665,~1.8721 may be responsible for this feature. \vskip 0.2in \centerline{A4. IRAS~F04103--2838} The optical spectrum of this object is ambiguous, spanning the boundary between H~II galaxies and LINERs (VKS). Interestingly, this galaxy has the largest $f_{25}/f_{60}$ among all of the LINERs and H~II galaxies of our combined sample ($f_{25}/f_{60}$ = 0.30). No obvious broad Pa$\alpha$ emission is visible in the infrared spectrum of this galaxy, but the Pa$\alpha$ profile presents a clear blue wing. A conservative fit to this wing emission gives $F$(Pa$\alpha_{\rm bl}$) $\approx$ 1.5 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$ with $\Delta V_{\rm FWHM} \approx$ 3,700~km~s$^{-1}$. Contamination from H$_2$~$\lambda\lambda$1.8665,~1.8721 appears unlikely in this object considering the weaknesses of the other H$_2$ lines. The feature near 1.96 $\mu$m is considerably stronger than H$_2$ $\lambda$2.122, suggesting that [Si~VI] $\lambda$1.962 may be present in this LINER. However, contamination from Br$\delta$ makes this statement uncertain. \vskip 0.2in \centerline{A5. IRAS~F04313--1649} This is the most distant source of our sample. There is no high-resolution optical spectrum of this object that could be used to determine the optical spectral type. This galaxy is relatively faint at near-infrared wavelengths. No obvious broad Pa$\alpha$ emission is detected in this object. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.2 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$~2,300~km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A6. IRAS~F05024--1941} The optical spectrum of this object is that of a Seyfert~2 galaxy (VKS), but it has relatively `cool' {\em IRAS} 25-to-60 $\mu$m color ($f_{25}/f_{60}$ = 0.13). This object is faint at near-infrared wavelengths. No obvious broad Pa$\alpha$ emission is detected in this object within the uncertainties of our data. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.3 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$~2,500~km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A7. IRAS~F05156--3024} The optical spectrum of this object is that of a Seyfert~2 galaxy (VKS), but it presents `cool' {\em IRAS} 25-to-60 $\mu$m color ($f_{25}/f_{60}$ = 0.089). A weak wing is seen extending blueward from the base of Pa$\alpha$ in this object. A conservative fit to this wing emission gives $F$(Pa$\alpha_{\rm bl}$) $\approx$ 0.5 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$ with $\Delta V_{\rm FWHM} \approx$ 2,400~km~s$^{-1}$. However, contamination from H$_2$~$\lambda\lambda$1.8665,~1.8721 may be responsible for this feature. \vskip 0.2in \centerline{A8. IRAS~F05189--2524} This object is one of six ULIGs from the BGS sample of Sanders et al. (1988a) included in the present study. This is the warmest object of this set ($f_{25}/f_{60}$ = 0.25). The optical spectrum of this object is that of a Seyfert~2 galaxy (Veilleux et al. 1995). The K-band spectrum of this object clearly shows the presence of broad ($\Delta V_{\rm FWHM} \approx$ 2,600~km~s$^{-1}$) emission at Pa$\alpha$. This spectrum also illustrates very well the difficulty in detecting the broad component to Br$\gamma$. This explains the negative results of Goldader et al. (1995) on this object. The J-band spectrum confirms the presence of broad emission in the recombination lines. Both Pa$\beta$ and He~I $\lambda$1.0832 present broad profiles. Note, however, that the broad line parameters derived from the J-band spectrum are considered less accurate than those derived from Pa$\alpha$ because both Pa$\beta$ and He~I $\lambda$1.0832 are weaker than Pa$\alpha$, and the intensities and profiles of both lines are affected by strong atmospheric absorption features. Based on the broad Pa$\alpha$ flux listed in Table 3 and the assumption that a BLR with one-third the intensity of the observed H$\alpha$ emission would have been detected in the optical spectrum of VKS, a lower limit to the color excess in the BLR of $E(B-V)_{\rm bl} >$~2.8~mag is derived. \vskip 0.2in \centerline{A9. IRAS~F08201+2801} An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.2 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 6,600~km~s$^{-1}$, is derived in this optically classified H~II galaxy. \vskip 0.2in \centerline{A10. IRAS~F08572+3915 NW} This object is part of the original set of 10 ULIGs from the BGS sample (Sanders et al. 1988a). The optical spectrum of this relatively `warm' object ($f_{25}/f_{60}$ = 0.23) is ambiguous, spanning the boundary between H~II galaxies and LINERs (Veilleux et al. 1995; VKS). Our near-infrared spectrum does not reveal any obvious signs of an obscured AGN (BLR or [Si~VI] feature) at near-infrared wavelengths. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 1 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 6,700~km~s$^{-1}$, is derived. A comparison of the H$_2$~$\lambda$1.958 and $\lambda$2.122 fluxes suggests the presence of [Si~VI]~$\lambda$1.962 emission in this galaxy, but this conclusion is very uncertain because H$_2$~$\lambda$1.958 is clearly blended with Br$\delta$. \vskip 0.2in \centerline{A11. IRAS~F10091+4704} This powerful LINER/H~II galaxy presents very strong H$_2$ 1--0 S(3)~$\lambda$1.958 and 1--0 S(5) $\lambda$1.835 lines relative to Pa$\alpha$. Weak [Fe~II] $\lambda$1.644 is also detected in this object. The relatively poor signal-to-noise ratio of the spectrum near Pa$\alpha$ (in part due to corrections for atmospheric absorption features at those wavelengths) prevents us from making a strong statement on the presence of a BLR in this object. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.1 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 5,800~km~s$^{-1}$, is derived. We cannot make any statement on the existence of [Si~VI]~$\lambda$1.962 in this object because the critical H$_2$ line at 2.122~\micron\ is not included within the spectral range of our spectrum. \vskip 0.2in \centerline{A12. IRAS~F10190+1322} This low-redshift H~II galaxy shows no obvious signs of obscured BLRs or [Si~VI] $\lambda$1.962 emission. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.3 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 7,300~km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A13. IRAS~F10594+3818} This object is optically classified as an H~II galaxy (VKS). It shows no obvious signs of obscured BLRs or [Si~VI] $\lambda$1.962 emission. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.6 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 7,300~km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A14. IRAS~F11028+3130} Our near-infrared spectrum of this LINER is essentially featureless. An upper limit on the total Pa$\alpha$ flux of $F$(Pa$\alpha$) $<$~0.04 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 500~km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A15. IRAS~F11095--0237} The optical spectrum of this object is that of a LINER. Our infrared spectrum does not show any evidence for broad Pa$\alpha$ or [Si~VI] $\lambda$1.962 emission. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.6 $\times$ 10$^{-14}$ ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 6,500~km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A16. IRAS~F11180+1623} There is no obvious signs of broad emission in the Pa$\alpha$ profile of this LINER/H~II galaxy. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.2 $\times$ 10$^{-14}$ ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 6,500~km~s$^{-1}$, is derived. We cannot make any strong statement on the existence of [Si~VI]~$\lambda$1.962 in this object because the H$_2$ $\lambda$2.122 feature is close to the edge of the spectrum and is strongly affected by atmospheric absorption features. \vskip 0.2in \centerline{A17. IRAS~F11223--1244} There is no obvious signs for broad Pa$\alpha$ emission in this `cool' ($f_{25}/f_{60}$ = 0.11) Seyfert 2 galaxy. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.4 $\times$ 10$^{-14}$ ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 7,500~km~s$^{-1}$, is derived. We cannot make any strong statement on the existence of [Si~VI] $\lambda$1.962 in this galaxy because our spectrum does not extend longward enough in wavelength to include the critical H$_2$~$\lambda$2.122 line. An upper limit to the broad Pa$\alpha$ flux of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.3 $\times$ 10$^{-14}$~erg~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 5,000~km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A18. IRAS~F11506+1331} The optical spectrum of this object is ambiguous, spanning the boundary between H~II galaxies and LINERs (VKS). The near-infrared spectrum is dominated by the strong Pa$\alpha$ feature. The relatively weak H$_2$ lines are typical of those observed in optically classified H~II galaxies (cf.~Fig.~1). There is no obvious signs of AGN activity in this galaxy (BLR or [Si~VI] $\lambda$1.962 feature). An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 1 $\times$ 10$^{-14}$ ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 6,300~km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A19. IRAS~F11582+3020} No broad Pa$\alpha$ emission is detected in this LINER within the uncertainties of our measurements. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.2 $\times$ 10$^{-14}$ ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 3,600~km~s$^{-1}$, is derived. The spectral range of our spectrum prevents us from measuring the strength of H$_2$~$\lambda$2.122 and comparing it with that of H$_2$~$\lambda$1.958. The feature at 1.96~\micron\ is also too faint for detailed profile analysis. Therefore, no statement can be made on the existence of [Si~VI] $\lambda$1.962 in this galaxy. \vskip 0.2in \centerline{A20. IRAS~F12018+1941} The optical spectral type of this galaxy (LINER) was derived using the measurements of Armus et al. (1989). There is no evidence for [Si~VI] $\lambda$1.962 or broad Pa$\alpha$ emission in this object. The continuum near Pa$\alpha$ appears slightly convex, but this may be due to contaminants redward of Pa$\alpha$. A very conservative upper limit to the broad Pa$\alpha$ flux in this galaxy was derived by assuming that the convexity in the continuum is due entirely to broad Pa$\alpha$ emission: $F$(Pa$\alpha_{\rm bl}$) $<$ 0.5 $\times$ 10$^{-14}$ ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 8,300~km~s$^{-1}$. \vskip 0.2in \centerline{A21. IRAS~F12032+1707} This object shares many resemblances with F10091+4704 (cf.~A11). Both of these galaxies are powerful (log[L$_{\rm IR}$/L$_\odot$] $>$ 12.5) ULIGs with LINER characteristics at optical wavelengths and with strong H$_2$ emission lines relative to Pa$\alpha$. The [Fe~II] $\lambda$1.644 feature is also detected in both galaxies, and none of these objects show any obvious signs for broad Pa$\alpha$. The quality of our spectrum of F12032+1707 is slightly better than that of F10091+4704, therefore allowing us to derive a strong upper limit to the flux from broad Pa$\alpha$: $F$(Pa$\alpha_{\rm bl}$) $<$ 0.6 $\times$ 10$^{-14}$ ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 7,800~km~s$^{-1}$. No statement can be made about the presence of [Si~VI] $\lambda$1.962 in this object because the H$_2$ $\lambda$2.122 is redshifted out of our spectral range. \vskip 0.2in \centerline{A22. IRAS~F12112+0305} The optical spectrum of this BGS ULIG shows the signatures of a LINER. The relatively low redshift of this object shifts the Pa$\alpha$ line into a region of the spectrum which is affected by atmospheric absorption features. Nevertheless, the excellent correction for these features (as evident from the flat and featureless continuum in this region) allows us to derive a strong upper limit to the flux from broad Pa$\alpha$: $F$(Pa$\alpha_{\rm bl}$) $<$ 0.7 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$ assuming $\Delta V_{\rm FWHM} \approx$ 6,000~km~s$^{-1}$. The H$_2$ lines in this object are all very faint relative to Pa$\alpha$. \vskip 0.2in \centerline{A23. IRAS~F12447+3721} The infrared spectrum of this H~II galaxy is similar to that of F15206+3342, another H~II galaxy in the sample of VSK: strong narrow emission from Pa$\alpha$ is visible superposed on a faint continuum; the lines from H$_2$ are barely or not detected. An upper limit to the broad Pa$\alpha$ flux of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.6 $\times$ 10$^{-14}$~erg~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 2,700~km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A24. IRAS~F13106--0922} The optical spectrum of this object is ambiguous, spanning the boundary between H~II galaxies and LINERs (VKS). No obvious signs of [Si~VI] $\lambda$1.962 or broad Pa$\alpha$ emission are detected in this object. An upper limit to the broad Pa$\alpha$ flux of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.3 $\times$ 10$^{-14}$~erg~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 5,000~km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A25. IRAS~F13305--1739} The infrared spectrum of this `warm' ($f_{25}/f_{60}$ = 0.34) Seyfert~2 galaxy presents broad ($\Delta V_{\rm FWHM} \approx$ 3,000~km~s$^{-1}$) Pa$\alpha$ emission but very little H$_2$ emission. Based on the broad Pa$\alpha$ flux listed in Table 3 and the assumption that a BLR with one-third the intensity of the observed H$\alpha$ emission would have been detected in the optical spectrum of VKS, a lower limit to the color excess in the BLR of $E(B-V)_{\rm bl} >$~0.8~mag is derived. Comparison of the H$_2$~$\lambda$1.958 and $\lambda$2.122 fluxes also suggests the presence of [Si~VI]~$\lambda$1.962 emission in this galaxy. \vskip 0.2in \centerline{A26. Mrk~273 = IRAS~F13428+5608} This relatively `cool' ($f_{25}/f_{60}$ = 0.10) ULIG from the original BGS sample (Sanders et al. 1988a) show Seyfert 2 characteristics at optical wavelengths (Kim et al. 1998). Our J- and K-band spectra of this object show no clear evidence for broad emission. The continuum near Pa$\alpha$ appears slightly convex, but this may be due to our correction for atmospheric absorption in this spectral region. A very conservative upper limit to the broad Pa$\alpha$ flux in this galaxy was derived by assuming that the convexity in the continuum is due entirely to broad Pa$\alpha$ emission: $F$(Pa$\alpha_{\rm bl}$) $<$ 3.7 $\times$ 10$^{-14}$ ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 7,600~km~s$^{-1}$. An upper limit to the broad Pa$\beta$ and He~I $\lambda$1.083 fluxes was also derived from the J-band spectrum: $F$(He~I$_{\rm bl}$) $<$ 1.5 $\times$ 10$^{-14}$ ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 7,500~km~s$^{-1}$. Comparison of the H$_2$~$\lambda$1.958 and $\lambda$2.122 fluxes suggests the presence of [Si~VI]~$\lambda$1.962 emission in this galaxy, but the evidence is marginal because the H$_2$ $\lambda$1.958/H$_2$ $\lambda$2.122 flux ratio is only $\sim$ 1.4 and contamination from Br$\delta$ makes the flux of H$_2$ $\lambda$1.958 uncertain. \vskip 0.2in \centerline{A27. IRAS~F13443+0802 NE} The infrared spectrum of this H~II galaxy presents very weak H$_2$ emission lines and does not show any evidence for broad Pa$\alpha$ emission. An upper limit to the broad Pa$\alpha$ flux of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.5 $\times$ 10$^{-14}$~erg~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$ 4,800~km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A28. IRAS~F13443+0802 SW} This is the galaxy in our sample with the lowest apparent Pa$\alpha$ flux. Nevertheless, the S/N in the continuum is sufficient to put strong constraints on any broad component to Pa$\alpha$. A weak blue wing is detected at the base of Pa$\alpha$ in this Seyfert. A Gaussian fit to this feature gives $F$(Pa$\alpha_{\rm bl}$) $\approx$ 0.2 $\times$ 10$^{-14}$ erg~s$^{-1}$~cm$^{-2}$ with a $\Delta V_{\rm FWHM} \approx$ 2,400~km~s$^{-1}$. Part of this feature may be attributed to H$_2$~$\lambda\lambda$1.8665,~1.8721. The line profile of H$_2$~$\lambda$1.957 appears unusually broad ($\Delta V_{\rm FWHM} \approx$ 1,250~km~s$^{-1}$), but it is uncertain. The signal-to-noise of our spectrum near H$_2$ $\lambda$2.122 is rather poor and prevents us from making a strong statement on the relative intensity of H$_2$ $\lambda$1.958 and H$_2$ $\lambda$2.122. \vskip 0.2in \centerline{A29. IRAS~F13454--2956} This Seyfert 2 galaxy has an unusually `cool' 25-to-60 $\mu$m IRAS flux ratio of only 0.03. There is no obvious signs of broad emission in the profile of Pa$\alpha$. The excellent signal-to-noise ratio of our near-infrared spectrum allows us to put strong constraints on any broad component to Pa$\alpha$: $F$(Pa$\alpha_{\rm bl}$) $\approx$ 0.4 $\times$ 10$^{-14}$ erg~s$^{-1}$~cm$^{-2}$, assuming a $\Delta V_{\rm FWHM} \approx$ 6,600~km~s$^{-1}$. Comparison of the H$_2$~$\lambda$1.958 and $\lambda$2.122 fluxes suggests the presence of [Si~VI]~$\lambda$1.962 emission in this galaxy. \vskip 0.2in \centerline{A30. IRAS~F13509+0442} The near-infrared spectrum of this H~II galaxy presents strong Pa$\alpha$ but weak H$_2$. No obvious broad Pa$\alpha$ emission is visible in our spectrum. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.8 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$~4,000 km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A31. IRAS~F14053--1958} The only near-infrared emission line detected in this Seyfert~2 galaxy is narrow Pa$\alpha$. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.2 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$~2,300 km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A32. IRAS~F14070+0525} This object is the most powerful ULIG in our sample. The infrared spectrum of this Seyfert~2 galaxy is dominated by narrow Pa$\alpha$. The [Fe~II] $\lambda$1.644 feature is also present at considerably fainter flux levels. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.2 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$~3,800 km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A33. IRAS~F14394+5332} Several emission lines from H~I and H$_2$ are detected in the near-infrared spectrum of this Seyfert 2 galaxy. A weak blue wing is detected at the base of Pa$\alpha$. A Gaussian fit to this feature gives $F$(Pa$\alpha_{\rm bl}$) $\approx$ 2 $\times$ 10$^{-14}$ erg~s$^{-1}$~cm$^{-2}$ with a $\Delta V_{\rm FWHM} \approx$ 6,300~km~s$^{-1}$. Part of this feature may be attributable to H$_2$~$\lambda\lambda$1.8665,~1.8721. There is no evidence for strong [Si~VI] emission in this object. \vskip 0.2in \centerline{A34. IRAS~F15250+3609} The optical spectrum of this BGS ULIG is ambiguous, straddling the boundary between H~II galaxies and LINERs (VKS). There is no obvious signs for an obscured broad-line region or strong [Si~VI] $\lambda$1.962 emission in this galaxy. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 1.4 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$~5,500 km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A35. IRAS~F16156+0146} A blue wing is detected at the base of Pa$\alpha$ in this optical Seyfert 2. A Gaussian fit to this feature gives $F$(Pa$\alpha_{\rm bl}$) $\approx$ 0.7 $\times$ 10$^{-14}$ erg~s$^{-1}$~cm$^{-2}$ with a $\Delta V_{\rm FWHM} \approx$ 5,600~km~s$^{-1}$. Part of this feature may be attributable to H$_2$~$\lambda\lambda$1.8665,~1.8721. \vskip 0.2in \centerline{A36. IRAS~F16300+1558} This optically classified LINER presents strong narrow Pa$\alpha$ and weak H$_2$ $\lambda$1.958 and [Fe~II] $\lambda$1.644. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 0.3 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$~3,500 km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A37. IRAS~F17208--0014} This BGS ULIG is optically classified as an H~II galaxy (Veilleux et al. 1995). There is no obvious broad Pa$\alpha$ emission in this object within the uncertainties of the data. Errors in correcting for the presence of atmospheric absorption features near Pa$\alpha$ increases the noise in this region. A very conservative Gaussian fit of the emission under narrow Pa$\alpha$ gives $F$(Pa$\alpha_{\rm bl}$) $\approx$ 2 $\times$ 10$^{-14}$ erg~s$^{-1}$~cm$^{-2}$ with a $\Delta V_{\rm FWHM} \approx$ 4,700~km~s$^{-1}$. \vskip 0.2in \centerline{A38. IRAS~F22491--1808} This H~II galaxy is part of the original BGS sample of ULIGs (Sanders et al. 1988). The near-infrared spectrum of this object presents only very weak H$_2$ emission and no evidence for strong [Si~VI] or broad Pa$\alpha$ emission within the uncertainties of the data. Errors in correcting for the presence of atmospheric absorption features near Pa$\alpha$ increases the noise in this region. A very conservative Gaussian fit of the emission under narrow Pa$\alpha$ gives $F$(Pa$\alpha_{\rm bl}$) $\approx$ 0.4 $\times$ 10$^{-14}$ erg~s$^{-1}$~cm$^{-2}$ with a $\Delta V_{\rm FWHM} \approx$ 6,400~km~s$^{-1}$. \vskip 0.2in \centerline{A39. IRAS~F23365+3604} The infrared spectrum of this optical LINER shows no evidence for an obscured broad-line region or strong [Si~VI] $\lambda$1.962 emission. An upper limit of $F$(Pa$\alpha_{\rm bl}$) $<$ 1 $\times$ 10$^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, assuming $\Delta V_{\rm FWHM} \approx$~3,300 km~s$^{-1}$, is derived. \vskip 0.2in \centerline{A40. IRAS~F11058--1131 (LIG)} Optical spectropolarimetry of this luminous infrared galaxy by Young et al. (1993) reveals a hidden BLR, seen in scattered, polarized light. Young et al. measure $\Delta V_{\rm FWHM} \approx$ 7,600~km~s$^{-1}$ and $\Delta V_{\rm FWZI} \approx$ 16,800~km~s$^{-1}$ for polarized H$\alpha$. The infrared spectrum of this Seyfert 2 galaxy is equally stunning: broad ($\Delta V_{\rm FWHM} \approx$ 7,200 ~km~s$^{-1}$) wings are clearly visible at Pa$\alpha$, extending to $\sim$ $\pm$ 10,000~km~s$^{-1}$ at zero intensity and beautifully confirming the results of Young et al. (1993). The ratio of Pa$\alpha_{\rm bl}$ emission to scattered H$\alpha_{\rm bl}$ emission is $\sim$ 8, nearly two orders of magnitude larger than the value predicted from case B recombination. If we make the rather safe assumption that a BLR with one-third the intensity of the observed H$\alpha$ emission would have been detected in the optical spectrum of Osterbrock \& de Robertis (1985), we derive a lower limit to the color excess towards the BLR based on Pa$\alpha_{\rm bl}$/H$\alpha_{\rm nl}$ of $E(B-V)_{\rm bl}$ $>$ 1.4. \clearpage \input{table1_v4.tex} \input{table2_v4.tex} \input{table3_v4.tex} \clearpage \normalsize
\section{#1}} \renewcommand{\theequation}{\thesection.\arabic{equation}} \title{\large \bf Drinfeld Basis And a Nonclassical Free Boson Representation of Twisted Quantum Affine Superalgebra ${U_q[osp(2|2)^{(2)}]}$} \author{{\large Wen-Li Yang~$^1$ and Yao-Zhong Zhang~$^2$}\\ \\ {\small $^1$ Institute of Modern Physics, Northwest University, Xian 710069, China}\\ {\small $^2$ Department of Mathematics, University of Queensland, Brisbane, Qld 4072, Australia}} \date{} \maketitle \vspace{10pt} \begin{abstract} We derive from the super RS algebra the Drinfeld basis of the twisted quantum affine superalgebra ${U_q[osp(2|2)^{(2)}]}$ by means of the Gauss decomposition technique. We explicitly construct a nonclassical level-one representation of ${U_q[osp(2|2)^{(2)}]}$ in terms of two $q$-deformed free boson fields. \vskip 2cm \noindent{\bf Keywords:} Drinfeld basis, quantum affine superalgebras, free boson representation. \end{abstract} \vspace{10pt} \sect{Introduction\label{intro}} The algebraic analysis method based on infinite dimensional non-abelian symmetries has proved extremely successful in both formulating and solving lower dimensional integrable systems \cite{Is88}\cite{Jm94}. The key elements of this method are infinite-dimensional highest weight representations of quantum (super) algebras and vertex operators. Drinfeld bases of quantum (super) algebras are of great importance in constructing the infinite-dimensional representations and vertex operators.The Drinfeld bases of quantum affine bosonic algebras were given by Drinfeld \cite{Dri88}. For the case of quantum affine superalgebras, the Drinfeld bases have been known only for ${U_q[gl(m|n)^{(1)}]}$ \cite{Yam96,Zha97,Cai98} and ${U_q[osp(1|2)^{(1)}]}$ \cite{Gou98,Din99}. In this paper, we derive the Drinfeld basis of the twisted quantum affine superalgebra ${U_q[osp(2|2)^{(2)}]}$ by using the super version of the Reshetikhin-Semenov-Tian-Shansky (RS) algebra \cite{Res90} the Gauss decomposition techinque of Ding-Frenkel \cite{Din93}. Moreover, we explicitly construct a nonclassical level-one representation of ${U_q[osp(2|2)^{(2)}]}$ by means of two $q$-deformed free boson fields. \section{Drinfeld Basis of ${U_q[osp(2|2)^{(2)}]}$} The symmetric Cartan matrix of the twisted affine superalgebra $osp(2|2)^{(2)}$ is $(a_{ij})$ with $a_{11}=a_{22}=1,~ a_{12}=a_{21}=-1$. Twisted quantum affine superalgebra ${U_q[osp(2|2)^{(2)}]}$ is a $q$-analogue of the universal enveloping algebra of $osp(2|2)^{(2)}$ and is generated by the Chevalley generators $\{E_i,~F_i,~K^{\pm 1}_i | i=0,1\}$. The ${\bf Z}_2$-grading of the Chevalley generators is $[E_i]=[F_i]=1,~ [K_i]=0,~ i=0,1$. The defining relations are \begin{eqnarray} &&K_iK_j=K_jK_i,\nonumber\\ &&K_jE_iK^{-1}_j=q^{a_{ij}}E_i,~~~~ K_jF_iK^{-1}_j=q^{-a_{ij}}F_i,\nonumber\\ &&\{E_i,F_j\}=\delta_{ij}\frac{K_i-K^{-1}_i}{q-q^{-1}} \end{eqnarray} plus $q$-Serre relations which we omit. Here $\{X,Y\}\equiv XY+YX$. ${U_q[osp(2|2)^{(2)}]}$ is a ${\bf Z}_2$-graded quasi-triangular Hopf algebra endowed with the coproduct $\Delta$, counit $\epsilon$ and antipode $S$ given by \begin{eqnarray} &&\Delta(E_i)=E_i\otimes 1+K_i\otimes E_i,~~~~ \Delta(F_i)=F_i\otimes K^{-1}_i+1\otimes F_i,\nonumber\\ &&\Delta(K_i)=K_i\otimes K_i,~~~~\epsilon(E_i)=\epsilon(F_i)=0,~~~~\epsilon(K_i)=1,\nonumber\\ &&S(E_i)=-K^{-1}_iE_i,~~~~ S(F_i)=-F_i K_i,~~~~ S(K_i)=K^{-1}_i,~~~i=0,1. \end{eqnarray} The antipode $S$ is a ${\bf Z}_2$-graded algebra anti-homomorphism, i.e. for homogeneous elements $a,b\in {U_q[osp(2|2)^{(2)}]}$, we have $S(ab)=(-1)^{[a][b]}S(b)S(a)$. We now give our definition of ${U_q[osp(2|2)^{(2)}]}$ in terms of Drinfeld generators. \begin{Definition} ${U_q[osp(2|2)^{(2)}]}$ is an associative algebra with unit 1 and the Drinfeld generators: $X^\pm(z)$ and $\psi^\pm(z)$, a central element $c$ and a nonzero complex parameter $q$. $\psi^\pm(z)$ are invertible. The gradings of the generators are: $[X^\pm(z)]=1$ and $[\psi^\pm(z)]=[c]=0$. The relations are given by \begin{eqnarray} &&\psi^\pm(z)\psi^\pm(w)=\psi^\pm(w)\psi^\pm(z),\nonumber\\ &&\psi^+(z)\psi^-(w)=\frac{(z_+q+w_-)(z_-+w_+q)} {(z_++w_-q)(z_-q+w_+)}\psi^-(w)\psi^+(z),\nonumber\\ &&\psi^\pm(z)^{-1}X^+(w)\psi^\pm(z)=\frac{z_\pm+w q}{z_\pm q+w} X^+(w),\nonumber\\ &&\psi^\pm(z)X^-(w)\psi^\pm(z)^{-1}=\frac{z_\mp+w q}{z_\mp q+w} X^-(w),\nonumber\\ &&(z+wq^{\pm 1})X^\pm(z)X^\pm(w)+(zq^{\pm 1}+w)X^\pm(w)X^\pm(z) =0,\nonumber\\ &&\{X^+(z), X^-(w)\}=\frac{1}{(q-q^{-1})zw}\left[ \delta(\frac{w}{z}q^{c})\psi^+(w_+) -\delta(\frac{w}{z}q^{-c})\psi^-(z_+)\right].\label{current} \end{eqnarray} \end{Definition} Expand the currents in the form \begin{eqnarray} &&X^\pm(z)=\sum_{n\in{\bf Z}} X^\pm_n\;z^{-n-1},\nonumber\\ &&\psi^\pm(z)=\sum_{n\in{\bf Z}}\psi^\pm_n\;z^{-n} =K^{\pm 1}\exp\left(\pm(q-q^{-1})\sum_{n>0}H_{\pm n}z^{\mp n}\right). \end{eqnarray} In terms of modes $\{H_n\,|\,n\in {\bf Z}-\{0\},~ X^\pm_n\,|\,n\in{\bf Z}\}$ and $K$, the defining relations of ${U_q[osp(2|2)^{(2)}]}$ are given by \begin{eqnarray} &&c:~{\rm central~ element},\nonumber\\ &&[K, H_n]=0,~~~~KX^\pm_n K^{-1}=q^{\pm 1}X^\pm_n,\nonumber\\ &&[H_n, H_m]=\delta_{n+m,0}(-1)^n\frac{[n]_q[nc]_q}{n},~~~n\neq 0,\nonumber\\ &&[H_n,X^\pm_m]=\pm\frac{1}{n}(-1)^n[n]_q q^{\mp \frac{c}{2}|n|}X^\pm_{n+m},~~~n\neq 0,\nonumber\\ &&X^\pm_{n+1}X^\pm_m+q^{\pm 1}\,X^\pm_mX^\pm_{n+1} +q^{\pm 1}\,X^\pm_nX^\pm_{m+1}+X^\pm_{m+1}X^\pm_n=0,\nonumber\\ &&\{X^+_n, X^-_m\}=\frac{1}{q-q^{-1}}\left(q^{\frac{c}{2}(n-m)} \psi^+_{n+m}-q^{\frac{c}{2}(m-n)}\psi^-_{n+m}\right).\label{mode} \end{eqnarray} Here and throughout $[i]_q=(q^i-q^{-i})/(q-q^{-1})$. The Chevalley generators are obtained by the formulae: \begin{eqnarray} &&K_1=K,~~~~E_1=X^+_0,~~~~F_1=X^-_0,\nonumber\\ &&K_0=q^cK^{-1},~~~~E_0=X^-_1K^{-1},~~~~F_0=-KX^+_{-1}.\label{simple} \end{eqnarray} In terms of the Chevalley generators, the Drinfeld generators can be built up recursively by \begin{eqnarray} &&H_1=q^\frac{c}{2}K^{-1}(X^+_0X^-_1+X^-_1X^+_0),\nonumber\\ &&H_{-1}=q^{-\frac{c}{2}}K(X^+_{-1}X^-_0+X^-_0X^+_{-1}),\nonumber\\ &&X^\pm_{n+1}=\mp q^{\pm \frac{c}{2}}[H_1,X^\pm_n],~~~~ X^\pm_{-n-1}=\mp q^{\pm\frac{c}{2}}[H_{-1},X^\pm_{-n}],~~n\geq 0 \end{eqnarray} plus the formulae for $H_n,~H_{-n}~(n>0)$ given as follows \begin{equation} H_{\pm n}=\pm\frac{1}{q-q^{-1}}\sum_{p_1+2p_2+\cdots+np_n=n} \frac{(-1)^{\sum\,p_i-1} \;(\sum\,p_i-1)!} {p_1!\cdots p_n!}(K^{\mp 1}\psi^\pm_{\pm 1})^{p_1} \cdots (K^{\mp 1}\psi^\pm_{\pm n})^{p_n},\\ \end{equation} where \begin{equation} \psi^\pm_{\pm n}=\pm(q-q^{-1})q^{\mp\frac{c}{2}(n-2)} \{X^\pm_{\pm n\mp 1},X^\mp_{\pm 1}\},~~n>0. \end{equation} \section{Derivation of Drinfeld Basis from Super RS Algebra} Let us recall the definition of the super RS algebra. Let $R(z)\in End(V\otimes V)$, where $V$ is a ${\bf Z}_2$ graded vector space, be a R-matrix which satisfies the graded Yang-Baxter equation \begin{equation}\label{rrr} R_{12}(z)R_{13}(zw)R_{23}(w)=R_{23}(w)R_{13}(zw)R_{12}(z). \end{equation} We introduce \cite{Zha97,Gou98} \begin{Definition}\label{rs}: The super RS algebra $U({\cal R})$ is generated by invertible L-operators $L^\pm(z)$, which obey the relations \begin{eqnarray} R({z\over w})L_1^\pm(z)L_2^\pm(w)&=&L_2^\pm(w)L_1^\pm(z)R({z\over w}),\nonumber\\ R({z_+\over w_-})L_1^+(z)L_2^-(w)&=&L_2^-(w)L_1^+(z)R({z_-\over w_+}),\label{super-rs} \end{eqnarray} where $L_1^\pm(z)=L^\pm(z)\otimes 1$, $L_2^\pm(z)=1\otimes L^\pm(z)$ and $z_\pm=zq^{\pm {c\over 2}}$. For the first formula of (\ref{super-rs}), the expansion direction of $R({z\over w})$ can be chosen in $z\over w$ or $w\over z$, but for the second formula, the expansion direction must only be in $z\over w$. \end{Definition} The multiplication rule for the tensor product is defined by \begin{equation} (a\otimes b)(a'\otimes b')=(-1)^{[b][a']}\,(aa'\otimes bb'), \end{equation} for homogeneous elements $a,~ b,~ a'$, $b'$ of ${U_q[osp(2|2)^{(2)}]}$. In the following we apply the super RS algebra to derive the Drinfeld basis of ${U_q[osp(2|2)^{(2)}]}$. We take $R({z\over w})$ to be the R-matrix associated to the 3-dimensional representation $V$ of ${U_q[osp(2|2)^{(2)}]}$. Let $v_1,~v_2,~v_3$ be the basis vectors of $V$ with the ${\bf Z}_2$-grading $[v_1]=[v_3]=0$ and $[v_2]=1$. It can be shown that the R-matrix has the following form: \begin{equation} R({z\over w})=\left( \begin{array}{ccccccccc} 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ 0 & a & 0 & b & 0 & 0 & 0 & 0 & 0\\ 0 & 0 & d & 0 & c & 0 & r & 0 & 0\\ 0 & f & 0 & a & 0 & 0 & 0 & 0 & 0\\ 0 & 0 & g & 0 & e & 0 & c & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & a & 0 & b & 0\\ 0 & 0 & s & 0 & g & 0 & d & 0 & 0\\ 0 & 0 & 0 & 0 & 0 & f & 0 & a & 0\\ 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \end{array} \right), \label{r12} \end{equation} where \begin{eqnarray} &&a=\frac{q(z-w)}{zq^2-w},~~~~b=\frac{w(q^2-1)}{zq^2-w},~~~~ c=-\frac{q^{1/2}w(q^2-1)(z-w)}{(zq^2-w)(zq+w)},\nonumber\\ &&d=\frac{q(z-w)(z+qw)}{(zq^2-w)(zq+w)},~~~~ e=a-\frac{zw(q^2-1)(q+1)}{(zq^2-w)(zq+w)},\nonumber\\ &&f=\frac{z(q^2-1)}{zq^2-w},~~~~ g=-\frac{q^{1/2}z(q^2-1)(z-w)}{(zq^2-w)(zq+w)},\nonumber\\ &&r=\frac{w^2(q-1)(q+1)^2}{(zq^2-w)(zq+w)},~~~~ s=\frac{z^2(q-1)(q+1)^2}{(zq^2-w)(zq+w)}. \end{eqnarray} As in the non-super case \cite{Din93}, $L^\pm(z)$ allow the unique Gauss decomposition \begin{equation} L^\pm(z)=\left ( \begin{array}{ccc} 1 & 0 & 0\\ e^\pm_1(z) & 1 & 0\\ e^\pm_{3,1}(z) & e^\pm_2(z) & 1 \end{array} \right ) \left ( \begin{array}{ccc} k^\pm_1(z) & 0 & 0\\ 0 & k^\pm_2(z) & 0\\ 0 & 0 & k^\pm_3(z) \end{array} \right ) \left ( \begin{array}{ccc} 1 & f^\pm_1(z) & f^\pm_{1,3}(z) \\ 0 & 1 & f^\pm_2(z)\\ 0 & 0 & 1 \end{array} \right ),\label{l+-} \end{equation} where $e^\pm_{i,j}(z),~f^\pm_{j,i}(z)$ and $k^\pm_i(z) ~(i>j)$ are elements in the super RS algebra and $k^\pm_i(z)$ are invertible; and $e_i^\pm(z)\equiv e^\pm_{i+1,i}(z),~f_i^\pm(z)\equiv f^\pm_{i,i+1}(z)$. Let us define \begin{eqnarray} X^+_i(z)&=&f^+_{i}(z_+)-f^-_{i}(z_-),\nonumber\\ X^-_i(z)&=&e^-_{i}(z_+)-e^+_{i}(z_-). \end{eqnarray} By the definition of the super RS algebra and the Gauss decomposition formula (\ref{l+-}), and after tedious calculations parallel to those of the $U_q[osp(1|2)^{(1)}]$ case, we arrive at \begin{eqnarray} k^\pm_i(z)k^\pm_j(w)&=&k^\pm_j(w)k^\pm_i(z),~~i,j=1,2,3,\nonumber\\ k^+_1(z)k^-_1(w)&=&k^-_1(w)k^+_1(z),\nonumber\\ k^+_3(z)k^-_3(w)&=&k^-_3(w)k^+_3(z),\nonumber\\ \frac{z_\pm-w_\mp}{z_\pm q^2-w_\mp} k^\pm_1(z)k^\mp_2(w)&=& \frac{z_\mp-w_\pm}{z_\mp q^2-w_\pm} k^\mp_2(w)k^\pm_1(z),\nonumber\\ \frac{(z_\mp-w_\pm)(z_\mp+w_\pm q)}{(z_\mp q^2-w_\pm)(z_\mp q+w_\pm)} k^\pm_1(z)k^\mp_3(w)^{-1}&=& \frac{(z_\pm-w_\mp)(z_\pm+w_\mp q)}{(z_\pm q^2-w_\mp)(z_\pm q+w_\mp)} k^\mp_3(w)^{-1}k^\pm_1(z),\nonumber\\ \frac{(z_\pm-w_\mp q^2)(z_\pm q+w_\mp)}{(z_\pm q^2-w_\mp)(z_\pm+w_\mp q)}k^\pm_2(z)k^\mp_2(w)&=& \frac{(z_\mp-w_\pm q^2)(z_\mp q+w_\pm)}{(z_\mp q^2-w_\pm) (z_\mp+w_\pm q)}k^\mp_2(w)k^\pm_2(z),\nonumber\\ \frac{z_\pm-w_\mp}{z_\pm q^2-w_\mp} k^\pm_2(z)^{-1}k^\mp_3(w)^{-1}&=& \frac{z_\mp-w_\pm}{z_\mp q^2-w_\pm} k^\mp_3(w)^{-1}k^\pm_2(z)^{-1},\nonumber \end{eqnarray} \begin{eqnarray} k^\pm_1(z)X^-_1(w)k^\pm_1(z)^{-1}&=&\frac{z_\pm q^2-w}{q(z_\pm-w)} X^-_1(w),\nonumber\\ k^\pm_1(z)^{-1}X^+_1(w)k^\pm_1(z)&=&\frac{z_\mp q^2-w}{q(z_\mp-w)} X^+_1(w),\nonumber\\ k^\pm_2(z)X^-_1(w)k^\pm_2(z)^{-1}&=&\frac{(z_\pm q^2-w)(z_\pm+wq)} {q(z_\pm-w)(z_\pm q+w)}X^-_1(w),\nonumber\\ k^\pm_2(z)^{-1}X^+_1(w)k^\pm_2(z)&=&\frac{(z_\mp q^2-w)(z_\mp+wq)} {q(z_\mp-w)(z_\mp q+w)}X^+_1(w),\nonumber\\ k^\pm_3(z)X^-_1(w)k^\pm_3(z)^{-1}&=&\frac{z_\pm+wq}{z_\pm q+w} X^-_1(w),\nonumber\\ k^\pm_3(z)^{-1}X^+_1(w)k^\pm_3(z)&=&\frac{z_\mp+wq}{z_\mp q+w} X^+_1(w),\nonumber\\ k^\pm_1(z)X^-_2(w)k^\pm_1(z)^{-1}&=&\frac{z_\pm q+w}{z_\pm+wq} X^-_2(w),\nonumber\\ k^\pm_1(z)^{-1}X^+_2(w)k^\pm_1(z)&=&\frac{z_\mp q+w}{z_\mp+wq} X^+_2(w),\nonumber\\ k^\pm_2(z)X^-_2(w)k^\pm_2(z)^{-1}&=&\frac{(z_\pm-wq^2)(z_\pm q+w)} {q(z_\pm-w)(z_\pm+wq)}X^-_2(w),\nonumber\\ k^\pm_2(z)^{-1}X^+_2(w)k^\pm_2(z)&=&\frac{(z_\mp-wq^2)(z_\mp q+w)} {q(z_\mp-w)(z_\mp+wq)}X^+_2(w),\nonumber\\ k^\pm_3(z)X^-_2(w)k^\pm_3(z)^{-1}&=&\frac{z_\pm-wq^2}{q(z_\pm-w)} X^-_2(w),\nonumber\\ k^\pm_3(z)^{-1}X^+_2(w)k^\pm_3(z)&=&\frac{z_\mp-wq^2}{q(z_\mp-w)} X^+_2(w),\nonumber \end{eqnarray} \begin{eqnarray} (z-wq^2)X^+_1(z)X^+_2(w)+q(z-w)X^+_2(w)X^+_1(z) &=&0,\nonumber\\ q(z-w)X^-_1(z)X^-_2(w)+(z-wq^2)X^-_2(w)X^-_1(z) &=&0,\nonumber\\ (z+wq^{\pm 1})X^\pm_1(z)X^\pm_1(w)+(zq^{\pm 1}+w)X^\pm_1(w)X^\pm_1(z) &=&0,\nonumber\\ (z+wq^{\pm 1})X^\pm_2(z)X^\pm_2(w)+(zq^{\pm 1}+w)X^\pm_2(w)X^\pm_2(z) &=&0,\nonumber \end{eqnarray} \begin{eqnarray} \{X^-_1(w), X^+_1(z)\}&=&(q-q^{-1})\left[-\delta(\frac{z}{w}q^c)k^+_2(z_+) k^+_1(z_+)^{-1}\right.\nonumber\\ & &~~~~~~~~\left.+\delta(\frac{z}{w}q^{-c})k^-_2(w_+)k^-_1(w_+)^{-1}\right],\nonumber\\ \{X^-_2(w), X^+_2(z)\}&=&(q-q^{-1})\left[\delta(\frac{z}{w}q^c)k^+_3(z_+) k^+_2(z_+)^{-1}\right.\nonumber\\ & &~~~~~~~~\left.-\delta(\frac{z}{w}q^{-c})k^-_3(w_+)k^-_2(w_+)^{-1}\right],\nonumber\\ \{X^-_2(w), X^+_1(z)\}&=&(q-q^{-1})q^{-{1\over 2}} \left[-\delta(-\frac{z}{w}q^{c-1})k^+_2(z_+)k^+_1(z_+)^{-1}\right.\nonumber\\ & &~~~~~~~~\left.+\delta(-\frac{z}{w}q^{-c-1})k^-_3(w_+)k^-_2(w_+)^{-1}\right],\nonumber\\ \{X^-_1(w), X^+_2(z)\}&=&(q-q^{-1})q^{1\over 2}\left[\delta(-\frac{z}{w}q^{c+1}) k^+_2(-z_+q)k^+_1(-z_+q)^{-1}\right.\nonumber\\ & &~~~~~~~~\left.-\delta(-\frac{z}{w}q^{-c+1}) k^-_3(-w_+q^{-1})k^-_2(-w_+q^{-1})^{-1}\right], \label{x1-x2} \end{eqnarray} where \begin{equation} \delta(z)=\sum_{l\in {\bf Z}}\,z^l \end{equation} is a formal delta function which enjoys the following properties: \begin{equation} \delta(\frac{z}{w})=\delta(\frac{w}{z}),~~~~~~ \delta(\frac{z}{w})f(z)=\delta(\frac{z}{w})f(w). \end{equation} Defining the algebraic homomorphism \begin{eqnarray} &&X^\pm(z)=z(q-q^{-1})\left[X^\pm_1(z)+X^\pm_2(-zq^{-1})\right],\nonumber\\ &&\psi^-(z)=(1+q^{-{1\over 2}}-q^{1\over 2})\phi_1(z)-\phi_2(-zq^{-1}),\nonumber\\ &&\psi^+(z)=\psi_1(z)-(1+q^{{1\over 2}}-q^{-{1\over 2}})\psi_2(-zq^{-1}), \end{eqnarray} where $\phi_i(z)=k^+_{i+1}(z)k^+_i(z)^{-1},~~ \psi_i(z)=k^-_{i+1}(z)k^-_i(z)^{-1},~~i=1,2$, then we obtain from (\ref{x1-x2}), the current commutation relations (\ref{current}). \section{Level Zero Representation} We consider the evaluation representation $V_x$ of ${U_q[osp(2|2)^{(2)}]}$, where $V$ is the 3-dimensional graded vector space with basis vectors $v_1,~v_2$ and $v_3$. Let $e_{ij}$ be the $3\times 3$ matrix satisfying $e_{ij}v_k=\delta_{jk}v_i$. In the homogeneous gradation, the Chevalley generators of ${U_q[osp(2|2)^{(2)}]}$ are represented on $V_x$ by \begin{eqnarray} E_1&=&e_{12}-e_{23},~~~~F_1=e_{21}+e_{32},~~~~K_1=q^{e_{11}-e_{33}},\nonumber\\ E_0&=&xq^{-1}(-e_{21}+e_{32}),~~~~F_0=x^{-1}q(e_{12}+e_{23}),~~~~ K_0=q^{-e_{11}+e_{33}}. \end{eqnarray} Let $V_x^{*S}$ denote the dual module of $V_x$, defined by $\pi_{V^{*}}(a)=\pi_{V}(S(a))^{st}$. On $V^{*S}_x$, the Chevalley generators are given by \begin{eqnarray} E_1&=&-(q^{-1}e_{21}+e_{32}),~~~~F_1=qe_{12}-e_{23},~~~~ K_1=q^{-e_{11}+e_{33}},\nonumber\\ E_0&=&-xq^{-1}(e_{12}+q^{-1}e_{23}),~~~~F_0=x^{-1}q(-e_{21}+qe_{32}),~~~~ K_0=q^{e_{11}-e_{33}}. \end{eqnarray} The following proposition can be proved by induction. \begin{Proposition}: The Drinfeld generators are represented on $V_x$ by \begin{eqnarray} H_m&=&-x^m\frac{[m]_q}{m}\left[-(-1)^mq^{-m}e_{11}+\left(1-(-1)^mq^m\right)e_{22} +e_{33}\right],\nonumber\\ X^+_m&=&-x^m\left(-(-1)^me_{12}+q^{-m}e_{23}\right),\nonumber\\ X^-_m&=&x^m\left((-1)^me_{21}+q^{-m}e_{32}\right),~~~~~~ K=q^{e_{11}-e_{33}}, \end{eqnarray} and on $V^{*S}_x$ by \begin{eqnarray} H_m&=&-(-1)^mx^m\frac{[m]_q}{m}q^{-m}\left[e_{11}+\left(1-(-1)^mq^{-m}\right)e_{22} -(-1)^mq^{-m}e_{33}\right],\nonumber\\ X^+_m&=&-(-1)^mx^mq^{-m}\left(q^{-m-1}e_{21}+(-1)^me_{32}\right),\nonumber\\ X^-_m&=&(-1)^mx^mq^{-m}\left(q^{-m+1}e_{12}-(-1)^me_{23}\right),~~~~~~ K=q^{-e_{11}+e_{33}}. \end{eqnarray} \end{Proposition} \section{Nonclassical Free Boson Realization at Level One} We use the notation similar to that of \cite{Awa94,Kim97,Zha98}. Let us introduce two sets of bosonic oscillators $\{a_n,\;c_n,\;Q_{a},\; Q_{c}|n\in{\bf Z}\}$ which satisfy the commutation relations \begin{eqnarray} &&[a_n, a_m]=\delta_{n+m,0}(-1)^n\frac{[n]_q^2}{n},~~~~~ [a_0, Q_{a}]=1,\nonumber\\ &&[c_n, c_m]=\delta_{n+m,0}\frac{[n]_q^2}{n},~~~~~ [c_0, Q_{c}]=1.\label{oscilators} \end{eqnarray} The remaining commutation relations are zero. Introduce the $q$-deformed free boson fields \begin{eqnarray} &&a(z;\kappa)=Q_{a}+a_0\ln z -\sum_{n\neq 0}\frac{a_n}{[n]_q}q^{\kappa |n|}z^{-n},\nonumber\\ &&c(z)=Q_{c}+c_0\ln z-\sum_{n\neq 0}\frac{c_n}{[n]_q} z^{-n} \end{eqnarray} and set \begin{equation} a_\pm(z)=\pm(q-q^{-1})\sum_{n>0}a_{\pm n}z^{\mp n}\pm a_0\ln q. \end{equation} Then \begin{Theorem}\label{free-boson}: The Drinfeld generators of $U_q[osp(2|2)^{(2)}]$ at level one have the following nonclassical realization by the free boson fields \begin{eqnarray} &&\psi^{\pm}(z)=e^{a_\pm(z)},\nonumber\\ &&X^{\pm}(z)=:e^{\pm a(z;\mp\frac{1}{2})}\;Y^{\pm}(z):F^\pm,\label{boson} \end{eqnarray} where $F^+=q^{1/2}+q^{-1/2},~~F^-=1/(q-q^{-1})$ and \begin{eqnarray} &&Y^+(z)=e^{c(q^{1/2}z)}+e^{-c(-q^{-1/2}z)},\nonumber\\ &&Y^-(z)=e^{c(-q^{1/2}z)}+ e^{-c(q^{-1/2}z)}. \end{eqnarray} \end{Theorem} This free boson realization is nonclassical in the sense that $X^-(z)$ defined by (\ref{boson}) does not have classical or $q\rightarrow 1$ limit. \noindent{\it Proof.} We prove this theorem by checking that the bosonized currents (\ref{boson}) satisfy the defining relations (\ref{current}) of $U_q[osp(2|2)^{(2)}]$ with $c=1$. It is easily seen that the first two relations in (\ref{current}) are true by construction. The third and fourth ones follow from the definition of $X^\pm(z)$ and the commutativity between $a_n$ and $c_n$. So we only need to check the last two relations in (\ref{current}). We write \begin{equation} Z^\pm(z)=:e^{\pm a(z;\mp\frac{1}{2})}: . \end{equation} We obtain the operator products \begin{eqnarray} Z^\pm(z)Z^\pm(w)&=& (z+q^{\mp 1}w)\;:Z^\pm(z)Z^\pm(w):,\nonumber\\ Z^+(z)Z^-(w)&=& (z+w)^{-1}\;:Z^+(z)Z^-(w):.\nonumber\\ Y^\pm(z)Y^\pm(w)&=&\pm q^{1/2} (z-w):e^{c(\pm q^{1/2}z)}e^{c(\pm q^{1/2}w)}: \mp q^{-1/2} (z-w):e^{- c(\mp q^{-1/2}z)}e^{- c(\mp q^{-1/2}w)}:\nonumber\\ & & \pm \frac{q^{-1/2}}{z+q^{-1}w}:e^{c(\pm q^{1/2}z)}e^{- c(\mp q^{-1/2}w)}: \mp \frac{q^{1/2}}{z+qw}:e^{-c(\mp q^{-1/2}z)}e^{c(\pm q^{1/2}w)}:,\nonumber\\ Y^+(z)Y^-(w)&=&q^{1/2} (z+w):e^{c(q^{1/2}z)}e^{c(-q^{1/2}w)}: -q^{-1/2} (z+w):e^{-c(-q^{-1/2}z)}e^{-c(q^{-1/2}w)}:\nonumber\\ & & +\frac{q^{-1/2}}{z-q^{-1}w}:e^{c(q^{1/2}z)}e^{-c(q^{-1/2}w)}: -\frac{q^{1/2}}{z-qw}:e^{-c(-q^{-1/2}z)}e^{c(-q^{1/2}w)}:.\nonumber \end{eqnarray} Then the second last relation in (\ref{current}) is easily seen to be true, and as to the last relation we have \begin{eqnarray} \{X^+(z), X^-(w)\}&=&\frac{q^{1/2}+q^{-1/2}}{q-q^{-1}}:Z^+(z)Z^-(w):\nonumber\\ & &\left[\left(\frac{q^{-1/2}}{(z+w)(z-q^{-1}w)}+\frac{q^{1/2}}{(w+z) (w-qz)}\right):e^{c(q^{1/2}z)}e^{-c(q^{-1/2}w)}:\right.\nonumber\\ & &\left.-\left(\frac{q^{1/2}}{(z+w)(z-qw)}+\frac{q^{-1/2}}{(w+z) (w-q^{-1}z)}\right):e^{-c(-q^{-1/2}z)}e^{c(-q^{1/2}w)}:\right]\nonumber\\ &=&\frac{1}{(q-q^{-1})zw}:Z^+(z)Z^-(w):\nonumber\\ & &\left[\left(\delta(-\frac{w}{z})-\delta(\frac{w}{z}q^{-1})\right) :e^{c(q^{1/2}z)}e^{-c(q^{-1/2}w)}:\right.\nonumber\\ & &\left.-\left(\delta(-\frac{w}{z})-\delta(\frac{w}{z}q)\right) :e^{-c(-q^{-1/2}z)}e^{c(-q^{1/2}w)}:\right]\nonumber\\ &=&\frac{1}{(q-q^{-1})zw} \left(\delta(\frac{w}{z}q) -\delta(\frac{w}{z}q^{-1})\right):Z^+(z)Z^-(w):\nonumber\\ &=&\frac{1}{(q-q^{-1})zw} \left[\delta(\frac{w}{z}q)\psi^+(wq^{1/2}) -\delta(\frac{w}{z}q^{-1})\psi^-(wq^{-1/2})\right]. \end{eqnarray} This completes the proof. \section*{Acknowledgement} Y.-Z.Z would like to thank Australia Research Council IREX programme for an Asia-Pacific Link Award and Institute of Modern Physics of Northwest University for hospitality.The financial support from Australian Research Council large, small and QEII fellowship grants is also gratefully acknowledged. \vskip.3in
\section{Introduction and Overview} The phenomenon of $CP$ violation, so far observed only in the neutral kaon system, can be accommodated by a complex phase in the Cabibbo-Kobayashi-Maskawa (CKM) quark-mixing matrix~\cite{CKM}. Whether this phase is the correct, or only, source of $CP$ violation awaits experimental confirmation. $B$ meson decays, in particular charmless $B$ meson decays, will play an important role in verifying this picture. The decays $B \rightarrow \pi^+\pi^-$ and $B\rightarrow \rho^+\pi^-$, dominated by the $b \rightarrow u$ tree diagram (Fig.~\ref{fig:feynman}(a)), can be used to measure $CP$ violation due the interference between $B^0-\bar B^0$\ mixing and decay. However, theoretical uncertainties due to the presence of the $b\to dg$ penguin diagram (Fig.~\ref{fig:feynman}(b)) (``{\em Penguin Pollution}'') make it difficult to extract the angle $\alpha$ of the unitarity triangle from $B \rightarrow \pi^+\pi^-$ alone. Additional measurements of $B^\pm \rightarrow \pi^\pm \pi^0$, $B,\bar{B} \rightarrow \pi^0\pi^0$, or a flavor tagged proper-time dependent full Dalitz plot fit for $B\to\pi^+\pi^-\pi^0$ and the use of isospin symmetry may resolve these uncertainties~\cite{isospin}\cite{grossman_quinn}\cite{snyder-quinn}. Alternatively, measurement of $CP$ violation due to the interference of $B\to \pi^\pm\chi_{c0}$ and $B\to\pi^\pm\rho^0$ in $B^\pm\rightarrow \pi^+\pi^-\pi^\pm$~\cite{bediaga} may provide information about the angle $\gamma$ of the unitarity triangle. Neither flavor tagging nor a measurement of the proper-time before decay of the $B$ meson is required in this case. Extraction of the angle $\gamma$ from this measurement is subject to theoretical uncertainties due to Penguin Pollution. However, factorization predicts this to be less severe here than in $B\to\pi^+\pi^-$ due to at least partial cancelation of the gluonic penguin contribution among short distance operators with different chirality. $B\to K\pi $\ decays are dominated by the $b \rightarrow sg$ gluonic penguin diagram, with additional contributions from $b \rightarrow u$ tree and color-allowed electroweak penguin (Fig.~\ref{fig:feynman}(d)) processes. Interference between the penguin (Fig.~\ref{fig:feynman}(b),(d)) and spectator (Fig.~\ref{fig:feynman}(a),(c)) amplitudes can lead to direct $CP$ violation, which would manifest itself as a rate asymmetry for decays of $B$ and $\bar{B}$~mesons. Several methods of measuring the angle $\gamma$ using only decay rates of $B\to K\pi,~\pi\pi$ processes were also proposed~\cite{triangles}. This is particularly important, as $\gamma $\ is the least known parameter of the unitarity triangle and is likely to remain the most difficult to determine experimentally. The ratios $R={\cal B}(B\to K^{\pm}\pi^{\mp})/{\cal B}(B^{\pm}\to K^0\pi^{\pm}) $ ~\cite{fleischermannel}, and $R_\star ={\cal B}(B^\pm\to K^{0}\pi^{\pm})/ 2{\cal B}(B^{\pm}\to K^\pm\pi^{0}) $ ~\cite{neubert-rosner}, were recently suggested as a way to constrain $\gamma$. Electroweak penguins and final state interactions (FSI) in $B\to K\pi$ decays can significantly affect the former method~\cite{weyers}, whereas the latter method requires knowledge of the ratio $|(T+C)/P|_s$ of spectator to penguin amplitudes in $b\to s$ transitions. Uncertainties due to FSI and electroweak penguins are eliminated using isospin and fierz-equivalence of certain short distance operators. Studies of $B$ decays to $KK$ final states can provide useful limits on FSI effects~\cite{gronaukk}. \begin{figure}[hbp] \centering \leavevmode \epsfxsize=6.5in \epsffile{wuerthwein421fig12.eps} \caption{The dominant decay processes are expected to be (a) external W-emission, (b) gluonic penguin, (c) internal W-emission, (d) external electroweak penguin.} \label{fig:feynman} \end{figure} $B$ decays to $\eta^\prime K^0_s$, $\rho^0 K^0_s$, and $\phi K^0_s$ may allow for future measurements of $\sin 2\beta$, $\beta$ being the third angle of the unitarity triangle. This is of interest because one probes the interference between the amplitudes for $b\to s$ penguin and $B^0-\bar B^0$\ mixing, rather than $b\to c$ tree and $B^0-\bar B^0$\ mixing, as done in the more ubiquitous $B\to\psi K^0_s$ decay. It has been argued~\cite{btosCP} that a variety of new physics scenarios would affect the CP violating phase of the $b\to s$ penguin only, leaving the phases of $B^0-\bar B^0$\ mixing and $b\to u$ tree amplitudes unchanged. Such new physics scenarios would thus lead to a difference between proper-time dependent $CP$ violation as measured for example in $B$ decays to $\eta^\prime K^0_s$ as compared to $B\to\psi K^0_s$. The present paper presents preliminary CLEO results on two-body charmless hadronic decays of $B$ mesons into final states containing two pseudo-scalar mesons ($B\to PP$), or a pseudo-scalar and a vector meson ($B\to PV$). Section II discusses the analysis technique that is common to all of these analyses. Results on $B\to PP$ and $B\to PV$ are presented in Sections III. Section IV discusses possible implications of some of the measurements presented. \section{Data Analysis Technique} The data set used in this analysis is collected with the CLEO II and CLEO II.5 detectors at the Cornell Electron Storage Ring (CESR). Roughly $2/3$ of the data is taken at the $\Upsilon$(4S) (on-resonance) while the remaining $1/3$ is taken just below $B\bar{B}$ threshold. The below-threshold sample is used for continuum background studies. The on-resonance sample contains 5.8~million $B\bar{B}$ pairs for all final states except $\rho^+h^-, K^{\star +}h^-$ ($h^+$ being a charged kaon or pion), and $\rho^0K^0_s$. For those final states a total of 7.0~million $B\bar{B}$ pairs was used. This is an $80\% $ increase in the number of $B\bar{B}$ pairs over the published analyses~\cite{prl98}. In addition, we have re-analyzed the CLEO II data set with improved calibration constants and track-fitting algorithm allowing us to extend our geometric acceptance and track quality requirements. This has lead to an overall increase in reconstruction efficiency of $10 - 20$ $\%$ as compared to the previously published analyses. The CLEO detector has been decommissioned for a major detector and accelerator upgrade. Preliminary results based on the full data set of roughly 10~million $B\bar{B}$ pairs are expected to be ready for the summer conferences in 1999. CLEO II and CLEO II.5 are general purpose solenoidal magnet detectors, described in detail elsewhere~\cite{detector}. In CLEO II, the momenta of charged particles are measured in a tracking system consisting of a 6-layer straw tube chamber, a 10-layer precision drift chamber, and a 51-layer main drift chamber, all operating inside a 1.5 T superconducting solenoid. The main drift chamber also provides a measurement of the specific ionization loss, $dE/dx$, used for particle identification. For CLEO II.5 the 6-layer straw tube chamber was replaced by a 3-layer double sided silicon vertex detector, and the gas in the main drift chamber was changed from an argon-ethane to a helium-propane mixture. Photons are detected using a 7800-crystal CsI(Tl) electromagnetic calorimeter. Muons are identified using proportional counters placed at various depths in the steel return yoke of the magnet. Charged tracks are required to pass track quality cuts based on the average hit residual and the impact parameters in both the $r-\phi$ and $r-z$ planes. Candidate $K^0_S$ are selected from pairs of tracks forming well measured displaced vertices. Furthermore, we require the $K^0_S$ momentum vector to point back to the beam spot and the $\pi^+\pi^-$ invariant mass to be within $10$~MeV, two standard deviations ($\sigma$), of the $K^0_S$\ mass. Isolated showers with energies greater than $40$~MeV in the central region of the CsI calorimeter and greater than $50$~MeV elsewhere, are defined to be photons. Pairs of photons with an invariant mass within $25$~MeV ($\sim 2.5\sigma$) of the nominal $\pi^0$ mass are kinematically fitted with the mass constrained to the $\pi^0$ mass. To reduce combinatoric backgrounds we require the lateral shapes of the showers to be consistent with those from photons. To suppress further low energy showers from charged particle interactions in the calorimeter we apply a shower energy dependent isolation cut. Charged particles are identified as kaons or pions using $dE/dx$. Electrons are rejected based on $dE/dx$ and the ratio of the track momentum to the associated shower energy in the CsI calorimeter. We reject muons by requiring that the tracks do not penetrate the steel absorber to a depth greater than seven nuclear interaction lengths. We have studied the $dE/dx$\ separation between kaons and pions for momenta $p \sim 2.6$~GeV$/c$\ in data using $D^{*+}$-tagged $D^0\rightarrow K^- \pi^+$ decays; we find a separation of $(1.7\pm 0.1)~\sigma$ for CLEO II and $(2.0\pm 0.1)~\sigma$ for CLEO II.5. We calculate a beam-constrained $B$ mass $M = \sqrt{E_{\rm b}^2 - p_B^2}$, where $p_B$ is the $B$ candidate momentum and $E_{\rm b}$ is the beam energy. The resolution in $M$\ ranges from 2.5 to 3.0~${\rm MeV}/{\it c}^2$, where the larger resolution corresponds to decay modes with a high momentum $\pi^0$. We define $\Delta E = \sum_i E_i - E_{\rm b}$, where $E_i$ are the energies of the daughters of the $B$ meson candidate. The resolution on $\Delta E$ is mode-dependent. For final states without $\pi^0$'s the $\Delta E$ resolution for CLEO II(II.5) is $\sim$ 20$-$26(17$-$22)MeV. For final states with a high momentum $\pi^0$ the $\Delta E$ resolution is worse approximately by a factor of two and becomes asymmetric because of energy loss out of the back of the CsI crystals. The energy constraint also helps to distinguish between modes of the same topology. For example, $\Delta E$ for $B \rightarrow K^+ \pi^-$, calculated assuming $B \rightarrow \pi^+\pi^-$, has a distribution that is centered at $-42$~MeV, giving a separation of $1.6(1.9)\sigma$ between $B \rightarrow K^+ \pi^-$ and $B \rightarrow \pi^+\pi^-$ for CLEO II(II.5). In addition, $\Delta E$ is very powerful in distinguishing $B\to K^{\star 0}\pi^+$ from $B\to\rho^0\pi^+$, especially if the positive track from the vector meson is of low momentum. We accept events with $M$\ within $5.2-5.3$~$\rm {GeV/c^2}$. The fiducial region in $\Delta E$ depends on the final state. For $B\to PP$ we use $|\Delta E|<200(300)$~MeV for decay modes without (with) a $\pi^0$\ in the final state. The selection criteria for $B\to PV$ are listed in Table II. This fiducial region includes the signal region, and a sideband for background determination. We have studied backgrounds from $b\to c$\ decays and other $b\to u$\ and $b\to s$\ decays and find that all are negligible for $B$ decays to two pseudo-scalar mesons. In contrast, some of the $B$ decays to a pseudo-scalar and a vector meson have significant backgrounds from $b\to c$ as well as other charmless $B$ decays. We discuss these in more detail below in Section III. However, the main background in all analyses arises from $e^+e^-\to q\bar q$\ (where $q=u,d,s,c$). Such events typically exhibit a two-jet structure and can produce high momentum back-to-back tracks in the fiducial region. To reduce contamination from these events, we calculate the angle $\theta_S$ between the sphericity axis\cite{shape} of the candidate tracks and showers and the sphericity axis of the rest of the event. The distribution of $\cos\theta_S$\ is strongly peaked at $\pm 1$ for $q\bar q$\ events and is nearly flat for $B\bar B$\ events. We require $|\cos\theta_S|<0.8$\ which eliminates $83\%$\ of the background for all final states except those including $\eta^\prime$ or $\phi$. For the latter final states a looser cut of $|\cos\theta_S|<0.9$ is used. Using a detailed GEANT-based Monte-Carlo simulation~\cite{geant} we determine overall detection efficiencies (${\cal E}$) ranging from a few $\%$ to $53\%$ in $B\to K^+\pi^-$. Efficiencies are listed for all decay modes in the tables in Section III. We estimate systematic errors on the efficiencies using independent data samples. Additional discrimination between signal and $q\bar q$\ background is provided by a Fisher discriminant technique as described in detail in Ref.~\cite{phd}. The Fisher discriminant is a linear combination ${\cal F}\equiv \sum_{i=1}^{N}\alpha_i y_i$\ where the coefficients $\alpha_i$ are chosen to maximize the separation between the signal and background Monte-Carlo samples. The 11 inputs, $y_i$, are $|\cos\theta_{cand}|$ (the cosine of the angle between the candidate sphericity axis and beam axis), the ratio of Fox-Wolfram moments $H_2/H_0$~\cite{fox}, and nine variables that measure the scalar sum of the momenta of tracks and showers from the rest of the event in nine angular bins, each of $10^\circ$, centered about the candidate's sphericity axis. Some of the analyses (final states including $\eta^\prime$ or $\phi$) use $|\cos\theta_B|$ (the angle between the $B$ meson momentum and beam axis) instead of $H_2/H_0$ as one of the inputs to the Fisher discriminant. We perform unbinned maximum-likelihood (ML) fits using $\Delta E$, $M$, ${\cal F}$, $|\cos\theta_B|$ (if not used as input to ${\cal F}$) and $dE/dx$ (where applicable) as input information for each candidate event to determine the signal yields. Resonance masses ($\eta^\prime$ and vector resonances) and helicity angle of the vector meson are also used as input information in the fit where applicable. In each of these fits the likelihood of the event is parameterized by the sum of probabilities for all relevant signal and background hypotheses, with relative weights determined by maximizing the likelihood function ($\cal L$). The probability of a particular hypothesis is calculated as a product of the probability density functions (PDFs) for each of the input variables. Further details about the likelihood fit can be found in Ref.~\cite{phd}. The parameters for the PDFs are determined from independent data and high-statistics Monte-Carlo samples. We estimate a systematic error on the fitted yield by varying the PDFs used in the fit within their uncertainties. These uncertainties are dominated by the limited statistics in the independent data samples we used to determine the PDFs. The systematic errors on the measured branching fractions are obtained by adding this fit systematic in quadrature with the systematic error on the efficiency. In decay modes for which we do not see statistically significant yields, we calculate $90\%$\ confidence level (C.L.) upper limit yields by integrating the likelihood function \begin{equation} {\int_0^{N^{UL}} {\cal L}_{\rm max} (N) dN \over \int_0^{\infty} {\cal L}_{\rm max} (N) dN} = 0.90 \nonumber \end{equation} where ${\cal L}_{\rm max}(N)$ is the maximum $\cal L$\ at fixed $N$\ to conservatively account for possible correlations among the free parameters in the fit. We then increase upper limit yields by their systematic errors and reduce detection efficiencies by their systematic errors to calculate branching fraction upper limits given in Table I and IV. \section{Results} Given the enormous number of results to summarize in this Section, we choose to show figures only for those decay modes for which we observe statistically significant yields, and no branching fraction measurements have previously been published. Additional figures for preliminary updates on previously published branching fraction measurements can be found elsewhere.~\cite{vancouver} The figures we show are contour plots of $-2\ln{\cal L}$ for the ML fit as well as projection plots for some of the fit inputs. The curves in the contour plots represent the $n\sigma$ contours , which correspond to the increase in $-2\ln{\cal L}$ by $n^2$. Contour plots do not have systematic errors included. The statistical significance of a given signal yield is determined by repeating the fit with the signal yield fixed to be zero and recording the change in $-2\ln{\cal L}$. For the projection plots we apply additional cuts on all variables used in the fit except the one displayed. These additional cuts suppress backgrounds by an order of magnitude at signal efficiencies of roughly $50\%$. Overlaid on these plots are the projections of the PDFs used in the fit, normalized according to the fit results multiplied by the efficiency of the additional cuts. All results shown are preliminary. Not all published analyses~\cite{prl98} have been updated yet. \subsection{$B$ Decays to Two Pseudo-scalar Mesons} Table I lists the preliminary CLEO results for $B$ decays to two pseudo-scalar mesons. Not all possible final states with two pseudo-scalar mesons have been updated yet. For published results please refer to Ref.~\cite{prl98}. Figure~\ref{fig:contourkpi0} illustrates a contour plot for the ML fit to the signal yield ($N$) in the track $\pi^0$ final state. The dashed curve marks the $3\sigma$ contour. To further illustrate the fit, Figure~\ref{fig:mass_hpz} shows $M$ ($\Delta E$) projections as defined above. Events in Figure~\ref{fig:mass_hpz} are required to be more likely to be kaons than pions according to $dE/dx$. We find statistically significant signals for the decays $B \to K^\pm \pi^\mp$, $B^\pm \to K^\pm \pi^0$, $B^\pm \to K^0_S \pi^\pm$, as well as the two $B\to\eta^\prime K$ decays. The corresponding branching fractions are listed in Table~\ref{tab:pp}. Table~\ref{tab:pp} also shows $90\%$ confidence level upper limits for all the decay modes where we do not measure statistically significant yields. \def\ {\ } \def\ {\ } \def\ {\ } \begin{table} \begin{center} \caption Summary of preliminary CLEO results for $B$ decays to two pseudo-scalar mesons. Yield and efficiencies in decay modes including $\eta^\prime$ refer to $\eta^\prime\to\eta\pi^+\pi^-, \eta\to\gamma\gamma$ ($\eta^\prime\to\rho\gamma$) decays.} \vskip 0.2cm \begin{tabular}{lllll} \hline Mode & Eff (\%) & Yield & Signif & BR/UL ($10^{-5}$) \\ \hline \ {$K^\pm\pi^\mp$} & \ {$53\pm 5$} & \ {$43.1_{-8.2}^{+9.0}$} & \ {$>6\sigma$} & \ {$1.4\pm 0.3 \pm 0.2$} \\ \ {$K^\pm\pi^0$} & \ {$42\pm 4$} & \ {$38.1_{-8.7}^{+9.7}$} & \ {$>6\sigma$} & \ {$1.5\pm 0.4\pm 0.3$} \\ \ {$K^0\pi^\pm$} & \ {$15\pm 2$} & \ {$12.3_{-3.9}^{+4.7}$} & \ {$>5\sigma$} & \ {$1.4\pm 0.5 \pm 0.2$} \\ \ {$\eta^\prime K^\pm$} & \ {$5(11)$} & \ {$18.4(50.2)$} & \ {$>6\sigma $} & \ {$7.4^{+0.8}_{-1.3}\pm 1.0$} \\ \ {$\eta^\prime K^0$} & \ {$1.6(3.4)$} & \ {$5.4(12.7)$} & \ {$>5\sigma $} & \ {$5.9^{+1.8}_{-1.6}\pm 0.9$} \\ \hline \ {$\pi^\pm\pi^\mp$} & \ {$53\pm 5$} & \ {$11.5^{+6.3}_{-5.2}$} & \ {$<3\sigma$} & \ {$<0.84$}\\ \ {$\pi^\pm\pi^0$} & \ {$42\pm 4$} & \ {$14.9^{+8.1}_{-6.9}$} & \ {$<3\sigma$} & \ {$<1.6$}\\ \ {$\eta^\prime \pi^\pm$} & \ {$5(11)$} & \ {$1.0(0.0)$} & \ {} & \ {$<1.2$} \\ \hline \ {$K^\pm K^\mp$} & \ {$53\pm 5$} & \ {$0.0_{-0.0}^{+1.6}$} & & \ {$<0.23$}\\ \ {$K^\pm K^0$} & \ {$15\pm 2$} & \ {$1.8_{-1.4}^{+2.6}$} & & \ {$<0.93$}\\ \hline \end{tabular} \label{tab:pp} \end{center} \end{table} \begin{figure}[htbp] \centering \leavevmode \epsfxsize=3.0in \epsffile{wuerthwein421fig1.ps} \caption{Contour of the $-2\ln{\cal L}$ for the ML fit to $N_{K^\pm \pi^0}$ and $N_{\pi^\pm\pi^0}$ for $B^\pm \rightarrow K^\pm \pi^0$ and $B^\pm\rightarrow \pi^\pm\pi^0$. } \label{fig:contourkpi0} \end{figure} \begin{figure}[htbp] \centering \leavevmode \epsfxsize=3.0in \epsffile{wuerthwein421fig2.eps} \caption{Projection plots in $B^\pm \to K^\pm \pi^0$. $\Delta E$ for $B^\pm\to K^\pm\pi^0$ is centered around $-42$MeV because we use pion mass when calculating the energy of a track. Only events for which the candidate track is more likely to be a kaon than a pion according to $dE/dx$ enter these figures.} \label{fig:mass_hpz} \end{figure} \subsection{$B$ Decays to a Pseudo-scalar and a Vector Meson} Helicity conservation dictates that the polarization of the vector in $B\to PV$ is purely longitudinal (helicity = 0 state). The kinematics of these decays (assuming two-body decay of the vector) therefore results in a final state with two energetic particles and one soft particle. The pseudo-scalar $P$ is always very energetic, with a momentum range from 2.3 to 2.8 GeV. On the other hand the decay daughters from the vector meson have a very wide momentum range. While the more energetic particle has momentum between 1.0 and 2.8 GeV, the soft particle can have momentum as low as 200 MeV. The backgrounds from $\mbox{$B\bar{B}$}$ events are potentially dangerous as they may peak in either or both of the $M$ and $\Delta E$ distributions. There are two types of $\mbox{$B\bar{B}$}$ backgrounds that can contribute to $PV$: $b \rightarrow c$ processes and other rare b processes. Among the $B \rightarrow P V$ modes we are searching for, $B\rightarrow \pi^{\pm} V$ and $B\rightarrow K^{\pm} V$ can be well separated, using the dE/dx information of the very energetic $\pi^{+}$ or $K^{+}$ and the separation in $\Delta E$, just like the $B \rightarrow P P$ modes. Crosstalk of two kinds exist among $PV$ modes. First, $\rho\leftrightarrow K^\star$ misidentification is possible for track $\pi^0$ as well as two track decays of the $\rho$ or $K^\star$ if the fast particle is misidentified due to the limited particle ID for fast tracks. Crosstalk among $B^{+}\rightarrow h^{+} \rho^{0}$ and $B^{+}\rightarrow h^{+} K^{*0}$ can be controlled to a level of 20$\%$ or less just by requirements on dE/dx (2 $\sigma$) of the decay daughters of the vector meson. Further separation is achieved by using $\Delta E$ and resonance mass of the vector as inputs to the likelihood fit. Second, it is possible to swap a slow momentum pion from the vector with a slow momentum pion from the other $B$. This is particularly severe for slow momentum $\pi^0$, as the fake/real $\pi^0$ ratio is about a factor 20 worse for the slow pions than the fast pions from the vector. In such cases we impose helicity requirement to remove the region with soft $\pi^0$. Using data (doubly charged vector candidates) and Monte Carlo we determine the remaining backgrounds from other rare b processes to be small effects that we correct for. The dominant $b \rightarrow c$ background for $PV(h^+ h^+ h^-)$ is $B^{+}\rightarrow \bar{D}^0 \pi^{+}$, where $\bar{D}^0 \rightarrow\pi^{+}\pi^{-}$ or $\bar{D}^0\rightarrow K^{+}\pi^{-}$. This particular background has exactly the same final state particles as the $PV(h^+ h^+ h^- )$ signal and therefore peaks both in $M$ at 5.28 GeV and in $\Delta E$ at 0.0 GeV. Other $b \rightarrow c$ processes $B^{+}\rightarrow\bar{D}^0\rho^{+}$ and $B^{0}\rightarrow D^{*-}\pi^{+}$ will have peak structure in $M$, but not in $\Delta E$ due to the missing soft particle. Because of the large $b \rightarrow c$ branching ratio, the contribution from these processes needs to be highly suppressed. We apply a $\bar{D}^0$ (30MeV) veto to all possible $h^{+} h^{-}$ combinations in $PV(h^+ h^+ h^-)$ modes. Similarly the $b \rightarrow c$ background for $PV(h^+ h^- \pi^{0}$) are $B^{0}\rightarrow D^- \pi^{+}$ where $D^{-}\rightarrow \pi^{-}\pi^{0}$ or $B^{0}\rightarrow D^{0} \pi^{0}$ where $D^{0}\rightarrow K^{-}\pi^{+}$. However, their contribution is negligible due to the branching ratios involved. Finally, there are potentially backgrounds from non-resonant $B$ decays to three-body final states. We test for such backgrounds in data by allowing a non-resonant signal contribution in the fit, as well as by determining the fit yield in bins of helicity angle. Neither of these tests shows any evidence of non-resonant contributions to any of our final states. The increase of the error on the fitted yield due to possible non-resonance contributions is accounted for as part of our systematic errors. \subsubsection{First Observation of $B^{\pm} \rightarrow \pi^{\pm}\rho^{0}$} We select separate $h^+ \rho^{0}$ and $h^+ K^{\star 0}$ samples as discussed above and in Table II. We then fit for the $B^{+}\rightarrow\pi^{+}\rho^{0}$ and $B^{+}\rightarrow K^{+}\rho^{0}$ components in this $h^{+} \rho^{0}$ sample, as well as a $B\to\pi^+K^{\star 0}$ reflection, averaging over charge conjugate modes. Similarly we select a $h^{+} K^{*0}$ sample and fit for the $B^{+}\rightarrow\pi^{+}K^{*0}$ and $B^{+}\rightarrow K^{+}K^{*0}$, as well as a $B^{+}\rightarrow \pi^{+}\rho^{0}$ reflection. We do not attempt a simultaneous fit to the $h^+ \rho^{0}$ and $h^+ K^{\star 0}$ samples at this point as this would require us to model the full momentum dependence of $\Delta E$, $dE/dx$, and resonance mass in order to separate $\rho$ and $K^*$ contributions. The variables $M$, ${\cal F}$, E($\pi\pi\pi$) $-$ E$_{b}$ (E($\pi K \pi$) $-$ E$_{b}$), dE/dx of $h$ in $B\rightarrow h \rho^{0}$ ($B\rightarrow h K^{*0}$), Mass of $\rho^{0}$ ($K^{*0}$) candidate and cos($\rho^{0}$ ($K^{*0}$) helicity angle) are used to form probability density function (PDF) to perform the ML fit for $B^{\pm}\rightarrow h^{\pm} \rho^{0}$ ($B^{\pm}\rightarrow h^{\pm} K^{*0}$) sample. We do not use $dE/dx$ for the daughters of the vector meson in the fit. Efficiencies and results are summarized in Table IV. A significant signal in $B^{\pm} \rightarrow \pi^{\pm}\rho^{0}$ is observed. The contour and projection plots are shown in Fig.~\ref{pirho0}. \vskip 1.0cm \begin{figure}[hbp] \centering \leavevmode \epsfxsize=2.0in \epsfysize=2.5in \epsffile{wuerthwein421fig3.ps} \epsfxsize=2.0in \epsfysize=2.5in \epsffile{wuerthwein421fig4.ps} \epsfxsize=2.0in \epsfysize=2.5in \epsffile{wuerthwein421fig5.ps} \caption{Contour and projection plot for $B^{\pm} \rightarrow \pi^{\pm} \rho^{0}$.} \label{pirho0} \end{figure} \begin{table}[hhh] \label{selection} \caption{Event Selection for $B\to PV$ decays} \begin{center} \begin{tabular}{|c|c|c|c|} \hline Sample & E $-$ Ebeam & Resonance Mass Window & Cos(Resonance Helicity Angle) \\ \hline $h^{\pm}\rho^{0}$ & $|$E($\pi\pi\pi$)$-$Ebeam$|<$100MeV & 200MeV & $-$0.9 $-$ 0.9 \\ \hline $h^{\pm} K^{*0}$ & $|$E($\pi K\pi$)$-$Ebeam$|<$100MeV & 75MeV & $-$0.9 $-$ 0.9 \\ \hline $h^{\pm}\rho^{\mp}$ & $|$E($\pi\pi\pi^{0}$)$-$Ebeam$|<$150MeV & 200MeV & 0.0 $-$ 0.9 \\ \hline $h^{\pm} K^{*\mp}(K^{\mp}\pi^{0})$ & $|$E($\pi K\pi^{0}$)$-$Ebeam$|<$300MeV & 200MeV & 0.1 $-$ 1.0 \\ \hline $h^{\pm} K^{*\mp}(K^0_s\pi^\mp)$ & $|$E($\pi K\pi^{0}$)$-$Ebeam$|<$200MeV & 200MeV & $-$0.86 $-$ 1.0 \\ \hline \end{tabular} \end{center} \end{table} The contribution of $b\rightarrow c$ and other related rare $B$ processes are small but not negligible. They are evaluated using about 25 million generic b$\mbox{$\bar{\mbox{b}}$}$ Monte Carlo events, and specific Monte Carlo samples for all the rare $B$ processes mentioned in this paper. The dominant contributions are listed in Table III. All other contributions are negligible. \begin{table}[hbp] \label{Correction} \caption{Contributions to the $B^{\pm}\rightarrow \pi^{\pm}\rho^{0}$ Yield from Non-continuum physics backgrounds} \begin{center} \begin{tabular}{|c|c|} \hline Decay Process & Contribution to $B^\pm\rightarrow \pi^\pm\rho^{0}$ Yield \\ \hline b $\rightarrow$ c & 0.9$\pm$0.7 \\ \hline $B^{0}\rightarrow \pi^{+}\rho^{-}$ & 0.7$\pm$0.3 \\ \hline $B^{0}\rightarrow K^{+}\rho^{-}$ & 0.1$\pm$0.1 \\ \hline $B^{0}\rightarrow \pi^{+} K^{*0}$ & 0.3$\pm$0.2 \\ \hline \hline TOTAL & 2.0$\pm$0.8 \\ \hline \end{tabular} \end{center} \end{table} The final $B^{\pm}\rightarrow \pi^{\pm}\rho^{0}$ yield after background subtraction is: 26.1$^{+9.1}_{-8.0}$ events, leading to a branching fraction measurement of ${\cal B}(B^+\to\pi^+\rho^0) = (1.5\pm 0.5\pm 0.4)\times 10^{-5}$. This is the first observed hadronic $b \rightarrow u$ transition. \subsubsection{First Observation of $B^{0} \rightarrow \pi^{\pm}\rho^{\mp}$} As discussed above, the $\pi^0$ daughter of the $\rho$ has a bi-modal momentum distribution due to the longitudinal polarization (helicity = 0) of the $\rho$. The ratio of real to fake $\pi^0$ is roughly $1/2$ for the low and $10/1$ for the high momentum $\pi^0$ region. This leads to largely increased backgrounds from all sources as well as multiple entries per event in the low momentum $\pi^0$ region. In addition, the charged pion tends to be fast for the slow $\pi^0$ region, thus leading to increased $K^{\star +}\leftrightarrow\rho^+$ misidentification. In contrast, the only drawback of the fast $\pi^0$ region over the three track sample is a factor two degraded $\Delta E$ resolution. We therefore choose to use only the half of the sample that has a high momentum $\pi^0$ in our fits in the two track $\pi^0$ final state at this point. Besides this, the same likelihood fits are made as described for the three track final state. Efficiencies and results are summarized in Table IV. The crossfeed rates from other $PV$ modes as well as $b\to c$ decay backgrounds are negligible. A significant signal in $B^{0} \rightarrow \pi^{\pm}\rho^{\mp}$ is observed at a branching fraction of ${\cal B}(B^0\to \pi^\mp\rho^\pm) = (3.5^{+1.1}_{-1.0}\pm 0.5)\times 10^{-5}$. Note that we do not tag the flavor of the $B$ in the present analysis. The measured branching fraction is therefore the sum of $B^0\to\rho^+\pi^-$ and $B^0\to\rho^-\pi^+$. In addition, averaging over charge conjugate states is as always implied. \vskip 1.0cm \begin{figure}[hbp] \centering \leavevmode \epsfxsize=2.0in \epsfysize=2.5in \epsffile{wuerthwein421fig6.ps} \epsfxsize=2.0in \epsfysize=2.5in \epsffile{wuerthwein421fig7.ps} \epsfxsize=2.0in \epsfysize=2.5in \epsffile{wuerthwein421fig8.ps} \caption{Contour and projection plot $B^{0} \rightarrow \pi^{\pm} \rho^{\mp}$.} \label{pirho} \end{figure} \subsubsection{Evidence for $B^{0} \rightarrow \pi^{-} K^{*+}$} We search for $B^{0} \rightarrow \pi^{-} K^{*+}$ with submodes $K^{*+}\rightarrow K^{0}_{S}\pi^{+}$ and $K^{*+}\rightarrow K^{+}\pi^{0}$. Due to the large combinatoric and physics backgrounds in the soft $\pi^{0}$ region, we only select the hard $\pi^{0}$ region for the $K^+\pi^0$ decay of the $K^*$. Backgrounds other than those from continuum are negligible. Event selections are presented in Table II. Efficiencies and results are summarized in Table IV. The individual branching ratios obtained in the two $K^{*+}$ submodes are consistent, and we combine the two submodes to arrive at an average branching ratio of ${\cal B}(B^{0} \rightarrow \pi^{-} K^{*+})$ = (2.2 $^{+0.8\ +0.4}_{-0.6\ -0.5}$) $\times$ 10$^{-5}$ which is 5.9$\sigma$ from zero. We note that the statistical significance depends largely on the two track $K^0_s$ final state, which has less background and larger efficiency than the two track $\pi^0$ final state. In contrast to the two observed $B\to\rho\pi$ decays, the one dimensional projections of the fit (see Fig.~\ref{fig:pikstar}) are somewhat less than inspiring, and a simple event count in the mass plot does result in an excess of only $2.4\sigma$. However, goodness of fit ($21\% CL$), and likelihood per event distributions are perfectly consistent with expectations from Monte Carlo. The most likely signal events have signal likelihoods consistent with what one may expect from signal Monte Carlo, rather than the background data taken below $B\bar B$ threshold. In addition, we generated 25000 distinct Monte Carlo background samples in the $K^0_s\pi^+\pi^-$ final state. Each of these samples has the same number of events as our actual data in this final state. We perform a likelihood fit to each of these 25000 samples and record signal yield and significance as reported by each fit. We find that none of these background samples leads to a reported yield or significance as large as found in data. We therefore conclude that our result is exceedingly unlikely to be due to a background fluctuation. \vskip 1.0cm \begin{figure}[hbp] \centering \leavevmode \epsfxsize=2.0in \epsfysize=2.0in \epsffile{wuerthwein421fig9.eps} \epsfxsize=2.0in \epsfysize=2.0in \epsffile{wuerthwein421fig10.eps} \epsfxsize=2.0in \epsfysize=2.0in \epsffile{wuerthwein421fig11.eps} \caption{Contour and projection plot $B^{0} \rightarrow \pi^{\pm} K^{*\mp}$.} \label{fig:pikstar} \end{figure} \begin{table} \begin{center} \caption \bf Summary of CLEO results for $B$ decays to a pseudo-scalar and a vector mesons ($PV$ modes)} \vskip 0.2cm \begin{tabular}{lllll} \hline Mode & Eff (\%) & Yield & Signif & BR/UL ($10^{-5}$) \\ \hline {$\pi^{\pm}\rho^{0}$} & {$30 \pm 3$} & {$26.1^{+9.1}_{-8.0} $} & {5.2$\sigma$} & {$1.5\pm 0.5 \pm 0.4$} \\ {$\pi^{\pm}\rho^{\mp}$} & {$12 \pm 1$} & {$28.5^{+8.9}_{-7.9} $} & {5.6$\sigma$} & {$3.5^{+1.1}_{-1.0} \pm 0.5$} \\ {$\pi^{\pm} K^{*\mp}(K^{0}_{S}\pi^{\mp})$} & {$7 \pm 1$} & {$10.8^{+4.3}_{-3.5} $} & {5.2$\sigma$} & {$2.3^{+0.9}_{-0.7} \pm 0.3$} \\ {$\pi^{\pm} K^{*\mp}(K^{\mp}\pi^{0})$} & {$4.1 \pm 0.4$} & {$5.7^{+4.3}_{-3.2} $} & {2.5$\sigma$} & {$2.0^{+1.5\ +0.3}_{-1.1\ -0.4}$} \\ {$\pi^{\pm} K^{*\mp}$} & & & {5.9$\sigma$} & {$2.2^{+0.8\ +0.4 }_{-0.6\ -0.5} $} \\ {$\pi^\pm K^{\star 0}(K^{+} \pi^{-})$} & {$18 \pm 2$} & {$12.3^{+5.7}_{-4.7} $} & {$<$ 3$\sigma$} & {$<$ 2.7} \\ {$K^{\pm}\rho^{0}$} & {$28 \pm 2$} & {$14.8^{+8.8}_{-7.7} $} & {$<$ 3$\sigma$} & {$<$ 2.2} \\ {$K^{0}\rho^{0}$} & {$10 \pm 1$} & {$8.2^{+4.9}_{-3.9} $} & {$<$ 3$\sigma$} & {$<$ 2.7} \\ {$K^{\pm}\rho^{\mp}$} & {$11 \pm 1$} & {$8.3^{+6.3}_{-5.0} $} & {$<$ 3$\sigma$} & {$<$ 2.5} \\ {$K^{\pm}\phi$} & {$26 \pm 3$} & & & {$<$ 0.59} \\ {$K^{0}\phi$} & {$7 \pm 1$} & & & {$<$ 2.8} \\ {$\pi^{\pm}\phi$} & {$26 \pm 3$} & & & {$<$ 0.40} \\ {$\pi^{0}\phi$} & {$17 \pm 2$} & & & {$<$ 0.54} \\ {$K^{\pm} K^{*\mp}(K^{0}_{S} \pi^{\mp})$} & {$7 \pm 1$} & {$0.0^{+0.9}_{-0.0} $} & & {$<$ 0.8} \\ {$K^{\pm} K^{*\mp}(K^\mp \pi^{0})$} & {$4.1 \pm 0.4$} & {$0.0^{+1.3}_{-0.0} $} & & {$<$ 1.7} \\ {$K^{\pm} K^{*\mp}$} & & & & {$<$ 0.6} \\ {$K^{+} K^{*0}(K^{+} \pi^{-})$} & {$18 \pm 2$} & {$0.0^{+2.1}_{-0.0}$} & & {$<$ 1.2} \\ \hline \end{tabular} \label{tab:pv} \end{center} \end{table} \section{Discussion of our Results} Let us start by summarizing some of the more striking features seen in the data. First of all, we see no evidence for $B\to K\bar K$ decays in either $B\to PP$ or $B\to PV$. Such decays would proceed either via highly suppressed $W-$exchange (e.g. $B\to K^+K^-$) and $b\to d$ penguin diagrams (e.g. $B\to K^0_sK^\pm$, $B\to K^0_sK^0_s$) or via final state rescattering (FSI). Given that our upper limits for some of these decays are an order of magnitude smaller than at least some of the branching fractions we measure it seems fair to neglect FSI when trying to understand the dominant contributions to charmless hadronic $B$ decays. Second, we see no evidence for $B\to \pi\pi$ decays while we observe both $B\to K\pi$ as well as $B\to\rho\pi$ decays. We try to make sense out of this in Section~\ref{ss-model-pp} in the context of isospin and factorization. Third, we are so far unable to measure the branching fraction for any of the $B\to\rho K$ decay modes, despite the fact that we have measured $B\to\rho\pi$ and $B\to K\pi$, and at least one of the $B\to K^\star\pi$ decay modes. This is in full agreement with factorization predictions. Factorization predicts destructive (constructive) interference between penguin operators of opposite chirality for $B\to\rho K$ ($B\to K\pi$), leading to a rather small (large) penguin contribution in these decays. In addition, factorization and CVC predict that only the left-handed penguin operator contributes in $B\to K^{\star +}\pi^-$. Destructive interference of penguin operators is therefore not expected in this decay mode. Fourth, we want to note that the measured ratio $R_\rho = {\cal B}(B^0\to\rho^\pm\pi^\mp)/{\cal B}(B^+\to\rho^0\pi^+)$ is much smaller than naively expected. In $B\to\rho\pi$ decays the $\rho$ can either come from the upper or lower vertex, and it is generally believed that upper vertex $\rho$ production clearly dominates due to favorable form factors as well as decay constants. In addition, $B^+\to\rho^0\pi^+$ is further suppressed by a factor two because only the $u\bar u$ part of the $\rho^0$ wave function contributes. The present CLEO measurement of ${\cal B}(B^0\to\rho^\pm\pi^\mp)$ is the sum of upper and lower vertex $\rho$ production. It is therefore rather surprising that the measured $R_\rho = 2.3\pm 1.3$ is not significantly larger than two. Measurements of $B\to\rho^+\pi^0$ as well as a flavor tagged measurement of $B\to\rho^+\pi^-$ would help to clarify the situation in $B\to\rho\pi$ decays. It remains to be seen whether or not such measurements are within reach using the full CLEO data set. Finally, maybe the most striking observation in our data are the large branching fractions measured for charged as well as neutral $B$ decays to $\eta^\prime K$. Violation of a sum-rule proposed by Lipkin~\cite{lipkin} seems to indicate that a significant flavor singlet contribution is needed to explain these rates. The literature is full~\cite{etaprime-puzzle} of attempts to explain this apparent discrepancy, the most interesting of which is the suggestion that R-parity violating couplings may explain the large $\eta^\prime K$ as well as the stringent limit on $\phi K$~\cite{rparity}. The latter is particularly amusing as one of the relevant couplings ($\lambda^\prime_{323}$) would also be present in $B_s-$mixing~\cite{rparity-mixing} and could therefore lead to a different value for $\gamma$ as inferred from $B\to K\pi$ decays and the limit on $\Delta m_s/\Delta m_d$ in the context of the usual analysis of the $\rho -\eta $ plane ~\cite{rho-eta}. \subsection{Understanding the non-observation of $B\to\pi\pi$} \label{ss-model-pp} Most theoretical predictions lead us to expect a branching fraction for $B\to\pi^+\pi^-$ at a level of $1-2\times 10^{-5}$.~\cite{all} Instead, the central value and 90$\%$ confidence level upper limit presented here are $4$ and $8\times 10^{-6}$. With results like this a natural question to ask is ``{\em How small can ${\cal B}(B\to\pi^+\pi^-)$ be?}''. Let us start our answer by describing a data based factorization prediction. Assuming factorization, and neglecting W-exchange, penguin annihilation, and electroweak penguin diagrams one may expect the following expressions for the decay amplitudes:~\cite{zeppenfeld} \begin{equation} \begin{array}{ccccc} \sqrt{2}A^{\pm 0} &=& -(T+C) \\ A^{+-} &=& -(T+P) &=& -|T|e^{i\gamma}\times (1- |P/T|e^{i\alpha}) \\ \sqrt{2}A^{00} &=& P-C \\ \end{array} \label{eq:amp} \end{equation} Superscripts $+,-,0$ indicate the charge of the final state pions, and $T,C,P$ stand for external and internal W-emission, and gluonic penguin diagrams respectively (Fig. 1(a), Fig. 1(c), and Fig. 1(b)). We can arrive at ``data based factorization estimates'' of these amplitudes if we identify $C = a_2/a_1 \times T$ and use $a_2/a_1 = 0.21\pm 0.14$ from measurements in $B\to D$ decays~\cite{a2/a1}. We then estimate $T$ using factorization and the CLEO measurement ${\cal B}(B\to\pi l\nu) = (1.8\pm 0.5)\times 10^{-4}$~\cite{pilnu} as follows: \begin{equation} \begin{array}{ccccccccc} T &\sim & \sqrt{6}\pi f_\pi & \times & a_1 & \times & \sqrt{d\Gamma(B\to\pi l\nu)/dq^2 |_{q^2 = m_{\pi}^2} \over \Gamma(B\to\pi l\nu)} & \times & \sqrt{{\cal B}(B\to\pi l\nu)} \\ &\sim & 1.0\mathrm{GeV} &\times &(1.0\pm 0.1) &\times & (0.27\pm 0.05)/\mathrm{GeV} &\times & (0.0135\pm 0.0022)\\ &\sim & \multicolumn{7}{l}{(3.6\pm 0.9)\times 10^{-3}}\\ \end{array} \label{eq:factorization} \end{equation} The dominant error here is due to the spread among a variety of theoretical models for the $q^2$ dependence of the form factor~\cite{lkg}. We do not assign any error due to a possible breakdown of the factorization hypothesis. Throughout this paper we express the absolute size of amplitudes in units of $\sqrt{\mathrm{Branching\ Fraction}}$. The decay $B^+\to K^0_s\pi^+$ has three down type quarks in the final state. Inspection of Figure 1 shows that this final state can only be reached via penguin diagrams, or final state rescattering. Furthermore, the electroweak penguin contribution to this decay is color suppressed, rather than the color allowed one shown in Figure 1(d). It is therefore reasonable to estimate $P$ from the measured ${\cal B}(B\to K^0\pi^\pm)$ corrected by CKM and SU(3) breaking factors. Using these numbers we arrive at $|T/P|_d = 5.0\pm 2.3$. This leads to the factorization predictions ${\cal B}(B^0\to\pi^+\pi^-) = (8\pm 5)\times 10^{-6}$, ${\cal B}(B^+\to\pi^+\pi^0) = (10\pm 5)\times 10^{-6}$, and ${\cal B}(B^0\to\pi^0\pi^0) \sim $ few $\times 10^{-6}$. The last of these three estimates is not very meaningful given the errors on the quantities that enter. We assume maximum destructive interference ($\cos\alpha = 1$). Ignoring the penguin contribution (i.e. $\cos\alpha = 0$) leads to a prediction of ${\cal B}(B^0\to\pi^+\pi^-) = (13.0\pm 6.5)\times 10^{-6}$. As an aside, we can calculate $|T/P|_s = 0.26\pm 0.08$. This means that CP violating rate asymmetries as large as $50\%$ are in principle possible for decays like $B\to K^+\pi^-$ if the relevant weak and strong phases are close to $\pm\pi/2$. In addition to these factorization estimates, it is quite illustrative to look at the isospin decomposition of $B\to\pi\pi$:~\cite{isospin} \begin{equation} \begin{array}{ccc} \sqrt{2}A^{\pm 0} &=& 3 A_{3/2}^\gamma\times e^{i\delta}\\ A^{+-} &=& { A_{3/2}^\gamma \times e^{i\delta} + A_{1/2}^\gamma} + { A_{1/2}^\beta} \\ \sqrt{2}A^{00} &=& { 2 A_{3/2}^\gamma \times e^{i\delta} - A_{1/2}^\gamma} -{ A_{1/2}^\beta} \\ \end{array} \label{eq:isoamp} \end{equation} Here the subscripts $1/2, 3/2$ indicate the two different isospin amplitudes. Note that only the $A_{1/2}$ amplitude has any contribution from $b\to d$ penguins, whereas the $A_{3/2}$ amplitude is a pure $b\to u$ transition. We indicate this by making the dependence on weak ($\beta,\ \gamma$) and strong interaction phases ($\delta $) explicit. \footnote{We ignore a possible strong phase difference between penguin and tree contribution to $A_{1/2}$.} Using the factorization estimates above, it is easy to show that $|A_{1/2}^\gamma|,\ |A_{3/2}^\gamma|$, and $|A_{1/2}^\beta|$ are of the same order of magnitude. Equation~\ref{eq:isoamp} shows that ${\cal B}(B\to\pi^\pm\pi^0)$ can be estimated without making any assumptions about strong or weak phases. However very little can be said about the relative size of $B\to\pi^0\pi^0$ versus $B\to\pi^+\pi^-$ without making such assumptions about relative phases. Common prejudice assumes $\delta << 1$ and therefore $B\to \pi^+\pi^-\ >>\ B\to\pi^0\pi^0$ due to the destructive interference between $A_{3/2}$ and $A_{1/2}$ in $B\to\pi^0\pi^0$. However, as we allow for $\delta$ to increase towards $\pi$ we not only decrease (increase) $B\to\pi^+\pi^- (\pi^0\pi^0)$ but also increase the size of the ``penguin pollution'' in any future attempt of measuring $\sin 2\alpha$ via time dependent CP violation in $B\to\pi^+\pi^-$. We are thus in the amusing situation that we would like $B\to\pi^0\pi^0$ to be large to make the Gronau, London isospin decomposition~\cite{isospin} experimentally feasible. Though at the same time, we can only hope for $\delta << 1 $ (i.e. vanishingly small $B\to\pi^0\pi^0$) to avoid destructive interference between the two $b\to u$ pieces in the amplitude for $B\to \pi^+\pi^-$. We conclude that our present data is still consistent with factorization predictions for $B\to\pi^+\pi^-$. However, $B\to\pi^+\pi^-$ could be significantly smaller than predicted by factorization if the strong interaction phase between isospin amplitudes is non-zero. \subsection{Comment on Neubert-Rosner bound on $\gamma$} \label{ss-bound-gamma} As previously mentioned, the ratio $R_\star ={\cal B}(B^\pm\to K^{0}\pi^{\pm})/ 2{\cal B}(B^{\pm}\to K^\pm\pi^{0}) $ ~\cite{neubert-rosner}, may be used to constrain $\cos\gamma$ if $R_\star \neq 1$. Our measurement of this ratio is $R_\star = 0.47\pm 0.24$. The relevant equation for bounding $\cos\gamma$ out of the paper by Neubert and Rosner~\cite{neubert-rosner} is: \begin{equation} \cos\gamma = \delta_{EW} - ((1-\sqrt{R_{\star}})/\epsilon_{3/2})/\cos\phi + O(\epsilon_{3/2}^2) \label{eq:eq1} \end{equation} The parameter $\epsilon_{3/2}$ is defined in terms of experimentally measurable quantities below. It is essentially given by the ratio of $b\to u$ tree and $b\to s$ gluonic penguin amplitudes. The $O(\epsilon_{3/2}^2)$ terms were shown to be small in Ref.~\cite{neubert}. Here, $\delta_{EW} = 0.63\pm 0.15$ is the theoretically calculated contribution from electroweak penguin operators~\cite{neubert-rosner}. The dominant uncertainty in Equation~\ref{eq:eq1} is the unknown strong phase $\cos\phi$. Taking the extreme values of $0$ and $\pi$ for this phase we thus arrive at an excluded region for $\cos\gamma$ rather than an actual measurement. To be conservative, one may choose values for $\delta_{EW}$ such as to minimize this excluded region: \begin{equation} 0.48 + |1-\sqrt{R_{\star}}|/\epsilon_{3/2} \ge \cos\gamma \ge 0.78 - |1-\sqrt{R_{\star}}|/\epsilon_{3/2} \label{eq:bound1} \end{equation} The structure of this is obviously to exclude values for $\cos\gamma$ near $\cos\gamma = \delta_{EW}$ if $X \equiv |1-\sqrt{R_{\star}}|/\epsilon_{3/2}> 0.15$. The size of the exclusion region is determined by the central value of $X$ as well as its error. The variable $X$ defined here is given in terms of measurable quantities up to small uncertainties due to non-factorizable SU(3) breaking: \begin{equation}\begin{array}{ccc} X \equiv |1-\sqrt{R_\star}|/\epsilon_{3/2}) & = & |1 - a/b|\times a/c \\ a & =: & \sqrt{{\cal B}(B\to K^0\pi^+)} \\ & = & (3.74\pm 0.72)\times 10^{-3} \\ b & =: & \sqrt{2\times{\cal B}(B\to K^+\pi^0)} \\ & = & (5.48\pm 0.91)\times 10^{-3} \\ c & =: & |V_{us}/V_{ud}| f_K/f_\pi \times \sqrt{2\times{\cal B}(B\to \pi^+\pi^0)} \\ & = & (0.95\pm 0.31\ (0.91\pm 0.18))\times 10^{-3} \\ \end{array} \label{eq:eq2} \end{equation} The two different values for $c$ are obtained using either the most likely value for ${\cal B}(B\to\pi^+\pi^0)$ based on the preliminary CLEO results (including statistical and systematic errors added in quadrature), or a weighted average of the latter with theoretical predictions based on factorization~\cite{neubert-rosner}. When calculating $X$ from these numbers we additionaly increase $c$ to conservatively account for theoretical uncertainties due to non-factorizable SU(3) breaking ($ ``f_K/f_\pi'' \equiv 1.33$ rather than the experimental value of $1.2$). The resulting values for $X$ are $1.15\pm 0.62 $ and $1.20\pm 0.56$ respectively for the two different values for $c$. In the following we will use $X = 1.20\pm 0.56$. A number of comments are in order at this point. First, the central value for $X$ leads to a physical value for $\cos\gamma$ via Equation~\ref{eq:eq1} only if the strong phase $\phi\sim 0$. In that case, the measured $X$ then prefers rather large values of $|\gamma |\sim 120^\circ$. Such values of $\gamma$ are generally not the favored ones as they would tend to imply $B_s$ mixing to be smaller than the present limits and/or $f_{B_s}\sqrt{B_{B_s}}/f_{B_d}\sqrt{B_{B_d}}$ to be at the large end of the generally assumed range. It was pointed out by He, Hou, and Yang~\cite{hou} that a number of other charmless hadronic $B$ decay results from CLEO also suggest $|\gamma |> 90^\circ$. Second, only $10-15\%$ of a Gaussian with mean $0.48 + X$ and $\sigma = \sigma_X$ ly within the physically allowed region for $\cos\gamma$. Calculating a bound in this case isn't all that meaningful. Instead one may consider $\cos\phi < 0$ to be ruled out at $\sim 90\%$ confidence level. Using the usual procedure of calculating one-sided confidence levels based on the area inside the physical region only, results in the bound $\cos\gamma \le 0.33\ @\ 90\%$ confidence level. Third, the experimental errors on $X$ are large, roughly 1/4 of the physically allowed region total. It is fair to say that the only reason why we may deduce a non-zero exclusion region for $\cos\gamma$ from present measurements is because our present central value for $X$ indicates a prefered value for $\cos\gamma $ that is far away from $\cos\gamma = \delta_{EW}$. This is in contrast to some of the recent analyses of the $\rho - \eta$ plane~\cite{rho-eta} which tend to favor $\cos\gamma \sim \delta_{EW}$. \section{Conclusion} In summary, we have measured branching fractions for three of the four exclusive $B\to K\pi$ decays, as well as the two $B\to\eta^\prime K$ decays, while only upper limits could be established for all other $B$ decays to two pseudo-scalar mesons. In addition, we have observed two of the four $B\to\rho\pi$ decays, as well as one of the four $B\to K^{\star}\pi$ decays. We do not observe significant yields for $B$ decays to $\rho K$, $K^\star K$, $\phi \pi$ or $\phi K$. The pattern of observed decays is broadly consistent with expectations from factorization. We see significant contributions from both $b\to u$ as well as $b\to s$ transitions. In addition, the Neubert-Rosner bound derived from present CLEO data on charmless hadronic $B$ decays indicates $\cos\gamma < 0.33\ @\ 90\%$ confidence level. This is in slight disagreement with some of the more aggressive analyses of the $\rho -\eta$ plane found in the literature which prefer larger values of $\cos\gamma$. Many thanks to our colleagues at CLEO for many stimulating discussions as well as the experimental work that made this paper possible. Further thanks go to A.~Ali, J.-M.~G\'erard, M.~Gronau, W.-S.~Hou, M.~Neubert, J.~L.~Rosner, and H.~Yamamoto for discussions on topics related to Section IV. We gratefully acknowledge the effort of the CESR staff in providing us with excellent luminosity and running conditions.
\section{Introduction} \par The investigation on the electromagnetic (EM) masses of the low-lying mesons has a long history, however, which is mainly on evaluating the EM masses of pseudoscalar $\pi$ and $K$-mesons (see Ref. \cite{GLY} and references therein). In Ref. \cite{GLY}(hereafter referred as paper I), we have presented a systematic investigation on the EM mass splittings of the low-lying mesons including $\pi$, $K$, $a_1$, $K_1$, and $K^*$ in the framework of the U(3)$_L\times$U(3)$_R$ chiral theory of mesons \cite{Li95}. The present paper is the continuation of paper I. We are to calculate the EM self-energies of the vector $\rho$, $\omega$, and $\phi$ mesons to one-loop order and $O(\alpha_{\rm EM})$. Chiral quark model is originated by Weinberg \cite{W79}, and developed by Manohar and Georgi \cite{MG84}. In Ref. \cite{LYL91}, the vector meson( $\omega-$meson) is firstly introduced into this model in order to study quark spin contents in chiral soliton model. The author of Ref. \cite{Li95} further extended it to include the low-lying 1$^-$(vector) and 1$^+$(axial-vector) mesons (called as U(3)$_L\times$U(3)$_R$ chiral theory of mesons). The U(3)$_L\times$U(3)$_R$ chiral theory of mesons has been investigated extensively \cite{GLY,Li97,LGY98,Li98} and its theoretical results agree well with the data. Chiral perturbation theory (ChPT), which is expanded in powers of derivatives of the meson fields, is rigorous and phenomenologically successful in describing the physics of the pseudoscalar mesons at very low energies \cite{GL85}. In Ref. \cite{WY98}, starting from the U(3)$_L\times$U(3)$_R$ chiral theory of mesons, and by using path integral method to integrate out the vector and axial-vector resonances, the authors have derived the chiral coupling constants of ChPT ($L_1$, $L_2$, $L_3$, $L_9$, and $L_{10}$). The results are in good agreement with the experimental values of the $L_i$ at $\mu=m_\rho$ in ChPT. Therefore, the QCD constraints discussed in Ref. \cite{EGLPR} are met by this theory. It has been pointed out in \cite{Li95} that vector meson dominance (VMD)\cite{KLZ,Sak69} has been introduced into the U(3)$_L\times$U(3)$_R$ chiral theory of mesons. This means that the electromagnetic interaction of the mesons has been well established, which makes it possible to evaluate the EM self-energies of the low-lying mesons systematically. In paper I, the logarithmic divergences coming from meson-loop diagrams which contribute to the EM mass splittings of $\pi$, $K$, $a_1$, $K_1$, and $K^*$ mesons have been factorized by using the intrinsic parameter $g$ of the theory (see paper I for details). However, when this method is extended to the cases of $\rho$, $\omega$, and $\phi$-mesons, the circumstances will be complicated. The Feynman diagrams contributing to the EM self-energies of $\rho$-mesons can be divided into three kinds, which have been shown in Fig. 1. Fig. 1.1 is tree diagram, and it contributes the finite result; Fig. 1.2 involves only logarithmic divergences, which could be evaluated by using the same method in paper I. In the framework of the U(3)$_L\times$U(3)$_R$ chiral theory of mesons, the explicit one-loop Feynman diagrams which belong to Fig. 1.2 will be drawn in Figs. 8, 9, and 10; It will meet some difficulties in computing Fig. 1.3 because the quadratic or higher order divergences will emerge in the Feynman integrations of these loop diagrams. Therefore, it is unsuitable to factorize these divergences by using $g$ in which only the logarithmic one is involved. However, if watching these Feynman diagrams carefully, one will find that the contribution from Fig. 1.3 could be ignored. Considering one-loop correction to the $\rho-\gamma$ vertex (Fig. 2), and comparing Fig. 1.3 with Fig. 2, one will see that the contribution of Fig. 1.3 to EM self-energies of $\rho$-mesons should vanish if the one-loop renormalized $\rho-\gamma$ vertex is used in Fig. 1.1. The expressions of VMD in the U(3)$_L\times$U(3)$_R$ chiral theory of mesons have been derived in Ref. \cite{Li95}, which can be read as \begin{eqnarray} \frac{e}{f_\rho}\{-\frac{1}{2}F^{\mu \nu}\rho^0_{\mu \nu}+A^\mu J_\mu^\rho\},\\ \frac{e}{f_\omega}\{-\frac{1}{2}F^{\mu \nu}\omega_{\mu\nu}+A^\mu J_\mu^\omega\},\\ \frac{e}{f_\phi}\{-\frac{1}{2}F^{\mu \nu}\phi_{\mu \nu}+A^\mu J_\mu^\phi\}, \end{eqnarray} where \begin{eqnarray} \frac{1}{f_\rho}=\frac{1}{2}g,\;\;\;\frac{1}{f_\omega}=\frac{1}{6}g,\;\;\; \frac{1}{f_\phi}=-\frac{1}{3\sqrt{2}}g. \end{eqnarray} $J_\mu^\rho$, $J_\mu^\omega$, and $J_\mu^\phi$ are the corresponding hadronic currents for $\rho^0$, $\omega$, and $\phi$ mesons respectively. By employing the $\rho-\gamma$ vertex in eq. (1), the decay width of $\rho\rightarrow e^+ e^-$ is \begin{equation} \Gamma(\rho\rightarrow e^+ e^-)=\frac{4\pi\alpha_{\rm EM}^2}{f_\rho^2}\frac{m_\rho}{3}. \end{equation} Here $\alpha_{\rm EM}=\frac{e^2}{4\pi}$. When $g$=0.39 \footnote{if setting $g$=0.39 instead of $g$=0.35, we will get much better fit for the data than that in Ref. \cite{Li95}, and $g$=0.39 has been set in paper I and Refs. \cite{Li97,LGY98,Li98}.}, the numerical result of $\Gamma(\rho\rightarrow e^+ e^-)$ is 6.53 keV, which is in good agreement with the experimental data 6.77$\pm$0.32 keV \cite{PDG98}. It is expected that one-loop corrections to the $\rho-\gamma$ vertex are very small, and can be ignored. It is enough to evaluate the contributions from Fig. 1.1 and 1.2 to EM self-energies of $\rho$-mesons. The effect of Fig. 1.3 is included by renormalizing the $\rho-\gamma$ vertex. In fact, Fig. 1.3 is not the 1PI (one particle irreducible) diagram for the one-loop EM self-energies of $\rho$-meson. Thus, it is straightforward to use the method developed in paper I to factorize the logarithmic divergences from mesonic loop diagrams, then to get a finite result of $\rho^0-\rho^\pm$ EM mass difference. The difficulties in calculating the EM self-energies of $\omega$ and $\phi$-mesons can be circumvented similarly. It has been pointed out by Sakurai \cite{Sak69} that there are two ways of writing down VMD. In eqs. (1)-(3) (called as VMD1), the photon-meson coupling have two approaches: the first one is the direct coupling of photon and mesons ($A^\mu J^i_\mu$, $i$=$\rho$, $\omega$, and $\phi$), the second one is the interaction through the neutral vector mesons including $\rho^0$, $\omega$, and $\phi$. The other representation of VMD (called as VMD2) \footnote{In Ref. \cite{OPTW}, the two representations of VMD have been referred as VMD1 and VMD2.}, in which the photon-mesons coupling must be through the neutral vector mesons, is expressed as follows \begin{eqnarray} -\frac{e m_\rho^2}{f_\rho}\rho^0_\mu A^\mu+\frac{1}{2}(\frac{e m_\rho}{f_\rho})^2 A_\mu A^\mu,\\ -\frac{e m_\omega^2}{f_\omega}\omega_\mu A^\mu+\frac{1}{2}(\frac{e m_\omega}{f_\omega})^2 A_\mu A^\mu,\\ -\frac{e m_\phi^2}{f_\phi}\phi_\mu A^\mu+\frac{1}{2}(\frac{e m_\phi}{f_\phi})^2 A_\mu A^\mu. \end{eqnarray} As proved by Kroll, Lee, and Zumino \cite{KLZ}, and by Sakurai \cite{Sak69}, VMD1 and VMD2 are equivalent in describing the electromagnetic interaction of the mesons. Fortunately, in the framework of the U(3)$_L\times$U(3)$_R$ chiral theory of mesons and employing VMD2, the loop diagram (Fig. 1.3) contributing to the EM self-energies of the vector mesons which contains quadratic or higher order divergences will disappear automatically, and only the logarithmic divergence is involved in the rest loop diagrams. Our calculations show that, without considering Fig. 1.3, contributions to EM self-energies of $\rho$-mesons by using VMD1 are entirely equivalent to ones by using VMD2 directly. The purposes of the present paper are twofold: 1.~The investigation of EM self-energies of the vector mesons has been concerned by particle physicists. Here, we are to extend the method used in paper I to calculate the one-loop EM self-energies of the low-lying mesons in the case of vector mesons including $\rho$, $\omega$, and $\phi$-mesons, which makes the investigation in paper I much more complete and systematic. 2.~An interesting effect, EM masses anomaly of the massive Yang-Mills particles, has been revealed in Ref. \cite{YG98}, which states that the non-Abelian gauge structure of the massive Yang-Mills particles makes the EM mass of the neutral $K^*$(892) larger than that of the charged one. This claim will be re-examined in the cases of $\rho$ meson and other massive Yang-Mills particles elaborately. The paper is organized as follows. In Sec. 2, the equivalence of VMD1 and VMD2 is discussed briefly. As examples, we show that, in the framework of the U(3)$_L\times$U(3)$_R$ chiral theory of mesons, VMD1 and VMD2 could reduce the same results of the pion EM form factor and the EM mass splitting of $\pi^\pm-\pi^0$, and the well-known result of $m_{\pi^\pm}^2-m_{\pi^0}^2$ obtained by Das, Guralnik, Mathur, Low, and Young \cite{DGMLY} is reproduced. Sec. 3, the EM self-energies of the vector mesons including $\rho$, $\omega$,and $\phi$ are evaluated, and the EM mass anomaly of the massive Yang-Mills particles is discussed. Sec. 4, we give the summary of the results. \section{VMD1 and VMD2} VMD predates the standard model. It has not been derived from the standard model directly, but nevertheless enjoys phenomenological supports in describing hadronic electromagnetic interactions. In the U(3)$_L\times$U(3)$_R$ chiral theory of mesons, the basic lagrangian of the theory is (Hereafter we use the notations in paper I) \begin{eqnarray} \lefteqn{{\cal L}=\bar{\psi}(x)(i\gamma\cdot\partial+\gamma\cdot v +e_0Q\gamma\cdot A+\gamma\cdot a\gamma_{5} -mu(x))\psi(x)}\nonumber \\ & &+{1\over 2}m^{2}_{1}(\rho^{\mu}_{i}\rho_{\mu i}+ \omega^{\mu}\omega_{\mu}+a^{\mu}_{i}a_{\mu i}+f^{\mu}f_{\mu})\nonumber\\ & &+{1\over2}m^2_2(K_{\mu}^{*a}K^{*a\mu}+K_1^{\mu}K_{1\mu})\nonumber \\ & &+\frac{1}{2}m^2_3(\phi_{\mu} \phi^{\mu}+f_s^{\mu}f_{s\mu}) +{\cal L}_{\rm EM} \end{eqnarray} where $u(x)=exp[i\gamma_{5} (\tau_{i}\pi_{i}+\lambda_a K^a+\eta +\eta^{\prime})]$($i$=1,2,3 and $a$=4,5,6,7), $a_{\mu}=\tau_{i}a^{i}_{\mu}+ \lambda_a K^a_{1\mu}+(\frac{2}{3} +\frac{1}{\sqrt{3}} \lambda_8)f_{\mu}+( \frac{1}{3}-\frac{1}{\sqrt{3}} \lambda_8)f_{s\mu}, v_{\mu}=\tau_{i}\rho^{i}_{\mu}+\lambda_a K_{\mu}^{*a}+(\frac{2}{3}+ \frac{1}{\sqrt{3}}\lambda_8)\omega_{\mu}+(\frac{1}{3}- \frac{1}{\sqrt{3}} \lambda_8)\phi_{\mu}$, $A_\mu$ is the photon field, $Q$ is the electric charge operator of $u$, $d$ and $s$ quarks, $\psi$ is quark-fields, $m$ is a parameter related to the quark condensate, and ${\cal L}_{\rm EM}$ is the kinetic lagrangian of photon field. Thus, the use of path integral method to integrate out the quark fields will reduce the effective lagrangian of the mesons. Since the photon field $A_\mu$ has been introduced into eq. (9), it is very natural to deduce the explicit expressions of VMD in the mesonic effective lagrangian, which are just the eqs.(1)-(3). The full lagrangian which describes the interaction of photon and the neutral vector mesons could be expressed as follows, \begin{equation} {\cal L}_{\rm VMD1}=-\frac{1}{4}F_{\mu\nu}F^{\mu\nu}- \frac{1}{4}\rho^0_{\mu\nu}\rho^{0\mu\nu}+\frac{1}{2}m_\rho^2\rho^0_\mu \rho^{0\mu}+\rho^{0\mu} J_\mu^\rho+ \frac{e}{f_\rho}(-\frac{1}{2}F^{\mu \nu}\rho^0_{\mu \nu}+A^\mu J_\mu^\rho). \end{equation} For simplicity, here we only discuss the interactions between $\rho^0$ and the photon. VMD1 is manifestly gauge invariant for the neutral vector meson $\rho^0$ coupling with the conserved current in strong interaction. As pointed out by Sakurai \cite{Sak69}, there is an equivalent way of writing down a gauge invariant expression for hadronic electromagnetic interaction, which is \begin{equation} {\cal L}_{\rm VMD2}=-\frac{1}{4}F_{\mu\nu}F^{\mu\nu}- \frac{1}{4}\rho^0_{\mu\nu}\rho^{0\mu\nu}+\frac{1}{2}m_\rho^2\rho^0_\mu \rho^{0\mu}+\rho^{0\mu} J_\mu^\rho-\frac{e m_\rho^2}{f_\rho}\rho^0_\mu A^\mu+\frac{1}{2}(\frac{e m_\rho}{f_\rho})^2 A_\mu A^\mu. \end{equation} From VMD1 to VMD2, one can use the following transformation, \begin{eqnarray} \rho^0_\mu\rightarrow\rho^0_\mu-\frac{e}{f_\rho}A_\mu,\nonumber\\ A_\mu\rightarrow\sqrt{1-(\frac{e}{f_\rho})^2}~A_\mu,\\ e\rightarrow e~ \sqrt{1-(\frac{e}{f_\rho})^2}.\nonumber \end{eqnarray} It seems that the mass term of the photon field in eq. (11) violates gauge invariance of the electromagnetic interaction. However, the $\rho\cdot A$ term in this equation contributes an imaginary mass term of the photon, which exactly cancels the contribution from mass term of the photon. To illustrate this, one could calculate the modification of the photon propagator using the new interaction lagrangian including the new mass term of the photon (Fig. 3). From Fig. 3, one will obtain \begin{eqnarray} iD(k^2)=\frac{-i}{k^2}+\frac{-i}{k^2}\frac{-i e m_\rho}{f_\rho} \frac{-i}{k^2-m_\rho^2}\frac{-i e m_\rho^2}{f_\rho}\frac{-i}{k^2}+...\nonumber\\ +\frac{-i}{k^2}\frac{i e^2 m_\rho^2}{f_\rho^2}\frac{-i}{k^2}+...\nonumber\\ =\frac{-i}{k^2}+\frac{-i}{k^2}\frac{i e^2 m_\rho^2}{f_\rho^2} \frac{k^2}{k^2-m_\rho^2}\frac{-i}{k^2}+...\nonumber \end{eqnarray} Using the operator identity \begin{eqnarray*} \frac{1}{A-B}=\frac{1}{A}+\frac{1}{A}B\frac{1}{B} +\frac{1}{A}B\frac{1}{A}B\frac{1}{A}+..., \end{eqnarray*} we have \begin{eqnarray} iD(k^2)=\frac{-i}{k^2-\frac{e^2 m_\rho^2} {f_\rho^2}\frac{k^2}{k^2-m_\rho^2}} \longrightarrow \frac{-i}{k^2}(1-\frac{e^2}{f_\rho^2}) \end{eqnarray} for small $k^2$. Thus, we are simply left with a change in EM coupling constant $e$ $$ e^2\longrightarrow e^2 (1-\frac{e^2}{f_\rho^2}). $$ In the following, as an example, we show the equivalence of VMD1 and VMD2 by calculating pion EM form factor. In Ref. \cite{Li95}, the interaction ${\cal L}_{\rho\pi\pi}$ has been derived \begin{eqnarray} {\cal L}_{\rho\pi\pi}=\frac{2}{g}\epsilon_{ijk}\rho_\mu^i\pi^j\partial^\mu \pi^k-\frac{2}{\pi^2f_\pi^2 g}[(1-\frac{2c}{g})^2-4\pi^2 c^2]\epsilon_{ijk}\rho_\mu^i\partial_\nu\pi^j\partial^\mu\partial^\nu\pi^k, \end{eqnarray} where $c=\frac{f_\pi^2}{2 g m_\rho^2}$, and $f_\pi$ is the decay constant of pion with value $0.186$ GeV. Thus, the pion EM form factor $F_\pi(k^2)$ can be defined through the amplitude of process $\gamma\longrightarrow \pi^+\pi^-$ \begin{equation} {\cal M}^\mu_{\gamma\rightarrow\pi^+\pi^-}=-e(q^+-q^-)^\mu F_\pi(k^2), \end{equation} here $q$ and $k$ are momentum of pion and the virtual photon respectively. From VMD1 (as shown in Fig. 4.1), we have \begin{equation} F_\pi(k^2)=(1+\frac{k^2}{m_\rho^2-k^2})\frac{f_{\rho\pi\pi}(k^2)}{f_\rho}, \end{equation} where \begin{eqnarray*} f_{\rho\pi\pi}(k^2)=\frac{2}{g}\{1+\frac{k^2}{2\pi^2 f_\pi^2}[(1-\frac{2c}{g})^2-4\pi^2 c^2]\}. \end{eqnarray*} Note that the second term in eq. (16) due to the direct interaction of $\rho$ and $\gamma$ vanishes at $k^2=0$, and it is therefore very important to take into account the effect of the contact $J_\mu^\rho A^\mu$ interaction. From VMD2 (as shown in Fig. 4.2), one obtains \begin{equation} F_\pi(k^2)=\frac{m_\rho^2}{m_\rho^2-k^2}\frac{f_{\rho\pi\pi}(k^2)}{f_\rho}. \end{equation} It is obvious that the two approaches lead to the identical result from eq. (16) and eq. (17). Now, we show that VMD1 and VMD2 are equivalent in evaluating the EM mass splitting of pions. In paper I, the EM mass difference between $\pi^\pm$ and $\pi^0$ in the chiral limit has been obtained in terms of VMD1. Here we will re-examine this calculation by employing VMD2. The Feynman diagrams contributing to $(m_{\pi^\pm}^2-m_{\pi^0}^2)_{\rm EM}$ have been shown in Fig. 5 and Fig. 6, which correspond to ones from VMD1 and VMD2 respectively (The diagrams which receive contributions from the interaction ${\cal L}_{\rho\pi\pi}$ have been neglected because their contributions vanish in the chiral limit. This point has been demonstrated in paper I). Since there are no the direct vertices of $\pi$ and $\gamma$, the number of the Feynman diagrams contributing to EM self-energies of pions in terms of VMD2 has been much more decreased. Comparing Fig. 5 and Fig. 6, one will find that if the effective propagators in Fig. 7.1 and Fig. 7.2 are equivalent, the contributions of Fig. 5 and Fig. 6 must be the same to the EM self-energies of pion-mesons. Note that in Fig. 7.1, the $\rho-\gamma$ vertex comes from VMD1, which is momentum dependent; in Fig. 7.2, the $\rho-\gamma$ vertex is from VMD2. It is easy to prove that Fig. 7.1 and Fig. 7.2 both contribute \begin{equation} \frac{-i}{k^2}\frac{e^2}{f_\rho^2}[\frac{m_\rho^4}{(k^2-m_\rho^2)^2} (g_{\mu\nu}-\frac{k_\mu k_\nu}{k^2})+a\frac{k_\mu k_\nu}{k^2}], \end{equation} where $a$ is the gauge parameter. Because there is always a factor $\frac{e}{f_\rho}$ in the coupling of the photon with the current $J_\mu^\rho$ in VMD1, this factor has been considered in calculating Fig. 7.1. Eq. (18) has been obtained by Lee and Nieh in Ref. \cite{LN68}. Therefore, it is not surprising that the result of EM self-energies of pions in terms of VMD1 is the same as one in terms of VMD2. The straightforward calculation of Fig. 6 leads to \begin{equation} (m_{\pi^\pm}^2-m_{\pi^0}^2)_{\rm EM} =i{e^2 \over f_\pi^2} {\int \frac{d^4 k}{(2\pi)^4} (D-1)m_\rho^4} {{(F^2+{k^2 \over 2\pi^2})} \over {k^2 (k^2-m_\rho^2)^2 }} [1+{\gamma^2 \over g^2}{{F^2+{k^2 \over 2\pi^2}} \over {k^2 -m_a^2} }]. \end{equation} This is just eq. (34) in paper I, which is derived in terms of VMD1. The gauge independence of $(m_{\pi^\pm}^2-m_{\pi^0}^2)_{\rm EM}$ can be proven in the same way as that in paper I. In deriving eq. (19) or eq. (34) in paper I, the calculations on EM-masses are up to the fourth order covariant derivatives in effective lagrangians \cite{Li95}. In the remainder of this section, we find that the well-known result of $(m_{\pi^\pm}^2-m_{\pi^0}^2)_{\rm EM}$ given by Das et al \cite{DGMLY} can be reproduced if the EM self-energies of pions receives the contributions only from the second order derivative terms. In this case, the interaction lagrangians ${\cal L}_{\rho\rho\pi\pi}$ and ${\cal L}_{\rho\pi a}$ [eqs.(18)(19) in paper I] will be simplified as follows \begin{eqnarray*} & &{\cal L}_{\rho\rho\pi\pi}=\frac{2F^2}{g^2 f_\pi^2}\rho^i_\mu \rho^{j\mu}(\pi^2\delta_{ij}-\pi_i\pi_j),\;\;\;\;\; {\cal L}_{\rho\pi a}=-\frac{2F^2}{f_\pi g^2}\rho^i_{\mu}\epsilon_{ijk}\pi^k a^{j\mu}. \end{eqnarray*} Thus, in the chiral limit, the EM mass difference of pions is \begin{eqnarray} (m_{\pi^\pm}^2-m_{\pi^0}^2)_{\rm EM} =\frac{i3e^2}{f_\pi^2} \int\frac{d^4k}{(2\pi)^4}\frac{m_\rho^4 F^2}{k^2(k^2-m_\rho^2)^2} (1+\frac{F^2}{g^2(k^2-m_a^2)}). \end{eqnarray} The Feynman integration in eq.(20) is finite. So it is straightforward to get the result of $(m_{\pi^\pm}^2-m_{\pi^0}^2)_{\rm EM}$ after performing this integration, which is \begin{equation} (m_{\pi^\pm}^2-m_{\pi^0}^2)_{\rm EM} =\frac{3\alpha_{\rm EM}m_\rho^4}{8\pi f_\pi^2} \{\frac{2F^2}{m_\rho^2}-\frac{2F^4}{g^2(m_a^2-m_\rho^2)}(\frac{1}{m_\rho^2} +\frac{1}{m_a^2-m_\rho^2}{\rm log}{\frac{m_\rho^2}{m_a^2}})\}. \end{equation} Because we only consider the second order derivative terms in the lagrangian, the relation between $m_a$ and $m_\rho$ is $m_a^2=\frac{F^2}{g^2} +m_\rho^2$ instead of eq.(4) in paper I. Thus, we can get \begin{equation} (m_{\pi^\pm}^2-m_{\pi^0}^2)_{\rm EM} =\frac{3\alpha_{\rm EM}}{4\pi}\frac{m_a^2 m_\rho^2}{m_a^2-m_\rho^2} {\rm log}{\frac{m_a^2}{m_\rho^2}}. \end{equation} When substituting the relation $m_a^2=2 m_\rho^2$, which can be derived from the Weinberg sum rules \cite{Wb67}, into eq.(21), we have \begin{equation} (m_{\pi^\pm}^2-m_{\pi^0}^2)_{\rm EM} =\frac{3{\rm log}2}{2\pi}\alpha_{\rm EM} m_\rho^2, \end{equation} which is exactly the result obtained by Das et al. \cite{DGMLY}, and serves as the leading term of eq.(19). \section{EM self-energies of the vector mesons and EM mass anomaly of the massive Yang-Mills particles} The subject of EM contributions of vector mesons is in fact very old. The early attempts are in \cite{LN68,S64,C69}, and the recent are in \cite{BG96,AA99}. In Ref. \cite{GLY,YG98}, the EM masses of $K^*(892)$ have been evaluated in the framework of the U(3)$_L\times$U(3)$_R$ chiral theory of mesons. In this section, we will deal with the EM self-energies of $\rho$, $\omega$, and $\phi$-mesons to one-loop order and $O(\alpha_{\rm EM})$. The one-loop Feynman diagrams contributing to the EM self-energies of $\rho$-meson are only from the massive Yang-Mills (MYM)self-interactions and Gauging Wess-Zumino-Witten (GWZW) anomaly interactions in the effective lagrangian of the U(3)$_L\times$U(3)$_R$ chiral theory of mesons \cite{Li95}. This case is the same as that of $K^*(892)$-meson \cite{YG98}. From Ref. \cite{Li95}, MYM lagrangian related to $\rho$-meson is \begin{equation} {\cal L}_{\rm MYM}=-{1 \over 8} Tr (\partial_\mu \rho_\nu-\partial_\nu \rho_\mu- \frac{i}{g}[\rho_\mu, \rho_\nu]-\frac{i}{g}[a_\mu, a_\nu])^2+ {\rm mass\; terms\; of }\; \rho. \end{equation} Thus, the three-point and four-point MYM interaction vertices which contribute to EM self-energies of $\rho$-meson can be read from eq. (24) \begin{eqnarray} {\cal L}_{\rm MYM}^{(3)}=\frac{1}{g}\epsilon_{ijk} \partial_\mu\rho^i_\nu\rho^{j\mu}\rho^{k\nu},\\ {\cal L}_{\rm MYM}^{(4)}=\frac{1}{g^2}(\rho^i_\mu\rho^i_\nu \rho^{j\mu}\rho^{j\nu}-\rho^i_\mu \rho^{i\mu}\rho^j_\nu\rho^{j\nu}). \end{eqnarray} The GWZW anomaly lagrangian in the present theory contributing to EM self-energies of $\rho$-meson is \cite{Li95} \begin{equation} {\cal L}_{\rm GWZW}^{\omega\rho\pi}=-\frac{3}{\pi^2 g^2 f_\pi} \epsilon^{\mu\nu\alpha \beta}\partial_\mu\omega_\nu\rho_\alpha^i \partial_\beta\pi^i. \end{equation} Now, combining VMD1 [eqs.(1),(2)] or VMD2 [eqs. (6), (7)], one can evaluate the one-loop EM self-energies of $\rho$-meson in the standard way presented in paper I. The corresponding diagrams are shown in Figs. 8, 9, and 10. Note that all the diagrams will contribute to EM self-energies of $\rho$-meson if we use VMD1 while only Figs. 8.3, 9.3 and 10.3 give contributions if using VMD2. In the following, we calculate these diagrams separately. From Fig. 8, the EM mass square difference corresponding to the contributions of three-point MYM interactions can be calculated directly, which is \begin{eqnarray} & &m^2_{\rm EM}(\rho^\pm)_{\rm MYM}^{(3)}= i e^2 \int\frac{d^4 k}{(2\pi)^4}\frac{m_\rho^4}{k^2(k^2-2 p\cdot k)(k^2-m_\rho^2)^2}\nonumber\\ & &\times [3k^2+4 m_\rho^2-\frac{(k^2)^2}{m_\rho^2}-4\frac{(k\cdot p)^2}{k^2}+2k\cdot p +\frac{\langle(k\cdot\rho^{\underline{i}})(k\cdot\rho^{\underline{i}}) \rangle}{\langle~\rho^{\underline{i}}\rho^{\underline{i}}~\rangle} (9+\frac{k^2}{m_\rho^2}-\frac{2k\cdot p}{k^2})], \end{eqnarray} where $p$ is 4-momentum of $\rho$-meson, and $p^2=m_\rho^2$; $\langle~\rho^{\underline{i}}\rho^{\underline{i}}~\rangle=\int d^4 x \langle~\rho|\rho_\mu^{\underline{i}}(x)\rho^{\underline{i}\mu}(x)|\rho~\rangle$, and $\underline{i}=1,~2$. After performing the Feynman integrations in eq. (28), we have \begin{equation} m^2_{\rm EM}(\rho^\pm)_{\rm MYM}^{(3)}=-e^2 m_\rho^2 [-\frac{3}{4} \chi_\rho+\frac{1}{16\pi^2}(\frac{11\pi}{4\sqrt{3}}-\frac{37}{12})]. \end{equation} The logarithmic divergence is involved in $\chi_\rho$, which has been factorized by using the intrinsic parameter $g$ of the theory in paper I, \begin{eqnarray} \chi_\rho=\frac{1}{g^2}+\frac{1}{32\pi^2}+\frac{1}{16\pi^2}{\rm log}\frac{f_\pi^2}{6(g^2 m_\rho^2-f_\pi^2)}. \end{eqnarray} $g=0.39$ has been fixed in paper I. Substituting the experimental values of $f_\pi$ and $m_\rho$ into eq. (29), we obtain the numerical result \begin{equation} m^2_{\rm EM}(\rho^\pm)_{\rm MYM}^{(3)}=-3.36\times 10^{-4}~{\rm GeV^2}. \end{equation} Similarly, the contribution to EM self-energies of $\rho$-meson (Fig. 9) from four-point MYM interaction is \begin{equation} m^2_{\rm EM}(\rho^\pm)_{\rm MYM}^{(4)}=-ie^2 \frac{9m_\rho^4}{4}\int \frac{d^4 k}{(2\pi)^4}\frac{1}{k^2(k^2-m_\rho^2)^2}. \end{equation} It is remarkable that eq. (32) is free of the divergence and independent of $g$, so it is straightforward to get \begin{eqnarray} m^2_{\rm EM}(\rho^\pm)_{\rm MYM}^{(4)}=-\frac{9e^2}{64 \pi^2 m_\rho^2}. \end{eqnarray} Numerically \begin{equation} m^2_{\rm EM}(\rho^\pm)_{\rm MYM}^{(4)}=-7.75\times 10^{-4} ~{\rm GeV^2}. \end{equation} The contribution to the EM self-energies of $\rho$-mesons from GWZW anomaly can be evaluated by using eq. (27) [Fig. 10]. However, it is easy to see that the contribution of Fig. 10 leads to the same shift of the charged and neutral $\rho$-meson masses, so it has no contribution to the EM mass difference of $\rho^\pm-\rho^0$. The EM self-energies of $\rho$-meson in this case is easily obtained \begin{eqnarray} m^2_{\rm EM}(\rho^0)_{\rm GWZW}=m^2_{\rm EM}(\rho^\pm)_{\rm GWZW}= \frac{e^2 m_\rho^4}{128 \pi^6 g^2 f_\pi^2} =9.96\times 10^{-5}~{\rm GeV^2}. \end{eqnarray} There is also no divergences involved in Fig. 10. We take $m_\rho^2=m_\omega^2$ in deriving eq. (35), and its numerical result is small. Different from the case of $K^*(892)$-meson, the direct $\rho^0$-photon coupling which comes from VMD can bring the tree diagram contributing to EM masses of $\rho^0$, which has been shown in Fig. 1.1. It is easy to evaluate its contribution \begin{equation} m^2_{\rm EM}(\rho^0)_{\rm Tree}=-\frac{e^2 g^2}{4}m_\rho^2. \end{equation} Numerically \begin{equation} m^2_{\rm EM}(\rho^0)_{\rm Tree}=-2.06\times 10^{-3}~{\rm GeV^2}. \end{equation} Totally, $\rho^0-\rho^\pm$ EM mass difference is \begin{eqnarray} (m_{\rho^0}^2-m_{\rho^\pm}^2)_{\rm EM}=-9.49\times 10^{-4}~{\rm GeV^2},\\ (m_{\rho^0}-m_{\rho^\pm})_{\rm EM}=-0.62 ~{\rm MeV}. \end{eqnarray} The light quark mass terms of the QCD lagrangian are $$ {\cal L}_m=-m_u \bar{u} u-m_d \bar{d} d -m_s \bar{s}s- ... $$ It is important to note that the isospin violating piece of ${\cal L}_m$, $-\frac{1}{2}(m_d-m_u)(\bar{d} d-\bar{u} u)$, transforms as $\Delta I=1$ under the isospin subgroup. The operator which produces the $\pi^\pm-\pi^0$ (or $\rho^\pm-\rho^0$) mass difference must have $\Delta I=2$, therefore $(m_{\pi^\pm}-m_{\pi^0})_{\rm qm}= (m_{\rho^\pm}-m_{\rho^0})_{\rm qm}=0$ at the leading order in quark mass expansion (The subscript qm denotes the contribution from quark mass). This has been verified in the case of pions, as is well known that the $\pi^\pm-\pi^0$ mass difference is almost entirely electromagnetic in origin. The experimental value of $m_{\rho^0}-m_{\rho^\pm}$ \cite{PDG98} has a large error bar \begin{equation} m_{\rho^0}-m_{\rho^\pm}=0.1\pm0.9 ~{\rm MeV}. \end{equation} Recent result obtained by ALEPH Collaboration \cite{ALEPH97} is \begin{equation} m_{\rho^0}-m_{\rho^\pm}=0.0\pm1.0 ~{\rm MeV}. \end{equation} Comparing with eqs. (40) and (41), one will find that our prediction [eq. (39)] is in agreement with the measurements. This means that EM mass correction to the mass difference of $\rho$-meson is important, although its value is much smaller than the corresponding one of pions. The EM self-energies of $\omega$-meson can also be calculated. The tree level contribution from $\omega\longrightarrow\gamma\longrightarrow \omega$, which is \begin{equation} m^2_{\rm EM}(\omega)_{\rm Tree}=-\frac{e^2 g^2 m_\omega^2}{36}=-2.37\times 10^{-4}~{\rm GeV^2}. \end{equation} Besides the contribution from tree diagram, eq. (27) also contributes one-loop diagrams to the EM self-energies of $\omega$-meson (as shown in Fig. 11), and the loop diagrams give the finite contributions, \begin{equation} m^2_{\rm EM}(\omega)_{\rm 1-LOOP}=\frac{9e^2m_\rho^2m_\omega^2}{128 \pi^6 g^2 f_\pi^2}=4.62\times 10^{-4}~{\rm GeV^2}. \end{equation} Thus the numerical result of the total EM self-energies of $\omega$-meson to one-loop order and $O(\alpha_{\rm EM})$ is \begin{eqnarray} m^2_{\rm EM}(\omega)_{\rm total}=2.25\times 10^{-4}~{\rm GeV^2},\\ (m_\omega)_{\rm EM}=0.14~{\rm MeV}. \end{eqnarray} In the framework of the U(3)$_L\times$U(3)$_R$ chiral theory of mesons, the vector mesons are written as \begin{eqnarray*} v_{\mu}=\tau_{i}\rho^{i}_{\mu}+\lambda_a K_{\mu}^{*a}+(\frac{2}{3}+ \frac{1}{\sqrt{3}}\lambda_8)\omega_{\mu}+(\frac{1}{3}- \frac{1}{\sqrt{3}}\lambda_8)\phi_{\mu}, \end{eqnarray*} which means that the $\omega$-meson is free of $s\bar{s}$, $\phi$-meson is pure $s\bar{s}$, and the mixing of $\omega$ and $\phi$-meson is ideal. This is implied by the Okubo-Zweig-Iizuka (OZI) rule \cite{OZI}. In the large $N_C$ limit, the calculation of the EM self-energies of $\phi$-meson will be much simpler, only the tree level contribution from $\phi\longrightarrow \gamma\longrightarrow \phi$. Therefore we have \begin{eqnarray} m^2_{\rm EM}(\phi)=-\frac{e^2 g^2 m_\phi^2}{18}=-8.06\times 10^{-4}~{\rm GeV^2},\\ (m_\phi)_{\rm EM}=-0.40 ~{\rm MeV}. \end{eqnarray} Empirically, the U(3)$_L\times$U(3)$_R$ symmetry is broken due to the strong U(1) anomaly. Therefore the corrections from the next leading order corrections of large $N_C$ expansion to the processes which are related to the $\omega$ and $\phi$ mixing are non-trivial. However, this is beyond the scope of the present paper. In Ref. \cite{YG98}, the EM masses of $K^*(892)$-meson has been calculated, especially, one effect called as EM mass anomaly of massive Yang-Mills particles has been revealed. In the case of $\rho$-meson, it is easy to find that this conclusion also hold from eqs. (31) and (34) because the three-point and four-point MYM interactions contribute the negative values to the EM masses of $\rho^\pm$. Actually, this effect can also be seen in evaluating the EM self-energies of the axial-vector mesons $a_1$ and $K_1$ for these mesons are also introduced into the U(3)$_L\times$U(3)$_R$ chiral theory of mesons as the massive Yang-Mills particles. In paper I, EM masses of $a_1$ and $K_1$ have been calculated. Here we extract the contributions coming from the interaction lagrangians related to the three-point and four-point MYM vertices, \begin{eqnarray*} m_{\rm EM}^2(a^\pm)^{(3)}_{\rm MYM}=-0.002688~{\rm GeV^2},\\ m_{\rm EM}^2(a^\pm)^{(4)}_{\rm MYM}=-0.000648~{\rm GeV^2}, \end{eqnarray*} for $a_1$-meson, and \begin{eqnarray*} & &[m_{\rm EM}^2(K_1^\pm)-m_{\rm EM}^2(K_1^0)]^{(3)}_{\rm MYM} =-0.003474~{\rm GeV^2},\\ & &[m_{\rm EM}^2(K_1^\pm)-m_{\rm EM}^2(K_1^0)]^{(4)}_{\rm MYM} =-0.000781~{\rm GeV^2}, \end{eqnarray*} for $K_1$ meson. Here, it is obvious that the contributions from the MYM self-interaction make the EM masses of the neutral mesons larger than that of the charged one for the massive Yang-Mills particles. It is expected that this is a non-trivial and unusual effect for the massive Yang-Mills particles because it is contrary to the common knowledge of the hadron's EM-masses, such as $m({\rm neutron})_{\rm EM}<m({\rm proton})_{\rm EM}$, $m(\pi^0)_{\rm EM}<m(\pi^+)_{\rm EM}$, and $m(K^0)_{\rm EM}<m(K^+)_{\rm EM}$ \cite{GL82}. The studies in Refs. \cite{YG98,GY98} show that the experiment favors to support this EM masses anomaly effect. However, the experimental information for the EM masses of the vector and axial-vector mesons is rather poor, so this effect needs to be further investigated. \section{Summary} In terms of two equivalent representations of VMD (VMD1 and VMD2), the method developed in paper I to calculate the one-loop EM self-energies of the low-lying mesons has been naturally extended to the case of the vector sector including $\rho$, $\omega$, and $\phi$-mesons. Because all the parameters of the theory have been fixed previously, this is a parameter free study in the present paper. The theoretical result of the EM mass difference of $\rho$ meson is in agreement with the measurements, which means that the EM contribution is important to the $\rho^0-\rho^\pm$ mass difference. The EM self-energies of $\omega$ and $\phi$-mesons which make the shifts for the mass of the mesons are also evaluated. It has been shown that the effect of EM mass anomaly of the massive Yang-Mills particles, which has been revealed in Ref. \cite{YG98}, holds in the case of the $\rho$-meson. An elaborate analysis of the cases of $a_1$ and $K_1$ also supports this conclusion. This interesting effect should be further investigated, and also be tested by the experiments in the future. \begin{center} {\bf ACKNOWLEDGMENTS} \end{center} This work is partially supported by NSF of China through Chen Ning Yang and the Grant LWTZ--1298 of Chinese Academy of Sciences.
\section{Introduction and basic equations} \label{1sec} It is well-known that to investigate a dilute Bose gas of particles with an arbitrary strong repulsion (the strong-coupling regime), one should go beyond the Bogoliubov approach~\cite{Bog1} (the weak-coupling case) and treat the short-range boson correlations in a more accurate way. An ordinary manner of doing so is the use of the Bogoliubov model with the ``dressed", or effective, interaction potential containing ``information" on the short-range boson correlations (see Ref. \cite{Lee}). Below it is demonstrated that this manner leads to a loss of the thermodynamic consistency. To overcome this trouble, we propose a new way of investigating the strong-coupling regime which concerns the reduced density matrix of the second order~(the 2-matrix) and is based on the variational method. The 2-matrix for the many-body system of spinless bosons can be represented as~\cite{Bog2}: \begin{equation} \rho_2({\bf r}_1^{\prime},{\bf r}_2^{\prime};{\bf r}_1,{\bf r}_2) = {F_2({\bf r}_1,{\bf r}_2;{\bf r}_1^{\prime},{\bf r}_2^{\prime}) \over N(N-1)}, \label{1} \end{equation} where the pair correlation function is given by \begin{equation} F_2({\bf r}_1,{\bf r}_2;{\bf r}_1^{\prime},{\bf r}_2^{\prime})= \langle \psi^{\dagger}({\bf r}_1) \psi^{\dagger}({\bf r}_2) \psi ({\bf r}_2^{\prime})\psi ({\bf r}_1^{\prime})\rangle. \label{2} \end{equation} Here $\psi ({\bf r})$ and $\psi^{\dagger}({\bf r})$ denote the boson field operators. Recently it has been found~\cite{Chern1,Chern2} that for the uniform system with a small depletion of the zero-momentum state the correlation function (\ref{2}) can be written in the thermodynamic limit as follows: \begin{eqnarray} F_2({\bf r}_1,{\bf r}_2;{\bf r}_1^{\prime},{\bf r}_2^{\prime})&=& n_0^2\,\varphi^*(r)\,\varphi(r^{\prime})\nonumber\\ &&+2n_0\int \frac{d^3q}{(2\pi)^3}\, n_q\,\varphi_{{\bf q}/2}^*({\bf r}) \varphi_{{\bf q}/2}({\bf r}^{\prime})\nonumber\\ &&\times\exp\{ i {\bf q}({\bf R}^{\prime}-{\bf R})\}, \label{3} \end{eqnarray} where ${\bf r}={\bf r}_1 - {\bf r}_2, \;{\bf R}=({\bf r}_1 +{\bf r}_2)/2$ and similar relations take place for ${\bf r}^{\prime}$ and ${\bf R}^{\prime}$, respectively. In Eq.~(\ref{3}) $n_0=N_0/V$ is the density of the particles in the zero-momentum state, $n_q=\langle a_{\bf q}^{\dagger}a_{\bf q}\rangle$ stands for the distribution of the noncondensed bosons over momenta. Besides, $\varphi(r)$ is the wave function of a pair of particles being both condensed. In turn, $\varphi_{{\bf q}/2}({\bf r})$ denotes the wave function of the relative motion in a pair of bosons with the total momentum $\hbar{\bf q}$, this pair including one condensed and one noncondensed particles. So, Eq.~(\ref{3}) takes into account the condensate-condensate and supracondensate-condensate pair states and is related to the situation of a small depletion of the zero-momentum one-boson state. For the wave functions $\varphi(r)$ and $\varphi_{{\bf p}} ({\bf r})$ we have \begin{eqnarray} \varphi(r)\!=\!1+\psi(r),\ \varphi_{{\bf p}}\!({\bf r})\!=\! \sqrt{2}\cos({\bf p}{\bf r})+\psi_{{\bf p}}({\bf r})\;(p\not=0) \label{4} \end{eqnarray} with the boundary conditions $ \psi(r) \to 0$ and $\psi_{{\bf p}}({\bf r}) \to 0$ for $r \to \infty.$ The functions $\psi(r)$ and $\psi_{{\bf p}}({\bf r})$ can explicitly be expressed in terms of the Bose operators $a_{{\bf p}}^{\dagger}$ and $a_{{\bf p}}$ \cite{Chern1}: \begin{eqnarray} \widetilde{\psi}(k) &=&\int\psi(r)\exp(-i{\bf k}{\bf r})\,d^3r= \langle a_{{\bf k}}\,a_{-{\bf k}}\rangle/n_0, \label{6}\\ \widetilde{\psi}_{\bf p}({\bf k}) &=&\int\psi_{{\bf p}}({\bf r})\exp(-i{\bf k}{\bf r})\,d^3r \nonumber\\ &&=\sqrt{\frac{V}{2 n_0}} \frac{\langle a^{\dagger}_{2{\bf p}} a_{{\bf p}+{\bf k}} a_{{\bf p}-{\bf k}}\rangle}{n_{2p}}. \label{44} \end{eqnarray} Having in our disposal the distribution function $n_{k}$ and the set of the pair wave functions $\varphi(r)$ and $\varphi_{{\bf p}} ({\bf r})$, we are able to calculate the main thermodynamic quantities of the system of interest. In particular, the mean energy per particle is expressed in terms of $n_k$ and $g(r)$ via the well-known formula \begin{eqnarray} \varepsilon=\int \frac{d^3k}{(2\pi)^3} T_k \frac{n_k}{n}+ \frac{n}{2}\int g(r) \Phi(r) d^3r, \label{7} \end{eqnarray} where $T_k=\hbar^2 k^2/2m$ is the one-particle kinetic energy, $n=N/V$ stands for the boson density and the relation \begin{equation} g(r)=F_2({\bf r}_1,{\bf r}_2;{\bf r}_1,{\bf r}_2)/n^2. \label{8} \end{equation} is valid for the pair distribution function $g(r)$. \section{The Bogoliubov model} \label{2sec} The starting point of our investigation is the weak-coupling regime which implies weak spatial correlations of particles and, thus, is characterized by the set of the inequalities \begin{equation} |\psi(r)| \ll 1, \;\quad |\psi_{{\bf p}}({\bf r})| \ll 1\;. \label{9} \end{equation} Specifically, the Bogoliubov model corresponds to the choice~\cite{Chern1,Chern2} \begin{equation} |\psi(r)| \ll 1, \;\quad \psi_{{\bf p}}({\bf r}) = 0\;. \label{10} \end{equation} Besides, owing to a small depletion of the Bose condensate $(n-n_0)/n$ we have for the one-particle density matrix $F_{1}(r)=\langle\psi^{\dagger}({\bf r}_{1})\psi({\bf r}_{2})\rangle$: $$ \left|\frac{F_{1}(r)}{n}\right|= \left|\int\frac{d^3k}{(2\pi)^3}\frac{n_k}{n} \exp(i{\bf k}{\bf r})\right|\leq \frac{n-n_{0}}{n}\ll 1. $$ So, investigating the Bose gas within the Bogoliubov scheme, we have two small quantities: $\psi(r)$ and $F_{1}(r)/n$. This enables us to write Eq.~(\ref{8}) with the help of (\ref{3}) as follows: \begin{equation} g(r)=1+2\psi(r)+\frac{2}{n} \int\frac{d^3k}{(2\pi)^3}n_k\exp(i{\bf k}{\bf r}), \label{11} \end{equation} where we restricted ourselves to the terms linear in $\psi(r)$ and $F_{1}(r)/n$ and put $\psi^*(r)=\psi(r)$ because the pair wave functions can be chosen as real quantities. Equations for $\widetilde{\psi}(k)$ and $n_k$ can be found varying the mean energy (\ref{7}) with Eq. (\ref{11}) taken into account. However, previously one should realize an important point, namely: $n_k$ and $\widetilde{\psi}(k)$ can not be independent variables. Indeed, when there is no interaction between particles, there are no spatial particle correlations either. So, $\widetilde{\psi}(k)=0$ and, since the zero-temperature case is considered, all the bosons are condensed, $n_k=0$. While ``switching on" the interaction results in appearing the spatial correlations and condensate depletion: $\widetilde{\psi}(k)\not=0$ together with $n_k\not=0$. In the framework of the Bogoliubov scheme $\widetilde{\psi}(k)$ is related to $n_k$ by the expression \begin{equation} n_k (n_k+1)=n_0^2\widetilde{\psi}^2(k). \label{12} \end{equation} Indeed, the canonical Bogoliubov transformation~\cite{Bog1} implies that \begin{equation} a_{{\bf k}}=u_{k}\alpha_{{\bf k}}+v_{k} \alpha^{\dagger}_{-{\bf k}},\quad a^{\dagger}_{{\bf k}}=u_{k}\alpha^{\dagger}_{{\bf k}}+ v_{k}\alpha_{-{\bf k}}, \label{12a} \end{equation} where \begin{equation} u_{k}^2-v_{k}^2=1. \label{13} \end{equation} At zero temperature $\langle\alpha^{\dagger}_{{\bf k}} \alpha_{{\bf k}} \rangle=0$ and, using Eqs.~(\ref{6}) and (\ref{12a}) we arrive at \begin{equation} n_k=v_{k}^2, \quad \widetilde{\psi}(k)=u_{k}v_{k}/n_0. \label{14} \end{equation} With Eqs.~(\ref{13}) and (\ref{14}) one can readily obtain Eq. (\ref{12}). Now, let us show that all the results on the thermodynamics of a weak-coupling Bose gas can be derived for the Bogoliubov scheme with variation of the mean energy (\ref{7}) under the conditions (\ref{11}) and (\ref{12}). Inserting Eq. (\ref{11}) into Eq. (\ref{7}) and, then, varying the obtained expression, we arrive at \begin{equation} \delta\varepsilon=\int \frac{d^3k}{(2\pi)^3} \left\{\Bigl(T_k+ n\widetilde\Phi(k)\Bigr)\frac{\delta n_k}{n} + n\widetilde\Phi(k) \delta\widetilde\psi(k)\right\}. \label{15} \end{equation} Relation (\ref{12}) connecting $\widetilde\psi(k)$ with $n_k$ results in \begin{equation} \delta\widetilde{\psi}(k)=\frac{(2n_k+1)\delta n_k}{2n_0^2 \widetilde{\psi}(k)}+ \frac{\widetilde{\psi}(k)}{n_0}\int\frac{d^3q}{(2\pi)^3}\delta n_q, \label{16} \end{equation} where the equality \begin{equation} n=n_0+\int\frac{d^3k}{(2\pi)^3}\;n_k \label{17} \end{equation} is taken into consideration. Setting $\delta\varepsilon=0$ and using Eqs.~(\ref{15}) and (\ref{16}), we derive the following expression: \begin{eqnarray} -2\,T_k\widetilde{\psi}(k) &=&\frac{n^2}{n^2_0} \widetilde{\Phi}(k)(1+2n_k) +2n\,\widetilde{\psi}(k)\nonumber\\ &&\times\left(\widetilde{\Phi}(k)+\frac{n}{n_0} \int\frac{d^3q}{(2\pi)^3}\widetilde{\Phi}(q)\widetilde{\psi}(q)\right). \label{18} \end{eqnarray} Here one should realize that Eq.~(\ref{18}) is able to yield results being accurate only to the leading order in $(n-n_0)/n$ because the used expression for $g(r)$ given by Eq. (\ref{11}) is valid to the next-to-leading order~\cite{note4}. So, Eq.~(\ref{18}) should be rewritten as \begin{equation} -2\,T_k \widetilde{\psi}(k)= \widetilde{\Phi}(k)(1+2n_k)+2n\,\widetilde{\psi}(k)\Phi(k). \label{19} \end{equation} Equation (\ref{19}) is an equation of the Bethe-Goldstone type or, in other words, the in-medium Schr\"odinger equation for the pair wave function. As $2\widetilde{\Phi}(k)(n_k+n \widetilde{\psi}(k))$ is the product of the Fourier transforms of $\Phi(r)$ and $n(g(r)-1)$, we can rewrite Eq.~(\ref{19}) in the more customary form \begin{equation} \frac{\hbar^2}{m}\nabla^2\varphi(r)=\Phi(r) +n\int\Phi(|{\bf r}-{\bf y}|)\Bigl( g(y)-1\Bigr)d^3y. \label{20} \end{equation} The structure of Eq.~(\ref{20}) is discussed in the papers~\cite{Chern2,Shan1}. Here we only remark that the right-hand side (r.h.s.) of Eq. (\ref{20}) is the in-medium potential of the boson-boson interaction in the weak-coupling approximation. The system of equations (\ref{12}) and (\ref{19}) can easily be solved, which leads to the familiar results~\cite{Bog1}: \vspace*{-2mm} \begin{eqnarray} &&n_k=\frac{1}{2}\left(\frac{T_k+n\widetilde{\Phi}(k)}{\sqrt{T_k^2+ 2nT_k\widetilde{\Phi}(k)}}-1\right),\nonumber\\[-1mm] &&\widetilde{\psi}(k)=-\frac{ \widetilde{\Phi}(k)}{2\sqrt{T_k^2+2nT_k\widetilde{\Phi}(k)}}\,. \label{21} \end{eqnarray} \section{A dilute Bose gas within the Bogoliubov model} \label{3sec} As it was mentioned, the aim of our paper is investigation of the case of a dilute Bose gas with an arbitrary strong repulsion between bosons. So, considering a dilute Bose gas in the weak-coupling approximation can be a good exercise providing us with useful information. Let us investigate the thermodynamics of a dilute Bose gas within the Bogoliubov model. With Eqs.~(\ref{7}), (\ref{11}) and (\ref{21}) we derive \begin{eqnarray} \varepsilon&=&\frac{n}{2}\widetilde{\Phi}(0) +\frac{1}{2n}\int\frac{d^3k}{(2\pi)^3}\nonumber\\ &&\times\left(\sqrt{T_k^2+2nT_k\widetilde{\Phi}(k)}-T_k - n\widetilde{\Phi}(k)\right). \label{21a} \end{eqnarray} The well-known argument of Landau (see the footnote in Ref.~\cite{Bog1} and discussion in Ref.~\cite{Lee}) testifies that the properties of dilute quantum gases are ruled by the scattering length. Within the Bogoliubov model this length is usually assumed to be equal to $m\widetilde{\Phi}(0)/4\pi \hbar^2$. If so, when expanding $\varepsilon$ in powers of the boson density $n$, one could replace $\widetilde{\Phi}(k)$ by $\widetilde{\Phi}(0)$ in Eq. (\ref{21a}), introducing the low-momentum approximation. However, this leads to a divergency because at large $k$ the integrand behaves as $-n^2\widetilde{\Phi}^{2}(k)/2T_k$. To properly calculate the integral in Eq.~(\ref{21a}), we should rewrite Eq. (\ref{21a}) in the following form: \begin{eqnarray} \varepsilon=\frac{n}{2}\left(\widetilde{\Phi}(0)- \int\frac{d^3k}{(2\pi)^3} \frac{\widetilde{\Phi}^2(k)}{2T_k}\right)+I, \label{21b} \end{eqnarray} where \begin{eqnarray} I&=&\frac{1}{2n}\int\limits_{0}^{\infty}dk\frac{4\pi k^2}{(2\pi)^3} \nonumber \\ &&\times\Bigl(\sqrt{T_k^2+2nT_k\widetilde{\Phi}(k)} -T_k - n\widetilde{\Phi}(k)+ \frac{n^2\widetilde{\Phi}^2(k)}{2T_k}\Bigr). \nonumber \end{eqnarray} Now, substituting $k=(2 m n y)^{1/2}/\hbar$ in the integral, we obtain the expression \begin{eqnarray} I&=&\frac{\sqrt{2}}{4\pi^2}\left(\frac{m n}{\hbar^2}\right)^{3/2} \int\limits_{0}^{\infty} dy\biggl\{y\sqrt{y+ 2\widetilde{\Phi}(\sqrt{2m n y}/\hbar)}\nonumber\\ &&-y^{3/2} -\widetilde{\Phi}(\sqrt{2m n y}/\hbar)\,y^{1/2}+ \frac{\widetilde{\Phi}^2(\sqrt{2m n y}/\hbar)}{2\sqrt{y}} \biggr\} \nonumber \end{eqnarray} which at sufficiently small $n$ may be rewritten as \begin{eqnarray} I&=&\frac{\sqrt{2}}{4\pi^2} \left(\frac{m n}{\hbar^2}\right)^{3/2} \nonumber \\ &&\times\int\limits_{0}^{\infty} dy \Bigl\{ y\sqrt{y+2\widetilde{\Phi}(0)} -y^{3/2}-\widetilde{\Phi}(0)y^{1/2} +\frac{\widetilde{\Phi}^2(0)}{2\sqrt{y}}\Bigr\}. \nonumber \end{eqnarray} The derived integral is readily calculated. The result is given by \begin{equation} I=\frac{8}{15\pi^2} \left(\frac{m n}{\hbar^2}\right)^{3/2} \widetilde{\Phi}^{5/2}(0). \label{21e} \end{equation} In turn, the first term in the r.h.s. of Eq.~(\ref{21b}) can be represented as \begin{equation} \frac{n}{2}\left(\widetilde{\Phi}(0)- \int\frac{d^3k}{(2\pi)^3} \frac{\widetilde{\Phi}^2(k)}{2T_k}\right)= \frac{n}{2} \int \varphi^{(0)}(r) \Phi(r) d^3r, \label{21f} \end{equation} where $\varphi^{(0)}$ is the solution of Eq.~(\ref{20}) in the limit $n \to 0$. This is nothing else but the Schr\"odinger equation in the Born approximation. According to the relations (\ref{21e}) and (\ref{21f}) we has to conclude that the scattering length in the case of interest is expressed in the form \begin{equation} a_B=\frac{m}{4 \pi \hbar^2} \int \varphi^{(0)}(r) \Phi(r) d^3r. \label{21g} \end{equation} One can easily be convinced that $\widetilde{\Phi}(0)$ can not be represented only in terms of $a_B$ and, hence, the dependence on the shape of the interaction potential appears in the series expansion for the mean energy in the first correction to the term $2\pi \hbar^2 a_B n/m$. To rewrite our result for $\varepsilon$ in a graphic form, we introduce one more characteristic length $b>0$ which obeys the relation \begin{equation} b=-\frac{m}{4\pi \hbar^2}\int \psi^{(0)}(r) \Phi(r) d^3r= \frac{m}{4\pi \hbar^2}\int \frac{d^3 k}{(2\pi)^3} \frac{\Phi^2(k)}{2T_k}, \label{21h} \end{equation} where $\psi^{(0)}(r)=\varphi^{(0)}(r)-1$. Further, with the help of Eqs.~(\ref{21b})-(\ref{21h}), we arrive at \begin{equation} \varepsilon=\frac{2\pi \hbar^2 a_B n}{m} \left\{1+\frac{128}{15\sqrt{\pi}}\sqrt{n a_B^3}\left(1+ \frac{5b}{2 a_B}\right)+\cdots \right\}, \label{21i} \end{equation} here the condition $b \ll a_B$ is of use. It is not difficult to see that the expression (\ref{21h}) taken with negative sign is the next correction to the scattering length calculated within the Born approximation. As to the Eq.~(\ref{21g}), it is related to the next-to-Born approximation. Stress that in the Bogoliubov model the energy term $n\widetilde{\Phi}(0)/2$ is treated as the major one~\cite{Bog1}, which implies that the condition $b \ll a_B$ is fulfilled. This qualitative criterion can be written as \begin{equation} \widetilde{\Phi}(0) \gg \int \frac{d^3k}{(2\pi)^3} \frac{\widetilde{\Phi}^2(k)}{2T_k}. \label{21j} \end{equation} Beyond this inequality the model may be thermodynamically unstable. In particular, the opposite case $$ \widetilde{\Phi}(0) < \int \frac{d^3k}{(2\pi)^3} \frac{\widetilde{\Phi}^2(k)}{2T_k} $$ leads to the negative scattering length (\ref{21g}) which at sufficiently low densities results in $-\partial^{2}E/\partial V^{2}=\partial p/\partial V > 0$. Thus, investigated within the Bogoliubov model, the thermodynamics of a dilute Bose gas is ruled by the scattering length only in the zero-density limit. While the next-to-leading term in the series expansion given by Eq. (\ref{21i}) depends on the shape of the interaction, which is expressed in appearance of the additional characteristic length $b$. This conclusion differs from the results of papers~\cite{Lee} according to which the series expansion for $\varepsilon$ taken to the same order as that of Eq. (\ref{21i}), is fully determined by the scattering length. To clarify the situation concerning this difference, we should go to the strong-coupling regime. \section{The strong-coupling regime} \label{4sec} Now, after the detailed investigations of the Bogoliubov model within the scheme proposed, we are able to demonstrate that the investigation of the strong-coupling case based on the Bogoliubov model with the effective boson-boson interaction, results in a loss of the thermodynamic consistency. Indeed, as it was shown in the previous section, any calculating scheme using the basic relations of the Bogoliubov model (\ref{11}), (\ref{12}) conclusively leads to Eqs.~(\ref{19})-(\ref{21}) provided this scheme does yield the minimum of the mean energy. In this case Eqs.~(\ref{19})-(\ref{21}) certainly includes the quantity $\Phi(r)$ which is the ``bare" interaction potential appearing in Eq. (\ref{7}). The use of the Bogoliubov model with the effective interaction potential substituted for $\Phi(r)$ can in no way disturb the relations given by Eqs. (\ref{11}) and (\ref{12}). And Eq.~(\ref{7}) is the same in both the weak- and strong-coupling regimes. Thus, any attempts of replacing $\Phi(r)$ by the effective ``dressed" potential without modifications of Eqs. (\ref{11}) and (\ref{12}) results in a calculating procedure which does not really provide the minimum of the mean energy. It is nothing else but a loss of the thermodynamic consistency. We remark that we do not mean, of course, that the t-matrix approach or the pseudopotential method can not be applied in the quantum scattering problem. It is only stated that the usual way of combining the ladder diagrams with the random phase approximation faces the trouble mentioned above. Though our present investigation is limited to the consideration of the many-boson systems, the derived result gives a hint that the similar situation is likely to take place in the Fermi case, too. In this connection it is worth noting the problem associated with the lack of self-consistency of the standard method of treating the dilute Fermi gas~\cite{Fetter}. The strong-coupling regime is characterized by significant spatial correlations. So, Eq.~(\ref{10}) resulting in Eq. (\ref{11}) is not relevant for an arbitrary strong repulsion between bosons at small separations when we have $\psi(0)=-1, \quad \psi_{{\bf p}}(0)= -\sqrt{2}$ (see Refs.~\cite{Chern1,Chern2}). Therefore, to investigate the strong-coupling regime, Eq.~(\ref{11}) should be abandoned in favor of Eq. (\ref{3}). Expression (\ref{3}) is accurate to the next-to-leading order in $(n-n_0)/n$. So, using Eqs. (\ref{3}) and (\ref{8}), we can write \begin{equation} g(r)=\varphi^2(r)+\frac{2}{n}\int\frac{d^3q}{(2\pi)^3} n_q \left(\varphi^2_{{\bf q }/2}({\bf r})-\varphi^2(r) \right). \label{25} \end{equation} Let us now perturb $\widetilde{\psi}(k)$ and $n(k)$. Working to the first order in the perturbation and keeping in mind conditions (\ref{12}) and (\ref{25}), from Eq. (\ref{7}) we derive: \begin{equation} -2\,T_k \widetilde{\psi}(k)= \widetilde{U}(k)(1+2n_k)+2n\,\widetilde{\psi}(k) \widetilde{U}^{\prime}(k) \label{26} \end{equation} with \begin{equation} \widetilde{U}(k)=\int \varphi(r) \Phi(r) \exp(-i{\bf k}{\bf r}) d^3r \label{27} \end{equation} and \begin{eqnarray} \widetilde{U}^{\prime}(k)= \int\Bigl(\varphi_{{\bf k}/2}^2({\bf r}) -\varphi^2(r)\Bigr)\,\Phi(r)\,d^3r. \label{28} \end{eqnarray} Using Eqs.~(\ref{27}), (\ref{28}) as well as the relation $\psi_{\bf k}({\bf r}) \to \sqrt{2}\psi(r)\;(k \to 0)$ (see the boundary conditions (\ref{4}))~\cite{note1}, we obtain $\widetilde{U}(0)\not=\widetilde{U}^{\prime}(0).$ This implies that the system of Eqs.~(\ref{12}) and (\ref{26}) is not able to yield the relation $n_k\propto 1/k\;(k \to 0)$ following from the ``$1/k^2$" theorem of Bogoliubov for the zero temperature~\cite{Bog3}. Indeed, let us assume $n_k \to \infty$ for $k \to 0.$ Then, from Eq.~(\ref{12}) at $n=n_0$ we find $n|\widetilde{\psi}(k)|/n_k \to 1$ when $k \to 0.$ On the contrary, Eq.~(\ref{26}) gives $n|\widetilde{\psi}(k)|/n_k \to \widetilde{U}(0)/\widetilde{U}^{\prime}(0)\not=1$ for $k \to 0.$ So, consideration of the Bose gas based on Eqs.~(\ref{3}) and (\ref{12}) does not produce satisfactory results. Nevertheless, it is worth noting that Eq.~(\ref{26}) has an important peculiarity which differentiate it from Eq.~(\ref{19}) in an advantageous way. The point is that in both the limits $n\to 0$ and $k \to \infty$ Eq.~(\ref{26}) is reduced to \begin{equation} -\frac{\hbar^2}{m}\,\nabla^2\,\varphi(r)+\Phi(r)\varphi(r)=0. \label{28aa} \end{equation} As it is seen, this is the exact ``bare" (not in-medium) Schr\"odinger equation, other than its Born approximation following from Eq. (\ref{20}). Thus, we can expect the line of our investigation to be right. As it was shown in the previous paragraph, an approach adequate for a dilute Bose gas with an arbitrary strong interaction can not be constructed without modifications of Eq.~(\ref{12}). This is also in agreement with a consequence of the relation \begin{equation} |\langle a_{{\bf k}}\,a_{-{\bf k}}\rangle|^2 \leq \langle a_{{\bf k}}\,a_{{\bf k}}^{\dagger}\rangle \langle a_{-{\bf k}}^{\dagger}\,a_{-{\bf k}}\rangle \label{29} \end{equation} resulting from the inequality of Cauchy-Schwarz-Bo\-go\-liu\-bov~\cite{Bog3} $$ |\langle \widehat{A}\widehat{B}\rangle|^{2} \leq \langle\widehat{A}\widehat{A}^{\dagger}\rangle \langle\widehat{B}^{\dagger}\widehat{B}\rangle. $$ With Eqs. (\ref{6}) and (\ref{29}) one can easily derive $n_0^2 \widetilde{\psi}^2(k) \leq n_k(n_k+1)$. Thus, it is reasonable to assume that Eq.~(\ref{12}) takes into account only the condensate-condensate channel and ignores the supracondensate-condensate ones. Now the question arises how to find corrections to the r.h.s. of Eq.~(\ref{12}). At present we have no regular procedure allowing us to do this in any order of $(n-n_0)/n$. However, there exists an argument which makes it possible to realize the first step in this direction. The matter is that the alterations needed have to produce the equation for $\widetilde{\psi}_{\bf p}({\bf k})$ which is reduced to the equation for $\widetilde{\psi}(k)$ in the limit $p \to 0.$ Though this requirement does not uniquely determine the corrections to Eq.~(\ref{12}), it turns out to be significantly restrictive. In particular, even the simplest variant of correcting Eq.~(\ref{12}) in this way, leads to promising results. Indeed, this variant is specified by the expression \begin{eqnarray} n_k(n_k+1)= n_0^2\,\widetilde{\psi}^2(k) +2 n_0\int \frac{d^3q}{(2\pi)^3}\,n_q \widetilde{\psi}^2_{{\bf q}/2}({\bf k}). \label{30} \end{eqnarray} Eq.~(\ref{30}) is valid to the next-to-leading order in $(n-n_0)/n$. So, we may rewrite it as \begin{equation} n_k(n_k+1)=n^2\widetilde{\psi}^2\!(k)+2n\!\int\! \frac{d^3q}{(2\pi)^3} n_q\!\left(\!\widetilde{\psi}^2_{{\bf q }/2} ({\bf k})-\widetilde{\psi}^2(k)\!\right)\!. \label{31} \end{equation} Perturbing $\widetilde{\psi}(k)$ and $n_k$ and bearing in mind conditions (\ref{25}) and (\ref{31}), Eq.~(\ref{7}) gives Eq.~(\ref{26}) again. However, now $\widetilde{U}^{\prime}(k)$ obeys the new relation \begin{eqnarray} \widetilde{U}^{\prime}(k)&=& \int\Bigl(\varphi_{{\bf k}/2}^2({\bf r}) -\varphi^2(r)\Bigr)\,\Phi(r)\,d^3r \nonumber\\ &&-\int\frac{d^3q}{(2\pi)^3} \frac{\widetilde{U}(q) \bigl(\widetilde{\psi}^2_{{\bf k}/2}({\bf q})- \widetilde{\psi}^2(q)\bigr)}{\widetilde{\psi}(q)} \label{33} \end{eqnarray} which significantly differs from Eq. (\ref{28}). Indeed, the choice of the pair wave functions as real quantities implies that operating with integrands in Eqs. (\ref{27}) and (\ref{33}), one can exploit $\psi_{{\bf p}}({\bf r})-\sqrt{2} \psi(r) \propto p^2$ at small $p$~\cite{note3}. For $k \to 0$ this provides $\widetilde{U}^{\prime}(k)-\widetilde{U}(k)= t_k= c\,k^4 +\cdots$. Note that for $k \to \infty$ we have $t_k \to -\widetilde{U}(0)$. Similar to Eq.~(\ref{19}), Eq.~(\ref{26}) can yields results correct only to the leading order in $(n-n_0)/n$. So, it has to be solved together with Eq. (\ref{12}) where $n_0^2$ should be replaced by $n^2$, rather than with Eq. (\ref{31}). This leads to the following relations: \begin{eqnarray} &&n_k=\frac{1}{2}\left(\frac{\widetilde{T}_k+n\widetilde{U}(k)} {\sqrt{\widetilde{T}_k^2+2n\widetilde{T}_k\widetilde{U}(k)}} -1\right),\label{35a}\\ &&\widetilde{\psi}(k)=-\frac{ \widetilde{U}(k)}{2\sqrt{\widetilde{T}_k^2+2n\widetilde{T}_k \widetilde{U}(k)}}, \label{36} \end{eqnarray} where $\widetilde{T}_k=T_k+nt_k$, with the limit $\widetilde{T}_k/T_k \to 1$ at $k \to \infty$. For $k\to0$ Eq.~(\ref{35a}) gives $n_k \simeq (\sqrt{n\,m\,\widetilde{U}(0)}/ \hbar k-1)/2$, which is fully consistent with the ``$1/k^2$" theorem of Bogoliubov for the zero temperature~\cite{Bog3}. As it is seen, the strong-coupling regime is more complicated than the Bogoliubov one because we do not know the quantity $\widetilde{U}(k)$ {\it ab initio}. To find it, one should solve Eqs. (\ref{27}) and (\ref{36}) in a self-consistent manner. Equations (\ref{27}) and (\ref{36}) lead to one more interesting relation \begin{equation} \widetilde{U}(k)=\widetilde{\Phi}(k)- \frac{1}{2}\int\frac{d^3q}{(2\pi)^3} \frac{\widetilde{\Phi}(|{\bf k} -{\bf q}|)\widetilde{U}(q)}{\sqrt{\widetilde{T}^2_q+ 2n\widetilde{T}_q\widetilde{U}(q)}} \label{37} \end{equation} which can be called the in-medium Lippmann-Schwinger equation for the scattering amplitude. To obtain the expansion for the energy at low densities, we must solve Eq.~(\ref{37}) at $n \to 0$. Let us rewrite it in the form $$ \widetilde{U}(k)=\widetilde{\Phi}(k)- \frac{1}{2}\int \frac{d^3q}{(2\pi)^3} \frac{\widetilde{\Phi}(|{\bf k} -{\bf q}|)\widetilde{U}(q)}{T_q}-I_1, $$ where for $I_1$ we have $$ I_1=\frac{1}{2}\!\int\!\frac{d^3q}{(2\pi)^3}\!\left\{ \frac{\widetilde{\Phi}(|{\bf k} -{\bf q}|)\widetilde{U}(q)}{\sqrt{\widetilde{T}^2_q+ 2n\widetilde{T}_q\widetilde{U}(q)}}- \frac{\widetilde{\Phi} (|{\bf k}-{\bf q}|)\widetilde{U}(q)}{T_q}\!\right\}. $$ Operating with $I_1$ in the same manner as we dealt with $I$ in the section~\ref{3sec} and taking into account that $t_k=0$ at $k=0$, for $n \to 0$ we derive \begin{equation} I_1= - \alpha \widetilde{\Phi}(k), \quad \alpha=\frac{\sqrt{n m^3}}{\pi^2\hbar^3}\widetilde{U}^{3/2}(0). \label{38} \end{equation} From Eqs.~(\ref{37}) and (\ref{38}) it now follows that \begin{eqnarray} \widetilde{U}(k)-\widetilde{U}^{(0)}(k)&=& \alpha\widetilde{\Phi}(k) -\int \frac{d^3q}{(2\pi)^3} \frac{\widetilde{\Phi}(|{\bf k} -{\bf q}|)}{2T_q}\nonumber \\ &&\times\Bigl(\widetilde{U}(q) -\widetilde{U}^{(0)}(q)\Bigr). \label{39} \end{eqnarray} Here $\widetilde{U}^{(0)}(k)=\int \varphi^{(0)}(r) \Phi(r) \exp(-i{\bf k}{\bf r})d^3r$ but now $\varphi^{(0)}(r)$ obeys Eq.~(\ref{28aa}) rather than Eq.~(\ref{20}) taken in the limit $n \to 0$ like in the section~\ref{3sec}. Let us introduce the new quantity $\widetilde{\xi}(q)=-(\widetilde{U}(q)- \widetilde{U}^{(0)}(q))/2T_q$. Then, for its Fourier transform $\xi(r)$ we obtain \begin{equation} -\frac{\hbar^2}{m}\nabla^2\Bigl(\alpha + \xi(r)\Bigr) +\Phi(r)\Bigl(\alpha+\xi(r)\Bigr)=0, \label{40} \end{equation} here $\xi(r) \to 0$ when $r \to \infty$. Comparing Eq.~(\ref{40}) with Eq. (\ref{28aa}), we find $\xi(r)=\alpha \psi^{(0)}(r)$. Hence, for $n \to 0$ we have \begin{equation} \widetilde{U}(k) \simeq \widetilde{U}^{(0)}(k)\left(1 +\gamma(k,n) \frac{8}{\sqrt{\pi}}\sqrt{na^3}\right). \label{41} \end{equation} Here $\gamma(k,n) \to 1$ when $n \to 0$. Besides, the relation $\widetilde{U}^{(0)}(0)=4\pi \hbar^2 a/m$ is used in Eq.~(\ref{41}), where $a$ is the scattering length. Having in our disposal Eq.~(\ref{41}), we are able to calculate the expansion in powers of $n$ for the condensate depletion and energy of a dilute Bose gas with an arbitrary strong interparticle potential. Considering the condensate depletion $(n-n_0)/n= 1/(2\pi)^3\int_{0}^{+\infty}dk\, 4\pi k^2 n_k/n$, with the help of Eq. (\ref{35a}) we obtain \begin{equation} \frac{n-n_0}{n}=\frac{8}{3\sqrt{\pi}}\sqrt{n a^3}+\cdots. \label{42} \end{equation} Notice that according to Eq. (\ref{41}) one can expect that among the omitted terms in Eq.~(\ref{42}) there is one proportional to $n a^3$. The most simple way of deriving the expansion for the mean energy per particle is based on using the chemical potential which, in the presence of the Bose condensate, is given by \begin{equation} \mu=\frac{1}{\sqrt{n_0}}\int d^3 r^{\prime} \Phi(|{\bf r}-{\bf r}^{\prime}|) \langle \psi^{\dagger}({\bf r}^{\prime}) \psi({\bf r}^{\prime}) \psi({\bf r})\rangle. \label{43}\end{equation} This formula follows from the well-known expression $ \delta \Omega = \langle\delta \left(\hat H-\mu \hat N\right) \rangle, $ where $\delta\Omega$ is an infinitesimal change of the grand canonical potential, and relation~(see Ref. \cite{Bog3}) $$ \frac{\partial \Omega(N_0,\mu,T)}{\partial N_0}=0. $$ Using the specific expressions for the scattering parts of the condensate-condensate and supracondensate-condensate pair wave functions \cite{Chern1} given by Eqs. (\ref{6}) and (\ref{44}) one can represent Eq.~(\ref{43}) in the following form: \begin{equation} \mu=n_0 \widetilde{U}(0)+ \sqrt{2}\int \frac{d^3q}{(2\pi)^3}\;n_q\; \widetilde{U}_{{\bf q}/2}({\bf q}/2), \label{45}\end{equation} here $$ \widetilde{U}_{{\bf p}}({\bf k})= \int \varphi_{\bf p}({\bf r}) \Phi(r) \exp(-i{\bf k}{\bf r}) d^3 r. $$ Now, for $n \to 0$ (see the procedure of calculating the integral $I$ in the section~\ref{2sec} one can rewrite Eq.~(\ref{45}) as \begin{equation} \mu=n\widetilde{U}(0)\left(1+\frac{n-n_0}{n}+ \cdots\right). \label{46}\end{equation} Inserting Eqs. (\ref{41}) and (\ref{42}) into Eq. (\ref{46}), we arrive at \begin{equation} \mu=\frac{4\pi \hbar^2 a n}{m}\left(1+\frac{32}{3\sqrt{\pi}} \sqrt{n a^3}+\cdots\right). \label{47} \end{equation} This result for the chemical potential implies, due to the basic thermodynamic formula $\mu=\partial(\varepsilon n)/\partial n$, the following expansion for the mean energy per particle: \begin{equation} \varepsilon=\frac{2 \pi \hbar^2 a n}{m}\left(1+ \frac{128}{15\sqrt{\pi}}\sqrt{n a^3}+\cdots\right). \label{48}\end{equation} As it is seen, the relation (\ref{48}) coincides with the well-known result of the approach~\cite{Lee} being reduced to the Bogoliubov model with the ``dressed'' interaction. It is not a surprise because according to the conclusions of the section~\ref{2sec}, we know that the numerical factor $128/(15\sqrt{\pi})$ appears in the series expansion for the mean energy per particle within the Bogoliubov model (see Eq. (\ref{21i})). Replacing the bare interaction potential by the ``dressed'' one results in replacing the scattering length $a_B$ in Eq.(\ref{21i}) by its exact value $a$. The only problem of doing so concerns the parameter $b$. Indeed, it follows from Eq. (\ref{21h}) that substituting the hard-sphere potential $\widetilde{U}(0)=4\pi\hbar^2 a/m$ for $\widetilde{\Phi}(k)$ leads to the familiar divergency~(see, e.g. Ref. \cite{Fetter}, p. 314). This obstacle has been overcome with the help of the well-known argument of Landau~(see the footnote in the paper~\cite{Bog1}) stating that the thermodynamics of dilute quantum gases is only ruled by the vacuum scattering amplitude. According to this reasoning one can expect that dependence on the shape of the interaction potential should not appear in the first orders of the density series expansion of the thermodynamic quantities. So, various regularizing procedures, more or less speculative, have been worked out in order to exclude this divergence~(together with the parameter $b$). On the contrary, there are no problems like this within the approach of the present paper. Here Eq.~(\ref{48}) is derived on the solid theoretical basis rather than with the help of Landau's argument. In spite of its reasonable character, it needed to be corroborated, and the results of this paper given by Eqs.~(\ref{42}), (\ref{46}) and (\ref{48}) have proved the validity of Landau's argument beyond any inconsistencies and divergencies. In the weak-coupling case when $|\Phi(r)| \ll 1$, the energy per particle calculated within our scheme is expressed by Eq.~(\ref{48}) with $a$ replaced by $a_B$. So, the appearance of the parameter $b$ in the results of the section \ref{2sec} is an artifact following from the neglect of scattering in the supracondensate-condensate pair wave channel. The divergence mentioned in the previous paragraph is not typical of the strong-coupling perturbation theory for the many-boson systems but results from, say, the weak-coupling spirit of the approach of Ref. \cite{Lee}. A simple way to be convinced of this is to consider the spatial boson correlations. Taken to the lowest-order with respect to the density, the structural factor (see the last paper in Ref.~\cite{Lee}) is of the form \begin{equation} S(k)=\frac{T_k}{\sqrt{T_k^2+2nT_k \widetilde{U}^{(0)}(k)}}. \label{49} \end{equation} By definition we have \begin{equation} g(r)=1+\frac{1}{n} \int \frac{d^3k}{(2\pi)^3} (S(k)-1) \exp(i{\bf k}{\bf r}). \label{50} \end{equation} Using Eqs. (\ref{49}) and (\ref{50}), for $n \to 0$ one can readily find \begin{equation} g(r) \to 1+2\psi^{(0)}(r), \label{51} \end{equation} where $\psi^{(0)}(r)$ obeys Eq.~(\ref{28aa}). This result answers the approximation (\ref{11}) while $\psi^{(0)}(r)$ is not related to the weak-coupling regime and obeys the exact ``bare'' Schr\"odinger equation. In the situation $\Phi(r) \to \infty$ for $r \to 0$ one has $\psi^{(0)}(r=0)=-1$, which implies, according to Eq. (\ref{51}), $g(r=0) \to -1$ for $n \to 0.$ It is not consistent with the physical sense of $g(r)$ and has nothing to do with the strong-coupling case corresponding to Eq. (\ref{25}) when for $n \to 0$ $$ g(r) \to \left(1+\psi^{(0)}(r)\right)^2. $$ Notice that the zero-density limits for the thermodynamic quantities of a strongly interacting dilute Bose gas were first found in the Bogoliubov original paper~\cite{Bog1}: $$ (n-n_0)/n\to 0,\ g(r)\to(\varphi^{(0)}(r))^{2},\ \varepsilon/n\to \widetilde{U}^{(0)}(0)/2. $$ At last, we remark that due to the incorrect picture of the spatial boson correlations found in papers~\cite{Lee}, one can expect significant alterations for the spectrum of the elementary excitations too. However, to clarify these corrections we should conclusively solve the problem concerning relation between the momentum distribution and scattering parts of the pair wave functions. Indeed, it has been mentioned that there exist various possibilities of generalizing Eq.~(\ref{12}) so as to obtain the equation for $\widetilde{\psi}_{\bf p}({\bf k})$ being reduced to the equation for $\widetilde{\psi}(k)$ in the limit $p \to 0.$ These possibilities result in the same series expansions for the thermodynamic quantities (\ref{42}), (\ref{47}) and (\ref{48}) but produce different data for the long-range spatial boson correlations. Here we limited ourselves to considering the most simple variant of generalizing Eq. (\ref{12}), which makes it possible to investigate only the thermodynamics of a strongly interacting Bose gas. The interesting and important problem of the spectrum of the elementary excitations is thus beyond the scope of this paper and will be the subject of the future investigations. \section{Conclusion} \label{5sec} Concluding let us take notice of the important points of this paper once more. It was demonstrated that thermodynamically consistent calculations based on Eqs. (\ref{11}) and (\ref{12}) conclusively result in Eqs.~(\ref{19})-(\ref{21}). Therefore, using the Bogoliubov model with the ``dressed" interaction does not provide the satisfactory solution of the problem of the strong-coupling Bose gas. As it was shown, when investigating this subject, one should go beyond the Bogoliubov scheme. To do this, we developed the approach reduced to the system of Eqs.~(\ref{27}), (\ref{33}), (\ref{35a}) and (\ref{36}). These equations leading to the in-medium Lippmann-Schwinger equation (\ref{37}), reproduce the familiar results (\ref{42}), (\ref{46}) and (\ref{48}) for the condensate depletion, chemical potential and mean energy but yield completely different picture of the spatial boson correlations. This difference should manifest itself in the next orders of the density series expansions for the thermodynamic quantities and in the excitation spectrum as well. This work was supported by the RFBR Grant No. 97-02-16705.
\section{Introduction} The investigation of $\phi$-meson production in both hadronic \cite{Ast,CrB,Obe,Ber95,Wur96,Disto,Ell95,Gut97,Loche,Buzat,Klemp,Mul94,Rekal,Tito1,Kondr} and electromagnetic \cite{Willi,Tito2} processes at low and intermediate energies has attracted much attention in recent years. A major motivation for those studies is the hope that one can obtain information about the amount of hidden strangeness in the nucleon. Indications for a possibly significant $\bar ss$ component in the nucleon have been found, among others, in the analysis of the so-called $\pi N$ $\Sigma$ term \cite{Don86} and from the EMC measurement of deep inelastic polarized $\mu p$ scattering (``nucleon-spin crisis'') \cite{EMC88}. In the context of $\phi$-meson production processes one expects \cite{Ell89,Hen92} that a large amount of hidden strangeness in the nucleon would manifest itself in reaction cross sections that significantly exceed the values estimated from the OZI rule \cite{OZI}. This phenomenological rule states that reactions involving disconnected quark lines are forbidden. In the naive quark model the nucleon has no $\bar ss$ content, whereas the $\phi$-meson is an ideally mixed pure $\bar ss$ state. Thus, in this case, the OZI rule implies vanishing nucleon-nucleon-$\phi$ ($NN\phi$) coupling and, accordingly, a negligibly small production of $\phi$-mesons from nucleons (or anti-nucleons) by electromagnetic or (non-strange) hadronic probes. In practice there is a slight deviation from ideal mixing of the vector mesons, which means that the $\phi$-meson has a small $\bar uu + \bar dd$ component. Thus, even if the OZI rule is strictly enforced, there is a non-zero coupling of the $\phi$ to the nucleon, although the coupling is very small. Its value can be used to calculate lower limits for corresponding cross sections. For example, under similar kinematic conditions to cancel out phase space effects, one expects cross section ratios of reactions involving the production of a $\phi$- and an $\omega$-meson, respectively, to be \cite{Lip76} \begin{equation} R = {\sigma (A+B\rightarrow \phi X) \over \sigma (A+B\rightarrow \omega X)} \approx \tan ^2 (\alpha_V)\ , \label{Ratio} \end{equation} where $A$, $B$ and $X$ are systems that do not contain strange quarks. Here, $\alpha_V \equiv \theta_V - \theta_{V(ideal)}$ is the deviation from the ideal $\omega - \phi$ mixing angle. With the value $\alpha_V \cong 3.7^o$ \cite{PDG} one gets the rather small ratio of $R = 4.2 \times 10^{-3}$. Recent experiments on antiproton-proton ($\bar pp$) annihilation at rest at the LEAR facility at CERN revealed, however, considerably larger branching ratios for various $\phi$ production channels \cite{Ast,CrB,Obe}. Many of the measured ratios $\sigma(\bar pp \rightarrow \phi X) / \sigma(\bar pp \rightarrow \omega X)$ are about $100 \times 10^{-3}$ or even larger (cf. the compilation of data given in Ref.~\cite{Ell95}), which means that they exceed the estimate from the OZI rule by more than one order of magnitude. Significant deviations from the OZI rule were also found in the reactions $\bar p p\rightarrow \phi\phi$ \cite{Ber95} and $pd \rightarrow ^3$He$\phi$ \cite{Wur96}. These observed large $\phi$-production cross sections were interpreted by some authors as a clear signal for an intrinsic $\bar ss$ component in the nucleon \cite{Ell95,Gut97}. However, this explanation is still controversial. There is an alternative approach aimed at understanding these large cross sections solely by the strong rescattering effects in the final state \cite{Loche,Buzat,Klemp,Mul94}. In this case $\phi$ production occurs via two-step processes with intermediate states such as $\bar KK$, $\bar KK^*$, $\bar \Lambda \Lambda$, etc., where each step is allowed by the OZI rule. Corresponding quantitative calculations have, indeed, demonstrated that the resulting production cross sections are sufficiently large to agree with the experiments without a violation of the OZI rule \cite{Loche,Buzat,Klemp,Mul94}. In this connection, the production of $\phi$-mesons in proton-proton ($pp$) collisions is of special interest. In principle, the production cross section can be used for a direct determination of the $NN\phi$ coupling strength. Any appreciable $NN\phi$ coupling in excess of the value given by the OZI rule could be seen as evidence for hidden strangeness in the nucleon. Of course, there is also an alternative picture: one in which the coupling of the $\phi$-meson to the nucleon does not occur via possible $\bar ss$ components in the nucleon, but via intermediate states with strangeness \cite{Geiger,Meissner}. Specifically, this means that the $\phi$-meson couples to the nucleon via virtual $\Lambda K$, $\Sigma K$, etc. states. Corresponding model calculations have shown, however, that such processes give rise to (effective) $NN\phi$ coupling constants comparable to the OZI values and therefore should not play a role in drawing conclusions concerning hidden strangeness in the nucleon. Accordingly, one expects that cross section ratios $\sigma(pp\rightarrow pp\phi)/ \sigma(pp\rightarrow pp\omega)$ should provide a clear sign for a possible OZI violation. Indeed, data presented recently by the DISTO collaboration \cite{Disto} indicate that this ratio, after correcting for the phase space effects, is about eight times larger than the OZI estimate (\ref{Ratio}). The present work focuses on the $pp \rightarrow pp\phi$ process. We discuss the experimental prerequisites that would enable one to disentangle competing reaction mechanisms for $\phi$-production in nucleon-nucleon ($NN$) collisions and to extract the value of the $NN\phi$ coupling strength. In this context we shall show that the angular distribution of the produced $\phi$-mesons plays a crucial role. In particular, we will demonstrate that the almost isotropic angular distribution seen in the data from the DISTO collaboration \cite{Disto} indicates that the $NN\phi$ coupling is very small. In the absence of a set of data sufficient to permit an accurate determination of the $NN\phi$ coupling strength, we impose certain reasonable assumptions and carry out a combined analysis of the recent $\phi$-meson production data of the DISTO collaboration \cite{Disto} and the data of $\omega$-meson production from SATURNE \cite{Saclay}. This analysis yields a range of values for $g_{NN\phi}$ that is compatible with the OZI rule. We describe the $pp \rightarrow pp\phi$ reaction within a relativistic meson-exchange model. (See Ref.~\cite{Nak1} for the details of the formalism.) The transition amplitude is calculated in Distorted Wave Born Approximation, where the $NN$ final state interaction is taken into account explicitly. The final state interaction is known to be very important at near-threshold energies. For the $NN$ interaction we employ the model Bonn B as defined in Table A.1 of Ref.~\cite{MHE87}. This model reproduces the $NN$ phase shifts up to the pion-production threshold as well as the deuteron properties \cite{MHE87}. As in our previous work \cite{Nak1}, we do not consider the initial state interaction explicitly. At the corresponding high incident energies the $NN$ interaction is a slowly varying function of energy. Its main effect is to lead to an overall reduction of the $pp\rightarrow pp\phi$ cross section \cite{Hanhart}, as has been shown explicitly for the case of the reaction $pp \rightarrow pp\eta$ by Batini\'c et al. \cite{BSL}. In the present model this effect of the initial state interaction is accounted for by an appropriate adjustment of the (phenomenological) form factors at the hadronic vertices. In the next section we discuss possible basic production currents that can contribute to the reaction $pp\rightarrow pp\phi$. Using SU(3) flavor symmetry and imposing the OZI rule, we calculate meson-meson-meson and nucleon-nucleon-meson coupling constants that are relevant for the $\phi$-production currents and then estimate the corresponding contributions to the total cross section. These reveal that the $\phi\rho\pi$ meson-exchange current is the dominant one. In the same section we also give a short outline of our formalism and introduce the free parameters of our model. In Sect.\ III we describe in detail the procedure for the combined analysis of $\phi$- and $\omega$-production data. We conclude with a summary of our results in the last section. \section{Production currents} In the case of the reaction $pp\rightarrow pp\omega$, the dominant production mechanisms ---namely, the nucleonic and $\omega\rho\pi$ mesonic currents, as depicted in Fig.~\ref{fig1}---can be identified without involved considerations. This is because of the relatively large $NN\omega$ and $\omega\rho\pi$ coupling strengths. The $\omega\rho\pi$ meson-exchange current depends also on the $NN\rho$ and $NN\pi$ couplings, which are likewise large. Therefore this production mechanism is by far the dominant one among the possible exchange currents \cite{Nak1}. For $\phi$-meson production, however, the situation is much less clear {\it a priori} and requires careful consideration. In order to determine the importance of various possible meson-exchange currents, we first estimate systematically the coupling strengths of the hadronic vertices that enter into these $\phi$-production exchange currents. This is done using effective Lagrangians, assuming SU(3) flavor symmetry, and imposing the OZI rule. We note that in this scheme the SU(3) symmetry breaking is introduced through the use of the physical masses of the hadrons in calculating the observables. In the present work we are interested in three-meson vertices involving at least one $\phi$-meson. We, then, have $VVP$, $VPP$ and $VVV$ vertices ($V =$ vector meson, $P =$ pseudo-scalar meson). The SU(3) effective Lagrangian has the form \begin{eqnarray} \nonumber {\cal L}_{MMM} & = & g_{888}\left[-(1-\beta)Tr([M_8,M_8]M_8) + \beta Tr(\{M_8,M_8\}M_8) \right] \\ \nonumber & + & g_{88s}\beta Tr(\{M_8,M_8\}M_s) + g_{8s8}\beta Tr(\{M_8,M_s\}M_8) \\ & + & g_{sss}\beta Tr(\{M_s,M_s\}M_s) \ , \label{SU3LM} \end{eqnarray} where $M_8$ ($M_s$) stands for the SU(3) meson-octet (-singlet) matrix, and $\beta$ is the parameter specifying the admixture of the $D$-type ($\beta =1$) and $F$-type ($\beta = 0$) couplings. The OZI condition relates the basic coupling constants of the SU(3) octet and singlet members in the Lagrangian, $g_{sss}=g_{88s}=g_{8s8}=\sqrt{{2\over 3}} \ g_{888}$, so that there is only one free coupling constant. This one is then fixed by a fit to some appropriate observables such as decay rates. The $VPP$ and $VVV$ vertices involve only $F$-type coupling if we require charge conjugation invariance. This leads to the vanishing of all $VPP$ and $VVV$ couplings involving either the $\omega$- or $\phi$-meson relevant to the present work. The $D$-type coupling leads to G-parity violating vertices and will be ignored. In contrast, the $VVP$ vertices involve only $D$-type coupling. In Ref.~\cite{Durso} essentially the same scheme as proposed here has been used to describe radiative meson decays. Therefore, we use the model parameters determined in that paper. The effective Lagrangian used in Ref.~\cite{Durso} can easily be cast into the form given by Eq.(\ref{SU3LM}). The resulting coupling constants $g_{VVP}$ at those vertices relevant for the meson-exchange currents of interest are tabulated in Table~\ref{tab1}. In order to estimate the magnitude of the various meson-exchange currents one needs not only the $VVP$ couplings, but the corresponding $NNV$ and $NNP$ couplings as well. Thus, in addition to the $NN\pi$ and $NN\rho$ couplings, the empirically poorly known $NN\omega$, $NN\phi$, $NN\eta$ and $NN\eta '$ couplings are needed. We estimate these coupling strengths using the same scheme used for determining the $VVP$ couplings, i.e., the SU(3) effective Lagrangian \begin{eqnarray} \nonumber {\cal L}_{BBM} & = & g_8\left[-(1-\beta)Tr([\bar B,B]M_8) + \beta Tr(\{\bar B,B\}M_8) \right] \\ & + & g_s \beta Tr(\{\bar B,B\}M_s) \ , \label{SU3LB} \end{eqnarray} plus the OZI rule. This latter condition yields $g_s = (3-4\beta)g_8/\sqrt{6}$. In the above Lagrangian $B$ stands for the SU(3) baryon-octet matrix. The $D$-type coupling in the SU(3) Lagrangian is not allowed for the $NNV$ couplings if one requires spin-independence for $BB\omega$ and $BB\phi$ couplings within the identically flavored baryons, $B=\Sigma, \Lambda$. Using the value of the $\omega - \phi$ mixing angle from Ref.~\cite{PDG} and $g_8 (= g_{NN\rho})$ from Ref.~\cite{MHE87}, we find $g_{NN\omega} \cong 9$ and $g_{NN\phi} \cong -0.6$ for the $NN\omega$ and $NN\phi$ vector-coupling constants, respectively. The $NNv$ tensor-coupling constants are simply chosen to be $f_{NNv} = \pm 0.5 \ g_{NNv}$, where $v \equiv \omega, \ \phi$ (as distinct from $V$, which stands for vector mesons in general). This choice of $f_{NNv}$ is consistent with values from an SU(3) estimate \cite{Dover} as well from other sources \cite{Meissner,MHE87}. For the $NN\eta$ coupling constant we take the value used in $NN$ scattering analysis \cite{MHE87}, $g_{NN\eta}=6.14$. This value, together with the $\eta - \eta'$ mixing angle of $\theta_P \cong -9.7^o$, as suggested by the quadratic mass formula, and the $NN\pi$ coupling constant, $g_{NN\pi}= 13.45$, leads to the ratio $D/F \equiv \beta / (1 - \beta) \cong 1.43$ -- which is close to the value of $D/F = 1.58 \pm 0.07$ extracted from a systematic analysis of semileptonic hyperon decays \cite{Bourquin}. The $NN\eta'$ coupling constant is related to the $NN\eta$ coupling constant by $g_{NN\eta'} = - g_{NN\eta} \tan(\alpha_P)$, where $\alpha_P \equiv \theta_P - \theta_{P(ideal)} \cong -45^o$ denotes the deviation from the pseudoscalar ideal mixing angle. We find $g_{NN\eta'} \cong 6.1$. Once all the relevant coupling constants have been determined, we can estimate the relative importance of various $VVP$-exchange currents to the $\phi$-meson production. In corresponding test calculations it turned out that the $\phi\rho\pi$-exchange current is by far the dominant mesonic current. The {\it combined} contribution of all other exchange currents to the total cross section is about two orders of magnitude smaller. We have also examined contributions from meson-exchange currents involving other heavy mesons, in particular, the $\phi\phi f_1$- and $\phi\omega f_1$-exchange currents, whose coupling constants (upper limits) can be estimated from the observed decay of $f_1 \rightarrow \phi + \gamma$ \cite{PDG}. We find a negligible contribution to the $\phi$-meson production. Furthermore the $\phi\phi\sigma$- and $\phi\omega\sigma$-exchange currents also turned out to be negligible. Finally we note that, as in the case of $\omega$ production, there are no experimental indications of any of the known isospin-1/2 $N^*$ resonances decaying into $N\phi$. With the above considerations, the $v$-meson ($v=\omega, \phi$) production current $J^\mu$ is given by the sum of the nucleonic and $v\rho\pi$ meson-exchange currents, $J^\mu = J^\mu_{nuc} + J^\mu_{mec}$, as illustrated diagrammatically in Fig.~\ref{fig1}. Explicitly, the nucleonic current is defined as \begin{equation} J^\mu_{nuc} = \sum_{j=1,2}\left ( \Gamma^\mu_j iS_j U + U iS_j \Gamma^\mu_j \right ) \ , \label{nuc_cur} \end{equation} with $\Gamma^\mu_j$ denoting the $NNv$ vertex and $S_j$ the nucleon (Feynman) propagator for nucleon $j$. The summation runs over the two interacting nucleons, 1 and 2. $U$ stands for the meson-exchange $NN$ potential. It is, in principle, identical to the driving potential used in the construction of the $NN$ interaction \cite{MHE87}, except that here meson retardation effects are retained following the Feynman prescription. Eq.(\ref{nuc_cur}) is illustrated in Fig.~\ref{fig1}a. The structure of the $NNv$ vertex, $\Gamma^\mu_j$, required in Eq.(\ref{nuc_cur}) for the production is obtained from the Lagrangian density \begin{equation} {\cal L}(x) = - \bar\Psi(x) \left( g_{NNv} [\gamma_\mu -{\kappa_v\over 2m_N}\sigma_{\mu\nu}\partial^\nu ] V^\mu(x) \right) \Psi(x) \ , \label{NNv} \end{equation} where $\Psi(x)$ and $V^\mu(x)$ stand for the nucleon and vector-meson field, respectively. $g_{NNv}$ denotes the vector coupling constant and $\kappa_v \equiv f_{NNv}/g_{NNv}$, with $f_{NNv}$ the tensor coupling constant. $m_N$ denotes the nucleon mass. As in most meson-exchange models of interactions, each hadronic vertex is furnished with a form factor in order to account for, among other things, the composite nature of the hadrons involved. In this spirit the $NNv$ vertex obtained from the above Lagrangian is multiplied by a form factor. The theoretical understanding of this form factor is beyond the scope of the present paper; we assume it to be of the form \begin{equation} F_{vNN}(l^2) = \frac{\Lambda_{Nv}^4} {\Lambda_{Nv}^4 + (l^2-m_N^2)^2} \ , \label{formfactorN} \end{equation} where $l^2$ denotes the four-momentum squared of either the incoming or outgoing off-shell nucleon. It is normalized to unity when the nucleon is on its mass shell, i.e., when $l^2 = m_N^2$. The $v\rho\pi$ vertex required for constructing the meson-exchange current, $J^\mu_{mec}$ (Fig.~\ref{fig1}b), is derived from the Lagrangian density \begin{equation} {\cal L}_{v\rho\pi}(x) = \frac{g_{v\rho\pi}} {\sqrt{m_v m_\rho}} \varepsilon_{\alpha\beta\nu\mu} \partial^\alpha \vec \rho^\beta(x) \cdot \partial^\nu \vec \pi(x) V^\mu(x) \ , \label{pirhoomega} \end{equation} where $\varepsilon_{\alpha\beta\nu\mu}$ denotes the Levi-Civita antisymmetric tensor with $\varepsilon_{0123}=-1$. The $v\rho\pi$ vertex obtained from the above Lagrangian is multiplied by a form factor which is taken to be of the form \begin{equation} F_{v\rho\pi}(q_\rho^2, q_\pi^2) = \left ( \frac{\Lambda_{Mv}^2 - x m_\rho^2} {\Lambda_{Mv}^2 - q_\rho^2} \right ) \left ( \frac{\Lambda_{Mv}^2 - m_\pi^2} {\Lambda_{Mv}^2 - q_\pi^2} \right ) \ . \label{formfactorM} \end{equation} It is normalized to unity at $q_\pi^2 = m_\pi^2$ and $q_\rho^2 = x m_\rho^2$. The parameter $x$ (= 0 or 1) is introduced in order to allow for different normalization points as explained later. The form factor given above differs from the one used in \cite{Nak1} in that it uses a common cutoff parameter $\Lambda_{Mv}$ for both $\pi$ and $\rho$-meson instead of separate cutoff masses. The meson-exchange current is then given by \begin{equation} J^\mu_{mec} = [\Gamma^\alpha_{NN\rho}(q_\rho)]_1 iD_{\alpha\beta}(q_\rho) \Gamma^{\beta\mu}_{v\rho\pi}(q_\rho, q_\pi, k_v) i\Delta(q_\pi) [\Gamma_{NN\pi}(q_\pi)]_2 + (1\leftrightarrow 2) \ , \label{mec_cur} \end{equation} where $D_{\alpha\beta}(q_\rho)$ and $\Delta(q_\pi)$ stand for the $\rho$- and $\pi$-meson (Feynman) propagators, respectively. The vertices $\Gamma$ involved are self-explanatory. Both the $NN\rho$ and $NN\pi$ vertices, $\Gamma^\alpha_{NN\rho}$ and $\Gamma_{NN\pi}$, are taken consistently with the $NN$ potential used to generate the $NN$ final state interaction. Our model for vector-meson production described above contains five parameters: two for the mesonic current (the coupling constant $g_{v\rho\pi}$ and the cutoff parameter $\Lambda_{Mv}$) and three for the nucleonic current (the coupling constants $g_{NNv}$ and $f_{NNv}$, and the cutoff parameter $\Lambda_{Nv}$). In Ref.~\cite{Nak1} we pointed out that the angular distribution of the emitted $\omega$-mesons is a sensitive quantity for determining the {\it absolute amount} of nucleonic as well as mesonic current contributions, in addition to the relative sign between the two contributions. This applies also to the case of $\phi$-meson production since, as we have argued above, the dominant production mechanisms are exactly the same for both processes. Specifically, this means that the knowledge of the angular distribution allows one to fix the cutoff parameter in the mesonic current since the coupling constant $g_{v\rho\pi}$ can be extracted from the relevant measured partial decay widths \cite{PDG}. The nucleonic current, however, involves three free parameters. In this case knowledge of the angular distribution allows one to determine only the product of the $NNv$ coupling constants and the form factor. Consequently one cannot extract a unique value for $g_{NNv}$ directly from the analysis; further constraints---such as the energy dependence of the angular distribution---are needed. As discussed in \cite{Nak1}, the energy dependence will impose some constraint on the form factor. Also, in the energy region far from the threshold, the tensor-to-vector coupling ratio $\kappa_v = f_{NNv}/g_{NNv}$ influences the energy dependence of the total cross section. Another constraint may be imposed by spin polarization observables. In the case of $pp$ bremsstrahlung producing hard photons, it is known that the reaction is dominated by the magnetization current, to which the tensor coupling of the $NN\gamma$ vertex contributes. Therefore we would expect that the spin observables in $pp \rightarrow ppv$ reactions become more sensitive to the tensor coupling $f_{NNv}$ for energetic vector mesons. \section{Application } As mentioned in the introduction, the determination of the $NN\phi$ coupling strength is of special interest in the study of vector-meson production since its magnitude is usually associated with the amount of hidden strangeness in the nucleon. In this section we utilize our model to obtain some information on this coupling strength. However, as pointed out in the previous section, the angular distribution alone is not sufficient for determining $g_{NN\phi}$ uniquely. We therefore choose to perform a combined analysis of $\phi$- and $\omega$-meson production; i.e., to use the available data from both reactions to extract $g_{NN\phi}$. A combined analysis of $\phi$ and $\omega$ production also allows us to address the important issue concerning the violation of the OZI rule. Before doing so some considerations about the existing experimental data are in order. So far only very few precision data of vector-meson production in $NN$ collisions are available. First there are total cross sections for the reaction $pp\rightarrow pp\omega$ in the energy range $T_{lab} \cong 1.89$ to 1.98 GeV from Saclay \cite{Saclay}. In addition, there are angular distributions of the emitted meson for both $\omega$- and $\phi$-meson production (although with no absolute normalization) and the ratio of the total cross sections $\sigma_\phi / \sigma_\omega \equiv \sigma(pp \rightarrow pp\phi) / \sigma(pp\rightarrow pp\omega)$ at $T_{lab} = 2.85$ GeV from the DISTO collaboration \cite{Disto}. All the other presently available data \cite{Flam} are from the 1970's and not very accurate. Furthermore these data lie in an energy range far above the vector-meson production thresholds. At the corresponding excess energies the $NN$ interaction in the final state is already dominated by inelastic processes. Such processes are not accounted for in the $NN$ model that we employ, which is valid only for energies below the pion-production threshold (i.e., excess energies below $\approx 140$ MeV), and therefore we do not consider those high-energy data in our analysis. Since the excess energy in the $\omega$-meson production channel of the DISTO measurement ($Q = 319$ MeV) is already beyond the energy range where we trust our model, we do not use the measured production ratio directly. Instead we interpolate the total cross section for $\omega$-meson production from the existing data \cite{Saclay,Flam} and use this value, and the measured ratio, to fix the absolute normalization of the measured $\phi$-meson angular distribution \cite{Disto}. Our estimate yields $\sigma_\omega \sim 0.7 \times 10^2 ~\mu$b at $T_{lab}=2.85$ GeV. This value of $\sigma_\omega$, combined with the measured ratio of $\sigma_\phi / \sigma_\omega = (3.7 \pm 0.5) \times 10^{-3}$ \cite{Disto}, leads to $\sigma_\phi \sim 0.26~ \mu$b. Inevitably, such an interpolation is subject to uncertainties. We have, therefore, carried out the same analysis as reported below, but starting from a $\phi$-production cross section that is smaller/larger by about 40\%. We arrived at basically the same conclusions and therefore refrain from showing the corresponding results here. Since we have only one $\phi$-meson angular distribution and five $\omega$-meson total cross section data available in the range of applicability of our model, we are forced to impose some constraints on the model in order to reduce the number of free parameters. To this end we make the following assumptions: \begin{itemize} \item[1)] The same cutoff parameter $\Lambda_M\equiv \Lambda_{M\omega} = \Lambda_{M\phi}$ (cf. Eq.(\ref{formfactorM})) is used for the meson-exchange currents for $\omega$- and $\phi$-meson production. This is a reasonable choice since the off-shell particles at the $v\rho\pi$ vertex are the same in both production processes. Likewise, the cutoff parameter $\Lambda_{Nv}$ (cf. Eq.(\ref{formfactorN})) in the nucleonic current is assumed to be the same for $\omega$- and $\phi$-meson production, i.e., $\Lambda_N\equiv \Lambda_{N\omega}=\Lambda_{N\phi}$. \item[2)] The $NN\omega$ vector coupling constant is given by the value obtained from SU(3) flavor symmetry and imposing the OZI rule, i.e., $g_{NN\omega} = 3 g_{NN\rho} \cos(\alpha_V)$, where $\alpha_V \equiv \theta_V - \theta_{V(ideal)}$ is the deviation from the ideal $\omega - \phi$ mixing angle. With $\alpha_V \cong 3.7^o$ \cite{PDG} and the $NN\rho$ coupling constant of $g_{NN\rho} = 2.3 - 3.36$ \cite{rhocoup}, this yields a value of $g_{NN\omega} \cong (9 \pm 2)$, which is close to the value of $g_{NN\omega}\cong 11$ obtained in a recent $NN$ scattering analysis \cite{Janssen}. The uncertainty here comes from the uncertainties involved in $g_{NN\rho}$ and $\alpha_V$. \item[3)] The tensor-to-vector coupling ratio, $\kappa_v = f_{NNv}/g_{NNv}$, is the same for the $\omega$- and $\phi$-mesons: $\kappa \equiv\kappa_\omega = \kappa_\phi$. This relation is also suggested by SU(3) symmetry. Furthermore we choose the parameter $\kappa$ to be fixed beforehand. We consider the values $\kappa = \pm 0.5$, which covers a rather ample range. Unlike the case of the $\rho$-meson, we do not expect $f_{NNv} = \kappa_v g_{NNv}$ to be large, especially for the $\omega$-meson. This is supported by an estimate \cite{Dover}, in which SU(3) is applied to the sum $f_{NNv} + g_{NNv}$ rather than to $f_{NNv}$ alone, which is motivated by the success of SU(6) in predicting the magnetic moments of the baryons. Other investigations \cite{Meissner,MHE87} also support small values of $\kappa_v$. \end{itemize} These assumptions reduce the number of parameters of the model to be fixed to a total of five: the cutoff parameters $\Lambda_M$ and $\Lambda_N$, and the coupling constants $g_{\phi\rho\pi}$, $g_{\omega\rho\pi}$, and $g_{NN\phi}$. As mentioned in the previous section, the first two coupling constants can be extracted from the measured branching ratios. Specifically, the coupling constant of $g_{\phi\rho\pi}=-1.64$ is determined directly from the measured decay width of $\phi \rightarrow \rho + \pi$ \cite{PDG}. The coupling constant $g_{\omega\rho\pi}$, however, cannot be determined directly (as in the case of $g_{\phi\rho\pi}$) since $\omega \rightarrow \rho + \pi$ is energetically forbidden. We therefore extract it indirectly from the radiative decay width of $\omega \rightarrow \pi + \gamma$, assuming vector meson dominance; we obtain $g_{\omega\rho\pi}=10$. The signs of these couplings are inferred from SU(3) symmetry considerations. We note that these coupling constants are extracted at different kinematics: $g_{\phi\rho\pi}$ is determined at $q^2_\rho = m^2_\rho$ and $q^2_\pi = m^2_\pi$, whereas $g_{\omega\rho\pi}$ is extracted at $q^2_\rho = 0$ and $q^2_\pi = m^2_\pi$. The corresponding form factor (cf. Eq.(\ref{formfactorM})) should, therefore, be normalized accordingly, i.e., $x=1$ and $x=0$ for the $\phi\rho\pi$ and $\omega\rho\pi$ vertex form factor, respectively. With the coupling constants at the three-meson-point vertices fixed, we are then left with three free parameters which may be adjusted to reproduce the $\phi$- and $\omega$-meson production data. We are now prepared to apply the model to the reactions $pp\rightarrow pp\omega$ and $pp\rightarrow pp\phi$. Before doing any calculation, however, we note that the nucleonic current contribution to the $\phi$-meson production should be rather small, as the measured angular distribution shown in Fig.~\ref{fig2} is more or less isotropic. Recall that the angular distribution alone is sufficient to establish the magnitude of both the nucleonic and mesonic currents uniquely, and that the mesonic current yields a flat angular distribution, whereas the nucleonic current gives an approximately $ \cos^2(\theta)$ dependence~\cite{Nak1}. For the determination of the free parameters $\Lambda_M$, $\Lambda_N$ and $g_{NN\phi}$ we proceed in the following way: we assume that the angular distribution of the $\phi$-meson resembles the solid curve in Fig.~\ref{fig2}. We then determine the required contributions from both the nucleonic and mesonic currents. The mesonic current involves only one free parameter: namely the cutoff mass $\Lambda_M (\equiv\Lambda_{M\phi}=\Lambda_{M\omega})$ of the $\phi\rho\pi$ vertex form factor (cf. Eq.(\ref{formfactorM})). That is fixed by the requirement of reproducing the angular distribution. We obtain $\Lambda_M = 1450$ MeV. Turning now to $\omega$ production, we require that our model describe the total cross section data from SATURNE \cite{Saclay}. Since we assumed the $NN\omega$ vector coupling to be $g_{NN\omega} = 9$ (and $\kappa\equiv\kappa_\omega=\kappa_\phi$ is fixed to the two values mentioned above), one can determine the cutoff parameter $\Lambda_N (\equiv\Lambda_{N\omega}=\Lambda_{N\phi})$ in Eq.(\ref{formfactorN})---the only remaining free parameter---from the $\omega$-production cross section. Owing to the destructive interference between the nucleonic and mesonic currents, we find, in general, two possible values of $\Lambda_N$ for given $g_{NN\omega}$ and $\kappa$. The resulting values are listed in Table~\ref{tab2} and the corresponding cross sections are shown in Fig.~\ref{fig3}. As pointed out in \cite{Nak1}, in contrast to the angular distribution, the total cross section is unable to constrain uniquely the absolute contribution of the nucleonic and mesonic currents. As can be seen from Fig.~\ref{fig3}, the energy dependence of the total cross section may be useful to further constrain the ratio $\kappa$ and/or the form factor at the production vertex. Unfortunately, no data is available at energies that are well above the threshold, yet still low enough for the present model to be applicable. It should be mentioned that effects of the finite width of the $\omega$-meson---which influence considerably the energy dependence of the total cross section close to the threshold energy \cite{Wilkin}---haven been taken into account in the results shown in Fig.~\ref{fig3}. This is done by folding the calculated cross section with the Breit-Wigner mass distribution of the $\omega$-meson. Once the cutoff parameter $\Lambda_N$ is fixed, one can return to the angular distribution of the $\phi$-meson production in Fig.~\ref{fig2} and adjust the $NN\phi$ coupling constant $g_{NN\phi}$ to reproduce the amount of the nucleonic current contribution previously determined. Since we have two values of $\Lambda_N$ for each chosen value of $\kappa$, we get four values of $g_{NN\phi}$ corresponding to the four possible combinations. The values for the $NN\phi$ coupling constant thus extracted are compiled at the bottom of Table~\ref{tab2}. They range from $g_{NN\phi}=-0.19$ to $g_{NN\phi}=-0.90$. As already acknowledged above, the lack of a more complete and accurate set of data prevents us from achieving a more accurate determination of $g_{NN\phi}$. The values of $g_{NN\phi}$ thus obtained may be compared with those resulting from SU(3) flavor symmetry considerations and imposition of the OZI rule, \begin{equation} g_{NN\phi} = - 3 g_{NN\rho} \sin(\alpha_V) \cong -(0.60 \pm 0.15) \ , \label{OZI} \end{equation} where the factor $\sin(\alpha_V)$ is due to the deviation from the ideal $\omega - \phi$ mixing. The numerical value is obtained using the values $g_{NN\rho} = 2.63 - 3.36$ \cite{rhocoup} and $\alpha_V \cong 3.7^o$ \cite{PDG}. The error bar quoted is due to the uncertainty involved in $g_{NN\rho}$ and $\alpha_V$. Comparing the value (Eq.(\ref{OZI})) with those obtained in the present work we conclude that the $\phi$-production data can be described with $NN\phi$ coupling constants that are compatible with the OZI value. It should be mentioned that although our analysis of the existing DISTO and SATURNE data yields $NN\phi$ coupling constants which are compatible with the OZI value, it is necessary to introduce a violation of the OZI rule at the $v\rho\pi$ vertices ($v=\phi , \omega$) in the meson-exchange current. This can easily be verified by calculating, e.g., the $\phi\rho\pi$ coupling constants at $q_\rho^2=0$ using the form factor given by Eq.(\ref{formfactorM}) with the cutoff parameter $\Lambda_M$ extracted from the $\phi$-meson angular distribution data. We obtain $g_{\phi\rho\pi}(q_\rho^2=0) = g_{\phi\rho\pi}(q_\rho^2=m_\rho^2)\times F_{v\rho\pi}(q_\rho^2=0,q_\pi^2)=-1.18$. This value is almost a factor of 2 larger than the corresponding OZI value of $g_{\phi\rho\pi}(q_\rho^2=0) = - g_{\omega\rho\pi}(q_\rho^2=0) \tan(\alpha_V) \cong -0.65$ given in Table~\ref{tab1}. We found it impossible to describe the data without introducing this OZI violation (independently of the considerations about the $NN\phi$ coupling constant) for the following reasons: A sufficiently large contribution from the mesonic current in the $\phi$-meson production, as demanded by the measured angular distribution, can only be obtained with the $\phi\rho\pi$ coupling constant extracted from the measured decay width of $\phi \rightarrow \rho + \pi$ \cite{PDG}, i.e., $g_{\phi\rho\pi} = -1.64$. For the coupling constant extracted from radiative meson decay, an unrealistically large cut-off mass $\Lambda_M$ in excess of $3 GeV$ would be required. On the other hand, with $g_{\phi\rho\pi} = -1.64$ and the corresponding OZI value of $g_{\omega\rho\pi} = -g_{\phi\rho\pi}/ \tan(\alpha_V) = 25$ for the $\omega\rho\pi$ coupling constant, we have little chance of describing the energy dependence of the $\omega$-meson total cross section; it simply rises much too strongly with the energy. In this context let us mention that a knowledge of the angular distribution of the $\omega$-meson in the energy region where the model is applicable is of particular importance. Corresponding data would enable us to pin down the contributions for the mesonic current and thereby impose much more stringent constraints on the parameters of the model relevant to a possible violation of the OZI rule at the $v\rho\pi$ vertices. It would be very useful to clarify this point, specifically because an analysis of the radiative decay widths of the vector mesons within the vector-meson dominance model, involving the very same coupling constants, led to $v\rho\pi$ coupling constants which satisfy the OZI rule to within $15-20\%$ \cite{Durso}. Furthermore it is certainly desirable to test the present model in describing other independent reactions processes (e.g., $\phi$- and $\omega$-meson photoproduction). We plan such investigations in the near future. Finally, we wish to make a remark on the ``naive'' OZI estimation as expressed by Eq.(\ref{Ratio}). This equation is simply a consequence of assuming that the $\phi$- and $\omega$-meson production amplitudes differ from each other only by their coupling strengths, which are assumed to be related by SU(3) flavor symmetry plus the OZI rule. Differences in the kinematics (besides trivial phase-space effects) induced, for example, by the mass difference between the $\omega$- and $\phi$ mesons, interference effects between the nucleonic and meson-exchange currents, etc., are completely ignored. Our model offers the possibility to test explicitly the validity of this assumption and thus the reliability of Eq.~(\ref{Ratio}). For this purpose we carried out a (full) model calculation where we imposed the OZI rule to relate the relevant coupling constants; i.e., $g_{\phi\rho\pi} / g_{\omega\rho\pi} = - \tan(\alpha_V)$ and $g_{NN\phi} / g_{NN\omega} = - \tan(\alpha_V)$. In addition, we used the same form factors at the $\omega$- and $\phi$-meson production vertices in the mesonic and the nucleonic current, respectively. It turned out that such a calculation yields results that are qualitatively very similar to the simple estimate of Eq.(\ref{Ratio}), indicating that the above-mentioned differences in the kinematics, etc., do not induce any significant deviation from Eq.(\ref{Ratio}). In fact, the cross section ratio taken at the same excess energy is roughly $3 \times 10^{-3}$, or about 30\% smaller than the ''naive`` OZI estimate. \section{Summary} We have investigated the reaction $pp \rightarrow pp\phi$ using a relativistic meson-exchange model. We find that the nucleonic and $\phi\rho\pi$ exchange currents are the two dominant sources contributing to $\phi$-production in this reaction, and that they interfere destructively. Since these two reaction mechanisms give rise to distinct angular distributions, measurements of this observable can yield valuable information on the magnitude of the nucleonic current and, specifically, on the $NN\phi$ coupling constant. In fact, the flat angular distribution exhibited by the data from DISTO collaboration \cite{Disto} indicates that the contribution from the nucleonic current must be very small. This, in turn, implies that the value of the $NN\phi$ coupling constant must also be small. Indeed, our semi-quantitative combined analysis of those $\phi$-meson production data and the data for $\omega$-meson production from SATURNE yields values of $g_{NN\phi}$ which are compatible with the OZI rule. {\bf Acknowledgments} J.D. acknowledges the hospitality of the Institute f\"ur Kernphysik, Forschungzentrum J\"ulich. One of the authors (C.H.) is grateful for financial support by the COSY FFE--Project No. 41324880. \vfill \eject
\section{Introduction} Many Diophantine problems can be reduced to determining points on elliptic curves. A fascinating example is from Bremner et al \cite{bgn}, where finding possible representations of $n$ as \begin{displaymath} n = ( x + y + z) \left ( \frac{1}{x} + \frac{1}{y} + \frac{1}{z} \right ) \end{displaymath} is shown to be equivalent to finding points of infinite order on the elliptic curve \begin{displaymath} E_n: u^2 = v^3 + (n^2-6n-3) v^2 + 16n v \end{displaymath} The rank of $E_n$ can be estimated from the L-series of the curve, assuming the Birch \& Swinnerton-Dyer conjecture, but this just provides evidence of existence. For an actual representation, we need to explicitly find a rational point. If the curve has rank greater than 1, it is usually fairly simple to find a point by a trivial search. For rank 1 curves, however, this is often not feasible. The L-series computations can be extended to give an estimate for the height of a rational point of the curve. If this is large, then a simple search will take far too long. Recently, Silverman \cite{silv} produced a very nice and effective procedure to determine such points. Based on a preprint of this paper, the author coded the method in UBASIC, and used it on several families of elliptic curves from Diophantine problems. This usage has shown that the method starts to become time-consuming for heights greater than about 10 in the Silverman normalisation - 20 in the alternative (from now on all heights will be in the Silverman normalisation). If the curve has a rational point of order 2, we can use a 4-descent procedure to calculate the point, which we have found to be effective up to heights of about 15. By effective we mean that a point can usually be found in a matter of minutes on a 200MHz PC with a program written in UBASIC. If one is willing to wait several hours, the height barrier can obviously be increased. It should be noted that Silverman's method works just as well on curves with no rational points of order 2, and so is more general. \section{Four-descent} The following set of formulae describe the algebra necessary to perform a general four-descent procedure for the elliptic curve \begin{equation} y^2 = x^3 + a x^2 + b x \end{equation} so the curve is assumed to have at least one point of order 2. This discussion is purely concerned with computing a rational point, and not with determining the rank, and so is not as general as Cremona's {\bf mwrank} program, described in \cite{crem}. We can assume without loss of generality that $x = du^2/v^2$ and $y=duw/v^3$, with $d$ squarefree and $(d,v)=1$, giving \begin{equation} d^2 w^2 = d^3 u^4 + d^2 a u^2 v^2 + d b v^4 \end{equation} which implies that $d|b$, so that $b=de$. There are thus a finite set of possible values for d, which we can work out easily from the factors of $b$, unless $b$ is very large and difficult to factor. Remember that $d$ can be negative. Consider first the simpler quadratic \begin{equation} h^2 = d f^2 + a f g + e g^2 \end{equation} and look for a solution $(h_0,f_0,g_0)$ either by searching or by more advanced methods, and assume $g_0 \ne 0$. If we find a solution, we can parameterise as follows. Define $x=f/g$ and $y=h/g$, so that the equation is \begin{displaymath} y^2=d x^2+a x+e \end{displaymath} then the line through $(x_0,y_0)=(f_0/g_0,h_0/g_0)$, with gradient $m$, meets the curve again where \begin{equation} x=\frac{a + d x_0 + m (m x_0 - 2 y_0)}{m^2-d} \end{equation} Assuming $m=p/q$ is rational, we simplify this to \begin{equation} \frac{a g_0 q^2 + d f_0 q^2 + p (f_0 p - 2 h_0 q)}{g_0 (p^2 - d q^2)} \end{equation} To solve our original problem we look for values of the parameters giving $f/g=u^2/v^2$, which leads to the two quadratics \begin{equation} k u^2 = g_0 p^2 - g_0 d q^2 \end{equation} and \begin{equation} k v^2 = f_0 p^2 - 2 h_0 p q + (a g_0 + d f_0)q^2 \end{equation} with $k$ squarefree. The possible values of $k$ are those which divide the resultant of the two quadratics, which means that $k$ must divide the determinant \begin{center} \begin{math} \begin{array}{|cccc|} g_0&0&-d g_0&0 \\ 0&g_0&0&-d g_0 \\ f_0&-2 h_0&a g_0+d f_0&0 \\ 0&f_0&-2 h_0&a g_0 + d f_0 \end{array} \end{math} \end{center} which reduces to $g_0^4(a^2-4b)$. Let $k_0$ be a squarefree divisor (again possibly negative). We search the first of these quadratics to find a solution $(u_0,p_0,q_0)$. If we find one, another simple line-quadratic intersection analysis characterises the solutions to the first quadratic as \begin{equation} p = 2 u_0 k_0 r s - p_0 (g_0 s^2 + k_0 r^2) \end{equation} and \begin{equation} q = q_0 ( g_0 s^2 - k_0 r^2) \end{equation} Substitute these into $k_0 * (7)$, giving the quartic \begin{equation} z^2 = z_1 r^4 + z_2 r^3 s + z_3 r^2 s^2 + z_4 r s^3 + z_5 s^4 \end{equation} with \begin{equation} z_1 = k_0^3 (a g_0 q_0^2 + d f_0 q_0^2 + f_0 p_0^2 - 2 h_0 p_0 q_0) \end{equation} \begin{equation} z_2 = 4 u_0 k_0^3 (h_0 q_0 - f_0 p_0) \end{equation} \begin{equation} z_3 = 2 k_0^2 (f_0 (2 u_0^2 k_0 - d g_0 q_0^2 + g_0 p_0^2) - a g_0^2 q_0^2) \end{equation} \begin{equation} z_4 = -4 u_0 g_0 k_0^2 (f_0 p_0 + h_0 q_0) \end{equation} \begin{equation} z_5 = g_0^2 k_0 (a g_0 q_0^2 + d f_0 q_0^2 + f_0 p_0^2 + 2 h_0 p_0 q_0) \end{equation} This quartic can then be searched for possible solutions. Having found one, the various transformations eventually lead back to a point on the curve. \section{COMPUTING} The above formulae form the basis of a simple code written by the author in UBASIC. This system is fast, simple, free, and runs on a multitude of PC's, old and new. The major advantage, however, is that large numbers can be dealt with by one single statement at the start of the program - the rest of the code is standard Basic. Constructing a UBASIC code also leads to a structure which can easily be translated into Fortran, C, C++, etc using the many multiple-precision packages available in these languages. The basic structure of the algorithm is \begin{flushleft} find divisors d of b for s1 = s1a to s1b \hspace{2em}test if $h^2=df^2+afg+eg^2$ soluble with $|f|+|g|=s1$ \hspace{2em}if $(d,h_0,f_0,g_0)$ a solution then \hspace{2em}find divisors k of $g_0(a^2-4b)$ \hspace{2em}for s2 =2 to s2b \hspace{4em}test if $k u^2 = g_0 p^2 - g_0 d q^2$ with $|p|+|q|=s2$ \hspace{4em}if $(k_0,u_0,p_0,q_0)$ a solution, then \hspace{4em}form quartic equation from equations (10) to (15) \hspace{4em}test if quartic is soluble, and if it is \hspace{4em}for s3=2 to s3b \hspace{6em}test if quartic is square for $|r|+|s|=s3$ \hspace{6em}if it is, determine point on curve, and stop \hspace{4em}next s3 \hspace{2em}next s2 next s1 \end{flushleft} If the curve has a rational point of moderate height, the above search-based procedure works very well. We have found the method to be effective for points with height of up to about 15, if we select s2b=99 and s3b=199. It is impossible to predict in advance the best choice of s1a and s1b. For larger heights, the quartic search needs a larger value of s3b and can take a considerable time. In many cases, the quartic is insoluble so searching is futile. Cremona describes how to test this, based on ideas from Birch and Swinnerton-Dyer \cite{bsd}, and we have implemented this test as a precursor to searching. This method is probably the most advanced mathematically in the whole procedure. It is also a vital time-saving procedure since the majority of quartics are not soluble. Cremona also describes various methods for searching the quartic which reduce the time taken. We have chosen NOT to implement these, partly to keep the code reasonably understandable by amateurs, and partly because UBASIC has restrictions on the number of variables. These methods are only effective time-savers if s3b is large. For investigating families of curves where $a$ and $b$ are functions of $n$, we found the simple search for $(h_0,f_0,g_0)$ could fail to find any solutions in the specified range. In such a case we can adapt the search as follows. If \begin{displaymath} h^2 = d f^2 +a f g + b g^2 / d \end{displaymath} then \begin{displaymath} d(2h)^2 = ( 2df+ ag)^2 - (a^2-4b)g^2 \end{displaymath} and if we factor $a^2-4b = \alpha \beta^2$ with $\alpha$ squarefree, \begin{equation} dH^2 = F^2 - \alpha G^2 \end{equation} with $F = 2df+ag , \; G=\beta g, \; H = 2h$. From a $(F_0,G_0,H_0)$ solution, we can recover integer solutions to the original equation as $f_0=\beta F_0 - a G_0 , \; g_0 = 2 d G_0, \; h_0 = d \beta H_0$. In equation (6), it is clear that $(\pm u_0,\pm p_0, \pm q_0)$ satisfy the equation. We do not, however, need to consider all 8 possible combinations of sign. From (11) to (15), the sign of $u_0$ is irrelevant if we allow negative values of r or s. Similarly, $-p_0$ and $-q_0$ lead to the same quartic as $p_0$ and $q_0$, and the other two possible combinations give the same quartic. Thus there are only two essentially different quartics, and it is easy to show the relationship between them implies that they are both soluble or both insoluble. It is worthwhile to search both (if soluble), as the coefficients are different so the fixed search range might give a solution for one but not the other. At several points in our code, we need to factor numbers to find divisors. This is done by an extremely simple-minded search procedure, without recourse to any modern factorisation techniques. So far, this has not had a severe effect on the performance of the code, but one could easily devise an elliptic curve where the factorisation time was dominant. The current code has the advantage that is is short. Finally, it should be noted that the method could easily find a torsion point of the curve, and not a point of infinite order. Since the latter is usually wanted, the code checks that the point found is not a torsion point from a list of x-values provided by the user. \section{Further Descent} As stated, the above method begins to become time-consuming at certain heights. In such cases, several workers have resorted to impressive further descents to determine rational points. A classic example is the work of Bremner and Cassels on the curve $y^2=x^3+px$, see \cite{bc}. These methods seem to be very problem-dependent and to involve manipulations in algebraic number fields. The author was interested in a general method of higher descent which does not involve anything more than rational arithmetic - especially for use by the many amateurs interested in Diophantine equations. The following method is based on a remarkably simple idea. The author cannot believe this has not been thought of before, but can find no direct reference to the idea. The problem in the 4-descent for large heights is the determination of $(r,s)$ values giving a square quartic. Suppose \begin{equation} z_1r^4+z_2r^3s+z_3r^2s^2+z^4rs^3+z_5s^4=(u_1r^2+u_2rs+u_3s^2)(v_1r^2+v_2rs +v_3s^2) \end{equation} with $u_1,u_2,u_3,v_1,v_2,v_3$ integers, then we can consider the following two quadratics \begin{equation} u_1r^2+u_2rs+u_3s^2=k_1m^2 \end{equation} \begin{equation} v_1r^2+v_2rs+v_3s^2=k_1t^2 \end{equation} with $k_1$ squarefree. As before $k_1$ divides the resultant, which is \begin{equation} u_1^2v_3^2-u_1u_2v_2v_3-2u_1u_3v_1v_3+u_1u_3v_2^2+ u_2^2v_1v_3-u_2u_3v_1v_2+u_3^2v_1^2 \end{equation} Suppose we find a solution $(k_1,t_1,r_1,s_1)$ to the second quadratic, then we can parameterise as \begin{equation} r=i^2k_1r_1-2ijk_1t_1+j^2(r_1v_1+s_1v_2) \end{equation} \begin{equation} s=s_1(i^2k_1-j^2v_1) \end{equation} which, when substituted into $k_1*(18)$, requires the following quartic to be square \begin{equation} c_1 i^4 + c_2 i^3 j + c_3 i^2 j^2 + c_4 i j^3 + c_5 j^4 \end{equation} with \begin{equation} c_1=k_1^3(r_1^2 u_1+r_1 s_1 u_2 + s_1^2 u_3) \end{equation} \begin{equation} c_2=-2 k_1^3 t_1 ( 2 r_1 u_1 + s_1 u_2) \end{equation} \begin{equation} c_3=k_1^2 (4 k_1 t_1^2 u_1 + 2 r_1^2 u_1 v_1 + 2 r_1 s_1 u_1 v_2 + s_1^2 (u_2 v_2 - 2 u_3 v_1)) \end{equation} \begin{equation} c_4=- 2 k_1^2 t_1 (2 r_1 u_1 v_1 + s_1 ( 2 u_1 v_2 - u_2 v_1 ) ) \end{equation} \begin{equation} c_5=k_1 (r_1^2 u_1 v_1^2 + r_1 s_1 v_1 ( 2 u_1 v_2 - u_2 v_1 ) + s_1^2 (u_1 v_2^2 - v_1 (u_2 v_2 - u_3 v_1))) \end{equation} This quartic can be tested for solubility and then searched. \section{Computing the 8-descent} The basic structure of the algorithm is obviously very similar to the 4-descent structure. \bigskip \begin{flushleft} find divisors d of b for s1 = s1a to s1b \hspace{2em}test if $h^2=df^2+afg+eg^2$ soluble with $|f|+|g|=s1$ \hspace{2em}if $(d,h_0,f_0,g_0)$ a solution then \hspace{2em}find divisors k of $g_0(a^2-4b)$ \hspace{2em}for s2 = 2 to s2b \hspace{4em}test if $k u^2 = g_0 p^2 - g_0 d q^2$ with $|p|+|q|=s2$ \hspace{4em}if $(k_0,u_0,p_0,q_0)$ a solution, then \hspace{4em}form quartic equation from equations (10) to (15) \hspace{4em}test if quartic soluble, and if so \hspace{4em}test if quartic can be factored into 2 integer quadratics \hspace{4em}if so, find possible values of $k_1$ \hspace{4em}for s3=2 to s3b \hspace{6em}test if $k_1 t^2 = v_1 r^2 + v_2 rs + v_3 s^2$ with $|r|+|s| \le s3$ \hspace{6em}if $(k_1,t_1,r_1,s_1)$ a solution, form quartic (20) \hspace{6em}test if quartic soluble, and if it is \hspace{6em}for s4=2 to s4b \hspace{8em}test if $c_1i^4+c_2i^3j+c_3i^2j^2+c_4ij^3+c_5j^4$ square \hspace{8em}if it is, use various transformations to find point and stop \hspace{6em}next s4 \hspace{4em}next s3 \hspace{2em}next s2 next s1 \end{flushleft} \medskip The factorisation of the quartic is accomplished by considering the equations: \begin{center} \begin{math} \begin{array}{ccl} z_1&=&u_1 v_1 \\ z_5&=&u_3 v_3 \\ z_2&=&u_1 v_2 + v_1 u_2 \\ z_4&=&u_3 v_2 + v_3 u_2 \\ z_3&=&u_1 v_3 + u_2 v_2 + u_3 v_1 \end{array} \end{math} \end{center} We thus have to factorise $z_1$ and $z_5$. For each possible splitting, we solve the third and fourth equations for $u_2$ and $v_2$. If the solutions are integers we test whether the 6 values satisfy the last equation. We found that quadratics which themselves split into 2 linear factors lead to torsion points, so we test whether the quadratics have rational roots, and reject those which do. \section{Numerical Examples} In this section we give some examples of the use of the 8-descent method, with some specimen timings on either a 200MHz or 300MHz PC. It is clear that timings are dependant on the search parameters, so we specify the values of (s2b,s3b,s4b). s1a is set to 2 and we search until a solution is found. \bigskip (a) Congruent numbers N are integers which can be the area of a rational right-angled triangle. Finding the sides of such a triangle is equivalent to finding a point of infinite order on the elliptic curve $y^2=x^3-N^2x$. One of the most famous of these numbers is $N=157$, since it forms the basis of an impressive diagram in Chapter 1 of Koblitz's book \cite{kob}, where the sides involve numbers with 20-30 digits. An L-series calculation gives the height of a point as $27.3$. Running the program with s2b=s3b=99 and s4b=199 finds the point with x-coord \begin{displaymath} \frac{-1661 \, 3623 \, 1668 \, 1852 \, 6754 \, 0804}{28 \, 2563 \, 0694 \, 2511 \, 4585 \, 8025} \end{displaymath} in 18.8 secs (200MHz). With s4b=99 it takes only 7.0 secs, while for s4b=499 it takes 31.5 secs. But, with s4b=599 the time goes down to 28.3 secs, and if s4b=699 the time is only 9.0 secs. The variation is due to the number of quartics which need to be searched. Up to 499 it takes 22 quartics, but 599 needs only 4, while 699 finds a point on the first. The important point is that it takes less than 1 minute to find such a large point. Even an ancient 80387 machine finds this point in less than 5 minutes. The author has used this technique in a search for actual solutions to the congruent number problem for $1 \le N \le 1999$. The current method has led to the completion of a table of solutions for $[1,499]$. The largest height encountered was $51.15$ for $N=367$. Searching with s2b=s3b=99 and s4b=9999, the following solution was found on a 300MHz machine in 17595 secs. \begin{displaymath} x = \frac{-367 (4 \, 9695 \, 3629 \, 6085 \, 1360 \, 8777)^2} {(163 \, 8216 \, 8821 \, 6485 \, 0643 \, 1464)^2} \end{displaymath} It is perfectly possible that the modular form Heegner-point method of Elkies \cite{elk} could find this point much faster, but this method is much more difficult for the non-expert to understand. Elkies method also does not generalise to other families of curves. Since $x^3-N^2x=x(x-N)(x+N)$, we can change the origin to produce the two equivalent elliptic curves $y^2=x^3 \pm 3Nx^2 + 2N^2x$. They may be equivalent mathematically, but their performance computationally is not the same. The author has found several solutions to the congruent number problem from these curves having been unsuccessful with the original curve. This also happens in other families of elliptic curves with 3 rational points of order 2. \bigskip (b) The paper of Bremner et al mentioned in the introduction has a representation of $N=564$ in its abstract. The L-series computation predicts a point with height $38.01$. A 300MHz PC found a solution in 4497 secs. with s2b=s3b=99 and s4b=9999. This solution leads to a representation with \medskip x = 32736 87951 95203 44322 22320 98479 60433 77911 47254 01060 y = 53 58267 18225 66098 96868 10234 90522 46809 90105 26717 z = - 1158 25525 22781 02629 66659 36639 59067 36616 11576 01937 \bigskip (c) The paper of Bremner and Cassels describes the finding of a point on $y^2=x^3+877x$. The L-series predicts a height of $48.0$. A 300MHz PC finds the following x-coordinate in 51874 seconds with s2b=s3b=99 and s4b=12999. \begin{displaymath} x = \frac{877 \, ( 7884 \, 1535 \, 8606 \, 8390 \, 0210 )^2} {(6 \, 1277 \, 6083 \, 1879 \, 4736 \, 8101 )^2} \end{displaymath} \bigskip (4) The final example is included for historical reasons, since it was the first time that the author tried the idea of factorising the quartic. Since this was the start of the current research, clearly the computer programs used in the previous examples were not available. The computations were done by simple searches and algebraic manipulation with Derive. The problem comes from the diophantine problem of finding an integer triangle with base/altitude = n. For $n=79$, we consider the equation $y^2=x^3+6243x^2+x$. The L-series computations suggested rank 1, but with a height of over 40 for a rational point. The 2-isogenous curve is $y^2=x^3-12486x^2+38975045x$, which was indicated to have a point with height about 20. The author selected to try $d=5$, which meant looking first for solutions of $h^2=5f^2-12486fg+7795009g^2$. A very simple search program quickly finds $f=93, g=1, h=2584$, which means (6) and (7), are \begin{displaymath} k u^2 = p^2 - 5 q^2 \end{displaymath} \begin{displaymath} k v^2 = 93 p^2 - 5168 pq -12021 q^2 \end{displaymath} where $k=\pm 1$. We selected $k=1$. It is possible to parameterise the first as $p=r^2+5s^2$,$q=2rs$, which gives \begin{displaymath} v^2=93r^4-10336r^3s-47154r^2s^2-51680rs^3+2325s^4 \end{displaymath} which Derive fairly easily factors as \begin{displaymath} v^2=(3r^2-340rs-775s^2)(31r^2+68rs-3s^2) \end{displaymath} The two quadratic factors form the basis of (18) and (19), and we find that $k_1|158$. Picking $k_1=1$, we found several $(r,s)$ solutions to $3r^2-340rs-775s^2=t^2$, which lead to parameterisations for r and s, but which lead to insoluble quartics. We then tried $k_1=158$, and found the solution $r=9,s=-1,t=4$, which gives the parameterisation \begin{displaymath} \frac{r}{s} = \frac{-(1422i^2+1264ij+367j^2)}{158i^2-3j^2} \end{displaymath} and hence to the quartic \begin{displaymath} 316^2(74892 i^4 +154840 i^3j+123789i^2j^2+45916ij^3+6725j^4) \end{displaymath} which has to be square. Thus quartic was everwhere soluble, so a search was started which quickly found the solution $i=151, j=-158$, which lead back to a point on the 2-isogenous curve with \begin{displaymath} x=\frac{2836 8499 3467 6319 5139 0020}{4689 8490 9449 9234 0041} \end{displaymath} and thus to a point on the original curve with \begin{displaymath} x=\frac{2654 7926 1289 1944 1996 8505 1867 1143 3025} {1705 4187 5947 7256 7676 9862 5643 5806 2336} \end{displaymath} For interested readers, this point leads to the triangle with sides \begin{math} \begin{array}{crr} \, & \, & \, \\ A =& 1465 86997 18477 82318 35321& 97194 40069 87886 57474 85658 \\ \, & \, & 64108 26213 28674 16311 64960 \\ \, & \, & \, \\ B =& 892 76765 34887 48588 76033 & 62942 70957 75073 78277 30811 \\ \, & \, & 86659 99941 08625 53894 71249 \\ \, & \, & \, \\ C =& 573 59536 91823 05619 55378 & 66267 79319 29261 59738 76797 \\ \, & \, & 12797 54707 31247 71171 08209 \end{array} \end{math} \section{Further Work} The reasonable level of success of this method for several families of elliptic curves provided the impetus to write this report, but there is still much work to be done. \begin{enumerate} \item Translate the UBASIC code into a compiled high-level language so that the method can be run on non PC's, especially UNIX workstations. \item Try further descents on the last quartic produced. The author has experimented with this idea, but initial results are disappointing in that the method seems to be finding multiples of a generator rather than the generator itself. Since the multiples have much larger height it is currently no benefit to try these extra descents. \item Try to underpin the method with some theory. The method was developed by trying an idea, seeing it work, and improving it. The method does not always work, but does provide another tool in the investigator's toolbox. \end{enumerate} \newpage
\section{Introduction} The interactions between globular proteins are rather poorly understood but it seems clear that many of the attractive interactions are highly directional \cite{durbin96,visser92,neal99}; two protein molecules must not only be close to each other to attract each other but they must also be correctly oriented. An example is the attraction between hydrophobic patches on the surfaces of globular proteins; only if the proteins are oriented so that these parts of their surfaces face each other is there an attraction. This attraction is both highly directional and short ranged \cite{visser92}. Hydrogen bonds, which are highly directional, are known to have a dramatic effect on the behaviour of simple molecules such as water. For example, ice's diamond-like lattice and very low density are imposed by the tetrahedral arrangement of the hydrogen bonds formed by water molecules. It is striking that much of the language used to describe the interactions between globular proteins \cite{durbin96,visser92,neal99} is the same as that used in describing hydrogen-bonding fluids \cite{blanca,nezbeda90,sj}. Therefore, we believe it is very worthwhile to explore the phase behaviour of a simple model of globular proteins which includes directional attractions. We find phase diagrams of qualitatively the same form as observed in experiment. In particular, we find metastable transitions between dilute and dense fluid phases. Solutions of globular proteins, like simple molecules, exhibit two types of phase transition: fluid-to-fluid, analogous to a vapour-liquid transition, and fluid-to-solid. The behaviour of globular proteins differs from that of simple molecules in one important respect: the transition between a dilute and a dense fluid phase is metastable with respect to solidification, \cite{broide91,muschol97} whereas, simple molecules such as argon and water have equilibrium liquid phases. The proteins crystallise in a variety of different lattice types. This is suggestive of directional attractions as at low temperature these impose the lattice. For example, a water molecule bonds to four molecules arranged tetrahedrally around it. In a diamond lattice but not a face-centred-cubic lattice each water molecule is surrounded by four others arranged tetrahedrally, and so water freezes into a solid with a diamond not a face-centred-cubic lattice. Isotropic attractions of any range have a relatively weak effect on the solid phase; the solid is face-centred cubic as it is for hard spheres. Globular proteins are not perfectly spherical, of course, and this may also have an effect on the lattice type. Previous work \cite{rosenbaum96,tenwolde97,poon97,sear99b} has ignored any directionality in the attraction and modeled the protein molecules with a potential consisting of a hard core and a short-range isotropic attraction. The phase behaviour of such a potential is qualitatively similar to experiment and to the behaviour we find for our model, with the exception of the solid phase. We will return to this point when we discuss our results in section 5 but the interaction between proteins almost certainly includes both isotropic and directional attractions \cite{neal99}. The work is presented as follows. In the next section we specify the simplest model of a protein we can devise and which includes directional attractions. It does not include isotropic attractions and all the directional attractions are of the same strength. This model is very similar but not identical to very simple models of water \cite{nezbeda90,ghonasgi93,duda98,vega98}. Given its simple nature and our lack of knowledge of the interaction between globular proteins our analysis will be purely qualitative. Nevertheless, we can describe the experimental phase behaviour and the phase diagrams we calculate have intrinsic interest as some of the first calculations which include the fluid and solid phases when the attractions are highly directional. For related work see Ref. \citen{vega98}. The theories for the fluid and solid phases are derived in sections 3 and 4, respectively. We restrict ourselves to a brief summary of the theory for the fluid phase as it is well established, it is based on the formalism derived by Wertheim \cite{wertheim84,wertheim86b,wertheim86,wertheim87,jackson88,chapman88}. The theory for the solid phase is new, it is a cell theory, like that of Vega and Monson \cite{vega98}. Example phase diagrams are shown and discussed in section 5. We finish with a discussion of the connection between the interactions and both the fluid-fluid transition and the stability or otherwise of the solid phase. \section{Model} Our model is not new, it is of a type which has been widely employed to model hydrogen-bonding molecules such as water \cite{nezbeda90} and alcohols \cite{blanca}. Indeed, the idea of a potential consisting of a repulsion plus a sticky patch goes back to Boltzmann \cite{boltzmann}. Specifically, we will consider the conical-site model of Chapman, Jackson and Gubbins \cite{jackson88,chapman88}. A schematic of the model is shown in Fig. 1. The potential is a pair potential $\phi$ which is a sum of two parts: a hard-sphere repulsion, $\phi_{hs}$, and a set of sites which mediate short-range, directional attractions. The sites come in pairs: a site on one particle binds only to its partner on another particle. So, the number of sites, $n_s$, is constrained to be an even number. The two sites of a pair are numbered consecutively so that an odd-numbered site, $i$, binds only to the even-numbered site, $i+1$. This is the only interaction between the sites, an odd-numbered site, $i$, does not interact at all with sites other than the $(i+1)$th site. Each site has a specific orientation, see Fig. 1. The orientation of site number $i$ is specified by means of a unit vector ${\bf u}_i$. We can write the interaction potential between a pair of particles as \begin{equation} \phi(r_{12},\Omega_1,\Omega_2)=\phi_{hs}(r_{12}) +\sum_i^{'} \left[ \phi_{ii+1}(r_{12},\Omega_1,\Omega_2), + \phi_{ii+1}(r_{12},\Omega_2,\Omega_1)\right], \label{pot} \end{equation} where the dash on the first sum denotes that it is restricted to odd values of $i$. The interactions between the sites on the two particles are $\phi_{ii+1}(r_{12},\Omega_1,\Omega_2)$, which is the interaction between site $i$ on particle 1 and site $i+1$ on particle 2, and $\phi_{ii+1}(r_{12},\Omega_2,\Omega_1)$, which is the interaction between site $i$ on particle 2 and site $i+1$ on particle 1. These are functions of $r_{12}$, $\Omega_1$ and $\Omega_2$, which are the scalar distance between the centres of particles 1 and 2, the orientation of particle 1 and the orientation of particle 2, respectively. The particle is rigid, but not axially symmetric, so its position is completely specified by the position of its centre and its orientation $\Omega$, which may be expressed in terms of the three Euler angles \cite{allen87}. The hard-sphere potential, $\phi_{hs}$, is given by \begin{equation} \phi_{hs}(r)= \left\{ \begin{array}{ll} \infty & ~~~~~~ r \le \sigma\\ 0 & ~~~~~~ r > \sigma\\ \end{array}\right. , \label{hspot} \end{equation} where $\sigma$ is the hard-sphere diameter. The conical-site interaction potential $\phi_{ii+1}$ is given by \cite{jackson88} \begin{equation} \phi_{ii+1}(r_{12},\Omega_1,\Omega_2)= \left\{ \begin{array}{ll} -\epsilon & ~~~~~~ r_{12}\le r_c ~~~~~ \mbox{and} ~~~~~ \theta_{1i}\le\theta_c ~~~~~ \mbox{and} ~~~~~ \theta_{2i+1}\le\theta_c\\ 0 & ~~~~~~ \mbox{otherwise}\\ \end{array}\right. , \label{sitepot} \end{equation} where $\theta_{1i}$ is the angle between a line joining the centres of the two particles and the unit vector ${\bf u}_i$ of particle 1, and $\theta_{2i+1}$ is the angle between a line joining the centres of the two particles and the unit vector ${\bf u}_{i+1}$ of particle 2. The conical-site potential depends on two parameters: the range, $r_c$, and the maximum angle at which a bond is formed, $\theta_c$. Of course, as the attractions are directional, $\theta_c$ will be small, no more than about $30^{\circ}$. The attractions are also short ranged, $r_c$ no more than 10\% larger than $\sigma$. The angles between the site orientations, the vectors ${\bf u}_i$ will determine which solid lattice is formed. This is like water, where the diamond-like lattice is enforced by the tetrahedral arrangements of the hydrogen bonds. For simplicity, we will take the sites to be arranged such that they are compatible with a simple cubic lattice. Then if we express the unit vectors ${\bf u}_i$ in Cartesian coordinates, $(x,y,z)$, then when we have four sites, $n_s=4$, the set of vectors ${\bf u}_1=(1,0,0)$, ${\bf u}_2=(-1,0,0)$, ${\bf u}_3=(0,1,0)$ and ${\bf u}_4=(0,-1,0)$ would describe our model. For six sites then we add two additional sites at orientations ${\bf u}_5=(0,0,1)$ and ${\bf u}_6=(0,0,-1)$. Note that this arrangement of sites is perfect for a cubic lattice, in general we would expect the angles between the attractions to be not quite perfectly aligned with the angles in the solid phase, These mismatches will tend to destabilise the solid phase so our model is an extreme case; a general arrangement of sites will tend to have a less stable solid phase. \section{Theory for the fluid phase} The theory for the fluid phase of particles interacting via a hard-core and directional attractions mediated by sites is well established. Wertheim developed highly accurate perturbation theories for models with one \cite{wertheim84,wertheim86b} and two sites \cite{wertheim86,wertheim87}. By perturbation theory we mean a theory which calculates the difference in free energy between a fluid of particles interacting via the full potential and that of a fluid of particles interacting just via the hard-sphere part of the potential. Chapman, Jackson and Gubbins, by assuming additivity of the effect of the sites on the free energy, generalised the perturbation theory to potentials with more than two sites \cite{jackson88,chapman88}. The theory has been extensively tested against the results of computer simulation \cite{jackson88,ghonasgi93,duda98,vega98}. Thus, the perturbation theory gives for the Helmholtz free energy per particle of the fluid phase, $a_f$, \cite{jackson88} \begin{equation} \beta a_f (\eta,T)=\beta a_{hs}(\eta) + n_s\left[\ln X+\frac{1}{2}(1-X)\right], \label{af} \end{equation} where $a_{hs}$ is the Helmholtz free energy per particle of a fluid of hard spheres. We use an accurate expression derived from the equation for the pressure of Carnahan and Starling \cite{carnahan69,hansen86} for $a_{hs}$. The volume fraction $\eta=\rho(\pi/6)\sigma^3$ is a reduced density; $\rho$ is the number density of particles. $\beta=1/kT$, where $k$ is Boltzmann's constant and $T$ is the temperature. $X$ is the fraction of sites which are {\em not} bonded to another site. As all site-site interactions are equivalent the fraction of each type of site which is not bonded is the same. The fraction of sites which are bonded and the fraction which are not bonded must, of course, add up to one. Thus we can simply write down a mass-action equation for $X$, \cite{jackson88} \begin{equation} 1=X+\rho X^2 Kg_{hs}^c(\eta)\exp(\beta\epsilon), \label{ma} \end{equation} where $g_{hs}^c$ is the contact value of pair distribution function of a fluid of hard spheres. It can be obtained from Carnahan and Starling's expression for the pressure of hard spheres using the virial equation \cite{hansen86}. The volume of phase space (both translational and orientational coordinates) over which a bond exists is $K$ \cite{jackson88}, \begin{equation} K=\pi\sigma^2(r_c-\sigma)(1-\cos\theta_c)^2. \label{kdef} \end{equation} See Refs. \citen{wertheim84,wertheim86b,wertheim86,wertheim87,jackson88,chapman88,sear96,sj} for a detailed discussion and derivation of Eqs. (\ref{af}) and (\ref{ma}). The rightmost term in Eq. (\ref{ma}) is the fraction of sites which are bonded. This is equal to the fraction of sites which are unbonded and so are available to form a bond, $X$, times the probability of such a site forming a bond. This probability is the density of other sites available for bonding, $\rho X$, times a Boltzmann factor $\exp(\beta\epsilon)$ and integrated over the phase volume for which a bond exists. The factor of $g_{hs}^c$ is required at high density to take account of the effect of surrounding particles pushing particles together. As for the free energy, Eq. (\ref{af}), it is easy to check that, for $n_s=2$, it gives the correct low density limit \cite{sear96}. The mass-action equation, Eq. (\ref{ma}), is a quadratic equation for $X$ and the physical solution is \begin{equation} X=\frac{2}{1+[1+4\rho Kg_{hs}^c\exp(\beta\epsilon)]^{1/2}}, \label{x} \end{equation} Note that Eqs. (\ref{ma}) and (\ref{x}) are not quite the same as the equivalent equations in Refs. \citen{wertheim84,wertheim86b,wertheim86,wertheim87,jackson88,chapman88,sear96,sj}. In those references $\exp(\beta\epsilon)$ is replaced by $\exp(\beta\epsilon)-1$. As $\beta\epsilon$ is quite large, five or more, the difference between the two is very small. Also, our $K$ is $4\pi$ times the $K_{AB}$ of Ref. \citen{jackson88}. The pressure of the fluid phase, $p_f$, can be found by differentiating the free energy, $a_f$: \begin{eqnarray} \beta p_f&=&\rho^2\left(\frac{\partial\beta a_f}{\partial\rho}\right)\\ &=&\beta p_{hs}+n_s\rho^2\left(\frac{\partial X}{\partial \rho}\right) \left(\frac{1}{X}-\frac{1}{2}\right) \end{eqnarray} where $p_{hs}$ is the pressure of a fluid of hard spheres, and the derivative of $X$ may be obtained by taking the derivative of Eq. (\ref{x}). The chemical potential of the fluid phase $\mu_f=a_f+p_f/\rho$. The state of our single component fluid is specified by the ratio of the site energy to the thermal energy, $\beta\epsilon$, and the density, $\rho$, or volume fraction, $\eta$. If there is phase coexistence at any temperature then the two coexisting densities can be obtained from the requirement that the chemical potential and pressure be equal in the two phases. This requirement yields a pair of coupled nonlinear equations which may be solved to give the two coexisting densities. \subsection{The second virial coefficient} In experiments on proteins, the strength of the attractions is often assessed by measuring the second virial coefficient, $B_2$, via a scattering experiment. The second virial coefficient is equal to the coefficient of the term linear in density in the Helmholtz free energy per particle \cite{hansen86}. From Eq. (\ref{af}) we see that it will be the sum of the second virial coefficient of hard spheres, $B_2^{hs}=(2\pi/3)\sigma^3$, and a contribution from the sites. This contribution of the sites may be obtained by first setting $g_{hs}^c$ to one in Eq. (\ref{x}); one is its value to lowest order in density. Then we expand out Eq. (\ref{x}) for $X$ as a density expansion and truncate after the linear term. This truncated density expansion for $X$ is then substituted in Eq. (\ref{af}), and the result expanded out in density. The linear term is the second virial coefficient. The result is \begin{equation} B_2= B_2^{hs} - \frac{n_s}{2}K\exp(\beta\epsilon). \label{b2} \end{equation} \section{Theory for the solid phase} At low temperature, solidification is driven by the attractive interactions, not packing effects as it is with hard spheres \cite{hoover68} or with particles with only one or two sites \cite{sear95}. In the solid phase the position of a particle's centre of mass and its orientation are constrained by the need for it to always be in such a position and orientation that none of the bonds it makes with the neighbouring particles are broken. Like hard spheres the particle is constrained by its neighbours but unlike hard spheres the constraint is not only due to the requirement that the particle not overlap with one of its neighbours but also due to the requirement that bonds not be broken, We will use a cell theory to describe the free energy of the solid phase of our model \cite{buehler51,sear98}. This means we will approximate the energy by the energy when all bonds are formed, the ground-state energy, and the entropy by the logarithm of the phase volume available to a particle when it is constrained to lie in the solid, forming these $n_s$ bonds with its neighbours. Vega and Monson \cite{vega98} have used a cell theory to describe the solid phase of a very similar model, a simple model of water. They avoid a couple of the approximations used here at the cost of not having an analytical free energy. Comparison with the results of computer simulation show that their cell theory is quite accurate. Within a cell theory for a solid phase, the Helmholtz free energy per particle, $a_s$, is given by \begin{equation} \beta a_s=-\ln q_P \label{adef} \end{equation} where $q_P$ is the partition function of a single particle trapped in a cage formed by the requirements that all its $n_s$ sites bond to neighbouring particles, and that its hard core not overlap with any of these neighbours. In order for these bonds to not be broken the particle must always be within $r_c$ of the surrounding particles. This fixes the lattice constant, $a$, at a little less than $r_c$. It is a little less as when the particle moves off the lattice site it will be moving towards some of its neighbours and away from others. Thus it can explore regions where it is further than $a$ from some of its neighbours. Of course, a particle cannot move within $\sigma$ of any of its neighbours due to the hard-sphere interaction. Thus the particle can move a distance $a-\sigma$ in the direction of any of its neighbours. The exact value of the maximum lattice constant for which the particle can move about, constrained by the hard-sphere interactions, without breaking any bond, is difficult to estimate; as is the volume available to the centre of mass of the particle \cite{buehler51}. Therefore, we approximate the lattice constant $a$ by $r_c$ and the volume to which the particle is restricted by $(r_c-\sigma)^3$. The requirement that no bonds be broken also severely restricts the orientations of the particle. When a non-axially symmetric particle is free to rotate it explores an angular phase space of $8\pi^2$. However, in the solid its rotations will be restricted to those which are small enough not to violate the requirement that the orientations of its site vectors are within $\theta_c$ of the lines joining the centre of the particle with the those of the neighbouring particles. Again the exact value of angular space available to the particle is complex, and it also depends on the position of the particle. We approximate this angular space by assuming that each of the three angular degrees of freedom can vary independently over a range of $2\theta_c$. The normalised angular space available to a particle in the solid phase is then $(2\theta_c)^3/8\pi^2=\theta_c^3/\pi^2$. The energy per particle is, of course, $-(n_s/2)\epsilon$, and so the partition function, $q_P$, is then just the volume available to the centre of mass of the particle times the angular space available times $\Lambda^{-1}\exp[(n_s/2)\beta\epsilon]$, where $\Lambda^{-1}$ is the integral over the momentum degrees of freedom. Thus, we have for $q_P$, \begin{equation} q_P=(r_c-\sigma)^3\left(\frac{\theta_c^3}{\pi^2}\right) \Lambda^{-1}\exp\left(\frac{n_s}{2}\beta\epsilon\right) \end{equation} Inserting this expression for $q_P$ into Eq. (\ref{adef}), \begin{equation} \beta a_s=-3\ln\left(\frac{r_c}{\sigma}-1\right) -\ln\left(\frac{\theta_c^3}{\pi^2}\right) -\frac{n_s}{2}\beta\epsilon =\beta \mu_s, \label{mus} \end{equation} where we have neglected a term $\ln(\Lambda/\sigma^3)$. This is the free energy at a lattice constant of $r_c$. The maximum possible density of a simple-cubic lattice is when the lattice constant $a=\sigma$, then the density is $\sigma^{-3}$. This density corresponds to a volume fraction $\eta=\pi/6$. When the lattice constant is $r_c$, the density is $r_c^{-3}$ and the volume fraction is $(\pi/6)(\sigma/r_c)^3$. On compression, the free energy of solid phase increases very rapidly and diverges when $a=\sigma$. We are interested in finding coexistence between the solid phase and the fluid phase at low temperature, when our assumption that no bonds are broken in the solid phase will be accurate. Then the pressure at coexistence will be low and the solid will be near the its minimum possible density, $r_c^{-3}$. The chemical potential $\mu_s=a_s+p_s/\rho$ where $p_s$ is the pressure and $\rho$ is the density. At low pressure $p_s/\rho$ contributes a negligible amount to the chemical potential, which enables us to equate $a_s$ and $\mu_s$ as we have done in Eq. (\ref{mus}). The coexisting fluid density at the fluid-solid transition is then found by equating the chemical potentials in the two phases. The density of the coexisting solid phase, when the temperature is low enough that solidification is driven by the attractive interactions not packing effects, is assumed constant at $r_c^{-3}$. Essentially, the requirement that the pressure be equal in the two phases disappears due to the fact that in the solid phase the pressure is very sensitive to density. This is best seen in the schematic diagram of the free energy shown in Fig. 2. The free energy has very narrow minimum at a density of $r_c^{-3}$; at higher densities it rapidly increases as the maximum density of the simple-cubic lattice is approached and at lower densities it is much higher as some of the bonds are broken. The pressure (= minus the slope of the free energy with respect to the volume) then varies very rapidly. See Fig. 11 of Ref. \citen{vega98} which shows the results of computer simulations for a similar solid phase; the pressure varies from almost zero to $20kT/\sigma^{-3}$ over a range in volume fraction of about 0.01. Now, the more rapid the variation the smaller the displacement in density that is required to vary the pressure of the solid phase to equal that in the fluid phase. By fixing the density of the solid phase to be $r_c^{-3}$ and then determining the density of coexisting fluid phase by just equating the chemical potentials in the two phases, we are neglecting this displacement in density. \section{Phase behaviour} The only phase transition of hard spheres is a fluid-solid transition \cite{hoover68}. Spheres which have only one site and so form only pairs of spheres, also only have a fluid-solid transition \cite{sear95}. When there are two sites, linear chains of particles form, again there is only a fluid-solid transition \cite{malanoski97}. All these transitions are driven by packing effects, hard spheres pack more efficiently in the solid than in the fluid phase. When there are more than two sites, there is a fluid--fluid transition, according to the free energy of the fluid phase, \cite{notecp} Eq. (\ref{af}). Also, when there are more than two sites a fluid-solid transition driven by the interactions between sites not the packing of the hard cores, is possible. Indeed at low temperatures the sites will drive the solidification, they will impose the lattice type. It will be the one which maximises the number of sites which can interact. This low temperature solid is well described by the theory of section 4. We will only plot results at low temperature when this theory is accurate. Nezbeda and Iglesias-Silva \cite{nezbeda90} were the first to observe a fluid--fluid transition in a model of the type we are considering here. Their model has four sites and was developed as a simple model of water. Figures 3 to 5 are example phase diagrams of four, Figs. 3 and 4, and six site, Fig. 5, model proteins. In each case, as $n_s>2$, the fluid phase free energy predicts a transition between a dilute and a dense fluid phase. The values of $\beta\epsilon$, $\eta$ and $B_2$ at the three critical points, $(\beta\epsilon)^{cp}$, $\eta^{cp}$ and $B_2^{cp}$, respectively, are tabulated in Table 1. However, only in Fig. 4 does this transition appear in the equilibrium phase diagram. In Figs. 3 and 5 the fluid--fluid transition is metastable. This means that at equilibrium the transition does not occur, it is preempted by the fluid-solid transition, but if the fluid can be cooled sufficiently far into the two-phase region that it crosses the metastable fluid--fluid coexistence curve then a fluid--fluid transition might be observed. Whether this can be done or not depends on the dynamics of crystallisation \cite{debenedetti,tenwolde97,sear99b}. A metastable fluid-fluid transition was also predicted by Vega and Monson \cite{vega98}. Experiments on globular proteins have found such metastable fluid--fluid transitions \cite{broide91,muschol97}; the crystallisation of proteins is often slow, taking several days, which allows the protein solution to be cooled into a region of the phase diagram where the fluid phase separates into two fluid phases of differing densities. Our toy model of a protein has three parameters: the number of sites, $n_s$, the range of the attraction between sites, $r_c$, and the angle over which two sites interact, $2\theta_c$. Comparing Figs. 3 and 4 we see that decreasing $\theta_c$ destabilises the dense fluid phase (the `liquid') with respect to the solid phase leaving only a fluid-solid transition in the phase diagram. Comparing Figs. 4 and 5 we see that increasing the number of sites has the same effect. Increasing the range of attraction has a weaker effect, although our theory predicts that increasing the range actually tends to drive a fluid-fluid transition metastable. This is opposite to what is found for isotropic attractions where increasing the range stabilises the dense fluid phase, resulting in the appearance of an equilibrium fluid-fluid transition, The particles have six degrees of freedom, three translational and three rotational. In order for a particle to form a bond with another particle both its position and orientation must be within small bounds. The distance between the centre of mass of the particle and that of the particle to which it is bonding must lie within the range $\sigma$ to $r_c$. The orientation of the site vector ${\bf u}_i$ must also lie within $\theta_c$ of the line joining the two centres of mass. From Eq. (\ref{kdef}) we see that in order for a bond to be formed between a pair of particles one of the translational degrees of freedom must be restricted to the small range $r_c-\sigma$ (the other two correspond to motion over a sphere of radius $\simeq\sigma$ and so are much less restricted), and two of the orientational degrees of freedom must be restricted (the other one corresponds to rotation about the axis of the bond and is unrestricted). In the solid phase all six degrees of freedom are restricted to small ranges, $\simeq r_c-\sigma$ for the translational degrees of freedom and $\simeq \theta_c$ for the orientational degrees of freedom. However, in the solid phase the particle forms $n_s$ bonds. In the fluid phase each bond that forms requires the restriction of one translational and two orientational degrees of freedom. So, in the solid phase when we go from four to six sites then the entropy does not change by much as in both cases all six degrees of freedom are restricted. Our theory, Eq. (\ref{mus}), actually predicts that the entropy is identical for four and for six sites but this is a consequence of the approximations we have made. But of course on adding two extra sites the energy per particle decreases by $\epsilon$. Thus the extra sites reduce the chemical potential of the solid phase, this then coexists with a lower density fluid phase: a fluid phase at a density below the lower of the two fluid-phase densities of the fluid-fluid transition. The fluid-fluid transition is then metastable. On examining Table 1 we see that neither the critical volume fraction, $\eta^{cp}$, nor the value of the second virial coefficient at the critical temperature, $B_2^{cp}$, are sensitive to the value of $\theta_c$. However, $\eta^{cp}$ is sensitive and $B_2^{cp}$ is very sensitive to the number of sites. Broide {\it et al.} \cite{broide91} estimate that the volume fraction of the globular protein $\gamma$-crystallin in the solution at the critical point is approximately 20\%, whereas that of lyzozyme is approximately 15\%. To give some idea of what we might expect $\eta^{cp}$ to be, its value in the van-der-Waals fluid is 0.13. The van-der-Waals fluid is a fluid of hard spheres with a very long-range attraction \cite{hansen86,sear95}. Both our six-site model and short-range isotropic attractions yield critical volume fractions above 0.13 but our four-site model predicts a lower volume fraction. Thus, the experiments rule out modeling the attractions as just four sites, at least for $\gamma$-crystallin. This is not a surprise, we expect there to be an isotropic attraction present and are neglecting it here purely for simplicity. As for the second virial coefficient at the critical temperature. It is equal to $-3.5\sigma^3$ for a van-der-Waals fluid. For short-range isotropic attractions it is less negative than this \cite{hagen94}. Indeed as the range of an isotropic attraction tends to zero the fluid-fluid transition crosses the fluid-solid transition and becomes a solid-solid transition and the value of the second virial coefficient at the critical point of the solid-solid transition tends to towards its value for hard spheres, $(2/3)\pi\sigma^3$ \cite{bolhuis94}. Correspondingly, $\eta^{cp}$ tends towards the close-packing volume fraction for hard spheres, $\simeq0.74$. The second virial coefficient of solutions of globular proteins can be measured via a scattering experiment. Measuring it under conditions near those for which the metastable fluid-fluid transition has its critical point would shed some light on the nature of the interactions. For instance, a large and negative second virial coefficient would rule out isotropic attractions as the dominant cause of the fluid-fluid transition. A small and negative second virial coefficient would however be ambiguous; both short-range isotropic attractions and our six-site model would be consistent with this finding. Our approximation for the free energy of the solid phase, Eq. (\ref{mus}) is quite crude. We can estimate the accuracy of the phase diagrams we have calculated by replacing our approximation of the lattice constant $a$ by $r_c$ which overestimates $a$, by the more conservative estimate $a=(r_c+\sigma)/2$ \cite{vega98}. With this approximation the trends we observe do not change however the range of parameter values over which the fluid-fluid transition is metastable does change. For example, with the parameter values used in Fig. 3 and approximating $a$ by $(r_c+\sigma)/2$ the fluid-fluid transition is just stable. It is stable for values of $\theta_c\ge0.28$ whereas with the approximation $a=r_c$ the fluid-fluid transition is only stable for $\theta_c\ge0.43$. By stable we mean that the critical point of the fluid-fluid transition does not lie within the fluid-solid coexistence region. The approximation $a=(r_c+\sigma)/2$ also, of course, implies a denser solid phase than $a=r_c$. \section{Conclusion} A simple model of a globular protein molecule, incorporating highly directional interactions, has been proposed. Example phase diagrams, Figs. 3 to 5, have been calculated. As observed by Vega and Monson \cite{vega98} highly directional attractions can have phase diagrams which are similar to those of short-range isotropic attractions; there is a fluid-fluid transition but it is metastable. We have also been able to show that if the number of sites is not too large, then directional attractive interactions can induce a stable fluid-fluid transition, see Fig. 4. As far as we are aware this is the first demonstration which shows a fluid-fluid transition driven by directional attractions which is definitely stable with respect to a fluid-solid transition. There have been several calculations of fluid-fluid transitions driven entirely by directional attractions with no isotropic attraction present \cite{nezbeda90,ghonasgi93,duda98} but only Vega and Monson \cite{vega98} considered the solid phase. Clearly, our potential with only directional attractions is as much a caricature of the interaction between globular proteins as is a potential with only a short-range isotropic attraction. It does have one advantage, however, in that, unlike isotropic attractions, the solid phase is not necessarily either face-centred-cubic or hexagonal-close-packed. In our model the lattice is determined by the arrangement of the sites. We chose an arrangement which allowed all four or six sites to form bonds in a simple-cubic lattice but we could equally well choose arrangements of sites which can bond in, say, a body-centred-cubic lattice. Globular proteins form solids with a variety of different solid lattices. If the interaction between globular proteins can be represented without too much error by a pairwise additive potential then the limited amount that we know of interactions between globular proteins suggests that a potential with a hard-core, and both isotropic and directional attractions is the most appropriate one to choose. Such potentials have been used with considerable success to predict the phase behaviour of hydrogen-bonding liquids \cite{blanca,gilvillegas97}. Our model does raise one issue: that of differences between the many globular proteins. Previous work \cite{rosenbaum96,tenwolde97,poon97,sear99b} on models with isotropic attractions inherently assumes that all globular proteins are alike. This supposition is supported by the presence of metastable fluid-fluid phase transitions for several different globular proteins. However, the ease with which a globular protein crystallises differs enormously from protein to protein. Indeed the fact that some have never been crystallised provides much of the motivation behind the work. If we vary the orientations of the site vectors, ${\bf u}_i$, then we dramatically change the stability of the solid phase. If the vectors are arranged in way which is not compatible with any solid lattice then at low temperatures there will be no solid phase (except at extremely high osmotic pressures). If in the solid phase not all the bonds can be made but they can in the fluid phase then the fluid will not solidify as this will increase the energy, and at low temperature the energetics dominates. We speculate that the proteins which have resisted efforts to crystallise them may have directional attractions which are not compatible with any solid lattice. If this is so then crystallising them will be all but impossible without minimising the strength of these directional attractions which will act to stabilise the fluid phase.
\section{Introduction} \label{sect:intro} \subsection{Physical motivation} \label{subsect:phys} Many concepts in physics rely on the idea of the normal modes (NMs) of a conservative system. Thus one speaks of energy eigenstates, molecular orbitals, energy bands, transitions between states and excitation energies (at the first quantized level), or of propagators in a Feynman diagram (at the second quantized level). These concepts for {\it quantum} and {\it interacting} (i.e., nonlinear) systems are rooted in eigenfunction expansions in terms of the NMs of {\it classical} and {\it linear} systems; e.g., the photon propagator in QED depends on the free classical electromagnetic (EM) field being expandable in plane waves. Such expansions are possible because conservative systems are associated with hermitian operators. On the other hand, if energy can escape to the outside, the system would be open and nonconservative, the associated mathematical operators are not hermitian in the usual sense. The eigenfunctions are then quasinormal modes (QNMs) with complex frequencies $\omega$. Figure 1a shows schematically the EM spectrum observed outside a linear optical cavity of length $a$, due to emission by a broad-band source, e.g., fluorescent dye molecules. The resonances, spaced by $\Delta \omega \approx \pi c/a$ (where $c$ is the speed of light), are QNMs, which are characteristic of the cavity rather than of the source. The resonance width $\gamma$ is determined (in the absence of absorption) by the amount of leakage. In the limit of zero leakage, these QNMs reduce to the NMs of a conservative cavity. It would be both interesting and useful if the physics of open systems could be discussed in terms of the discrete QNMs, in other words, an eigenfunction expansion for the dynamcis. This task is nontrivial because there are very few known results for eigenfunction expansion in nonconservative, non-hermitian systems. Yet the formulas we show in this article will look extremely familiar --- and that is precisely the point of this paper: the familiar formalism, with very little alteration, apply to a wide class of these systems. The problem of EM waves escaping from an open cavity is brought into sharp focus by advances in cavity QED: in Fabry-Perot cavities \cite{fab1,fab2}, in superconducting microwave cavities \cite{mic1,mic3}, in semiconductor heterostructures \cite{semi1} and in microspheres which confine glancing rays by total internal reflection \citeaffixed{mie2,sando}{see, e.g.,}. First, the spectrum is dominated by the resonances. Figure 1b shows an experimental spectrum observed from a microsphere, similar in essence to Figure 1a. Second, the decay rate of an excited atom or molecule is enhanced (suppressed) if the emitted radiation falls on (away from) the resonances \cite{mic1,fab1,fab2,barnes} --- the atom or the molecule of dimension $\mbox{$\sim 0.1 $ nm}$ ``knows" about its environment, on a scale of $\mbox{$\sim 1 \, \mu$m}$, even before any photon is emitted. Gravitational waves can be produced by matter falling into a black hole, and may soon be observed \citeaffixed{abram}{see, e.g.,}. The curvature of the intervening space forms an effective potential akin to a cavity, from which the gravitational waves eventually escape. Numerical simulations \cite{vish,det,smarr,stark,anninos} show that the amplitude is dominated, at intermediate times, by a ringing signal (Figure 2) $\sum a_j e^{-i\omega_j t}$, where $j$ labels the QNMs, and each complex frequency $\omega_j$ is characteristic of the background geometry rather than of the emitting source. In these and many other examples, the resonances or QNMs often dominate. But how complete are the QNMs? In mathematical terms, this means, first of all, whether any function $\phi(x)$ can be expanded in terms of the QNMs $f_j(x)$: \begin{equation} \phi(x) = \sum_j a_j f_j(x) \label{eq:expand1} \end{equation} \noindent More importantly, do the resonances represent the dynamics exactly: \begin{equation} \Phi(x,t) = \sum_j a_j f_j(x) e^{-i\omega_j t} \label{eq:expandt} \end{equation} \noindent for all $t \ge 0$? These expansion should make no reference to the environment or external bath. One would wish to establish conditions under which these expansions are valid, and in circumstances where they are not, to characterize the ``remainder". A projection formula, i.e., some sort of inner product, is needed to obtain the coefficients $a_j$ from $\phi(x)$. Then, initial value problems become formally trivial: take the initial data, project out $a_j$, and evolve by $a_j \mapsto a_j e^{-i\omega_j t}$. Moreover, to the extent that similarities can be established with the conservative case, one should be able to transcribe the tools of mathematical physics, and to seek a parallel formalism --- including, e.g., second quantization. These developments should lead to a unifying view of QNMs in different systems. This article discusses recent progress towards these goals. Open systems are a special class of dissipative systems, which have been studied by many authors, including \citeasnoun{feyn}, \citeasnoun{uller}, \citeasnoun{cald}. In all cases, the bath degrees of freedom are first included, and then eliminated from the path integral or equations of motion. Similarly, an open cavity (linear dimension $a$) can be embedded in a ``universe" (linear dimension $\Lambda \rightarrow \infty$), and the totality is a conservative system with the modes of the universe (MOU) forming a continuum ($\Delta \omega \propto \Lambda^{-1} \rightarrow 0$). A rigorous theory of lasing in 1-d cavities has been developed using the MOU \cite{lang1,lang2} and the ideas can be generalized to higher dimensions, e.g., optics in microspheres \citeaffixed{ching1,ching2}{see, e.g.,}, including nonlinear optical processes such as spontaneous and stimulated Brillouin scattering \cite{bs,sbs}. Since each eigenfunction or QNM is a resonance, the expansions (\ref{eq:expand1}) and (\ref{eq:expandt}) in effect expresses the idea of resonance domination. In cavity QED, resonance domination was first discussed by \citeasnoun{purcell}. He proposed that in the Fermi golden rule, the the density of states per unit volume $\rho_0(\omega) = \omega^2/(\pi^2c^3)$ should be replaced by $\rho(\omega) \sim D/(2\gamma V)$ for a $D$-fold degenerate QNM of width $\gamma$ in a cavity of volume $V$. This leads to an enhancement factor of $K = \rho/\rho_0 \sim (1/8\pi)DQ (\lambda^3/V)$ for spontaneous emission, where $\lambda$ is the wavelength of light emitted and $Q$ is the quality factor of the cavity. Purcell's argument can be made more precise using the MOU \cite{ching1,ching2}: the photon modes are re-distributed --- enhanced at the resonance and depleted away from the resonances. These works, in company with much earlier works in metastable states in nuclear physics \cite{gamow,siegert} typically use {\em some} resonances for an {\em approximate} description. Our focus will be on the use of {\em all} the discrete resonances to obtain, if possible, an {\em exact} description --- and where this fails, to characterize the remainder. \subsection{The mathematical problem} \label{subsect:math} Waves in open systems may be governed by the Schr\"odinger equation (e.g., nuclear physics), the wave equation (e.g., optics) or the Klein-Gordon (KG) equation with a potential $V$ (e.g., each angular momentum sector of linearized gravitational waves \cite{chand}). The wave equation can be mapped exactly into the KG equation, and as far as time-independent poblems are concerned, the KG equation can be mapped into the Schr\"odinger equation by relabeling $\omega^2 \rightarrow \omega$. Thus, many properties are similar (although we shall also highlight the key differences), and we shall focus on the scalar wave equation, on a certain domain ${\cal R}$, \begin{equation} D \, \Phi( {\bf r} , t ) \equiv \left[ \rho( {\bf r} ) \partial_t^2 - \nabla^2 \right] \Phi ( {\bf r} , t) = 0 \label{eq:wave3d} \end{equation} \noindent This describes the scalar model of EM ($\rho =$ dielectric constant, and henceforth $c=1$) or elastic vibrations ($\rho =$ density). In this article, we shall for the most part consider the 1-d version ($\nabla^2 \mapsto \partial_x^2$), restricted to a half line $0 \le x < \infty$; the ``cavity" is the interval ${\cal R} = [0,a]$. A node is imposed at the origin (a totally reflecting mirror), and waves escape through the point $x=a$ (a partially transmitting mirror) to the rest of the ``universe" in $(a, \infty)$. This model describes a laser cavity \cite{lang1}, as well as gravitational radiation from a stellar object \cite{price}, or the transverse vibrations of a string \cite{dekker4,laistring}. For conservative systems, one usually imposes $\Phi =0$ on the boundary $\partial {\cal R}$ ($x=a$ in the 1-d case) \footnote{The Dirichlet condition can be replaced by the Neumann condition.}. The eigenfunctions or NMs are factorized solutions $\Phi( {\bf r}, t) = f( {\bf r} ) e^{-i\omega t}$. The nodal condition on $\partial {\cal R}$ implies that ${\cal R}$ is {\it closed}, so that energy cannot escape. Mathematically, the operator $-\nabla^2$ is hermitian; thus $\omega$ is real and the eigenfunctions $\{ f \}$ form a complete orthogonal set. In contrast, if $\Phi$ satisfies the outgoing wave condition on $\partial {\cal R}$; this defines an {\em open} system. The eigenfunctions, labeled by an index $j$, satisfy (in 1-d) $\partial_x^2 f_j (x) = -\omega_j^2 \rho(x) f_j(x)$, with $\mbox{Im } \omega_j < 0$. Note a doubling of modes compared with the conservative case: If $\omega_j$ is an eigenfrequency, then so is $\omega_{-j} \equiv -\omega_j^*$, but since $\omega_{-j}^2 \neq \omega_j^2$, $f_{-j}(x)$ and $f_j(x)$ are linearly independent. The question then is whether these $\{ f_j \}$ are complete in the sense of (\ref{eq:expand1}) and (\ref{eq:expandt}). \section{Completeness} \label{sect:comp} Even though QNMs do not always capture all the physics, it is still useful to first establish the baseline case in which the QNMs are complete on ${\cal R} = [0,a]$. This holds provided two conditions are satisfied: (a) $\rho (x)$ has a step or stronger discontinuity at $x=a$ (discontinuity condition); (b) $\rho(x)=1$ (or other constant value) for $x>a$ (``no-tail" condition). These two conditions can be understood heuristically. A discrete representation of $\phi(x)$ such as (\ref{eq:expand1}) could at best be valid over a finite interval, for which a discontinuity provides a natural demarcation. Moreover, a dynamical expansion such as (\ref{eq:expandt}) would require that waves are not affected by the outside $(a,\infty)$; this can only be the case if $\rho$ is trivial outside, and thus outgoing waves are not scattered back. We now sketch the main elements of the proof of completeness, which will also provide the framework for understanding other contributions when the QNMs are not complete. The Green's function $G(x,y;t)$, defined by $D G(x,y;t) = \delta(x-y) \delta(t)$, with $G(x,y;t)=0$ for $t \leq 0$. Its fourier transform ${\tilde G}(x,y;\omega)$ is given explicitly as ${\tilde G}(x,y,\omega) = f(\omega,x) g(\omega,y) / W(\omega)$ for $0 \leq x \leq y$. Here $f$ and $g$ are homogeneous solutions to the time-independent wave equation at frequency $\omega$, with $f$ satisfying the left nodal boundary condition ($f(\omega,0) = 0$) and $g$ satisfying the right outgoing wave boundary condition ($g(\omega,x) \propto e^{i\omega x}$ for $x \rightarrow \infty$). Their Wronskian is $W$, and the combination $fg/W$ is independent of normalization convention. For $t \ge 0$, $G(x,y;t)$ is evaluated by the inverse Fourier transform with the contour of integration closed by a large semicircle in the lower half $\omega$-plane. In general, there are three different contributions. \noindent {\em Large semicircle --- prompt response} The contribution along the large semicircle ($|\omega| \rightarrow \infty$) gives the short time or prompt response. For large $t$, the factor $e^{-i\omega t}$ provides sufficient damping in the lower half $\omega$-plane to control the asymptotic behavior and it can be shown \cite{bachelot,lly1} that the prompt response always vanishes for $t > t_p(x,y)$, where $t_p(x,y)$ is the geometric optics transit time between $x$ and $y$. If the discontinuity condition is satisfied, one has a stronger result: by a WKB estimate, this contribution vanishes for all $ t \ge 0$. It is natural that the $|\omega| \rightarrow \infty$ behavior should be sensitive to short spatial length scales, in particular a discontinuity. \noindent {\em Singularities in $g$ --- late-time tail} The function $f(\omega,x)$ is obtained by integrating the time-independent wave equation, in which $\omega$ appears analytically, through a {\em finite} distance from $0$ to $x$. Crudely speaking, this is a finite combination of analytic functions of $\omega$, so $f(\omega,x)$ is guaranteed to be analytic in $\omega$ \cite{newton}. On the other hand, $g(\omega,x)$ could have discontinuities in $\omega$ (typically a cut on the negative $\mbox{Im } \omega$ axis), because it is obtained by integrating through an infinite interval starting from $x = \infty$. However, if the ``no tail" condition is satisfied, then one can impose the right boundary condition for $g(\omega,x)$ at $x=a^+$; one then again integrates only through a finite distance, thus guaranteeing the analyticity of $g$ as well, and removing this contribution. \noindent {\em Zeros of W --- QNM contributions} Finally, there are contributions from the zeros of $W(\omega)$. At a zero $\omega_j$ of $W$, the functions $f$ and $g$ are linearly dependent: $f_j(x) \equiv f(\omega_j , x) = C_j g(\omega_j , x)$; thus $f_j$ satisfies {\em both} the left nodal and the right outgoing-wave boundary conditions, and is an eigenfunction or QNM with eigenfrequency $\omega_j$. In general, all three contributions are present, and correspond nicely with the main features of the time domain signal \cite{taillett,tail}, as shown in the spacetime diagram in Figure 3. The contribution from the semicircle in the $\omega$-plane leads to the initial transients --- a wave propagating directly from the source point $y$ to the observation point $x$ (rays $1a, 1b$). The zeros of $W$ give rise to the QNM ringing at intermediate times, which are caused by repeated scattering from the potential (ray $2$). The cut in the $\omega$-plane, especially its tip near $\omega=0$, then gives the late-time tail: waves propagate from the source point $y$ to a distant point $x'$, are scattered by $\rho(x')$, and return to the observation point $x$ (ray $3$). We now focus on the case where the discontinuity and ``no-tail" conditions are satisfied. At each zero $\omega_j$, the residue is related to $dW(\omega_j)/d\omega$, which can be evaluated using the defining equations for $f$ and $g$ to be \begin{eqnarray} -\frac{1}{C_j} \frac{dW(\omega_j)}{d\omega} &=& 2 \omega_j \int_0^{a^+}\mbox{\Large{}} \rho(x) f_j(x)^2 dx + i f_j(a^+)^2 \nonumber \\ & \equiv & \langle f_j | f_j \rangle \label{eq:norm} \end{eqnarray} \noindent Using the normalization (and phase) convention $ \langle f_j | f_j \rangle = 2\omega_j$, one then obtains \begin{equation} G(x,y;t) = \frac{i}{2} \sum_j \frac{1}{\omega_j} f_j(x) f_j(y) e^{-i\omega_j t} \label{eq:grep} \end{equation} \noindent The initial conditions $G(x,y;t=0)=0$, $\rho(x) \partial_t G(x,y;t=0) = \delta(x-y)$ lead to\footnote{Subject to the validity of term-by-term differentiation and the limit $t \rightarrow 0$.} \begin{eqnarray} \frac{i}{2} \sum_j \frac{1}{\omega_j} f_j(x) f_j(y) &=& 0 \label{eq:id1} \\ \frac{ \rho(x) }{2}\sum_j f_j(x) f_j(y) &=& \delta(x-y) \label{eq:id2} \end{eqnarray} \noindent for $x,y \in {\cal R}$. The identity (\ref{eq:id2}) has an obvious analog in conservative systems; the factor $1/2$ accounts for the doubling of modes. On the other hand (\ref{eq:id1}) becomes vacuous in the limit $\epsilon \rightarrow 0$. The heart of the eigenfunction expansion is contained in (\ref{eq:grep}), (\ref{eq:id1}) and (\ref{eq:id2}). The identity (\ref{eq:id1}) obviously leads to (\ref{eq:expand1}), while (\ref{eq:grep}) will be seen to lead to (\ref{eq:expandt}). \section{Time-independent problems} \label{sect:timeindep} This simplest and most useful applications concern time-independent problems, especially time-independent perturbation theory. \subsection{Perturbation} \label{subsect:pert1} Let $\rho (x) = \rho_0 (x) + \mu V(x)$, where $|\mu |\: \ll 1$, and $V(x)$ is nonzero only in the cavity. The system with $\rho_0 (x)$ is exactly solvable and its QNMs form a complete set for $x \in [0,a]$ in the sense of (\ref{eq:grep}), (\ref{eq:id1}) and (\ref{eq:id2}). Thus, it is straightforward to develop a {\em discrete} representation for the frequency shifts $\omega _j = \omega ^{(0)}_j + \mu \omega ^{(1)}_j + \mu ^{2} \omega ^{(2)}_j + \cdots $ : \begin{equation} \omega ^{(1)}_j = - {\omega ^{(0)}_j \over 2} V_{jj} \label{eq:shift1a} \end{equation} \begin{equation} \omega ^{(2)}_j = {\omega ^{(0)}_j \over 4}\Bigl\{2(V_{jj})^2 + \sum^{}_{i\neq j}\frac{ V_{ji} V_{ij} \omega ^{(0)^2}_j} { \omega ^{(0)}_i(\omega ^{(0)}_j - \omega ^{(0)}_i)} \Bigr\} \label{eq:shift2a} \end{equation} \noindent etc., where the matrix elements are $V_{ji} = \int_0^a f_j^{(0)}V f_i^{(0)} dx$ These complex formulas give the shifts in the resonance positions and the widths --- thus they contain twice the information compared with their superficially similar conservative counterparts. The wavefunctions can be calculated in a parallel way \cite{lly2}. The first-order correction has been known for a long time for the Schr\"{o}dinger equation, but here the normalization condition (see (\ref{eq:norm})) is stated through a surface term rather than via regularization \cite{zel}, which is less convenient computationally. The first-order result has been generalized to the EM case \cite{per1}, and in fact applies also to systems without discontinuities. To illustrate, let $\rho(x) = 1 + n^2 + \mu \Theta(x-b)$ for $x<a$ and $\rho(x) = 1$ for $x>a$ (i.e. a ``dielectric rod" with a ``bump"). Figure 4 shows the QNM positions in the $\omega$-plane for $n^2 = 4.0$, $\mu=0.4$, compared with $\mu=0$. The two are quite different, in that for the former case $\mbox{Im $\omega_j $}$ are oscillatory in the mode number $j$. The perturbative results based on (\ref{eq:shift1a}) and (\ref{eq:shift2a}) are also shown in Figure 4; although these formulas mimic the case of {\em real \/} frequencies for conservative systems, they nevertheless give the correct shifts, even for the imaginary parts. The sum over intermediate states is discrete, at a spacing $\Delta \omega \sim \pi / a$; thus the effective expansion parameter is $\mu / |\Delta \omega| \sim \mu a / \pi$, a perspective not apparent in the MOU approach. \subsection{Physical examples} \label{subsect:microdroplet} Some interesting applications relate to optics; the interface between two media naturally gives a discontinuity, and vacuum outside the system naturally satisfies the ``no tail" condition. Consider a microsphere of radius $a$ and refractive index $n$. Glancing rays suffer total internal reflection and are confined, but evanescent waves cause some leakage, and this makes the system an open one. In practice, experiments are often done on droplets falling in air, slightly oblate (ellipticity $e \sim 10^{-3} - 10^{-2}$) due to viscous drag. The breaking of spherical symmetry lifts the degeneracy among the $2l+1$ members of a multiplet. This problem has been treated by brute force numerical calculation \cite{hill}, but the perturbative formalism allows an analytic expression \cite{per1} \begin{equation} \frac{\omega_m^{(1)}}{\omega^{(0)}} = -\frac{e}{6} \left[ 1 - \frac{3m^2}{l(l+1)} \right] \label{eq:split} \end{equation} \noindent in which $m$ is the azimuthal quantum number, and $e = (r_p - r_e)/a$, where $r_p$ and $r_e$ are respectively the polar and equatorial radii, and $a = (r_p r_e^2)^{1/3}$. This perturbative formula has been used to interpret spectroscopic measurements and to determine the ellipticity $e$, either by measuring the splitting between two neigboring lines ($m$ and $m+1$) \cite{split1}, or by determining the total spread of the multiplet (from $m=0$ to $m= \pm l$) \cite{split2}. Two other developments along this line should be mentioned. In one interesting experiment \cite{arnold}, a quadrupole field $E_Q$ imposed on a levitated droplet caused a spheroidal distortion with an ellipticity $e \propto E_Q/s$, where $s$ is the surface tension. The splitting of the multiplet was determined spectroscopically and the ellipticity $e$ was then found from (\ref{eq:split}), allowing $s$ to be determined, down to length scales of tens of $\mu$m. (Hitherto, surface tension is known only macroscopically, at length scales of say 1 mm or more.) In another experiment \cite{prec}, light is launched into a slightly oblate microdroplet, propagating along great circles inclined at an angle $\theta$ to the equatorial plane. In photon language, the angular momentum ${\vec L}$ satisfies $m/l = L_z/L = \cos \theta$, and precesses at the group angular velocity $\Omega = d \omega_m/ dm = \omega |e| m / [l(l+1)] $. The precession is observed as variations in intensity when a particular spot at an angle $\theta$ is observed. The observed period of typically 50 ps then allows the ellipticity to be determined. On a very different length scale, the perturbation results have also been applied to gravitational waves. The QNM spectrum of linearized gravitational waves propagating on a Kerr or Schwarzschild background has been extensively studied analytically \cite{chand} and numerically \cite{vish,det,smarr,stark}, and their complex frequencies are known in detail \cite{leaver1,guinn,leaver2,andersson,nollert}. However, black holes are perturbed by interactions with its surroundings (e.g., a massive accretion disk). The shifts may provide a probe of the intervening spacetime curvature. This situation is obviously one to which perturbation theory can be applied. The first-order perturbation theory is expressed by the KG analog to (\ref{eq:shift1a}), further generalized to allow for a ``tail" \cite{dirtybh}.\footnote{With a ``tail", the QNMs are not complete, but it is not surprising that this does not matter for the first-order shift.} We here present some results for a Schwarzschild black hole perturbed by a thin static shell of matter \cite{dirtybh}. Figure 5 shows, in the complex $\omega$-plane, the locus of the lowest QNM for $l=1$ scalar waves propagating on a Schwarzschild background, as a light shell ($0.01$ times the mass of the hole) is placed at different positions $r_s$. The first-order perturbative results (dashed lines) capture the qualitative features of the exact results (solid lines). \section{Dynamics} \label{sect:timedep} Time-dependent problems require two independent sets of initial data: $\phi \equiv \Phi(x,t=0)$ and ${\hat \phi} \equiv \rho(x) \partial_t \Phi (x,t=0)$. \footnote{These are assumed to vanish at infinity, representing waves emitted at a finite time in the past.} One is thus led to consider the {\em simultaneous} expansion of a {\em pair} of functions $(\phi, {\hat \phi})$ \begin{equation} \pmatrix{ \phi(x) \cr {\hat \phi}(x) \cr } = \sum_j a_j \pmatrix{ 1 \cr -i \omega_j \rho(x) \cr } f_j(x) \label{eq:expand2} \end{equation} \noindent using the {\it same} coefficients $a_j$'s for both components. The factor $\rho(x)$ turns the second component into the conjugate momentum. The use of two components is natural \cite{fesh,unruh,ford}, but here the outgoing wave condition \footnote{The last condition is specified at $a^+$ because the system may have discontinuities at $x=a$.} ${\hat \phi} (x=a^+) = - \phi ' (x=a^+)$ links them in a novel way. Compared with the case of NMs, twice the degrees of freedom ($a_j$ and $a_{-j}$) are used to satisfy twice the conditions ($\phi$ and ${\hat \phi}$). In the conservative limit, $\omega_{-j}^2 = \omega_j^2$ and $f_{-j}(x) = f_j(x)$. Thus, upon grouping pairs of terms, $\phi(x) = \sum_{j>0} b_j f_j(x)$, ${\hat \phi}(x) = \rho(x) \sum_{j>0} c_j f_j(x)$, where $b_j = a_j + a_{-j}$, and $c_j = -i\omega_j (a_j - a_{-j})$. So in this limit, the simultaneous expansion of two functions $(\phi, {\hat \phi})$ using the full set of NMs $j= \pm1, \pm2, \cdots$ is equivalent to the separate and independent expansion of $\phi$ and ${\hat \phi} / \rho$ using only the NMs $j=1,2, \cdots$. The solution to the initial value problem is then \begin{equation} \Phi (x,t) = \int_0^{\infty} \left[ G(x,y;t) {\hat \phi} (y) + \partial _t G(x,y;t) \rho(y) \phi (y) \right] dy \label{eq:dyn} \end{equation} \noindent From (\ref{eq:dyn}), one can then obtain (\ref{eq:expand2}) and also (\ref{eq:expandt}) with $a_j$ given by \begin{eqnarray} a_j = \frac{i}{2 \omega_j} \bigg\{ \int_0^{a^+} &&\left[ f_j(y) {\hat \phi}(y) + {\hat f}_j(y) \phi(y) \right] dy \nonumber \\ && + \, f_j(a) \phi(a) \bigg\} \label{eq:aj} \end{eqnarray} \noindent All reference to the outside of the cavity is removed, because the integral on $(a,\infty)$ can be collapsed to the surface term, by making use of the outgoing condition on the initial data and the retarded condition on $G$. The elimination of the outside is the crucial step in obtaining a self-contained description of the cavity. The projection formula (\ref{eq:aj}) suggests the definition of a generalized inner product, and with it a linear space structure analogous to that for conservative systems \cite{twocomp1,twocomp2}. Consider the linear space of outgoing function pairs $ (\phi, {\hat \phi})$, denoted as ket vectors $| \mbox{\boldmath $\phi$} \rangle = (\phi(x) , {\hat \phi}(x) )^{\rm T}$. For a QNM, ${\hat f}_j(x) = -i\omega_j \rho(x) f_j(x)$, where $\omega_j$ is the eigenvalue. Time-evolution can be written in Schr\"odinger form $\partial_t | \mbox{\boldmath $\phi$} \rangle \, = \, -i \cal{H} | \mbox{\boldmath $\phi$} \rangle$, where \begin{equation} \cal{H} = \mbox{$i$} \pmatrix{ 0 & \rho(x)^{-1} \cr \partial_x^2 & 0 } \label{eq:htwocomp} \end{equation} \noindent The first component of the Schr\"odinger equation reproduces the identification of ${\hat \phi}$ as $\rho(x) \partial_t \Phi$. The eigenfunctions, or QNMs, can now be defined simply by $\cal{H} \, | \mbox{\boldmath $f_j$} \rangle = \mbox{$\omega_j$} \, | \mbox{\boldmath $f_j$} \rangle$. The projection (\ref{eq:aj}) suggests a generalized inner product between two vectors $| \mbox{\boldmath $\phi$} \rangle$ and $| \mbox{\boldmath $\psi$} \rangle$: \begin{equation} \langle \mbox{\boldmath $\psi$}| \mbox{\boldmath $\phi$} \rangle \, = \, i \left[ \int_0^{a^+ } ( \psi {\hat \phi} + {\hat \psi} \phi ) dx \, + \, \psi(a) \phi(a) \right] \label{eq:inner} \end{equation} \noindent which is symmetric and linear in both the bra and ket vectors (rather than conjugate linear in the bra vector). The inner product of a QNM with itself agrees exactly with the generalized norm (\ref{eq:norm}). Moreover, the coefficients (\ref{eq:aj}) for the eigenfunction expansion can now be written compactly as \begin{equation} a_j = \frac{\langle \mbox{\boldmath $f$}_j | \mbox{\boldmath $\phi$} \rangle} {\langle \mbox{\boldmath $f$}_j | \mbox{\boldmath $f$}_j \rangle} = \frac{1}{2 \omega_j} \langle \mbox{\boldmath $f$}_j | \mbox{\boldmath $\phi$} \rangle \label{eq:ajcomp} \end{equation} Under the generalized inner product, the Hamiltonian is symmetric: $\langle \mbox{\boldmath $\psi$} | \mbox{ {\large \{}} \cal{H} | \mbox{\boldmath $\phi$} \rangle \mbox{ {\large \}}} = \langle \mbox{\boldmath $\phi$} | \mbox{ {\large \{}} \cal{H} | \mbox{\boldmath $\psi$} \rangle \mbox{ {\large \}}} \equiv \langle \mbox{\boldmath $\psi$} | \cal{H} | \mbox{\boldmath $\phi$} \rangle$. In the proof of this statement, there is an integration by parts; surface terms are incurred because the functions do not vanish at the end-points, but these are compensated exactly by the surface terms in the definition of the inner product. There is thus an almost complete parallel with conservative systems, and the mathematical structure is in place to carry over essentially all the familiar tools based on eigenfunction expansions. For example, the symmetry of ${\cal H}$ (similar to the hermitian property in the conservative case) immediately leads to the orthogonality of the QNMs and hence the uniqueness of the expansion (\ref{eq:expand2}). The only exception is the lack of a positive-definite norm, and with it a simple probability interpretation --- hardly surprising since probability (or energy) is not conserved in ${\cal R}$. For example, the perturbation theory can now be obtained by an almost trivial transcription of textbook derivations. The formalism generalizes to a full line ($-\infty < x < \infty$); the expansion is valid in an interval $[a_1,a_2]$, provided (a) there are at least two discontinuities, with the extreme ones at $a_1$ and $a_2$; (b) there is ``no tail" on either side, and (c) the outgoing condition is imposed at the two ends $a_1^-$ and $a_2^+$. Much the same formalism applies to the KG equation. Discontinuities must be present in the potential $V(x)$, and the ``no tail" condition is that $V(x)=0$ on the ``outside" \cite{kgcomplett,kgcomp}. Absorption can also be included; e.g., in the scalar model of EM, the complex and frequency-dependent dielectric function ${\tilde \rho}(x,\omega)$ must satisfy the usual Kramers-Kronig dispersion relation \cite{lly2}. To apply to cavity QED, it is necessary to (a) go from free fields to interacting fields, and (b) go from classical fields to quantized fields. Here we indicate the rudiments of this program. Consider a nonlinear wave equation: $D \Phi + \lambda(x) \Phi^3 = 0$ where $\lambda(x)$ has support only within ${\cal R}$ (interaction occurs only inside the cavity). Expanding this in terms of the eigenfunctions of noninteracting theory, $( d/dt + i\omega_j) a_j = -i/(2\omega_j) \sum_{klm} \lambda_{jklm} a_k a_l a_m$ where $\lambda_{jklm} = \int_0^{a^+} \lambda f_j f_k f_l f_m dx$ Time-dependent perturbation of the linear theory (i.e., adding a time-dependent piece to $\rho(x)$) can be handled in very much the same way. The point is that the dynamics is now expressed through {\em discrete} variables $a_j(t)$ Since $\phi(x)$ and ${\hat \phi}(x)$ satisfy $\left[ \phi(x) , {\hat \phi}(y) \right] = i\delta (x-y)$ at equal times, one obtains from (\ref{eq:aj}) commutation relations for $[a_j, a_k^+]$. The commutators differ slightly from $\delta_{ij}$, the difference being a measure of the dissipation. Feynman rules can be developed in the standard way. This description, unlike the MOU approach, will be particularly useful when the cavity is only slightly leaky and the connection with a totally enclosed cavity is to be emphasized. Applications to emission, absorption and other quantum phenomena are being developed. An important result will be to put idea of \citeasnoun{purcell} on resonance enhancement of decay rates on a firm footing, by writing it as the contribution of one discrete QNM, in a precise manner. \section{Non-resonant contributions} \label{sect:gen} We have so far concentrated on the baseline case in which the discontinuity and ``no tail" conditions are satisfied, and the QNMs are complete. However, in other cases there are non-resonant contributions, and a brief discussion is given in this Section. Gravitational radiation from Schwarzschild black holes is described by the KG equation; the potential $V(x)$, related to the background metric, is continuous and goes asymptotically as $x^{-3} \log x$ as $x \rightarrow +\infty$ (apart from a centrifugal barrier), violating both the discontinuity and the ``no tail" condition. We indicate briefly how this situation affects the dynamics \cite{taillett,tail}. First, because there is no discontinuity, there is a prompt signal due to the contribution from the large semicircle. Secondly, the auxiliary function $g(\omega,x)$ has to be integrated from $x \rightarrow +\infty$. This leads to a cut on the negative $\mbox{Im $\omega$}$ axis, controlled by the spatial asymptotics of $V(x)$. The cut extends all the way to $\omega=0$, and determines the late time dynamics. Generically, if $V(x) \sim x^{-\alpha} (\log x)^{\beta}$, then $\phi(x,t) \sim t^{-\mu} (\log t)^{\nu}$, where $\mu$ and $\nu$ are determined by $\alpha$, $\beta$ and the angular momentum $l$. For example, for $\beta = 0$ or $1$, in general $\mu = 2l+\alpha$ and $\nu = \beta$. It turns out, interestingly, that the case of the Schwarzschild black hole ($\alpha = 3$, $\beta=1$) is exceptional for $l > 0$: the leading term $t^{-(2l+\alpha)}\log t$ vanishes, and the dynamics is dominated by the next leading term $t^{-(2l+\alpha)}$. Figure 2 shows precisely such a power-law tail. However, the signal at intermediate times would still be dominated by the sum over QNMs, so that in practice, a restricted and approximate notion of completeness still holds. For the Schr\"{o}dinger equation, outgoing waves are defined by $\phi(x) \sim e^{ikx}$ as $x \rightarrow \infty$, where $k = \sqrt{2\omega}$ instead of $k = \omega$, and moreover $k>0$. Thus a continuum contribution cannot be avoided, and the system decays by a power-law $t^{-\alpha}$ at large times, due to the contribution near threshold ($k \approx 0$). This phenomena was first noted by \citeasnoun{khafin} and \citeasnoun{winter}. In particular, \citeasnoun{winter} considered a 1-d well bounded by an impenetrable wall and a delta-function potential and showed that the wavefunction inside the well decays as $t^{-3/2}$ as large $t$. However, if the potential is unbounded from below, no finite threshold exists, and the large $t$ behavior is again given by the QNMs \cite{sy1}. This turns out to apply to some models of mini-superspace (in which $x$ is essentially the scale factor of the universe) \cite{sy2}, and there is the intriguing possibility that wavefunction of the universe at the end of the quantum era is given by the lowest QNM, independent of cosmological initial conditions. The power-law decay in the case of the Schr\"odinger equation comes from the threshold, whereas the power-law in the KG case comes from the spatial tail of $V(x)$; the origins of the two are entirely different. Despite these complications, in time-independent problems, one is free to relabel $\omega^2 \mapsto \omega$, so the Schr\"{o}dinger case should not be different. An application in this restricted time-independent sense is given in \citeasnoun{per4}. \section{Conclusion} \label{sect:diss} We have considered {\em linear}, {\em classical \/} waves in an {\em open} system of a certain type, and sketched a formalism in terms of an QNM expansion, in a manner analogous to the description of conservative systems by its NMs. Dissipation is contained in these discrete QNMs themselves, especially in Im $\omega_j$. Remarkably, apart from this feature, almost nothing needs to be changed, and the familiar tools for hermitian systems can be carried over. These results are nontrivial, in that if either the discontinuity condition or the ``no tail" condition is violated, the QNMs would not be complete. The most significant feature in these cases is a power-law tail in the large $t$ behavior. With these mathematical concepts in place, the way is now open for formulating QED phenomena in an open cavity, using a {\em discrete} basis.
\section{Introduction} The distribution of element abundances in galaxies is an important fossil record of their formation and evolution. Of primary importance is the metallicity (the mass fraction of all elements heavier than He), which contains important information about the past star formation history, which could well be strongly influenced by mergers and interactions. For early-type galaxies, the metallicity is an excellent estimator of the luminosity, or even the mass. Also very important are metallicity gradients, which are among the few parameters that can tell us something about the orbital structure in galaxies. In recent years the quality of observational data has become so good, that one can also start thinking of measuring abundances of individual elements in external galaxies. Individual abundances will greatly help to understand their chemical evolution. In particular, one will be able to understand better the way in which the ISM of galaxies is enriched by metals, and what the relevant time-scales are. In the last 2 decades it has become clear that the abundance distribution in stars is not always the same as in the Sun. This was discovered first in individual stars in our galaxy (see Wheeler, Sneden \& Truran 1989) and later in integrated spectra of elliptical galaxies (Peletier 1989, Worthey, Faber \& Gonz\'alez 1992). In this review I will discuss what we currently know about abundance ratios in galaxies, and what we can learn from this about the formation and evolution of galaxies. This paper is in part based on the excellent paper by Worthey (1998), but also includes some new high-quality data, which might shed new light on some issues in this rapidly evolving field. The paper starts in Section 2 discussing some global relations for galaxies as a function of luminosity or velocity dispersion. In Section 3 the central abundances and abundance ratios are discussed, and in Section 4 their gradients. In Section 5 it is briefly discussed how we can understand non-solar abundance ratios. The need for stellar models with non-solar abundance ratios is mentioned in Section 6, after which some conclusions are given. \section{Global Relations} It has been known for some time that the stellar populations of early-type galaxies are strongly linked to their other properties. Two examples are the colour-magnitude relation (see e.g. Sandage \& Visvanathan 1978) and the relation between Mg$_2$ and velocity dispersion ($\sigma$) (Terlevich et al. 1981). Both relations can be understood well if the average metallicity of a galaxy is larger when the galaxy is brighter. The usefulness of these relations in understanding galaxies strongly increased when Schweizer \& Seitzer (1992) showed that the residuals of the colour/Mg$_2$ -- $\sigma$ relation correlate with a parameter indicating recent mergers. This correlation implies that the scatter for undisturbed early-type galaxies is smaller than the amount due to observational uncertainties, while colours of disturbed galaxies are somewhat bluer for a given $\sigma$, due to increased star formation during the merger. This interpretation was confirmed by the work of Bower, Lucey \& Ellis (1992), who found a very low scatter in the colour-magnitude relation (CMR) in the Coma cluster, showing no sign of any recent star formation in the early-type galaxies in this cluster. Later, using the more sensitive H$\gamma$ absorption line index, Caldwell et al. (1993) showed that many early-type galaxies in the SW part of the cluster show signs of small amounts of young stars. Recently, A. Terlevich (1998) redid the study of Bower et al. (1992) with a significantly larger sample in the Coma cluster. He finds that the intrinsic scatter in the elliptical galaxies in $U-V$ is 0.036 mag (consistent with Bower et al.). No difference was found in the slope of the CMR between ellipticals and S0 galaxies. The slope of the CMR was also the same in different areas of the cluster. Most galaxies blue-ward of the CMR are either late-type galaxies (Andreon et al. 1996) or seem to have late-type morphologies. In the outer parts of the cluster the residuals about the CMR become somewhat bluer, in agreement with the the results of Caldwell et al. (1993). These detailed colour studies indicate that the colour of an early-type galaxy is determined by its luminosity to a high accuracy. The same can be said of the Mg$_2$ absorption line (Bender, Burstein \& Faber 1993, Schweizer \& Seitzer 1992). Although a colour of a galaxy, and also an absorption line, depends on many parameters, like metallicity, age and Initial Mass Function (IMF) slope, the fact that the CMR has the same slope across various factors of 10 in luminosity implies that the relations described above are almost certainly driven by metallicity: fainter galaxies have lower metallicities, and because of that bluer colours. This behaviour has been successfully reproduces in Galactic enrichment models (e.g. Arimoto \& Yoshii 1987). Starting with a proto-galactic cloud, several generations of stars are formed, until the rate of Supernovae is so large that all the gas is expulsed from the galaxy, and the star formation stops. Being able to model the CMR has been very important for our understanding of galaxy formation. We are now however in a position that we can go one step further, and ask ourselves how the abundances of individual elements varies as a function of $\sigma$ or luminosity. We know that the strength of the Mg$_2$ feature depends strongly on the Mg abundance (Worthey 1998), so the fact that there is a good correlation between Mg$_2$ and $\sigma$ or luminosity tells us that the Mg-abundance increases with that parameter. Colours don't contain much information about individual element-abundances, except that blue colours, through line blanketing, depend very much on the abundance of Fe-peak elements, which would imply that the Fe-elements would be a strong function as well of $\sigma$. Do we know whether the abundances of other elements also correlate strongly with $\sigma$? High-quality measurements of elements other than Mg are scarce, because of the high signal-to-noise required, and the difficulty associated with calibrating indices like $\langle$Fe$\rangle$ and H$\beta$ onto the Lick system (Faber et al. 1985, Gorgas et al. 1993, Worthey et al. 1994). Recently Fisher, Franx \& Illingworth (1996) and J\o rgensen (1997) published $\langle$Fe$\rangle$ and $\sigma$ data for a reasonably large number of galaxies. Although in both cases the scatter in the $\langle$Fe$\rangle$ -- $\sigma$ relation is large, it is smaller in the data of J\o rgensen (1997). In Fig.~1a we show her Figure 2, displaying Mg$_2$, $\langle$Fe$\rangle$ and H$\beta$ as a function of $\sigma$. If, as she claims, the scatter is larger than the instrumental scatter, it would mean that $\langle$Fe$\rangle$ cannot be a simple function of metallicity, like Mg$_2$, but that there has to be a second parameter, most likely the Mg/Fe abundance ratio. This parameter cannot be the age, because the galaxies in J\o rgensen (1997) are early-type galaxies in clusters, which fall well onto their CMR. Her interpretation however might not be correct. Recently, Kuntschner (1998) published some high-quality data of the Fornax cluster (see also Kuntschner \& Davies 1998). As a comparison we have plotted them in Fig.~1b. He showed that for $\log~\sigma_0$ $\ge$ 1.9 the scatter in Fe3', an index very similar to $\langle$Fe$\rangle$, and in H$\beta$ is smaller or comparable to the scatter in Mg$_2$. This is a very important result, implying that in Fornax $\langle$Fe$\rangle$, just like Mg$_2$, depends only on $\sigma$, not on a second parameter. Since Kuntschner's signal-to-noise is much larger than J\o rgensen's, it looks as if the observational scatter in J\o rgensen (1997) has been underestimated. If this is not the case, it would mean that in some galaxy clusters the stellar populations are affected by a second parameter, while in others (Fornax) this would not be the case. \begin{figure}[h] \begin{center} \mbox{\epsfxsize=13cm \epsfbox{peletier1.ps}} \caption{Line strength vs. $\sigma$ relations from J\o rgensen (1997, left) for 9 clusters and Kuntschner (1998, right) for the Fornax cluster. In the figure on the right open symbols denote S0 galaxies, while ellipticals are indicated by filled circles. Fe3' is a pseudo-magnitude index containing the Fe features at 4383, 5270 and 5335 \AA\ , and H$\beta$' is the H$\beta$ index in magnitudes.} \end{center} \end{figure} \section{Abundance Ratios in the Centres of Galaxies} \subsection{Mg and Fe} The largest dataset of nuclear line strengths in ellipticals and bulges has been the Lick sample, which was finally published by Trager et al. in 1998. It consists of measurements of almost 400 galaxies. The observations were performed with the Lick IDS, the same instrument with which the stars defining the Lick system have been observed. Worthey (1998) shows $\langle$Fe$\rangle$ vs. Mg$_2$ for a subset of this sample in his Figure 1. In this Figure no difference can be seen between bulges of S0s and ellipticals. Generally, objects with central velocity dispersion below about 200 km/s seem to have solar Mg/Fe ratios, while the others are over-abundant in Mg. Since the Lick IDS detector suffered from several instrumental problems, the errors in the individual measurements are larger than one would like. For that reason I have made a different compilation, with data from more recent papers with high quality line strength measurements, and shown that in Fig.~2. It includes a sample of spiral bulges (Jablonka, Martin \& Arimoto 1996) subdivided in a group of early-type bulges (type 0-2) and later type objects (type 3-5), a sample of bulges of S0 galaxies (Fisher et al. 1996), a mixed sample of ellipticals and S0 galaxies (Kuntschner 1998), a sample of mainly faint elliptical galaxies (Halliday 1998) and two samples of bright ellipticals. The models plotted here are from Vazdekis et al. (1996). The conclusion from Fig.~2 is again that bulges and ellipticals are indistinguishable. The models of Vazdekis et al. (1996) are compatible with those of Worthey (1994), and also for these models galaxies start deviating from their locus at Mg$_2$ $>$ 0.25. Bulges of Sa-Sc spirals, which were not included in the Lick sample, seem to have solar Mg/Fe ratios. Although there might be some small systematic offsets between the individual samples, they generally agree well with each other.. There might be a few objects, which do not follow the general trend. NGC~4458 and NGC~4464 (large filled dots, from Halliday 1998) have a very low Mg/Fe ratio, compared to other galaxies with the same Mg$_2$. Both galaxies are faint ellipticals; NGC~4458 rotates very slowly, and has a small kinematically decoupled core (Simien \& Prugniel 1998), while NGC~4464 has a v/$\sigma$ of about 0.5 for an ellipticity of 0.3, so (v/$\sigma$)$^*$ is close to 1 (Davies et al. 1983), so that it is probably an oblate, rotating elliptical. \begin{figure}[h] \begin{center} \mbox{\epsfxsize=10cm \epsfbox{peletier2.ps}} \caption{Literature compilation of Mg$_2$ vs. $\langle$Fe$\rangle$. Plotted are central values. Also added are models with solar abundance ratios of Vazdekis et al. (1996). The lines represent models of constant metallicity (resp. Z=0.008, 0.02 and 0.05 from left to right), along which age increases to the right. JMA = Jablonka et al. (19960; CD = Carollo, Danziger \& Buson (1993); FFI = Fisher et al. (1996), and DSP = Davies et al. (1993).} \end{center} \end{figure} The situation for our own Galactic Bulge is in agreement with external galaxies. It is estimated from high-resolution spectra that the mean [Fe/H] in Baade's window is somewhat lower than solar (--0.25, McWilliam \& Rich 1994), but that [Mg/Fe] there is 0.3 -- 0.4. For such a Mg-abundance, one would expect the Mg$_2$ index to be about 0.29, assuming an age of 17 Gyr, if one uses the Vazdekis et al. (1996) models, or less, if the bulge would be younger. Assuming that the bulge is indeed old, it will lie in Fig.~2 in the region where galaxies are over-abundant in Mg w.r.t. Fe. \subsection{Other elements} Worthey (1998) extensively summarises our knowledge about the relative abundance of metals other than Mg in giant ellipticals and in the Galactic Bulge. Although it appears that the situation concerning the elements Sc, V and Ti is very complicated and confusing, there are about half a dozen other elements for which we know something about their behaviour in giant elliptical galaxies. In Table~1 I have schematically summarised our knowledge about them. All the information has been obtained from various Lick indices, by comparing measured line strengths with the values that one expects based on stellar population models. As Worthey (1998) mentions, there are several differences in element abundances between giant ellipticals and our Bulge, which implies that there must have been differences in their formation processes. Especially in the case of N this is striking: while in the Bulge the abundance of CN is generally lower than solar, in giant ellipticals this is the opposite. It seems that [C/Fe] $\approx$ 0 in ellipticals (from the C$_2$4668 line) and in our Bulge (Worthey 1998), so that it is thought that N is depleted in our Galaxy and overabundant in giant ellipticals. Peculiar is also that [O/Fe] $\approx$ 0 (McWilliam \& Rich 1994), since one would expect that O, an $\alpha$-element, would follow Mg. McWilliam \& Rich (1994) however warn us that the only stars for which they could measure O are at the tip of the RGB, where their abundances might not be representative of the Bulge because of metal-enrichment in the star itself. More measurements of O-abundances would be very welcome. \begin{table}[h] \begin{center} \caption{Abundance ratios for some important elements in giant ellipticals, the Bulge, and the galactic halo} \begin{tabular}{l|ccc} \multicolumn{4}{c}{~} \\ \tableline ~ & \multicolumn{3}{c}{[X/Fe] in} \\ Element X & Giant Ell. & Gal. Bulge & Gal. Halo \\ \tableline C & Solar & Solar ? & Solar \\ N & $>$0 & $<$0 & $\le$0 \\ O & Unknown & Solar ? & $>$0 \\ Mg & $>$0 & $>$0 & $>$0 \\ Na & $>$0 & $>$0 & $<$0 \\ Ca & $\le$0 ? & Solar & $>$0 \\ \tableline \end{tabular} \end{center} \end{table} Here I would like to revisit our ideas about the [Ca/Fe] abundance ratio in giant ellipticals. The Lick system has two indices which can be used to measure the Ca abundances in galaxies: Ca 4227 and possibly Ca 4455. Both are faint, narrow features that are difficult to measure in giant ellipticals, because of the large correction for velocity broadening that one has to apply to measure them. Vazdekis et al. (1997), using high signal to noise spectra of three giant early-type galaxies, found that the measured Ca 4227 in all of them was much lower than expected from their and also Worthey (1994)'s stellar population model. Their Fig.~13 nicely illustrates how enormous the velocity dispersion correction is here. An independent confirmation of this result comes from observations of the NIR Ca II triplet (CaT) of the same 3 galaxies. We observed them using 2d-FIS, an Integral Field Spectrograph on the 4.2m WHT at La Palma, using a fibre bundle to feed the light from the Cassegrain focus to the slit of the ISIS double spectrograph (Peletier et al. 1999). Details about the instrument are given in Arribas, Mediavilla, \& Rasilla (1991). In Fig.~3 our central measurements are plotted. Using the system of D\'\i az, Terlevich \& Terlevich (1989) to define the band-passes, it was found that the CaT equivalent width is less strong than predicted by the models of Vazdekis et al. (1996) with solar Ca/Fe ratios (see also Fig.~3). Although those models were based on the stellar library of D\'\i az et al. (1989), which doesn't fully cover the range of metallicities of giant ellipticals, the main conclusions will not change. Also, the same result is obtained if the models of Garc\'\i a-Vargas, Moll\'a \& Bressan (1998) are used. Our observations of the CaT however are in good agreement with Terlevich, D\'\i az \& Terlevich (1990). How to explain this apparent {\sl under-abundance} of Ca? It is possibly that the effect is entirely caused by a problem in calculating the models. Idiart, Thevenin \& de Freitas Pacheco (1997) and also Borges et al. (1995) point out that most models calculate integrated spectra by summing linear combinations of observed spectra of standard stars, assuming that [Ca/Fe] = 0 for those standard stars. Idiart et al. however obtained a separate library of standard stars, determined Ca, Mg and Fe abundances for each star, and calculated integrated indices using those individual abundances. Applying their models (Fig.~3) they find that [Ca/Fe] in the three galaxies is solar. This is however still peculiar, since Ca is an $\alpha$-element, which properties should follow closely those of Mg. Since the library of Idiart et al. (1997) also is rather limited, it is very important to obtain a stellar library of the size of the Lick library in the region of the CaT, to be able to interpret this important line index, which can also be used to constrain the IMF in galaxies. Together with the group at the Universidad Complutense in Madrid we are currently working on providing such a library (Cenarro et al. 1999, in preparation). \begin{figure}[h] \begin{center} \mbox{\epsfxsize=13cm \epsfbox{peletier3.ps}} \caption{Models of the Ca T compared with the observations in the central aperture of 1.7$''$ (from Peletier et al. 1999). The drawn lines are models of Vazdekis et al. (1996), with metallicities of Z=0.008, 0.02 and 0.05, and with age ranging from 1 to 17 Gyr from left to right. The dashed lines are the models of Idiart et al. (1997) and Borges et al. (1995). In the left figure the thick dashed lines indicate models with solar [Mg/Fe], while for the thin dashed line [Mg/Fe]=0.30.} \end{center} \end{figure} \section{Line Strength Gradients in Galaxies} Line strength gradients have been presented by various authors (e.g. Gonz\'alez 1993, Davies, Sadler \& Peletier 1993, Carollo, Danziger \& Buson 1993, Fisher, Franx \& Illingworth 1995, 1996, Vazdekis et al. 1997), but only for the Mg$_2$ and Mg$_b$ indices reasonably high quality measurements are available in the literature for a sufficiently large sample, and possibly also for $\langle$Fe$\rangle$ and H$\beta$. Gradients in elliptical galaxies in Fe and Mg indices are generally following tracks with constant [Mg/Fe], which means that for many galaxies the gradients are steeper than the line linking galaxy nuclei (e.g. Worthey et al. 1992, Davies et al. 1993). New, excellent quality data by Halliday (1998, Figure~4) confirm this also for low luminosity ellipticals. Current data seems to imply that Mg/Fe within all galaxies is constant. The behavior of Mg vs. Fe seems to be the same in bulges and disks. Fisher et al. (1996) find that in S0 galaxies the radial gradients in Mg$_2$ and $\langle$Fe$\rangle$ are smaller along the major axis than on the minor axis, implying that the gradients in the disk are smaller than those in the bulge. However, in the Mg$_2$ vs. $\langle$Fe$\rangle$ diagram the galaxies have the same slope, on both major and minor axis. \begin{figure} \begin{center} \mbox{\epsfxsize=13cm \epsfbox{peletier4.ps}} \caption{Measurements of Mg$_b$ and $\langle$Fe$\rangle$ for both the major and minor axes of the sample of Halliday (1998) are compared with the models of Worthey (1994). Major axis measurements are given by crossed symbols and minor axis by open squares. Central measurements for both axes are encircled. Typical errors for Mg$_b$ and $\langle$Fe$\rangle$ are given in the lower right-hand corner of each plot, the larger errors indicating errors for the outer parts of each galaxy and smaller errors, for the galaxy centre. The model grids consist of dashed lines connecting points of equal metallicity and solid lines, points of equal age; values of age are given in Gyr and metallicity, in units of [Fe/H]. A guide plot is given in lower right where lines corresponding to particular ages and metallicities are labelled.} \end{center} \end{figure} Information about abundance gradients in other elements is limited. Vazdekis et al. (1997) present gradients in about 20 lines for the three above mentioned galaxies. They find that the radial gradients can be explained well by gradients in the overall metallicity. Much more work however has to be done to investigate gradients in, e.g., bulges and smaller elliptical galaxies. Recently it has become possible to make reliable line strength maps in two dimensions, using Integral Field Spectrography. This offers an exciting range of new possibilities. For example, in galaxies with kinematically decoupled cores it can be investigated whether for example Mg/Fe in the decoupled core is different from the ratio in the rest of the galaxy, giving clues to the origin of the decoupled core. In Peletier et al. (1999) one galaxy with a decoupled core is included, the Sombrero galaxy, which has a rapidly rotating inner disk (e.g. Wagner, Bender \& Dettmar 1988). In Fig.~5 we show the Mg$_2$ and $\langle$Fe$\rangle$ maps in an inner field of 8.2$''$ $\times$ 11$''$. The continuum intensity is shown as well. The figure shows that Mg is enhanced in the inner disk, compared to the bulge, and that Fe is almost not enhanced (see also Emsellem et al. 1996). The noise however in the $\langle$Fe$\rangle$ image is still so large that it cannot be established whether the Mg/Fe ratio itself in the inner disk has gone up. This galaxy shows, together with other cases in the literature (e.g. NGC~4365, Surma \& Bender 1995, NGC~7626, Davies et al. 1993) that the central Mg abundance in galaxies with kinematically decoupled cores increases sharply, but up to now Mg/Fe seems to remain constant. \begin{figure}[h] \begin{center} \mbox{\epsfysize=8cm \epsfbox{peletier5.ps}} \caption{Absorption line maps of Mg$_2$ and $\langle$Fe$\rangle$ in the inner region of the Sombrero galaxy (field size is 8.2$''$ $\times$ 11$''$, from Peletier et al. 1999). On the left a reconstructed continuum map is shown. } \end{center} \end{figure} The inner disk also shows an increased H$\gamma$ line strength, while the CaT has a similar behaviour as $\langle$Fe$\rangle$ (Peletier et al. 1999). Although these data are still rather noisy for this purpose, it is expected that much better data will become available soon from the new, wide-field (33$''$ $\times$ 41$''$) integral field spectrograph SAURON (PIs R. Bacon, P.T. de Zeeuw and R.L. Davies) on the WHT. \section{What determines the Abundance Ratios in Galaxies?} From the previous sections we can draw the following conclusions: \begin{itemize} \item Small galaxies (Mg$_{2,c}$ $<$ 0.22, $\sigma_c$ $<$ 150 km/s) have solar Mg/Fe ratios, large galaxies (Mg$_{2,c}$ $>$ 0.28, $\sigma_c$ $>$ 225 km/s are overabundant in Mg. \item There is no difference in the Mg/Fe ratio between ellipticals and bulges of the same velocity dispersion (or Mg$_{2,c}$ value). \item Within a galaxy Mg/Fe appears to remain constant. \end{itemize} It is thought that the reason the Mg/Fe ratio from galaxy to galaxy varies, is that the ratio of the number of supernovae Type II vs. Type Ia can vary. SNe Type II occur in massive stars, and produce large amounts of light elements. SNe Type Ia come from binary accretion onto white dwarfs, and produce relatively much more Fe-peak elements (see e.g. Pagel 1998). Since the lifetime of the progenitors of SNe Type II is very short, there is a period of a few times 10$^8$ years from initial star formation in which enrichment through SNe Type II dominates (Worthey 1998). The most popular scenarios to vary this ratio of the two SNe types as a function of galaxy size (or velocity dispersion) are \begin{itemize} \item the formation time-scale, that should be shorter in large galaxies \item a variation in the IMF, such that large galaxies have a relatively larger fraction of massive stars. \end{itemize} Although it is difficult to reject the first scenario, there are various reasons why I would prefer the second. If the Mg/Fe ratio would be determined by the formation time-scale, then bulges and ellipticals would have had to form the majority of their stars on the same time-scale. Clearly this would predict that disks, in which the star formation is slow, would have lower Mg/Fe ratios. Although no good measurements of disks are available at present, the measurements by Fisher et al. (1996), which show no difference in Mg/Fe between major and minor axis in S0 galaxies, do not favour this scenario. Central disks (e.g. in the Sombrero) have high, rather than low Mg/Fe ratios. Secondly, it would be very difficult to form the brightest galaxies (with high Mg/Fe ratios) in an hierarchical way, on time-scales of Gyrs, since SNe Type Ia would lower the Mg/Fe of the gas, and make it very difficult to reach large Mg/Fe ratios. For example, the central disk of the Sombrero should have formed very fast from gas that was not very metal rich before the last merger event. The second option seems more favourable, although it also has its difficulties. If Mg/Fe would be a function just of one variable, e.g., velocity dispersion (or escape velocity, as suggested by Franx \& Illingworth 1990), then this scenario would not be able to explain the line strength gradients in galaxies, since it seems that for the same Mg abundance Mg/Fe in the outer parts of bright galaxies is generally lower than Mg/Fe in the centres of faint galaxies (gradients are steeper than the line connecting nuclei). To also be able to explain the gradients we will modify this scenario: the IMF, which mainly determines the enrichment of the elements, has to be dependent only on the mass of the galaxy as a whole. How realistic is this scenario? Although Elmegreen (these proceedings) claims that the IMF is generally universal, there are indications in some starburst galaxies that the IMF there is skewed towards high mass stars (Kennicutt 1998, p. 71). At present we can't confirm, nor rule out this second scenario. It is however capable of explaining the current observations rather easily, much better than the first scenario. The hot gas fraction, both in clusters of galaxies and in individual objects, might further constrain chemical enrichment models of galaxies. The measurements in the ISM of elliptical galaxies however indicate very low Fe-abundances (0.1 -- 0.4 times solar, Arimoto et al. 1997), which seems inconsistent with the measurements from stellar spectra. For that reason we have to wait until the abundances from X-ray emssion lines are better understood (see also Barnes, these proceedings). \section{The Need for Better Stellar Models} Since abundance ratios in galaxies can only be obtained through detailed comparisons with stellar models, it is crucial that the models are up to date. At present, there are several stellar population models that predict integrated line strength indices on the Lick system. The models do not differ too much from each other (Worthey 1994, Bruzual \& Charlot (see Leitherer et al. 1996), Vazdekis et al. 1996, Tantalo et al. 1996, Borges et al. 1995). Most of them use the stellar tracks of the Padova group (Bressan, Chiosi \& Fagotto 1994). None of them however takes into account the fact that if the abundance ratios in the stars are non-solar, the stellar parameters, like effective temperature and gravity, might be different. There have been some papers studying how stellar isochrones change as a function of the Mg/Fe ratio for solar or larger metallicities (Weiss, Peletier \& Matteucci 1995, Barbuy 1994). The conclusion of Weiss et al. is that the isochrones basically do {\sl not} change, if the total metallicity (i.e. the mass fraction of elements heavier than He) remains constant. This is consistent with Barbuy (1994). If this result is reliable, isochrones with non-solar abundance ratios are not necessary, as long as the abundance ratios are taken into account in the fitting functions, which are used to calculate the line indices. This result is in agreement with Salaris, Chieffi \& Straniero (1993) for lower metallicities. However, in a new paper Salaris \& Weiss (1998), using new opacities, find that for Z=0.01 (their largest metallicity) $\alpha$-enhancement, even while keeping the total metallicity constant, does change properties like the Main Sequence Turnoff and the RGB colour significantly. It is important to find out why this result is in contradiction with previous work. The results of Idiart et al. (1997), who showed that very different Ca-abundances are obtained for models of integrated stellar populations if the Ca-abundance of individual stars is taking into account in the fitting functions, indicate that accurate fitting functions are crucial. It is not expected that the models for $\langle$Fe$\rangle$ and Mg$_2$ will change much, since in Weiss et al. (1995) it is shown that if one takes solar neighbourhood stars to calculate the fitting functions for Mg$_2$ and $\langle$Fe$\rangle$, the integrated indices are very similar to the ones that one obtains when determining the fitting functions from Galactic Bulge stars. It is clear however, that the next generation of stellar population model will have to include fitting functions determined from as many stars as possible, using abundance ratios calculated for each star individually. Using 8m telescopes, it should be possible to obtain spectra of stars covering the parameter space of temperature, gravity, metallicity and some abundance ratios, that would be required to do this. \section{Summary} I discuss the evidence for abundance ratios in galaxies, supplementing the recent review by Worthey (1998). My main conclusions are: \begin{itemize} \item The scatter for early-type galaxies in the relations between $\langle$Fe$\rangle$ and H$\beta$ vs. velocity dispersion is small, comparable to the scatter in the Mg$_2$ and colour vs. $\sigma$ relation. \item Small galaxies (Mg$_{2,c}$ $<$ 0.22, $\sigma_c$ $<$ 150 km/s) have solar Mg/Fe ratios, large galaxies (Mg$_{2,c}$ $>$ 0.28, $\sigma_c$ $>$ 225 km/s are overabundant in Mg. There is no difference in the Mg/Fe ratio between ellipticals and bulges of the same velocity dispersion (or Mg$_{2,c}$ value). Within a galaxy Mg/Fe appears to remain constant. \item We know very little known about the behaviour of other elements (see Worthey 1998), although [Ca/Fe] in giant ellipticals appears to be solar, contrary to what one would expect from an $\alpha$-element. \item Improved stellar population models, using stellar evolutionary tracks with non-solar abundance ratios and fitting functions using abundance ratios determined for each standard star separately, would be very welcome to calculate accurate abundance ratios. \end{itemize} \acknowledgments I like to thank Claire Halliday, Alejandro Terlevich and Harald Kuntschner for communicating results in advance of publication, and John Beckman for organising an interesting meeting.
\section{Introduction} In this article we will report some observation that we have made in gauge invariant model defined on a lattice with an action which contains two parallel plaquettes \cite{sav}. This action is non-Abelian generalization of the $Z_2$ gauge spin system which was introduced as a lattice realization of random surfaces with gonihedric action \cite{sav1}. It is defined as a product of gauge group variables over two parallel plaquettes belonging to a given three-dimensional cube of the lattice \cite{sav} \begin{equation} H_{gonihedric} = \frac{n}{g^{2}} \sum_{\{ parallel~plaquettes \}} \{ 1- (1/n^2) Re~ Tr U_{plaq} \cdot Re ~Tr U_{plaq} \} , \label{linear} \end{equation} where $U_{plaq}$ is a product of gauge group elements $U_{ij}$ defined on a link $<i,j>$ over all sites of a plaquette and $U_{ij} = exp(i g A_{\mu} a)$. As one can see the action contains a product of two traces taken over two parallel plaquettes belonging to a 3D-cube. The summation is over all pairs of parallel plaquettes \footnote{The model with an action which is defined as a sum of contributions over all closed loops made up with six non-repeated links has been considered in \cite{carlo}. }. It is easy to see that for smooth classical fields the action reduces to the original action for non-Abelian fields as it takes place for the Wilson action \cite{wils,creutz} \begin{equation} H = \frac{2n}{g^{2}} \sum_{\{plaquettes \}} (1- (1/n) Re~Tr U_{plaq} ) . \label{area} \end{equation} Indeed the matrix product of four $U's$ around plaquette $\mu,\nu = 1,2$ becomes $$ Re~Tr U_{plaq} = Re~Tr~ exp (i g a^2 F_{12} +...) $$ and after expanding fields around the center $x_{\mu}$ of the 3D-cube we will get $$ 1 - (1/n^2)(n - \frac{g^2 a ^4 }{2} Tr F^{2}_{12}(x_{\mu} + \frac{a}{2}\delta_{\mu 3}))(n - \frac{g^2 a ^4 }{2} Tr F^{2}_{12}(x_{\mu} + \frac{a}{2} \delta_{\mu 3})) \approx \frac{g^2 a ^4}{n} Tr F^{2}_{12} $$ therefore it reproduces the classical action for pure Yang-Mills fields. Thus on a classical level both models (\ref{linear}) and (\ref{area}) are equivalent. But on a quantum level it is not obvious at all. The reason is that the vacuum structure of the model (\ref{linear}) is drastically different \cite{sav,weg}. In addition to the trivial vacuum configuration when all plaquette variables are equal to one \begin{equation} \frac{1}{2}Tr~U_{plaq} = + 1, \label{plus} \end{equation} there are $2^{dN}$ different vacuum configurations \cite{sav,weg} with frustrated plaquettes $\frac{1}{2}Tr~U_{plaq}=-1$. One of these vacuua can be easily described as a configuration with all plaquettes frustrated \begin{equation} \frac{1}{2}Tr~U_{plaq} = -1. \label{minus} \end{equation} It is obvious that it is impossible to connect vacuum configurations (\ref{plus}) and (\ref{minus}) by a gauge transformation. The riach vacuum structure inherited from the "parent" $Z_{2}$ gauge spin system with the Hamiltonian \cite{sav} \begin{equation} H_{gonihedric}= -\frac{1}{g^{2}} \sum_{\{parallel~plaquettes\}} (\sigma\sigma\sigma\sigma)~(\sigma\sigma\sigma\sigma) \label{parallel} \end{equation} In both models (\ref{parallel}) and (\ref{linear}) we have the same vacuum structure: towers of frustrated plaquettes describe different vacuua \cite{weg}. The difference between $SU(2)$ and $Z_2$ systems is that in the model (\ref{linear}) we have a continuous Lie gauge group and therefore continuous fluctuations of gauge fields around these vacuua. The problem of our main concern is the physics in these vacuua and its relevance to the continuum limit. It is also interesting to understand: what is the remnant of this riach vacuum structure in the continuum limit. Is there any dilaton-like field which describes different vacuua in the continuum limit or not? And if there is a continuum limit in all these vacuua, then what is the difference between them? In the next section we will present results of simple analytical consideration of the model and the results of the Monte Carlo simulations for the $SU(2)$ gauge group which are performed on three-dimensional lattices for an obvious simplification \cite{kostas,teper}. \section{High and low temperature expansion of average action} As usually, all observables can be constructed as a product of gauge variables over closed loops and we shall consider first the average action. For the Wilson action (\ref{area}) a simple one-plaquette average is equal to \cite{creutz} $$ P ~=~ <1-\frac{1}{2} Re~Tr~ U_{plaq}>~ =~ 1 - \frac{\beta}{4}+... ~~~when~~~~\beta \rightarrow 0 $$ and $$ P ~=~ \frac{3}{4\beta}+... ~~~~~~~when~~~~~~ \beta \rightarrow \infty. $$ Let us now consider two-plaquette action (\ref{linear}) which, as we shall see, has a different high temperature expansion. At high temperatures the average action is equal to: $$ PP~=~ <1-\frac{1}{4} Re~Tr ~U_{plaq}\cdot Re~Tr ~U_{plaq}> ~=~1 - \frac{\beta}{16} +... $$ and the low temperature expansion is the same as for the standard action $$ PP = \frac{3}{4\beta}+... $$ On Figure 1 one can see the behavior of average action in both models which we got by Monte Carlo simulations. In the simulations we used the Heat Bath algorithm \cite{kenpen}. The new action has indeed smaller values at high temperatures and the slope is four times smaller. \begin{figure} \centerline{\hbox{\psfig{figure=fig1.ps,height=8cm,angle=-90}}} \caption[fig1]{The behaviour of the average action in two different models. The new action has four times smaller values at high temperatures.} \label{fig1} \end{figure} One can see also crossover from high to low temperatures. In the thermal cycles we have seen small hysteresis loop, but the measurement of susceptibility demonstrates that the maximum does not scale with the volume ( see Figure 2,3). Thus we do not see any phase transition at $\beta \approx 3.5$. \begin{figure} \centerline{\hbox{\psfig{figure=fig2a.ps,height=8cm,angle=-90}}} \caption[fig2a]{The thermal cycle of the two-plaquette action.} \label{fig2a} \end{figure} \begin{figure} \centerline{\hbox{\psfig{figure=fig2b.ps,height=8cm,angle=-90}}} \caption[fig2b]{Susceptibility versus $\beta$ for two volumes $8^3$,$24^3$. The maximum of susceptibility does not scale with the volume.} \label{fig2b} \end{figure} \section{Loop averages and vacuum structure} In the model with a one-plaquette action (\ref{area}) the high temperature expansion of the loop averages is equal to \cite{creutz} $$ <W(C)>~ =~ W(I,J)~ =~ (\frac{\beta}{4})^{IJ},~~~~~~~~\beta \rightarrow 0, $$ and if one defines $W(I,J) \approx exp(-\sigma A)$, where $A$ is the area $IJ$ of the loop $C$ then the string tension is equal to $$ \sigma ~\approx ~\frac{1}{a^2} ln (\frac{4}{\beta}). $$ In the two-plaquette model (\ref{linear}) the loop averages demonstrate peculiar behaviour. Indeed at high temperature the loop average \begin{equation} <W(C)> = Z^{-} \cdot \int dU~ e^{-S}~~ \frac{1}{2}~ Tr_{C}~ U_{ij} \end{equation} is equal to zero for $SU(2)$ gauge group, because there are no low order lattice diagrams (proportional, let us say, to area) which can contribute to the expectation value of $W(C)$ (see Figures 4 and 5). We have $absolute$ confinement in this model. It is similar to the limit $n \rightarrow \infty$ of the model with one-plaquette action $$ \sigma = \frac{1}{a^2} ln \frac{2n^2}{\beta} \rightarrow \infty $$ and thus $<W(C)> =\approx exp(-\sigma A) \rightarrow 0$\footnote{The situation changes for large gauge groups where there exist high temperature diagrams which contribute to one-plaquette average. Specifically for the $SU(3)$ group the 3D-box diagram on a top of plaquette gives nonzero contribution.}. The Monte-Carlo simulations of $SU(2)$ group demonstrate that this behaviour takes place until $\beta \ \approx 3.5 - 4$ as one can see on Figures 4 and 5. At these temperatures there appears proliferation of large lattice diagrams which cover almost the whole lattice and $<W(C)>$ gets nonzero value. At low temperatures the nontrivial vacuum structure begins to play a central role. To see that let us consider the situation when two opposite plaquettes inside a 3D cube are frustrated. As it is easy to see from the action the energy of the new configuration will be the same. Indeed the energy is minimal when $\frac{1}{4}Tr U_{plaq}\cdot Tr U^{paral}_{plaq} =1$, therefore we have two possibilities ether $$ \frac{1}{2}Tr~ U_{plaq} = \frac{1}{2}Tr~ U^{paral}_{plaq} =+1 $$ or $$ \frac{1}{2}Tr~ U_{plaq} = \frac{1}{2}Tr~ U^{paral}_{plaq} =-1. $$ On the whole lattice one can build towers of frustrated plaquettes so that the new configuration will have the same energy as the trivial vacuum $\frac{1}{2}Tr~ U_{plaq} = +1$. Thus the vacuum has global degeneracy as it was in the corresponding $Z_2$ gauge spin system \cite{sav}. This actually means that we have many vacuua in which the gauge field $A_{\mu}$ is very far from perturbative value. \begin{figure} \centerline{\hbox{\psfig{figure=fig3a.eps,height=8cm,angle=00 }}} \caption[fig3a]{Schematic multivacuum structure in the model with two-plaquette action. This figure shows the absolute confinement at high temperatures $\beta \leq 3.5$ and gauge nonequivalent vacuua at low temperatures.} \label{fig3a} \end{figure} \begin{figure} \centerline{\hbox{\psfig{figure=fig3b.ps,height=8cm,angle=-90}}} \caption[fig3b]{ The behaviour of the mean value for the $1 \times 1$ and $2 \times 2$ loops for two-plaquette action. It represents the evolution of the averages starting from low temperature with $\frac{1}{2}Tr~U_{plaq} = +1$ to high temperature and then back to low one. As one can see the system fluctuates at pseudocritical point and then falls into vacuum with $Tr~U_{plaq} \approx 0$ } \label{fig3b} \end{figure} This complicated vacuum structure is reflected on Figures 4 and 5 which show multivacuum structure at low temperature. The top curve corresponds to the vacuum in which all plaquettes are equal to plus one, $\frac{1}{2}Tr ~U_{plaq}=1$, and the bottom curve corresponds to the opposite case when all plaquettes are frustrated $\frac{1}{2}Tr~U_{plaq}=-1$. It is also not difficult to construct explicitly vacuum configurations in which only part of the lattice plaquettes are frustrated. The vacuum with all plaquettes frustrated, $\frac{1}{2}Tr~U_{plaq}=-1$, corresponds to a gauge field which is far from the perturbative value $A_{\mu}=0$ in the whole lattice. Two natural questions arize:\\ i) is there a sensible continuum limit in all these vacuua in the limit $a \rightarrow 0, \beta \rightarrow \infty$ ?\\ ii) and if there is a continuum limit, then what is difference between them ? What type of parameter ("coupling constant") or field ("dilaton") may describe them? \section{String tension in different vacuua} In this paragraph we will study the behaviour of string tension in the limit $\beta \rightarrow \infty$. We choose this quantity because it can characterize the system in the continuum limit if it exists. As usually we define the string tension by the behaviour of W(L,L) for large L as: $$ W(L,L) \approx e^{-(\sigma L^2 + m L + Const)} $$ To calculate string tension we shall follow the procedure: \\ i) calculate W(L,L) for different values of $L \leq 6$ at a given $\beta $ (in our simulations the lattice volume is $16^3$ and the coupling constant is in the range $6 < \beta < 10$), \\ ii)then fit the quantity $-ln \vert W(L,L)\vert $ using second order polynomials and extract $\sigma $ for each $\beta $, \\ iii)because string tension scales as $\sigma = \sigma_{phys} a^2$ and $\beta = 2/g^2 a$ we shall plot the ratio $\beta \sqrt{\sigma} = 2 \sqrt{\sigma_{phys}}/ g^{2}$ versus $1/\beta$ (see Figure 6 ). \\ The last ratio converges to a finite limit when $\beta \rightarrow \infty$. We expect that this limit can be reached within the finite size corrections which are of order $O(\frac{1}{\beta})$. On Figure 6 we also depicted string tension in the vacuum $<\frac{1}{2}Tr~U_{plaq}> = +1$ calculated for the larger lattice $24^3$ in order to see the variation with volume. The convergence is better for small $\beta$. Until now we have been calculating string tension in the vacuum $<\frac{1}{2}Tr~U_{plaq}> = +1$ when $\beta \rightarrow \infty$. In the new vacuua we have frustrated plaquettes and the quantity $\frac{1}{2}Tr~U_{plaq}$~ is less than one. It remains constant during Monte Carlo simulations for large $\beta \geq 6$ and we conclude that these vacuua are well separated \footnote{When we start from a given vacuum configuration $<\frac{1}{2}Tr~U_{plaq}> \neq +1$ it never drifts into other vacuua when $\beta \geq 6$.}. To explore vacuum structure we have to calculate $W(L,L)$ for these vacuua as well. A typical result is shown in Table 1, where we choose $\beta = 10$, the volume is equal to $16^3$ and L is smaller than six. By "vacuum $-1/3$" we mean the vacuum which has the mean value $ <\frac{1}{2} Tr~U_{plaq}> = -1/3$. We observe that the absolute value $\vert W(L,L) \vert $ is the same for two vacuua +1 and -1, as well for the vacuua +1/3 and -1/3. The absolute value $\vert W(L,L) \vert $ for odd L in the vacuua $+1/3$ or $-1/3$ is one third of its value in the vacuua $+1$ or $- 1$. Because of this relation, $\vert {W_{1}(L,L)} \vert = \frac {1}{3} \vert {W_{1/3}(L,L)} \vert$ we have \footnote{By $W_{-1} (L,L)$ we mean the mean value of W(L,L) in the vacuum $Tr~U_{plaq}= -1$ for example.} $$ -ln\vert W_{1/3} (L,L) \vert = -ln \vert W_{1} (L,L) \vert + ln(3) =\sigma L^2 + m L + ln(3) + Const $$ and the string tension is the same for all these vacuua. Thus we can argue that string tension does indeed scale in all these vacuua and we can speculate that one can distinguish them by the scalar operator $\frac{1}{2}<Tr~U_{plaq}>$-gluon like condensate \cite{gs}. One of us (K.F.) wishes to thank the EU for partial financial support (TMR project FMRX-CT97-0122). \begin{figure} \centerline{\hbox{\psfig{figure=fig4.ps,height=8cm,angle=-90}}} \caption[fig4]{String tension versus inverse $\beta$ in the standard vacuum $Tr~U_{plaq} = +1$.} \label{fig4} \end{figure} \begin{center} \begin{tabular}{c} \hline \hline TABLE 1\\ \hline \hline \\The values of $<W(L,L)>$ in different vacuua for $\beta =10$ and volume $16^3$ :\\ \end{tabular} \end{center} \begin{tabular}{||c|c|c|c|c|c|c||} \hline \hline L&vacuum +1&vacuum -1& vacuum +1/3&vacuum- 1/3\\ \hline \hline 1&0.946281(4)&-0.946277(3)&0.315429(3)&-0.315427(4) \\ \hline 2&0.83866(2)&0.83863(2)&0.83865(1)&0.83864(2) \\ \hline 3&0.71894(4)&-0.71889(4)&0.23968(3)&-0.23967(4) \\ \hline 4&0.60273(7)&0.60265(8)&0.60269(8)&0.60265(9) \\ \hline 5&0.49675(13)&-0.49656(12)&0.16561(9)&-0.16567(10) \\ \hline 6&0.40381(19)&0.40352(20)&0.40361(19)&0.40369(21) \\ \hline \hline \end{tabular}
\section*{Introduction} Recently, statistical mechanical approaches to the problems of information science attracted much attention of the researchers who are working in the field of information science. Among these, a particular interest has been given to the techniques by which one tries to reconstruct an image from its corrupted version, {\it e.g.} sent by a defective fax, a fickle {\it e}-mail {\it et cetera}, since any data transmission through a channel is in principle affected by some kind of noise. In the mathematical engineering fields, the traditional way to obtain the optimal recovered image has been regarded as a sort of optimization problems. In this framework, one first constructs the energy (cost) function so that this function represents the distance between the original image and the recovered one as properly as possible; then, one minimizes it using suitable heuristic methods like {\it simulated annealing} \cite{Kirkpatrick83}. In fact, Geman and Geman \cite{Geman84} succeeded in constructing a method of image restoration using simulated annealing, and they discussed in detail the properties of its convergence including the optimal annealing schedule. Successful results in this direction have been reached by means of the usual techniques of disordered spin systems, assuming that each spin is naturally associated to a pixel or bit. In the language of the disordered spin systems, the optimization problems we just mentioned are naturally translated into the search of the ground state a system possessing many local minima of order ${\exp}(N)$. In contrast, Marroquin {\it et al.}\cite{Marroquin87} found that the temperature of the system plays an important role for the image recovering process. From the statistical mechanical point of view, each recovered image can be regarded as the equilibrium state of a random spin systems. Marroquin {\it et al.}\cite{Marroquin87} investigated the effect of the temperature on the quality of image restoration by computer simulation and found the optimality of finite temperature image restoration. Recently, this finite temperature effects on image restoration was checked in a more careful way by Pryce and Bruce\cite{PB}, although these works were restricted to numerical simulations. In the context of the convolutional error-correcting codes, Ruj$\acute{\rm a}$n\cite{Rujan} proposed the finite temperature decoding in which we regard the sign of the local magnetization at a specific temperature (this temperature is well known as the {\it Nishimori temperature}\cite{Nishi} in the field of spin-glasses) as the correct bit. Recently Nishimori and Wong\cite{NW} pointed out that the optimal restoration of an image is also obtained at some specific temperature and showed that image restoration(IR) and error-correcting codes theory (ECC) can be treated within a single framework. Indeed, to the usual IR Hamiltonian, ferromagnetic and random field terms, they added a spin-glass term borrowed from the ECC theory\cite{Sourlas} used for {\it parity-check}. They could exactly solve the infinite range spin model and find the optimal values of the temperature and the field (referred to as {\it hyper-parameters} from now on) at which the best retrieval quality is achieved. However, their works are restricted to the case of Ising spin systems and in this sense they are able to restore black/white pictures. On the other hand, there remain many open questions about the restoration of multi-color images or, somehow equivalently, gray toned images. This kind of problem has been also widely studied in the context of neural network with multi-state neurons, able to store and retrieve gray-scaled patterns (see \cite{Bolle} and references therein). For our purposes, we therefore map the set of the pixels onto $q$-state (chiral) Potts spins, with a ferromagnetic Hamiltonian in the presence of a random field (conventional IR) and, further on, a glass term (ECC-{\it like} term). The choice of the chiral Potts Hamiltonian is motivated by the fact that it exhibits the same gauge invariance as the Ising glass, although a work for the usual random Potts model is under consideration. Here, we show that, as in the Ising case\cite{NW}, the presence of the glass term significantly increases the quality of the reconstructed image. We should mention that several remarkable studies about IR using the Potts model have been made by several authors. However, their works mostly depend on the computer simulations. In addition, their methods (mean field annealing\cite{Zhang}, cluster algorithm \cite{Tanaka96,Tanaka97}, {\it etc.}) are devoted to the restorations at zero temperature. Therefore, it seems that there exist a lot of open questions about IR using the Potts model, especially, about the performance of the finite temperature restoration. In the next section, we will introduce our model within the image restoration theory and adopt the overlap as a measure of the restoration quality. In Sec. III, we will discuss the infinite range model and give the exact expression for the overlap as a function of the temperature and the external field, thus obtaining a relation between the temperature of source image and that of restoration temperature. We shall also see the improvement of the restoration quality by adding the glassy term. Finally, in Sec. IV, guided by the infinite range results, we will give explicit and realistic examples of image reconstructions for three and eight gray-scale picture. \section{The model and IR formulation \label{model} } As already mentioned in Introduction, we choose to represent pixels of a gray-scaled image by means of $q$-component Potts spin variables. The usual Potts Hamiltonian $H=-\sum \delta_{\sigma_i\sigma_j}$ admits a complex representation \cite{Stephen} by means of the following identity \begin{equation} \delta_{\sigma_i\sigma_j}=\frac{1}{q}\sum_{r=0}^{q-1} (\sigma_i)^r (\sigma_j)^{q-r} \end{equation} where each spin takes on one of the $q$ roots of unity \begin{eqnarray} {\sigma}_{i}={\exp} \left( \frac{2 \pi i}{q}K_{i} \right)\,\,\,\,\,\,\, \mbox{($K_{i}=0,{\cdots},q-1$)}. \label{potts-spin} \end{eqnarray} From now on, we will use the notation $\{\xi\}$ for the original pixels and $\{\sigma\}$ for the variables of the recovering process. Let us now send the original image through a noise channel not only by the form of ${\xi_i^r}$ itself but also by the following products ${\xi}_{i}^r{\xi}_{j}^{r*}={\xi}_{i}^r{\xi}_{j}^{q-r}$. Without loss of generality we raised the spins and their products to some power $r$, since this corresponds only to a rotation in the complex circle. The reasons of this choice will be clear soon. For this expression, the output $(\{\tau^{(r)}\},\{J^{(r)}\})$ is stochastically determined by the channel. For instance, in the case of a Gaussian channel (GC) the output function $P_{\rm out}(\{J^{(r)}\},\{\tau^{(r)}\}|\{\xi\})$ is given by \begin{eqnarray} P_{\rm out}(\{J^{(r)}\},\{\tau^{(r)}\}|\{\xi\}) & = & \frac{1}{(2\pi J)^{N_{B}/2}}\frac{1}{(2\pi \tau)^{N/2}} \nonumber \\ \mbox{} & \times & {\exp}\left[ -\frac{1}{2J^{2}} \sum_{(ij)}\sum_{r=0}^{q-1} (J_{ij}^{(r)}-J_{0}{\xi}_{i}^r{\xi}_{j}^{q-r}) (J_{ij}^{(r)*}-J_{0}{\xi}_{i}^{q-r}{\xi}_{j}^{r}) \right] \nonumber \\ \mbox{} & \times & {\exp}\left[ -\frac{1}{2\tau^{2}} \sum_{i}\sum_{r=0}^{q-1} ({\tau}_{i}^{(r)}-{\tau}_{0}{\xi}_{i}^{r}) ({\tau}_{i}^{(r)*}-{\tau}_{0}{\xi}_{i}^{q-r}) \right], \label{gc} \end{eqnarray} where $J_{ij}^{(r)}$ and $\tau_i^{(r)}$ are complex numbers which satisfy \begin{equation} \left(J_{ij}^{(r)}\right)^*\,=\,J_{ij}^{(q-r)} \,\,\,\,\, \left(\tau_{i}^{(r)}\right)^*\,=\,\tau_{i}^{(q-r)} \label{condition} \end{equation} in order to insure the realness of the sums in (\ref{gc}). Obviously, if a noise-free transmission could be achieved, we would obtain ${\tau}_{i}^{(r)}={\xi}_{i}^r$ and $J_{ij}^{(r)}={\xi}_{i}^{q-r}{\xi}_{j}^{r}$. The conditional probability $P(\{\sigma\}|\{J^{(r)}\},\{\tau^{(r)}\})$, which is probability that the source sequence is $\{\sigma\}$ provided that the outputs are $\{J\}$ and $\{\tau\}$, according to the Bayes theorem reads \begin{eqnarray} P(\{\sigma\}|\{J^{(r)}\},\{\tau^{(r)}\})\,{\sim}\, {\exp}\left( \frac{{\beta}_{J}}{q}\sum_{(ij)}\sum_{r=1}^{q-1} J_{ij}^{(r)}{\sigma}_{i}^{(r)}{\sigma}_{j}^{(q-r)} + \frac{h}{q} \sum_{i}\sum_{r=1}^{q-1}{\tau}_{i}^{(r)}{\sigma}_{j}^{(q-r)} \right)P_{d}(\sigma) \label{condprob} \end{eqnarray} where $P_{d}(\sigma)$ is a model of the prior distribution $P_{s}(\xi)$, that is, \begin{eqnarray} P_{d}(\sigma)\,{\equiv}\, {\exp}\left( \frac{\beta_{d}}{q} \sum_{(ij)}\sum_{r=1}^{q-1} {\sigma}_{i}^{(r)}{\sigma}_{j}^{(q-r)} \right). \label{modelP} \end{eqnarray} Our choice of the above prior distribution (\ref{modelP}) is due to the assumption that in real world, images should be locally smooth. From this point of view, the distribution (\ref{modelP}) is suitable because it gives high probability if the nearest neighboring sites take a same value. For the Ising model, in order to get the restored pixels out of the average quantities, the pixel at site $i$ (to be denoted as $\Sigma_i$) is naturally taken as the sign of the local magnetization. This means that the restored pixel is chosen as $\Sigma=+1$ ($\Sigma=-1$) if the spin points upward (downward) on average at the equilibrium. For our model, instead, since the value of the local magnetization is not simply confined to the interval $[-1,1]$ but it runs all over the complex circle, we introduce the following generalized restored variable \begin{eqnarray} \Sigma_{i} \left(\langle \sigma_i \rangle \right) \,=\,\exp \left[ i\sum_{\alpha=0}^{q-1} \frac{2\pi}{q}\,{\alpha}\,\Xi_{\alpha}({\theta}_{i}) \right] \label{newsigma} \end{eqnarray} with \begin{eqnarray} {\Xi}_{\alpha}(x)= {\Theta}\left( x-\frac{2\pi}{q}{\alpha} +\frac{\pi}{q} \right) - {\Theta}\left( x-\frac{2\pi}{q}{\alpha} -\frac{\pi}{q} \right), \label{defTheta} \end{eqnarray} $\Theta$ being the usual step function, and \begin{eqnarray} {\theta}_{i}={\tan}^{-1} \left( \frac{\langle {\rm Re}\,[{\sigma}_{i}]\rangle} {\langle {\rm Im}\,[{\sigma}_{i}]\rangle} \right). \label{angle} \end{eqnarray} In simpler words, $\Sigma_i$ is the closest spin on the circle to the value of the local magnetization $\langle \sigma_i \rangle \equiv \langle {\rm Re} \,[\sigma_i] \rangle +i \,\langle {\rm Im}\, [\sigma_i] \rangle$. For $q=2$, it is straightforward to check that (\ref{newsigma}) reduces to a sign function up to a normalization constant. The quantities $\langle {\rm Re}\, [{\sigma}_{i}] \rangle$ and $\langle {\rm Im}\, [{\sigma}_{i}] \rangle$ are the average over the Boltzmann distribution ${\rm e}^{-{\cal H}_{\rm eff}}$ with the following effective Hamiltonian \cite{NS} \begin{eqnarray} {\cal H}_{\rm eff} -\frac{\beta_{J}}{q} \sum_{(ij)}\sum_{r=1}^{q-1} J_{ij}^{(r)}({\sigma}_{i})^{r}({\sigma}_{j})^{q-r} -\frac{\beta_{d}}{q}\sum_{(ij)}\sum_{r=1}^{q-1} ({\sigma}_{i})^{r}({\sigma}_{j})^{q-r} -h\sum_{i}\sum_{r=1}^{q-1} {\tau}_{i}^{(r)} {\sigma}_{j}^{q-r}. \label{effHam} \end{eqnarray} The condition (\ref{condition}) gives the above Hamiltonian the same spin gauge-symmetry as Ising spin glass, thus suppressing the spontaneous magnetization at low temperature which is present in the usual random Potts model. For the restoration purposes, the random field term aligns the spins according to the corrupted picture, whereas the ferromagnetic term insures the smoothness, by suppressing the isolated pixels within one small cluster. Therefore, a balance between ${\beta}_{d}$ and $h$) will helps us to reconstruct the original picture well. The first term, instead, has been recently introduced in the problem of image restoration by Nishimori and Wong \cite{NW} and this term has been well known as the {\it parity check codes} in the field of error-correcting codes. Obviously, this term carries much more information about the original picture than the other two terms. Therefore, the performance of the image recovery is expected to be improved by this term. As a measure of the restoration quality, we shall adopt the following overlap $M$ \begin{eqnarray} M & = & \left[ \frac{1}{q}\sum_{r=0}^{q-1} \xi_i^{q-r} \Sigma_{i}^r \right]_{ \{ \xi, J, \tau \} } \,\equiv\, \frac{1}{q}\sum_{r=0}^{q-1} \sum_{\xi}\sum_{J}\sum_{\tau} P_{\rm out}(\{J^{(r)}\},\{\tau^{(r)}\}|\{\xi\})P({\xi}) {\xi}_{i}^{q-r} {\Sigma}_{i}^{r} \label{overlap} \end{eqnarray} in which $(1/q)\sum_{r=0}^{q-1}\xi_{i}^{q-r}{\Sigma}_{i}^{r}$ at each single site gives $1$ if the original spin is in the same state as the restored one, and zero otherwise. Here the dependence on the local magnetization is buried in the angle $\theta_i$, give by Eq. (\ref{angle}), and the sum over all the sites is understood. The main goal of this paper is to maximize the overlap $M$ as a function of the temperatures ($\beta_J$ and $\beta_d$) and the external field $h$ (referred to as an estimate of the {\it hyper-parameters}). In the next section, we will start with an exactly solvable model, that is, an infinite range version of the Potts spin-glass. \section{Mean field solution \label{meanfield}} We will now investigate the performance of our model within the mean field approximation, {\it viz.} each spin is influenced by all the others. As the source image, we will consider a ferromagnetic state generated by a Boltzmann distribution at some finite temperature $T_s$. For the sake of simplicity, we will restrict ourselves to the case of $q=3$, although the results can be generalized to any value of $q$. We thus assume that the original set of pixels $\{\xi\}$ is generated by a ferromagnetic 3-state Potts Hamiltonian with probability \begin{eqnarray} P({\xi})=\frac{1}{{\cal Z}_{s}(\beta_{s})} \,{\exp}\left[ \frac{\beta_{s}}{2N} \sum_{i<j} \left( {\xi}_{i}{\xi}_{j}^{*}+{\xi}_{i}^{*}{\xi}_{j} \right) \right] \end{eqnarray} where ${\cal Z}_{s}(\beta_{s})$ is a normalization constant and $\beta_s$ is the inverse source temperature. According to the conditional probability, the observable are computed as \begin{eqnarray} \left[\langle f \rangle\right]_{ \{\xi, J, \tau \}} \,=\, \sum_{\xi}\sum_{J}\sum_{\tau} P(\{J^{(r)}\},\{\tau^{(r)}\}|\{\xi\})P(\xi) \frac{{\rm Tr}_{\sigma} f\,{\rm e}^{-{\cal H}_{\rm eff}}}{\cal Z} \end{eqnarray} with \begin{eqnarray} {\cal Z}\,{\equiv}\,{\rm Tr}_\sigma{\exp}\left(-{\cal H}_{\rm eff}\right). \end{eqnarray} It is rather straightforward to average out the disorder by means of the well-known replica trick \cite{SK} and, assuming replica symmetry ansatz and isotropy (no dependence on $r$), the saddle point equations for the order parameters are given by \begin{eqnarray} [\langle{\sigma}_{i}^{r}\rangle]{\equiv}m & = & \frac{1}{{\cal Z}_{s}} \sum_{\xi}{\rm e}^{{\beta}_{s}(m_{s}^{(1)}{\rm Re}[\xi] +m_{s}^{(2)}{\rm Im}[\xi])}\int\frac{du}{\sqrt{\pi}}\frac{dv}{\sqrt{\pi}} {\rm e}^{-u^{2}-v^{2}}\frac{Z_{\rm cos}(\xi)}{Z(\xi)} \label{sigma} \\ \mbox{}[{\rm Re}[{\xi}_{i}]\langle{\sigma}_{i}^{r}\rangle]{\equiv}t_{1} & = & \frac{1}{{\cal Z}_{s}} \sum_{\xi}{\rm Re}[\xi]{\rm e}^{{\beta}_{s}(m_{s}^{(1)}{\rm Re}[\xi] +m_{s}^{(2)}{\rm Im}[\xi])}\int\frac{du}{\sqrt{\pi}}\frac{dv}{\sqrt{\pi}} {\rm e}^{-u^{2}-v^{2}}\frac{Z_{\rm cos}(\xi)}{Z(\xi)} \\ \mbox{}[{\rm Im}[{\xi}_{i}]\langle{\sigma}_{i}^{r}\rangle]{\equiv}t_{2} & = & \frac{1}{{\cal Z}_{s}} \sum_{\xi}{\rm Im}[\xi]{\rm e}^{{\beta}_{s}(m_{s}^{(1)}{\rm Re}[\xi] +m_{s}^{(2)}{\rm Im}[\xi])}\int\frac{du}{\sqrt{\pi}}\frac{dv}{\sqrt{\pi}} {\rm e}^{-u^{2}-v^{2}}\frac{Z_{\rm cos}(\xi)}{Z(\xi)} \\ \mbox{}[\langle{\sigma}_{i}^{r}\rangle \langle {\sigma}_{j}^{q-r} \rangle]{\equiv}Q & = & \frac{1}{{\cal Z}_{s}} \sum_{\xi}{\rm e}^{{\beta}_{s}(m_{s}^{(1)}{\rm Re}[\xi] +m_{s}^{(2)}{\rm Im}[\xi])} \nonumber \\ \mbox{} & \times & \int\frac{du}{\sqrt{\pi}}\frac{dv}{\sqrt{\pi}} {\rm e}^{-u^{2}-v^{2}} \frac{1}{Z^{2}(\xi)}[Z_{\rm cos}^{2}(\xi)+Z_{\rm sin}^{2}(\xi)] \label{Q} \end{eqnarray} Here $m_{s}^{(1)}$ and $m_{s}^{(2)}$ are simply the real and imaginary components of the source magnetization, {\it viz} the usual non-random Potts model \cite{Stephen} mean field equations \begin{eqnarray} [{\rm Re}[{\xi}_{i}]]\,{\equiv}\, m_{s}^{(1)} & = & \frac{1}{{\cal Z}_{s}} \left( {\rm e}^{\beta_{s}m_{s}^{(1)}} -{\rm e}^{-\frac{1}{2}\beta_{s}m_{s}^{(1)}}{\cosh} \left[ (\sqrt{3}/2) {\beta}_{s}m_{s}^{(2)} \right] \right) \\ \mbox{}[ {\rm Im} [\xi_{i}] ] \,{\equiv}\, m_{s}^{(2)} & = & \frac{1}{ {\cal Z}_{s}} \sqrt{3}\,{\rm e}^{-\frac{1}{2}\beta_{s}m_{s}^{(1)}} {\sinh}\left[ (\sqrt{3}/2) {\beta}_{s}m_{s}^{(2)} \right] \end{eqnarray} and \setcounter{saveeqn}{\value{equation} \begin{eqnarray} {\cal Z}_{s} & \,=\, & {\rm e}^{{\beta}_{s}m_{s}^{(1)}} +2{\rm e}^{-\frac{1}{2}{\beta}_{s}m_{s}^{(1)}} {\cosh}\left[ (\sqrt{3}/{2})\, {\beta}_{s}m_{s}^{(2)} \right] \\ Z_{\rm cos}(\xi) & \,=\, & {\rm e}^{U(\xi)}- {\rm e}^{-\frac{1}{2}U(\xi)}{\cosh}V \\ Z_{\rm sin}(\xi) & \,=\, & \sqrt{3}\,{\rm e}^{-\frac{1}{2}U(\xi)} {\sinh}V \\ Z(\xi)& \,=\, & {\rm e}^{U(\xi)} +2{\rm e}^{-\frac{1}{2}U(\xi)}{\cosh}V \end{eqnarray} \setcounter{equation}{\value{saveeqn} with \setcounter{saveeqn}{\value{equation} \begin{eqnarray} U & = & u\left[ \frac{\beta_{J}^{2}J^{2}}{q^{2}}Q +{\tau}^{2}h^{2} \right]^{\frac{1}{2}} +\frac{\beta_{d}}{q}m +\frac{\beta_{J}J_{0}}{q} \left[ t_{1}{\rm Re}[\xi] +t_{2}{\rm Im}[\xi] \right] +{\tau}_{0}h{\rm Re}[\xi] \\ V & = & \frac{\sqrt{3}}{2}\frac{\beta_{J}J}{q} Q^{\frac{1}{2}}v. \end{eqnarray} \setcounter{equation}{\value{saveeqn} Finally, the overlap $M$ is expressed as the following weighted average \begin{equation} M \,=\, \frac{1}{{\cal Z}_{s}} \sum_{\xi}{\rm e}^{{\beta}_{s}(m_{s}^{(1)}{\rm Re}[\xi] +m_{s}^{(2)}{\rm Im}[\xi])} \int_{{\cal S}(\xi)}\frac{du}{\sqrt{\pi}}\frac{dv}{\sqrt{\pi}} {\rm e}^{-u^{2}-v^{2}} \label{OVERLAP} \end{equation} receiving contributions from the following $q=3$ regions in the complex circle \setcounter{saveeqn}{\value{equation} \begin{eqnarray} {\cal S}(1) & = & \left\{ u,v\,{\Biggr |}\, -\frac{\pi}{3}\,{\leq}\,{\tan}^{-1}\frac{Z_{\rm sin}}{Z_{\rm cos}}\,{\leq}\, \frac{\pi}{3} \,\bigcap \, Z_{\rm cos}>0 \right\} \\ {\cal S}({\rm e}^{\frac{2\pi i}{3}}) & = & {\Biggr \{} u,v\,{\Biggr |}\, \left( \frac{\pi}{3}\,{\leq}\,{\tan}^{-1}\frac{Z_{\rm sin}}{Z_{\rm cos}}\, {\leq}\, \frac{\pi}{2} \,\bigcap \, Z_{\rm cos} \,{\geq}\, 0, Z_{\rm sin} > 0 \right) \nonumber \\ \mbox{} & \mbox{} & \bigcup \left(-\frac{\pi}{2}\,{\leq}\,{\tan}^{-1} \frac{Z_{\rm sin}}{Z_{\rm cos}}\,{\leq}\, 0 \,\bigcap\, Z_{\rm cos}\,{\leq}\, 0, Z_{\rm sin} \,{\geq}\, 0 \right) {\Biggr \}} \\ {\cal S}({\rm e}^{\frac{4\pi i}{3}}) & = & {\Biggr \{} u,v \,{\Biggr |}\, \left( \frac{\pi}{2}\,{\leq}\,{\tan}^{-1} \frac{Z_{\rm sin}}{Z_{\rm cos}}\,{\leq}\, -\frac{\pi}{3}\,\bigcap\, Z_{\rm cos} > 0, Z_{\rm sin} \,{\geq}\, 0 \right) \nonumber \\ \mbox{} & \mbox{} & \, \bigcup\, \left(0\,{\leq}\,{\tan}^{-1}\frac{Z_{\rm sin}}{Z_{\rm cos}}\,{\leq}\, \frac{\pi}{2}\, \bigcap\,Z_{\rm cos} < 0, Z_{\rm sin} \,{\leq}\, 0 \right) {\Biggr \}}. \end{eqnarray} \setcounter{equation}{\value{saveeqn} We first assume that the exchange term is absent (${\beta}_{J}=0$) \cite{Nishi83}, that is no redundancy is fed into the channel. In this case, the saddle point equations (\ref{sigma}-\ref{Q}) are drastically simplified and the overlap (\ref{OVERLAP}) simply reads \begin{equation} M = \frac{{\rm e}^{\beta_{s}m_{s}}} {{\cal Z}_{s}} {\rm Erf}\left[ -\sqrt{2}\frac{{\beta}_{d}m+{\tau}_{0}h}{\tau h} \right] + \frac{{\rm e}^{-\frac{1}{2}}} {{\cal Z}_{s}} \left\{ 1-{\rm Erf}\left[ -\sqrt{2}\frac{{\beta}_{d}m-{\tau}_{0}h}{\tau h} \right] \right\} \end{equation} with the magnetization given by \[ m = \frac{{\rm e}^{{\beta}_{s}m_{s}}} {{\cal Z}_{s}} \int\frac{du}{\sqrt{\pi}} {\rm e}^{-u^{2}} \frac{1-{\exp}[-\frac{3}{2}(u\tau h +{\beta}_{d}m/q+{\tau}_{0}h)]} {1+2{\exp}[-\frac{3}{2}(u\tau h +{\beta}_{d}m/q+{\tau}_{0}h)]} \] \begin{equation} \mbox{} + \frac{2{\rm e}^{-\frac{1}{2}{\beta}_{s}m_{s}}} {{\cal Z}_{s}} \int\frac{du}{\sqrt{\pi}} {\rm e}^{-u^{2}} \frac{1-{\exp}[-\frac{3}{2}(u\tau h +{\beta}_{d}m/q-{\tau}_{0}h/2)]} {1+2{\exp}[-\frac{3}{2}(u\tau h +{\beta}_{d}m/q-{\tau}_{0}h/2)]} \end{equation} where we defined ${\rm Erf}(x)\,{\equiv}\,\int_{x}^{\infty}{\rm e}^{-x^{2}}dx /\sqrt{\pi}$. The problem is thus reduced to a one-dimensional model, corresponding to an Ising model in which the length of spin turns out to be $(+1,-1/2)$ instead of $(+1,-1)$. This is not surprising if one thinks that the fluctuations along the imaginary axis are governed only by the glassy term, meanwhile the magnetic field acts along the real direction. In Fig. \ref{q3MTd_inf}, we plotted the overlap $M$ as a function of $T_{d}$ for the some values of $h$. It is straightforward to check that the maximum value of the overlap $M_{\rm max}$ does not depend on magnetic field $h$, since at the stationary point $\left(\partial M/\partial \beta_d=0\right)$ $m$ is proportional to the magnetic field \begin{eqnarray} \frac{1}{2} {\beta}_{s}m_{s}= \frac{1}{3}m{\beta}_{d} \frac{\tau_{0}}{\tau^{2}h} +\frac{1}{4} \frac{{\tau}_{0}^{2}}{\tau^{2}}. \label{relation} \end{eqnarray} This feature holds also for the Ising case, although the stationary equation (\ref{relation}) is simpler \begin{eqnarray} {\beta}_{s}m_{s}=m{\beta}_{d}\frac{{\tau}_{0}}{{\tau}^{2}h}. \end{eqnarray} Expression (\ref{relation}) is thought to be valid only for the infinite range model, as confirmed in the next section by numerical results in $d=2$. Now we set the decoding temperature at the optimal value, that is $M(T_d^{\rm opt})\equiv M_{\rm max}$ and we switch the exchange interaction ($\beta_J \neq 0$) as depicted in Fig. \ref{q3Mbeta_inf}. We notice that also a small amount of redundancy highly improves the value of the overlap $M_{\rm max}$ which quickly increases and slowly decreases, after the peak, meanwhile the exchange term becomes dominant on the ferromagnetic one. \section{Monte Carlo simulations for real world picture \label{Montecarlo}} Although for the mere restoration aims it is not wise to smoothen two points far away from each other, we shall see that the infinite range model provides a useful guide for the more interesting case of real-world pictures, since the results remain qualitatively similar. We thus carried out Monte Carlo simulations for realistic pictures with short-range effective Hamiltonian. In this case, the ferromagnetic term will be concerned only with the points within the range of interaction and two points far away will not influence each other. It would be extremely interesting to study the restoration quality as a function of the interaction radius, but this goes beyond the aim of the present work and we limit ourselves to a first nearest-neighbors interaction Hamiltonian. Therefore let us consider a simple $q=3$ gray scale level picture (upper left of Fig. \ref{q3noise15}), where each pixels has been randomly flipped to another value with some probability, say $p=0.15$ (upper right of Fig. \ref{q3noise15} ). The curves shown in Fig.\ref{q3MTd} are the result of the restoration process without the glassy term, that is $\beta_J=0$, at different values of the ratio $H=h/\beta_d$. Here the maximum value of the overlap is achieved around $H_{\rm max}{\equiv}h/{\beta}_{d}\,{\sim}\,0.6$ and $T_{d}\,{\sim}\,0.2$ and the corresponding restored image is drawn in the lower left of Fig. \ref{q3noise15}. Adding the glassy term at $H_{\rm max}$ fixed improves drastically the value of the overlap and the quality of the restored image (lower right of Fig. \ref{q3noise15}), drawn at the peak of the Fig. \ref{q3Mbeta}. The same procedure is repeated in the presence of higher noise, $p=0.30$, at the same $H_{\rm max}$ and $\beta_{J;{\rm max}}$ and the results of the restoration are shown in Figs. \ref{q3noise30}. Finally, we applied the same algorithm to an eight gray-scale level picture with $20\%$ and $30\%$ of noise, upper images in Figs. \ref{q8noise20} and \ref{q8noise30}. The results without exchange term are shown in Fig. \ref{q8MTd}. Once again we find a maximum for some values of $T_d$ and $H$ and the corresponding restored images are in Fig. \ref{q8Mbeta}. \section{Conclusions} In this paper, we investigated the possibility of gray-scaled image restoration using chiral random Potts model. We exactly solved the infinite ranged version thus deriving the explicit expression of the overlap as a function of the estimates of the hyper-parameters $h$, ${\beta}_{d}$ and ${\beta}_{J}$. In the absence of the glassy term, we obtained the exact relation between the restoration temperature ${\beta}_{d}$ and the source temperature ${\beta}_{s}$ which gives the maximum value of the overlap. This seems a highly non-trivial result because it is natural for us to assume that the best recovery of the image should be achieved for ${\beta}_{d}={\beta}_{s}$, as it turns out to be true for the Ising case\cite{NW}. The Monte Carlo results on real pictures confirmed the expected high improvement due to the presence of the redundancy, {\it i.e.} the glassy term. However, so far, in our prescription to recover a corrupted image at the best restoration values, one is supposed to know the original data. In other words, the receiver has to meet the sender at least once to find the optimal restoration values. Only after that, the other receivers will be able to get an optimal restoration for the same image, provided that the channels remain, at least qualitatively, unchanged. In this sense, it would be extremely useful to provide some {\it a priori} criteria (the receiver will not be supposed to meet the sender) for the optimum values of the hyper-parameters, once that some intrinsic characteristics ({\it e.g.} temperature) of the original image are known. Therefore, in order to check if the relation (\ref{relation}) still holds down to two dimensions, we restored $q=3$ ferromagnetic snapshots generated at some known temperature. However, so far we have not yet obtained reliable results and detailed investigations in this direction will be reported in a forth coming paper. \section*{acknowledgement} We thank Prof. Hidetoshi Nishimori for useful discussions and showing us his paper prior to publication. We also thank Prof. Kazuyuki Tanaka for useful advice and fruitful discussions. The authors were supported by JSPS-Royal Society/British Council Anglo-Japanese Scientific Cooperation Programme. One of the authors (D.M.C.) started this work under JSPS grant No.P96215.
\section{Introduction} In spite of many experimental and theoretical efforts (for a recent review see \cite{arneodo}), the origin of the nuclear EMC effect has not yet been fully clarified, and the problem as to whether the quark distributions of nucleons undergo deformations due to the nuclear medium remains open. Understanding the origin of the EMC effect would be of great relevance in many respects; consider, for example, that most QCD sum rules and predictions require the knowledge of the neutron quark distributions, which can only be extracted from nuclear experiments; this implies, from one side, a reliable knowledge of various nuclear quantities, such as the nucleon removal energy and momentum distributions, and, from the other side, a proper treatment of the lepton-nucleus reaction mechanism, including the effect of final state interaction (FSI) of the leptoproduced hadrons with the nuclear medium. Since the $Q^2$ and $x$-dependences of the EMC effect is smooth, the measurements of the nuclear quark distributions in inclusive deep inelastic scattering (DIS) processes have not yet established enough constraints to distinguish between different theoretical approaches. In order to progress in this field, one should go beyond inclusive experiments, e.g. by considering semi-inclusive experiments in which another particle is detected in coincidence with the scattered electron. Most of theoretical studies in this field concentrated on the process $D(e,e'N)X$, where $D$ denotes the deuteron, $N$ a nucleon, and $X$ the undetected hadronic state. Current theoretical models of this process are based upon the impulse approximation (IA) (also called {\it the spectator model} ), according to which $X$ results from DIS on one of the two nucleons in the deuteron, with $N$ recoiling without interacting with $X$ and being detected in coincidence with the scattered electron (for an exhaustive review see \cite{fs}). The model has been improved by introducing FSI \cite{fsi}, as well as by considering deviations from the spectator model by assuming that the detected nucleon originates from quark hadronisation \cite{dieper,ciofisim}. The semi-inclusive process on the deuteron $D(e,e'N)X$, on which experimental data will soon be available \cite{semiexp}, could not only clarify the origin of the EMC effect, but, as illustrated in \cite{simula}, could also provide more reliable information on the neutron structure function. The spectator model has also been extended to complex nuclei by considering the process $A(e,e'N)X$, and by assuming that DIS occurs on a nucleon of a correlated pair, with the second nucleon $N$ recoiling and being detected in coincidence with the scattered electron \cite{ciofisim}. In the present paper two new types of semi-inclusive processes on complex nuclei will be considered, namely: i) the process $A(e,e'(A-1))X$, in which DIS occurs on a mean-field, low-momentum nucleon, and the nucleus $(A-1)$ recoils with low momentum and low excitation energy and is detected in coincidence with the scattered electron (note that for $A=2$ such a process coincides with the process $D(e,e'N)X$ discussed previously); ii) the process $A(e,e'N{_2}(A-2))X$, in which DIS occurs on a high momentum nucleon $N_1$ of a correlated pair, and the nucleon $N_2$ and the nucleus $A-2$ recoil with high and low momenta, respectively, and are detected in coincidence with the scattered electron. It will be shown that these processes exhibit a series of very interesting features which could in principle provide useful insight on the following basic issues: $i)$ the nature and the relevance of FSI in DIS; $ii)$ the validity of the spectator mechanism leading to the cross section (\ref{crosa-1}); $iii)$ the medium induced modifications of the nucleon structure function; $iv)$ the origin of the $EMC$ effect. For the above reasons, the semi-inclusive processes we will consider are worth being theoretically analysed, even though their experimental investigation represents a difficult task. It should be emphasised, in this respect, that the first version of the present paper\cite{nasharchiv} was motivated by the discussions on the feasibility of an electron-ion collider, where the detection of various nuclear fragments resulting from DIS, could in principle be possible \cite{hera,gsi}. Our paper is organised as follows: in Section II the basic nuclear quantities which enter the problem, viz. the one-body and two-body nuclear Spectral Functions are briefly discussed; the cross section for the process $A(e,e'(A-1))X$ is presented in Section III, where the possibilities offered by the process to experimentally check the validity of the spectator mechanism and the properties of the structure function of a mean-field, {\it weakly bound} nucleon, are discussed; the cross section for the process $A(e,e'N{_2}(A-2))X$, and how this process can be used to investigate the spectator model and the properties of the structure function of a {\it deeply bound} nucleon, are discussed in Section IV; the {\it local} EMC effect, i.e. the separate contribution to the EMC effect of nucleons having different binding in the nucleus, is discussed in Section V; the Summary and Conclusions are presented in Section VI. Appendix A contains the derivation of the cross sections for both processes. \section{The nuclear spectral function} In order to make clear the nuclear physics aspects underlying the above processes, few basic concepts about the relationships between the nucleon momentum distributions in the parent nucleus $A$ and the excitation energy of daughter nuclei $(A-1)$ and $(A-2)$ in semi-inclusive processes, will be recalled. The nucleon Spectral Function $P_{N_1}(|\vec p_1|,E)$ represents the joint probability to have in the parent nucleus a nucleon with momentum $|\vec p_1|$ and removal energy $E$ \begin{eqnarray} && P_{N_1}(|\vec p_1|,E)= \langle \Psi_A^0\,|\, a^+_{\vec p_1}\delta \left ( E- ( H_A-E_A^0) \right ) \, a_{\vec p_1}\,|\Psi_A^0\rangle =\nonumber\\[1mm] && \sum_f \left | \langle \vec p_1, \Psi_{A-1}^f\,|\Psi^0_A \rangle \right |^2\delta \left (E-( E_{A-1}^f-E_A^0)\right), \label{spectr1} \end{eqnarray} where $a^+_{\vec p_1}$ and $a_{\vec p_1}$ are creation and annhilitation operators, $H_A$ is the nuclear Hamiltonian, $E_A^0$ ($\Psi_A^0$)is the ground state energy (wave function) of $A$, and $E_{A-1}^f=E_{A-1}^0 + E_{A-1}^*$ ($\Psi_{A-1}^f$)is the intrinsic energy (wave function) of $A-1$, whose ground state energy is $E_{A-1}^0$. Thus, the nucleon removal energy $ E = E_{A-1}^f - E_A^0 = M_{A-1} +M - M_A + E^*_{A-1}$ (where $M_i$ is the mass of system $i$) is the energy required to remove a nucleon from $A$ leaving $(A-1)$ with excitation energy $E^*_{A-1}$. A common representation of the spectral function is as follows (omitting unnecessary here indices and summations) \cite{ciofispec} \be P_{N_1}^A(|\vec p_1|,E) = P_0(|\vec p_1|,E)+P_1(|\vec p_1|,E) \label{spectrrap} \ee where \be P_0(|\vec p_1|,E)\label{eq1} = \sum\limits_{\alpha < F}n^A_\alpha(|\vec p_1|)\delta(E-\varepsilon_\alpha) \label{P0} \ee and \be P_1(|\vec p_1|,E) = {1 \over (2 \pi)^3} {1 \over 2 J_0 + 1} \sum_{M_0 \sigma} ~ \sum_{f \neq \alpha} \left | \int d \vec{r} ~ e^{i\vec{p_1} \cdot \vec{r}} ~ G_{f0} (\vec{r}) \right |^2 ~ \delta[E - (E_{A-1}^f - E_A)] \label{P1} \ee In the above equations $F$ denotes the Fermi level, $n^A_\alpha(|\vec p_1|)$ is the momentum distribution of a bound shell model state with eigenvalue $\varepsilon_\alpha > 0$, and $G_{f0}$ is the overlap between the wave functions of the ground state of the parent $A$ and the state $f$ of the daughter $(A-1)$ (see for details ref.~\cite{ciofi-fs,ciofi1,mar}). The quantity $P_0(|\vec p_1|,E)$, represents the shell model contribution to the Spectral Function, where the occupation numbers of the shell model states below the Fermi sea are given by $N_{\alpha}= \int {{d\vec p_1}n^A_\alpha(|\vec p_1|)} < 1$, whereas $P_1(|\vec p_1|,E)$ provides the contribution from correlations, which deplete the shell model states $\alpha < F$. The so called Momentum Sum Rule links the spectral function to the nucleon momentum distribution, viz. \be n^A(|\vec p_1|)=\int\limits_{E_{\rm min}}^\infty\,P_{N_1}^A(|\vec p_1|,E) dE = \sum\limits_{\alpha < F}n^A_\alpha(|\vec p_1|) + ~ \sum_{f \neq \alpha} \left | \int d \vec{r} ~ e^{i\vec{p_1} \cdot \vec{r}} ~ G_{f0} (\vec{r}) \right |^2~, \label{eq2} \ee where $E_{min}=E_{A-1}-E_A$. It can therefore be seen that $ n^A_0(|\vec p_1|) \equiv \displaystyle\sum\limits_{\alpha < F} n^A_\alpha(|\vec p_1|)= \int\limits_{E_{\rm min}}^\infty\,P^A_0(|\vec p_1|,E) dE , $ represents the momentum distribution in the parent, when the daughter is either in the ground state or in hole states of the parent, whereas $ n^A_1( |\vec p_1|) \equiv n^A( |\vec p_1|)-n^A_0( |\vec p_1|)= \int\limits_{E_{\rm min}}^\infty\,P^A_1(|\vec p_1|,E) dE $ represents the momentum distribution in the parent, when the daughter is left in highly excited states, with at least one particle in the continuum; this means that $n^A_0( |\vec p_1|)$ is the momentum distribution of weakly bound (shell-model) nucleons, while $n^A_1( |\vec p_1|)$ is the momentum distributions of deeply bound nucleons generated by N-N correlations. A realistic model for the latter leads to the following form of the corresponding spectral function $P^A_1( |\vec p_1|,E)$ \cite{ciofi-fs,ciofi1,mar} \be && P^A_{1}( |\vec p_1|,E)= \label{sfcorr}\\[3mm] && \int d^3k_{cm} n^A_{rel}\left (|\vec p_1 -\vec p_{cm}/2|\right ) n^A_{cm} (|\vec p_{cm}|) \delta \left [ E -E_{thr}^{(2)} - \frac{(A-2)}{2M(A-1)}\cdot \left ( \vec p_1 - \frac{(A-1)\vec p_{cm}}{(A-2)} \right )^2\right ], \nonumber \ee where $n^A_{rel}$ and $n^A_{cm}$ are, respectively, the relative and Center of Mass momentum distributions of a correlated pair. It has been shown \cite{ciofi1} that such a model satisfactorily reproduces the nuclear spectral functions calculated within many-body approaches with realistic $NN$ interaction and describes fairly well the quasi elastic inclusive $A(e,e')X$ processes. Within the spectator model, the process $A(e,e'(A-1))X$ is directly proportional to the one-nucleon spectral function, whereas the process $A(e,e'N(A-2))X$ is proportional to the two-nucleon spectral function, which is defined as follows \begin{eqnarray} && P_{N_1N_2}(\vec p_2,\vec p_1,E^{(2)})= \langle \Psi_A^0\,|\, a^+_{\vec p_1}a^+_{\vec p_2} \delta \left ( E^{(2)} - ( H_{A-2} - E_A) \right ) \, a_{\vec p_2}a_{\vec p_1}\,|\Psi_A^0\rangle =\nonumber\\[1mm] && \sum_f \left | \langle \vec p_1,\vec p_2, \Psi_{A-2}^f\,|\Psi^0_A \rangle \right |^2 \delta \left (E^{(2)} - ( E_{A-2}^f - E_A)\right), \label{spectr2} \end{eqnarray} where $E^{(2)}= E^{(2)}_{th} + E_{A-2}^*$ is the two-nucleon removal energy, $ E_{A-2}^*$ is the intrinsic excitation energy of the $A-2$ system, and $E^{(2)}_{th} = 2M + M_{A-2} -M_A$ the two- nucleon break-up threshold. If one adheres to the model leading to Eq. (\ref{sfcorr}), the correlated part of the two-nucleon spectral function can be written as follows\cite{ciofisim}: \begin{eqnarray}&& P_{N_1N_2}(\vec p_1,\vec p_2,E^{(2)})= \,n^A_{cm}(|\vec P_{A-2}|)\, n^A_{rel.}(|\vec p_2 +\vec P_{A-2}/2 |) \delta \left ( E^{(2)}- E^{(2)}_{th}\right ) \label{ster} \end{eqnarray} \section{The {\bf \it A\lowercase{(e,e'}(A-1))X} process} In Impulse Approximation, the process $A(e,e'(A-1))X$ (depicted Fig.\ref{fig1}a), represents the absorption of the the virtual photon by a quark of a shell-model nucleon, followed by the recoil of the nucleus $A-1$ in a low momentum ,${\vec P_{A-1}}$, and low excitation energy, $E_{A-1}^*$, state ($E_{A-1}^* \simeq 0$ or $\simeq$ shell-model hole state energy of the target);the scattered electron and the nucleus $(A-1)$ are detected in coincidence . The aim for studying such a process is twofold: \begin{itemize} \item [ i)] to investigate the nature of the final state interaction (FSI) of the hit quark with the surrounding nuclear medium; as a matter of fact, the observation of a nucleus $(A-1)$ in the ground state (or in a low shell model excited states) would represent obvious evidence that the leptoproduced hadrons propagated through the nucleus $(A-1)$ without strong FSI. Therefore, the number of observed $(A-1)$ systems and its variation with $A$ could provide important information on e.g. the hadronization length in the medium ; \item [ii)] to investigate the $A$-dependence of possible medium induced modifications of the DIS structure function of weakly bound nucleons. \end{itemize} In IA the differential cross section in the laboratory system has the following form (see Appendix A)~\cite{nasharchiv} \begin{eqnarray} &&\!\!\!\!\!\! \sigma^A_1 (x_{Bj},Q^2,\vec P_{A-1})\equiv\sigma^A_1= \frac{d\sigma^A}{d x_{Bj} d Q^2 d \vec P_{A-1}}\nonumber\\&& = K^A( x_{Bj},Q^2,y_A,z_1^{(A)}) z_1^{(A)} F_2^{N/A}(x_A,Q^2,p_1^2)\,n^A_0(|\vec P_{A-1}|), \label{crosa-1} \end{eqnarray} where: $Q^2 =-q^2= -(k_e-k_e')^2 = \vec q^{\,\,2} - \nu^2=4 {\cal E}_e {\cal E}_e' sin^2 {\theta \over 2}$ is the 4-momentum transfer (with $\vec q = \vec k_e - \vec k_{e'}$, $\nu= {\cal E}_e - {\cal E}_e' $ and $ \theta \equiv \theta_{\widehat{\vec k_e \vec k_{e'}}}$); $ x_{Bj} = Q^2/2M\nu $ is the Bjorken scaling variable; $p_1 \equiv(p_{10},\vec {p_1})$, with $\vec {p_1} \equiv - \vec P_{A-1} $, is the four momentum of the nucleon; $F_2^{N/A}$ is the DIS structure function of the nucleon $N$ in the nucleus $A$; $n_0^A(|\vec P_{A-1}|)$ is the 3-momentum distribution of the bound nucleon; $K^A( x_{Bj},Q^2,y_A, z_1^{(A)}) $ is the following kinematical factor \begin{eqnarray} && K^A( x_{Bj},Q^2,y_A,z_1^{(A)})= \frac{4\alpha^2}{Q^4} \frac {\pi}{x_{Bj}}\cdot \left( \frac{y}{y_A}\right)^2 \left[\frac{y_{A}^2}{2} + (1-y_A) - \frac{p_1^2x_{Bj}^2 y_A^2}{z_1^{(A)2}Q^2}\right ]~, \label{ka} \end{eqnarray} and \begin{eqnarray} && y=\nu/{\cal E}_e ~, \,\,\,\, y_A = (p_1\cdot q)/(p_1\cdot k_e) \label{ydef}\\ && x_A = {x_{Bj} \over z_1^{(A)}}, \quad z_1^{(A)} = {p_1 \cdot q \over M \nu}~. \label{adef} \end{eqnarray} Nuclear effects in Eq. (\ref{crosa-1}) are generated by the nucleon momentum distribution $n_0^A(|\vec P_{A-1}|)$, and by the quantities $y_A$ and $z_1^{(A)}$, which differ from the corresponding quantities for a free nucleon ($y=\nu/{\cal E}_e$ and $z_1^{(N)}=1$), if the off mass shellness of the nucleon ($p_1^2\neq M^2$ ) generated by nuclear binding is taken into account. Equation (\ref{crosa-1}) is valid for finite values of $Q^2$, and for $A=2$ agrees with the expression used in refs. \cite{simula,fs} (note, that in ref. \cite{simula} the quantity $D^N=K^A/K^N$ has been used, $K^N$ being the quantity (\ref{ka}) for a free nucleon, which will be discussed later on). In this paper, we follow the usual procedure consisting of disregarding the explicit dependence of $F_2^{N/A}$ upon $p_1^2$, and choose the form of $F_2^{N/A}$ to be the same as for the free nucleon; within such an approach, the effect of the nuclear medium will be considered within two main models: $i)$ {\underline{\sl the x-rescaling model}}, which directly follows from the convolution formula of inclusive scattering, leading to energy conservation at the hadronic vertex in Fig.1, i.e. \begin{equation} p_{10}= M_A - \sqrt{ (M_{A-1} + E_{A-1}^*)^2 + \vec P_{A-1}^2}~, \label{p10} \end{equation} which, when placed in eqs. (\ref{ydef}) and (\ref{adef}), leads to the following structure function for a bound nucleon \begin{equation} F_2^{N/A}(x_A,Q^2,p_1^2)=F_2^{N/A}\left (\frac{x_{Bj}}{z_1^{(A)}},Q^2 \right) \label{f2na} \end{equation} with \begin{equation} z_1^{(A)} = (p_{10}+|\vec P_{A-1} |\eta \cos\theta_{\widehat{\vec P_{A-1} \vec q} }) /M~, \label{zan} \end{equation} and \begin{equation} \eta = |\vec q|/\nu\,=\sqrt{1+\frac{4M^2x_{Bj}^2}{Q^2}}~. \label{eta} \end{equation} Since the $(A-1)$ system is detected in a low excited state ($E_{A-1}^* \simeq 0$) and with low momentum ($|\vec P_{A-1}|<<M_{A-1}$), Eq. (\ref{p10}) can be safely replaced by \begin{equation} p_{10} \simeq (M-E_{min})- {|\vec P_{A-1}|^2 \over 2M_{A-1}}~, \label{p10a} \end{equation} Eq. (\ref{zan}) then becomes \begin{equation} z_1^{(A)} \simeq 1 - {E_{min} \over M} - {|\vec P_{A-1}|^2 \over 2 M M_{A-1}} + \frac{\eta}{M} |\vec P_{A-1}| cos \theta_{\widehat{\vec P_{A-1} \vec q}}~ \label{zac} \end{equation} and for a heavy nucleus, for which the recoil term in (\ref{zac}) is negligibly small, one has \begin{equation} z_1^{(A)} \simeq 1 - {E_{min} \over M} + \frac{\eta}{M} |\vec P_{A-1}| cos \theta_{\widehat{\vec P_{A-1} \vec q}}~. \label{zaf} \end{equation} Moreover, being ${E_{min} \over M} << 1$, it can be concluded that the structure functions (\ref{f2na}) will exhibit almost no $A$-dependent effects, apart from the case of the few nucleon systems ($A$=2,3,4), for which the recoil term in (\ref{zac}) cannot be disregarded. In the Bjorken limit ( $Q^2\to\infty,\,\, \nu\to\infty$, $x_{Bj}={\rm const,\,\, \nu\sim\ |\vec q| }$), $\eta\to 1$) \begin{equation} z_1^{(A)}= (p_{10}+|\vec P_{A-1}| \cos \theta_{\widehat{\vec P_{A-1} \vec q}})/M \label{zab}. \end{equation} Note that Eq. (\ref{zab}) can also be written as ( $E_{A-1}^* = 0$ in the processes we are considering) \begin{equation} z_1^{(A)}=\frac{M_A}{M} -\frac{M_{A-1}z_{A-1}}{M} \label{z1} \end{equation} where \begin{equation} z_{A-1}= \frac{ \sqrt{\vec P_{A-1}^2 +M_{A-1}^2} - |\vec P_{A-1}| \cos\theta_{\widehat{\vec P_{A-1} \vec q} } }{M_{A-1}} \end{equation} is the light cone momentum of the $A-1$ recoiling nucleus. Eq. (\ref{z1}) is nothing but the energy conservation of the process \begin{equation} \nu +M_A=\sqrt{M_X^2+(\vec p_1+\vec q)^2}+\sqrt{M_{A-1}^2+{\vec p_1}^2} \label{conserv} \end{equation} in the Bjorken limit, where $M_X$ is the invariant mass of the produced hadronic state $X$; in the case of the deuteron, the term $\displaystyle\frac{|E_{min}|}{M}$ can be disregarded, so that $M_A/M \simeq 2$ and the well known relation $z_1^{(2)}=2-z_2$, where $z_2 = (\sqrt{|\vec p_2|^2 +M^2}-|\vec p_2|\cos \theta_{\widehat{\vec p_2 \vec q}})/M$, and $\vec p_2$ is the momentum of the recoiling nucleon, is recovered. $ii)$ {\underline{\sl the $Q^2-$rescaling model}} \cite{close}, which is based on the idea of a medium modification of the $Q^2-$evolution equations of $QCD$, leading to \begin{equation} F_2^{N/A}(x,Q^2)=F_2^{N}(x, \xi_A (Q^2)Q^2)~, \label{f2nar} \end{equation} where the $Q^2$ dependence of the quantity $\xi_A (Q^2)$ is determined so as to satisfy the $QCD$ evolution equations on both sides of (\ref{f2nar}), with the additional hypothesis that the quark confinement radius for a bound nucleon ($\lambda_A$) is larger than that for a free nucleon ($\lambda_N$), according to the ansatz \begin{equation} {\lambda_A^2 \over \lambda_N^2} = {\mu_N^2 \over \mu_A^2} = \xi_A (\mu_A^2)~, \label{lamb} \end{equation} where $\mu_A$ and $\mu_N$ are the lower momentum cutoffs for the bound and free nucleons, respectively. The following relation can then be obtained \begin{equation} \xi_A(Q^2)= \left (\lambda_A^2 \over \lambda_N^2 \right ) ^{ ln( Q^2/ \Lambda_{QCD}^2) \over ln( \mu_A^2 /\Lambda_{QCD}^2) }~, \label{csia} \end{equation} where $\Lambda_{QCD}$ is the universal $QCD$ scale parameter. To sum up, in the $Q^2$-rescaling model an explicit $A$ dependence is provided by Eq. (\ref{csia}) whereas, in the $x$-rescaling model, the $A$-dependence of $F_2^{N/A}$ is generated implicitly by the momentum $|\vec P_{A-1}|$ of the detected $A-1$ system (cf. Eq. (\ref{zac})). We will now discuss a series of processes, which could in principle provide useful insight on the following basic issues: $i)$ the nature and the relevance of FSI in DIS; $ii)$ the validity of the spectator mechanism leading to the cross section (\ref{crosa-1}); $iii)$ the medium induced modifications of the nucleon structure function; $iv)$ the origin of the $EMC$ effect. \subsection{Checking the spectator mechanism in the semi-inclusive process $A(e,e'(A-1))X$} The validity of the spectator mechanism could experimentally be checked in the following way. Let us consider the cross section (\ref{crosa-1}) for two different nuclei $A$ and $A'$, and the same values of $x_{Bj}$, $Q^2$ and $|\vec P_{A-1}|=|\vec P_{A'-1}|$. Consider now the ratio \begin{eqnarray} && R(x_{Bj},Q^2,|\vec P_{A-1}|,z_1^{(A)},z_1^{(A')},y_A ,y_{A'})= \frac{\sigma^A_1 ( x_{Bj},Q^2,|\vec P_{A-1}|,z_1^{(A)},y_A ) } {\sigma^{A'}_1 ( x_{Bj},Q^2,|\vec P_{A-1}|,z_1^{(A')},y_{A'} ) }= \nonumber \\ & = & \frac{K^A}{K_{A'}} \frac{z_1^{(A)} F_2^{N/A}(x_A,Q^2,p_1^2)}{z_1^{(A')} F_2^{N/A'}(x_{A'},Q^2,p_1^2) } \frac{n_0^A(|\vec P_{A-1}|)}{n_0^{A'}(|\vec P_{A-1}|)}~, \label{ratioa-1} \end{eqnarray} with $y_A$ and $z_1^{(A)}$ defined in Eqs. (\ref{ydef}) and (\ref{adef}), respectively. For reasons that would be clear later on, our aim is to get rid as much as possible of the various $A$ and $A'$ dependencies appearing in (\ref{ratioa-1}) , except the ones provided by the nucleon momentum distributions. The dependence upon $A$ and $A'$ is contained in the quantities $z_1^{(A)}$, $x_A$, $K^A$, and $n_0^A(|{\vec P_{A-1}}|)$; in order to get rid of the $A$-dependence due to the first three quantities let us consider coplanar kinematics, i.e. \be && y_A=y\cdot\frac{ p_{10}+\eta |\vec P_{A-1}|\cos \theta_{ \widehat{ \vec P_{A-1} \vec q} } } { p_{10}+\eta |\vec P_{A-1}|\cos \theta_{ \widehat{ \vec P_{A-1} \vec k_e} } }~, \quad \label{add1} \ee with \be && \cos\theta_{ \widehat{ \vec P_{A-1} \vec q}} = - \cos( \theta_{\widehat {\vec P_{A-1} \vec k_e}} +\theta_{ \widehat {\vec q \vec k_e}});\quad \cos\theta_{ \widehat {\vec q \vec k_e} } = \left (1+\frac{Mx_{Bj}} {{\cal E}_k}\right )/\eta~. \label{add2} \ee In the Bjorken limit $\eta\to 1$, $z_1^{(A)}= (p_{10}+|\vec P_{A-1}| \cos \theta_{\widehat{\vec P_{A-1} \vec q}})/M$ (cf. Eq. (\ref{zan})), $\theta_{ \widehat{ \vec P_{A-1} \vec q}}\to \theta_{ \widehat{\vec P_{A-1} \vec k_e}}$, $y_A\to y$, and \begin{eqnarray} && K^A( x_{Bj},Q^2,y_A,z_1^{(A)}) \rightarrow K^N( x_{Bj},Q^2,y) = \frac{4\alpha^2}{Q^4} \frac {\pi}{x_{Bj}}\cdot \left[\frac{y^2}{2} + 1 - y - { Q^2 \over 4 {\cal E}_e^2} \right ]~, \label{kap} \end{eqnarray} where $K^N$ is nothing but the trivial kinematic factor appearing in the DIS $eN$- inclusive cross section; the cross section (\ref{crosa-1}) thus becomes \begin{eqnarray} &&\!\!\!\!\!\! \left ( \frac{d\sigma^A}{d x_{Bj} d Q^2 d \vec P_{A-1}}\right )_{Bj} = K^N( x_{Bj},Q^2,y) z_1^{(A)} F_2^{N/A}(x_{Bj}/z_1^{(A)}, Q^2)\,n^A_0(|\vec p_{A-1}|), \label{crossBj} \end{eqnarray} with the $A$-dependence now appearing only in $ F_2^{N/A}$, $n^A_0(|\vec p_{A-1}|)$ and $z_1^{(A)}$. The latter dependence, however, can be eliminated by considering that Eq. (\ref{zac}) reduces (due to $E_{min}/M <<1$ ) to $z_1^{(A)} \simeq 1 - {|\vec P_{A-1}|^2 \over 2 M M_{A-1}} + \frac{ |\vec P_{A-1} | } {M} cos \theta_{\widehat{ \vec P_{A-1} \vec q}}~,$ so that by by fixing $|\vec P_{A-1}|$, and properly changing $\theta_{ \widehat{\vec P_{A-1} \vec q} }$, the condition $z_1^{(A)} \sim z_1^{(A')}$ can easily be achieved. As a result, the Bjorken limit of Eq.(\ref{ratioa-1}) becomes \begin{eqnarray} R_{Bj}(x_{Bj}/z_1^{(A)},Q^2,|\vec P_{A-1}|,A,A') & = & \frac {F_2^{N/A}(x_{Bj}/z_1^{(A)},Q^2)} {F_2^{N/A'}(x_{Bj}/z_1^{(A')},Q^2)} \frac{n_0^A(|\vec P_{A-1}|)}{n_0^{A'}(|\vec P_{A-1}|)} \rightarrow \\ \nonumber & \rightarrow & \frac{n_0^A(|\vec P_{A-1}|)}{n_0^{A'}(|\vec P_{A-1}|)}\equiv R(| {\vec P_{A-1}}\ |), \label{eq5} \end{eqnarray} where the last step is strictly valid only within the $x$-rescaling model, for in the $Q^2$-rescaling model the additional $A$ and $A'$-dependences appearing in $F_2^{N/A}(x, Q^2) =F_2^N(x, \xi_A(Q^2)Q^2)$ does not cancel out, being different in the numerator and the denominator; such a dependence, however, is overwhelmed by the $A$-dependence of $n_0(\left | {\vec P_{A-1}}\right |)$ as it will be shown later on. We have thus obtained that in the Bjorken limit the $A$ dependence of the ratio $R$ is entirely governed by the $A$ dependence of the nucleon momentum distribution $n_0^A(|\vec P_{A-1}|)$. Since the latter exhibits a strong $A$ dependence for low values of $|\vec P_{A-1}|$, a plot of $R$ versus $|\vec P_{A-1}|$ should reproduce the behaviour of $n_0^A(|\vec P_{A-1}|)$ which is fairly well known, so that the experimental observation of such a behaviour would represent a stringent test of the spectator mechanism independently of the model for $F_2^{N/A}$. Fig. 2 illustrates the expected behaviour of the ratio (\ref{eq5}) for $A=2$ and different values of $A'$. The measurement of the quantity $R$ shown in Fig. 2 would imply the detection, in coincidence with scattered electrons, of backward recoiling , with momentum $\vec P_{A-1}$, protons, deuterons, $^3He$ and $^{12}C$ nuclei resulting from the processes $D(e,e'p)X$, $^3H(e,e'D)X$, $^4He(e,e'^3He)X$ and $^{12}C(e,e'^{11} C)X$, respectively, with the DIS scattering processes supposed to occur on a neutron. \footnote { Note that the condition $z_1^{(A)}/z_1^{(A')} =1$ cannot be achieved if both $\theta_{\widehat{\vec P_{A-1} \vec q}}$ are fixed, so that in Fig.2 $z_1^{(A)}/z_1^{(A')}$ is a function of $P_{A-1}$; however the $P_{A-1}$-dependence of the quantity $z_1^{(A)}F_2^{N/A}(x_{Bj}/z_1^{(A)})/ z_1^{(A')}F_2^{N/A'}(x_{Bj}/z_1^{(A')})$ is at most of the order 5 percents and the dependence of $R(|\vec P_{A-1}|)$ upon $|\vec P_{A-1}|$ is entirely provided by the momentum distributions.} Since the results presented in Fig.2 were obtained in the Bjorken limit, where $K^A=K^{A'}=K^N$, let us analyse at which value of $Q^2$ such an equality is fulfilled. To this end, in Fig. 3 the ratio $K^A/K^N$ is shown vs. $Q^2$ for of $A=4$. It can be seen that at $Q^2 \simeq 5 GeV^2$ $K^A$ and $K^N$ differ by $5\%$ only. The cross sections corresponding to the processes considered in Fig.2 are presented in Fig.4. From the results we have exhibited, it is clear that the observation of recoiling nuclei in the ground state, with a $|\vec P_{A-1}|$-dependence similar to the one predicted by the momentum distributions, would represent a stringent check of the spectator mechanism, which, in turns, would indicate the absence of significant FSI between the lepto-produced hadronic states and the nuclear medium. The experimental observation of $(A-1)$ nuclei in the ground states would represent a strong indication that the hadronization length is larger than the effective nuclear dimension, since if the hit quark hadronizes inside the nucleus, the latter is expected to be strongly excited. Of particular relevance, in this respect, would be the processes $^3He (^3H) (e,e'D)X$, for if FSI plays an important role, the weakly bound final state deuteron will easily break down. It is clear, therefore, that the experimental observation of the exclusive process $A(e,e'(A-1)_{gr})X$ is strong evidence of the smallness of FSI. Although recent calculations \cite{fsia} and experimental data \cite{adams} seem to indicate that FSI on a complex nucleus are small in semi inclusive DIS, particularly when the low momentum hadrons are detected backward, the situation is not clearly settled, and therefore the observation of protons and deuterons emitted backward in the processes $D(e,e'p)X$, $^3He(e,e'D)X$ with a $|\vec P_{A-1}|$ dependence shown in Fig.2, would represent strong indication of the absence of FSI. The situation here is different from the usually investigated semi-inclusive DIS processes $A(e,e'N)X$ where the detected nucleon can originate not only from a correlated pair, as originally suggested\cite{fs}, but from competitive processes as well, such as nucleon current and target fragmentations \cite{dieper,ciofisim}. Let us now discuss the possibility to obtain information on the nucleon structure function of a weakly bound nucleon by means of the process $A(e,e'(A-1))X$. \subsection{Investigating the structure functions of weakly bound nucleons by the process $A(e,e'(A-1))X$} Consider the following quantity \begin{equation} R^A(x_{Bj},x_{Bj}',Q^2,|\vec P_{A-1}|) \equiv \frac{ \sigma_1^A (x_{Bj}, Q^2,|\vec P_{A-1}|,z_1^{(A)},y_A)} { \sigma_1^A (x_{Bj}',Q^2,|\vec P_{A-1}|,z_1^{(A)},y_A)} \label{x1x2} \end{equation} which represents the ratio between the cross section (\ref{crosa-1}) on the nucleus $A$ considered at two different values of the Bjorken scaling variable. It is clear that all terms of (\ref{crosa-1}), but the nucleon structure functions, cancel out in the ratio, and one has \begin{equation} R^A(x_{Bj},x_{Bj}',z_1^{(A)},Q^2) = { x_{Bj}' \over x_{Bj} } \frac{ F_2^{N/A} \left ( { x_{Bj} \over z_1^{(A)} } , Q^2 \right ) } { F_2^{N/A} \left ( { x_{Bj}' \over z_1^{(A)}} , Q^2 \right ) } \label{x1x2x} \end{equation} in the $x-$rescaling approach, and \begin{equation} R^A( x_{Bj}, x_{Bj}', Q^2 ) = { x_{Bj}' \over x_{Bj} } { F_2^{N/A } \left ( x_{Bj}, \xi_A (Q^2) Q^2 \right ) \over F_2^{N/A} \left ( x_{Bj}', \xi_A (Q^2) Q^2 \right ) } = constant~, \label{x1x2q} \end{equation} in the $Q^2-$rescaling model. Eqs. (\ref{x1x2x}) and (\ref{x1x2q}) will in general exhibit a different $|\vec P_{A-1}|$ dependence: Eq. (\ref{x1x2q}) will be a $|\vec P_{A-1}|$-independent constant different for different nuclei, whereas Eq. (\ref{x1x2x}) will depend both upon $A$ and $|\vec P_{A-1}|$, due to the dependence of $z_1^{(A)}$ upon $|\vec P_{A-1}|$ (cf. Eq. (\ref{zan}) with $\eta=1$). Let us consider the ratio (\ref{x1x2}) for the deuteron and for a complex nucleus; placing (\ref{p10a}) in (\ref{zan}), one obtains $z_1^{(2)} \simeq 1 - {{\cal E}_D \over M} - \frac{ |\vec P_{A-1}|^2}{2M^2} +\frac{|\vec P_{A-1}|}{M} cos \theta_{ \widehat {\vec P_{A-1} \vec q} }$~ and a strong $|\vec P_{A-1}|$ dependence will originate from the recoil and the angle-dependent terms; for a complex nucleus, on gets $z_1^{(A)} \simeq 1 - { E_{min} \over M} +\frac{|\vec P_{A-1}|}{M} cos \theta_{\widehat{\vec P_{A-1}\vec q}}$, which appreciably differs from unity only for $\theta_{\widehat{\vec P_{A-1} \vec q} }=180^o$ and/or large values of $|\vec P_{A-1}|$. Thus, the $|\vec P_{A-1}|$-dependence of the ratio (\ref{x1x2x}) can be changed by varying the dependence of $z_1^{(A)}$ upon $|\vec P_{A-1}|$; in such a way, ${x_{Bj} \over z_1^{(A)}} \neq {x_{Bj}'\over z_1^{(A)}}$ and $R^A$ will differ from a constant. The ratio (\ref{x1x2x}), for $A=2$ and $A=40$, is shown in Figs.\ref{fig5} and \ref{fig6} in correspondence of two values of the emission angle $\theta_{\widehat{\vec P_{A-1} \vec q} }$ of the nucleus $A-1$ ( $\theta_{\widehat{\vec P_{A-1} \vec q} } =90^o$ and $180^o$), and $x_{Bj}=0.2$ and $x_{Bj}'=0.5$. It can indeed be seen that: $i)$ in the $Q^2$-rescaling model the ratio is independent of $|\vec P_{A-1}|$, $ii)$ when $\theta_{\widehat{\vec P_{A-1} \vec q} } =90^o$, the $x-$rescaling model predicts a $|\vec P_{A-1}|$-independent ratio for $^{40}Ca$ $(z_1^{(40)}\simeq 1)$ and a strongly $|\vec P_{A-1}|$-dependent ratio for $D$ $(z_1^{(2)}\simeq 1 - \frac{ |\vec P_{A-1}|^2}{2M^2} )$; when $\theta_{\widehat{\vec P_{A-1} \vec q} }=180^o$, also the ratio for $^{40}Ca$ becomes strongly $|\vec P_{A-1}|$-dependent, for, now, $z_1^{(40)}\simeq 1 - \frac{ |\vec P_{A-1}|}{M}$. To sum up, it can be seen that the semi-inclusive process allows one to choose a variety of kinematical conditions which enhance various aspects of the problem. We have seen in Section III-B that, due to the small values of $|\vec {P_{A-1}}|$ and $E_{min}$, the process $A(e,e'(A-1))X$ on a complex nucleus is characterized by $z_1^{(A)}\simeq 1$ when $\theta_{\widehat {\vec P_{A-1} \vec q}}=90^o$; as a result, the off-mass-shell dependence of $F_2^{N/A}$ disappears (cf. full curve in Fig. \ref{fig5}); the off--mass--shell dependence of $F_2^{N/A}$ can on the contrary be enhanced if $\theta_{\widehat{ \vec P_{A-1} \vec q}}=180^o$, for an appreciable contribution from the last term of Eq. (\ref{zac}) is now generated; if so, however, the ratio (\ref{x1x2x}) for a complex nucleus will not appreciably differ from that of the deuteron (cf. full curves in Fig.\ref{fig6}), since off-mass-shell effects are solely due to the nucleon momentum $| \vec P_{A-1}|$, and not to medium effects provided by e.g. the nucleon binding ($E_{min}/M << 1$). Possible modifications of $F_2^{N/A}$ due to medium effects, will be discussed in the next Section. \section{The {\bf \it A\lowercase{(e,e'} $N_2$(A-2))X} process} In the previous Section we have discussed the case of weakly bound, non-correlated nucleons. In the present section we will investigate the semi inclusive processes occurring on a strongly correlated nucleon pair. To this end, let us consider the process depicted in Fig. 1 (b), which represents the absorption of the virtual photon by a correlated nucleon $N_1$ (with high momentum $|\vec P_{A-1}|$), followed by the emission of the partner nucleon $N_2$ ( with momentum $\vec p_2$), and by the recoil of the $(A-2)$ system, with low momentum $\vec P_{A-2}=-(\vec P_{A-1} + \vec p_2)$ and low excitation energy. The experimental investigation of such a process would require the coincidence detection of the scattered electron, the nucleon $N_2$ and the system $(A-2)$. The differential cross section of the process reads as follows \begin{eqnarray} \sigma_2 ^A(x_{Bj}, & Q^2 &,\vec P_{A-2}, \vec p_2) \equiv \frac{d\sigma^A}{d x d Q^2 d \vec P_{A-2} d \vec p_{2}} = \nonumber\\[1mm] & & K^A( x_{Bj},Q^2,y_A,z_1^{(A)}) z_1^{(A)} F_2^{N/A}(x_{Bj}/z_1^{(A)},Q^2) n^A_{cm}(|\vec P_{A-2}|)\, n^A_{rel.}(|\vec p_2 +\vec P_{A-2}/2 |) \label{crosa-2} \ee where $P_{N_1,N_2}$ is the two-nucleon spectral function, defined in Section II, and $K^A$, $y$, $y^A$, $x^A$, and $z_1^A$ are defined by Eqs. (\ref{ka}), (\ref{ydef}), (\ref{adef}), and (\ref{zan})with \begin{equation} \vec p_1 = -\vec P_{A-1} = (\vec p_2 + \vec P_{A-2})~. \label{queva} \end{equation} and \begin{equation} p_{10}=M_A-\sqrt{M^2+\vec p_2^2} -\sqrt{M_{A-2}^2+\vec P_{A-2}^2}. \label{aa2} \end{equation} In the Bjorken limit $\eta=1$ one has \begin{equation} z_1^{(A)}=\frac{M_A}{M}-z_2-\frac{M_{A-2}}{M}z_{A-2}, \label{aa3} \end{equation} where \begin{equation} z_2=\frac{\sqrt{M^2+\vec p_2^2} - |\vec p_2 | \cos \theta_{\widehat{\vec p_2 \vec q}}}{M} \label{aa4} \end{equation} and \begin{equation} z_{A-2}=\frac{\sqrt{M_{A-2}^2+\vec P_{A-2}^2}-|\vec P_{A-2}| \cos \theta_{\widehat{\vec P_{A-2}\vec q}}}{M_{A-2}} \label{aa5} \end{equation} are the light cone momentum fraction of nucleon $N_2$ and nucleus $(A-2)$, respectively. Note that Eq. (\ref{aa3}) is nothing but the energy conservation of the process \begin{equation} \nu +M_A=\sqrt{M_x^2+(\vec p_1+\vec q)^2} + \sqrt{M^2+\vec p_2^2} + \sqrt{M_{A-2}^2+\vec P_{A-2}^2} \label{conserv1} \end{equation} in the Bjorken limit. In the non relativistic approximation one obtains \begin{equation} z_1^{(A)} \simeq 1 - {E \over M} - {|\vec P_{A-1}|^2 \over 2 (A-1)M^2} + \frac{|\vec P_{A-1}|}{M} cos \theta_{\widehat{\vec P_{A-1} \vec q} }~, \label{zac21} \end{equation} with $ E = (E_{th}^{(2)} + E_{A-1}^*)$. Due to the small value of $\vec k_{cm} = - \vec P_{A-2}$, we can write $E_{A-1}^* \simeq {(A-2) \over 2M(A-1)} |\vec P_{A-1}|^2$, and by considering that $\vec P_{A-1} = -(\vec p_2 + \vec P_{A-2})$ (cf. (\ref{queva}) ), one gets: \begin{equation} E = E_{th}^{(2)} + {(A-2) \over 2 M (A-1)} \left [ |\vec p_2|^2 + \left ( {A-1 \over A-2} \right )^2 \vec P_{A-2}^2 + 2 { A-1 \over A-2 } |\vec p_2| |\vec P_{A-2}| cos \theta_{ \widehat {\vec p_2 \vec P_{A-2}}} \right ] ~. \label{30a} \end{equation} It should be stressed that the nucleon structure function $F_2^{N/A}(x_{Bj}/z_1^{(A)},Q^2)$ appearing in Eqs. (\ref{crosa-1}) and (\ref{crosa-2}) reflects different physical situations, for in the first case $F_2^{N/A}$ represents the quark distribution in a weakly bound, quasi-free nucleon, whereas, in the second case, it represents the quark distribution in a strongly bound nucleon, which, in principle, can undergo, because of binding, off-mass-shell deformations (see, for instance refs. \cite{kukhanna,mthomas}). Therefore, if the nucleon structure function could be extracted from the cross section (\ref{crosa-2}) and compared with the one obtained from the cross section (\ref{crosa-1}), a direct comparison of nucleon structure functions for weakly bound and deeply bound nucleons could, for the first time, be carried out. It should be pointed out that, since $y_A$ depends upon the high momentum $|\vec p_2|$, the factor $ K^A(x_{Bj},Q^2,y_A)$ may strongly differ from $ K^N(x_{Bj},Q^2,y)$, unless one of the two following kinematical conditions are chosen: $i)$ small values of $x_{Bj}$; $ii)$ the Bjorken limit. We found that at $Q^2=20 GeV^2/c^2 $ and $x_{Bj}=0.05$, the direction of the momentum transfer $\vec q$ coincides, in the frame where the target is at rest, with the electron beam direction ( $ \theta_{\widehat {\vec k \vec q}}\approx 2^0$ ); in this case, $y_A\simeq y$ and $K^A \simeq K^N$ (our numerical estimates show that $K^A/K^N$ varies from $0.99$ at $|\vec p_2|=350$ MeV/c to $0.96$ at $|\vec p_2|=1$ GeV/c); adopting realistic figures for an electron-ion collider, i.e. ${\cal E}_e \approx 5 $ GeV, $T_N$ = (kinetic energy per nucleon) $\approx 25$ GeV \cite{gsi} in its laboratory system, the chosen values of $Q^2$ and $x_{Bj}$ correspond to ${\cal E}_e'\approx 2\,GeV;\quad \theta_{\widehat {\vec k \vec k'}} \approx 90^0$ (in the nucleus rest frame they would correspond to ${\cal E}_e\sim 260$ GeV, $ {\cal E}_{e'}\approx 50$ GeV; $\theta_{\widehat {\vec k \vec k'}} \approx 2^0$). \subsection{Checking the spectator mechanism in the semi-inclusive process $A(e,e'N_2(A-2))X$} The validity of eq. (\ref{crosa-2}) can experimentally be tested by taking advantage of the observation ~\cite{pan} that for high values of $|\vec P_{A-1}|$ the nucleon momentum distribution for a complex nucleus turns out to be the rescaled momentum distribution of the deuteron, with very small $A$ dependence (unlike what happens for the low momentum part of $n(k)$ (cf. Fig. 2)). Let us therefore consider the following ratio, where $|\vec P_{A-2}|=|\vec P_{A'-2}|$: \begin{eqnarray} R(x_{Bj},Q^2,\vec P_{A-2},\vec p_2,z_1^{(A)},z_1^{(A')} )\, & \equiv\,& \displaystyle\frac{\sigma_2^A(x_{Bj},Q^2,\vec P_{A-2},|\vec p_2|)} {\sigma_2^{A'}(x_{Bj},Q^2,\vec P_{A-2},|\vec p_2|)} \nonumber\\[1mm] & = & \frac{z_1^{(A)}}{z_1^{(A')}} \frac{F_2^{N/A}(x_{Bj}/z_1^{(A)},Q^2)}{F_2^{N/A'}(x_{Bj}/z_1^{(A')},Q^2)} \cdot \frac{n^A_{rel.}(|\vec p_{rel.}|)}{n^{A'}_{rel.}(|\vec p_{rel.}|)} \cdot \frac{n^A_{cm}({ |\vec P_{A-2}| })}{n^{A'}_{cm}({ |\vec P_{A-2}| })}~, \label{ratio2} \end{eqnarray} where $ |\vec p_{rel}| =|\vec p_2 +\vec P_{A-2}/2 |$. If $|\vec P_{A-2}|$ is fixed, then, provided $F_2^{N/A} = F_2^{N/A'}$, the ratio, measured at $p_{rel} \geq 2 - 3$ $fm^{-1}$, would be roughly a constant, since $n^A_{rel.}\propto n^{D}$ for any $A$. The condition $F_2^{N/A} = F_2^{N/A'}$ can be achieved by properly choosing, for $A$ and $A'$, the values of $\vec p_2$ and $\vec P_{A-2}$ appearing in (\ref{30a}), so as to make $z_1^{(A)} \simeq z_1^{(A')}$, i.e. $F_2^{N/A} \simeq F_2^{N/A'}$ (note, moreover, that for large values of $| \vec P_{A-1}|$ and large values of $A$, the dependence of $z_1^{(A)}$ upon $A$ is unessential). To summarize, the cross-section (\ref{crosa-2}) should be measured for the systems $A$ and $A'$ at the same values of $x_{Bj}$, $Q^2$ and $\vec P_{A-2}$, changing the values of the angle $\theta_{\widehat {\vec p_2 \vec P_{A-2} }}$ and $|\vec p_2|$ so as to vary $|\vec p_{rel}| = |\vec p_2 + \vec P_{A-2}/2|$, keeping $z_1^{(A)}=z_1^{(A')}$. If Eq. (\ref{crosa-2}) is basically correct, the ratio (\ref{ratio2}) plotted versus $|\vec p _{rel}| \geq 2 - 3$ fm$^{-1}$ should exhibit (as shown in Fig. \ref{fig7})the same deuteron-like behaviour for any two nuclei in the range, say, $2<A<208$. If such a deuteron -- like behaviour of eq. (\ref{ratio2}) is experimentally found, it would represent a stringent test of the spectator mechanism. A word of caution is however in order here: the FSI between the nucleon $N_2$ and the nucleus $(A-2)$ will presumably affect the ratio (\ref{ratio2}). Calculations of the FSI within the Glauber multiple scattering approach are in progress and will be reported elsewhere; preliminary results indicate that in the region of the considered kinematics, the replacement of the undistorted two-body Spectral Function with the distorted one, mainly affects the absolute value of the ratio (\ref{ratio2}). \subsection{Investigating the structure functions of deeply bound nucleons by the process $A(e,e'N_2(A-2))X$} As in the case of the $A(e,e'(A-1))X$ process, in order to investigate the structure function of a (deeply) bound nucleon, we have to figure out experimentally measurable quantities which could provide information on $F_2^{N/A}$ without contaminations from the nucleon momentum distributions, or other momentum dependent terms. To this end, we use the analog of the ratio (33) which, within the convolution model, assumes the following form \begin{eqnarray} R(x_{Bj},x_{Bj}',Q^2,|\vec P_{A-2}|,|\vec p_2)|\, & \equiv\,& \displaystyle\frac{\sigma_2^A(x_{Bj},Q^2,|\vec P_{A-2}|,|\vec p_2|)} {\sigma_2^{A}(x_{Bj}',Q^2,|\vec P_{A-2}|,|\vec p_2|)} \nonumber\\[1mm] & = & \frac{x_{Bj}'}{x_{Bj}} \frac{F_2^{N/A}(x_{Bj}/z_1^{(A)},Q^2)}{F_2^{N/A}(x_{Bj}'/z_1^{(A)},Q^2)}~. \label{ratios} \end{eqnarray} It should be pointed out that, although the r.h.s. of eqs. (\ref{x1x2x}) and (\ref{ratios}) look formally the same, they differently depend upon the nucleon binding, for, we reiterate, in Eq. (\ref{x1x2x}) one has (cf. Eq. (\ref{zac})) \begin{equation} z_1^{(A)} \simeq 1 - {E_{min} \over M} - {|\vec P_{A-1}|^2 \over 2 (A-1)M^2} + \frac{|\vec P_{A-1}|}{M} cos \theta_{\widehat {\vec P_{A-1} \vec q }}~. \label{zacn} \end{equation} with $\vec P_{A-1} \equiv -\vec {p_1}$, whereas in Eq. (\ref{ratios}), one has (cf. Eq. (\ref{zacn}) ) \begin{equation} z_1^{(A)} \simeq 1 - {E \over M} - {|\vec P_{A-1}|^2 \over 2 (A-1)M^2} - \frac{|\vec P_{A-1}|}{M} cos \theta_{\widehat {\vec p_1 \vec q}}~, \label{zac2} \end{equation} with $\vec P_{A-1} = -(\vec p_2 + \vec {P_{A-2}})$ and $E$ given by Eq. (\ref{30a}). We have seen in Section 3.2 that, due to the small values of $|\vec {P_{A-1}}|$ and $E_{min}$, the process $A(e,e'(A-1))X$ on a complex nucleus is characterized by $z_1^{(A)}\simeq 1$, when $\theta_{\widehat {\vec P_{A-1} \vec q}}=90^o$, with the result that the off-mass-shell dependence of $F_2^{N/A}$ disappears (cf. the full line for $^{40}Ca$ in Fig.\ref{fig5}); the off--mass--shell dependence of $F_2^{N/A}$ can be enhanced if $\theta_{\widehat{ \vec P_{A-1} \vec q}}=180^o$, for an appreciable contribution from the last term of Eq. (\ref{zac2}) is generated; if so, however, the ratio (\ref{x1x2x}) for a complex nucleus will not appreciably differ from that of the deuteron (cf. the full curves in in Fig.\ref{fig6}), since off-mass-shell effects are solely due to the nucleon momentum $| \vec P_{A-1}|$, with no contribution from nucleon binding ($E_{min}/M << 1$); a totally different situation is expected to occur in the process $A(e,e'N_2(A-2))X$; as a matter of fact, in this case the ``binding term'' ${ E/M}$ in Eq. (\ref{zac2}) will generate an appreciable contribution to $z_1^{(A)}$, due to the large value of $|\vec p_2|$ associated to nucleon-nucleon correlations (or, equivalently, to high values of the removal energy $E$). Thus, in order to check whether the structure for a deeply bound nucleon would dynamically differ from the one for a weakly bound one, the ratios (\ref{x1x2x}) and (\ref{ratios}) for a given nucleus should be plotted versus the same value of $z_1^{(A)}$; in such a way, the off-mass-shell dependence of $F_2^{N/A}$ is quantitatively the same, but it originates from different contributions to $z_1^{(A)}$, viz. the momentum $\vec P_{A-1}$, for the weakly bound nucleon, and the binding effect $E$, for a deeply bound nucleon. If a different behaviour of the two ratios is found, this would represent strong evidence that the structure functions for weakly and deeply bound nucleons are different. Here, again, the $N_2-(A-2)$ FSI should be taken into account, although its effect is expected to be canceled in the ratio (\ref{ratios}). Another possibility to investigate the nucleon structure functions would be to analyze the following ratio \begin{equation} R_2=\frac { \sigma_2 ^A } { \sigma_1^D } =\frac{ z_1^{(A)}\, F_2^{N/A} ( x_A, Q^2, p_1^2 )\, n_{cm}^A( |\vec P_{A-2}| )\,n_{rel.}^A (|\vec p_2 + \vec P_{A-2}|)} {z_1^{(D)} \,F_2^{N/D} ( x_D, Q^2, p_1^2 ) \, n^D ( |\vec P_{A-1}| = |\vec p_2 + \vec P_{A-2}| ) } \label{ratio3}~. \end{equation} \\ \noindent The results for $A=12$ are presented in Fig.\ref{fig8}, in correspondence of two values of $|\vec p_2|$, for fixed values of the following kinematical variables: $|\vec P_{A-2}| = 50 MeV/c$, $\theta_{ \widehat{p_2q}} = 90^0$, $\theta_{ \widehat{p_2P_{A-2}}} = 180^0$, and $Q^2=20 GeV^2/c^2$; the full and dashed curves refer to the $x$- and $Q^2$-rescaling models, respectively. Let us first analyze the results predicted by the first model, viz. $F_2^{N/A} (x_A,Q^2,p_1^2) \rightarrow F_2^{N/A} (x_{Bj}/z_1^{(A)},Q^2,p_1^2)$ with $z_1^{(A)}$ given by (\ref{zac21}). The value of the three-momentum of the $(A-1)$ fragment (a nucleon) in the $A(e,e'(A-1))X$ cross section off the deuteron, has been chosen the same as the three-momentum of the interacting nucleon $\vec P_{A-1}$ in the case of the $A(e,e',N_2(A-2))X$ process off $^{12}C$; by this choice, the removal energy which appears in $z_1^{(A)}$ (\ref{zac2}) is almost equal to the recoil energy appearing in $z_D$, so that $z_1^{(A)} \simeq z_1^{D}$; by this way one should expect a constant behaviour of $R_2$ (note that $K_A \simeq K_D$, for in both cases one has to do with the same values of $\vec P_{A-1}$); the deviation from a constant exhibited by the full lines in Fig.\ref{fig8} is due to the fact that, with the chosen kinematics, $z_1^{(D)} > z_1^{(A)}$. Again, the observation of a behaviour different from the one presented in Fig.\ref{fig8} would indicate a dependence of $F_2^{N/A}$ upon the binding energy. Let us now consider the prediction by the $Q^2$-rescaling model. For the latter, we have considered the model of Ref. \cite{mthomas}, where the renormalization scale associated to the momentum of a bound nucleon is given by its invariant mass, $p_1^2 \not= M^2$ . Such an assumption leads to the ansatz $F_2^{N/A}(x_A,Q^2,p_1^2)= F_2^N(x_A,\xi_A(Q^2,p_1^2)Q^2)$ with $\xi_A(Q^2,p_1^2)= { (M^2/p_1^2) } ^{(\alpha(p_1^2))/ (\alpha(Q^2))}$, i. e. to an explicit dependence upon the off--shellness of the nucleon . Since the invariant mass of a deeply bound nucleon strongly differs from $M^2$, the ratio (\ref{ratio3}) gets the strong $x_{Bj}$ and $|\vec p_2|$ dependence shown in Fig. 8. \section{The local $EMC$ effect} In the binding model ($x-$ rescaling) of the $EMC$ effect, the slope of the ratio $R(x_{Bj},Q^2)$=$F_2^A(x_{Bj},Q^2)/(A F_2^N(x_{Bj},Q^2))$ is generated by the average value of the nucleon removal energy $<E>$: the larger the value of $<E>$, the stronger the $EMC$ effect\cite{fs}. Since $NN$ correlations produce high values of $E$, and therefore strongly affect the value of $R$ \cite{simo}, it would be extremely interesting to measure the so-called {\it local EMC} (LEMC) effect, i.e., the separate contribution to the ratio $R$ of the weakly and deeply bound nucleons. Several calculations of the local $EMC$ effect appeared \cite{loc1,loc2}, and attempts have also been made to compare them with experimental data on neutrino-nucleus DIS \cite{loc3}, but the comparison was not conclusive due to the apparently very large contaminations of the data from non nuclear effects, like e.g. quark fragmentation. The semi-inclusive $A(e,e'(A-1))X$ and $A(e,e',N_2(A-2))X$ processes, offer the possibility to investigate the LEMC effect. As a matter of fact, let us consider the cross sections (\ref{crosa-1}) and (\ref{crosa-2}) for a nucleus $A$ and the cross section (\ref{crosa-1}) for the deuteron, integrated over a certain interval of $\vec P_{A-1}$, with $\vec P_{A-1} = - \vec p_1$ in (\ref{crosa-1}), and $\vec P_{A-1} = - ( \vec p_2 + \vec P_{A-1} )$, in (\ref{crosa-2}). The following two quantities \begin{equation} R_0(x_{Bj},Q^2) = { \int_a^b \sigma_1^A( x_{Bj},Q^2,\vec P_{A-1})d \vec P_{A-1} \over \int_a^b \sigma_1^D( x_{Bj},Q^2,\vec P_{A-1}) d \vec P_{A-1}}~, \label{r0} \end{equation} \begin{equation} R_1(x_{Bj},Q^2) = { \int_a^b \sigma_2^A( x_{Bj},Q^2,\vec P_{A-2}, \vec p_2) d \vec P_{A-2} d \vec p_2 \over \int_a^b \sigma_1^D( x_{Bj},Q^2,\vec P_{A-1}) d \vec P_{A-1}} \label{r1} \end{equation} will therefore provide the LEMC effect, for they represent the contribution from weakly bound (\ref{r0}) and strongly bound (\ref{r1}) nucleons, respectively \cite{loc2}. Since the calculation of Eq. (\ref{r1}) is a bit involved, we will consider a more restricted type of LEMC, namely the separate contributions of the EMC effct from the various shells of a complex nucleus, i.e. the separate contribution of the various shells to the ratio $R_0$ \cite{loc1}. This means that we will assume that the energy resolution in the process $A(e,e'(A-1))X$ is such, that the contribution to the ratio $R_0$ due to the ground state, and to the excited states corresponding to the hole state of the target, can experimentally be separated. In what follows the $^{12}C$ nucleus will be considered assuming that DIS occured on a neutron; this means that the final nucleus to be detected is $^{11}C$ in the ground state (deep inelastic scattering on a $p$-shell neutron) and in an excited state with excitation energy of about $20MeV$ (deep inelastic scattering on a $s$-shell neutron). We have therefore calculated the ratio(\ref{r0}) using realistic Hartree-Fock momentum distributions for the $s$ and $p$ shells with single-particle energies $\epsilon_{0s}$= 36 $MeV$ and $\epsilon_{0p}$= 16 $MeV$. The results are presented in Fig.\ref{fig9}, where the usual inclusive EMC ratio, i.e. Eq.(\ref{r0}) integrated over the full space, is compared with the separate contribution from the $s$ and $p$ shells; it can be seen that, in agreement with \cite{loc2}, the $s$ shell exhibits a stronger EMC effect, but since in $^{12}C$ there are 4 $s$ shell and 8 $p$-shell nucleons, the total EMC effect is less. In what follows, we will consider the ratio $R_0$ integrated in a restricted space, viz $0<{|\vec P_{A-1}|}<2$ ${fm}^{-1}$ and $0{^o}<\theta_{\widehat {\vec P_{A-1}\vec q}} <20^o$ (the nucleus (A-1) is emitted forward) and $160{^o}<\theta_{\widehat {\vec P_{A-1}\vec q}} <180^o$ (the nucleus (A-1) is emitted backward); in Fig.\ref{fig10} the forward and backward ratios are compared with the full inclusive ratio, and it can be seen that the latter results from the sum of two almost equal contributions. In what follows, only backward emission will be considered, for this is expected to be less affected by FSI between the (A-1) nucleus and the hadrons resulting from quark hadronisation. The semi-inclusive backward ratio is shown in Fig.\ref{fig11} together with the separate contributions from the $s$ and $p$ shells; it can be seen that not only the shell contributions are well separated, but that the LEMC effect is much larger than the usual EMC effect. In order to give a flavor of the order of magnitude of the cross sections involved, these are presented in Fig. \ref{fig12} . \section{Summary and Conclusions} In the present paper, two new types of semi-inclusive DIS processes of leptons off complex nuclei, have been investigated. The first one, the process $A(e,e'(A-1))X$ , represents DIS on a shell model, low momentum and low removal energy nucleon, followed by the coherent, low momentum recoil, of the spectator nucleus $(A-1)$ in the ground, or in a low energy excited state; the second one, the processs $A(e,e'N_2 (A-2))X$, represents DIS on a nucleon $N_1$ of a correlated pair, followed by the emission of the high momentum nucleon $N_2$ of the pair, and the low momentum spectator nucleus $(A-2)$ in the ground, or in a low energy excited state. The experimental investigation of these processes would imply the coincidence detection of $e'$ and $(A-1)$, in the first case, and $e'$, $N_2$ and $(A-2)$, in the second case, respectively . We have demonstrated that both processes can provide relevant information on the following topics: $i)$ the relevance and nature of the FSI between the hadronic jet with the nuclear medium; $ii)$ the validity of the spectator model; $iii)$ the off-shell deformation of the nucleon structure function in the nuclear medium and the A-dependence of the ratio of the n/p structure functions; $iv)$ the origin of the EMC effect. As a matter of facts : $i)$ if nuclei $(A-1)$ and $(A-2)$ are detected in coincidence with the scattered electron,this is a clear signal of the absence of FSI; at the same time, the amount of observed nuclei, i.e. the cross section, will of course depend upon the FSI, therefore the investigation of its absolute value and its dependence upon $A$, would allow one to investigate the nature of the FSI, e.g. the hadronisation lenght of the hit quark; $ii)$ by a proper choice of the kinematics, the ratio of the cross section $ \sigma [A(e,e'(A-1))X]$ to the cross section $ \sigma [D(e,e'N)X]$, measured versus $|\vec P_{A-1}|$ = $|\vec p_N| \equiv |\vec p|$, at a fixed value of the Bjorken scaling variable $x_{Bj}$, has been shown to depend , within the Spectator model approach, only upon the low momentum part of the nucleon momentum distributions $n_A(|\vec p|)$ and $n_D(|\vec p|)$, and since these sharply differ for $|\vec p|\leq 1 fm^{-1}$ , the ratio should exhibit a strong $|\vec p|$ dependence (cf. Fig. \ref{fig2}), whose experimental observation would represent a stringent check of the validity of the spectator model. At the same time, the ratio of the cross section $ \sigma [A(e,e'N_2 (A-2))X]$ to the cross section $ \sigma [D(e,e'N)X]$ measured versus $|\vec p_{rel}| =|\vec p_2 +\vec P_{A-2}/2 |$ for fixed value of $|\vec P_{A-2}|$ and fixed value of $x_{Bj}$ , has been shown to depend only upon the relative momentum distributions $n^A_{rel}(|\vec p_{rel}|)$ and $n^D(|\vec p_{rel}|)$, so that the ratio should exhibit a $|\vec p_{rel}|$ dependence similar for all values of A, for $n^A_{rel} \sim n^D$ for $|\vec p_{rel}| \geq 2 fm^{-1}$ (cf. Fig. 7); again, the experimental observation of such a scaling behaviour would also represent a stringent test of the Spectator model mechanism; $iii)$ it has been shown that by a proper choice of the kinematics, the ratio of the cross sections for the same nucleus but at two different values of $x_{Bj}$, becomes independent of the nuclear quantities, being determined only by the nucleon structure function; it has therefore been demonstrated, in the case of the process $A(e,e'(A-1))X$, that such a ratio could provide significant information on different models of the structure function of weakly bound nucleons (cf. Figs. 5 and 6). Eventually (cf. Fig. 8) it has been shown that the ratio of the cross section for the process $A(e,e'N_2 (A-2))X$ to the deuteron cross section, could provide information on the {\it binding energy dependence} of the nucleon structure functions; $iv)$ the local EMC effect has been investigated (cf. Figs. 9-12), pointing out that that the processes $A(e,e'(A-1))X$ and $A(e,e'N_2(A-2))X$ integrated over a proper value of the momenta of the detected particles ( $A-1$, $N_2$ and $A-2$) will provide, for the first time, the separate contribution to the EMC ratio of the weakly and deeply bound nucleons, thus providing a stringent check of the binding model (x-rescaling) of the EMC effect. Detailed calculations have been performed for the process $A(e,e'(A-1))X$, demonstrating that in the binding model, the inclusive $EMC$ effect results from the cancellation of two large contributions from the forward and backward emitted $(A-1)$ nuclei (cf. Fig. 10); therefore, a significant check of the binding model could be provided by the measurement of the backward ratio which exhibits a 60 percent deviation from unity instead of the 10 percent deviation of the usual inclusive EMC effect (cf. Fig. 11). In closing, we would like to point out that the results we have exhibited have been obtained with non relativistic momentum distributions and spectral functions. Calculations for the two- and three-body systems including relativistic effects by a full covariant Bethe-Salpeter approach and by light-cone spectral functions, respectively, will be presented elsewhere ( \cite{bssemi}, \cite{lcsemi}). \vskip 6mm \leftline{\Large{\bf Acknowledgements}} \vskip 4mm \noindent We gratefully acknowledge Boris Kopeliovitch and Daniele Treleani for useful comments and discussions. L.P.K. thanks INFN Sezione di Perugia for warm hospitality and financial support. S.S. thanks the TMR programme of the European Commission ERB FMR-CT96-008, INFN Sezione di Perugia and Department of Physics, University of Perugia for partial financial support.
\section{Introduction} \setcounter{equation}{0} \label{sec1} Trace maps have been introduced to calculate the spectrum of 1D Kronig-Penney and related models on non-periodic structures like the Fibonacci chain \cite{KKT,Siggia}. In the sequel, these maps have found a broad variety of applications, ranging from electronic and phononic structure to 1D scattering and transport (cf.\ \cite{Kohmoto,Iochum,BJK,Bellissard} and references therein), from a spin in a magnetic field to kicked two-level systems \cite{Luck1,Graham}, and from classical to quantum spin systems \cite{Luck,Sutherland,You}. Within only a couple of years, the literature on these subjects thus became widespread and hard to survey. On the other hand, there is a considerable amount of activity on the mathematical aspects of trace maps, ranging from the systematic treatment of substitution rules \cite{Allouche}, through computer algebraic investigations \cite{Ali} to powerful analytic and algebraic results \cite{Peyriere,Iguchi,Belli1,Kramer,Frank,Sire91}, and to the treatment of trace maps as dynamical systems as such \cite{Holzer,Roberts}. Unfortunately, several if not many of the exact results have not found their way to the explicit treatment of physical systems, partially because they escaped notice and partially because they require some non-trivial mathematics like, e.g., the general gap labeling theorem for Schr\"odinger operators \cite{Belli1,Simon}. It is this gap between the theory of trace maps and their application which we would like to bridge, and we hope that the present article might prove useful for that. It is important to find a systematic, but simple formulation of the trace maps. By this we mean a formulation that is close to that of the Fibonacci case (being the best known) but brings no artificial complications for the generalizations (like singularities induced by coordinate systems). It turns out that the formulation of \cite{Allouche} is very appropriate. Here, all two-letter substitution rules (and we will only consider those) lead to trace maps that are mappings $F$ from $\mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}}^3$ to $\mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}}^3$ where the component mappings are polynomials with integral coefficients only. No poles are present as in alternative formulations with rational functions (cf.\ \cite{You}), and invariants can be described more easily. Therefore, we will follow this path and occasionally link it to other formulations. Let us now briefly outline how the remainder of this article is organized. In Sec.\ 2, we start with some basic definitions and properties of trace maps, using the algebraic approach via the free group of two generators \cite{Peyriere,Kramer}. We then formulate the invariance and transformation properties of the Fricke character $I(x,y,z) = x^2+y^2+z^2-2xyz-1$ \cite{Fricke}, a proof of which is given in the Appendix. We illustrate the results with several examples and discuss, in Sec.\ 3, the generalized Fibonacci chains in some more detail. Sec.\ 4 deals with the application to electronic structure, where we apply the gap labeling theorem to the generalized Fibonacci class. For a subclass of models, we also present an intuitive, direct approach which does not rely on the abstract result and, additionally, gives an indexing scheme of the gaps of the approximants. In Sec.\ 5, we very briefly describe the application to a kicked two-level system which follows a quasiperiodic SU(2) dynamics. Sec.\ 6 then deals with the classical 1D Ising model with coupling constants and magnetic fields varying according to the non-periodic chain. Here, the trace map gives rise to a direct iteration of the partition function which allows at least an effective numerical treatment of the case of non-commuting transfer matrices. This is followed by the concluding remarks in Sec.\ 7 and an Appendix where we prove the general transformation law of the Fricke character $I(x,y,z)$ as given in Eq.~(\ref{2.16}). \section{Trace maps and their properties} \setcounter{equation}{0} \label{sec2} To explain what a trace map of a two-letter substitution rule is, and what its main properties are in simple but general terms, it is the best to start from the {\em free group} generated by two generators $a, b$, i.e., from \begin{equation} \label{2.1} F_2 := <\!<a,b>\!> . \end{equation} Its elements consists of all possible finite words $w = w(a,b,a^{-1},b^{-1})$. $F_2$ is an infinite group, multiplication is formal composition of words, and the empty word, $e$, is the neutral element. A {\em substitution rule} $\varrho$ now is a mapping of $F_2$ into itself, where - for obvious reasons - we are only interested in homomorphisms, i.e., in $\varrho$'s with the property \begin{equation} \label{2.2} \varrho(w_1w_2) = \varrho(w_1)\varrho(w_2) . \end{equation} Here, the multiplication is the group multiplication in $F_2$. To specify a homomorphism, it is therefore sufficient to give the images of the two generators, \begin{equation} \label{2.3} {\varrho} : \begin{array}{ccc} a & \rightarrow & w_a \\ b & \rightarrow & w_b \end{array} \end{equation} where $w_a$ and $w_b$ are words in $a,b,a^{-1},b^{-1}$, for details we refer to \cite{Magnus,Kramer}. The concatenation of two homomorphisms, $\varrho_2 \circ \varrho_1$, is well defined. However, one has to {\em insert} the rule of $\varrho_1$ into that of $\varrho_2$ which is a bit against the usual practice. Therefore, we follow \cite{Peyriere} and define the multiplication \begin{equation} \label{2.31} \varrho_1\varrho_2 = \varrho_1 \cdot \varrho_2 := \varrho_2 \circ \varrho_1 . \end{equation} With this multiplication, the set of homomorphisms becomes a monoid with the identity as neutral element. We call this monoid $\Theta_2 = \mbox{Hom}(F_2)$. A special role is played by the invertible homomorphisms $\varrho$, i.e., by {\em automorphisms}. They form a group $\Phi_2$, that is also finitely (but not freely) generated \cite{Nielsen1,Neumann,Magnus} \begin{equation} \label{2.4} \Phi_2 := <\!<U,\sigma,P>\!> \end{equation} where the generators are defined by \vspace{2mm} \begin{center} \begin{tabular}{|c|c|c|c|} \hline {\rule[-3mm]{0mm}{8mm}} $w$ & $U(w)$ & $\sigma(w)$ & $P(w)$ \\ \hline \rule[-1mm]{0mm}{6mm} $a$ & $ab$ & $a^{-1}$ & $b$ \\ \rule[-3mm]{0mm}{6mm} $b$ & $b$ & $b$ & $a$ \\ \hline \end{tabular} \end{center} \vspace{2mm} For the relations amongst the generators, we refer to p.\ 164 of \cite{Magnus}. Automorphisms transform the {\em group commutator} ${\bm K}(a,b) := aba^{-1}b^{-1}$ to a conjugate either of itself or of its inverse \begin{equation} \label{2.6} \varrho({\bm K}(a,b)) = {\bm K}(\varrho(a),\varrho(b)) = w \cdot [{\bm K}(a,b)]^{\pm1} \cdot w^{-1} , \end{equation} where $w$ is an element of $F_2$ and $({\bm K}(a,b))^{-1} = {\bm K}(b,a)$. This can directly be verified by means of the generators, but has also been shown independently \cite{Nielsen2}. As we will see, this does not extend to $\varrho \not\in \Phi_2$. For the characterization of a substitution rule, it is useful to define the corresponding {\em substitution matrix}. First, one maps the elements of $F_2$ to lattice points in the square lattice, $\mbox{\sf Z\zr{-0.45}Z}^2$, by simply adding the powers of $a$ and the powers of $b$ separately in the word considered. This powercounting induces a homomorphism from $\Phi_2 = \mbox{Aut}(F_2)$ to Gl$(2,\mbox{\sf Z\zr{-0.45}Z}) = \mbox{Aut}(\mbox{\sf Z\zr{-0.45}Z}^2)$ \cite{Magnus} and, even more, from the monoid $\Theta_2 = \mbox{Hom}(F_2)$ to Hom$(\mbox{\sf Z\zr{-0.45}Z}^2) = \mbox{Mat}(2,\mbox{\sf Z\zr{-0.45}Z})$, the latter being the monoid of 2x2-matrices with integral elements. Note that Gl$(2,\mbox{\sf Z\zr{-0.45}Z})$ consists of precisely those matrices ${\bm R} \in \mbox{Hom}(\mbox{\sf Z\zr{-0.45}Z}^2)$ with det$({\bm R}) = \pm1$, and it is this determinant which determines the exponent in Eq.~(\ref{2.6}). We call ${\bm R}_{\varrho}$ the substitution matrix of the substitution rule $\varrho$, if we can read the number of $a$'s and $b$'s in the words $w_a$ and $w_b$ rowwise, i.e., \begin{equation} \label{2.7} {\bm R}_{\varrho} = \left( \begin{array}{ll} \#_a(w_a) & \#_b(w_a) \\ \#_a(w_b) & \#_b(w_b) \\ \end{array} \right) \end{equation} Note that often the transpose of ${\bm R}_{\varrho}$ is used. With our definition of the product $\varrho\sigma$, see Eq.~(\ref{2.31}), we have indeed \begin{equation} \label{2.9} {\bm R}_{\varrho\sigma} = {\bm R}_{\varrho} \cdot {\bm R}_{\sigma} , \end{equation} and, e.g., the Fibonacci rule reads $a \rightarrow b,\hspace{1mm} b \rightarrow ba \hspace{1mm}$ and leads to the matrix \begin{equation} \label{2.8} {\bm R}_{\varrho} = \left( \begin{array}{ll} 0 & 1 \\ 1 & 1 \\ \end{array} \right). \end{equation} Of course, different substitution rules can result in the same substitution matrix and the kernel of the homomorphism from $\Theta_2$ to Mat$(2,\mbox{\sf Z\zr{-0.45}Z})$ contains all inner automorphisms of $\Phi_2$: if $\varrho$ acts as $\varrho(a) = w a w^{-1}$ and $\varrho(b) = w b w^{-1}$, one has ${\bm R}_{\varrho} = \mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}}$. The kernel of the mapping from $\Phi_2$ to Gl$(2,\mbox{\sf Z\zr{-0.45}Z})$ consists of precisely these inner automorphisms \cite{Nielsen1} while there are further elements in the general case. So far, we have only set up the algebraic background. To come to the trace maps, we have to interpret the substitution in the free group as a substitution --- and thus a recursion --- in the group Sl$(2,\mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}})$ of 2x2 complex unimodular matrices. We simply insert two matrices ${\bm A}, {\bm B}$ for the letters $a, b$ and continue. But Sl$(2,\mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}})$ matrices have very special properties as a consequence of the Cayley-Hamilton theorem of linear algebra, namely \begin{equation} \label{2.10} \begin{array}{rcl} {\bm A}^2 & = & \mbox{tr}({\bm A}){\bm A} - \mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}} \\ {\bm A} + {\bm A}^{-1} & = & \mbox{tr}({\bm A})\mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}} \\ {\bm A}^n & = & U_{n-1}(x){\bm A} - U_{n-2}(x)\mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}} \end{array} \end{equation} where $x = \frac{1}{2}\mbox{tr}({\bm A})$ and $U_n(x)$ are Chebyshev's polynomial of the second kind \cite{Abramowitz} defined by: \begin{equation} \label{2.11} \begin{array}{ccl} U_{-1}(x) & \equiv & 0, \hspace{5mm} U_0(x) \equiv 1 \\ U_{n+1}(x)& = & 2xU_n(x) - U_{n-1}(x) \; . \\ \end{array} \end{equation} For later use, we run the recursion also backwards and formally obtain polynomials $U_n$ for all $n \in \mbox{\sf Z\zr{-0.45}Z}$. Also, Eq.~(\ref{2.10}) is then valid for all integers $n$. We now define \begin{equation} \label{2.12} x = \frac{1}{2}\mbox{tr}({\bm A}), \hspace{5mm} y = \frac{1}{2}\mbox{tr}({\bm B}), \hspace{5mm} z = \frac{1}{2}\mbox{tr}({\bm A}{\bm B}). \end{equation} Then, as a result of Eq.~(\ref{2.10}) and of the general identities $\mbox{tr}({\bm A} + {\bm B}) = \mbox{tr}({\bm A}) + \mbox{tr}({\bm B})$ and $\mbox{tr}({\bm A}{\bm B}) = \mbox{tr}({\bm B}{\bm A})$, we find that the traces after one iteration of an arbitrary substitution rule $\varrho$ can be expressed by $x, y, z$ again \cite{Allouche}. This is the corresponding trace map \begin{equation} \label{2.13} F_{\varrho} : \hspace{2mm} \left( \begin{array}{c} x \\ y \\ z \end{array} \right) \hspace{2mm} \rightarrow \hspace{2mm} \left( \begin{array}{c} f_{\varrho}(x,y,z) \\ g_{\varrho}(x,y,z) \\ h_{\varrho}(x,y,z) \\ \end{array} \right) \end{equation} where $f_{\varrho}, g_{\varrho}, h_{\varrho} \in \mbox{\sf Z\zr{-0.45}Z}[x,y,z]$ are three polynomials with integral coefficients. All trace maps defined that way are thus $C^{\infty}$ mappings of $\mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}}^3$ into itself - and interesting dynamical systems as such. Furthermore we find \begin{equation} \label{2.14} F_{\varrho\sigma} = F_{\varrho} \circ F_{\sigma}, \end{equation} compare also \cite{Peyriere}. The mapping from the substitution rules to the trace maps is a homomorphism between monoids. For the mapping from $\Phi_2$ to Diff($\mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}}^3$), the kernel is known, compare \cite{Roberts}. The mapping induces also a homomorphism from Gl(2,$\mbox{\sf Z\zr{-0.45}Z}$) to the trace maps, the kernel of which is $\{\pm\mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}}\}$. One of the most celebrated properties of trace maps is the invariance of the polynomial \begin{equation} \label{2.15} I(x,y,z) = x^2 + y^2 + z^2 - 2xyz - 1 \end{equation} under all trace maps that stem from an automorphism \cite{KKT,Ali,Peyriere,Kramer}. In our formulation, it follows either immediately from the Nielsen theorem (Eq.~(\ref{2.6})) (observing that ${\bm A} \in \mbox{Sl}(2,\mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}})$ implies $\mbox{tr}({\bm A}) = \mbox{tr}({\bm A}^{-1})$) or, alternatively, from the generators $U, \sigma, P$ in Eq.~(\ref{2.3}) and their trace maps by simple substitution into Eq.~(\ref{2.15}): \vspace{2mm} \begin{center} \begin{tabular}{|c|c|l|} \hline {\rule[-3mm]{0mm}{8mm}} $\varrho$ & ${\bm R}_{\varrho}$ & $\hspace{17mm}F_{\varrho}$ \\ \hline $U$ \rule[-8mm]{0mm}{18mm} & $ \left( \begin{array}{cc} 1 & 1 \\ 0 & 1 \\ \end{array} \right) $ & $ \left( \begin{array}{c} x \\ y \\ z \\ \end{array} \right) \hspace{2mm} \rightarrow \hspace{2mm} \left( \begin{array}{c} z \\ y \\ 2yz-x \\ \end{array} \right) $ \\ \hline $\sigma$ \rule[-8mm]{0mm}{18mm} & $ \left( \begin{array}{cc} -1 & 0 \\ 0 & 1 \\ \end{array} \right) $ & $\left( \begin{array}{c} x \\ y \\ z \\ \end{array} \right) \hspace{2mm} \rightarrow \hspace{2mm} \left( \begin{array}{c} x \\ y \\ 2xy-z \\ \end{array} \right)$ \\ \hline $P$ \rule[-8mm]{0mm}{18mm} & $ \left( \begin{array}{cc} 0 & 1 \\ 1 & 0 \\ \end{array} \right) $ & $ \left( \begin{array}{c} x \\ y \\ z \\ \end{array} \right) \hspace{2mm} \rightarrow \hspace{8mm} \left( \begin{array}{c} y \\ x \\ z \\ \end{array} \right) $ \\ \hline \end{tabular} \end{center} \vspace{2mm} This remarkable property appeared in the Fibonacci case in \cite{KKT}, but the theory of Fricke characters, where $I(x,y,z)$ belongs to, is much older (cf.\ \cite{Kramer} and references therein): It goes back to the last century \cite{Fricke}. But this is not the end of the story because one can show that the Fricke character follows a specific transformation law under {\em any} substitution rule, not just under automorphisms. Indeed, for $\varrho \in \mbox{Hom}(F_2)$, one finds \begin{equation} \label{2.16} I(F_{\varrho}(x,y,z)) = P_{\varrho}(x,y,z) \cdot I(x,y,z) \end{equation} where $P_{\varrho}(x,y,z) \in \mbox{\sf Z\zr{-0.45}Z}[x,y,z]$ is a polynomial with integral coefficients, called the transformation polynomial of $\varrho$. This relation was conjectured on the base of computer algebraic manipulations \cite{Ali} and proved shortly after \cite{Peyriere}. In the Appendix, we give an independent and slightly more complete proof of it. This relation, together with the general properties mentioned above, has a number of strong consequences. The reader is invited to find some himself. We list the following: \begin{itemize} \item The variety ${\cal M}_0 = \{ (x,y,z) \in \mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}}^3 \mbox{\hspace{2mm}}|\mbox{\hspace{2mm}} I(x,y,z) = 0 \}$ is an invariant surface for all trace maps. \item $(1,1,1)$ is a fixed point and $\{ (1,-1,-1), (-1,1,-1), (-1,-1,1) \}$ is an invariant set of all trace maps. \item If $(0,0,0)$ is not a fixed point of $F_{\varrho}$, its image lies on ${\cal M}_0$. \item If $\varrho \in \Phi_2$, $(0,0,0)$ is a fixed point of $F_{\varrho}$. Conversely, $F_{\varrho}(0,0,0) = (0,0,0)$ implies $P_{\varrho}(0,0,0) = 1$ and $\nabla P_{\varrho}(0,0,0) = (0,0,0)$, but not $P_{\varrho} \equiv 1$. \item $P_{\varrho}(0,0,0)$ is either $0$ or $1$. \item If $P_{\varrho}$ is constant, we have either $P_{\varrho} \equiv 1$ or $P_{\varrho} \equiv 0$. \item If $\varrho$ is of finite order, $\varrho \in \Phi_2$ and $P_{\varrho} \equiv 1$. Furthermore, one has $\varrho^{n}=e$ with $n=1,2,3,4$, or $6$ and $F_{\varrho}^{m}=\mbox{Id}$ with $m=1,2$, or $3$. \item The substitution $\varrho$ is invertible if and only if $P_{\varrho} \equiv 1$. \item The substitution $\varrho$ has nontrivial kernel if and only if $P_{\varrho} \equiv 0$. \item The substitution $\varrho$ is injective but not onto if and only if $P_{\varrho} \not\equiv$ const. \item The set $\{F_{\varrho}\; |\; \varrho\!\in\!\Phi_{2}\}$ of invertible trace maps is a group isomorphic to PGl(2,$\mbox{\sf Z\zr{-0.45}Z}$). \end{itemize} We cannot give the proofs here, many of which can be found in the work of Peyri\`ere and coworkers, compare \cite{Peyriere2} and references therein. Let us only remark that one takes profit from using trace orbits of SU(2) matrices and some standard arguments from calculus. Instead, we would like to remark that, for all trace maps, the dynamics is completely integrable on ${\cal M}_0$. Any triple $(x,y,z)$ can be obtained as traces of diagonal matrices, cf.\ Appendix. Those in turn can be written as diag$(\exp{(i\alpha)},\exp{(-i\alpha)})$, $\alpha \in \mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}}$, wherefore the multiplication is easily done by counting the exponents of the matrices involved. For any approximant, this can be expressed by the eigenvalues of the substitution matrix ${\bm R}_{\varrho}$ and is a straightforward generalization of the corresponding property of the Fibonacci chain. This way, one sets up an interesting relation to the theory of pseudo Anosov maps, see \cite{Roberts,MacKay} for details. \section{The generalized Fibonacci chains and new invariants} \setcounter{equation}{0} \label{sec3} In this chapter, we will illustrate some of the properties in the class of generalized Fibonacci chains. The latter are defined by the substitution rules \begin{equation} \label{3.1} \varrho^{(k,\ell)} : \hspace{2mm} \left( \begin{array}{l} a \rightarrow b \\ b \rightarrow b^{\ell}a^k \end{array} \right) , \hspace{5mm} {\bm R}_{\varrho}^{(k,\ell)} = \left( \begin{array}{cc} 0 & 1 \\ k & \ell \end{array} \right) \end{equation} where $k, \ell \in\mbox{\sf Z\zr{-0.45}Z}$. If the context is clear, we will frequently drop the index $(k,\ell)$. ${\bm R}_{\varrho}$ is unimodular if and only if $k = \pm1$. This coincides with $\varrho$ being an automorphism, i.e., invertible. From det$({\bm R}_{\varrho}) = -k$ we also see that $k \not= 0$ gives non-singular cases. What are the corresponding trace maps? A one-page calculation results in \begin{equation} \label{3.2} F_{\varrho} : \hspace{2mm} \left( \begin{array}{c} x \\ y \\ z \end{array} \right) \hspace{2mm} \rightarrow \hspace{2mm} \left( \begin{array}{c} y \\ g(x,y,z) \\ h(x,y,z) \\ \end{array} \right) \end{equation} with $g(x,y,z) = U_{k-1}(x)U_{\ell-1}(y) \cdot z - U_{k-1}(x)U_{\ell-2}(y) \cdot x - U_{k-2}(x)U_{\ell-1}(y) \cdot y + U_{k-2}(x)U_{\ell-2}(y)$ and $h(x,y,z) = g(x,y,z)|_{\ell \rightarrow \ell+1}$. This is equivalent to the formulation of \cite{You}. There, the iteration is given as a difference equation of the form $y_{n+1} = r(y_n,y_{n-1},y_{n-2})$ with rational function $r$. The latter has singularities as a pure consequence of the coordinates chosen. Therefore, Eq.~(\ref{3.2}) is advantageous. Another page of calculation establishes the transformation polynomial to be \begin{equation} \label{3.3} P_{\varrho}(x,y,z) = (U_{k-1}(x))^2. \end{equation} The cases $k = 1$ and $k = -1$ yield $P_{\varrho} \equiv 1$ (and hence the invariance of $I(x,y,z)$) which we know already from the invertibility of $\varrho$ in these cases (independently of $\ell$). The substitution rule $\varrho$ has nontrivial kernel for $k = 0$ (e.g., the kernel contains $a^{-\ell}b$) wherefore $P_{\varrho} \equiv 0$ is the necessary consequence. The remaining cases ($k \not\in \{-1,0,1\}, \ell \in \mbox{\sf Z\zr{-0.45}Z}$) belong to the class of injective, but non-invertible substitutions. Many properties of Sec.\ 2 can therefore be demonstrated within this class of examples. One interesting question is the existence of further invariants. That this indeed is possible can be seen from the case $\ell = 1, k = 2$ where we find \begin{equation} \label{3.4} F_{\varrho} : \hspace{2mm} \left( \begin{array}{c} x \\ y \\ z \end{array} \right) \hspace{2mm} \rightarrow \hspace{2mm} \left( \begin{array}{c} y \\ 2xz - y\\ 4xyz -2x^2 - 2y^2 + 1 \\ \end{array} \right). \end{equation} As can be checked directly, this leaves the polynomial $H$ invariant, \begin{equation} \label{3.5} H(x,y,z) = (4x^2 - 1)y - 2xz \end{equation} This could be expected from the close relationship of Eq.~(\ref{3.4}) to the period doubling map \cite{Luck2,Ali,Bovier} where an analogous invariant exists. $H$ is different from $I$ and does in fact foliate the single invariant sheet $\{ I(x,y,z) = 0 \}$, see Fig.\ 1. A closer look shows that the 1D lines on $\{ I = 0 \}$ are an immediate consequence of ${\bm R}_{\varrho}$ having an eigenvalue of modulus one. The motion of the dynamical system on $\{ I = 0 \}$ is integrable and related to torus maps. But if there is a direction neither contracted nor expanded, one could expect another invariant of the trace map $F_{\varrho}$ which foliates at least the surface $\{ I = 0 \}$ and, eventually, even the whole $\mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}}^3$. Now, our substitution matrix ${\bm R}_{\varrho}$ has eigenvalues $\lambda_{\pm} = \frac{1}{2} (\ell \pm \sqrt{\ell^2 + 4k})$. If they are rational numbers, they are automatically integers. This happens if and only if \begin{equation} \label{3.6} k = m \cdot \ell + m^2 \end{equation} with $m$ integer, where we find $\lambda_+ = \ell + m$ and $\lambda_- = -m$. Eq.~(\ref{3.4}) corresponds to $m = 1$, and it turns out that it is not a singular case. Indeed, for $k = \ell + 1$ in Eq.~(\ref{3.2}), we find the amazingly simple invariant \begin{equation} \label{3.7} H(x,y,z) = U_{\ell+1}(x) \cdot y - U_{\ell}(x) \cdot z . \end{equation} This is a generalization of Eq.~(\ref{3.5}), for a proof one needs the identities $U_{\ell}^2 - U_{\ell-1}U_{\ell+1} \equiv 1$ and $U_{\ell}U_{\ell-1} - U_{\ell+1}U_{\ell-2} \equiv U_1$. Also for the case $m = -(\ell+1)$, Eq.~(\ref{3.7}) gives the right answer, only $\lambda_+$ and $\lambda_-$ are interchanged. The remaining cases are $m = -1$ and $m = 1 - \ell$. Both result in $k = 1 - \ell$ and are equivalent. From the extension of the recursion, Eq.~(\ref{3.2}), to negative indices, we know that $U_{-(n+2)} = -U_n$ for $n \in \mbox{\sf Z\zr{-0.45}Z}$. By means of the identity $U_{l-1}^2 - U_lU_{l-2} \equiv 1$ one can show the invariance of the polynomial \begin{equation} \label{3.8} \tilde{H}(x,y,z) = U_{\ell-1}(x) \cdot y - U_{\ell-2}(x) \cdot z = -(U_{-\ell-1}(x) \cdot y - U_{-\ell}(x) \cdot z) . \end{equation} This again is a very simple expression and resembles that of Eq.~(\ref{3.7}). We think that no further invariants exist in the class of models investigated here but could not finally settle this question. We can combine the two examples shown formulating $k=1\pm\ell$ and $H^{(\pm)}(x,y,z) = \pm(U_{\pm(\ell+1)}(x)\cdot y - U_{\pm \ell}(x) \cdot z)$. This covers all cases in the class of generalized Fibonacci chains where the corresponding torus map has a unimodular eigenvalue. Let us close this section with a short remark why invariants are $\mbox{important}$. The mappings are three dimensional and could show the dynamics of 3D discrete dynamical systems. If an invariant exists, they locally look like 2D systems and inherit the well understood properties of them like period doubling routes to chaos etc., compare the Appendix of \cite{Roberts}. Now, the Fricke character is an invariant precisely for the trace maps of automorphisms, but many interesting trace maps do not stem from automorphisms. Nevertheless, period doubling can occur (as in the so-called period doubling map). The existence of invariants like $H$ or $\tilde{H}$ can explain why --- and a further investigation is in progress. \section{Electronic spectra and gap labeling} \setcounter{equation}{0} \label{sec4} Amongst the applications of trace maps to physical systems, that to electronic spectra is perhaps the most widespread one \cite{KKT,Fujiwara}. The transfer matrices attached to the different intervals describe the propagation of the amplitudes of the wave function through the chain. In the case of the continuous Schr\"odinger equation, the transfer matrices for plane waves belong to SU(1,1) \cite{BJK,Wurtz,Sueto}. The corresponding trace map acts in $\mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.05}{\sf R}}^3$ and allows the description of the spectrum, while the underlying matrix system can be used to identify the states. The best way to proceed is to tackle the full Schr\"odinger equation \begin{equation} \label{4.1} (-\frac{\hbar^2}{2m}\Delta + V(x))\Psi(x) = E\Psi(x) \end{equation} where $V(x)$ is the potential on the chain. But things are a little bit complicated because several quantities become energy dependent. Especially the very invariant $I(x,y,z)$ shows a significant dependence on the energy of the incident wave. Therefore, most authors restrict themselves to the tight-binding case, where Eq.~(\ref{4.1}) is replaced by the discrete version (in suitable units) \begin{equation} \label{4.2} ({\bm {\cal H}}\Psi)_n = \Psi_{n+1} + \Psi_{n-1} + V_n\Psi_n = E\Psi_n, \end{equation} where $I(x,y,z) = \frac{1}{4}(V_2-V_1)^2$. $V_1, V_2$ are the potentials on the two different intervals. Although one misses some typical features of the continuous case \cite{BJK}, many generic properties of the spectra remain the same, e.g., in both cases many spectra are Cantor sets with zero Lebesgue measure (see \cite{Belli1} for a criterion). The price one has to pay for going from the full Schr\"odinger equation to the tight-binding approximation is that in the latter case one needs a whole one parameter set of Hamiltonians which correspond to the single Hamiltonian in the Schr\"odinger equation \cite{Belli4}. Nevertheless, looking at the gap labeling of the spectra, we can make the restriction to the tight-binding case for simplicity without loss of information. We emphasize that all results can be extended to the continuous equation, compare \cite{Belli4}. For the tight-binding Hamiltonian of Eq.~(\ref{4.2}), the transfer matrices read: \begin{equation} \label{4.3} {\bm T}_i = \left( \begin{array}{cr} E-V_i & -1 \\ 1 & 0 \end{array} \right), \end{equation} and, therefore, \begin{equation} \label{4.4} \begin{array}{cclcl} x & = & \frac{1}{2}\mbox{tr}({\bm T}_1) & = & \frac{1}{2}(E-V_1) \\ & & & & \\ y & = & \frac{1}{2}\mbox{tr}({\bm T}_2) & = & \frac{1}{2}(E-V_2) \\ & & & & \\ z & = & \frac{1}{2}\mbox{tr}({\bm T}_2\cdot{\bm T}_1) & = & \frac{1}{2}(E-V_1)(E-V_2)-1 . \end{array} \end{equation} One could now think of a specific substitution rule $\varrho$ and investigate the spectra by iterative techniques. Here, we want to follow a slighty different path. First we want to present the famous gap labeling theorem formulated by Bellissard and coworkers. Because it uses non-trivial mathematics we will interpret it in simple words and apply the more concrete version to the well-known case of the generalized Fibonacci sequences. Afterwards, we will give an intuitive way of understanding how the gap labeling theorem works. Bellissard and coworkers proved the following theorem using techniques of $C^*$ algebras and K-Theory \cite{Belli1}: Let ${\bm {\cal H}}$ be a Hamiltonian as in (\ref{4.2}) with a bounded potential and ${\cal A}_{\bm {\cal H}}$ the $C^*$ algebra generated by the family of operators obtained from ${\bm {\cal H}}$ by translation. The possible values of the Integrated Density of States (IDOS) on the gaps of the spectrum are given by the image of the $K_0$ group of the $C^*$ algebra ${\cal A}_{\bm {\cal H}}$ of the Hamiltonian ${\bm {\cal H}}$ under the trace per unit volume, $\tau_*(K_0({\cal A}_{\bm {\cal H}}))$. This is a general result no matter what the dimension $D$ is. Thinking of the 1D case (\ref{4.2}) with a potential that takes on only finitely many different values, the gap labeling theorem tells us that the possible values of the IDOS on the gaps are given by all possible frequencies of all possible words in the infinite chain of the substitution $\varrho$. But one can express all these frequencies of words with length greater than one by the frequencies of the words of length one and two. This leads to the "concrete" gap labeling theorem \cite{Belli1}: Let ${\bm {\cal H}}$ be a Hamiltonian as in (\ref{4.2}) with the potential given by a primitive substitution, $\varrho$, on a finite alphabet, $A$. The possible values of the IDOS on the gaps in the spectrum are given by the $\mbox{\sf Z\zr{-0.45}Z}(\lambda^{-1})$ module (resp.\ the part contained in [0,1]), constructed by the components of the normalized eigenvectors $v^{(1)}$ and $v^{(2)}$ to the maximal common eigenvalue $\lambda$ of the substitution matrix for one-letter words ${\bm M}^{(1)}_{\varrho}$ and for two-letter words ${\bm M}^{(2)}_{\varrho}$. The substitution $\varrho$ is called irreducible if, for any pair of letters $a,b$ from the alphabet $A$, the word $\varrho^{k}(a)$ contains $b$ for some $k$. If $k$ can be chosen independently of the letters $a,b$, then $\varrho$ is called {\em primitive}. To guarantee the existence of a (half-)infinite word as a fixed point of $\varrho$, one considers a substitution which generates an infinite chain from every letter of the alphabet $A$. There must be at least one letter, $b \in A$ say, so that $\varrho(b)$ begins with $b$, and this letter $b$ must appear in every possible chain. Let us now apply this theorem. For the calculation of ${\bm M}^{(1)}_{\varrho}$ and ${\bm M}^{(2)}_{\varrho}$, we need the substitution of all two-letter words. Let $\varrho(w) = a_0a_1a_2...a_n$ be a substitution of a word $w$, then (with the usual notation, cf.\ \cite{Belli1}) \begin{equation} \label{4.4a} \varrho_N(w) = (a_0a_1...a_{N-1})(a_1a_2...a_N)...(a_{m-1}a_m...a_{m+N-2}) \end{equation} is the substitution of an $N$-letter word ($m$ being the total length of $\varrho(a_0)$ obtained by the powercounting described in Sec.\ 2). In particular, the two-letter substitution is: \mbox{$\varrho_2(w) = (a_0a_1)(a_1a_2)...(a_{m-1}a_m)$}. Our example is the recursion \begin{equation} \label{4.5} \begin{array}{ccl} a & \rightarrow & b \\ b & \rightarrow & b^{\ell}a^k \\ \end{array}, \end{equation} which leads to the matrix sytem: ${\bm T}_{n+1} = {\bm T}_{n-1}^k \cdot {\bm T}_n^{\ell}$, compare \cite{BJK,You}. The matrix ${\bm M}_{\varrho}^{(1)}$ is nothing but the transpose of the substitution matrix ${\bm R}_{\varrho}$, i.e., \begin{equation} \label{4.7} {\bm M}_{\varrho}^{(1)} = \left( \begin{array}{cc} 0 & k \\ 1 & \ell \end{array} \right), \hspace{5mm} \begin{array}{l} m_a = 1 \\ m_b = k+\ell \end{array} \end{equation} with eigenvalues \begin{equation} \label{4.8} \lambda_{\pm} = \frac{1}{2} (\ell \pm \sqrt{\ell^2 + 4k}). \end{equation} Let us introduce $D = k(k+\ell-1)$ and exclude the case $D=0$ ($\varrho$ is reducible for $k=0$ and does not increase the length of any word for $k+\ell=1$). Then, we can write the normalized eigenvector to $\lambda_+$ as \begin{equation} \label{4.9} v^{(1)} = \frac{1}{D} \left( \begin{array}{ccc} k(k+\ell) & - & k\lambda_+ \\ -k & + & k\lambda_+ \end{array} \right). \end{equation} Here, normalization is a statistical one, i.e., $\sum_i v_i = 1$. Also, ${\bm M}_{\varrho}^{(2)}$ is easy to calculate. We find \begin{equation} \label{4.10} \begin{array}{ccc} aa & \rightarrow & bb \\ ab & \rightarrow & b^{\ell+1}a^k \\ ba & \rightarrow & b^{\ell}a^kb \\ bb & \rightarrow & b^{\ell}a^kb^{\ell}a^k \end{array} \hspace{5mm} \Rightarrow \hspace{5mm} \begin{array}{ccl} \varrho(aa) & \rightarrow & (bb) \\ \varrho(ab) & \rightarrow & (bb) \\ \varrho(ba) & \rightarrow & (bb)^{\ell-1}(ba)(aa)^{k-1}(ab) \\ \varrho(bb) & \rightarrow & (bb)^{\ell-1}(ba)(aa)^{k-1}(ab) \end{array} \end{equation} which means \begin{equation} \label{4.11} {\bm M}_{\varrho}^{(2)} = \left( \begin{array}{cccc} 0 & 0 & k-1 & k-1 \\ 0 & 0 & 1 & 1 \\ 0 & 0 & 1 & 1 \\ 1 & 1 & \ell-1 & \ell-1 \end{array} \right) . \end{equation} The eigenvalues are \begin{equation} \label{4.12} \lambda \in \{ 0, 0, \frac{1}{2} (\ell \pm \sqrt{\ell^2 + 4k}) \} . \end{equation} The (statistically) normalized eigenvector to $\lambda_+$ reads \begin{equation} \label{4.13} v^{(2)} = \frac{1}{D} \left( \begin{array}{ccc} (k+\ell)(k-1) & - & (k-1)\lambda_+ \\ (k+\ell) & - & \lambda_+ \\ (k+\ell) & - & \lambda_+ \\ -(2k+\ell) & + & (k+1)\lambda_+ \end{array} \right), \end{equation} because of $1/\lambda_+ = (\lambda_+-\ell)/k$, and $\lambda_+^2 = k+\ell\lambda_+$. Now the calculation of the frequency module has to be done. This requires some care. First, we observe that the components of $v^{(1)}$ can be obtained by integral linear combinations of components of $v^{(2)}$, and the latter are all of the form \mbox{$\frac{1}{D}[(m((k+\ell)-\lambda_+) + n(-(2k+\ell)+(k+1)\lambda_+)], \hspace{1mm} m,n \in \mbox{\sf Z\zr{-0.45}Z}$}. This can be rewritten as \begin{equation} \label{4.13a} \frac{1}{D}(\tilde{\mu} + \tilde{\nu}\lambda_+), \hspace{5mm} \tilde{\mu}, \tilde{\nu} \in \mbox{\sf Z\zr{-0.45}Z} \end{equation} but now with the additional constraints \begin{equation} \label{4.13b} \begin{array}{ccc} \tilde{\mu} + \tilde{\nu} & \equiv & 0 \hspace{1mm} (D) \\ \tilde{\mu} + (k+\ell)\tilde{\nu} & \equiv & 0 \hspace{1mm} (D) \end{array}. \end{equation} This is necessary to guarantee $m,n \in \mbox{\sf Z\zr{-0.45}Z}$ in the preceeding expression. Remember that $1/\lambda_+ = (\lambda_+-\ell)/k$. It is easy to check that multiplication with $(\lambda_+-\ell)$ leads to new numbers $\tilde{\mu}', \tilde{\nu}'$ which also fulfil Eq.~(\ref{4.13b}). Consequently, the $\mbox{\sf Z\zr{-0.45}Z}(\lambda_+^{-1})$ module turns out to be \begin{equation} \label{4.14} \tau_*(K_0({\cal A}_{\bm {\cal H}})) = \{ \frac{1}{D}\frac{\tilde{\mu} + \tilde{\nu}\lambda_+}{k^p} \mbox{\hspace{2mm}}|\mbox{\hspace{2mm}} \tilde{\mu},\tilde{\nu},p \in \mbox{\sf Z\zr{-0.45}Z}, \;\; \begin{array}{ccc} \tilde{\mu} + \tilde{\nu} & \equiv & 0 \hspace{1mm} (D) \\ \tilde{\mu} + (k+\ell)\tilde{\nu} & \equiv & 0 \hspace{1mm} (D) \end{array} \}. \end{equation} for the continuous Schr\"odinger equation and $\tau_*(K_0({\cal A}_{\bm {\cal H}})) \cap [0,1]$ for the tight-binding case. One has to keep in mind that these are only all {\em possible} values the IDOS can take. The gap labeling theorem does not tell us whether all gaps are really open. The most prominent example where gaps are closed systematically is the Thue-Morse sequence \cite{Luck2,Belli5}. But there, the eigenvalues $\lambda_{\pm}$ are integers, and one has to check whether the frequency module really requires all linear combinations of the components of $v^{(1)}$ and $v^{(2)}$ or not. The analogous question arises for the generalized Fibonacci chains with Eq.~(\ref{3.6}) as condition, but we cannot go into details here. The special subclass of the metallic Fibonacci sequences $(k=1, \ell \in \mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.05}{\sf N}})$ has the module \begin{equation} \label{4.14a} \{ \frac{1}{\ell}(\tilde{\mu} + \tilde{\nu}\lambda_+) \mbox{\hspace{2mm}}|\mbox{\hspace{2mm}} \tilde{\mu},\tilde{\nu} \in \mbox{\sf Z\zr{-0.45}Z}, \hspace{5mm} \tilde{\mu}+\tilde{\nu} \equiv 0 \hspace{1mm} (\ell) \} , \end{equation} whereas the largest eigenvalues of the substitution matrices are given by the metallic means: \begin{equation} \label{4.15} \lambda_+ = \frac{1}{2}(\ell + \sqrt{\ell^2+4}) = [\ell;\ell,\ell,\ell,...]. \end{equation} Looking at these chains one is able to give a more heuristic but intuitive way of understanding how the gap labeling theorem works. Let us introduce the generalized Fibonacci numbers, $f_n^{(\ell)}$, by \begin{equation} \label{4.16} f_{n+1}^{(\ell)} = \ell \cdot f_n^{(\ell)} + f_{n-1}^{(\ell)}, \hspace{5mm} f_0^{(\ell)} = 0, \hspace{3mm} f_1^{(\ell)} = 1 . \end{equation} This choice of initial conditions will prove useful shortly. We observe that two consecutive numbers are coprime because the greatest common divisor obeys gcd$(f_n^{(\ell)},f_{n+1}^{(\ell)}) = \mbox{gcd}(f_n^{(\ell)},\ell \cdot f_n^{(\ell)}+f_{n-1}^{(\ell)}) = \mbox{gcd}(f_n^{(\ell)},f_{n-1}^{(\ell)})$, i.e., coprimality is inherited from the initial conditions. If we now consider the series of periodic approximants starting from $b$, i.e., \mbox{$b \rightarrow b^{\ell}a \rightarrow (b^{\ell}a)^{\ell}b$} etc., we know that these approximants are optimal in the following sense: a given approximant of length $g_n^{(\ell)}$, where \begin{equation} \label{4.17} g_n^{(\ell)} = f_n^{(\ell)} + f_{n-1}^{(\ell)}, \end{equation} is closer to the limit structure than any given other approximant of length $L \leq g_n^{(\ell)}$. One therefore expects that the structure of the IDOS of this approximant is as close to the limit IDOS as possible with approximants up to that length. In particular, the values of the IDOS on the gaps converge rapidly. It is tempting to select series of IDOS plateaus in the sequence of approximants in such a way that the corresponding gaps belong to each other (by quantum numbers, strength, symmetry, etc.). Then, if one can find a labeling, the limit $n \rightarrow \infty$ should reproduce the result of the general theorem described above. The chains with the metallic means facilitate this procedure. Here, gaps in consecutive approximants can be attached to one another by the following simple rule. Given a gap in the $n$-th approximant. Then, in the step from $n$ to $n+1$, one always grabs the closest IDOS value possible. If this is not unique (which happens only occasionally) one takes both possibilities as two branches. One branch will actually correspond to a new series of gaps, but that is hard to see in general and does not matter for the limit. Now comes the trick: the gaps in the $n$-th approximant lead to IDOS values of the form \begin{equation} \label{4.18} \frac{m}{f_{n}^{(\ell)}+ f_{n-1}^{(\ell)}}, \hspace{5mm} 0 \leq m \leq f_n^{(\ell)} +f_{n-1}^{(\ell)}, \end{equation} which is an exact result from Bloch theory for the periodic approximants. But, $m$ can be written as \begin{equation} \label{4.19} m = \mu \cdot f_n^{(\ell)} + \nu \cdot f_{n-1}^{(\ell)}, \hspace{5mm} \mu, \nu \in \mbox{\sf Z\zr{-0.45}Z}, \end{equation} because $f_n^{(\ell)}$ and $f_{n-1}^{(\ell)}$ are coprime. Taking $(\mu,\nu)$ as label, it turns out that Eq.~(\ref{4.19}) selects series of IDOS plateaus --- and hence gaps --- that follow the rule mentioned before! Furthermore, it is easy to see that neither possible gaps are missed nor index pairs are missing (as long as $\mu \cdot f_{n}^{(\ell)} + \nu \cdot f_{n-1}^{(\ell)} \in [0,f_n^{(\ell)}+f_{n-1}^{(\ell)}]$). It now is straightforward to calculate the limit points which gives the set \begin{equation} \label{4.20} \{ \frac{\lambda_+-1}{\ell}(\mu + \nu \frac{1}{\lambda_+}) \mbox{\hspace{2mm}}|\mbox{\hspace{2mm}} \mu, \nu \in \mbox{\sf Z\zr{-0.45}Z} \} \cap [0,1] . \end{equation} A similar argument was recently given for the original Fibonacci chain \cite{Liu} where this procedure is most simple. A simple substitution shows that Eq.~(\ref{4.20}) {\em precisely} reproduces the result obtained above in Eq.~(\ref{4.14a}), including the modulo condition. The labels obtained from Eq.~(\ref{4.20}) are slightly handier as can be seen in Fig.\ 2, where, as an example, we show the IDOS of the first approximants of the 'octonacci' chain, ($k = 1, \ell = 2$), and the index pairs $(\mu,\nu)$ starting with the approximants of length 3 and 7. Let us finally remark that the coincidence of the abstract and the concrete approach has a strong consequence: As far as we calculated, all gaps of the approximants involved here are open (for suitable choices of potentials). But from the second approach, it is meaningful to conjecture that all gaps are open for the class of metallic Fibonacci chains. For a proof, one could try pertubative arguments along the lines of \cite{Luck2} because these chains are quasiperiodic with embedding dimension 2, compare \cite{BJK}. This will be given elsewhere. \section{Some properties of kicked two-level systems} \setcounter{equation}{0} \label{sec5} As already mentioned in the Introduction, matrices and trace maps derived from two-letter substitution rules find applications in a variety of problems. Here, we briefly describe the case of SU(2) dynamics. The latter appears naturally in the solution of Schr\"odinger's equation for an electron or another particle with spin $s = 1/2$ in a magnetic field $\vec{B}$ \cite{Haken}. \begin{equation} \label{5.1} i \hbar \dot{\Psi} = \mu_B \vec{B} {\vec{\bm \sigma}} \Psi \end{equation} with Bohr's magneton $\mu_B = \frac{\hbar}{2m_0}$, Pauli's spin matrices ${\bm \sigma}_i$, a spinor $\Psi$ of two components, and \begin{equation} \label{5.2} \vec{B} {\vec{\bm \sigma}} = \left( \begin{array}{cc} B_3 & B_1 - i B_2 \\ B_1 + i B_2 & -B_3 \end{array} \right). \end{equation} We have written the time-dependent equation in Eq.~(\ref{5.1}) because the stationary case with constant magnetic field is completely integrable \cite{Haken} and hence less interesting than the truly time-dependent one. Here we are interested in a special class of time-dependent fields that allow the application of the recursive transfer matrix techniques, where transfer is now in {\em time}. Rewriting Eq.~(\ref{5.1}) slightly, we obtain the structural core as \begin{equation} \label{5.3} \dot{\Psi} = -i {\bm {\cal H}}(t) \Psi \end{equation} with hermitian and traceless operator ${\bm {\cal H}}(t)$. The (formal) solution is \begin{equation} \label{5.4} \Psi(t) = {\bm U}(t,t_0) \Psi(t_0) \end{equation} with \begin{equation} \label{5.5} {\bm U}(t,t_0) = \mbox{T}[\exp{(-i \int_{t_0}^t {\bm {\cal H}}(\tau)d\tau)}]. \end{equation} The time evolution operator ${\bm U}$ is a time-dependent SU(2) matrix, and T denotes time-odering \cite{Field}. Let us now think of a time sequence following the generalized Fibonacci chain, cf.~(\ref{4.5}). With ${\bm U}_n := {\bm U}(t_n,t_0)$ we obtain \begin{equation} \label{5.6} {\bm U}_{n+1} = {\bm U}_{n-1}^k \cdot {\bm U}_n^{\ell} \end{equation} for the SU(2) dynamics. The treatment of this matrix system gives full information but goes beyond the scope of the present article. For a summary of what can be done with that we refer to \cite{Kramer}. Here, we follow \cite{Sutherland} and consider the trace systems only which gives insight into the behaviour of certain observables. Let us formulate it for the Fibonacci case ($k=\ell=1$), the extension to the other examples is immediate. To be explicit in the formulas, we will consider $\delta$-kicks on the time intervals, more precisely, at the beginning of them. Any SU(2) matix ${\bm A}$ can be written in the form \begin{equation} \label{5.7} {\bm A} = \exp{(-i a {\vec{n}}{\vec{\bm \sigma}})} = \cos{(a)} \mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}} - i \sin{(a)} {\vec{n}}{\vec{\bm \sigma}} \end{equation} with $a \in \mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.05}{\sf R}}^+$ and ${\vec{n}}$ a unit vector (hence, $({\vec{n}}{\vec{\bm \sigma}})^2 = \mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}}$). From the relation \begin{equation} \label{5.8} [ {\vec{n}}{\vec{\bm \sigma}} , {\vec{n}}'{\vec{\bm \sigma}} ] = 2i \epsilon_{k \ell m} n_k n_{\ell}' {\bm \sigma}_m \end{equation} we can conclude that transfer matrices of the elementary intervals certainly commute if the kicks are in the same direction. But then, the real traces run on $\{ I = 0 \}$ and stay bounded, $|x| \leq 1, |y| \leq 1, |z| \leq 1$. The orbit belongs to the pseudo Anosov system, compare \cite{MacKay,Roberts}. The question now is what happens off this surface. Consider the kicks $\delta(t) a_j {\vec{n}}_j{\vec{\bm \sigma}}, \hspace{1mm} j = 0,1$ for the two elementary time intervals. We find \begin{equation} \label{5.9} {\bm U}^{(j)} = \exp{(-i \int_0^{t_j}\delta(t)a_j{\vec{n}}_j{\vec{\bm \sigma}} dt)} = \cos{(a_j)} \mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}} - i \sin{(a_j)} {\vec{n}}_j{\vec{\bm \sigma}} \end{equation} with \begin{equation} \label{5.10} ({\vec{n}}_0{\vec{\bm \sigma}}) \cdot ({\vec{n}}_1{\vec{\bm \sigma}}) = ({\vec{n}}_0{\vec{n}}_1) \mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}} + i ({\vec{n}}_0 \times {\vec{n}}_1) {\vec{\bm \sigma}} . \end{equation} We can work out the product ${\bm U}^{(2)} = {\bm U}^{(0)}{\bm U}^{(1)}$ and obtain \begin{eqnarray} \label{5.11} {\bm U}^{(2)} & = & \{\cos{(a_0)}\cos{(a_1)} - \sin{(a_0)}\sin{(a_1)} {\vec{n}}_0{\vec{n}}_1\} \mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}} \\ & & + i \{{\vec{n}}_0\times{\vec{n}}_1 \sin{(a_0)}\sin{(a_1)} - {\vec{n}}_0 \sin{(a_0)}\cos{(a_1)} - {\vec{n}}_1 \cos{(a_0)}\sin{(a_1)}\} {\vec{\bm \sigma}} \nonumber \end{eqnarray} This gives the initial conditions for the traces as \begin{equation} \label{5.12} \begin{array}{l} x_0 = \cos{(a_0)}, \hspace{4mm} x_1 = \cos{(a_1)}, \\ x_2 = \cos{(a_0)}\cos{(a_1)} - \sin{(a_0)}\sin{(a_1)} ({\vec{n}}_0{\vec{n}}_1) . \end{array} \end{equation} With some further calculations one can establish the formula \begin{equation} \label{5.13} I(x_0,x_1,x_2) = [({\vec{n}}_0{\vec{n}}_1)^2 - 1]\cdot(\sin{(a_0)}\sin{(a_1)})^2 \end{equation} which shows $-1 \leq I \leq 0$, as it must be for SU(2). One can now start from $I = -1$ and go to $I = 0$ by an adiabatic change of parameters. This way, one could try to follow (for the Fibonacci case, say,) the period doubling route to chaos, described recently in \cite{Roberts}. It is an interesting question whether one could experimentally find the {\em conservative} feigenvalues predicted theoretically. \section{Classical 1D Ising model with non-commuting transfer matrices} \setcounter{equation}{0} \label{sec6} Let us consider a linear chain of $N$ Ising spins \mbox{$\sigma_{j}\in\{\pm 1\}$}, $j=1,\ldots ,N$, with the energy of a configuration $\sigma = (\sigma_{1},\sigma_{2},\ldots ,\sigma_{N})$ being given by \cite{Baxter} \begin{equation} E(\sigma)\;\; = \;\; - \sum_{j=1}^{N} \left( J_{j,j+1} \sigma_{j} \sigma_{j+1}\: +\: H_{j} \sigma_{j} \right) , \label{6.1} \end{equation} where we choose periodic boundary conditions, i.e., $\sigma_{N+1} = \sigma_{1}$. The canonical partition function $Z^{(N)}$ of this system is obtained as a sum over all possible configurations $\sigma$ as follows \begin{equation} Z^{(N)}\;\; = \;\; \sum_{\sigma} \prod_{j=1}^{N} \exp \left( K_{j,j+1} \sigma_{j}\sigma_{j+1} + h_{j} \sigma_{j} \right) \label{6.2} \end{equation} with $K_{j,j+1} = J_{j,j+1}/k_{B} T$ and $h_{j} = H_{j}/k_{B} T$. Here, $T$ is the temperature and $k_{B}$ denotes Boltzmann's constant. The corresponding free energy per site $F^{(N)}$ is given by \mbox{$F^{(N)} = - \frac{1}{N} \log Z^{(N)}$} where we absorbed the factor $1/(k_{B}T)$ into the definition in order to obtain a dimensionless quantity. Eq.~(\ref{6.2}) can be rewritten in the following way (see \cite{Baxter}) \begin{eqnarray} Z^{(N)} & = & \sum_{\sigma}\; {\bm {\cal T}}^{(1,2)}(\sigma_{1},\sigma_{2})\: {\bm {\cal T}}^{(2,3)}(\sigma_{2},\sigma_{3})\: \cdots {\bm {\cal T}}^{(N,1)}(\sigma_{N},\sigma_{1}) \nonumber \\ & = & \mbox{tr}\left( {\bm {\cal T}}^{(1,2)} {\bm {\cal T}}^{(2,3)} \cdots {\bm {\cal T}}^{(N,1)} \right) \label{6.3} \\ & = & \mbox{tr}\left( {\bm {\cal T}}^{(N)} \right) \nonumber \end{eqnarray} where the elementary transfer matrix ${\bm {\cal T}}^{(j,j+1)}$ is given by the $2\!\times\! 2$ matrix \cite{Baxter} \begin{equation} {\bm {\cal T}}^{(j,j+1)} \;\; = \;\; \left( \begin{array}{ll} \exp\left( K_{j,j+1} + \frac{h_{j} + h_{j+1}}{2} \right) & \exp \left( - K_{j,j+1} \right) \\ \exp \left( - K_{j,j+1} \right) & \exp\left( K_{j,j+1} - \frac{h_{j} + h_{j+1}}{2} \right) \end{array} \right) \;\; , \label{6.4} \end{equation} while ${\bm {\cal T}}^{(N)}$ denotes the transfer matrix of $N$ sites. Note that the transfer matrices at different sites in general do not commute and therefore cannot be diagonalized simultaneously. So far we have not specified the coupling constants $K_{j,j+1}$ and the magnetic fields $h_{j}$. A particularly interesting case is that of coupling constants $K_{j,j+1} = 1$ and quasiperiodic fields $h_j$. The ground state properties have recently been analyzed exactly \cite{Sire}. Here, we now want to consider non-periodic chains where $K_{j,j+1}$ {\em and} $h_j$ are modulated according to the two-letter substitution rule $\varrho^{(k,\ell)}$ (\ref{3.1}) but taking only two possible values. To be more precise, consider a particular word $w_{n}$ ($w_{0}=a$, $w_{1}=b$, $w_{n+1}=\varrho^{(k,\ell)}(w_{n})$) with length $N_{n}=f^{(k,\ell)}_n+kf^{(k,\ell)}_{n-1}$, where the $f^{(k,\ell)}_{n}$ are defined (generalizing Eq.~(\ref{4.16})) by \begin{equation} f^{(k,\ell)}_{n+1}\;\; = \;\; \ell f^{(k,\ell)}_{n}\: +\: k f^{(k,\ell)}_{n-1} \hspace{1cm} f_{0}^{(k,\ell)} = 0, \hspace{5mm} f_{1}^{(k,\ell)} = 1 \; . \label{6.5} \end{equation} To $w_{n}$ we associate an Ising spin chain with $N_{n} = N_{n}^{(k,\ell)}$ sites in the following way: We choose $K_{j,j+1} = K^{(i)}$ and $h_{j} = h^{(i)}$ ($i=0,1$) where $i=0$ if the $j^{\mbox{th}}$ letter in the word $w_{n}$ is an $a$ and $i=1$ otherwise. Hence the chains we consider are characterized by four real parameters $K^{(i)}$ and $h^{(i)}$, $i=0,1$, and the two numbers $k, \ell \in\mbox{\sf Z\zr{-0.45}Z}$ which determine the substitution rule. Denoting the transfer matrix of the chain associated with the word $w_{n}$ by ${\bm {\cal T}}_n$, one has the following recursive equation for the transfer matrices \begin{equation} {\bm {\cal T}}_{n+1}\;\; =\;\; {\bm {\cal T}}_{n}^{\,\ell}\: {\bm {\cal T}}_{n-1}^{\, k} \label{6.6} \end{equation} with initial conditions \begin{equation} {\bm {\cal T}}_{i}\;\; =\;\; \left( \begin{array}{ll} \exp\left( K^{(i)} + h^{(i)}\right) & \exp\left( -K^{(i)}\right) \\ \exp\left( -K^{(i)}\right) & \exp\left( K^{(i)} - h^{(i)}\right) \end{array} \right) \hspace{1cm} i=0,1 \label{6.7} \end{equation} corresponding to the words $w_{0}=a$ and $w_{1}=b$. Now, we are almost in the position to use the trace map (\ref{3.2}) to obtain a recursive relation for the partition function and hence for the free energy of our chains. In order to do so, we define unimodular matrices $\tilde{\bm {\cal T}}_{n} = \frac{1}{d_{n}} {\bm {\cal T}}_{n}$ where $d_{n} = \sqrt{\det {\bm {\cal T}}_{n}}$ and use the trace map (\ref{3.2}) for \begin{equation} x_{n} \;\; = \;\; \frac{1}{2}\mbox{tr}\left( \tilde{\bm {\cal T}}_{n} \right) \;\; = \;\; \frac{1}{2d_{n}} Z_{n} \label{6.8} \end{equation} which yields a recursion relation for the partition function $Z_{n}=\mbox{tr} ({\bm {\cal T}}_{n})$ of the chain associated with the word $w_{n}$, compare \cite{Luck4} for an equivalent formulation with a nested iteration. The determinants are simply given by (see Eq.~(\ref{6.5})) \begin{eqnarray} d_{n+1} & = & d_{n}d_{n-1} \nonumber \\ & = & (d_{0})^{kf^{(k,\ell)}_{n-1}} (d_{1})^{f^{(k,\ell)}_{n}} \label{6.9} \end{eqnarray} with $d_{i}=\sqrt{2\sinh (2K^{(i)})}$, $i=0,1$. In particular, all $d_{n}$ are non-zero if $d_{0}$ and $d_{1}$ are different from zero, i.e. if both coupling constants $K^{(i)}$, $i=0,1$, do not vanish. Thus the trick we used to be able to apply the trace map (\ref{3.2}) to the transfer matrices of the Ising chain was to split the recursion relation (\ref{6.6}) for the transfer matrices into two parts: one for the trace of the transfer matrices and a rather trivial one for their determinant. Explicitly, the trace map (\ref{3.2}) becomes \begin{equation} \left( \begin{array}{c} x_{n} \\ y_{n} \\ z_{n} \end{array} \right) \hspace{4mm} \longrightarrow \hspace{4mm} \left( \begin{array}{c} x_{n+1} \\ y_{n+1} \\ z_{n+1} \\ \end{array} \right) \hspace{2mm} = \hspace{2mm} \left( \begin{array}{c} y_{n} \\ g(x_{n},y_{n},z_{n}) \\ h(x_{n},y_{n},z_{n}) \\ \end{array} \right), \label{6.10} \end{equation} compare Eq.~(\ref{3.2}), with starting values \begin{eqnarray} x_{0} & = & \frac{\mbox{tr}\left({\bm {\cal T}}_{0}\right)}{2d_{0}} \hspace{4mm} =\hspace{4mm} \frac{\exp K^{(0)} \cosh h^{(0)}} {\left(2\sinh (2K^{(0)})\right)^{1/2}} \nonumber \\ y_{0} & = & \frac{\mbox{tr}\left({\bm {\cal T}}_{1}\right)}{2d_{1}} \hspace{4mm} =\hspace{4mm} \frac{\exp K^{(1)} \cosh h^{(1)}} {\left(2\sinh (2K^{(1)})\right)^{1/2}} \label{6.11} \\ z_{0} & = & \frac{\mbox{tr}\left({\bm {\cal T}}_{0}{\bm {\cal T}}_{1}\right)}{2d_{0}d_{1}} \nonumber \\ & = & \frac{\exp (K^{(0)}+K^{(1)}) \cosh (h^{(0)}+h^{(1)}) + \exp(-(K^{(0)}+K^{(1)}))} {2\left(\sinh (2K^{(0)})\sinh (2K^{(1)})\right)^{1/2}}. \nonumber \end{eqnarray} Iterating the trace map thus corresponds to performing the thermodynamic limit $n\rightarrow\infty$. In this way the trace map (\ref{3.2}) provides an efficient way to compute the free energy in the infinite size limit numerically. Unfortunately, an analytic solution to the recursion relation is only known for the case that the starting point lies on the invariant variety ${\cal M}_0 = \{ (x,y,z) \in \mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}}^3 \mbox{\hspace{2mm}}|\mbox{\hspace{2mm}} I(x,y,z) = 0 \}$ which turns out to be equivalent to the requirement that the transfer matrices ${\bm {\cal T}}_{0}$ and ${\bm {\cal T}}_{1}$ commute. For this case, the solution is trivial since one only counts the relative frequencies of the two different couplings in the chain. It should be added that recursion relations for the magnetization (one-point function) can be derived by similar means. So far, we have not been able to extend this procedure to higher correlation functions. A renormalization group analysis (see \cite{NelsonFisher}) of the field-free case ($h^{(0)}=h^{(1)}=0$) gives the substitution matrix as linearization at the zero-temperature fixed point and hence the thermal exponent $\alpha=1$ as one would expect. Further work along these directions is in progress. \section{Concluding remarks} \setcounter{equation}{0} \label{sec7} Having described structure and applications of trace maps of two-letter substitution rules, we will briefly address further developments. It seems that the mathematical aspects are understood fairly well. Nevertheless, the trace maps as dynamical systems on their own right show interesting features that deserve further exploration: orbit structure, nontrivial symmetries, reversibility, period-doubling, and pseudo Anosov structure, compare \cite{Roberts} for some results. The physical applications could still profit from simply soaking up known exact results. A prominent example is the gap labeling where the combination of algebraic methods, generalized Bloch theory and pertubation theory improves the understanding of spectra and wave functions. For the latter, the extension to matrix maps should also be studied in more detail. A similar remark holds true of the other applications. Kicked two-level systems and spin systems are not yet brought to the understanding possible. Another, rather restrictive point is the treatment of two-letter rules: more than that is desirable. Several results are known for $n$-letter substitution rules, compare \cite{Belli1,Graham,Peyriere2,Luck3,Franz}, but the generality of the findings is lacking. Many simple properties do not extend, e.g., trace maps are higher dimensional and investigation of invariants is much more complicated. Only the gap labeling theorem still applies in full generality, wherefore further work in this direction will be needed. \section*{Acknowledgements} \label{sec8} The authors are grateful to A.\ van Elst, P.\ Kramer, J.\ A.\ G.\ Roberts, C.\ Sire, and F.\ Wijnands for discussions and helpful comments. Financial support from Deutsche Forschungsgemeinschaft and Alfried Krupp von Bohlen und Halbach Stiftung is thankfully acknowledged. M.\ B.\ would like to thank the Department of Mathematics at the University of Melbourne for hospitality where part of this work was done. It is a pleasure to dedicate this article to Peter Kramer on the occasion of his 60th birthday and to thank him this way for the support over the years. \section*{Appendix} \renewcommand{\theequation}{A.\arabic{equation}} \setcounter{equation}{0} \label{app} If $F_{\varrho}$ is the trace map of a two-letter substitution rule $\varrho \in \mbox{Hom}(F_2)$, then there is a uniquely defined transformation polynomial $P_{\varrho} \in \mbox{\sf Z\zr{-0.45}Z}[x,y,z]$ such that the Fricke character $I(x,y,z) = x^2 + y^2 + z^2 - 2xyz - 1$ shows the transformation law \begin{equation} \label{app.1} I(F_{\varrho}(x,y,z)) = P_{\varrho}(x,y,z) \cdot I(x,y,z) \end{equation} Here, the standard basis is used, i.e., $x = \frac{1}{2}\mbox{tr}({\bm A}), y = \frac{1}{2}\mbox{tr}({\bm B}), z = \frac{1}{2}\mbox{tr}({\bm A}{\bm B})$ with ${\bm A}, {\bm B} \in \mbox{Sl}(2,\mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}})$. Instead of the complex numbers, one can also use any other algebraically complete field. For the proof of this remarkable equation, the variety ${\cal M}_0 = \{(x,y,z) \in \mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}}^3 \mbox{\hspace{2mm}}|\mbox{\hspace{2mm}} I(x,y,z) = 0 \}$ plays an important role. We first observe that \begin{equation} \label{app.2} \mbox{tr}({\bm K}({\bm A},{\bm B})-\mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}}) = 4 \cdot I(x,y,z) \end{equation} for unimodular 2x2-matrices, where ${\bm K}({\bm A},{\bm B}) = {\bm A}{\bm B}{\bm A}^{-1}{\bm B}^{-1}$ is the group commutator. This follows straightforwardly from the Cayley-Hamilton theorem. We observe next that any triple $(x,y,z) \in {\cal M}_0$ can be seen as traces of diagonal matrices since we can simply choose ${\bm A} = \mbox{diag}(\alpha,\alpha^{-1}), \hspace{2mm} {\bm B} = \mbox{diag}(\beta^{\varepsilon},\beta^{-\varepsilon})$, with $\alpha = x + \sqrt{x^2 + 1}, \hspace{2mm} \beta = y + \sqrt{y^2 + 1}$, and $\varepsilon = \pm1$. Then, we find from the condition $I(x,y,z) = 0$ (a quadratic equation in $z$), the relation $2z = \alpha\beta^{\varepsilon} + 1/(\alpha\beta^{\varepsilon})$, i.e., $z = \frac{1}{2}\mbox{tr}({\bm A}{\bm B})$. The two choices of $\varepsilon$ cover the two possible solutions of the quadratic equation. Observing that $[{\bm A},{\bm B}] = 0$ implies ${\bm K}({\bm A},{\bm B}) = \mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}}$, we see from Eq.~(\ref{app.2}) that $I(x,y,z) = 0$ implies $I(F_{\varrho}(x,y,z)) = 0$. On the other hand, we know that \begin{equation} \label{app.3} P_{\varrho}(x,y,z) = \frac{I(F_{\varrho}(x,y,z))}{I(x,y,z)} = \frac{\mbox{tr}({\bm K}(\varrho({\bm A}),\varrho({\bm B}))-\mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}})} {\mbox{tr}({\bm K}({\bm A},{\bm B})-\mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}})} . \end{equation} Thus, $P_{\varrho}$ is the quotient of two polynomials with integral coefficients. The denominator, $I(x,y,z)$, is irreducible in $\mbox{\sf Z\zr{-0.45}Z}[x,y,z]$, and the numerator polynomial vanishes at least on the set ${\cal M}_0$. From this, we can conclude that the denominator divides the numerator. Since this can been seen in an elementary way, we will briefly give the argument. Consider a polynomial $Q \in \mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}}[x,y,z]$ that vanishes at least on ${\cal M}_0$. If $Q$ is of less than second order in $x, y$, or $z$, we can only have $Q \equiv 0$. To show this, assume, without loss of generality, $Q$ linear in $z$, i.e., $Q = \xi(x,y) + \eta(x,y) \cdot z$ with polynomials $\xi$ and $\eta$. For arbitrary $x, y \in \mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}}$ we can find $z_{\pm} = z_{\pm}(x,y)$ so that $(x,y,z_+) \in {\cal M}_0$ and $(x,y,z_-) \in {\cal M}_0$. Consequently, $Q$ must also vanish on these points, which gives \begin{equation} \label{app.4} z_+ \cdot \eta(x,y) = z_- \cdot \eta(x,y). \end{equation} But we can choose $(x,y)$ such that $z_+(x,y) \not= z_-(x,y)$ wherefore $\eta(x,y) = 0$. Furthermore, there is a whole $\varepsilon$-disc in $\mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}}^2$ around $(x,y)$ with $z_+(x,y) \not= z_-(x,y)$. On this disc, $\eta$ vanishes due to Eq.~(\ref{app.4}). Thus $\eta \equiv 0$ on $\mbox{\zr{0.1}\rule{0.04em}{1.6ex}\zr{-0.30}{\sf C}}^2$ and, from $Q(x,y,z_+) = 0$, also $\xi \equiv 0$. Hence, $Q \equiv 0$ under the conditions stated. Let us now consider $Q = I(F_{\varrho}(x,y,z))$, which vanishes at least on ${\cal M}_0$. If $Q$ is only linear (or less) in $z$, say, the above argument applies and we have $Q \equiv 0$ --- giving $P_{\varrho} \equiv 0$ as well. But, if $Q$ is at least quadratic in $z$, we can apply Euler's division algorithm because the ring $\mbox{\sf Z\zr{-0.45}Z}[x,y,z]$ is factorial \cite{Lang} and the denominator starts with $z^2$. We find \begin{equation} \label{a5} Q = P_{\varrho} \cdot I + r \end{equation} with $r$ a polynomial of degree $\leq 1$ in $z$. But $r$ obviously vanishes at least on ${\cal M}_0$, wherefore we must have $r \equiv 0$ according to the above given argument. We have thus rigorously and completely established that $I(x,y,z)$ devides $I(F_{\varrho}(x,y,z))$. Now, it remains to be stated that $\mbox{\sf Z\zr{-0.45}Z}[x,y,z]$ is factorial wherefore $P_{\varrho}$ is again a polynomial with integral coefficients. This proves Eq.~(\ref{app.1}). \vspace{3mm} \parindent15pt
\section*{Acknowledgments} C. M. thanks the Theory Group at SLAC for their kind hospitality and Prof.~C.~Pajares of the University of Santiago de Compostela, the Director of the research project which partially financed this work. We would also like to thank Markus Diehl for conversations.
\section{Introduction} PNC experiments in atomic physics provide an important possibility to deduce informations on the Standard Model independent of high-energy physics experiments. The recent LEP experiments \cite{l1}, that yield extremely accurate values for Z-boson properties, correspond to the resonant process. This means that all the nonresonant corrections are strongly suppressed. Hence this might not be the most convincing way for the search of all types of ``new physics'' beyond the minimal Standard Model, e.g., for the existence of a second Z-boson etc. The observation of ``new physics`` in atomic physics experiments is most probable related to processes beyond the tree level, for which electroweak radiative corrections are taken into account. This question was thoroughly discussed by many authors, in particular in \cite{l2} - \cite{l6}. Atomic PNC experiments are usually performed with heavy atoms (Cs, Tl, Pb, Bi) due to the strong enhancement of PNC effects with increasing nuclear charge number $Z$, which was first noticed by Bouchiat and Bouchiat \cite{l7}. However, the electrons involved in various PNC processes in these atoms are loosely bound valence electrons and the corresponding momentum transfer is much smaller than the squared rest mass of the electron: \begin{equation} q^{2}\ll m_{e}^{2} \;. \end{equation} The situation is different in highly charged ions (HCI), where $q^{2}\approx m_{e}^{2}$. This peculiar property of HCI was a reason for many experimental and theoretical efforts to ``test'' QED in strong fields. In particular experimental \cite{l8} - \cite{l11} and theoretical \cite{l12} - \cite{l17} attempts were made recently to verify the Lamb shift in one-electron heavy ions up to the second order in the fine structure constant $\alpha$. These attempts are yet uncomplete from both sides. Therefore there would be even a deeper reason to test the validity of the Standard Model in strong fields. We should emphasize here that the concept of the strong field is not constrained to the large momentum transfer in some particular processes. The latter often occurs in high energy physics collision experiments. During the collision period the particles are also influenced by the strong field. However, for the tightly bound electron in HCI this strong field is present for a relatively long time, defined by the lifetime of the corresponding electronic state. In this paper we show that the test of the Standard Model (or the search for ``new physics'' beyond the Standard Model) is possible in experiments with HCI. Thus PNC experiments in HCI can open a new field of research independent of the successes of PNC experiments in neutral atoms. We calculate electroweak radiative corrections to the matrix element $F_{0}=\\ \langle ns_{1/2}|{\hat H_{\rm PNC}}|n'p_{1/2}\rangle$, which represents the basis for most parity nonconserving processes studied in atomic physics of highly charged ions. Here the operator ${\hat H}_{\rm PNC}$ indicates the parity nonconserving relativistic effective atomic Hamiltonian at the tree level. The electroweak radiative corrections will be provided by $\langle ns_{1/2}|{\hat H}_{\rm PNC}^{\rm rad}|n'p_{1/2}\rangle\,\equiv \, F_{\rm SF}^{\rm rad}$. Previously radiative electroweak corrections were calculated for the case of low momentum transfer ($q^{2}=0$) or for the low field case, that is valid for neutral atoms. We are following the work by Lynn and Sandars \cite {l6} who represented their results in the factorized form \begin{equation} {\hat H}_{\rm PNC}^{\rm rad}= {\delta} {A_{\rm PNC}^{\rm rad}} \:{\hat H_{\rm PNC}} \end{equation} where the factor $\delta A_{\rm PNC}^{\rm rad}$ is independent of electron variables. This factorization is possible only in the low field case. Thus the quantity corresponding to $F_{\rm SF}^{\rm rad} $ in the low field case should be defined as \begin{equation} F_{\rm LF}^{\rm rad}=\delta A_{\rm PNC}^{\rm rad} \langle ns_{1/2}|\hat{H}_{\rm PNC}|n'p_{1/2}\rangle \;. \end{equation} The deviation of the function $f^{\rm rad}\equiv \frac{F_{\rm SF}^{\rm rad}(Z)-F_{0}(Z)}{ F_{\rm LF}^{\rm rad}(Z)-F_{0}(Z)}$ from unity for large $Z$ values will manifest the existence of strong field effects for electroweak radiative corrections in the Standard Model. The electroweak radiative corrections to the matrix element $\langle ns_{1/2}|{\hat H}_{\rm PNC}|n'p_{1/2}\rangle$ can be partitioned into the corrections $\langle ns_{1/2}|{\hat H}_{\rm PNC}^{\rm rad}|n'p_{1/2}\rangle$ to the PNC operator and to corrections to the wave functions. The corrections to the wave functions were not considered in \cite{l6}. Moreover, the full treatment requires the evaluation of radiative corrections to the total expression of the PNC atomic amplitude, including the PNC matrix element and the photon emission(absorption) matrix element. In this paper we will concentrate on radiative corrections to the PNC matrix element. In section 2 of this paper we analyse the influence of strong fields on different electroweak corrections. In section 3 we evaluate electroweak radiative loop corrections to the PNC operator. In section 4 the loop corrections to wave functions are evaluated. Section 5 contains the discussion of the numerical results and conclusions. \section{Analysis of electroweak radiative corrections in highly charged ions} In this paper we will consider only the nuclear spin-independent part of ${\hat H}_{\rm PNC}$: \begin{equation} {\hat H_{\rm PNC}}=A_{\rm PNC}\,{\gamma_{5}}{\rho_{N}(r)} \; , \end{equation} where $\gamma_{5}$ is the Dirac matrix and $\rho_{N}(r)$ is the nuclear density. In the original Bouchiat formulation \cite{l7} the constant $A_{\rm PNC}$ reads \begin{equation} A_{\rm PNC}=\frac{G_{F}Q_{W}}{2\sqrt{2}} \; , \end{equation} where $G_{F}$ is the Fermi constant and $Q_{W}$ is the ``weak charge'' of the nucleus. At the tree level $Q_{W}$ is given by \begin{equation} Q_{W}=-N+Z(1-4s^{2}) \; , \end{equation} where $ s^{2}=\sin^{2}\theta_{W}$, $\theta_{W}$ is the Weinberg angle, $N$ is the number of neutrons in the nucleus. In the following we will also utilize the equivalent Sandars definition of $A_{\rm PNC}$ \cite{l18}: \begin{equation} {A_{\rm PNC}}=\frac {\pi \alpha } {4M_{Z}^{2}} P_{W} \; , \end{equation} where \ $\alpha $ is the fine structure constant and $ M_{Z}$ is the mass of Z-boson. At the tree level we have \begin{equation} P_{W}=\frac{-N+Z(1-4s^{2})}{s^{2}(1-s^{2})}\;. \end{equation} According to Lynn and Sandars \cite{l6}, all the electroweak radiative corrections can be divided in two classes. The first class, called ``oblique'' corrections, corresponds to the Feynman graphs depicted in Fig. 1. These corrections can be incorporated into the running coupling constants, dependent on $q^{2}$. To include the ``oblique'' corrections into the PNC calculations for neutral atoms, Lynn and Sandars employ the running fine structure constant $\alpha^{*}(q^{2}=0)$, the sine of the Weinberg angle ${s^{*}}(q^{2}=0)$ and the mass of the Z-boson $ M^{2}_{Z}(q^{2}=0)$. The Fermi constant $G_{F}$ does not enter the Sandars description of PNC effects. In the case of atomic experiments $\alpha^{*}(q^{2}=0)=\alpha$, where $\alpha$ is the standard atomic value, $s$ and $M_{Z}^{*}$ are obtained from the LEP values \cite{l1} by scaling to $q^{2}=0$. The remarkable feature of Sandars description is that for all heavy elements of experimental interest $P_{W}$ is very close to $-{\frac{16}{3}}N$ and is weakly dependent on $s^{2}$ \cite{l18}. Therefore it is convenient to introduce the quantity \begin{equation} {\tilde P}_{W}=-{\frac {3}{16N}} P_{W} \;. \end{equation} Then the ''oblique'' radiative corrections can be included in ${\tilde {P}}_{W}$ \begin{equation} \tilde{P}_{W}={\tilde P}_{W}^{*}\,{(1+\delta_{P}^{M})} \; , \end{equation} where ${\tilde P}_W^{*}$ is defined from Eq. (8) with $s^{2}=s^{* 2}(q^{2}=0)=0.2394$. The correction $\delta _{P}^{M}$ results as \cite{l6} \begin{equation} \delta _{P}^{M}=\frac {M_{Z}^{2}} {[{M_{Z}^{*}}(q^{2}=0)]^{2}}-1 \simeq 0.0880 \;. \end{equation} Returning to our problem for HCI, we emphasize that we are interested to follow the change of the running constants in the interval from $q^{2}=0$ to $q^{2}=m^{2}_{e}=(0.5\ {\rm MeV})^{2}$. This interval is $10^{5}$ times smaller than the interval from $q^{2}=0$ to $q^{2}=M_{Z}^{2}=(91\ {\rm GeV})^{2}$. Therefore the value $\delta _{P}^{M}$ should be considered as field independent. Our goal is to search for the field dependent radiative corrections and to compare them with the value given by Eq. (11). In section 3 of this paper we will evaluate loop corrections directly, using the extension of the Furry picture of QED for tightly bound electrons. In the Furry picture the electrons are considered from the beginning in the external field of the nucleus. The Feynman rules for QED in the Furry picture can be found, for example, in \cite{l19}. To represent the basic atomic PNC matrix element in the Furry picture we have to consider first the Feynman graph corresponding to the exchange of a Z-boson between the atomic electron and the quark. Due to the vector current conservation the Z-boson coupling to the quarks in the case of an atom transforms to the Z-boson coupling to nucleons and to the nucleus. Thus the quark line in the Feynman graph should be replaced by the nuclear line. Moreover, the interaction of the electron with the nucleus via the exchange of a Z-boson can be replaced by the interaction with the electroweak external field described by Eq. (4). This corresponds to the Feynman graph depicted in Fig. 2. Now we can draw the Feynman graphs corresponding to loop corrections in the Furry picture of the Standard Model for tightly bound electrons. We suppose that the main contribution to the difference between the cases of $q^{2}=m^{2}_{e}$ (HCI) and $q^{2}=0$ (neutral atoms) arises from electron loops, since the heavier particle loops should be less sensitive to the strong field effect. In this respect we have to consider the graphs presented in Fig. 3 a)-d). In the evaluation of these graphs we utilize the Uehling approximation. Then the bound electron loop is approximated by the first term of the expansion in powers of the external potential (cf. Fig. 4 a)-d)). The Uehling approximation is justified even for tightly bound electrons in QED. The evaluation of electron loop corrections corresponding to Fig. 4 a), b) will be performed in section 3. In section 4 we will investigate the corrections corresponding to Fig. 4 c), d), i.e. the corrections to the wave functions. The other type of electroweak radiative corrections, called ``specific'' \cite{l6}, corresponds to Feynman graphs displayed in Figs. 5 and 6. Fig. 5 represents the contribution of the electron weak anapole moment, Fig. 6 corresponds to the vertices that describe the electromagnetic renormalization of the Z-boson coupling. According to our analysis only these graphs can contribute to the difference between the strong field and low field cases. Lynn and Sandars \cite{l6} present the electroweak radiative corrections to ${\tilde { P}}_{W}$ in the form: \begin{eqnarray} {\tilde { P}_{W}}&=&{\tilde { P}}_{W}^{*}(1+{\delta }^{M}_{P})+{\delta }^{\rm anapole}_{P} +{\delta }_{P}^{\rm vertex } \end{eqnarray} where ${\tilde { P}_{W}^{*}}$ and $\delta^{M}_{P}$ are defined by Eqs. (10) and (11), $\delta_{P}$ are specific radiative corrections to $\tilde { P}_{W}$. In Eq. (12) we omitted small field-independent ''specific'' corrections given by the ''box'' Feynman graphs \cite{l6}. We present the results of our calculations in the form \begin{equation} \Delta_{\rm rad}=\Delta{\tilde {P}_{W}}+\delta_{P}^{\rm w.f.} \end{equation} with \begin{eqnarray} \Delta {\tilde {P}_{W}} &=&\delta_{P}^{\rm loop-op} (f_{\rm loop-op}^{\rm rad}-1)\nonumber \\& &+\delta_{P}^{\rm anapole} (f_{\rm anapole}^{\rm rad}-1)+ \delta_{P}^{\rm vertex} (f_{\rm vertex}^{\rm rad}-1) \;, \end{eqnarray} where $\Delta {\tilde P_{W}} $ is the difference between the corrections for HCI and neutral atoms, $\delta_{P}^{\rm loop-op.}$, $\delta_{P}^{\rm anapole}$ and $\delta_{P}^{\rm vertex }$ denote the $q^{2}=0$ limit for the different radiative corrections to the PNC operator, and the functions $f^{\rm rad}$ are defined in the Introduction. Actually in this paper only the term $\delta_{P}^{\rm loop-w.f.}$, the most important after $\delta_{P}^{M}$, will be calculated numerically. This term represents the loop corrections to the wave functions. \section {Loop corrections to the PNC operator} We begin with the evaluation of the corrections of Fig. 4 a). To write down the corresponding matrix element we use the standard Feynman rules formulated for the QED of tightly bound electrons \cite{l19}. These rules are easily extended to the Standard Model calculations. The S-matrix element corresponding to the Feynman graph of Fig. 4 a) is given by \begin{eqnarray} S &=&(-i)^{2}g^{2}\int{d^{4}x_{1}d^{4}x_{2}d^{4}x_{3}}\,{\bar{\psi}} _{ns_{1/2}}(x_{1}){\gamma^{\mu }(\gamma_{5}-\eta)}{\psi_{n'p_{1/2}}}(x_{1}) \nonumber \\ && \times D_{\mu \nu}^{Z}(x_{1}-x_{2}) {\rm Tr}[{\gamma^{\nu}(\gamma_{5}-\eta)}S^{0}(x_{2}-x_{3}) {\gamma^{\lambda}}S^{0}(x_{3}-x_{2})]A_{\lambda}^{\rm ext}(x_{3}) \; , \end{eqnarray} where $\eta =1-4\sin^{2}\theta_{W}$, Tr corresponds to the Dirac matrices, that enter the electron loop. The Z-boson propagator $D^{Z}_{\mu \nu}$ in momentum space can be expressed as \begin{equation} D^{Z}_{\mu \nu}(k)=-\frac{{4\pi}{i} \,g_{\mu \nu}}{k^{2}-M^{2}_{Z}+i0} \;. \end{equation} We will use the expression for $D^{Z}_{\mu \nu}(x_{1}-x_{2})$ in coordinate space \begin{equation} D^{Z}_{\mu \nu }(x_{1}-x_{2})= \int{\frac{d\omega}{2\pi}} \: e^{-i \omega(t_{1}-t_{2})} {D^{Z}_{\mu \nu}} (\vec x_{1}-\vec x_{2},\omega) \end{equation} where \begin{equation} D^{Z}_{\mu \nu }({\vec x_{1}}-{\vec x_{2}},\omega )=-i 4\pi g_{\mu \nu } \int{\frac{d^{3} k}{2\pi}} \frac { e^{i {\vec k}\ ({\vec x}_{1}-{\vec x}_{2})} } {\omega^{2}-{\vec k}^{2}-M_{Z}^{2}+i 0} \;. \end{equation} The external electromagnetic field $A_{\nu}^{\rm ext}(x)$ reads \begin{equation} A^{\rm ext}_{\nu}(x)= {\delta_{\nu 0}}\,eU(\vec{x}) \end{equation} where $U(\vec {x})$ is the electric field of the nucleus (point-like or extended). $S^{0}(x-y)$ is the free electron propagator \begin{equation} S^{0}(x-y)= {\frac{1}{(2\pi)^{4}}} \int d^{4}p\, S^{0}(p) \,e^{-i p(x-y)} \end{equation} with \begin{equation} S^{0}(p)=i {\frac{p\!\!\!/+m_{e}}{p^{2}-m^{2}_{e}}}\;. \end{equation} The wave functions ${\bar \psi }_{ns_{1/2}}, {\psi_{n'p_{1/2}}}$ are the eigenvectors of the Dirac equation for the electron in the field of the nucleus \begin{equation} \psi_{n}(x)= e^{-iE_{n}t} \, \psi_{n} (\vec x) \; , \end{equation} where $E_{n}$ are the corresponding Dirac eigenvalues. \ The Standard Model constant $g$ is equal to \begin{equation} g^{2}=\frac {e^{2}}{s^{2}} \; , \end{equation} where $e$ is the electron charge. We employ the pseudoeuclidean metric with the usual metric tensor $g_{\mu \nu }$. $\gamma_{\mu }$, $\gamma_{5}$ are the usual Dirac matrices. The $S$-matrix element is connected with the amplitude by the relation \begin{equation} S_{if}={2\pi}i \, \delta (E_{i}-E_{f}) \, M_{if} \end{equation} where $M_{if}$ is the amplitude and $E_{i}$, $E_{f}$ are the initial and final state energies of the system. Transforming to the momentum space in expression (15) and integrating over the frequency variables, we obtain \begin{equation} M_{if}=-{\frac {(1-4s^{2})}{16c^{2}s^{2}}} {\frac{1}{8\pi^{2}}} \int d^3p_{1} d^3p_{2}\, {\bar \psi_{ns_{1/2}}} (\vec p_{1})\, { \frac {eU(\vec q\,)} {\vec q^{\,2}+M_{Z}^{2}} }\, \gamma_{0} \gamma_{5}\, \Pi (0,{\vec q}^{\,2}) {\psi_{n'p_{1/2}}} (\vec p_{2}) \end{equation} where ${\vec q}={\vec p}_{1}-{\vec p}_{2}$. In the expression (25) we retain only the parity violation terms. These terms contain the $\gamma_{5}$ matrix in one of the vertices connected with Z-boson. The vertex connected with the loop yields a zero result due to the identities: \begin{eqnarray} {\rm Tr} (\gamma_{\mu } \gamma_{\nu } \gamma_{5})&=&0\;,\\ {\rm Tr} (\gamma_{\mu } \gamma_{\nu } \gamma_{\alpha } \gamma_{\beta } \gamma_{5})&=& i \epsilon_{\mu \nu \alpha \beta } \;, \end{eqnarray} where $\epsilon_{\mu \nu \alpha \beta }$ is the unit antisymmetrical tensor. This tensor appears in the combination with the symmetrical product $p_{\mu }p_{\nu }$, so that \begin{equation} \epsilon_{\mu \nu \alpha \beta } \,p_{\mu } p_{\nu }=0\; . \end{equation} The function $\Pi ({\vec q}^{\:2})$ is divergent and should be renormalized. We shall use from the very beginning the known renormalized expression $\Pi_{R}(q^{2})$ (cf. for example \cite{l20}) \begin{equation} \Pi_{R} (q^{2}) ={ \frac{2i e^{2}} {2\pi^{2}} } q^{2} \int\limits_{0}^{1} dx \: x(1-x) \ln \left[1-\frac{q^{2}}{m^{2}_{e}} x(1-x)\right] \;. \end{equation} Since all the integrations in Eq. (25) after the insertion of the renormalized expression (29) are convergent, we can omit ${\vec q}^{\:2}$ in the denominator. In the case of the pure Coulomb potential with \begin{equation} U(\vec q\,)=4\pi \frac {eZ}{{\vec q}^{\:2}} \end{equation} we obtain \begin{eqnarray} M_{if} &=& {\frac {(1-4s^{2})} {2c^{2}s^{2}} } {\frac {e^{3}Z} {(2\pi )^{2}} } \frac{1} {M^{2}_{Z}} \int\limits_{0}^{1} dx \: x(1-x) \nonumber \\ & & \int d^3p_{1} d^3p_{2}\, {\bar \psi}_{ns_{1/2}} (\vec p_{1}) \gamma_{0} \gamma_{5} \ln \left[1+\frac{{\vec q}^{\:2}}{m^{2}_{e}} x(1-x)\right] \psi_{n'p_{1/2}} (\vec p_{2}) \;. \end{eqnarray} We consider first the low field limit of Eq. (31). Then ${\vec q}^{\:2}/m^{2}_{e}\ll 1$ and we can write \begin{eqnarray} M_{if}&=&- {\frac {(1-4s^{2})} {16c^{2}s^{2}} } {\frac {e^{2}Z} {\pi^{3}} } \frac{e^{2}} {m^{2}_{e}M^{2}_{Z}} \int \limits_{0}^{1}dx \: {x^{2}(1-x)^{2}} \nonumber\\ & & \times\left[ \int d^3p_{1} d^3p_{2}\, {\varphi }_{ns_{1/2}}^{+} ({\vec p}_{1}) {{\vec q}^{\:2}} {\chi }_{n'p_{1/2}} ({\vec p}_{2}) + \int d^3p_{1} d^3p_{2}\, {\chi }_{ns_{1/2}}^{+} ({\vec p}_{1}) {{\vec q}^{\:2}} {\varphi }_{n'p_{1/2}} ({\vec p}_{2}) \right] \,, \end{eqnarray} where $\varphi ,\chi $ are the upper and lower components of the Dirac wave function, respectively. Transforming to the coordinate representation we obtain \begin{eqnarray} M_{if}&=& {\frac {(1-4s^{2})} {16c^{2}s^{2}} } {\frac {e^{2}Z} {\pi^{3}} } \frac{e^{2}(2\pi )^{3}} {m^{2}_{e}M^{2}_{Z}} \int \limits_{0}^{1}dx \: {x^{2}(1-x)^{2}} \nonumber\\ & & \times\left[ \int d^3r_{1} d^3r_{2}\, {\varphi }_{ns_{1/2}}^{+} ({\vec r}_{1}) (\vec\nabla_{1}+\vec\nabla_{2})^{2} \delta ({\vec r}_{1}) \delta ({\vec r}_{2}) {\chi }_{n'p_{1/2}} ({\vec r}_{2}) \right. \nonumber\\ & & + \left. \int d^3r_{1} d^3r_{2}\, {\chi }_{ns_{1/2}}^{+} ({\vec r}_{1}) (\vec \nabla_{1}+\vec\nabla_{2})^{2} \delta ({\vec r}_{1}) \delta ({\vec r}_{2}) {\varphi }_{n'p_{1/2}} ({\vec r}_{2}) \right] \;. \end{eqnarray} Finally we find: \begin{eqnarray} M_{if}& =& \frac {1-4s^{2}} {12c^{2}s^{2}} \frac {e^{4}Z}{\pi } \frac{\pi } {m^{2}M^{2}_{Z}} \int \limits_{0}^{1}dx \; x^{2}(1-x)^{2} \left[ \nabla^{2} {\varphi }_ {ns_{1/2}}^ {+} (0) {\chi }_ {n'p_{1/2}} (0) \right. \nonumber\\ & & + {\varphi }_ {ns_{1/2}}^ {+} (0) \nabla^ {2} {\chi }_{n'p_{1/2}} (0)+2 \vec\nabla {\varphi }_ {ns_{1/2}}^ {+} (0) \cdot \vec\nabla {\chi }_{n'p_{1/2}} (0) \nonumber\\ & & +\left. {\chi }_{n'p_{1/2}} ^{+} (0) \nabla^{2} {\varphi }_ {n'p_{1/2}} (0) +2 \vec\nabla {\chi }_ {ns_{1/2}}^ {+} (0) \cdot \vec\nabla {\varphi }_{n'p_{1/2}} (0) \right] \;. \end{eqnarray} Now we compare Eq. (34) with the matrix element of ${\hat H}_{\rm PNC}$ given by Eq. (4). We can write the latter in the form \begin{equation} ({\hat H}_{\rm PNC})_{if}= { \frac {\pi \alpha } {4M_{Z}^{2}} } P_{W} \varphi_{ns_{1/2}}^{+}(0)\,\chi_{n'p_{1/2}}(0)\; . \end{equation} This comparison leads to the estimate \begin{equation} \delta_{P}^{\rm loop-op}\approx \frac{1}{15\pi } \left({-\frac {Z} {N}}\right) (1-4s^{2}) \alpha (\alpha Z)^{2} \;. \end{equation} Now we evaluate the correction of Fig. 4b. The corresponding matrix element reads \begin{equation} S = (-\imath ) e^2 \int d^{4}x_{1}d^{4}x_{2}d^{4}x_{3} \, \bar \psi_{ns1/2} (x_{1}) \gamma^{\mu } \psi_{n'p1/2} (x_{1}) D_{\mu \nu }^{\gamma } (x_{1}-x_{2}) \Pi_{R}(x_{2},x_{3}) Z^{\nu } (x_{3}) \end{equation} where $D^{\gamma }_{\mu \nu }(x_{1}-x_{2})$ is the photon propagator in Feynman gauge \begin{equation} D^{\gamma }_{\mu \nu } (x_{1}-x_{2})= \int{\frac{d\omega}{2\pi}} \: e^{-i \omega(t_{1}-t_{2})} {D^{\gamma }_{\mu \nu}} (\vec x_{1}-\vec x_{2},\omega) \end{equation} with \begin{equation} D_{\mu \nu }^{\gamma } (\vec x_{1}-\vec x_{2},\omega)= -i 4\pi \,g_{\mu\nu} \int {\frac {d^{3}k} {(2\pi )^{3}} } \frac {e^{i\vec k({\vec x}_{2}-{\vec x}_{1})}} {\omega^{2}-{\vec k}^{2}+i \epsilon} \;, \end{equation} and $Z_{\nu }$ is the external electroweak field defined by Eq. (4) \begin{equation} Z_{\nu }^{e}=\delta_{\nu 0} {\rm A}_{\rm PNC} \gamma_{5} \,\rho_{N} (\vec x\,) \;. \end{equation} It turns out that the matrix element (37) is exactly zero. Returning from Fig. 4 b) to Fig. 3 b) we expand the bound electron loop in powers of the external potential (19). This expansion will contain an increasing number of Dirac matrices $\gamma^\alpha$ together with one matrix $\gamma_5$ from Eq. (40). The trace of the product of an arbitrary number of Dirac matrices can be reduced to traces of lower products. Then, using the Eqs. (26), (27) we will obtain a zero result for an arbitrary term of the bound electron loop expansion. Thus, the correction of Fig. 3 b) is absent for the nuclear spin-independent part of $\hat H_{\rm PNC}$. \section{Loop corrections to wave functions} We start with the investigation of the graphs Fig. 4 c), d). Unlike the graphs 4 a), b) these graphs are reducible \cite{l19}. This means that the initial state of the system (the ``reference'' state) can be found among the intermediate states. The presence of the reference state in the sums over intermediate states leads to singularities that have to be avoided. For the solution of the ``reference state'' problem for the diagonal matrix element (i.e. the energy correction) the adiabatic approach of Gell-Mann and Low \cite{l21}, modified by Sucher \cite{l22} is used most frequently \cite{l19}. The extention of this approach to the nondiagonal matrix element can be most naturally formulated within the framework of the line profile QED theory \cite{l23}. Actually the graphs in Fig. 4 do not correspond to any amplitude, since the amplitude should describe some process in an atom. Still the graphs Fig. 4 a), b) are irreducible and Eq. (24) formally can be applied to them as well. \ \ In the case of the graphs 4 c), d) we have to remember that the PNC matrix element enters necessarily in some complex amplitude describing the atomic process. In the simplest case it can be the process of photon emission by an atomic electron in one-electron ions. We will consider the situation when only two levels of opposite parity $ns_{1/2}$ and $n'p_{1/2}$ are mixed by the electroweak interaction. Actually this situation does not occur in one-electron HCI, but can be found in two-electron ions \cite{l19}, \cite{l24}, \cite{l25}. \ \ Instead of the graph of Fig. 4 c) we have now to consider the graph in Fig. 7 a). The corresponding S-matrix element is given by \begin{eqnarray} S = (-i )^{3} e^{2} \int d^{4}x_{1} \, d^{4}x_{2} \, d^{4}x_{3} \, {\bar \psi }_{n''s_{1/2}}(x_{1}) \gamma^{\mu } {\rm A}_{\mu }^{(\omega )*}(x_{1}) \nonumber \\ S_{e} (x_{1},x_{2}) \gamma^{\nu } {\rm A}^{{\rm ext}}_{\nu } (x_{2}) S_{e} (x_{2},x_{3}) \gamma^{\lambda } Z_{\lambda }^{(e)} (x_{3}) \psi_{ns_{1/2}}(x_{3}) \;, \end{eqnarray} where ${\rm A}^{(\omega )*}_{\mu } (x)$ is the wave function of the emitted photon \begin{equation} {\rm A}_{\mu }^{(\omega )*}(x)= \sqrt { \frac {4\pi } {2V\omega } } {\vec e}^{\ (\omega )*} \, e^{-i (\omega t-\vec k \vec x)} \;. \end{equation} $\vec e$ is the polarization vector, $\omega$, $\vec k$ are the frequency and the wave vector of the photon. The external electromagnetic potential ${\rm A}_{\nu }^{{\rm ext}}(x)$ is \begin{equation} {\rm A}^{{\rm ext}}_{\nu } (x)= \delta_{\nu 0} {\rm V}_{U} (x) \;, \end{equation} where ${\rm V}_{U}(x)$ is the Uehling potential \cite{l26}, \cite{l27} \begin{equation} {\rm V}_{U}(x)= { \frac {2e^{3}Z} {3\pi x} } \int^{\infty }_{1} e^{-2mxy} \left( 1+ \frac{1}{2 y^{2}} \right) \frac {\sqrt{y^{2}-1}} {y^{2}} \, dy \;. \end{equation} $S_{e}(x_{1},x_{2})$ denotes the electron propagator in the external field \cite{l27} \begin{equation} S_{e}(x_{1},x_{2})= {\frac {1} {2\pi i} } \int d\omega ' \; e^{i \omega '(t_{1}-t_{2})} \sum_{m} {\frac { \psi_{m}(\vec x_{1}) {\bar \psi}_{m} (\vec x_{2}) } { E_{m} (1-i 0) +\omega^{,} }}\; . \end{equation} \ The sum over $m$ in Eq. (45) is extended over the complete spectrum of the Dirac Hamiltonian for the electron in the field of the nucleus. \ The integration over the time variables in Eq. (41) with the help of formula (24) yields \ \begin{equation} {\rm M}_{n''s_{1/2};ns_{1/2}}= -\imath e^{2} \sum_{m',m''} \frac { \langle n''s_{1/2}| {\vec e} {\vec {\rm A}^{(\omega )*}}| m' \rangle \langle m'| {\rm V}_{U}| m'' \rangle \langle m''| H_{\rm PNC}| ns_{1/2} \rangle } { (E_{m'}-E_{ns_{1/2}})(E_{m''}-E_{ns_{1/2}})} \,. \end{equation} \ There are singular terms in the sums over $m',m''$ when $m',m''=ns_{1/2}$. These singularities can be avoided by the use of the line profile theory \cite{l23}. Actually in the framework of this theory the singular terms should be omitted, but some additional terms containing derivatives of the potentials with respect to the energy can arise. In our case, due to the independence of the potentials ${\rm V}_{U},H_{\rm PNC}$ on the energy, these additional terms are absent. \ Remembering now that we assumed that only one level of opposite parity $n'p_{1/2}$ is close to the initial level $ns_{1/2}$, we set $m'=n'p_{1/2}$. Then \begin{equation} {\rm M}_{n''s_{1/2};ns_{1/2}}= e^{2} \frac { \langle n''s_{1/2}| \vec e \vec {\rm A}^{(\omega )*}| n'p_{1/2}\rangle } { E_{n'p_{1/2}}-E_{ns_{1/2}} } \sum_{ m\ne ns_{1/2}} \frac { \langle n'p_{1/2}| {\rm V}_{U}| m\rangle \langle m| H_{\rm PNC}| ns_{1/2}\rangle } { E_{m}-E_{ns_{1/2}} } \; . \end{equation} \ Performing the same calculations for the graph in Fig. 4d) (i.e. refering to the graph Fig. 8b)) and using the same assumptions we would obtain the expression \begin{equation} {\rm M}_{n''s_{1/2};ns_{1/2}}= e^{2} \frac { \langle n''s_{1/2}| \vec e \vec {\rm A}^{(\omega )*}| n'p_{1/2}\rangle } { E_{n'p_{1/2}}-E_{ns_{1/2}} } \sum_{ m\ne ns_{1/2}} \frac { \langle n'p_{1/2}| H_{\rm PNC}| m\rangle \langle m|{\rm V}_{U}| ns_{1/2}\rangle } { E_{m}-E_{ns_{1/2}} }\; . \end{equation} \ The parts of the expressions (47) and (48) containing the sums over $m$ yield evidently the corrections to the wave functions in the PNC matrix element. \ \ Note that we can consider the graph in Fig. 7a as a radiative loop correction to the photon emission matrix element in the parity violating photon emission amplitude. \ \ Moreover, apart from the graphs in Fig. 7 we have, in principle, to consider also the graph in Fig. 8 which describes exclusively the radiative corrections to the photon emission. However, in this paper we will restrict ourselves only to the corrections to the PNC matrix element. \ \ Using Eq. (41) and remembering that in the low field limit the energy denominators in Eqs. (47) and (48) are of the order $m(\alpha Z)^{2}$, we obtain the estimate \begin{equation} \delta_{P}^{\rm loop-w.f.}\approx \alpha (\alpha Z)^{2} \;. \end{equation} Comparing (49) with the estimate (36) we conclude, that the corrections to the wave functions are dominant. \section{Numerical results} \ In this paper we provide numerical results only for the leading loop corrections corresponding to the Feynman graphs in Fig. 4 c), d). \ \ These leading corrections are the corrections to wave functions discussed in section 4. The loop correction corresponding to Fig. 4 a) is suppressed by the factor $(1-4s^{2})$ in Eq. (25) and the correction corresponding to Fig. 4b is absent for the nuclear spin-independent $\hat H_{\rm PNC}$. The electron anapole moment correction is suppressed again by the factor $(1-4s^{2})$. The vertex corrections do not contain this suppression and should be compiled together with the loop corrections of Fig. 4 c), d). The vertex corrections will be treated separately in a subsequent paper. \ \ Thus, we retain here only the last term in Eq. (13). We performed the calculations for the PNC matrix element including the Uehling potential in the Dirac equation for the atomic electrons. This equation was solved numerically with the computer code published in \cite{l28}. Then we subtracted the same matrix element calculated without the Uehling potential. \ \ The results are listed in Table 1. The value of $A_M$ given in the second column is: $A_M =$ Atomic Weight. In the third column the nuclear radius is given. The fourth column in this Table represents the values for the PNC matrix element without the loop correction, the fifth column is the same matrix element calculated with the Uehling potential taken into account. \ \ An extended nucleus with an uniform charge distribution is employed throughout all the calculations. The nuclear radius $R$ is taken to be \begin{equation} R=1.2\,A_M^{1/3}\;{\rm fm} \;. \end{equation} \ In Table 2 we present the values of $\delta_{P}^{\rm loop-w.f.}$ for different $Z$ values. For $Z=92$ this correction is 7 times smaller than main correction (11) that is insensitive to the strong field. From QED calculations we know, that the vacuum polarization corrections are strongly sensitive to the field, i.e., results obtained in the low field approximation and extrapolated to the strong field case differ from the accurate strong field calculation by 100\% and more. \ \ From Tables 1, 2 we can deduce, that the strong field effect for the electroweak radiative corrections in HCI exceeds 10\symbol{"25} for $Z=92$. This is an order of magnitude higher than could be expected from the simple extrapolation of the Lynn and Sandars \cite{l6} values. The results obtained here demonstrate that the experiments with HCI would provide the possibility to test the Standard Model in the strong field. \ \ The most likely candidate for future PNC experiments with HCI is the He-like uranium ion \cite{l24}, \cite{l25}, \cite{l29}. The theory developed in the present paper allows for the evaluation of all the electroweak radiative corrections for this ion as well. \section{Acknowledgments} \looseness=-1 \ The authors are grateful to Dr. M. G. Kozlov for fruitful discussions. The work of I. B. and L. L. was supported by the Russian Fund for Fundamental Investigations, grant $N\stackrel{0}{=}99-02-18526$, the work of I. B. also by the Administration of St. Petersburg, grant $N\stackrel{0}{=}M-97-2.2d-974$. G.S. acknowledges support by the BMBF, the DFG, and by GSI (Darmstadt). \newpage \begin{table} \begin{center} \begin{tabular}{rrrrr} Z & $A_{M}$ & $R_{\rm nucl}$ & PNC & PNC(Uehl.) \\ \hline 1 & 1.007 & 1.212 & .1954019E-17 & .1954049E-17 \\ 2 & 4.001 & 1.921 & -.7939763E-15 & -.7940027E-15 \\ 3 & 6.939 & 2.307 & -.8095202E-14 & -.8095641E-14 \\ 4 & 9.010 & 2.517 & -.3254761E-13 & -.3255015E-13 \\ 5 & 10.807 & 2.675 & -.9200810E-13 & -.9201771E-13 \\ 6 & 12.007 & 2.770 & -.1960428E-12 & -.1960690E-12 \\ 7 & 14.002 & 2.916 & -.4258223E-12 & -.4258927E-12 \\ 8 & 15.995 & 3.048 & -.8348039E-12 & -.8349702E-12 \\ 9 & 18.994 & 3.228 & -.1698600E-11 & -.1698998E-11 \\ 10 & 20.173 & 3.293 & -.2637995E-11 & -.2638717E-11 \\ 20 & 40.069 & 4.140 & -.9367160E-10 & -.9374468E-10 \\ 30 & 65.363 & 4.874 & -.1014421E-08 & -.1015930E-08 \\ 40 & 91.198 & 5.446 & -.5932577E-08 & -.5946879E-08 \\ 50 & 118.662 & 5.945 & -.2623394E-07 & -.2632746E-08 \\ 60 & 144.207 & 6.345 & -.9515140E-07 & -.9562729E-07 \\ 70 & 173.001 & 6.742 & -.3259331E-06 & -.3281342E-06 \\ 80 & 200.546 & 7.082 & -.1044648E-05 & -.1053974E-05 \\ 82 & 207.155 & 7.159 & -.1325790E-05 & -.1338273E-05 \\ 90 & 231.989 & 7.434 & -.3371690E-05 & -.3410852E-05 \\ 92 & 234.993 & 7.466 & -.4167152E-05 & -.4218238E-05 \\ 92 & 238.000 & 7.498 & -.4249630E-05 & -.4301619E-05 \\ \end{tabular} \end{center} \caption{ \label{tab1} Loop correction to the wave functions in the PNC matrix element for $n=n'=2$. The values of the matrix elements are given in eV. The nuclear radius is given in units of fm. } \end{table} \newpage \begin{table} \begin{center} \begin{tabular}{rrrr} Z & $A_{M}$ & $R_{\rm nucl}$ & $\delta_{\rm P}^{\rm loop w.f.}$ \\ \hline 1 & 1.007 & 1.212 & .1528E-04 \\ 2 & 4.001 & 1.921 & .3332E-04 \\ 3 & 6.939 & 2.307 & .5414E-04 \\ 4 & 9.010 & 2.517 & .7790E-04 \\ 5 & 10.807 & 2.675 & .1044E-03 \\ 6 & 12.007 & 2.770 & .1339E-03 \\ 7 & 14.002 & 2.916 & .1653E-03 \\ 8 & 15.995 & 3.048 & .1992E-03 \\ 9 & 18.994 & 3.228 & .2343E-03 \\ 10 & 20.173 & 3.293 & .2736E-03 \\ 20 & 40.069 & 4.140 & .7801E-03 \\ 30 & 65.363 & 4.874 & .1488E-02 \\ 40 & 91.198 & 5.446 & .2410E-02 \\ 50 & 118.662 & 5.945 & .3564E-02 \\ 60 & 144.207 & 6.345 & .5001E-02 \\ 70 & 173.001 & 6.742 & .6753E-02 \\ 80 & 200.546 & 7.082 & .8927E-02 \\ 82 & 207.155 & 7.159 & .9415E-02 \\ 90 & 231.989 & 7.434 & .1161E-01 \\ 92 & 234.993 & 7.466 & .1223E-01 \\ 92 & 238.000 & 7.498 & .1225E-01 \\ \end{tabular} \end{center} \caption{\label{tab2} The wave function contribution to $\Delta_{\rm rad}$ (Eq. (13)). } \end{table} \newpage
\section{Introduction} Due to its correlation in time and location, the $\gamma$-ray burst GRB~980425 has a high probability of being associated with SN~1998bw (Galama et al. 1998). This connection is supported by the association of a radio source with SN~1998bw that requires rapid expansion (Waxman \& Loeb 1998) and probably relativistic expansion (Kulkarni et al. 1998). Recent corrections to the BeppoSAX positions and record of time variability show that the supernova falls in the error box (Piro, et al. 1998) of a time-variable X-ray source (Pian, et al. 1998). From optical spectra, SN~1998bw was classified as a SN~Ic by Patat and Piemonte (1998). What sets SN~1998bw apart from other Type Ic (SN~Ic) are higher expansion velocities as indicated by the Si II and Ca H and K lines ($\approx$ 30 to 50$\%$ higher at maximum light than SN1994I and SN1983V; Clochiatti \& Wheeler 1997), the red colors at maximum light, and the large intrinsic brightness. A peak luminosity of $1.3\pm 0.6\times10^{43}$ erg~s$^{-1}$ can be inferred from the redshift of the host galaxy ($z=0.0085$, Tinney et al. 1998) and the reddening ($A_V=0.2^m$, Schlegel et al. 1998), if we assume $H_o = $ 67 km/s/Mpc. The uncertainties are rather large as the host galaxy is not yet fully in the Hubble flow, so the peculiar velocity may be of the order of 300 to 400 km~s$^{-1}$, $H_o$ is known only to an accuracy of $\approx 10 \%$, and $A_V$ may vary by $\approx 0.1^m$. The radio source associated with SN~1998bw is the brightest ever observed for a supernova (Kulkarni, et al. 1998). The properties of SN~1998bw suggest that it was a ``hypernova" event (Paczy\'nski 1997). Based on their light curve calculations, Iwamoto et al. (1998) and Woosley, Eastman \& Schmidt (1998) d erived explosion energies of 20-50 foe and 22 foe ($1 foe = 10^{51} erg$) ejecta masses of 12-15 $M_\odot$ and 6 $M_\odot$, and $^{56}$Ni masses of 0.6-0.8 and 0.5 $M_\odot$, respectively. Guided by the deduced properties of more traditional core collapse supernovae and SN~Ic in particular, we have investigated whether asphericity can provide an alternative explanation for SN~1998bw that allows its optical properties, at least, to remain in the range of ``normal" SN~Ic. \section{Polarization of Supernovae} Both spectral analyses and light curve calculations support the picture that SN~II, SN~Ib and SN~Ic form a sequence involving core collapse with successively smaller H and He envelopes (Clochatti \& Wheeler 1997). The analysis of spectra and light curves gives, however, essentially no insight into the geometry of the expanding envelope. Polarization, on the other hand, provides a unique tool to explore asymmetries. For the last several years we have engaged in a program to obtain spectropolarimetry of as many supernovae as possible from McDonald Observatory (Wang et al. 1996). Linear polarization of $\approx 1 \% $ seems to be typical for SN~II (Wang, Wheeler \& H\"oflich 1998). There is a trend, however, for the observed polarization to increase in core-collapse supernovae with decreasing envelope mass, e.g. from SN~II to SN~Ic (Wang et al. 1998). SN 1987A, with a hydrogen envelope of about 10\ifmmode M_\odot\else M$_\odot$\fi, had a polarization of $\approx 0.5 \%$; for SN~1993J, with a hydrogen envelope mass of only a few tenths of \ifmmode M_\odot\else M$_\odot$\fi, the observed linear polarization was as high as $\approx 1.5 \% $; for the SN~Ic 1997X, with no hydrogen and little or no helium envelope, the polarization was even higher, perhaps several percent (Wang et al. 1998; Wang \& Wheeler 1998). The suggestion is that the closer one looks to the collapsing core itself, the larger the polarization and the larger the asymmetry. This trend, while tentative, clearly points toward the interpretation that the explosion itself is strongly asymmetric. From theoretical calculations of scattering dominated atmospheres, this size of polarization, $\gta 1 $ \%, requires axis ratios of the order of 2 or 3 to 1, requiring the ejecta to be highly aspherical. The luminosity $L(\Theta )$ will vary by a factor of $\approx $ 2 as the line of sight varies from the equator to the pole (H\"oflich 1991, H\"oflich et al. 1995). Given the ubiquitous presence of polarization in core collapse supernovae and especially SN~Ic, inclusion of asphericity effects in SN~Ib/c may prove to be critical to their understanding (Wang et al. 1998). Polarization was been observed in SN1998bw (Kay et al. 1998) at the level of about $0.5\%$. Asymmetries thus cannot be neglected in a complete model for this event. An obvious question is whether the asymmetries are directly connected to the fact that it produced a $\gamma$-ray burst. \section{Type Ib/c Supernovae and Gamma-Ray Bursts} Wang \& Wheeler (1998) examined the question of whether there is a general connection between supernovae and $\gamma$-ray bursts. They concluded that SN~Ia could be excluded from any such correlation at the $4\sigma$ level, that the sample of sufficiently well-studied SN~II was too small to reach meaningful conclusions, but that a correlation with SN~Ib/c could not be ruled out. Table I gives the list of candidate SN~Ib/c -- $\gamma$-ray burst\ correlations suggested by Wang \& Wheeler. \vspace{1cm} \centerline{\bf Table 1 - Supernovae of Type Ib/c and GRB Associations} \begin{table}[h] \hspace{1.5cm} \begin{tabular}{|c|c|c|c|c|} \hline SN & Type & Dates & RA (2000.0) & DEC (2000.0) \\ \hline SN 1994at & Ib/c & $\sim$961009 & 17.1 & -1.0 \\ GRB 960925 & & & 29.37 & -13.9 \\ \hline SN 1996N & Ib & $\sim$960310 & 54.7 & -26.3 \\ GRB 960221& & & 47.8 & -31.24 \\ \hline SN 1997B& Ic & $\sim$970104 & 88.3 & -17.9 \\ GRB 971218& & & 97.75 & -21.73 \\ \hline SN 1997dq& Ib & $\sim$971105 & 175.2 & +11.5 \\ GRB 971013& & & 167.0 & +2.7 \\ \hline SN 1997ef& ? & $\sim$971205 & 119.3 & +49.6 \\ GRB 971152& & & 84.6 & +41.7 \\ \hline SN 1997ei& Ic & $\sim$971223 & 178.5 & +58.5 \\ GRB 971120& & & 155.7 & +76.4 \\ \hline SN 1998bw & ? & 980424-980427 & 19 35 03.3 & -52 50 44.8\\ GRB 980425& & 980425.909 & 19 35 21 & -52 52 19\\ \hline \end{tabular} \end{table} Kippen et al. (1998) found no evidence for a connection between supernovae and $\gamma$-ray bursts\ and Graziani, Lamb, \& Marion (1998) concluded that not all SN~Ib/c could be associated with $\gamma$-ray bursts, although a fraction might be. With the small sample of SN~Ib/c and partial sky coverage, it will be difficult to resolve this issue with statistics alone. Woosely, Eastman, \& Schmidt (1998) suggested that SN~1998cy might be associated with GRB~9970514. Unlike the suggestions in Table I, SN~1998cy shows both narrow and broad features of hydrogen. This event was even brighter than SN~1998bw (Schmidt, 1998). If it is correlated with a $\gamma$-ray burst\ it represents another type of exploding object. When it was first observed, SN~1997ef was noted to be an object of unprecedented spectral features, although it bore some resemblance to SN~Ic. The similarity of SN~1997ef to SN~1998bw was pointed out by Garnavich (1997). Wang \& Wheeler found that SN~1997ef was just beyond the $2\sigma$ BATSE error box for GRB~971115. A model for the light curve of SN~1997ef has been presented by Iwamoto et al. (1999) based on a carbon/oxygen core exploding with a ``normal" kinetic energy. At this meeting, Nomoto reported a revised calculation based on a ``hypernova" scenario with in excess of 10 foe of kinetic energy that better matched the line profiles. The question of whether this event was polarized and hence asymmetric also arises. Garnavich, Jha, \& Kirshner (1998) report that SN 1998ey was ``identical" to SN~19978ef two weeks after discovery and similar to SN~1998bw. They report no correlation with a BATSE trigger, but again, BATSE only surveys a portion of the sky at one instant and so the constraint on a single event with the uncertainty in explosion much greater than the orbit of CGRO is weak. \section{Description of the Concept and Numerical Methods} We have investigated the question of whether the optical properties of SN~1998bw can be explained based on an asymmetric explosion with otherwise normal amounts of kinetic energy and total ejecta and \noindent\ mass as opposed to the unprecedentedly large values of these quantities required in the ``hypernova" models of Iwamoto et al. (1998) and Woosely, Eastman, \& Schmidt (1998). For the initial setup, we use the chemical and density structures of spherical C/O cores of Nomoto and Hashimoto (1988). These structures are scaled to adjust the total mass of the ejecta. This is an approximation, but the details of the chemical profiles are not expected to effect the light curves. The explosion models are calculated using a one-dimensional radiation-hydro code which includes a detailed nuclear network. The code also simultaneously solves for the radiation transport via moment equations. Photon redistribution and thermalization is based on detailed NLTE-models. Several hundred frequency groups are used to calculate monochromatic light curves, frequency-averaged Eddington factors, and opacity means. A Monte Carlo scheme is used for $\gamma$-rays. For details, see H\"oflich, Wheeler \& Thielemann (1999) and references therein. Aspherical density structures are constructed based on the original spherical density distribution. For the simple models presented here, we impose the asymmetry after the ejecta has reached the homologous expansion phase. We generate an asymmetric configuration by preserving the mass fraction per steradian from the spherical model, but imposing a different law of homologous expansion as a function of the angle $\Theta$ from the equatorial plane. For typical density structures, a higher energy deposition along the polar axis results in oblate density structures. Such an energy pattern may be produced if jet-like structures are formed during the central core collapse as suggested by Wang \& Wheeler (1998). In contrast, a prolate density structure would be produced if more energy is released in the equatorial region than in the polar direction. For more details, see H\"oflich, Wang \& Wheeler (1999). The emissivity along a given isodensity contour in the asymmetric models is taken to be constant and equal to the corresponding equivalent velocity-weighted mean layer in the spherical model. The bolometric and broad band light curves are constructed by convolving the spherical light curves with the photon redistribution functions, L$(\Theta)$/L(mean), that are calculated by the Monte Carlo code for polarization (H\"oflich 1991, H\"oflich et al. 1995). Typical conditions at the photosphere, and therefore the colors, are expected to be similar in both the spherical and aspherical models because the energy flux, $F(\Theta)$, is found to be similar both in the spherical and aspherical configuration to within $\approx 40 \% $. Stationarity is assumed to calculate the photon redistribution functions since the geometry does not change during the typical diffusion time scale. We assume implicitly the same mean diffusion time scales for both the spherical and aspherical configurations. This mostly effects the very early phases of the light curve when the hydrodynamical time scales are short. For more details, see H\"oflich, Wang \& Wheeler (1998). \section{Results} We construct models in such a way a way that at day 20 the axis ratio at the photosphere is 2. In comparison to the spherical model, the homology expansion parameters are a factor of $\approx 2.2 $ larger along the pole for oblate ellipsoids and a factor of $\approx 1.5$ larger for prolate ellipsoids along the equator. We first calculated aspherical light curves based on the C/O core CO21 which gives a good representation of the BVRI light curves of the SN~Ic 1994I (Iwamoto et al. 1994). The ejecta mass is $0.8 M_\odot $ and the explosion energy is $E_{kin} = 10^{51} erg $. A mass of $0.08~ M_\odot $ of $^{56}Ni$ is ejected. This model failed to produce the peak brightness by a factor of 2, gave too short a rise time by about 5 days, and shows blue color indices at maximum light. To boost the total luminosity to the level of observation we increased the amount of ejected $^{56}Ni$ to 0.2 $M_\odot$. This quantity of nickel is still below the estimates for $^{56}Ni$ of 0.3 $M_\odot$ in the bright SN~II 1992am (Schmidt et al. 1997). As shown in Fig.1, the time of maximum light is rather insensitive to asphericity effects. The need to delay the time to maximum and to produce the red color at maximum light suggested the need to increase the ejecta mass with an appropriate increase in the kinetic energy to provide the observed expansion. We thus computed a series of models with $M_{ej}=2 M_\odot$ $E_{kin}=2\times10^{51}$ erg and $M_{Ni} = 0.2$\ifmmode M_\odot\else M$_\odot$\fi. \begin{figure}[t] \psfig{figure=wheeler.fig1.ps,width=13.0cm,rwidth=12.0cm,clip=,angle=270} \caption{Directional dependence of the bolometric light curve for oblate (left) and prolate (right) ellipsoids. The luminosity of the corresponding spherical model is shown as $L(mean)$. In addition, the instantaneous $\gamma-ray$ deposition is shown. $P(0^o)$ is the polarization at maximum light ($ P(\Theta) \approx P(0^o) \times cos^2\Theta$).} \end{figure} Asphericity of the amplitude assumed here can change the luminosity over a range of roughly 2 magnitudes (Fig. 1). For oblate ellipsoids, the luminosity is enhanced along the pole whereas for prolate density structures the enhancement occurs in the equatorial direction. Combined with the polarization properties, this provides a clear separation between oblate and prolate geometries as $P$ always goes to 0 if the structure is seen pole-on and $P$ increases towards lower latitudes (H\"oflich 1991). \begin{figure}[t] \psfig{figure=wheeler.fig2.ps,width=13.0cm,rwidth=12.0cm,clip=,angle=270} \caption{Comparison of bolometric and broad-band light curves as observed for SN1998bw with those of an oblate ellipsoid seen at high angle with respect to the equator assuming a distance of 36 Mpc and $A_V=0.2^m$.} \end{figure} Observations of the polarization of SN~1998bw 23 days after the explosion show little polarization ($< 1 \% $, Patat et al. 1998). By day 58, the intrinsic polarization was reported to be $0.5 \%$ (Kay et al. 1998). Polarization data on SN~Ic is rare, but this value is less than seen in some SN~Ic and related events (see above). SN~1998bw was also rather bright. This combination implies oblate geometries if asymmetry is involved. For the comparison between the observed and theoretical light curves as shown in Fig. 2, we have used the relative calibration of Woosley, Eastman, \& Schmidt (1998) for the ``bolometric light curve". The broad-band data was obtained from Galama et al (1998). To account for the obervations with our asymmtric model, the object must be seen from an angle of $\ge 60^o$ from the equator. The same conclusion can be drawn independently from the detected, but relatively small linear polarization. This model thus suggests that the maximum polarization of SN~1998bw was greater than the $0.5 \%$ measured by Kay et al. (1998). Overall, the broad-band light curves agree with the data within the uncertainties. The intrinsic color excess B-V matches the observations within $0.1^m$ and, after the initial rise of $\approx 7$ days, the agreement in each band is better than $0.3^m$. The main discrepancy with the observations occurs during the initial rise when the diffusion time scales are much longer than the expansion time. Under these conditions, our approximation for redistribution of the energy of a spherical model breaks down since the diffusion time scale is long compared to the hydrodynamical time scale. The decline after maximum is slightly too steep in the models both in the bolometric and broad band light curves. This is likely to be related to the energy generation in the envelope by $\gamma$-ray\ deposition or to the change in the escape probability of low energy photons. The decline rate immediately after peak can be reduced by increasing the amount of radioactive $^{56}Ni$ by $\approx 40 \%$. In our models, the escape probability for $\gamma$-rays increased rapidly between day 20 and 80 from 3\% to 50\%. An alternative means to flatten the light curve is to reduce the increase in escape probability. This can be achieved by a modification restricted to the inner layers of the ejecta because the escape probability is determined by those layers. Either the expansion velocity of the inner layers can be reduced or the density gradient might be steeper. Both are expected for strongly aspherical explosions. \section{Discussion and Conclusions} The ``hypernovae" models for SN~1998bw provide interesting fits to the data, but these models have some problems. Although the fit of Iwamoto et al. (1998) of the light curve is excellent with errors $\leq 0.3^m $ over 40 days, the spectra show absorption lines that are too narrow by a factor of 2 to 3 indicating too narrow a range of formation in velocity space. This may be related to the high envelope mass. In the lower mass models of Woosley, Eastman, \& Schmidt (1998), the computed color indices (B-V, V-R, V-I) are too red at all epochs. The hypernova models presented to date are spherically symmetric and hence make no pretense of accounting for the observed polarization, but this is a critical feature and cannot be ignored. We have shown that the high apparent luminosity of SN~1998bw may be understood within the framework of ``classical" SNIc by invoking asymmetry of the ejecta of the degree required to account for the polarization. Even with the invocation of significant asymmetry, our model for SN~1998bw remains at the bright end of the scale for normal SN~Ic. We note that the luminosity of SN~1998bw may be uncertain by a factor of 2 due to non-Hubble motion within the cluster and uncertainties in the Hubble constant and reddening. For a model with an ejected mass of 2 $M_\odot$, an explosion energy of $2\times 10^{51}$ erg, and a $^{56}Ni$- ejection of $0.2 M_\odot$, both the bolometric and broad-band light curves are rather well reproduced by an oblate ellipsoid with an axis ratio of 2 to 1 which is observed within $30\deg$ of the symmetry axis. This angle for the line of sight is consistent with the low (but still significant) polarization observed for SN~1998bw. In a Lagrangian frame, the polar expansion velocity is a factor of 2 larger than the mean velocity. This is also in agreement with the rather large expansion velocities seen in SN~1998bw. Neither this model nor the hypernova models give an obvious explanation of the $\gamma$-ray burst\ nor the especially bright radio emission. Woosley, Eastman, \& Schmidt (1998) have analyzed the possibility of $\gamma $-ray bursts in the framework of spherical models. Even with their explosion energies of more than 20 foe they showed that the $\gamma$-ray burst\ associated with SN~1998bw/GRB~980425 cannot be explained by the acceleration of matter to relativistic speeds at shock-breakout. In our picture, the specific energy released in the polar region is comparable. We want to stress, however, that the asymmetry in the energy distribution in the model presented here is set after homologeous expansion has been established. Because the early hydrodynamical evolution will tend to wipe out asymmetries, the initial anisotropy in the energy distribution as the explosion is initiated after core collapse is expected to be significantly higher. We have shown that SN~1998bw may be understood within the framework of ``classical" core-collapse supernovae rather than by a ``hypernova," but the actual model parameters must be regarded as uncertain both because of the model assumptions and the uncertainty in the observed luminosity. In light of the good fits, however, SN~1998bw may indeed be a ''hypernova." Continuous measurements of the polarization and the velocity of $^{56}Co$ lines are critical to unreveal the nature and geometry of this object. If the late-time tail tracks the \co\ decay line, then the ejected \noindent\ mass could be determined. This might prove the simplest discriminant between the ``hypernova" models and models based on significant asymmetry. There remains the broader issue of the possible connection of supernovae and $\gamma$-ray bursts. If some of the $\gamma$-ray bursts\ at large redshifts are driven by supernovae, then the associated energy flow must be strongly collimated and aspherical core collapse may be a natural way to produce such jets. The basic purpose of this work is to underline the fact that asymmetries must be taken into account in a complete consideration of core collapse supernovae, of SN~1998bw in particular, and, by extension of $\gamma$-ray bursts. \acknowledgements We thank Ken Nomoto for providing us with the monochromatic light curve data in digital form and Titus Galama, Brian Schmidt, Ken Nomoto, Stan Woosely, Martin Rees, Ralph Wijers, Dieter Hartmann, and Peter M\'esz\'aros for valuable conversations. This research was supported in part by NSF Grant AST 9528110, NASA Grant NAG 5-2888, and a grant from the Texas Advanced Research Program.
\section{Introduction} Clearly the missing-mass problem is central to astronomy today since all attempts to understand the growth of structure in the universe, as well as the dynamics of galaxies and their clusters, require an understanding of the nature and location of this dominant mass component (\cite{car94}). Because most of the inner-halo galactic missing-mass is probably in compact objects, attempts to detect it have been based upon quasar-microlensing and star-microlensing which can detect objects of any mass but cannot detect a smooth distribution of matter. The first published quasar-microlensing reports were from lensed quasars by \cite{van89} and \cite{irw89}, although certainly the high optical depth Q0957+561A,B quasar-microlensing reported by Vanderriest et al. was known by \cite{sch86} who did not presume to discuss microlensing until their measured time delay difference of 1.1 years between the A and B quasar images was confirmed. A secure image time delay difference is required to allow correction for intrinsic fluctuations of the source brightness at $0.1\--0.3$ year planetary-mass periods by subtraction of the image light curves matched at the time their light was lensed. The 1.1 year \cite{sch86} delay value has been confirmed and is now generally accepted, \cite{kun97}. Star-microlensing searches at low optical depth in directions of large star densities toward the Large Magellanic Cloud (LMC) and the Galactic Bulge resulted in detections (\cite{alc93}, \cite{aub93}, \cite{uda93}). However the interpretation of the microlensing statistics from the two kinds of programs has produced disagreement. The quasar-microlensing in Q2237 was first to be analyzed, and earliest reports suggesting solar mass stars (\cite{wam90}, \cite{cor91}, \cite{yee92}) were later amended by \cite{ref93} who demonstrated that the apparently rapid fluctuations detected in the observations could indicate a dominant population of terrestrial mass microlenses, but called for more accurate lightcurves over longer time spans to make the indication conclusive. Interpretation of the Q0957 microlensing was delayed by a long controversy about the time delay, but the critical fact that rapid microlensing is observed was already noted by \cite{sch91}. Exhaustive statistical analyses of the Q0957 data caused \cite{sch95} to again conclude that rapid microlensing is observed, and a power spectrum by \cite{tho96}, also shown in \cite{sch96}, led to the Schild conclusion that the microlensing is caused by ``rogue planets ... likely to be the missing mass.'' A primordial origin for such micro-brown-dwarf (MBD) objects as the baryonic galactic dark matter had already been proposed by \cite{gib96}, who predicted a condensation mass of about $10^{-7}$ $M_{\sun}$ for the relatively inviscid gas formed at photon decoupling from the cooling viscous plasma. Recent best estimates from fossil turbulence theory by \cite{gib00a}bc of $10^{-6}$ $M_{\sun}$ for the ``Primordial Fog Particle'' (PFP) mass are closer to the ``rogue planet'' value found by Schild 1996 for Q0957, and are compatible with results for Q2237 as tentatively interpreted by \cite{ref93}. The \cite{gib96} theory predicts simultaneous fragmentation of the primordial gas into $10^{6}$ $M_{\sun}$ proto-globular-cluster (PGC) clumps as expected from the \cite{jns02} theory (but for different reasons), adding yet another source of unaccounted for undersampling error to the star-microlensing searches. Microlensing of stars has to date produced many detections, but their interpretation is clouded by uncertainties in the models of the Halo of the Galaxy. The original theoretical work that justified these searches (\cite{pac86}, \cite{gri91}) did not take into account that microlensing of LMC stars might be produced by dark matter or foreground stars on the near side of the LMC itself, as pointed out by \cite{sah94}. This produces uncertainties in the interpretation of star-microlensing so large that parallax observations from one or more satellites may be needed to accurately compute the locations and parameters of the lenses (\cite{gau97}). Moreover, it was anticipated in the MACHO project experimental design (\cite{grt91}) that any Halo dark matter objects would have mass $(10^{-3}-10^{-1}) \, M_{\sun}$ much larger than $10^{-7} M_{\sun}$, so search frequencies and strategies prevented detection of lower mass objects for the first several years. In all MACHO/EROS calculations of exclusion estimates a spatially uniform distribution is assumed, rather than the highly intermittent distributions expected ($\S$\ref{sec2}) if the Halo dark matter mass is dominated by hydrogenous planetoids (PFPs, MBDs) in dark PGC clumps. A small fraction of the LMC stars sampled may intersect PGCs, but even these have little chance of any PFP microlensing within practical observational periods from the extremely intermittent PFP number density $n_p$ we show will occur within most PGCs due to their accretional cascades. Whereas the available statistics of star-microlensing do not appear to have convincingly produced a dark matter detection (\cite{ker98}), they have sometimes, but not consistently, been presented as precluding the population discovered in quasar-microlensing. The purpose of the present manuscript is to compare these two microlensing approaches to the missing-mass detection problem, with particular reference to the PGC clumps of PFP planetoids proposed by \cite{gib96} as the dominant inner-halo component of the galaxy missing-mass based on nonlinear (non-Jeans) fluid mechanical theory (\S \ref{sec1}). We show in our Conclusions (\S \ref{secc}) that results from quasar-microlensing observations (\S \ref{seca}) and star-microlensing observations (\S \ref{secb}) are actually compatible, considering the intermittent lognormal number density expected for these objects as they cluster in a nonlinear cascade to form larger clusters, larger objects and finally, for a small percentage, stars (\S \ref{sec2}). \section{Quasar-microlensing observations} \label{seca} The first dark matter conclusion based upon microlensing was by \cite{kee82}, who confirmed from 80 years of archived images the remarkable concentration of bright quasars near bright galaxies noted by \cite{bur79} and \cite{arp80}. Keel found QSO surface densities $\sigma$ increased by more than two orders of magnitude at separations $R \approx$ 10 kpc on the galaxy plane above background values with separations $R \ge$ 100 kpc, but showed low amplitudes of quasar brightness fluctuations (less than 1 magnitude or 17\%) at stellar-mass frequencies. Keel therefore concluded that ``if gravitational amplification is involved in producing an excess of bright QSOs near galaxies, ordinary halo stars are not responsible.'' The effective radius of the bright QSO density distribution was 17 kpc, ``slightly larger than globular cluster systems of the Galaxy and M31.'' From this evidence, the mass of the microlensing objects must either be much larger than $M_{\sun}$ or much smaller, and concentrated within the $10^{21}$ m (32 kpc) inner-galaxy-halo range observed for globular clusters around galaxies. The directly measured microlensing (by \cite{van89}) for Q0957 for a 415 day time delay was not analyzed for missing-mass implications for many years, possibly in part because the time delay would remain controversial. The detected microlensing in Q2237, first reported in \cite{irw89} and initially interpreted as due to masses in the range $0.001 M_{\sun} < M < 0.1 M_{\sun}$, was reinterpreted in terms of a statistical microlensing theory by \cite{ref93}, applicable to the case of resolved accretion disks, giving possible microlensing masses as small as $10^{-7} M_{\sun}$. With time delay issues for the Q0957 system resolved in favor of the original (\cite{sch86}) time delay value, the microlensing was discussed by \cite{sch94}, 1995, who removed the intrinsic source fluctuations of the A image by subtracting the B image shifted 404 days forward, revealing a continuously fluctuating pattern in the difference signal, due only to microlensing, that would be evidence of the missing-mass. A Fourier power spectrum of these microlensing fluctuations in \cite{sch95}, and updated in \cite{sch96}, shows that the microlensing power spectral area is dominated by a broad peak whose center frequency corresponds to a microlensing mass of $10^{-5.5} M_{\sun}$ ($4.4\times10^{24}$ kg). The dominant microlens mass might be smaller than this due to clumping, but not much larger. To understand why quasar-microlensing and star-microlensing experiments seem to give different results, it is necessary to understand how the higher optical depth in quasar-microlensing affects the outcome differently for microlensing masses with the intermittent spatial distribution expected for any primordial population of small self-gravitating masses. In quasar-microlensing, two or more images of the same quasar are observed, and a normal galaxy, acting as a lens, will produce unit optical depth to microlensing. The definition of optical depth is given precisely in \cite{schn92}. It is formulated for an unresolved source, and is effectively the probability that a microlensing event is underway. The surface optical depths for the (A,B) images of the Q0957 system are (0.3, 1.3), which means that for image B it is virtually certain that at least one microlensing event is underway at any time. For image A, the probability is only about 30\% that at any statistical moment a microlensing event is underway. These probabilities are independent of the microlensing mass, which can be estimated from the event duration. A more complex situation, more favorable for the detection of microlensing, exists for the Q0957 system, because the Einstein rings of all microlensing masses are substantially smaller than the size of the luminous quasar accretion disk. \cite{sch96} used the historic record of Q0957 brightness fluctuations, in combination with the appropriate \cite{ref93} statistical theory, to estimate the size of the quasar source as 6 times larger than the stellar-microlens Einstein ring. For a half-solar mass typical microlens, presumed to be present from the lens galaxy's spectrum, this gives an accretion disk diameter of $3 \times 10^{15}$ m. More importantly, this result means that at any time there are $6^2$ independent lines of sight to the quasar B image that each has a probability of 1.3 of hosting a statistically independent stellar-microlensing test, or a net probability of $1.3 \times 6^2$ stellar mass microlenses at any moment, and the observed brightness fluctuations on a 30-year time scale should simply reflect the Poisson statistics of this number of occupied stellar microlensing sites. The luminous matter described above is known to exist from the lens galaxy's light, whose spectrum is dominated by stars of approximately half-solar mass. From the amplitudes and durations of its stellar-microlensing events we have made inferences about the area of the luminous quasar source. But \cite{sch96} emphasizes the microlensing signature of a dominant second population of objects, whose 10 to 100 day fluctuation pattern corresponds to a microlens mass of $10^{-6}$ $M_{\sun}$. For these objects to be the missing-mass, there would need to be at least $10^{6}$ of them for each half-solar mass star with the usual assumption that the lens galaxy is dominated by the missing-mass at the 90\% level or more. Consistent with the estimates for the quasar source luminous area given above, the line of sight to image B should have $1.3 \times 36 \times 10^{7}$ micro-brown-dwarfs, and image A should have $0.3 \times 36 \times 10^{7}$, out of $\approx 10^{17}$ in a galaxy. The brightness fluctuations observed should reflect the statistical distribution in the numbers of these microlenses on time scales appropriate for such small particles (10 to 100 days). The pattern of brightness fluctuations in each quasar image should be continuous and should be symmetrical (there should be equal positive and negative fluctuations), as observed in both the A and B images, and in their phased difference. We discuss in \S \ref{sec2} the implications of this large number of statistical tests on estimates of the statistical parameters of a primordial population of missing-mass objects. A reanalysis of the time delay and microlensing in this lens system by \cite{pel98} leads to a smaller source size, but this does not affect the conclusion that many MBDs would be projected against the luminous quasar source at all times. It appears likely that the microlensing interpretation is beginning to confront the real quasar source structure, and that the simple, uniformly bright, luminous disk source models are becoming inadequate. It is beyond the scope of the present paper to model the hydro-gravitational processes of quasar accretion, but it is not unreasonable to speculate that such regions of large turbulence dissipation rate $\varepsilon$, high gas density $\rho$, and enormous luminosity could rapidly produce a large number of powerful point sources of light. Stars form in turbulent regions with mass $M_T = (\varepsilon^6 \rho^{-5} G^{-9})^{1/4}$ in times $\tau_G = (\rho G)^{-1/2}$, \cite{gib96}, with $M_T$ limited only by radiation pressure differences that decrease as the ambient luminosity increases (the Eddington limit with small ambient luminosity is ${M_E}_0 \approx 100 M_{\sun}$). The luminosity of such quasar superstars should increase rapidly with $M_T \gg {M_E}_0$. The Tully-Fisher relationship, where spiral galaxy luminosity increases with rotation rate, is another candidate for enhanced turbulence causing an excess of large stars with large luminosities. \section{Star-microlensing observations} \label{secb} The star-microlensing, or MACHO/EROS/OGLE experiments, are operating in a very different optical depth realm, and are therefore affected by the detection statistics in profoundly different ways. These experiments were justified by the statistical inference of \cite{pac86}, later confirmed by \cite{gri91}, that if the missing-mass consists of stellar objects the optical depth for microlensing by a star in the Galaxy's Halo of a star in the disk of the Large Magellanic cloud would be $\approx (1/2) \times10^{-6}$. Thus, by observing up to $8 \times 10^6$ LMC stars nightly, the probability of detecting a microlensing event would hopefully be above unity. Early reports of the detection of such MACHO/EROS/OGLE events and claims of the detection of the Halo missing-mass had to be corrected by \cite{sah94}, who demonstrated that microlenses in the LMC halo were about as important as microlenses in the Galaxy's Halo, and the events would be indistinguishable to available experiments. To date these experiments have reported several LMC events (hundreds of events toward the Galaxy bulge) but no detection of the missing-mass, partly because of the ambiguity about whether the LMC microlenses are in the Halo of the Galaxy or in the LMC itself, and partly, we suggest, because the likelihood of MACHO clumping and intermittency have been neglected for the $10^{-6} M_{\sun}$ mass range. The statistics of star-microlensing are profoundly different than those for quasar-microlensing because of the very different optical depth regimes. For star-microlensing, on any night, of order $2 \times 10^6$ stars are sampled, and so there are 2\,000\,000 independent statistical tests asking ``is a solar-mass microlensing event underway.'' Since an event for solar-mass microlenses lasts about 30 days, a new sample is available monthly. However, it is not an independent sample since it involves the same objects only slightly displaced by their relative velocities. The time required to assure complete temporal independence of the sample is a few million years and spatial independence of the sample would require a different set of stars elsewhere on the sky. If the missing-mass is not approximately solar but rather terrestrial (i.e., $10^{-6} M_{\sun}$), then events should be of duration only hours, and most would have escaped detection in the main star-microlensing searches. However, some higher frequency searches for such events have been undertaken (\cite{alc98}, \cite{ren98}) and the claim made that non-detection constitutes proof that they do not exist. No explanation is given for the fact that rapid microlensing is observed in quasar monitoring, \cite{sch96}, and we conclude that star-microlensing is unable to detect and study the missing matter of galaxies. Why? In \S \ref{sec2} we show that the failure of star-microlensing to detect the missing Galactic Halo matter is readily understood from the sparse number of independent samples and because of the expected non-Gaussian distributions of the PFP masses of the microlensing objects within PGC clumps as well as the likely non-Gaussian distributions of the PGC clumps. Our purpose is to show that any massive, primordial, population of small condensed objects would be expected to have entered a nonlinear gravitational cascade of particle aggregation, and the resulting highly intermittent lognormal number density distribution would leave too much uncertainty in the star-microlensing statistics for any conclusions to be drawn about the mean number density $n_p$ of $10^{-6} M_{\sun}$ objects from limited star-microlensing searches. Because quasar-microlensing involves optical depths and detection probabilities at least a factor of $10^6$ higher, it offers greater prospects for robust sampling of mean masses and mean densities of such small dark-matter objects. \cite{ker98} have also suggested inner-outer Halo population differences might contribute to the failure of star microlensing searches to find the baryonic dark matter. \section{Nonlinear hydro-gravitational condensation of the primordial gas} \label{sec1} Star microlensing surveys have not aggressively searched for planetoids as the missing Galactic mass because the existence of such objects is impossible by the generally accepted \cite{jns02} theory of gravitational instability. However, Jeans's linear acoustic theory is subject to vast errors, especially for very strongly turbulent and very weakly turbulent flows and for weakly collisional materials such as neutrinos with enormous diffusivities $D$, \cite{gib00a}bc. Cold dark matter theories for galaxy formation from nonbaryonic dark matter condensations fail because the diffusivity $D$ for such material is so large that the diffusive Schwarz scale is larger than the corresponding Jeans scale, $L_{SD} \equiv ({D^2}/\rho G)^{1/4} \gg L_J \equiv V_S / (\rho G)^{1/2}$, where $L_J$ values in CDM theories are made small enough for galaxy size CDM halos to form by assuming the nonbaryonic dark matter is cold to reduce the sound speed $V_S$. However, no structures can form by gravity on scales smaller than the largest Schwarz scale according to \cite{gib96}, and in this case the largest Schwarz scale will be $L_{SD}$. For a weakly collisional nonbaryonic fluid, the criterion $L_{SD} \gg L_J$ is equivalent to ${m_p}^{3/2}/\sigma \gg (kT\rho/G)^{1/2}$, where $m_p$ is the particle mass, $\sigma$ is the collision cross section, and $k$ is Boltzmann's constant. Substitution of $\rho$ and $T$ values from the plasma epoch, when CDM halos were presumably formed, give values about $10^{-12}$ $\rm kg^{3/2} m^{-2}$. Typically assumed CDM particle masses ${m_p} \ge 10^{-27}$ kg would require $\sigma$ values $\approx 10^{-28} \, \rm m^2$ for the equality to fail, but realistic $\sigma$ values for CDM must be many orders of magnitude less for its particles to escape detection, $< \approx 10^{-43} \, \rm m^2$. Observed outer-halo galaxy scales of $10^{22}$ m (324 kpc) suggest from $L_{SD}$ that the nonbaryonic dark matter has weakly collisional particles with $m_p$ values closer to $10^{-35}$ kg, and this material makes only a small ($\approx 1\%$) contribution to the inner-halo galaxy density, \cite{gib00b}. The case of strong turbulence was first recognized by \cite{cha51}, who suggested that a ``turbulent pressure'' $p_{T} = \rho V_{T}^2$ should simply be added to the gas pressure $p$ in the Jeans length scale definition $L_J \equiv V_S/(\rho G)^{1/2} \approx (p/\rho ^2 G)^{1/2}$, where $V_S \approx (p/\rho)^{1/2}$ is the sound velocity, $p$ is pressure, $\rho$ is density, and $G$ is Newton's constant of gravitation. This argument is flawed for two reasons: firstly because pressure without gradients offers no resistance to gravitational forces (the basic flaw in Jean's theory), and secondly because the turbulent velocity is a function of length scale $L$. Substituting $p_T$ for $p$ in the expression for $L_J$ rather than adding and using the Kolmogorov expression $V_{T} \sim (\varepsilon L)^{1/3}$ gives the turbulent Schwarz scale $L_{ST} \equiv \varepsilon ^{1/2}/ (\rho G)^{3/4}$ of \cite{gib96}, where $\varepsilon$ is the viscous dissipation rate of the strong turbulence. If the dissipation rate $\varepsilon \le \rho \nu G$, where $\nu$ is the kinematic viscosity of the gas, then the flow is non-turbulent at the scale of gravitational domination and is controlled by viscous forces at the viscous Schwarz scale $L_{SV} \equiv (\gamma \nu/\rho G)^{1/2}$, where $\gamma$ is the rate-of-strain of the flow, \cite{gib96}. In gas where $L_{ST} \gg L_J$, Jeans mass stars cannot form because the turbulence forces overwhelm gravitational forces at all scales smaller than $L_{ST}$. Such a situation may occur due to supernova shock turbulence, in starburst regions and in cold molecular clouds, and due to gravitational infall turbulence near galaxy and PGC cores with $\varepsilon \ge RT(\rho G)^{1/2}$ for strongly turbulent gas regions on scales larger than $L_{ST}$, where $V_S^2 \approx RT$ and $R$ is the gas constant. The case of interest in the present paper has been overlooked by standard cosmological models; that is, the hot gasses of the early universe emerging from the plasma epoch with weak or nonexistent turbulence, where both $L_{SV}$ and $L_{ST} \ll L_J$. The physical significance of the Jeans scale $L_J$ is found by deriving the maximum length scale of pressure equilibrium in a fluctuating density field with self gravitational forces. Pressure equilibrium occurs if the gravitational free fall time $\tau_G \equiv (\rho G)^{-1/2}$ is greater than or equal to the time required for acoustic propagation of the resulting adjustment in the pressure field on scale $L$; that is, $\tau_G \ge \tau_P \equiv L/V_S$. Solving gives $L \le L_J$. Thus $L_J$ is not a minimum scale of gravitational instability as usually supposed, but is a maximum length scale of acoustical pressure equilibrium. This shows that there should be fragmentation of the primordial gas at both the Jeans scale $L_J \approx 10^4 L_{ST}$ and the weakly turbulent or nonturbulent Schwarz scales $L_{ST} \approx L_{SV}$, \cite{gib96}. From the ideal gas law $p = \rho R T$, the pressure can adjust to density increases and decreases due to gravitational condensations and void formations on scales $L_J \ge L \ge L_{ST}$ keeping the temperature nearly constant without radiant heat transfer. Voids forming at scales larger than $L_J$ cannot adjust their pressure fast enough to maintain constant temperature, causing a temperature decrease within the voids. Radiant heating of such voids larger than $L_J$ by the surrounding warmer gas therefore accelerates their formation in the expanding proto-galaxy, causing fragmentation of the primordial gas into proto-globular-cluster (PGC) blobs with mass of order $10^{36}$ kg, or $10^{6} M_{\sun}$. Since this instability has nothing to do with the Jeans theory, the scale $L_{IC} \equiv (RT/\rho G)^{1/2}$ has been termed the initial condensation scale, \cite{gs99a}b. Because pressure forces propagate by electromagnetic radiation in plasmas the sound speed in the plasma epoch is nearly the light speed $c$, giving Jeans scales $L_J$ larger than the Hubble scale of causal connection $L_H \equiv ct$ and therefore no possibility of structure formation in the baryonic component of matter by the Jeans theory. However, we have seen that the Jeans scale is generally not the criterion for gravitational structure formation, but instead gives the maximum scale of pressure equilibrium by acoustic propagation. Thus we can conclude that the small inhomogeneities in the CMB observed by COBE and other experiments are not acoustic as often assumed, but are more likely remnants of the first formation of structure due to gravity. The power spectrum of COBE temperature fluctuations peaks at a length scale ${L_{H}}_{FS} \ll L_H$ much smaller than the horizon, or Hubble, scale $L_H$ at the time of photon decoupling, suggesting the first structure formation must have occurred previously to match the spectral peak scale to a smaller horizon scale ${L_{H}}_{FS}$ existing at some earlier time of first structure, \cite{gib00b}. Formation of proto-super-cluster to proto-galaxy mass fragments from void formations was possible in the plasma epoch, but with slight increases in density from gravitational condensation because the universe age $t$ was at all times less than the gravitational free fall time $\tau_G \equiv (\rho G)^{-1/2}$ and because of non-baryonic diffusive compensation, \cite{gs99b}. Expansion of the universe enhances formation of voids and retards condensation of fragments, but amplitudes of both are damped by rapid diffusive density compensation as the non-baryonic component fills the minima and exits the maxima formed in the baryonic component. Fragmentation and fractionation of the primordial plasma was first triggered when the Hubble scale of causal connection $L_H \equiv ct$ decreased to scales less than the viscous or weakly turbulent Schwarz length scales $L_{SV} \approx L_{ST}$, \cite{gib96}. Remarkably, the baryonic matter density $\rho(t)$ from Einstein's equation at $t = 10^{12}$ s of $10^{-17}$ kg $\rm m^{-3}$, \cite{win72}, matches the density of globular star clusters (GCs). We take this $\rho$ value to be that of the primordial gas at $t = 10^{13}$ s, and further evidence of structure formation beginning at about $t = 10^{12}$ s. Observations of $\rho$ $\approx 10^{-17}$ kg $\rm m^{-3}$ in dim and luminous globular clusters suggests the internal evolution of PGCs has been quite gentle and nondisruptive so that dark PGCs with PFPs in less advanced stages of their accretional cascades may be considered metastable, with average densities $\rho_{PGC} \approx \rho_{IC}$ and masses $M_{PGC} \approx M_{IC}$ that depart slowly from these initial values. Additional evidence suggesting the time of first structure was about $t_{FS} \approx 10^{12}$ s, or $30 \, 000$ years, is that this is the time when the supercluster mass of $10^{46}$ kg matches the Hubble mass $\rho(t) (ct)^3$, \cite{gib97a}. Taking the horizon scale at this time equal to the viscous Schwarz scale $L_{SV} \equiv (\gamma \nu / \rho G)^{1/2}$ with $\gamma \approx t^{-1}$ gives a kinematic viscosity $\nu$ of $5 \times 10^{26}$ $\rm m^2$ $\rm s^{-1}$, which matches the photon viscosity $\nu \equiv l_C c$ obtained from the mean free path for Thomson scattering $l_C = 1/n_e \sigma_T$ with free electrons of the plasma times the speed of light $c$, where $n_e$ is the free electron number density and $\sigma_T$ is the Thomson scattering cross section, \cite{gib99b}. This enormous viscosity (a million times that of the Earth's upper mantle) gives a horizon scale Reynolds number $\rm Re \equiv c^2 t/\nu \approx 200$ which is slightly above critical, consistent with evidence from COBE of weak turbulence in the primordial plasma, and precludes any interpretation of the observed CMB temperature anisotropies as ``sonic peaks'' since any such powerful plasma sound would be promptly damped by viscosity, even if its unknown source could be identified, \cite{gib00b}. Evidence of proto-supercluster to proto-galaxy formation in the plasma epoch can also be inferred from the advanced morphology of superclusters, with $10^{24}$ m voids bounded by ``great walls'' as shown by deep galaxy maps, and by galaxy rotation rates of order $10^{-15}$ rad $\rm s^{-1}$ consistent from angular momentum conservation as the galaxies expand with our estimated time of proto-galaxy fragmentation at $\approx 10^{13}$ s, with angular velocity $\Omega$ and strain rate $\gamma$ both near $10^{-13}$ rad $\rm s^{-1}$. Thus, we assume that the neutral primordial gas emerging from the plasma epoch was fragmented into $\approx 10^{41}$ kg proto-galaxy blobs of hydrogen and helium with a very uniform density $\rho \approx 10^{-17}$ kg $\rm m^{-3}$ reflecting the time $t_{FS} \approx 10^{12}$ s of first fragmentation and proto-superclusters, and $\Omega \approx 10^{-13}$ rad $\rm s^{-1}$. Turbulent mixing scrambles density fluctuations from large scales to small to produce nonacoustic density nuclei that move with the fluid, and these trigger gravitational condensation on maxima and void formation at minima. Condensation on acoustic density nuclei; that is, density maxima moving with the sound speed as envisaged by Jeans, is not a realistic description of gravitational structure formation under any circumstances, as shown by \cite{gib96}. The smallest scale $\rho$ fluctuation produced by turbulent mixing is at the Batchelor length scale $L_B \equiv (D/\gamma)^{1/2}$, \cite{gib91}, corresponding to a local equilibrium between convective enhancement of $\nabla \rho$ by the rate-of-strain $\gamma$ and smoothing of $\nabla \rho $ by the molecular diffusivity $D$, near points of maximum and minimum $\rho$ moving with the fluid velocity $\bf \vec v$, \cite{gib68} and \cite{gs99a}. The Jeans theory fails because it relies on the Euler momentum equations that neglect viscous forces, and because it employs linear perturbation stability analysis, which neglects the nonlinear inertial-vortex forces that dominate turbulent flows, \cite{gib99a}. Viscous forces are necessary to stabilize the rapidly expanding flow of the early universe for times $t$ before structures form, with rate-of-strain $\gamma$ values of order $t^{-1}$, \cite{gib99b}. Expanding inviscid flows are absolutely unstable (\cite{lan59}, \cite{gib00b}), so a major result of the COsmic Background Experiment (COBE) was a demonstration that temperature fluctuations $\delta T/T$ were $\le 10^{-5}$ rather than $\ge 10^{-2}$ to be expected if plasma epoch flows before transition to neutral gas were fully turbulent. In the absence of structure to produce buoyancy forces, only viscous forces (small Reynolds numbers) can explain the absence of turbulence that is manifest in the homogeneity of the cosmic microwave background (CMB) temperature fluctuations. Linearity and the Jeans theory cannot be salvaged by arguing that $\delta \rho / \rho$ is small so that very long times would be required for the gravitational structure formation to ``go nonlinear''. The time required to form structure by gravity is $\approx \tau_G \equiv (\rho G)^{-1/2}$ where $\tau_G$ is the gravitational free fall time, and nearly independent of the magnitude of the dominant density fluctuations on scales larger than the largest Schwarz scale. The mass of such a fluctuation $M'(t) \approx M'(0) exp[2 \pi (t/\tau_G)^2]$, \cite{gib99b}. Thus, the $10^{-3}$ difference between the observed CMB density fluctuations and the larger values expected for turbulence only causes a 45\% increase in the time required for $M'(t)$ to reach planetoidal values. For $\rho = 10^{-17}$ kg $\rm m^{-3}$, $\tau_G = 3.9 \times 10^{13}$ s or $1.2 \times 10^6$ years. Starting from $10^{-17}$ kg $\rm m^{-3}$, the time to reach density $\rho = 10^{0}$ kg $\rm m^{-3}$ is $3.1 \times 10^6$ years ($10^{14}$ s) and $< 10\%$ more to reach $10^{3}$ kg $\rm m^{-3}$ and quasi-hydrostatic equilibrium. This first condensation process of the universe has a profound effect on everything else that happens afterwards. The dynamics shifts from that of a collisional gas to that of weakly collisional warming gas blobs in an increasingly viscous, cooling, gaseous medium, evolving slowly toward a cold dark metastable state of no motion as the baryonic dark matter of galaxies, as observed by \cite{sch96}. The sites with the highest likelihood of instability to form the first stars and quasars are the cores of PGCs and the cores of PGs, at this same time $ t \approx 10^{14}$ s with $z \approx 264$. A large population of quasars at extremely high redshifts $z$ is indicated by dim point sources of hard X-rays detected by the Chandra satellite, \cite{mus00}. A survey of star-forming galaxies for $3.8 \le z \le 4.5$ after improved dust extinction corrections shows no indication of a peak in the (star formation rate) SFR/comoving-volume and therefore a monotonic increase in SFR/proper-volume with increasing $z$ toward a peak at some high $z \ge 4.5$ value, \cite{sti99}, presumably also at $z \approx 264$, compared to a peak at $z = 1 \-- 2$ assumed by many authors following the ``collapse'' of gas into CDM halos to form galaxies at $z \approx 5$ for the standard CDM model. Galaxies never collapse after their formation at $z \approx 1000$ in the present scenario. Clusters of galaxies observed at z = 3.09, \cite{sti00}, also present no difficulties for our scenario of first structure growth. The kinematic viscosity of the neutral primordial gas formed from the cooling plasma at decoupling is dramatically less than that of the plasma, with $\nu \approx 5 \times 10^{12}$ $\rm m^2$ $\rm s^{-1}$, \cite{gib99b}. Thus the viscous Schwarz scale $L_{SV} = (5 \times 10^{12} \times 10^{-13} / 10^{-17} \times 6.7 \times 10^{-11})^{1/2}$ = $2.7 \times 10^{13}$ m, and the viscous Schwarz mass $M_{SV} \equiv L_{SV}^3 \rho = (2.7 \times 10^{13})^3 \times 10^{-17}$ is $2 \times 10^{23}$ kg, or $10^{-7} M_{\sun}$, which we take to be the minimum expected mass of a primordial fog particle. The minimum PFP mass increases to $\approx 5 \times 10^{24}$ kg if fossil turbulence effects are included, taking $\gamma_o = 10^{-12}$ $\rm s^{-1}$ as a fossil vorticity turbulence remnant of the time of first structure $10^{12}$ s, \cite{gib00b}c. The weak turbulence levels possible from the COBE observations increase the condensation mass to maximum $M_{ST} \le 10^{25}$ kg values that are still in the small-planetary-mass range. The formation of these gaseous planetoids represents the first true condensation by gravity since the Big Bang, where the mass density $\rho$ increases with time. Whatever small turbulence may have existed at decoupling will now be damped by buoyancy forces as the PFPs begin their gradual evolution from hot gas blobs to freezing cold liquid-solid Neptunes at $z \approx 3$, with a gentle, complex, accretional cascade to form stars that results in practical invisibility of these PFPs to star-microlensing searches. Frozen hydrogenous planetoids are stable to evaporation at the temperature of the present universe for masses larger than about $10^{-8} M_{\sun}$, \cite{der92}. No account has been made for possible increases in PFP atmosphere size or increased PFP number density on their fraction of supernova energy absorbed with increasing $z$, even though $\Lambda \ne 0$ and ``dark energy'' inferences depend on only 30\% dimming of supernova Type Ia events at the larger $z$ values to justify claims of accelerations in the universe expansion. \section{Intermittent lognormal number density of hydrogenous, primordial, small-planetary-mass objects} \label{sec2} Star-microlensing estimates (\cite{alc98}) of the contribution to the missing-mass of the Galaxy Halo by compact Halo objects in various mass ranges have assumed that the objects are uniformly distributed in space. This assumption is highly questionable, especially for objects in the small-planetary-mass range $M_p \approx 10^{-6} M_{\sun}$. If such objects are sufficiently compact to cause microlensing, and sufficiently numerous to dominate the Galaxy mass, they must be primordial, and consist of H-He in primordial proportions. They comprise the initial material of construction for the first small stars in PGCs, and an important stage in the process of such star formation. Non-baryonic materials are too diffusive to become compact, and no cosmological model has produced this much baryonic mass that is not H-He. Starting from a nearly uniform distribution of $10^{6} M_{\sun}$ PGC-mass clumps of hot-gas PFPs in the proto-galaxy (PG), all of the PFPs must have participated to some extent in nonlinear, gravitational, accretion cascades culminating for luminous GCs in the formation of the first small stars simultaneously, giving an abrupt end to the ``dark ages'' of the universe at $z \approx 264$, versus $5$ much later as usually assumed, with the first large stars forming simultaneously with reevaporation and the first strong turbulence at PGC-cores near PG-cores as Population III super-stars, with their supernovas, black holes, and proto-quasars, and probably one or less for each proto-galaxy so that the universe never reionized as often assumed. Those PGCs in the outer regions of PGs should cool more rapidly and be less successful in accretion cascades of their PFPs than those nearer PG-cores. Pairs of PFPs should form first, and then pairs of pairs, etc., producing a complex array of nested clumps within each PGC, and large voids in between. Large drag forces from the fairly dense, weakly turbulent, inter-PFP gas should produce small PFP terminal velocities from its drag compared to virial velocities, and the large number of neighboring PFPs and their clumps should tend to randomize PFP velocity directions. Such an accretion process produces a highly non-uniform PFP number density $n_p$, with the potential for large undersampling errors in star-microlensing estimates of the mean number density $\langle n_p \rangle$ of such objects in the Halo from the resulting small numbers of independent samples of $n_p$. A sample consists of several observations of each star for times longer than the expected microlensing time for a number of stars sufficient to give at least one event, where the sampling time required depends on how many objects are expected and whether the Galaxy Halo consists of a uniform distribution of objects in the sampled mass range or a nonuniform distribution. It takes longer to acquire a sample if the probability density function (pdf) of $n_p$ is nonuniform and much longer if it is extremely intermittent. An independent sample is one taken from stars separated from other samples by distances larger than the largest size of clumps of such objects or any clumps of such clumps. If the same stars are used for each sample, one must allow time between samples so that the sampled field of objects can move away and be replaced by an independent field. To avoid undersampling error, one must account for nonuniformity of the objects sampled and take sufficient numbers of independent samples to achieve an acceptable level of statistical uncertainty in the estimated mean $\langle n_p \rangle$. \subsubsection{Clustering models for the accretion of primordial-fog-particles (PFPs)} Detailed modeling of the accretion of PFP objects within the proto-globular-cluster (PGC) initial condensation masses is an important topic, but is rather complex and therefore beyond the scope of the present paper. Rates of PFP merging depend on relative velocities, frictional damping, and degrees of condensation of the PFP gaseous atmospheres. A gas epoch existed before $z = 3$ with no possibility for condensation of the primordial gas to liquid or solid states because universe temperatures were above hydrogen and helium dew point and freezing point temperatures. Thus rates of merging of the large, entirely gaseous PFPs existing then should have been more rapid than the colder, more compact PFPs to appear later with condensation to liquid and solid states in the colder universe. Frictional and tidal forces between gassy PFPs were significant, reducing velocities and merging rates and permitting radiative cooling and further condensation toward more collisionless PFP states. Principles of collisionless dynamics (\cite{bin87}) describing star interactions often may not apply to PFP interactions, particularly in the early stages of PFP merging when large scale gas atmospheres about the objects existed. Neither must stellar virial theorems apply to PFPs because frictional forces of PFP extended gaseous atmospheres will reduce their relative velocities below virial values if they have been exposed to radiation, tidal, or convective heating flows that can cause re-evaporation. As an exercise, we can compare star clustering with PFP clustering. It is well known from stellar dynamics that the third star in a triple must lie at least three times the radius of a binary away from the center. An accretion cascade of triples to form triples of triples, etc., within a uniform cluster of stars, would result in an ever decreasing average mass density of the star cluster, rather than a constant or increasing density as expected for frictionally interacting PFPs within a proto-globular-cluster (PGC) as they accrete in a bottom-up cascade to form globular cluster stars. If there were $n$ stages in a tripling cascade and we assumed the third member of the $k$-th triple cluster must be four times the radius $R_k$ separating the associated binary of $(k-1)$-th clusters, where $1 \le k \le n$, then the mass of the $n$-th cluster is $M_n = 3^n M_{star}$, and the radius $R_n = (4m)^n d$, with star size $d$ and binary star separation $R_{bs} = md, m \gg 1$. This would result in a monotonic decrease in the mass density of star clusters $M_n = 3^n M_{star}/(4m)^{3n} {R_{bs}}^3$ with increasing $n$. If the initial mass density of unclustered stars is $\rho_{0} = M_{star}/(4md)^3 \approx \rho_{star}/(4m)^3$, where $\rho_{star}$ is the star density, then $\rho_n /\rho_{star} = 3^n (\rho_0/\rho_{star})^{3n}$, which also decreases rapidly with increasing values of $n$ since $\rho_0/\rho_{star} \ll 1$. Such a star tripling cascade is possible, but is slow. Times between collisions $t_{col} = 1/(\sigma_{star} v n)$ are large. For example, in a virialized globular cluster with $\sigma_{star} \approx 10^{18}$ $\rm m^2$, $v \approx 8 \times 10^3$ $\rm m \> s^{-1}$, and $n \approx 10^{-51}$ $\rm m^{-3}$ we find $t_{col} \approx 10^{29}$ s, which is $\approx 3 \times 10^{11}$ larger than the age of the universe. On the other hand, times $t_{PFP}$ for collisions of PFPs $t_{PFP} \approx L_{sep}(\rho_{PFP}/\rho_{0})^{2/3}/v$ are much shorter, where $L_{sep} = (M_{PFP}/\rho_{0})^{1/3}$ is the PFP separation distance, $\rho_{0}$ is the mass density of the PGC, $v$ is the relative velocity, and $\rho_{PFP}$ is the PFP density. For the same relative velocity $v \approx 8 \times 10^3$ $\rm m \> s^{-1}$, for $\rho_{PFP}/\rho_{0} = 10$, $\rho_{0} = 10^{-17}$ $\rm kg \, m^{-3}$, and $L_{sep} = 2 \times 10^{13}$ m, we find $t_{PFP} \approx 10^{10}$ s, which is much less than the age of the universe of $10^{13}$ s when PFPs were first formed. Relative velocity values for PFPs then were certainly much less than virial values $(2GM/R)^{1/2}$, with collision times as much as a few times the universe age as a lower bound based on $ v \approx \gamma R$ from the universe expansion value $\gamma = 1/t$. As another exercise, consider the commonly assumed virial theorem for collisionless objects such as stars. The virial theorem applied to PFPs within a PGC soon after fragmentation would incorrectly equate the kinetic energy per unit mass $v^2/2$ of each PFP to its potential energy per unit mass $GM/R$ within the PGC of mass $M$ and radius $R$. By Hegge's law (\cite{bin87}), the outer orbital velocity of stars, from the application of the virial theorem to a star cluster, must exceed the dispersion velocity of the stars in the cluster, or else any internal hierarchical star structures will be destroyed. Hegge's law is appropriate to star clusters because stars are virtually collisionless. However, frictional and tidal forces between interacting PFPs prevent virialization and enhance merging. For the purposes of the present paper, it suffices to postulate that the aggregation cascade at every stage from PFP to star mass within a PGC is nonlinear and approximately self-similar, leading to clumping and increased intermittency of the PFP number density as the number of stages increases. It is not known what fraction of the gas of a proto-galaxy fragments into proto-globular-clusters simultaneous with PGC fragmentation into PFPs. Since there are only about 200 globular clusters (GCs) in the Milky Way Halo, and this number is typical for nearby galaxies, one might think that PGC formation were a rare event since about a million PGCs are required to equal the mass of a proto-galaxy and only 0.02\% are observed in our nearby galaxies close enough to sample. A more likely scenario is that $\approx 10^{6}$ PGCs formed per galaxy as expected, but most PGCs have remained dark with so few stars that they have escaped detection, either in the bright, dusty, galaxy core where the clustering of stars is more difficult to discern, or in the dark halo where external triggers of star formation are rare and most PGCs are still only partially evolved. Some fraction of the original PGCs of a galaxy will be disrupted to form the interstellar medium of the disk and core where rogue PFPs will generally dominate the ISM mass. Dynamical constraints on dark clusters of objects such as PGCs are examined by \cite{car99}, and our value of $[M_C / M_{\sun} , R_C /pc ] = [10^6 , 8]$ is well within their range of values not excluded. An authoritative book on globular cluster systems suggests that observations of these systems have ``outstripped theoretical work'' on how the systems and the globular clusters themselves are formed, especially young globular clusters, \cite{ash98}. The accumulation of evidence supports early suggestions of these authors that galaxy merger-induced starbursts are favorable environments for globular cluster formation, as dramatically confirmed in the Hubble Space Telescope (HST) images of the merger galaxy NGC 3256 where more than 1000 bright blue objects have been identified as young globular clusters, \cite{zep99}. Our suggestion is that the bright system of young GCs observed within $10^{20}$ m of the central region of the NGC 3256 merger can be readily explained as a small fraction of the many dark, primordial, PGC-PFP systems available in the merging galaxies that have been brought out of cold storage by tidal forces of the merger process accelerating the accretion of PFPs to form stars. With the recent high resolution and infrared capabilities of the HST, many dense star clusters near the core of the Milky Way have been revealed. These clusters have masses of order $10^6 M_{\sun}$ expected for PGC objects formed at the initial condensation $L_{IC}$ scale at decoupling. Densities up to $10^{-13}$ kg $\rm m^{-3}$ are indicated, with massive $100 M_{\sun}$ stars reflecting the large $L_{ST}$ scales caused by the high turbulence levels from the frequent supernovas and powerful radiation and winds. Supernovas, radiation and winds are also mechanisms for increasing PGC density for PGCs close to a galaxy core because they increase friction by producing higher gas densities in the inter-PFP medium within evolving PGCs. Super-star clusters with mass $10^6 M_{\sun}$ and $\rho \approx 10^{-15}$ kg $\rm m^{-3}$ are reported by \cite{oco94} in the core of the dwarf galaxy NGC 1569, observed from the HST even before its repair, confirming the 1985 claim of Arp and Sandage that these objects were not stars but super-star-clusters. \cite{mar97} conclude that the brightest of the several super-star-clusters NGC 1569A in the NGC 1569 galaxy is actually a superposition of two clusters separated by a distance close to their sizes, so the clusters are apparently in orbit. They are identified as ``young globular star clusters''. Similar young globular star clusters are reported by \cite{ho96} at the center of NGC 1705, an amorphous galaxy, by \cite{hol96} in the interacting galaxy 1275, and by \cite{wat96} in the starburst galaxy NGC 253. In the two colliding galaxies NGC 4038/4039 termed The Antennae, a population of over 700 blue pointlike objects were identified by \cite{whi95} as young globular clusters formed as a result of the merger, and are therefore possibly dark PGCs brought out of cold storage. The size of the objects is given as 18 pc, giving a density of $10^{-17}$ kg $\rm m^{-3}$ that is very close to the density at the time of PGC formation from primordial gas, and represents a typical density of Halo GCs. Thus we have evidence that dark proto-globular-clusters (DPGCs) exist in abundance in a variety of galaxy types. If DPGCs consist of PFPs, no significant fraction of their original number of PFPs have apparently evaporated by collisionless dynamics mechanisms. Instead, the GC mass-density $\rho$ changes observed are to larger rather than smaller values. The original condensation mass $10^6 M_{\sun}$ has not changed. For PGCs formed in the outer regions of proto-galaxies (PGs) that were most likely to have maximum turbulence and minimum gas density and thus maximum PFP mass and separation and most rapid cooling to a collisionless state, a more stellar-like dynamics of clustering might seem appropriate with consequent decrease in the average PGC mass density. However, we find no evidence that GC densities outside galaxy core regions ($\approx 10^{20}$ m) deviate appreciably from $\rho \approx 10^{-17}$ kg $\rm m^{-3}$, suggesting that PGCs are gas stabilized by the possibility of reevaporation of their PFPs. The PGCs themselves presumably dispersed, along with the luminous GCs, from locations within an initial maximum galaxy diameter of $10^{20}$ m to the present size of $10^{21}$ m, partly due to the expansion of the universe and perhaps partly from their more collisionless dynamics compared to those near galaxy cores. Initial Hubble flow velocity differences at PG scales after decoupling were $\delta v_r = \gamma d = 10^{-13} \times 10^{20} = 10^7$ m $\rm s^{-1}$, giving rapid expansion of the galactic scale protovoids. The average mass density of galaxies is certainly substantially less today, with $\rho \approx 10^{-21}$ kg $\rm m^{-3}$, than the average mass density of $\rho \approx 10^{-17}$ kg $\rm m^{-3}$ for the PGs discussed in \S \ref{sec1}. Evidence for the existence of Halo dark matter in the form of cold $\rm H_2$ gas clouds clumped in dark clusters is claimed from measurements of diffuse $\gamma$-ray flux, \cite{dep99}. Similarly, cold self-gravitating hydrogen clouds explain ``extreme scattering events'' of compact radio quasars, \cite{wal98}. Both of these gas-cloud interpretations are consistent with and support our postulated dark metastable PGC-PFPs as the inner-galaxy-halo dark-matter, with their inter-PFP-gas and PFP-atmospheres, except that we require that most of the PGC mass be in the form of mostly-frozen solid-liquid PFPs as the source of this gas that resists changes in the average PGC mass density by shifts in the equilibrium gas density. For cold $\rm H_2$ clouds to exist completely as gas in clumps with PGC densities and sizes, huge turbulence dissipation rates of $\varepsilon \ge 3.4 \times 10^{-6}$ $\rm m^2$ $\rm s^{-3}$ would be required to prevent their condensation to form stars at $L_{ST}$ scales, compared to primordial $\varepsilon$ values of only $\approx 10^{-13}$ $\rm m^2$ $\rm s^{-3}$ estimated at the plasma-gas transition from the measured CMB temperature fluctuations leading to PFPs, \cite{gib99b}. No source of kinetic energy is available to produce or sustain such turbulence, and without such a source to prevent gravitational condensation the turbulence would dissipate in a few million years and the PGC mass gas cloud would collapse in a powerful star-burst of super-star formation. \subsubsection{Intermittent lognormal $n_p$ probability density function} The average separation distance $L_p$ between small-planetary-objects (PFPs) for a galaxy at present is about $ (M_p/\rho_{halo})^{1/3} \approx 7\times 10^{14}$ m for a galactic density $\rho \approx 10^{-21}$ $\rm kg \> m^{-3}$, with object size $r_p \approx (M_p/\rho_p)^{1/3} \approx 4\times 10^{6}$ m for a density $\rho_p \approx 3 \times 10^{3}$ $\rm kg \> m^{-3}$ corresponding to condensed H-He at halo temperatures, giving approximately 8 decades for the cascade from separation to accretion scales of $M_p$ mass objects. The number density $n_p (r)$ of such planetary objects per unit volume becomes an increasingly intermittent random variable (rv) as the averaging volume size $r$ decreases. If we assume most of the nonlinear gravitational cascade is self-similar, then the probability density function of $n_p (r)$ will be lognormal with an intermittency factor $I_p (R/r) \equiv \sigma ^{2}_{ln [n_p (r) / n_p (R)] }$, where $\sigma ^2$ is the variance about the mean, that increases as the range of the cascade $R/r$ increases as \begin{equation} \sigma ^{2}_{ln [ n_p (r) / n_p (R)] } \approx \mu_p ln (R/r) \label{eqz} \end{equation} where $R$ is the largest scale of the cascade, $r_p \le r$ is the smallest, and $\mu_p$ is a universal constant of order one. The expression is derived by expressing $n_p (r) / n_p (R)$ as the product of a large number of breakdown coefficients $b_m = n_p (k^m r) / n_p (k^{m+1}r)$ for the $0 \le m \le l$ intermediate stages of the cascade with scale ratio $k$ sufficiently large for the random variables $b_m$ to be independent. Taking the natural logarithm of the product \begin{equation} {n_p (r) \over n_p (R)} = {n_p (r) \over n_p (kr)} {n_p (kr) \over n_p (k^2 r)} {n_p (k^2 r) \over n_p (k^3 r)} ... \\ {n_p (k^{l-1}r) \over n_p (R)} = b_1 b_2 ... b_l \end{equation} gives \begin{equation} ln [{n_p (r) / n_p (R)}] = ln (b_1) + ln (b_2) + ... + ln (b_l). \end{equation} By the central limit theorem, the random variable $ln [{n_p (r) / n_p (R)}]$ is Gaussian, or normal, because it is the sum of many identically distributed, independent random variables $ln (b_m)$, which makes the random variable $[{n_p (r) / n_p (R)}]$ lognormal. The variance of a normal rv formed from a sum of rvs is the sum of the variances of its components. Since there are $l$ breakdown coefficients $b_m$, and $R/r = k^l$, then \begin{equation} \sigma ^{2}_{ln [ n_p (r) / n_p (R)] } \approx l \times \sigma ^{2}_{ln (b_m) } = {ln (R/r) \over ln (k)} \sigma ^{2}_{ln (b_m) } \end{equation} which gives Eq. (\ref{eqz}) and its constant \begin{equation} \mu_p \approx {\sigma ^{2}_{ln (b_m) } \over ln (k)}. \end{equation} The universal constant $\mu_p$ is always positive because $\sigma ^{2}_{ln (b_m) }$ is positive and $k > 1$. If the intermittency factor $I_p$ is large, then estimation of the mean number density $\langle n_p \rangle = n_p (R)$ from a small number of samples becomes difficult. A single sample gives an estimate of the mode, or most probable value, of a distribution, and the mean to mode ratio $G$ (the Gurvich number) for a lognormal $n_p$ distribution is \begin{equation} G \equiv \frac{{\langle n_p \rangle}_{mean}}{{\langle n_p \rangle}_{mode}} = exp (3I_p/2). \label{eqa} \end{equation} Taking $\mu_p \approx 1/2$, \cite{ber94}, and substituting $R/r_p = 7 \times 10^{14} / 4 \times 10^{6}$ gives $I_p = 9.5$. Thus, $G = 1.5 \times 10^6$ from Eq. (\ref{eqa}), which is the probable undersampling error for $\langle n_p \rangle$ from a single sample. Increasing $M_p$ has no effect on the intermittency factor $I_p$ or the undersampling error $G$ as long as the aggregation cascade is complete since the ratio $R/r_p$ is unaffected by $M_p$. The intermittent lognormal distribution function for $n_p (r)$ inferred above is generic for nonlinear cascades over a wide range of scales and masses and with a wide range of physical mechanisms, since it is based on the central limit theorem which gives nearly Gaussian distributions even when the summed component random variables are not identically distributed and are not perfectly independent as long as there are many components. \subsubsection{Implications for star-microlensing versus quasar-microlensing studies} The EROS and MACHO collaborations (\cite{alc97}, \cite{ren98}, \cite{alc98}) estimate they should have encountered about 20 star-microlensing events in their two years of observing with time periods corresponding to $M_p \approx 10^{-6} M_{\sun}$, assuming such objects are uniformly distributed in the Halo with $\langle n_p \rangle R^3 M_p \approx M_{Halo}$. Because they encountered none, they conclude that the contribution of $M_p$ objects can be no more than 1/20 of $M_{Halo}$. However, since $n_p$ is expected to be a lognormal random variable (LNrv) with intermittency factor $I_p = 9.5$, the most that can be said from this information is that the modal mass $\langle n_p \rangle _{mode} R^3 M_p$ of such objects is less than $M_{Halo}/20$. Since the mean mass $\langle n_p \rangle R^3 M_p \approx G \langle n_p \rangle _{mode} R^3 M_p \le M_{Halo} G/20 $, this is no constraint at all on $M_p \approx 10^{-6} M_{\sun}$ as the mass of Halo objects dominating $M_{Halo}$ because $G \approx 10^{6} \gg 20$. About ${2G}/{20}$ y of observing time, with no detections, would be needed to make the claimed exclusion, or 150,000 years. Furthermore, 150,000 years assumes all LMC stars are seen through PGCs even though these cover a small fraction of the sky and the PGC density itself is likely to be an intermittent lognormal. As mentioned, another serious problem with the EROS/MACHO sampling method is a lack of independence of the samples. Both collaborations are constrained because the solid angles occupied by the stars of the Large Magellanic Cloud are smaller than those expected for dark proto-globular-clusters (PGCs) of PFPs. Assuming a PGC diameter of $4.6 \times 10^{17}$ m at a distance of $10^{21}$ m, each PGC occupies a fraction $1.7 \times 10^{-8}$ of the sky, or 1.7\% for a million PGCs per galaxy. Therefore, most MACHO/EROS stars are likely to be in areas between PGCs, so they probably will see nothing, as they have done. In order to give independent samples of the Halo mass, stars chosen for star-microlensing tests must be separated by angular distances larger than the separation between the largest clumps. The time for a dark PGC to drift its diameter in the Halo is about $10^{12}$ s. Three hundred independent samples are required to achieve 50\% accuracy with 95\% confidence in estimating the expected value of a LNrv with $I_p = 9.5$ (\cite{bak87}), requiring $\approx 10^7$ y of observations toward the LMC to achieve this accuracy. This is very impractical. A million PGCs in the Galaxy Halo should occupy about 2\% of the sky, so if they were uniformly distributed only 2\% of the LMC stars would intersect a PGC. This in itself would justify a claim that PFPs fail as the missing mass based on the assumption (questioned herein) that particle density $n_p$ distribution function is uniform. We have shown that even if all the LMC stars were seen through PGCs, the indicated $\langle n_p \rangle$ would still be underestimated since the Gurvich factor for any PGC is likely to be much larger than 200 ($G \approx 10^6$). Quasar microlensing is much less affected by intermittency effects since as shown in \S \ref{seca} the number of microlensing masses seen projected in front of the quasar's luminous disk is likely to be of order $10^{8}$. Because no model as yet exists for the rapid brightness fluctuations observed, we do not yet know the role of structure in the quasar accretion disk and the distribution of implied masses. \section{Conclusions} \label{secc} Quasar-microlensing signals have continuous fluctuations at short time scales indicating the lensing galaxy mass is dominated by small-planetary-mass objects (\cite{sch96}). From a new theory of hydrodynamic gravitational condensation (\cite{gib96}$\-- 2000$) these objects are identified as PFP ``primordial fog particles'' that fragmented within PGC fragments within proto-galaxy blobs of neutral gas emerging from the plasma epoch at 300\,000 years. The proto-galaxies emerged embedded in a nested foam of proto-supercluster to proto-galaxy structures formed previously within the superviscous Big Bang plasma starting at about 30\,000 years, with the non-baryonic component diffusing to fill the voids and damp gravitational formation of $\rho$ fluctuations with amplitudes larger than the $\delta \rho / \rho \approx 8 \times 10^{-5}$ peak level observed, (\S \ref{sec1}). Observations of quasars, galaxies, and galaxy clusters at large redshifts support predictions of the theory that proto-superclusters formed at $z \approx 5700$, proto-galaxies at $z \approx 1000$, and the first stars, super-stars, proto-quasars, as well as the \cite{sch96} ``rogue planets'', near $z \approx 260$, (\S \ref{sec1}). The large population of clumped, frozen, hydrogenous planetoids predicted can masquerade as dark energy, quintessence, and the cosmological constant by causing high $z$ supernova dimming as an unanticipated time dependent radiation energy sink and overestimates of the cluster baryon fraction because of similar unanticipated gas clumping and cooling effects. Because the baryonic matter is mostly sequestered as PGC clumps of PFPs before the appearance of quasars, the Gunn-Peterson paradox of the missing intergalactic gas in quasar spectra is resolved. Failure of star-microlensing studies to observe equivalent microlensing events from the Bulge and LMC is not in conflict with the quasar-microlensing observations because fragmentation of primordial gas as PGCs and the nonlinear clustering of PFP fragments within PGCs during their accretion to form stars produces extremely intermittent lognormal distributions of the object number density $n_p$, therefore increasing the minimum number of independent samples required for statistically significant estimates of the mean density $\langle n_p \rangle$ by star-microlensing far beyond practical limits. Sample calculations (\S \ref{sec1}, \cite{gib96}, 1997ab, 1999ab, 2000abc) give an estimated primordial fog particle (PFP, MBD) condensation mass range $M_{PFP} \approx 10^{-7} \-- 10^{-6} M_{\sun}$, supporting the \cite{sch96} quasar-microlensing conclusion that the lens galaxy mass of Q0957+561A,B is dominated by ``rogue planets.'' The EROS/MACHO star-microlensing exclusion of such micro-brown-dwarf objects as the missing Halo mass (\cite{alc98}) is attributed to the unwarranted assumption that the spatial distribution of such objects is uniform rather than extremely intermittent. \acknowledgments
\section{Introduction} The Virasoro master equation (VME) summarizes the general Virasoro construction$^{\rf{9},\rf{13}}$ on affine Lie algebra, \begin{equation} T(z)=L^{ab}:J_a(z)J_b(z): \hspace{.0666in}, \hspace{.3in} a,b=1,...,\textup{dim}g\ . \label{VMET} \end{equation} See Ref.~\rf{2} for a review of the VME and irrational conformal field theory. Subsequent developments in this subject include semiclassical examples of conformal blocks$^{\rf{18}}$ in irrational conformal field theory, unification$^{\rf{19}-\rf{22}}$ with the general non-linear sigma model and the construction of the stress tensors of twisted sectors of cyclic permutation orbifolds$^{\rf{1}}$ from copies of the general Virasoro construction (\ref{VMET}). Our starting point in this paper is the orbifold affine algebra$^{\rf{1}}$ at integer order $\lambda$, which was recently found in the twisted sectors of cyclic permutation orbifolds and includes affine Lie algebra when $\lambda=1$. Following the suggestion of Ref.~\rf{1}, we consider here the general Virasoro construction on orbifold affine algebra \begin{equation} \hat{T}(z)=\sum_{r=0}^{\lambda-1}{\mathcal{L}}_r^{ab}:\hat{J}_a^{(r)}(z)\hat{J}_b^{(-r)}(z):,\hspace{.2in}a,b=1,...,\textup{dim}g, \hspace{.2in} \lambda\in\Z^+ \end{equation} where $\hat{J}_a^{(r)}(z)$, $r=0,...,\lambda -1$ are the orbifold currents. When $\lambda=1$, the orbifold currents $\hat{J}_a^{(0)}(z)=J_a(z)$ satisfy affine Lie algebra and $\hat{T}(z)$ reduces to $T(z)$ in (\ref{VMET}). The stress tensor $\hat{T}(z)$ is conformal when ${\mathcal{L}}^{ab}_r$ satisfies the {\it{orbifold Virasoro master equation}} (OVME) at order $\lambda$, and the OVME contains the conformal field theories of the VME when $\lambda=1$. At higher $\lambda$, the OVME contains large classes of stress tensors of twisted sectors of conventional permutation orbifolds, and many other constructions. The generic construction of the OVME is apparently a twisted sector of an orbifold or a Ramond (spin) sector because the ground state conformal weights $\hat{\Delta}_0$ of the constructions \begin{equation} L(m\geq 0)|0\rangle=\delta_{m,0}\hat{\Delta}_0|0\rangle,\hspace{.2in}\hat{\Delta}_0\neq 0 \end{equation} are generically non-zero. New unitary solutions of the OVME are obtained however, with irrational central charge and/or irrational conformal weights, for which we have so far found no conventional orbifold interpretation. \section{Background} \setcounter{equation}{0} \subsection{Orbifold affine algebra} We consider the \textit{orbifold affine algebra}$^{\rf{1}}$ $g_\lambda$ \namegroup{jalg} \alpheqn \begin{equation} [\hat{J}^{(r)}_a(m\+\frac{r}{\lambda}),\hat{J}^{(s)}_b(n\+\frac{s}{\lambda})]\=if_{ab}^{ \ \ c}\hat{J}^{(r+s)}_c(m\+n\+\frac{r+s}{\lambda})\+\hat{G}_{ab}(m\+\frac{r}{\lambda})\delta_{m+n+\frac{r+s}{\lambda},0} \end{equation} \begin{equation} m,n\in\Z,\hspace{.3in}a,b=1,...,\textup{dim}g,\hspace{.3in} r,s = 0,..,\lambda-1 \end{equation} \begin{equation} \hat{J}_a^{(r)}(m+\frac{r}{\lambda})|0\rangle=0\textrm{\ \ when\ \ }m+\frac{r}{\lambda}\geq 0 \end{equation} \reseteqn where the order $\lambda$ of $g_\lambda$ is any positive integer and $g_{\lambda=1}$ is affine Lie algebra$^{\rf{10}-\rf{6}}$. The relations in (\ref{jalg}) are understood with the periodicity condition \begin{equation} \hat{J}_a^{(r\pm\lambda)}(m+\frac{r\pm\lambda}{\lambda})=\hat{J}_a^{(r)}(m\pm 1+\frac{r}{\lambda}) \label{percond} \end{equation} and the state $|0\rangle$ is called the ground state of $g_\lambda$. The quantities ${f_{ab}}^c$ are the structure constants of any semisimple Lie algebra $g=\oplus_Ig_I$, and the metric $\hat{G}_{ab}$ is \begin{equation} \hat{G}_{ab}=\oplus_I\hat{k}_I\eta_{ab}^I. \label{gdef} \end{equation} Here, $\eta^I_{ab}$ is the Killing metric of $g_I$ and $\hat{k}_I=\lambda k_I$ is the level of the associated orbifold affine subalgebra. For applications we restrict ourselves to simple compact $g$, \begin{equation} \hat{G}_{ab}=\hat{k}\eta_{ab},\hspace{.2in}\hat{k}=\lambda k,\hspace{.2in}\hat{x}=\frac{2\hat{k}}{\psi^2}=\lambda x,\hspace{.2in}x=\frac{2k}{\psi^2} \label{xdef} \end{equation} where $\psi$ is the highest root of $g$ and unitarity$^{\rf{1}}$ requires that the invariant level $x$ be a positive integer. We will also need the operator product form$^{\rf{1}}$ of $g_\lambda$ \alpheqn \begin{equation} \hat{J}^{(r)}_a(z)\hat{J}^{(s)}_b(w)= \frac{\hat{G}_{ab}\delta_{r+s,0\hspace{.05in}mod\hspace{.04in}\lambda}}{(z-w)^2} +\frac{if_{ab}^{ \ \ c}\hat{J}_c^{(r+s)}(w)}{(z-w)} + O((z-w)^0) \label{JJOPE} \end{equation} \begin{equation}\hat{J}^{(r)}_a(z)=\sum_{m\in\Z}\hat{J}^{(r)}_a(m+\frac{r}{\lambda})z^{-1-m-r/\lambda} \end{equation} \reseteqn where $\hat{J}^{(r)}_a(z)$ are the local orbifold currents. \subsection{Subalgebras of $g_\lambda$} \label{sec-sub} The algebra $g_\lambda$ contains a set of regularly embedded subalgebras $g_\eta\subset g_{\lambda}$, \alpheqn \begin{equation} \hat{\hat{J\hspace{.045in}^{}}}\hspace{-.05in}^{(r)}_a(m+\frac{r}{\eta})=\hat{J}^{(\lambda r/\eta)}_a(m+\frac{\lambda r/\eta}{\lambda}),\hspace{.25in}r=0,...,\eta-1,\hspace{.25in}\eta\in{\Z}^+,\hspace{.25in}\frac{\lambda}{\eta}\in{\Z}^+ \end{equation} \begin{eqnarray} [\hat{\hat{J\hspace{.045in}^{}}}\hspace{-.05in}^{(r)}_a(m+\frac{r}{\eta}),\hat{\hat{J\hspace{.045in}^{}}}\hspace{-.05in}^{(s)}_b(n+\frac{s}{\eta})] \hspace{-.02in} = \hspace{-.02in} i f_{ab}^{\ \ c} \hat{\hat{J\hspace{.045in}^{}}}\hspace{-.05in}^{(r+s)}_c(m\!+\!n\!+\!\frac{r+s}{\eta}) \!+\! \hat{G}_{ab}(m\!+\!\frac{r}{\eta})\delta_{m+n+\frac{r+s}{\eta},0} \end{eqnarray} \reseteqn which is isomorphic to the set of order $\eta$ orbifold affine algebras, taken at levels $\{\hat{k}_I=\lambda k_I\}$. The special case $\eta=1$ is known as the integral affine subalgebra$^{\rf{4},\rf{5},\rf{1}}$, whose generators \begin{equation} \hat{\hat{J\hspace{.045in}^{}}}\hspace{-.05in}^{(0)}_a(m)=\hat{J}^{(0)}_a(m) \end{equation} form an affine Lie algebra at levels $\{\hat{k}_I\}$. Related subalgebras include $h_{\eta}\subset g_{\lambda}$, $\lambda/\eta\in{\Z}^+$ where $h\subset g$ is any Lie subalgebra of $g$. There are many other subalgebras of $g_\lambda$, for example the algebra generated by \begin{equation} \hat{J}_a^{(\mu n\hspace{.05in}mod\hspace{.05in}\lambda)}(\lfloor\frac{\mu n}{\lambda}\rfloor+\frac{\mu n\hspace{.05in}mod\hspace{.05in}\lambda}{\lambda}),\hspace{.2in} n\in{\Z}, \hspace{.1in}\mu=0,1,2,... \hspace{.1in} \end{equation} where $\lfloor x \rfloor$ is the integer less than or equal to $x$. When $\mu=\frac{\lambda}{\eta}\in\Z^+$, these are the generators \begin{equation} \hat{\hat{J\hspace{.045in}^{}}}\hspace{-.05in}^{(r)}_a(m+\frac{r}{\eta})=\hat{J}^{(\mu r)}_a(m+\frac{\mu r}{\lambda}),\hspace{.5in}r=0,...,\eta-1 \end{equation} of the $g_\eta$ subalgebras above. \subsection{Normal ordering} In Ref. \rf{1}, the (mother theory) normal-ordering convention \begin{eqnarray} :\hat{A}^{(r)}(z)\hat{B}^{(s)}(z):_M & \equiv & \sum_{m=-\infty}^{\infty}z^{-\Delta(A)-\Delta(B)-m-\frac{r+s}{\lambda}} [\sum_{p\leq-\Delta(A)}\hat{A}^{(r)}(p+\frac{r}{\lambda}) \hat{B}^{(s)}(m-p+\frac{s}{\lambda}) \nonumber\\ & &+\sum_{p>-\Delta(A)}\hat{B}^{(s)}(m-p+\frac{s}{\lambda})\hat{A}^{(r)}(p+\frac{r}{\lambda})] \end{eqnarray} was used for the product of any two integer-moded orbifold principal primary fields, including the orbifold currents with (mother theory) conformal weights $\Delta(A)=\Delta(B)=1$. We will use instead the OPE normal-ordering convention \begin{equation} :\hat{A}^{(r)}(z)\hat{B}^{(s)}(z): \hspace{.1in}\equiv \ \oint_z\frac{(dx)}{x-z}\hat{A}^{(r)}(x)\hat{B}^{(s)}(z),\hspace{.5in} (dx)\equiv\frac{dx}{2\pi i} \end{equation} where the $x$ contour does not encircle the origin. In the case of the orbifold currents, the two conventions are related by \begin{equation} :\hat{J}_a^{(r)}(z)\hat{J}_b^{(s)}(z): \hspace{.1in}\equiv \hspace{.1in} :\hat{J}_a^{(r)}(z)\hat{J}_b^{(s)}(z):_M -\frac{irf_{ab}^{\hspace{.11in}c}\hat{J}_c^{(r+s)}(z)}{\lambda z} +\frac{r(\lambda-r)}{2\lambda^2z^2}\hat{G}_{ab}\delta_{r+s,0\hspace{.05in}mod\hspace{.04in}\lambda} \end{equation} and we note that the modes of the OPE normal-ordered bilinears satisfy \alpheqn \begin{equation} :\hat{J}_a^{(r)}(z)\hat{J}_b^{(-r)}(z): \hspace{.1in}= \hspace{.1in} \sum_{m\in\Z} :\hat{J}_a^{(r)}\hat{J}_b^{(-r)}:(m) \ z^{-m-2} \end{equation} \begin{equation} :\hat{J}_a^{(r)}\hat{J}_b^{(-r)}:(m\geq 0)|0\rangle=\frac{r(\lambda-r)}{2\lambda^2}\hat{G}_{ab}\delta_{m,0}|0\rangle,\hspace{.3in}r=0,...,\lambda-1 \label{eq:delta} \end{equation} \reseteqn because $:\hat{J}_a^{(r)}\hat{J}_b^{(-r)}:_M(m\geq 0)$ annihilates the ground state. The current bilinears also have a symmetry (see App. A) \begin{equation}:\hat{J}_{(a}^{(r)}(z)\hat{J}_{b)}^{(s)}(z): \hspace{.1in}=\hspace{.1in}:\hat{J}_{(a}^{(s)}(z)\hat{J}_{b)}^{(r)}(z): \label{keyident} \end{equation} which will play an important role in the general construction below. (The symmetry holds for $:JJ:_M$ as well.) \section{General Virasoro Construction} \setcounter{equation}{0} \subsection{Summary of the computation} Following the suggestion of Ref. \rf{1}, we consider the candidate stress tensors \namegroup{stresst} \alpheqn \begin{equation} \hat{T}(z)=\sum_{r=0}^{\lambda-1}{\mathcal{L}}_r^{ab}:\hat{J}_a^{(r)}(z)\hat{J}_b^{(-r)}(z): \hspace{.1in} =\sum_{m\in{\Z}}L(m)z^{-m-2} \label{stress} \end{equation} \begin{equation} {\mathcal{L}}_r^{ab}={\mathcal{L}}_r^{ba},\hspace{.3in}r=0,...,\lambda-1 \label{orange} \end{equation} \reseteqn where $\{{\mathcal{L}}_r^{ab}\}$ is a set of $\lambda$ ``inverse inertia tensors''. Then, following Ref.$\hspace{.05in}$\rf{9}, we require that $\hat{T}$ satisfies the Virasoro algebra \begin{equation} \hat{T}(z)\hat{T}(w)=\frac{\hat{c}/2}{(z-w)^4}+\frac{2\hat{T}(w)}{(z-w)^2}+\frac{\partial_w\hat{T}(w)}{(z-w)}+O((z-w)^0) \hspace{.1in}. \label{viralg} \end{equation} This computation (see App. A) results in the restriction on ${\mathcal{L}}_r^{ab}$ \namegroup{restric} \alpheqn \begin{equation} {\mathcal{F}}_r^{ab} = {\mathcal{F}}_r^{ac} \hat{G}_{cd} {\mathcal{F}}_r^{db} - \frac{1}{2} \sum^{\lambda-1}_{s=0} {\mathcal{F}}_s^{cd} [ {\mathcal{F}}_{r+s}^{ef}f_{ce}^{\ \ a}f_{df}^{\ \ b} + f_{ce}^{\ \ f}f_{df}^{\ \ (a} {\mathcal{F}}_r^{b) e}] \label{restrica} \end{equation} \begin{equation} {\mathcal{F}}_r^{ab} \equiv {\mathcal{L}}_r^{ab}+{\mathcal{L}}_{\lambda-r}^{ab}, \hspace{.3in} r=0,...,\lambda-1\label{restricb} \end{equation} \reseteqn (where we have defined ${\mathcal{L}}_{r \pm \lambda}^{ab}\equiv{\mathcal{L}}_r^{ab}$) and the central charge, \begin{equation} \hat{c}=2 \hat{G}_{ab} \sum_{r=0}^{\lambda-1}{\mathcal{L}}_r^{ab}. \label{cc} \end{equation} The form of the restriction (\ref{restric}), in terms of the combination ${\mathcal{F}}_r^{ab}$, is a consequence of the symmetry (\ref{keyident}). Any solution of the system (\ref{restric}) gives a conformal stress tensor $\hat{T}$, but this system is problematic because it has more unknowns than equations. \subsection{Equivalent Solutions} The system (\ref{restric}) determines only the combination ${\mathcal{F}}_r^{ab}$ in (\ref{restricb}). This means that (\ref{restric}) has a ``gauge invariance'' under the gauge transformation \namegroup{gaug} \alpheqn \begin{equation} {\mathcal{L}}_r^{ab}\rightarrow ({\mathcal{L}}_r^{'})^{ab}={\mathcal{L}}_r^{ab}+\Lambda_r^{ab},\hspace{.3in} {\mathcal{L}}_{\lambda-r}^{ab}\rightarrow ({\mathcal{L}}_{\lambda-r}^{'})^{ab}={\mathcal{L}}_{\lambda-r}^{ab}-\Lambda_r^{ab} \hspace{.3in} \label{gauge} \end{equation} \begin{equation} {\mathcal{F}}_r^{ab} \rightarrow({\mathcal{F}}_r^{'})^{ab}= {\mathcal{F}}_r^{ab} \end{equation} \begin{equation} r=1,...,\lfloor\frac{\lambda-1}{2}\rfloor \end{equation} \reseteqn for any set of symmetric matrices $\Lambda_r^{ab}=\Lambda_r^{ba}$: If ${\mathcal{L}}_r^{ab}$ is a solution of (\ref{restric}) then ${{\mathcal{L}}_r^{'}}^{ab}$ is also a solution. Using the symmetry (\ref{keyident}) again, we see that the gauge transformation (\ref{gaug}) is also an invariance of the stress tensor \[ \hat{T}=\left\{\begin{array}{cl}{\mathcal{L}}_0^{ab}:\hat{J}_a^{(0)}\hat{J}_b^{(0)}: + {\mathcal{L}}_{\lambda/2}^{ab}:\hat{J}_a^{(\lambda/2)}\hat{J}_b^{(\lambda/2)}: + {\displaystyle\sum_{r=1}^{\frac{\lambda}{2}-1}} {\mathcal{F}}_r^{ab}:\hat{J}_a^{(r)}\hat{J}_b^{(-r)}:\hspace{.1in} \textrm{for}\ \lambda\hspace{.08in}\textrm{even}\\ {\mathcal{L}}_0^{ab}:\hat{J}_a^{(0)}\hat{J}_b^{(0)}: + {\displaystyle\sum_{r=1}^{\frac{\lambda-1}{2}}} {\mathcal{F}}_r^{ab} :\hat{J}_a^{(r)}\hat{J}_b^{(-r)}:\hspace{1.6in} \textrm{for}\ \lambda\hspace{.08in}\textrm{odd} \end{array}\right. \] \vspace{-.4in} \begin{equation}\end{equation} and similarly for the central charge. This tells us that all solutions in any given gauge orbit are physically equivalent, and we are entitled to choose a gauge. \subsection{The OVME} The most convenient gauge choice is \begin{equation} {\mathcal{L}}_r^{ab} = {\mathcal{L}}_{\lambda-r}^{ab}, \hspace{.1in} r=1,...,\lfloor\frac{\lambda-1}{2}\rfloor \end{equation} (choose $\Lambda_r^{ab}=({\mathcal{L}}_{\lambda-r}^{ab}-{\mathcal{L}}_r^{ab})/2$) because this choice preserves the form of ${\mathcal{L}}_r^{ab}$ given for the twisted sectors of cyclic orbifolds in Ref. \rf{1}. In this gauge we have \begin{equation} {\mathcal{F}}_r^{ab}=2{\mathcal{L}}_r^{ab} \end{equation} and the Virasoro condition (\ref{viralg}) is summarized by the system \namegroup{OVMEgrp} \alpheqn \begin{equation} \hat{T}=\sum_{r=0}^{\lambda-1}{\mathcal{L}}_r^{ab}:\hat{J}_a^{(r)}\hat{J}_b^{(-r)}: \label{stresstensor}\end{equation} \begin{equation} {\mathcal{L}}_r^{ab} = 2 {\mathcal{L}}_r^{ac} \hat{G}_{cd} {\mathcal{L}}_r^{db} - \sum^{\lambda-1}_{s=0} {\mathcal{L}}_s^{cd} [{\mathcal{L}}_{r+s}^{ef} f_{ce}^{\ \ a}f_{df}^{\ \ b} + f_{ce}^{\ \ f}f_{df}^{\ \ (a} {\mathcal{L}}_r^{b) e}],\hspace{.15in} 0\leq r\leq\lfloor\frac{\lambda}{2}\rfloor \label{OVME} \end{equation} \begin{equation} {\mathcal{L}}_r^{ab}={\mathcal{L}}_{r \pm \lambda}^{ab}={\mathcal{L}}_{\lambda \pm r}^{ab}, \hspace{.3in}\lfloor\frac{\lambda}{2}\rfloor<r\leq\lambda\label{period} \end{equation} \begin{equation} \hat{c}=2 \hat{G}_{ab} \sum_{r=0}^{\lambda-1}{\mathcal{L}}_r^{ab}. \end{equation} \reseteqn In what follows, we will refer to (\ref{OVME},c) as the {\textit{orbifold Virasoro master equation}} (OVME). As a check at $\lambda=1$, we define \begin{equation} L^{ab}\equiv{\mathcal{L}}^{ab}_0, \hspace{.3in} J_a\equiv\hat{J}_a^{(0)},\hspace{.3in}T\equiv\hat{T} \end{equation} and the OVME reduces to the Virasoro master equation$^{\rf{9},\rf{13}}$ (VME): \namegroup{VME} \alpheqn \begin{equation} T=L^{ab}:J_aJ_b: \label{VMEaa} \end{equation} \begin{equation} L^{ab} = 2 L^{ac} G_{cd} L^{db} - L^{cd} L^{ef} f_{ce}^{\ \ a}f_{df}^{\ \ b} - L^{cd} f_{ce}^{\ \ f}f_{df}^{\ \ (a}L^{b) e} \label{VMEa} \end{equation} \begin{equation} G_{ab}=\oplus_I k_I\eta_{ab}^I,\hspace{.2in} c=2G_{ab}L^{ab} \label{gabdef} \end{equation} \reseteqn as it should since $g_{\lambda=1}$ is affine Lie algebra. The solutions of the OVME are in one-to-one correspondence with the conformal stress tensors $\hat{T}$ in (\ref{stresstensor}), and we find \begin{equation} L(m\geq 0)|0\rangle=\delta_{m,0}\hat{\Delta}_0|0\rangle,\hspace{.2in}\hat{\Delta}_0 = \hat{G}_{ab} \sum_{r=0}^{\lambda -1} {\mathcal{L}}_r^{ab} \frac{r (\lambda - r)}{2 \lambda^2}\label{delta} \end{equation} where $\hat{\Delta}_0$ is the conformal weight of the ground state. This conformal weight is generically nonzero so that the generic construction of the OVME is like a twisted sector of an orbifold or a Ramond spin sector. The ground state conformal weight may occasionally vanish, e.g. the CFT's of the VME at $\lambda=1$ have $\hat{\Delta}_0=0$. Using the constraints (\ref{period}) to pull back the ${\mathcal{L}}$'s into the fundamental range\footnote{For $\lambda=1$ and $2$ the fundamental range in (\ref{frange}) is the original range in (\ref{orange}).}, \begin{equation} {\mathcal{L}}_r^{ab}, \hspace{.1in}r=0,...,\lfloor\frac{\lambda}{2}\rfloor \label{frange} \end{equation} the OVME has the explicit form: \underline{For even $\lambda$} \namegroup{ovmee} \alpheqn \begin{equation} \hat{T} = {\mathcal{L}}_0^{ab} :\hat{J}^{(0)}_a \hat{J}^{(0)}_b: + {\mathcal{L}}_{\lambda/2}^{ab} :\hat{J}_a^{(\lambda/2)} \hat{J}_b^{(\lambda /2)}: + 2 \sum_{r=1}^{\frac{\lambda}{2}-1}{\mathcal{L}}_r^{ab}:\hat{J}_a^{(r)}\hat{J}_b^{(-r)}:\end{equation} \begin{eqnarray} {\mathcal{L}}_r^{ab} &\hspace{-.15in}=\hspace{-.15in} & 2 {\mathcal{L}}_r^{ac} \hat{G}_{cd} {\mathcal{L}}_r^{db} - (2 \hspace{-.075in}\sum^{\frac{\lambda}{2}-r-1}_{s=0} \hspace{-.05in} {\mathcal{L}}_s^{cd} {\mathcal{L}}_{r+s}^{ef}+\hspace{-.1in} \sum^{\frac{\lambda}{2}}_{s=\frac{\lambda}{2}-r} \hspace{-.05in} {\mathcal{L}}_s^{cd} {\mathcal{L}}_{\lambda-r-s}^{ef} + \sum^{r}_{s=0} {\mathcal{L}}_s^{cd} {\mathcal{L}}_{r-s}^{ef}-2{\mathcal{L}}_0^{cd}{\mathcal{L}}_r^{ef})f_{ce}^{\ \ a}f_{df}^{\ \ b} \nonumber\\ && - ({\mathcal{L}}_0^{cd} + {\mathcal{L}}_{\lambda/2}^{cd} + 2 \sum_{s=1}^{\frac{\lambda}{2}-1}{\mathcal{L}}_s^{cd} )f_{ce}^{\ \ f}f_{df}^{\ \ (a} {\mathcal{L}}_r^{b) e},\hspace{.4in}r=0,...,\frac{\lambda}{2} \label{ovmee}\end{eqnarray} \begin{equation} \hat{c}=2 \hat{G}_{ab}{\mathcal{L}}_0^{ab} + 2 \hat{G}_{ab} {\mathcal{L}}_{\lambda/2}^{ab} + 4 \hat{G}_{ab}\sum_{r=1}^{\frac{\lambda}{2}-1}{\mathcal{L}}_r^{ab}\end{equation} \begin{equation} \hat{\Delta}_0 = \hat{G}_{ab} {\mathcal{L}}_{\lambda/2}^{ab} + \hat{G}_{ab} \sum_{r=1}^{\frac{\lambda}{2}-1} {\mathcal{L}}_r^{ab} \frac{r (\lambda - r)}{\lambda^2} .\end{equation} \reseteqn \underline{For odd $\lambda$} \namegroup{ovmeo} \alpheqn \begin{equation} \hat{T} = {\mathcal{L}}_0^{ab} :\hat{J}^{(0)}_a \hat{J}^{(0)}_b: + 2 \sum_{r=1}^{\frac{\lambda-1}{2}}{\mathcal{L}}_r^{ab}:\hat{J}_a^{(r)}\hat{J}_b^{(-r)}: \end{equation} \begin{eqnarray} {\mathcal{L}}_r^{ab} & = & 2 {\mathcal{L}}_r^{ac} \hat{G}_{cd} {\mathcal{L}}_r^{db} - (2\sum^{\frac{\lambda-1}{2}-r}_{s=1} {\mathcal{L}}_s^{cd} {\mathcal{L}}_{r+s}^{ef} +\sum^{\frac{\lambda-1}{2}}_{s=\frac{\lambda-1}{2}+1-r} {\mathcal{L}}_s^{cd} {\mathcal{L}}_{\lambda-r-s}^{ef} +\sum^{r}_{s=0} {\mathcal{L}}_s^{cd} {\mathcal{L}}_{r-s}^{ef})f_{ce}^{\ \ a}f_{df}^{\ \ b} \nonumber\\ & & - ({\mathcal{L}}_0^{cd} + 2 \sum_{s=1}^{\frac{\lambda-1}{2}}{\mathcal{L}}_s^{cd} )f_{ce}^{\ \ f}f_{df}^{\ \ (a} {\mathcal{L}}_r^{b) e},\hspace{.4in}r=0,...,\frac{\lambda-1}{2} \label{ovmeo} \end{eqnarray} \begin{equation} \hat{c}=2 \hat{G}_{ab} {\mathcal{L}}_0^{ab} + 4 \hat{G}_{ab} \sum_{r=1}^{\frac{\lambda-1}{2}}{\mathcal{L}}_r^{ab} \end{equation} \begin{equation} \hat{\Delta}_0 = \hat{G}_{ab} \sum_{r=1}^{\frac{\lambda-1}{2}} {\mathcal{L}}_r^{ab} \frac{r (\lambda - r)}{\lambda^2} .\end{equation} \reseteqn In these relations, we have used the convention that a sum is zero when its lower limit is greater than its upper limit. The forms (\ref{ovmee}) and (\ref{ovmeo}) of the OVME are useful for computational purposes, while the form (\ref{OVME}) is useful for studying general properties of the system. \section{Properties of the OVME} \setcounter{equation}{0} \subsection{Counting} The OVME is a set of \begin{equation} n(g,\lambda)=(\lfloor \frac{\lambda}{2} \rfloor +1)\frac{\textup{dim}g (\textup{dim}g +1)}{2} \label{numbeqns} \end{equation} coupled quadratic equations for the same number of unknowns (\ref{frange}). The number of physically inequivalent solutions to the OVME expected$^{\rf{2}}$ at each level $\hat{k}$ is therefore \begin{equation} N(g,\lambda)\approx 2^{n(g,\lambda)-{\scriptsize \textup{dim}}g} \label{numbsolns} \end{equation} where we have subtracted the degrees of freedom associated with the Lie $g$ covariance$^{\rf{2}}$ \begin{equation} {\mathcal{L}}_r^{'ab}={\mathcal{L}}_r^{cd}(\omega^{-1})_c^{\ a}(\omega^{-1})_d^{\ b},\ \ \omega \in Aut(g) \end{equation} of the OVME. As examples of (\ref{numbsolns}), one finds \alpheqn \begin{eqnarray} &&N(SU(2),\lambda=1)= 8,\hspace{.466in} N(SU(3),\lambda=1)\approx\frac{1}{4} \ \textrm{billion}\\ &&N(SU(2),\lambda=2)=512,\hspace{.3in} N(SU(3),\lambda=2)\approx 18 \ \textrm{quintillion} \end{eqnarray} \reseteqn where $\lambda=1$ is the VME and the growth with $\lambda$ at fixed $g$ is exponential. It is clear that, as in the VME, most of the solutions of the OVME will be new and (because the OVME is a large set of coupled quadratic equations) the new constructions will have generically irrational conformal weights and central charges. \subsection{${\mathcal{L}}_{r}^{ab}$ independent of $r$} For any $\lambda$, we consider the simple consistent ansatz \begin{equation} {\mathcal{L}}_r^{ab}(\hat{G})={\mathcal{L}}^{ab}(\hat{G}), \end{equation} ($\hat{G}$ is defined in (\ref{gdef})) which collects all inverse inertia tensors independent of $r$. Then the OVME simplifies to a rescaled VME \begin{equation} {\mathcal{L}}^{ab}(\hat{G}) = \lambda[2 {\mathcal{L}}^{ac}(\hat{G}) G_{cd} {\mathcal{L}}^{db}(\hat{G}) - {\mathcal{L}}^{cd}(\hat{G}) {\mathcal{L}}^{ef}(\hat{G}) f_{ce}^{\ \ a}f_{df}^{\ \ b} - {\mathcal{L}}^{cd}(\hat{G}) f_{ce}^{\ \ f}f_{df}^{\ \ (a}{\mathcal{L}}^{b) e}(\hat{G})] \label{rindepOVME} \end{equation} where $G_{ab}$ is defined in (\ref{gabdef}). Inspection of (\ref{rindepOVME}) gives \namegroup{rindep} \alpheqn \begin{equation} \hat{T}={\mathcal{L}}^{ab}(\hat{G})\sum_{r=0}^{\lambda-1}:\hat{J}_a^{(r)}\hat{J}_b^{(-r)}: \label{rindepT} \end{equation} \begin{equation} {\mathcal{L}}^{ab}(\hat{G})=\frac{1}{\lambda}L^{ab}(G) \label{rindeptensor} \end{equation} \begin{equation} \hat{c}=\lambda c= 2 \lambda G_{ab}L^{ab}(G), \hspace{.2in} \hat{\Delta}_0=\frac{\hat{c}}{24}(1-\frac{1}{\lambda^2}) \end{equation} \reseteqn where $L^{ab}(G)$ is any solution of the VME (\ref{VMEa}) on semisimple affine Lie $g$. The constructions (\ref{rindep}) were given in Ref.~\rf{1}. They are the stress tensors of the twisted sectors of all the cyclic permutation orbifolds \begin{equation} \frac{A\stackrel{\lambda\hspace{.05in}times}{\times...\times}A}{\Z_{\lambda}}\label{corb} \end{equation} which can be constructed from $\lambda$ copies of any affine-Virasoro construction (\ref{VMEaa}). In what follows, we will call (\ref{rindep}) the set of \textit{cyclic constructions}.\footnote{The stress tensors of the twisted sectors of the $S_n$ permutation orbifolds can also be constructed$^{\rf{25},\rf{1},\rf{23},\rf{24}}$ as sums of commuting copies of the cyclic constructions (\ref{rindep}). This is because every element of $S_n$ can be expressed as a product of disjoint cyclic permutations.} Because they operate in cyclic permutation orbifolds, the cyclic constructions form a Virasoro subalgebra of the orbifold Virasoro algebra$^{\rf{1}}$ \alpheqn \begin{equation} \hat{T}^{(r)}(z)\hat{T}^{(s)}(w) = \frac{(\hat{c}/2)\delta_{r+s,0\hspace{.05in}mod\hspace{.03in}\lambda}}{(z-w)^4} + \frac{2 \hat{T}^{(r+s)}(w)}{(z-w)^2} + \frac{\partial_w\hat{T}^{(r+s)}(w)}{z-w}+O((z-w)^0) \end{equation} \begin{equation}\hat{T}^{(r)}(z)=\frac{1}{\lambda}L^{ab}(G)\sum_{s=0}^{\lambda-1}:\hat{J}_a^{(s)}(z)\hat{J}_b^{(r-s)}(z):\end{equation} \reseteqn where the zero-twist component $\hat{T}^{(0)}$ of the extended algebra is the cyclic construction $\hat{T}$ in (\ref{rindep}). Among the cyclic constructions, we note in particular the \textit{orbifold affine-Sugawara construction} $\hat{T}_{g_\lambda}$, \namegroup{oassemi} \alpheqn \begin{equation} \hat{T}_{g_\lambda}=\sum_{r=0}^{\lambda-1}({\mathcal{L}}_r^{ab})_{g_\lambda}:\hat{J}_a^{(r)}\hat{J}_b^{(-r)}: \end{equation} \begin{equation} ({\mathcal{L}}_{r}^{ab})_{g_\lambda}(\hat{G})=\frac{L^{ab}_g(G)}{\lambda}=\oplus_I\frac{\eta^{ab}_I}{2\hat{k}_I+\lambda Q_I} \end{equation} \begin{equation} \hat{c}_{g_\lambda}=\lambda c_g= \lambda \sum_I \frac{2\hat{k}_I\textup{dim}g_I}{2\hat{k}_I+\lambda Q_I} ,\hspace{.2in}\hat{\Delta}^{g_\lambda}_0=\frac{\hat{c}_{g_\lambda}}{24}(1-\frac{1}{\lambda^2}) \end{equation} \reseteqn where $Q_I$ is the quadratic Casimir of $g_I$ and $c_g$ is the central charge of the affine-Sugawara construction$^{\rf{6},\rf{8}-\rf{14}}$ at levels $\{k_I\}$ on semisimple $g$. The form of these constructions$^{\rf{4},\rf{5},\rf{1}}$ on simple $g$ is: \namegroup{oassimple} \alpheqn \begin{equation} ({\mathcal{L}}_{r}^{ab})_{g_\lambda}(\hat{k})=\frac{L^{ab}_g(k)}{\lambda}=\frac{\eta^{ab}}{2\hat{k}+\lambda Q_g} \end{equation} \begin{equation} \hat{c}_{g_\lambda}=\frac{\lambda\hat{x}\textup{dim}g}{\hat{x}+\lambda \tilde{h}_g}=\lambda c_g(x),\hspace{.2in}c_g(x)=\frac{x\textup{dim}g}{x+\tilde{h}_g} \end{equation} \begin{equation} \hat{\Delta}^{g_\lambda}_0=\frac{\hat{c}_{g_\lambda}}{24}(1-\frac{1}{\lambda^2}). \label{oas} \end{equation} \reseteqn Here $\tilde{h}_g=Q_g/\psi^2$ is the dual Coxeter number of $g$, $x\in\Z^+$ is the invariant level in (\ref{xdef}) and $c_g(x)$ is the central charge of the affine-Sugawara construction on simple $g$. The orbifold affine-Sugawara constructions include the affine-Sugawara constructions $T_g$ at $\lambda=1$, and we shall see below that the orbifold affine-Sugawara constructions play the same fundamental role in the OVME that the affine-Sugawara constructions play in the VME. \subsection{Constructions on subalgebras} The OVME at order $\lambda$ has solutions ${\mathcal{L}}(\lambda;h_\eta)$ given by \namegroup{sub} \alpheqn \[{\mathcal{L}}^{ab}_r(\lambda;h_\eta)=\left\{\begin{array}{l}{\mathcal{L}}_s^{AB}(h_\eta), \hspace{.5in}\textrm{if}\ \exists s\in [0,\eta -1]\ \textrm{s.t.}\ r=\frac{\lambda}{\eta}s\\ 0,\hspace{1in}\textrm{otherwise}\ \end{array}\right.\] \vspace{-.4in} \begin{equation}\end{equation} \vspace{-.2in} \begin{equation}A,B=1,...,\textup{dim}h,\hspace{.2in}r=0,...,\lambda-1,\hspace{.2in}\eta\in\Z^+,\hspace{.2in}\frac{\lambda}{\eta}\in{\Z^+} \end{equation} \begin{equation} \hat{c}(\lambda;h_\eta) = \hat{c}(h_\eta),\hspace{.2in} \ \hat{\Delta}_0(\lambda;h_\eta)=\hat{\Delta}_0(h_\eta) \end{equation} \reseteqn where ${\mathcal{L}}(h_\eta)$ is any solution to the OVME on $h_\eta$, $h\subset g$ (with central charge $\hat{c}(h_\eta)$ and ground state conformal weight $\hat{\Delta}_0(h_\eta)$). The order $\eta$ solutions appear at order $\lambda$ because $h_\eta\subset g_\lambda$ (see Section~\ref{sec-sub}). The subalgebra constructions (\ref{sub}) include in particular the orbifold affine-Sugawara construction on $h_\eta\subset g_\lambda$, \namegroup{th} \alpheqn \begin{equation} \hat{T}_{h_\eta}(\lambda;h_\eta)=\frac{\eta^{AB}}{2\hat{k}+\eta Q_h}\sum_{r=0}^{\eta-1}:\hat{J}_A^{(\frac{\lambda}{\eta}r)}\hat{J}_B^{(-\frac{\lambda}{\eta}r)}:,\hspace{.2in} A,B=1,...,\textup{dim}h \end{equation} \begin{equation} \hat{c}_{h_\eta}=\frac{r\eta\hat{x}\textup{dim}h}{r\hat{x}+\eta \tilde{h}_h} ,\hspace{.2in} \hat{\Delta}^{h_\eta}_0=\frac{\hat{c}_{h_\eta}}{24}(1-\frac{1}{\eta^2}) \end{equation} \reseteqn for simple $g$ and simple $h$ where $r$ is the index of embedding of $h\subset g$. We also mention the case $h_\eta=g_1$ of (\ref{sub}), which collects all constructions on the integral affine subalgebra: \alpheqn \begin{equation} {\mathcal{L}}_r^{ab}(\lambda;g_1) = \delta_{r,0} L^{ab}(\hat{G}) \end{equation} \begin{equation} \hat{T}(\lambda;g_1) = L^{ab}(\hat{G}):\hat{J}_a^{(0)}\hat{J}_b^{(0)}:\hspace{.05in} , \hspace{.3in} \hat{c}(\lambda;g_1) = 2L^{ab}(\hat{G})\hat{G}_{ab}, \hspace{.3in} \hat{\Delta}_0(\lambda;g_1)=0 \end{equation} \reseteqn where $L^{ab}(\hat{G})$ is any solution to the VME (\ref{VMEa}) with $G_{ab}$ replaced by $\hat{G}_{ab}$ in (\ref{gdef}). These stress tensors are isomorphic to the stress tensors of the VME and the ground state conformal weights vanish, as expected, for all these constructions because ${\mathcal{L}}_0^{ab}$ does not contribute to $\hat{\Delta}_0$ in (\ref{delta}). \subsection{K-conjugation covariance} \label{ksec} At order $\lambda$, the OVME exhibits \textit{K-conjugation covariance} through the orbifold affine-Sugawara construction $\hat{T}_{g_\lambda}$, which includes the familiar K-conjugation covariance$^{\rf{6},\rf{8},\rf{7},\rf{15},\rf{9}}$ through the affine-Sugawara construction when $\lambda=1$. This means (see App. A) that the K-conjugate inertia tensor $\tilde{{\mathcal{L}}}$ \begin{equation} \tilde{{\mathcal{L}}}_r^{ab}=({\mathcal{L}}_r^{ab})_{g_\lambda}-{\mathcal{L}}_r^{ab} \end{equation} is a solution of the OVME when ${\mathcal{L}}$ is a solution, and \alpheqn \begin{equation} \tilde{\hat{T}\hspace{.035in}}=\hat{T}_{g_\lambda}-\hat{T},\hspace{.2in}\tilde{\hat{c}\hspace{.018in}}=\hat{c}_{g_{\lambda}}-\hat{c},\hspace{.2in}\tilde{\hat{\Delta\hspace{.01in}}}_0=\hat{\Delta}^{g_\lambda}_0-\hat{\Delta}_0, \end{equation} \begin{equation} \hat{T}(z)\tilde{\hat{T}\hspace{.025in}}(w)=O((z-w)^0) \label{commute} \end{equation} \reseteqn where $\tilde{\hat{T}\hspace{.035in}}$ is the corresponding K-conjugate stress tensor. Drawing on experience with the VME, K-conjugation covariance tells us that the conformal field theory corresponding to $\hat{T}$ can be considered as a gauge theory$^{\rf{29},\rf{30},\rf{2},\rf{19}-\rf{22}}$ in which the cyclic permutation orbifold of $\lambda$ copies of WZW is gauged by the K-conjugate partner $\tilde{\hat{T}\hspace{.035in}}\hspace{.02in}$of $\hspace{.05in}\hat{T}$. The OVME has another covariance, which we call $Aut({\Z}_\lambda)$ covariance, that has no analog in the VME: For any automorphism of ${\Z}_\lambda$, the inertia tensor ${\mathcal{L}}^{'}$ \begin{equation}{\mathcal{L}}_r^{'ab}={\mathcal{L}}_{\phi(r)}^{ab},\hspace{.3in} \hat{c}^{'} =\hat{c}, \hspace{.3in}\phi\in Aut({\Z}_\lambda) \end{equation} is a solution when ${\mathcal{L}}$ is a solution. $Aut({\Z}_\lambda)$ covariance relates conformal constructions with the same central charge but generically different conformal weights. \subsection{Coset constructions} At order $\lambda$, consider the orbifold affine-Sugawara construction $\hat{T}_{h_\eta}\equiv\hat{T}_{h_\eta}(\lambda;h_\eta)$ on $h_\eta\subset g_\lambda$ where $g$ and $h$ are semisimple. (For simple $g$ and $h$, this construction is given in (\ref{th})). The K-conjugate partner of $\hat{T}_{h_\eta}$ is the general $g_{\lambda}/h_{\eta}$ coset construction, \begin{equation} \hat{T}_{g_\lambda/h_\eta}=\tilde{\hat{T}\hspace{.035in}} = \hat{T}_{g_\lambda}-\hat{T}_{h_\eta}, \hspace{.3in}\hat{c}_{g_\lambda/h_\eta}=\tilde{\hat{c}\hspace{.018in}}=\hat{c}_{g_\lambda}-\hat{c}_{h_\eta} \label{coset} \end{equation} which includes the ordinary coset constructions$^{\rf{6},\rf{8},\rf{7}}$ at $\lambda=1$. The special case of (\ref{coset}) given by Ka\u{c} and Wakimoto$^{\rf{4},\rf{5}}$ is $g_\lambda/g_{\eta=1}$, that is $\eta=1$ and $h=g$ with \begin{equation} \hat{c}_{g_\lambda/g_{\eta=1}}=\hat{c}_{g_\lambda/g_1}=\lambda c_g(x)-c_g(\lambda x). \end{equation} The extension to $g_{\lambda}/h_{\eta=1}$ was given in Ref.~\rf{1}. It was conjectured in Ref.~\rf{1} that the coset construction $g_{\lambda}/h_{\eta=1}$ corresponds to the twisted sectors of a different kind of orbifold \begin{equation} \frac{(\frac{g\stackrel{\lambda\hspace{.05in}times}{\times...\times}g} {h_D(\lambda)})}{{\Z}_{\lambda}} \end{equation} where $h_D(\lambda)$ is the diagonal subalgebra of $h \times...\times h$. \subsection{Nests} Repeated K-conjugation on nested orbifold affine subalgebras $g_\lambda\supset h^{(1)}_{\eta_1}\supset...\supset h^{(n)}_{\eta_n}$ gives the \textit{orbifold affine-Sugawara nests}, which generalize the affine-Sugawara nests$^{\rf{34},\rf{16},\rf{2}}$ of the VME on $g\supset h^{(1)}\supset...\supset h^{(n)}$. The orbifold affine-Sugawara constructions and the coset constructions are the lowest nests, and the first non-trivial nests are \alpheqn \begin{equation} \hat{T}_{g_\lambda/h_\eta/h^\prime_{\eta^\prime}} = \hat{T}_{g_\lambda} - (\hat{T}_{h_\eta}-\hat{T}_{h^\prime_{\eta^\prime}}) \end{equation} \begin{equation} \hat{c}_{g_\lambda/h_\eta/h^\prime_{\eta^\prime}} = \hat{c}_{g_\lambda} - (\hat{c}_{h_\eta}-\hat{c}_{h^\prime_{\eta^\prime}}). \end{equation} \reseteqn More generally one has the \textit{orbifold affine-Virasoro nests}, including \alpheqn \begin{equation}\hat{T}^\#_{g_\lambda/h_\eta/h^\prime_{\eta^\prime}} = \hat{T}_{g_\lambda} - (\hat{T}_{h_\eta}-\hat{T}^\#_{h^\prime_{\eta^\prime}}) \end{equation} \begin{equation}\hat{c}^\#_{g_\lambda/h_\eta/h^\prime_{\eta^\prime}} = \hat{c}_{g_\lambda} - (\hat{c}_{h_\eta}-\hat{c}^\#_{h^\prime_{\eta^\prime}})\end{equation} \reseteqn where $\hat{T}^\#_{h^\prime_{\eta^\prime}}$ is an arbitrary construction on the subalgebra $h^\prime_{\eta^\prime}$. These nests generalize the affine-Virasoro nests$^{\rf{16},\rf{2}}$ familiar at $\lambda=1$. In what follows, we refer to the solutions discussed above (the cyclic constructions, the subalgebra constructions on $h_\eta\subset g_\lambda$ and the orbifold affine-Virasoro nests) as ``known'' solutions of the OVME and all other solutions will be called ``new''. \section{The Lie $g$-Invariant Constructions} Many consistent ans$\ddot{\textrm{a}}$tze$^{\rf{16},\rf{2}}$ and subans$\ddot{\textrm{a}}$tze can be found for the OVME, as for the VME. In this section we concentrate on the Lie $g$-invariant constructions on simple $g$, whose abundance is a surprising feature of the OVME. The generalization of the ``graph theory ansatz'' on $SO(n)$, familiar$^{\rf{28}}$ at $\lambda=1$, is also given in App.~B. \subsection{Group-invariant ansatz} \setcounter{equation}{0} We consider the group-invariant ansatz \begin{equation}\psi^2{\mathcal{L}}_r^{ab}=l_r \eta^{ab},\ \ \ r=0,...,\lfloor\frac{\lambda}{2}\rfloor \label{ginvar} \end{equation} ($\eta^{ab}$ is the inverse Killing metric of $g$) which collects all \textit {Lie $g$-invariant constructions} in the OVME on simple $g$. We know that this ansatz includes at least the trivial construction ${\mathcal{L}}=0$, the orbifold affine-Sugawara construction $\hat{T}_{g_\lambda}$, the particular subalgebra constructions $\hat{T}_{g_{\eta}}$, the coset constructions $g_\lambda /g_\eta$ and all orbifold affine-Sugawara nests of the form $g_\lambda\supset g_\eta \supset...\supset g_{\eta\prime}$. Among these only two constructions survive at $\lambda=1$, namely the trivial construction and the affine-Sugawara construction on $g$. We shall see that the number of new Lie $g$-invariant constructions increases rapidly with $\lambda$. Substitution of (\ref{ginvar}) into (\ref{OVME}) gives the reduced OVME of the ansatz: \namegroup{OVMELie} \alpheqn \begin{equation} \psi^2 \hat{T}=\sum_{r=0}^{\lambda-1}l_r\eta^{ab}:\hat{J}_a^{(r)}\hat{J}_b^{(-r)}: \end{equation} \begin{equation} l_r = \hat{x} l_r^{\ 2} + \tilde{h}_g\sum^{\lambda-1}_{s=0} l_s (2 l_r - l_{r+s}),\hspace{.2in}0\leq r\leq\lfloor\frac{\lambda}{2}\rfloor \label{g} \end{equation} \begin{equation} l_r=l_{r\pm\lambda}=l_{\lambda\pm r},\hspace{.2in}\lfloor\frac{\lambda}{2}\rfloor<r\leq\lambda \end{equation} \begin{equation} \hat{c}=\hat{x}\textup{dim}g(\sum_{r=0}^{\lambda-1}l_r),\hspace{.2in}\hat{\Delta}_0=\hat{x}\textup{dim}g\sum_{r=0}^{\lambda-1}l_r\frac{r(\lambda-r)}{4\lambda^2} \end{equation} \reseteqn where $\psi$ is the highest root of $g$ and $\hat{x}=\lambda x,\ x\in\Z^+$. For computational purposes we give also the forms of the reduced OVME in the fundamental range: \underline{For even $\lambda$} \namegroup{ge} \alpheqn \begin{equation} \psi^2\hat{T} = l_0\eta^{ab}:\hat{J}_a^{(0)}\hat{J}_b^{(0)}:+l_{\lambda/2}\eta^{ab}:\hat{J}_a^{(\lambda/2)}\hat{J}_b^{(-\lambda/2)}:+2\sum_{r=1}^ {\frac{\lambda}{2}-1}l_r\eta^{ab}:\hat{J}_a^{(r)}\hat{J}_b^{(-r)}: \end{equation} \vspace{-.2in} \begin{eqnarray} l_r & = & \hat{x} l_r^{\ 2} - \tilde{h}_g(2\sum^{\frac{\lambda}{2}-r-1}_{s=0} l_sl_{r+s} +\sum^{\frac{\lambda}{2}}_{s=\frac{\lambda}{2}-r} l_s l_{\lambda-r-s} +\sum^{r}_{s=0} l_s l_{r-s}-2l_0l_r)\nonumber\\ & &\ \ +2\tilde{h}_gl_r(l_0 + l_{\lambda/2} + 2 \sum_{s=1}^{\frac{\lambda}{2}-1}l_s ),\hspace{.4in}r=0,...,\frac{\lambda}{2} \label{geb} \end{eqnarray} \vspace{-.2in} \begin{equation} \hat{c}=\hat{x}\textup{dim}g(l_0+l_{\lambda/2}+2\sum_{r=1}^{\frac{\lambda}{2}-1}l_r),\hspace{.2in}\hat{\Delta}_0=\hat{x}\textup{dim}g(\frac{l_{\lambda/2}}{16}+\sum_{r=1}^{\frac{\lambda}{2}-1}l_r\frac{r(\lambda-r)}{2\lambda^2}). \end{equation} \reseteqn \underline{For odd $\lambda$} \namegroup{go} \alpheqn \begin{equation} \psi^2\hat{T} = l_0\eta^{ab}:\hat{J}_a^{(0)}\hat{J}_b^{(0)}:+2\sum_{r=1}^{\frac{\lambda-1}{2}}l_r\eta^{ab}:\hat{J}_a^{(r)}\hat{J}_b^{(-r)}: \end{equation} \vspace{-.2in} \begin{eqnarray} l_r & = & \hat{x} l_r^{\ 2} - \tilde{h}_g(2\sum^{\frac{\lambda-1}{2}-r}_{s=1} l_sl_{r+s} +\sum^{\frac{\lambda-1}{2}}_{s=\frac{\lambda-1}{2}+1-r} l_s l_{\lambda-r-s} +\sum^{r}_{s=0} l_s l_{r-s})\nonumber\\ & &\ \ +2\tilde{h}_gl_r(l_0 + 2 \sum_{s=1}^{\frac{\lambda-1}{2}}l_s ),\hspace{.4in}r=0,...,\frac{\lambda-1}{2} \label{gob} \end{eqnarray} \vspace{-.2in} \begin{equation} \hat{c} =\hat{x}\textup{dim}g(l_0+2\sum_{r=1}^{\frac{\lambda-1}{2}}l_r),\hspace{.2in}\hat{\Delta}_0=\hat{x}\textup{dim}g(\sum_{r=1}^{\frac{\lambda-1}{2}}l_r\frac{r(\lambda-r)}{2\lambda^2}). \end{equation} \reseteqn We see that the group-invariant ansatz (\ref{ginvar}) is a consistent ansatz, giving $\lfloor \frac{\lambda}{2} \rfloor +1$ coupled quadratic equations and unknowns. At each level $\hat{x}$, it follows that there are \begin{equation} N(\lambda) = 2^{\lfloor \frac{\lambda}{2} \rfloor + 1} \end{equation} Lie $g$-invariant solutions at order $\lambda$. In fact, these solutions are organized into $N(\lambda)$ level families$^{\rf{2}}$, according to the high-level behavior$^{\rf{28},\rf{2}}$ \begin{equation} l_r=\frac{\theta_r}{\hat{x}}+O(\hat{x}^{-2}),\hspace{.3in}\theta_r\in\{0,1\} ,\hspace{.3in}r=0,...,\lfloor\frac{\lambda}{2}\rfloor \end{equation} of each level family. Only four of these level families are ``known'' at prime $\lambda$, and so there will be \begin{equation} N_{new}(\lambda)=2^{\lfloor\frac{\lambda}{2}\rfloor+1}-4, \hspace{.3in} \lambda \ {\textrm{prime}} \end{equation} new Lie $g$-invariant level families at these orders. The high-level expansion$^{\rf{28},\rf{2}}$ of the Lie $g$-invariant level families also shows that each level family is unitary for $\hat{x}=\lambda x$, $x\in\Z^+$, at least down to some finite radius of convergence $\hat{x}_0=\lambda x_0$. Experience$^{\rf{2}}$ with the VME shows that $x_0$ is usually quite low, for example $x_0=1$ or $2$ (see also Subsecs. \ref{lambdafive} and \ref{lambdasix}). We turn now to finding exact solutions of Eqs.~(\ref{geb}) and (\ref{gob}). In solving these equations it was useful to note that only the $r=0$ equations have $l_0$ dependence on their right hand sides, which effectively reduces the number of equations by one. Then we were able to factorize$^{\rf{16},\rf{2}}$ linear combinations of the $r\neq0$ equations through $\lambda=6$. We find no new solutions \footnote{For example the four solutions at $\lambda=2$ are the trivial construction, the orbifold affine-Sugawara construction, the affine-Sugawara construction on the integral affine subalgebra and a Ka\u{c}-Wakimoto coset.} for $\lambda=1$, 2, 3 and 4. The results for $\lambda=5$ and $6$ are reported below. \subsection{Irrational conformal weights at $\lambda = 5$} \label{lambdafive} At $\lambda=5$ there are $2^{2+1}=8$ level families, four of which are new: \alpheqn \begin{eqnarray} l_0 & = & \frac{1}{2(\hat{x} + 5 \tilde{h}_g)} + \frac{\theta}{2(\hat{x} + \tilde{h}_g)}\\ l_1 & = & \frac{1}{2(\hat{x} + 5 \tilde{h}_g)}(1 + \eta\sqrt{\frac{\hat{x} + 5\tilde{h}_g}{\hat{x} + \tilde{h}_g}})\\ l_2 & = & \frac{1}{2(\hat{x} + 5 \tilde{h}_g)}(1 - \eta\sqrt{\frac{\hat{x} + 5\tilde{h}_g}{\hat{x} + \tilde{h}_g}})\\ \hat{c} &=& \frac{\hat{x}}{2} \ \textup{dim}g(\frac{5}{\hat{x} + 5 \tilde{h}_g} + \frac{\theta }{\hat{x} + \tilde{h}_g}), \hspace{.3in} \eta, \theta= \ \pm 1\ . \label{eq:c} \end{eqnarray} \reseteqn K-conjugation takes $(\theta,\eta) \rightarrow (-\theta,-\eta)$ and all these constructions are unitary \footnote{It also follows that the high-level expansions of these level families are convergent down to and including level $x=1$.} for $\hat{x} =5x$, $x \in \Z^+$ because the ${\mathcal{L}}$'s are real$^{\rf{2}}$. The central charges (\ref{eq:c}) are rational, but the conformal weights of the ground states of these constructions are generically irrational \begin{equation} \hat{\Delta}_0 = \frac{\hat{x} \textup{dim}g}{10} (\frac{1}{\hat{x} + 5 \tilde{h}_g} - \frac{\eta}{5 \sqrt{(\hat{x} + 5 \tilde{h}_g)(\hat{x} + \tilde{h}_g)}}) \label{cw5} \end{equation} and other conformal weights will be generically irrational because the ${\mathcal{L}}$'s are irrational. We will discuss this set of new constructions further in Sec. 6. \subsection{Unitary irrational central charge at $\lambda=6$} \label{lambdasix} At $\lambda =6$ there are $2^{3+1}=16$ level families, four of which are new: \alpheqn \begin{eqnarray} l_0 & = & \frac{1}{2(\hat{x} + 6 \tilde{h}_g)} ( 1 + \eta \frac{\tilde{h}_g(- \hat{x}^2 + 3 \hat{x} \tilde{h}_g - 6 \tilde{h}_g^2)}{(\hat{x} + \tilde{h}_g) \alpha}) + \frac{\theta}{2(\hat{x} + \tilde{h}_g)} \\ l_1 & = & \frac{1}{2(\hat{x} + 6 \tilde{h}_g)} ( 1 + \eta \frac{- \hat{x}^2 - 5 \hat{x} \tilde{h}_g + 18 \tilde{h}_g^2}{\alpha}) \\ l_2 & = & \frac{1}{2(\hat{x} + 6 \tilde{h}_g)} ( 1 + \eta \frac{\hat{x}^2 + 3 \hat{x} \tilde{h}_g - 6 \tilde{h}_g^2}{\alpha}) \\ l_3 & = & \frac{1}{2(\hat{x} + 6 \tilde{h}_g)} ( 1 + \eta \frac{\hat{x}^2 + \hat{x} \tilde{h}_g - 18 \tilde{h}_g^2}{\alpha})\\ \alpha &\equiv& \sqrt{\hat{x}^4 + 2 \hat{x}^3 \tilde{h}_g -19 \hat{x}^2 \tilde{h}_g^2 + 12 \hat{x} \tilde{h}_g^3 + 36 \tilde{h}_g^4}, \hspace{.3in} \eta, \theta= \pm 1. \label{alpha} \end{eqnarray} \reseteqn As in the previous case, K-conjugation takes $(\theta,\eta)\rightarrow (-\theta,-\eta)$. Using the fact that the dual Coxeter number $\tilde{h}_g$ is positive for all compact Lie algebras, we have checked that $\alpha$, and hence ${\mathcal{L}}_r^{ab}$, are real for all levels $\hat{x}=6x$, $x\in\Z^+$ of all compact Lie algebras. It follows that these constructions are also unitary\footnote{The high-level expansions of these level families are also convergent through level $x=1$.} for these levels and algebras. For these constructions, both the central charge and the ground state conformal weight are generically irrational: \alpheqn \begin{eqnarray} \hat{c}&=&\frac{\hat{x}}{2} \ \textup{dim}g(\frac{6}{\hat{x}+ 6\tilde{h}_g} + \eta \frac{\hat{x}^3 - 3 \hat{x}^2 \tilde{h}_g + 6 \hat{x}\tilde{h}_g^2}{\alpha (\hat{x}+6\tilde{h}_g)(\hat{x}+\tilde{h}_g)}+ \frac{\theta}{\hat{x} +\tilde{h}_g}) \label{c6}\\ \hat{\Delta}_0&=& \frac{\hat{x} \ \textup{dim}g}{288(\hat{x} + 6 \tilde{h}_g)} (35 + \eta \frac{15 \hat{x}^2 + 7 \hat{x} \tilde{h}_g - 78 \tilde{h}_g^2}{\alpha}) \end{eqnarray} \reseteqn due to the quantity $\alpha$ in (\ref{alpha}). For certain levels the central charge is rational (for example level $\hat{x}=6x=6$ of $SU(2)$ and $SU(3)$). The lowest irrational central charges for each of the simple Lie algebras occur at $\theta=\eta=-1$ and are listed below: \vspace{.3in} \hspace{1.766in}\begin{tabular}{r|l|l} &$\hat{x}$&$\hspace{.9in}\hat{c}$\\ \hline $SU(2)$ & $12 $&$ \hspace{.2in} \frac{9}{14}(5-\frac{1}{\sqrt{2}}) \approx 2.7597$ \\ $SO(3)$ & $12 $&$ \hspace{.2in}\frac{9}{14}(5-\frac{1}{\sqrt{2}}) \approx 2.7597$ \\ $SP(1)$ & $12 $&$ \hspace{.2in}\frac{9}{14}(5-\frac{1}{\sqrt{2}}) \approx 2.7597$ \\ $E_6$ & $6 $&$ \hspace{.2in}5- \frac{19}{\sqrt{601}} \approx 4.2249$ \\ $E_7$ & $6 $&$ \hspace{.2in}\frac{35}{8} - \frac{161}{8 \sqrt{769}} \approx 3.6492$ \\ $E_8$ & $6 $&$ \hspace{.2in}\frac{10}{3}-\frac{68}{3\sqrt{1471}} \approx 2.7423$ \\ $F_4$ & $6 $&$ \hspace{.2in}\frac{26}{5} - \frac{26}{5 \sqrt{46}} \approx 4.4330$ \\ $G_2$ & $6 $&$ \hspace{.2in}\frac{21}{5} - \frac{21}{5\sqrt{41}} \approx 3.5440$ \\ \end{tabular} \[\] At fixed $g$, the central charges increase monotonically with $\hat{x}$, and, for the classical Lie algebras (e.g. $SU(n)$), the central charges increase monotonically with $n$. The lowest unitary irrational central charge in this family of constructions is therefore \begin{equation} \hat{c}( (E_8)_{\hat{x}=6},\lambda=6)=\frac{10}{3}-\frac{68}{3\sqrt{1471}} \approx 2.7423 \label{eq:c2} \end{equation} and we note that level $\hat{x}=6$ of $E_8$ corresponds, via the orbifold induction procedure$^{\rf{1}}$, to the historic level $x=1$ of $E_8$. The value (\ref{eq:c2}) is less than the lowest known unitary irrational central charge among the cyclic permutation orbifolds (\ref{corb}) at $\lambda=6$, \begin{equation} \hat{c}=6 \ c(SU(3)_5)^{\#}_{D(1)})=6 \ (2)(1-\frac{1}{\sqrt{61}}) \approx 10.4635 \end{equation} where $c(SU(3)_5)^{\#}_{D(1)})$ is the lowest known$^{\rf{2}}$ unitary irrational central charge in the VME. The central charge in (\ref{eq:c2}) is also less than $\hat{c}=6\times 1=6$, which is the lower bound on possible unitary irrational central charge of any order $6$ cyclic permutation orbifold. \section{Discussion: The Lie $h$-invariant constructions} \setcounter{equation}{0} The solutions above for $\lambda \leq 6$ show that the Lie $g$-invariant constructions of the OVME are closely related to the Lie $h$-invariant constructions$^{\rf{17},\rf{2}}$, with $h \subset g$, which have been studied in the VME. Following Ref.~\rf{17}, the Lie $h$-invariant constructions of the OVME are those constructions on $g_{\lambda}$ whose inertia tensors are Lie $h$-invariant \begin{equation} \delta {\mathcal{L}}_r^{ab}={\mathcal{L}}_r^{c(a}f_{cd}^{\ \ b)} \psi^d=0, \hspace{.3in} r=0,...,\lfloor \frac{\lambda}{2} \rfloor \end{equation} where $\psi^a$ parameterizes the Lie group $H \subset G$ near the origin. Then we know from Ref.~\rf{17} that these constructions occur in Lie $h$-invariant quartets, octets, etc., and the currents of the Lie $h$-invariant stress tensors are either $(1,0)$ operators (associated to a global $h$-symmetry) or $(0,0)$ operators (associated to a local $h$-symmetry). In the case of the OVME, Lie $h$ symmetry guarantees at least a global or local $h_1$ symmetry, although $h_{\eta}$ symmetry can also appear. For example, the general coset construction on $g_{\lambda}/h_{\eta}$ lives in the Lie $h$-invariant quartet \begin{picture}(350,185)(0,0) \put(135,150){$\hat{T}_{g_\lambda /h_\eta}$} \put(170,155){\vector(1,0){95}} \put(205,160){$+\hat{T}_{h_\eta}$} \put(275,150){$\hat{T}_{g_\lambda}$} \put(150,140){\vector(0,-1){80}} \put(150,60){\vector(0,1){80}} \put(125,105){$K_{g_\lambda}$} \put(165,140){\vector(4,-3){110}} \put(271,60){\vector(-4,3){110}} \put(223,105){$K_{g_\lambda /h_\eta}$} \put(283,140){\vector(0,-1){80}} \put(283,60){\vector(0,1){80}} \put(290,105){$K_{g_\lambda}$} \put(145,45){$\hat{T}_{h_\eta}$} \put(270,50){\vector(-1,0){100}} \put(205,55){$+\hat{T}_{h_\eta}$} \put(280,45){$0$} \put(100,15){Figure 1: The simplest Lie $h$-invariant quartet} \end{picture}\\ where $K_{g_{\lambda}}$ is the usual K-conjugation $(\tilde{\hat{T}\hspace{.035in}} = \hat{T}_{g_\lambda}-\hat{T})$ through the orbifold affine-Sugawara construction on $g_{\lambda}$ and $K_{g_{\lambda}/h_{\eta}}$ is (a trivial example of) another K-conjugation through the coset construction itself ($\tilde{\hat{T}\hspace{.035in}} = \hat{T}_{g_\lambda/ h_\eta}-\hat{T}$ with $\hat{T}=0$ or $\hat{T}_{g_\lambda/ h_\eta})$. The general Lie $h$-invariant quartet has the form \begin{picture}(350,185)(0,0) \put(120,150){$\hat{T}(h_\eta \subset g_\lambda)$} \put(190,155){\vector(1,0){50}} \put(200,160){$+\hat{T}_{h_\eta}$} \put(250,150){$\hat{T}(h_\eta \subset g_\lambda)+\hat{T}_{h_\eta}$} \put(150,140){\vector(0,-1){80}} \put(150,60){\vector(0,1){80}} \put(125,105){$K_{g_\lambda}$} \put(165,140){\vector(4,-3){110}} \put(271,60){\vector(-4,3){110}} \put(223,105){$K_{g_\lambda /h_\eta}$} \put(283,140){\vector(0,-1){80}} \put(283,60){\vector(0,1){80}} \put(290,105){$K_{g_\lambda}$} \put(100,45){$\hat{T}_{g_\lambda}-\hat{T}(h_\eta \subset g_\lambda)$} \put(244,50){\vector(-1,0){53}} \put(204,55){$+\hat{T}_{h_\eta}$} \put(250,45){$\hat{T}_{g_\lambda/ h_\eta}-\hat{T}(h_\eta \subset g_\lambda)$} \put(100,15){Figure 2: The general Lie $h$-invariant quartet} \end{picture}\\ where $\hat{T}(h_{\eta} \subset g_{\lambda})$ is any locally Lie $h$-invariant construction on $g_{\lambda}$ (so that $\hat{T}$ commutes with the currents of $h_{\eta}$). The K-conjugation $K_{g_\lambda /h_\eta}$ through the coset construction relates the two locally-invariant constructions in the quartet, the other two constructions being globally invariant. For the Lie $g$-invariant ansatz, one should read $h_{\eta} \rightarrow g_{\eta}$. We have checked that the two new sets of constructions at $\lambda=5$ and $6$ are Lie $g$-invariant quartets with the form \begin{picture}(350,185)(0,0) \put(103,150){$\eta=+1,\ \theta=-1$} \put(191,155){\vector(1,0){50}} \put(200,160){$+\hat{T}_{g_1}$} \put(245,150){$\eta=+1,\ \theta=+1$} \put(150,140){\vector(0,-1){80}} \put(150,60){\vector(0,1){80}} \put(125,105){$K_{g_\lambda}$} \put(165,140){\vector(4,-3){110}} \put(271,60){\vector(-4,3){110}} \put(223,105){$K_{g_\lambda /g_1}$} \put(283,140){\vector(0,-1){80}} \put(283,60){\vector(0,1){80}} \put(290,105){$K_{g_\lambda}$} \put(103,45){$\eta=-1,\ \theta=+1$} \put(244,48){\vector(-1,0){53}} \put(204,55){$+\hat{T}_{g_1}$} \put(250,45){$\eta=-1,\ \theta=-1$} \put(100,15){Figure 3: New Lie $g$-invariant quartets at $\lambda=5,6$} \end{picture}\\ so that $K_{g_\lambda /g_1}$ is conjugation through the Ka\u{c}-Wakimoto coset construction in both these cases. The constructions with $\theta=-1$ have a local $g$-invariance while the constructions with $\theta=1$ have only a global $g$-invariance. To check these conclusions for all the Lie $g$-invariant constructions, we have evaluated the $\hat{T}\hat{J}^{(0)}$ OPE in this case to find \alpheqn \begin{equation} \hat{T}(z)\hat{J}_a^{(0)}(w)=M({\mathcal{L}})_a^{\ b} (\frac{1}{(z-w)^2}+\frac{\partial_w}{z-w})\hat{J}_b^{(0)}+O((z-w)^0) \end{equation} \begin{equation} M({\mathcal{L}})_a^{\ b}=\Delta\delta_a^{\ b}, \hspace{.3in} \Delta \equiv \hat{x}l_0+\tilde{h}_g\sum_{r=0}^{\lambda-1}l_r \label{DEL} \end{equation} \reseteqn where the matrix $M({\mathcal{L}})_a^{\ b}$ is defined for all ${\mathcal{L}}$ in App. A. We have also checked that \begin{equation} \Delta^2=\Delta \label{nil} \end{equation} using the reduced OVME (\ref{g}) of the Lie $g$-invariant ansatz. The result (\ref{nil}) verifies that the currents $\hat{J}_a^{(0)}$ of the integral affine subalgebra $g_{\eta=1}$ are either $(1,0)$ or $(0,0)$ operators in all Lie $g$-invariant constructions, as they should be. The specific identifications in Fig.~$3$ follow easily from the central charges in (\ref{eq:c}) and (\ref{c6}). Although all the Lie $g$-invariant constructions have a local or global symmetry associated to $g_{\eta=1}$, it is also possible to have larger symmetries associated to $h_\eta=g_{\eta}$, $\eta\geq 2$. Indeed, there is a quartet at $\lambda=4$ associated to $g_{\eta=2}$. One remaining question is the nature of the extra (perhaps discrete) symmetry of ${\mathcal{L}}_r^{ab}$ which dictates the larger symmetry. In the case of the $\lambda=5$ quartet, one sees further structure because there are only two independent central charges in (\ref{eq:c}). In fact, we may write these central charges as \[ \hat{c}=\frac{1}{2}(\hat{c}_{g_{\lambda=5}}+\theta \hat{c}_{g_{\eta=1}})=\left\{\begin{array}{cl} \frac{1}{2}(\hat{c}_{g_5}+\hat{c}_{g_1})\\ \frac{1}{2}\hat{c}_{g_5/g_1}\\ \end{array}\right. \] \vspace{-.4in} \begin{equation} \label{cquart} \end{equation} which shows that the two $g_1$-local theories $(\theta=-1)$ have the same central charge, and moreover that this central charge is exactly half of the central charge of the Ka\u{c}-Wakimoto coset. Local Lie $h$-invariant constructions with $c=\frac{1}{2}c_{g/h}$, called the self $K_{g/h}$-conjugate constructions \hspace{-.05in}$^{\rf{17},\rf{2}}$, are known from the VME, where $\tilde{\hat{T}\hspace{.035in}}$ and $\hat{T}$ in $\tilde{\hat{T}\hspace{.035in}} = \hat{T}_{g_\lambda}-\hat{T}$ are automorphically equivalent inertia tensors (related by a transformation in $Aut(h)$). The mechanism here, although similar, is not the same: In the present case, conjugation by $K_{g_5/g_1}$ $(\eta\rightarrow -\eta)$ is an example of the $Aut(\Z_\lambda)$ covariance of the OVME (see Subsec.~\ref{ksec}), so that the two constructions with $\hat{c}=\frac{1}{2}\hat{c}_{g_5/g_1}$ have generically different conformal weights. The central charges in (\ref{eq:c}) (or (\ref{cquart})) can also be written as \begin{equation} \hat{c}=5(\frac{c_g(x)}{2})+\theta (\frac{c_g(5x)}{2}) \label{orbc5}, \hspace{.2in} c_g(x)=\frac{x \textup{dim}g}{x+\tilde{h}_g}, \hspace{.3in} x\in\Z^+ \end{equation} where $c_g(x)$ is the central charge of the affine-Sugawara construction on $g$. Since conventional orbifoldization does not change the central charge of a theory, the form (\ref{orbc5}) suggests that the new constructions at $\lambda=5$ might be twisted sectors of orbifolds which start with copies of a conformal field theory whose central charge is $c=c_g/2$. Conformal field theories with $c=c_g/2$ are in fact known (the self K-conjugate constructions$^{\rf{28},\rf{2}}$ at $\lambda=1$), but these occur only for $\textup{dim}g= \textup{even}$, leaving us without a conventional orbifold interpretation for the new constructions at $\lambda=5$. For the new constructions at $\lambda=6$, we have no suggestion at present for a conventional orbifold interpretation. \section{Extensions} \setcounter{equation}{0} In this section we consider the ``Feigin-Fuchs'' and the inner-automorphic deformations of the general Virasoro construction on orbifold affine algebra. For $\lambda\ge 3$, further deformations by the full antisymmetric part of the current bilinears $:\hat{J}_{[a}\hat{J}_{b]}:$ may also be possible. \subsection{$\hat{c}$-changing deformations} The ``Feigin-Fuchs'' extension is \alpheqn \begin{equation} \hat{T}(z)=\sum_{r=0}^{\lambda-1}{\mathcal{L}}_r^{ab}:\hat{J}_a^{(r)}(z)\hat{J}_b^{(-r)}(z):+D^a\partial\hat{J}_a^{(0)}(z) \end{equation} \begin{equation} {\mathcal{L}}_r^{ab} = 2 {\mathcal{L}}_r^{ac} \hat{G}_{cd} {\mathcal{L}}_r^{db} - \sum^{\lambda-1}_{s=0}{\mathcal{L}}_s^{cd} [ {\mathcal{L}}_{r+s}^{ef} f_{ce}^{\ \ a}f_{df}^{\ \ b} + f_{ce}^{\ \ f}f_{df}^{\ \ (a} {\mathcal{L}}_r^{b) e}]+if_{cd}^{\ \ (a}{\mathcal{L}}_r^{b)c}D^d \nonumber \label{FFOVME} \end{equation} \begin{equation} 0\leq r\leq\lfloor\frac{\lambda}{2}\rfloor \end{equation} \begin{equation} {\mathcal{L}}_r^{ab}={\mathcal{L}}_{r \pm \lambda}^{ab}={\mathcal{L}}_{\lambda \pm r}^{ab}, \hspace{.3in}\lfloor\frac{\lambda}{2}\rfloor<r\leq\lambda \end{equation} \begin{equation} D^aM({\mathcal{L}})_a^{\ b}=D^b, \hspace{.3in}\hat{c}=2 \hat{G}_{ab} (\sum_{r=0}^{\lambda-1}{\mathcal{L}}_r^{ab}-6D^aD^b) \end{equation} \reseteqn where the matrix $M({\mathcal{L}})_a^{\ b}$ is given in (\ref{MN}). The ground state conformal weight $\hat{\Delta}_0$ is still given by (\ref{delta}), and these generalized $\hat{\textup{c}}$-changing deformations include the familiar $c$-changing deformations$^{\rf{26},\rf{9}}$ at $\lambda=1$. \subsection{$\hat{c}$-fixed deformations} Similarly, the generalized $\hat{\textup{c}}$-fixed deformations have the form \alpheqn \begin{equation} \hat{T}(z)=\sum_{r=0}^{\lambda-1}{\mathcal{L}}_r^{ab}:\hat{J}_a^{(r)}(z)\hat{J}_b^{(-r)}(z):+d^a \frac{\hat{J}_a^{(0)}(z)}{z} + \frac{1}{2z^2}\hat{G}_{ab}d^ad^b \end{equation} \begin{equation} d^aM({\mathcal{L}})_a^{\ b}=d^b \end{equation} \begin{equation} \hat{c}=2 \hat{G}_{ab} \sum_{r=0}^{\lambda-1}{\mathcal{L}}_r^{ab}, \hspace{.3in} \hat{\Delta}_0 = \frac{\hat{G}_{ab}}{2}(\sum_{r=0}^{\lambda -1} {\mathcal{L}}_r^{ab} \frac{r (\lambda - r)}{\lambda^2}+d^a d^b) \end{equation} \reseteqn and the OVME (\ref{OVMEgrp}) holds for ${\mathcal{L}}_r^{ab}$. These deformations describe inner automorphic twists or spectral flow in the twisted sectors (see also Subsec. \hspace{-.1in} \ref{doubtwist}), a phenomenon which is familiar$^{\rf{6},\rf{26},\rf{9}}$ at $\lambda=1$. \subsection{Deformation of the Lie $g$-invariant constructions} The Lie $g$-invariant constructions of Secs. $5$ and $6$ allow a large class of both the $\hat{c}$-changing and $\hat{c}$-fixed deformations. In this case we find that \alpheqn \begin{equation} f_{cd}^{\ \ (a}{\mathcal{L}}_r^{b)c} = 0 \label{FFdef} \end{equation} \begin{equation} M({\mathcal{L}})_a^{\ b}=\Delta \delta_a^{\ b} \ \rightarrow \ D^a(\Delta -1)=0 \label{FFeigen} \end{equation} \begin{equation} \Delta^2=\Delta\ \rightarrow \ \Delta\in\{0,1\} \label{FFeigenc} \end{equation} \reseteqn where $\Delta$ is given in (\ref{DEL}). The result (\ref{FFdef}) tells us that the extra term in the generalized OVME ({\ref{FFOVME}) vanishes, so that the reduced OVME (\ref{g}) is maintained for $\hat{c}$-fixed or $\hat{c}$-changing deformations of any Lie $g$-invariant construction. We have recalled in (\ref{FFeigenc}) that all currents $\hat{J}_a^{(0)}$ of any Lie $g$-invariant construction are either (1,0) or (0,0) operators. Then (\ref{FFeigen}) tells us that one may deform any Lie $g$-invariant construction by arbitrary $D_a$ or $d_a$ for any $\hat{J}_a^{(0)}$ which is a (1,0) current (global $g$ symmetry) of the construction. In particular the orbifold affine-Sugawara construction may be deformed by arbitrary $D$ or $d$, a fact which is familiar$^{\rf{26}}$ for the affine-Sugawara constructions at $\lambda=1$. \subsection{The doubly-twisted affine algebra} \label{doubtwist} The $\hat{c}$-fixed deformation of the orbifold affine-Sugawara construction for simple $g$ \namegroup{sflow} \alpheqn \begin{equation} \hat{T}_{g_\lambda}(z;d)=\hat{T}_{g_\lambda}(z)+d^a \frac{\hat{J}_a^{(0)}(z)}{z} + \frac{1}{2z^2}\hat{k}\eta_{ab}d^ad^b,\ \ \forall d \end{equation} \begin{equation} \hat{c}_{g_\lambda}(d)=\lambda c_{g_\lambda}(x),\hspace{.2in}\hat{\Delta}_0^{g_\lambda}(d)=\frac{\hat{c}_{g_\lambda}}{24}(1-\frac{1}{\lambda^2})+\frac{1}{2}\hat{k}\eta_{ab}d^ad^b \end{equation} \reseteqn (see Eqs. (\ref{oassemi}) and (\ref{oassimple})) describes spectral flow in the twisted sectors of the WZW permutation orbifolds \begin{equation} \frac{g(d)\stackrel{\lambda\hspace{.05in}times}{\times...\times}g(d)}{{\Z}_{\lambda}} \end{equation} where each copy of $g$ has been twisted by the same arbitrary vector $d$, using the known $\lambda=1$ form$^{\rf{26}}$ of (\ref{sflow}). Following Ref.~\rf{26}, the spectral flow (\ref{sflow}) is equivalent to introducing inner-automorphically twisted orbifold currents. In the Cartan-Weyl basis, these currents satisfy the ``doubly-twisted'' affine algebra \namegroup{dta} \alpheqn \begin{equation} [\hat{H}_A^{(r)}(m+\frac{r}{\lambda};d),\hat{H}_B^{(s)}(n+\frac{s}{\lambda};d)] =\hat{k}(m+\frac{r}{\lambda})\eta_{AB}\delta_{m+n+\frac{r+s}{\lambda},0} \end{equation} \begin{equation} [\hat{H}_A^{(r)}(m+\frac{r}{\lambda};d),\hat{E}_\alpha^{(s)}(n+\frac{s-d\cdot\alpha}{\lambda};d)] =\alpha_A\hat{E}_\alpha^{(r+s)}(m+n+\frac{r+s-d\cdot\alpha}{\lambda};d) \end{equation} \begin{equation} [\hat{E}_\alpha^{(r)}(m+\frac{r-d\cdot\alpha}{\lambda};d),\hat{E}_\beta^{(s)}(n+\frac{s-d\cdot\beta}{\lambda};d)]\hspace{2.5in} \end{equation} \[\hspace{.8in} =\left\{\begin{array}{lll} N(\alpha,\beta)\hat{E}_{\alpha+\beta}^{(r+s)}(m+n+\frac{r+s-d\cdot(\alpha+\beta)}{\lambda};d) && \textrm{if}\ (\alpha+\beta)\in\Delta\\ \alpha\cdot\hat{H}^{(r+s)}(m+n+\frac{r+s}{\lambda};d)+\hat{k}(m+\frac{r-d\cdot\alpha}{\lambda})\delta_{m+n+\frac{r+s}{\lambda},0} && \textrm{if}\ (\alpha+\beta)=0\\ 0 && \textrm{otherwise}\\ \end{array}\right. \] \( [\hat{H}_A^{(r)}(m+\frac{r}{\lambda};d)-\delta_{m+\frac{r}{\lambda},0}kd_A]|0\rangle=\hat{E}_\alpha^{(r)}(m+\frac{r-d\cdot\alpha}{\lambda};d)|0\rangle \nonumber =0 {\textrm{\ \ when\ \ }}m+\frac{r}{\lambda}\geq0 \) \begin{equation} A,B=1,...,{\textrm{rank}}g,\hspace{.4in}\alpha,\beta\in\Delta \end{equation} \begin{equation} m,n\in\Z,\hspace{.3in}r,s=0,...,\lambda-1,\hspace{.3in}\lambda\in\Z^+ \end{equation} \reseteqn for all $d_A$, where $\hat{H}_A$ and $\hat{E}_\alpha$ are the Cartan and root operators respectively. The relations (\ref{dta}) are understood with the periodicity conditions \alpheqn \begin{equation} \hat{H}_A^{(r\pm\lambda)}(m+\frac{r\pm\lambda}{\lambda};d)=\hat{H}_A^{(r)}(m\pm 1+\frac{r}{\lambda};d) \end{equation} \begin{equation} \hat{E}_\alpha^{(r\pm\lambda)}(m+\frac{r\pm\lambda-d\cdot\alpha}{\lambda};d)=\hat{E}_\alpha^{(r)}(m\pm1+\frac{r-d\cdot\alpha}{\lambda};d) \end{equation} \reseteqn in analogy to the conditions (\ref{percond}). The doubly-twisted algebra (\ref{dta}) reduces to the orbifold affine algebra (\ref{jalg}) when $d=0$ and it reduces to inner-automorphically twisted$^{\rf{32},\rf{26},\rf{31}}$ affine Lie algebra when $\lambda=1$. More generally, the doubly-twisted algebra exhibits an intricate interplay between the discrete orbifold twist (outer automorphism) and the continuous spectral flow (inner automorphism). The doubly-twisted algebra can be obtained by the orbifold induction procedure$^{\rf{1}}$ from the spectral flow at $\lambda=1$ given in Ref.~\rf{26}. The relations are \alpheqn \begin{eqnarray} \hat{H}_A^{(r)}(m+\frac{r}{\lambda};d) &\equiv&\hat{J}_A^{(r)}(m+\frac{r}{\lambda})+kd_A\delta_{m+\frac{r}{\lambda},0}\nonumber\\ &=&J_A(\lambda m+r;d)_{BOOST}\nonumber\\ &=&J_A(\lambda m+r)+kd_A\delta_{\lambda m+r,0} \end{eqnarray} \begin{eqnarray} \hat{E}_\alpha^{(r)}(m+\frac{r-d\cdot\alpha}{\lambda};d) &\equiv&\hat{E}_\alpha^{(r)}(m+\frac{r}{\lambda})\nonumber\\ &=&E_\alpha(\lambda m+r-d\cdot\alpha;d)_{BOOST}\nonumber\\ &=&E_\alpha(\lambda m+r) \end{eqnarray} \reseteqn where the currents $J_A(m;d)_{BOOST},\ E_\alpha(m-d\cdot\alpha;d)_{BOOST}$ are the inner automorphically-twisted currents of Ref.~\rf{26} and $J_A(m),\ E_\alpha(m)$ satisfy affine Lie algebra. \bigskip \noindent {\bf Acknowledgements} We thank L. Loveridge and C. Schweigert for helpful discussions. J.~E and M.~B.~H. were supported in part by the Director, Office of Energy Research, Office of Basic Energy Sciences, of the U.S. Department of Energy under Contract DE-AC03-76F00098 and in part by the National Science Foundation under grant PHY95-14797.
\section*{Abstract} In the quantization scheme which weakens the hermiticity of a Hamiltonian to its mere ${\cal PT}$ invariance the superposition $V(x) = x^2+ Ze^2/x$ of the harmonic and Coulomb potentials is defined at the purely imaginary effective charges ($Ze^2=if$) and regularized by a purely imaginary shift of $x$. This model is quasi-exactly solvable: We show that at each excited, $(N+1)-$st harmonic-oscillator energy $E=2N+3$ there exists not only the well known harmonic oscillator bound state (at the vanishing charge $f=0$) but also a normalizable $(N+1)-$plet of the further elementary Sturmian eigenstates $\psi_{\{n\}}(x)$ at eigencharges $f=f_{\{n\}}>0$, $n = 0, 1, \ldots, N$. Beyond the smallest multiplicities $N$ we recommend their perturbative construction. \vspace{9mm} \noindent PACS 03.65.Ge, 03.65.Fd \begin{center} \end{center} \newpage \section{Introduction} Schr\"{o}dinger equation for all the asymptotically harmonic oscillators in one dimension, \begin{equation} \left [-\,\frac{d^2}{dx^2} + x^2+2\alpha\,x -E + {\cal O} (1/x) \right ] \, \psi(x) = 0, \label{one} \end{equation} need not necessarily be kept defined just on the real axis of coordinates. Indeed, its available physical asymptotic solutions \begin{equation} \psi (x) \sim \exp\left[-\frac{x^2}{2}- \alpha\, x + b\, {\rm ln} \,(x) + {\cal O} (1/x) \right]\ ,\ \ \ \ \ b =(E+\alpha^2-1)/2 \label{polansatz} \end{equation} are normalizable not only on the real intervals $x > x_a\gg 1$ and $x < -x_a$ but also in their complex vicinity defined by the respective formulae \begin{equation} |x| \in (x_a, \infty), \ \ \ \ {\rm arg}\ x \in (-\pi/4, \pi/4)\ \ {\rm and}\ \ {\rm arg}\ x \in (-5\pi/4, -3\pi/4). \label{wedges} \end{equation} One may connect these two complex wedges by a contour of integration which is arbitrarily deformed within the domain of analyticity of the potential in question. In particular, one may choose the purely harmonic $V(x)=x^2$ and move the real axis of coordinates $x$ to a parallel line at a distance $a$. The new, shifted potential $V(x) = x^2 + 2aix - a^2$ loses its hermiticity and preserves only a certain symmetry with respect to a simultaneous change of the parity ${\cal P}$ ($x\to -x$) and of the time ordering ${\cal T}$ (which means the mere complex conjugation $i \to -i$ in time-independent cases), {\em without any change in the discrete spectrum itself}. Similar paradoxes seem to have inspired a deeper analysis of the various ${\cal PT}$ symmetric phenomenological models. Bessis and Zinn-Justin \cite{Bessis} related some of them to the so called Lee-Yang zeros in field theory \cite{Itzykson} and Bender with coauthors paid attention to their possible role in the parity breaking \cite{BM}, phase transitions \cite{BBprl} and quantum electrodynamics \cite{BMb}. ${\cal PT}$ symmetry does not prove less exciting on a purely methodical level. Its use may range from the fundamental problems of the ambiguities of quantization \cite{Meis} and of the exact solvability and Darboux transformations \cite{Cannata} up to the questions of convergence of perturbation expansions \cite{myctyri}. In this context, there also appeared an amazing discovery \cite{BBjpa} of the so called quasi-exact (which means incomplete \cite{Ushveridze}) solvability of quartic oscillators. In contrast to its unsolvable three-dimensional counterpart the latter quartic model did not contain the Coulombic component $ e^2/x $ \cite{Classif}. In all the one-dimensional Hermitean models this term is routinely being omitted due to its strongly singular character in the origin. In the ${\cal PT}-$symmetric non-hermitean setting, nevertheless, a purely imaginary shift {\em could} regularize this singular term in principle. This was our main inspiration. Even after the screening $x \to x-ic$ the persistent imaginary part of forces $V(x) \sim 1/(x-i\,c)$ would still violate unitarity and cause the predominance of unstable, resonant bound states. At this point we may recall the above-mentioned strategy which tries to compensate the instabilities by the constraint of ${\cal PT}$ symmetry. This re-defines the electric charge. With real $f$ in its effective ${\cal PT}-$symmetric value $Z e^2\equiv if$ the nontrivial interaction model need not even contain the cubic and quartic terms. Its Schr\"{o}dinger equation with the three real parameters $a$, $f$ and $c$ reads \begin{equation} \left[-\,\frac{d^2}{dx^2} + x^2+2iax + i\frac{f}{x-i\,c}\right]\, \psi(x) = E \psi(x), \ \ \ \ \ \psi(\pm \infty) = 0, \label{SE} \end{equation} and gives, presumably, a real and discrete spectrum of energies $E$. In what follows, we shall re-examine eq. (\ref{SE}) from the point of view of its possible quasi-exact solvability. We shall be able to show that its elementary $(N+1)-$plets of Sturmian eigenstates exist at any $N=0, 1, \ldots$ (Section 2) and that their explicit construction (mediated by the vanishing of an $(N+1) \times (N+1)-$dimensional secular determinant, see below) may be significantly facilitated via perturbative techniques (Section 3). A summary of our results will be outlined in Section 4. \section{Quasi-exact solutions} \subsection{Taylor series and its termination} Equations (\ref{one}) and (\ref{polansatz}) clarify the structure of all the normalizable solutions $\psi(x)$ of eq. (\ref{SE}) near $x = \pm \infty$. In the light of the identity $1/(x-ic) \equiv (x+ic)/(x^2+c^2)$ our model remains smooth and regular at all the finite coordinates and everywhere off the complex pole of $V(x)$ at $x = ic$. In the vicinity of this complex point it is convenient to demand that our wave functions vanish, $\psi (x) \approx x-ic$. Having in mind the possible deformations of the contour of integration, such a requirement is immediately inspired by the universal and widely accepted regularization of Schr\"{o}dinger equations near their strong singularities \cite{Frank}. On this basis let us now try to solve our present Schr\"{o}dinger bound state problem (\ref{SE}) by means of the following elementary harmonic-oscillator-like ansatz \begin{equation} \psi (x) = (c + ix)\, \exp\left[-\frac{x^2}{2}- iax\right]\, \varphi(x), \ \ \ \ \ \ \ \ \ \varphi(x) = \sum_{n=0}^{N} \,h_n \,(ix)^n. \label{ansatz} \end{equation} Such a terminating Taylor-series assumption fixes the energy (cf. eq. (\ref{polansatz})), \begin{equation} E=2N+a^2+3. \label{energy} \end{equation} Its insertion in our differential Schr\"{o}dinger bound state problem (\ref{SE}) leads to the equivalent $N+1$ recurrence relations $$ A_nh_{n-1}+B_nh_{n}+C_n h_{n+1} + D_n h_{n+2}=0, \ \ \ \ \ \ \ \ \ \ \ \ \ n = 0, 1, \ldots, N. $$ Keeping in mind that $h_{-1} = h_{N+1}=h_{N+2}=0$ we easily derive the values of the coefficients, $$ D_n=c(n+1)(n+2), \ \ \ \ C_n=(n+1)(n+2-2ac), $$ $$ B_n = -2a(n+1) - 2c(N+1-n)-f, \ \ \ \ \ \ A_n = -2(N+1-n). $$ They may be arranged in a square matrix with four diagonals, $$ Q = \left( \begin{array}{cccccc} B_{0} & C_0& D_0 & & & \\ A_1&B_{1} & C_1& D_1 && \\ & A_2&\ddots & \ddots& \ddots & \\ &&\ddots&\ddots&\ddots&D_{N-2}\\ &&&A_{N-1}&B_{N-1}&C_{N-1}\\ &&&&A_N&B_N \end{array} \right). $$ In notation with the row vectors $\vec{h}= (h_0, h_1, \ldots, h_N)$ our recurrences may be then re-interpreted as a non-hermitean matrix problem $$ Q\,\vec{h}^T=0. $$ Normalization $h_N=1$ implies its immediate compact and unique solution \begin{equation} h_{N-k-1}=\frac{1}{2^{k+1}\,(k+1)!} \det \left( \begin{array}{ccccc} B_{N-k} & C_{N-k}& D_{N-k} & & \\ A_{N-k+1}&\ddots & \ddots& \ddots & \\ &\ddots&\ddots&\ddots&D_{N-2}\\ &&A_{N-1}&B_{N-1}&C_{N-1}\\ &&&A_N&B_N \end{array} \right) \label{defi} \end{equation} for all the relevant indices $k = 0, 1, \ldots, N-1$. At the remaining, ``redundant" value of $k = N$ the left-hand-side symbol vanishes identically since, due to our assumptions, $h_{-1} \equiv 0$. With the linear charge-dependence in $B_n=B_n(f)=B_n(0)-f$ and in $Q = Q(f)=Q(0)-fI$ this is equivalent to the current secular equation \begin{equation} \det\left[ Q(0) - f I\right] = 0. \label{seculare} \end{equation} As a polynomial constraint of the $(N+1)-$st degree this determines an $(N+1)-$plet of eigencouplings $f=f_{\{k\}}$. They are complex in general. In the spirit of our introductory remarks it only remains for us to demonstrate that all these roots are real (i.e., not breaking the overall ${\cal PT}-$symmetry of our model), in a certain sub-domain of parameters at least. \subsection{Full ${\cal PT}$ symmetry in the large-screening domain} For convenience let us shift the variable $f = X+(N+2)(c-a)$ and look at Table~1. It lists the explicit secular equations at the first few lowest integers $N$. From the Table we may infer that all the zeros $X_{\{n\}}$ and/or $f_{\{n\}}$ of eq. (\ref{seculare}) are just functions of the single parameter $d = c+a$. This is not surprising. The change of the value of $a$ is just a shift of the integration path within the domain of analyticity of our potential. We may demonstrate the explicit form of this type of invariance of our differential eq. (\ref{SE}) algebraically: The equation remains the same after we compensate the change of the coordinate $x \to x+i\delta$ by the simultaneous shift $a \to a+\delta$ and $c \to c - \delta$ of our pair of parameters. The values of $f$ and $d$ remain unchanged. In our subsequent considerations we shall put $a=0$ without any loss of generality, therefore. Assuming that the strong Coulomb singularity lies off the real line, $c \neq 0$, we may introduce $\lambda = 1/c$ and re-scale $Y = X/c$. Our secular equation (\ref{seculare}) may be re-phrased as corresponding to the asymmetric linear algebraic problem \begin{equation} \left( \begin{array}{cccccc} -Y-N & 2\lambda& 2 & & & \\ -2N\lambda&-Y-N+2 & 6\lambda& 6 && \\ & (2-2N)\lambda&\ddots & \ddots& \ddots & \\ &&\ddots&\ddots&\ddots&N(N-1)\\ &&&-4\lambda&-Y+N-2&N(N+1)\lambda\\ &&&&-2\lambda&-Y+N \end{array} \right) \left( \begin{array}{c} h_0\\ h_1\\ h_2\\ \vdots\\ h_{N-1}\\ h_N \end{array} \right) = 0 . \label{linear} \end{equation} In the large-displacement limit $\lambda \to 0$ we get an exactly solvable case which determines the real eigencharges $f_{\{n\}}$. This may be proved easily since in this limit our ``Hamiltonian" $\lambda Q = H- Y\,I$ becomes an upper triangular matrix. This leads to the following closed formula $$ \left (Y_{\{0\}}, Y_{\{1\}}, \ldots, Y_{\{N\}} \right ) \to \left (Y^{[0]}_{\{0\}}, Y^{[0]}_{\{1\}}, \ldots, Y^{[0]}_{\{N\}} \right ) = \left (N, N-2, \ldots , -N+2, -N\right ) $$ which determines the real and approximately equidistant spectrum of eigencharges $f=f_{\{n\}} (c)= 2(n+1)c+{\cal O}(1/c)$ with $n = 0, 1,\ldots, N$. \section{Sturmian bound states at the general $N$} Once we fix the integer $N$ and recall the nonlinear algebraic definition (\ref{seculare}) of our $(N+1)$ eigencouplings $f = f_{\{n\}}$ we immediately imagine that the explicit construction of our Sturmians becomes more and more complicated for the larger dimensions $N$. Even if we extend, accordingly, the list of our secular polynomials in Table 1 we must necessarily resort to the purely numerical methods when trying to determine their exact roots $f$. The task is easy for the first few integers $N$ only. At the large $N$, our numerical algorithms may still start from the above equidistant asymptotic estimates and make use of the expected smooth change of the roots $f$ with the decrease of the parameter $|d|<\infty$. Indeed, the boundary of the domain $d> d_{critical}(N)$ where all our roots $f$ stay real is only slowly growing with $N$ since $d_{critical}(1)=2$, $d_{critical}(2)\approx 2.9865$, $d_{critical}(3)\approx 3.765$ etc. All this indicates that in practice a perturbative evaluation of the eigencharges could prove almost as efficient as their direct numerical determination, especially beyond the smallest $N$. \subsection{Perturbation expansions with $\lambda<\lambda_{critical}$} The inspection of the (extended) Table 1 reveals that \begin{equation} Y(\lambda)=Y^{[0]} + \lambda^2\,Y^{[2]} + \lambda^4 Y^{[4]} + \ldots\ . \label{eners} \end{equation} This means that all the odd perturbation corrections vanish identically, $Y^{[2k+1]}=0$. In the similar spirit we may also write \begin{equation} \vec{h}=\vec{h}(\lambda) = \vec{h}^{[0]} + \lambda\,\vec{h}^{[1]} + \lambda^2 \vec{h}^{[2]} + \ldots\ . \label{coeff} \end{equation} In order to appreciate the possible merits of such an approach let us return to our asymmetric eq. (\ref{linear}) and notice that at $\lambda=0$ it decays into a direct sum of the two linear equations. Each of them couples only $h_n$'s with the same parity of the subscript $n$. This is a pleasant simplification. For example, the five-dimensional unperturbed eigenvalue problem at $N=4$ decays into the separate two- and three-dimensional sub-equations. We may omit the redundant superscripts~$^{[0]}$ and display the latter subset for illustration, \begin{equation} \left( \begin{array}{ccc} -Y-4 & 2& 0 \\ 0 & -Y & 12\\ 0& 0& -Y+4 \end{array} \right) \left( \begin{array}{c} h_0\\ h_2\\ h_4 \end{array} \right) = 0. \label{trinear} \end{equation} Relaxing our above normalization convention and working in the integer arithmetics (i.e., in full precision, without any round-off errors) its first solution $ \vec{h}_{\{a\}}=(1,0,0)$ is found for the eigencharge $Y=Y_{\{a\}}=-4$ while $\vec{h}_{\{b\}}=(1,2,0)$ is obtained at the vanishing $Y_{\{b\}}=0 $ and $\vec{h}_{\{c\}}=(3,12,4)$ results from $Y_{\{c\}}=+4$. Returning to our matrix form $ Q\,\vec{h}^T=0$ of the differential Schr\"{o}dinger eq. (\ref{SE}) with ``Hamiltonian" $\lambda Q = H- Y\,I$ and decomposition $H=H^{[0]}+\lambda\,H^{[1]}$ at a general $N$ we may now solve it by means of the textbook Rayleigh-Schr\"{o}dinger perturbation theory \cite{Messiah}. In the present implementation of this recipe the $k-$th unknown perturbation corrections will be determined by the relation $$ \left (H^{[0]}-Y^{[0]}I\right ) \left ( \vec{h}^{[k]} \right )^T + \left (H^{[1]}-Y^{[1]}I\right ) \left ( \vec{h}^{[k-1]} \right )^T -\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ $$ $$ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ -Y^{[2]} \left ( \vec{h}^{[k-2]} \right )^T - \ldots -Y^{[k]} \left ( \vec{h}^{[0]} \right )^T=0. \label{RSE} $$ We have to solve it step-by-step, at all the subsequent perturbation orders ${\cal O}(\lambda^k) $ numbered by the integer $k=1, 2, \ldots$. For illustration let us pick up $N=2$ and $ Y^{[0]}=2$. In the first-order approximation with $k=1$ we normalize $h_N^{[1]}=0$ and drop all the superscripts $^{[1]}$ from $Y^{[1]}$ and $\vec{h}^{[1]}$. The perturbed equation (\ref{linear}) then acquires its ${\cal O}(\lambda)$ first-order form $$ \left( \begin{array}{rrr} -4 & 0&+ 2 \\ 0 & -2 & 0\\ 0& 0& 0 \end{array} \right) \left( \begin{array}{c} h_0\\ h_1\\ 0 \end{array} \right) + \left( \begin{array}{rrr} -Y & 2& 0 \\ -4 & -Y & 6\\ 0& -2& -Y \end{array} \right) \left( \begin{array}{c} 1\\ 0\\ 2 \end{array} \right) = 0. \label{twonear} $$ This implies that $Y = Y^{[1]}=0$ (as expected) and $\vec{h}=\vec{h}^{[1]}=(0,4,0)$. By the way, the latter result nicely illustrates a general feature: We might replace eq. (\ref{coeff}) by a component-by-component expansion in the powers of squares $\lambda^2$ in a way paralleling the expansion of charges $Y$ above. This reflects a symmetry of our model with respect to the change of sign of the parameter of screening $d$. \subsection{The role of asymmetry} We have established that our recurrences (\ref{linear}) provide a new, not entirely standard framework for application of perturbation theory. The manifestly non-hermitean two-diagonal structure of the underlying exactly solvable unperturbed Hamiltonians $H^{[0]}$ requires a modification of the formalism itself. The necessity of re-considering the standard concepts (say, of the model-space projector) seems to deserve a separate comment. The point is that due to the manifest difference between our original and transposed pseudo-Hamiltonians we must consider not only the ``direct" zero-order eigenvalue problem $ [Q(0)- f\,I] \left (\vec{h}^{[0]}\right )^T = 0$ (or rather equation $$ \left (H^{[0]} -Y_{\{\alpha\}}\,I\right ) \left ( \vec{h}_{\{\alpha\}}^{[0]}\right )^T= 0 $$ at a fixed subscript $\alpha = 0, 1, \ldots N$) but also its transposed, non-equivalent pendant $$ \left [ \left (H^{[0]}\right )^T-Y_{\{\alpha\}}\, I \right]\left ( \vec{g}_{\{\alpha\}}^{[0]}\right )^T = 0. $$ It is worth noticing that in a way paralleling eq. (\ref{defi}) above, all the left eigenvectors $\vec{g}=\vec{g}(\lambda)$ may be defined by closed formulae, $$ g_{k+1}=\frac{(N-k-1)!}{2^{k+1}\,(k+1)!} \det \left( \begin{array}{ccccc} B_{0} & C_{0}& D_{0} & & \\ A_{1}&\ddots & \ddots& \ddots & \\ &\ddots&\ddots&\ddots&D_{k-2}\\ &&A_{k-1}&B_{k-1}&C_{k-1}\\ &&&A_k&B_k \end{array} \right) . \label{defipr} $$ Their knowledge would simplify the large-order algorithm. Still, due to the unphysical, auxiliary character of all the left eigenvectors, it is shorter to generate and use them just in the zero order. Dropping their superscript $^{[0]}$ as redundant, we may return to our illustration (\ref{trinear}) and complement it by the left eigenvectors $\vec{g}_{\{a\}} =(4,-2,3)$, $\vec{g}_{\{b\}}=(0,1,-3)$ and $\vec{g}_{\{c\}} =(0,0,1)$. Incidentally, the overlaps $G = (g,h)$ form a diagonal matrix, $G_{11}=G_{33}=4$, $G_{22}=2$. This trivializes $F = G^{-1}$ needed in the model-space projectors $ P_{\{\alpha\}}= \vec{h}_{\{\alpha\}}^T F_{\alpha, \alpha} \vec{g}_{\{\alpha\}}. $ They are, counter-intuitively, non-diagonal but compatible with the usual completeness relation $ I = \sum_{\alpha,\beta}\, \vec{h}_{\{\alpha\}}^T F_{\alpha, \beta} \vec{g}_{\{\beta\}} $. \section{Concluding remarks} \subsection{Sturmians and the algebra $sl(2)$.} Our wave function (\ref{ansatz}) becomes purely real after the change of variables $x = -iy$ which rotates the axes by $\pi/2$. Of course, the exponential growth of our asymptotics (\ref{polansatz}) in the new artificial variable $y$ has no meaning at all and only the original coordinate $x$ remains physical. Still, the use of $y$ simplifies our ansatz (\ref{ansatz}). With $a=0$ it leads to a new differential Schr\"{o}{dinger equation $\hat{H}\hat{\varphi}(y) = f\,\hat{\varphi}(y)$ for real polynomials $\hat{\varphi}(y)= \varphi(-iy)$ themselves. In this notation our third Hamiltonian-like operator $\hat{H}$ is the quadratic function of generators of the complex $sl(2)$ Lie algebra, $$ \hat{H}={\cal J}^0 {\cal J}^{-} +2 {\cal J}^{+}-2c {\cal J}^{0} + (N+2) {\cal J}^{-} -2(N+c) $$ with $$ \left [ {\cal J}^{-} {\cal J}^{0}\right ] = {\cal J}^{-} = \frac{d}{dy}, $$ $$ \left [ {\cal J}^{-} {\cal J}^{+}\right ] = 2 {\cal J}^{0} = 2(y+c) \frac{d}{dy} -2N, $$ $$ \left [ {\cal J}^{0} {\cal J}^{+}\right ] = {\cal J}^{+} = (y+c)^2 \frac{d}{dy} -2(y+c)N. $$ This fits the general scheme \cite{Turbiner}, parallels closely the similar quartic-oscillator result of ref. \cite{BBjpa} (with the same algebra but different $\hat{H}$) and re-confirms our above conclusion that in spite of its non-hermiticity, our ${\cal PT}-$symmetric Schr\"{o}dinger eq. (\ref{SE}) is quasi-exactly solvable. In this context, we would like to emphasize that there exists a nice parallel between the Hermitean and ${\cal PT}-$symmetric quasi-exactly solvable models. Picking up the characteristic examples and adding our present results, we may now summarize and list the following four different possibilities and types of the quasi-exact constructions. \begin{itemize} \item The characteristic polynomial example $V(x) = \alpha\,x^2+\beta\,x^4+x^6$ with a constrained variability, say, of the coupling $\alpha$ leads to the existence of some $N$ elementary bound states at a finite multiplet of the binding energies in the hermitean case \cite{Singh}. \item With a constrained variability of the energy $E$ the non-polynomial hermitean potentials exemplified by $V(r) = x^2+\alpha/(1+\beta\,x^2)$ lead to the elementary Sturmian solutions at a finite $N-$plet of couplings $\alpha$ \cite{Flessas}. \item In the ${\cal PT}-$symmetric case, the complex quartic polynomial potential of ref. \cite{BBjpa}, $V(x) = \alpha\,x+\beta\, x^2+\gamma\,x^3-x^4$ with a constrained variability of $\alpha$, exhibits the quasi-exact solvability of the sextic-oscillator type. \item $V(x) = x^2 +i\alpha\,x+i\beta\,(x+i\gamma)/(x^2+\gamma^2)$ of the present model (\ref{SE}) is the ``missing" ${\cal PT}-$symmetric partner to the quasi-exact solvability of the non-polynomial Sturmian type. \end{itemize} \subsection{Complex charges} We have seen that the one-dimensional Schr\"{o}dinger equation becomes quasi-exactly solvable for Coulomb plus harmonic superpositions of potentials provided only that we regularize these forces in the ${\cal PT}-$symmetric manner. Multiplets of the Sturmian eigenstates acquire then an elementary polynomial form at certain purely imaginary couplings $Z e^2 = i\,f_{\{n\}}$ at $n = 0,1, \ldots, N$ and any $N = 0, 1, \ldots$. In a tentative physical support of these {\em complex} electric charges we might recall a few of their ${\cal PT}$ symmetric predecessors. A close connection exists with the Bessis' cubic force possessing a purely imaginary coupling $g = i\,f$. Its rigorous mathematical analysis has already been delivered, many years ago, by Calicetti et al \cite{Calicetti}. One may also mention an even closer parallel with the Bender's and Milton's electrodynamics which replaces, for several independent reasons \cite{BMb}, the charge $e$ itself (i.e., not its present square $e^2$) by a purely imaginary quantity. In a broader methodical context, our new phenomenological model (\ref{SE}) extends the family of potentials which are comparatively easily described in the language of analytic continuations \cite{Alvarez}. Its distinctive feature, in this setting, is the presence of a complex pole. Its most important formal merit is its quasi-exact solvability shared with the quartic model of ref. \cite{BBjpa}. In the latter comparison, one could emphasize the ``more natural" asymptotic behaviour of our present wave functions: The real axis of coordinates still lies within the wedges (\ref{wedges}) of the admissible analytic continuations. In the context of perturbative considerations it is amusing to notice that a ``nice" (which means real, discrete and bounded) character of spectrum of our present ${\cal PT}$ symmetric model may be viewed as a result of perturbation of {\em any} one of its {\em two} exactly solvable halves. \subsection*{Acknowledgement} Partly carried out within the frame of TMR - Network PECNO - ERB-FMRX-CT96-0057. \vspace{1cm} Table 1. The first six secular equations (\ref{seculare}) for eigencharges $f + (N+2)(a-c) \equiv X$, with a variable parameter $c+a\equiv d$ and abbreviation $d^2-N-3 \equiv h$. $$ \begin{array}{|c|c|} \hline N & \det Q(f)=0\\ \hline & \\ 0&-X=0\\ 1&{X}^{2}-h=0 \\ 2&-{X}^{3}+4\,h{X}+8\,d=0 \\ 3&{X}^{4}-10\,h{X}^{2}-48\,d X +9\,h^2-36 =0 \\ 4&-{X}^{5}+20\,h{X}^{3}+168\,d {X}^{2}-32\,\left (2\,h^2-9\right ) X-384\,h d=0 \\ 5&{X}^{6}-35\,h{X}^{4}-448\,d{X}^{3}+\left ( 259\,{h}^{2}-1296\right ){X}^{2}+ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \\ & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ + 3520\,h d X -225\,{h}^{3}+10000\,h+51200=0\\ & \\ \hline \end{array} $$ \newpage
\section{Introduction} Cosmological models of hierarchical merging in a cold dark matter universe are in some difficulty. High-resolution N-body simulations (Navarro et al. 1996a; NFW) have shown that the density profiles $\rho_{NFW}$ of virialized dark matter halos should have a universal shape of the form \begin{equation} \rho_{NFW} (r) = \frac{3 H_0^2}{8 \pi G} \frac{\delta_c }{(r/r_s)(1+r/r_s)^2} \end{equation} \noindent where $r_s$ is a characteristic length scale and $\delta_c$ a characteristic density enhancement. The two free parameters $\delta_c$ and $r_s$ can be determined from the halo concentration c and the virial mass $M_{200}$ \begin{eqnarray} \delta_c & = & \frac{200}{3} \frac{c^3}{\ln (1+c) - c/(1+c)} \\ r_s & = & \frac{R_{200}}{c} = \frac{1.63 \times 10^{-2}}{c} \left(\frac{M_{200}}{M_{\odot}} \right)^{1/3} h^{-2/3} kpc. \end{eqnarray} \noindent where $R_{200}$ is the virial radius, that is the radius inside which the average overdensity is 200 times the critical density of the universe and $M_{200}$ is the mass within $R_{200}$. For any particular cosmology there also exists a good correlation between c and $M_{200}$ which results from the fact that dark halo densities reflect the density of the universe at the epoch of their formation (NFW, Salvador-Sol\'{e} et al. 1998) and that halos of a given mass are preferentially assembled over a narrow range of redshifts. As lower mass halos form earlier, at times when the universe was significantly denser, they are more centrally concentrated. NFW have published concentrations for dark matter halos in the mass range of $3 \times 10^{11} M_{\odot} \leq M_{200} \leq 3 \times 10^{15} M_{\odot}$ which can be well fitted by the following power-law functions: \begin{eqnarray} c & = & 8.91 \times 10^2 \left(\frac{M_{200}}{M_{\odot}} \right)^{-0.14} \ \ \ \ \ {\it for} \ \ \ {\rm SCDM} \\ c & = & 1.86 \times 10^2 \left(\frac{M_{200}}{M_{\odot}} \right)^{-0.10} \ \ \ \ \ {\it for} \ \ \ {\rm CDM}\Lambda \nonumber \end{eqnarray} \noindent where SCDM denotes a standard biased cold dark matter model with $\Omega_0$=1, h=0.5, $\sigma_8$=0.65 and CDM$\Lambda$ denotes a low-density universe with a flat geometry and a non-zero cosmological constant, defined by $\Omega_0$=0.25, $\Omega_\Lambda$ = 0.75, h=0.75, $\sigma_8$=1.3. Note that the universal profile (equation 1) and the scaling relations (equation 4) have only been determined from simulations for halo masses as small as $M_{200}=10^{11}M_{\odot}$, but there is no reason to believe that these results would not be valid for halos which are, say, one order of magnitude lower in mass. In summary, dark matter halos represent a one-parameter family, with their density distribution being determined completely by their virial mass $M_{200}$. The universal character of dark matter profiles and the validity of the NFW-profile for different cosmogonies has been verified by many N-body calculations (e.g. Navarro et al. 1997, Cole \& Lacey 1997, Tormen et al. 1997, Tissera \& Dominguez-Tenreiro 1998, Jing 1999). In one of the highest resolution simulations to date, Moore et al. (1998, see also Fukushige \& Makino 1997) found good agreement with the NFW profile (equation 1) at and outside of $r_s$. Their simulations did however lead to a steeper innermost slope $\rho \sim r^{-1.4}$ which extends all the way down to their resolution limit of 0.01 $r_s$. On the analytical side, early spherically symmetric collapse models by Gunn \& Gott (1972) studied the collapse of a uniformly overdense region. Gott (1975) and Gunn (1977) investigated secondary infall onto already collapsed density perturbations and predicted $r^{-9/4}$ profiles. Fillmore \& Goldreich (1984) found self-similarity solutions for the secondary infall models. Hoffman \& Shaham (1985) took into account more realistic Gaussian initial conditions and predicted sharp central density peaks of the form $\rho \sim r^{-2}$. An updated version of these models by Krull (1999) abandoned self-similarity and explicitly took into account the hierarchical formation history. His models lead to excellent agreement with the NFW-profile in the radius range $0.5 r_s \leq r \leq 10 r_s$. \begin{figure}[h] \centerline{\psfig{figure=fig1.eps,height=8cm}} \caption{The density distributions of dark halos in a cold dark matter simulation of KKBP (thin solid lines) are compared with the Burkert profile (thick solid line) that provides a good fit to the observed rotation curves of dark matter-dominated dwarf galaxies. The thick dashed line shows the NFW profile which predicts too much dark matter mass inside $r_b$. The dot-dashed curve shows an isothermal profile with a finite density core which fails in the outer regions where it decreases as $r^{-2}$. } \end{figure} In a different series of very high-resolution models, using a new adaptive refinement tree N-body code, Kravtsov et al. (1998, KKBP) found significant deviations from the NFW profile or an even steeper inner power-law density distribution for $r \leq 0.5 r_s$. In this region their dark matter profiles show a substantial scatter around an average profile that is characterized by a shallow central cusp with $\rho \sim r^{-0.3}$. Although the scatter is large, this result is in clear contradiction to the simulations of Moore et al. (1998) with equally high central resolution. Figure 1 (adopted from Primack et al. 1998) compares the NFW-profile (dashed line) with the profiles of dark matter halos of KKBP (thin solid lines). \section{The dark matter halo of DDO 154} DDO 154 (Carignan \& Freeman 1988) is one of the most gas-rich galaxies known with a total H I mass of $2.5 \times 10^8 M_{\odot}$ and an inner stellar component of only $5 \times 10^7 M_{\odot}$. Recently, Carignan \& Purton (1998) measured the rotation curve of its extended H I disk all the way out to 21 optical disk scale lengths. As the rotation curve, even in the innermost regions, is almost completely dominated by dark matter, this galaxy provides an ideal laboratory for testing the universal density profile predictions of cosmological models. \begin{figure}[h] \centerline{\psfig{figure=fig2.eps,height=8cm}} \caption{The dark matter rotation curve of DDO 154 with error bars. The dotted line is a fit to the inner parts, adopting a NFW profile. The dashed curve shows a fit to the outer regions. The solid line shows the fit which is achieved with a two-component model where the lower dashed line shows the contribution of the dark baryonic component and the lower dot-dashed curve is the standard dark matter halo contribution adopting a NFW profile.} \end{figure} Figure 2 shows the dark matter rotation curve of DDO 154 and compares it with the NFW profile. Note that the maximum rotational velocity $v_{max} \approx$ 47 km/s is reached at a radius of $r_{max} \approx$ 4.5 kpc, beyond which the rotation decreases again. Fitting the inner regions (dotted line) has been known to pose a problem (Flores \& Primack 1994, Moore 1994, Burkert 1995). However an even larger problem exists in the outermost regions where far too much dark matter would be expected. The dashed line in figure 2 shows a fit to the outer regions. In this case, the dark matter excess in the inner regions is unacceptably large. We conclude that the well-studied dark matter rotation curve of DDO 154 is far from agreement with NFW profiles. We can also compare the observed location $r_{max}$ and the value $v_{max}$ of the observed maximum rotational velocity with predictions of a SCDM model. Adopting a NFW profile, the virial radius is determined by $R_{200}$= 0.5 c r$_{max}$. The virial mass is then given by the relation \begin{equation} 1.63 \times 10^{-2} \left(\frac{M_{200}}{M_{\odot}} \right)^{1/3} h^{-2/3} = \left(\frac{R_{200}}{\rm kpc} \right) = 0.5 c \left(\frac{r_{max}}{\rm kpc} \right) . \end{equation} Adopting the SCDM model with h=0.5 and inserting equation (4) for $c$ one obtains \begin{equation} M_{200}^{SCDM} = 9 \times 10^8 \left(\frac{r_{max}}{\rm kpc} \right)^{2.11} M_{\odot} \end{equation} For DDO 154 with $r_{max}$ = 4.5 kpc we find $M_{200} = 2.1 \times 10^{10} M_{\odot}$, $R_{200}$ = 72 kpc and $c=31.9.$ For these halo values, the predicted maximum rotational velocity would be \begin{equation} v_{max} = 0.465 \left( \frac{c}{ln(1+c)-c/(1+c)} \right)^{1/2} h \left(\frac{R_{200}}{\rm kpc} \right) = 60 \ {\rm km/s} \end{equation} \noindent which is a factor 1.3 larger than observed. Adopting instead the CDM$\Lambda$ model with h=0.75, a similar calculation leads to \begin{equation} M_{200}^{CDM \Lambda} = 3 \times 10^8 \left(\frac{r_{max}}{\rm kpc} \right)^{2.3} M_{\odot} \end{equation} \noindent and therefore to $M_{200} = 9.5 \times 10^9 M_{\odot}$, $R_{200}$=42 kpc, c=18.7 and $v_{max}$=44.4 km/s which is in excellent in agreement with the observations (47 km/s), especially if one notes that equation 4 has been verified only for viral masses of $M_{200} > 10^{11} M_{\odot}$. In summary, the radius and mass scale of DDO 154 as determined from the value and location of its maximum rotational velocity is in perfect agreement with the predictions of the currently most favoured cosmological models (CDM$\Lambda$). The inferred dark matter density distribution is however quite different. \section {The universality of observed dark matter mass profiles.} DDO 154 is not a peculiar case. Burkert (1995) showed that the dark matter rotation curves of four dwarf galaxies studied by Moore (1994) have the same shape which can be well described by the density distribution \begin{equation} \rho_B(r) = \frac{\rho_b}{(1+r/r_b)[1+(r/r_b)^2]}. \end{equation} KKBP extended this sample to ten dark matter-dominated dwarf irregular galaxies and seven dark matter-dominated low surface brightness galaxies. As shown in figure 3 {\it all} have the same shape, corresponding to a density distribution given by equation (9) and in contradiction to equation (1). \begin{figure}[h] \centerline{\psfig{figure=fig3.eps,height=9cm}} \caption{This figure, adopted from Primack et al. (1998) shows the dark matter rotation curves of (a) ten dwarf irregular and (b) seven low surface brightness galaxies. The solid line shows the profile proposed by KKBP which is nearly identical to $\rho_B$ (equation 9).} \end{figure} Equation 9 predicts a flat dark matter core, the origin of which is difficult to understand in the context of hierarchical merging (Syer \& White 1998) as lower-mass dark halos in general have high densities and therefore spiral into the center of the more diffuse merger remnant, generating high-density cusps. There is a fundamental difference in the kinematical properties of dark matter halos described by NFW profiles and Burkert profiles (equation 9). Assuming isotropy and spherical symmetry in the inner regions, the Jeans equation predicts a velocity dispersion profile $\sigma (r)$ of NFW halos that decreases towards the center as $\sigma \sim r^{1/2}$ (see also the simulations of Fukushige \& Makino 1997) whereas Burkert halos have isothermal cores with constant velocity dispersion. Again, non-isothermal, kinematically cold dark matter cores might be expected in hierarchical merging scenarios as the denser clumps that sink towards the center of merger remnants have on average smaller virial dispersions. \section{On the origin of dark matter cores} It is rather unsatisfying that numerical studies of dark matter halos using different techniques lead to different results. KKBP find dark matter halo profiles that, on average, can be well fitted by a Burkert profile (Fig. 1) and therefore also provide a good fit to observed rotation curves. The NFW-profiles, or even steeper inner density gradients (Moore et al. 1998) seem to be in clear conflict with observations (Fig. 2). Numerical resolution cannot be the answer as the N-body simulations of Moore et al. (1998) have enough resolution to determine the dark matter density distribution within 0.5 $r_s$, where any flattening should have been found. Instead, the authors find profiles that are even steeper than $r^{-1}$. One should note that the results of KKBP have not yet been reproduced by other groups, whereas dark profiles with central $r^{-1}$ profiles or even steeper cusps have been found independently in many studies using different numerical techniques. On the other hand, the high-resolution studies of KKBP sample galactic halos, whereas most of the other studies simulate halos of galactic clusters. Indeed, there exists observational evidence that the dark matter distribution in clusters of galaxies is well described by NFW profiles (Carlberg et al. 1997, McLaughlin 1999). One conclusion therefore could be that dark matter halos are not self-similar but that their core structure does depend on their virial mass. Because one is sampling different parts of the primordial CDM density fluctuation power spectrum, from which the initial conditions are derived, it is possible that the initial conditions could influence the final result. For example, low mass dark halos have initial fluctuations which result in virialized velocity dispersions that are nearly independent of mean density, and so a hierarchy of substructure would be expected to be nearly isothermal. For more massive halos typical of normal galaxies, the velocity dispersion varies as $\rho^{\frac{n-1}{3(n+3)}}$ where $n$ is the effective power-law power spectrum index $\delta \rho / \rho \propto M^{-\frac{n+3}{6}}$ and $n \approx -2 $ on galaxy scales but $n \approx -3$ for dwarfs. Low-mass dark halos could then indeed have isothermal, constant density cores whereas high-mass dark halos should contain non-isothermal, cold power-law cores. However, even in this case, one still has the problem that the simulations lead to a much greater dispersion of the inner radial profiles than expected from observed rotation curves. Additional effects might therefore be important that have not been taken into account in dissipationless cosmological simulations. \subsection{Secular dynamical processes} Cold dark matter cores with steep density cusps are very fragile and can easily be affected by mass infall and outflow. This has been shown, for example, by Tissera \& Dom\'{i}nguez-Tenreiro (1998) who included gas infall in their cosmological models and found even steeper power-laws in the central regions than predicted by purely dissipationless merging due to the adiabatic contraction of the dark component. In order to generate flat cores through secular processes, Navarro et al. (1996b) proposed a scenario where after a slow growth of a dense gaseous disk the gas is suddenly ejected. The subsequent expansion and violent relaxation phase of the inner dark matter regions leads to a flattening of the core. This model has been improved by Gelato \& Sommer-Larsen (1999) who applied it to DDO 154 and found that it is not easy to satisfactorily explain the observed rotation curve even for extreme mass loss rates. In fact, it is unlikely that DDO 154 lost any gas, given its large gas fraction. In addition, secular dynamical processes due to mass outflow would predict inner dark matter profiles that depend sensitively on the detailed physics of mass ejection and therefore should again show a wide range of density distributions, and these are not observed. \subsection{A second dark and probably baryonic component} The rotation curves of the galaxies shown in figure 3 clearly cannot be understood by including only the visible component. It may well be that some non-negligible and as yet undetected fraction of the total baryonic mass contributes to their dark component, in addition to the non-baryonic standard cold dark component that is considered in cosmological models. In fact, there is ample room for such a dark baryonic component. Primordial nucleosynthesis requires a baryonic density of $\Omega_b h^2 \approx 0.015 \pm 0.008$ (Kurki-Suonio et al. 1997, Copi et al. 1995), whereas the observed baryonic density for stellar and gaseous disks lies in the range of $\Omega_d \approx 0.004 \pm 0.002$ (Persic \& Salucci 1992). Moreover modelling of Lyman alpha clouds at $z\sim 2-4$ suggests that all of the baryons expected from primordial nucleosynthesis were present in diffuse form, to within the uncertainties, which may amount to perhaps a factor of 2 (Weinberg et al. 1997). Hence dark baryons are required at low redshift. These may be in the form of hot gas that must be mostly outside of systems such as the Local Group and rich galaxy clusters (Cen and Ostriker 1999). But an equally plausible possibility is that the dark baryons are responsible for a significant fraction of the mass in galaxy halos, as is motivated by arguments involving disk rotation curves and halo morphologies (cf Gerhard and Silk 1996; Pfenniger et al. 1994). This second dark baryonic component could be diffuse $H_2$ within the disks or some spheroidal distribution of massive compact baryonic objects (MACHOs), comparable to those that have been detected via gravitational microlensing events towards the Large Magellanic Cloud (LMC). It is difficult to reconcile the inferred typical MACHO lens mass of $\sim 0.5 \pm 0.3$ M$_{\odot}$, as derived from the first 2.3 years of data for 8.5 million stars in the LMC (Alcock et al. 1996), with ordinary hydrogen-burning stars or old white dwarfs (Bahcall et al. 1994, Hu et al. 1994, Carr 1994, Charlot \& Silk 1996). Brown dwarfs, substellar objects below the hydrogen-burning limit of 0.08 $M_{\odot}$ would be ideal candidates. Indeed, halo models can be constructed, e.g. by assuming a declining outer rotation curve, for which the most likely MACHO mass is 0.1 M$_{\odot}$ or less ( Honma \& Kan-ya 1999) with a MACHO contribution to the total dark mass of almost 100\%. Freese et al. (1999) have however shown by deep star counts using HST that faint stars and massive brown dwarfs contribute no more than 1\% of the expected total dark matter mass density of the Galaxy, ruling out such a low-mass population. A simple explanation of the MACHO mass problem has been presented by Zaritsky \& Lin (1997) and Zhao (1998), who argued that the MACHOs reside in a previously undetected tidal stream, located somewhere in front of the LMC. In this case the microlensing events would represent stellar objects in the outer regions of the LMC and would not be associated with a dominant dark matter component of the Milky Way. This solution is supported by the fact that all lensing events toward the LMC and SMC with known distances (e.g. the binary lensing event 98 LMC-9, or 98-SMC-1) appear to be a result of self-lensing within the Magellanic Clouds. However the statistics are abysmal: only two SMC events have been reported, and there are approximately 20 LMC events in all, of which two have known distances. Moreover one can only measure distances for binary events, and these are very likely due to star-star lensing. Finally, the SMC is known to be extended along the line of sight, thereby enhancing the probability of self-lensing. Thus, while evidence for a possible dark baryonic component in the outer regions of galaxies is small, such a component could still be the solution to the dark matter core problem. Burkert \& Silk (1997) showed that the observed rotation curve of DDO 154 could be reconciled with standard cosmological theories if, in addition to a standard dark matter halo with an NFW-profile (corrected for adiabatic contraction), a separate and centrally condensed dark baryonic component is introduced. The solid line in figure 2 shows the fit achieved with the 2-component dark model of Burkert and Silk adopting dark matter halo parameters which are in agreement with CDM$\Lambda$ models and using a spherically symmetric dark baryonic component with a physically reasonable density distribution that decreases monotonically with increasing radius. The required mass of the dark baryonic spheroid is $1.5 \times 10^{9} M_{\odot}$ which is 4-5 times the mass of the visible gaseous galactic disk and about 25\% the total mass of the non-baryonic dark component. The apparent universality of rotation curves would then suggest that the relative mix of the two dark components should in turn be universal. The origin of such a dark baryonic component which must have formed during an early dissipative condensation phase of baryons relative to the nondissipative, collisionless dark matter component is not understood. However this is not to say that it could not have occurred. Our understanding of early star formation and the primordial initial mass function is sufficiently primitive that this remains very much an open possibility. \section*{References}
\section{Pr\'esentation des r\'esultats} \label{intro} Soit $X$ un espace topologique. Un $X$-{\it groupo\"\i de \ } est un espace topologique $C$ muni de : a) deux applications continues $\alpha,\beta : C\to X$ ({\it source} et {\it but}). Pour $x,y\in X$, on note $C_x:=\alpha^{-1}(\{x\})$, $C^y:=\beta^{-1}(\{y\})$ et $C_x^y:= C_x\cap C^y$. Le groupo\"\i de \ $C$ est dit {\it transitif} si $C_x^y\not=\emptyset$ pour tout $x,y\in X$. b) une {\it composition} partiellement d\'efinie $ C\times_X C :=\{(c_1,c_2)\in C\mid \alpha(c_1)=\beta (c_2)\} \to \quad ,\quad (c_1,c_2) \mapsto c_1c_2 \kern 3 truecm $$ qui est continue, $C\times_X C$ \'etant un sous-espace de $C\times C$ avec la topologie produit. On demande que $\alpha(c_1c_2)=\alpha(c_2)$, $\beta(c_1c_2)=\beta(c_1)$ et que la composition soit associative. c) une application continue $i : X\to C$, associant \`a $x$ l'{\it unit\'e} $i_x\in C^x_x$ qui est \'el\'ement neutre \`a gauche et \`a droite pour la composition. d) une anti-involution continue $c\mapsto c^{-1}$ de $C$, envoyant $C^y_x$ sur $C^x_y$, telle que $cc^{-1}=i_{\alpha(c)}$ et $c^{-1}c=i_{\beta(c)}$. On suppose que $X$ est muni d'un point base $*\in X$. En vertu de d), l'espace $C^*_*$ est un groupe topologique que l'on note $\Omega_C$. \vskip .2 truecm On peut voir $C$ comme l'espace des morphismes (tous inversibles) d'une petite cat\'egorie topologique dont l'espace des objets est $X$. C'est le point de vue de \cite{Ma}. Si l'on pr\'ef\`ere les groupo\"\i des \ "sans objets" \cite[II.5]{Co}, on pr\'esentera $C$ comme un groupo\"\i de \ topologique dont l'espace des unit\'es est identifi\'e \`a $X$. Soit $G$ un groupe topologique et soit $\xi:= (E\fl{p}{} X)$ un {\it $G$-espace principal au dessus de $X$}. On entend par l\`a que $p$ est continue et que l'on s'est donn\'e une action continue libre $E\times G\to E$ telle que $p$ induise une bijection continue du quotient $E/G$ sur $X$. Par exemple, un $G$-fibr\'e principal sur $X$ est un $G$-espace au dessus de $X$ qui est localement trivial. On suppose que l'espace $E$ est point\'e par $\tilde * \in p^{-1}(*)$. Une {\it repr\'esentation} d'un $X$-groupo\"\i de \ $C$ sur un $G$-espace $\xi$ est une application continue $w: C\times_X E\to E$ (o\`u $C$ est vu au dessus de $X$ via $\alpha$) telle que, pour tout $c,d\in C$, $z\in E$ et $g\in G$, on ait \begin{enumerate} \item $p(w(c,z)) = \beta(c)$. \item $w(c\, d,z)=w(c,w(d,z))$. \item $w(c^{-1},(w(c,z))=z$. \item $w(c,z\cdot g) = w(c,z)\cdot g$. \end{enumerate} Par exemple, si $X=*$, alors $C$ est un groupe topologique et $w$ correspond \`a une repr\'esentation (= homomorphisme continu) de $C$ dans $G$. Plus g\'en\'eralement, si $w$ est une repr\'esentation de $C$ sur $\xi$, l'\'equation $w(c,\tilde *) = \tilde * \cdot h(c)$ d\'efinit un homomorphisme continu $h^w:\Omega_C\to G$ appel\'e l'{\it holonomie} de $w$. Les r\'esultats de cet article concernent l'existence et la classification de repr\'esentations d'un $X$-groupo\"\i de \ $C$ sur un $G$-fibr\'e principal $\xi$ donn\'es. Nos hypoth\`eses principales seront que $C$ est un $X$-groupo\"\i de \ {\it localement trivial} et $G$ un groupe \hbox{\sl SIN}\ (d\'efinitions ci-dessous). Soit $C$ un $X$-groupo\"\i de. Une {\it $C$-contraction} est une application continue $\rho : U_\rho\to C^*$, o\`u $U_\rho$ est un ouvert de $X$, telle que $\alpha(\rho(x))=x$. Soit ${\rm Cont\,}_C(X)$ l'ensemble des $C$-contractions. Le $X$-groupo\"\i de $C$ est dit {\it localement trivial} si $\{U_\rho\mid \rho\in{\rm Cont\,}_C(X)\}$ est un recouvrement ouvert de $X$ (\cite[p. 32]{Ma}; cette d\'efinition co\"\i ncide avec la terminologie de \cite{Eh} si $C$ est transitif). Si le recouvrement $\{U_\rho\}$ est de plus num\'erisable\footnote{Nous utilisons ``num\'erisable" pour \'equivalent fran\c cais du n\'eologisme anglais ``numerable" introduit par Dold [Do]. De m\^eme, un fibr\'e sera dit {\it num\'erisable} s'il admet un recouvrement num\'erisable d'ouverts trivialisants.}, c'est-\`a-dire s'il existe une partition de l'unit\'e $\{\mu_\rho : X\to [0,1]\}$ qui lui est subordonn\'ee, on dira que $C$ est un {\it $X$-groupo\"\i de \ localement trivial num\'erisable}. Plusieurs exemples naturels de tels groupo\"\i des sont pr\'esent\'es au paragraphe \ref{P:exp}. Reprenant les id\'ees de Ch. Ehresmann \cite{Eh}, nous d\'emontrerons au \S\ \ref{P:pthma} l'existence d'une {\it repr\'esentation universelle de $C$} et d'un principe de reconstruction~: \begin{Theorem}[Th\'eor\`eme A] \label{couniv} Soit $C$ un $X$-groupo\"\i de localement trivial num\'erisable. Alors~: a) l'application $\beta: C_*\to X$ est un $\Omega_C$-fibr\'e principal num\'erisable (que nous appellerons $\xi_C$). La formule $\tilde w(c,u):=c\, u$ d\'efinit une repr\'esentation de $C$ sur $\xi_C$. b) soit $\xi:= (E\fl{p}{} X)$ un $G$-espace principal muni d'une repr\'esentation $w$ de $C$. Alors $\xi$ est le fibr\'e associ\'e \`a $\xi_C$ $$E= C_* \times _{\Omega_C} G = C_* \times _{h^w} G \, ,$$ o\`u $\Omega_C$ agit \`a gauche sur $G$ via l'holonomie $h^w$ de $w$. En particulier, $\xi$ est un $G$-fibr\'e principal num\'erisable. De plus, la repr\'esentation $w$ est obtenue de $\tilde w$ par la formule $w(c,[u,g])=[\tilde w(c,u),g]=[cu,g]$. \end{Theorem} Rappelons qu'un $G$-fibr\'e principal num\'erisable $\xi$ sur $X$ est induit du fibr\'e universel $EG\to BG$ sur le classifiant de Milnor $BG$ par une application continue (unique \`a homotopie pr\`es) $\nu_\xi :X\to BG$. En particulier, d'apr\`es le th\'eor\`eme A, un $X$-groupo\"\i de localement trivial num\'erisable $C$ donnera une application $\nu_C:=\nu_{\xi_C}:X\to B\Omega_C$ induisant $\xi_C$. Le th\'eor\`eme d'existence d'une repr\'esentation de $C$ sur $\xi$, qui d\'ecoule facilement du th\'eor\`eme A, est le suivant~: \begin{Theorem}[Th\'eor\`eme d'existence] \label{counivex} Soit $C$ un $X$-groupo\"\i de localement trivial num\'erisable et $\xi$ un $G$-fibr\'e principal. Alors, $\xi$ admet une repr\'esentation de $C$ si et seulement si il est num\'erisable et s'il existe un homomorphisme continu $\phi : \Omega_C\to G$ tel que $B\phi\kern .7pt {\scriptstyle \circ} \kern 1pt \nu_C$ soit homotope \`a $\nu_\xi$. \end{Theorem} Ayant r\'esolu la question de l'existence, int\'eressons-nous \`a l'ensemble $\cala$ des repr\'esentations de $C$ sur $\xi$. Il est muni d'une action du {\it groupe de jauge} ${\cal G}$ de $\xi$. Les \'el\'ements de ${\cal G}$, les {\it transformations de jauge}, sont les automorphismes $G$-\'equi\-va\-riants de $E$ au dessus de ${\rm id}_X$. Pour $w\in\cala$ et $\chi\in{\cal G}$, on d\'efinit $w^\chi$ par $$w^\chi(c,z):=\chi^{-1}(w(c,\chi(z)).$$ ce qui donne une action \`a droite de ${\cal G}$ sur $\cala$. Le sous-groupe invariant ${\cal G}_1$ de ${\cal G}$ form\'e des transformations de jauge $\tilde *$ joue un r\^ole important car il agit librement sur $\cala$ (voir \ref{L:g1lib}). Les ensembles $\cala$ et ${\cal G}$ sont munis de la topologie compact-ouvert (CO-topologie). Nous d\'emontrerons, au \S\ \ref{unif} que ${\cal G}$ est un groupe topologique et que l'action $\cala\times{\cal G}\to\cala$ est continue, ceci sous l'hypoth\`ese que $G$ est \hbox{\sl SIN} . Rappelons qu'un groupe topologique est \hbox{\sl SIN} , si son \'el\'ement neutre admet un syst\`eme fondamental de voisinages qui sont $G$-invariants (par conjugaison; \hbox{\sl SIN}\ = small invariant neighbourhood). Par exemple, les groupes compacts, ab\'elien ou discrets sont \hbox{\sl SIN}\ (voir \cite{Pa}, pour la litt\'erature classique sur ces groupes). Nous montrerons que si $G$ est \hbox{\sl SIN}, la CO-topologie sur $\cala$ et ${\cal G}$ provient de structures uniformes et que toutes les op\'erations sont uniform\'ement continue (voir \S\ \ref{unif}). Pour \'etudier les quotients $\cala/{\cal G}_1$ et $\cala/{\cal G}$, introduisons l'espace ${\cal R} (\Omega_C,G)$ des repr\'esentations (i.e. homomorphismes continus) de $\Omega_C$ dans $G$, muni de la CO-topologie. Soit ${\cal R} (\Omega_C,G)_\xi$ le sous-espace des $\phi\in{\cal R} (\Omega_C,G)$ tels que l'application compos\'ee $n(\phi):=B\phi\kern .7pt {\scriptstyle \circ} \kern 1pt \nu_C$ soit homotope \`a $\nu_\xi$ (${\cal R} (\Omega_C,G)_\xi$ est une union de componates connexes par arc de ${\cal R} (\Omega_C,G)$). Au niveau des espaces classifiants, on d\'esigne par ${\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi$ l'espace des applications continues point\'ees de $f:X\to BG$ qui induisent $\xi$, muni de la CO-topologie. Un espace topologique $Z$ est dit {\it semi-localement contractile} si tout $z\in Z$ admet un voisinage $U_z$ dont l'inclusion $U_z\hookrightarrow Z$ est homotope \`a une application constante. Cette condition ne d\'epend que du type d'homotopie de $Z$. \begin{Theorem}[Th\'eor\`eme B] \label{gau} Soit $C$ un $X$-groupo\"\i de localement trivial num\'erisable et s\'epar\'e. Soit $\xi$ un $G$-fibr\'e principal num\'erisable sur $X$. On suppose que $G$ est \hbox{\sl SIN} , que $X$ est localement compact et que ${\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi$ est semi-localement contractile. Soit $h:\cala\to{\cal R} (\Omega_C,G)_\xi$ l'application qui \`a une repr\'esentation de $C$ $w$ associe son holonomie $h^w$. Alors $h$ est est un ${\cal G}_1$-fibr\'e principal. \end{Theorem} Nous ne savons pas si l'hypoth\`ese ``${\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi$ semi-localement contractile" est toujours v\'erifi\'ee. Observons que cette condition ne d\'epend que des types d'homotopie de $X$ et de $G$. Elle est vraie si, par exemple, $X$ est compact et $G$ est un groupe de Lie compact. En effet, $BG$ a alors le type d'homotopie d'un CW-complexe d\'enombrable, limite inductive quotients de vari\'et\'e de Stiefel \cite[\S\ 19.6]{St}. L'espace ${\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi$ est ainsi semi-localement contractile par \cite[lemmes 2 et 3 p. 277]{Mi1}. D'autre part, tout recouvrement d'un espace compact est num\'erisable, ce qui prouve le~: \begin{Corollaire}\label{gaucor} Soit $C$ un $X$-groupo\"\i de localement trivial et s\'epar\'e, avec $X$ compact. Soit $\xi$ un $G$-fibr\'e principal sur $X$ avec $G$ un groupe de Lie compact. Alors $h:\cala\to{\cal R} (\Omega_C,G)_\xi$ est est un ${\cal G}_1$-fibr\'e principal. \end{Corollaire} Le th\'eor\`eme B montre que le quotient $\cala/{\cal G}_1$ est hom\'eomorphe \`a ${\cal R} (\Omega_C,G)_\xi$. Pour d\'ecrire $\cala/{\cal G}$, on consid\`ere l'action \`a droite de $G$ sur ${\cal R} (\Omega_C,G)$ par conjugaison. Si $G$ est connexe par arc, cette action pr\'eserve ${\cal R} (\Omega_C,G)_\xi$. \begin{Theorem}[Th\'eor\`eme C] \label{gauser} Avec les hypoth\`eses du th\'eor\`eme B, l'appli\-ca\-tion $h$ induit un hom\'eomorphisme $\cala /{\cal G}_1\hfl{\cong}{} {\cal R} (\Omega_C,G)_\xi$. Si, de plus, $G$ est connexe par arc, elle induit un hom\'eomorphisme $\cala /{\cal G}\hfl{\cong}{} {\cal R} (\Omega_C,G)_\xi/G.$ \end{Theorem} Les th\'eor\`emes B et C pr\'esentent une analogie avec des r\'esultats de la th\'eorie de jauge qu'il serait int\'eressant d'\'etudier plus \`a fond. On sait que, si $\xi$ est un $G$-fibr\'e diff\'erentiable sur une vari\'et\'e compacte $X$ ($G$ groupe de Lie compact connexe), alors l'espace des connexions sur $\xi$ modulo ${\cal G}_1$ a le type d'homotopie faible de ${\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi$ \cite[Prop. 5.1.4]{DK}. On verra au \S\ \ref{clipamconn} qu'une connexion donne une repr\'esentation d'un groupo\"\i de \ ${\bf D}(X)$ construit \`a l'aide des chemins dans $X$ lisses par morceau. Pour l'instant, observons que les espaces $\calr(\Omega_C,G)_\xi$ et ${\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi$ sont reli\'es par une application continue $n:\calr(\Omega_C,G)_\xi\to{\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi$ (d\'etails en \ref{applicn}) dont on peut d\'ecrire la fibre homotopique lorsque $X$ est ``bien point\'e"~: \begin{Theorem}[Th\'eor\`eme D] \label{gauserthd} Supposons que l'on ait les hypoth\`eses du th\'eor\`eme B et que l'inclusion $\{*\}\subset X$ soit une cofibration. Alors, la fibre homotopique de $n : {\cal R} (\Omega_C,G)_\xi \to {\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi$ a le type d'homotopie faible de $\cala$. \end{Theorem} Le \S\ \ref{P:pthma} est consacr\'e \`a la preuve du th\'eor\`eme A et du th\'eor\`eme d'existence. Les \S\ \ref{unif} et \ref{P:gfibuniv} pr\'eparent aux preuves des th\'eor\`emes B,C et D qui sont donn\'ees dans le \S\ \ref{P:pthmbc}. Enfin, le \S\ \ref{P:exp} pr\'esente quelques exemples et applications. \paragraph{Remerciements : } Ce travail a b\'en\'efici\'e du support du Fonds National Suisse de la Recherche Scientifique. L'auteur tient \'egalement \`a remercier E. Dror-Farjoun, P. de la Harpe et R. Vogt pour d'utiles discussions. \section{Preuve des th\'eor\`emes A et d'existence} \label{P:pthma} \begin{Lemme} \label{rectrican} Soit $C$ un $X$-groupo\"\i de localement trivial num\'erisable. Soit $\xi:(E\hfl{p}{}X)$ un $G$-espace principal admettant une repr\'esentation $w$ de $C$. Alors, $\xi$ est un $G$-fibr\'e principal avec recouvrement trivialisant (num\'erisable) $\{U_\rho\mid\rho\in {\rm Cont\,}_C(X)\}$. Plus pr\'ecis\'ement, $\xi$ restreint \`a $U_\rho$ admet la trivialisation $\psi^w_\rho: p^{-1}(U_\rho)\to U_\rho\times G$ donn\'ee par \begin{equation}\label{deftriv1} \psi^w_\rho(z) := (p(z), \lambda_w(z)) \quad \hbox{ o\`u } \quad w(\rho(p(z)),z) = \tilde * \lambda_w(z).\end{equation} \end{Lemme} \noindent {\sc Preuve: \ } Il est banal que $\psi^w_\rho$ est $G$-equivariante et on v\'erifie que inverse de $\psi^w_\rho$ est $(\psi^w_\rho)^{-1}(x,g):=w(\rho(x)^{-1}, \tilde * )\cdot g$ \cqfd. \paragraph{Preuve du th\'eor\`eme A} Observons que l'application $\beta:C_*\to X$ est surjective si $C$ est localement trivial. En effet, si $x\in X$, il existe $\rho\in {\rm Cont\,}_C(X)$ telle que $x\in U_\rho$ et alors $x=\beta(\rho(x)^{-1})$. L'action de $\Omega_C$ sur $C_*$ est d\'efinie par $u\cdot c := u\, c$. Si $\beta(u)=\beta(\bar u)$, alors $\bar u = \bar u\cdot c$ avec $c:= (\bar u^{-1}\, u)$. L'application $\beta$ induit donc une bijection continue $\bar \beta :C_*/\Omega_C\to X$. Les trivialistions construites ci-dessous font que $\beta$ est ouverte et donc $\bar\beta$ est un hom\'eomorphisme. Pour $\rho\in {\rm Cont\,}_C(X)$, on d\'efinit la trivialisation $\hat \psi_\rho : \beta^{-1}(U_\rho)\to U_\rho\times \Omega_C$ de la mani\`ere suivante: \begin{equation} \label{deftriv2} \hat\psi_\rho(u) = (\beta(u), \rho(\beta (u))\, u). \end{equation} Elle admet pour inverse \begin{equation} \hat\psi^{-1}_\rho(x,c) = \rho(x)^{-1}c. \end{equation} La $\Omega_C$-equivariance de l'hom\'eomorphisme $\hat\psi_\rho$ est banale. Ceci montre que $\beta : C_*\to X$ est un fibr\'e $\Omega_C$-principal avec recouvrement trivialisant num\'erisable $\{U_\rho\}$. Le fait que la formule $\tilde w (c,u)= c\, u$ d\'efinisse une repr\'esentation de $C$ sur ce fibr\'e provient des axi\^omes de groupo\"\i de . Passons au point b) du th\'eor\`eme A. Soit $w$ une repr\'esentation de $C$ sur un $G$-espace principal $\xi:= (E\fl{p}{} X)$. Le Lemme \ref{rectrican} montre que $\xi$ est un fibr\'e principal num\'erisable. L'espace $E$ \'etant point\'e par $\tilde * \in p^{-1}(*)$, on d\'efinit $\Phi : C_* \times _{\Omega_C} G \to E$ par \begin{equation} \Phi(u,g) = w(u,\tilde * ) \cdot g \end{equation} a) {\it $\Phi$ est bien d\'efinie : } Soit $c\in \Omega_C$. On a $$\begin{array}{lcl} \Phi(u\, c,g) & = & w(u\, c,\tilde * ) \cdot g = w(u,w(c,\tilde * )) \cdot g = w(u, \tilde *\cdot h(c))\cdot g = \\[2pt] & = & w(u, \tilde *)\cdot (h(c) g) = \Phi(u, h(c)g).\end{array}$$ \vskip .2 truecm b) {\it $\Phi$ est $G$-equivariante : } \'evident. \vskip .2 truecm c) {\it $\Phi$ est surjective : } Soit $z\in E$. Soit $\rho\in {\rm Cont\,}_C(X)$ telle que $p(z)\in U_\rho$. Posons $u:=\rho(p(z))^{-1}\in C_*^{p(z)}$. On a ainsi $p(\Phi(u,1))= \beta(u)=p(z)$. Il existe donc un unique $g\in G$ tel que $\Phi(u,g)=\Phi(u,1)\cdot g = z$. \vskip .2 truecm d) {\it $\Phi$ est injective : } Supposons que $\Phi(u,g)=\Phi(\bar u,\bar g)$. On a donc $\bar u = u \, c$ avec $c:=u^{-1}\bar u\in\Omega_C$ et $$\begin{array}{lcl} \Phi(\bar u ,\bar g) & = & w( u\, c,\tilde * ) \cdot \bar g = w(u,w(c,\tilde * )) \cdot \bar g = \\[2pt] & = & w(u, \tilde *)\cdot (h(c) \bar g) \end{array}$$ Comme d'autre part $$\Phi(\bar u ,\bar g)=\Phi(u ,g) = w(u, \tilde *)\cdot g,$$ il s'en suit que $g=h(c) \bar g$. Dans $C_*\times _{\Omega_C} G$, on aura ainsi les \'egalit\'es $$(\bar u,\bar g) = (u\, c,\bar g) = (u, h(c)\bar g) = (u,g).$$ \vskip .2 truecm e) {\it $\Phi$ est un hom\'eomorphisme : } Soit $\rho\in {\rm Cont\,}_C(X)\}$. Avec les trivialisations $\psi_\rho$ et $\hat \psi_\rho$ introduites en \eqref{deftriv1} et \eqref{deftriv2}, l'application $$\Phi\rho:=\psi_\rho \kern .7pt {\scriptstyle \circ} \kern 1pt \Phi \kern .7pt {\scriptstyle \circ} \kern 1pt \hat \psi_\rho^{-1} : U_\rho\times(\Omega_C\times_{\Omega_C}G)\to U_\rho\times G$$ s'\'ecrit $\Phi\rho(x,(c,g)) = (x,h(c)\, g)$. L'application $\Phi_\rho$ est donc un hom\'eo\-mor\-phsime pour tout $\rho$, ce qui implique que $\Phi$ est un hom\'eomorphisme. \paragraph{Preuve du th\'eor\`eme d'existence} Supposons que $\xi$ admette une repr\'e\-sen\-tation $w$ de $C$. Par le point b) du th\'eor\`eme A, $\xi$ est obtenu du fibr\'e $\xi_C$ par la construction de Borel avec l'holonomie $h^w$. Cela implique que $\nu_\xi$ est homotope \`a $Bh^w\kern .7pt {\scriptstyle \circ} \kern 1pt \nu_C$. R\'eciproquement, supposons qu'il existe un homomorphisme continu $\phi : \Omega_C\to G$ tel que $\nu_\xi$ est homotope \`a $B\phi\kern .7pt {\scriptstyle \circ} \kern 1pt \nu_C$. L'espace total du fibr\'e induit $\nu_xi^*(EG)$ est de la forme $C_*\times_{\Omega_C} G$ et admet donc la repr\'esentation $w(c,[u,g])=[cu,g]$. Le fibr\'e $\nu_xi^*(EG)$ \'etant isomorphe \`a $\xi$, cela donne une repr\'esentation de $C$ sur $\xi$. \cqfd \section Structures uniformes sur ${\cal G}$ et $\cala$} \label{unif} Dans ce paragraphe, $\xi: E\hfl{p}{}X$ d\'esigne un $G$-fibr\'e principal num\'erisable et ${\cal G}$ son groupe de jauge, muni de la CO-topologie. Le fait que ${\cal G}$ est un groupe topologique est non-trivial, car $E$ n'est m\^eme pas suppos\'e localement compact. Pour d\'emontrer ce fait, nous allons, lorsque $G$ est \hbox{\sl SIN} , munir ${\cal G}$ d'une structure uniforme ${\bf U}_{\cal G}$, dont on montrera, si $X$ est s\'epar\'e, qu'elle induit la CO-tolopogie (Proposition \ref{compou}). La m\^eme strat\'egie sera utilis\'ee pour l'action de ${\cal G}$ sur $\cala$. Pour d\'ecrire ${\bf U}_{\cal G}$, consid\'erons l'application continue $\gamma :E\times_X E\to G$ d\'efinie par l'\'equation \begin{equation}\label{defgamma} y=z\cdot \gamma(y,z) \quad, \quad (y,z)\in E\times_X E. \end{equation} Soit ${\cal V}_G$ l'ensemble des ouverts $V$ de $G$ contenant l'\'el\'ement neutre et tels que $V=V^{-1}$. Pour $V\in{\cal V}_G$ et $K$ un compact de $X$, on d\'efinit $${\cal O}^{\cal G}({K,V}) :=\{(\chi,\tilde\chi)\in{\cal G}\times{\cal G}\mid \gamma(\chi(z),\tilde\chi(z))\in V\hbox{ \small pour tout } z\in p^{-1}(K)\}.$$ Comme ${\cal O}^{\cal G}({K_1,V_1})\cap{\cal O}^{\cal G}({K_2,V_2})$ contient ${\cal O}^{\cal G}({K_1\cup K_2,V_1\cap V_2})$, la famille ${\cal O}^{\cal G}(K,V)$ est un syst\`eme fondamental d'entou\-rages de la diagonale dans ${\cal G}\times{\cal G}$, d\'eterminant, par d\'efinition, la structure uniforme ${\bf U}_{\cal G}$ sur ${\cal G}$. \begin{Proposition} \label{unicalg} Si $G$ est un groupe {\it SIN}, le groupe de jauge ${\cal G}$, muni de la structure uniforme ${\bf U}_{\cal G}$, est un groupe topologique. De plus, ${\cal G}$ est alors {\it SIN}. \end{Proposition} \noindent {\sc Preuve: \ } Nous allons tout d'abord d\'emontrer que la multiplication ${\cal G}\times{\cal G}\to{\cal G}$ et le passage \`a l'inverse sont uniform\'ement continus. La topologie induite par ${\bf U}_{\cal G}$ fera donc de ${\cal G}$ un groupe topologique dont on v\'erifiera directement qu'il est \hbox{\sl SIN} . Soient $\chi_1$, $\chi_2$, $\tilde \chi_1$, $\tilde \chi_2$ des \'el\'ements de ${\cal G}$. Par d\'efinition de l'application $\gamma :E\times_X E\to G$, on a, pour $z\in E$~: $$\tilde\chi_1\kern .7pt {\scriptstyle \circ} \kern 1pt\tilde\chi_2(z) = \chi_1\kern .7pt {\scriptstyle \circ} \kern 1pt\chi_2(z) \cdot \gamma(\tilde\chi_1\kern .7pt {\scriptstyle \circ} \kern 1pt\tilde\chi_2(z),\chi_1\kern .7pt {\scriptstyle \circ} \kern 1pt\chi_2(z)) .$$ Par $G$-equivariance des transformations de jauge, on a~: $$\begin{array}{lcl} \tilde\chi_1\kern .7pt {\scriptstyle \circ} \kern 1pt\tilde\chi_2(z) &=& \tilde\chi_1\big(\chi_2(z)\cdot\gamma(\tilde\chi_2(z),\chi_2(z))\big)= \tilde\chi_1(\chi_2(z))\cdot\gamma(\tilde\chi_2(z),\chi_2(z)) = \\[2pt] &=& \chi_1(\chi_2(z))\cdot \gamma(\tilde\chi_1(\chi_2(z)),\chi_1(\chi_2(z))) \cdot\gamma(\tilde\chi_2(z),\chi_2(z)) \end{array}.$$ On en d\'eduit que $$\gamma(\tilde\chi_1\kern .7pt {\scriptstyle \circ} \kern 1pt\tilde\chi_2(z),\chi_1\kern .7pt {\scriptstyle \circ} \kern 1pt\chi_2(z))= \gamma(\tilde\chi_1\kern .7pt {\scriptstyle \circ} \kern 1pt\chi_2(z),\chi_1\kern .7pt {\scriptstyle \circ} \kern 1pt\chi_2(z)) \,\, \gamma(\tilde\chi_2(z),\chi_2(z)).$$ Soit $V\in{\cal V}_G$. Comme $G$ est un groupe topologique, il existe $W\in{\cal V}_G$ tel que $W\cdot W \subset V$. La condition $(\tilde\chi_1\kern .7pt {\scriptstyle \circ} \kern 1pt\tilde\chi_2,\chi_1\kern .7pt {\scriptstyle \circ} \kern 1pt\chi_2)\in{\cal O}^{\cal G}(K,V)$ sera vraie si $(\tilde \chi_1,\chi_1)$ et $(\tilde \chi_2,\chi_2)$ sont dans ${\cal O}^{\cal G}(K,W)$. Ceci d\'emontre la continuit\'e uniforme de la composition ${\cal G}\times{\cal G}\to{\cal G}$. Pour le passage \`a l'inverse, soient $\chi,\tilde\chi\in{\cal G}$. Les \'equations $$\chi(z)=z\cdot \gamma(\chi(z),z) \quad , \quad \tilde\chi(z)=z\cdot \gamma(\tilde\chi(z),z)$$ donnent $$\tilde\chi(z)=\chi(z)\cdot \gamma(\chi(z),z)^{-1} \cdot \gamma(\tilde\chi(z),z)$$ d'o\`u \begin{equation}\label{E:inv1} \gamma(\tilde\chi(z),\chi(z))= \gamma(\chi(z),z)^{-1}\, \gamma(\tilde\chi(z),z). \end{equation} Observons que $$\chi_1(\chi_2(z))=\chi_1(z)\cdot\gamma(\chi_2(z),z)= z\cdot\gamma(\chi_1(z),z)\cdot\gamma(\chi_2(z),z)$$ et donc \begin{equation}\label{E:inv2} \gamma(\chi_1\kern .7pt {\scriptstyle \circ} \kern 1pt\chi_2(z),z)= \gamma(\chi_1(z),z)\,\gamma(\chi_2(z),z). \end{equation} On en d\'eduit que $\gamma(\chi^{-1}(z),z)=\gamma(\chi(z),z)^{-1}$. En changeant $\chi,\tilde\chi$ en $\chi^{-1},\tilde\chi^{-1}$ dans \eqref{E:inv1}, on obtient ainsi \begin{equation}\label{E:inv3} \begin{array}{lcl} \gamma(\tilde\chi^{-1}(z),\chi^{-1}(z)) &=& \gamma(\chi(z),z)\, \gamma(\tilde\chi(z),z)^{-1}=\\[2pt] &=& \gamma(\chi(z),z)\,\gamma(\tilde\chi(z),\chi(z))^{-1}\,\gamma(\chi(z),z)^{-1}. \end{array}\end{equation} Soit $V\in{\cal V}_G$ et $K$ un compact de $X$. Comme $G$ est \hbox{\sl SIN} , il existe un voisinage invariant $W$ de l'\'el\'ement neutre contenu dans $V$. En rempla\c cant au besoin $W$ par $W\cup W^{-1}$, on peut supposer que $W=W^{-1}$. Gr\^ace \`a l'\'equation \eqref{E:inv3}, si $(\tilde\chi,\chi)\in{\cal O}^{\cal G}(K,W)$, alors $(\tilde\chi^{-1},\chi^{-1})\in{\cal O}^{\cal G}(K,V)$, ce qui prouve la continuit\'e uniforme de $\chi\mapsto\chi^{-1}$. Pour voir que ${\cal G}$ est \hbox{\sl SIN} , on utilise que tout voisinage de ${\rm id}_E$ contient un voisinage du type ${\cal H}(K,W):=\{\chi\mid (\chi,{\rm id}_E)\in{\cal O}^{\cal G}(K,W)\}$ o\`u $K$ est un compact de $X$ et $W$ un voisinage invariant de l'\'el\'ement neutre dans $G$. Par la formule \eqref{E:inv2}, on a, pour tout $\tilde\chi\in{\cal G}$ et tout $z\in E$ $$\gamma(\tilde\chi^{-1}\kern .7pt {\scriptstyle \circ} \kern 1pt\chi\kern .7pt {\scriptstyle \circ} \kern 1pt\tilde\chi(z),z)= \gamma(\tilde\chi(z),z)^{-1}\,\gamma(\chi(z),z)\, \gamma(\tilde\chi(z),z).$$ On en d\'eduit imm\'ediatement que ${\cal H}(K,W)$ est ${\cal G}$-invariant. \cqfd D\'efinissons $\hbox{\rm res\,} : {\cal G}\to G$ par la formule \begin{equation}\label{E:defres} \chi(\tilde *) = \tilde *\cdot\hbox{\rm res\,} (\chi). \end{equation} L'\'equation \ref{E:inv2} implique que $\hbox{\rm res\,}$ est un homomorphisme. Le noyau de $\hbox{\rm res\,}$ est \'evidemment ${\cal G}_1$. Le groupe $G$ est muni de sa structure uniforme naturelle~: une base d'entourages est donn\'ee par ${\cal O}^V:=\{(\tilde g ,g)\mid \tilde g g^{-1}\in V\}$. On obtient la m\^eme structure uniforme avec la condition $\tilde g^{-1}g\in V$ lorsque $G$ est \hbox{\sl SIN} . \begin{Proposition} \label{numer} Si $G$ est \hbox{\sl SIN} , l'homomrophisme $\hbox{\rm res\,}$ est uniform\'ement continu. Si $G$ est connexe par arc, $\hbox{\rm res\,}$ est surjectif. \end{Proposition} \noindent {\sc Preuve: \ } Pour tout $V\in{\cal V}_G$, on a $\hbox{\rm res\,}({\cal O}^{\cal G}(\{*\},V)\subset V$, ce qui prouve la continuit\'e uniforme de $\hbox{\rm res\,}$ (ceci ne semble pas utiliser que $G$ est \hbox{\sl SIN}\ mais rappelons que cette condition est n\'ecessaire pour que $G$ soit un groupe topologique par la proposition \ref{unicalg}). La preuve de la surjectivit\'e de $\hbox{\rm res\,}$ utilise que $\xi$ est num\'erisable. Soit $\mu :X\to{\bf R}$ une application continue telle que $\mu(*)\not = 0$ et dont le support est contenu dans un ouvert trivialisant $U$. Soit $\psi : p^{-1}(U)\to U\times G$ une trivialisation. Posons $\psi(z)=(p(z),\delta(z))$. En divisant $\mu$ par $\mu(*)$, on peut supposer que $\mu(*)=1$. Soit $g\in G$. Comme $G$ est connexe par arc, il existe un chemin continu $g(t)$ avec $g(0)=e$ et $g(1)=g$. On d\'efinit alors $\chi\in{\cal G}$ par $$\chi(z):=\left\{\begin{array}{lllll} \psi^{-1}(p(z),g(\mu(p(z))\,\delta(z))\ & \hbox{si } p(z)\in U\\ z & \hbox{sinon. }\end{array}\right.$$ Il est clair que $\hbox{\rm res\,} (\chi) = g$. \cqfd Nous allons maintenant munir l'espace $\cala$ des repr\'esentations de $C$ sur $\xi$ de la structure uniforme ${\bf U}_\cala$ dont une base d'entourages est donn\'ee par $${\cal O}^{\cal R}(L,V) :=\{(w,\tilde w)\in\cala\times\cala\mid \gamma(w(c,z),\tilde w(c,z))\in V \ \forall (c,z)\in L\times_X E\}$$ o\`u $V\in {\cal V}_G$ et $L$ est un un compact de $C$. \begin{Proposition} \label{unicala} L'action $\cala\times{\cal G}\to\cala$ est uniform\'ement continue. \end{Proposition} \noindent {\sc Preuve: \ } Soient $\chi,\tilde\chi\in{\cal G}$ et $w,\tilde w\in\cala$. Pour $(c,z)\in C\times_X E$, on d\'emontre, comme dans la preuve de la proposition \ref{unicala} que $\gamma(\tilde w^{\tilde\chi}(c,z),w^{\chi}(c,z))$ est le produit $\gamma_1\,\gamma_2\,\gamma_3$ avec $$\begin{array}{lcl} \gamma_1 &=& \gamma\big(\tilde\chi^{-1} (w(c,\chi(z))),\chi^{-1} (w(c,\chi(z)))\big) \\[2pt] \gamma_2 &=& \gamma(\tilde w(c,\chi(z)),w(c,\chi(z))) \\[2pt] \gamma_3 &=& \gamma(\tilde \chi(z),\chi(z))). \end{array} $$ Soit $L$ un compact de $C$ et $V\in{\cal V}_G$. Si $W\in{\cal V}_G$ est $G$-invariant et satisfait $W\cdot W\cdot W\subset V$, cela prouve, en utilisant la formule \eqref{E:inv3}, que si $(\tilde\chi,\chi)\in{\cal O}^{\cal G}(\alpha(L),W)\cap{\cal O}^{\cal G}(\beta(L),W)$ et $(\tilde w,w)\in{\cal O}^{\cal R}(L,W)$, alors $(\tilde w^{\tilde\chi},w^{\chi})\in{\cal O}^{\cal R}(L,V)$. \cqfd Enfin, l'epace ${\cal R} (\Omega_C,G)$ peut \^etre muni d'une structure uniforme ${\bf U}_{{\cal R}}$ ayant pour base d'entourages $${\cal O}^{{\cal R}}(K,V) :=\{(h,\tilde h)\in{\cal R} (\Omega_C,G)\times{\cal R} (\Omega_C,G)\mid \tilde h(c)h(c^{-1})\in V ,\ \forall c\in K\}$$ o\`u $V\in {\cal V}_G$ et $K$ est un un compact de $\Omega_C$. \begin{Proposition} \label{unihom} Si $G$ est {\it SIN}, l'action de $G$ sur ${\cal R} (\Omega_C,G)$ par conjugaison est uniform\'ement continue. \end{Proposition} \noindent {\sc Preuve: \ } Soit $V\in{\cal V}_G$. Soit $W\in{\cal V}_G$ tel que $W$ soit $G$-invariant et $W\cdot W\cdot W\subset V$. Soient $\varphi,\tilde\varphi\in{\cal R}(\Omega_C,G)$ et $g,\tilde g\in G$. Soit $L$ un compact de $\Omega_C$ et $c\in L$. La formule $$(\tilde g^{-1}\tilde\varphi(c)\tilde g)(g^{-1}\varphi(c) g)^{-1} = \tilde g^{-1}\tilde\varphi (c)(\tilde g g^{-1}) \tilde\varphi^{-1} (c) (\tilde\varphi (c)\varphi^{-1}(c)) \tilde g\,(\tilde g^{-1} g)$$ montre que $\tilde g\,\tilde g^{-1}\in W$ et $(\tilde\varphi ,\varphi)\in{\cal O}^{\cal R} (L,W)$ impliquent que \\ $(\tilde g^{-1}\tilde\varphi\tilde g, g^{-1}\varphi g)\in {\cal O}^{\cal R} (L,V)$. \cqfd Nous terminons ce paragraphe en montrant que les structures uniformes consid\'er\'ees induisent la CO-topologie. Rappelons que, par d\'efinition, une sous-base de la CO-topologie sur l'espace fonctionnel ${\rm map\,}(X,Y)$ est form\'ee des ensembles $CO(K,U):=\{f\in {\rm map\,}(X,Y)\mid f(K)\subset U\}$, o\`u $K$ parcourt l'ensemble des compacts de $X$ et $U$ celui des ouverts de $Y$. \begin{Proposition} \label{compou} Si $G$ est {\it SIN} et $X$ est un espace s\'epar\'e, les topologies sur ${\cal G}$, $\cala$ et ${\cal R}(\Omega_C,G)$ induites par les structures uniformes ${\bf U}_{\cal G}$, ${\bf U}_\cala$ et ${\bf U}_{\cal R}$ co\"\i ncident avec la CO-topologie. \end{Proposition} Nous aurons besoin d'un analogue du lemme de Lebesgue (voir aussi \cite[II.4.3]{Bo})~: \begin{Lemme}\label{Lebesgue} Soit $f:K\to E$ une application continue d'un compact dans l'espace total de $\xi$. Soit $U$ un ouvert de $E$ avec $f(K)\subset U$. Alors, il existe $V\in{\cal V}_G$ tel que $f(K)\cdot V\subset U$. \end{Lemme} \noindent {\sc Preuve: \ } Pour tout $z\in K$, il existe un ouvert $S_z$ de $E$ et $V_z\in{\cal V}_G$ tels que $f(z)\in S_z\subset U$ et $S_z$ est de la forme $\sigma(p(S_z))\times (f(z)\cdot V_z \cdot V_z)$, o\`u $\sigma$ est une section locale de $\xi$ au voisinage de $p(f(z))$. Comme $K$ est compact, on a $f(K)=\bigcup_{z\in K_0}S_z$ pour un sous-ensemble fini $K_0$ de $K$. Soit $V:=\bigcap_{z\in K_0}V_z\in{\cal V}_G$. Alors, pour tout $z\in K$, on a $f(z)\cdot V\subset U$. En effet, il existe $y\in K_0$ tel que $f(z)\in f(y) \dot V_y$ d'o\`u $$f(z)\dot V\subset f(y) \cdot V_y\cdot V \subset f(y) \cdot V_y\cdot V_y\subset U. \cqfd $$ \vskip .3 truecm\noindent{\sc Preuve de \ref{compou} : } L'affirmation pour ${\cal R}(\Omega_C,G)$ est un fait classique pour les applications dans un espace uniforme \cite[X.3.4, th\'eor\`eme 2]{Bo}. \vskip .3 truecm\noindent{\it Preuve pour ${\cal G}$ : } Soit $\chi\in{\cal G}$, $K$ un compact de $E$ et $U$ un ouvert de $E$ tels que $\chi(K)\subset U$. Par le lemme \ref{Lebesgue}, il existe $V\in{\cal V}_G$ tel que $\chi(K)\cdot V\subset U$. On a $\chi\in{\cal O}^{\cal G} (p(K),V)(\chi)\subset CO(K,U)$ (observons que $p(K)$ est compact puisque $X$ est s\'epar\'e). R\'eciproquement, soit $\chi\in{\cal G}$, $L$ un compact de $X$ et $V\in{\cal V}_G$. Il faut trouver un ouvert $\Sigma$ pour la CO-topologie tel que $\chi\in\Sigma\subset{\cal O}^{\cal G} (L,V)(\chi)$, o\`u ${\cal O}^{\cal G} (L,V)(\chi):=\{\tilde\chi\in{\cal G}\mid (\chi ,\tilde\chi)\in {\cal O}^{\cal G} (L,V)\}$. Comme $G$ est \hbox{\sl SIN} , il existe $W\in{\cal V}_G$ qui est $G$-invariant avec $W\subset V$. Consid\'erons tout d'abord le cas o\`u $L$ est contenu dans un ouvert $Y$ de $X$ au dessus duquel $\xi$ admet une section $\sigma$. L'ensemble $\chi(\sigma(L))\cdot W$ est un ouvert de $p^{-1}(L)$. Il existe donc un ouvert $U$ de $E$ tel que $\chi(\sigma(L))\cdot W=U\cap p^{-1}(L)$. Soit $\tilde\chi\in CO(\sigma(L),U)$. Si $z\in p^{-1}(L)$, il existe $g\in G$ tel que $z=\sigma(p(z))\cdot g$. La $G$-\'equivariance de $\tilde\chi$ donne, pour tout $u\in E$, la formule \begin{equation}\label{E:Geqchi} \gamma(\tilde\chi(u\cdot g),\chi(u\cdot g))= g^{-1}\,\gamma(\tilde\chi(u),\chi(u))\, g . \end{equation} Comme $W$ est $G$-\'equivariant, la formule \eqref{E:Geqchi} appliqu\'ee \`a $u:=\sigma(p(z))$ entra\^\i ne que $\gamma(\tilde\chi(z),\chi(z))\in W$, ce qui prouve que $\chi\in CO(\sigma(L),U)\subset{\cal O}^{\cal G} (L,V)(\chi)$. Dans le cas g\'en\'eral on recouvre $L$ par un nombre fini d'ouverts trivialisants, $L\subset Y_1\cup\cdots Y_n$, au dessus desquels on choisit des sections $\sigma_i$ de $\xi$. On construit comme ci-dessus les $\chi(\sigma_i(L_i))\cdot W\subset U_i$ et on aura $$\chi\in\bigcap_{i=1}^nCO (\sigma_i(L),U_i) \subset{\cal O}^{\cal G} (L,W)(\chi)\subset{\cal O}^{\cal G} (L,V)(\chi). \cqfd $$ \vskip .3 truecm\noindent{\it Preuve pour $\cala$ : } Soit $w\in\cala$. Soit $L$ un compact de $C\times_X E$ et $U$ un ouvert de $E$ tel que $w(L)\subset U$. Par le lemme \ref{Lebesgue}, il existe $V\in{\cal V}_G$ tel que $w(L)\cdot V\subset U$ et l'on a $w\in {\cal O}^{\cal R} (L,V)(w)\subset CO(L,U)$. R\'eciproquement, soit $w\in\cala$, $K$ un compact de $C$ et $V\in{\cal V}_G$. Comme $G$ est \hbox{\sl SIN} , il existe $W\in{\cal V}_G$ qui est $G$-invariant avec $W\subset V$. Consid\'erons tout d'abord le cas o\`u $\alpha(K)$ est contenu dans un ouvert $Y$ de $X$ au dessus duquel $\xi$ admet une section $\sigma$. L'ensemble $w(\sigma(\alpha(K)))\cdot W$ est un ouvert de $p^{-1}(\beta(K))$. Il existe donc un ouvert $U$ de $E$ tel que $w(\sigma(\alpha(K)))\cdot W=U\cap p^{-1}(L)$. On d\'emontre, comme dans la preuve pour ${\cal G}$ ci-dessus, que $w\in CO(\sigma(\alpha(K)),U)\subset{\cal O}^{\cal G} (K,V)(w)$. Le cas g\'en\'eral s'obtient aussi comme dans la preuve pour ${\cal G}$. \cqfd \section{Groupe de jauge et classifiant de Milnor} \label{P:gfibuniv} Les r\'esultats de ce paragraphe seront utilis\'es pour la preuve du th\'eor\`eme D et les techniques se retrouveront dans la preuve du th\'eor\`eme C. Soit $\xi:E\hfl{p}{} X$ et $\eta:A\hfl{\bar p}{} B$ deux $G$-fibr\'e principaux num\'erisables. Tous les espaces sont point\'es. D\'esignons par ${\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)$ l'espace des applications continues point\'ees $G$-\'equivariantes de $E$ dans $A$, muni de la CO-topologie. Le passage au quotient donne une application $q: {\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)\to{\rm Map\,}^{\scriptscriptstyle\bullet}(X,B)_\xi$ o\`u ${\rm Map\,}^{\scriptscriptstyle\bullet}(X,B)_\xi$ est l'espace des applications continues point\'ees $h:X\to B$ telles que $h^*\eta\approx\xi$, muni de la CO-topologie. Le groupe de jauge ${\cal G}_1$ de $\xi$ agit \`a droite sur ${\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)$ par pr\'e-composition et $q(f\kern .7pt {\scriptstyle \circ} \kern 1pt\chi)=q(f)$. \begin{Theorem}\label{T:gfhinv} Supposons que $X$ est localement compact, que ${\rm Map\,}^{\scriptscriptstyle\bullet}(X,B)_\xi$ est semi-localement contractile et que $G$ est \hbox{\sl SIN} . Alors, l'application $q: {\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)\to{\rm Map\,}^{\scriptscriptstyle\bullet}(X,B)_\xi$ est un ${\cal G}_1$-fibr\'e principal. \end{Theorem} Pour d\'emontrer le th\'eor\`eme \ref{T:gfhinv}, on utilise, dans l'esprit du \S\ \ref{unif}, une sous-base particuli\`ere de la CO-topologie sur ${\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)$. Consid\'erons l'ensemble ${\cal T}$ des applications $G$-equivariantes $\tau : q^{-1}(U_\tau)\to G$ o\`u $U_\tau$ est un ouvert de $B$. Comme $\eta$ est un $G$-fibr\'e principal, la collection $\{U_\tau\mid \tau\in{\cal T}\}$ est un recouvrement ouvert de $B$. Soit $f\in{\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)$. Soit $K$ un compact de $X$ et $\tau\in{\cal T}$ tel que $f(p^{-1}(K))\in q^{-1}(U_\tau)$. Soit encore $V\in{\cal V}_G$. On d\'efinit ${\bf W} (f,K,\tau,V)$ comme l'ensemble des $\tilde f\in{\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)$ telles que , pour tout $z\in p^{-1}(K)$ on ait $q\kern .7pt {\scriptstyle \circ} \kern 1pt\tilde f(z)\subset U_\tau$ et $\tau(\tilde f (z)) \in \tau( f (z))\cdot V$. \begin{Lemme}\label{T:deutop} Si $G$ est \hbox{\sl SIN} , les ensembles ${\bf W} (f,K,\tau,V)$ forment une sous-base de la CO-topologie sur ${\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)$. \end{Lemme} \noindent {\sc Preuve: \ } Appelons ${\bf T}$ la topologie engendr\'ee par les ${\bf W} (f,K,\tau,V)$ et ${\bf T}_{co}$ la CO-topologie. Pour $L$ un compact de $E$ et $S$ un ouvert de $A$, notons $CO(L,S):=\{h\in{\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)\mid h(L)\subset S\}$. Les ensembles $CO(L,S)$ forment la sous-base standard de ${\bf T}_{co}$. Soit ${\bf W} (f,K,\tau,V)$. Soit $\sigma:U_\tau\to A$ une section de $\eta$ restreint \`a $U_\tau$. Comme l'application quotient $q(f)\in{\rm Map\,}^{\scriptscriptstyle\bullet}(X,B)_\xi$ envoie $K$ dans $U_\tau$, le fibr\'e $\xi$ admet une section $\hat\sigma$ au dessus de $K$ telle que $f\kern .7pt {\scriptstyle \circ} \kern 1pt \hat \sigma = \sigma\kern .7pt {\scriptstyle \circ} \kern 1pt q(f)$. Soit $L:=\hat\sigma (K)$ et $S:=\sigma(U)\cdot W$ o\`u $W\in{\cal V}_G$ est $G$-invariant et contenu dans $V$. Soit $\tilde f\in CO(L,S)$. Si $z\in p^{-1}(K)$, on a $z=z_0\cdot g$ avec $z_0:=\hat\sigma(p(z))\in L$ et $$\tilde f(z) := \tilde f (z_0)\cdot g \in f(z_0)\cdot W\cdot g = f(z_0)\cdot g\cdot (g^{-1}\, W\, g) = f(z)\cdot W.$$ Ceci montre que $f\in CO(L,S)\subset {\bf W} (f,K,\tau,V)$ et donc ${\bf T} \subset {\bf T}_{co}$. R\'eciproquement, soient $L$ un compact de $E$ et $S$ un ouvert de $A$ avec $f(L)\subset S$. Supposons tout d'abord que $S=S(\tau,V):=\sigma(U_\tau)\times V$ pour $V\in{\cal V}_G$ et $\sigma$ la section de $\eta$ restreint \`a $U_\tau$ telle que $\tau\kern .7pt {\scriptstyle \circ} \kern 1pt \sigma (y)=e$, l'\'el\'ement neutre de $G$. On a donc $\tau\kern .7pt {\scriptstyle \circ} \kern 1pt f (L)\subset V$. Par l'analogue du lemme de Lebesgue \cite[II.4.3]{Bo}, il existe $W\in{\cal V}_G$ tel que pour $\tau(f(z))\cdot W\subset V$ pour tout $z\in K$. Il est alors clair que $f\in W(f,p(L),\tau ,W)\subset CO(L;S)$. Comme les ouverts $S(\tau,V)$ forment une base de la topologie de $A$, on aura, en g\'en\'eral $$CO(L,S)=\bigcap _{i=1}^m CO(L_i,S(\tau_i,V_i))\supset \bigcap _{i=1}^m {\bf W} (f,p(L_i),\tau_i ,W_i) \ni f. $$ On a ainsi prouv\'e ${\bf T} \supset {\bf T}_{co}$. \cqfd \paragraph{Preuve du th\'eor\`eme \ref{T:gfhinv}~: } \begin{ccote}\label{PP0} Principe de la d\'emonstration~: \rm on va montrer que~: \begin{enumerate} \item $q$ est continue, ${\cal G}_1$-\'equivariante et induit une injection de ${\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A) /{\cal G}_1$ dans ${\rm Map\,}^{\scriptscriptstyle\bullet}(X,B)_\xi$. \item l'action ${\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)\times{\cal G}_1\to{\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)$ est libre et continue. \item si $f_1,f_2\in{\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)$ satisfont $q(f_1)=q(f_2)$, les points 1) et 2) donnent un unique $\delta(f_1,f_2)\in {\cal G}_1$ tel que $f_2=f_1\kern .7pt {\scriptstyle \circ} \kern 1pt \delta(f_1,f_2)$. Ceci d\'efinit une application $\delta : {\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)\times_{{\rm Map\,}^{\scriptscriptstyle\bullet}(X,B)_\xi} {\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)\to{\cal G}_1$. On d\'emontre que $\delta$ est continue. \item l'application $q$ admet des sections locales continues. \end{enumerate} Les points ci-dessus permettent de d\'emontrer que $q$ est un ${\cal G}_1$-fibr\'e principal. En effet, pour construire des trivialisations locales, on choisit une section continue $s : T\to {\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)$ de $q$ au dessus d'un ouvert $T$ de ${\rm Map\,}^{\scriptscriptstyle\bullet}(X,B)_\xi$. La correspondance $(f,\chi)\to s(f)\kern .7pt {\scriptstyle \circ} \kern 1pt \chi$ donne une bijection continue ${\cal G}_1$-\'equivariante $\tau : T\times{\cal G}_1\to q^{-1}(T)$. Son inverse $\tau^{-1}(\tilde f) = (q(\tilde f), \delta(s(q(\tilde f)),\tilde f))$ \'etant continue par 3), l'application $\tau$ est un hom\'eo\-mor\-phisme. \end{ccote} \begin{ccote}\label{PP1} L'application $q: {\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)\to{\rm Map\,}^{\scriptscriptstyle\bullet}(X,B)_\xi$ est continue~: \rm Soit $f\in{\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)$. Soit $K$ un compact de $X$ et $U$ un ouvert de $B$ avec $q(f)\in CO(K,U)$. On peut \'ecrire $K=\bigcup_{i=1}^m K_i$ o\`u $K_i$ est un compact contenu dans le domaine de d\'efinition d'une section locale $\sigma_i$ de $\xi$. On a alors $$q\big( \bigcap_{i=1}^m CO(\sigma_i(K_i),p^{-1}(U)\big) \subset \bigcap_{i=1}^m CO(K_i,U) = CO(K,U). $$ \end{ccote} \begin{ccote}\label{PP2} L'action ${\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)\times{\cal G}_1\to{\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)$ est libre et continue~: \rm Il est banal que cette action est libre. Soit $(f,\chi)\in{\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)\times{\cal G}_1$. Consid\'erons un voisinage de $f\kern .7pt {\scriptstyle \circ} \kern 1pt\chi$ de la forme ${\bf W} (f\kern .7pt {\scriptstyle \circ} \kern 1pt\chi,K,\tau,V)$. Soit $W\in{\cal V}_G$ tel que $W\cdot W\subset V$. Comme dans la preuve de la proposition \ref{unicalg}, on v\'erifie que $\tilde f\kern .7pt {\scriptstyle \circ} \kern 1pt \tilde\chi\in {\bf W} (f\kern .7pt {\scriptstyle \circ} \kern 1pt\chi,K,\tau,V)$ si $\tilde f\in {\bf W} (f,K,\tau,W)$ et $(\tilde\chi,\chi)\in{\cal O}^{\cal G}(K,W)$. \end{ccote} \begin{ccote}\label{PP3} L'application $q$ induit une injection continue de ${\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)/{\cal G}_1$ dans\\ ${\rm Map\,}^{\scriptscriptstyle\bullet}(X,B)_\xi$~: \rm Il est clair que $q$ est ${\cal G}$-invariante. Supposons que $q(f_1)=q(f_2)=:f$. On a alors des uniques isomorphismes $\hat f_i : E\hfl{\approx}{} E(f^*\eta)$, ($i=1,2$). La composition $\phi:=\hat f_2^{-1}\kern .7pt {\scriptstyle \circ} \kern 1pt \hat f_1$ est un \'el\'ement de ${\cal G}$ et $f_1=f_2\kern .7pt {\scriptstyle \circ} \kern 1pt \phi$. (Observons que $q$ est \'evidemment surjective. Nous omettons ce fait car il est redonn\'e par le point \ref{PP5}. Cette \'economie sera avantageuse dans la preuve du th\'eor\`eme B). \end{ccote} \begin{ccote}\label{PP4} Continuit\'e de l'application $\delta$ : \rm Soient $f_1,f_2\in{\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)$ telles que $f_2=f_1\kern .7pt {\scriptstyle \circ} \kern 1pt \chi$. Il s'agit de montrer que $\chi$ d\'epend contin\^ument du couple $(f_1,f_2)$. Soient $K$ un compact de $X$ et $V\in{\cal V}_G$. Soit $W\in{\cal V}_G$ tel que $W\cdot W\subset V$. Supposons tout d'abord qu'il existe $\tau$ telle que $K\subset f_1^{-1}(U_\tau)\cap f_2^{-1}(U_\tau)$. Soient $\tilde f_i\in {\bf W} (f_i,K,\tau ,W)$ ($i=1,2$) avec $\tilde f_2 = \tilde f_1\kern .7pt {\scriptstyle \circ} \kern 1pt\tilde\chi$. Pour $z\in p^{-1}(K)$, on a $$\tau (\tilde f_2(z)) \in \tau (f_2(z))\cdot W = \tau (f_1(\chi(z)))\cdot W \in \tau (\tilde f_1(\chi(z)))\cdot W\cdot W .$$ D'autre part~: $$\tau (\tilde f_2(z)) = \tau (\tilde f_1(\tilde\chi(z)))= \tau (\tilde f_1(\chi(z)))\gamma(\tilde\chi(z),\chi(z)).$$ Comme $V=V^{-1}$, on aura $\tilde\chi (z)\in\chi (z) \cdot V$. Dans le cas g\'en\'eral, on utilise que $K$ est une r\'eunion finie de compacts $K_i$ tels que $K\subset f_1^{-1}(U_{\tau_i})\cap f_2^{-1}(U_{\tau_i})$. \end{ccote} \begin{ccote}\label{PP5} Construction de sections locales~: \rm Nous allons construire une section locale au dessus de chaque ouvert $T$ de ${\rm Map\,}^{\scriptscriptstyle\bullet}(X,B)_\xi$ dont l'inclusion $T\subset{\rm Map\,}^{\scriptscriptstyle\bullet}(X,B)_\xi$ est contractile. On suppose donc qu'il existe une application $H:I\times T\to {\rm Map\,}^{\scriptscriptstyle\bullet}(X,B)_\xi$, o\`u $I=[0,1]$, telle que $H(0,f)=f$ et $H(1,f)=f_1$. Comme $X$ est localement compact, $H$ donne naissance \`a une application continue $h:I\times T\times X\to B$ \cite[p. 261]{Du} telle que $h(t,f,*)=*$. D\'esignons par $h_t:T\times X\to A$ l'application continue $h_t(f,x)=h(t,f,x)$ et par ${\cal E}_t$ l'espace total du fibr\'e induit sur $T\times X$ par $h_t$~: ${\cal E}_t:=E(h_t^*\eta)$. Vu que $h_t(f,*)=*$, l'espace ${\cal E}_t$ est point\'e par $(*,\tilde *)$. Comme $\eta$ est num\'erisable, le rel\`evement des homotopies donne un isomorphisme de $G$-fibr\'es point\'es de $h_1^*\eta$ sur $h_0^*\eta$. D'autre part, on a des isomorphismes de $G$-fibr\'es point\'es $h_1^*\eta\approx T\times f_1^*\eta\approx T\times \xi$. Tout ceci forme un diagramme commutatif $$\begin{array}{cccccccccccccc} T\times E & \hfl{\approx}{} & T\times E(h_1^*\eta) &\hfl{\approx}{} & {\cal E}_1 &\hfl{\approx}{} & {\cal E}_0 &\hfl{}{} & A \\ \downarrow && \downarrow && \downarrow && \downarrow && \downarrow \\ T\times X & \hfl{\rm id}{} & T\times X &\hfl{\rm id}{} & T\times X &\hfl{\rm id}{} & T\times X &\hfl{h_0}{} & B. \end{array}$$ Comme ${\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)$ est munie de la CO-topologie, la ligne sup\'erieure d\'eter\-mine une application continue $T\to{\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,A)$ au dessus de l'inclusion $T\subset {\rm Map\,}^{\scriptscriptstyle\bullet}(X,B)_\xi$, c'est-\`a-dire une section locale continue au dessus de $T$. \end{ccote} Par \ref{PP5}, la d\'emonstration du th\'eor\`eme \ref{T:gfhinv} est ainsi termin\'ee. De la m\^eme mani\`ere, on d\'emontre le r\'esultat analogue pour les applications non-point\'ees $q: {\rm Map\,}_G(E,A)\to{\rm Map\,}(X,B)_\xi$~: \begin{Theorem}\label{T:gfhinvnp} Supposons que $X$ est localement compact, que ${\rm Map\,}(X,B)_\xi$ est semi-localement contractile et que $G$ est \hbox{\sl SIN} . Alors, l'application \\ $q: {\rm Map\,}_G(E,A)\to {\rm Map\,}(X,B)_\xi$ est un ${\cal G}$-fibr\'e principal. \end{Theorem} Le cas particulier o\`u $\eta$ est le fibr\'e de Milnor $EG\to BG$ est int\'eressant \`a cause de la proposition suivante, utilis\'ee pour d\'emontrer le th\'eor\`eme D~: \begin{Proposition}\label{T:mapegcontr} Soit $\xi$ un $G$-fibr\'e principal num\'erisable sur $X$. Alors, l'espace ${\rm Map\,}_G(E,EG)$ est contractile. Si, de plus, $X$ est localement compact et que l'inclusion $\{*\}\subset X$ soit une cofibration, ${\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,EG)$ est faiblement contractile (i.e. ses groupes d'homotopie sont ceux d'un point). \end{Proposition} \noindent {\sc Preuve: \ } La preuve habituelle que deux applications $G$-equivariantes $f_0,f_1 : E\to EG$ sont toujours homotopes fournit une homotopie {\it canonique} entre $f_0$ et $f_1$, qui d\'epend contin\^ument de $(f_0,f_1)$ (voir \cite[prop. 12.3]{Hu}). L'espace ${\rm Map\,}_G(E,EG)$ est donc contractile (convexe), de m\^eme que $EG = {\rm Map\,}_G(G,EG)$. Pour montrer l'assertion sur ${\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,EG)$, il suffit de montrer que la suite $${\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,EG)\hookrightarrow {\rm Map\,}_G(E,EG)\hfl{\rm ev}{} EG$$ est, avec nos hypoth\`eses sur $X$, une fibration de Hurewicz, o\`u ${\rm ev}$ est l'\'eva\-lua\-tion sur le point base $\tilde *$. Soit $f:A\to {\rm Map\,}_G(E,EG)$ une application continue et $F_{\tilde *}: [0,1]\times A\times \{\tilde *\}\to EG$ une homotopie de ${\rm ev}\kern .7pt {\scriptstyle \circ} \kern 1pt f$. Par passage au quotient, on obtient $\bar f:A\to {\rm Map\,}(X,BG)$ et $\bar F: [0,1]\times A\times \{*\}\to BG$. Comme $X$ est localement compact, ces applications induisent des applications continues $\bar f\,\check{} : A\times X\to BG$ et $\bar F_*\check{} : [0,1]\times A\times \{*\}\to BG$. Etant \'equivariante, l'application induite $f\,\check{} : A\times E\to EG$ est donc aussi continue (bien que $E$ ne soit pas suppos\'e localement compact). Comme $*\subset X$ est une cofibration, il existe une r\'etraction de $[0,1]\times X$ sur $\{0\}\times X \cup [0,1]\times \{*\}$ qui permet d'\'etendre $\bar F_*\check{}$ en $\bar F\check{} : [0,1]\times A\times X\to BG$. Ceci prouve que $ {\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)\to {\rm Map\,}(X,BG)\to BG$ est une fibration de Hurewicz. Par rel\^evement des homotopies, l'espace total du fibr\'e induit sur $[0,1]\times A\times X$ par $\bar F\check{}$ est $G$-hom\'eomorphe \`a $[0,1]\times A\times E $, ce qui produit une homotopie $F\tilde {} : [0,1]\times A\to {\rm Map\,}_G(E,EG)$ partant $f$ et au dessus de $F$. \cqfd \begin{ccote} \bf Remarque : \rm On d\'eduit de \ref{T:gfhinv}, \ref{T:gfhinvnp} et \ref{T:mapegcontr} que $\pi_{i}({\rm Map\,}^\bullet(X,BG)_\xi)\approx \pi_i(B{\cal G}_1) $ et $\pi_{i}({\rm Map\,}(X,BG)_\xi)\approx \pi_i(B{\cal G}) $. Nous ne savons pas, en g\'en\'eral, si ces isomorphismes sont induits par une application ${\rm Map\,}(X,BG)_\xi\to B{\cal G}$. Observons que des r\'esultats analogues ont \'et\'e obtenus dans d'autres contextes (\cite{DDK}, \cite[Prop. 5.1.4]{DK}). \end{ccote} \section{Preuve des th\'eor\`emes B, C et D} \label{P:pthmbc} \paragraph{Pr\'eparatifs : } \begin{ccote} Continuit\'e des foncteurs de Milnor. \rm Nous aurons besoin de savoir que les foncteurs $E$ et $B$ de Milnor sont continus, ce qui ne semble pas figurer dans la litt\'erature~: \end{ccote} \begin{Lemme}\label{L:contB} Soient ${\cal R} (F,G)$ l'espace des homomorphismes continus entre les groupes topologiques $F$ et $G$. Supposons que $F$ est s\'epar\'e. Alors les applications $$ E : {\cal R} (F,G)\to {\rm Map\,}_F^{\scriptscriptstyle\bullet}(EF,EG) \qquad ( \phi\ \mapsto \ E\phi )$$ et $$ B : {\cal R} (F,G) \to {\rm Map\,}^{\scriptscriptstyle\bullet}(BF,BG) \qquad ( \phi\ \mapsto \ B\phi )$$ sont continues (tous les espaces \'etant munis de la CO-topologie). \end{Lemme} \noindent {\sc Preuve: \ } Comme l'application ${\rm Map\,}_F^{\scriptscriptstyle\bullet}(EF,EG)\to {\rm Map\,}^{\scriptscriptstyle\bullet}(BF,BG)$ est continue (se d\'emontre comme \ref{PP1}), il suffit de donner une preuve pour l'application $E$. Rappelons que si $P$ est un groupe toplogique, les \'el\'ements de $EP$ sont repr\'esent\'es par des suites $(t_jp_j)$, o\`u $j\in{\bf N}$, $t_j\in [0,1]$ et $p_j\in P$. On d\'esigne par $\tau_i : EP\to [0,1]$ l'application $\tau_i((t_jp_j)):=t_i$ et par $\gamma_i: \tau_i^{-1}(]0,1])\to P$ l'application $\gamma_i((t_jp_j)):=p_i$ (on utilise les m\^emes notations $\tau_i$ et $\gamma_i$ pour tout groupe $P$). L'espace $EP$ est muni de la topologie la plus grossi\`ere telle que les applications $\tau_i$ et $\gamma_i$ soient continues. Une sous-base ${\cal S}$ de cette topologie est donc constitu\'ee par les ouverts du type \begin{enumerate} \item\label{ty1} $\tau_i^{-1}(J)$ o\`u $J$ est un ouvert de $[0,1]$. \item\label{ty2} $\gamma_i^{-1}(V)$ o\`u $V$ est un ouvert de $P$. \end{enumerate} Soit $\phi_0\in{\cal R} (F,G)$, $K$ un compact de $EF$ et $U$ un ouvert de $EG$ tel que $\phi_0(K)\subset U$. Soit $CO(K,U):=\{\alpha\in {\rm Map\,}_F^{\scriptscriptstyle\bullet}(EF,EG) \mid \alpha(K)\subset U\}$. Il faut montrer que $E^{-1}(CO(K,U))$ est un voisinage de $\phi_0$ dans ${\cal R} (F,G)$. Il suffit de le faire pour $U\in {\cal S}$ (voir \cite[X.3.4, remarque 2]{Bo}. Si $U$ est du type \ref{ty1} ci-dessus, cela ne pose pas de probl\`eme. En effet, $\tau_i\kern .7pt {\scriptstyle \circ} \kern 1pt E = \tau_i$, d'o\`u $E^{-1}(CO(K,U))={\cal R} (F,G)$. Si $U=\gamma_i^{-1}(V)$ avec $V$ un ouvert de $G$, on pose $K_i:=\gamma_i(K)$ ($\gamma_i$ est d\'efinie sur $K$ puisque $\tau_i\kern .7pt {\scriptstyle \circ} \kern 1pt E = \tau_i$). L'espace $K_i$ est compact puisque $F$ est s\'epar\'e. Comme $\gamma_i\kern .7pt {\scriptstyle \circ} \kern 1pt E\phi = \phi\kern .7pt {\scriptstyle \circ} \kern 1pt\gamma_i$, on a $\phi_0\in CO(K_i,V) \subset E^{-1}(CO(K,U))$. \cqfd \begin{ccote}\label{applicn} L'applcation $n :\calr(\Omega_C,G)_\xi\to{\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi$. \rm En choisissant une partition de l'unit\'e $\mu_\rho$ subordonn\'ee au recouvrement $U_\rho$ ($\rho\in {\rm Cont\,}_C(X)$), on d\'etermine, gr\^ace aux trivialisations $\hat\psi_\rho$ de \eqref{deftriv2}, une application classifiante $\nu_C:X\to B\Omega_C$ (voir \cite[ch. 4, prop. 12.1 et th. 12.2]{Hu}). Il est possible de faire ce choix de mani\`ere que $\nu_C$ soit point\'ee (le point base de $EF$, pour un groupe topologique $F$, est toujours la suite $(1e,0,0,\dots)$, o\`u $e$ \'el\'ement neutre de $F$; le point base de $BF$ est l'image de celui de $EF$). Pour cela, il faut tout d'abord avoir une partition de l'unit\'e d\'enombrable et trivialisante $\mu_i$ ($i=1,2,\dots $) telle que $\mu_1(*)=1$. Or, une telle partition existe car~: a) il existe une partition de l'unit\'e $\hat \mu_\rho$ et $\hat\rho\in {\rm Cont\,}_C(X)$ tels que $\hat \mu_{\hat\rho(x)}=1$ au voisinage de $*$. Pour voir cela, on choisit $\hat\rho$ tel que $\mu_{\hat\rho}(*)\not = 0$. On consid\`ere une fonction $\delta: [0,1]\to [0,1]$ telle que $\delta(t) = 0$ si $t\geq \mu_{\hat\rho}(*)/2$ et $\delta(t) > 0$ si $t < \mu_{\hat\rho}(*)/2$. On pose $$\mu'_\rho(x) := \left\{ \begin{array}{lll} \mu_{\hat\rho}(x) & \hbox{si}\ \rho = \hat\rho\\ \delta(\mu_{\hat\rho}(x))\,\mu_\rho(x) & \hbox{sinon.}\end{array}\right.$$ Avec ces d\'efinitions, $$\hat\mu_\rho (x):=\frac{\mu'_\rho(x)}{\sum_{\sigma}\mu'_\sigma(x)}$$ est une partition de l'unit\'e avec $\hat \mu_{\hat\rho(x)}=1$ au voisinage de $*$. \vskip .2 truecm b) le proc\'ed\'e de \cite[ch. 4, prop. 12.1]{Hu} fournit, \`a partir de la partition $\hat\mu_\rho$ une partition de l'unit\'e $\mu_i$ ($i=1,2,\dots $) telle que $\mu_1(x)=1$ au voisinage de $*$. On peut supposer que $\hat\rho(*)=i_{\hat\rho(*)}$. Sinon, $\hat\rho(*)=b\in\Omega_C$ et l'on remplace $\hat\rho$ par $b^{-1}\,\hat\rho$. Dans ces conditions, l'application $\nu_C$ obtenue par \cite[ch. 4, prop. 12.1]{Hu} est point\'ee. Ayant fix\'e $\nu_C\in {\rm map\,}(X,B\Omega_C)$ comme ci-dessus, on d\'efinit $n: \calr(\Omega_C,G)_\xi\to{\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi$ par $n(\phi):=B\phi\kern .7pt {\scriptstyle \circ} \kern 1pt \nu_C$. Comme $C$ est s\'epar\'e, l'application $\phi\mapsto B\phi$ est continue (lemme \ref{L:contB}), d'o\`u la continuit\'e de $n$. \end{ccote} \paragraph{Preuve du th\'eor\`eme B : } Par le th\'eor\`eme d'existence, on a $\cala=\emptyset$ si et seulement ${\cal R} (\Omega_C,G)_\xi = \emptyset$. Dans ce cas, le th\'eor\`eme B est banal. On suppose donc que $\cala\not =\emptyset$. Le th\'eor\`eme A implique que l'image de $h$ est dans ${\cal R} (\Omega_C,G)_\xi$. Le principe de la preuve du th\'eor\`eme B est alors le m\^eme que celui du th\'eor\`eme \ref{T:gfhinv} (voir \ref{PP0}). \begin{ccote} \label{hconti} $h$ est continue. \ \rm Soit $w\in\cala$. Soit $K$ un compact de $\Omega_C$, $U$ un ouvert de $G$ tels que $h^w(K)\subset U$. En utilisant une trivialisation locale de $\xi$ au voisinage de $*$, on peut trouver un ouvert $\tilde U$ de $E$ tel que $\tilde U \cap E_* = \tilde * \cdot U$. Alors $w\in CO(K\times\{\tilde *\},\tilde U)$ qui est un ouvert de $\cala$ et \\ $h(CO(K\times\{\tilde *\},\tilde U)\subset CO(K,U)$. \end{ccote} \begin{ccote} \label{L:g1lib} l'action $\cala\times{\cal G}_1\to\cala$ est continue et libre. \ \rm La continuit\'e, utilisant le fait que $G$ est \hbox{\sl SIN} , a \'et\'e d\'emontr\'ee dans la proposition \ref{unicala}. Soit $w\in\cala$ et $\chi\in{\cal G}$. Si $\chi\not ={\rm id}_E$, il existe $z\in E$ tel que $\chi(z)\not = z$. Soit $\theta\in C_{\scriptscriptstyle p(z)}^*$. Comme $\chi$ restreinte \`a $p^{-1}(*)$ est l'identit\'e, on aura $w^\chi (\theta,z)\not = w(\theta,z)$ ce qui montre l'assertion \ref{L:g1lib}. \end{ccote} \begin{ccote} ${\cal G}$-invariance de $h$ : \ \rm Soit $\chi\in{\cal G}$. Posons $\chi(\tilde *) = \tilde *\, g$, autrement dit~: $g:=\hbox{\rm res\,}(\chi)$. Pour $c\in\Omega_C$, on a $$\begin{array}{lcl} \tilde *\, h^{(w^\chi)}(c) & = & w^\chi(c,\tilde *) = \chi^{-1}(w(c,\tilde * \cdot g)) = \chi^{-1}(\tilde * \cdot h^w(c))\cdot g =\\[2pt] & = & \chi^{-1}(\tilde * )\cdot (h^w(c)\, g) = \tilde *\cdot (g^{-1}\, h^w(c)\, g) \end{array}$$ d'o\`u \begin{equation}\label{conju34} h^{(w^\chi)}(c)= \hbox{\rm res\,} (\chi)^{-1}\, h^w(c)\, \hbox{\rm res\,}(\chi) \end{equation} et $h$ induit des applications $$\bar h_1: \cala/{\cal G}_1\to{\cal R} (\Omega_C,G)_\xi \quad \hbox{ et } \quad \bar h: \cala/{\cal G}\to{\cal R} (\Omega_C,G)_\xi/G.$$ \end{ccote} \begin{ccote}\label{injbah1} {\it Injectivit\'e de $\bar h_1$ : } \rm Soient $w,\tilde w\in\cala$ telles que $h^w=h^{\tilde w}$. Soit $z\in E$. Choisissons $\rho\in {\rm Cont\,}_C(X)$ tel que $p(z)\in U_\rho$ et notons $\theta:=\rho(p(z))\in C_{p(z)}^*$. On d\'efinit \begin{equation}\label{trgau} \chi (z):= \tilde w (\theta^{-1},w(\theta,z)). \end{equation} Nous allons montrer que l'\'egalit\'e $h^w=h^{\tilde w}$entra\^\i ne que $\chi(z)$ ne d\'epend pas du choix de $\rho$. Soit $\bar \rho$ une autre $C$-contraction, donnant $\bar\theta$ et $\bar \chi (z)$. Rappelons que la fibre $p^{-1}(*)$ est identifi\'ee \`a $G$ par $g\mapsto \tilde *\cdot g$. Via cette identification, $G$ agit \`a gauche sur $p^{-1}(*)$ et, si $c\in\Omega_C$ et $y\in p^{-1}(*)$, on a $w(c,y)= h^w(c)\cdot y$. Avec ces conventions, on a $$\begin{array}{lcl} \bar\chi(z) &=& \tilde w (\bar\theta^{-1},w(\bar\theta,z)) = \tilde w (\bar\theta^{-1},w(\bar\theta\theta^{-1}\theta,z)) = \\[2pt]&=& \tilde w (\bar\theta^{-1},w(\bar\theta\theta^{-1}, w(\theta,z))) = \tilde w (\bar\theta^{-1},h^w(\bar\theta\theta^{-1})\cdot w(\theta,z)) = \\[2pt]&=& \tilde w (\theta^{-1}\theta\bar\theta^{-1},h^w(\bar\theta\theta^{-1})\cdot w(\theta,z)) = \\[2pt]&=& \tilde w (\theta^{-1}, \underbrace{h^{\tilde w}(\theta\bar\theta^{-1}) h^w(\bar\theta\theta^{-1})}_{= 1}\cdot w(\theta,z)) = \tilde w(\theta^{-1},w(\theta,z)) = \chi(z). \end{array}$$ On a ainsi d\'efini une application $\chi : E\to E$ qui, par la formule \eqref{trgau} est continue. Son inverse s'obtient en \'echangeant $w$ et $\tilde w$. Les formules $\chi(z\cdot g) = \chi(z)\cdot g$, pour $g\in G$ et $p(\chi(z))=p(z)$ sont banales. De plus, on a $\chi(\tilde *)=\tilde *$, d'o\`u $\chi\in{\cal G}_1$. Voyons maintenant que $\tilde w^\chi = w$. Soit $z\in E$ et $c\in C_{p(z)}$. Observons que, dans la la formule \eqref{trgau}, on n'utilise la $C$-contraction $\rho$ que pour garantir la continuit\'e. La d\'efinition de $\chi(z)$ ne n\'ecessite que l'\'el\'ement $\theta\in C_{p(z)}^*$ et $\chi(z)$ ne d\'epend pas de $\theta$. En choisissant $\theta$ pour la d\'efinition de $\chi(z)$ et $\theta c^{-1}$ pour celle de $\chi^{-1}(\tilde w(c,\chi(z))$, on aura $$\begin{array}{lcl} \tilde w^\chi(c,z) &=& \chi^{-1}(\tilde w (c,\chi(z))) = \chi^{-1}(\tilde w (c,\tilde w(\theta^{-1},w(\theta,z))) = \\[2pt]&=& \chi^{-1}(\tilde w (c\theta^{-1},w(\theta,z)) = w(c\theta^{-1},\tilde w(\theta c^{-1},\tilde w (c\theta^{- 1},w(\theta,z))= \\[2pt]&=& w(c\theta^{-1},w(\theta,z)) = w(c,z). \end{array}$$ \end{ccote} \begin{ccote} L'application $\delta : \cala\times_{\calr(\Omega_C,G)_\xi}\cala\to{\cal G}$ est uniform\'ement continue~: \rm Soit $V\in{\cal V}_G$ et $K$ un compact de $X$. Supposons d'abord que $K\subset U_\rho$ pour une $C$-contraction $\rho\in {\rm Cont\,}_C(X)$. Comme $C$ est s\'epar\'e, les sous-espaces $L$ et $L^{-1}$ de $C$ d\'efinis par $$L:= \{\rho(x)\mid x\in K\}\subset C^* \qquad \hbox{et} \qquad L^{-1}:= \{\rho(x)^{-1}\mid x\in K\}\subset C_*.$$ sont compacts. Soit $W\in{\cal V}_G$ tel que $W\cdot W\in V$. Soient $\tilde w_1,\tilde w_2\in \cala$. Il suit de \ref{injbah1} que $\delta$ satisfait, pour $z\in p^{-1}(K)$, \`a l'\'equation \begin{equation}\label{E:pfb111} \delta(\tilde w_1,\tilde w_2)(z) = \tilde w_1(\theta^{-1},\tilde w_2(\theta,z)) \end{equation} avec $\theta:=\rho(p(z))\in C_{p(z)}^*$. Il s'en suit que si $(\tilde w_1, w_1)\in {\cal O}^{\cal R} (L^{-1},W)$ et $(\tilde w_2, w_2)\in {\cal O}^{\cal R} (L,W)$, alors $(\delta(\tilde w_1,\tilde w_2), \delta(w_1,w_2))\in {\cal O}^{{\cal G}_1}(K,V)$. Dans le cas g\'en\'eral, on utilise que $K:=\bigcup_{\rho\in P} K_\rho$ o\`u $P$ est un ensemble fini dans ${\rm Cont\,}_C(X)$ et $K_\rho$ est un compact de $U_\rho$. On d\'efinit, comme ci-dessus, $L_\rho:=\rho(K_\rho)$ et $L_\rho^{-1}:=\rho(K_\rho)^{-1}$ et on aura $$\left.\begin{array}{l}\displaystyle (\tilde w_1, w_1)\in \bigcap_{\rho\in P}{\cal O}^{\cal R} (L_\rho^{-1},W)\\ \hbox{et}\\[1 pt] \displaystyle (\tilde w_2, w_2)\in \bigcap_{\rho\in P}{\cal O}^{\cal R} (L_\rho,W) \end{array}\right\} \ \Rightarrow \ (\delta(\tilde w_1,\tilde w_2), \delta(w_1,w_2))\in {\cal O}^{{\cal G}_1}(K,V).$$ \end{ccote} \begin{ccote}\label{trivlocthb} Sections locales : \rm Soit $T$ un ouvert de ${\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi$ tel que l'inclusion $T\subset{\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi$ soit contractile. Soit $S:=n^{-1}(T)$, o\`u $n:\calr(\Omega_C,G)_\xi\to{\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi$ est l'application d\'efinie en \ref{applicn}. Consid\'erons l'application compos\'ee \begin{equation}\label{defN1} N: \, S\times X \hfl{n\times {\rm id}}{} T\times X \hfl{{\rm ev}_X}{} BG \end{equation} Comme $X$ est localement compact, l'\'evaluation ${\rm ev}_X$ est continue et $N$ est continue. D\'esignons par ${\cal E}:= N^*EG$, l'espace total du $G$-fibr\'e principal sur $S\times X$ induit par $N$. Observons que $N$ est aussi la composition \begin{equation}\label{defN2} N : S\times X \hfl{{\rm id_S}\times \nu_C}{} S\times B\Omega_C \hfl{{\rm ev}_\Omega}{} BG \end{equation} de ${\rm id_S}\times \nu_C$ avec l'application d'\'evaluation (peut-\^etre non-continue) ${\rm ev}_\Omega$. Comme le $\Omega_C$-fibr\'e induit sur $X$ par $\nu_C$ est $\xi_C : C_*\hfl{\beta}{} X$, on obtient, en utilisant \eqref{defN2}, que l'espace ${\cal E}$ est le quotient de $S\times C_*\times G$ par la relation d'\'equivalence $(\phi ,ub,g)\sim (\phi ,u, \phi(b)g)$, o\`u $(\phi ,u,g)\in S\times C_*\times G$ et $b\in\Omega_C$. Consid\'erons ${\cal E}$ comme un espace au dessus de $X$ par l'application $(\phi ,u,g)\mapsto \beta (u)$ et d\'efinissons l'application continue $v: C\times _X {\cal E} \to {\cal E}$ par $v(c,(\phi ,u,g)):=(\phi ,cu,g)$. Comme l'application $S\to{\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi$ est homotope \`a une application constante, on montre, comme dans \ref{PP5}, qu'il existe un hom\'eomorphisme $G$-equivariant $S\times E \hfl{\approx}{}{\cal E}$ au dessus de ${\rm id}_{S\times X}$. Via cet hom\'eomorphisme et en composant avec la projection $S\times E\to E$, l'application $v$ donne une application continue $\hat v: C\times_X (S\times E)\to E$. A son tour, $\hat v$ d\'etermine une application continue $w : S\to {\rm map\,}(C\times_X E, E)$. On v\'erifie facilement que l'image de $w$ est dans $\cala$. La pr\'eservation des points base par les divers isomorphismes utilis\'es $\clubsuit$ entra\^\i ne que $w$ est une section locale de $h$ au dessus de $S$. \end{ccote} Ayant \'etabli les points \ref{hconti} \`a \ref{trivlocthb}, la d\'emonstration du th\'eor\`eme B se termine comme expliqu\'e dans \ref{PP0}. \cqfd \paragraph{Preuve du th\'eor\`eme C : } Soient $w,\tilde w\in\cala$ et $g\in G$ tels que $h^w=g^{-1}h^{\tilde w}g$. Comme $G$ est connexe par arc, il existe, par le lemme \ref{numer}, un \'el\'ement $\chi\in{\cal G}$ tel que $\chi(\tilde *)=\tilde * \cdot g$. L'\'equation \eqref{conju34} montre qu'alors $h^{w^\chi}= h^{\tilde w}$. Comme $\bar h_1$ est injective par \ref{injbah1}, les repr\'esentations ${w^\chi}$ et $\tilde w$ sont dans la m\^eme classe modulo ${\cal G}_1$. Cela prouve l'injectivit\'e de $\bar h$. Le diagramme commutatif \begin{equation}\label{hh1} \begin{array}{cccccc} \cala/{\cal G}_1 & \hfl{\bar h_1}{} & {\cal R} (\Omega_C,G)_\xi \\ \downarrow && \downarrow \\ \cala/{\cal G} & \hfl{\bar h}{} & {\cal R} (\Omega_C,G)_\xi/\hbox{\small G} \end{array}\end{equation} et le fait que $\bar h_1$ soit un hom\'eomorphisme (vu le th\'eor\`eme B) font que $\bar h$ est un hom\'eomorphisme. \paragraph{Preuve du th\'eor\`eme D :}\label{P:pthmd} \begin{Lemme}\label{appator} L'application $n$ est couverte par un morphisme de ${\cal G}_1$-fibr\'es principaux \begin{equation}\label{appatordiag} \begin{array}{cccccc} \cala\ \ & \hfl{\hat n}{} & {\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,EG) \\ \downarrow\! h && \downarrow\\ \calr(\Omega_C,G)_\xi & \hfl{n}{} & {\rm Map\,}^{\scriptscriptstyle\bullet}(X,BG)_\xi .\end{array}\end{equation} \end{Lemme} Le th\'eor\`eme D d\'ecoulera du lemme \ref{appator} puisque, i l'inclusion $\{*\}\subset X$ est une cofibration, l'espace ${\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,EG)$ est faiblement contractile (Proposition \ref{T:mapegcontr}). En fait, l'application $n$ est classifiante pour le ${\cal G}_1$-fibr\'e principal $h : \cala\to\calr(\Omega_C,G)_\xi$. \vskip .2 truecm {\sc Preuve du lemme \ref{appator} : } Le proc\'ed\'e pour fixer l'application $\nu_c$ vu en \ref{applicn} produit en fait un diagramme commutatif $$\begin{array}{cccc} C_* & \hfl{\hat \nu_C}{} & E\Omega_C\\ \downarrow\!\! \beta &&\downarrow\\ X & \hfl{\nu_C}{} & E\Omega_C\end{array}$$ o\`u $\hat \nu_C\in {\rm map\,}_G^{\scriptstyle\bullet}(C_*,E\Omega_C)$. L'application $\hat n\in {\rm map\,}_{\cal G}(\cala,{\rm Map\,}_G^{\scriptscriptstyle\bullet}(E,EG))$ est d\'efinie par $\hat n(w):=Eh^w\kern .7pt {\scriptstyle \circ} \kern 1pt \hat\nu_C$. Le diagramme \ref{appatordiag} est bien commutatif. Observons que $\hat n$ est obtenue par le proc\'ed\'e de \cite[ch. 4, prop. 12.1]{Hu} \`a l'aide de la partition de l'unit\'e $\hat \mu_\rho$ construite en \ref{applicn} et des trivialisations $\psi^w_\rho$ du lemme \ref{rectrican}. \section{Exemples et applications} \label{P:exp} \subsection{Le groupo\"\i de associ\'e \`a un fibr\'e principal} \label{SP:groassfp} Soit $\xi : E\hfl{p}{} X$ un $G$-fibr\'e principal. Soit $EE:=EE(\xi):=(E\times E)/G$, le quotient de $E\times E$ par l'action diagonale de $G$. D\'enotons par $\llangle{a}{b}$ l'orbite de $(a,b)$ dans $EE$. On fait de $EE$ un $X$-groupo\"\i de \ en posant $\alpha (\llangle{a_2}{a_1}) := p(a_1)$, $\beta (\llangle{a_2}{a_1}) := p(a_2)$, $i_x:=\llangle{a}{a}$ avec $p(a)=x$; la composition vient de la formule $\llangle{a_3}{a_2}\llangle{a_2}{a_1}=\llangle{a_3}{a_1}$, ou, plus g\'en\'eralement, $$\llangle{a_3}{a'_2}\llangle{a_2}{a_1}=\llangle{a_3\cdot \gamma(a'_2,a_2)}{a_1}$$ o\`u $\gamma$ est l'application d\'efinie en \eqref{defgamma}. L'inverse est \'evidemment donn\'e par $\llangle{a}{b}^{-1}=\llangle{b}{a}$. Le $X$-groupo\"\i de \ ainsi obtenu s'appelle le {\it $X$-groupo\"\i de \ associ\'e \`a $\xi$} \cite[p. 5]{Ma}. Le groupe structural $G$ de $\xi$ est isomorphe \`a $\Omega_{EE}$ par $g\mapsto \llangle{\tilde * \cdot g}{\tilde *}$. Observons que $EE$ est localement trivial num\'erisable si $\xi$ est num\'eridable. En effet, $x\mapsto \llangle{\tilde *}{\sigma(x)}$ est une $EE$-contraction lorsque $\sigma$ est une section locale de $\xi$. Le $X$-groupo\"\i de \ $EE$ a une repr\'esentation tautologique $v$ sur $\xi$ d\'etermin\'ee par l'application continue $v : (E\times E)\times E \to E$ d\'efinie par $v((a,b),z):=a\cdot\gamma(b,z)$; son holonomie est ${\rm id\,}_G$. Pour un $X$-groupo\"\i de \ $C$, d\'esignons par ${\rm Mor\,}(C,EE)$ l'ensemble des morphismes continus de $C$ dans $EE$ (au dessus de ${\rm id\,}_X$). \begin{Proposition}\label{reptoto} La composition avec la repr\'esentation tautologique donne une bijection de ${\rm Mor\,}(C,EE(\xi))$ sur $\cala$. \end{Proposition} \noindent {\sc Preuve: \ } La bijection inverse $w\mapsto \Psi_w$ est donn\'ee par $\Psi_w(c):=\llangle{w(c,z)}{z}$ ($z\in E_{\alpha(c)}$). La seule chose non-triviale \`a v\'erifier est que $\Psi_w$ est continu. Soit $c\in C$ et $T\ni \Psi_w(c)$ un ouvert de $EE$. Soit $z\in E_{\alpha(c)}$. Il existe \begin{itemize} \item $V$ un ouvert de $E$ contenant $w(c,z)$, \item $U$ un ouvert de $X$ contenant $x:=\alpha(c)=p(z)$ et $\sigma :U\to E$ une section continue locale de $\xi$ avec $\sigma(x)=z$ et \item $W\in {\cal V}_G$ \end{itemize} tels que $\pi (V\times (\sigma (U)\cdot W))\subset T$, o\`u $\pi$ d\'esigne la projection de $E\times E$ sur $EE$. Par continuit\'e de $w$, il existe $A$ un ouvert de $C$ contenant $c$ et $R$ un ouvert de $E$ contenant $z$ tels que $w(A\times R)\subset V$. Soient $U'$ un ouvert de $X$ et $W'\in{\cal V}_G$ tels que $x\in U'\subset U$, $W'\subset W$ et $\sigma(u')\cdot W'\subset R$. Soit $A':= A\cap \alpha^{-1}(U')$. On a $c\in A'$ et $\Psi_w(A')\subset \pi(V\times (\sigma(U')\cdot W')\subset T$, ce qui prouve la continuit\'e de $\Psi_w$ en $c$. \cqfd \subsection{Pr\'egroupo\"\i des} \label{SP:pregro} Comme nous le verrons plus loin, une fa\c con commode de d\'efinir un groupo\"\i de \ topologique est de partir d'une cat\'erorie topologique avec anti-involution (pr\'egroupo\"\i de ). Soit $X$ un espace topologique muni d'un point base $*\in X$. Un $X$-{\it pr\'egroupo\"\i de \ } est un espace topologique $\tilde C$ muni de deux applications continues $\alpha,\beta : \tilde C\to X$, d'une {\it composition} partiellement d\'efinie et d'une application $x\mapsto i_x$ de $X$ dans $\tilde C$ qui satisfont aux propri\'et\'es a), b) et c) de la d\'efinition du \S\ 1 d'un $X$-groupo\"\i de . En revanche, la condition d) consiste seulement en d') $\tilde C$ est muni d'une anti-involution continue $c\mapsto \bar c$, envoyant $\tilde C^y_x$ sur $\tilde C^x_y$. Une {\it repr\'esentation} d'un $X$-pr\'egroupo\"\i de \ $\tilde C$ sur un $G$-espace $\xi$ est une application continue $w: \tilde C\times_X E\to E$ (o\`u $\tilde C$ est vu au dessus de $X$ via $\alpha$) telle que, pour tout $c,d\in \tilde C$, $z\in E$ et $g\in G$, on ait \begin{enumerate} \item $p(w(c,z)) = \beta(c)$. \item $w(c\, d,z)=w(c,w(d,z))$. \item $w(\bar c,(w(c,z))=z$. \item $w(c,z\cdot g) = w(c,z)\cdot g$. \end{enumerate} Si $X$ est s\'epar\'e, un $X$-pr\'egroupo\"\i de \ d\'etermine un $X$-groupo\"\i de \ s\'epar\'e, avec la propri\'et\'e suivante : \begin{Proposition}\label{predefgro} Soit $\tilde C$ un $X$-pr\'egroupo\"\i de \ localement trivial num\'erisable avec $X$ s\'epar\'e. Alors, il esiste un unique $X$-groupo\"\i de \ s\'epar\'e localement trivial num\'erisable $C$ avec un morphisme continu surjectif $\tilde C\to C$ satisfaisant \`a la condition suivante~: toute repr\'esentation de $\tilde C$ sur un $G$-fibr\'e principal $\xi$ au dessus de $X$, avec $G$ s\'epar\'e, est induite par une unique repr\'esentation de $C$. \end{Proposition} La d\'emonstration de \ref{predefgro} utilise le lemme suivant~: \begin{Lemme}\label{grosep} Soit $C$ un $X$-groupo\"\i de \ localement trivial. Alors, $C$ est s\'epar\'e si et seulement si $X$ et $\Omega_C$ le sont. \end{Lemme} \noindent {\sc Preuve: \ } Supposons que $X$ et $\Omega_C$ soient s\'epar\'es (l'autre sens est banal). Comme $C$ est localement trivial, $\beta : C_*\to X$ est un $\Omega_C$-fibr\'e principal, par le th\'eorme A. L'espace $C_*$ est donc s\'epar\'e. Il en est \'evidemment de m\^eme pour $C^*$. On utilise alors que $C= C_*\times_{\Omega_C}C^*$ pour \'etablir que $C$ est s\'epar\'e. \cqfd \vskip .2 truecm\noindent {\sc Preuve de la proposition \ref{predefgro} : } Soit $\hat C$ l'ensemble quotient de $\tilde C$ par la relation d'\'equivalence engendr\'ee par $c\bar u u d \sim cd$, pour tout $c,d,u\in\tilde C$ avec $\beta (d) = \alpha(u) = \alpha (c)$. Il est clair que la structure (alg\'ebrique) de $X$-pr\'egroupo\"\i de \ descend sur $\hat C$ et que $\hat C$ est un $X$-groupo\"\i de (alg\'ebrique), avec $[c]^{-1} = [\bar c]$. La topologie sur $\hat C$ sera obtenue de la mani\`ere suivante~: consid\'erons l'ensemble ${\cal K}$ des paires $(K,T)$, o\`u \begin{itemize} \item $K$ est un sous-groupe normal de $\Omega_{\hat C}$. Le quotient de $\hat C$ par le sous-groupo\"\i de \ normal $N_K:=\{u K u^{-1}\mid u\in \hat C_*\}$ est alors un $X$-groupo\"\i de. \item $T$ est une topologie sur $\hat C/N_K$ qui fait de $\hat C/N_K$ un groupo\"\i de \ topologique et telle que la projection $\tilde C \to \hat C/N_K$ soit continue. \end{itemize} Les projections $\hat C \to \hat C/N_K$ ($(K,T)\in{\cal K}$) forment un syst\`eme projectif de $X$-groupo\"\i des \ au dessous de $\hat C$ (non-vide, car on peut prendre pour $T$ la topologie grossi\`ere). On munit $\hat C$ de la topologie de limite projective pour ce syst\`eme de projections. On v\'erifie que $\hat C$ est alors un $X$-groupo\"\i de, que $\tilde C \to \hat C$ est continue et que tout morphisme continu de $\tilde C$ dans un $X$-groupo\"\i de \ se factorise de fa\c con unique par l'un des quotient $\hat C/N_K$. Comme $\tilde C$ est localement trivial num\'erisable, $\hat C$ l'est aussi. Nous ignorons si la topologie ainsi obtenue sur $\hat C$ est la topologie quotient de celle de $\tilde C$. Comme $X$ est s\'epar\'e, l'adh\'erence de $\{i_*\}$ est contenue dans $\Omega_C$ o\`u elle constitue un sous-groupe ferm\'e. En quotientant $\hat C$ par le sous-groupo\"\i de \ normal engendr\'e par $\overline{\{i_*\}}$, on obtient, avec la topologie quotient, un $X$-groupo\"\i de \ $C$ \cite[Th. 2.15, p. 38]{Ma}. Comme $\overline{\{i_*\}}$ est ferm\'e dans $\Omega_C$, le groupe $\Omega_C$ est s\'epar\'e. On en d\'eduit que $C$ est s\'epar\'e par le lemme \ref{grosep}. Il est aussi clair que que tout morphisme continu de $\tilde C$ dans un $X$-groupo\"\i de \ s\'epar\'e factorise de fa\c con unique par $\tilde C\to C$. Soit $\tilde w : \tilde C\times E \to E$ une repr\'esentation de $\tilde C$ sur un $G$-fibr\'e principal $\xi$. Comme dans la proposition \ref{reptoto}, $\tilde w$ d\'etermine un morphisme continu $\Psi_{\tilde w}$ de $\tilde C$ dans $EE:=EE(\xi)$ tel que $\Psi_{\tilde w}(\bar c) = \Psi_{\tilde w}(c)^{-1}$. Puisque $EE$ est localement trivial et que $X$ et $G$ sont s\'epar\'es, l'espace $EE$ est s\'epar\'e par le lemme \ref{grosep}. Le morphisme $\Psi_{\tilde w}$ se factorise en un unique morphisme continu $\Psi_w:C\to EE$ correspondant, par la proposition \ref{reptoto}, \`a $w\in\cala$. \cqfd \subsection{Groupo\"\i des de chemins} \label{SP:groch} \begin{ccote}Chemins de Moore : \ \rm Soit $X$ un espace topologique s\'epar\'e. Soit $\tilde C = \tilde {\bf Ch}(X)$ la cat\'egorie des {\it chemins de Moore} dans $X$. Un chemin de Moore est un couple $(a,c)$ o\`u $a\in{\bf R}_{\geq 0}$ et $c : [0,a] \fl{}{} X$ est une application continue. La topologie sur $\tilde C$ est induite par l'application $(a,c) \mapsto c^{\sharp}$ de $\tilde C$ dans ${\rm map\,}([0,1],X)$, o\`u $c^{\sharp}(t)=c(at)$, avec la CO-topologie sur ${\rm map\,}([0,1],X)$). (Observons que $\tilde C$ n'est pas s\'epar\'e puisque tous les chemins constants ont m\^eme image dans ${\rm map\,}([0,1],X)$). Les applications $\alpha,\beta : C\to X$ sont donn\'ees par $\alpha(a,c)=c(a)$ et $\beta (a,c)=c(0)$. L'espace $C_x^y$ est donc l'ensemble des chemins allant de $y$ \`a $x$ (cette malencontreuse inversion est due \`a la convention usuelle de la r\`egle de composition des chemins). La composition $(a,c)=(a_2,c_2)(a_1,c_1)$, lorsque $\beta (a_1,c_1)=\alpha (a_2,c_2)$ est d\'efinie par $a:=a_2+a_1$ et $$ c(t) = \cases{ c_2(t) & si $t \leq a_2$ \cr c_1(t-a_2) & si $t \geq a_2$ \cr }$$ Cette composition est bien associative et l'unit\'e $i_x$ est donn\'ee par le chemin constant $[0,0] \fl{}{} \{x\}$. La d\'efinition de l'involution est donn\'ee par $\overline{(a,c)}:=(a,\bar c)$ o\`u $\bar c(t):=c(a-t)$. On v\'erifie facilement que ${\bf Ch}(X)$ est un pr\'egroupo\"\i de. Il est localement trivial si et seulement si $X$ est connexe par arc et semi-localement contractile. Le groupo\"\i de s\'epar\'e associ\'e \`a $\tilde {\bf Ch}(X)$ par la proposition \ref{predefgro} sera not\'e ${\bf Ch}(X)$ et appel\'e le {\it groupo\"\i de \ des chemins de Moore dans $X$}. \end{ccote} \begin{ccote} Le groupo\"\i de fondamental : \rm Soit ${\bf Ch}_1(X)\subset {\bf Ch}(X)$ l'ensemble des classes de chemins $(a,c)$ qui sont des lacets ($c(0)=c(a)$) et tels qu'il existe une homotopie $H:[0,a]\times [0,1]\to X$ telle que $H(0,s)=H(a,s)=c(0)$, $H(t,0)=c(t)$ et $H(t,1)=c(0)$. Les \'el\'ements de ${\bf Ch}_1(X)$ forment un sous-$X$-groupo\"\i de \ normal totalement intransitif de ${\bf Ch}(X)$. L'espace quotient $\pi(X)$ h\'erite donc d'une structure de $X$-groupo\"\i de \ et la projection ${\bf Ch}(X)\to\pi(X)$ est un morphisme continu \cite[Th. 2.15, p. 38]{Ma}. Il est clair qu'alg\'ebriquement, $\pi(X)$ s'identifie au groupo\"\i de \ fondamental de $X$ et $\Omega_{\pi(X)}$ au groupe fondamental $\pi_1(X,*)$ \cite[Ch. 1, \S\ 7]{Sp}. Cependant, $\pi(X)$ est ici muni d'une topologie. Il est localement trivial si $X$ est connexe par arc et semi-localement simplement connexe \cite{Sp}. Si, de plus, $X$ est localement connexe par arc, on peut montrer que $\Omega_{\pi(X)}=\Pi_1(X,*)$ est discret, que le $\Omega_{\pi(X)}$-fibr\'e principal du th\'eor\`eme A est le rev\^etement universel de $X$ et que la topologie sur $\pi(X)$ s'identifie \`a celle de \cite{BD}. Nous n'utiliserons pas ces r\'esultats. La proposition suivante est int\'eressante pour les $G$-fibr\'es principaux avec $G$ un groupe de Lie. \end{ccote} \begin{Proposition}\label{petitssgr} Soit $G$ un groupe topologique admettant un voisinage de son \'el\'ement neutre qui ne contienne aucun sous-groupe non-trivial. Alors, toute repr\'esentation de ${\bf Ch}(X)$ sur un $G$-espace principal $\xi$ au dessus de $X$ se factorise par une repr\'esentation du groupo\"\i de \ fondamental $\pi(X)$ de $X$. \end{Proposition} \noindent {\sc Preuve: \ } Soit $w\in{\cal R} ({\bf Ch}(X),\xi)$. On regarde $w$ comme un morphisme continu $w:{\bf Ch}(X)\to EE(\xi)$ par le lemme \ref{reptoto}. Il s'agit de montrer que ${\bf Ch}_1(X)\subset \ker w$. Pour $x\in X$, d\'esignons par ${\rm Vois\,}_x$ l'ensemle des voisinages ouverts de $x$. Observons que l'ensemble $\{\tilde {\bf Ch}(W)^x_x\mid W\in {\rm Vois\,}_x\}$ constitue un syst\`eme fondamental de voisinages ouverts de $i_x$ dans le mono\"\i de $\tilde {\bf Ch}(X)^x_x$. En choisissant $\tilde x\in E_x$, on obtient une holonomie $h^w_x : \tilde {\bf Ch}(X)^x_x \to G$ en $x$ qui est un morphisme continu de mono\"\i des avec anti-involution. Comme il existe un voisinage de l'\'el\'ement neutre dans $G$ qui ne contient aucun sous-groupe non-trivial et que chaque \'el\'ement de $\{\tilde {\bf Ch}(W)^x_x\mid W\in {\rm Vois\,}_x\}$ est un sous-mono\"\i de avec anti-involution de $\tilde {\bf Ch}(X)^x_x$, on en d\'eduit qu'il existe $U_x\in {\rm Vois\,}_x$ tel que $\tilde {\bf Ch}(U_x)^x_x\subset \ker h^w_x$. Soit ${\bf Ch}_{\rm loc}(X)$ le sous-groupo\"\i de \ normal de ${\bf Ch}(X)$ engendr\'e par l'image des $\tilde {\bf Ch}(U_x)^x_x$ pour tous les $x\in X$. Par ce qui pr\'ec\`ede, on a ${\bf Ch}_{\rm loc}(X)\subset \ker w$. Soit $\gamma\in {\bf Ch}_1(X)^x_x$ repr\'esent\'e par un lacet $c: [0,a]\to X$ en $x$. Par d\'efinition de ${\bf Ch}_1(X)$, il existe une homotopie de lacets $H:[0,a]\times [0,1]\to X$ entre $c$ et le lacet constant. Par l'argument habituel du nombre de Lebesgue, on peut d\'ecomposer $[0,a]\times [0,1]$ en petits rectangles $R_i$ ($1=1,\dots , N$) tels que $H(R_i)\subset U_{x(i)}$. Il s'en suit facilement que $\gamma\in {\bf Ch}_{\rm loc}(X)\subset \ker w$, ce qui prouve que ${\bf Ch}_1(X)\subset \ker w$. \cqfd \subsection{Chemins lisses par morceaux -- Connexions}\label{clipamconn} Soit $X$ une vari\'et\'e diff\'erentiable $C^1$ paracompacte. On d\'enotera par $\tilde {\bf D} = \tilde {\bf D}(X)$ l'espace des chemins $C^1$ par morceau sur $X$. En tant qu'ensemble, $\tilde {\bf D}$ est le sous-pr\'egroupo\"\i de \ de $\tilde {\bf Ch}X$ form\'e des chemins qui sont des compositions de chemins $C^1$. La topologie, plus fine, s'obtient via l'application $(a,c) \mapsto c^\sharp$ en topologisant l'ensemble ${\bf M}^\sharp$ des applications $C^1$ par morceau de $[0,1]$ dans $X$. Pour cela, soit $$P = \{t_0=0 < t_1 < \cdots < t_k=1\}$$ un partage de $[0,1]$. Soit $${\bf M}_P^\sharp := \{c \in {\bf M}^\sharp \mid c\,_{|\,[t_i,t_{i+1}]} \in C^1([t_i,t_{i+1}],X) \} $$ o\`u $C^1([a,b],X)$ est l'espace des applications $C^1$ de $[a,b]$ dans $X$ muni de la topologie $C^1$. On a une application ${\bf M}_P^\sharp \to C^1([0,1],X)^k$ donn\'e par $$c \mapsto \big( c\,_{|\,[t_0,t_1]}^\sharp ,c\,_{|\,[t_1,t_2]}^\sharp , \dots ,c\,_{|\,[t_{k-1},t_k]}^\sharp \big)$$ en \'etendant la d\'efinition de $c^\sharp$ \`a une chemin $c : [a,b]\fl{}{} X$ par $c^\sharp(t):=c(a+(b-a)t)$. L'application $(a,c) \mapsto c^\sharp$ induit une topologie (non-s\'epar\'ee) sur ${\bf M}_P^\sharp$. Si $P'$ est un partage plus fin que $P$ (i.e. $P \subset P'$), on v\'erifie que l'inclusion naturelle ${\bf M}_P^\sharp \subset {\bf M}_{P'}^\sharp$ est continue. La topologie sur ${\bf M}$ est, par d\'efinition, celle de limite inductive des ${\bf M}_P^\sharp$ pour tous les partages de $[0,1]$. Le groupo\"\i de s\'epar\'e associ\'e \`a $\tilde {\bf D}(X)$ sera not\'e ${\bf D}(X)$ et appel\'e le {\it groupo\"\i de \ des chemins lisses par morceaux dans $X$}. Si $X$ est connexe, alors ${\bf D}(X)$ est localement trivial num\'erisable. On a un morphisme \'evident de groupo\"\i des topologiques de ${\bf D}(X)$ dans ${\bf Ch}(X)$. \begin{Proposition}\label{trans} Soit $G$ un groupe de Lie et $\xi : E\hfl{p}{} X$ un $G$-fibr\'e principal diff\'erentiable $C^1$ au dessus d'une vari\'et\'e $X$. Soit $A$ une connexion sur $\xi$ \cite[Ch. II]{KN}. Alors, le transport parall\`ele associ\'e \`a $A$ d\'etermine une repr\'esentation de ${\bf D}(X)$ sur $\xi$. \end{Proposition} \noindent {\sc Preuve: \ } Par la proposition \ref{predefgro}, il suffit de voir qu'une connexion $A$ d\'efinit une repr\'esentation $w_A : \tilde {\bf D}(X)\times E \to E$ de $\tilde {\bf D}(X)$, le point $w_A(c,z)$ \'etant le r\'esultat du transport $A$-parall\`ele de $z$ au dessus de $c$. Les conditions 1. \`a 4. de la d\'efinition d'une repr\'esentation d\'ecoulent imm\'ediatement des propri\'et\'es classiques du transport parall\`ele \cite[Ch. II, prop. 3.2. et 3.3]{KN}. La seule chose \`a v\'erifier est que $w_A$ est continue en tout $(c,z)\in \tilde {\bf D}(X)\times E$. Vu la topologie sur $\tilde {\bf D}(X)$, il est suffisant de la faire pour $c: \in C^1([0,1],U)$ o\`u $U$ est un ouvert de $X$ trivialisant pour $\xi$ et domaine d'une carte. On peut donc supposer que $E=U\times G$ ou $U$ est un ouvert d'un espace euclidien et $z=(c(0),e)$. Le relev\'e horizontal $\tilde c $ de $c$ partant de $z$ s'\'ecrit alors $\tilde c (t) = (c(t), g(t))$ o\`u $g\in C^1([0,1],G)$ avec $g(0)=e$. Consid\'erons un voisinage ouvert $Q$ dans $E$ de $w_A(c,z)=\tilde c (1)= (c(1),g(1))$. Il s'agit de tourver un voisinage $T$ de $(c,z)$ dans $C^1([0,1],U)\times E$ tel que $w_A(T)\subset Q$. On peut supposer que $Q$ est de la forme $S\times (g(1)\cdot V)$ o\`u $V\in{\cal V}_G$ o\`u $S$ est un ouvert de $U$. Soit $\varepsilon >0$ tel que la boule ouverte $B(0,\varepsilon)$ de centre $0$ et de rayon $\varepsilon$ dans $T_eG={\rm Lie\,}(G)$, muni de la m\'etrique de Killing, est envoy\'ee diff\'eomorphiquement sur $W\in{\cal V}_G$ avec $W\cdot W\subset V$. Par \cite[Ch. II, prop. 1.1]{KN}, la connexion $A$ est donn\'ee par une $1$-forme $\gamma_A\in\Omega^1(E,{\rm Lie\,}(G))$ et l'on a $\gamma_A(\dot{\tilde c}(t)) = 0$ pour tout $t$ puisque $\tilde c$ est horizontal. Par continuit\'e de $\gamma_A$, il existe $\delta>0$ tel que $B(c(1),\delta)\subset S$ et \begin{equation}\label{norme1} \sup_{t\in [0,1]} \{\|c_1(t)-c(t)\|,\|\dot c_1(t)-\dot c(t)\|\}<\delta \quad \Rightarrow \quad \|\gamma_A(\dot{\bar c_1}(t))\|<\varepsilon \end{equation} o\`u ${\bar c_1}(t):=(c_1(t),g(t))$. Soit $h\in C^1([0,1],G)$ la courbe telle que $\hat c_1(t)\cdot h(t)$ soit horizontal. Par \cite[preuve du lemme p. 69]{KN}, la courbe $h$ satisfait $h(0)=e$ et $\dot h(t) = T_e R_{h(t)}(\gamma_A(\dot{\bar c_1}(t))$, o\`u $R_a$ est la translation \`a droite $g\mapsto ag$. On d\'eduit de \eqref{norme1} que $\ell (h)$ est $<\varepsilon$, o\`u $\ell(h)$ est la longueur de $h$ pour la m\'etrique riemannienne sur $G$ obtenue par translations \`a droite de la m\'etrique de Killing. L'exponentielle des rayons donnant des g\'eod\'esiques minimisantes pour cette m\'etrique riemannienne, on en d\'eduit que $h(1)\in W$. L'ouvert $T:=T_\delta\times (B(c(0),\delta)\times W)$ de $C^1([0,1],U)\times G$, o\`u $T_\delta$ est l'ouvert de $C^1([0,1],U)$ apparaissant dans \eqref{norme1}, contient $(c,z)$ et satisfait $w_A(T)\subset Q$. \cqfd Le langage des repr\'esentations de groupo\"\i des \ permet de bien poser le probl\`eme suivant~: quand est-ce que le transport parall\`ele associ\'e \`a une connexion sur un fibr\'e diff\'erentiable s'\'etend aux chemins $C^0$~? La r\'eponse est la suivante~: \begin{Proposition}\label{plate} Soient $G$, $\xi$ et $A$ comme dans la proposition \ref{trans}. Alors, la repr\'esentation $w_A\in{\cal R} ({\bf D}(X),\xi)$ induite par le transport parall\`ele de $A$ s'\'etend en un repr\'esentation de ${\bf Ch}(X)$ sur $\xi$ si et seulement si $A$ est une connexion plate. \end{Proposition} \noindent {\sc Preuve: \ } Par d\'efinition, $A$ est plate si et seulement si et seulement si $E$ est feuillet\'ee en vari\'et\'es horizontales \cite[II.9]{KN}. Il est clair que dans ce cas le transport parall\`ele de $A$ s'\'etend en un repr\'esentation $\bar w_A\in{\cal R}({\bf Ch}(X),\xi)$. R\'eciproquement, si une telle extension existe, comme le groupe de Lie $G$ n'a pas de petits sous-groupes, la proposition \ref{petitssgr} assure que $\bar w_A$ se factorise par le groupo\"\i de \ fondamental $\Pi(X)$. Le th\'eor\`eme de r\'eduction classique \cite[II, th. 7.1]{KN} montre qu'alors le fibr\'e $\xi_{\mid U} : p^{-1}(U)\to U$, au dessus de tout ouvert contractile $U$ de $X$, admet une r\'eduction de son groupe structural au groupe trivial et que $A$ restreinte \`a $p^{-1}(U)$ est plate. \cqfd \subsection{Th\'eorie de jauge sur graphes} \label{SP:lattice} Rappelons qu'un {\it graphe} (non-orient\'e) $\Gamma$ consiste en une paire d'ensembles $(S(\Gamma),A(\Gamma))$ (sommets et ar\^etes) avec deux applications $\alpha,\beta : A(\Gamma)\to S(\Gamma)$ et une involution $a\mapsto \bar a$ sur $A$ telle que $\alpha(\bar a) = \beta(a)$ et $\beta(\bar a)=\alpha(a)$. Par exemple, pour $n\in{\bf N}$, le graphe $[n]$ se d\'efinit par $S([n]):=\{0,1,\dots ,n\}$, $A([n]):=\{(i,j)\mid |i-j|=1\}$, $\alpha(i,j):=j$, $\beta(i,j):=i$ et $\overline{(i,j)}:=(j,i)$. Un {\it chemin} (de longueur $n$) dans $\Gamma$ est un morphisme de graphes de $[n]$ dans $\Gamma$. L'ensemble des chemins dans $\Gamma$ forment un $S(\Gamma)$-pr\'egroupo\"\i de \ $\widetilde {C_\Gamma}$, muni de la topologie discr\`ete. Les applications source et but, la composition et l'anti-involution sont d\'efinies comme pour les chemins de Moore dans un espace topologique (voir \S \ref{SP:groch}). Observons que $A(\Gamma)$ s'identifie naturellement au sous-ensemble de $\widetilde C_\Gamma$ form\'e des chemins de longueur 1. Le $S(\Gamma)$-groupo\"\i de associ\'e par la proposition \ref{predefgro} sera not\'e ${\bf C}(\Gamma)$. Il est \'egalement discret et s'identifie au groupo\"\i de \ fondamental de la r\'ealisation g\'eom\'etrique $|\Gamma |$ de $\Gamma$. De m\^eme, $\Omega_{{\bf C}(\Gamma)}$ s'identifie au groupe fondamental $\pi_1(|\gamma |,*)$. Nous supposerons que $|\Gamma |$ est connexe. Ce qui s'appelle en anglais un "$G$-valued lattice gauge field" sur $\Gamma$ correspond \`a un $G$-fibr\'e $\xi$ sur $S(\Gamma)$ muni d'une repr\'esentation $w\in{\cal R}({\bf C}(\Gamma),\xi)$. Comme $S(\Gamma)$ est discret, le fibr\'e $\xi$ est trivial. Une trivialisation peut en \^etre obtenue \`a \`a l'aide de $w$ en choisissant un arbre maximal $T$ dans $\Gamma$. En effet, $T$ donne une ${\bf C}(\Gamma)$-contraction sur tout $S(\Gamma)$. En fixant une trivialisation $E(\xi)=S(\Gamma)\times G$ de $\xi$, la repr\'esentation $w$ est d\'etermin\'ee par la donn\'ee d'une application $w_1: A(\Gamma)\to G$ telle que $w_1(\bar a)=w_1(a)^{-1}$. La formule reliant $w$ \`a $w_1$ est la suivante~: \begin{equation} w(a,(\alpha(a),g))=(\beta(a),w_1(a)\,g) \ , \ a\in A(\Gamma). \end{equation} Dans la litt\'erature sur le sujet, l'application $w_1$ est prise comme d\'efinition d'un "$G$-valued lattice gauge field" (\cite[Ch. 7]{Cr}, \cite[\S\ 3]{PS}). Si $\gamma$ est fini, le groupe $\pi_1(|\gamma |,*)$ est libre de rang $1-\chi(|\Gamma |)$, o\`u $\chi(|\Gamma |)$ est la caract\'eristique d'Euler de $|\Gamma |$. Le th\'eor\`emes B et C donnent ainsi des hom\'eomorphismes \begin{equation} {\cal R} ({\bf C}(\Gamma),\xi)/{\cal G}_1 \approx {\cal R} (\pi_1(|\gamma |,*),G) \approx G^{1-\chi(|\Gamma |)}. \end{equation} et \begin{equation}\label{gau2} {\cal R} ({\bf C}(\Gamma),\xi)/{\cal G} \approx G^{1-\chi(|\Gamma |)}\big/\hbox{\scriptsize conjugaison} . \end{equation} Sous certaines hypoth\`eses, la donn\'ee de $(\xi,w)$ d\'etermine un $G$-fibr\'e principal $\xi_w$ sur $|\Gamma |$ qui n'est pas trivial (voir, par exemple, \cite{PS}). Par \eqref{gau2}, on obtient une partition de $G^{1-\chi(|\Gamma |)}\big/\hbox{\scriptsize conjugaison}$ en fonction des classes d'isomorphisme de $\xi_w$ qu'il serait int\'eressant d'\'etudier. \subsection{Fibr\'es \'equivariants} \label{SP:fequi} Soit $\Gamma$ un groupe topologique agissant \`a gauche sur $X$. Le graphe de l'action $$C:=\{(y,\gamma,x)\in X\times\Gamma\times X\mid y=\gamma x\}$$ est un $X$-groupo\"\i de \ par les donn\'ees suivantes~: $\alpha(y,\gamma,x):=x$, $\beta(y,\gamma,x):=y$, $i_x:=(x,e,x)$, $(z,\gamma_2,y)(y,\gamma_1,x):=(z,\gamma_2\gamma_1,x)$ et $(y,\gamma,x)^{-1}:=(x,\gamma^{-1},y)$. Le groupe $\Omega_{C}$ est banalement isomorphe au groupe d'isotropie $\Gamma^{\{*\}}$ de $*$. Le $X$-groupo\"\i de \ $C$ est localement trivial si l'action de $\Gamma$ sur $X$ est transitive et si l'application $q:\Gamma\to X$ donn\'ee par $\gamma\mapsto \gamma\, *$ admet des sections locales continues. Cette application sera alors un $\Gamma^{\{*\}}$ (le fibr\'e $\xi_C$ du th\'eor\`eme A). On sait que cette situation se produit si, par exemple, $X$ est le quotient $\Gamma/\Gamma_0$ d'un groupe de Lie $\Gamma$ par un sous-groupe ferm\'e $\Gamma_0$ \cite[\S\ 7.5]{St}. Soit $\xi: E\hfl{p}{} X$ un $G$-fibr\'e principal sur $X$. Une repr\'esentation de $C$ sur $\xi$ est simplement une action \`a gauche de $\Gamma$ sur $E$ telle que la projection $p:E\to X$ soit \'equivariante. On parle de $G$-fibr\'e principal $\Gamma$-\'equivariant ou d'action de $\Gamma$ sur $\xi$. L'espace ${\cal R} (C,\xi)$ classe ces $\Gamma$-actions sur $\xi$ \`a conjugaison par une transformation de jauge pr\`es. Par le th\'eor\`eme B, ${\cal R} (C,\xi) \approx {\cal R} (\Gamma^{\{*\}},G)_\xi$. Remarquons que $B\times_{\Gamma^{\{*\}}} E\Gamma \approx B\Gamma^{\{*\}}$; dans le cas o\`u $G$ est ab\'elien, on retrouve ainsi le th\'eor\`eme A de [LMS]. Les relations avec d'autres approches de fibr\'es \'equivariants, comme par exemple [BH], restent \`a \'etudier. Voici quelques exemples : \begin{ccote} $\Gamma = SO(3)$ agissant sur $X=S^2$ et $G=S^1$. \rm On a $\Omega_{C}=S^1$ et le fibr\'e principal $\xi_{C}$ du th\'eor\`eme A est le fibr\'e tangent unitaire \`a $S^2$, dont la classe d'Euler est 2. Par le th\'eor\`eme d'existence, un $S^1$-fibr\'e principal sur $S^2$ admettra une $SO(3)$-action si et seulement si sa classe d'Euler est paire. Dans ce cas, l'espace ${\cal R} (S^1,S^1)$ \'etant discret (hom\'eomorphe \`a ${\bf Z}$ par le degr\'e), le th\'eor\`eme B implique qu'il y a exactement une classe d'action de $SO(3)$ sur $\xi$ \`a conjugaison par une transformation de jauge pr\`es. Observons qu'il n'y a aucun choix possible pour l'action sur la fibre au dessus de $*$~; cette action est d\'etermin\'ee par $\xi$. \end{ccote} \begin{ccote} $\Gamma = SU(2)$ agissant sur $X=S^2$ et $G=S^1$. \rm Le fibr\'e $\xi_{C}$ du th\'eor\`eme A est alors le fibr\'e de Hopf $S^3\to S^2$. On en d\'eduit que tout $S^1$-fibr\'e principal sur $S^2$ admet une $SU(2)$-action unique \`a conjugaison par une transformation de jauge pr\`es. \end{ccote} \begin{ccote} $\Gamma = SU(2)$ agissant sur $X=S^2$ et $G=SO(3)$. \rm Il y a deux $SO(3)$-fibr\'es principaux $\xi_0$ et $\xi_1$ sur $S^2$, $\xi_0$ \'etant le fibr\'e trivial. Tout deux associ\'es au fibr\'e de Hopf, ils admettent des $SU(2)$-actions mais seul $\xi_0$ admet des $SO(3)$-actions. Un homomorphisme continu de $S^1$ dans $SO(3)$ est diff\'erentiable, donc un \'el\'ement de ${\cal R}(S^1,SO(3))$ est un sous-groupe \`a un param\`etre dont l'image est un tore maximal de $SO(3)$. On en d\'eduit que si l'on identifie l'alg\`ebre de Lie $so(3)$ \`a ${\bf R}^3$ (quaternions purs), l'espace ${\cal R}(S^1,SO(3))$ est la r\'eunion des sph\`eres de rayons entiers $n=0,1,2,\dots$. Donc, ${\cal R}(C,\xi_0)/{\cal G}_1$ est hom\'eomorphe \`a la r\'eunion des 2-sph\`eres de rayons $2n$ ($n=0,1,2,\dots$) et ${\cal R}(C,\xi_1)/{\cal G}_1$ \`a celles de rayon $2n+1$. Quant \`a ${\cal R}(C,\xi_i)/{\cal G}$, ils sont tout deux discrets d\'enombrables. \end{ccote} \vskip .3 truecm\goodbreak
\section{Introduction} \label{sec-intro} It is a surprising fact that a series of at first sight unrelated phenomena in mathematics and physics are governed by the scheme of A-D-E Dynkin diagrams, such as simple Lie algebras, finite subgroups of ${\it{SL}}(2;\bbC)$, simple singularities of complex surfaces, quivers of finite type, modular invariant partition functions of ${\it{SU}}(2)$ WZW models and subfactors of Jones index less than four. Though a good understanding of the interrelations has not yet been achieved, this coincidence indicates that there are deep connections between these different fields which even seem to go beyond the A-D-E governed cases, e.g.\ finite subgroups of ${\it{SL}}(n;\bbC)$, modular invariants of ${\it{SU}}(n)$ WZW models, or (certain) ${\it{SU}}(n)_k$ subfactors of larger index. This paper is addressed to the relation between the (classifications of) modular invariants in conformal field theory and subfactors in operator algebras. In rational (chiral) conformal field theory one deals with a chiral algebra which possesses a certain finite spectrum of representations (or superselection sectors) $\pi_\lambda$ acting on a Hilbert space ${\cal{H}}_\lambda$. Its characters $\chi_\lambda(\tau)={\rm{tr}}_{{\cal{H}}_\lambda}({\rm{e}}^{2\pi{\rm{i}}\tau(L_0-c/24)})$, $\rm{Im}(\tau)>0$, $L_0$ being the conformal Hamiltonian and $c$ the central charge, transform unitarily under ``reparametrization of the torus'', i.e.\ there are matrices $S$ and $T$ such that \[ \chi_\lambda(- 1/\tau) = \sum_\mu S_{\lambda,\mu} \chi_\mu(\tau)\,, \qquad \chi_\lambda(\tau+1) = \sum_\mu T_{\lambda,\mu} \chi_\mu(\tau)\,, \] which are the generators of a unitary representation of the (double cover of the) modular group ${\it{SL}}(2;\bbZ)$ in which $T$ is diagonal.\footnote{More precisely, for current algebras the characters depend also on other variables than $\tau$, corresponding to Cartan subalgebra generators which are omitted here for simplicity. But these variables are responsible that one is in general dealing with the whole group ${\it{SL}}(2;\bbZ)$ rather than ${\it{PSL}}(2;\bbZ)$.} In order to classify conformal field theories, in particular extensions in a certain sense of a given theory, one searches for modular invariant partition functions $Z(\tau)=Z(-1/\tau)=Z(\tau+1)$ of the form \[ Z(\tau)= \sum_{\lambda,\mu} Z_{\lambda,\mu} \chi_\lambda(\tau) \chi_\mu(\tau)^* \,,\] where \begin{equation} Z_{\lambda,\mu} = 0,1,2,\ldots \,, \qquad\qquad Z_{0,0}=1 \,. \label{massmatZ} \end{equation} Here the label ``0'' refers to the ``vacuum'' representation, and the condition $Z_{0,0}=1$ reflects the physical concept of uniqueness of the vacuum state. The matrix $Z$ arising this way is called a modular invariant mass matrix. Mathematically speaking, the problem can be rephrased like this: Find all the matrices $Z$ in the commutant of the unitary representation of ${\it{SL}}(2;\bbZ)$ defined by $S$ and $T$ subject to the conditions in \erf{massmatZ}. In this paper we study this mathematical problem in the subfactor context. We start with a von Neumann algebra, more precisely a factor $N$ endowed with a system of braided endomorphisms. Such a braiding defines matrices $S$ and $T$ which provide a unitary representation of ${\it{SL}}(2;\bbZ)$ if it is non-degenerate. We then study embeddings $N\subset M$ in larger factors $M$ which are in a certain sense compatible with the braided system of endomorphisms. We show that such an embedding $N\subset M$ determines a modular invariant mass matrix in exactly the sense specified above. Longo and Rehren have studied nets of subfactors and defined a useful formula to extend a localized transportable endomorphism of the smaller to the larger observable algebra, realizing a suggestion in \cite{R}. Xu \cite{X1,X2} has worked on essentially the same construction applied to subfactors arising from conformal inclusions with the loop group construction of A. Wassermann \cite{W2}. Two of us systematically analyzed the Longo-Rehren extension for nets of subfactors on $S^1$ \cite{BE1,BE3}. As sectors, a reciprocity between extension and restriction of localized transportable endomorphisms was established, analogous to the induction-restriction machinery of group representations, and therefore the extension was called $\alpha$-induction in order to avoid confusion with the different sector induction. It was also noticed in \cite{BE1} that the extended endomorphisms leave local algebras invariant and hence $\alpha$-induction can also be considered as a map which takes certain endomorphisms of a local subfactor to endomorphisms of the embedding factor. This theory was applied to nets arising from conformal field theory models in \cite{BE2,BE3}, and it was shown that for all type I modular invariants of ${\it{SU}}(2)$ respectively ${\it{SU}}(3)$ there are associated nets of subfactors and in turn $\alpha$-induction gives rise to fusion graphs. In fact it was shown that that these graphs are the A-D-E Dynkin diagrams respectively their generalizations of \cite{DZ1,DZ2}, and this is no accident: The homomorphism property of $\alpha$-induction relates the spectrum of the fusion graphs to the non-zero diagonal entries of the modular invariant mass matrix. A few months after the work of Longo-Rehren, Ocneanu presented his theory of ``quantum symmetries'' of Coxeter graphs and gave lectures \cite{O7} one year later. He introduced a notion of a ``double triangle algebra'' and defined elements $p_j^\pm$ which we refer to as ``chiral generators'' as they were not specifically named there. Ocneanu's analysis has much in common with work of Xu \cite{X1} and two of us \cite{BE2,BE3} about subfactors of type E$_6$, E$_8$ and D$_{{\rm{even}}}$. The reason for this is that the same structures are studied from different viewpoints, as we will outline in this paper. We start with a fairly general setting which admits both constructions, $\alpha$-induction as well as Ocneanu's double triangle algebras and chiral generators. Namely, we consider a type III subfactor $N\subset M$ of finite index with a finite system of $N$-$N$ morphisms which includes the irreducible constituents of the dual canonical endomorphism. (A ``system of morphisms'' means essentially that, as sectors, the morphisms form a closed algebra under the sector ``fusion'' product, see Definition \ref{system} below.) Therefore the subfactor is in particular forced to have finite depth. The inclusion structure associates to the $N$-$N$ system automatically $N$-$M$, $M$-$N$ and $M$-$M$ systems. The typical situation is that the system of $M$-$M$ morphisms is the ``unknown part'' of the theory. As an easy reformulation of Ocneanu's idea from his work on Goodman-de la Harpe-Jones subfactors associated with Dynkin diagrams one can define the double triangle algebra for such a setting, and it provides a powerful tool to gain information about the ``unknown part'' from the ``known part'' of the theory. Namely, the double triangle algebra is a direct sum of intertwiner spaces equipped with two different product structures, and its center ${\cal{Z}}_h$ with respect to the ``horizontal product'' turns out to be isomorphic to the (in general non-commutative) fusion rule algebra of the $M$-$M$ system when endowed with the ``vertical product''. This kind of duality is the subfactor analogue to the group algebra with its pointwise and convolution products. Under the assumption that the $N$-$N$ system is braided there is automatically the notion of $\alpha$-induction, which extends $N$-$N$ to (possibly reducible) $M$-$M$ morphisms. (This notion does not even depend on the finite depth condition.) The braiding provides powerful tools to analyze the structure of the center ${\cal{Z}}_h$ at the same time, and the analysis is most conveniently carried out with a graphical intertwiner calculus which will be explained in detail in this paper. Besides the standard ``braiding fusion symmetries'' for wire diagrams representing intertwiners of the braided $N$-$N$ morphisms, we show that the theory of $\alpha$-induction gives rise naturally to an extended symmetry which we call ``intertwining braiding fusion relations''. This reduces all graphical manipulations representing the relations between intertwiners to easily visible purely topological moves, and it allows us to work without the ``sliding moves along walls'' involving ``quantum $6j$-symbols for subfactors'' which are the main technical tool in \cite{O7}. With a braiding on the $N$-$N$ system we can define chiral generators $p_\lambda^\pm$ in the center ${\cal{Z}}_h$, and our notion essentially coincides with Ocneanu's definition of elements $p_j^\pm$ given graphically in his A-D-E setup. We show that the decomposition of the $p_\lambda^\pm$'s into minimal central projections in ${\cal{Z}}_h$ corresponds exactly to the sector decomposition of the $\alpha$-induced sectors $[\alpha_\lambda^\pm]$, and therefore they can be naturally identified. As shown by Rehren \cite{R0}, a system of braided endomorphisms gives rise to S- and T-matrices which provide a unitary representation of the modular group ${\it{SL}}(2;\bbZ)$ whenever the braiding is non-degenerate. (Relations between modular S- and T-matrices and braiding data are also discussed in \cite{MS3,FG,FRS2}.) In terms of $\alpha$-induction we define a matrix $Z$ with entries $Z_{\lambda,\mu}=\langle \alpha_\lambda^+,\alpha_\mu^- \rangle$ for $N$-$N$ morphisms $\lambda,\mu$, where the brackets denote the dimension of the intertwiner space ${\rm{Hom}}(\alpha_\lambda^+,\alpha_\mu^-)$. As it corresponds to the ``vacuum'' in physical applications, we use the label ``0'' for the identity morphism ${\rm{id}}_N$, and hence our matrix $Z$ satisfies the conditions in \erf{massmatZ}, where now $Z_{0,0}=1$ is just the factor property of $M$. We show that $Z$ commutes with $S$ and $T$ and therefore $Z$ is a ``modular invariant mass matrix'' in the sense of conformal field theory if the braiding is non-degenerate. In fact, the non-degenerate case is the most interesting one, as in the ${\it{SU}}(n)_k$ examples in conformal field theory. We apply an argument of Ocneanu to our situation to show that in that case, due to the identification with chiral generators, both kinds of $\alpha$-induction together generate the whole $M$-$M$ fusion rule algebra. Moreover, the essential information about its representation theory (or equivalently, about the decomposition of the center ${\cal{Z}}_h$ with the vertical product into simple matrix algebras) is then encoded in the mass matrix $Z$: We show that the irreducible representations of the $M$-$M$ fusion rule algebra are labelled by pairs $\lambda,\mu$ with $Z_{\lambda,\mu}\neq 0$, and that their dimensions are given exactly by the number $Z_{\lambda,\mu}$. Consequently, the $M$-$M$ fusion rules are then commutative if and only if all $Z_{\lambda,\mu}\in\{0,1\}$. An analogous result has been claimed by Ocneanu for his A-D-E setting related to the modular invariant mass matrices of the ${\it{SU}}(2)$ WZW models of \cite{CIZ,Kt}. He has his own geometric construction of modular invariants sketched in the lectures but not included in the lecture notes \cite{O7}. Our construction is different and based on the results of \cite{BE3}, and it shows that the structural results do not depend on the very special properties of Dynkin diagrams and hold in a far more general context. We also analyze the representation of the $M$-$M$ fusion rule algebra arising from its left action on $M$-$N$ sectors. As corollaries of our analysis we find that the number of $N$-$M$ (or $M$-$N$) morphisms is given by the trace ${\rm{tr}}(Z)$, whereas the number of $M$-$M$ morphisms is given by ${\rm{tr}}(Z \,{}^{{\rm t}} \! Z)$. In a forthcoming publication we will further analyze and apply our construction to subfactors constructed by means of the level $k$ positive energy representations of the ${\it{SU}}(n)$ loop group theory. For these examples, the braiding is always non-degenerate and, moreover, the S- and T-matrices are the modular matrices performing the character transformations of the corresponding ${\it{SU}}(n)_k$ WZW theory. Therefore the construction of braided subfactors\footnote{We remark that our short-hand notion of a ``braided subfactor'' meaning a subfactor for which Assumptions \ref{set-fin} and \ref{set-braid} below hold is different from the notion used in \cite{L2.5}.} for these models yields non-diagonal modular invariants $Z$. E.g.\ for ${\it{SU}}(2)_k$ one can construct the subfactors in terms of local loop groups which recover the A-D-E modular invariants of \cite{CIZ,Kt}. In our setting also the ``type II'' or ``non-blockdiagonal'' invariants can be treated by dropping the chiral locality condition. (The chiral locality condition, expressing local commutativity of the extended chiral theory in the formulation of nets of subfactors \cite{LR}, implies ``$\alpha\sigma$-reciprocity'' \cite{BE1} which in turn forces the modular invariant to be of type I. Detailed explanation and non-local examples will be provided in \cite{BEK2}.) Thus this paper extends the known results on conformal inclusions \cite{X1,X2,BE2,BE3} and simple current extensions \cite{BE2,BE3} of ${\it{SU}}(n)_k$, and it generally relates (the classification of) modular invariants to (non-degenerately) braided subfactors. Furthermore our results prove two conjectures by two of us \cite[Conj.\ 7.1 \& 7.2]{BE3}. This paper is organized as follows. In Sect.\ \ref{sec-morsecbraid} we review some basic facts about morphisms, intertwiners, sectors and braidings, and we reformulate Rehren's result about S- and T-matrices arising from superselection sectors in our context of braided factors. In Sect.\ \ref{sec-graphcal} we establish the graphical methods for the intertwiner calculus we use in this paper. The abstract mathematical structure underlying the basic graphical calculus (Subsect.\ \ref{sec-basgraphcal}) is ``strict monoidal $C^*$-categories'' \cite{DR}. Graphical methods for calculations involving fusion and braiding have been used in various publications, see e.g.\ \cite{MS1,KR,Wi,FK,FG,Kf,J2}. However, for our purposes it turns out to be extremely important to handle normalization factors with special care, and to the best of our knowledge, a comprehensive exposition which applies to our framework has not been published somewhere. So we work out a ``rotation covariant'' intertwiner calculus here, based on a formulation of Frobenius reciprocity by Izumi \cite{I2}. We then define $\alpha$-induction for braided subfactors and use it to extend our graphical calculus conveniently. In Sect.\ \ref{sec-dta} we present the double triangle algebra and analyze its properties. In Sect.\ \ref{sec-aicpmodinv} we present our version of Ocneanu's graphical notion of chiral generators, and we show that it can be naturally identified with the $\alpha$-induced sectors. We then define the ``mass matrix'' $Z$ and show that it commutes with the S- and T-matrices of the $N$-$N$ system. Assuming now that the braiding is non-degenerate, we show that the $M$-$M$ fusion rule algebra is generated by the images of the two kinds ($+$ and $-$) of $\alpha$-induction. In Sect.\ \ref{sec-repMM} we decompose ${\cal{Z}}_h$ with the vertical product into simple matrix algebras which is equivalent to the determination of all the irreducible representations of the $M$-$M$ fusion rule algebra, and we show that their dimensions are given by the entries of the modular invariant mass matrix. Then we analyze the representation arising from the left action on $M$-$N$ sectors. In Sect.\ \ref{sec-concl} we finally conclude this paper with general remarks and comments and an outlook to the applications to subfactors arising from conformal field theory which will be treated in \cite{BEK2}. \section{Preliminaries} \label{sec-morsecbraid} \subsection{Morphisms and sectors} \label{sec-morsec} For our purposes it turns out to be convenient to make use of the formulation of sectors between different factors. We follow here (up to minor notational changes) Izumi's presentation \cite{I2,I3} based on Longo's sector theory \cite{L2}. Let $A$, $B$ be infinite factors. We denote by ${\rm{Mor}}(A,B)$ the set of unital $\ast$-homomorphisms from $A$ to $B$. We also denote ${\rm{End}}(A)={\rm{Mor}}(A,A)$, the set of unital $\ast$-endomorphisms. For $\rho\in{\rm{Mor}}(A,B)$ we define the statistical dimension $d_\rho=[B:\rho(A)]^{1/2}$, where $[B:\rho(A)]$ is the minimal index \cite{J,Ko}. A morphism $\rho\in{\rm{Mor}}(A,B)$ is called irreducible if the subfactor $\rho(A)\subset B$ is irreducible, i.e.\ if the relative commutant $\rho(A)'\cap B$ consists only of scalar multiples of the identity in $B$. Two morphisms $\rho,\rho'\in{\rm{Mor}}(A,B)$ are called equivalent if there exists a unitary $u\in B$ such that $\rho'(a)=u\rho(a)u^*$ for all $a\in A$. We denote by ${\rm{Sect}}(A,B)$ the quotient of ${\rm{Mor}}(A,B)$ by unitary equivalence, and we call its elements $B$-$A$ sectors. Similar to the case $A=B$, ${\rm{Sect}}(A,B)$ has the natural operations, sums and products: For $\rho_1,\rho_2\in{\rm{Mor}}(A,B)$ choose generators $t_1,t_2\in B$ of a Cuntz algebra ${\cal{O}}_2$, i.e.\ such that $t_i^*t_j=\del ij {\bf1}$ and $t_1t_1^*+t_2t_2^*={\bf1}$. Define $\rho\in{\rm{Mor}}(A,B)$ by putting $\rho(a)=t_1\rho_1(a)t_1^*+t_2\rho_2(a)t_2^*$ for all $a\in A$, and then the sum of sectors is defined as $[\rho_1]\oplus[\rho_2]=[\rho]$. The product of sectors comes from the composition of endomorphisms, $[\rho_1][\rho_2]=[\rho_1\circ\rho_2]$. We often omit the composition symbol ``$\circ$'', so $[\rho_1][\rho_2]=[\rho_1\rho_2]$. The statistical dimension is an invariant for sectors (i.e.\ equivalent morphisms have equal dimension) and is additive and multiplicative with respect to these operations. Moreover, for $[\rho]\in{\rm{Sect}}(A,B)$ there is a unique conjugate sector $\overline{[\rho]}\in{\rm{Sect}}(B,A)$ such that, if $[\rho]$ is irreducible, $\overline{[\rho]}$ is irreducible as well and $\overline{[\rho]}\times[\rho]$ contains the identity sector $[{\rm{id}}_A]$ and $[\rho]\times\overline{[\rho]}$ contains $[{\rm{id}}_B]$ precisely once. We choose a representative endomorphism of $\overline{[\rho]}$ and denote it naturally by $\co\rho$, thus $[\co\rho]=\overline{[\rho]}$. For conjugates we have $d_{\co\rho}=d_\rho$. As for bimodules one may decorate $B$-$A$ sectors $[\rho]$ with suffixes, ${}_B[\rho]_A$, and then we can multiply ${}_B[\rho]_A \times {}_A[\sigma]_B$ but not, for instance, ${}_B[\rho]_A$ with itself. For $\rho,\tau\in{\rm{Mor}}(A,B)$ we denote \[ {\rm{Hom}} (\rho,\tau) = \{ t\in B : t\, \rho(a) = \tau (a) \, t \,,\,\,\, a\in A \} \] and \[ \langle \rho, \tau \rangle = {\rm{dim}}\,{\rm{Hom}} (\rho,\tau) \,.\] If $[\rho]=[\rho_1]\oplus[\rho_2]$ then \[ \langle \rho, \tau \rangle = \langle \rho_1, \tau \rangle + \langle \rho_2, \tau \rangle \,. \] Note that if $\rho$ is irreducible then for $t,t'\in{\rm{Hom}} (\rho,\tau)$ it follows that $t^*t'$ is a scalar and then putting \begin{equation} t^*t'=\langle t,t' \rangle {\bf1}_B \label{inprint} \end{equation} defines an inner product on ${\rm{Hom}} (\rho,\tau)$. One often calls ${\rm{Hom}} (\rho,\tau)$ a ``Hilbert space of isometries'' in this case. If $\rho\in{\rm{Mor}}(A,B)$ with $d_\rho<\infty$ then $\co\rho\in{\rm{Mor}}(B,A)$ is a conjugate if there are isometries $r_\rho\in{\rm{Hom}}({\rm{id}}_A,\co\rho\rho)$ and ${\co r}_\rho\in{\rm{Hom}}({\rm{id}}_B,\rho\co\rho)$ such that \[ \rho(r_\rho)^* {\co r}_\rho = d_\rho^{-1} {\bf1}_B \qquad \mbox{and} \qquad \co\rho ({\co r}_\rho)^* r_\rho = d_\rho^{-1}{\bf1}_A \,, \] and in the case that $\rho$ is irreducible such isometries $r_\rho$ and ${\co r}_\rho$ are unique up to a common phase. If $C$ is another factor and $\sigma\in{\rm{Mor}}(C,A)$ and $\tau\in{\rm{Mor}}(C,B)$ are morphisms with finite statistical dimensions $d_\sigma,d_\tau<\infty$, and conjugate morphisms $\co\sigma\in{\rm{Mor}}(A,C)$, $\co\tau\in{\rm{Mor}}(B,C)$, respectively, then the ``left and right Frobenius reciprocity maps'', \[ \begin{array}{ll} {\cal{L}}_\rho: {\rm{Hom}}(\tau,\rho\sigma) \longrightarrow {\rm{Hom}}(\sigma,\co\rho\tau)\,,\qquad & t \longmapsto \sqrt{\displaystyle\frac{d_\rho d_\sigma}{d_\tau}} \, \co\rho(t)^* r_\rho \,,\\[.8em] {\cal{R}}_\rho: {\rm{Hom}}(\co\sigma,\co\tau\rho) \longrightarrow {\rm{Hom}}(\co\tau,\co\sigma\co\rho)\,,& s \longmapsto \sqrt{\displaystyle\frac{d_\rho d_\tau}{d_\sigma}} \, s^*\co\tau({\co r}_\rho)\,, \end{array} \] are anti-linear (vector space) isomorphisms with inverses \[ \begin{array}{ll} {\cal{L}}_\rho^{-1}: {\rm{Hom}}(\sigma,\co\rho\tau) \longrightarrow {\rm{Hom}}(\tau,\rho\sigma)\,,\qquad & x \longmapsto \sqrt{\displaystyle\frac{d_\rho d_\tau}{d_\sigma}} \, \rho(x)^* {\co r}_\rho \,,\\[.8em] {\cal{R}}_\rho^{-1}: {\rm{Hom}}(\co\tau,\co\sigma\co\rho) \longrightarrow {\rm{Hom}}(\co\sigma,\co\tau\rho)\,,& y \longmapsto \sqrt{\displaystyle\frac{d_\rho d_\sigma}{d_\tau}} \, y^*\co\sigma(r_\rho)\,, \end{array} \] respectively \cite{I2}. (See also \cite[Sect.\ 5]{FG} and \cite[App.\ A]{FRS2} for such formulae arising from superselection sectors.) Hence we have in particular Frobenius reciprocity \cite{I2,L3}, \[ \langle \tau,\rho\sigma \rangle = \langle \co\rho\tau, \sigma \rangle = \langle \co\rho, \sigma\co\tau \rangle = \langle \co\sigma\co\rho, \co\tau \rangle = \langle \co\sigma, \co\tau\rho \rangle = \langle \tau\co\sigma,\rho \rangle \,. \] If $\tau$ and $\sigma$ are irreducible then the Frobenius reciprocity maps are even (anti-linearly) isometric: With the inner products as in \erf{inprint} on the above intertwiner spaces we have $\langle t,t' \rangle = \langle {\cal{L}}_\rho(t'), {\cal{L}}_\rho(t) \rangle$ for $t,t'\in{\rm{Hom}}(\tau,\rho\sigma)$ and similarly $\langle s,s' \rangle = \langle {\cal{R}}_\rho(s'), {\cal{R}}_\rho(s) \rangle$ for $s,s'\in{\rm{Hom}}(\co\sigma,\co\tau\rho)$. The map $\phi_\rho:B\rightarrow A$ defined by \[ \phi_\rho(b) = r_\rho^* \, \co\rho (b) \, r_\rho \,, \qquad b\in B \] is completely positive, normal, unital $\phi_\rho({\bf1}_B)={\bf1}_A$ and satisfies \[ \phi_\rho (\rho(a_1)b\rho(a_2)) = a_1 \phi_\rho(b) a_2 \,, \qquad a_1,a_2\in A \,, \quad b\in B\,.\] The map is called the (unique) standard left inverse. The minimal conditional expectation for the subfactor $\rho(A)\subset B$ is given by $E_\rho=\rho\circ\phi_\rho$. Let now $\rho,\sigma,\tau$ as above be irreducible with standard left inverses $\phi_\rho,\phi_\sigma,\phi_\tau$, respectively, and let $t\in{\rm{Hom}}(\tau,\rho\sigma)$ be non-zero. Then $\phi_\rho (tt^*)\in{\rm{Hom}}(\sigma,\sigma)$ is a positive scalar and $\tilde{E}_\tau:B\rightarrow \tau(C)$ given by $\rho\circ\phi_\rho (tt^*)\tilde{E}_\tau(b) =\tau\circ\phi_\sigma\circ\phi_\rho (tbt^*)$ for all $b\in B$ is a conditional expectation for the subfactor $\tau(C)\subset B$. Since conditional expectations for irreducible subfactors are unique we conclude that \[ \phi_\tau(b) \, E_\rho (tt^*) = \phi_\sigma\circ\phi_\rho (tbt^*)\,,\qquad b\in B \] holds for any $t\in{\rm{Hom}}(\tau,\rho\sigma)$. Moreover, $t^*t'$ is a scalar for any $t,t'\in{\rm{Hom}}(\tau,\rho\sigma)$, $t^*t'=\langle t,t' \rangle {\bf1}_B$, and so is ${\cal{L}}_\rho(t)^*{\cal{L}}_\rho(t')$, in fact \[ \langle t,t' \rangle {\bf1}_A = \langle {\cal{L}}_\rho(t'), {\cal{L}}_\rho(t) \rangle {\bf1}_A \equiv {\cal{L}}_\rho(t')^*{\cal{L}}_\rho(t) = \frac{d_\rho d_\sigma}{d_\tau} \, r_\rho^* \co\rho(t't^*) r_\rho \] and this is \begin{equation} \phi_\rho (t't^*) = \frac{d_\tau}{d_\rho d_\sigma} \, \langle t,t' \rangle {\bf1}_A \,. \label{phiddd} \end{equation} Now let $N\subset M$ be an infinite subfactor of finite index. Let $\gamma\in{\rm{End}}(M)$ be a canonical endomorphism from $M$ into $N$ and $\theta=\gamma|_N\in{\rm{End}}(N)$. By $\iota\in{\rm{Mor}}(N,M)$ we denote the injection map, $\iota(n)=n\in M$, $n\in N$. Then $d_\iota=[M:N]^{1/2}$, and a conjugate ${\co\iota}\in{\rm{Mor}}(M,N)$ is given by ${\co\iota}(m)=\gamma(m)\in N$, $m\in M$. (These formulae could in fact be used to define the canonical and dual canonical endomorphism.) Note that $\gamma=\iota{\co\iota}$ and $\theta={\co\iota}\iota$, and there are isometries $w\equiv r_\iota\in{\rm{Hom}}({\rm{id}}_N,\theta)$ and $v\equiv {\co r}_\iota\in{\rm{Hom}}({\rm{id}}_M,\gamma)$ such that $w^*v = \gamma(v ^*)w = [M:N]^{-1/2}{\bf1}$. Moreover, we have the pointwise equality $M=Nv$, and for each $m\in M$ the decomposition $m=nv$ yields a unique element $n\in N$. Explicitly, $n=[M:N]^{1/2}w^*\gamma(m)$. Now let us consider a single factor $A$ and its sectors. For a set of irreducible sectors which is closed under conjugation and irreducible decomposition of products (a ``sector basis'' in the notation of \cite{BE1,BE2,BE3} in the case that the set is finite) it is often useful to choose one representative endomorphism for each sector. \begin{definition} \label{system} {\rm We call a subset ${\Delta}\subset{\rm{End}}(A)$ a {\sl system of endomorphisms} if it satisfies the following properties. \begin{enumerate} \item Each $\lambda\in {\Delta}$ is irreducible and has finite statistical dimension. \item Different elements in ${\Delta}$ are inequivalent, i.e.\ different as sectors. \item ${\rm{id}}_A\in {\Delta}$. \item For any $\lambda\in{\Delta}$, we have a morphism $\bar\lambda\in{\Delta}$ such that $[\bar\lambda]$ is the conjugate sector of $[\lambda]$. \item ${\Delta}$ is closed under composition and subsequent irreducible decomposition, i.e.\ for any $\lambda,\mu\in{\Delta}$ we have non-negative integers $N_{\lambda,\mu}^\nu$ with $[\lambda][\mu]=\sum_{\nu\in{\Delta}} N_{\lambda,\mu}^\nu [\nu]$ as sectors. \end{enumerate} }\end{definition} Note that we do not assume finiteness of ${\Delta}$ in this definition. The numbers $N_{\lambda\mu}^\nu=\langle\lambda\mu,\nu\rangle$ are called fusion coefficients. Frobenius reciprocity now reads $N_{\lambda,\mu}^\nu=N_{\co\lambda,\nu}^\mu=N_{\nu,\co\mu}^\lambda$, and associativity of the sector product yields $\sum_{\mu\in{\Delta}} N_{\lambda,\mu}^\nu N_{\rho,\sigma}^\mu =\sum_{\tau\in{\Delta}} N_{\lambda,\rho}^\tau N_{\tau,\sigma}^\nu$. The additivity and multiplicativity of the statistical dimension with respect to sector sums and products implies $\sum_{\nu\in{\Delta}} N_{\lambda,\mu}^\nu d_\nu = d_\lambda d_\mu$, $\lambda,\mu,\nu\in{\Delta}$. Defining matrices $N_\mu$ with entries $(N_\mu)_{\lambda,\nu}=N_{\lambda,\mu}^\nu$ gives $N_{\co\mu}$ as the transpose of $N_\mu$ and defines the ``regular representation'' of the sector products, $N_\lambda N_\mu = \sum_{\nu\in{\Delta}} N_{\lambda,\mu}^\nu N_\nu$, and the statistical dimension can be regarded as a one-dimensional representation or as a simultaneous eigenvector of all matrices $N_\mu$ with eigenvalues $d_\mu$ ($\lambda,\mu,\nu\in{\Delta}$). \subsection{Braided endomorphisms} \label{braid-prelim} Let $A$ again be an infinite factor and ${\Delta}$ a system of endomorphisms of $A$. In general the sector products are not commutative. If the sectors commute, then a ``systematic choice of unitary intertwiners'' in each space ${\rm{Hom}}(\lambda\mu,\mu\lambda)$, $\lambda,\mu\in{\Delta}$, is called a braiding (which need not exist in general). To be more precise, we give the following \begin{definition} \label{braid}{\rm We say that a system ${\Delta}$ of endomorphisms is {\sl braided} if for any pair $\lambda,\mu\in{\Delta}$ there is a unitary operator $\eps\lambda\mu\in{\rm{Hom}}(\lambda\mu,\mu\lambda)$ subject to initial conditions \begin{equation} \eps {{\rm{id}}_A}\mu=\eps\lambda{{\rm{id}}_A}={\bf1} \,, \label{ini} \end{equation} and whenever $t\in {\rm{Hom}} (\lambda,\mu\nu)$ we have the braiding fusion equations (BFE's) \begin{equation} \begin{array}{rl} \rho(t) \, \eps \lambda \rho &=\,\, \eps \mu \rho \, \mu (\eps \nu \rho) \, t \,, \\[.4em] t \, \eps \rho \lambda &=\,\, \mu (\eps \rho \nu) \, \eps \rho \mu \, \rho (t) \,, \\[.4em] \rho(t)^* \, \eps \mu\rho \, \mu( \eps \nu\rho ) &=\,\, \eps \lambda\rho \, t^* \,, \\[.4em] t^* \, \mu( \eps \rho\nu ) \, \eps \rho\mu &= \,\, \eps \rho\lambda \, \rho(t)^* \,, \end{array} \label{BFE} \end{equation} for any $\lambda,\mu,\nu\in{\Delta}$. }\end{definition} The unitaries $\eps\lambda\mu$ are called {\sl braiding operators} (or {\sl statistics operators}). Note that a braiding $\varepsilon\equiv\varepsilon^+$ always comes along with another ``opposite'' braiding $\varepsilon^-$, namely operators $\epsm \lambda\mu = (\epsp \mu\lambda )^*$, $\epsp \mu\lambda \equiv \eps \mu\lambda$, satisfy the same relations. The unitaries $\epsp \lambda\mu$ and $\epsm \lambda\mu$ are different in general but may coincide for some $\lambda$, $\mu$. Later we will also use the following notion of non-degeneracy of a braiding (cf.\ \cite{R0}). \begin{definition} \label{non-deg}{\rm We say that a braiding $\varepsilon$ on a system of endomorphisms ${\Delta}$ is {\sl non-degenerate}, if the following condition is satisfied: If some morphism $\lambda\in{\Delta}$ satisfies $\varepsilon^+(\lambda,\mu)=\varepsilon^-(\lambda,\mu)$ for all morphisms $\mu\in{\Delta}$, then we have $\lambda={\rm{id}}_A$. }\end{definition} We may also extend a given braiding from ${\Delta}$ in a well defined manner to all equivalent and sum endomorphisms as follows. We denote by $\Sigma({\Delta})$ the set of all endomorphisms $\lambda,\rho\in{\rm{End}}(A)$ given as $\lambda(a)=\sum_{i=1}^n t_i\lambda_i(a)t_i^*$ and $\rho(a)=\sum_{j=1}^m s_j\rho_j(a)s_j^*$ for all $a\in A$, where $t_i\in A$, $i=1,2\dots,n$, and $s_j\in A$, $j=1,2,\dots,m$, are Cuntz algebra generators, i.e.\ $t_i^*t_k=\del ik{\bf1}$ and $\sum_{i=1}^n t_it_i^*={\bf1}$, and similarly $s_j^*s_l=\del jl{\bf1}$ and $\sum_{j=1}^m s_js_j^*={\bf1}$, and $\lambda_i,\rho_j\in{\Delta}$. (Here $n,m\ge 1$.) For $\lambda,\rho$ as above we put \begin{equation} \eps \lambda\rho = \sum_{i=1}^n \sum_{j=1}^m s_j \rho_j (t_i) \, \eps {\lambda_i}{\rho_j} \, \lambda_i(s_j^*) t_i^* \,, \label{esum} \end{equation} and one can check that this definition is independent of the ambiguities in the choice of isometries $t_i\in{\rm{Hom}}(\lambda_i,\lambda)$ and $s_j\in{\rm{Hom}}(\rho_j,\rho)$. Note that in the case $n=m=1$ this reads \begin{equation} \eps {{\rm{Ad}}(u)\circ\lambda}{{\rm{Ad}}(q)\circ\rho} = q \rho (u) \, \eps \lambda\rho \, \lambda(q^*) u^* \label{eun} \end{equation} with some unitaries $u,q\in A$. Then for any sum endomorphisms $\lambda,\mu,\rho\in\Sigma({\Delta})$ the BFE's (\ref{BFE}) hold as well or, alternatively, we have the naturality equations \begin{equation} \rho(t) \, \eps \lambda \rho = \eps \mu \rho \, t \,, \qquad t \, \eps \rho \lambda = \eps \rho \mu \, \rho (t) \label{nat} \end{equation} whenever $t\in {\rm{Hom}}(\lambda,\mu)$. Using decompositions of products $\lambda\mu$, $\lambda,\mu\in\Sigma({\Delta})$ one can then easily show by use of the BFE's that \begin{equation} \eps {\lambda\mu}\rho = \eps \lambda\rho \, \lambda (\eps \mu\rho) \,, \qquad \eps \lambda{\mu\rho} = \mu (\eps \lambda\rho) \, \eps \lambda\mu \,. \label{comp} \end{equation} By plugging this in \erf{nat} we find that BFE's hold for endomorphisms in $\Sigma({\Delta})$ as well and \erf{nat} yields for $\eps \lambda\mu\in{\rm{Hom}}(\lambda\mu,\mu\lambda)$ the braid relation (or ``Yang-Baxter equation'') \begin{equation} \rho(\eps\lambda\mu) \, \eps\lambda\rho \, \lambda(\eps\mu\rho) = \eps\mu\rho \, \mu(\eps\lambda\rho) \, \eps\lambda\mu \,. \label{YBE} \end{equation} Now let ${\Delta}$ be a braided system of endomorphisms and let $\rho,\co\rho\in{\Delta}$ be conjugate morphisms. Denote by $r\equiv r_\rho\in{\rm{Hom}}({\rm{id}}_A,\bar\rho\rho)$ and $\co r\equiv {\co r}_\rho \in{\rm{Hom}}({\rm{id}}_A,\rho\bar\rho)$ isometries such that \[ \rho (r)^* \co r = \co\rho (\co r)^* r = d_\rho^{-1} {\bf1} \,,\] which are then unique up to a common phase.\footnote{If $\rho$ is not self-conjugate then we may choose $r_{\co\rho}={\co r}_\rho$ and ${\co r}_{\co\rho}=r_\rho$. However, if $\rho$ is self-conjugate, $\rho=\co\rho$, we do not have $r_\rho=\co r_{\rho}$ in general. This is only true for so-called ``real'' sectors, and for ``pseudo-real'' sectors we have $r_\rho=-\co r_{\rho}$.} Note that $\eps {\co\rho}\rho ^* \co r\in{\rm{Hom}}({\rm{id}}_A,\bar\rho\rho)$ is an isometry and hence $\eps {\co\rho}\rho ^* \co r = \omega_\rho r$ for some phase $\omega_\rho\in\Bbb{T}$ which is called the {\sl statistics phase} and is obviously independent of the common phase of $r$ and $\co r$. In fact $\omega_\rho$ is even independent of the choice of $\rho$ and $\co\rho$ within their sectors: If $\rho'={\rm{Ad}}\; u \circ \rho$ and $\co\rho'={\rm{Ad}}\; \co u \circ \co\rho$ for some unitaries $u,\co u\in A$, then it is easy to see that isometries $r'=\co u \co\rho(u)r\in{\rm{Hom}}({\rm{id}}_A,\bar\rho '\rho')$ and $\co r'=u\rho(\co u)\co r\in{\rm{Hom}}({\rm{id}}_A,\rho'\bar\rho')$ also fulfill $\rho (r')^* \co r' = \co\rho (\co r')^* r' = d_\rho^{-1} {\bf1}$. Now the braiding operator transforms as $\eps {\co\rho'}{\rho'} = u\rho(\co u)\eps{\co\rho}\rho \co\rho(u)^*\co u ^*$ and hence \[ \eps {\co\rho'}{\rho'} ^* \co r' = \co u\co\rho(u)\eps{\co\rho}\rho ^*\co r = \omega_\rho r' \,.\] The statistics phase can also be obtained by \[ \phi_\rho(\eps\rho\rho) = r^* \co\rho (\eps\rho\rho)r = \omega_\rho d_\rho^{-1} {\bf1} \,.\] (The number $\omega_\rho d_\rho^{-1}$ is usually called the {\sl statistics parameter}.) This is obtained from the initial condition and the BFE: \[ \rho(r) = \rho(r) \eps {{\rm{id}}_A}\rho = \eps {\co\rho}\rho \co\rho (\eps\rho\rho) r \,, \] but since $r^*\eps {\co\rho}\rho ^*=\omega_\rho\co r ^*$ we obtain \[ r^* \co\rho (\eps\rho\rho)r = r^* \eps {\co\rho}\rho ^* \rho(r) = \omega_\rho \co r ^* \rho (r) = \omega_\rho d_\rho^{-1} {\bf1} \,.\] Moreover we have $\omega_\rho=\omega_{\co\rho}$. This can be seen as follows. We have \[ r=r\eps \rho{{\rm{id}}_A} = \co\rho(\eps\rho\rho)\eps\rho{\co\rho} \rho(r) \,,\] hence $r^* \co\rho(\eps\rho\rho)=\rho(r)^*\eps\rho{\co\rho} ^*$, thus \[ \omega_\rho d_\rho^{-1} {\bf1} = \rho(r)^*\eps\rho{\co\rho} ^*r = \omega_{\co\rho}\rho(r)^* \co r = \omega_{\co\rho} d_\rho^{-1} \,, \] since $\eps \rho{\co\rho} ^* r = \omega_{\co\rho} \co r$ by definition. Therefore we have $\omega_\rho r^*=\co r^* \eps \rho{\co\rho} ^*$. Another application of the BFE yields $\eps\rho\rho \rho(\co r) = \rho(\eps\rho{\co\rho})^* \co r$, hence we have \[ \rho(\co r)^* \eps\rho\rho \rho (\co r) = \rho (\co r)^* \rho (\eps \rho{\co \rho})^* \co r = \omega_\rho \rho (r)^* \co r = \omega_\rho d_\rho^{-1} {\bf1} \,.\] Now let $\lambda,\mu,\nu\in{\Delta}$. Let $r\equiv r_\lambda\in{\rm{Hom}}({\rm{id}}_A,\co\lambda\la)$ and $\co r\equiv {\co r}_\lambda\in{\rm{Hom}}({\rm{id}}_A,\lambda\co\lambda)$ be isometries such that $\lambda(r)^*\co r=\co\lambda(\co r)^*r = d_\lambda^{-1}{\bf1}$. Let $t,t'\in{\rm{Hom}}(\lambda,\mu\nu)$. Recall that $\phi_\mu (t't^*)=d_\lambda d_\mu^{-1} d_\nu^{-1} t^*t'\in{\rm{Hom}}(\lambda,\lambda)$ is a scalar. We can now compute \[ \begin{array}{ll} \omega_\lambda d_\mu^{-1} d_\nu^{-1}\, t^*t' &= \omega_\lambda d_\lambda^{-1} \, \phi_\nu\circ\phi_\mu (t't^*) = \phi_\nu\circ\phi_\mu (t' \lambda(\co r)^* \eps \lambda\la \lambda (\co r) t^*) \\[.4em] &= \co r ^* \, \phi_\nu\circ\phi_\mu (t' \eps \lambda\la t^*) \, \co r = \co r ^* \, \phi_\nu\circ\phi_\mu ( \eps\lambda{\mu\nu} \lambda(t')t^*)\, \co r \\[.4em] &= \co r ^* \, \phi_\nu\circ\phi_\mu ( \eps\lambda{\mu\nu} t^*) \, t' \co r = \co r ^* t^* \, \phi_\nu\circ\phi_\mu ( \eps{\mu\nu}{\mu\nu}) \, t' \co r \\[.4em] &= \co r ^* t^* \, \phi_\nu\circ\phi_\mu ( \mu(\eps\mu\nu)\mu^2 (\eps\nu\nu) \eps\mu\mu \mu(\eps\nu\mu))\, t' \co r \\[.4em] &= \omega_\mu d_\mu^{-1} \, \co r ^* t^* \, \phi_\nu ( \eps\mu\nu \mu (\eps\nu\nu) \eps\nu\mu ) \, t' \co r \\[.4em] &= \omega_\mu d_\mu^{-1} \, \co r ^* t^* \, \phi_\nu ( \nu (\eps\nu\mu \eps\nu\nu \nu (\eps\mu\nu )\, t' \co r \\[.4em] &= \omega_\mu \omega_\nu d_\mu^{-1} d_\nu^{-1}\, \co r ^* t^* \, \eps\nu\mu \eps\mu\nu \, t' \co r \\[.4em] &= \omega_\mu \omega_\nu d_\mu^{-1} d_\nu^{-1}\, t^* \eps\nu\mu \eps\mu\nu \, t' \,, \end{array} \] where we finally could omit the $\co r$'s since $t^* \eps\nu\mu \eps\mu\nu t'\in{\rm{Hom}}(\lambda,\lambda)$ is a scalar. As $\eps\nu\mu \eps\mu\nu t'\in{\rm{Hom}}(\lambda,\mu\nu)$ we find $\omega_\lambda \langle t,t'\rangle = \omega_\mu \omega_\nu \langle t, \eps\nu\mu \eps\mu\nu t' \rangle$ for any $t,t'\in{\rm{Hom}}(\lambda,\mu\nu)$, and therefore we arrive at the important relation \begin{equation} \eps\nu\mu \eps\mu\nu \, t = \frac{\omega_\lambda}{\omega_\mu \omega_\nu} \, t \qquad \mbox{for all} \qquad t\in{\rm{Hom}}(\lambda,\mu\nu)\,. \label{mondiag} \end{equation} Decomposing $[\mu\nu]$ in all irreducible sectors $[\lambda]$ and choosing for each $\lambda\in{\Delta}$ some orthonormal bases of intertwiners $t_{\lambda;i}\in{\rm{Hom}}(\lambda,\mu\nu)$, $i=1,2,\dots,N_{\mu,\nu}^\lambda$, where $N_{\mu,\nu}^\lambda = \langle \lambda,\mu\nu\rangle$ as usual, we have $\sum_{\lambda\in{\Delta}}\sum_i t_{\lambda;i}t_{\lambda;i}^*={\bf1}$, and therefore we find by Eqs.\ (\ref{phiddd}) and (\ref{mondiag}), \[ \phi_\mu (\eps\nu\mu \eps\mu\nu)^* = \phi_\mu \left( \eps\nu\mu \eps\mu\nu \sum_{\lambda\in{\Delta}} \sum_i t_{\lambda;i}t_{\lambda;i}^* \right)^* = \sum_{\lambda\in{\Delta}} \frac{\omega_\mu \omega_\nu}{\omega_\lambda} N_{\mu,\nu}^\lambda \frac{d_\lambda}{d_\mu d_\nu} {\bf1} \,. \] One then defines a matrix $Y$ in terms of these numbers \cite{R0} (see also \cite{FG,FRS2}): \begin{equation} Y_{\mu,\nu} = \sum_{\lambda\in{\Delta}} \frac{\omega_\mu \omega_\nu}{\omega_\lambda} N_{\mu,\nu}^\lambda \, d_\lambda \,, \qquad \mu,\nu\in{\Delta} \,, \label{Ydef} \end{equation} i.e. $d_\mu d_\nu \phi_\mu (\eps\nu\mu \eps\mu\nu)^* = Y_{\mu,\nu}{\bf1}$. Then one has \[Y_{\lambda,\mu}=Y_{\mu,\lambda}=Y_{\co\lambda,\mu}^*=Y_{\co\lambda,\co\mu}\,.\] The first equality is obvious from \erf{Ydef}, so we only need to show $Y_{\lambda,\mu}=(Y_{\co\lambda,\mu})^*$. In fact, applying the BFE again yields $\co\lambda(\eps\lambda\mu) r_\lambda=\eps{\co\lambda}\mu ^* \mu(r_\lambda)$ and $r_\lambda^* \co\lambda (\eps\mu\lambda) = \mu(r_\lambda)^* \eps\mu{\co\lambda} ^*$. Hence \[ \begin{array}{ll} Y_{\lambda,\mu} {\bf1} &= \phi_\mu (Y_{\lambda,\mu}) = d_\lambda d_\mu (r_\mu^* \co\mu (r_\lambda^* \co\lambda (\eps\mu\lambda \eps\lambda\mu) r_\lambda)r_\mu)^* \\[.4em] &= d_\lambda d_\mu (r_\lambda^* r_\mu^* \co\mu(\eps \mu{\co\lambda} ^* \eps {\co\lambda}\mu ^*) r_\mu r_\lambda)^* = (r_\lambda ^* Y_{\co\lambda,\mu} r_\lambda)^* = (Y_{\co\lambda,\mu})^* {\bf1} \,. \end{array} \] Moreover, we have \[ Y_{\nu,\rho}Y_{\mu,\rho} =d_\rho \sum_\lambda N_{\mu,\nu}^\lambda Y_{\rho,\lambda}\,, \] since \[ \begin{array}{ll} Y_{\nu,\rho}Y_{\mu,\rho} {\bf1} &= d_\rho^2 d_\mu d_\nu \, \phi_\nu (\eps\rho\nu \phi_\mu(\eps\rho\mu\eps\mu\rho)\eps\nu\rho )^* \\[.4em] &= d_\rho^2 d_\mu d_\nu \, \phi_\nu\circ\phi_\mu (\mu(\eps\rho\nu) \eps\rho\mu \eps\mu\rho \mu(\eps\nu\rho) )^* \\[.4em] &= d_\rho^2 d_\mu d_\nu \sum_\lambda \sum_i \phi_\nu\circ\phi_\mu (\eps\rho{\mu\nu} \rho(t_{\lambda;i}t_{\lambda;i}^*) \eps{\mu\nu}\rho )^* \\[.4em] &= d_\rho^2 d_\mu d_\nu \sum_\lambda \sum_i \phi_\nu\circ\phi_\mu (t_{\lambda;i} \eps\rho\lambda \eps\lambda\rho t_{\lambda;i}^*)^* \\[.4em] &= d_\rho^2 d_\mu d_\nu \sum_\lambda \sum_i \phi_\mu (t_{\lambda;i}t_{\lambda;i}^*)^* \phi_\lambda (\eps\rho\lambda \eps\lambda\rho)^* = d_\rho \sum_\lambda N_{\mu,\nu}^\lambda Y_{\rho,\lambda} {\bf1} \,. \end{array} \] {}From now on we assume that the system ${\Delta}$ is finite. We define the complex number \[ z_{\Delta}= \sum_{\lambda\in{\Delta}} d_\lambda^2 \omega_\lambda \,, \] and if $z_{\Delta} \neq 0$ we put $c = 4 \arg (z_{\Delta})/ \pi$. Note that the $c$ is here only defined mod $8$ and we may make a choice. Let $C$ be the conjugation matrix with entries $C_{\lambda,\mu}=\del \lambda{\co\mu}$. Clearly, $C=C^*=C^{-1}$. We then have the following \begin{proposition} \label{ST0} Let ${\Delta}$ be finite system of endomorphisms with $z_{\Delta}\neq 0$. Then S- and T-matrices defined by \[ S_{\lambda,\mu} = |z_{\Delta}|^{-1} \, Y_{\lambda,\mu} \,, \qquad T_{\lambda,\mu} = {\rm{e}}^{-\pi{\rm{i}} c/12} \, \omega_\lambda \, \del \lambda\mu \,, \qquad \lambda,\mu\in{\Delta} \,, \] obey the partial Verlinde modular algebra $TSTST=S$, $CTC=T$, $CSC=S$ and $T^*T={\bf1}$. \end{proposition} To prove the proposition, we simply compute \[ \begin{array}{ll} \sum_\mu \omega_\lambda Y_{\lambda,\mu} \omega_\mu Y_{\mu,\nu} \omega_\nu &= \omega_\lambda \omega_\nu \sum_\mu \omega_\mu Y_{\lambda,\co\mu}^* Y_{\nu,\co\mu}^* = \omega_\lambda \omega_\nu \sum_{\mu,\sigma} \omega_\mu d_\mu N_{\lambda,\nu}^\sigma Y_{\co\mu,\sigma}^* \\[.4em] &= \omega_\lambda \omega_\nu \sum_{\mu,\rho,\sigma} \omega_\mu d_\mu N_{\lambda,\nu}^\sigma N_{\co\mu,\sigma}^\rho \frac{\omega_\rho}{\omega_\mu \omega_\sigma} d_\rho = \omega_\lambda \omega_\nu \sum_{\rho,\sigma} d_\rho^2 d_\sigma N_{\lambda,\nu}^\sigma \frac{\omega_\rho}{\omega_\sigma} \\[.4em] &= Y_{\lambda,\nu} \sum_\rho d_\rho^2 \omega_\rho = Y_{\lambda,\nu} z_{\Delta} \,, \end{array} \] hence $TSTST = {\rm{e}}^{-\pi{\rm{i}} c/4}|z_{\Delta}|^{-1} S z_{\Delta}=S$. The remaining relations $CTC=T$, $CSC=S$ and $T^*T={\bf1}$ are obvious. We define {\sl weight vectors} $y^\lambda$ with components $y^\lambda_\mu=Y_{\lambda,\mu}$ and {\sl statistics characters} $\chi_\lambda:{\Delta}\rightarrow\Bbb{C}$ with evaluations $\chi_\lambda(\mu)=d_\lambda^{-1}Y_{\lambda,\mu}$, $\lambda,\mu\in{\Delta}$. We have seen that the weight vectors $y^\lambda$ are simultaneous eigenvectors of the fusion matrices $N_\mu$ with eigenvalues $\chi_\lambda(\mu)$, $N_\mu y^\lambda = \chi_\lambda(\mu) y^\lambda$. Hence we obtain by computing inner products, \[ \chi_\mu(\rho) \langle y^\lambda,y^\mu \rangle = \langle y^\lambda , N_\rho y^\mu \rangle = \langle N_{\co\rho} y^\lambda,y^\mu \rangle = \chi_\lambda(\co\rho)^* \langle y^\lambda,y^\mu \rangle = \chi_\lambda(\rho) \langle y^\lambda,y^\mu \rangle \,.\] Therefore the eigenvectors are either orthogonal, $\langle y^\lambda,y^\mu \rangle =0$, or parallel, $d_\mu y^\lambda = d_\lambda y^\mu$ since then the characters are equal, $\chi_\lambda=\chi_\mu$. It is obvious that if some $\lambda\in{\Delta}$ is degenerate, i.e.\ has trivial monodromy with all other $\mu\in{\Delta}$, then $y^\lambda$ is parallel to the vector $y^0$. (Here and later we use the label ``$0$'' for the identity ${\rm{id}}_A\in{\Delta}$.) Note that we have $y^0_\mu=d_\mu$, and then $Y_{\lambda,\mu}=d_\lambda d_\mu$. Conversely, if $y^\lambda$ is parallel to $y^0$ we have seen that then necessarily $Y_{\lambda,\mu}=d_\lambda d_\mu$, hence \[ Y_{\lambda,\mu} = \sum_{\rho\in{\Delta}} \frac{\omega_\lambda \omega_\mu}{\omega_\rho} N_{\lambda,\mu}^\rho \, d_\rho = d_\lambda d_\mu = \sum_{\rho\in{\Delta}} N_{\lambda,\mu}^\rho \, d_\rho \,, \qquad \mu\in{\Delta} \,, \] and this is clearly only possible if all the eigenvalues $\omega_\lambda \omega_\mu \omega^{-1}_\rho$ of the monodromy are trivial, i.e.\ if $\lambda$ is degenerate. We conclude that a braiding on ${\Delta}$ is non-degenerate if and only if $\langle y^\lambda , y^0 \rangle = \del \lambda 0 w$, where $w=\sum_{\lambda\in{\Delta}} d_\lambda^2$ is the {\sl global index}. We now arrive at Rehren's result \cite{R0}. \begin{theorem} \label{ST} The following conditions are equivalent for a finite braided system of endomorphisms ${\Delta}$: \begin{enumerate} \item The braiding on ${\Delta}$ is non-degenerate. \item We have $w=|z_{\Delta}|^2$ and the matrices $S$ and $T$ obey the full Verlinde modular algebra \[ S^*S=T^*T={\bf1} \,,\quad (ST)^3=S^2 = C \,, \quad CTC=T \,, \] moreover $S$ diagonalizes the fusion rules (Verlinde formula): \[ N_{\lambda,\mu}^\nu = \sum_{\rho\in{\Delta}} \frac{S_{\lambda,\rho}S_{\mu,\rho} S_{\nu,\rho}^*}{S_{0,\rho}} \,. \] \end{enumerate} \end{theorem} Note that the implication ${\it 2.}\Rightarrow{\it 1.}$ is trivial since invertibility of $S$ implies that there is no vector $y^\lambda$ parallel $y^0$. So let us assume that the braiding is non-degenerate: $\langle y^\lambda , y^0 \rangle = \del \lambda 0 w$ for all $\lambda\in{\Delta}$. Then we can first check \[ \begin{array}{ll} w &= \sum_\mu \langle y^0,y^\mu \rangle d_\mu \omega_\mu^{-1} = \sum_{\mu,\nu} d_\nu Y_{\mu,\nu} d_\mu \omega_\mu^{-1} = \sum_{\mu,\nu,\lambda} d_\nu \frac{\omega_\mu\omega_\nu}{\omega_\lambda} N_{\mu,\nu}^\lambda d_\lambda d_\mu \omega_\mu^{-1} \\[.4em] &= \sum_{\mu,\nu,\lambda} d_\lambda d_\nu \frac{\omega_\nu}{\omega_\lambda} N_{\co\nu,\lambda}^\mu d_\mu = \sum_{\lambda,\nu} d_\lambda^2 \omega_\lambda^{-1} d_\nu^2 \omega_\nu \,, \end{array} \] thus $w=\left|\sum_{\lambda\in{\Delta}} d_\lambda^2 \omega_\lambda\right|^2\equiv |z_{\Delta}|^2$. Next we compute \[ \langle y^\lambda,y^\mu \rangle = \sum_\rho Y_{\lambda,\rho}^* Y_{\mu,\rho} = \sum_{\rho,\nu} N_{\co\lambda,\mu}^\nu Y_{\rho,\nu} d_\rho = \sum_\nu N_{\co\lambda,\mu}^\nu \langle y^0,y^\nu \rangle = N_{\co\lambda,\mu}^0 w = \del \lambda\mu w \,, \] hence $S^*S={\bf1}$. Similarly we observe that $\sum_\rho Y_{\lambda,\rho} Y_{\mu,\rho}= \sum_\rho Y_{\co\lambda,\rho}^* Y_{\mu,\rho}= \del {\co\lambda}\mu w$, giving $S^2=C$ which obviously commutes with $T$. Finally we check \[ \sum_\rho \frac{S_{\lambda,\rho}S_{\mu,\rho} S_{\nu,\rho}^*}{S_{0,\rho}} = w^{-1} \sum_\rho \frac{Y_{\lambda,\rho}Y_{\mu,\rho} Y_{\nu,\rho}^*}{d_\rho} = w^{-1} \sum_{\rho,\sigma} N_{\lambda,\mu}^\sigma Y_{\rho,\sigma} Y_{\nu,\rho}^* = \sum_\sigma N_{\lambda,\mu}^\sigma \del \nu\sigma = N_{\lambda,\mu}^\nu \,,\] proving the Verlinde identity. \begin{corollary} \label{sltzaction} If the braiding on ${\Delta}$ is non-degenerate, then the matrix $S$ and the diagonal matrix $T$ are the images $S=U({\cal{S}})$ and $T=U({\cal{T}})$ of canonical generators \[ {\cal{S}} = \left(\begin{array}{cc} 0 & -1 \\ 1 & 0 \end{array}\right), \qquad {\cal{T}} = \left(\begin{array}{cc} 1 & 1 \\ 0 & 1 \end{array}\right), \] in a unitary representation $U$ of the modular group\footnote{In the literature the name ``modular group'' is often reserved for ${\it{PSL}}(2;\bbZ)={\it{SL}}(2;\bbZ)/\Bbb{Z}_2$ rather than ${\it{SL}}(2;\bbZ)$. Clearly, we obtain a representation of ${\it{PSL}}(2;\bbZ)$ whenever the charge conjugation is trivial, $C={\bf1}$.} ${\it{SL}}(2;\bbZ)$ with dimension $|{\Delta}|$, the cardinality of ${\Delta}$. \end{corollary} \section{Graphical Intertwiner Calculus} \label{sec-graphcal} \subsection{Basic graphical intertwiner calculus} \label{sec-basgraphcal} We now introduce our conventions to represent and manipulate intertwiners graphically. We consider a braided system of endomorphisms ${\Delta}\subset{\rm{End}}(A)$ with $A$ a type III factor. Essentially we represent intertwiners by ``wire diagrams'' where the (oriented) wires represent endomorphisms $\lambda\in{\Delta}$. This works as follows. For an intertwiner $x\in{\rm{Hom}}(\lambda_1\lambda_2\cdots\lambda_n,\mu_1\mu_2\cdots\mu_m)$ we draw a (dashed) box with $n$ (downward) incoming wires labelled by $\lambda_1,\dots,\lambda_n$ and $m$ (downward) outgoing wires $\mu_1,\dots,\mu_m$ as in Fig.\ \ref{boxx}, $\lambda_i,\mu_j\in{\Delta}$. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(60,30) \allinethickness{0.3pt} \put(5,10){\dashbox{2}(50,10){$x$}} \put(10,30){\vector(0,-1){10}} \put(25,30){\vector(0,-1){10}} \put(50,30){\vector(0,-1){10}} \put(10,10){\vector(0,-1){10}} \put(25,10){\vector(0,-1){10}} \put(50,10){\vector(0,-1){10}} \put(3,25){\makebox(0,0){$\lambda_1$}} \put(18,25){\makebox(0,0){$\lambda_2$}} \put(57,25){\makebox(0,0){$\lambda_n$}} \put(3,5){\makebox(0,0){$\mu_1$}} \put(18,5){\makebox(0,0){$\mu_2$}} \put(57,5){\makebox(0,0){$\mu_m$}} \put(37.5,25){\makebox(0,0){$\cdots$}} \put(37.5,5){\makebox(0,0){$\cdots$}} \end{picture} \end{center} \caption{An intertwiner $x$} \label{boxx} \end{figure} Therefore the diagrammatic representation of $x$ does not only specify it as an operator, it even specifies the intertwiner space it is considered to belong to. (Note that the same operator can belong to different intertwiner spaces as e.g.\ the identity operator belongs to any ${\rm{Hom}}(\lambda,\lambda)$ with $\lambda$ varying.) If a morphism $\rho\in{\Delta}$ is applied to $x$, then $\rho(x)\in{\rm{Hom}}(\rho\lambda_1\lambda_2\cdots\lambda_n,\rho\mu_1\mu_2\cdots\mu_m)$ is represented graphically by adding a straight wire on the left as in Fig.\ \ref{rhoboxx}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(80,30) \allinethickness{0.3pt} \put(25,10){\dashbox{2}(50,10){$x$}} \put(10,30){\vector(0,-1){30}} \put(30,30){\vector(0,-1){10}} \put(45,30){\vector(0,-1){10}} \put(70,30){\vector(0,-1){10}} \put(30,10){\vector(0,-1){10}} \put(45,10){\vector(0,-1){10}} \put(70,10){\vector(0,-1){10}} \put(3,15){\makebox(0,0){$\rho$}} \put(23,25){\makebox(0,0){$\lambda_1$}} \put(38,25){\makebox(0,0){$\lambda_2$}} \put(77,25){\makebox(0,0){$\lambda_n$}} \put(23,5){\makebox(0,0){$\mu_1$}} \put(38,5){\makebox(0,0){$\mu_2$}} \put(77,5){\makebox(0,0){$\mu_m$}} \put(57.5,25){\makebox(0,0){$\cdots$}} \put(57.5,5){\makebox(0,0){$\cdots$}} \end{picture} \end{center} \caption{The intertwiner $\rho(x)$} \label{rhoboxx} \end{figure} Reflecting the fact that $x$ can also be considered as an intertwiner in ${\rm{Hom}}(\lambda_1\lambda_2\cdots\lambda_n\rho,\mu_1\mu_2\cdots\mu_m\rho)$ we can always add (or remove) a straight wire on the right as in Fig.\ \ref{boxxrho} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(80,30) \allinethickness{0.3pt} \put(5,10){\dashbox{2}(50,10){$x$}} \put(10,30){\vector(0,-1){10}} \put(25,30){\vector(0,-1){10}} \put(50,30){\vector(0,-1){10}} \put(10,10){\vector(0,-1){10}} \put(25,10){\vector(0,-1){10}} \put(50,10){\vector(0,-1){10}} \put(70,30){\vector(0,-1){30}} \put(3,25){\makebox(0,0){$\lambda_1$}} \put(18,25){\makebox(0,0){$\lambda_2$}} \put(57,25){\makebox(0,0){$\lambda_n$}} \put(3,5){\makebox(0,0){$\mu_1$}} \put(18,5){\makebox(0,0){$\mu_2$}} \put(57,5){\makebox(0,0){$\mu_m$}} \put(77,15){\makebox(0,0){$\rho$}} \put(37.5,25){\makebox(0,0){$\cdots$}} \put(37.5,5){\makebox(0,0){$\cdots$}} \end{picture} \end{center} \caption{The intertwiner $x$} \label{boxxrho} \end{figure} without changing the intertwiner as an operator. We say that intertwiners $x\in{\rm{Hom}}(\lambda_1\lambda_2\cdots\lambda_n,\mu_1\mu_2\cdots\mu_m)$ and $y\in{\rm{Hom}}(\nu_1\nu_2\cdots\nu_k,\rho_1\rho_2\cdots\rho_l)$, $\rho_j\in{\Delta}$, are {\sl diagrammatically composable} if $m=k$ and $\mu_i=\nu_i$ for all $i=1,2,\dots,m$. Then the composed intertwiner $yx\in{\rm{Hom}}(\lambda_1\lambda_2\cdots\lambda_n,\rho_1\rho_2\cdots\rho_l)$ is represented graphically by putting the wire diagram for $x$ on top of that for $y$ as in Fig.\ \ref{x-top-y}. We also call this graphical procedure composition of wire diagrams. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(80,50) \allinethickness{0.3pt} \put(5,30){\dashbox{2}(50,10){$x$}} \put(5,10){\dashbox{2}(50,10){$y$}} \put(10,50){\vector(0,-1){10}} \put(25,50){\vector(0,-1){10}} \put(50,50){\vector(0,-1){10}} \put(10,30){\vector(0,-1){10}} \put(25,30){\vector(0,-1){10}} \put(50,30){\vector(0,-1){10}} \put(10,10){\vector(0,-1){10}} \put(25,10){\vector(0,-1){10}} \put(50,10){\vector(0,-1){10}} \put(3,45){\makebox(0,0){$\lambda_1$}} \put(18,45){\makebox(0,0){$\lambda_2$}} \put(57,45){\makebox(0,0){$\lambda_n$}} \put(3,25){\makebox(0,0){$\mu_1$}} \put(18,25){\makebox(0,0){$\mu_2$}} \put(57,25){\makebox(0,0){$\mu_m$}} \put(3,5){\makebox(0,0){$\rho_1$}} \put(18,5){\makebox(0,0){$\rho_2$}} \put(57,5){\makebox(0,0){$\rho_l$}} \put(37.5,45){\makebox(0,0){$\cdots$}} \put(37.5,25){\makebox(0,0){$\cdots$}} \put(37.5,5){\makebox(0,0){$\cdots$}} \end{picture} \end{center} \caption{Product $yx$ of diagrammatically composable intertwiners $x$ and $y$} \label{x-top-y} \end{figure} Sometimes diagrammatic composability may be achieved by adding or removing straight wires on the right. Now let also $x'\in{\rm{Hom}}(\lambda_1'\lambda_2'\cdots\lambda_{n'}',\mu_1'\mu_2'\cdots\mu_{m'}')$ with $\lambda_i',\mu_j'\in{\Delta}$. The intertwining property of $x$ yields the identity $\mu_1\mu_2\cdots\mu_m \rho_1\rho_2\cdots\rho_l(x')x = x\lambda_1\lambda_2\cdots\lambda_n \rho_1\rho_2\cdots\rho_l(x')$, and this is diagrammatically given in Fig.\ \ref{xx'}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(240,50) \allinethickness{0.3pt} \put(5,30){\dashbox{2}(30,10){$x$}} \put(75,10){\dashbox{2}(30,10){$x'$}} \put(10,50){\vector(0,-1){10}} \put(30,50){\vector(0,-1){10}} \put(10,30){\vector(0,-1){30}} \put(30,30){\vector(0,-1){30}} \put(45,50){\vector(0,-1){50}} \put(65,50){\vector(0,-1){50}} \put(80,50){\vector(0,-1){30}} \put(100,50){\vector(0,-1){30}} \put(80,10){\vector(0,-1){10}} \put(100,10){\vector(0,-1){10}} \put(20,45){\makebox(0,0){$\cdots$}} \put(20,15){\makebox(0,0){$\cdots$}} \put(55,25){\makebox(0,0){$\cdots$}} \put(90,35){\makebox(0,0){$\cdots$}} \put(90,5){\makebox(0,0){$\cdots$}} \put(4,45){\makebox(0,0){$\lambda_1$}} \put(37,45){\makebox(0,0){$\lambda_n$}} \put(4,5){\makebox(0,0){$\mu_1$}} \put(37,5){\makebox(0,0){$\mu_m$}} \put(50,15){\makebox(0,0){$\rho_1$}} \put(60,8){\makebox(0,0){$\rho_l$}} \put(74,45){\makebox(0,0){$\lambda_1'$}} \put(107,45){\makebox(0,0){$\lambda_{n'}'$}} \put(74,5){\makebox(0,0){$\mu_1'$}} \put(107,5){\makebox(0,0){$\mu_{m'}'$}} \put(120,25){\makebox(0,0){$=$}} \put(135,10){\dashbox{2}(30,10){$x$}} \put(205,30){\dashbox{2}(30,10){$x'$}} \put(140,50){\vector(0,-1){30}} \put(160,50){\vector(0,-1){30}} \put(140,10){\vector(0,-1){10}} \put(160,10){\vector(0,-1){10}} \put(175,50){\vector(0,-1){50}} \put(195,50){\vector(0,-1){50}} \put(210,50){\vector(0,-1){10}} \put(230,50){\vector(0,-1){10}} \put(210,30){\vector(0,-1){30}} \put(230,30){\vector(0,-1){30}} \put(150,35){\makebox(0,0){$\cdots$}} \put(150,5){\makebox(0,0){$\cdots$}} \put(185,25){\makebox(0,0){$\cdots$}} \put(220,45){\makebox(0,0){$\cdots$}} \put(220,15){\makebox(0,0){$\cdots$}} \put(134,45){\makebox(0,0){$\lambda_1$}} \put(167,45){\makebox(0,0){$\lambda_n$}} \put(134,5){\makebox(0,0){$\mu_1$}} \put(167,5){\makebox(0,0){$\mu_m$}} \put(180,15){\makebox(0,0){$\rho_1$}} \put(190,8){\makebox(0,0){$\rho_l$}} \put(204,45){\makebox(0,0){$\lambda_1'$}} \put(237,45){\makebox(0,0){$\lambda_{n'}'$}} \put(204,5){\makebox(0,0){$\mu_1'$}} \put(237,5){\makebox(0,0){$\mu_{m'}'$}} \end{picture} \end{center} \caption{Vertical translation intertwiners $x$ and $x'$} \label{xx'} \end{figure} Thus we have some freedom in translating intertwiner boxes vertically without actually changing the represented intertwiner. The intertwiners we consider are (sums over) compositions of {\sl elementary intertwiners} arising from the unitary braiding operators $\eps\lambda\mu\in{\rm{Hom}}(\lambda\mu,\mu\lambda)$ and isometries $t\in{\rm{Hom}}(\lambda,\mu\nu)$. The wire diagrams and boxes we are dealing with are therefore compositions of ``elementary boxes'' representing the elementary intertwiners. We now have to introduce some normalization convention. First, the identity intertwiner ${\bf1}\equiv{\bf1}_A$ is naturally given by the ``trivial box'' with only straight wires of arbitrary labels. The next elementary intertwiner is $\rho_1\rho_2\cdots\rho_n(\eps\lambda\mu)$ for which we draw a box as in Fig.\ \ref{rrrepslm} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(170,20) \allinethickness{0.3pt} \put(0,0){\dashbox{2}(170,20){}} \put(10,20){\vector(0,-1){20}} \put(25,20){\vector(0,-1){20}} \put(60,20){\vector(0,-1){20}} \put(75,20){\vector(1,-1){20}} \put(95,20){\line(-1,-1){8}} \put(83,8){\vector(-1,-1){8}} \put(110,20){\vector(0,-1){20}} \put(125,20){\vector(0,-1){20}} \put(160,20){\vector(0,-1){20}} \put(5,10){\makebox(0,0){$\rho_1$}} \put(20,10){\makebox(0,0){$\rho_2$}} \put(55,10){\makebox(0,0){$\rho_n$}} \put(100,5){\makebox(0,0){$\lambda$}} \put(70,5){\makebox(0,0){$\mu$}} \put(115,10){\makebox(0,0){$\nu_1$}} \put(130,10){\makebox(0,0){$\nu_2$}} \put(165,10){\makebox(0,0){$\nu_m$}} \put(37.5,10){\makebox(0,0){$\cdots$}} \put(147.5,10){\makebox(0,0){$\cdots$}} \end{picture} \end{center} \caption{$\rho_1\rho_2\cdots\rho_n(\eps\lambda\mu)$} \label{rrrepslm} \end{figure} where the arbitrary labels $\nu_1,\dots,\nu_m$ are irrelevant and may be omitted. Similarly, the box of Fig.\ \ref{rrrt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(170,20) \allinethickness{0.3pt} \put(0,0){\dashbox{2}(170,20){}} \put(10,20){\vector(0,-1){20}} \put(25,20){\vector(0,-1){20}} \put(60,20){\vector(0,-1){20}} \put(85,20){\vector(0,-1){10}} \put(85,10){\vector(-1,-1){10}} \put(85,10){\vector(1,-1){10}} \put(110,20){\vector(0,-1){20}} \put(125,20){\vector(0,-1){20}} \put(160,20){\vector(0,-1){20}} \put(5,10){\makebox(0,0){$\rho_1$}} \put(20,10){\makebox(0,0){$\rho_2$}} \put(55,10){\makebox(0,0){$\rho_n$}} \put(90,15){\makebox(0,0){$\lambda$}} \put(70,5){\makebox(0,0){$\mu$}} \put(100,5){\makebox(0,0){$\nu$}} \put(85,5){\makebox(0,0){$t$}} \put(115,10){\makebox(0,0){$\nu_1$}} \put(130,10){\makebox(0,0){$\nu_2$}} \put(165,10){\makebox(0,0){$\nu_m$}} \put(37.5,10){\makebox(0,0){$\cdots$}} \put(147.5,10){\makebox(0,0){$\cdots$}} \end{picture} \end{center} \caption{$\protect\sqrt[4]{\frac{d_\mu d_\nu}{d_\lambda}} \rho_1\rho_2\cdots\rho_n(t)$ where $t\in{\rm{Hom}}(\lambda,\mu\nu)$ is an isometry} \label{rrrt} \end{figure} represents the elementary intertwiner $d_\mu^{1/4} d_\nu^{1/4} d_\lambda^{-1/4}\rho_1\rho_2\cdots\rho_n(t)$, where $t\in{\rm{Hom}}(\lambda,\mu\nu)$ is an isometry. We label the trivalent vertex in the box by $t$ since ${\rm{Hom}}(\lambda,\mu\nu)$ may be more than one-dimensional and so we have to specify the intertwiner. (Note that there would still be an ambiguity of a phase for the choice of an isometry even if ${\rm{Hom}}(\lambda,\mu\nu)$ is only one-dimensional.) Finally, the elementary intertwiners $\eps\lambda\mu ^*=\epsm\mu\lambda$ and $d_\mu^{1/4} d_\nu^{1/4} d_\lambda^{-1/4}\rho_1\rho_2\cdots\rho_n(t)^*$ are represented by Figs.\ \ref{rrrepslm*} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(170,20) \allinethickness{0.3pt} \put(0,0){\dashbox{2}(170,20){}} \put(10,20){\vector(0,-1){20}} \put(25,20){\vector(0,-1){20}} \put(60,20){\vector(0,-1){20}} \put(75,20){\line(1,-1){8}} \put(87,8){\vector(1,-1){8}} \put(95,20){\vector(-1,-1){20}} \put(110,20){\vector(0,-1){20}} \put(125,20){\vector(0,-1){20}} \put(160,20){\vector(0,-1){20}} \put(5,10){\makebox(0,0){$\rho_1$}} \put(20,10){\makebox(0,0){$\rho_2$}} \put(55,10){\makebox(0,0){$\rho_n$}} \put(100,5){\makebox(0,0){$\mu$}} \put(70,5){\makebox(0,0){$\lambda$}} \put(115,10){\makebox(0,0){$\nu_1$}} \put(130,10){\makebox(0,0){$\nu_2$}} \put(165,10){\makebox(0,0){$\nu_m$}} \put(37.5,10){\makebox(0,0){$\cdots$}} \put(147.5,10){\makebox(0,0){$\cdots$}} \end{picture} \end{center} \caption{$\rho_1\rho_2\cdots\rho_n(\eps\lambda\mu)^* =\rho_1\rho_2\cdots\rho_n(\epsm\mu\lambda)$} \label{rrrepslm*} \end{figure} and \ref{rrrt*}, i.e.\ they are obtained from the original boxes in Figs.\ \ref{rrrepslm} and \ref{rrrt} by vertical reflection and inversion of all the arrows. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(170,20) \allinethickness{0.3pt} \put(0,0){\dashbox{2}(170,20){}} \put(10,20){\vector(0,-1){20}} \put(25,20){\vector(0,-1){20}} \put(60,20){\vector(0,-1){20}} \put(75,20){\vector(1,-1){10}} \put(95,20){\vector(-1,-1){10}} \put(85,10){\vector(0,-1){10}} \put(110,20){\vector(0,-1){20}} \put(125,20){\vector(0,-1){20}} \put(160,20){\vector(0,-1){20}} \put(5,10){\makebox(0,0){$\rho_1$}} \put(20,10){\makebox(0,0){$\rho_2$}} \put(55,10){\makebox(0,0){$\rho_n$}} \put(90,5){\makebox(0,0){$\lambda$}} \put(70,15){\makebox(0,0){$\mu$}} \put(100,15){\makebox(0,0){$\nu$}} \put(87,16){\makebox(0,0){$t^*$}} \put(115,10){\makebox(0,0){$\nu_1$}} \put(130,10){\makebox(0,0){$\nu_2$}} \put(165,10){\makebox(0,0){$\nu_m$}} \put(37.5,10){\makebox(0,0){$\cdots$}} \put(147.5,10){\makebox(0,0){$\cdots$}} \end{picture} \end{center} \caption{$\protect\sqrt[4]{\frac{d_\mu d_\nu}{d_\lambda}} \rho_1\rho_2\cdots\rho_n(t)^*$ where $t\in{\rm{Hom}}(\lambda,\mu\nu)$ is an isometry} \label{rrrt*} \end{figure} Note that $\varepsilon\equiv\varepsilon^+$ represents overcrossing and $\varepsilon^-$ undercrossing of wires. We will consider intertwiners which are products of diagrammatically composable elementary intertwiners. In terms of wire diagrams we are correspondingly dealing with compositions of elementary boxes of Figs.\ \ref{rrrepslm}, \ref{rrrt}, \ref{rrrepslm*}, \ref{rrrt*} so that the wires with the same labels (and orientations) can and will be glued together in parallel and then we finally forget about the boundaries of the (dashed) boxes. Therefore, if a wire diagram represents some intertwiner $x$ then $x^*$ is represented by the diagram obtained by vertical reflection and reversing all the arrows. Note that our resulting wire diagrams are then composed only from straight lines, over- and undercrossings (in X-shape) and trivalent vertices (in Y-shape or inverted Y-shape). So far, we have considered only wires with downward orientation. We now introduce also the reversed orientation in terms of conjugation as follows: Reversing the orientation of an arrow on a wire changes its label $\lambda$ to $\co\lambda$. Also we will usually omit drawing a wire labelled by ${\rm{id}}\equiv{\rm{id}}_A$. For each $\lambda\in{\Delta}$ we fix (the common phase of) isometries $r_\lambda\in{\rm{Hom}}({\rm{id}},\co\lambda\la)$ and ${\co r}_\lambda\in{\rm{Hom}}({\rm{id}},\lambda\co\lambda)$ such that $\lambda(r_\lambda)^*{\co r}_\lambda=\co\lambda({\co r}_\lambda)^*r_\lambda=d_\lambda^{-1}{\bf1}$ and in turn for $\sqrt{d_\lambda} r_\lambda$ we draw one of the equivalent diagrams in Fig.\ \ref{risom}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(145,20) \allinethickness{0.3pt} \put(25,20){\vector(0,-1){10}} \put(25,10){\vector(-1,-1){10}} \put(25,10){\vector(1,-1){10}} \put(30,15){\makebox(0,0){${\rm{id}}$}} \put(10,5){\makebox(0,0){$\co\lambda$}} \put(40,5){\makebox(0,0){$\lambda$}} \put(50,10){\makebox(0,0){$=$}} \put(75,10){\vector(-1,-1){10}} \put(75,10){\vector(1,-1){10}} \put(60,5){\makebox(0,0){$\co\lambda$}} \put(90,5){\makebox(0,0){$\lambda$}} \put(100,10){\makebox(0,0){$=$}} \put(120,2){\arc{20}{3.142}{0}} \put(110,0){\line(0,1){2}} \put(130,0){\line(0,1){2}} \put(135,5){\makebox(0,0){$\lambda$}} \put(130,0){\vector(0,-1){0}} \end{picture} \end{center} \caption{Wire diagrams for $\protect\sqrt{d_\lambda} r_\lambda$} \label{risom} \end{figure} So the normalized isometries and their adjoints appear in wire diagrams as ``caps'' and ``cups'', respectively. The point is that with our normalization convention, the relation $\lambda(r_\lambda)^*{\co r}_\lambda=d_\lambda^{-1}{\bf1}$ (and its adjoint) gives a {\sl topological invariance} for intertwiners represented by wire diagrams, displayed in Fig.\ \ref{isoinv1}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(170,40) \allinethickness{0.3pt} \put(8,5){\makebox(0,0){$\lambda$}} \put(15,20){\vector(0,-1){20}} \put(25,20){\arc{20}{3.142}{0}} \put(45,20){\arc{20}{0}{3.142}} \put(55,40){\line(0,-1){20}} \put(70,20){\makebox(0,0){$=$}} \put(85,40){\vector(0,-1){40}} \put(92,5){\makebox(0,0){$\lambda$}} \put(100,20){\makebox(0,0){$=$}} \put(115,40){\line(0,-1){20}} \put(125,20){\arc{20}{0}{3.142}} \put(145,20){\arc{20}{3.142}{0}} \put(155,20){\vector(0,-1){20}} \put(162,5){\makebox(0,0){$\lambda$}} \end{picture} \end{center} \caption{A topological invariance for intertwiners represented by wire diagrams} \label{isoinv1} \end{figure} Note that then the wire diagrams in Fig.\ \ref{dla} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(80,20) \allinethickness{0.3pt} \put(22,19.84){\vector(1,0){0}} \put(20,10){\circle{20}} \put(5,10){\makebox(0,0){$\lambda$}} \put(40,10){\makebox(0,0){$=$}} \put(58,19.84){\vector(-1,0){0}} \put(60,10){\circle{20}} \put(75,10){\makebox(0,0){$\lambda$}} \end{picture} \end{center} \caption{Wire diagrams for the statistical dimension $d_\lambda$} \label{dla} \end{figure} represent the scalar $d_\lambda$ (i.e.\ the intertwiner $d_\lambda {\bf1}\in{\rm{Hom}}({\rm{id}},{\rm{id}})$). Also note the ``vertical Reidemeister move of type II'' in Fig.\ \ref{Reid2} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(130,40) \allinethickness{0.3pt} \put(0,20){\arc{50}{5.356}{0.927}} \put(40,20){\arc{50}{2.214}{2.418}} \put(40,20){\arc{50}{2.578}{3.705}} \put(40,20){\arc{50}{3.865}{4.069}} \put(8,5){\makebox(0,0){$\lambda$}} \put(15,0){\vector(-4,-3){0}} \put(25,0){\vector(4,-3){0}} \put(32,5){\makebox(0,0){$\mu$}} \put(45,20){\makebox(0,0){$=$}} \put(60,40){\vector(0,-1){40}} \put(70,40){\vector(0,-1){40}} \put(53,5){\makebox(0,0){$\lambda$}} \put(77,5){\makebox(0,0){$\mu$}} \put(85,20){\makebox(0,0){$=$}} \put(90,20){\arc{50}{5.356}{5.560}} \put(90,20){\arc{50}{5.720}{0.564}} \put(90,20){\arc{50}{0.724}{0.927}} \put(130,20){\arc{50}{2.214}{4.069}} \put(98,5){\makebox(0,0){$\lambda$}} \put(105,0){\vector(-4,-3){0}} \put(115,0){\vector(4,-3){0}} \put(122,5){\makebox(0,0){$\mu$}} \end{picture} \end{center} \caption{Unitarity of braiding operators as a vertical Reidemeister move of type II} \label{Reid2} \end{figure} is just the unitarity condition $\eps\lambda\mu ^* \eps\lambda\mu={\bf1}=\eps\mu\lambda\eps\mu\lambda ^*$. The BFE's yield another topological invariance, see Fig.\ \ref{wireBFE1} for the first equation \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(120,60) \allinethickness{0.3pt} \put(26.180,5){\arc{32.361}{3.142}{4.249}} \put(10,5){\vector(0,-1){5}} \put(28,24){\line(-2,-1){9.1}} \put(33.820,45){\arc{32.361}{0}{1.107}} \put(50,60){\line(0,-1){15}} \put(32,26){\line(2,1){9.1}} \put(30,60){\line(0,-1){35}} \put(30,25){\vector(0,-1){15}} \put(30,10){\vector(1,-1){10}} \put(30,10){\vector(-1,-1){10}} \put(30,5){\makebox(0,0){$t$}} \put(35,55){\makebox(0,0){$\lambda$}} \put(18,5){\makebox(0,0){$\mu$}} \put(42,5){\makebox(0,0){$\nu$}} \put(5,5){\makebox(0,0){$\rho$}} \put(60,25){\makebox(0,0){$=$}} \put(86.180,5){\arc{32.361}{3.142}{4.199}} \put(70,5){\vector(0,-1){5}} \put(90,25){\line(-2,-1){8.1}} \put(93.820,45){\arc{32.361}{0}{1.057}} \put(110,60){\line(0,-1){15}} \put(90,25){\line(2,1){8.1}} \put(90,60){\vector(0,-1){10}} \put(90,50){\line(1,-1){10}} \put(90,50){\line(-1,-1){10}} \put(90,45){\makebox(0,0){$t$}} \put(80,40){\line(0,-1){20}} \put(100,40){\line(0,-1){10}} \put(80,10){\line(0,1){10}} \put(100,10){\line(0,1){20}} \put(80,10){\vector(0,-1){10}} \put(100,10){\vector(0,-1){10}} \put(95,55){\makebox(0,0){$\lambda$}} \put(85,5){\makebox(0,0){$\mu$}} \put(105,5){\makebox(0,0){$\nu$}} \put(65,5){\makebox(0,0){$\rho$}} \end{picture} \end{center} \caption{The first braiding fusion equation} \label{wireBFE1} \end{figure} and Fig.\ \ref{wireBFE2} for the second equation. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(120,60) \allinethickness{0.3pt} \put(26.180,45){\arc{32.361}{2.034}{3.142}} \put(10,60){\line(0,-1){15}} \put(30,25){\line(-2,1){11.1}} \put(33.820,5){\arc{32.361}{5.176}{6.283}} \put(50,5){\vector(0,-1){5}} \put(30,25){\line(2,-1){11.1}} \put(30,60){\line(0,-1){33}} \put(30,23){\vector(0,-1){13}} \put(30,10){\vector(1,-1){10}} \put(30,10){\vector(-1,-1){10}} \put(30,5){\makebox(0,0){$t$}} \put(35,55){\makebox(0,0){$\lambda$}} \put(18,5){\makebox(0,0){$\mu$}} \put(42,5){\makebox(0,0){$\nu$}} \put(55,5){\makebox(0,0){$\rho$}} \put(60,25){\makebox(0,0){$=$}} \put(86.180,45){\arc{32.361}{2.034}{3.142}} \put(70,60){\line(0,-1){15}} \put(90,25){\line(-2,1){11.1}} \put(93.820,5){\arc{32.361}{5.176}{6.283}} \put(110,5){\vector(0,-1){5}} \put(90,25){\line(2,-1){11.1}} \put(90,60){\vector(0,-1){10}} \put(90,50){\line(1,-1){10}} \put(90,50){\line(-1,-1){10}} \put(90,45){\makebox(0,0){$t$}} \put(80,40){\line(0,-1){8}} \put(100,40){\line(0,-1){18}} \put(80,10){\line(0,1){18}} \put(100,10){\line(0,1){8}} \put(80,10){\vector(0,-1){10}} \put(100,10){\vector(0,-1){10}} \put(95,55){\makebox(0,0){$\lambda$}} \put(75,5){\makebox(0,0){$\mu$}} \put(95,5){\makebox(0,0){$\nu$}} \put(115,5){\makebox(0,0){$\rho$}} \end{picture} \end{center} \caption{The second braiding fusion equation} \label{wireBFE2} \end{figure} The third and fourth equations are obtained similarly by use of the co-isometry $t^*$; we leave it as an exercise to the reader to draw the corresponding wire diagrams. Up to conjugation they can also be obtained by changing over- to undercrossings in Figs.\ \ref{wireBFE1} and \ref{wireBFE2}. Finally, the braid relation, \erf{YBE}, represents graphically a vertical Reidemeister move of type III, presented in Fig.\ \ref{Reid3}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(140,60) \allinethickness{0.3pt} \put(15,60){\line(0,-1){15.858}} \put(25,30){\vector(1,-1){30}} \put(25,30){\line(-1,1){7.071}} \put(35,60){\line(1,-1){17.071}} \put(55,30){\line(0,1){5.858}} \put(55,30){\line(0,-1){5.858}} \put(47,12){\line(1,1){5.071}} \put(43,8){\vector(-1,-1){8}} \put(55,60){\line(-1,-1){8}} \put(43,48){\line(-1,-1){16}} \put(23,28){\line(-1,-1){5.071}} \put(15,15.858){\vector(0,-1){15.858}} \put(25,44.142){\arc{20}{2.356}{3.142}} \put(45,35.858){\arc{20}{5.498}{0}} \put(45,24.142){\arc{20}{0}{0.785}} \put(25,15.858){\arc{20}{3.142}{3.927}} \put(8,5){\makebox(0,0){$\rho$}} \put(30,5){\makebox(0,0){$\mu$}} \put(60,5){\makebox(0,0){$\lambda$}} \put(70,30){\makebox(0,0){$=$}} \put(85,60){\line(1,-1){37.071}} \put(125,15.858){\vector(0,-1){15.858}} \put(105,60){\line(-1,-1){8}} \put(93,48){\line(-1,-1){5.071}} \put(85,30){\line(0,1){5.858}} \put(85,30){\line(0,-1){5.858}} \put(95,10){\vector(1,-1){10}} \put(95,10){\line(-1,1){7.071}} \put(125,60){\line(0,-1){15.858}} \put(117,32){\line(1,1){5.071}} \put(113,28){\line(-1,-1){16}} \put(93,8){\vector(-1,-1){8}} \put(95,24.142){\arc{20}{2.356}{3.142}} \put(115,15.858){\arc{20}{5.498}{0}} \put(115,44.142){\arc{20}{0}{0.785}} \put(95,35.858){\arc{20}{3.142}{3.927}} \put(80,5){\makebox(0,0){$\rho$}} \put(110,5){\makebox(0,0){$\mu$}} \put(132,5){\makebox(0,0){$\lambda$}} \end{picture} \end{center} \caption{The braid relation as a vertical Reidemeister move of type III} \label{Reid3} \end{figure} The topological invariance gives us the freedom to write down the intertwiner algebraically from a given wire diagram: We can deform the wire diagram by finite sequences of the above moves and then split it in elementary wire diagrams --- in whatever way we decompose the wire diagrams into horizontal slices of elementary intertwiners, we always obtain the same intertwiner due to our topological invariance identities. Next we recall that we can write the statistics phase $\omega_\lambda$ as the intertwiner $d_\lambda r_\lambda^* \co\lambda(\eps\lambda\la)r_\lambda$. Therefore we obtain for $\omega_\lambda$ the wire diagram on the left-hand side of Fig.\ \ref{statph}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(120,40) \allinethickness{0.3pt} \put(10,20){\arc{20}{0.785}{5.498}} \put(24.142,20){\line(-1,1){7.071}} \put(22.142,18){\line(-1,-1){5.071}} \put(24.142,20){\line(1,-1){7.071}} \put(26.142,22){\line(1,1){5.071}} \put(24.142,34.142){\arc{20}{0}{0.785}} \put(24.142,5.858){\arc{20}{5.498}{6.283}} \put(34.142,40){\line(0,-1){5.858}} \put(34.142,5.858){\vector(0,-1){5.858}} \put(41.142,5){\makebox(0,0){$\lambda$}} \put(60,20){\makebox(0,0){$=$}} \put(110,20){\arc{20}{3.927}{2.357}} \put(95.858,20){\line(1,-1){7.071}} \put(97.858,22){\line(1,1){5.071}} \put(95.858,20){\line(-1,1){7.071}} \put(93.858,18){\line(-1,-1){5.071}} \put(95.858,34.142){\arc{20}{2.357}{3.142}} \put(95.858,5.858){\arc{20}{3.142}{3.927}} \put(85.858,34.142){\vector(0,1){5.858}} \put(85.858,0){\line(0,1){5.858}} \put(78.858,35){\makebox(0,0){$\lambda$}} \end{picture} \end{center} \caption{Statistics phase $\omega_\lambda$ as a ``twist''} \label{statph} \end{figure} The diagram on the right-hand side expresses that $\omega_\lambda$ can also be obtained as $d_\lambda \co\lambda(r_\lambda)^*\eps{\co\lambda}{\co\lambda} \co\lambda (r_\lambda)$. Note that we obtain the complex conjugate $\omega_\lambda^*$ by exchanging over- and undercrossings. Similarly, we recall that we can write a matrix element $Y_{\lambda,\mu}=Y_{\mu,\lambda}$ of Rehren's Y-matrix as $d_\lambda d_\mu \phi_\mu(\eps \lambda\mu \eps \mu\lambda)^* = d_\lambda d_\mu r_\mu^* \co\mu (\epsm \lambda\mu \epsm \mu\lambda) r_\mu$. Dividing by $d_\lambda$ we obtain $\chi_\lambda(\mu)$, the statistics character $\chi_\lambda$ evaluated on $\mu$, represented graphically by the wire diagram in Fig.\ \ref{statch}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(35,45) \allinethickness{0.3pt} \put(20,25){\arc{20}{5.012}{4.412}} \put(20,45){\line(0,-1){27}} \put(20,12){\vector(0,-1){12}} \put(10.12,27){\vector(0,1){0}} \put(5,20){\makebox(0,0){$\mu$}} \put(27,5){\makebox(0,0){$\lambda$}} \end{picture} \end{center} \caption{Rehren's statistics character $\chi_\lambda$ evaluated on $\mu$: $\chi_\lambda(\mu)$} \label{statch} \end{figure} We have drawn the circle $\mu$ symmetrically relative to the straight wire $\lambda$ because it does not make a difference whether we put the ``caps'' and ``cups'' for the isometry $r_\mu$ and its conjugate $r_\mu^*$ on the left or on the right due to the braiding fusion relations. As it is a scalar, we can write $Y_{\lambda,\mu}={\co r}_\mu^*Y_{\lambda,\mu}{\co r}_\mu$ and therefore its expression $d_\lambda d_\mu {\co r}_\mu^*r_\lambda^* \co\lambda (\epsm \mu\lambda \epsm \lambda\mu) r_\lambda {\co r}_\mu$ yields exactly the ``Hopf link'' as the wire diagram for the matrix element $Y_{\lambda,\mu}$, given by the left-hand side of Fig.\ \ref{Ymatrix}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(160,32) \allinethickness{0.3pt} \put(25,15){\arc{30}{5.742}{5.142}} \put(45,15){\arc{30}{2.601}{2.001}} \put(27,29.8){\vector(1,0){0}} \put(43,29.8){\vector(-1,0){0}} \put(5,25){\makebox(0,0){$\lambda$}} \put(65,25){\makebox(0,0){$\mu$}} \put(80,15){\makebox(0,0){$=$}} \put(115,15){\arc{30}{1.141}{0.541}} \put(135,15){\arc{30}{4.283}{3.683}} \put(117,29.8){\vector(1,0){0}} \put(137,29.8){\vector(1,0){0}} \put(95,25){\makebox(0,0){$\lambda$}} \put(155,25){\makebox(0,0){$\mu$}} \end{picture} \end{center} \caption{Matrix element $Y_{\lambda,\mu}$ of Rehren's Y-matrix as a ``Hopf link''} \label{Ymatrix} \end{figure} The equality to the right-hand side is just the relation $Y_{\lambda,\mu}=Y_{\lambda,\co\mu}^*$ together with the prescription of representing conjugates. Recall that if ${\Delta}$ is finite then the Y-matrix differs from the S-matrix just by an overall normalization factor $\sqrt w$, where $w$ is the global index. Often we consider intertwiners which are sums over intertwiners represented by the same wire diagram but the sum runs over one or more of the labels. Then we simply write the sum symbol in front of the diagram, we may similarly insert scalar factors. Now recall that for finite ${\Delta}$ the non-degeneracy of the braiding is encoded in the orthogonality relation $\langle y^0,y^\lambda \rangle = \del \lambda 0 w$. In terms of the statistics characters this reads $\sum_\mu d_\mu \chi_\lambda(\mu) =d_\lambda^{-1}\del \lambda 0 w=\del \lambda 0 w$. Graphically this can be represented as in Fig. \ref{ort0}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(195,45) \allinethickness{0.3pt} \put(10,22){\makebox(0,0){$\displaystyle\sum_{\mu\in{\Delta}}\,\,d_\mu$}} \put(45,25){\arc{20}{5.012}{4.412}} \put(45,45){\line(0,-1){27}} \put(45,12){\vector(0,-1){12}} \put(35.12,27){\vector(0,1){0}} \put(30,20){\makebox(0,0){$\mu$}} \put(52,5){\makebox(0,0){$\lambda$}} \put(70,24){\makebox(0,0){$=$}} \put(90,22){\makebox(0,0){$\displaystyle\sum_{\mu\in{\Delta}}$}} \put(120,25){\arc{20}{0}{6.283}} \put(110.12,27){\vector(0,1){0}} \put(105,20){\makebox(0,0){$\mu$}} \put(155,25){\arc{20}{5.012}{4.412}} \put(155,45){\line(0,-1){27}} \put(155,12){\vector(0,-1){12}} \put(145.12,27){\vector(0,1){0}} \put(140,20){\makebox(0,0){$\mu$}} \put(162,5){\makebox(0,0){$\lambda$}} \put(185,24){\makebox(0,0){$=\,\,\del \lambda 0 \, w$}} \end{picture} \end{center} \caption{Orthogonality relation for a non-degenerate braiding (``killing ring'')} \label{ort0} \end{figure} This kind of (graphical) relation has also been used more recently in \cite{TW,O6,KL} and was called a ``killing ring'' in \cite{O6}. Wire diagrams can also be used for intertwiners of morphisms between different factors. Let $A,B,C$ infinite factors, $\rho\in{\rm{Mor}}(A,B)$, $\sigma\in{\rm{Mor}}(C,B)$, $\tau\in{\rm{Mor}}(A,C)$ irreducible morphisms and $t\in{\rm{Hom}}(\rho,\sigma\tau)$ an isometry. Then Fig.\ \ref{tABC} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(60,40) \allinethickness{0.3pt} \dottedline{1}(15,25)(30,10)(45,25)(15,25) \put(30,30){\vector(0,-1){10}} \put(30,20){\vector(-1,-1){10}} \put(30,20){\vector(1,-1){10}} \put(10,30){\makebox(0,0){$B$}} \put(30,5){\makebox(0,0){$C$}} \put(50,30){\makebox(0,0){$A$}} \put(15,5){\makebox(0,0){$\sigma$}} \put(30,35){\makebox(0,0){$\rho$}} \put(45,5){\makebox(0,0){$\tau$}} \put(30,15){\makebox(0,0){$t$}} \end{picture} \end{center} \caption{The intertwiner $\protect\sqrt[4]{\frac{d_\sigma d_\tau}{d_\rho}}t$ as a triangle} \label{tABC} \end{figure} represents the intertwiner $d_\sigma^{1/4} d_\tau^{1/4} d_\rho^{-1/4} t$. Similarly we can draw a picture using a co-isometry. Along the lines of the previous paragraphs, we can similarly build up larger wire diagrams out of trivalent vertices involving different factors. We do not need the triangles with corners labelled by factors as we can also label the regions between the wires. So far we do not have a meaningful way to cross wires with differently labelled regions left and right, but all the arguments listed above which do not involve braidings can be used for intertwiners of morphisms between different factors exactly as proceeded above. Moreover, the diagrams may also involve wires where left and right regions are labelled by the same factor, i.e.\ these wires correspond to {\em endo}morphisms of some factor which may well form a braided system, and then one may have crossings for those wires. \subsection{Frobenius reciprocity and rotations} \label{graph2-prelim} Let $A,B,C$ be infinite factors, $\rho\in{\rm{Mor}}(A,B)$, $\tau\in{\rm{Mor}}(C,B)$, $\sigma\in{\rm{Mor}}(C,A)$ morphisms with finite statistical dimensions $d_\rho,d_\tau,d_\sigma<\infty$, respectively, and let $t\in{\rm{Hom}}(\tau,\rho\sigma)$. Then \[ {\cal{L}}_\rho(t) = \sqrt \frac{d_\rho d_\sigma}{d_\tau} \bar\rho(t)^* r_\rho \in {\rm{Hom}}(\sigma,\co\rho \tau) \] and \[ {\cal{R}}_\sigma(t) = \sqrt \frac{d_\rho d_\sigma}{d_\tau} t^* \rho({\co r}_\sigma) \in {\rm{Hom}}(\rho,\tau \co\sigma) \] are the images under left and right Frobenius maps. Displaying the intertwiners $d_\rho^{1/2} r_\rho^* \bar\rho(t)$ and $d_\sigma^{1/2} \rho({\co r}_\sigma)^*t$ graphically yields the identities in Figs.\ \ref{LFRt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(150,50) \allinethickness{0.3pt} \put(20,20){\dashbox{2}(40,20){$t$}} \put(40,50){\vector(0,-1){10}} \put(50,20){\vector(0,-1){20}} \put(30,20){\line(0,-1){5}} \put(10,15){\vector(0,1){35}} \put(20,15){\arc{20}{0}{3.142}} \put(5,45){\makebox(0,0){$\rho$}} \put(45,45){\makebox(0,0){$\tau$}} \put(55,5){\makebox(0,0){$\sigma$}} \put(80,30){\makebox(0,0){$=$}} \put(100,20){\dashbox{2}(40,20) {$\sqrt{\frac{d_\tau}{d_\sigma}} {\cal{L}}_\rho(t)^*$}} \put(110,40){\vector(0,1){10}} \put(130,50){\vector(0,-1){10}} \put(120,20){\vector(0,-1){20}} \put(105,45){\makebox(0,0){$\rho$}} \put(135,45){\makebox(0,0){$\tau$}} \put(125,5){\makebox(0,0){$\sigma$}} \end{picture} \end{center} \caption{Left Frobenius reciprocity for an intertwiner $t\in{\rm{Hom}}(\tau,\rho\sigma)$} \label{LFRt} \end{figure} and \ref{RFRt}, respectively. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(150,50) \allinethickness{0.3pt} \put(10,20){\dashbox{2}(40,20){$t$}} \put(30,50){\vector(0,-1){10}} \put(20,20){\vector(0,-1){20}} \put(40,20){\line(0,-1){5}} \put(60,15){\vector(0,1){35}} \put(50,15){\arc{20}{0}{3.142}} \put(15,5){\makebox(0,0){$\rho$}} \put(25,45){\makebox(0,0){$\tau$}} \put(65,45){\makebox(0,0){$\sigma$}} \put(80,30){\makebox(0,0){$=$}} \put(100,20){\dashbox{2}(40,20) {$\sqrt{\frac{d_\tau}{d_\rho}} {\cal{R}}_\sigma(t)^*$}} \put(110,50){\vector(0,-1){10}} \put(130,40){\vector(0,1){10}} \put(120,20){\vector(0,-1){20}} \put(125,5){\makebox(0,0){$\rho$}} \put(105,45){\makebox(0,0){$\tau$}} \put(135,45){\makebox(0,0){$\sigma$}} \end{picture} \end{center} \caption{Right Frobenius reciprocity for an intertwiner $t\in{\rm{Hom}}(\tau,\rho\sigma)$} \label{RFRt} \end{figure} These morphisms need not be irreducible. Taking them as products, we may replace any of them by bundles of wires. We call the linear isomorphisms $t\mapsto d_\rho^{1/2} r_\rho^* \bar\rho(t)$ and $t\mapsto d_\sigma^{1/2} \rho({\co r}_\sigma)^*t$ the left and right Frobenius rotations. Now let us assume that $t$ is isometric and labels a trivalent vertex of wires corresponding to irreducible morphisms $\rho,\tau,\sigma$. With the above ``transformation law'' we then have the identity of Fig.\ \ref{Ltight}, where the first equality is just a definition which gives us some prescription of ``tightening'' wires at trivalent vertices. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(170,40) \allinethickness{0.3pt} \put(20,20){\vector(0,-1){20}} \put(12.929,27.071){\line(1,-1){7.071}} \put(27.071,27.071){\vector(-1,-1){7.071}} \put(20,34.142){\arc{20}{2.356}{3.142}} \put(20,34.142){\arc{20}{0}{0.785}} \put(10,34.142){\vector(0,1){5.858}} \put(30,40){\line(0,-1){5.858}} \put(5,35){\makebox(0,0){$\rho$}} \put(35,35){\makebox(0,0){$\tau$}} \put(25,5){\makebox(0,0){$\sigma$}} \put(16,17){\makebox(0,0){$t$}} \put(50,20){\makebox(0,0){$:=$}} \put(70.858,20){\vector(0,1){20}} \put(80.858,20){\arc{20}{0.785}{3.142}} \put(95,40){\vector(0,-1){20}} \put(95,20){\line(-1,-1){7.071}} \put(95,20){\line(1,-1){7.071}} \put(95,5.858){\arc{20}{5.498}{0}} \put(105,5.858){\vector(0,-1){5.858}} \put(65,35){\makebox(0,0){$\rho$}} \put(100,35){\makebox(0,0){$\tau$}} \put(110,5){\makebox(0,0){$\sigma$}} \put(95,15){\makebox(0,0){$t$}} \put(125,20){\makebox(0,0){$=$}} \put(155,20){\vector(0,-1){20}} \put(147.929,27.071){\line(1,-1){7.071}} \put(162.071,27.071){\vector(-1,-1){7.071}} \put(155,34.142){\arc{20}{2.356}{3.142}} \put(155,34.142){\arc{20}{0}{0.785}} \put(145,34.142){\vector(0,1){5.858}} \put(165,40){\line(0,-1){5.858}} \put(140,35){\makebox(0,0){$\rho$}} \put(170,35){\makebox(0,0){$\tau$}} \put(160,5){\makebox(0,0){$\sigma$}} \put(156,30){\makebox(0,0){{\footnotesize ${\cal{L}}_\rho(t)^*$}}} \end{picture} \end{center} \caption{Left Frobenius reciprocity for a trivalent vertex labelled by an isometry} \label{Ltight} \end{figure} In fact, the label ${\cal{L}}_\rho(t)^*$ of the trivalent vertex makes sense since it is a co-isometry: Due to irreducibility of $\tau$ and $\sigma$, the map $t\mapsto{\cal{L}}_\rho(t)^*$ is isometric. Similarly, we get Fig.\ \ref{Rtight} (using irreducibility of $\tau$ and $\rho$). \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(170,40) \allinethickness{0.3pt} \put(20,20){\vector(0,-1){20}} \put(12.929,27.071){\vector(1,-1){7.071}} \put(27.071,27.071){\line(-1,-1){7.071}} \put(20,34.142){\arc{20}{2.356}{3.142}} \put(20,34.142){\arc{20}{0}{0.785}} \put(10,40){\line(0,-1){5.858}} \put(30,34.142){\vector(0,1){5.858}} \put(5,35){\makebox(0,0){$\tau$}} \put(35,35){\makebox(0,0){$\sigma$}} \put(25,5){\makebox(0,0){$\rho$}} \put(24,17){\makebox(0,0){$t$}} \put(50,20){\makebox(0,0){$:=$}} \put(104.142,20){\vector(0,1){20}} \put(94.142,20){\arc{20}{0}{2.356}} \put(80,40){\vector(0,-1){20}} \put(80,20){\line(-1,-1){7.071}} \put(80,20){\line(1,-1){7.071}} \put(80,5.858){\arc{20}{3.142}{3.927}} \put(70,5.858){\vector(0,-1){5.858}} \put(110,35){\makebox(0,0){$\sigma$}} \put(75,35){\makebox(0,0){$\tau$}} \put(65,5){\makebox(0,0){$\rho$}} \put(80,15){\makebox(0,0){$t$}} \put(125,20){\makebox(0,0){$=$}} \put(155,20){\vector(0,-1){20}} \put(147.929,27.071){\vector(1,-1){7.071}} \put(162.071,27.071){\line(-1,-1){7.071}} \put(155,34.142){\arc{20}{2.356}{3.142}} \put(155,34.142){\arc{20}{0}{0.785}} \put(145,40){\line(0,-1){5.858}} \put(165,34.142){\vector(0,1){5.858}} \put(140,35){\makebox(0,0){$\tau$}} \put(170,35){\makebox(0,0){$\sigma$}} \put(160,5){\makebox(0,0){$\rho$}} \put(156,30){\makebox(0,0){{\footnotesize ${\cal{R}}_\sigma(t)^*$}}} \end{picture} \end{center} \caption{Right Frobenius reciprocity for a trivalent vertex labelled by an isometry} \label{Rtight} \end{figure} Hence the prefactor in Figs.\ \ref{LFRt} and \ref{RFRt} is just such that it transforms isometries with natural normalization prefactors into co-isometries with natural normalization prefactors and, by taking adjoints, the other way round which gives the graphical identities given in Fig.\ \ref{RLtight*}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(240,40) \allinethickness{0.3pt} \put(20,40){\vector(0,-1){20}} \put(12.929,12.929){\vector(1,1){7.071}} \put(27.071,12.929){\line(-1,1){7.071}} \put(20,5.858){\arc{20}{3.142}{3.927}} \put(20,5.858){\arc{20}{5.498}{0}} \put(10,5.858){\line(0,-1){5.858}} \put(30,5.858){\vector(0,-1){5.858}} \put(5,5){\makebox(0,0){$\rho$}} \put(35,5){\makebox(0,0){$\tau$}} \put(25,35){\makebox(0,0){$\sigma$}} \put(15,23){\makebox(0,0){$t^*$}} \put(50,20){\makebox(0,0){$=$}} \put(80,40){\vector(0,-1){20}} \put(72.929,12.929){\vector(1,1){7.071}} \put(87.071,12.929){\line(-1,1){7.071}} \put(80,5.858){\arc{20}{3.142}{3.927}} \put(80,5.858){\arc{20}{5.498}{0}} \put(70,5.858){\line(0,-1){5.858}} \put(90,5.858){\vector(0,-1){5.858}} \put(65,5){\makebox(0,0){$\rho$}} \put(95,5){\makebox(0,0){$\tau$}} \put(85,35){\makebox(0,0){$\sigma$}} \put(80,10){\makebox(0,0){{\footnotesize ${\cal{L}}_\rho(t)$}}} \put(120,20){\makebox(0,0){and}} \put(160,40){\vector(0,-1){20}} \put(152.929,12.929){\line(1,1){7.071}} \put(167.071,12.929){\vector(-1,1){7.071}} \put(160,5.858){\arc{20}{3.142}{3.927}} \put(160,5.858){\arc{20}{5.498}{0}} \put(150,5.858){\vector(0,-1){5.858}} \put(170,5.858){\line(0,-1){5.858}} \put(145,5){\makebox(0,0){$\tau$}} \put(175,5){\makebox(0,0){$\sigma$}} \put(165,35){\makebox(0,0){$\rho$}} \put(165,23){\makebox(0,0){$t^*$}} \put(190,20){\makebox(0,0){$=$}} \put(220,40){\vector(0,-1){20}} \put(212.929,12.929){\line(1,1){7.071}} \put(227.071,12.929){\vector(-1,1){7.071}} \put(220,5.858){\arc{20}{3.142}{3.927}} \put(220,5.858){\arc{20}{5.498}{0}} \put(210,5.858){\vector(0,-1){5.858}} \put(230,5.858){\line(0,-1){5.858}} \put(205,5){\makebox(0,0){$\tau$}} \put(235,5){\makebox(0,0){$\sigma$}} \put(225,35){\makebox(0,0){$\rho$}} \put(220,10){\makebox(0,0){{\footnotesize ${\cal{R}}_\sigma(t)$}}} \end{picture} \end{center} \caption{Frobenius reciprocity for a trivalent vertex labelled by a co-isometry} \label{RLtight*} \end{figure} We may now use the replacement prescription three times, beginning with a trivalent vertex labelled by an isometry $t\in{\rm{Hom}}(\tau,\rho\sigma)$ and proceeding in a clockwise direction. Then we end up with a co-isometry $\Theta(t)^*\in{\rm{Hom}}(\co\sigma\co\rho,\co\tau)$ in the corner where we originally had the label $t$. In fact, \[ \Theta(t) = {\cal{R}}_\rho({\cal{L}}_\tau({\cal{R}}_\sigma(t))) = \sqrt{d_\rho d_\sigma d_\tau}\, r_\tau^* \co\tau (t^*\rho({\co r}_\sigma) {\co r}_\rho) \,. \] Similarly we can go in the counter-clockwise direction and then we obtain $\tilde\Theta(t)^*\in{\rm{Hom}}(\co\sigma\co\rho,\co\tau)$, where \[ \tilde\Theta(t) = {\cal{L}}_\sigma({\cal{R}}_\tau({\cal{L}}_\rho(t))) = \sqrt{d_\rho d_\sigma d_\tau}\, \co\sigma\co\rho({\co r}_\tau^* t^*) \co\sigma(r_\rho) r_\sigma \,, \] and in order to establish a well-defined rotation procedure we have to show that $\Theta(t)=\tilde\Theta(t)$. Now \[ \begin{array}{ll} \Theta(t)^* \tilde\Theta(t) &= \sqrt{d_\rho d_\sigma d_\tau} \, \co\tau({\co r}_\tau^* t^*) \Theta(t)^* \co\sigma(r_\rho) r_\sigma \\[.4em] &= d_\rho d_\sigma d_\tau \, \co\tau({\co r}_\tau^* t^*) \co\tau ({\co r}_\rho^* \rho({\co r}_\sigma^*)t) \co\tau\tau(\co\sigma(r_\rho) r_\sigma)r_\tau \\[.4em] &= d_\rho d_\sigma d_\tau \, \co\tau({\co r}_\tau^* t^* {\co r}_\rho^* \rho({\co r}_\sigma^*) \rho\sigma(\co\sigma(r_\rho) r_\sigma) t) r_\tau \\[.4em] &= d_\rho d_\tau \, \co\tau({\co r}_\tau^* t^* {\co r}_\rho^* \rho(r_\rho) t) r_\tau = d_\tau \, \co\tau({\co r}_\tau^*) r_\tau = {\bf1} \,, \end{array} \] hence $(\Theta(t)-\tilde\Theta(t))^*(\Theta(t)-\tilde\Theta(t))=0$, i.e.\ $\Theta(t)=\tilde\Theta(t)$. Thus a trivalent vertex labelled with an isometry $t\in{\rm{Hom}}(\tau,\rho\sigma)$ can equivalently be labelled with a co-isometry $\Theta(t)^*\in{\rm{Hom}}(\co\sigma\co\rho,\co\tau)$. So here we have established some ``rotation invariance'' of trivalent vertices (in standard inverted Y-shape or Y-shape) with a replacement prescription for the rotated labelling (co-) isometries. Next we turn to the rotation of crossings when we have a braiding. Assume we have a braided system of endomorphisms $\Delta\ni\lambda,\mu,\nu$ of some factor $A$. From the BFE we obtain $r_\lambda = \bar\lambda (\epsmp \mu\lambda) \epsmp \mu{\bar\lambda} \mu(r_\lambda)$. Applying $\lambda$ and multiplying by $d_\lambda \epspm\lambda\mu {\co r}_\lambda^*$ from the left yields \begin{equation} \epspm \lambda\mu = d_\lambda {\co r}_\lambda^* \lambda (\epsmp \mu{\bar\lambda}) \lambda\mu(r_\lambda)\,. \label{rotcreq} \end{equation} The BFE yields similarly $\lambda({\co r}_\mu)= \epspm \mu\lambda \mu(\epsmp{\co\mu}\lambda) {\co r}_\mu$, and by multiplying with $d_\mu \mu\lambda(r_\mu^*) \epspm\lambda\mu$ from the left we obtain \[ \epspm \lambda\mu = d_\mu \mu\lambda(r_\mu^*)\mu (\epsm{\co\mu}\lambda){\co r}_\mu \,,\] and therefore we have the graphical identity given in Fig.\ \ref{rotcross}, here displayed only for overcrossings. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(220,40) \allinethickness{0.3pt} \put(10.858,40){\line(0,-1){20}} \put(20.858,20){\arc{20}{0.785}{3.142}} \put(35,20){\line(-1,-1){7.071}} \put(35,20){\line(1,1){7.071}} \put(49.142,20){\arc{20}{3.926}{0}} \put(59.142,20){\vector(0,-1){20}} \put(25,40){\line(0,-1){5.858}} \put(33,22){\line(-1,1){5.071}} \put(37,18){\line(1,-1){5.071}} \put(35,34.142){\arc{20}{2.356}{3.142}} \put(45,5.858){\vector(0,-1){5.858}} \put(35,5.858){\arc{20}{5.498}{0}} \put(40,3.5){\makebox(0,0){$\mu$}} \put(64.142,5){\makebox(0,0){$\lambda$}} \put(80,20){\makebox(0,0){$=$}} \put(100,40){\line(0,-1){5.858}} \put(110,20){\line(-1,1){7.071}} \put(110,20){\line(1,-1){7.071}} \put(110,34.142){\arc{20}{2.356}{3.142}} \put(120,5.858){\vector(0,-1){5.858}} \put(110,5.858){\arc{20}{5.498}{0}} \put(120,40){\line(0,-1){5.858}} \put(112,22){\line(1,1){5.071}} \put(108,18){\line(-1,-1){5.071}} \put(110,34.142){\arc{20}{0}{0.785}} \put(100,5.858){\vector(0,-1){5.858}} \put(110,5.858){\arc{20}{3.142}{3.927}} \put(95,3.5){\makebox(0,0){$\mu$}} \put(125,5){\makebox(0,0){$\lambda$}} \put(140,20){\makebox(0,0){$=$}} \put(160.858,20){\vector(0,-1){20}} \put(170.858,20){\arc{20}{3.142}{5.498}} \put(183,22){\line(-1,1){5.071}} \put(187,18){\line(1,-1){5.071}} \put(199.142,20){\arc{20}{0}{2.356}} \put(209.142,40){\line(0,-1){20}} \put(185,20){\line(-1,-1){7.071}} \put(185,20){\line(1,1){7.071}} \put(195,40){\line(0,-1){5.858}} \put(185,34.142){\arc{20}{0}{0.785}} \put(175,5.858){\vector(0,-1){5.858}} \put(185,5.858){\arc{20}{3.142}{3.927}} \put(155.858,3.5){\makebox(0,0){$\mu$}} \put(180,5){\makebox(0,0){$\lambda$}} \end{picture} \end{center} \caption{Rotation of crossings} \label{rotcross} \end{figure} Then this procedure can even be iterated so that we obtain arbitrarily twisted crossings. Note that for the rotation of crossings we do not need any relabelling prescription as this is encoded in the BFE's. We now turn to the discussion of ``abstract pictures'' which admit different intertwiner interpretations according to Frobenius rotations. Let $A_1,A_2,...,A_\ell$ be factors equipped with sets ${\Delta}_{i,j}\subset{\rm{Mor}}(A_i,A_j)$, $i,j=1,2,...,\ell$, of irreducible, pairwise inequivalent morphisms with finite index such that $\bigsqcup_{i,j} {\Delta}_{i,j}$ is closed under conjugation and irreducible decomposition of products (whenever composable) as sectors, and in particular each ${\Delta}_{i,i}$ is a system of endomorphisms. Some of the systems ${\Delta}_{i,i}$ may be braided. We now consider ``labelled knotted graphs'' of the following form. On a finite connected and simply connected region in the plane we have a finite number of wires (i.e.\ images of piecewise $C^\infty$ maps from the unit interval into the region). Within the region there is a finite number of trivalent vertices (i.e.\ common endpoints of three wires) and crossings of two wires, and for the latter there is a notion of over- and undercrossing (i.e.\ for each crossing there is one wire ``on top of the other''). If wires are not closed (i.e.\ if their two endpoints do not coincide) then they are only allowed to have trivalent vertices or distinguished points on the boundary of the region as their endpoints. The wires meet each other only at the trivalent vertices and crossings, and they are directed and labelled by the morphisms in $\bigsqcup_{i,j} {\Delta}_{i,j}$ subject to the following rules. Crossings are only possible for wires with labelling morphisms in some ${\Delta}_{i,i}$ with braiding. Furthermore it must be possible to associate the factors $A_i$ to the free regions between the wires such that any wire labelled by some $\rho\in{\Delta}_{i,j}$ has the ``source'' factor $A_i$ on its left and the ``range'' factor $A_j$ on its right relative to the orientation (composition compatibility). We identify graphs which are transformed into each other by inversion of the orientation of a wire and simultaneous replacement of its label, say $\rho\in{\Delta}_{i,j}$, by the representative conjugate morphism $\bar\rho\in{\Delta}_{j,i}$. Finally, the trivalent vertices are labelled either by isometric or co-isometric intertwiners which are associated locally to one corner region of the trivalent vertex as follows. If $\tau\in{\Delta}_{i,j}$, $\rho\in{\Delta}_{k,j}$, $\sigma\in{\Delta}_{i,k}$ label the three wires of a trivalent vertex, $\tau$ is entering and, following counter-clockwise, $\rho$ and $\sigma$ are outgoing (as e.g.\ the trivalent vertex in Figs.\ \ref{Ltight} and \ref{Rtight}, possibly up to isotopy and rotation), then in the local corner region opposite to $\tau$ the label must either be an isometry $t\in{\rm{Hom}}(\tau,\rho\sigma)$ or a co-isometry $s^*\in{\rm{Hom}}(\co\sigma\co\rho,\co\tau)$. If the wires at a trivalent vertex have orientation different from this, the rule can be derived from the previous case by reversing orientations and simultaneous relabelling by conjugate morphisms. Now let ${\cal{G}}$ be such a labelled knotted graph as above. To interpret ${\cal{G}}$ as an intertwiner, we may put it in some ``Frobenius annulus'' as shown in Fig.\ \ref{annul} for an example.\footnote{Our notion of a Frobenius annulus is inspired by the annular invariance used in Jones' definition of a ``general planar algebra'' \cite{J2}.} \begin{figure}[htb] \begin{center} \unitlength 0.4mm \begin{picture}(150,115) \allinethickness{0.3pt} \put(5,10){\dashbox{2}(145,95){}} \put(55,50){\arc{30}{1.571}{4.712}} \put(85,50){\arc{30}{4.712}{1.571}} \put(55,35){\line(1,0){30}} \put(55,65){\line(1,0){30}} \put(115,65){\arc{60}{3.142}{0}} \put(40,65){\arc{30}{1.571}{3.142}} \put(40,65){\vector(0,1){50}} \put(25,65){\line(0,1){50}} \put(100,35){\arc{30}{4.712}{0}} \put(55,65){\vector(0,1){50}} \put(70,115){\vector(0,-1){50}} \put(55,0){\vector(0,1){35}} \put(70,0){\vector(0,1){35}} \put(85,0){\vector(0,1){35}} \put(50,65){\arc{20}{2.51}{3.142}} \put(90,65){\arc{20}{0}{0.632}} \put(90,35){\arc{20}{5.652}{0}} \put(50,35){\arc{20}{3.142}{3.773}} \put(115,65){\arc{30}{3.142}{0}} \put(90,65){\arc{20}{0}{0.632}} \put(25,35){\arc{30}{0}{3.142}} \put(85,0){\vector(0,1){35}} \put(10,35){\vector(0,1){80}} \put(100,35){\vector(0,-1){35}} \put(115,0){\line(0,1){35}} \put(130,65){\vector(0,-1){65}} \put(145,65){\vector(0,-1){65}} \put(40,50){\vector(1,0){0}} \put(100,50){\vector(-1,0){0}} \put(70,50){\makebox(0,0){${\cal{G}}$}} \put(4,110){\makebox(0,0){{\footnotesize $\rho_2$}}} \put(20,110){\makebox(0,0){{\footnotesize $\rho_1$}}} \put(47.5,110){\makebox(0,0){{\footnotesize $\rho_{12}$}}} \put(63,110){\makebox(0,0){{\footnotesize $\rho_{11}$}}} \put(77,110){\makebox(0,0){{\footnotesize $\rho_{10}$}}} \put(151,5){\makebox(0,0){{\footnotesize $\rho_9$}}} \put(136,5){\makebox(0,0){{\footnotesize $\rho_8$}}} \put(120,5){\makebox(0,0){{\footnotesize $\rho_7$}}} \put(50,5){\makebox(0,0){{\footnotesize $\rho_3$}}} \put(65,5){\makebox(0,0){{\footnotesize $\rho_4$}}} \put(80,5){\makebox(0,0){{\footnotesize $\rho_5$}}} \put(94,5){\makebox(0,0){{\footnotesize $\rho_6$}}} \end{picture} \end{center} \caption{A Frobenius annulus surrounding ${\cal{G}}$} \label{annul} \end{figure} A Frobenius annulus has labelled wires inside such that each of them meets an open end of a wire of ${\cal{G}}$ at one endpoint (labelled by $\rho_1$,...,$\rho_{12}$ in our example), matching the label and orientation of this wire, and this way all the open ends of the wires of ${\cal{G}}$ are either connected to the top or bottom of the outside square boundary of the annulus. No crossings or trivalent vertices are allowed in the annulus, but it may contain cups or caps. Gluing the wires together and forgetting about the boundary of ${\cal{G}}$ and the annulus, we will read the result as a wire diagram and therefore the annulus corresponds to a ``Frobenius choice'', deciding whether we will get a certain intertwiner or its image by certain Frobenius rotations, cf.\ Figs.\ \ref{LFRt} and \ref{RFRt} (and their adjoints). Reading vertically downwards, we may now have the problem that on a finite number of horizontal levels a finite number of singular points of crossings, trivalent vertices, cups and caps are exactly on the same level (or ``height'') so that we cannot time slice the diagram into stripes containing only one elementary intertwiner. Also some wires may have pieces going exactly horizontally. We now allow to make small vertical translations such that these crossings and trivalent vertices are put on slightly different levels and all wires obtain piecewise slopes, without letting wires touch or producing new crossings, but we may possibly produce some new cups or caps. In the latter case we can always arrange it so that even each new cup or cap appears on a distinct level. The trivalent vertices and crossings may not be in ``standard form'', i.e.\ in Y- or inverted Y shape respectively X-shape. In an ``$\epsilon$-neighborhood'' of a trivalent vertex, we now bend the wires so that the angles are arranged in standard form. Similarly we modify the crossings to bend them into an X-shape. Using for labels at trivalent vertices our replacement prescription by Frobenius reciprocity, we can obtain isometries as labels for trivalent vertices in inverted Y-shape, located on the bottom corner region, and co-isometries as labels for trivalent vertices in Y-shape, located on the top corner region. Again, these topological moves are allowed to produce at most new cups or caps, all on different levels so that the resulting diagram can be time sliced into stripes of elementary diagrams. Clearly, this procedure of deforming a labelled knotted graph in a Frobenius annulus into a regular wire diagram is highly ambiguous. However, the ambiguities in the above procedures are irrelevant: The ambiguities arising from the production of slopes of wires and different levels of certain elementary intertwiners are irrelevant due to the topological invariance of Fig.\ \ref{isoinv1} and the freedom of translating intertwiners vertically as shown in Fig.\ \ref{xx'}, and the ambiguities arising from rotations of the elementary intertwiners are irrelevant due to the rotation invariance of trivalent vertices and crossings, as we have established in Figs.\ \ref{Ltight}, \ref{Rtight}, \ref{RLtight*} and \ref{rotcross}. Now let ${\cal{G}}_1$ and ${\cal{G}}_2$ be two labelled knotted graphs as above which are defined on the same (connected, simply connected) region in the plane and have the same entering and outgoing wires at the same points with the same orientation, i.e.\ they have coinciding open ends so that they fit in the same Frobenius annuli. When embedded in some Frobenius annulus it may now happen that the corresponding intertwiners are the same, even if ${\cal{G}}_1$ and ${\cal{G}}_2$ are different. Because of the isomorphism property of Frobenius rotations it is clear that then ${\cal{G}}_1$ and ${\cal{G}}_2$ yield the same intertwiner through embedding in any Frobenius annulus. We can write down sufficient conditions for such equality in terms of some ``regular isotopy'': For given ${\cal{G}}_1$ and ${\cal{G}}_2$ as above choose a Frobenius annulus and regularize the pictures into two wire diagrams ${\cal{W}}_1$ and ${\cal{W}}_2$, respectively. We call ${\cal{G}}_1$ and ${\cal{G}}_2$ regularly isotopic if ${\cal{W}}_1$ can be transformed into ${\cal{W}}_2$ by the following list of moves: \begin{enumerate} \item Reversing orientation of some wires with simultaneous relabelling by conjugate morphisms, \item any horizontal translations of elementary intertwiners which may change slopes of wires but which do not let the wires meet or involve cups or caps, \item vertical translations of elementary intertwiners as in Fig.\ \ref{xx'}, \item topological moves as in Fig.\ \ref{isoinv1}, \item rotations of trivalent vertices and their labels as in Figs.\ \ref{Ltight}, \ref{Rtight}, \ref{RLtight*}, \item and for wires corresponding to a braided system ${\Delta}_{i,i}$ we additionally admit \begin{enumerate} \item vertical Reidemeister moves of type II as in Fig.\ \ref{Reid2}, \item moving crossings over and under trivalent vertices, cups and caps according to the BFE's (cf.\ Figs.\ \ref{wireBFE1} and \ref{wireBFE2} for the first two relations), \item vertical Reidemeister moves of type III for crossings (cf.\ Fig.\ \ref{Reid3} for overcrossings), \item rotations of crossings (cf.\ Fig. \ref{rotcross} for overcrossings). \end{enumerate} \end{enumerate} Thus the ambiguity in the regularization procedure means in particular that from one graph we can only obtain wire diagrams that can be transformed into each other by these moves. It is easy to see that regular isotopy is an equivalence relation for knotted labelled graphs. Moreover, for closed labelled knotted graphs (i.e.\ without open ends) which are then embedded in a trivial annulus, the local rotation invariance of the elementary intertwiners ends up in a total rotation invariance: We can rotate the picture freely, the rotated graph is always regularly isotopic to the original one and we will always end up with the same scalar (times ${\bf1}_{A_i}$, where $A_i$ is the factor associated to the outside region).\footnote{For a single kind of wire corresponding to a braided system, this invariance is similar to the complex number-valued regular isotopy invariant of knotted graphs obtained in \cite{MO}.} Let us finally consider an intertwiner $x\in{\rm{Hom}}(\rho,\rho)$ with $\rho\in{\rm{Mor}}(A,B)$ irreducible. Then clearly $x$ is a scalar: $x=\xi{\bf1}_B$, $\xi\in\Bbb{C}$. Hence we have the identity $d_\rho \xi{\bf1}_B \equiv d_\rho x = d_\rho {\co r}_\rho^* x {\co r}_\rho$, and this is graphically the left-hand side in Fig.\ \ref{cut}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(160,30) \allinethickness{0.3pt} \put(5,15){\makebox(0,0){$d_\rho$}} \put(15,10){\dashbox{2}(10,10){$x$}} \put(20,30){\vector(0,-1){10}} \put(20,10){\vector(0,-1){10}} \put(25,26){\makebox(0,0){$\rho$}} \put(25,4){\makebox(0,0){$\rho$}} \put(40,15){\makebox(0,0){$=$}} \put(50,10){\dashbox{2}(10,10){$x$}} \put(65,20){\arc{20}{3.142}{0}} \put(65,10){\arc{20}{0}{3.142}} \put(75,10){\line(0,1){10}} \put(75,17){\vector(0,1){0}} \put(79,5){\makebox(0,0){$\rho$}} \put(105,15){\makebox(0,0){$\longleftrightarrow$}} \put(135,10){\line(0,1){10}} \put(135,17){\vector(0,1){0}} \put(145,20){\arc{20}{3.142}{0}} \put(145,10){\arc{20}{0}{3.142}} \put(150,10){\dashbox{2}(10,10){$x$}} \put(131,5){\makebox(0,0){$\rho$}} \end{picture} \end{center} \caption{Two intertwiners of the same scalar value} \label{cut} \end{figure} On the other hand, application of the left inverse yields $d_\rho \phi_\rho(x)=d_\rho r_\rho^*\co\rho(x)r_\rho=d_\rho\xi{\bf1}_A$, which is a different intertwiner of the same scalar value, and it is represented graphically by the right-hand side in Fig.\ \ref{cut}. Thus the left- and right-hand side in Fig.\ \ref{cut} represent the same scalar. If we consider closed wire diagrams and are only interested in the scalars they represent, then we therefore have a ``regular isotopy on the 2-sphere''. \subsection{$\alpha$-Induction for braided subfactors} \label{sec-aindbsf} We now consider $\alpha$-induction of \cite{BE1,BE2,BE3} in the setting of braided subfactors. Here we work with a type III subfactor $N\subset M$, equipped with a braided system ${\Delta}\subset{\rm{End}}(N)$ in the sense of Definition \ref{system} such that for the injection map $\iota:N\to M$, the sector $[\co\iota\iota]$ decomposes into a finite sum of sectors of morphisms in ${\Delta}$. (Here $\co\iota$ denotes any choice of a representative morphism for the conjugate sector of $[\iota]$.) Note that since elements in ${\Delta}$ have by definition finite statistical dimension, it follows that the injection map has finite statistical dimension and thus the subfactor $N\subset M$ has finite index. But also note that we did neither assume the finite depth condition on $N\subset M$ (we did not assume finiteness of ${\Delta}$) nor non-degeneracy of the braiding at this point. As usual, we denote the canonical endomorphism $\iota\co\iota\in{\rm{End}}(M)$ by $\gamma=\iota\co\iota$, the dual canonical endomorphism $\co\iota\iota\in{\rm{End}}(N)$ by $\theta=\co\iota\iota$ and ``canonical'' isometries by $v\in M$ and $w\in N$, more precisely, we have $v\in{\rm{Hom}}({\rm{id}}_M,\gamma)$ and $w\in{\rm{Hom}}({\rm{id}}_N,\theta)$ such that $w^*v=\gamma(v^*)w=[M:N]^{-1/2}{\bf1}$. Recall that we have pointwise equality $M=Nv$. With a braiding $\varepsilon$ on ${\Delta}$ and its extension to $\Sigma({\Delta})$ as in Subsection \ref{braid-prelim} we can define the $\alpha$-induced $\alpha^\pm_\lambda$ for $\lambda\in{\Sigma({\Delta})}$ exactly as in \cite{LR,BE1}, namely we define \[ \alpha_\lambda^\pm = \co\iota^{\,-1} \circ {\rm{Ad}} (\epspm \lambda\theta) \circ \lambda \circ \co\iota \,. \] Then $\alpha^+_\lambda$ and $\alpha^-_\lambda$ are morphisms in ${\rm{Mor}}(M,M)$ with the properties $\alpha^\pm_\lambda\circ\iota=\iota\circ\lambda$, $\alpha^\pm_\lambda(v)=\epspm \lambda\theta ^* v$, $\alpha_{\lambda\mu}^\pm=\alpha_\lambda^\pm \alpha_\mu^\pm$ if also $\mu\in{\Sigma({\Delta})}$, and clearly $\alpha_{{\rm{id}}_N}^\pm={{\rm{id}}}_M$. Note that the first property yields immediately $d_{\alpha_\lambda^\pm}=d_\lambda$ by the multiplicativity of the minimal index \cite{L2.5}. We also obtain easily that $\overline{\alpha_\lambda^\pm}=\alpha_{\co\lambda}^\pm$, since we obtain $r_\lambda=\epspm \theta{\co\lambda\la}\theta(r_\lambda)$, and similarly ${\co r}_\lambda=\epspm \theta {\lambda\co\lambda} \theta({\co r}_\lambda)$ easily from \erf{nat}. Multiplying both relations by $v$ from the right yields $r_\lambda v= \alpha_{\co\lambda}^\pm \alpha_\lambda^\pm(v) r_\lambda$ and ${\co r}_\lambda v= \alpha_\lambda^\pm \alpha_{\co\lambda}^\pm(v) {\co r}_\lambda$, hence $r_\lambda\in{\rm{Hom}}({\rm{id}}_M,\alpha_{\co\lambda}^\pm\alpha_\lambda^\pm)$, ${\co r}_\lambda\in{\rm{Hom}}({\rm{id}}_M,\alpha_\lambda^\pm\alpha_{\co\lambda}^\pm)$ as $M=Nv$, thus we can put $R_{\alpha_\lambda^\pm}=\iota(r_\lambda)$, ${\co R}_{\alpha_\lambda^\pm}=\iota({\co r}_\lambda)$ as R-isometries for the $\alpha$-induced morphisms, i.e.\ $\overline{\alpha_\lambda^\pm}=\alpha_{\co\lambda}^\pm$. Note also that the definition of $\alpha^\pm_\lambda$ does not depend on the choice of the representative morphism $\co\iota$ for the conjugate sector of $[\iota]$ due to the transformation properties of the braiding operators, \erf{eun}. Though the local net structure for $N(I)\subset M(I)$ is assumed in \cite{LR,BE1}, we need only an assumption of a braiding for the definition of $\alpha^\pm_\lambda$. We, however, have to be careful, because we do not assume the chiral locality condition $\eps \theta\canr \gamma(v) = \gamma(v)$ in this paper. (The name ``chiral locality'' is motivated from the treatment of extensions of chiral observables in conformal field theory in the setting of nets of subfactors \cite{LR}, where the extended net is shown to satisfy local commutativity if and only if the condition $\eps \theta\canr \gamma(v) = \gamma(v)$ is met \cite[Thm.\ 4.9]{LR}.) Some theorems in \cite{BE1,BE2,BE3} do depend on the chiral locality condition and are {\sl not} true in this more general setting of $\alpha$-induction. Namely, with $\eps \theta\canr \gamma(v) = \gamma(v)$ it was easily derived \cite[Lemma 3.5]{BE1} by using the BFE that then ${\rm{Hom}}(\alpha_\lambda^\pm,\alpha_\mu^\pm)={\rm{Hom}}(\iota\lambda,\iota\mu)$ for $\lambda,\mu\in{\Sigma({\Delta})}$. As a surprising corollary (cf.\ \cite[Cor.\ 3.6]{BE1}) one found by putting $\lambda=\mu={\rm{id}}_N$ that $\iota$, thus the subfactor $N\subset M$, was irreducible which had not been assumed. Another corollary was then the ``main formula'' \cite[Thm.\ 3.9]{BE1}, giving $\langle\alpha_\lambda^\pm,\alpha_\mu^\pm\rangle=\langle\iota\lambda,\iota\mu\rangle =\langle\theta\lambda,\mu\rangle$ by Frobenius reciprocity. (Moreover, in the framework of {\it nets of} subfactors ${\cal{N}}\subset{\cal{M}}$, where the braidings arise from the transportability of localized endomorphisms, a certain reciprocity formula $\langle \alpha_\lambda^\pm,\beta \rangle=\langle \lambda,\sigma_\beta\rangle$, called ``$\alpha\sigma$-reciprocity'', between localized transportable endomorphisms $\lambda$ and $\beta$ of the smaller respectively the larger net was established; here $\sigma$-restriction is essentially $\sigma_\beta=\co\iota\beta\iota$.) Without chiral locality, these results are in general not true: The subfactor $N\subset M$ is neither forced to be irreducible, nor does the main formula hold, however, we always have the inequality $\langle\alpha_\lambda^\pm,\alpha_\mu^\pm\rangle\le\langle\theta\lambda,\mu\rangle$, since only the ``$\ge$'' part of the proof of \cite[Thm.\ 3.9]{BE1} uses chiral locality. It is a simple application of the braiding fusion equation and does not involve chiral locality that for $\lambda,\mu,\nu\in{\Sigma({\Delta})}$ we have the (equivalent) relations \cite[Lemma 3.25]{BE1} \begin{equation} \alpha^\mp_\rho(Q) \epspm \lambda\rho = \epspm \mu\rho Q \,,\qquad Q \epspm \rho\lambda = \epspm \rho\mu \alpha^\pm_\rho (Q) \label{anat1} \end{equation} whenever $Q\in{\rm{Hom}}(\iota\lambda,\iota\mu)$. Let $a\in{\rm{Mor}}(M,N)$ be such that $[a]$ is a subsector of $[\mu\bar\iota]$ for some $\mu\in{\Sigma({\Delta})}$. Hence $a\iota\in{\Sigma({\Delta})}$. Similarly, let $\co b \in{\rm{Mor}}(N,M)$ be such that $[\bar b]$ is a subsector of $[\iota{\co\nu}]$ for some ${\co\nu}\in{\Sigma({\Delta})}$. If $T\in{\rm{Hom}}(\bar b,\iota{\co\nu})$ is an isometry we put \[ \Epspm \lambda {\bar b} = T^* \epspm \lambda{\co\nu} \alpha^\pm_\lambda (T) \,, \qquad \Epspm {\bar b}\lambda = (\Epsmp \lambda {\bar b})^* \,. \] Note that the definition is independent of the choice of $T$ and ${\co\nu}$ in the following sense: If also $S\in{\rm{Hom}}(\bar b,\iota{\co\tau})$ is an isometry for some ${\co\tau}\in{\Sigma({\Delta})}$ then $ST^*\in{\rm{Hom}}(\iota{\co\nu},\iota{\co\tau})$ and therefore \[ \Epspm \lambda {\bar b} = S^*ST^* \epspm \lambda{\co\nu} \alpha^\pm_\lambda (T) = S^* \epspm \lambda{{\co\tau}} \alpha^\pm_\lambda (ST^*T) = S^* \epspm \lambda{{\co\tau}} \alpha^\pm_\lambda (S) \,. \] Similarly one easily checks that $\Epspm \lambda {\bar b}$ is unitary. \begin{proposition} \label{stat} Let $\lambda\in{\Sigma({\Delta})}$, let $a\in{\rm{Mor}}(M,N)$ be such that $[a]$ is a subsector of $[\mu\bar\iota]$ for some $\mu\in{\Sigma({\Delta})}$ and let $\co b \in{\rm{Mor}}(N,M)$ be such that $[\bar b]$ is a subsector of $[\iota{\co\nu}]$ for some ${\co\nu}\in{\Sigma({\Delta})}$. Then we have \begin{equation} \epspm \lambda {a\iota} \in {\rm{Hom}}(\lambda a, a \alpha_\lambda^\pm) \,,\qquad \Epspm \lambda{\bar b} \in {\rm{Hom}}(\alpha^\pm_\lambda \bar b , \bar b\lambda) \,. \label{genstat} \end{equation} \end{proposition} \begin{proof} The first relation in \erf{genstat} is trivial on $N$, so we only need to show it for $v$ since $M=Nv$. Note that $a(v)\in{\rm{Hom}}(a\iota,a\iota\theta)$, therefore \erf{BFE} yields \[ a(v) \epspm \lambda {a\iota} = a\iota(\epspm \lambda\theta) \epspm \lambda {a\iota} \lambda(a(v)) \,, \] hence \[ a \circ\alpha^\pm_\lambda (v) = a\iota(\epspm \lambda\theta ^*) a(v) = \epspm \lambda {a\iota} \lambda (a(v)) \epspm \lambda {a\iota} ^* = {\rm{Ad}}\;\epspm \lambda{a\iota}\circ\lambda\circ a (v) \,. \] For the second relation we use the fact that $TT^*\in{\rm{Hom}}(\iota{\co\nu},\iota{\co\nu})$ for $T\in{\rm{Hom}}(\bar b,\iota{\co\nu})$: \[ \begin{array}{ll} \Epspm \lambda {\bar b} \alpha_\lambda^\pm \bar b(n) &= T^* \epspm \lambda{\co\nu} \alpha_\lambda^\pm(TT^*{\co\nu}(n)T) = T^* \epspm \lambda{\co\nu} \lambda{\co\nu}(n) \alpha^\pm_\lambda (T) \\[.4em] &= T^* {\co\nu}\lambda(n) \epspm \lambda{\co\nu} \alpha^\pm_\lambda (T) = \bar b \lambda (n) \Epspm \lambda {\bar b} \end{array} \] for all $n\in N$. \end{proof} Due to Prop.\ \ref{stat} we can now draw the pictures in Fig.\ \ref{NMbraid} for the operators $\epspm \lambda {a\iota}$ and $\Epspm \lambda {\bar b}$. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(220,40) \allinethickness{0.3pt} \put(10,40){\line(0,-1){5.858}} \put(16.5,23.5){\vector(1,-1){0}} \put(20,20){\line(-1,1){7.071}} \put(20,34.142){\arc{20}{2.356}{3.142}} \allinethickness{2.0pt} \put(20,20){\line(1,-1){7.071}} \put(30,0){\line(0,1){5.858}} \put(30,0){\vector(0,-1){0}} \put(20,5.858){\arc{20}{5.498}{0}} \allinethickness{1.0pt} \put(30,40){\line(0,-1){5.858}} \put(22,22){\line(1,1){5.071}} \put(18,18){\line(-1,-1){5.071}} \put(20,34.142){\arc{20}{0}{0.785}} \put(10,0){\line(0,1){5.858}} \put(10,0){\vector(0,-1){0}} \put(20,5.858){\arc{20}{3.142}{3.927}} \put(5,3.5){\makebox(0,0){$a$}} \put(36,4){\makebox(0,0){$\alpha_\lambda^+$}} \put(5,36){\makebox(0,0){$\lambda$}} \put(35,35){\makebox(0,0){$a$}} \put(50,20){\makebox(0,0){$;$}} \allinethickness{0.3pt} \put(70,40){\line(0,-1){5.858}} \put(76.5,23.5){\vector(1,-1){0}} \put(78,22){\line(-1,1){5.071}} \put(80,34.142){\arc{20}{2.356}{3.142}} \allinethickness{2.0pt} \put(82,18){\line(1,-1){5.071}} \put(90,0){\line(0,1){5.858}} \put(90,0){\vector(0,-1){0}} \put(80,5.858){\arc{20}{5.498}{0}} \allinethickness{1.0pt} \put(90,40){\line(0,-1){5.858}} \put(80,20){\line(1,1){7.071}} \put(80,20){\line(-1,-1){7.071}} \put(80,34.142){\arc{20}{0}{0.785}} \put(70,0){\line(0,1){5.858}} \put(70,0){\vector(0,-1){0}} \put(80,5.858){\arc{20}{3.142}{3.927}} \put(65,3.5){\makebox(0,0){$a$}} \put(96,4){\makebox(0,0){$\alpha_\lambda^-$}} \put(65,36){\makebox(0,0){$\lambda$}} \put(95,35){\makebox(0,0){$a$}} \put(110,20){\makebox(0,0){$;$}} \allinethickness{2.0pt} \put(130,40){\line(0,-1){5.858}} \put(136.5,23.5){\vector(1,-1){0}} \put(140,20){\line(-1,1){7.071}} \put(140,34.142){\arc{20}{2.356}{3.142}} \allinethickness{0.3pt} \put(140,20){\line(1,-1){7.071}} \put(150,0){\line(0,1){5.858}} \put(150,0){\vector(0,-1){0}} \put(140,5.858){\arc{20}{5.498}{0}} \allinethickness{1.0pt} \put(150,40){\line(0,-1){5.858}} \put(150,40){\vector(0,1){0}} \put(142,22){\line(1,1){5.071}} \put(138,18){\line(-1,-1){5.071}} \put(140,34.142){\arc{20}{0}{0.785}} \put(130,0){\line(0,1){5.858}} \put(140,5.858){\arc{20}{3.142}{3.927}} \put(125,4){\makebox(0,0){$b$}} \put(156,4){\makebox(0,0){$\lambda$}} \put(125,35){\makebox(0,0){$\alpha_\lambda^+$}} \put(155,35.5){\makebox(0,0){$b$}} \put(170,20){\makebox(0,0){$;$}} \allinethickness{2.0pt} \put(190,40){\line(0,-1){5.858}} \put(196.5,23.5){\vector(1,-1){0}} \put(198,22){\line(-1,1){5.071}} \put(200,34.142){\arc{20}{2.356}{3.142}} \allinethickness{0.3pt} \put(202,18){\line(1,-1){5.071}} \put(210,0){\line(0,1){5.858}} \put(210,0){\vector(0,-1){0}} \put(200,5.858){\arc{20}{5.498}{0}} \allinethickness{1.0pt} \put(210,40){\line(0,-1){5.858}} \put(210,40){\vector(0,1){0}} \put(200,20){\line(1,1){7.071}} \put(200,20){\line(-1,-1){7.071}} \put(200,34.142){\arc{20}{0}{0.785}} \put(190,0){\line(0,1){5.858}} \put(200,5.858){\arc{20}{3.142}{3.927}} \put(185,4){\makebox(0,0){$b$}} \put(216,4){\makebox(0,0){$\lambda$}} \put(185,35){\makebox(0,0){$\alpha_\lambda^-$}} \put(215,35.5){\makebox(0,0){$b$}} \end{picture} \end{center} \caption{Wire diagrams for $\epsp \lambda {a\iota}$, $\epsm \lambda {a\iota}$, $\Epsp \lambda {\bar b}$, $\Epsm \lambda {\bar b}$, respectively} \label{NMbraid} \end{figure} The pictures for their conjugates $\epsmp {a\iota}\lambda$ and $\Epsmp {\bar b}\lambda$ are as usual obtained by horizontal reflection and inversion of arrows of the pictures in Fig.\ \ref{NMbraid}. \begin{lemma} \label{Enat} Let $\co a,\co b\in{\rm{Mor}}(M,N)$ be such that $[\co a]$ and $[\co b]$ are subsectors of $[\iota\co\mu]$ and $[\iota{\co\nu}]$ for some $\co\mu,{\co\nu}\in{\Sigma({\Delta})}$, respectively. Whenever $Y\in{\rm{Hom}}(\co a,\co b)$ we have \[ \alpha_\rho^\mp(Y) \, \Epspm {\co a}\rho = \Epspm {\co b}\rho \, Y \,,\qquad Y \, \Epspm \rho{\co a} = \Epspm \rho {\co b} \, \alpha_\rho^\pm (Y) \,. \] \end{lemma} \begin{proof} Let $S\in{\rm{Hom}}(\co a,\iota\co\mu)$ and $T\in{\rm{Hom}}(\co b,\iota\co\nu)$ be isometries. Then $\Epspm{\co a}\rho=\alpha_\rho^\mp(S)^* \epspm {\co\mu}\rho S$ and $\Epspm \rho{\co b}= T^* \epspm \rho{\co\nu} \alpha_\rho^\pm(T)$. Now $TYS^*\in{\rm{Hom}}(\iota\co\mu,\iota\co\nu)$. Inserting this in \erf{anat1} yields the statement. \end{proof} In order to establish a symmetry for ``moving crossings over trivalent vertices'' we can now state the following \begin{proposition} \label{intbfe} Let $\lambda,\rho\in{\Sigma({\Delta})}$, let $a,b\in{\rm{Mor}}(M,N)$ be such that $[a]$ and $[b]$ are subsectors of $[\mu\bar\iota]$ and $[\nu\bar\iota]$ for some $\mu,\nu\in{\Sigma({\Delta})}$ and let $\co a,\co b \in{\rm{Mor}}(N,M)$ be conjugates, respectively. Whenever $t\in{\rm{Hom}}(\lambda,a\bar b)$, $x\in{\rm{Hom}}(a,\lambda b)$ and $Y\in{\rm{Hom}}(\co a,\co b \lambda)$, we have the intertwining braiding fusion equations (IBFE's): \begin{eqnarray} \rho(t) \, \epspm \lambda\rho &=& \epspm {a\iota}\rho \, a(\Epspm {\bar b}\rho) \, t \,, \label{ibfe1} \\ t \, \epspm \rho\lambda &=& a (\Epspm \rho{\bar b}) \, \epspm \rho{a\iota} \, \rho (t) \,, \label{ibfe2} \\ \rho(x) \, \epspm {a\iota} \rho &=& \epspm \lambda\rho \, \lambda(\epspm {b\iota}\rho) \, x \,, \label{ibfe3} \\ x \, \epspm \rho {a\iota} &=& \lambda (\epspm \rho{b\iota}) \, \epspm \rho\lambda \, \rho (x) \,,\label{ibfe4} \\ \alpha_\rho^\mp(Y) \, \Epspm {\co a}\rho &=& \Epspm {\co b}\rho \, \co b(\epspm \lambda\rho) \, Y \,, \label{ibfe5} \\ Y \, \Epspm \rho{\co a} &=& \bar b (\epspm \rho\lambda) \, \Epspm \rho{\bar b} \, \alpha_\rho^\pm (Y) \label{ibfe6} \,. \end{eqnarray} \end{proposition} \begin{proof} Since $[\co b]$ must be a subsector of $[\iota\co\nu]$ for $\co\nu\in{\Sigma({\Delta})}$ a conjugate of $\nu$, there is an isometry $T\in{\rm{Hom}}(\bar b,\iota{\co\nu})$. Note that then $a(T)\in{\rm{Hom}}(a\bar b,a\iota{\co\nu})$. Hence by naturality and Proposition \ref{stat} we compute \[ \begin{array}{ll} \epspm \rho{a\bar b} &= a(T^*) \epspm \rho{a\iota{\co\nu}} \rho a (T) = a(T^*) a (\epspm \rho{\co\nu}) \epspm \rho{a\iota} \rho a(T) \\[.4em] &= a(T^*) a (\epspm \rho {\co\nu}) a\alpha_\rho^\pm(T) \epspm \rho{a\iota} = a( \Epspm \rho{\bar b}) \epspm \rho{a\iota} \,, \end{array} \] and hence also $\epspm {a\bar b}\rho = \epspm {a\iota}\rho a(\Epspm {\bar b}\rho$. We also obtain $\epspm {\lambda b\iota}\rho=\epspm\lambda\rho \lambda(\epspm{b\iota}\rho)$ and $\epspm \rho{\lambda b\iota}=\lambda(\epspm \rho{b\iota}) \epspm \rho\lambda$ by \erf{comp}. Note that $x\in{\rm{Hom}}(a\iota,\lambda b\iota)$ by restriction. Eqs.\ (\ref{ibfe1})-(\ref{ibfe4}) follow now by naturality, \erf{nat}. Next, we note that $T\in{\rm{Hom}}(\bar b \lambda,\iota{\co\nu}\lambda)$, and hence $\Epspm \rho{\co b \lambda}=T^* \epspm \rho{\co\nu\lambda} \alpha_\rho^\pm(T)$. Therefore \[ \begin{array}{ll} \Epspm \rho{\co b \lambda} &= T^* \co\nu(\epspm \rho\lambda) \epspm\rho{\co\nu} \alpha_\rho^\pm(T) = \co b(\epspm \rho\lambda) T^* \epspm \rho{\co\nu} \alpha_\rho^\pm(T) \\[.4em] & = \co b(\epspm \rho\lambda) \Epspm \rho{\co b} \,, \end{array} \] and hence also $\Epspm {\co b \lambda}\rho = \Epspm {\co b}\rho \co b (\epspm \lambda\rho)$. Now Eqs.\ (\ref{ibfe5}) and (\ref{ibfe6}) follow from Lemma \ref{Enat}. \end{proof} These IBFE's can be nicely visualized in diagrams. We display \erf{ibfe1} in Fig.\ \ref{IBFE1} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(120,60) \allinethickness{0.3pt} \put(26.180,5){\arc{32.361}{3.142}{4.249}} \put(10,5){\vector(0,-1){5}} \put(28,24){\line(-2,-1){9.1}} \put(33.820,45){\arc{32.361}{0}{1.107}} \put(50,60){\line(0,-1){15}} \put(32,26){\line(2,1){9.1}} \put(30,60){\line(0,-1){35}} \put(30,25){\vector(0,-1){15}} \allinethickness{1.0pt} \put(40,0){\vector(-1,1){10}} \put(30,10){\vector(-1,-1){10}} \put(30,5){\makebox(0,0){$t$}} \put(35,55){\makebox(0,0){$\lambda$}} \put(18,4.5){\makebox(0,0){$a$}} \put(42,5){\makebox(0,0){$b$}} \put(5,3.5){\makebox(0,0){$\rho$}} \put(60,25){\makebox(0,0){$=$}} \allinethickness{0.3pt} \put(86.180,5){\arc{32.361}{3.142}{4.199}} \put(70,5){\vector(0,-1){5}} \put(93.820,45){\arc{32.361}{0}{1.057}} \put(110,60){\line(0,-1){15}} \put(90,60){\vector(0,-1){10}} \allinethickness{1.0pt} \put(100,40){\vector(-1,1){10}} \put(90,50){\line(-1,-1){10}} \put(90,45){\makebox(0,0){$t$}} \put(80,40){\line(0,-1){20}} \put(100,40){\line(0,-1){10}} \put(80,10){\line(0,1){10}} \put(100,10){\line(0,1){20}} \put(80,10){\vector(0,-1){10}} \put(100,10){\line(0,-1){10}} \allinethickness{2.0pt} \put(90,25){\line(2,1){8}} \put(90,25){\line(-2,-1){8}} \put(95,55){\makebox(0,0){$\lambda$}} \put(85,4.5){\makebox(0,0){$a$}} \put(105,5){\makebox(0,0){$b$}} \put(65,3.5){\makebox(0,0){$\rho$}} \put(90,19){\makebox(0,0){$\alpha_\rho^-$}} \end{picture} \end{center} \caption{The first intertwining braiding fusion equation (overcrossings)} \label{IBFE1} \end{figure} and \erf{ibfe6} in Fig.\ \ref{IBFE2}, both for overcrossings. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(120,60) \allinethickness{2.0pt} \put(26.180,45){\arc{32.361}{2.034}{3.142}} \put(10,60){\line(0,-1){15}} \put(30,25){\line(-2,1){11.1}} \allinethickness{0.3pt} \put(33.820,5){\arc{32.361}{5.176}{6.283}} \put(50,5){\vector(0,-1){5}} \put(30,25){\line(2,-1){11.1}} \allinethickness{1.0pt} \put(30,27){\vector(0,1){33}} \put(30,23){\line(0,-1){13}} \put(20,0){\vector(1,1){10}} \allinethickness{0.3pt} \put(30,10){\vector(1,-1){10}} \put(3,55){\makebox(0,0){$\alpha_\rho^+$}} \put(30,3){\makebox(0,0){$Y$}} \put(35,55){\makebox(0,0){$a$}} \put(18,5){\makebox(0,0){$b$}} \put(42,5){\makebox(0,0){$\lambda$}} \put(55,4){\makebox(0,0){$\rho$}} \put(60,25){\makebox(0,0){$=$}} \allinethickness{2.0pt} \put(86.180,45){\arc{32.361}{2.034}{3.142}} \put(70,60){\line(0,-1){15}} \allinethickness{0.3pt} \put(93.820,5){\arc{32.361}{5.176}{6.283}} \put(110,5){\vector(0,-1){5}} \put(80,30){\line(-2,1){1.1}} \put(100,20){\line(2,-1){1.1}} \allinethickness{1.0pt} \put(90,50){\vector(0,1){10}} \put(80,40){\vector(1,1){10}} \put(80,40){\line(0,-1){8}} \put(80,10){\line(0,1){18}} \put(80,10){\line(0,-1){10}} \allinethickness{0.3pt} \put(100,40){\line(-1,1){10}} \put(100,40){\line(0,-1){18}} \put(100,10){\line(0,1){8}} \put(100,10){\vector(0,-1){10}} \put(90,25){\line(-2,1){10}} \put(90,25){\line(2,-1){10}} \put(90,43){\makebox(0,0){$Y$}} \put(95,55){\makebox(0,0){$a$}} \put(75,5){\makebox(0,0){$b$}} \put(95,5){\makebox(0,0){$\lambda$}} \put(115,4){\makebox(0,0){$\rho$}} \put(63,55){\makebox(0,0){$\alpha_\rho^+$}} \end{picture} \end{center} \caption{The sixth intertwining braiding fusion equation (overcrossings)} \label{IBFE2} \end{figure} We leave the remaining diagrams as a straightforward exercise to the reader. Note that the IBFE's give us the freedom to move wires with label $\rho$ and $\alpha_\rho^\pm$ freely over trivalent vertices which involve one $N$-$N$ wire and two $N$-$M$ wires. Unitarity of operators $\Epspm \lambda{\co b}$ yields a ``vertical Reidemeister move of type II'' similar to Fig.\ \ref{Reid2}. We can now also easily elaborate the rotation behavior of mixed crossings displayed in Fig.\ \ref{NMbraid} (and consequently their conjugates). Crucial for this is the fact that $R_{\alpha_\lambda^\pm}=\iota(r_\lambda)\equiv r_\lambda$ and ${\co R}_{\alpha_\lambda^\pm}=\iota({\co r}_\lambda)\equiv {\co r}_\lambda$ can be used as R-isometries for the $\alpha$-induced morphisms as $R_{\alpha_\lambda^\pm}\in{\rm{Hom}}({\rm{id}}_M, \overline{\alpha^\pm_\lambda}\alpha^\pm_\lambda)$ and ${\co R}_{\alpha_\lambda^\pm}\in{\rm{Hom}} ({\rm{id}}_M,\alpha^\pm_\lambda\overline{\alpha^\pm_\lambda})$ satisfy $\alpha^\pm_\lambda(R_{\alpha_\lambda^\pm})^*{\co R}_{\alpha_\lambda^\pm} =d_\lambda^{-1}{\bf1}_M$ and $\overline{\alpha^\pm_\lambda}({\co R}_{\alpha_\lambda^\pm})^* \co R_{\alpha_\lambda^\pm} =d_\lambda^{-1}{\bf1}_M$ and $d_{\alpha^\pm_\lambda}=d_\lambda$. First we notice that we have \[ \epspm \lambda{a\iota} = d_\lambda \, {\co r}_\lambda \, \lambda(\epsmp {a\iota}{\co\lambda}) \, \lambda a(r_\lambda) \] by \erf{rotcreq}. Now let $R_a\in{\rm{Hom}}({\rm{id}}_M,\co a a)$ and ${\co r}_a\in{\rm{Hom}}({\rm{id}}_N, a\co a)$ be isometries such that $a(R_a)^*{\co r}_a=d_a^{-1} {\bf1}_N$ and $\co a({\co r}_a)^*R_a=d_a^{-1}$, and otherwise we keep the notations as in Prop.\ \ref{intbfe}. From \erf{ibfe2} we obtain $a(\Epsmp {\co a}\lambda){\co r}_a = \epspm \lambda{a\iota} \lambda({\co r}_a)$. Hence we have \[ \begin{array}{ll} \epspm \lambda{a\iota} &= d_a \, \epspm \lambda{a\iota}\, \lambda a(R_a)^* \lambda({\co r}_a) = d_a \, a\alpha_\lambda^\pm (R_a)^* \, \epspm \lambda{a\iota} \, \lambda({\co r}_a) \\[.4em] &= d_a \, a\alpha_\lambda^\pm (R_a)^* \, a(\Epsmp {\co a}\lambda) \, {\co r}_a \,. \end{array} \] Next we compute, using again \erf{rotcreq}, \[ \begin{array}{ll} \Epspm \lambda{\co b} &= T^*\,\epspm \lambda{\co\nu} \, \alpha_\lambda^\pm(T) = d_\lambda \, T^*\,{\co r}_\lambda \lambda(\epsmp {\co\nu}{\co\lambda}) \lambda\co\nu(r_\lambda) \, \alpha_\lambda^\pm (T) \\[.4em] &= d_\lambda \, {\co r}_\lambda^* \alpha_\lambda^\pm( \alpha_{\co\lambda}^\pm(T)^* \epsmp {\co\nu}{\co\lambda}T) \alpha_\lambda^\pm \co b(r_\lambda) = d_\lambda \, {\co r}_\lambda^* \alpha_\lambda^\pm( \Epsmp {\co b}{\co\lambda}) \alpha_\lambda^\pm \co b(r_\lambda) \,. \end{array} \] Finally, as \erf{ibfe2} yields ${\co r}_a^* a(\Epspm \lambda{\co a}=\lambda({\co r}_a)^* \epsmp {a\iota}\lambda$, we obtain \[ \Epspm \lambda{\co a}= d_a \, \co a({\co r}_a)^* \, \co a a (\Epspm \lambda{\co a}) R_a = d_a \, \co a\lambda({\co r}_a)^* \, \co a(\epsmp {a\iota}\lambda) R_a \,. \] Drawing for $R_{\alpha_\lambda^\pm}=\iota(r_\lambda)$ and ${\co R}_{\alpha_\lambda^\pm}=\iota({\co r}_\lambda)$ caps of the wires $\alpha_\lambda^\pm$, these relations yield graphically the analogues of Fig.\ \ref{rotcross}. We conclude that we can include the crossings of Fig.\ \ref{NMbraid} consistently in our ``rotation covariant'' graphical framework. \section{Double Triangle Algebras for Subfactors} \label{sec-dta} We now formulate Ocneanu's construction \cite{O7} for a subfactor with finite index and finite depth rather than for bi-unitary connections and bimodules arising from Goodman-de la Harpe-Jones subfactors associated to A-D-E Dynkin diagrams in order to apply it in a more general context. From now on we work with $N\subset M$ satisfying the following \begin{assumption} \label{set-fin}{\rm Let $N\subset M$ be a type III subfactor with finite index. We assume that we have a system of endomorphisms ${}_N {\cal X}_N\subset{\rm{Mor}}(N,N)\equiv{\rm{End}}(N)$ in the sense of Definition \ref{system} such that for the injection map $\iota:N\to M$, the sector $[\theta]=[\bar\iota \iota]$ decomposes into a sum of sectors of morphisms in ${}_N {\cal X}_N$. We choose sets of morphisms ${}_N {\cal X}_M\subset{\rm{Mor}}(M,N)$, ${}_M {\cal X}_N\subset{\rm{Mor}}(N,M)$ and ${}_M {\cal X}_M\subset{\rm{Mor}}(M,M)\equiv{\rm{End}}(M)$ consisting of representative endomorphisms of irreducible subsectors of sectors of the form $[\lambda\co\iota]$, $[\iota\lambda]$ and $[\iota\lambda\co\iota]$, $\lambda\in{}_N {\cal X}_N$, respectively. (We may and do choose ${\rm{id}}_M$ in ${}_M {\cal X}_M$ as the endomorphism representing the trivial sector.) We also assume that ${}_N {\cal X}_N$ is finite. Consequently, the set ${\cal{X}}= {}_N {\cal X}_N \sqcup {}_N {\cal X}_M \sqcup {}_M {\cal X}_N \sqcup {}_M {\cal X}_M$ is finite. }\end{assumption} Note that Assumption \ref{set-fin} implies that representative morphisms for all irreducible sectors appearing in decompositions of powers $[\gamma^k]$ ($[\theta^k]$) of Longo's (dual) canonical endomorphism are contained in ${}_M {\cal X}_M$ (${}_N {\cal X}_N$). In other words, the set ${\cal{X}}$ contains at least the morphisms corresponding to the (equivalence classes of) bimodules arising from this subfactor through the Jones tower, and therefore we may call an ${\cal{X}}$ which does not contain any other morphisms a {\sl minimal choice}. We conclude that finiteness of ${}_N {\cal X}_N$ in Assumption in \ref{set-fin} automatically implies that the subfactor $N\subset M$ has finite depth. We used sectors instead of bimodules in view of our ``identification'' of chiral generators with $\alpha$-induced sectors below. Therefore we need a sector approach in order to define $\alpha$-induction since its definition involves $\co\iota^{\,-1}$, and hence we work with factors of type III. (We do not need hyperfiniteness of $M$ for our purposes.) We now use the graphical calculus presented in Section \ref{sec-graphcal}. In the graphical method of \cite{O3} (and \cite[Chapter 12]{EK3}), factors, bimodules (morphisms), and intertwiners are represented with trivalent vertices, edges, and triangles, respectively, and this is where the name ``double triangle algebra'' comes from. However, here (as in \cite{O6,O7}) these three kinds of objects are represented by regions, wires, and trivalent vertices, respectively, though the labels for regions are omitted for notational simplicity. For $\cal X$ in Assumption \ref{set-fin}, we define the {\sl double triangle algebra} $\dta$ with two multiplications $*_h$ and $*_v$ as follows. As a linear space, we set \[ \dta=\bigoplus_{a,b,c,d\in{}_N {\cal X}_M} {\rm{Hom}}(a \bar b, c \bar d) \,. \] This is a finite dimensional complex linear space. An element in $\dta$ is presented graphically as in Fig.\ \ref{dta1} under the interpretation in Section \ref{sec-graphcal} with the convention of reading the diagram from the top to the bottom. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(40,40) \allinethickness{1.0pt} \put(10,10){\line(1,0){20}} \put(10,30){\line(1,0){20}} \allinethickness{0.3pt} \put(20,30){\vector(0,-1){20}} \put(7,30){\makebox(0,0){$a$}} \put(33,30){\makebox(0,0){$b$}} \put(7,10){\makebox(0,0){$c$}} \put(33,10){\makebox(0,0){$d$}} \put(17,20){\makebox(0,0){$\lambda$}} \put(20,34){\makebox(0,0){$s^*$}} \put(20,6){\makebox(0,0){$t$}} \end{picture} \end{center} \caption{An element in $\protect\dta$} \label{dta1} \end{figure} (A general element in $\dta$ is a linear combination of this type of element.) We can interpret the same diagram with the convention of reading the diagram from the left to the right or, equivalently, keeping the top-to-bottom convention but putting the diagram in a suitable Frobenius annulus. Then the resulting intertwiner is in \[ \dtap=\bigoplus_{a,b,c,d\in{}_N {\cal X}_M} {\rm{Hom}}(\bar c a,\bar d b) \,. \] The isomorphism of these two spaces is given by application of two Frobenius rotations, and we can use this isomorphism to identify $\dta$ and $\dtap$. By our convention of the normalization in Section \ref{sec-graphcal}, the diagram of Fig.\ \ref{dta1} represents an element $d_a^{1/4} d_b^{1/4} d_c^{1/4} d_d^{1/4} d_\lambda^{-1/2}ts^*$ in the block ${\rm{Hom}}(a\bar b, c\bar d)$, where $s\in{\rm{Hom}} (\lambda,a\bar b)$ and $t\in{\rm{Hom}}(\lambda, c\bar d)$ are isometries and $\lambda\in{}_N {\cal X}_N$. Similarly we may use elements in $\dta$ which are graphically represented as in Fig.\ \ref{dta2} with isometries $S\in{\rm{Hom}}(\beta,\bar c a)$, $T\in{\rm{Hom}}(\beta,\bar d b)$ and $\beta\in{}_M {\cal X}_M$. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(40,40) \allinethickness{1.0pt} \put(10,10){\line(0,1){20}} \put(30,10){\line(0,1){20}} \allinethickness{2.0pt} \put(10,20){\vector(1,0){20}} \put(10,33){\makebox(0,0){$a$}} \put(30,33){\makebox(0,0){$b$}} \put(10,7){\makebox(0,0){$c$}} \put(30,7){\makebox(0,0){$d$}} \put(20,24){\makebox(0,0){$\beta$}} \put(5,20){\makebox(0,0){$S^*$}} \put(35,20){\makebox(0,0){$T$}} \end{picture} \end{center} \caption{An element in $\protect\dta$} \label{dta2} \end{figure} Note that elements of the form in Fig.\ \ref{dta1}, or equivalently of the form in Fig.\ \ref{dta2}, span $\dta$ linearly. Our graphical convention is as follows. We use thin, thick, and very thick wires for $N$-$N$ morphisms, $N$-$M$ morphisms, and $M$-$M$ morphisms, respectively, analogous to the convention \cite{O7}. We call them $N$-$N$ wires, and so on. We label $N$-$N$ morphisms with Greek letters $\lambda,\mu,\nu,\dots$, $N$-$M$ morphisms with Roman letters $a,b,c,d,\dots$, and $M$-$M$ morphisms with Greek letters $\beta,\beta', \beta'',\dots$. We orient $N$-$N$ or $M$-$M$ wires but we put no orientations on $N$-$M$ wires since it is clear from the context whether we mean an $N$-$M$ morphism $a$ or an $M$-$N$ morphism $\bar a$. We simply put a label $a$ for an unoriented thick wire for both. Note that, whatever we consider, $\dta$ or $\dtap$, the same intertwiner (as an operator) may appear in different blocks of the double triangle algebra, e.g.\ the identity ${\rm{id}}_N$ is an element in any ${\rm{Hom}}(a \bar b,a \bar b)$, $a,b\in{}_N {\cal X}_M$. The graphical notation is particularly useful in order to avoid this kind of confusion because diagrams as in Figs.\ \ref{dta1} and \ref{dta2} always specify also the associated block in addition to the intertwiner as an operator. The {\sl horizontal product} $*_h$ on $\dta$ is defined as in Fig.\ \ref{prod}. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(150,40) \allinethickness{1.0pt} \put(10,10){\line(1,0){20}} \put(10,30){\line(1,0){20}} \put(50,10){\line(1,0){20}} \put(50,30){\line(1,0){20}} \put(110,10){\line(1,0){30}} \put(110,30){\line(1,0){30}} \allinethickness{0.3pt} \put(20,30){\vector(0,-1){20}} \put(60,30){\vector(0,-1){20}} \put(120,30){\vector(0,-1){20}} \put(130,30){\vector(0,-1){20}} \put(7,30){\makebox(0,0){$a$}} \put(33,30){\makebox(0,0){$b$}} \put(7,10){\makebox(0,0){$c$}} \put(33,10){\makebox(0,0){$d$}} \put(47,30){\makebox(0,0){$a'$}} \put(73,30){\makebox(0,0){$b'$}} \put(47,10){\makebox(0,0){$c'$}} \put(73,10){\makebox(0,0){$d'$}} \put(40,20){\makebox(0,0){$*_h$}} \put(93,20){\makebox(0,0){$=\del b{a'} \del d{c'}$}} \put(107,30){\makebox(0,0){$a$}} \put(143,30){\makebox(0,0){$b'$}} \put(107,10){\makebox(0,0){$c$}} \put(143,10){\makebox(0,0){$d'$}} \put(125,14){\makebox(0,0){$d$}} \put(125,26){\makebox(0,0){$b$}} \put(17,20){\makebox(0,0){$\lambda$}} \put(57,20){\makebox(0,0){$\mu$}} \put(117,20){\makebox(0,0){$\lambda$}} \put(133,20){\makebox(0,0){$\mu$}} \put(20,34){\makebox(0,0){$s^*$}} \put(20,6){\makebox(0,0){$t$}} \put(61,34){\makebox(0,0){${s'}^*$}} \put(60,6){\makebox(0,0){$t'$}} \put(120,34){\makebox(0,0){$s^*$}} \put(120,6){\makebox(0,0){$t$}} \put(132,34){\makebox(0,0){${s'}^*$}} \put(130,6){\makebox(0,0){$t'$}} \end{picture} \end{center} \caption{The horizontal product $*_h$ on $\protect\dta$} \label{prod} \end{figure} The meaning of the right-hand side is as follows. The product is by definition zero if the labels of the open ends of the wires facing each other do not match. If they match, we glue the wires of the two diagrams together as in Fig.\ \ref{prod} and interpret it as an intertwiner. It belongs to the block of the double triangle algebra which is specified by the four remaining open ends of the new diagram. This is a horizontal version of the composition of intertwiners described in Section \ref{sec-graphcal}. We also can represent this horizontal product in terms of elements in Fig.\ \ref{dta2}. This is described in Fig.\ \ref{prod2}, because the convention of Section \ref{sec-graphcal} means that this product is just the composition of the intertwiners in $\dtap$, and this composition is realized by taking the inner product of the two intertwiners in the right-hand side in Fig.\ \ref{prod2}. \begin{figure}[htb] \begin{center} \unitlength 0.7mm \begin{picture}(200,40) \allinethickness{1.0pt} \put(10,10){\line(0,1){20}} \put(30,10){\line(0,1){20}} \put(60,10){\line(0,1){20}} \put(80,10){\line(0,1){20}} \put(170,10){\line(0,1){20}} \put(190,10){\line(0,1){20}} \allinethickness{2.0pt} \put(10,20){\line(1,0){20}} \put(20,20){\vector(1,0){0}} \put(60,20){\line(1,0){20}} \put(70,20){\vector(1,0){0}} \put(170,20){\line(1,0){20}} \put(180,20){\vector(1,0){0}} \put(10,7){\makebox(0,0){$c$}} \put(30,7){\makebox(0,0){$d$}} \put(60,7){\makebox(0,0){$c'$}} \put(80,7){\makebox(0,0){$d'$}} \put(170,7){\makebox(0,0){$c$}} \put(190,7){\makebox(0,0){$d'$}} \put(10,33){\makebox(0,0){$a$}} \put(30,33){\makebox(0,0){$b$}} \put(60,33){\makebox(0,0){$a'$}} \put(80,33){\makebox(0,0){$b'$}} \put(170,33){\makebox(0,0){$a$}} \put(190,33){\makebox(0,0){$b'$}} \put(20,24){\makebox(0,0){$\beta$}} \put(70,24){\makebox(0,0){$\beta'$}} \put(180,24){\makebox(0,0){$\beta$}} \put(6,20){\makebox(0,0){$S^*$}} \put(34,20){\makebox(0,0){$T$}} \put(56,20){\makebox(0,0){${S'}^*$}} \put(84,20){\makebox(0,0){$T'$}} \put(166,20){\makebox(0,0){$S^*$}} \put(194,20){\makebox(0,0){$T'$}} \put(45,20){\makebox(0,0){$*_h$}} \put(125,20){\makebox(0,0){$=\displaystyle\del b{a'} \del d{c'} \del \beta{\beta'}\sqrt{\frac{d_b d_d}{d_\beta}} \langle S',T \rangle$}} \end{picture} \end{center} \caption{The horizontal product presented in another way} \label{prod2} \end{figure} We similarly define the {\sl vertical product} $*_v$ on $\dta$ by composing two diagrams vertically, but with extra coefficients as in Fig.\ \ref{prod3}. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(170,50) \allinethickness{1.0pt} \put(10,15){\line(0,1){20}} \put(30,15){\line(0,1){20}} \put(50,15){\line(0,1){20}} \put(70,15){\line(0,1){20}} \put(140,10){\line(0,1){30}} \put(160,10){\line(0,1){30}} \allinethickness{2.0pt} \put(10,25){\vector(1,0){20}} \put(50,25){\vector(1,0){20}} \put(140,20){\vector(1,0){20}} \put(140,30){\vector(1,0){20}} \put(10,38){\makebox(0,0){$a$}} \put(30,38){\makebox(0,0){$b$}} \put(10,12){\makebox(0,0){$c$}} \put(30,12){\makebox(0,0){$d$}} \put(50,38){\makebox(0,0){$a'$}} \put(70,38){\makebox(0,0){$b'$}} \put(50,12){\makebox(0,0){$c'$}} \put(70,12){\makebox(0,0){$d'$}} \put(140,43){\makebox(0,0){$a'$}} \put(160,43){\makebox(0,0){$b'$}} \put(140,7){\makebox(0,0){$c$}} \put(160,7){\makebox(0,0){$d$}} \put(137,25){\makebox(0,0){$a$}} \put(163,25){\makebox(0,0){$b$}} \put(20,29){\makebox(0,0){$\beta$}} \put(60,29){\makebox(0,0){$\beta'$}} \put(150,34){\makebox(0,0){$\beta'$}} \put(150,24){\makebox(0,0){$\beta$}} \put(40,25){\makebox(0,0){$*_v$}} \put(108,25){\makebox(0,0){$=\hphantom{}\del a{c'}\del b{d'} \sqrt{d_a d_b}$}} \end{picture} \end{center} \caption{The vertical product $*_v$ in $\protect\dta$} \label{prod3} \end{figure} The meaning of the right-hand side is as before. Note that the definitions of horizontal and vertical products are not completely symmetric due to the extra coefficients we chose. This choice is somewhat arbitrary but it just turns out to be useful for our purposes. Namely, with this definition of the products, the minimal central projections of $(\dta,*_h)$ have simple and useful composition rules with respect to the vertical product $*_v$, see Theorem \ref{double2} below. We clearly also have a $*$-structure for the horizontal product obtained by vertical reflection of the diagram, adjoining labels for trivalent vertices and reversing orientations of wires. Analogously, a $*$-structure for the vertical product comes from horizontal reflection. The basic idea is that the 90-degree rotation is something like a ``Fourier transform'' which transforms the two products into each other, similar to the situation of the group algebra of a finite or compact group. For each $\beta,\lambda,a,b$ we choose orthonormal bases of isometries $T_{\bar b,a}^{\beta;i} \in{\rm{Hom}}(\beta, \bar b a)$, $i=1,2,...,N_{\bar b,a}^\beta$, and $t_{a,\bar b}^{\lambda;j} \in{\rm{Hom}}(\lambda, a \bar b)$ $j=1,2,...,N_{a\bar b}^\lambda$, so that \begin{equation} \sum_{\beta\in{}_M {\cal X}_M} \sum_{i=1}^{N_{\bar b,a}^\beta} T_{\bar b,a}^{\beta;i}(T_{\bar b,a}^{\beta;i})^*={\bf1}_M \qquad\mbox{and}\qquad \sum_{\lambda\in{}_N {\cal X}_N} \sum_{j=1}^{N_{a,\bar b}^\lambda} t_{a,\bar b}^{\lambda;j}(t_{a,\bar b}^{\lambda;j})^*={\bf1}_N \label{totsum} \end{equation} for all $a,b\in{}_N {\cal X}_M$. Then it is easy to see that the elements in Fig.\ \ref{munit} \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(232,40) \allinethickness{1.0pt} \put(85,10){\line(0,1){20}} \put(105,10){\line(0,1){20}} \allinethickness{2.0pt} \put(85,20){\vector(1,0){20}} \put(85,33){\makebox(0,0){$a$}} \put(105,34){\makebox(0,0){$b$}} \put(85,6){\makebox(0,0){$c$}} \put(105,7){\makebox(0,0){$d$}} \put(73,20){\makebox(0,0){$(T_{\bar c,a}^{\beta;i})^*$}} \put(115,19){\makebox(0,0){$T_{\bar d,b}^{\beta;j}$}} \put(95,24){\makebox(0,0){$\beta$}} \put(30,20){\makebox(0,0){$e_{\beta;c,a,i}^{d,b,j}= \displaystyle\frac{\sqrt{d_\beta}}{\sqrt[4]{d_a d_b d_c d_d}}$}} \put(125,16){\makebox(0,0){,}} \allinethickness{1.0pt} \put(210,10){\line(1,0){20}} \put(210,30){\line(1,0){20}} \allinethickness{0.3pt} \put(220,30){\vector(0,-1){20}} \put(207,33){\makebox(0,0){$a$}} \put(233,34){\makebox(0,0){$b$}} \put(207,6){\makebox(0,0){$c$}} \put(233,7){\makebox(0,0){$d$}} \put(220,36){\makebox(0,0){$(t_{a, \bar b}^{\lambda;i})^*$}} \put(220,4){\makebox(0,0){$t_{c, \bar d}^{\lambda;j}$}} \put(224,20){\makebox(0,0){$\lambda$}} \put(170,20){\makebox(0,0){$f_{\lambda;c,d,j}^{a,b,i}= {\displaystyle\sqrt{\frac{d_\lambda}{d_a d_b d_c d_d}}}$}} \end{picture} \end{center} \caption{Matrix units $e_{\beta;c,a,i}^{d,b,j}$ for $(\protect\dta,*_h)$ and $f_{\lambda;c,d,j}^{a,b,i}$ for $(\protect\dta,*_v)$} \label{munit} \end{figure} form bases of $\dta$ which constitute complete systems of matrix units $(\dta, *_h)$ respectively $(\dta,*_v)$. Thus for each of the two multiplications the double triangle algebra is a direct sum of full matrix algebras. The two different bases are transformed into each other by a unitary transformation with coefficients given by the $6j$-symbols for subfactors of \cite{O3} (see \cite[Chapter 12]{EK3} for the basic properties of ``quantum $6j$-symbols''), but this will not be exploited here. \begin{definition}\label{defe-beta} {\rm For each $\beta\in{}_M {\cal X}_M$ we define an element $e_\beta=\sum_{a,b,i} e_{\beta;b,a,i}^{b,a,i}\in\dta$. Graphically, this element is given by the left-hand side in Fig.\ \ref{e-beta}. We use the convention shown on the right-hand side in Fig.\ \ref{e-beta} to represent this element. }\end{definition} \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(173,40) \allinethickness{1.0pt} \put(60,10){\line(0,1){20}} \put(80,10){\line(0,1){20}} \allinethickness{2.0pt} \put(60,20){\vector(1,0){20}} \put(60,33){\makebox(0,0){$a$}} \put(80,33){\makebox(0,0){$a$}} \put(60,7){\makebox(0,0){$b$}} \put(80,7){\makebox(0,0){$b$}} \put(48,20){\makebox(0,0){$(T_{\bar b,a}^{\beta;i})^*$}} \put(90,19){\makebox(0,0){$T_{\bar b,a}^{\beta;i}$}} \put(70,24){\makebox(0,0){$\beta$}} \put(20,18){\makebox(0,0){$\displaystyle \sum_{a,b,i} \sqrt{\frac{d_\beta}{d_a d_b}}$}} \put(120,19){\makebox(0,0){$=:$}} \allinethickness{1.0pt} \put(150,10){\line(0,1){20}} \put(170,10){\line(0,1){20}} \allinethickness{2.0pt} \put(150,20){\vector(1,0){20}} \put(150,33){\makebox(0,0){$a$}} \put(170,33){\makebox(0,0){$a$}} \put(150,7){\makebox(0,0){$b$}} \put(170,7){\makebox(0,0){$b$}} \allinethickness{0.3pt} \put(170,20){\arc{5}{1.57}{4.71}} \put(150,20){\arc{5}{4.71}{7.85}} \put(160,24){\makebox(0,0){$\beta$}} \put(137,16){\makebox(0,0){$\displaystyle \sum_{a,b}$}} \end{picture} \end{center} \caption{The minimal central projection $e_\beta$} \label{e-beta} \end{figure} Due to the summation over $i=1,2,...,N_{\bar b,a}^\beta$, the definition is independent of the choice of the intertwiner bases as different orthonormal bases are related by a unitary matrix. We will use such a graphical convention whenever we have a sum over internal ``fusion channels'' of two corresponding trivalent vertices together with prefactors which renormalize the trivalent vertices to isometries. Note that we obtain a prefactor, as displayed in Fig.\ \ref{trick} for an example, when we turn around the small arcs at trivalent vertices. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(180,20) \allinethickness{1.0pt} \put(0,10){\line(1,0){15}} \put(15,10){\line(1,-1){10}} \dottedline{2}(25,0)(45,0) \put(55,10){\line(-1,-1){10}} \put(55,10){\line(1,0){15}} \allinethickness{0.3pt} \put(15,10){\line(1,1){10}} \dottedline{2}(25,20)(45,20) \put(55,10){\line(-1,1){10}} \put(37,20){\vector(1,0){0}} \put(5,15){\makebox(0,0){$a$}} \put(65,15){\makebox(0,0){$a$}} \put(35,5){\makebox(0,0){$b$}} \put(35,15){\makebox(0,0){$\lambda$}} \put(15,10){\arc{5}{3.142}{0.785}} \put(55,10){\arc{5}{2.356}{0}} \put(93,10){\makebox(0,0){$=\;\displaystyle\frac{d_\lambda}{d_b}$}} \allinethickness{1.0pt} \put(110,10){\line(1,0){15}} \put(125,10){\line(1,-1){10}} \dottedline{2}(135,0)(155,0) \put(165,10){\line(-1,-1){10}} \put(165,10){\line(1,0){15}} \allinethickness{0.3pt} \put(125,10){\line(1,1){10}} \dottedline{2}(135,20)(155,20) \put(165,10){\line(-1,1){10}} \put(147,20){\vector(1,0){0}} \put(115,15){\makebox(0,0){$a$}} \put(175,15){\makebox(0,0){$a$}} \put(145,5){\makebox(0,0){$b$}} \put(145,15){\makebox(0,0){$\lambda$}} \put(125,10){\arc{5}{5.498}{3.142}} \put(165,10){\arc{5}{0}{3.927}} \end{picture} \end{center} \caption{Turning around small arcs yields a prefactor} \label{trick} \end{figure} Here the dotted parts mean that there might be expansions as given in the following lemma or later even be braiding operators in between; it is just important that the small arcs at corresponding trivalent vertices denote the same summation over internal fusion channels. \begin{lemma} \label{idexp} The identity of Fig.\ \ref{exp-id} holds. Analogous identities hold if $a,b,\beta$ are replaced by wires of other type (in a compatible way). \end{lemma} \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(90,40) \allinethickness{1.0pt} \put(10,10){\line(1,0){20}} \put(10,30){\line(1,0){20}} \put(60,10){\line(0,1){20}} \put(80,10){\line(0,1){20}} \allinethickness{2.0pt} \put(60,20){\vector(1,0){20}} \put(20,34){\makebox(0,0){$a$}} \put(20,6){\makebox(0,0){$b$}} \put(60,33){\makebox(0,0){$a$}} \put(80,33){\makebox(0,0){$a$}} \put(60,7){\makebox(0,0){$b$}} \put(80,7){\makebox(0,0){$b$}} \put(70,24){\makebox(0,0){$\beta$}} \allinethickness{0.3pt} \put(80,20){\arc{5}{1.57}{4.71}} \put(60,20){\arc{5}{4.71}{7.85}} \put(44,18){\makebox(0,0){$\displaystyle =\sum_\beta$}} \end{picture} \end{center} \caption{The identity with expansion using $\beta$} \label{exp-id} \end{figure} \begin{proof} With the normalization convention as in Fig.\ \ref{e-beta}, this is just the expansion of the identity in \erf{totsum}, and this certainly holds as well using similar expansions with other intertwiner bases. \end{proof} Note that the identity in Fig.\ \ref{exp-id} may, for example, also appear rotated by 90 degrees as we can put the left- and right-hand sides in some Frobenius annulus as described in Subsect.\ \ref{graph2-prelim}. As we have already indicated, the horizontal product is essentially the composition of intertwiners in $\dtap$. The main point of the double triangle algebra is the following. Suppose we have complete information on the fusion rules of $N$-$N$, $N$-$M$, $M$-$N$ morphisms in ${\cal{X}}$ and their $6j$-symbols. We can define the algebra $\dta$ in terms of matrix elements $f_{c,d,j}^{\lambda;a,b,i}$ and determine their composition with respect to the horizontal product without any information of the $M$-$M$ morphisms. Then we can {\sl find} $M$-$M$ sectors and determine their fusion rules by the following theorem which generalizes a result for Goodman-de la Harpe-Jones subfactors in \cite{O7} in a straightforward manner. \begin{theorem} \label{double2} For any $\beta\in{}_M {\cal X}_M$ the element $e_\beta\in\dta$ of Definition \ref{defe-beta} is a minimal central projection with respect to the horizontal product, and all minimal central projections arise in this way in a bijective correspondence. Furthermore, we have\footnote{Note that the fusion coefficients with dimension prefactors as in \erf{dddN} coincide with the structure constants used for $C$-algebras \cite{BI}.} \begin{equation} e_\beta *_v e_{\beta'} = \sum_{{\beta}''\in{}_M {\cal X}_M} \frac{d_\beta d_{\beta'}}{d_{\beta''}} N_{\beta,\beta'}^{\beta''} \, e_{\beta''} \label{dddN} \end{equation} for all $\beta,\beta'\in{}_M {\cal X}_M$. In particular, the center ${\cal{Z}}_h$ of $\dta$ with respect to the horizontal product is closed under the vertical product. \end{theorem} \begin{proof} That each $e_\beta$ is a minimal central projection and that all minimal central projections arise in this way is obvious from the description of the matrix units. The vertical product $e_\beta *_v e_{\beta'}$ is given graphically by the left-hand side of Fig\ \ref{v-prod1}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(211,50) \allinethickness{0.3pt} \put(10,23){\makebox(0,0){$\displaystyle \sum_{a,b,c}\; d_b$}} \put(50,15){\arc{5}{1.571}{4.712}} \put(30,15){\arc{5}{4.712}{1.571}} \put(50,35){\arc{5}{1.571}{4.712}} \put(30,35){\arc{5}{4.712}{1.571}} \allinethickness{1.0pt} \put(30,5){\line(0,1){40}} \put(50,5){\line(0,1){40}} \allinethickness{2.0pt} \put(30,15){\line(1,0){20}} \put(30,35){\line(1,0){20}} \put(42,15){\vector(1,0){0}} \put(42,35){\vector(1,0){0}} \put(26,3){\makebox(0,0){$c$}} \put(27,25){\makebox(0,0){$b$}} \put(26,47){\makebox(0,0){$a$}} \put(54,3){\makebox(0,0){$c$}} \put(53,25){\makebox(0,0){$b$}} \put(54,47){\makebox(0,0){$a$}} \put(40,10){\makebox(0,0){$\beta$}} \put(40,40){\makebox(0,0){$\beta'$}} \put(84,21){\makebox(0,0){$=\displaystyle \sum_{a,b,c,\atop\beta'',\beta''',\beta''''} d_b$}} \allinethickness{0.3pt} \put(125,25){\arc{5}{1.571}{4.712}} \put(110,25){\arc{5}{4.712}{1.571}} \put(165,25){\arc{5}{1.571}{4.712}} \put(150,25){\arc{5}{4.712}{1.571}} \put(205,25){\arc{5}{1.571}{4.712}} \put(190,25){\arc{5}{4.712}{1.571}} \put(180,35){\arc{5}{1.571}{4.712}} \put(135,35){\arc{5}{4.712}{1.571}} \put(180,15){\arc{5}{1.571}{4.712}} \put(135,15){\arc{5}{4.712}{1.571}} \allinethickness{1.0pt} \put(110,5){\line(0,1){40}} \put(125,10){\line(0,1){30}} \put(135,10){\line(0,1){30}} \put(130,40){\arc{10}{3.142}{0}} \put(130,10){\arc{10}{0}{3.142}} \put(205,5){\line(0,1){40}} \put(180,10){\line(0,1){30}} \put(190,10){\line(0,1){30}} \put(185,40){\arc{10}{3.142}{0}} \put(185,10){\arc{10}{0}{3.142}} \allinethickness{2.0pt} \put(110,25){\line(1,0){15}} \put(119.5,25){\vector(1,0){0}} \put(135,15){\line(1,0){10}} \put(135,35){\line(1,0){10}} \put(142,15){\vector(1,0){0}} \put(142,35){\vector(1,0){0}} \put(145,30){\arc{10}{4.712}{0}} \put(145,20){\arc{10}{0}{1.571}} \put(150,20){\line(0,1){10}} \put(150,25){\line(1,0){15}} \put(159.5,25){\vector(1,0){0}} \put(170,30){\arc{10}{3.142}{4.712}} \put(170,20){\arc{10}{1.571}{3.142}} \put(165,20){\line(0,1){10}} \put(190,25){\line(1,0){15}} \put(199.5,25){\vector(1,0){0}} \put(170,15){\line(1,0){10}} \put(170,35){\line(1,0){10}} \put(177,15){\vector(1,0){0}} \put(177,35){\vector(1,0){0}} \put(106,3){\makebox(0,0){$c$}} \put(106,47){\makebox(0,0){$a$}} \put(209,3){\makebox(0,0){$c$}} \put(209,47){\makebox(0,0){$a$}} \put(123,3){\makebox(0,0){$c$}} \put(123,47){\makebox(0,0){$a$}} \put(192,3){\makebox(0,0){$c$}} \put(192,47){\makebox(0,0){$a$}} \put(145,10){\makebox(0,0){$\beta$}} \put(145,40){\makebox(0,0){$\beta'$}} \put(170,10){\makebox(0,0){$\beta$}} \put(170,40){\makebox(0,0){$\beta'$}} \put(117.5,30){\makebox(0,0){$\beta'''$}} \put(157.5,30){\makebox(0,0){$\beta''$}} \put(197.5,30){\makebox(0,0){$\beta''''$}} \put(139,25){\makebox(0,0){$b$}} \put(176,25){\makebox(0,0){$b$}} \end{picture} \end{center} \caption{The vertical product $e_\beta *_v e_{\beta'}$} \label{v-prod1} \end{figure} We can use the expansion of Lemma \ref{idexp} for the two parallel wires $\beta$ and $\beta'$ in the middle. Now note that the horizontal unit is given by ${\bf1}_h=\sum_\beta e_\beta$. Therefore, by multiplying ${\bf1}_h$ from the left and from the right, we obtain the diagram on the right-hand side of Fig.\ \ref{v-prod1}. Reading the diagram from left to right, we observe that intertwiners in ${\rm{Hom}}(\beta''',\beta'')$ and ${\rm{Hom}}(\beta'',\beta'''')$ are involved here. Hence we first obtain a factor $\del {\beta'''}{\beta''} \del {\beta''}{\beta''''}$. Next, we can use the trick of Fig.\ \ref{trick} to turn around the small arcs at the trivalent vertices involving $a,b,\beta'$. This yields a factor $d_\beta'/d_b$. This way we see that the diagram on the right-hand side of Fig.\ \ref{v-prod1} represents the same element of the $\dta$ as the diagram in. Fig.\ \ref{v-prod2}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(206,50) \allinethickness{0.3pt} \put(14,23){\makebox(0,0){$\displaystyle \sum_{a,b,c,\beta''} d_{\beta'}$}} \put(40,12.5){\arc{5}{4.712}{1.571}} \put(60,12.5){\arc{5}{1.571}{4.712}} \put(75,27.5){\arc{5}{4.712}{1.571}} \put(165,27.5){\arc{5}{1.571}{4.712}} \put(90,12.5){\arc{5}{4.712}{1.571}} \put(150,12.5){\arc{5}{1.571}{4.712}} \put(110,27.5){\arc{5}{4.712}{1.571}} \put(130,27.5){\arc{5}{1.571}{4.712}} \put(180,12.5){\arc{5}{4.712}{1.571}} \put(200,12.5){\arc{5}{1.571}{4.712}} \dottedline{4}(65,0)(65,50)(175,50)(175,0)(65,0) \allinethickness{1.0pt} \put(40,5){\line(0,1){40}} \put(60,10){\line(0,1){5}} \put(65,10){\arc{10}{1.571}{3.142}} \put(65,15){\arc{10}{3.142}{4.712}} \put(65,5){\line(1,0){20}} \put(65,20){\line(1,0){5}} \put(75,25){\line(0,1){2.5}} \put(70,25){\arc{10}{0}{1.571}} \put(90,10){\line(0,1){12.5}} \put(85,10){\arc{10}{0}{1.571}} \put(85,22.5){\arc{10}{4.712}{0}} \put(75,27.5){\line(1,0){10}} \put(165,25){\line(0,1){2.5}} \put(170,25){\arc{10}{1.571}{3.142}} \put(170,20){\line(1,0){5}} \put(180,10){\line(0,1){5}} \put(175,10){\arc{10}{0}{1.571}} \put(175,15){\arc{10}{4.712}{0}} \put(155,5){\line(1,0){20}} \put(155,27.5){\line(1,0){10}} \put(150,10){\line(0,1){12.5}} \put(155,10){\arc{10}{1.571}{3.142}} \put(155,22.5){\arc{10}{3.142}{4.712}} \put(200,5){\line(0,1){40}} \allinethickness{2.0pt} \put(40,12.5){\line(1,0){20}} \put(52,12.5){\vector(1,0){0}} \put(99.5,12.5){\vector(1,0){0}} \put(144.5,12.5){\vector(1,0){0}} \put(122,27.5){\vector(1,0){0}} \put(87,45){\vector(1,0){0}} \put(157,45){\vector(1,0){0}} \put(65,35){\line(1,0){5}} \put(65,45){\line(1,0){40}} \put(65,40){\arc{10}{1.571}{4.712}} \put(75,27.5){\line(0,1){2.5}} \put(70,30){\arc{10}{4.712}{0}} \put(105,17.5){\arc{10}{0}{1.571}} \put(105,40){\arc{10}{4.712}{0}} \put(110,17.5){\line(0,1){22.5}} \put(90,12.5){\line(1,0){15}} \put(110,27.5){\line(1,0){20}} \put(135,17.5){\arc{10}{1.571}{3.142}} \put(135,40){\arc{10}{3.142}{4.712}} \put(130,17.5){\line(0,1){22.5}} \put(135,12.5){\line(1,0){15}} \put(170,35){\line(1,0){5}} \put(135,45){\line(1,0){40}} \put(175,40){\arc{10}{4.712}{1.571}} \put(165,27.5){\line(0,1){2.5}} \put(170,30){\arc{10}{3.142}{4.712}} \put(180,12.5){\line(1,0){20}} \put(192,12.5){\vector(1,0){0}} \put(36,3){\makebox(0,0){$c$}} \put(36,47){\makebox(0,0){$a$}} \put(204,3){\makebox(0,0){$c$}} \put(204,47){\makebox(0,0){$a$}} \put(75,8){\makebox(0,0){$c$}} \put(70,16){\makebox(0,0){$a$}} \put(165,8){\makebox(0,0){$c$}} \put(170,16){\makebox(0,0){$a$}} \put(85,22.5){\makebox(0,0){$b$}} \put(155,22.5){\makebox(0,0){$b$}} \put(97.5,17.5){\makebox(0,0){$\beta$}} \put(95,40){\makebox(0,0){$\beta'$}} \put(142.5,17.5){\makebox(0,0){$\beta$}} \put(145,40){\makebox(0,0){$\beta'$}} \put(50,17.5){\makebox(0,0){$\beta''$}} \put(120,32.5){\makebox(0,0){$\beta''$}} \put(190,17.5){\makebox(0,0){$\beta''$}} \end{picture} \end{center} \caption{The vertical product $e_\beta *_v e_{\beta'}$} \label{v-prod2} \end{figure} Now let us look at the part of this picture inside the dotted box. Reading it from the left, this part can be read for fixed $a$ and $c$ as $\sum_{i,k} T_i T_{\beta,\beta'}^{\beta'';k} (T_{\beta,\beta'}^{\beta'';k})^* T_i^*$, and the sum over $i$ runs over a full orthonormal basis of isometries $T_i$ in the Hilbert space ${\rm{Hom}}(\beta,\co c a\co{\beta'})$ since we have the summation over $b$. Next we look at the part inside the dotted box of the diagram in Fig.\ \ref{v-prod3}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(206,50) \allinethickness{0.3pt} \put(16.5,23){\makebox(0,0){$\displaystyle \sum_{a,c,\beta'',\beta'''} d_{\beta'}$}} \put(40,12.5){\arc{5}{4.712}{1.571}} \put(60,12.5){\arc{5}{1.571}{4.712}} \put(90,22.5){\arc{5}{4.712}{1.571}} \put(150,22.5){\arc{5}{1.571}{4.712}} \put(75,12.5){\arc{5}{4.712}{1.571}} \put(165,12.5){\arc{5}{1.571}{4.712}} \put(110,32.5){\arc{5}{4.712}{1.571}} \put(130,32.5){\arc{5}{1.571}{4.712}} \put(180,12.5){\arc{5}{4.712}{1.571}} \put(200,12.5){\arc{5}{1.571}{4.712}} \dottedline{4}(65,0)(65,50)(175,50)(175,0)(65,0) \allinethickness{1.0pt} \put(40,5){\line(0,1){40}} \put(60,10){\line(0,1){5}} \put(65,10){\arc{10}{1.571}{3.142}} \put(65,15){\arc{10}{3.142}{4.712}} \put(65,5){\line(1,0){5}} \put(65,20){\line(1,0){5}} \put(70,15){\arc{10}{4.712}{0}} \put(75,10){\line(0,1){5}} \put(70,10){\arc{10}{0}{1.571}} \put(170,15){\arc{10}{3.142}{4.712}} \put(170,20){\line(1,0){5}} \put(180,10){\line(0,1){5}} \put(175,10){\arc{10}{0}{1.571}} \put(175,15){\arc{10}{4.712}{0}} \put(170,5){\line(1,0){5}} \put(165,10){\line(0,1){5}} \put(170,10){\arc{10}{1.571}{3.142}} \put(200,5){\line(0,1){40}} \allinethickness{2.0pt} \put(40,12.5){\line(1,0){20}} \put(52,12.5){\vector(1,0){0}} \put(99.5,22.5){\vector(1,0){0}} \put(144.5,22.5){\vector(1,0){0}} \put(82,12.5){\vector(1,0){0}} \put(162,12.5){\vector(1,0){0}} \put(122,32.5){\vector(1,0){0}} \put(87,45){\vector(1,0){0}} \put(157,45){\vector(1,0){0}} \put(65,35){\line(1,0){20}} \put(65,45){\line(1,0){40}} \put(65,40){\arc{10}{1.571}{4.712}} \put(90,17.5){\line(0,1){12.5}} \put(85,17.5){\arc{10}{0}{1.571}} \put(85,30){\arc{10}{4.712}{0}} \put(105,27.5){\arc{10}{0}{1.571}} \put(105,40){\arc{10}{4.712}{0}} \put(110,27.5){\line(0,1){12.5}} \put(75,12.5){\line(1,0){10}} \put(90,22.5){\line(1,0){15}} \put(135,22.5){\line(1,0){15}} \put(110,32.5){\line(1,0){20}} \put(135,27.5){\arc{10}{1.571}{3.142}} \put(135,40){\arc{10}{3.142}{4.712}} \put(130,27.5){\line(0,1){12.5}} \put(155,12.5){\line(1,0){10}} \put(155,35){\line(1,0){20}} \put(135,45){\line(1,0){40}} \put(175,40){\arc{10}{4.712}{1.571}} \put(150,17.5){\line(0,1){12.5}} \put(155,17.5){\arc{10}{1.571}{3.142}} \put(155,30){\arc{10}{3.142}{4.712}} \put(180,12.5){\line(1,0){20}} \put(192,12.5){\vector(1,0){0}} \put(36,3){\makebox(0,0){$c$}} \put(36,47){\makebox(0,0){$a$}} \put(204,3){\makebox(0,0){$c$}} \put(204,47){\makebox(0,0){$a$}} \put(69,8){\makebox(0,0){$c$}} \put(69,17){\makebox(0,0){$a$}} \put(171,8){\makebox(0,0){$c$}} \put(171,17){\makebox(0,0){$a$}} \put(87,6){\makebox(0,0){$\beta'''$}} \put(153,6){\makebox(0,0){$\beta'''$}} \put(101,17){\makebox(0,0){$\beta$}} \put(101,40){\makebox(0,0){$\beta'$}} \put(139,17){\makebox(0,0){$\beta$}} \put(139,40){\makebox(0,0){$\beta'$}} \put(50,17.5){\makebox(0,0){$\beta''$}} \put(120,37.5){\makebox(0,0){$\beta''$}} \put(190,17.5){\makebox(0,0){$\beta''$}} \end{picture} \end{center} \caption{The vertical product $e_\beta *_v e_{\beta'}$} \label{v-prod3} \end{figure} Here, since we introduced the sum over $\beta'''$, the part can be similarly read for fixed $a$ and $c$ as $\sum_{j,k} S_j T_{\beta,\beta'}^{\beta'';k} (T_{\beta,\beta'}^{\beta'';k})^* S_j^*$, where the sum over $j$ runs over another orthonormal basis of isometries $S_i$ in the Hilbert space ${\rm{Hom}}(\beta,\co c a\co{\beta'})$. Since such bases $\{T_i\}$ and $\{S_j\}$ are related by a unitary matrix transformation (this is essentially ``unitarity of $6j$-symbols''), we conclude that the diagrams in Figs.\ \ref{v-prod2} and \ref{v-prod3} represent the same element in $\dta$. We now see that we first obtain a factor $\del {\beta''}{\beta'''}$. Next we can turn around the small arcs at the outer two trivalent vertices involving $\beta,\beta'$ and $\beta'''=\beta''$ so that we obtain a factor $d_\beta/d_{\beta''}$. Then, by ``stretching'' the diagram a bit, we can read the diagram for fixed $a,c,\beta''$ as \[ \begin{array}{l} \displaystyle\sum_{i,j,m=1}^{N_{\co c,a}^{\beta''}} \sum_{k,l=1}^{N_{\beta,\beta'}^{\beta''}} \frac{d_\beta d_{\beta'}}{d_{\beta''}} T_{\co c,a}^{\beta'';i} (T_{\co c,a}^{\beta'';i})^* T_{\co c,a}^{\beta'';j} (T_{\beta,\beta'}^{\beta'';l})^* T_{\beta,\beta'}^{\beta'';k} (T_{\beta,\beta'}^{\beta'';k})^* T_{\beta,\beta'}^{\beta'';l} (T_{\co c,a}^{\beta'';j})^* T_{\co c,a}^{\beta'';m} (T_{\co c,a}^{\beta'';m})^* \\[.4em] \qquad\qquad\qquad = \displaystyle\sum_{i=1}^{N_{\co c,a}^{\beta''}} \frac{d_\beta d_{\beta'}}{d_{\beta''}} N_{\beta,\beta'}^{\beta''}\, T_{\co c,a}^{\beta'';i} (T_{\co c,a}^{\beta'';i})^* \,. \end{array} \] Now proceeding with the summations over $a,c,\beta''$ yields the statement. \end{proof} Now consider the vector space with basis elements $[\beta]$, $\beta\in{}_M {\cal X}_M$ which we can endow with a product through $[\beta][\beta']=\sum_{\beta''}N_{\beta,\beta'}^{\beta''}[\beta'']$. We call the algebra defined this way the $M$-$M$ fusion rule algebra. Similarly we define the $N$-$N$ fusion rule algebra using morphisms in ${}_N {\cal X}_N$. \begin{definition} \label{Phi}{\rm We define a linear map $\Phi$ from the $M$-$M$ fusion rule algebra to ${\cal Z}_h$ by linear extension of $\Phi([\beta])=e_\beta/d_\beta$. }\end{definition} Theorem \ref{double2} now says that this map $\Phi$ is an isomorphism from the $M$-$M$ fusion rule algebra onto $({\cal{Z}}_h, *_v)$. Note that $({\cal{Z}}_h, *_v)$ is a non-unital subalgebra of $(\dta, *_v)$. The unit ${\bf1}_v$ of $(\dta, *_v)$ is given by ${\bf1}_v=\sum_\lambda f_\lambda$, where $f_\lambda=\sum_{a,b,j} f_{\lambda;a,b,j}^{a,b,j}$ whereas the unit of $({\cal{Z}}_h, *_v)$ is given by $e_0$. \begin{definition} \label{traces}{\rm We define two linear functionals $\varphi_h$ and $\tau_v$ on $\dta$ corresponding to the two product structures $*_h$ and $*_v$ by linear extension of \begin{equation} \begin{array}{rl} \varphi_h (e_{\beta;c,a,i}^{d,b,j}) &= \del ab \del cd \del ij \, d_a d_c d_\beta / w^2 \,,\\[.4em] \tau_v (f_{\lambda;c,d,j}^{a,b,i}) &= \del ac \del cd \del ij \, d_\lambda \,. \end{array} \label{eqstates} \end{equation} }\end{definition} Applied to an element in Fig.\ \ref{dta1} (Fig.\ \ref{dta2}) the functional $\varphi_h$ ($\tau_v$) can be characterized graphically as in Fig.\ \ref{hori-st} \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(240,40) \allinethickness{1.0pt} \put(30,10){\line(0,1){20}} \put(50,10){\line(0,1){20}} \put(135,20){\circle{20}} \allinethickness{2.0pt} \put(30,20){\vector(1,0){20}} \put(125,20){\vector(1,0){20}} \put(30,33){\makebox(0,0){$a$}} \put(50,33){\makebox(0,0){$b$}} \put(30,7){\makebox(0,0){$c$}} \put(50,7){\makebox(0,0){$d$}} \put(40,24){\makebox(0,0){$\beta$}} \put(135,33){\makebox(0,0){$a$}} \put(135,7){\makebox(0,0){$c$}} \put(135,24){\makebox(0,0){$\beta$}} \put(10,20){\makebox(0,0){$\varphi_h:$}} \put(25,20){\makebox(0,0){$S^*$}} \put(55,20){\makebox(0,0){$T$}} \put(120,20){\makebox(0,0){$S^*$}} \put(150,20){\makebox(0,0){$T$}} \put(87,20){\makebox(0,0){$\longmapsto \; \del ab \del cd \, \displaystyle\frac{d_a d_c}{w^2}$}} \put(200,20){\makebox(0,0){$= \del ab \del cd \, \displaystyle\frac{(d_a d_c)^{3/2}d_\beta^{1/2}}{w^2} \, \langle S,T \rangle$}} \end{picture} \end{center} \caption{The horizontal functional $\varphi_h$} \label{hori-st} \end{figure} (Fig.\ \ref{vert-st}). \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(240,40) \allinethickness{1.0pt} \put(30,10){\line(1,0){20}} \put(30,30){\line(1,0){20}} \put(145,20){\circle{20}} \allinethickness{0.3pt} \put(40,30){\vector(0,-1){20}} \put(145,30){\vector(0,-1){20}} \put(27,30){\makebox(0,0){$a$}} \put(53,30){\makebox(0,0){$b$}} \put(27,10){\makebox(0,0){$c$}} \put(53,10){\makebox(0,0){$d$}} \put(37,20){\makebox(0,0){$\lambda$}} \put(142,20){\makebox(0,0){$\lambda$}} \put(40,35){\makebox(0,0){$s^*$}} \put(40,5){\makebox(0,0){$t$}} \put(145,35){\makebox(0,0){$s^*$}} \put(145,5){\makebox(0,0){$t$}} \put(10,20){\makebox(0,0){$\tau_v:$}} \put(130,20){\makebox(0,0){$a$}} \put(160,20){\makebox(0,0){$b$}} \put(90,20){\makebox(0,0){$\longmapsto \;\del ac \del bd\, \displaystyle \sqrt{d_a d_b}$}} \put(205,20){\makebox(0,0){$= \del ab \del cd \, d_a d_b d_\lambda^{1/2} \, \langle s,t \rangle$}} \end{picture} \end{center} \caption{The vertical functional $\tau_v$} \label{vert-st} \end{figure} Therefore these functionals correspond to closing the open ends of a diagram with prefactors as in the middle part of Figs.\ \ref{hori-st} and \ref{vert-st}. Recall that the global index of ${}_N {\cal X}_N$ is given by $w=\sum_{\lambda\in{}_N {\cal X}_N} d_\lambda^2$. Note that we have sector decompositions $[a\iota]=\sum_\lambda \langle \lambda,a\iota \rangle [\lambda]$ and hence $d_a d_\iota=\sum_\lambda \langle \lambda,a\iota \rangle d_\lambda$ for any $a\in{}_N {\cal X}_M$. Using Frobenius reciprocity $\langle \lambda,a\iota \rangle = \langle \lambda\co\iota,a \rangle$ we obtain similarly $d_\lambda d_\iota=\sum_a \langle \lambda,a\iota \rangle d_a$. Hence $w=\sum_\lambda d_\lambda^2 = \sum_{\lambda,a} \langle \lambda,a\iota \rangle d_\lambda d_a/d_\iota = \sum_a d_a^2$. Similarly we obtain $w=\sum_\beta d_\beta^2$ (cf.\ \cite{O3}). \begin{lemma} \label{tr-val} We have $\varphi_h(e_\beta)=d_\beta^2/w$. In particular, the functional $\varphi_h$ is a faithful state on $(\dta,*_h)$. The functional $\tau_v$ is a (un-normalized) faithful trace on $(\dta,*_v)$. \end{lemma} \begin{proof} By Definition \ref{traces} and Fig.\ \ref{e-beta}, we compute \[ \varphi_h(e_\beta) = \sum_{a,b\in{}_N {\cal X}_M} N_{\bar a, b}^\beta d_a d_b d_\beta w^{-2} = \sum_{a\in{}_N {\cal X}_M} \left(\sum_{b\in{}_N {\cal X}_M} N_{a,\beta}^b d_b\right) d_\beta d_a w^{-2} = d_\beta^2 w^{-1}\,.\] Since the horizontal unit ${\bf1}_h$ is given by ${\bf1}_h=\sum_\beta e_\beta$ we find that $\varphi({\bf1}_h)=1$. As $\varphi_h$ sends off-diagonal matrix units to zero and the diagonal ones to strictly positive numbers, this proves that $\varphi_h$ is a faithful state. Obviously also $\tau_v$ sends off-diagonal matrix units (with respect to $*_v$) to zero and the diagonal ones to strictly positive numbers, and hence it is a strictly positive functional but it is not normalized. The trace property $\tau_v(xy)=\tau_v(yx)$ is clear from the definition of $\tau_v$ using matrix units for $x$ and $y$. \end{proof} For $\tau_v$ we could have gained analogous properties as for $\varphi_h$ by replacing the scalar $d_\lambda$ in \erf{eqstates} by $d_a d_b d_\lambda/w^2$ (and by multiplying the scalars in Fig.\ \ref{vert-st} also by $d_a d_b/w^2$). However, we chose a different normalization on each matrix unit in order to turn $\tau_v$ into a trace on $(\dta,*_v)$. Later we want to study the center $({\cal{Z}}_h,*_v)$ which is, as we have seen, a subalgebra of $(\dta,*_v)$. Therefore $\tau_v$ provides a faithful trace on $({\cal{Z}}_h,*_v)$ but it has in general different weightings on its simple summands. To construct from $\tau_v$ a trace which sends one-dimensional projections to one will in particular be possible in the case that ${}_N {\cal X}_N$ is non-degenerately braided, see Subsect.\ \ref{sec-dfa} below. This is also the case in the following most basic example of the double triangle algebra. Let $N$ be a type III factor and $G$ a finite group acting freely on $N$. Consider the subfactor $N\subset N\rtimes G=M$. Then (with the minimal choice for ${\cal{X}}$) the double triangle algebra $\dta$ for this subfactor is just the group algebra of $G$. That is, the double triangle algebra is spanned by the group elements linearly. The horizontal product is given by the group multiplication. By Proposition \ref{double2} we conclude that the minimal central projections in $\dta$ and thus irreducible $M$-$M$ sectors are labelled by the irreducible representations of $G$. (Of course, this identification of the $M$-$M$ sectors is well-known for that example.) The functional $\tau_v$ gives the standard trace on the group algebra, and the vertical product corresponds to the ordinary tensor product of group representations. \section{$\alpha$-Induction, Chiral Generators and Modular Invariants} \label{sec-aicpmodinv} \subsection{Relating $\alpha$-induction to chiral generators} \label{sec-iden} We will now define chiral generators for braided subfactors and prove that the concepts of $\alpha$-induction and chiral generators are essentially the same. For the rest of this paper deal with the following \begin{assumption} \label{set-braid} {\rm In addition to Assumption \ref{set-fin} we now assume that the system ${}_N {\cal X}_N$ is braided. }\end{assumption} With the braiding we have now the notion of $\alpha$-induction in the sense of Subsect.\ \ref{sec-aindbsf}. From now on we are also dealing with crossings of $N$-$N$ wires and mixed crossings introduced in Subsect.\ \ref{sec-aindbsf}. We now present chiral generators as our version of a definition Ocneanu originally introduced for systems of bimodules arising from A-D-E Dynkin diagrams in \cite{O7}. The construction of the chiral generator is similar to the ``Ocneanu projection'' in the tube algebra \cite{O6} (see also \cite{EK4}) and also related to Izumi's analysis \cite{I3} of the tube algebra in terms of sectors for the Longo-Rehren inclusion \cite{LR}. \begin{definition} \label{chiral1}{\rm For any $\lambda\in{}_N {\cal X}_N$, we define an element $p^+_\lambda\in\dta$ by the diagram on the left-hand side of Fig.\ \ref{chiral-proj} and call it a {\sl chiral generator}. Similarly, we also define $p^-_\lambda$ by exchanging over- and undercrossings. }\end{definition} \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(215,40) \allinethickness{1.0pt} \put(10,17){\makebox(0,0){$\displaystyle\sum_{a,b}$}} \put(30,5){\line(0,1){8}} \put(30,17){\line(0,1){18}} \put(50,5){\line(0,1){8}} \put(50,17){\line(0,1){18}} \allinethickness{2.0pt} \put(30,15){\line(1,0){20}} \put(42,15){\vector(1,0){0}} \allinethickness{0.3pt} \put(30,20){\arc{10}{1.571}{4.712}} \put(50,20){\arc{10}{4.712}{1.571}} \put(30,25){\arc{5}{1.571}{4.712}} \put(50,25){\arc{5}{4.712}{1.571}} \put(25,35){\makebox(0,0){$a$}} \put(55,35){\makebox(0,0){$a$}} \put(25,5){\makebox(0,0){$b$}} \put(55,5){\makebox(0,0){$b$}} \put(40,21){\makebox(0,0){$\alpha_\lambda^+$}} \put(70,20){\makebox(0,0){$=$}} \allinethickness{1.0pt} \put(80,17){\makebox(0,0){$\displaystyle\sum_{a,b}$}} \put(100,5){\line(0,1){18}} \put(100,27){\line(0,1){8}} \put(120,5){\line(0,1){18}} \put(120,27){\line(0,1){8}} \allinethickness{2.0pt} \put(100,25){\line(1,0){20}} \put(112,25){\vector(1,0){0}} \allinethickness{0.3pt} \put(100,20){\arc{10}{1.571}{4.712}} \put(120,20){\arc{10}{4.712}{1.571}} \put(100,15){\arc{5}{1.571}{4.712}} \put(120,15){\arc{5}{4.712}{1.571}} \put(95,35){\makebox(0,0){$a$}} \put(125,35){\makebox(0,0){$a$}} \put(95,5){\makebox(0,0){$b$}} \put(125,5){\makebox(0,0){$b$}} \put(110,19){\makebox(0,0){$\alpha_\lambda^+$}} \put(140,20){\makebox(0,0){$=$}} \allinethickness{0.3pt} \put(150,17){\makebox(0,0){$\displaystyle\sum_{a,b,\nu}$}} \put(185,10){\line(0,1){7}} \put(185,30){\line(0,-1){7}} \put(186,20.5){\vector(1,0){0}} \put(185,10){\vector(0,-1){0}} \put(185,30){\arc{20}{0}{3.14}} \put(185,30){\arc{5}{0}{3.14}} \put(175,30){\arc{5}{0}{3.14}} \put(195,30){\arc{5}{0}{3.14}} \put(185,10){\arc{5}{3.14}{6.28}} \allinethickness{1.0pt} \put(170,10){\line(1,0){30}} \put(170,30){\line(1,0){30}} \put(170,35){\arc{10}{1.571}{3.142}} \put(200,35){\arc{10}{0}{1.571}} \put(170,5){\arc{10}{3.142}{4.712}} \put(200,5){\arc{10}{4.712}{0}} \put(195,22){\makebox(0,0){$\lambda$}} \put(181,14){\makebox(0,0){$\nu$}} \put(160,5){\makebox(0,0){$b$}} \put(210,5){\makebox(0,0){$b$}} \put(160,35){\makebox(0,0){$a$}} \put(210,35){\makebox(0,0){$a$}} \put(180,34){\makebox(0,0){$b$}} \put(190,34){\makebox(0,0){$b$}} \end{picture} \end{center} \caption{A chiral generator $p^+_\lambda$} \label{chiral-proj} \end{figure} Note that we do {\sl not} assume the non-degeneracy of the braiding for the definition $p^+_\lambda$. We obtain the diagram in the middle from the one on the left-hand side in Fig.\ \ref{chiral-proj} by applying two IBFE's. This way we obtain two twists in the semi-circular thin wires which correspond to the label $\lambda$ but they give complex conjugate phases so that their effects cancel out. The diagram on the right-hand side is obtained by Lemma \ref{idexp} and application of the IBFE, and this shows that our definition coincides with Ocneanu's notion given in his setting. Since $\alpha_\lambda^\pm \iota=\iota\lambda$ we find that each irreducible subsector $[\beta]$ of $[\alpha_\lambda^\pm]$ is the equivalence class of some $\beta\in{}_M {\cal X}_M$ if $\lambda\in{}_N {\cal X}_N$. Therefore we have the sector decomposition $[\alpha_\lambda^\pm]=\sum_{\beta\in{}_M {\cal X}_M}\langle\beta,\alpha_\lambda^\pm\rangle [\beta]$, and we can consider $[\alpha_\lambda^\pm]$ as an element of the $M$-$M$ fusion algebra. The relation between the sector decomposition of $[\alpha_\lambda^\pm]$ and the chiral generator is clarified by the following result. \begin{theorem} \label{identify} For any $\lambda\in{}_N {\cal X}_N$, we have $d_\lambda^{-1} p^\pm_\lambda = \sum_{\beta\in{}_M {\cal X}_M} d_\beta^{-1} \langle\beta,\alpha_\lambda^+\rangle e_\beta$, and consequently $p^\pm_\lambda= d_\lambda \Phi([\alpha^\pm_\lambda])$. In particular, $p^\pm_\lambda$ is in the center ${\cal{Z}}_h$. \end{theorem} \begin{proof} We only show the statement for the $+$-sign; the other case is analogous. First we fix $a,b\in{}_N {\cal X}_M$ and $\lambda\in{}_N {\cal X}_N$. For each $\beta\in{}_M {\cal X}_M$ we choose orthonormal bases of isometries $T_{\co b a}^{\beta;i}\in{\rm{Hom}}(\beta,\co b a)$, $i=1,2,...,N_{\co b,a}^\beta$, so that $\sum_{\beta,i} T_{\co b a}^{\beta;i} (T_{\co b a}^{\beta;i})^* = {\bf1}_M$. Using Frobenius reciprocity, we obtain an orthonormal basis of isometries ${\cal{L}}^{-1}_b(T_{\co b a}^{\beta;i})=d_a^{1/2} d_b^{1/2} d_\beta^{-1/2} b(T_{\co b a}^{\beta;i})^* {\co r}_b\in{\rm{Hom}}(a,b\beta)$. Here we chose an isometry ${\co r}_b\in{\rm{Hom}}({\rm{id}}_N,b\co b)$ such that there is an isometry $R_b\in{\rm{Hom}}({\rm{id}}_M,\co b b)$ subject to relations $b(R_b)^*{\co r}_b=d_b^{-1}{\bf1}_N$ and $\co b({\co r}_b)^*R_b=d_b^{-1}{\bf1}_M$, as usual. Choosing also orthonormal bases of isometries $V_{\beta;\ell}\in{\rm{Hom}}(\beta,\alpha_\lambda^+)$, $\ell=1,2,...,\langle \beta, \alpha_\lambda^+ \rangle$, for each $\beta\in{}_M {\cal X}_M$ (so that $\sum_{\beta,\ell} V_{\beta;\ell} V_{\beta;\ell}^*={\bf1}_M$) we find that $\{ b(V_{\beta;\ell}) {\cal{L}}^{-1}_b(T_{\co b a}^{\beta;i}) \}_{\beta,i,\ell}$ gives an orthonormal basis of isometries of ${\rm{Hom}}(a,b\alpha_\lambda^+)$. Finally, using Proposition \ref{stat}, we find that putting \[ s_{\beta;\ell,i}=\epsp \lambda{b\iota} ^* b(V_{\beta;\ell}){\cal{L}}^{-1}_b(T_{\co b a}^{\beta;i}) = \sqrt\frac{d_a d_b}{d_\beta} \epsp \lambda{b\iota} ^* b(V_{\beta;\ell} (T_{\co b a}^{\beta;i})^*){\co r}_b \] defines an orthonormal basis of isometries $\{s_{\beta;\ell,i}\}_{\beta,i,\ell}$ of ${\rm{Hom}}(a,\lambda b)$. Then we have for any $\ell=1,2,...,\langle \beta,\alpha_\lambda^+ \rangle$ by the elementary relations for the intertwiners $R_b,{\co r}_b$ the following identity: \[ \begin{array}{ll} T_{\co b a}^{\beta;i}(T_{\co b a}^{\beta;i})^* &= d_b^2 \, \co b ({\co r}_b)^* \, \co b b(T_{\co b a}^{\beta;i} V_{\beta;\ell}^*) \, R_b R_b^* \, \co b b(V_{\beta;\ell} (T_{\co b a}^{\beta;i})^*) \, \co b({\co r}_b) \\[.5em] &= \displaystyle\frac{d_\beta d_b}{d_a} \, \co b (s_{\beta;\ell,i} \epsp \lambda{b\iota} ^*) \, R_bR_b^* \, \co b( \epsp \lambda{b\iota} s_{\beta;\ell,i}^* )\,. \end{array} \] The second line yields graphically exactly the diagram in Fig.\ \ref{inter1} where we read the diagram from the left to the right in order to interpret it as an intertwiner in $\dtap$. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(120,50) \put(55,25){\line(1,0){10}} \put(45,35){\vector(1,-1){10}} \put(85,25){\vector(1,0){10}} \put(95,25){\line(1,1){10}} \allinethickness{1.0pt} \put(35,5){\line(1,0){20}} \put(35,35){\line(1,0){10}} \put(45,35){\line(1,1){10}} \put(65,35){\line(0,-1){7}} \put(65,15){\line(0,1){7}} \put(55,15){\arc{20}{0}{1.57}} \put(55,35){\arc{20}{4.71}{6.28}} \put(115,5){\line(-1,0){20}} \put(115,35){\line(-1,0){10}} \put(105,35){\line(-1,1){10}} \put(85,35){\line(0,-1){7}} \put(85,15){\line(0,1){7}} \put(95,15){\arc{20}{1.57}{3.14}} \put(95,35){\arc{20}{3.14}{4.71}} \allinethickness{2.0pt} \put(65,25){\vector(1,0){10}} \put(75,25){\line(1,0){10}} \put(32,5){\makebox(0,0){$b$}} \put(32,35){\makebox(0,0){$a$}} \put(65,45){\makebox(0,0){$b$}} \put(59,21){\makebox(0,0){$\lambda$}} \put(55,34){\makebox(0,0){$s_{\beta;\ell,i}$}} \put(118,5){\makebox(0,0){$b$}} \put(118,35){\makebox(0,0){$a$}} \put(85,45){\makebox(0,0){$b$}} \put(91,21){\makebox(0,0){$\lambda$}} \put(75,20){\makebox(0,0){$\alpha_\lambda^+$}} \put(95,34.5){\makebox(0,0){$s_{\beta;\ell,i}^*$}} \put(13,22){\makebox(0,0){$\displaystyle\frac{d_\beta} {\sqrt{d_\lambda d_a d_b}}$}} \end{picture} \end{center} \caption{Diagram for $T_{\co b a}^{\beta;i}(T_{\co b a}^{\beta;i})^*$} \label{inter1} \end{figure} Now let us take on both sides first the summation over $i=1,2,...,N_{\co b,a}^\beta$. Then the left-hand side gives exactly the ${\rm{Hom}}(\co b a,\co b a)$ part of $e_\beta$ (in $\dtap$) as defined in Definition \ref{defe-beta}. Next we divide by $d_\beta$ and we proceed with the summation over $\ell=1,2,...,\langle\beta,\alpha_\lambda^+\rangle$ and $\beta\in{}_M {\cal X}_M$. On the left-hand side we obtain the ${\rm{Hom}}(\co b a,\co b a)$ part of $\sum_\beta d_\beta^{-1} \langle\beta,\alpha_\lambda^+\rangle e_\beta$ this way, and this is exactly the ${\rm{Hom}}(\co b a,\co b a)$ part of $\Phi([\alpha_\lambda^+])$. On the right-hand side we now have a summation over the full basis $\{s_{\beta;\ell,i}\}_{\beta,i,\ell}$ of ${\rm{Hom}}(a,\lambda b)$. Therefore we can use the graphical convention of Fig.\ \ref{e-beta} to put a small semi-circle around the wire labelled by $\lambda$ at the two trivalent vertices. This gives us a factor $\sqrt{d_a d_b/d_\lambda}$ so that only a factor $d_\lambda^{-1}$ remains from the original prefactor in Fig.\ \ref{inter1}. Thus, by repeating the above procedure for all $a,b\in{}_N {\cal X}_M$ and making finally the summation over $a,b\in{}_N {\cal X}_M$, we obtain on the left the full $\Phi([\alpha_\lambda^+])$ whereas the right-hand side gives graphically the diagram in Fig.\ \ref{ident}. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(115,50) \put(50,25){\line(1,0){10}} \put(40,35){\vector(1,-1){10}} \put(80,25){\vector(1,0){10}} \put(90,25){\line(1,1){10}} \put(40,35){\arc{5}{5.49}{9.42}} \put(100,35){\arc{5}{0}{3.93}} \allinethickness{1.0pt} \put(30,5){\line(1,0){20}} \put(30,35){\line(1,0){10}} \put(40,35){\line(1,1){10}} \put(60,35){\line(0,-1){7}} \put(60,15){\line(0,1){7}} \put(50,15){\arc{20}{0}{1.57}} \put(50,35){\arc{20}{4.71}{6.28}} \put(110,5){\line(-1,0){20}} \put(110,35){\line(-1,0){10}} \put(100,35){\line(-1,1){10}} \put(80,35){\line(0,-1){7}} \put(80,15){\line(0,1){7}} \put(90,15){\arc{20}{1.57}{3.14}} \put(90,35){\arc{20}{3.14}{4.71}} \allinethickness{2.0pt} \put(60,25){\vector(1,0){10}} \put(70,25){\line(1,0){10}} \put(27,5){\makebox(0,0){$b$}} \put(27,35){\makebox(0,0){$a$}} \put(60,45){\makebox(0,0){$b$}} \put(54,21){\makebox(0,0){$\lambda$}} \put(113,5){\makebox(0,0){$b$}} \put(113,35){\makebox(0,0){$a$}} \put(80,45){\makebox(0,0){$b$}} \put(86,21){\makebox(0,0){$\lambda$}} \put(70,20){\makebox(0,0){$\alpha_\lambda^+$}} \put(11,22){\makebox(0,0){$\displaystyle \sum_{a,b}\; \frac{1}{d_\lambda}$}} \end{picture} \end{center} \caption{The image $\Phi([\alpha^+_\lambda])=\sum_\beta d_\beta^{-1} \langle\beta,\alpha_\lambda^+\rangle e_\beta$} \label{ident} \end{figure} The diagram on the left-hand side in Fig.\ \ref{chiral-proj} is obtained from Fig.\ \ref{ident}, up to the factor $d_\lambda$, by a topological move. \end{proof} Note that it was not clear from the definition that the chiral generators are in the center ${\cal{Z}}_h$, but Theorem \ref{identify} proves this centrality as it states that $p_\lambda^\pm$ is a linear combination of $e_\beta$'s. Also note that if $\alpha^\pm_\lambda$ is irreducible then $p^\pm_\lambda$ is a (horizontal) projection, however, if $\alpha^\pm_\lambda$ is not irreducible, then $p_\lambda^\pm$ is a sum over projections with weight coefficients arising from the nature of the isomorphism $\Phi$ in Definition \ref{Phi}. Two of us \cite[Subsection 3.3]{BE3} established a relative braiding between the two kinds of $\alpha$-induction, which holds in a fairly general context. (It does neither depend on chiral locality nor even on finite depth.) Theorem \ref{identify} now shows that Ocneanu's relative braiding \cite{O7} is a special case of the analysis in \cite[Subsection 3.3]{BE3}. {}From Theorem \ref{identify} and the homomorphism property of $\alpha$-induction \cite[Lemma 3.10]{BE1}, we obtain immediately the following \begin{corollary} \label{homom} The chiral generators $p^\pm_\lambda$ are in ${\cal{Z}}_h$. For $\lambda,\mu\in{}_N {\cal X}_N$, we have \[ p^\pm_\lambda *_v p^\pm_\mu = \sum_{\nu\in{}_N {\cal X}_N} \frac{d_\lambda d_\mu}{d_\nu} N_{\lambda,\mu}^\nu \, p^\pm_\nu \,. \] \end{corollary} Note that this corollary shows that the $M$-$M$ fusion rule algebra contains two representations of the $N$-$N$ fusion rule algebra. \subsection{Modular invariants for braided subfactors} \label{sec-modular} We will now show that a notion of ``modular invariant'' arises naturally for a braided subfactor. We first note that under Assumption \ref{set-braid}, we have matrices $Y=(Y_{\lambda,\mu})$ and $T=(T_{\lambda,\mu})$ for the system ${\Delta}={}_N {\cal X}_N$ as in Subsection \ref{braid-prelim}. We recall that in the case that the braiding is non-degenerate, the matrix $S=w^{-1/2}Y$ is unitary and the matrices $S$ and (the diagonal) $T$ obey the Verlinde modular algebra by Theorem \ref{ST}. Motivated by the results of \cite{BE3} we now construct a certain matrix $Z$ commuting with $Y$ and $T$ such that it is a ``modular invariant mass matrix'' in the usual sense of conformal field theory whenever the braiding is non-degenerate. \begin{definition} \label{modular1} {\rm For a system $\cal X$ satisfying Assumption \ref{set-braid}, we define a matrix $Z$ with entries $Z_{\lambda,\mu}=\langle \alpha_\lambda^+,\alpha_\mu^- \rangle$, $\lambda,\mu\in{}_N {\cal X}_N$. }\end{definition} As $Z_{\lambda,\mu}$ is by definition a dimension and since $\alpha_{{\rm{id}}_N}^\pm={\rm{id}}_M$ is irreducible by virtue of the factor property of $M$, the matrix elements obviously satisfy the conditions in \erf{massmatZ} for $\lambda,\mu\in{}_N {\cal X}_N$, where the label ``0'' refers as usual to the identity morphism ${\rm{id}}_N\in{}_N {\cal X}_N$. We relate the definition of $Z$ to the chiral generators by the following \begin{theorem} \label{modular1+} We have the identity \begin{equation} Z_{\lambda,\mu} = \frac w{d_\lambda d_\mu} \, \varphi_h(p_\lambda^+ *_h p_\mu^-) \,, \qquad \lambda,\mu\in{}_N {\cal X}_N \,. \label{Zvarphipp} \end{equation} Therefore the number $Z_{\lambda,\mu}$ is graphically represented as in Fig.\ \ref{Zgraph}. \end{theorem} \begin{figure}[htb] \begin{center} \unitlength 0.7mm \begin{picture}(115,50) \allinethickness{0.3pt} \put(34,23){\makebox(0,0){$Z_{\lambda,\mu} \;= \;\displaystyle \sum_{b,c} \;\;\frac{d_b d_c}{w d_\lambda d_\mu}$}} \put(80,5){\line(0,1){10}} \put(80,45){\line(0,-1){10}} \put(100,5){\line(0,1){8}} \put(100,45){\line(0,-1){8}} \put(80,45){\arc{5}{0}{3.142}} \put(100,45){\arc{5}{0}{3.142}} \put(80,5){\arc{5}{3.142}{0}} \put(100,5){\arc{5}{3.142}{0}} \allinethickness{2.0pt} \put(80,15){\line(0,1){20}} \put(100,17){\line(0,1){16}} \put(80,23){\vector(0,-1){0}} \put(100,27){\vector(0,1){0}} \allinethickness{1.0pt} \put(77,5){\line(1,0){26}} \put(77,45){\line(1,0){26}} \put(77,15){\line(1,0){1}} \put(77,35){\line(1,0){1}} \put(82,15){\line(1,0){21}} \put(82,35){\line(1,0){21}} \put(77,40){\arc{10}{1.571}{4.712}} \put(103,40){\arc{10}{4.712}{1.571}} \put(77,10){\arc{10}{1.571}{4.712}} \put(103,10){\arc{10}{4.712}{1.571}} \put(90,2){\makebox(0,0){$c$}} \put(90,48){\makebox(0,0){$c$}} \put(90,12){\makebox(0,0){$b$}} \put(90,38){\makebox(0,0){$b$}} \put(74,25.4){\makebox(0,0){$\alpha_\lambda^+$}} \put(107,24.6){\makebox(0,0){$\alpha_\mu^-$}} \end{picture} \end{center} \caption{Graphical representation of $Z_{\lambda,\mu}$} \label{Zgraph} \end{figure} \begin{proof} {}From Theorem \ref{identify} we obtain \[ \sum_{\beta\in{}_M {\cal X}_M}\frac 1{d_\beta} \langle \alpha^+_\lambda,\beta \rangle e_\beta = \frac 1{d_\lambda} p_\lambda^+ \,. \] Hence \[ \sum_{\beta\in{}_M {\cal X}_M} \frac 1{d_\beta^2} \langle \alpha^+_\lambda,\beta \rangle \langle \alpha^-_\mu,\beta\rangle e_\beta = \frac 1{d_\lambda d_\mu} \, p_\lambda^+ *_h p_\mu^- \,. \] Application of the horizontal state $\varphi_h$ of Definition \ref{traces} and multiplication by $w$ yields \erf{Zvarphipp} since $[\alpha_\lambda^+]$ and $[\alpha_\mu^-]$ decompose into sectors $[\beta]$ with $\beta\in{}_M {\cal X}_M$, and by Lemma \ref{tr-val}. Now the right-hand side of\ \erf{Zvarphipp} is given graphically by the diagram on the left in Fig.\ \ref{Z1}, \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(215,60) \allinethickness{1.0pt} \put(15,30){\makebox(0,0){$\displaystyle\sum_{b,c}\; \frac{d_b d_c}{w d_\lambda d_\mu}$}} \put(40,10){\line(0,1){23}} \put(40,37){\line(0,1){13}} \put(60,15){\line(0,1){18}} \put(60,37){\line(0,1){8}} \put(45,10){\arc{10}{1.571}{3.142}} \put(45,50){\arc{10}{3.142}{4.712}} \put(67.5,15){\arc{15}{0}{3.142}} \put(67.5,45){\arc{15}{3.142}{0}} \put(90,10){\arc{10}{0}{1.571}} \put(90,50){\arc{10}{4.712}{0}} \put(75,15){\line(0,1){30}} \put(95,10){\line(0,1){40}} \put(45,5){\line(1,0){45}} \put(45,55){\line(1,0){45}} \allinethickness{2.0pt} \put(40,35){\line(1,0){20}} \put(52,35){\vector(1,0){0}} \put(77,35){\line(1,0){16}} \put(87,35){\vector(1,0){0}} \allinethickness{0.3pt} \put(40,30){\arc{10}{1.571}{4.712}} \put(60,30){\arc{10}{4.712}{1.571}} \put(40,25){\arc{5}{1.571}{4.712}} \put(60,25){\arc{5}{4.712}{1.571}} \put(75,30){\arc{10}{1.571}{4.212}} \put(95,30){\arc{10}{5.212}{1.571}} \put(75,25){\arc{5}{1.571}{4.712}} \put(95,25){\arc{5}{4.712}{1.571}} \put(36,55){\makebox(0,0){$c$}} \put(99,5){\makebox(0,0){$b$}} \put(67.5,48.5){\makebox(0,0){$c$}} \put(67.5,12.5){\makebox(0,0){$b$}} \put(50,30){\makebox(0,0){$\alpha_\lambda^+$}} \put(85,30){\makebox(0,0){$\alpha_\mu^-$}} \put(110,30){\makebox(0,0){$=$}} \allinethickness{1.0pt} \put(135,30){\makebox(0,0){$\displaystyle\sum_{b,c}\; \frac{d_a d_b}{w d_\lambda d_\mu}$}} \put(160,10){\line(0,1){15.5}} \put(160,29.5){\line(0,1){20.5}} \put(180,15){\line(0,1){10.5}} \put(180,29.5){\line(0,1){15.5}} \put(165,10){\arc{10}{1.571}{3.142}} \put(165,50){\arc{10}{3.142}{4.712}} \put(187.5,15){\arc{15}{0}{3.142}} \put(187.5,45){\arc{15}{3.142}{0}} \put(210,10){\arc{10}{0}{1.571}} \put(210,50){\arc{10}{4.712}{0}} \put(195,15){\line(0,1){30}} \put(215,10){\line(0,1){40}} \put(165,5){\line(1,0){45}} \put(165,55){\line(1,0){45}} \allinethickness{2.0pt} \put(160,27.5){\line(1,0){20}} \put(172,27.5){\vector(1,0){0}} \put(162,32.5){\line(1,0){16}} \put(168,32.5){\vector(-1,0){0}} \allinethickness{0.3pt} \put(160,22.5){\arc{10}{1.571}{4.712}} \put(180,22.5){\arc{10}{4.712}{1.571}} \put(160,17.5){\arc{5}{1.571}{4.712}} \put(180,17.5){\arc{5}{4.712}{1.571}} \put(160,37.5){\arc{10}{2.071}{4.712}} \put(180,37.5){\arc{10}{4.712}{1.071}} \put(160,42.5){\arc{5}{1.571}{4.712}} \put(180,42.5){\arc{5}{4.712}{1.571}} \put(156,30){\makebox(0,0){$c$}} \put(184,30){\makebox(0,0){$c$}} \put(199,30){\makebox(0,0){$b$}} \put(211,30){\makebox(0,0){$b$}} \put(170,22.5){\makebox(0,0){$\alpha_\lambda^+$}} \put(170,38.5){\makebox(0,0){$\alpha_\mu^-$}} \end{picture} \end{center} \caption{The scalar $w d_\lambda^{-1} d_\mu^{-1} \varphi_h(p_\lambda^+ *_h p_\mu^-)$} \label{Z1} \end{figure} and we can slide around the trivalent vertices to obtain the diagram on the right-hand side. Without changing the scalar value we can now open the outer wire labelled by $b$ and close it on the other side, as in Fig.\ \ref{cut}. This way we obtain the picture in Fig.\ \ref{Zgraph} up to a 90 degree rotation, but a rotation is irrelevant for the scalar values. \end{proof} We remark that we can apply Lemma \ref{idexp} to replace the two horizontal wires labelled by $b$ by a summation over a thin wire $\nu$, and this way we obtain an equivalent diagram from Fig.\ \ref{Zgraph} for the matrix elements $Z_{\lambda,\mu}$, which only consists of thin ($N$-$N$) wires $\lambda,\mu,\nu$ and thick ($N$-$M$) wires $b,c$ but which does not involve very thick ($M$-$M$) wires labelled by $\alpha$-induced morphisms $\alpha_\lambda^+,\alpha_\mu^-$. \begin{theorem} \label{modular2} The matrix $Z$ of Definition \ref{modular1} commutes with the matrices $Y$ and $T$ of the system ${}_N {\cal X}_N$. \end{theorem} \begin{proof} Using the diagram for the matrix elements $Y_{\nu,\lambda}$ in Fig.\ \ref{Ymatrix}, the sum $\sum_\lambda Y_{\nu,\lambda}Z_{\lambda,\mu}$ can be represented by the diagram on the left-hand side of Fig.\ \ref{[Y,Z]1}. \begin{figure}[htb] \begin{center} \unitlength 0.7mm \begin{picture}(174,50) \allinethickness{0.3pt} \put(12,23){\makebox(0,0){$\displaystyle \sum_{b,c,\lambda} \;\frac{d_b d_c}{w d_\mu}$}} \put(40,5){\line(0,1){10}} \put(40,45){\line(0,-1){5.5}} \put(40,35.5){\line(0,-1){15.5}} \put(60,5){\line(0,1){8}} \put(60,45){\line(0,-1){23}} \put(40,45){\arc{5}{0}{3.142}} \put(60,45){\arc{5}{0}{3.142}} \put(40,5){\arc{5}{3.142}{0}} \put(60,5){\arc{5}{3.142}{0}} \put(40,32.5){\arc{10}{1.971}{1.171}} \put(40,22){\vector(0,-1){0}} \put(60,34){\vector(0,1){0}} \put(44.8,34){\vector(0,1){0}} \allinethickness{2.0pt} \put(40,15){\line(0,1){5}} \put(60,17){\line(0,1){1}} \allinethickness{1.0pt} \put(37,5){\line(1,0){26}} \put(37,45){\line(1,0){26}} \put(37,15){\line(1,0){1}} \put(37,20){\line(1,0){1}} \put(42,15){\line(1,0){21}} \put(42,20){\line(1,0){21}} \put(37,25){\arc{10}{1.571}{3.142}} \put(63,25){\arc{10}{0}{1.571}} \put(37,40){\arc{10}{3.142}{4.712}} \put(63,40){\arc{10}{4.712}{0}} \put(32,25){\line(0,1){15}} \put(68,25){\line(0,1){15}} \put(37,10){\arc{10}{1.571}{4.712}} \put(63,10){\arc{10}{4.712}{1.571}} \put(50,2){\makebox(0,0){$c$}} \put(50,48){\makebox(0,0){$c$}} \put(50,12){\makebox(0,0){$b$}} \put(50,24){\makebox(0,0){$b$}} \put(36,25){\makebox(0,0){$\lambda$}} \put(64,24){\makebox(0,0){$\mu$}} \put(50,32.5){\makebox(0,0){$\nu$}} \allinethickness{0.3pt} \put(105,23){\makebox(0,0){$= \;\displaystyle \sum_{a,b,c,\lambda} \;\frac{d_a d_b d_c}{w d_\mu d_\nu}$}} \put(145,5){\line(0,1){10}} \put(145,45){\line(0,-1){5.5}} \put(145,35.5){\line(0,-1){15.5}} \put(165,5){\line(0,1){8}} \put(165,45){\line(0,-1){23}} \put(145,45){\arc{5}{0}{3.142}} \put(165,45){\arc{5}{0}{3.142}} \put(145,5){\arc{5}{3.142}{0}} \put(165,5){\arc{5}{3.142}{0}} \put(145,32.5){\arc{10}{4.712}{1.171}} \put(137,37.5){\line(1,0){8}} \put(137,27.5){\line(1,0){6}} \put(137,37.5){\arc{5}{4.712}{1.571}} \put(137,27.5){\arc{5}{4.712}{1.571}} \put(145,22){\vector(0,-1){0}} \put(165,34){\vector(0,1){0}} \put(149.8,34){\vector(0,1){0}} \allinethickness{2.0pt} \put(145,15){\line(0,1){5}} \put(165,17){\line(0,1){1}} \allinethickness{1.0pt} \put(142,5){\line(1,0){26}} \put(142,45){\line(1,0){26}} \put(142,15){\line(1,0){1}} \put(142,20){\line(1,0){1}} \put(147,15){\line(1,0){21}} \put(147,20){\line(1,0){21}} \put(142,25){\arc{10}{1.571}{3.142}} \put(168,25){\arc{10}{0}{1.571}} \put(142,40){\arc{10}{3.142}{4.712}} \put(168,40){\arc{10}{4.712}{0}} \put(137,25){\line(0,1){15}} \put(173,25){\line(0,1){15}} \put(142,10){\arc{10}{1.571}{4.712}} \put(168,10){\arc{10}{4.712}{1.571}} \put(133,32.5){\makebox(0,0){$a$}} \put(155,2){\makebox(0,0){$c$}} \put(155,48){\makebox(0,0){$c$}} \put(135,46){\makebox(0,0){$b$}} \put(155,12){\makebox(0,0){$b$}} \put(155,24){\makebox(0,0){$b$}} \put(142,24){\makebox(0,0){$\lambda$}} \put(169,24){\makebox(0,0){$\mu$}} \put(155,32.5){\makebox(0,0){$\nu$}} \end{picture} \end{center} \caption{Commutation of $Y$ and $Z$} \label{[Y,Z]1} \end{figure} Using Lemma \ref{idexp} and also the trick to turn around the small arcs given in Fig.\ \ref{trick}, we obtain the right-hand side of Fig.\ \ref{[Y,Z]1}. We can now slide around the lower trivalent vertex of the wire $\nu$ to obtain the left-hand side of Fig.\ \ref{[Y,Z]2}. \begin{figure}[htb] \begin{center} \unitlength 0.7mm \begin{picture}(185,50) \allinethickness{0.3pt} \put(16,23){\makebox(0,0){$\displaystyle \sum_{a,b,c,\lambda} \;\frac{d_a d_b d_c}{w d_\mu d_\nu}$}} \put(50,5){\line(0,1){10}} \put(50,45){\line(0,-1){6}} \put(50,35){\line(0,-1){15}} \put(70,5){\line(0,1){8}} \put(70,45){\line(0,-1){6}} \put(70,22){\line(0,1){13}} \put(50,45){\arc{5}{0}{3.142}} \put(70,45){\arc{5}{0}{3.142}} \put(50,5){\arc{5}{3.142}{0}} \put(70,5){\arc{5}{3.142}{0}} \put(42,37){\line(1,0){36}} \put(42,37){\arc{5}{4.712}{1.571}} \put(78,37){\arc{5}{1.571}{4.712}} \put(50,27){\vector(0,-1){0}} \put(70,31){\vector(0,1){0}} \put(58,37){\vector(-1,0){0}} \allinethickness{2.0pt} \put(50,15){\line(0,1){5}} \put(70,17){\line(0,1){1}} \allinethickness{1.0pt} \put(47,5){\line(1,0){26}} \put(47,45){\line(1,0){26}} \put(47,15){\line(1,0){1}} \put(47,20){\line(1,0){1}} \put(52,15){\line(1,0){21}} \put(52,20){\line(1,0){21}} \put(47,25){\arc{10}{1.571}{3.142}} \put(73,25){\arc{10}{0}{1.571}} \put(47,40){\arc{10}{3.142}{4.712}} \put(73,40){\arc{10}{4.712}{0}} \put(42,25){\line(0,1){15}} \put(78,25){\line(0,1){15}} \put(47,10){\arc{10}{1.571}{4.712}} \put(73,10){\arc{10}{4.712}{1.571}} \put(60,23){\makebox(0,0){$a$}} \put(60,2){\makebox(0,0){$c$}} \put(60,48){\makebox(0,0){$c$}} \put(40,46){\makebox(0,0){$b$}} \put(80,46){\makebox(0,0){$b$}} \put(60,12){\makebox(0,0){$b$}} \put(46,25){\makebox(0,0){$\lambda$}} \put(74,24){\makebox(0,0){$\mu$}} \put(60,32.5){\makebox(0,0){$\nu$}} \allinethickness{0.3pt} \put(110,23){\makebox(0,0){$= \;\displaystyle \sum_{a,b,c,\rho} \;\frac{d_a d_b d_c}{w d_\mu d_\nu}$}} \put(172,5){\line(0,1){6}} \put(172,15){\line(0,1){20}} \put(172,45){\line(0,-1){6}} \put(172,45){\arc{5}{0}{3.142}} \put(172,5){\arc{5}{3.142}{0}} \put(140,13){\line(1,0){8}} \put(180,13){\line(-1,0){23}} \put(140,13){\arc{5}{4.712}{1.571}} \put(180,13){\arc{5}{1.571}{4.712}} \put(140,37){\line(1,0){10}} \put(155,37){\line(1,0){25}} \put(140,37){\arc{5}{4.712}{1.571}} \put(180,37){\arc{5}{1.571}{4.712}} \put(172,27){\vector(0,1){0}} \put(166,13){\vector(1,0){0}} \put(162,37){\vector(-1,0){0}} \allinethickness{2.0pt} \put(150,37){\line(1,0){5}} \put(152,13){\line(1,0){1}} \allinethickness{1.0pt} \put(160,5){\line(1,0){15}} \put(160,45){\line(1,0){15}} \put(145,40){\arc{10}{3.142}{0}} \put(145,10){\arc{10}{0}{3.142}} \put(175,40){\arc{10}{4.712}{0}} \put(175,10){\arc{10}{0}{1.571}} \put(160,40){\arc{10}{3.142}{4.712}} \put(160,10){\arc{10}{1.571}{3.142}} \put(140,10){\line(0,1){30}} \put(180,10){\line(0,1){30}} \put(150,10){\line(0,1){25}} \put(150,39){\line(0,1){1}} \put(155,10){\line(0,1){25}} \put(155,39){\line(0,1){1}} \put(158,24){\makebox(0,0){$c$}} \put(137,25){\makebox(0,0){$a$}} \put(184,25){\makebox(0,0){$a$}} \put(182,46){\makebox(0,0){$b$}} \put(182,4){\makebox(0,0){$b$}} \put(164,17){\makebox(0,0){$\rho$}} \put(147,25){\makebox(0,0){$b$}} \put(176,24){\makebox(0,0){$\mu$}} \put(164,33){\makebox(0,0){$\nu$}} \end{picture} \end{center} \caption{Commutation of $Y$ and $Z$} \label{[Y,Z]2} \end{figure} Next, we can use Lemma \ref{idexp} to replace the two parallel horizontal wires with labels $a$ and $b$ by a summation over a thin wire $\rho$. Similarly, but the other way round, we can then use Lemma \ref{idexp} to replace the summation over the wire with label $\lambda$ by two straight horizontal wires with labels $b$ and $c$. This way we obtain the right-hand side of Fig.\ \ref{[Y,Z]2}. Now it should be clear how to proceed: We slide around the upper trivalent vertex of the wire $\mu$ counter-clockwise. Then we see that the result gives us the diagram for $\sum_\rho Z_{\nu,\rho} Y_{\rho,\mu}$, rotated by 90 degrees. This proves $YZ=ZY$. Next we show commutativity of $Z$ with $T$. We have to show $\omega_\lambda Z_{\lambda,\mu}=Z_{\lambda,\mu}\omega_\mu$. Using the graphical expression for the statistics phase $\omega_\lambda$ on the left-hand side of Fig.\ \ref{statph}, we can represent $\omega_\lambda Z_{\lambda,\mu}$ by the left-hand side of Fig.\ \ref{[T,Z]}. \begin{figure}[htb] \begin{center} \unitlength 0.7mm \begin{picture}(169,50) \allinethickness{0.3pt} \put(13.5,23){\makebox(0,0){$\displaystyle \sum_{b,c} \;\frac{d_b d_c}{w d_\lambda d_\mu}$}} \put(45,5){\line(0,1){10}} \put(45,45){\line(0,-1){4.5}} \put(45,37.5){\line(0,-1){17.5}} \put(55,5){\line(0,1){8}} \put(55,45){\line(0,-1){23}} \put(45,45){\arc{5}{0}{3.142}} \put(55,45){\arc{5}{0}{3.142}} \put(45,5){\arc{5}{3.142}{0}} \put(55,5){\arc{5}{3.142}{0}} \put(41,37.5){\arc{8}{0.9}{6.283}} \put(45,25){\vector(0,-1){0}} \put(55,29){\vector(0,1){0}} \allinethickness{2.0pt} \put(45,15){\line(0,1){5}} \put(55,17){\line(0,1){1}} \allinethickness{1.0pt} \put(37,5){\line(1,0){26}} \put(37,45){\line(1,0){26}} \put(37,15){\line(1,0){6}} \put(37,20){\line(1,0){6}} \put(47,15){\line(1,0){16}} \put(47,20){\line(1,0){16}} \put(37,25){\arc{10}{1.571}{3.142}} \put(63,25){\arc{10}{0}{1.571}} \put(37,40){\arc{10}{3.142}{4.712}} \put(63,40){\arc{10}{4.712}{0}} \put(32,25){\line(0,1){15}} \put(68,25){\line(0,1){15}} \put(37,10){\arc{10}{1.571}{4.712}} \put(63,10){\arc{10}{4.712}{1.571}} \put(50,2){\makebox(0,0){$c$}} \put(50,48){\makebox(0,0){$c$}} \put(50,12){\makebox(0,0){$b$}} \put(50,24){\makebox(0,0){$b$}} \put(40,28){\makebox(0,0){$\lambda$}} \put(60,27){\makebox(0,0){$\mu$}} \allinethickness{0.3pt} \put(105,23){\makebox(0,0){$= \;\displaystyle \sum_{b,c} \;\frac{d_b d_c}{w d_\lambda d_\mu}$}} \put(140,5){\line(0,1){10}} \put(136,34.5){\line(1,0){2}} \put(142,34.5){\line(1,0){21}} \put(140,20){\line(0,1){18.5}} \put(147,5){\line(0,1){8}} \put(147,22){\line(0,1){4.5}} \put(155,20){\line(0,1){6.5}} \put(163,34.5){\arc{5}{1.571}{4.712}} \put(155,20){\arc{5}{3.142}{0}} \put(140,5){\arc{5}{3.142}{0}} \put(147,5){\arc{5}{3.142}{0}} \put(136,38.5){\arc{8}{1.571}{6.283}} \put(151,26.5){\arc{8}{3.142}{0}} \put(140,25){\vector(0,-1){0}} \put(147,26){\vector(0,1){0}} \allinethickness{2.0pt} \put(140,15){\line(0,1){5}} \put(147,17){\line(0,1){1}} \allinethickness{1.0pt} \put(132,5){\line(1,0){26}} \put(132,45){\line(1,0){26}} \put(132,15){\line(1,0){6}} \put(132,20){\line(1,0){6}} \put(142,15){\line(1,0){16}} \put(142,20){\line(1,0){16}} \put(132,25){\arc{10}{1.571}{3.142}} \put(158,25){\arc{10}{0}{1.571}} \put(132,40){\arc{10}{3.142}{4.712}} \put(158,40){\arc{10}{4.712}{0}} \put(127,25){\line(0,1){15}} \put(163,25){\line(0,1){15}} \put(132,10){\arc{10}{1.571}{4.712}} \put(158,10){\arc{10}{4.712}{1.571}} \put(143.5,2){\makebox(0,0){$c$}} \put(145,48.5){\makebox(0,0){$b$}} \put(152,12){\makebox(0,0){$b$}} \put(167,27){\makebox(0,0){$c$}} \put(135,28){\makebox(0,0){$\lambda$}} \put(151,24){\makebox(0,0){$\mu$}} \end{picture} \end{center} \caption{Commutation of $T$ and $Z$} \label{[T,Z]} \end{figure} We now start to rotate the upper oval consisting of the thick wires $b$ and $c$ in a clockwise direction. This way we obtain the right-hand side of Fig.\ \ref{[T,Z]}. It should now be clear that, if we continue rotating to a full rotation by 360 degrees, then we remove the twist from the wire $\lambda$ whereas we obtain a twist in the wire $\mu$ which is of the type displayed on the right-hand side of Fig.\ \ref{statph}, thus representing $\omega_\mu$. Hence $TZ=ZT$. \end{proof} The following is now immediate by Thm.\ \ref{ST}, which states that in the non-degenerate case matrices $S=w^{-1/2}Y$ and $T$ provide a unitary representation of the modular group ${\it{SL}}(2;\bbZ)$. \begin{corollary} \label{modular6} If the braiding on ${}_N {\cal X}_N$ is non-degenerate, then the matrix $Z$ defined in Definition \ref{modular1} is a modular invariant mass matrix. \end{corollary} In conformal field theory the ${\it{SL}}(2;\bbZ)$ action arises from a ``reparametrization of the torus'', and in the parameter space $S$ corresponds to a 90 degree rotation and $T$ to twisting the torus. Note that this action is nicely reflected in the proof of Thm.\ \ref{modular2}. \subsection{Generating property of $\alpha$-induction} \label{sec-genaind} We now show that both kinds of $\alpha$-induction generate the whole $M$-$M$ fusion rule algebra (or the sector algebra in our terminology of \cite{BE1,BE2,BE3}) in the case that the $N$-$N$ system is non-degenerately braided. That is, from now on we work with the following \begin{assumption} \label{set-nondeg}{\rm In addition to Assumption \ref{set-braid}, we now assume that the braiding on ${}_N {\cal X}_N$ is non-degenerate in the sense of Definition \ref{non-deg}. }\end{assumption} With Assumption \ref{set-nondeg} we can now use the ``killing ring'', the orthogonality relation of Fig.\ \ref{ort0}, and this turns out to be a powerful tool in the graphical framework. The following theorem states in particular that any minimal central projection $e_\beta$ of $(\dta,*_h)$ appears in the linear decomposition of some $p_\lambda^+ *_v p_\mu^-$. Such a generating property of $p_j^\pm$'s has also been noticed by Ocneanu in the setting of the lectures \cite{O7}. We can apply his idea of the proof (which is not included in the notes \cite{O7}) to our situation without essential change. \begin{theorem} \label{generating} Under Assumption \ref{set-nondeg}, we have $\sum_{\lambda,\mu\in{}_N {\cal X}_N} p_\lambda^+ *_v p_\mu^- = w{\bf1}_h$ in $\dta$, and consequently \begin{equation} \sum_{\lambda,\mu\in{}_N {\cal X}_N} d_\lambda d_\mu [\alpha^+_\lambda] [\alpha^-_\mu] = w \sum_{\beta\in{}_M {\cal X}_M} d_{\beta} [\beta] \end{equation} in the $M$-$M$ fusion rule algebra. In particular, for any $\beta\in{}_M {\cal X}_M$ the sector $[\beta]$ is a subsector of $[\alpha^+_\lambda][\alpha^-_\mu]$ for some $\lambda,\mu\in{}_N {\cal X}_N$. \end{theorem} \begin{proof} The sum $\sum_{\lambda,\mu} p^+_\lambda *_v p^-_\mu$ is given graphically by the left-hand side of Fig.\ \ref{gen1}. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(191,50) \allinethickness{1.0pt} \put(13,22){\makebox(0,0){$\displaystyle\sum_{a,b,c,\lambda,\mu}d_b$}} \put(40,0){\line(0,1){8}} \put(40,12){\line(0,1){38}} \put(60,0){\line(0,1){8}} \put(60,12){\line(0,1){38}} \allinethickness{2.0pt} \put(40,10){\line(1,0){20}} \put(52,10){\vector(1,0){0}} \put(42,30){\line(1,0){16}} \put(52,30){\vector(1,0){0}} \allinethickness{0.3pt} \put(40,15){\arc{10}{1.571}{4.712}} \put(60,15){\arc{10}{4.712}{1.571}} \put(40,20){\arc{5}{1.571}{4.712}} \put(60,20){\arc{5}{4.712}{1.571}} \put(40,35){\arc{10}{2.071}{4.712}} \put(60,35){\arc{10}{4.712}{1.071}} \put(40,40){\arc{5}{1.571}{4.712}} \put(60,40){\arc{5}{4.712}{1.571}} \put(35,47){\makebox(0,0){$a$}} \put(65,47){\makebox(0,0){$a$}} \put(35,25){\makebox(0,0){$b$}} \put(65,25){\makebox(0,0){$b$}} \put(35,3){\makebox(0,0){$c$}} \put(65,3){\makebox(0,0){$c$}} \put(50,16){\makebox(0,0){$\alpha_\lambda^+$}} \put(50,36){\makebox(0,0){$\alpha_\mu^-$}} \put(95,22){\makebox(0,0){$=\;\displaystyle\sum_{a,b,c,\lambda,\mu,\nu}d_b$}} \allinethickness{1.0pt} \put(130,0){\arc{10}{3.142}{4.712}} \put(130,50){\arc{10}{1.571}{3.142}} \put(180,0){\arc{10}{4.712}{0}} \put(180,50){\arc{10}{0}{1.571}} \put(130,5){\line(1,0){50}} \put(130,45){\line(1,0){50}} \allinethickness{0.3pt} \put(155,13){\vector(0,-1){8}} \put(155,45){\line(0,-1){28}} \put(155,5){\arc{5}{3.142}{0}} \put(155,45){\arc{5}{0}{3.142}} \put(135,45){\line(0,-1){25}} \put(175,45){\line(0,-1){25}} \put(140,15){\line(1,0){30}} \put(140,20){\arc{10}{1.571}{3.142}} \put(170,20){\arc{10}{0}{1.571}} \put(135,45){\arc{5}{0}{3.142}} \put(175,45){\arc{5}{0}{3.142}} \put(150,15){\vector(1,0){0}} \put(145,45){\line(0,-1){10}} \put(165,35){\vector(0,1){10}} \put(150,30){\line(1,0){3}} \put(157,30){\line(1,0){3}} \put(150,35){\arc{10}{1.571}{3.142}} \put(160,35){\arc{10}{0}{1.571}} \put(145,45){\arc{5}{0}{3.142}} \put(165,45){\arc{5}{0}{3.142}} \put(121,3){\makebox(0,0){$c$}} \put(189,3){\makebox(0,0){$c$}} \put(121,47){\makebox(0,0){$a$}} \put(189,47){\makebox(0,0){$a$}} \put(150,48){\makebox(0,0){$c$}} \put(160,48){\makebox(0,0){$c$}} \put(140,41){\makebox(0,0){$b$}} \put(170,41){\makebox(0,0){$b$}} \put(148,20){\makebox(0,0){$\lambda$}} \put(163,26){\makebox(0,0){$\mu$}} \put(159,10){\makebox(0,0){$\nu$}} \end{picture} \end{center} \caption{The sum $\sum_{\lambda,\mu} p^+_\lambda *_v p^-_\mu$} \label{gen1} \end{figure} By using Lemma \ref{idexp} for the two parallel vertical wires $c$ on the bottom and the IBFE moves we obtain the right-hand side of Fig.\ \ref{gen1}. For the summation over the thin wire $\lambda$ we can use Lemma \ref{idexp} again to obtain the left-hand side of Fig.\ \ref{gen2}. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(216,50) \allinethickness{1.0pt} \put(13,22){\makebox(0,0){$\displaystyle\sum_{a,b,c,\mu,\nu}d_b$}} \put(35,50){\line(0,-1){30}} \put(95,50){\line(0,-1){30}} \put(40,40){\line(0,-1){15}} \put(90,40){\line(0,-1){15}} \put(45,20){\line(1,0){40}} \put(45,25){\arc{10}{1.571}{3.142}} \put(85,25){\arc{10}{0}{1.571}} \put(40,15){\line(1,0){50}} \put(45,40){\arc{10}{3.142}{4.712}} \put(85,40){\arc{10}{4.712}{0}} \put(40,0){\arc{10}{3.142}{4.712}} \put(40,20){\arc{10}{1.571}{3.142}} \put(90,0){\arc{10}{4.712}{0}} \put(90,20){\arc{10}{0}{1.571}} \put(40,5){\line(1,0){50}} \put(45,45){\line(1,0){40}} \allinethickness{2.0pt} \put(65,17){\line(0,1){1}} \allinethickness{0.3pt} \put(65,13){\vector(0,-1){8}} \put(65,45){\line(0,-1){23}} \put(65,5){\arc{5}{3.142}{0}} \put(65,45){\arc{5}{0}{3.142}} \put(55,45){\line(0,-1){10}} \put(75,35){\vector(0,1){10}} \put(60,30){\line(1,0){3}} \put(67,30){\line(1,0){3}} \put(60,35){\arc{10}{1.571}{3.142}} \put(70,35){\arc{10}{0}{1.571}} \put(55,45){\arc{5}{0}{3.142}} \put(75,45){\arc{5}{0}{3.142}} \put(31,3){\makebox(0,0){$c$}} \put(99,3){\makebox(0,0){$c$}} \put(31,47){\makebox(0,0){$a$}} \put(99,47){\makebox(0,0){$a$}} \put(60,48){\makebox(0,0){$c$}} \put(70,48){\makebox(0,0){$c$}} \put(50,25){\makebox(0,0){$b$}} \put(73,26){\makebox(0,0){$\mu$}} \put(69,10){\makebox(0,0){$\nu$}} \allinethickness{1.0pt} \put(123,22){\makebox(0,0){$=\displaystyle\sum_{a,b,c,\mu,\nu}d_b$}} \put(150,50){\line(0,-1){30}} \put(210,50){\line(0,-1){30}} \put(155,40){\line(0,-1){15}} \put(205,40){\line(0,-1){15}} \put(160,20){\line(1,0){40}} \put(160,25){\arc{10}{1.571}{3.142}} \put(200,25){\arc{10}{0}{1.571}} \put(155,15){\line(1,0){50}} \put(160,40){\arc{10}{3.142}{4.712}} \put(200,40){\arc{10}{4.712}{0}} \put(155,0){\arc{10}{3.142}{4.712}} \put(155,20){\arc{10}{1.571}{3.142}} \put(205,0){\arc{10}{4.712}{0}} \put(205,20){\arc{10}{0}{1.571}} \put(155,5){\line(1,0){50}} \put(160,45){\line(1,0){40}} \allinethickness{2.0pt} \put(180,17){\line(0,1){1}} \allinethickness{0.3pt} \put(180,13){\vector(0,-1){8}} \put(180,45){\line(0,-1){15}} \put(180,22){\line(0,1){4}} \put(180,5){\arc{5}{3.142}{0}} \put(180,45){\arc{5}{0}{3.142}} \put(170,45){\line(0,-1){4}} \put(163,33){\vector(0,1){12}} \put(175,36){\line(1,0){3}} \put(182,36){\line(1,0){3}} \put(175,41){\arc{10}{1.571}{3.142}} \put(185,32){\arc{8}{4.712}{1.571}} \put(168,28){\line(1,0){17}} \put(168,33){\arc{10}{1.571}{3.142}} \put(170,45){\arc{5}{0}{3.142}} \put(163,45){\arc{5}{0}{3.142}} \put(146,3){\makebox(0,0){$c$}} \put(214,3){\makebox(0,0){$c$}} \put(146,47){\makebox(0,0){$a$}} \put(214,47){\makebox(0,0){$a$}} \put(175,48){\makebox(0,0){$c$}} \put(185,48){\makebox(0,0){$c$}} \put(166.5,40){\makebox(0,0){$b$}} \put(194,31){\makebox(0,0){$\mu$}} \put(184,10){\makebox(0,0){$\nu$}} \end{picture} \end{center} \caption{The sum $\sum_{\lambda,\mu} p^+_\lambda *_v p^-_\mu$} \label{gen2} \end{figure} Now we can slide around the right trivalent vertex of the wire $\mu$, and this yields the right-hand side of Fig.\ \ref{gen2}. Next we can use the trick of Fig.\ \ref{trick} to turn around the small arcs from the wire $\mu$ to the wire $b$. This yields a factor $d_\mu/d_b$. Then we can proceed with the summation over $b$, using Lemma \ref{idexp} once more, and this gives us the left-hand side of Fig.\ \ref{gen3}. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(207,50) \allinethickness{1.0pt} \put(13,22){\makebox(0,0){$\displaystyle\sum_{a,c,\mu,\nu}d_\mu$}} \put(35,50){\line(0,-1){30}} \put(95,50){\line(0,-1){30}} \put(40,40){\line(0,-1){15}} \put(90,40){\line(0,-1){15}} \put(45,20){\line(1,0){40}} \put(45,25){\arc{10}{1.571}{3.142}} \put(85,25){\arc{10}{0}{1.571}} \put(40,15){\line(1,0){50}} \put(45,40){\arc{10}{3.142}{4.712}} \put(85,40){\arc{10}{4.712}{0}} \put(40,0){\arc{10}{3.142}{4.712}} \put(40,20){\arc{10}{1.571}{3.142}} \put(90,0){\arc{10}{4.712}{0}} \put(90,20){\arc{10}{0}{1.571}} \put(40,5){\line(1,0){50}} \put(45,45){\line(1,0){40}} \allinethickness{2.0pt} \put(65,17){\line(0,1){1}} \allinethickness{0.3pt} \put(65,13){\vector(0,-1){8}} \put(65,45){\line(0,-1){15.5}} \put(65,22){\line(0,1){3.5}} \put(65,5){\arc{5}{3.142}{0}} \put(65,45){\arc{5}{0}{3.142}} \put(60,27.5){\line(1,0){10}} \put(66,27.5){\vector(-1,0){0}} \put(60,37.5){\line(1,0){3}} \put(67,37.5){\line(1,0){3}} \put(60,32.5){\arc{10}{1.571}{4.7122}} \put(70,32.5){\arc{10}{4.712}{1.571}} \put(31,3){\makebox(0,0){$c$}} \put(99,3){\makebox(0,0){$c$}} \put(31,47){\makebox(0,0){$a$}} \put(99,47){\makebox(0,0){$a$}} \put(50,24){\makebox(0,0){$c$}} \put(80,31.5){\makebox(0,0){$\mu$}} \put(69,10){\makebox(0,0){$\nu$}} \allinethickness{1.0pt} \put(125,22){\makebox(0,0){$=\;\displaystyle\sum_{a,c}\;\frac w{d_c}$}} \put(150,50){\line(0,-1){30}} \put(200,50){\line(0,-1){30}} \put(155,40){\line(0,-1){15}} \put(195,40){\line(0,-1){15}} \put(160,20){\line(1,0){30}} \put(160,25){\arc{10}{1.571}{3.142}} \put(190,25){\arc{10}{0}{1.571}} \put(155,15){\line(1,0){40}} \put(160,40){\arc{10}{3.142}{4.712}} \put(190,40){\arc{10}{4.712}{0}} \put(155,0){\arc{10}{3.142}{4.712}} \put(155,20){\arc{10}{1.571}{3.142}} \put(195,0){\arc{10}{4.712}{0}} \put(195,20){\arc{10}{0}{1.571}} \put(155,5){\line(1,0){40}} \put(160,45){\line(1,0){30}} \put(146,3){\makebox(0,0){$c$}} \put(204,3){\makebox(0,0){$c$}} \put(146,47){\makebox(0,0){$a$}} \put(204,47){\makebox(0,0){$a$}} \put(175,25){\makebox(0,0){$c$}} \end{picture} \end{center} \caption{The sum $\sum_{\lambda,\mu} p^+_\lambda *_v p^-_\mu$} \label{gen3} \end{figure} Now we observe that the summation over $\mu$ provides a killing ring, and hence we obtain a factor $w\del \nu 0$. The normalization convention for the small arcs yields another factor $1/d_c$, and hence we get exactly the right-hand side of Fig.\ \ref{gen3}. The circular wire $c$ cancels the factor $1/d_c$, and thus we are left exactly with the global index $w$ times a summation over two straight horizontal wires, and the latter is exactly the horizontal unit ${\bf1}_h=\sum_\beta e_\beta$. The rest is application of the isomorphism $\Phi$. \end{proof} We remark that the non-degeneracy of the braiding played an essential role in the proof. In fact there are counter-examples showing that the generating property does not hold in general if the braiding is degenerate (e.g.\ the finite group case discussed in Section 4.2 of \cite{BE1} serves as such an example). \section{Representations of the $M$-$M$ Fusion Rule Algebra} \label{sec-repMM} \subsection{Irreducible representations of the $M$-$M$ fusion rules} \label{sec-dfa} We next study in detail the algebra $({\cal{Z}}_h,*_v)$ or, equivalently, the $M$-$M$ fusion rule algebra in the case that the $N$-$N$ system is non-degenerately braided. Note that the Assumption \ref{set-braid} implies in particular that the $N$-$N$ fusion rules algebra is Abelian. However, the $M$-$M$ fusion rules are in general non-commutative, and therefore so is the center $({\cal{Z}}_h,*_v)$. We are now going to decompose $({\cal{Z}}_h,*_v)$ in simple matrix algebras. Note that such a decomposition of $({\cal{Z}}_h,*_v)$ is equivalent to the determination of the irreducible representations of the $M$-$M$ fusion rule algebra. We need some preparation. As in the graphical setting for the double triangle algebra, we can consider the diagram in Fig.\ \ref{Ombcts} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(70,40) \allinethickness{1.0pt} \put(17,0){\line(0,1){15}} \put(63,0){\line(0,1){15}} \put(27,15){\arc{20}{3.142}{4.712}} \put(53,15){\arc{20}{4.712}{0}} \put(27,15){\arc{10}{1.571}{4.712}} \put(53,15){\arc{10}{4.712}{1.571}} \put(27,10){\line(1,0){26}} \put(27,20){\line(1,0){1}} \put(32,20){\line(1,0){21}} \put(27,25){\line(1,0){1}} \put(32,25){\line(1,0){21}} \allinethickness{0.3pt} \put(30,10){\line(0,1){10}} \put(30,25){\line(0,1){15}} \put(50,10){\line(0,1){8}} \put(50,27){\line(0,1){13}} \put(30,30.5){\vector(0,-1){0}} \put(50,34.5){\vector(0,1){0}} \allinethickness{2.0pt} \put(50,22){\line(0,1){1}} \put(30,20){\line(0,1){5}} \put(67,4){\makebox(0,0){$a$}} \put(40,7){\makebox(0,0){$c$}} \put(40,16){\makebox(0,0){$b$}} \put(30,5.8){\makebox(0,0){$t$}} \put(50,5){\makebox(0,0){$s$}} \put(25,33.5){\makebox(0,0){$\lambda$}} \put(55,32.5){\makebox(0,0){$\mu$}} \put(3,20){\makebox(0,0){$\displaystyle\sum_a$}} \end{picture} \end{center} \caption{The vector $\Omega_{b,c,t,s}^{\lambda,\mu}\in{\cal{H}}_{\lambda,\mu}$} \label{Ombcts} \end{figure} as a vector $\Omega_{b,c,t,s}^{\lambda,\mu}\in{\cal{H}}_{\lambda,\mu}$, where ${\cal{H}}_{\lambda,\mu}$ is the vector space ${\cal{H}}_{\lambda,\mu} = \bigoplus_{a\in{}_N {\cal X}_M} {\rm{Hom}}(\lambda\co\mu,a\co a)$, $\lambda,\mu\in{}_N {\cal X}_N$. Here $b,c\in{}_N {\cal X}_M$, and $t\in{\rm{Hom}}(\lambda,b\co c)$ and $s\in{\rm{Hom}}(\co\mu,c\co b)$ are isometries labelling the two trivalent vertices in Fig.\ \ref{Ombcts}. It is important to notice that we do not allow coefficients depending on $a$: The same isometries $t,s$ are used in each block ${\rm{Hom}}(\lambda\co\mu,a\co a)$ of ${\cal{H}}_{\lambda,\mu}$. We next define the subspace $H_{\lambda,\mu}\subset{\cal{H}}_{\lambda,\mu}$ spanned by such vectors: \[ H_{\lambda,\mu} = {\rm span} \{ \Omega_{b,c,t,s}^{\lambda,\mu} \,\, | \,\, b,c\in{}_N {\cal X}_M\,,\,\, t\in{\rm{Hom}}(\lambda,b\co c)\,,\,\,s\in{\rm{Hom}}(\co\mu,c\co b) \} \,. \] Take two such vectors $\Omega_{b,c,t,s}^{\lambda,\mu}$ and $\Omega_{b',c',t',s'}^{\lambda,\mu}$. We define an element $|\Omega_{b',c',t',s'}^{\lambda,\mu}\rangle \langle\Omega_{b,c,t,s}^{\lambda,\mu}|\in\dta$ by the diagram in Fig.\ \ref{OmOmdta}. \begin{figure}[htb] \begin{center} \unitlength 0.7mm \begin{picture}(70,50) \allinethickness{1.0pt} \put(17,0){\line(0,1){10}} \put(63,0){\line(0,1){10}} \put(27,10){\arc{20}{3.142}{4.712}} \put(53,10){\arc{20}{4.712}{0}} \put(27,10){\arc{10}{1.571}{4.712}} \put(53,10){\arc{10}{4.712}{1.571}} \put(27,40){\arc{20}{1.571}{3.142}} \put(53,40){\arc{20}{0}{1.571}} \put(27,40){\arc{10}{1.571}{4.712}} \put(53,40){\arc{10}{4.712}{1.571}} \put(27,5){\line(1,0){26}} \put(27,15){\line(1,0){1}} \put(32,15){\line(1,0){21}} \put(27,20){\line(1,0){1}} \put(32,20){\line(1,0){21}} \put(17,50){\line(0,-1){10}} \put(63,50){\line(0,-1){10}} \put(27,30){\line(1,0){1}} \put(32,30){\line(1,0){21}} \put(27,35){\line(1,0){1}} \put(32,35){\line(1,0){21}} \put(27,45){\line(1,0){26}} \allinethickness{0.3pt} \put(30,5){\line(0,1){10}} \put(30,20){\line(0,1){10}} \put(50,5){\line(0,1){8}} \put(50,22){\line(0,1){6}} \put(30,23){\vector(0,-1){0}} \put(50,27){\vector(0,1){0}} \put(30,45){\line(0,-1){10}} \put(50,45){\line(0,-1){8}} \allinethickness{2.0pt} \put(50,17){\line(0,1){1}} \put(30,15){\line(0,1){5}} \put(50,33){\line(0,-1){1}} \put(30,35){\line(0,-1){5}} \put(67,4){\makebox(0,0){$a'$}} \put(40,2){\makebox(0,0){$c'$}} \put(40,11){\makebox(0,0){$b'$}} \put(30,1.8){\makebox(0,0){$t'$}} \put(50,1.8){\makebox(0,0){$s'$}} \put(67,46){\makebox(0,0){$a$}} \put(40,48){\makebox(0,0){$c$}} \put(40,39){\makebox(0,0){$b$}} \put(30,48.2){\makebox(0,0){$t^*$}} \put(50,48.2){\makebox(0,0){$s^*$}} \put(25,25){\makebox(0,0){$\lambda$}} \put(55,24){\makebox(0,0){$\mu$}} \put(3,25){\makebox(0,0){$\displaystyle\sum_{a,a'}$}} \end{picture} \end{center} \caption{The element $|\Omega_{b',c',t',s'}^{\lambda,\mu}\rangle \langle\Omega_{b,c,t,s}^{\lambda,\mu}|\in\protect\dta$} \label{OmOmdta} \end{figure} (This notation will be justified by Lemma \ref{innerprod} below.) We now choose orthonormal bases of isometries $t_{b,\co c}^{\lambda;i}\in{\rm{Hom}}(\lambda,b\co c)$, $i=1,2,...,N_{b,\co c}^\lambda$, for each $\lambda,b,c$ and put $\Omega_\xi^{\lambda,\mu} =\Omega_{b,c,{t_{b,\co c}^{\lambda;i}},{t_{c,\co b}^{\co\mu;j}}}^{\lambda,\mu}$ with some multi-index $\xi=(b,c,i,j)$. Varying $\xi$, we obtain a generating set of $H_{\lambda,\mu}$ which will, however, in general not be a basis as the vectors $\Omega_\xi^{\lambda,\mu}$ may be linearly dependent in $H_{\lambda,\mu}$. Let $\Phi_j^{\lambda,\mu}\in H_{\lambda,\mu}$, $j=1,2$, any two vectors. We can expand them as $\Phi_j^{\lambda,\mu}=\sum_\xi c_j^\xi \Omega^{\lambda,\mu}_\xi$ with $c_j^\xi\in\Bbb{C}$, but note that this expansion is not unique. We now define an element $|\Phi_1^{\lambda,\mu}\rangle\langle\Phi_2^{\lambda,\mu}|\in\dta$ by \begin{equation} |\Phi_1^{\lambda,\mu}\rangle\langle\Phi_2^{\lambda,\mu}| = \sum_{\xi,\xi'} c_1^\xi (c_2^{\xi'})^* |\Omega_\xi^{\lambda,\mu}\rangle\langle\Omega_{\xi'}^{\lambda,\mu}| \,, \label{|P><P|} \end{equation} and a scalar $\langle\Phi_2^{\lambda,\mu},\Phi_1^{\lambda,\mu}\rangle\in\Bbb{C}$, \begin{equation} \langle\Phi_2^{\lambda,\mu},\Phi_1^{\lambda,\mu}\rangle = \frac 1{d_\lambda d_\mu} \, \tau_v (|\Phi_1^{\lambda,\mu}\rangle\langle\Phi_2^{\lambda,\mu}|) \,. \label{<P,P>} \end{equation} \begin{lemma} \label{innerprod} \erf{|P><P|} extends to a sesqui-linear map $H_{\lambda,\mu}\times H_{\lambda,\mu}\rightarrow{\cal{Z}}_h$ which is positive definite: If $|\Phi^{\lambda,\mu}\rangle\langle\Phi^{\lambda,\mu}|=0$ for some $\Phi^{\lambda,\mu}\in H_{\lambda,\mu}$ then $\Phi^{\lambda,\mu}=0$. Consequently, \erf{<P,P>} defines a scalar product turning $H_{\lambda,\mu}$ into a Hilbert space. \end{lemma} \begin{proof} As in particular $\Phi_j\in{\cal{H}}_{\lambda,\mu}$, we can write $\Phi_j=\bigoplus_a (\Phi_j)_a$ with $(\Phi_j)_a\in{\rm{Hom}}(\lambda\co\mu,a\co a)$ according to the direct sum structure of ${\cal{H}}_{\lambda,\mu}$, $j=1,2$. Assume $\Phi_1=0$. Then clearly $(\Phi_1)_a=0$ for all $a$. Now the ${\rm{Hom}}(a\co a,a' \co{a'})$ part of $|\Phi_1^{\lambda,\mu}\rangle\langle\Phi_2^{\lambda,\mu}|\in\dta$ is given by $(\Phi_1)_{a'} (\Phi_2)_a^*$, hence $|\Phi_1^{\lambda,\mu}\rangle\langle\Phi_2^{\lambda,\mu}|=0$. A similar argument applies to $\Phi_2$, and hence the element $|\Phi_1^{\lambda,\mu}\rangle\langle\Phi_2^{\lambda,\mu}|\in\dta$ is independent of the linear expansions of the $\Phi_j$'s. Therefore \erf{|P><P|} defines a sesqui-linear map $H_{\lambda,\mu}\times H_{\lambda,\mu}\rightarrow\dta$. Now assume $|\Phi_1^{\lambda,\mu}\rangle\langle\Phi_1^{\lambda,\mu}|=0$. Then in particular $(\Phi_1)_a (\Phi_1)_a^*=0$ for all $a\in{}_N {\cal X}_M$, and hence $\Phi_1=0$, proving strict positivity. That the sesqui-linear form $\langle\cdot,\cdot\rangle$ on $H_{\lambda,\mu}$ is non-degenerate follows now from positive definiteness of $\tau_v$. It remains to show that $|\Phi_1^{\lambda,\mu}\rangle\langle\Phi_2^{\lambda,\mu}|\in{\cal{Z}}_h$. But this is clear since any element of the form in Fig.\ \ref{dta1} can be ``pulled through'' the diagram in Fig.\ \ref{OmOmdta} by using the IBFE's. \end{proof} \begin{lemma} \label{keyrel} We have the identity in Fig.\ \ref{keyrelpict} for intertwiners in ${\rm{Hom}}(\lambda'\co{\mu'},\lambda\co\mu)$, $\lambda,\mu,\lambda',\mu'\in{}_N {\cal X}_N$. \end{lemma} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(195,65) \allinethickness{1.0pt} \put(10,30){\makebox(0,0){$\displaystyle\sum_a\; d_a$}} \put(27,20){\line(0,1){25}} \put(73,20){\line(0,1){25}} \put(37,45){\arc{20}{3.142}{4.712}} \put(63,45){\arc{20}{4.712}{0}} \put(37,45){\arc{10}{1.571}{4.712}} \put(63,45){\arc{10}{4.712}{1.571}} \put(37,20){\arc{20}{1.571}{3.142}} \put(63,20){\arc{20}{0}{1.571}} \put(37,20){\arc{10}{1.571}{4.712}} \put(63,20){\arc{10}{4.712}{1.571}} \put(37,40){\line(1,0){26}} \put(37,50){\line(1,0){1}} \put(42,50){\line(1,0){21}} \put(37,55){\line(1,0){1}} \put(42,55){\line(1,0){21}} \put(37,10){\line(1,0){1}} \put(42,10){\line(1,0){21}} \put(37,15){\line(1,0){1}} \put(42,15){\line(1,0){21}} \put(37,25){\line(1,0){26}} \allinethickness{0.3pt} \put(40,40){\line(0,1){10}} \put(40,55){\line(0,1){10}} \put(60,40){\line(0,1){8}} \put(60,57){\line(0,1){8}} \put(40,0){\line(0,1){10}} \put(60,0){\line(0,1){8}} \put(40,58){\vector(0,-1){0}} \put(60,62){\vector(0,1){0}} \put(40,3){\vector(0,-1){0}} \put(60,7){\vector(0,1){0}} \put(40,25){\line(0,-1){10}} \put(60,25){\line(0,-1){8}} \allinethickness{2.0pt} \put(60,52){\line(0,1){1}} \put(40,50){\line(0,1){5}} \put(60,13){\line(0,-1){1}} \put(40,15){\line(0,-1){5}} \put(50,37){\makebox(0,0){$c'$}} \put(50,46){\makebox(0,0){$b'$}} \put(40,36.8){\makebox(0,0){$t'$}} \put(60,36.8){\makebox(0,0){$s'$}} \put(23,17){\makebox(0,0){$a$}} \put(50,29){\makebox(0,0){$c$}} \put(50,19){\makebox(0,0){$b$}} \put(40,29){\makebox(0,0){$t^*$}} \put(60,29){\makebox(0,0){$s^*$}} \put(35,5){\makebox(0,0){$\lambda$}} \put(65,4){\makebox(0,0){$\mu$}} \put(35,60){\makebox(0,0){$\lambda'$}} \put(65,60){\makebox(0,0){$\mu'$}} \put(125,32.5){\makebox(0,0){$=\;\del \lambda{\lambda'} \, \del \mu{\mu'} \; \langle\Omega_{b,c,t,s}^{\lambda,\mu}, \Omega_{b',c',t',s'}^{\lambda,\mu}\rangle$}} \allinethickness{0.3pt} \put(175,0){\line(0,1){65}} \put(185,0){\line(0,1){65}} \put(175,30.5){\vector(0,-1){0}} \put(185,34.5){\vector(0,1){0}} \put(170,10){\makebox(0,0){$\lambda$}} \put(190,10){\makebox(0,0){$\mu$}} \end{picture} \end{center} \caption{An identity in ${\rm{Hom}}(\lambda'\co{\mu'},\lambda\co\mu)$} \label{keyrelpict} \end{figure} \begin{proof} Using Lemma \ref{idexp} we can replace the left-hand side of Fig.\ \ref{keyrelpict} by the left-hand side of Fig.\ \ref{proofkey1}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(205,65) \allinethickness{1.0pt} \put(10,30){\makebox(0,0){$\displaystyle\sum_{\nu,a}\; d_a$}} \put(27,15){\line(0,1){20}} \put(93,15){\line(0,1){20}} \put(32,15){\arc{10}{1.571}{3.142}} \put(88,15){\arc{10}{0}{1.571}} \put(32,35){\arc{10}{3.142}{4.712}} \put(88,35){\arc{10}{4.712}{0}} \put(47,20){\arc{10}{1.571}{4.712}} \put(73,20){\arc{10}{4.712}{1.571}} \put(32,40){\line(1,0){56}} \put(32,10){\line(1,0){16}} \put(52,10){\line(1,0){36}} \put(47,15){\line(1,0){1}} \put(52,15){\line(1,0){21}} \put(47,25){\line(1,0){26}} \allinethickness{0.3pt} \put(37,40){\line(0,1){5}} \put(83,40){\line(0,1){5}} \put(37,40){\arc{5}{3.142}{0}} \put(83,40){\arc{5}{3.142}{0}} \put(62,55){\vector(1,0){0}} \put(47,45){\arc{20}{3.142}{4.712}} \put(73,45){\arc{20}{4.712}{0}} \put(47,55){\line(1,0){1}} \put(52,55){\line(1,0){21}} \put(50,40){\line(0,1){25}} \put(70,40){\line(0,1){13}} \put(70,57){\line(0,1){8}} \put(50,0){\line(0,1){10}} \put(70,0){\line(0,1){8}} \put(50,58){\vector(0,-1){0}} \put(70,62){\vector(0,1){0}} \put(50,3){\vector(0,-1){0}} \put(70,7){\vector(0,1){0}} \put(50,25){\line(0,-1){10}} \put(70,25){\line(0,-1){8}} \allinethickness{2.0pt} \put(70,13){\line(0,-1){1}} \put(50,15){\line(0,-1){5}} \put(60,37){\makebox(0,0){$c'$}} \put(45,44){\makebox(0,0){$b'$}} \put(75,44){\makebox(0,0){$b'$}} \put(50,36.8){\makebox(0,0){$t'$}} \put(70,36.8){\makebox(0,0){$s'$}} \put(23,17){\makebox(0,0){$a$}} \put(60,29){\makebox(0,0){$c$}} \put(60,19){\makebox(0,0){$b$}} \put(50,29){\makebox(0,0){$t^*$}} \put(70,29){\makebox(0,0){$s^*$}} \put(45,5){\makebox(0,0){$\lambda$}} \put(75,4){\makebox(0,0){$\mu$}} \put(45,60){\makebox(0,0){$\lambda'$}} \put(75,60){\makebox(0,0){$\mu'$}} \put(60,60){\makebox(0,0){$\nu$}} \put(118,30){\makebox(0,0){$=\;\displaystyle\sum_{\nu,\rho,\tau}\; d_\nu$}} \allinethickness{1.0pt} \put(142,20){\line(0,1){25}} \put(198,20){\line(0,1){25}} \put(147,20){\arc{10}{1.571}{3.142}} \put(193,20){\arc{10}{0}{1.571}} \put(147,45){\arc{10}{3.142}{4.712}} \put(193,45){\arc{10}{4.712}{0}} \put(152,32.5){\arc{15}{1.571}{4.712}} \put(188,32.5){\arc{15}{4.712}{1.571}} \put(147,50){\line(1,0){13}} \put(180,50){\line(1,0){13}} \put(147,15){\line(1,0){6}} \put(157,15){\line(1,0){3}} \put(180,15){\line(1,0){13}} \put(152,25){\line(1,0){1}} \put(157,25){\line(1,0){3}} \put(188,25){\line(-1,0){8}} \put(152,40){\line(1,0){8}} \put(188,40){\line(-1,0){8}} \put(160,47.5){\arc{5}{4.712}{0}} \put(160,22.5){\arc{5}{4.712}{0}} \put(160,42.5){\arc{5}{0}{1.571}} \put(160,17.5){\arc{5}{0}{1.571}} \put(180,47.5){\arc{5}{3.142}{4.712}} \put(180,22.5){\arc{5}{3.142}{4.712}} \put(180,42.5){\arc{5}{1.571}{3.142}} \put(180,17.5){\arc{5}{1.571}{3.142}} \put(162.5,17.5){\line(0,1){5}} \put(162.5,42.5){\line(0,1){5}} \put(177.5,17.5){\line(0,1){5}} \put(177.5,42.5){\line(0,1){5}} \allinethickness{0.3pt} \put(147,20){\arc{20}{1.571}{3.142}} \put(193,20){\arc{20}{0}{1.571}} \put(147,45){\arc{20}{3.142}{4.712}} \put(193,45){\arc{20}{4.712}{0}} \put(162.5,45){\line(1,0){15}} \put(162.5,20){\line(1,0){15}} \put(172,55){\vector(1,0){0}} \put(162.5,45){\arc{5}{4.712}{1.571}} \put(162.5,20){\arc{5}{4.712}{1.571}} \put(177.5,45){\arc{5}{1.571}{4.712}} \put(177.5,20){\arc{5}{1.571}{4.712}} \put(172,45){\vector(1,0){0}} \put(172,20){\vector(1,0){0}} \put(147,55){\line(1,0){6}} \put(157,55){\line(1,0){36}} \put(147,10){\line(1,0){6}} \put(157,10){\line(1,0){36}} \put(137,20){\line(0,1){25}} \put(203,20){\line(0,1){25}} \put(155,50){\line(0,1){15}} \put(185,50){\line(0,1){3}} \put(185,57){\line(0,1){8}} \put(155,0){\line(0,1){15}} \put(185,0){\line(0,1){8}} \put(185,12){\line(0,1){1}} \put(155,58){\vector(0,-1){0}} \put(185,62){\vector(0,1){0}} \put(155,3){\vector(0,-1){0}} \put(185,7){\vector(0,1){0}} \put(155,40){\line(0,-1){15}} \put(185,40){\line(0,-1){13}} \allinethickness{2.0pt} \put(185,23){\line(0,-1){6}} \put(155,25){\line(0,-1){10}} \put(164,52){\makebox(0,0){$c'$}} \put(176,52){\makebox(0,0){$c'$}} \put(159,47){\makebox(0,0){$t'$}} \put(181,47){\makebox(0,0){$s'$}} \put(160,37){\makebox(0,0){$c$}} \put(180,37){\makebox(0,0){$c$}} \put(160,29){\makebox(0,0){$b$}} \put(180,29){\makebox(0,0){$b$}} \put(164,13.5){\makebox(0,0){$b'$}} \put(176,13.5){\makebox(0,0){$b'$}} \put(152,44){\makebox(0,0){$t^*$}} \put(188,44){\makebox(0,0){$s^*$}} \put(150,5){\makebox(0,0){$\lambda$}} \put(190,4){\makebox(0,0){$\mu$}} \put(150,60){\makebox(0,0){$\lambda'$}} \put(190,60){\makebox(0,0){$\mu'$}} \put(170,60){\makebox(0,0){$\nu$}} \put(170,41){\makebox(0,0){$\rho$}} \put(170,24){\makebox(0,0){$\tau$}} \end{picture} \end{center} \caption{The identity in ${\rm{Hom}}(\lambda'\co{\mu'},\lambda\co\mu)$} \label{proofkey1} \end{figure} Next we can slide one of the trivalent vertices of the wire $\nu$ around the wire $a$. Using the identity of Fig.\ \ref{trick}, we obtain a factor $d_\nu/d_a$, and we can now proceed with the summation over $a$, again using Lemma \ref{idexp}. Using also Lemma \ref{idexp} for the parallel wires $c$, $c'$ as well as $b$ and $b'$, we obtain the right-hand side of Fig.\ \ref{proofkey1}. Using now Lemma \ref{idexp} once again for the wires $\rho$, $\tau$, we can pull the wire $\nu$ over the middle expansion. The summation over $\nu$ yields a killing ring which disconnects the picture into two halves, one is an intertwiner in ${\rm{Hom}}(\lambda',\lambda)$ and the other in ${\rm{Hom}}(\co{\mu'},\co\mu)$. Hence we obtain a factor $\del \lambda{\lambda'} \del \mu{\mu'}$, and we conclude that the left-hand side in Fig.\ \ref{keyrelpict} represents a scalar intertwiner $\del \lambda{\lambda'} \del \mu{\mu'} \zeta{\bf1}_N \in{\rm{Hom}}(\lambda\co\mu,\lambda\co\mu)$, $\zeta\in\Bbb{C}$. To compute that scalar, we can start again on the left-hand side of Fig.\ \ref{keyrelpict}, now putting $\lambda'=\lambda$ and $\mu'=\mu$. The diagram on the left-hand side of Fig.\ \ref{proofkey2} clearly represents an intertwiner of the same scalar value $\zeta$. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(225,65) \allinethickness{1.0pt} \put(13,32.5){\makebox(0,0){$\displaystyle\sum_a\; \frac{d_a}{d_\lambda d_\mu}$}} \put(32,5){\line(0,1){55}} \put(83,20){\line(0,1){25}} \put(37,5){\arc{10}{0}{3.142}} \put(37,60){\arc{10}{3.142}{0}} \put(47,5){\arc{10}{3.142}{4.712}} \put(47,60){\arc{10}{1.571}{3.142}} \put(73,45){\arc{20}{4.712}{0}} \put(47,45){\arc{10}{1.571}{4.712}} \put(73,45){\arc{10}{4.712}{1.571}} \put(73,20){\arc{20}{0}{1.571}} \put(47,20){\arc{10}{1.571}{4.712}} \put(73,20){\arc{10}{4.712}{1.571}} \put(47,40){\line(1,0){26}} \put(47,50){\line(1,0){1}} \put(52,50){\line(1,0){21}} \put(47,55){\line(1,0){1}} \put(52,55){\line(1,0){21}} \put(47,10){\line(1,0){1}} \put(52,10){\line(1,0){21}} \put(47,15){\line(1,0){1}} \put(52,15){\line(1,0){21}} \put(47,25){\line(1,0){26}} \allinethickness{0.3pt} \put(75,57){\arc{10}{3.142}{4.712}} \put(58,57){\arc{16}{3.142}{4.712}} \put(75,8){\arc{10}{1.571}{3.142}} \put(58,8){\arc{16}{1.571}{3.142}} \put(50,8){\line(0,1){2}} \put(50,55){\line(0,1){2}} \put(98,8){\line(0,1){49}} \put(103,8){\line(0,1){49}} \put(93,57){\arc{10}{4.712}{0}} \put(95,57){\arc{16}{4.712}{0}} \put(93,8){\arc{10}{0}{1.571}} \put(95,8){\arc{16}{0}{1.571}} \put(58,0){\line(1,0){37}} \put(58,65){\line(1,0){37}} \put(75,3){\line(1,0){18}} \put(75,62){\line(1,0){18}} \put(50,40){\line(0,1){10}} \put(70,40){\line(0,1){8}} \put(98,30.5){\vector(0,-1){0}} \put(103,34.5){\vector(0,1){0}} \put(50,25){\line(0,-1){10}} \put(70,25){\line(0,-1){8}} \allinethickness{2.0pt} \put(70,52){\line(0,1){1}} \put(50,50){\line(0,1){5}} \put(70,13){\line(0,-1){1}} \put(50,15){\line(0,-1){5}} \put(60,37){\makebox(0,0){$c'$}} \put(60,46){\makebox(0,0){$b'$}} \put(50,36.8){\makebox(0,0){$t'$}} \put(70,36.8){\makebox(0,0){$s'$}} \put(27,8){\makebox(0,0){$a$}} \put(60,29){\makebox(0,0){$c$}} \put(60,19){\makebox(0,0){$b$}} \put(50,29){\makebox(0,0){$t^*$}} \put(70,29){\makebox(0,0){$s^*$}} \put(93,31){\makebox(0,0){$\mu$}} \put(108,32.5){\makebox(0,0){$\lambda$}} \put(138,32.5){\makebox(0,0){$\longleftrightarrow\;\displaystyle\sum_a\; \frac{d_a}{d_\lambda d_\mu}$}} \allinethickness{1.0pt} \put(166,15){\line(0,1){35}} \put(224,15){\line(0,1){35}} \put(169,15){\arc{6}{0}{3.142}} \put(169,50){\arc{6}{3.142}{0}} \put(221,15){\arc{6}{0}{3.142}} \put(221,50){\arc{6}{3.142}{0}} \put(182,15){\arc{20}{3.142}{4.712}} \put(208,15){\arc{20}{4.712}{0}} \put(182,15){\arc{10}{1.571}{4.712}} \put(208,15){\arc{10}{4.712}{1.571}} \put(182,50){\arc{20}{1.571}{3.142}} \put(208,50){\arc{20}{0}{1.571}} \put(182,50){\arc{10}{1.571}{4.712}} \put(208,50){\arc{10}{4.712}{1.571}} \put(182,10){\line(1,0){26}} \put(182,20){\line(1,0){1}} \put(187,20){\line(1,0){21}} \put(182,25){\line(1,0){1}} \put(187,25){\line(1,0){21}} \put(182,40){\line(1,0){1}} \put(187,40){\line(1,0){21}} \put(182,45){\line(1,0){1}} \put(187,45){\line(1,0){21}} \put(182,55){\line(1,0){26}} \allinethickness{0.3pt} \put(185,10){\line(0,1){10}} \put(185,25){\line(0,1){15}} \put(205,10){\line(0,1){8}} \put(205,27){\line(0,1){11}} \put(185,30.5){\vector(0,-1){0}} \put(205,34.5){\vector(0,1){0}} \put(185,55){\line(0,-1){10}} \put(205,55){\line(0,-1){8}} \allinethickness{2.0pt} \put(205,22){\line(0,1){1}} \put(185,20){\line(0,1){5}} \put(205,43){\line(0,-1){1}} \put(185,45){\line(0,-1){5}} \put(164,9){\makebox(0,0){$a$}} \put(195,6){\makebox(0,0){$c'$}} \put(195,16){\makebox(0,0){$b'$}} \put(185,5){\makebox(0,0){$t'$}} \put(205,5){\makebox(0,0){$s'$}} \put(195,58){\makebox(0,0){$c$}} \put(195,49){\makebox(0,0){$b$}} \put(185,60){\makebox(0,0){$t^*$}} \put(205,60){\makebox(0,0){$s^*$}} \put(180,32.5){\makebox(0,0){$\lambda$}} \put(210,31.5){\makebox(0,0){$\mu$}} \end{picture} \end{center} \caption{Computation of the scalar $\zeta$} \label{proofkey2} \end{figure} We can now use the move of Fig.\ \ref{cut} which does not change the scalar value: We open the wire $a$ on the left and close it on the right. The resulting diagram is regularly isotopic to the diagram on the right-hand side of Fig.\ \ref{proofkey2}. Thus we are left with exactly the diagram for $d_\lambda^{-1}d_\mu^{-1}\tau_v (|\Omega_{b',c',t',s'}^{\lambda,\mu}\rangle \langle\Omega_{b,c,t,s}^{\lambda,\mu}|)$. This proves the lemma. \end{proof} The following is now immediate by the definition of the vertical product. \begin{corollary} \label{matunits} Let $\Phi_j^{\lambda,\mu}\in H_{\lambda,\mu}$ and $\Psi_j^{\lambda',\mu'}\in H_{\lambda',\mu'}$, $j=1,2$. Then we have \begin{equation} |\Phi_1^{\lambda,\mu}\rangle\langle\Phi_2^{\lambda,\mu}| *_v |\Psi_1^{\lambda',\mu'}\rangle\langle\Psi_2^{\lambda',\mu'}| \;=\; \del \lambda{\lambda'} \, \del \mu{\mu'} \, \langle\Phi_2^{\lambda,\mu},\Psi_1^{\lambda,\mu}\rangle \;|\Phi_1^{\lambda,\mu}\rangle\langle\Psi_2^{\lambda,\mu}| \end{equation} in the double triangle algebra. \end{corollary} Whenever $H_{\lambda,\mu}\neq\{0\}$ we can choose an orthonormal basis $\{ E_i^{\lambda,\mu} \}_{i=1}^{{\rm{dim}} H_{\lambda,\mu}}$. Then Lemma \ref{innerprod} and Corollary \ref{matunits} tell us that $\{\,|E_i^{\lambda,\mu}\rangle\langle E_j^{\lambda,\mu}|\,\}_{\lambda,\mu,i,j}$ forms a set of non-zero matrix units in $({\cal{Z}}_h,*_v)$. However, we do not know yet whether this is a complete set. \begin{lemma} \label{miniunit} Let $\pi_{\lambda,\mu}(e_\beta)\Omega_{b,c,t,s}^{\lambda,\mu}\in{\cal{H}}_{\lambda,\mu}$ denote the vector which is given graphically by the diagram in Fig.\ \ref{piebOm}, where $\lambda,\mu\in{}_N {\cal X}_N$, $b,c\in{}_N {\cal X}_M$, and $t\in{\rm{Hom}}(\lambda,b\co c)$, $s\in{\rm{Hom}}(\co\mu,c\co b)$ are isometries. Then in fact $\pi_{\lambda,\mu}(e_\beta)\Omega_{b,c,t,s}^{\lambda,\mu}\in H_{\lambda,\mu}$. \end{lemma} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(86,50) \allinethickness{1.0pt} \put(32,0){\line(0,1){30}} \put(78,0){\line(0,1){30}} \put(42,30){\arc{20}{3.142}{4.712}} \put(68,30){\arc{20}{4.712}{0}} \put(42,30){\arc{10}{1.571}{4.712}} \put(68,30){\arc{10}{4.712}{1.571}} \put(42,25){\line(1,0){26}} \put(42,35){\line(1,0){1}} \put(47,35){\line(1,0){21}} \put(42,40){\line(1,0){1}} \put(47,40){\line(1,0){21}} \allinethickness{0.3pt} \put(45,25){\line(0,1){10}} \put(45,40){\line(0,1){10}} \put(65,25){\line(0,1){8}} \put(65,42){\line(0,1){8}} \put(45,43){\vector(0,-1){0}} \put(65,47){\vector(0,1){0}} \put(32,10){\arc{5}{4.712}{1.571}} \put(78,10){\arc{5}{1.571}{4.712}} \allinethickness{2.0pt} \put(32,10){\line(1,0){46}} \put(65,37){\line(0,1){1}} \put(45,35){\line(0,1){5}} \put(57,10){\vector(1,0){0}} \put(27,3){\makebox(0,0){$a$}} \put(82,3){\makebox(0,0){$a$}} \put(82,22){\makebox(0,0){$a'$}} \put(55,22){\makebox(0,0){$c$}} \put(55,31){\makebox(0,0){$b$}} \put(55,4){\makebox(0,0){$\beta$}} \put(45,20.8){\makebox(0,0){$t$}} \put(65,20){\makebox(0,0){$s$}} \put(40,46){\makebox(0,0){$\lambda$}} \put(70,45){\makebox(0,0){$\mu$}} \put(10,25){\makebox(0,0){$\displaystyle\sum_{a,a'}\,d_{a'}$}} \end{picture} \end{center} \caption{The vector $\pi_{\lambda,\mu}(e_\beta) \Omega_{b,c,t,s}^{\lambda,\mu}\in{\cal{H}}_{\lambda,\mu}$} \label{piebOm} \end{figure} \begin{proof} Using Lemma \ref{idexp} and also the trick of Fig.\ \ref{trick}, we can draw the diagram on the left-hand side in Fig.\ \ref{proofpieb1} for $\pi_{\lambda,\mu}(e_\beta)\Omega_{b,c,t,s}^{\lambda,\mu}$. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(237,55) \allinethickness{1.0pt} \put(11,25){\makebox(0,0){$\displaystyle\sum_{a,a',\nu}\; d_\beta$}} \put(25,0){\line(0,1){20}} \put(30,20){\arc{10}{3.142}{4.712}} \put(30,25){\line(1,0){2.5}} \put(105,0){\line(0,1){20}} \put(100,20){\arc{10}{4.712}{0}} \put(97.5,25){\line(1,0){2.5}} \put(52.5,15){\line(1,0){25}} \put(52.5,20){\arc{10}{1.571}{3.142}} \put(77.5,20){\arc{10}{0}{1.571}} \put(47.5,20){\line(0,1){10}} \put(82.5,20){\line(0,1){10}} \put(42.5,30){\arc{10}{4.712}{0}} \put(87.5,30){\arc{10}{3.142}{4.712}} \put(37.5,30){\arc{10}{3.142}{4.712}} \put(92.5,30){\arc{10}{4.712}{0}} \put(37.5,35){\line(1,0){5}} \put(92.5,35){\line(-1,0){5}} \put(32.5,25){\line(0,1){5}} \put(97.5,25){\line(0,1){5}} \allinethickness{2.0pt} \put(32.5,25){\line(1,0){2.5}} \put(35,20){\arc{10}{4.712}{0}} \put(40,5){\line(0,1){15}} \put(45,5){\arc{10}{1.571}{3.142}} \put(45,0){\line(1,0){40}} \put(85,5){\arc{10}{0}{1.571}} \put(90,5){\line(0,1){15}} \put(95,20){\arc{10}{3.142}{4.712}} \put(95,25){\line(1,0){2.5}} \put(67,0){\vector(1,0){0}} \allinethickness{0.3pt} \dottedline{4}(20,20)(110,20) \put(55,15){\line(0,1){40}} \put(75,15){\line(0,1){28}} \put(75,47){\line(0,1){8}} \put(40,35){\line(0,1){5}} \put(90,35){\line(0,1){5}} \put(45,40){\arc{10}{3.142}{4.712}} \put(85,40){\arc{10}{4.712}{0}} \put(45,45){\line(1,0){8}} \put(57,45){\line(1,0){28}} \put(40,35){\arc{5}{3.142}{0}} \put(90,35){\arc{5}{3.142}{0}} \put(32.5,25){\arc{5}{3.142}{0}} \put(97.5,25){\arc{5}{3.142}{0}} \put(67,45){\vector(1,0){0}} \put(55,33){\vector(0,-1){0}} \put(75,37){\vector(0,1){0}} \put(20,3){\makebox(0,0){$a$}} \put(110,3){\makebox(0,0){$a$}} \put(65,5){\makebox(0,0){$\beta$}} \put(65,12){\makebox(0,0){$c$}} \put(55,10){\makebox(0,0){$t$}} \put(75,10){\makebox(0,0){$s$}} \put(59,28){\makebox(0,0){$\lambda$}} \put(71,26.5){\makebox(0,0){$\mu$}} \put(65,50){\makebox(0,0){$\nu$}} \put(34,38){\makebox(0,0){$a'$}} \put(46,38){\makebox(0,0){$b$}} \put(84,38){\makebox(0,0){$b$}} \put(97,38){\makebox(0,0){$a'$}} \put(130,25){\makebox(0,0){$=\;\displaystyle\sum_{a,a',\nu}\; d_\beta$}} \allinethickness{1.0pt} \put(150,0){\line(0,1){30}} \put(155,30){\arc{10}{3.142}{4.712}} \put(230,0){\line(0,1){30}} \put(225,30){\arc{10}{4.712}{0}} \put(177.5,15){\line(1,0){25}} \put(177.5,20){\arc{10}{1.571}{3.142}} \put(202.5,20){\arc{10}{0}{1.571}} \put(167.5,20){\arc{10}{4.712}{0}} \put(212.5,20){\arc{10}{3.142}{4.712}} \put(220,30){\arc{10}{3.142}{4.712}} \put(160,30){\arc{10}{4.712}{0}} \put(155,35){\line(1,0){5}} \put(225,35){\line(-1,0){5}} \put(165,25){\line(0,1){5}} \put(215,25){\line(0,1){5}} \put(165,25){\line(1,0){2.5}} \put(212.5,25){\line(1,0){2.5}} \allinethickness{2.0pt} \put(162.5,25){\line(1,0){2.5}} \put(162.5,20){\arc{10}{3.142}{4.712}} \put(157.5,5){\line(0,1){15}} \put(162.5,5){\arc{10}{1.571}{3.142}} \put(162.5,0){\line(1,0){55}} \put(217.5,5){\arc{10}{0}{1.571}} \put(222.5,5){\line(0,1){15}} \put(217.5,20){\arc{10}{4.712}{0}} \put(215,25){\line(1,0){2.5}} \put(192,0){\vector(1,0){0}} \allinethickness{0.3pt} \dottedline{4}(145,20)(235,20) \put(180,15){\line(0,1){40}} \put(200,15){\line(0,1){28}} \put(200,47){\line(0,1){8}} \put(157.5,35){\line(0,1){5}} \put(222.5,35){\line(0,1){5}} \put(162.5,40){\arc{10}{3.142}{4.712}} \put(217.5,40){\arc{10}{4.712}{0}} \put(162.5,45){\line(1,0){15.5}} \put(182,45){\line(1,0){35.5}} \put(157.5,35){\arc{5}{3.142}{0}} \put(222.5,35){\arc{5}{3.142}{0}} \put(165,25){\arc{5}{3.142}{0}} \put(215,25){\arc{5}{3.142}{0}} \put(192,45){\vector(1,0){0}} \put(180,33){\vector(0,-1){0}} \put(200,37){\vector(0,1){0}} \put(145,3){\makebox(0,0){$a$}} \put(235,3){\makebox(0,0){$a$}} \put(190,5){\makebox(0,0){$\beta$}} \put(190,12){\makebox(0,0){$c$}} \put(180,10){\makebox(0,0){$t$}} \put(200,10){\makebox(0,0){$s$}} \put(184,28){\makebox(0,0){$\lambda$}} \put(196,26.5){\makebox(0,0){$\mu$}} \put(190,50){\makebox(0,0){$\nu$}} \put(165,38){\makebox(0,0){$a'$}} \put(171,15){\makebox(0,0){$b$}} \put(209,15){\makebox(0,0){$b$}} \put(217,38){\makebox(0,0){$a'$}} \end{picture} \end{center} \caption{The vector $\pi_{\lambda,\mu}(e_\beta) \Omega_{b,c,t,s}^{\lambda,\mu}\in{\cal{H}}_{\lambda,\mu}$} \label{proofpieb1} \end{figure} Now let us look at the part of this picture above the dotted line. In a suitable Frobenius annulus, this part can be read for fixed $\nu$ and $a$ as $\sum_i \lambda\co\mu(t_i)\epsm\nu{\lambda\co\mu}t_i^*$, and the sum runs over a full orthonormal basis of isometries $t_i$ in the Hilbert space ${\rm{Hom}}(\nu,b\co\beta\co a)$ since we have the summation over $a'$. Next we look at the part above the dotted line on the right-hand side of Fig.\ \ref{proofpieb1}. This can be similarly read for fixed $\nu$ and $a$ as $\sum_j \lambda\co\mu(s_j)\epsm\nu{\lambda\co\mu}s_j^*$, where the sum runs over another full orthonormal basis of isometries $s_j\in{\rm{Hom}}(\nu,b\co\beta\co a)$. Since such bases $\{t_i\}$ and $\{s_j\}$ are related by a unitary matrix transformation (this is again just ``unitarity of $6j$-symbols''), the left- and right-hand side represent the same vector in ${\cal{H}}_{\lambda,\mu}$. Then, using again Lemma \ref{idexp} and also the trick of Fig.\ \ref{trick}, we conclude that the vector $\pi_{\lambda,\mu}(e_\beta)\Omega_{b,c,t,s}^{\lambda,\mu}$ can be represented by the diagram on the left-hand side of Fig.\ \ref{proofpieb2}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(217,50) \allinethickness{1.0pt} \put(32,0){\line(0,1){30}} \put(98,0){\line(0,1){30}} \put(42,30){\arc{20}{3.142}{4.712}} \put(88,30){\arc{20}{4.712}{0}} \put(42,30){\arc{10}{3.142}{4.712}} \put(88,30){\arc{10}{4.712}{0}} \put(37,25){\line(0,1){5}} \put(93,25){\line(0,1){5}} \put(42,25){\arc{10}{1.571}{3.142}} \put(88,25){\arc{10}{0}{1.571}} \put(42,20){\line(1,0){46}} \put(42,35){\line(1,0){11}} \put(57,35){\line(1,0){31}} \put(42,40){\line(1,0){11}} \put(57,40){\line(1,0){31}} \allinethickness{0.3pt} \put(55,20){\line(0,1){15}} \put(55,40){\line(0,1){10}} \put(75,20){\line(0,1){13}} \put(75,42){\line(0,1){8}} \put(55,43){\vector(0,-1){0}} \put(75,47){\vector(0,1){0}} \put(48,20){\arc{5}{0}{3.142}} \put(82,20){\arc{5}{0}{3.142}} \dottedline{4}(42,0)(88,0)(88,27.5)(42,27.5)(42,0) \allinethickness{2.0pt} \put(48,20){\line(0,-1){5}} \put(82,20){\line(0,-1){5}} \put(58,15){\arc{20}{1.571}{3.142}} \put(72,15){\arc{20}{0}{1.571}} \put(58,5){\line(1,0){14}} \put(75,37){\line(0,1){1}} \put(55,35){\line(0,1){5}} \put(67,5){\vector(1,0){0}} \put(27,3){\makebox(0,0){$a$}} \put(65,31){\makebox(0,0){$a'$}} \put(65,17){\makebox(0,0){$c$}} \put(79,24){\makebox(0,0){$b$}} \put(51,24){\makebox(0,0){$b$}} \put(65,9){\makebox(0,0){$\beta$}} \put(55,15.8){\makebox(0,0){$t$}} \put(75,15){\makebox(0,0){$s$}} \put(50,46){\makebox(0,0){$\lambda$}} \put(80,45){\makebox(0,0){$\mu$}} \put(10,25){\makebox(0,0){$\displaystyle\sum_{a,a'}\,d_{a'}$}} \put(140,12){\makebox(0,0){$\longleftrightarrow \displaystyle\sum_{c',i,j}\,{\rm{coeff}}_{(c',i,j)}$}} \allinethickness{0.3pt} \dottedline{4}(170,0)(216,0)(216,27.5)(170,27.5)(170,0) \allinethickness{1.0pt} \put(170,20){\line(1,0){46}} \allinethickness{0.3pt} \put(183,20){\line(0,1){7.5}} \put(203,20){\line(0,1){7.5}} \put(183,22){\vector(0,-1){0}} \put(203,25.5){\vector(0,1){0}} \put(182,11){\makebox(0,0){$t_{a',\co{c'}}^{\lambda;i}$}} \put(204,11){\makebox(0,0){$t_{c',\co{a'}}^{\co\mu;j}$}} \put(175,24){\makebox(0,0){$a'$}} \put(211,24){\makebox(0,0){$a'$}} \put(187,24){\makebox(0,0){$\lambda$}} \put(199,24){\makebox(0,0){$\mu$}} \put(193,16.5){\makebox(0,0){$c'$}} \end{picture} \end{center} \caption{The vector $\pi_{\lambda,\mu}(e_\beta) \Omega_{b,c,t,s}^{\lambda,\mu}\in{\cal{H}}_{\lambda,\mu}$} \label{proofpieb2} \end{figure} Now let us look at the part of the diagram inside the dotted box. In a suitable Frobenius annulus, this can be interpreted as an intertwiner in ${\rm{Hom}}(\lambda\co\mu,a'\co{a'})$. But any element in this space can be written as a linear combination of elements constructed from basis isometries $t_{a',\co{c'}}^{\lambda;i}$, $t_{c',\co{a'}}^{\co\mu;j}$, as indicated in the dotted box on the right-hand side of Fig.\ \ref{proofpieb2}. The coefficients in its linear expansion depend only on $c',i,j$ for fixed $a',\beta,b,c,t,s$, but certainly not on $a$. This shows that $\pi_{\lambda,\mu}(e_\beta)\Omega_{b,c,t,s}^{\lambda,\mu}$ is a linear combination of $\Omega_\xi^{\lambda,\mu}$'s, thus $\pi_{\lambda,\mu}(e_\beta)\Omega_{b,c,t,s}^{\lambda,\mu}\in H_{\lambda,\mu}$. \end{proof} The map $\Omega_{b,c,t,s}^{\lambda,\mu}\mapsto\pi_{\lambda,\mu}(e_\beta) \Omega_{b,c,t,s}^{\lambda,\mu}$ defines clearly a linear map $\pi_{\lambda,\mu}(e_\beta):H_{\lambda,\mu}\rightarrow{\cal{H}}_{\lambda,\mu}$ since it is just a linear intertwiner multiplication on each ${\rm{Hom}}(\lambda\co\mu,a\co a)$ block. From Lemma \ref{miniunit} we now learn that $\pi_{\lambda,\mu}(e_\beta)$ is in fact a linear operator on $H_{\lambda,\mu}$. With the definition of the vertical product we now immediately obtain the following \begin{corollary} \label{decompieb} With orthonormal bases $\{ E_i^{\lambda,\mu} \}_{i=1}^{{\rm{dim}} H_{\lambda,\mu}}$ of each $H_{\lambda,\mu}$ we have \begin{equation} | E_i^{\lambda,\mu}\rangle\langle E_j^{\lambda,\mu}| *_v e_\beta *_v | E_k^{\lambda',\mu'}\rangle\langle E_l^{\lambda',\mu'}| \;=\; \del \lambda{\lambda'} \, \del \mu{\mu'} \, \langle E_j^{\lambda,\mu}, \pi_{\lambda,\mu} (e_\beta) E_k^{\lambda,\mu}\rangle \; | E_i^{\lambda,\mu}\rangle\langle\ E_l^{\lambda,\mu}| \,. \end{equation} \end{corollary} Since ${\cal{Z}}_h$ is spanned by the $e_\beta$'s, we obtain a map $\pi_{\lambda,\mu}:{\cal{Z}}_h\rightarrow B(H_{\lambda,\mu})$ by linear extension, and we obtain similarly the following \begin{corollary} The map $\pi_{\lambda,\mu}:{\cal{Z}}_h\rightarrow B(H_{\lambda,\mu})$ is a representation of $({\cal{Z}}_h,*_v)$. \end{corollary} We now tackle the problem of completeness of the system of matrix units. \begin{definition} \label{character}{\rm For $\lambda,\mu\in{}_N {\cal X}_N$ we define the {\sl vertical projector} $q_{\lambda,\mu}\in\dta$ by \begin{equation} q_{\lambda,\mu} =\frac{\sqrt{d_\lambda d_\mu}}{w^2} \sum_\xi \, |\Omega_\xi^{\lambda,\mu} \rangle\langle \Omega_\xi^{\lambda,\mu}| \,. \label{qlmeq} \end{equation} }\end{definition} \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.7mm \begin{picture}(85,50) \allinethickness{1.0pt} \put(15,25){\makebox(0,0){$\displaystyle\sum_{a,b,c,d} \frac{d_b d_c}{w^2}$}} \put(37,0){\line(0,1){10}} \put(83,0){\line(0,1){10}} \put(47,10){\arc{20}{3.142}{4.712}} \put(73,10){\arc{20}{4.712}{0}} \put(47,10){\arc{10}{1.571}{4.712}} \put(73,10){\arc{10}{4.712}{1.571}} \put(47,40){\arc{20}{1.571}{3.142}} \put(73,40){\arc{20}{0}{1.571}} \put(47,40){\arc{10}{1.571}{4.712}} \put(73,40){\arc{10}{4.712}{1.571}} \put(47,5){\line(1,0){26}} \put(47,15){\line(1,0){1}} \put(52,15){\line(1,0){21}} \put(47,20){\line(1,0){1}} \put(52,20){\line(1,0){21}} \put(37,50){\line(0,-1){10}} \put(83,50){\line(0,-1){10}} \put(47,30){\line(1,0){1}} \put(52,30){\line(1,0){21}} \put(47,35){\line(1,0){1}} \put(52,35){\line(1,0){21}} \put(47,45){\line(1,0){26}} \allinethickness{0.3pt} \put(50,5){\line(0,1){10}} \put(50,20){\line(0,1){10}} \put(70,5){\line(0,1){8}} \put(70,22){\line(0,1){6}} \put(50,23){\vector(0,-1){0}} \put(70,27){\vector(0,1){0}} \put(50,45){\line(0,-1){10}} \put(70,45){\line(0,-1){8}} \put(50,5){\arc{5}{3.142}{0}} \put(70,5){\arc{5}{3.142}{0}} \put(50,45){\arc{5}{0}{3.142}} \put(70,45){\arc{5}{0}{3.142}} \allinethickness{2.0pt} \put(70,17){\line(0,1){1}} \put(50,15){\line(0,1){5}} \put(70,33){\line(0,-1){1}} \put(50,35){\line(0,-1){5}} \put(33,4){\makebox(0,0){$d$}} \put(60,2){\makebox(0,0){$c$}} \put(60,11){\makebox(0,0){$b$}} \put(33,46){\makebox(0,0){$a$}} \put(60,48){\makebox(0,0){$c$}} \put(60,39){\makebox(0,0){$b$}} \put(45,25){\makebox(0,0){$\lambda$}} \put(75,24){\makebox(0,0){$\mu$}} \end{picture} \end{center} \caption{A vertical projector $q_{\lambda,\mu}$} \label{qlamu} \end{figure} This is given graphically in Fig.\ \ref{qlamu}. (Clearly, we can use Lemma \ref{idexp} twice to obtain an equivalent picture which does not involve pieces of very thick wires corresponding to $\alpha_\lambda^+$ and $\alpha_\mu^-$.) We are now ready to prove the main result of this section. \begin{theorem} \label{character2} Under Assumption \ref{set-nondeg}, the vertical projector $q_{\lambda,\mu}$ is either zero or a minimal central projection in $({\cal{Z}}_h, *_v)$. We have mutual orthogonality $q_{\lambda,\mu} *_v q_{\lambda',\mu'}=\del \lambda{\lambda'}\del \mu{\mu'} q_{\lambda,\mu}$ and the vertical projectors sum up to the multiplicative identity of $({\cal Z}_h,*_v)$: $\sum_{\lambda,\mu\in{}_N {\cal X}_N} q_{\lambda,\mu}=e_0$. Moreover, $q_{\lambda,\mu}=0$ whenever $Z_{\lambda,\mu}=0$ and otherwise the simple summand $q_{\lambda,\mu} *_v {\cal{Z}}_h$ is a full $Z_{\lambda,\mu}\times Z_{\lambda,\mu}$ matrix algebra, where $Z_{\lambda,\mu}$ is the $(\lambda,\mu)$-entry of the modular invariant mass matrix of Definition \ref{modular1}. \end{theorem} \begin{proof} It follows from Corollary \ref{matunits} that $q_{\lambda,\mu} *_v q_{\lambda',\mu'}=0$ unless $\lambda=\lambda'$ and $\mu=\mu'$. We now show that $\sum_{\lambda,\mu} q_{\lambda,\mu}=e_0$. (We denote $e_0\equiv e_{{\rm{id}}_M}$.) The sum is given graphically by the left-hand side in Fig.\ \ref{sumqlamu1}. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(205,50) \allinethickness{1.0pt} \put(15,25){\makebox(0,0){$\displaystyle\sum_{a,b,c,d,\lambda,\mu} \frac{d_b d_c}{w^2}$}} \put(37,0){\line(0,1){10}} \put(83,0){\line(0,1){10}} \put(47,10){\arc{20}{3.142}{4.712}} \put(73,10){\arc{20}{4.712}{0}} \put(47,10){\arc{10}{1.571}{4.712}} \put(73,10){\arc{10}{4.712}{1.571}} \put(47,40){\arc{20}{1.571}{3.142}} \put(73,40){\arc{20}{0}{1.571}} \put(47,40){\arc{10}{1.571}{4.712}} \put(73,40){\arc{10}{4.712}{1.571}} \put(47,5){\line(1,0){26}} \put(47,15){\line(1,0){1}} \put(52,15){\line(1,0){21}} \put(47,20){\line(1,0){1}} \put(52,20){\line(1,0){21}} \put(37,50){\line(0,-1){10}} \put(83,50){\line(0,-1){10}} \put(47,30){\line(1,0){1}} \put(52,30){\line(1,0){21}} \put(47,35){\line(1,0){1}} \put(52,35){\line(1,0){21}} \put(47,45){\line(1,0){26}} \allinethickness{0.3pt} \put(50,5){\line(0,1){10}} \put(50,20){\line(0,1){10}} \put(70,5){\line(0,1){8}} \put(70,22){\line(0,1){6}} \put(50,23){\vector(0,-1){0}} \put(70,27){\vector(0,1){0}} \put(50,45){\line(0,-1){10}} \put(70,45){\line(0,-1){8}} \put(50,5){\arc{5}{3.142}{0}} \put(70,5){\arc{5}{3.142}{0}} \put(50,45){\arc{5}{0}{3.142}} \put(70,45){\arc{5}{0}{3.142}} \allinethickness{2.0pt} \put(70,17){\line(0,1){1}} \put(50,15){\line(0,1){5}} \put(70,33){\line(0,-1){1}} \put(50,35){\line(0,-1){5}} \put(33,4){\makebox(0,0){$d$}} \put(60,2){\makebox(0,0){$c$}} \put(60,11){\makebox(0,0){$b$}} \put(33,46){\makebox(0,0){$a$}} \put(60,48){\makebox(0,0){$c$}} \put(60,39){\makebox(0,0){$b$}} \put(45,25){\makebox(0,0){$\lambda$}} \put(75,24){\makebox(0,0){$\mu$}} \allinethickness{1.0pt} \put(115,25){\makebox(0,0){$=\;\displaystyle\sum_{a,b,c,d,\lambda,\mu,\nu,\rho} \frac{d_b d_c}{w^2}$}} \put(152,0){\arc{10}{3.142}{4.712}} \put(195,0){\arc{10}{4.712}{0}} \put(152,50){\arc{10}{1.571}{3.142}} \put(195,50){\arc{10}{0}{1.571}} \put(152,5){\line(1,0){43}} \put(152,45){\line(1,0){43}} \allinethickness{0.3pt} \put(155,5){\line(0,1){5}} \put(155,45){\line(0,-1){5}} \put(169,20){\line(1,0){13}} \put(169,30){\line(1,0){13}} \put(155,5){\arc{5}{3.142}{0}} \put(155,45){\arc{5}{0}{3.142}} \put(192,45){\arc{5}{0}{3.142}} \put(192,5){\arc{5}{3.142}{0}} \put(182,45){\arc{5}{0}{3.142}} \put(182,5){\arc{5}{3.142}{0}} \put(167,45){\arc{5}{0}{3.142}} \put(167,5){\arc{5}{3.142}{0}} \put(165,10){\arc{20}{3.142}{4.712}} \put(165,40){\arc{20}{1.571}{3.142}} \put(182,10){\arc{20}{4.712}{0}} \put(182,40){\arc{20}{0}{1.571}} \put(192,5){\line(0,1){5}} \put(192,45){\line(0,-1){5}} \put(167,5){\line(0,1){40}} \put(182,5){\line(0,1){13}} \put(182,45){\line(0,-1){13}} \put(182,22){\line(0,1){6}} \put(182,27){\vector(0,1){0}} \put(167,23){\vector(0,-1){0}} \put(176.5,30){\vector(1,0){0}} \put(172.5,20){\vector(-1,0){0}} \put(187,24){\makebox(0,0){$\mu$}} \put(162,25){\makebox(0,0){$\lambda$}} \put(153,19){\makebox(0,0){$\rho$}} \put(153,31){\makebox(0,0){$\nu$}} \put(144,5){\makebox(0,0){$d$}} \put(144,45){\makebox(0,0){$a$}} \put(203,5){\makebox(0,0){$d$}} \put(203,45){\makebox(0,0){$a$}} \put(174.5,2){\makebox(0,0){$c$}} \put(174.5,48){\makebox(0,0){$c$}} \put(163,9){\makebox(0,0){$b$}} \put(163,41){\makebox(0,0){$b$}} \put(186,9){\makebox(0,0){$b$}} \put(186,41){\makebox(0,0){$b$}} \end{picture} \end{center} \caption{The sum $\sum_{\lambda,\mu} q_{\lambda,\mu}$} \label{sumqlamu1} \end{figure} A twofold application of Lemma \ref{idexp} yields the right-hand side in Fig.\ \ref{sumqlamu1}. Applying Lemma \ref{idexp} twice again, we obtain the left-hand side of Fig.\ \ref{sumqlamu2}. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(240,60) \allinethickness{1.0pt} \put(24,30){\makebox(0,0){$\displaystyle \sum_{a,b,c,d\atop\mu,\nu,\rho,\tau,i,j} \frac{d_c\sqrt{d_\nu d_\rho}}{w^2\sqrt{d_a d_d}}$}} \put(57,0){\line(0,1){5}} \put(110,0){\line(0,1){5}} \put(57,60){\line(0,-1){5}} \put(110,60){\line(0,-1){5}} \put(62,5){\arc{10}{3.142}{4.712}} \put(105,5){\arc{10}{4.712}{0}} \put(77,15){\line(0,1){30}} \put(82,15){\line(0,1){30}} \put(72,15){\arc{10}{0}{1.571}} \put(72,45){\arc{10}{4.712}{0}} \put(62,55){\arc{10}{1.571}{3.142}} \put(105,55){\arc{10}{0}{1.571}} \put(62,10){\line(1,0){10}} \put(62,50){\line(1,0){10}} \put(87,15){\arc{10}{1.571}{3.142}} \put(87,45){\arc{10}{3.142}{4.712}} \put(87,10){\line(1,0){18}} \put(87,50){\line(1,0){18}} \allinethickness{0.3pt} \put(65,10){\line(0,1){40}} \put(92,50){\arc{5}{0}{3.142}} \put(92,10){\arc{5}{3.142}{0}} \put(65,30){\arc{5}{4.712}{1.571}} \put(102,30){\arc{5}{1.571}{4.712}} \put(102,10){\line(0,1){40}} \put(92,10){\line(0,1){18}} \put(92,50){\line(0,-1){18}} \put(65,30){\line(1,0){10}} \put(82,30){\line(1,0){20}} \put(65,38){\vector(0,-1){0}} \put(65,18){\vector(0,-1){0}} \put(102,42){\vector(0,1){0}} \put(102,22){\vector(0,1){0}} \put(89,30){\vector(1,0){0}} \put(92,22){\vector(0,1){0}} \allinethickness{2.0pt} \put(79,30){\line(1,0){1}} \put(97,35){\makebox(0,0){$\mu$}} \put(60,20){\makebox(0,0){$\rho$}} \put(60,40){\makebox(0,0){$\nu$}} \put(107,20){\makebox(0,0){$\rho$}} \put(107,40){\makebox(0,0){$\nu$}} \put(71,26){\makebox(0,0){$\tau$}} \put(52,5){\makebox(0,0){$d$}} \put(52,55){\makebox(0,0){$a$}} \put(115,5){\makebox(0,0){$d$}} \put(115,55){\makebox(0,0){$a$}} \put(84,6){\makebox(0,0){$c$}} \put(76,53){\makebox(0,0){$b$}} \put(97,14){\makebox(0,0){$b$}} \put(97,46){\makebox(0,0){$b$}} \put(66.5,56.5){\makebox(0,0){$(t_{a\co b}^{\nu;i})^*$}} \put(102,56.5){\makebox(0,0){$t_{a\co b}^{\nu;i}$}} \put(65,4){\makebox(0,0){$t_{d\co b}^{\rho;j}$}} \put(101,4){\makebox(0,0){$(t_{d\co b}^{\rho;j})^*$}} \allinethickness{1.0pt} \put(144,30){\makebox(0,0){$= \displaystyle\sum_{a,b,c,d,\atop\mu,\nu,\rho,\tau,i,j} \frac{d_c\sqrt{d_\nu d_\rho}}{w^2\sqrt{d_a d_d}}$}} \put(177,0){\line(0,1){5}} \put(230,0){\line(0,1){5}} \put(177,60){\line(0,-1){5}} \put(230,60){\line(0,-1){5}} \put(182,5){\arc{10}{3.142}{4.712}} \put(225,5){\arc{10}{4.712}{0}} \put(197,15){\line(0,1){30}} \put(202,15){\line(0,1){30}} \put(192,15){\arc{10}{0}{1.571}} \put(192,45){\arc{10}{4.712}{0}} \put(182,55){\arc{10}{1.571}{3.142}} \put(225,55){\arc{10}{0}{1.571}} \put(182,10){\line(1,0){10}} \put(182,50){\line(1,0){10}} \put(207,15){\arc{10}{1.571}{3.142}} \put(207,45){\arc{10}{3.142}{4.712}} \put(207,10){\line(1,0){18}} \put(207,50){\line(1,0){18}} \allinethickness{0.3pt} \put(185,10){\line(0,1){40}} \put(210,10){\arc{5}{3.142}{0}} \put(217,10){\arc{5}{3.142}{0}} \put(185,30){\arc{5}{4.712}{1.571}} \put(222,30){\arc{5}{1.571}{4.712}} \put(213.5,38){\arc{7}{3.142}{0}} \put(222,10){\line(0,1){40}} \put(217,10){\line(0,1){18}} \put(217,32){\line(0,1){6}} \put(210,10){\line(0,1){28}} \put(185,30){\line(1,0){10}} \put(202,30){\line(1,0){6}} \put(212,30){\line(1,0){10}} \put(185,38){\vector(0,-1){0}} \put(185,18){\vector(0,-1){0}} \put(222,42){\vector(0,1){0}} \put(222,22){\vector(0,1){0}} \put(194,30){\vector(1,0){0}} \put(210,18){\vector(0,-1){0}} \allinethickness{2.0pt} \put(199,30){\line(1,0){1}} \put(207,40){\makebox(0,0){$\mu$}} \put(180,20){\makebox(0,0){$\rho$}} \put(180,40){\makebox(0,0){$\nu$}} \put(227,20){\makebox(0,0){$\rho$}} \put(227,40){\makebox(0,0){$\nu$}} \put(191,26){\makebox(0,0){$\tau$}} \put(173,5){\makebox(0,0){$d$}} \put(173,55){\makebox(0,0){$a$}} \put(235,5){\makebox(0,0){$d$}} \put(235,55){\makebox(0,0){$a$}} \put(213.5,14){\makebox(0,0){$c$}} \put(193,46){\makebox(0,0){$b$}} \put(220,15){\makebox(0,0){$b$}} \put(205,55){\makebox(0,0){$b$}} \put(186.5,56.5){\makebox(0,0){$(t_{a\co b}^{\nu;i})^*$}} \put(222,56.5){\makebox(0,0){$t_{a\co b}^{\nu;i}$}} \put(185,4){\makebox(0,0){$t_{d\co b}^{\rho;j}$}} \put(221,4){\makebox(0,0){$(t_{d\co b}^{\rho;j})^*$}} \end{picture} \end{center} \caption{The sum $\sum_{\lambda,\mu} q_{\lambda,\mu}$} \label{sumqlamu2} \end{figure} We can now slide the upper trivalent vertex of the wire $\mu$ around to obtain the right-hand side of Fig.\ \ref{sumqlamu2}. Next we can use the trick of Fig.\ \ref{trick} to turn around the small arcs at the trivalent vertices of the wire $\mu$, yielding a factor $d_\mu/d_c$. This gives the right-hand side of Fig.\ \ref{sumqlamu2}. Since we have a summation over $c$, we can again use Lemma \ref{idexp}, and this gives us the left-hand side of Fig.\ \ref{sumqlamu3}. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(240,60) \allinethickness{1.0pt} \put(24,30){\makebox(0,0){$\displaystyle \sum_{a,b,d\atop\mu,\nu,\rho,\tau,i,j} \frac{d_\mu\sqrt{d_\nu d_\rho}}{w^2\sqrt{d_a d_d}}$}} \put(57,0){\line(0,1){5}} \put(110,0){\line(0,1){5}} \put(57,60){\line(0,-1){5}} \put(110,60){\line(0,-1){5}} \put(62,5){\arc{10}{3.142}{4.712}} \put(105,5){\arc{10}{4.712}{0}} \put(77,15){\line(0,1){30}} \put(82,15){\line(0,1){30}} \put(72,15){\arc{10}{0}{1.571}} \put(72,45){\arc{10}{4.712}{0}} \put(62,55){\arc{10}{1.571}{3.142}} \put(105,55){\arc{10}{0}{1.571}} \put(62,10){\line(1,0){10}} \put(62,50){\line(1,0){10}} \put(87,15){\arc{10}{1.571}{3.142}} \put(87,45){\arc{10}{3.142}{4.712}} \put(87,10){\line(1,0){18}} \put(87,50){\line(1,0){18}} \allinethickness{0.3pt} \put(65,10){\line(0,1){40}} \put(65,30){\arc{5}{4.712}{1.571}} \put(102,30){\arc{5}{1.571}{4.712}} \put(93.5,38){\arc{7}{3.142}{0}} \put(93.5,22){\arc{7}{0}{3.142}} \put(102,10){\line(0,1){40}} \put(97,22){\line(0,1){6}} \put(97,32){\line(0,1){6}} \put(90,22){\line(0,1){16}} \put(65,30){\line(1,0){10}} \put(82,30){\line(1,0){6}} \put(92,30){\line(1,0){10}} \put(65,38){\vector(0,-1){0}} \put(65,18){\vector(0,-1){0}} \put(102,42){\vector(0,1){0}} \put(102,22){\vector(0,1){0}} \put(74,30){\vector(1,0){0}} \put(90,32){\vector(0,-1){0}} \allinethickness{2.0pt} \put(79,30){\line(1,0){1}} \put(87,20){\makebox(0,0){$\mu$}} \put(60,20){\makebox(0,0){$\rho$}} \put(60,40){\makebox(0,0){$\nu$}} \put(107,20){\makebox(0,0){$\rho$}} \put(107,40){\makebox(0,0){$\nu$}} \put(71,26){\makebox(0,0){$\tau$}} \put(52,5){\makebox(0,0){$d$}} \put(52,55){\makebox(0,0){$a$}} \put(115,5){\makebox(0,0){$d$}} \put(115,55){\makebox(0,0){$a$}} \put(84,6){\makebox(0,0){$b$}} \put(73,46){\makebox(0,0){$b$}} \put(66.5,56.5){\makebox(0,0){$(t_{a\co b}^{\nu;i})^*$}} \put(102,56.5){\makebox(0,0){$t_{a\co b}^{\nu;i}$}} \put(65,4){\makebox(0,0){$t_{d\co b}^{\rho;j}$}} \put(101,4){\makebox(0,0){$(t_{d\co b}^{\rho;j})^*$}} \allinethickness{1.0pt} \put(144,30){\makebox(0,0){$= \displaystyle\sum_{a,b,\nu,i,j} \frac 1{w d_a}$}} \put(177,0){\line(0,1){5}} \put(230,0){\line(0,1){5}} \put(177,60){\line(0,-1){5}} \put(230,60){\line(0,-1){5}} \put(182,5){\arc{10}{3.142}{4.712}} \put(225,5){\arc{10}{4.712}{0}} \put(199,15){\line(0,1){30}} \put(208,15){\line(0,1){30}} \put(194,15){\arc{10}{0}{1.571}} \put(194,45){\arc{10}{4.712}{0}} \put(182,55){\arc{10}{1.571}{3.142}} \put(225,55){\arc{10}{0}{1.571}} \put(182,10){\line(1,0){12}} \put(182,50){\line(1,0){12}} \put(213,15){\arc{10}{1.571}{3.142}} \put(213,45){\arc{10}{3.142}{4.712}} \put(213,10){\line(1,0){12}} \put(213,50){\line(1,0){12}} \allinethickness{0.3pt} \put(188,10){\line(0,1){40}} \put(219,10){\line(0,1){40}} \put(188,28){\vector(0,-1){0}} \put(219,32){\vector(0,1){0}} \put(183,30){\makebox(0,0){$\nu$}} \put(224,30){\makebox(0,0){$\nu$}} \put(173,5){\makebox(0,0){$a$}} \put(173,55){\makebox(0,0){$a$}} \put(235,5){\makebox(0,0){$a$}} \put(235,55){\makebox(0,0){$a$}} \put(195,30){\makebox(0,0){$b$}} \put(212,30){\makebox(0,0){$b$}} \put(188.5,56.5){\makebox(0,0){$(t_{a\co b}^{\nu;i})^*$}} \put(220,56.5){\makebox(0,0){$t_{a\co b}^{\nu;i}$}} \put(187,4){\makebox(0,0){$t_{a\co b}^{\nu;j}$}} \put(219,4){\makebox(0,0){$(t_{a\co b}^{\nu;j})^*$}} \end{picture} \end{center} \caption{The sum $\sum_{\lambda,\mu} q_{\lambda,\mu}$} \label{sumqlamu3} \end{figure} As we have a prefactor $d_\mu$, the summation over $\mu$ provides a killing ring, and only $\tau={\rm{id}}_N$ survives it: We obtain a factor $w\del \tau 0$. Now our picture starts to collapse. The factor $\del \tau 0$ yields, with the normalization convention as in Fig.\ \ref{e-beta}, a factor $d_\nu^{-1} \del \nu\rho$. Since our picture is now disconnected into two parts which represent intertwiners in ${\rm{Hom}}(a,d)$, they are scalars and we obtain a factor $\del ad$. This gives us the right-hand side of Fig.\ \ref{sumqlamu3}. Therefore we are now left with a sum over scalars times two straight vertical wires labelled by $a$, representing a scalar intertwiner in ${\rm{Hom}}(a\co a,a\co a)$. The scalar value of each connected part of the picture is $\del ij \sqrt{d_\nu d_b / d_a}$, therefore we can compute the prefactor as \[ \frac 1{w d_a} \sum_{b,\nu} \sum_{i,j=1}^{N_{a\co b}^\nu} \left( \sqrt{\frac{d_\nu d_b}{d_a}} \del ij \right)^2 = \frac 1{w d_a^2} \sum_{b,\nu} d_b N_{a,\co b}^\nu d_\nu = \frac 1{w d_a} \sum_b d_b^2 = \frac 1{d_a} \,. \] Thus we are left with a sum over two vertical straight wires with label $a$ and prefactor $d_a^{-1}$. This is $e_0$. Next, we can expand each vector $\Omega_\xi^{\lambda,\mu}\in H_{\lambda,\mu}$, in an orthonormal basis as \[ \Omega_\xi^{\lambda,\mu} = \sum_{i=1}^{{\rm{dim}} H_{\lambda,\mu}} \langle E_i^{\lambda,\mu},\Omega_\xi^{\lambda,\mu}\rangle E_i^{\lambda,\mu} \,.\] Inserting this in \erf{qlmeq} yields \[ q_{\lambda,\mu} = \frac{\sqrt{d_\lambda d_\mu}}{w^2} \sum_{i,j}^{{\rm{dim}} H_{\lambda,\mu}} \sum_\xi \, \langle E_i^{\lambda,\mu},\Omega_\xi^{\lambda,\mu} \rangle \langle \Omega_\xi^{\lambda,\mu}, E_j^{\lambda,\mu} \rangle \; |E_i^{\lambda,\mu} \rangle\langle E_j^{\lambda,\mu}| \,. \] Now using $\sum_{\lambda,\mu} q_{\lambda,\mu}=e_0$ and Corollary \ref{matunits} we compute \[ \begin{array}{ll} \del ij \; |E_i^{\lambda,\mu} \rangle\langle E_j^{\lambda,\mu}| &= \sum_{\lambda',\mu'} |E_i^{\lambda,\mu} \rangle\langle E_i^{\lambda,\mu}| *_v q_{\lambda',\mu'} *_v |E_j^{\lambda,\mu} \rangle\langle E_j^{\lambda,\mu}| \\[.5em] &= \displaystyle\frac{\sqrt{d_\lambda d_\mu}}{w^2} \sum_\xi \, \langle E_i^{\lambda,\mu},\Omega_\xi^{\lambda,\mu} \rangle \langle \Omega_\xi^{\lambda,\mu}, E_j^{\lambda,\mu} \rangle \; |E_i^{\lambda,\mu} \rangle\langle E_j^{\lambda,\mu}| \,, \end{array} \] hence \[ q_{\lambda,\mu} = \sum_{i=1}^{{\rm{dim}} H_{\lambda,\mu}} |E_i^{\lambda,\mu} \rangle\langle E_i^{\lambda,\mu}| \,. \] Thus $q_{\lambda,\mu}$ is a projection and we also have $e_0=\sum_{\lambda,\mu}\sum_{i=1}^{{\rm{dim}} H_{\lambda,\mu}} |E_i^{\lambda,\mu} \rangle\langle E_i^{\lambda,\mu}|$. Hence for any $\beta\in{}_M {\cal X}_M$ we find \[ e_\beta = e_0 *_v e_\beta *_v e_0 = \sum_{\lambda,\mu}\sum_{i,j=1}^{{\rm{dim}} H_{\lambda,\mu}} \langle E_i^{\lambda,\mu}, \pi_{\lambda,\mu} (e_\beta) E_j^{\lambda,\mu}\rangle \; | E_i^{\lambda,\mu}\rangle\langle\ E_j^{\lambda,\mu}| \] by Corollary \ref{decompieb}. Thus each $e_\beta$ can be expanded in our matrix units, and since ${\cal{Z}}_h$ is spanned by the $e_\beta$'s we conclude that $\{ | E_i^{\lambda,\mu}\rangle\langle\ E_j^{\lambda,\mu}| \}_{\lambda,\mu,i,j}$ is a complete system of matrix units. It follows that the non-zero vertical projectors are minimal central projections in $({\cal{Z}}_h,*_v)$, and that the simple summand $q_{\lambda,\mu} *_v {\cal{Z}}_h$ is a full ${\rm{dim}} H_{\lambda,\mu} \times {\rm{dim}} H_{\lambda,\mu}$ matrix algebra. It remains to show ${\rm{dim}} H_{\lambda,\mu}=Z_{\lambda,\mu}$. The dimension of $H_{\lambda,\mu}$ can be counted as \[ {\rm{dim}} H_{\lambda,\mu} = \sum_{i=1}^{{\rm{dim}} H_{\lambda,\mu}} \langle E_i^{\lambda,\mu}, E_i^{\lambda,\mu} \rangle = \sum_{i=1}^{{\rm{dim}} H_{\lambda,\mu}} \frac 1{d_\lambda d_\mu}\, \tau_v(| E_i^{\lambda,\mu}\rangle\langle E_i^{\lambda,\mu}|) = \frac 1{d_\lambda d_\mu}\, \tau_v (q_{\lambda,\mu}) \,. \] Now $d_\lambda^{-1} d_\mu^{-1} \tau_v (q_{\lambda,\mu})$ is given graphically in Fig.\ \ref{tauqlm}. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(108,50) \allinethickness{1.0pt} \put(18,25){\makebox(0,0){$\displaystyle\sum_{a,b,c}\, \frac{d_a d_b d_c}{w^2 d_\lambda d_\mu}$}} \put(52,5){\line(0,1){5}} \put(98,5){\line(0,1){5}} \put(62,10){\arc{20}{3.142}{4.712}} \put(88,10){\arc{20}{4.712}{0}} \put(62,10){\arc{10}{1.571}{4.712}} \put(88,10){\arc{10}{4.712}{1.571}} \put(62,40){\arc{20}{1.571}{3.142}} \put(88,40){\arc{20}{0}{1.571}} \put(62,40){\arc{10}{1.571}{4.712}} \put(88,40){\arc{10}{4.712}{1.571}} \put(47,5){\arc{10}{0}{3.142}} \put(47,45){\arc{10}{3.142}{0}} \put(103,5){\arc{10}{0}{3.142}} \put(103,45){\arc{10}{3.142}{0}} \put(42,5){\line(0,1){40}} \put(108,5){\line(0,1){40}} \put(62,5){\line(1,0){26}} \put(62,15){\line(1,0){1}} \put(67,15){\line(1,0){21}} \put(62,20){\line(1,0){1}} \put(67,20){\line(1,0){21}} \put(52,45){\line(0,-1){5}} \put(98,45){\line(0,-1){5}} \put(62,30){\line(1,0){1}} \put(67,30){\line(1,0){21}} \put(62,35){\line(1,0){1}} \put(67,35){\line(1,0){21}} \put(62,45){\line(1,0){26}} \allinethickness{0.3pt} \put(65,5){\line(0,1){10}} \put(65,20){\line(0,1){10}} \put(85,5){\line(0,1){8}} \put(85,22){\line(0,1){6}} \put(65,23){\vector(0,-1){0}} \put(85,27){\vector(0,1){0}} \put(65,45){\line(0,-1){10}} \put(85,45){\line(0,-1){8}} \put(65,5){\arc{5}{3.142}{0}} \put(85,5){\arc{5}{3.142}{0}} \put(65,45){\arc{5}{0}{3.142}} \put(85,45){\arc{5}{0}{3.142}} \allinethickness{2.0pt} \put(85,17){\line(0,1){1}} \put(65,15){\line(0,1){5}} \put(85,33){\line(0,-1){1}} \put(65,35){\line(0,-1){5}} \put(38,10){\makebox(0,0){$a$}} \put(75,2){\makebox(0,0){$c$}} \put(75,11){\makebox(0,0){$b$}} \put(75,48){\makebox(0,0){$c$}} \put(75,39){\makebox(0,0){$b$}} \put(60,25){\makebox(0,0){$\lambda$}} \put(90,24){\makebox(0,0){$\mu$}} \end{picture} \end{center} \caption{The number $d_\lambda^{-1} d_\mu^{-1} \tau_v(q_{\lambda,\mu})$} \label{tauqlm} \end{figure} By the IBFE's we can pull out the circle with label $a$ which gives us another factor $d_a$. We can therefore proceed with the summation over $a$, and this yields a factor $w$, the global index, and then we are left exactly with the picture in Fig.\ \ref{Zgraph}. \end{proof} Note that we learn from the proof that putting ${\rm{Tr}}_v(z)=\sum_{\lambda,\mu}d_\lambda^{-1}d_\mu^{-1}\tau_v(q_{\lambda,\mu}*_v z)$ for $z\in{\cal{Z}}_h$ gives a matrix trace ${\rm{Tr}}_v$ on $({\cal{Z}}_h,*_v)$ which sends the minimal projections to one. Next we have learnt that for all $\lambda,\mu$ with $Z_{\lambda,\mu}\neq 0$, the $\pi_{\lambda,\mu}$'s are the irreducible representations of $({\cal{Z}}_h,*_v)$ and hence the $\pi_{\lambda,\mu}\circ\Phi$'s are the irreducible representations of the $M$-$M$ fusion rule algebra. \begin{corollary} \label{commutativity} Under Assumption \ref{set-nondeg}, the $M$-$M$ fusion rule algebra is commutative if and only if $Z_{\lambda,\mu}\in\{0,1\}$ for all $\lambda,\mu\in{}_N {\cal X}_N$. \end{corollary} \begin{corollary} \label{MM-number} Under Assumption \ref{set-nondeg}, the total number of morphisms in ${}_M {\cal X}_M$ is equal to ${\rm{tr}}(Z \,{}^{{\rm{t}}}\!Z)=\sum_{\lambda,\mu\in{}_N {\cal X}_N} Z^2_{\lambda,\mu}$. \end{corollary} \subsection{The left action on $M$-$N$ sectors} \label{sec-actMN} The decomposition of $({\cal{Z}}_h,*_v)$ into simple matrix algebras is equivalent to the irreducible decomposition of the ``regular representation'' (up to multiplicities given as the dimensions) of the $M$-$M$ fusion rule algebra, i.e.\ the representation obtained by its action on itself as a vector space. There is another representation of the $M$-$M$ fusion rule algebra, namely the one obtained by its (left) action on the $M$-$N$ sectors. This is what we study in the following. We define the vector space $K$ by $K=\bigoplus_{a\in{}_N {\cal X}_M}{\rm{Hom}}({\rm{id}}_N,a\co a)$. Note that each block consists just of scalar multiples of the isometries ${\co r}_a$ but we need the explicit form of $K$. We define basis vectors $v_{\co a}\in K$ corresponding to $d_a^{-1/2}{\co r}_a$ in each block ${\rm{Hom}}({\rm{id}}_N,a\co a)$. We can display each $v_{\co a}$ graphically by a thick wire ``cap'' with label $a\in{}_N {\cal X}_M$ together with a prefactor $1/d_a$. We furnish $K$ with a Hilbert space structure by putting $\langle v_{\co a},v_{\co b} \rangle=\del ab$. For each $a\in{}_N {\cal X}_M$ we define a vector $\varrho(e_\beta)v_{\co a}$ by putting \begin{equation} \varrho(e_\beta)v_{\co a} = d_\beta \sum_b N_{\beta,\co a}^{\co b} \, v_{\co b} \,. \label{rhodef} \end{equation} We can display the right-hand side graphically as in Fig.\ \ref{picrhodef}. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(125,40) \allinethickness{1.0pt} \put(5,16){\makebox(0,0){$\displaystyle\sum_b$}} \put(20,0){\line(0,1){30}} \put(40,0){\line(0,1){30}} \put(30,30){\arc{20}{3.142}{0}} \allinethickness{2.0pt} \put(20,12.5){\line(1,0){20}} \put(32,12.5){\vector(1,0){0}} \allinethickness{0.3pt} \put(20,12.5){\arc{5}{4.712}{1.571}} \put(40,12.5){\arc{5}{1.571}{4.712}} \put(16,33){\makebox(0,0){$a$}} \put(16,4){\makebox(0,0){$b$}} \put(44,4){\makebox(0,0){$b$}} \put(30,17.5){\makebox(0,0){$\beta$}} \allinethickness{1.0pt} \put(70,16){\makebox(0,0){$=\;\displaystyle\sum_b \; \frac 1{d_b}$}} \put(105,0){\arc{20}{3.142}{0}} \put(105,30){\arc{20}{3.142}{0}} \put(105,25){\arc{20}{0}{3.142}} \put(95,25){\line(0,1){5}} \put(115,25){\line(0,1){5}} \allinethickness{2.0pt} \put(95,27.5){\line(1,0){20}} \put(107,27.5){\vector(1,0){0}} \allinethickness{0.3pt} \put(95,27.5){\arc{5}{4.712}{1.571}} \put(115,27.5){\arc{5}{1.571}{4.712}} \put(93,36){\makebox(0,0){$a$}} \put(91,4){\makebox(0,0){$b$}} \put(119,4){\makebox(0,0){$b$}} \put(105,32){\makebox(0,0){$\beta$}} \put(117,19){\makebox(0,0){$b$}} \end{picture} \end{center} \caption{The element $\varrho(e_\beta) v_{\co a} \in K$} \label{picrhodef} \end{figure} The left- and right-hand side in Fig.\ \ref{picrhodef} are the same because both sides are scalar multiples of the isometry ${\co r}_a$ in each block ${\rm{Hom}}({\rm{id}}_N,a\co a)$. The map $\varrho(e_\beta):v_{\co a}\mapsto\varrho(e_\beta) v_{\co a}$ clearly defines a linear operator on $K$ for each $\beta\in{}_M {\cal X}_M$, and we can extend the map $e_\beta\mapsto\varrho(e_\beta)$ linearly to ${\cal{Z}}_h$. Graphically, this action of ${\cal{Z}}_h$ is quite similar to the vertical product. (Note that there also appears a factor $d_a$ cancelling the $d_a^{-1}$ in the definition of $v_{\co a}$ when gluing the picture for $v_{\co a}$ on top of that for $e_\beta$.) We observe that the map $\varrho:e_\beta\mapsto\varrho(e_\beta)$ extends linearly to a representation of $({\cal{Z}}_h,*_v)$ as we can compute for $\beta,\beta'\in{}_M {\cal X}_M$ as follows: \[ \begin{array}{ll} \varrho(e_\beta) (\varrho(e_{\beta'}) v_{\co a}) &= \varrho(e_\beta) \left(d_\beta \sum_b N_{\beta',\co a}^{\co b} \, v_{\co b} \right) = d_\beta d_{\beta'} \sum_{b,c} N_{\beta,\co b}^{\co c} N_{\beta',\co a}^{\co b} v_{\co c} \\[.4em] &= d_\beta d_{\beta'} \sum_{\beta'',c} N_{\beta,\beta'}^{\beta''} N_{\beta'',\co a}^{\co c} v_{\co c} = d_\beta d_{\beta'} \sum_{\beta'',c} d_{\beta''}^{-1} N_{\beta,\beta'}^{\beta''} \varrho(e_{\beta''}) v_{\co a} \\[.4em] &= \varrho (e_\beta *_v e_{\beta'}) v_{\co a} \,, \end{array} \] where we used associativity of the sector product in the third equality. Consequently, $\varrho(q_{\lambda,\mu})$ is a projection onto a subspace, and $\varrho|_{\varrho(q_{\lambda,\mu})K}$ is a subrepresentation. \begin{lemma} \label{Kdecom} We have $K=\bigoplus_{\lambda\in{}_N {\cal X}_N}K_\lambda$, where $K_\lambda=\varrho(q_{\lambda,\lambda})K$. \end{lemma} \begin{proof} The vector $\varrho(q_{\lambda,\mu})v_{\co a}\in K$ is given graphically by the left-hand side of Fig.\ \ref{rhoqlmv}. \allinethickness{0.3pt} \begin{figure}[htb] \begin{center} \unitlength 0.7mm \begin{picture}(200,70) \allinethickness{1.0pt} \put(15,30){\makebox(0,0){$\displaystyle\sum_{b,c,d} \frac{d_b d_c}{w^2}$}} \put(37,0){\line(0,1){15}} \put(83,0){\line(0,1){15}} \put(47,15){\arc{20}{3.142}{4.712}} \put(73,15){\arc{20}{4.712}{0}} \put(47,15){\arc{10}{1.571}{4.712}} \put(73,15){\arc{10}{4.712}{1.571}} \put(47,55){\arc{20}{1.571}{3.142}} \put(73,55){\arc{20}{0}{1.571}} \put(47,55){\arc{10}{1.571}{4.712}} \put(73,55){\arc{10}{4.712}{1.571}} \put(47,10){\line(1,0){26}} \put(47,20){\line(1,0){1}} \put(52,20){\line(1,0){21}} \put(47,25){\line(1,0){1}} \put(52,25){\line(1,0){21}} \put(37,65){\line(0,-1){10}} \put(83,65){\line(0,-1){10}} \put(47,45){\line(1,0){1}} \put(52,45){\line(1,0){21}} \put(47,50){\line(1,0){1}} \put(52,50){\line(1,0){21}} \put(47,60){\line(1,0){26}} \put(42,65){\arc{10}{3.142}{4.712}} \put(78,65){\arc{10}{4.712}{0}} \put(42,70){\line(1,0){36}} \allinethickness{0.3pt} \put(50,10){\line(0,1){10}} \put(50,25){\line(0,1){20}} \put(70,10){\line(0,1){8}} \put(70,27){\line(0,1){16}} \put(50,33){\vector(0,-1){0}} \put(70,37){\vector(0,1){0}} \put(50,60){\line(0,-1){10}} \put(70,60){\line(0,-1){8}} \put(50,10){\arc{5}{3.142}{0}} \put(70,10){\arc{5}{3.142}{0}} \put(50,60){\arc{5}{0}{3.142}} \put(70,60){\arc{5}{0}{3.142}} \allinethickness{2.0pt} \put(70,22){\line(0,1){1}} \put(50,20){\line(0,1){5}} \put(70,48){\line(0,-1){1}} \put(50,50){\line(0,-1){5}} \put(33,9){\makebox(0,0){$d$}} \put(60,7){\makebox(0,0){$c$}} \put(60,16){\makebox(0,0){$b$}} \put(33,61){\makebox(0,0){$a$}} \put(60,63){\makebox(0,0){$c$}} \put(60,54){\makebox(0,0){$b$}} \put(45,35){\makebox(0,0){$\lambda$}} \put(75,34){\makebox(0,0){$\mu$}} \put(115,30){\makebox(0,0){$=\;\displaystyle\sum_{b,c,d,i,j} \;\frac{\del \lambda\mu}{w^2}$}} \allinethickness{1.0pt} \put(137,0){\line(0,1){15}} \put(183,0){\line(0,1){15}} \put(147,15){\arc{20}{3.142}{4.712}} \put(173,15){\arc{20}{4.712}{0}} \put(147,15){\arc{10}{1.571}{4.712}} \put(173,15){\arc{10}{4.712}{1.571}} \put(147,55){\arc{20}{1.571}{3.142}} \put(173,55){\arc{20}{0}{1.571}} \put(147,55){\arc{10}{1.571}{4.712}} \put(173,55){\arc{10}{4.712}{1.571}} \put(147,10){\line(1,0){26}} \put(147,20){\line(1,0){1}} \put(152,20){\line(1,0){21}} \put(147,25){\line(1,0){1}} \put(152,25){\line(1,0){21}} \put(137,65){\line(0,-1){10}} \put(183,65){\line(0,-1){10}} \put(147,45){\line(1,0){1}} \put(152,45){\line(1,0){21}} \put(147,50){\line(1,0){1}} \put(152,50){\line(1,0){21}} \put(147,60){\line(1,0){26}} \put(142,65){\arc{10}{3.142}{4.712}} \put(178,65){\arc{10}{4.712}{0}} \put(142,70){\line(1,0){36}} \allinethickness{0.3pt} \put(150,10){\line(0,1){10}} \put(170,10){\line(0,1){8}} \put(150,25){\line(0,1){2}} \put(150,45){\line(0,-1){2}} \put(155,32){\line(1,0){10}} \put(155,38){\line(1,0){10}} \put(158,32){\vector(-1,0){0}} \put(162,38){\vector(1,0){0}} \put(155,27){\arc{10}{3.142}{4.712}} \put(165,27){\arc{10}{4.712}{0}} \put(155,43){\arc{10}{1.571}{3.142}} \put(165,43){\arc{10}{0}{1.571}} \put(150,60){\line(0,-1){10}} \put(170,60){\line(0,-1){8}} \allinethickness{2.0pt} \put(170,22){\line(0,1){1}} \put(150,20){\line(0,1){5}} \put(170,48){\line(0,-1){1}} \put(150,50){\line(0,-1){5}} \put(133,9){\makebox(0,0){$d$}} \put(160,7){\makebox(0,0){$c$}} \put(160,16){\makebox(0,0){$b$}} \put(133,61){\makebox(0,0){$a$}} \put(160,63){\makebox(0,0){$c$}} \put(160,54){\makebox(0,0){$b$}} \put(145,31){\makebox(0,0){$\lambda$}} \put(175,38){\makebox(0,0){$\lambda$}} \put(150,3.5){\makebox(0,0){$t_{b,\co c}^{\lambda;i}$}} \put(170,4){\makebox(0,0){$t_{c,\co b}^{\co\lambda;i}$}} \put(150,64.5){\makebox(0,0){$(t_{b,\co c}^{\lambda;i})^*$}} \put(171,64.5){\makebox(0,0){$(t_{c,\co b}^{\co\lambda;i})^*$}} \end{picture} \end{center} \caption{The vector $\varrho(q_{\lambda,\mu})v_{\co a}\in K$} \label{rhoqlmv} \end{figure} Now note that the upper part of the diagram represents an intertwiner in ${\rm{Hom}}({\rm{id}}_N,\lambda\co\mu)$. Therefore it vanishes unless $\lambda=\mu$ and then it must be a scalar multiple of ${\co r}_\lambda$. Hence we can insert a term ${\co r}_\lambda {\co r}_\lambda^*$ which corresponds graphically to the disconnection of the wires as on the right-hand side in Fig.\ \ref{rhoqlmv} and multiplication by $d_\lambda^{-1}$. Then the factor $d_b d_c/d_\lambda$ disappears because of the normalization convention for trivalent vertices with small arcs, and we are left exactly with the right-hand side of Fig.\ \ref{rhoqlmv}. It follows in particular that $\varrho(q_{\lambda,\mu})K=0$ unless $\lambda=\mu$. The claim follows now since the vertical projectors sum up to $e_0$ and $\varrho(e_0)$ is the identity on $K$. \end{proof} We are now ready to prove the following \begin{theorem} \label{varrhodecom} The representation $\varrho$ of $({\cal{Z}}_h,*_v)$ on $K$ obtained by \erf{rhodef} is unitarily equivalent to the direct sum over the irreducible representations $\pi_{\lambda,\lambda}$: \begin{equation} \varrho \simeq \bigoplus_{\lambda\in{}_N {\cal X}_N} \pi_{\lambda,\lambda} \,. \end{equation} Consequently, the representation $\varrho\circ\Phi$ of the $M$-$M$ fusion rule algebra which is obtained by the action on the $M$-$N$ sectors arising from ${}_M {\cal X}_N$ decomposes into irreducibles as $\varrho\circ\Phi\simeq\bigoplus_\lambda \pi_{\lambda,\lambda}\circ\Phi$. \end{theorem} \begin{proof} For $b,c\in{}_N {\cal X}_M$ and isometries $t\in{\rm{Hom}}(\lambda,b\co c)$ and $s\in{\rm{Hom}}(\co\lambda,c\co b)$ we define a vector $k_{b,c,t,s}^\lambda\in K$ by the diagram in Fig.\ \ref{kbcts}. \begin{figure}[htb] \begin{center} \unitlength 0.6mm \begin{picture}(69,40) \allinethickness{1.0pt} \put(17,0){\line(0,1){15}} \put(63,0){\line(0,1){15}} \put(27,15){\arc{20}{3.142}{4.712}} \put(53,15){\arc{20}{4.712}{0}} \put(27,15){\arc{10}{1.571}{4.712}} \put(53,15){\arc{10}{4.712}{1.571}} \put(27,10){\line(1,0){26}} \put(27,20){\line(1,0){1}} \put(32,20){\line(1,0){21}} \put(27,25){\line(1,0){1}} \put(32,25){\line(1,0){21}} \allinethickness{0.3pt} \put(30,10){\line(0,1){10}} \put(30,25){\line(0,1){9}} \put(50,10){\line(0,1){8}} \put(50,27){\line(0,1){7}} \put(35,39){\line(1,0){10}} \put(38,39){\vector(-1,0){0}} \put(35,34){\arc{10}{3.142}{4.712}} \put(45,34){\arc{10}{4.712}{0}} \allinethickness{2.0pt} \put(50,22){\line(0,1){1}} \put(30,20){\line(0,1){5}} \put(67,4){\makebox(0,0){$a$}} \put(40,7){\makebox(0,0){$c$}} \put(40,16){\makebox(0,0){$b$}} \put(30,5.8){\makebox(0,0){$t$}} \put(50,5){\makebox(0,0){$s$}} \put(25,33.5){\makebox(0,0){$\lambda$}} \put(4,20){\makebox(0,0){$\displaystyle\sum_a$}} \end{picture} \end{center} \caption{The vector $k_{b,c,t,s}^\lambda\in K$} \label{kbcts} \end{figure} Using again intertwiner bases, we also put $k_\xi^\lambda=k_{b,c,t_{b,\co c}^{\lambda;i},t_{c,\co b}^{\co\lambda;j}}$ with some multi-index $\xi=(b,c,i,j)$. It follows from the right-hand side in Fig.\ \ref{rhoqlmv} that $K_\lambda\subset {\rm span} \{k_\xi^\lambda\,|\,\xi=(b,c,i,j)\}$. Conversely, we obtain by Lemma \ref{keyrel} that $\varrho(q_{\mu,\mu}) k_\xi^\lambda=0$ unless $\lambda=\mu$, hence $K_\lambda = {\rm span} \{k_\xi^\lambda\,|\,\xi=(b,c,i,j)\}$. With $\lambda=\mu$, closing the wires on the bottom and on the top on both sides of Fig.\ \ref{keyrelpict} yields \[ \langle k_\xi^\lambda,k_{\xi'}^\lambda \rangle = d_\lambda \langle \Omega_\xi^{\lambda,\lambda},\Omega_{\xi'}^{\lambda,\lambda} \rangle \,.\] Hence linear extension of $\Omega_\xi^{\lambda,\lambda}\mapsto d_\lambda^{-1/2} k_\xi^\lambda$ defines a unitary operator $U_\lambda:H_{\lambda,\lambda}\rightarrow K_\lambda$. Note that $U$ means multiplication by ${\co r}_\lambda$ from the right in each block ${\rm{Hom}}(\lambda\co\lambda,a\co a)$ and this corresponds graphically to closing the open ends of the wires $\lambda$ in Fig.\ \ref{Ombcts} and multiplying by $d_\lambda^{-1/2}$. Therefore we find \[ U \left[\pi_{\lambda,\lambda}(e_\beta) \Omega_\xi^{\lambda,\lambda} \right] = d_\lambda^{-1/2} \varrho_\lambda(e_\beta) k_\xi^\lambda = \varrho_\lambda(e_\beta) U \left[ \Omega_\xi^{\lambda,\lambda} \right] \,,\] where $\varrho_\lambda=\varrho|_{K_\lambda}$. Thus $\varrho_\lambda\simeq\pi_{\lambda,\lambda}$. \end{proof} Since the dimension of $K$ is the cardinality of ${}_N {\cal X}_M$ we immediately obtain the following \begin{corollary} \label{NM-number} Under Assumption \ref{set-nondeg}, the total number of morphisms in ${}_N {\cal X}_M$ (or, equivalently, in ${}_M {\cal X}_N$) is equal to ${\rm{tr}}(Z)=\sum_{\lambda\in{}_N {\cal X}_N} Z_{\lambda,\lambda}$. \end{corollary} \section{Conclusions and Outlook} \label{sec-concl} We have analyzed braided type III subfactors and shown that in the non-degenerate case the system of $M$-$M$ system is entirely generated by $\alpha$-induction, including in particular the subsectors of Longo's canonical endomorphism $\gamma$. We established that in that case the essential structural information about the $M$-$M$ fusion rules is encoded in the modular invariant mass matrix $Z$. Our setting applies in particular to ${\it{SU}}(n)$ loop group subfactors $\pi^0({\it{L}_I\it{SU}(n)})''\subset\pi^0({\it{L}_I\it{G}})''$ of conformal inclusions ${\it{SU}}(n)_k\subset G_1$ and $\pi_0({\it{L}_I\it{SU}(n)})''\subset\pi_0({\it{L}_I\it{SU}(n)})''\rtimes_\sigma\Bbb{Z}_m$ which were analyzed by $\alpha$-induction in \cite{BE2,BE3}. Here $\pi^0$ denotes the level $1$ vacuum representation of the loop group ${\it{LG}}$, $\pi_0$ the level $k$ representation of ${\it{LSU}(n)}$, $I\subset S^1$ is an interval, and $\sigma$ is a ``simple current''. The braiding here arises from the localized transportable endomorphisms of the net of local algebras $A(I)=\pi_0({\it{L}_I\it{SU}(n)})''$. Since it follows from Wassermann's work \cite{W2} that these endomorphisms obey the ${\it{SU}}(n)_k$ fusion rules and from the conformal spin-statistics theorem \cite{GL} that the statistics phases are given by $\omega_\lambda={\rm{e}}^{2\pi{\rm{i}} h_\lambda}$ with $h_\lambda$ denoting the ${\it{SU}}(n)_k$ conformal dimensions, it follows that the S- and T-matrices from the braiding coincide with the well-known S- and T-matrices which transform the conformal characters. Therefore Theorem \ref{generating} shows in particular that Condition 4 in Proposition 5.1 in \cite{BE3} holds in the setting of conformal inclusions, and in turn it proves Conjecture 7.1 in \cite{BE3}. It also follows that in the setting of Proposition 5.1 in \cite{BE3}, the sum of $e_\beta$ for ``marked vertices'' $[\beta]$ (the $M$-$M$ sectors arising from the positive energy representations of the ambient theory) correspond to the projections appearing in the decomposition of $\sum_{\lambda,\mu} p_\lambda^+ *_h p_\mu^-$, the ``ambichiral projector'' in Ocneanu's language. Similarly, the results of this paper also prove Conjecture 7.2 in \cite{BE3}. Theorem \ref{generating} shows in particular that there are {\sl no} counter-examples for conformal inclusions where the $M$-$M$ sectors arising from the conformal inclusion subfactor are not generated by the mixed $\alpha$-induction (cf.\ \cite{X2}). Xu made some computation in \cite{X1} (see also \cite{BE2}) to find an example with non-commutative fusion rules of ($M$-$M$) sectors generated by the image of only one ``positive'' induction for subfactors arising from conformal inclusions. By Corollary \ref{commutativity}, it is at least very easy to find examples of a non-commutative entire $M$-$M$ fusion rule algebra. The D$_4$ case mentioned in \cite[Subsection 6.1]{BE3} is one such example. In fact, the whole D$_{2n}$ series arising from simple current extension of $SU(2)_{4n-4}$ also give examples of non-commutative $M$-$M$ fusion rule algebras. Such non-commutativity for D$_{{\rm{even}}}$ has been also pointed out in the setting of \cite{O7} (though not in the context of conformal inclusions or simple current extensions). We will present the details and more analysis about ${\it{SU}}(n)_k$ loop group subfactors, including the treatment of all ${\it{SU}}(2)$ modular invariants, in a forthcoming publication \cite{BEK2}. Our treatment can now also incorporate the type II invariants which were not considered in \cite{BE2,BE3}, because we dropped the chiral locality condition which automatically forces the mass matrix $Z$ to be type I, i.e.\ block-diagonal. Let us remark that we could also have defined $Z_{\lambda,\mu}$ with exchanged $\pm$-signs in Def.\ \ref{modular1}, and this would correspond to replacing $Z$ by the transposed mass matrix ${}^{{\rm{t}}}\! Z$. It is not hard to see that all our calculations go through with ${}^{{\rm{t}}}\! Z$ as well. That means $\alpha$-induction for a (non-degenerately) braided subfactor determines actually two modular invariant mass matrices $Z$ and ${}^{{\rm{t}}}\! Z$, and it is not clear to us at present whether they can in fact be different in our general setting. (We have $Z={}^{{\rm{t}}}\! Z$ for all ${\it{SU}}(2)$ and ${\it{SU}}(3)$ modular invariants). A notion of subequivalent paragroups was introduced in \cite{K4}. Since ${}_N {\cal X}_N$ and ${}_M {\cal X}_M$ are equivalent systems of endomorphisms by definition, $\alpha$-induction produces an example of a subequivalent paragroup. That is, for $\lambda\in{}_N {\cal X}_N$, the subfactors $\alpha^\pm_\lambda(M)\subset M$ are subequivalent to $\lambda(N)\subset N$. Various examples in \cite{K4} arise from this construction. Indeed, the most fundamental example in \cite{K4} comes from the Goodman-de la Harpe-Jones subfactor \cite[Section 4.5]{GHJ} with index $3+\sqrt3$. In our current setting, this example comes from the conformal inclusion $SU(2)_{10}\subset SO(5)_1$ and shows that the two paragroups with principal graph E$_6$ are subequivalent to the paragroup with principal graph A$_{11}$. As a corollary of a rigidity theorem presented by Ocneanu in Madras in January 1997, there are only finitely many paragroups with global index below a given upper bound. This implies that for a given paragroup we have only finitely many subequivalent paragroups since their global indices are less than or equal to the global index of the given paragroup. In the context of modular invariants, a simple argument of Gannon \cite{G1} shows $\sum_{\lambda,\mu} Z_{\lambda,\mu} \le 1/S_{0,0}^2$, which in turn implies that there are only finitely many modular invariant mass matrices $Z$ for a given unitary representation of ${\it{SL}}(2;\bbZ)$, where the S-matrix satisfies the standard relations $S_{0,\lambda}\ge S_{0,0}>0$. As for a non-degenerately braided system of morphisms this bound coincides with the global index, $w=1/S_{0,0}^2$, and in view of the relations between modular invariants and subfactors elaborated in this paper, it is natural to expect that these two finiteness arguments are not completely unrelated. We consider a good understanding of the connections between these two arguments to be highly desirable. Let us finally remark that in a recent paper of Rehren \cite{R4} the embedding of left and right chiral observables in a $2D$ conformal field theory are studied. Such embeddings give rise to subfactors and in turn to coupling matrices which are invariant mass matrices if the Fourier transform matrix of the chiral fusion rules is modular. As these subfactors are quite different from ours which appear in a framework considering chiral observables only, the relation between the two approaches also calls for a coherent understanding. \vspace{0.5cm} \begin{footnotesize} \noindent{\it Acknowledgement.} Part of this work was done during visits of the third author to the University of Wales Swansea and the University of Wales Cardiff, a visit of the second author to the University of Tokyo, visits of all the three to Universit\`a di Roma ``Tor Vergata'' and visits of the first two authors to the Australian National University, Canberra. We thank R.\ Longo, L.\ Zsido, J.\ E.\ Roberts, D.\ W.\ Robinson and these institutions for their hospitality. We would like to thank S.\ Goto for showing us a preliminary manuscript of \cite{O7}, M.\ Izumi for explaining \cite{I3}, T.\ Kohno, H.\ Murakami, and T.\ Ohtsuki for helpful explanations on topological invariants, and J.\ E.\ Roberts for his comments. Y.K. thanks A.\ Ocneanu for various conversations on \cite{O7} at the Fields Institute in 1995. We acknowledge the financial support of the Australian National University, CNR (Italy), EPSRC (U.K.), the EU TMR Network in Non-Commutative Geometry, Grant-in-Aid for Scientific Research, Ministry of Education (Japan), the Kanagawa Academy of Science and Technology Research Grants, the Universit\`a di Roma ``Tor Vergata'', and the University of Wales. \end{footnotesize} \begin{footnotesize}
\section{Introduction} The problem of bulge formation is in rapid evolution: not only many scenarios, dynamical processus, formation theories have been proposed and studied, but there has been great progress in observation of bulges: on ages, metallicities, structures, etc... One can cite in particular: \begin{itemize} \item Detailed age study of individual globular clusters, with colour-magnitude diagrams of individual stars, with the high spatial resolution of HST (in the Milky Way bulge, and in Local Group galaxies) \item More extinction-free studies of bulge structure through near-infrared imaging, made recently possible at large scale (wide extragalactic surveys, COBE for the Milky Way) \item Morphological studies of galaxies at high redshift: this has been extensively developped in this meeting, and is the privileged tool to tackle galaxy evolution in situ. \end{itemize} There are very good recent reviews on the subject, namely in proceedings of the STSci workshop held in 1998 (``When and how bulges form ?'', ed. Carollo, Fergusson \& Wyse), Renzini (1999), Silk \& Bouwens (1999), Carlberg (1999) or the Annual Review on ``Galactic Bulges'' by Wyse, Gilmore \& Franx (1997). The main scenarios proposed for the fornation of bulges are: \begin{itemize} \item Monolithic formation (Eggen, Lynden-Bell \& Sandage 1962), or very early dissipative collapse at the beginning of galaxy formation. This assumes that the gas experiences violent 3D star formation, so quickly that it had no time to settle into a disk. This scenario was first proposed to explain the almost spherical old metal-poor stellar halo and the subsequent formation of disks with different thickness, that were thought to be an age sequence following the progressive gas settling. This is now proposed for the central bulge, the disk being acquired later. \item Secular dynamical evolution: the time-scale of these processes are longer than the dynamical time (i.e. secular), but however smaller than the Hubble time. They are due to the dynamical interaction between the various components, disk, bulge, halo. Gravitational instabilities, such as bars and spirals, are able to transfer efficiently angular momentum, and produce radial mass flows towards the center. Due to vertical resonances, stars in the center are elevated above the plane, and contribute to bulge formation (Combes et al. 1990). \item galaxy interactions, through merger and mass accretion. It is well known that major mergers between spirals can result in the formation of an elliptical galaxy (Toomre \& Toomre 1972). Similarly, the accretion by a spiral galaxy of a dwarf companion could contribute to the formation of a spheroid at the center, assumimg that this minor merger has not destroyed the disk. \end{itemize} In fact, all three of these main scenarii certainly occur, the question is to estimate the relative role of each of them, which is related to the time-scale of bulge formation. All these processes may be included in the general frame of hierachical galaxy formation. Disks are supposed to form through gas cooling in a dark halo. Cooling runs progressively from the center to the outer parts (case of continuous gas infall), and produces an inside-out formation of disks. Either this cooling is first violent at the center, due to the short dynamical time-scales, and the absence of a stabilising heavy stellar object, and a spheroid can form first; or secular dynamical evolution could afterwards transfer angular momentum, and bring mass to the center. In all scenarii, the speed of evolution and rate of star formation depends strongly on environment. Galaxy interactions are both responsible of the merger scenario, and also trigger or boost bar formation and secular evolution. Bulges at the center of galaxies accumulate all stars formed, and in any case they are expected to be older than the outer disk, and mainly with much more scattered properties (ages, abundances). \smallskip We will first briefly review the constraints and clues brought by observations related to bulge formation, and then examine each scenario respectively. \section{Clues from observations} Observations of our own bulge is impeded by dust extinction, confusion (crowding), contamination by foreground stars, that make data on the Milky Way very uncertain. There exist in the literature a certain number of prejudices, like the idea that ``bulges are old and metal-rich, and small versions of ellipticals'', that are not today completely confirmed: the reality is not so simple, as clearly reviewed by Wyse et al (1997). \subsection{ Metallicity} In the Milky Way, the mean metallicity of the bulge is the same as that of the disk in the solar neighborhood, but in contrast to the disk, the bulge has a very wide scatter. This characteristic allows the metallicity distribution of the bulge to be explained by a closed box chemical model, contrary to the disk: the latter has a very narrow distribution of metallicities, that gives rise to the well-known G-dwarf problem (its solution requires other processes, like gas infall, etc..). In the Andromeda bulge also, super metal-rich globular clusters are the exception (Jablonka et al 1998). In the Milky Way bulge, there is no correlation between age, color, abundance and kinematics, which could have given clues about the origins (e.g. Rich 1997). There does not seem to have been a starburst, since there is no excess of $\alpha$ elements (Mc William \& Rich 1994). \subsection {Age} As for ages, however, we should not make the confusion: {\it the bulge has not necessarily the age of its stars}. In particular the bulge could have formed recently from old stars. It is true namely if the bulge has formed from bars. Thanks to the clear overall structure of the MW given by COBE-DIRBE, we know directly now that a bar is present, while it has long be assumed from gas kinematics only (e.g. Peters 1975). The peanut/boxy bulge of the MW has asymmetries due to a bar seen in perspective (e.g. Blitz \& Spergel 1991, Zhao et al 1996). It is even difficult to distinguish what is bar and what is bulge (Kuijken 1996). Since strong bars are only a transient phase (see below) it is easy to extrapolate how the stars presently in the bar will form the bulge. \begin{figure} \psfig{figure=bulge99_f1.ps,width=12cm,bbllx=1cm,bblly=5cm,bburx=19cm,bbury=20cm,angle=0} \caption{The $B-K$ color in surface-brightness of the central disk, versus that of the bulge, from de Jong (1996). The various symbols indicate morphological types, as described at the bottom left. The dashed line means equality. } \label{fig1} \end{figure} If the bulges are thought to be older than disks, it might be due to the implicit average over the whole disk. In fact, when bulges and inner disks are compared, they have the same integrated colors (Peletier \& Balcells 1996, de Jong 1996, see fig \ref{fig1}). But there exist clear radial gradients of colors and metallicities. Spiral galaxies become bluer with increasing radius. Moreover, colors correlate well with surface brightness. These colors and their gradients can best be explained with history of star formation, and dust reddening is not dominant, as shown by de Jong (1996): outer parts are younger than central regions. These observations support an inside-out galaxy formation. Alternatively, dynamical secular evolution can also account for these observations, through angular momentum transfer (cf Tsujimoto et al 1995). \subsection{ Are bulges similar or not to giant ellipticals ?} Bulges are spheroids that follow the same fundamental plane as elliptical galaxies. They also follow the same luminosity-metallicity relation (Jablonka et al 1996). But they are oblate rotaters (flattened by rotation), while giant ellipticals are not rotating. However, low-luminosity ellipticals also rotate (Davies et al 1983). There might be a continuity between bulges, lenticulars and ellipticals. Most of the latters have been observed with compact disks, e.g. Rix \& White 1992, Rix et al. 1999). There is also continuity in the light profiles. They can be fitted as an exponential of $r^{1/n}$, while $n$ is a function of luminosity ($n=$4 for Ellipticals, $\infty$ for Bulges, Andredakis et al 1995). \subsection{ High-redshift galaxies } From the CFRS and LDSS, surveying galaxies at redshifts below 1, there is very little evolution of the luminosity function of red galaxies (Lilly et al. 1995, 1998). Their disk scale-length stays constant, up to $z=1$ (which could mean either no-evolution, or stationary merging). On the contrary, there is substantial evolution in the blue objects. Does it mean that a large fraction of bulges/spheroids are already formed? (or mean age of stars $\propto$ age of Universe). From HST images of the CFRS galaxies, Shade et al (1995, 96) conclude that red galaxies have large bulge-to-disks (B/D) ratios, and that most blue galaxies are interacting (also forming bulges, nucleated galaxies). Steidel et al (1996) from their sample of z $\sim$ 3 galaxies find in average objects smaller than today (see also Bouwens et al 1998). If violent starbursts are forming at high redshift, like the ultra-luminous IR galaxies observed nearby, they should dominate the sub-mm surveys. Already a large fraction of the cosmic IR background has been resolved into sources (Hughes et al. 1998). If these objects are forming the spheroids (bulges and ellipticals), then their epoch of formation is relatively recent $z < 2$ (Lilly et al. 1999). From the high spatial resolution of HST, it becomes now possible to study the morphology of high-redshift galaxies, and address the evolution of the Hubble sequence. At least, it is possible to determine the concentration (C) and asymmetry (A) to classify galaxies (Abraham et al 1996, 99). If there is a consensus among the various studies, it is on the considerable increase of perturbed and interacting galaxies ($\sim$ 40\% objects interacting). But on the concentration, or bulge-to-disk ratios, there is no unanimity. Abraham et al (1998) do not find evolution in the B/D ratio, and tend to favour monolithic bulge formation. Marleau \& Simard (1998) find on the contrary many more disk-like faint galaxies in the Hubble Deep Fields, i.e. that there are less objects with high bulge-to-disk ratios at high redshift (with respect to $z=0$). \section{Monolithic dissipative formation} This kind of violent and intense star formation is different to what is seen today in galactic disks: it occurs in three-dimension (Elmegreen 1999). In the deep potential well of a giant galaxy center, almost entirely due to the dark halo, there cannot be SF self-regulation by blow-out (supernovae, winds, pressure) as in disks or dwarfs today (Meurer et al 1997). Most of the time, in the ultra-luminous galaxies of the nearby universe, the starburst occurs in nuclear rings or disks (Downes \& Solomon 1998). This could be different in the early universe, since there are no heavy disks formed. The threshold for SF in 3D can be estimated from the virial theorem (Spitzer 1942) as a critical central volumic gas density $\rho_c = \rho_{vir} /(1+\beta/2)$ for gaseous potential $\beta GM_{gas}/R$. The star formation rate SFR is then $SFR \propto \rho_{vir}/t_{dyn} \propto \rho_{vir}^{3/2}$ (an equivalent of Schmidt law). In localised dense regions, the SFR could be very large, and an extremely clumpy distribution of stars is expected. Simulations illustrating this process have been done by Noguchi (1998). \section{Secular evolution} A wide series of N-body simulations, supported by observations, have established the various phases of this secular dynamical evolution (e.g. Sellwood \& Wilkinson 1993, Pfenniger 1993, Martinet 1995, Buta \& Combes 1996). They can be described as: \begin{itemize} \item Gravitational instabilities spontaneously form bars, spirals, which transfer angular-momentum through the disk. Both stars and gas lose momentum to the wave. Radial gas flows in particular produce large central mass concentration. \item Due to vertical resonances, the bar thickens in the center (box or peanut-shape, Combes et al 1990), and contributes to bulge formation (see fig \ref{fig2}). This scenario is compatible with the observed correlation between scale-lengths of bulges and disks (Courteau et al 1996). \item Sufficient mass accumulation in the center (1-5\% of the total disk mass) perturbs the rotation curve, changes the orbit precession rates, and destroy the bar, by creating perpendicular orbits $x2$. The galaxy has now become a hot stable system, with a central spheroid. \item Through gas infall, the disk can become unstable again and reform a bar (with a different pattern speed) \end{itemize} \begin{figure} \psfig{figure=bulge99_f2.ps,width=17cm,bbllx=15mm,bblly=5mm,bburx=10cm,bbury=27cm,angle=-90} \caption{Edge-on projections of a bar obtained in N-body simulations, for three orientations of the bar with respect to the line of sight (90, 45 and 0$^\circ$).} \label{fig2} \end{figure} \section{Accretion and mergers} Ellipticals can form by the mergers of two spirals (e.g. NGC 7252 prototype); but the most general case is the formation by the merger of several smaller objects. That ellipticals have accreted several smaller companions during their life is supported by the frequency of shells (as much as 50\% according to Schweizer \& Seitzer 1988). Bulges could similarly be the result of minor mergers; even today, there exist numerous companions around giant spirals. In hierachical cosmological models, it is easy to compute analytically the history of dark halo formation, and their merging rate (Kauffmann et al 1993, Baugh et al 1996). But the fate of the dissipative visible matter is still not well-known. According to some star formation and feedback parameters that fit the colour-magnitude relation of ellipticals in clusters, Kauffmann \& Charlot (1998) have shown that massive galaxies must have formed continuously, and must have assembled only recently, in order to fit observed redshift distributions of K-band luminosity at $z=1$. If the bulge comes from accretion, observations of our own MW put constraints on the metallicity of the objects accreted (they should be relatively high). This put strong limits on the fraction of the bulge that has been accreted recently (Unavane et al 1996). \section{Combination of all scenarii} In a hierachical scenario, there are many uncertainties about the physical processes controlling the baryonic matter. Feedback must prevent efficient conversion of gas into stars, to represent observations. There remains much freedom to determine it more quantitatively, as well as the distribution of angular momentum, etc... The simplest recipe is to assume that the angular momentum of the matter is not redistributed, and that the gas cools at the radius where the cooling time becomes shorter than the dynamical time (Mo et al 1998; van den Bosch 1999). This represents gas infall as cooling flows in dark haloes, and the formation of disks as an inside-out process. There are however many hypotheses possible: for instance, that the mass of the disk is a fixed fraction of the halo mass, or that the angular momentum of the disk is a fixed fraction of the halo angular momentum; also that the disk is exponential in shape, and is stable (which gives a condition on specific momentum). The efficiency of galaxy formation ($\epsilon_{gf}$) is also a free parameter, as a function of redshift; it could vary proportionally to the Hubble constant, to ensure the Tully-Fisher relation (van den Bosch 1999). The specific angular momentum of the galaxies are supposed to be acquired by tides (Fall \& Efstathiou 1980). These scenarios, coupled with sufficient feedback (due to star-formation) can explain high-surface brightness galaxies (HSB) and even LSB, if a range of angular momentum is assumed (Dalcanton et al 1997). The following constraints can be satisfied: 1) the Tully-Fisher scaling relation; 2) the density-morphology relation (Dressler 1980); 3) the surface-density/size relation $\mu$ - R$_d$ (generalisation of Kormendy 1977). A general feature found by van den Bosch (1999), whatever the cooling/feedback hypotheses, is that the present observed disks lie very close to their stability limit (with respect to strong gravitational instabilities), which suggests a self-regulated formation, through dynamical processes: the disk alone is unstable until a sufficient bulge stabilises it. The gas infall could then alternatively favor bulge or disk formation according to the bulge-to-disk mass ratio it encounters at a given time. This could explain the observed coupling between bulges and disks. \begin{figure} \psfig{figure=bulge99_f3.ps,width=12cm,bbllx=1cm,bblly=45mm,bburx=17cm,bbury=24cm,angle=0} \caption{ Comparison of the bulge-to-total ratios (B/T) between observations (histograms) and model predictions (lines), from Bouwens et al (1999). At $z=0$, the data are from Peletier \& Balcells (1996, PB) and de Jong (1996, DJ). At high $z$ from Schade et al (1996) and Lilly et al (1996). In all frames, the models are: {\bf I} = secular evolution; {\bf II} = simultaneous formation; {\bf III} = early bulge formation. } \label{fig3} \end{figure} Bouwens et al. (1999) have recently confronted several bulge formation scenarii to the observations. At $z=0$, it appears almost impossible to distinguish between early and late bulge formation (cf fig \ref{fig3}). At high redshift, it becomes possible, essentially because a large fraction of the star formation occurs at the observed $z$. For instance, in fig \ref{fig3} at top-left, the model of simultaneous formation (of disk and bulge) departs from the two others, since star-formation makes the bulge momentarily brighter. Could the star formation history (SFH) constrain the period of formation of bulges and ellipticals? Since spheroids contain 30\% (Schechter \& Dressler 1987), or 66\% (Fukugita et al 1998) of the stars in the Universe, they cannot form too early, according to the Madau et al. (1996) SFH curve. However, this constraint is only a lower-limit on the formation time, since the spheroids could be recently formed from old stars. \section{Conclusions} Bulges are very heterogeneous structures, with large scatter in colors, metallicities and ages. At large luminosities, bulges become similar to elliptical galaxies, at low luminosities, they resemble more spiral disks (or bars). There exist objects with a whole range of properties ensuring continuity between these two extremes, suggesting evolution or the combination of several processes. Bulge formation is certainly a combination between three main scenarii: a fraction of bulges could have formed early (at first collapse); then secular dynamical evolution enrich them; in parallel, according to environment, accretion and minor mergers contribute to raise their mass. Big spheroids (SOs, Es) can only form through major/minor mergers. Present disks have been (re)-formed recently. To be more quantitative, and precise the time-scale of bulge formation, a privileged solution is to observe and classify the morphology of galaxies as a function of redshift, together with detailed balance of the comoving volumic density of galaxies of each type. This is necessary to avoid confusion between a no-evolution model, at a given type, or a stationary evolution flow along the Hubble sequence. \begin{moriondbib} \bibitem{} Abraham R., Tanvir N.R., Santiago B.X. et al. 1996, \mnras {279}{L47} \bibitem{} Abraham R., Ellis R.S., Fabian A.C., Tanvir N.R., Glazebrook K. 1999, \mnras {303}{641} \bibitem{} Andredakis Y.C., Peletier R.F., Balcells M., 1995, \mnras {275} {874} \bibitem{} Baugh C.M., Cole S., Frenk C.S., 1996, \mnras {283}{1361} \bibitem{} Blitz L., Spergel D.N., 1991, \apj {379}{631} \bibitem{} Bouwens R.J., Cayon L., Silk J., 1999, {\it preprint} (astro-ph/9812193) \bibitem{} Bouwens R.J., Broadhurst T., Silk J. 1998, \apj {506} {557} \bibitem{} Buta R., Combes F. 1996, {\it Fundamental of Cosmic Physics} {\bf 17}, 95-282 \bibitem{} Carlberg R.G., 1999, in "When and How do Bulges Form and Evolve?", ed. by C.M. Carollo, H.C. Ferguson \& R.F.G. Wyse (Cambridge University Press), (astro-ph/9903373) \bibitem{} Combes F., Debbasch F., Friedli D., Pfenniger D., 1990, \aa {233} {82} \bibitem{} Courteau S., de Jong R.S., Broeils A.H., 1996, \apj {457}{73} \bibitem{} Dalcanton J.J., Spergel D.N., Summers F.J., 1997, \apj {482}{659} \bibitem{} Davies R.L., Efstathiou G., Fall S.M., Illingworth G., Schechter P.L., 1983, \apj {266} {41} \bibitem{} de Jong R.S., 1996, \aa {313} {377} \bibitem{} Downes D., Solomon P.M., 1998, \apj {507}{615} \bibitem{} Dressler A., 1980, \apj {236}{351} \bibitem{} Eggen O.J., Lynden-Bell D., Sandage A., 1962, \apj {136} {748} \bibitem{} Elmegreen B.G., 1999, \apj {517} {May 20} \bibitem{} Fall S.M., Efstathiou G., 1980, \mnras {193}{189} \bibitem{} Fukugita M., Hogan C.J., Peebles P.J.E., 1998, \apj {503}{518} \bibitem{} Hughes D.H., Serjeant S., Dunlop J. et al.: 1998, \nat {394} {241} \bibitem{} Jablonka P., Martin P., Arimoto N., 1996, \aj {112} {1415} \bibitem{} Jablonka P., Bica, E., Bonatto C. et al., 1998, \aa {335} {867} \bibitem{} Kauffmann G., Charlot S., 1998, \mnras {297} {L23} \bibitem{} Kauffmann G., White S.D.M., Guiderdoni B., 1993, \mnras {264}{201} \bibitem{} Kormendy J., 1977, \apj {218} {333} \bibitem{} Kuijken K. 1996, in {\it Unsolved Problems of the Milky Way}, IAU Symp. 169, ed. L. Blitz \& P. Teuben, p. 71 (Kluwer) \bibitem{} Lilly S.J., Tresse L., Hammer F., Crampton D., LeFevre O., 1995, \apj {455}{108} \bibitem{} Lilly S.J., LeFevre O., Hammer F., Crampton D., 1996, \apj {460}{L1} \bibitem{} Lilly S.J., Shade D., Ellis R. et al., 1998, \apj {500}{75} \bibitem{} Lilly S.J., Eales S.A., Gear W.K. et al. 1999, in "When and How do Bulges Form and Evolve?", ed. by C.M. Carollo, H.C. Ferguson \& R.F.G. Wyse (Cambridge University Press), (astro-ph/9903157) \bibitem{} Madau P., Ferguson H.C., Dickinson M.E. et al. 1996, \mnras {283}{1388} \bibitem{} Marleau F.R., Simard L., 1998, \apj {507}{585} \bibitem{} Martinet L., 1995, {\it Fundamental of Cosmic Physics} {\bf 15}, 341-440 \bibitem{} McWilliam A., Rich R.M., 1994, \apjs {91}{749} \bibitem{} Meurer G.R., Heckman T.M., Lehnert M.D., Leitherer C., Lowenthal J., 1997, \aj {114} {54} \bibitem{} Mo H.J., Mao S., White S.D.M., 1998, \mnras {295}{319} \bibitem{} Noguchi M., 1999, {\it preprint} (astro-ph/9806355) \bibitem{} Peletier R.F., Balcells M., 1996, \aj {111} {2238} \bibitem{} Peters W.L., 1975, \apj {196} {617} \bibitem{} Pfenniger D., 1993, in {\it Physics of Nearby Galaxies: Nature or Nurture}, ed. T.X. Thuan, C. Balkowski, J. Tran Thanh Van, Ed. Frontieres, p. 519 \bibitem{} Renzini A., 1999, in "When and How do Bulges Form and Evolve?", ed. by C.M. Carollo, H.C. Ferguson \& R.F.G. Wyse (Cambridge University Press), (astro-ph/9902108) \bibitem{} Rich M., 1997, in {\it ``The Central Regions of the Galaxy and Galaxies''}, IAU Symp. 184, Kyoto (Kluwer Pub) \bibitem{} Rix H-W., White S.D.M., 1992, \mnras {254} {389} \bibitem{} Rix H-W., Carollo C.M., Freeman K., 1999, \apj {513} {L25} \bibitem{} Schade D., Lilly S.J., Crampton D. et al., 1995, \apj {451}{L1} \bibitem{} Schade D., Crampton D., Hammer F., LeFevre O., Lilly S.J., 1996, \mnras {278}{95} \bibitem{} Schechter P.L., Dressler A., 1987, \aj {94}{563} \bibitem{} Schweizer F., Seitzer P. 1988, \apj {328}{88} \bibitem{} Sellwood J.A., Wilkinson A., 1993, {\it Rep. Prog. Phys.} {\bf 56}, 173 \bibitem{} Silk, J., Bouwens R.J., 1999, in {\it Proceedings of ``Galaxy Evolution: Connecting the distant Universe with the local fossil record"}, Rencontres de Meudon, 21-25 sep 98, ed. M. Spite, Kluwer (astro-ph/9812057) \bibitem{} Spitzer L., 1942, \apj {95}{329} \bibitem{} Steidel C.C., Giavalisco M., Dickinson M., Adelberger K.L. 1996, \aj {112} {352} \bibitem{} Toomre A., Toomre J., 1972, \apj {178} {623} \bibitem{} Tsujimoto T., Yoshii Y., Nomoto K., Shigeyama T., 1995, \aa {302} {704} \bibitem{} Unavane M., Wyse R.F.G., Gilmore G., 1996, \mnras {278}{727} \bibitem{} van den Bosch F.C., 1999, \apj {507} {601} \bibitem{} Wyse R.F.G., Gilmore G., Franx M., 1997, {\it Ann. Rev. Astron. Astrophys.} {\bf 35}, 637 \bibitem{} Zhao H., Rich R.M., Spergel D.N., 1996 \mnras {282} {175} \end{moriondbib} \vfill \end{document}
\section{INTRODUCTION} NGC 2808 (C0911-646) is a prominent southern galactic globular cluster situated at low galactic latitude $(l=282.^{\circ }19,b=-11.^{\circ }25).$ \ It was first noticed by Harris (1974, 1975) that the horizontal branch (HB) morphology is very unusual, with the red HB (RHB) stars clearly separate from a group of very blue HB (BHB) stars. In a follow-up study (Harris 1978) the photometry was placed on a more secure basis and he suggested that the most plausible explanation for the two groups of stars was one whereby mass segregation of some $0.05-0.1M_{\odot }$ had occurred. His observations were confirmed in subsequent color magnitude diagram (CMD) studies by Ferraro et al. (1990) and by Byun \& Lee (1993). \ Gratton \& Ortolani (1986) and Buonanno, Corsi, \& Fusi Pecci (1989) extended the CMD to below the main sequence (MS) turn-off. \ A multi-color study by Alcaino et al. (1990) (hereafter A90) also reached below the MS but did not include HB stars; all these deep studies used small-format CCDs and single fields thus few evolved stars were included in their CMDs. \ Clement \& Hazen (1989) found only two RR Lyrae variables despite the evident richness of the total HB population. Sosin et al. (1997) (hereafter S97) used the WFPC2 on the Hubble Space Telescope to prepare a deep CMD for the center of NGC 2808 in the $B$ and $V$ bands. They also observed with an ultraviolet (F218W) filter, providing for the first time good color and hence temperature resolution for the BHB stars. The BHB was found to extend to far fainter magnitudes in the $V,B-V$ CMD than was previously known, well below the level of the MS turn-off. \ This long blue tail has two extra gaps, which are narrow and well-defined, their width corresponding to $\sim 0.01M_{\odot }$ from comparison with HB evolutionary models (Dorman, Rood \& O'Connell 1993). They discuss possible explanations for the HB multi-modality, which include mass loss, composition variations, and dynamical effects, but conclude that none of these explanations are completely satisfactory. \ This discussion has been extended \ by Ferraro et al. (1998)\ to other clusters that also have extended blue HB\ . \ Finally, Ferraro et al. (1997) (hereafter F97) present a $V,V-I$ CMD of a 4x4 arcmin region near NGC 2808, which reaches to $V\sim 23.5$ and demonstrate that the MS has dispersion greater on the redward side of the ridge-line than on the blue side. This is interpreted as being due to a large fraction ($\sim 24\%$) of binaries. \ The CMD morphology exhibited by clusters such as NGC 2808 challenges our understanding of the advanced stages of the evolution of low-mass stars, and how much the dense stellar environment of globular clusters affects these processes. \ \ As the extreme example, NGC 2808 is especially important. \ The present observations complement those made with HST WFPC2, by covering the whole cluster field at a higher photometric accuracy than that achieved by earlier ground-based studies. The faintest magnitude in the less crowded regions of the cluster is deeper than the HST photometry, however\ interpretation of the CMD here is complicated by the crowding, by the significant contamination due to field stars, and by variable extinction. \ \section{OBSERVATIONS} A series of CCD images centered on NGC 2808 was taken on the night of 1996 December 12, using the 4-m Blanco telescope at prime-focus. They consisted of $V$ band (6 x 5s, 5 x 50s, 5 x 300s) and $B$ band (4 x 10s, 4 x 60s, 4 x 400s) integrations using SITe 2048 CCD \#4, and filters from set \#1. The CCD was read out through all four amplifiers simultaneously with an Arcon controller which delivered a read time of 35s, thus observing efficiency was over 75\% despite the many short exposures. The 6-element prime focus corrector, which includes prisms providing correction for atmospheric dispersion, produces excellent quality images over an area several times greater than the 15 arcmin square field provided by the CCD. The pixel size of 24 microns corresponds to 0.432 arcsec and produces rather poorly sampled images even in the worse than median seeing of this night (image fwhm 1.1 arcsec). Additionally, the CCD is not flat, however to first order the curvature was compensated by use of a weak negative lens as dewar window. The particular lens installed slightly under-corrected the curvature, causing a slow degradation in focus as a function of field angle. The night was photometric. NGC 2808 was also observed on 1996 December 17 with the CTIO 0.9-m telescope and Tektronix 2048 \#3 CCD, through 3 x 3 inch $B$ and $V$ filters (set Tek 2). This CCD was also read out through all four amplifiers with an Arcon controller. Three sets of exposures were taken with differing integration times (30s, 100s and 300s in $V$; 50s, 150s and 450s in $B$), these frames were used for checking the photometric zeropoints, as described in detail below. Seeing was 1.0-1.2 arcsec and the night was photometric. \ Pixel size at the 0.9-m telescope is 0.40 arcsec. The frames were processed in the usual way, with overscan subtraction and trimming, followed by subtraction of a zero level exposure and removal of the instrument signature by division using twilight sky flat-field exposures. The zero and flat calibration frames were each made up from five individual exposures, combined so as to remove stars from the flats and cosmic rays from both flats and zeroes. \ Tests at the telescope using the illuminated dome ``white-spot'' were used to measure the CCD linearity and determine the shutter characteristics. The CCDs were measured to be linear to better than 0.5\% (laboratory tests demonstrate linearity at the 0.1\% level) over their operating range. Differential non-linearity between different CCD amplifiers is much easier to measure, and is at the 0.1\% level or better for both the CCDs used here. \ The shutters produce offsets of +65 ms (0.9-m) and +50 ms (4-m) at their center which must be added to the nominal exposure time, and a gradient such that by the corners of the CCDs the offset is near zero. A mean offset was added to all exposure times, but since all exposures were longer than 5 seconds (for which the error induced amounts to $\pm 0.5\%$) the gradient was disregarded. \section{DATA\ REDUCTIONS} The individual exposures from the 4-m data-set were moved onto common centers and combined to produce short, medium and long exposures in each color. Integration times for the short exposures did not saturate stars at the tip of the cluster red giant branch (RGB) and the resolution is sufficient that bright RGB stars can be measured to the cluster center. However this is a dense cluster, and on the longer exposures the cluster center is hopelessly saturated, for these the central regions were blobbed out with all pixels within a radius of 150 arcsec of the cluster center being set to invalid. \ Reductions then proceeded using DAOPHOT II and ALLSTAR (Stetson 1987, 1995), with the exception that a stand-alone program was written to more critically identify PSF stars than the PICK routine in DAOPHOT. Due to the slight variation in image quality as a function of radius, the PSF used was a linearly varying Moffat function. Between 25000 and 40000 stars were measured on each frame, following two passes through the FIND algorithm. Subtracted frames were clean, in particular the bright star residuals did not differ appreciably with position, demonstrating that the linearly varying PSF was correctly accounting for the spatial variations in image shape. \ Stars were carefully matched between $V$ and $B$ frames, which included accounting for a slight (approximately 0.05\%) difference in scale. The 0.9-m reductions were carried out in almost identical manner to those for the 4-m. The Moffat-function PSF chosen was allowed to vary quadratically as a function of radius, since this telescope is a classical Cassegrain and coma is obtrusive by the corners of the CCD. \ Stars further than 1000 pixels from the CCD center were not measured because of this degradation in image quality. \section{CALIBRATIONS} The 4-m observations used the standard 4-inch square PFCCD $B$ and $V$ filters, for which color terms are accurately known. These were able to be checked, and extinction coefficients measured, by observation of the Landolt (1992) fields SA 98 and PG 1047 prior to the NGC 2808 observations, and the PG 0918 field subsequently. The PG 1047 field was at airmass two, the others at airmass similar to the cluster. Each field contains several stars with very accurate photometry, and a total of 30 stars were measured here, using star aperture and sky annuli radii of 14, 18 and 28 pixels respectively. The fitted color equations are\newline $V=v+0.019(b-v)$\newline $B-V=0.901(b-v)$\newline with extinction coefficients $K_{V}=0.13$ and $K_{B}=0.26$. Standard errors of the $V$ and $B-V$ zeropoints are 0.002 and 0.003 mag. respectively. Considering now the 0.9-m data-set, a total of 75 standards from Landolt (1992) were measured in these two filters, and also with I filter number 31. The standards were measured using digital aperture photometry, with star aperture and sky annuli of 14, 18 and 28 pixels respectively. The fitted color equations are\newline $V=v+0.014(b-v)$\newline $B-V=0.904(b-v)$\newline $V-I=1.008(v-i)$\newline with extinction coefficients $K_{V}=0.111$, $K_{B}=0.215$, and $K_{I}=0.044$% . Standard errors of the zeropoints are 0.002 mag. in each of $V,B-V$ and $% V-I$. A discussion of the same telescope-filter-CCD combination is given by Walker (1995 ), from more extensive standard star observations than here. A conclusion from that study is that errors arising from ignoring small non-linearities in these color equations are negligible for $V$ and $V-I$, while for $B-V$ the non-linearities amount to $\pm 0.01$ mag or less for $% -0.3<B-V<1.8$. It was further supposed from the similarity of the 4-m and 0.9-m CCDs and filter sets that a similar conclusion could be made for the 4-m data. There is no evidence from the present data sets that color equation non-linearities exceed these very small levels, and as such they will be ignored. \ A caveat is that there are no Landolt standards observed here with colors in the range $-0.2<B-V<0.15$ and we have to assume that the color equation is well-behaved in this region. \ This is unfortunately precisely the color range of the BHB\ stars. \ With this assumption, we conclude that the zeropoints and color equations determined for observations on these two nights should allow transformation to the standard system to an accuracy of $\sim 0.01$ mag in both $V$ and $B-V$. \subsection{NGC 2808 Local Standards} Local standards were chosen on the 0.9m frames, and on the short-exposure 4-m frames, using the criteria of having no significant contaminating stars within 8 arcsec radius, and be bright but not saturated. Most of these stars are of necessity several arcmin distant from the cluster center. However we also included stars that had been measured by A90 which are rather closer to the cluster center. This included six stars with photoelectric photometry from Alcaino \& Liller (1986). On the short 4-m frames two of these stars are saturated, for the remaining four we find that in the mean, in the sense Walker-Alcaino\newline $\Delta V=0.01\pm 0.02$\newline $\Delta (B-V)=0.01\pm 0.01$\newline Similar results are found from the 0.9-m data, \newline $\Delta V=-0.01\pm 0.02$ (n=6 short), $-0.01\pm 0.03$ (n=4 med), $-0.01\pm 0.04$ (n=3 long)\newline $\Delta (B-V)=-0.01\pm 0.01$ (n=6), $-0.01\pm 0.01$ (n=4), $0.00\pm 0.02$ (n=3).\newline These results are very satisfactory, however there is a strong caveat. High contrast displays of the NGC 2808 frames clearly show that there are many faint stars within the r=14 pixel star aperture for the stars as close to the cluster center as the A90 standards. Thus despite the good agreement for these stars, aperture photometry through relatively large apertures may well be systematically incorrect. \ The new local standards set up here (Table 1) are all more distant from the cluster and are consequently much less crowded than the A90 stars. Tests were made by subtracting the psf-fitted faint stars adjacent to the local standards, then comparing photometry through a range of apertures for these frames and the originals. This showed that the new local standards have no discernible mean zero-point offset ($\ll 0.01$ mag) in $V$ and $B-V$ when photometry with and without faint companions are compared. However offsets for the A90 stars are typically a few 0.01 mag when this test is performed. An extensive set of photoelectric standards were provided by Harris (1978) in his major study of NGC 2808. \ Most of these stars are too distant to be on the CCD frames here. Nine stars, with $V$ magnitudes between 13.5 and 17.5 do overlap. Excluding one very discrepant star, the mean zeropoint difference between the psf photometry here and Harris' photoelectric photometry is, in the sense Walker-Harris,\newline $\Delta V=-0.01\pm 0.02$ (se), (n=8) \newline $\Delta (B-V)=0.01\pm 0.02$ (se). \newline We made two further aperture photometry comparisons. Firstly, we compared photometry for approximately 20 stars in the magnitude range $V=15-18$ mag that were measured with good signal to noise on all the 0.9-m frames. No systematic differences are found between the results from the short, medium and long exposures, and indeed most individual comparisons agree to within 0.01-0.02 mag per star. Secondly, we also compared aperture photometry for the same 20 stars between the 4-m short exposures and the 0.9-m long exposures. \ The mean differences are $\ 0.00\pm 0.01$ in both $V$ and $B-V$% . \ We calculate the formal error by adding in quadrature the mean errors for the calibrations (linearity, flat-fielding, shutter), the errors in color equations, zeropoints and extinction, and finally the errors in the system of local standards. The latter group of errors are always difficult to estimate and are usually dominated by systematic effects. Formally the errors in $V$ and $B-V$ are little more than 0.01 mag, but we will double these and assume that the NGC 2808 photometric system is accurate to $\pm 0.02$ mag\ in each of $V$ and $B-V$ for the purposes of the following discussion. \ The short, medium and long exposure frames were tied together via photometry of many (typically several hundred) overlap stars. \subsection{\protect\bigskip Comparison of HB\ Photometry, and Variable Stars% } \ At this stage we checked all available blue HB photometry of NGC 2808, since (see below) accurate photometry of the HB components is critical when photometrically determining the cluster reddening. \ In particular, we looked at the position of the RHB clump, and the color difference between the RHB clump and the BHB at $V=V(HB)+1.0$ to test whether the color differences between the various studies considered of a zeropoint shift only, or whether they also differ in scale. \ This comparison is listed in Table 2. \ Apparently only the Byun \& Lee (1993) photometry, which has never been published in detail, shows both scale and zeropoint differences when compared to the present results. The other studies show only zeropoint shifts to within the errors. Unfortunately these zeropoint shifts are rather large. \ Applying shifts to the Ferraro et al. (1990) photometry of $\Delta V=+0.11,\Delta (B-V)=-0.05$ mags produces an excellent match for the whole CMD, including the MS, with the exception that the brightest RGB stars are 0.02-0.03 mag redder here. \ For the S97 photometry, shifts of $\Delta V=0,\Delta (B-V)=+0.05$ mag are required. \ In this case, after these shifts are applied agreement between the two CMD's is excellent everywhere, including the brightest RGB stars. \ The only search for variable stars in NGC 2808 is that by Clement \& Hazen (1989), who from B-band photographic photometry identified two RR\ Lyrae variables (V12 with period 0.30577705 days and\ V6 with period 0.5389687 days), a Cepheid (V10 with period 1.76528 days) and two longer period variables (V1, V11). \ The latter two stars are confirmed to be very red, with photometry here of \ $V=13.10,B-V=1.85$ and $V=15.11,B-V=1.27$ respectively. \ \ V1 is at the tip of the giant branch, whereas V11 is well down the RGB. \ For V11, identification is not secure from the small-scale Clement \& Hazen (1989) chart, and there is a very red star with $% V=14.64,B-V=1.74$ only 3 arcsec N which may be the true variable. \ This latter star lies well to the red of the RGB and is probably a field star. \ There are a few stars which by virtue of their position in the CMD are RR\ Lyrae candidates, but they are not numerous by comparison to the total HB population. \section{\protect\bigskip METALLICITY\ AND REDDENING} Ferraro et al. (1990) provide a list of metallicity and reddening estimates for NGC 2808. The range of $[Fe/H]$ is from -0.96 to -1.48, and of $E(B-V)$ from 0.21 to 0.34. These are not all strictly comparable, and the widely used compilation by Zinn (1985) lists $[Fe/H]=-1.37\pm 0.09$ and $% E(B-V)=0.22 $. Rutledge et al. (1997a,b) measured the Calcium triplet feature for many galactic globular clusters; for NGC 2808 their mean value is based on 20 stars and on the Zinn scale is $[Fe/H]=-1.36\pm 0.05$, in excellent agreement with Zinn (1985). The situation for the NGC\ 2808 reddening is not very satisfactory, with a wide range of estimates in the recent literature. \ A90, via Vandenberg and Bell (1985) isochrones, found $E(B-V)=0.28$ at an adopted $[Fe/H]=-1.27$, while S97, using the Holtzman et al. (1995) filter zeropoints and color equations, find that they can only simultaneous fit their UV and visual CMD's with relatively small values for the reddening, between $E(B-V)=0.09$ and 0.16. \ Schlegel, Finkbeiner \& Davis (1998) find $E(B-V)=0.23$, while Burstein \& Heiles (1984) give $E(B-V)=0.17$. The Schlegel et al. (1998) values, plotted for a radius of 1 deg. around NGC 2208 show rather variable reddening, ranging from $E(B-V)=0.17$ to 0.29, also note that their zeropoint calibration produces reddenings systematically 0.02 mag greater than those of Burstein \& Heiles (1984). We will proceed by comparing the colors of the HB features and the RGB with those for other clusters. \ The position of the latter is determined both by the reddening and by the overall metallicity $[M/H].$ \ There are no high-dispersion spectroscopic analyses available for NGC\ 2808, or indeed for the comparison clusters we choose below, which would provide $[M/H].$ \ We make do with measurements based on the infrared Calcium triplet (Rutledge et al. 1997a,b), strictly $[Ca/H]$, and earlier measurements (see Zinn \& West 1984, Zinn 1985) denoted as $[Fe/H]$ but based on a number of indicators. \ This is an unsatisfactory situation, as discussed for example by Rutledge et al. (1997b), unlikely to be improved without more high dispersion spectroscopic analyses for stars in many globular clusters. \ \ If at all possible, reddening should be determined \ by methods that have little sensitivity to the metal abundance. \ \ Unfortunately, the NGC 2808 instability strip boundaries are not well defined thus precluding the use of the color of the blue edge which Walker (1998) confirms to have constant $B-V$ over a wide range of metallicity, nor are there accurate light curves and colors for the RR Lyraes which would allow the use of Sturch's method (Sturch 1966, Walker 1990). \ Instead we are forced mostly to metallicity-sensitive methods. \ Sarajedini \& Layden (1997) extend their method for simultaneously determining metallicity and reddening from a $(V,V-I)$ CMD to the $(V,B-V)$ case. Using several galactic globular clusters with well-determined values for these two quantities as standards they provide a calibration based on the color of the RGB at the level of the HB\footnote{% The level of the HB used is the mean magnitudes of the RR\ Lyraes if that is known, else the estimated mean magnitude of the HB at the color of the instability strip. \ For clusters with red HB morphology, the mean magnitude of the RHB\ clump is utilised (A. Sarajedini, private communication).}, and the magnitude difference between the RGB and the HB at $(B-V)_{0}=1.1$ and 1.2. \ The fiducial cluster NGC\ 1851 with HB level measured at the RR\ Lyrae position, and 47 Tucanae with HB\ measured from the red HB clump are the two calibrating clusters expected to bracket NGC 2808 in metallicity, so we will for comparison purposes measure the HB level both from the RHB\ stars and at the position of the RR\ Lyraes. \ Since NGC\ 2808 has very few stars in the vicinity of the instability strip, there is difficulty in estimating $V_{HB}$. \ We will use our observations of NGC\ 1851 (Walker 1992, 1998) and NGC\ 6362 (Brocato et al. 1999, Walker 1999), both of which have $<V_{RR}>-V_{RHB}=-0.08\pm 0.01$, where for NGC\ 6362 we use the bluer RHB clump stars only, due to the increase in $V$ magnitude range for the redder RHB\ clump stars.\ Thus for NGC\ 2808 $<V_{RR}>=16.22$ and $% <V_{RHB}>=16.30.$ \ \ The vertical extent of the RHB clump is 0.2 mag, so the ZAHB level for this feature is $V=16.40\pm 0.02$. \ Although these various HB levels are all very similar, the slope of the NGC 2808 RGB at the HB level is approximately $\Delta B-V)/\Delta V=-0.15$ and 2.5 mags further up the RGB\ $\Delta (B-V)/\Delta V=-0.45.$ \ Consequently, small errors in the placement of the HB have little influence on $(B-V)_{0,g}$ but will strongly affect any index involving the upper \ RGB. An example is the RGB shape calibration provided by Mighell, Sarajedini \& French (1998), who derive expressions for the color difference between the HB\ and positions 2 mags $(S_{-2.0})$ and 2.5 mags $(S_{-2.5})$ brighter, as a function of [Fe/H]. \ After iterating to find consistent values from the Sarajedini \& Layden (1997) \ fitting equations, we find $E(B-V)=0.16\pm 0.02$ and $% [Fe/H]=-1.05\pm 0.1$ (internal errors), using $<V_{RR}>$ to represent the HB level, and $E(B-V)=0.18\pm 0.02$ and $[Fe/H]=-1.15\pm 0.1$ (internal errors), using $V_{RHB}$ to represent $V_{HB}.$ \ Forcing $[Fe/H]$ to a more metal poor value consistent with the spectroscopy would increase the reddening by $\sim 0.03$ mag, very close to the $E(B-V)=0.20\pm 0.02$ obtained from the Zinn \& West (1984) $(B-V)_{0,g}$ calibration. \ Since the method uses a very similar $(B-V)_{0,g}$ relation to that of Zinn \& West (1984) this result is likely telling us that the the determination of the shape of the RGB is responsible for forcing this more metal-rich result. \ \ A \ more metal-rich value is also found from $S_{-2.0}$, $[Fe/H]=-1.10$ and $% -1.17$ using $V_{HB}=16.22$ and 16.30 respectively. \ \ This may be suggesting that the overall metal abundance $[M/H]$ is higher than indicated by $[Fe/H],$ assuming $[Ca/Fe]=0,$ since it is $[M/H]$ rather than $[Fe/H]$ that determines the location and shape of the RGB. \ \ We can also compare the CMD to that of NGC 1851 (Walker 1992, 1998), which has $E(B-V)=0.02\pm 0.02$, and $[Fe/H]=-1.29\pm 0.07$, \ the latter very similar to that for NGC 2808. \ The single Ca triplet measurement from an integrated spectrum of the cluster (Armandroff \& Zinn 1988) implies $% [Fe/H]=-1.16\pm 0.07$ (Da Costa \& Armandroff 1995) which is more consistent with comparisons between positions of RGB's for various clusters. \ Despite this uncertainty we will continue with the comparison, since NGC\ 1851, with bimodal HB, shares the characteristic of unusual HB\ structure with NGC 2808.\ \ \ NGC\ 1851 has a very tight clump of RHB\ stars centered at $% V=16.15,B-V=0.66,$ whereas the NGC\ 2808 RHB clump is more diffuse, and is centered near $V=16.30,B-V=0.83.$ Forcing these to match shows that the NGC 2808 stars are $0.17\pm 0.03$ redder than the NGC 1851 stars, therefore $% E(B-V)=0.19\pm 0.04$ . \ Comparing the two \ RGB's gives $E(B-V)=0.19\pm 0.03 $ with no correction applied for the differential difference. In principle the BHB stars can also be overlaid, which has the advantage that intrinsic color differences due to possibly different metallicity will be minimized. However there are some cautions. For these very blue stars the photometric color calibration may not be as secure as for cooler stars, as discussed section 4. \ Additionally, any cluster-cluster variation in the luminosity- temperature relationship for these stars means that they will follow different loci in the CMD. \ Borissova et al. (1997) show that the NGC 6229 and M3 sequences do not agree well, despite near-identical reddening and metallicity, and the same is found here when comparing the NGC 1851 and NGC 2808 BHB stars. Brocato et al. (1998) give a detailed discussion of this problem. \ More consistent results appear to be obtained if the comparison is restricted to the color of the BHB where the sequence begins its rapid decrease in ($V$) brightness, which occurs at $% (B-V)=0.10\pm 0.02$ for NGC\ 1851 and at $B-V=0.30\pm 0.05$ for NGC\ 2808, therefore $E(B-V)=0.22\pm 0.05$ for NGC 2808. \ \ As a consistency check we generated a fiducial from the Borissova et al. (1997) CMD of NGC 6229. This cluster has HB morphology similar to NGC 2808, with well determined and small reddening of $E(B-V)=0.01$, and $[Fe/H]=-1.44$. Their CMD covers the RGB and the HB only, so no comparison is possible for the turn-off region. We normalized the $V$ magnitudes at the RGB, and shifted the NGC 6229 color zeropoint by 0.19 mag to the red, i.e. we have assumed $E(B-V)=0.20$. We find that the RGB of NGC 6229 is approximately 0.02 mag bluer than that of NGC 2808 at the level of the HB, consistent with the nominal difference of approximately 0.1 dex in the metallicities. The slope of the NGC 6229 RGB is steeper, also consistent this cluster being somewhat more metal poor than NGC 2808. We can also compare with the CMD for NGC\ 6362 (Brocato et al. 1999), which has $[Fe/H]=-1.18\pm 0.06$ (Rutledge et al. 1997b) and with slightly more metal rich (0.1 dex) earlier estimates, all on the Zinn scale. \ Reddening, by a variety of methods, is $E(B-V)=0.05\pm 0.02.$ \ NGC\ 6362 is one of the most metal-rich clusters with halo kinematics, and has a well-populated HB\ with many BHB stars, but without the extensive tail of hot BHB\ stars displayed by NGC 2808. \ With offsets applied to NGC\ 6362 fiducials of $% \Delta V=0.92$ and $\Delta (B-V)=0.15$ mag the fit to the HB is excellent, however the NGC\ 6362 RGB\ lies 0.04 mag redder than the NGC 2808 RGB. \ This is consistent with NGC\ 2808 being 0.2 dex more metal poor than NGC 6362. In summary, there seems little doubt that the NGC\ 2808 \ reddening is near $% E(B-V)=0.20\pm 0.02.$ The metal abundance, based on Calcium triplet measurements (Rutledge at al. 1997b) is $[Fe/H]=-1.36\pm 0.05$, on the Zinn scale. \ The shape of the RGB\ is only marginally consistent with these values, suggesting an overall more metal-rich $[M/H]$ and a slightly smaller value for the reddening. \ There is difficulty matching both the HB\ and the RGB-MS in a comparison with NGC 1851, but the inconsistency is greatly reduced if the calcium triplet-based metallicity $[Fe/H]=-1.16$ is adopted for NGC\ 1851. \section{COLOR\ MAGNITUDE\ DIAGRAM\ MORPHOLOGY} The CMD morphology has been well-illustrated in the several previous studies of this cluster, as discussed in the introduction, and we have made comparisons with other clusters above. \ Of most interest is the HST study by S97 which for the first time clearly shows the presence of multiple groups of stars on the HB, and a prominent blue straggler sequence. However even with HST the crowding in the small field near the cluster center is sufficiently great that this CMD only reaches to $V=21$. We therefore begin by preparing a composite CMD from the three sets of exposures. It should be stressed that this diagram (Figure 2), containing 33464 stars, is primarily meant to illustrate the morphology of the various sequences and does not represent in any way the relative numbers of stars in each, since the short exposures reached to the cluster center for the brightest stars whereas for the longer exposures the central regions of the cluster were not measured. It also became apparent that there were significant reddening variations across the field, as the position of the color of the MS turn-off varied when the data were plotted in spatial (eg 512x512 pixels) sections. As an aside, in principle it would have been preferable to use the centroid of the clump of RHB stars to ascertain reddening, but there were insufficient such stars in the outer parts of the frame to provide an accurate centroid. Photometry in the outer parts of the frame were adjusted to match the mean of the center 1024x1024 pixels, the absolute values of the adjustments averaged 0.028 mag for $B-V$ and were all less than 0.055 mag. Variations in reddening across the field are not unexpected given the low galactic latitude and the evidence for clumpy reddening in the region from the Schlegel et al. (1998) values. The latter are consistent with the present results, showing slightly higher reddening to the NE of NGC 2808 and slightly lower reddening in the other three orthogonal directions. The resultant CMD, which has been corrected for differential reddening, clearly shows the principal sequences, contaminated by a very significant number of field stars. When preparing this diagram it became clear that crowding of this highly condensed cluster is the major factor in preventing accurate ground- based photometry. Even so, the main sequence is well defined and here extends to 4 magnitudes fainter than the turn-off. The HB morphology is, of course, highly unusual, and we confirm the gaps discovered by S97. Since their HB terminated near $V=21$ $B-V=-0.1$ which was their magnitude limit, it is important to see whether or not this really is the end of the HB. The magnitude limit of the present data is $B=25$ and $V=24$ and the HB end is seen near $V=21.2$, thus this $is$ indeed the end. There is a range in color for the BHB\ stars between magnitudes 16-18 that is likely due to star-star differential reddening, also responsible for broadening other vertical sequences with small photometric errors, such as the RGB. \ The most prominent gap in the HB is the region of the RR Lyraes. The RHB clump is well defined and is separated by a luminosity gap from stars at the base of the AGB. The subgiant branch and RGB appear normal, with the RGB terminating near $B-V=1.9$.\ The RGB\ bump, arising from a hiatus in the rate of stellar evolution up the RGB as a consquence of the hydrogen burning shell passing through a discontinuity left by the maximum penetration of the convective envelope (Iben 1968), is clearly visible in the CMD. \ Cassisi \& Salaris (1997) present a critical theoretical discussion of the positions of the ZAHB and the RGB\ bump, deriving $\Delta V_{HB}^{bump}=1.083+1.380[M/H]+0.231[M/H]^{2} $, where $[M/H]$ is the global heavy metal abundance, which can be approximated as (Salaris, Chieffi \& Straniero 1993) as $[M/H]=[Fe/H]+\log (0.638\cdot f+0.362)$, where $\log f=[\alpha /Fe]$, the enhancement factor for the $\alpha $ elements. \ They also find that the position of the bump depends on the mass of the model, and derive an increase of $\Delta V_{HB}^{bump}$ of almost 0.024 mag for an increase of 1 Gyr in age of the cluster. \ In an observational evaluation, Saviane et al. (1998) \ approximate the relation as \ $\Delta V_{HB}^{bump}=0.66[Fe/H]+0.87.$ \ The HB reference magnitude in both cases is that for the ZAHB in the vicinity of the RR\ Lyraes. \ We use the comparison \ with NGC\ 1851 and 6362 (see above) to estimate $V_{ZAHB}$ for NGC\ 2808, thus obtaining $\Delta V_{HB}^{bump}=0.00\pm 0.07$, and with the S98 calibration $[Fe/H]=-1.32\pm 0.11$. \ With the Cassisi \& Salaris (1997) formulation, we find $[M/H]=-0.93\pm 0.07.$ \ Since there is no measured value of $[\alpha /Fe]$ for NGC 2808, we will adopt a value of $0.30\pm 0.05$% , whereupon $[Fe/H]=-1.14\pm 0.08.$ \ As pointed out by Cassisi \& Salaris (1997), their theoretical scale agrees well with high-dispersion spectroscopic measurements of cluster metallicities, which tend to give values somewhat more metal-rich than the Zinn scale, the Zinn-scale estimate of $[Fe/H]=-1.36\pm 0.05$ transforms to $[Fe/H]=-1.11\pm 0.03$ on the Carretta \& Gratton (1997) high dispersion spectroscopic scale (Rutledge et al. 1997b). \ So in this admittably round-about fashion, the position of the NGC\ 2808 RGB\ bump is found to be in excellent agreement with the Cassisi \& Salaris (1997) models, assuming the $[\alpha /Fe]$ ratio is normal. \ This would also argue against enhanced $[\alpha /Fe]$ being responsible for the shallower than expected RGB, as discussed above. \section{RADIAL POPULATION GRADIENTS\qquad} The measurement of the relative numbers of stars as a function of radius in a cluster as rich as NGC 2808 is a difficult task, and one best suited to HST. Here we have made a few comparisons between groups of stars of roughly equivalent brightness in order to minimize magnitude-dependent completeness corrections. The extent of the difficulty with the present data set is evident by comparing the ratio of AGB+RGB stars brighter than $V=15$ with the RGB stars with $V=15$ to 17.5, using only the short exposure frames. Outside $r=150$ pixels\ (one pixel = 0.432 arcsec) the ratio faint:bright is constant at 8.8, whereas from 50:150 pixels it is 5.4, and inside 50 pixels it is 1.5. We find no significant differences between the distribution of RHB and BHB stars within the errors, comparing only the brightest clump ($% V<18.5$) of BHB stars; there are $2.1\pm 0.5$ times as many RHB stars as these BHB stars in the $150-300$ pixel radii annulus and this ratio is preserved, within a progressively larger error, at greater distances. For the two fainter clumps of BHB stars the number counts would appear to indicate a deficiency of such stars at small radii, compared to the RHB or brighter BHB stars, but a comparison between ratios of MS stars with the same $V$ magnitude shows that the number counts of stars as faint as this are seriously incomplete in the inner annuli. \ We do find that the HB star distribution in the outer annuli (ie at distances greater than 300 pixels from the cluster center, is similar to that seen at the cluster center by S97. This would argue for there being no gross mass differences between the sub-groups on the HB, or else the cluster is dynamically mixed. The latter could be tested by comparing the radial distribution of the BS stars to that for other cluster stars. \section{BINARIES\qquad} The F97 field is about 6 arcmin south and 2 arcmin east of the cluster center, and their CMD shows a redwards broadening of the MS that can be interpreted as being due to binaries, with some $24\pm 4\%$ of apparently single MS stars being in such a category. As cautioned by F97 these may be unresolved optical binaries, some of which must be expected due to the high star density, rather than true physical binaries. Unfortunately it is difficult to directly test between the two options, as both mass segregation of the binary systems (e.g. Rubenstein \& Bailyn 1999) and the effects of image blending will act to concentrate such stars towards the center of the cluster. F97 consider, from artificial star tests, that no more than 10\% of the candidate binaries are optical in character. We first compare the two extremes. Stars nearer than 350 pixels (2.52 arcmin) from the cluster center (note that no stars from the long exposures were measured this close) are plotted in Figure 3. Of the 3566 stars in this CMD, only 300 are on the MS ($V>20$) and these clearly suffer much photometric scatter. However there is very little field star contamination and the evolved star sequences, with the exception of the fainter BHB stars, are well populated. There are at least 25 BS stars. But from this CMD nothing can be said about the spread of MS stars. We can contrast this (Figure 4) with a CMD made up by selecting stars more than 1000 pixels from the cluster center. This shows a well defined cluster MS, few evolved stars, and the expected large number of field stars. A careful examination of the distribution of stars about a MS fiducial shows no evidence for any significant binary population. Turning now to fields at intermediate radii, we plot the CMD between radii of $600-800$ pixels (Figure 5) and $800-1000$ pixels (Figure 6). It is visually obvious from the CMDs, especially Figure 5, that there is a redwards spread of the MS. We therefore need to carefully test the stellar image crowding in these annuli in order to differentiate between the optical and physical binary hypotheses. The number of measured stars in each CMD is 9119 and 6753; thus there are 96 and 167 pixels per star respectively. With image fwhm of under 2.5 pixels, this degree of crowding does not seem excessive, and indeed with 1 arcsec seeing the image crowding for these data becomes severe only well inside the $r=600$ pixels radius. To put this initial impression on a more quantitative basis, artificial stars were added to the long-exposure frames using the ADDSTAR routine of DAOPHOT (Stetson 1987). As long as the brightness of the added stars exceeded the MS magnitude limit of $V\sim 23.5$ by more than one magnitude, the percentage of stars recovered in both annuli exceeded 95\%. It is therefore clear, as was concluded by F97, that optical binaries make only a minor contribution to the observed distribution. Our next test was to restrict the analysis to only the stars in the $600-800$ pixel annulus. Note that this annulus is slightly closer in the mean to the cluster center (5 arcmin) than the F97 field, We have already purged the star list of stars with poor profile fit and elliptical shape parameters, as determined by the DAOPHOT PSF fitting, and have also cut-off stars with standard errors (as produced by ALLSTAR) greater than twice the median standard error of all stars in the sample, in one magnitude wide bins. For stars on the MS, the cutoff corresponds to a standard error $\sigma (B-V)=(0.06,0.08,0.10,0.15)$ mag. at $V=(20,21,22,23) $ mag. respectively. \ We call this set of stars ``sample one'', and it corresponds to what is shown in Figure 5. We then further excluded all stars with standard errors greater than 0.030 mag, which had the effect of removing all stars fainter than $V=22.6$ as well as the significant number of stars with color errors between 0.03 mag and the cut-offs given above. The remaining stars are ``sample two''. The radial distributions of the stars in the two samples differ negligibly, with less than one pixel difference in mean distance from the cluster center. \ We then compared the color distributions across the same MS fiducial, from $V=20.0-22.0$, normalizing the two distributions at the peak. The normalizing factor (1.16) is not too different from unity, and the resulting distributions are plotted in Figure 7. \ Bluewards of the peak the two curves are nearly identical. The redward distributions are very different, with the{\it \ low error }plot (sample 2) showing substantially fewer stars and a distribution close to symmetrical across the peak, especially if the field star contributions are taken into account. \ Given the ADDSTAR experiments discussed above, some of small number of excess stars (the red side of the sample two distribution) may be optical binaries, but even if all the outliers are physical binaries they contribute no more than $10\%$ of the total number of stars measured on the MS. The distribution is however, equally consistent with a MS binary proportion of zero. These experiments show that within the annulus under consideration, the apparent binary population is due to a non-Gaussian distribution of color errors about the MS fiducial. An explanation could be that the most likely color of any contaminating star is redder than a given MS star, since fainter stars are either cool cluster MS stars or very red foreground dwarfs, or faint compact galaxies. \ We therefore conclude that there is no clear evidence from the MS color distribution to support the existence of a large proportion of physical binaries in NGC 2808, at least outside a radius of a few arcmin. \section{\protect\bigskip FINAL\ REMARKS} NGC\ 2808 has the most unusual HB\ morphology of any galactic GC, and is an important fiducial for any explanation of the late stages of stellar evolution in general and in high-density GC in particular. \ In this paper, based on photometrically secure observations, we determine the cluster reddening to be $E(B-V)=0.20\pm 0.02$ and show that this is consistent with the spectroscopically determined metallicity. \ Small differences in the values for $E(B-B)$ and $[Fe/H]$ via different methods may only reflect calibration difficulties, but may also be indicative of differences in $% [\alpha /Fe]$ between the comparison clusters chosen to have similar $[Fe/H]$ to NGC 2808. \ Significant progress in this regard requires high dispersion spectroscopy in order to determine total metal abundance $[M/H]$. \ \ The CMD presented confirms and extends recent HST results, although there is an 0.05 mag difference in the $B-V$ photometric zeropoint. \ No significant differences in radial gradients between various groups of stars on the CMD are found, while the presence of a large fraction of MS binaries is not confirmed. \ However the observations here do not extend into the core region of this very dense cluster, and studies of gradients and binary fraction to the very center of NGC 2808, using HST, would be of considerable interest.
\section{#1}} \defT$_{e\!f\!f}\;${$T_{\mathrm{eff}}$ } \defT$_*\;${T$_*\;$} \defM$_v\;${M$_v\;$} \begin{document} \thesaurus{emission line, Be: 08.05.2; evolution: 08.05.3; rotation: 08.18.1} \title{Differences in the fractions of Be stars in galaxies.} \author{Andr\'e Maeder$^1$ \and Eva K.\ Grebel$^{2,3,}$\thanks{Hubble Fellow} \and Jean--Claude Mermilliod$^4$} \institute{Observatoire de Gen\`eve, CH-1290 Sauverny, Switzerland, e--mail:<EMAIL> \and UCO/Lick Observatory, University of California, Santa Cruz, CA 95064, USA \and University of Washington, Department of Astronomy, Box 351580, Seattle, WA 98195, USA \and Institut d'Astronomie, UniL, CH--1290 Chavannes--des--Bois, Switzerland} \date{Received date; accepted date} \maketitle \markboth{A. Maeder et al.: Differences of the fractions of Be stars in galaxies.}{A. Maeder et al.: Differences of the fractions of Be stars in galaxies.} \begin{abstract} The number ratios $Be/(B+Be)$ of Be to B--type stars in young, well studied clusters of the Galaxy, the LMC and SMC are examined. In order to disentangle age and metallicity effects we choose clusters in the same age interval and for which reliable photometric and spectroscopic data are available. Number counts are made for various magnitude intervals, and the results are found to be stable with respect to this choice. In the magnitude interval $M_{\mathrm{V}} = -5$ to -1.4 (i.e. O9 to B3) we obtained a ratio $Be/(B+Be)$ = 0.11, 0.19, 0.23, 0.39 for 21 clusters located in the interior of the Galaxy, the exterior of the Galaxy, the LMC and the SMC, respectively. Various hypotheses for these differences are examined. An interesting possibility is that the average rotation is faster at low metallicities as a result of star formation processes. The much higher relative N--enrichment found by Venn et al. (\cite{vencar}) in A--type supergiants of the SMC, compared to galactic supergiants, also strongly supports the presence of more rotational mixing at low metallicities. We discuss whether high rotational mixing may be the source of primary nitrogen in the early chemical evolution of galaxies. \end{abstract} \section{Introduction} About 50\% of the B stars in the SMC cluster NGC 330 were found to be Be stars by Feast (\cite{fea}). This high fraction, compared to 10 to 20\% in the Milky Way, was confirmed by subsequent studies done by Grebel (\cite{greric, grerob}), Mazzali et al. (\cite{maz}), and Keller et al. (\cite{kel2}). A pronounced difference between the Be star content in the Magellanic Cloud clusters and in the Milky Way was also found by Grebel et al. (\cite{grerobwil}). The Be phenomenon is closely related to fast rotation. Suggestions were made in the past that Be stars occur near the end of the main sequence (MS) phase; however, extended studies in clusters have shown the occurrence of Be stars from the zero--age sequence to the end of the MS (Mermilliod \cite{mer}). Mermilliod has also found that the fraction of Be stars with respect to normal B stars is low in very young clusters, that it reaches a maximum for clusters with turnoff in the range of O9 to B3 and that it declines after that. Evidences for an age effect are also well shown by Fig. 5 from Grebel (\cite{gre}). Be stars are generally slightly shifted to the red of the MS band (Grebel et al. \cite {grerob}; cf. also Fig. 1 below). This effect is consistent with the effect of rotational reddening found by Maeder and Peytremann (\cite{maepey}) and also with the effect of a circumstellar envelope, which contributes to a reddening as well as to the formation of the emission lines (Gehrz et al.\ 1994, Cot\'e \& Waters 1987). A physical model of the radiation-driven wind in rotating stars has been proposed by Bjorkman and Cassinelli (\cite{bjo}). This model, the so-called wind compressed disk (WCD) model, predicts that the isotropic wind particles travel along trajectories that go through the equatorial plane, where shocks occur. The gas is thus confined there and forms a dense equatorial disk. Owocki et al. (\cite{owoc}) have shown that gravity darkening, as predicted by the von Zeipel theorem, leads to a very different wind morphology, where the disk is more difficult to form. The application of the radiative wind theory to rotating stars leads to an expression (cf. Maeder \cite{mae99a}) for the mass loss rates $\dot{M} (\vartheta)$ as a function of the colatitude $\vartheta$, this expression shows that there are two main effects influencing the anisotropic mass loss: ~a) the ``$g_{\mathrm{eff}}$''--effect, i.e. the higher gravity and T$_{e\!f\!f}\;$ at the pole of a rotating star (due to von Zeipel's theorem) enhances the polar mass loss; b) the ``opacity--effect'', i.e. the higher opacity, and the higher force multipliers, due to the lower temperature at the equator favour an equatorial ejection and the formation of a ring. The B[e] stars show both polar and equatorial ejections (Zickgraf \cite{zick}). The equatorial ejection is strongly favoured by the so--called bi--stability effect (Lamers \cite{lam}), i.e. a jump of the opacity near 20'000${\degr}$K, i.e. close to spectral type B2 where a maximum of the fraction of Be stars is observed. Models of rotating stars, including hydrostatic distorsion, shear mixing, meridional circulation, enhanced mass loss rates, loss of angular momentum etc. have been constructed recently (Meynet \cite{mey99}; Maeder \cite{mae99b}). These models show the high influence of rotation on massive star evolution. The main reason is that shear mixing, which is the most efficient mixing process, is favoured by a high thermal diffusivity as observed in massive stars (Maeder \cite{mae97}); also, other effects such as meridional circulation (Maeder \& Zahn \cite{maezah}) and anisotropic mass loss are quite significant in massive stars. In view of the large consequences of rotation, it is particularly useful to examine whether the Be fraction is systematically higher in clusters of lower metallicities, a possibility also suspected by Grebel et al.\ (\cite{grericboe}, \cite{gre}) and Mazzali et al. (\cite{maz}). Section 2 establishes the number data of Be stars in an ensemble of clusters of the Galaxy, LMC and SMC. The relation of the results with metallicity is examined in Section 3, together with abundance indications related to rotation. Section 4 gives the conclusions. \section{Number counts of Be stars in clusters} \subsection{Cluster data} Since the fraction of Be stars is rapidly changing with cluster ages (Mermilliod \cite{mer}), we must carefully separate ages and metallicity effects. We select clusters in the age range of $\log$ age = 7.00 to 7.40 which have their turnoff around the maximum of the Be star distribution. The comparison of clusters of various ages in different galaxies may lead to confusion between age and metallicity effects. For the Milky Way the clusters were selected from the cluster data base WEBDA (Mermilliod \cite{mer99}). This data base contains homogeneous photometric data with consistent estimates of reddening, distance moduli and ages. WEBDA also provides indications on known binaries, stellar peculiarities and Be star identifications based on standard spectroscopic criteria. The data base includes the recent survey of stars with H$\alpha$ emission by Kohoutek and Wehmeyer(\cite{koh}). Some 11 well studied clusters of the Galaxy are in the age interval considered (Table 1). On the basis of their galactocentric distances, the clusters are separated in two groups: one outside and one inside the solar location. \begin{figure}[htbp] \centerline{\hbox{\psfig{figure=8572_fig1.eps,width=\linewidth}}} \caption{Example of colour--magnitude diagram used for the number counts of Be and B stars in the case of NGC 884 from the WEBDA data base (Mermilliod \cite{mer99}). The limits of the various magnitude intervals are shown: $M_{\mathrm{V}}$ = -4, -2; $M_{\mathrm{V}}$ = -5, -2; $M_{\mathrm{V}}$ = -5, -1.4. Be stars are represented by black dots. The isochrone corresponding to log age = 7.1 is represented by a continuous line, and the top of the binary sequence upwards shifted by 0.75 mag. is shown by a broken line.}\label{Fig1} \end{figure} For the LMC the photometric data, the age determinations and the identifications of Be stars come from Grebel (\cite{gre}), Dieball and Grebel (\cite{die}), Keller et al. (\cite{kel2}), and Grebel and Chu (in preparation). A distance modulus of 18.50 is taken in agreement with current determinations (cf. Cole \cite{col}). Nine LMC clusters (listed in Table 1) are in the considered age interval. For the SMC the typical cluster NGC 330 is analysed with the data from Grebel et al. (\cite{grerob}). Unfortunately there is no other cluster with reliable Be star identification in or even near the considered age range. The cluster NGC~346 of the SMC, which contains a few Be stars, is much too young and has a turnoff far out of the age range where the maximum of Be stars occurs. Kudritzki et al. (\cite{kud}) give an age of $2.6 \cdot 10^6$~yr for NGC 346, which is consistent with the presence of stars up to about 100 M$_\odot$ (Massey et al. \cite{mas}). This is in agreement with other authors (Cassatella et al. \cite{cas}; Haser et al. \cite{has}) who notice the presence of O3 stars in NGC~346. \setlength{\tabcolsep}{4mm} \begin{table*}[htbp] \caption[]{Cluster data} \begin{flushleft} \begin{tabular}{lrrrrrrrrl} \noalign{\smallskip} \hline \noalign{\smallskip} \multicolumn{1}{r}{} & \multicolumn{1}{r}{} & \multicolumn{1}{r}{} & \multicolumn{2}{c}{$M_{\mathrm{V}}$ = -5,\,-1.4} & \multicolumn{2}{c}{$M_{\mathrm{V}}$ = -5,\,-2} & \multicolumn{2}{c}{$M_{\mathrm{V}}$ = -4,\,-2} & \multicolumn{1}{r}{}\\[2mm] Cluster & $\log$(age)\hspace{-3mm} & DM\hspace{1mm} & B+Be\hspace{-2mm} & Be & B+Be\hspace{-2mm} & Be & B+Be\hspace{-2mm} & Be & Ref.\\ \noalign{\smallskip} \hline \noalign{\smallskip} {\bf SMC} &&&&&&&&&\\ NGC 330 & 7.30 & 18.90 & 128 & 50 & 89 & 39 & 81 & 37 & Grebel et al. 1996\\[3mm] {\bf LMC} &&&&&&&&&\\ Hodge 301 & 7.30 & 18.50 & 44 & 10 & 25 & 9 & 25 & 9 & Grebel \& Chu in prep\\ KMHK 1019 & 7.20 & 18.50 & 11 & 2 & 1 & 1 & --- & --- & Dieball \& Grebel 1998\\ NGC 1818 A & 7.40 & 18.50 & 94 & 34 & 58 & 22 & 51 & 19 & Grebel 1997\\ NGC 1818 B & 7.40 & 18.50 & 7 & 3 & 4 & 2 & 4 & 2 & Grebel 1997\\ NGC 1948 & 7.40:\hspace{-1mm} & 18.50 & 101 & 11 & 71 & 9 & 64 & 9 & Keller et al. 1999\\ NGC 2004 & 7.40 & 18.50 & 130:\hspace{-1mm} & 25 & 96:\hspace{-1mm} & 21 & 82:\hspace{-1mm} & 17 & Grebel 1998\\ NGC 2006 & 7.30 & 18.50 & 35 & 10 & 21 & 5 & 21 & 5 & Dieball \& Grebel 1998\\ NGC 2100 & 7.40:\hspace{-1mm} & 18.50 & 67 & 19 & 49 & 16 & 43 & 15& Keller et al. 1999\\ SL 538 & 7.20 & 18.50 & 46 & 11 & 29 & 9 & 28 & 9 & Dieball \& Grebel 1998\\ TOTAL &&& 535 & 125 & 354 & 94 & 318 & 85 & \\[3mm] {\bf Galaxy ext.} && {$\mathrm{R_{gc}}(kpc)$}\hspace{-5mm} &&&&&&&\\ NGC 457 & 7.30 & 9.92 & 28 & 4 & 17 & 4 & 13 & 1 & Mermilliod 1999\\ NGC 581 & 7.20 & 9.54 & 8 & 1 & 5 & 0 & 4 & 0 & Mermilliod 1999\\ NGC 663 & 7.20 & 9.61 & 35 & 12 & 23 & 12 & 21 & 10 & Mermilliod 1999\\ NGC 869 & 7.10 & 9.60 &42 & 3 & 33 & 3 & 28 & 2 & Mermilliod 1999\\ NGC 884 & 7.10 & 9.91 & 28 & 6 & 24 & 6 & 18 & 3 & Mermilliod 1999\\ NGC 957 & 7.15 & 9.41 & 15 & 4 & 13 & 4 & 12 & 3 & Mermilliod 1999\\ NGC 2439 & 7.30 & 10.45 & 31 & 5 & 21 & 5 & 21 & 5 & Mermilliod 1999\\ NGC 7160 & 7.20 & 8.22 & 3 & 1 & 2 & 1 & 2 & 1 & Mermilliod 1999\\ TOTAL &&& 190 & 36 & 138 & 35 & 119 & 25 &\\[3mm] {\bf Galaxy int.} &&&&&&&&&\\ NGC 3293 & 7.20 & 7.70 & 37 & 1 & 32 & 1 & 25 & 1 & Mermilliod 1999\\ NGC 3766 & 7.40 & 7.44 & 42 & 10 & 25 & 8 & 24 & 8 & Mermilliod 1999\\ NGC 4755 & 7.20 & 7.07 & 47 & 3 & 27 & 3 & 23 & 3 & Mermilliod 1999\\ TOTAL &&& 126 & 14 & 84 & 12 & 72 & 12 &\\ \hline \end{tabular} \end{flushleft} \end{table*} \subsection{Number counts of Be stars in luminosity intervals} We choose to proceed to number counts in given intervals of magnitude $M_{\mathrm{V}}$ rather than in given ranges of spectral type and T$_{e\!f\!f}\;$ since spectral types are not available for all clusters. Also, we noticed that high rotation strongly modifies the average T$_{e\!f\!f}\;$ and very little the average stellar luminosities (Maeder and Peytremann \cite{maepey}). Furthermore, massive stars of the same mass, but different metallicites, have rather large differences in T$_{e\!f\!f}\;$, while this is not the case for the luminosities (Schaller et al. \cite{scha}). In order to test the independence of the results with respect to the chosen $M_{\mathrm{V}}$ range we have done number counts in three intervals of $M_{\mathrm{V}}$ centered on the domain O9 to B3, i.e. $M_{\mathrm{V}}$ = -4, -2; $M_{\mathrm{V}}$ = -5, -2; $M_{\mathrm{V}}$ = -5, -1.4 (the limit $M_{\mathrm{V}}$ = -1.4 corresponds to the spectral type B3 according to the calibration of Zorec and Briot (1991) for main sequence B--type stars). Fig. \ref{Fig1} shows an example of the HR diagram used for the number counts in the case of NGC 884 ($\chi$ Persei cluster) with the various magnitude intervals considered. Table 1 shows the number counts for the 19 clusters in the four different locations considered. The 2nd column gives the $\log$ of the age, the 3rd column gives the distance moduli in the SMC and LMC, the galactocentric distance for galactic clusters is given in the 3rd column as well; the following columns give the numbers of B plus Be stars, and those of Be stars only in the indicated magnitude intervals. The reference for the data is in the last column.\\ The average number ratios $Be/(B+Be)$ stars in the four zones and for the different magnitude intervals considered are given in Table 2. We clearly notice a systematic trend for smaller $Be/(B+Be)$ number ratios in the sequence of the four groups considered from SMC to LMC to the galactic exterior and interior. The total difference in the $Be/(B+Be)$ ratios amounts to a factor of about three, which is important and also systematic in relation with the metallicities as discussed below. \subsection{Possible relations with Z} We examine the possible relation between the fraction of Be stars and the local metallicity Z in the regions where these stars were formed. Let us notice that the suggestion to determine whether the frequency of Be stars is somehow related to the metal abundance was also made by Grebel et al. (1992) and Mazzali et al. (1996). \begin{table}[htbp] \caption[]{Number ratios of Be to B+Be stars in the Magellanic Clouds and Milky Way.} \begin{flushleft} \begin{tabular}{lllll} \noalign{\smallskip} \hline \noalign{\smallskip} \multicolumn{1}{l}{} & \multicolumn{1}{l}{$M_{\mathrm{V}}$} & \multicolumn{3}{c}{$\frac{Be}{B+Be}$}\\ \noalign{\smallskip} \cline{3-5} \noalign{\smallskip} \multicolumn{1}{l}{} & \multicolumn{1}{l}{} & \multicolumn{1}{l}{-5,\,-1.4} & \multicolumn{1}{l}{-5,\,-2} & \multicolumn{1}{l}{-4,\,-2}\\ \noalign{\smallskip} \hline \noalign{\smallskip} SMC & 128 & .39 & .44 & .46\\ LMC & 535 & .23 & .27 & .27\\ Galaxy ext. & 190 & .19 & .25 & .21\\ Galaxy int. & 126 & .11 & .14 & .17\\ \noalign{\smallskip} \hline \noalign{\smallskip} \end{tabular} \end{flushleft} \end{table} We use the recent data on the metallicity Z for the various zones. For the solar metallicity we take a value of Z$_\odot$ = 0.018, which is consistent with recent solar models and heliosismological data (Brun et al. 1998). For the chemical gradient in the Galaxy the current values range between $\Delta[O/H]/kpc = -0.07$ (cf. Shaver et al. \cite{sha}; Gummersbach et al. \cite{gum}) and -0.05 (cf. Vilchez and Esteban \cite{vil}). The group of clusters towards the anticenter has an average galactocentric distance $R_{gc}$ of about 1.5 kpc larger than that of the Sun (taken to be $R_{gc} = 8$ kpc), while this average difference is 0.6 kpc for the group of clusters towards the galactic interior. Thus we consider that the average local metallicities for the exterior and interior groups are respectively Z\,=\,0.014 and Z\,=\,0.020 for a galactic gradient of -0.07. For a gradient of -0.05 the Z-values would be 0.015 and 0.019 respectively, i.e. not very different from the latter values. For the young population in the LMC the range of estimated metallicities is rather large. For example, Olszewski et al. (\cite{ols}) give [Fe/H] values in the range -0.3 to -\,0.42, which corresponds to Z\,=\,0.010 to 0.008. Jasniewicz and Th\'evenin (1994) find values from [Fe/H] = -0.4 for NGC 1818 to -0.55 for NGC 2004, which correspond to respectively Z\,=\,0.008 and 0.006. Bica et al. (\cite{bic}) obtained values Z\,=\,0.004 to 0.005. Luck et al. (\cite{luck}) find a mean [Fe/H] = $-0.34\pm0.15$ for LMC field Cepheids. In this context, an average value of 0.007 seems appropriate. For the SMC field Reiterman et al. (\cite{rei}), Spite et al. (\cite{spi}), Grebel and Richtler (\cite{greric}) give values corresponding to Z\,=\,0.001 to 0.003. For NGC 330, Hilker et al. (\cite{hil}) give a value Z\,=\,0.002, and the results of Hill (\cite{hill}) correspond to 0.003 to 0.004 and those of Oliva et al. (\cite{oli}) to 0.001. Luck et al. (\cite{luck}) find a value corresponding to Z=0.004. Gonzalez and Wallerstein (1999 \cite{gon}) find Z = 0.002. In this context an average value Z\,=\,0.002 for the SMC is appropriate. Fig. 2 shows the relation between the fraction $Be/(B+Be)$ and the local average metallicity. The trend is quite clear for the various magnitude intervals considered. For the sample of the 4 locations there seems to be a clear increase of the fraction of the relative number of Be stars with the local average initial metallicity. \begin{figure}[htbp] \centerline{\hbox{\psfig{figure=8572_fig2.eps,width=\linewidth}}} \caption{Relation between the number ratio Be/(B+Be) and the local metallicity for the 4 groups of clusters considered in Table 2. To test the validity of the results, the number counts were made in different magnitude intervals, the dots refer to counts made in the magnitude interval $M_{\mathrm{V}}$ = -5, -1.4, the crosses to the interval -5, -2 and the triangles to the interval -4, -2.}\label{Fig2} \end{figure} \section{Interpretation of the results} There are various possibilities to interpret the above results. \noindent \parbox[t]{.6cm}{1.}\parbox[t]{8cm}{Statistical effects related to the intermittence of the Be phenomenon.}\\[2mm] \parbox[t]{.6cm}{2.}\parbox[t]{8cm}{The visibility of the Be phenomenon is enhanced at low metallicities.}\\[2mm] \parbox[t]{.6cm}{3.}\parbox[t]{8cm}{Shell ejection is favoured at low metallicities.}\\[2mm] \parbox[t]{.6cm}{4.}\parbox[t]{8cm}{Differences in the distribution of the rotational velocities.}\\ The first possibility is very unlikely. In the Milky Way Be stars have generally been observed over a longer period than in the LMC or SMC. Therefore, the intermittent Be star have a higher chance to be detected in the Milky Way. Consequently, the correction of this possible effect would increase the trend observed in Fig. 2 rather than reduce it. Indeed, in the SMC cluster NGC 330 the Be phenomenon has been found to be intermittent in a number of stars re-observed over the period of several years (Grebel 1995, Keller et al. 1999). The second hypothesis also seems difficult to sustain, because at higher metallicities the ejected shell and circumstellar material would be more opaque and contain more dust, thus their visibility through emission lines as well as by their infrared continuum would be enhanced. Thus, the correction of such an effect (if feasible) would increase the observed trend. The same kind of remark holds for the third hypothesis. Models of rotating stars with mass loss (Maeder \cite{mae99a}) show two effects: a) polar ejection is favoured in rotating stars by the higher T$_{e\!f\!f}\;$ at the pole; b) equatorial ejection of a shell is favoured by larger opacities. Thus, at lower Z, it is not expected that shell ejection is favoured in the equatorial regions of a rotating star, because the opacities are lower. The possible effect of metallicity on the terminal velocities, which may follow from the wind-compressed disk model (Bjorkman and Cassinelli \cite{bjo}; Grebel \cite{gre}), is found to be negligible (Maeder \cite{mae99a}). Thus we are left with the possibility that the higher fraction of Be stars at lower Z is the signature of more fast rotators at lower Z. Obviously, direct measurements in nearby galaxies are very needed and envisaged in order to confirm or reject this suggestion. For NGC 330 measurements of projected rotational velocities ($v\,\sin\,i$) by Mazzali et al. (\cite{maz}) and Keller and Bessell (\cite{kel1}) show values of up to 400 km s$^{-1}$. Furthermore, a recent study of nitrogen abundances done by Venn (\cite{ven}) offers a strong support to the possibility of faster rotation in NGC 330 of the SMC. Indeed, several authors have found that B-- and A--type supergiants in the Galaxy, LMC and SMC show nitrogen enhancements not predicted by standard evolutionary models (cf. Lennon et al. \cite{len}; Fitzpatrick and Bohannan \cite{fit}; Venn \cite{ven}, \cite{vencar}). The current interpretation of these surface enhancements requires some additional mixing processes, possibly rotational mixing (cf. Langer \cite{lan}; Meynet \cite{mey}; Maeder and Zahn \cite{maezah}). New rotating models show consistently some N--enrichments already present for medium rotation from the end of the MS onwards (Meynet \cite{mey}). A remarkable point is that part of the N--enhancement observable in B-- and A--type supergiants could be of primary origin, due to the fact that some new $^{12}$C resulting from the 3$\alpha$ burning is rotationally diffused into the H--burning shell, where the CNO cycle converts this ``new $^{12}$C'' into $^{14}$N, which is thus of primary origin. Another interesting fact recently found by Venn (\cite{vencar}) is that the degree of N enhancement for A supergiants is much higher in the SMC than in the Milky Way. For galactic supergiants the typical enhancement of [N/H] is about a factor of 2, while in the SMC it reaches a factor of 10--20 with respect to the standard local average [N/H] in the SMC. As a matter of fact, converting the $^{12}$C entirely into $^{14}$N by the end of the CN cycle (which is dominant) would lead to an increase of [N/H] by a factor of 4--5, but not as high as 20. Thus the high [N/H] enhancement observed for SMC supergiants might be a sign of primary nitrogen. We do not know, the only way to clarify is to measure the sum of CNO elements and check whether it is the same or not in supergiants as in MS stars. Anyhow, whether this is primary or secondary nitrogen, the point for now is that this higher [N/H] is the signature of much more mixing in the SMC than in the Galaxy, a fact which is quite consistent with the above suggestion of faster rotation for massive stars in the SMC. The possibility of faster rotation in the SMC has also been mentioned by Venn (\cite{vencar}) and the results presently available on the Be star fractions as a function of Z clearly support this view. \hspace{0.1mm} \section{Conclusions} The main result is that the relative fraction of Be stars with respect to all B stars in the spectral interval O9 to B3 is increasing for lower metallicities Z. This fact together with the much larger N-excesses observed in A-type supergiants of the SMC than in galactic supergiants strongly supports the suggestion that there are more fast rotators among massive stars at lower Z. Of course, direct observations of large numbers of $v\hspace{0.3mm}\sin{\hspace{-0.5mm}i}$ in LMC and SMC clusters are very much needed in order to further substantiate the above results. We may wonder about the origin of the possible higher rotation velocities at low metallicities. This origin is likely related to some metallicity effects in the process of star formation. There are many possibilities, for example we may notice that a lower Z implies less dust and ions in star forming regions. The coupling of the magnetic field to the matter is then weaker, the ambipolar diffusion of the magnetic field should proceed faster, thus leading to less angular momentum losses by the central contracting body. Another possibility is that during pre-main sequence evolution, the weaker opacities at lower Z are leading to an earlier disappearance of the external convective zones and thus to less magnetic coupling between the forming star and its surroundings. Also the importance and the survival lifetime of the accretion disk may increase with metallicity, thus favouring the dissipation of angular momentum at higher Z. Of course, numerical models are needed to examine these various tentative suggestions. Looking ahead we may also mention that the consequence of faster rotation at lower Z may be considerable, with large differences in evolutionary tracks, lifetimes and nucleosynthesis. In particular we know that current models are unable to account for the occurrence of the large quantity of red supergiants in the SMC for which more mixing would be needed (cf. Langer and Maeder \cite{lan}). In the context of nucleosynthesis the possible formation of primary nitrogen at low Z is interesting in relation with the suggestion that primary nitrogen is needed in the early chemical evolution of galaxies (cf. Matteucci and Fran\c{c}ois \cite{mat}; Pagel \cite{pag}). Appropriate stellar models would bring new insight into the interpretation of star populations in low Z galaxies, like blue compact galaxies or galaxies at cosmological distances like in the Hubble Deep Field. \begin{acknowledgements} We thank Dr. Kohoutek for having provided a printed and an e-version of his catalogue to one of us (J.-C. M.). EKG gratefully acknowledges support by NASA through grant HF-01108.01-98A from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. AM expresses his thanks to Dr. Hans Zinnecker for nice discussions on the possible origin of faster rotation at lower metallicities. \end{acknowledgements}
\section{Introduction} \label{sec:introduction} The early years of the next millennium will hopefully be remembered for the birth of Gravitational-wave Astronomy. With several large scale interferometers (LIGO, VIRGO, GEO600, TAMA) under construction, and the continued improvement of the technology for cryogenic resonant-mass detectors (ALLEGRO, AURIGA, EXPLORER, NAUTILUS), there are many reasons to be optimistic at the present time. However, the interpretation of data from the new generation of detectors will heavily depend on accurate ``templates'' of gravitational wave-forms. For a given astrophysical source of gravitational radiation, construction of such templates involves fully non-linear or perturbative approximations to Einstein's field equations. In both instances, the construction of appropriate initial data constitutes a fundamental issue. It is absolutely necessary that initial data represent a ``realistic'' stage of the astrophysical system under consideration. The early years of Numerical Relativity were in part characterized by studies aimed at constructing initial data for Einstein's equations, namely data that satisfies the Hamiltonian and momentum constraints. Of particular interest was obtaining solutions to the constraints which represent black hole binaries \cite{cook}. These initial data studies highlighted the importance that Lichnerowicz-York's conformal approach \cite{york79} plays in facilitating solving the constraints. In the perturbative arena, initial data sets also play a fundamental role if a connection with systems of astrophysical relevance is to be made. Examples of perturbative studies where astrophysically consistent initial data is needed are the point-particle \cite{particle} and close-limit approximations \cite{price94} to black hole coalescences. For the close-limit approximation in particular, this issue is crucial. The focus is then on the late stage of the merger, when the binary system can be approximated as a single, perturbed black hole. In the case of perturbations of relativistic stars, most of the studies \cite{allen} have not considered initial data with direct connection to a given astrophysical situation. In other words, the focus has so far been on investigating how the star reacts to generic perturbations. The goal of this paper is to provide a mechanism for generating initial data for perturbations of a relativistic star. This data should ideally represent astrophysical situations of relevance to gravitational-wave detectors. Inspired by the success that Lichnerowicz-York's conformal approach has enjoyed for solving Einstein's constraints, we base our methodology on ``linearizing Lichnerowicz-York's procedure''. By doing so, we take advantage of the prescription for knowing which pieces of information, among the metric and its ``velocity'', are fixed by the constraints and which are freely specifiable. Moreover, we inherit the machinery used in the past for the construction of initial data sets representing binary systems. That is, the procedure described in this paper provides a natural framework for obtaining initial data in connection with the close-limit approximation to neutron star mergers \cite{nsmergers}. As with black hole binaries, the close-limit to neutron star collisions deals with the late stages of the merger, at the point in which the systems can be approximated as a single neutron star ``dressed'' with perturbations. \section{Linearization of Lichnerowicz-York's Conformal Approach} \label{sec:linearization} Given a spacetime with 4-metric $\hat g_{\mu\nu}$ and a foliation of this spacetime with Eulerian observers having a 4-velocity $\hat n^\mu$, the constraints in Einstein's fields equations can be written in a 3+1 (or ADM) form \cite{adm} as \begin{eqnarray} \hat R + \hat K^2 - \hat K_{ij} \hat K^{ij} & = 16\, \pi\, \hat \sigma\hspace{0.2 true in} &\hbox{(Hamiltonian)} \label{eq:ham} \\ \widehat \nabla_{j}(\hat K^{ij} - \hat h^{ij} \hat K) & = 8\,\pi \hat j^{i}.\hspace{0.2 true in} &\hbox{(Momentum)} \label{eq:mom} \end{eqnarray} Above, $\hat h_{ij}$ is the 3-metric ($\hat h_{\mu\nu} = \hat g_{\mu\nu} + \hat n_\mu\,\hat n_\nu$) and $\hat K_{ij}$ the extrinsic curvature of the time-like hypersurfaces in the foliation, with $\hat K = \hat h^{ij}\hat K_{ij}$ and $\hat R$ the scalar curvature. Furthermore, $\widehat\nabla_{i}$ is the covariant derivative associated with the 3-metric $\hat h_{ij}$, and $\hat \sigma$ and $\hat j^i$ are the energy and momentum densities of the matter sources. Greek letters denote spacetime indexes and Latin letters spatial indices, and we use units in which $G = c = 1$. Contrary to the common practice, we use ``hats'' to denote physical space since we will be mostly working in the conformal space. For a perfect fluid, the stress-energy tensor is given by \begin{equation} \hat T_{\mu\nu} = (\hat\rho+\hat p)\,\hat u_\mu\,\hat u_\nu + \hat p\,\hat g_{\mu\nu}, \end{equation} where $\hat\rho$ and $\hat p$ are respectively the total mass-energy density and pressure of the fluid measure by an observer with 4-velocity $\hat u^\mu$. The energy and momentum densities appearing in the constraints are obtained from \begin{eqnarray} \hat\sigma & = \hat T_{\mu\nu}\,\hat n^\mu\,\hat n^\nu & = (\hat \rho + \hat p)\, \hat\gamma^2 - \hat p \label{eq:sigma}\\ \hat j^\mu & = -\hat T_{\nu\alpha} \,\hat n^\nu \, \hat h^{\alpha\mu} & = (\hat \rho + \hat p)\,\hat\gamma\,\hat h^{\mu\nu}\hat u_\nu \label{eq:j} \end{eqnarray} where $\hat\gamma \equiv -\hat n_\mu\,\hat u^\mu$ is the relativistic boost factor. The fundamental virtue of Lichnerowicz-York's conformal approach \cite{york79} for solving the Hamiltonian and momentum constraints is that it provides a concrete recipe for singling out which four ``pieces'' among the twelve components $(\hat h_{ij},\,\hat K_{ij}$) are to be solved from Eqs.~(\ref{eq:ham}) and (\ref{eq:mom}). The starting point is the Ansatz \begin{equation} \hat h_{ij} = \phi^4 h_{ij}, \label{eq:anzats} \end{equation} where the conformal metric $h_{ij}$ is assumed to be known. Thus, for the metric, the piece that is fixed by the constraints (Hamiltonian) is the conformal factor $\phi$. The other three quantities fixed by the constraints (momentum) involve the extrinsic curvature. The idea here is to decompose the extrinsic curvature into its trace, tracefree-transverse and tracefree-longitudinal parts. To achieve this, the extrinsic curvature is first split into \begin{equation} \hat K^{ij} = \hat A^{ij} + \frac{1}{3} \hat h^{ij} \hat K. \label{eq:kij} \end{equation} Before decomposing the tracefree part $\hat A^{ij}$ into its transverse and longitudinal parts, the following conformal transformation is applied: \begin{equation} \hat A^{ij} = \phi^{-10} A^{ij}. \label{eq:aij} \end{equation} The exponent in the conformal transformation is motivated by the fact that this transformation possesses the following property: \begin{equation} \widehat\nabla_j\hat A^{ij} = \phi^{-10} \nabla_j A^{ij}, \end{equation} with $\nabla_{i}$ covariant differentiation associated with the background metric $h_{ij}$. This property simplifies the conformal transformation of the divergence of the extrinsic curvature in the momentum constraint (\ref{eq:mom}). Since, as will become clear below, the trace of the extrinsic curvature $\hat K$ is not fixed by the constraints, no conformal transformation is imposed on the trace of the extrinsic curvature ($\hat K = K$). Once the conformal transformation is applied, the next step is to decompose $A_{ij}$ into its transverse and longitudinal parts, namely \begin{equation} A^{ij} = A^{ij}_{*} + (\l\,W)^{ij}, \end{equation} where \begin{eqnarray} (\l\,W)^{ij} & = & 2\,\nabla^{(i}W^{j)} - \frac{2}{3} h^{ij} \nabla_{k} W^{k}, \\ \nabla_j A^{ij}_* & = & 0. \end{eqnarray} With the above conformal transformations and transverse-longitudinal decompositions, the Hamiltonian and momentum constraints become \begin{eqnarray} 8 \nabla^i \nabla_{i} \phi - R\, \phi + A_{ij} A^{ij} \phi^{-7} - \frac{2}{3} K^2 \phi^5 + 16\,\pi \sigma\, \phi^{-3} & = & 0 \label{eq:hamc}\\ \nabla^j \nabla_{j} W^{i} + \frac{1}{3} \nabla^i \nabla_j W^j + R^{i}{}_{j} W^j - \frac{2}{3} \phi^6\, \nabla^{i} K - 8\,\pi\, j^{i}& =& 0\,, \label{eq:momc} \end{eqnarray} where $R^{i}{}_{j}$ is the 3-Ricci tensor of the conformal space and $R$ its trace. In deriving Eqs.~(\ref{eq:hamc}) and (\ref{eq:momc}), the following conformal transformations for the energy and momentum densities were used: \begin{eqnarray} \hat\sigma & = & \phi^{-8}\,\sigma \label{eq:csigma}\\ \hat j^i & = & \phi^{-10}\, j^i. \label{eq:cj} \end{eqnarray} To summarize, the Hamiltonian constraint fixes the conformal factor $\phi$ and the momentum constraint determines the generator $W^i$ of the longitudinal part of the conformal, traceless part of the extrinsic curvature. The freely specifiable data in this coupled set of equations are the conformal metric $h_{ij}$, the trace of the extrinsic curvature $K$, the source functions $(\sigma,\, j^i)$, and the divergence free, traceless part of the extrinsic curvature $A^{ij}_*$, which is hidden in $A_{ij}$ in Eq.~(\ref{eq:hamc}). Thus far, we have just reviewed Lichnerowicz-York's treatment of the initial data problem. We now introduce our {\em first} assumption. The initial data, $(\hat h_{ij},\, \hat A^{ij},\, \hat K,\, \hat\sigma,\, \hat j^i)$, are assumed to be close to a given background, i.e., \begin{eqnarray} \hat h_{ij} & = & \hat h_{ij}^{(0)} + \hat h_{ij}^{(1)}\\ \hat A^{ij} & = & \hat A^{ij}_{(0)} + \hat A^{ij}_{(1)}\\ \hat K & = & \hat K_{(0)} + \hat K_{(1)}\\ \hat \sigma & = & \hat \sigma_{(0)} + \hat \sigma_{(1)}\\ \hat j^i & = & \hat j^i_{(0)} + \hat j^i_{(1)}, \end{eqnarray} where ${}^{(0)}$ labels background quantities and ${}^{(1)}$ first-order perturbations. The corresponding decomposition of the conformal space quantities yields: \begin{eqnarray} h_{ij} & = & h_{ij}^{(0)} + h_{ij}^{(1)} \\ \phi & = & \phi^{(0)} + \phi^{(1)} \\ A^{ij}_{*} & = & A^{ij}_{*(0)} + A^{ij}_{*(1)}\\ W^i & = & W^i_{(0)} + W^i_{(1)} \\ K & = & K^{(0)} + K^{(1)} \\ \sigma & = & \sigma^{(0)} + \sigma^{(1)} \\ j^i & = & j^i_{(0)} + j^i_{(1)}, \end{eqnarray} where \begin{eqnarray} \hat h_{ij}^{(0)} & = & \phi^4_{(0)}\, h_{ij}^{(0)}\\ \hat h_{ij}^{(1)} & = & \phi^4_{(0)}\, h_{ij}^{(1)} + 4\,\phi^3_{(0)}\,\phi_{(1)}\,h_{ij}^{(0)}\\ \hat A^{ij}_{(0)} & = & \phi^{-10}_{(0)}\, A^{ij}_{(0)}\\ \hat A^{ij}_{(1)} & = & \phi^{-10}_{(0)}\, A^{ij}_{(1)} - 10\,\phi^{-11}_{(0)}\,\phi_{(1)}\,A^{ij}_{(0)}\\ \hat K_{(0)} & = & K_{(0)}\\ \hat K_{(1)} & = & K_{(1)}\\ \hat \sigma_{(0)} & = & \phi^{-8}_{(0)}\,\sigma_{(0)}\\ \hat \sigma_{(1)} & = & \phi^{-8}_{(0)}\,\sigma_{(1)} - 8\,\phi^{-9}_{(0)}\,\phi_{(1)}\,\sigma_{(0)}\\ \hat j^i_{(0)} & = & \phi^{-10}_{(0)}\,j^i_{(0)}\\ \hat j^i_{(1)} & = & \phi^{-10}_{(0)}\,j^i_{(1)} -10\,\phi^{-11}_{(0)}\,\phi_{(1)}\, j^i_{(0)}. \end{eqnarray} Although all scalars, vectors and tensors are expanded in order of smallness, it is important to stress again that all but $\phi$ and $W^i$ are freely specifiable, and this property is independent of the order of the perturbation. At this point we introduce our {\em second} assumption, which is that the perturbations of the conformal background vanish on the initial data slice; that is, $h_{ij}^{(1)} = 0$. The primary motivation for this choice is the simplification of the coupled system of constraint equations. We shall later discuss the physical relevance, as well as the implications and restrictions, that this assumption imposes on the class of initial data that one has access to with our procedure. Using $h_{ij}^{(1)} = 0$, together with the above perturbative expansions in the coupled elliptic system (\ref{eq:hamc}) and (\ref{eq:momc}), we obtain \begin{eqnarray} 8 \nabla^j \nabla_{j} \phi_{(1)} - \left[ R + 7 A_{(0)}^{ij} A^{(0)}_{ij}\,\phi_{(0)}^{-8} + \frac{10}{3} K^2_{(0)}\,\phi^{4}_{(0)} + 48\,\pi\, \sigma_{(0)} \,\phi^{-4}_{(0)} \right]\phi_{(1)} && \nonumber\\ +\, 2\,A^{(0)}_{ij} A^{ij}_{(1)}\,\phi^{-7}_{(0)} - \frac{4}{3} K_{(0)} K_{(1)}\,\phi^{5}_{(0)} - 16\,\pi\, \sigma_{(1)}\,\phi^{-3}_{(0)} & = & 0 \label{eq:haml} \\ \nabla^j \nabla_{j} W^{i}_{(1)} + \frac{1}{3} \nabla^i \nabla_j W^j_{(1)} + R^{i}{}_j W^j_{(1)} - \frac{2}{3} \phi^6_{(0)} \nabla^{i} K_{(1)} - \frac{12}{3} \phi^5_{(0)}\,\phi_{(1)}\nabla^{i} K_{(0)} - 8\,\pi \, j^{i}_{(1)} &=& 0 \label{eq:moml} \end{eqnarray} where $R$, $R_{ij}$ and $\nabla_i$ refer to the background. In writing Eqs.~(\ref{eq:haml}) and (\ref{eq:moml}), we have used the fact that the zeroth-order quantities satisfy the constraints. Notice that, as in the non-linear case, the constraints remain coupled. \section{The Initial Data Problem for Relativistic Stellar Perturbations} We now focus on constructing initial data sets for which the background is a static and spherically symmetric stellar model, with 4-metric given by \begin{equation} ds^2 = -e^{2\nu}\,dt^2 + e^{2\lambda}\,d\hat r^2 + \hat r^2\,d\theta^2 + \hat r^2 \sin^2 \theta\,d \varphi^2, \label{eq:4metric} \end{equation} where the metric coefficients $\nu$ and $\lambda$ are functions of the radial coordinate $\hat r$ only. Einstein's equations for this background reduce to solving three equations. The first equation defines the ``mass inside radius $\hat{r}$; \begin{equation} \frac{dm}{d\hat r} = 4\,\pi\,\hat r^2\,\hat\rho_{(0)} \ . \label{eq:mass} \end{equation} Here $m$ is a function of $\hat r$ that is related to the metric function $\lambda$ by \begin{equation} e^{-2\lambda} \equiv 1 - \frac{2\,m}{\hat r}. \end{equation} Equation~(\ref{eq:mass}) is directly obtained from the Hamiltonian constraint. The second equation belongs to Einstein's evolution equations and reads \begin{equation} \frac{d\nu}{d\hat r} = \frac{e^{2\lambda}}{\hat r^2} \left(m + 4\,\pi\,\hat r^3\, \hat p_{(0)}\right). \label{eq:nu} \end{equation} Finally, conservation of momentum yields the condition for hydrostatic equilibrium: \begin{equation} \frac{d\hat p_{(0)}}{d\hat r} = - (\hat \rho_{(0)}+\hat p_{(0)})\,\frac{d\nu}{d\hat r} \label{eq:press} \end{equation} The above stellar structure system of equations (the Tolman-Oppenheimer-Volkoff equations) must be supplemented with an equation of state. For simplicity, we use the polytropic equation of state $\hat p = \kappa\,\hat\rho^{\Gamma}$, where $\kappa$ and $\Gamma$ are the adiabatic constant and index, respectively. The adiabatic index $\Gamma$ is related to the polytropic index $n$ by $\Gamma = 1+1/n$. In the specific example provided later, we use $\Gamma = 2$ ($n = 1$). We now assume that the zero-order quantities ($\hat \rho_{(0)},\, \hat p_{(0)},\, \lambda,\, \nu$) have been obtained from solving Eqs.~(\ref{eq:mass}), (\ref{eq:nu}) and (\ref{eq:press}). The next step is to solve for the first-order perturbations from the linearized constraints (\ref{eq:haml}) and (\ref{eq:moml}). To facilitate this task, it is convenient to perform a coordinate transformation and bring the 3-metric in (\ref{eq:4metric}) into the isotropic, conformally flat form: \begin{equation} ds^2 = \phi^4_{(0)} ( dr^2 + r^2\,d\theta^2 + r^2 \sin^2 \theta\,d \varphi^2), \label{eq:cmetric} \end{equation} where the conformal factor is given by \begin{equation} \phi_{(0)} = \left(\frac{\hat r}{r} \right)^{1/2} \label{eq:cfac} \end{equation} and the transformation of the radial-coordinate is obtained from \begin{equation} \frac{dr}{d\hat r} = e^{\lambda}\, \frac{r}{\hat r}. \label{eq:drdr} \end{equation} The static nature of the background spacetime and the gauge choice of a vanishing shift vector imply $K^{(0)}_{ij} = j^i_{(0)} = 0$ for this background spacetime. This leads to considerable simplifications in the following, but the procedure we describe for constructing perturbative initial data can easily be extended to include also time-dependent, spherically symmetric background spacetimes. We introduce here our {\em third} assumption, which is that $A^{ij}_{*(1)} = K_{(1)} = 0$. The vanishing of the transverse-traceless part of the perturbation to the extrinsic curvature has direct implications to the gravitational radiation content of the initial data. In a way this assumption can be viewed as ``minimizing'' the amount of gravitational waves in the initial spacetime. That this may not be desirable is obvious, but in order to be able to assign this part of the free data physically correct values we need a detailed knowledge of the past history of the system. Such information requires long term, nonlinear evolutions and is far beyond our present capabilities. The present assumption is convenient in that it simplifies the calculations considerably. Furthermore, we are unlikely to overestimate the amount of gravitational radiation emerging from true physical systems if we base our estimates on the present approach. The vanishing of the trace of the extrinsic curvature to first order is less restrictive physically. It simply implies (using also the fact that $K_{(0)} = 0$) that the slicing of the perturbed spacetime is maximal to first order. With those further assumptions, the linearized constraints (\ref{eq:haml}) and (\ref{eq:moml}) decouple and take the form \begin{eqnarray} \nabla^j \nabla_{j} \phi_{(1)} & = & 6\,\pi\, \sigma_{(0)} \,\phi^{-4}_{(0)} \phi_{(1)} -2\,\pi\,\phi^{-3}_{(0)}\sigma_{(1)} \label{eq:haml2} \\ \nabla^i \nabla_{i} U &= &\nabla_i V^i \label{eq:u}\\ \nabla^j \nabla_{j} V^i & = & 8\,\pi\,j^i_{(1)} \label{eq:v}. \label{eq:moml2} \end{eqnarray} where we have decomposed the vector $W^i_{(1)}$ in Eq.~(\ref{eq:moml}) following \cite{bowen,bowen-york} as \begin{equation} W^i_{(1)} = V^i - \frac{1}{4}\nabla^iU . \label{eq:wsplit} \end{equation} From now on, we will drop the label ${}_{(1)}$ in $j^i_{(1)}$ and $W^i_{(1)}$ since the zero-order values for these quantities vanish. The set of equations~(\ref{eq:haml2}-\ref{eq:moml2}) constitute a coupled set of elliptic equations in three dimensions, expressed e.g., in the coordinates $(r,\theta,\varphi)$ of the background space. Due to the spherically symmetric nature of the background, these linearized equations allow for a separation of variables. Specifically, we can apply a spherical harmonic decomposition of form: \begin{eqnarray} \phi_{(1)}(r,\theta,\varphi) &=& \sum_{lm} \phi_{(1)}(r)\,Y_{lm}(\theta,\varphi)\\ \sigma_{(1)}(r,\theta,\varphi) &=& \sum_{lm} \sigma_{(1)}(r)\,Y_{lm}(\theta,\varphi)\\ U(r,\theta,\varphi) & = & \sum_{lm} U(r)\,Y_{lm}(\theta,\varphi) \label{eq:uscalar}\\ V^i(r,\theta,\varphi) & = & \sum_{lm} V_1(r)\,e^i_1(\theta,\varphi) + r\,V_2(r)\,e^i_2(\theta,\varphi) + r\,V_3(r)\,e^i_3(\theta,\varphi) \label{eq:vvec} \\ W^i(r,\theta,\varphi) & = & \sum_{lm} W_1(r)\,e^i_1(\theta,\varphi) + r\,W_2(r)\,e^i_2(\theta,\varphi) + r\,W_3(r)\,e^i_3(\theta,\varphi) \label{eq:wvec} \\ j^i(r,\theta,\varphi) & = & \sum_{lm} J_1(r)\,e^i_1(\theta,\varphi) + r\,J_2(r)\,e^i_2(\theta,\varphi), + r\,J_3(r)\,e^i_3(\theta,\varphi), \label{eq:jvec} \end{eqnarray} where \begin{eqnarray} e^i_1 & = & (Y_{lm},\, 0,\, 0) \nonumber \\ e^i_2 & = & \left(0,\, \frac{1}{r^2}\,\partial_\theta Y_{lm},\, \frac{1}{r^2\,\sin^2\theta}\,\partial_\varphi Y_{lm}\right) \\ e^i_3 & = & \left(0,\, -\frac{1}{r^2\,\sin\theta}\,\partial_\varphi Y_{lm},\, \frac{1}{r^2\,\sin\theta}\,\partial_\theta Y_{lm}\right). \nonumber \end{eqnarray} Here $e^i_1$ and $e^i_2$ are the basis vectors of even-parity perturbations and $e^i_3$ is the basis for odd-parity perturbations. In the above expressions and what follows, it is understood that the radial functions are for a given $(l,m)$. These indices have been suppressed for economy in notation. We note that in terms of the radial functions, Eq.~(\ref{eq:wsplit}) reduces to \begin{eqnarray} W_1 & = & V_1 - \frac{1}{4}\,\frac{d}{dr}U \nonumber\\ W_2 & = & V_2 - \frac{1}{4}\,U \label{eq:ww}\\ W_3 & = & V_3 \nonumber. \end{eqnarray} After separation of variables, the system of equations~(\ref{eq:haml2}-\ref{eq:moml2}) is rewritten as a system of coupled radial elliptic equations: \begin{eqnarray} \frac{1}{r^2}\frac{d}{dr}\left(r^2\frac{d}{dr}\phi_{(1)}\right) - \left[\frac{\l(\l+1)}{r^2} + 6\,\pi\, \sigma_{(0)} \,\phi^{-4}_{(0)}\right]\,\phi_{(1)} & = & -2\,\pi\,\phi^{-3}_{(0)}\,\sigma_{(1)} \label{eq:hamr} \\ \frac{1}{r^2}\frac{d}{dr}\left(r^2\frac{d}{dr}U\right) - \frac{\l(\l+1)}{r^2}\,U & = & \frac{1}{r^2}\frac{d}{dr}\left(r^2\,V_1\right) - \frac{\l(\l+1)}{r}\,V_2 \label{eq:uu}\\ \frac{1}{r^2}\frac{d}{dr}\left(r^2\frac{d}{dr}V_1\right) - \frac{[\l(\l+1)+2]}{r^2}\,V_1 + 2\,\frac{\l(\l+1)}{r^2}\,V_2 & = & 8\,\pi\,J_1 \label{eq:v1} \\ \frac{1}{r^2}\frac{d}{dr}\left(r^2\frac{d}{dr}V_2\right) - \frac{\l(\l+1)}{r^2}\,V_2 + \frac{2}{r^2}\,V_1 & = & 8\,\pi\,J_2 \label{eq:v2}\\ \frac{1}{r^2}\frac{d}{dr}\left(r^2\frac{d}{dr}V_3\right) - \frac{\l(\l+1)}{r^2}\,V_3 & = & 8\,\pi\,J_3 \label{eq:v3}. \end{eqnarray} Equations~(\ref{eq:hamr}-\ref{eq:v3}) fully characterize, for each $(l,m)$ harmonic, initial data $(\phi_{(1)},U,V_{i})$ to first perturbative order, once the background conformal factor $\phi_{(0)}$ and density $\sigma_{(0)}$ as well as the fluid perturbations $\sigma_{(1)}$ and $J_{i}$ are specified. Outside the sources, the solutions to the above equations are $U = u\,r^{a}$, $V_1 = v_1\,r^{b}$, $V_2 = v_2\,r^{c}$ and $V_3 = v_3\,r^{d}$ with $u$, $v_1$, $v_2$ and $v_3$ constants. In order to have regular solutions for $r \rightarrow \infty$, one needs $a = d = -(\l+1)$, $b = c = -l$ and $v_1 = \l\,v_2$, where we assume that $\l \ge 1$. As we shall later see, the corresponding interior solutions for head-on and inspiral close-limit collisions exhibit the same scaling; that is, $V_1 = \l\,V_2$ as well as $J_1 = \l\,J_2$. With this assumption, the system of equations (\ref{eq:uu}-\ref{eq:v3}) reduces to \begin{eqnarray} \frac{1}{r^2}\frac{d}{dr}\left(r^2\frac{d}{dr}U\right) - \frac{\l(\l+1)}{r^2}\,U & = & \frac{d}{dr}V_1 - \frac{(\l-1)}{r}\,V_1 \label{eq:uu2}\\ \frac{1}{r^2}\frac{d}{dr}\left(r^2\frac{d}{dr}V_1\right) - \frac{\l(\l-1)}{r^2}\,V_1 & = & 8\,\pi\,J_1 \label{eq:vv2}\\ \frac{1}{r^2}\frac{d}{dr}\left(r^2\frac{d}{dr}V_3\right) - \frac{\l(\l+1)}{r^2}\,V_3 & = & 8\,\pi\,J_3 \label{eq:vv3}. \end{eqnarray} \section{Correspondence with Regge-Wheeler variables} \label{sec:regge} Before we proceed to present examples of initial data sets constructed using the above approach, we want to establish the correspondence between our variables and the standard Regge-Wheeler variables. This is relevant since the Regge-Wheeler notation (and the associated gauge) is customarily used in perturbative evolutions for spherical stellar models \cite{allen}. To this end, consider a spatial tensor $T_{ij}$ such that $T_{ij}(\hat r,\,\theta,\,\varphi) = T_{ij}^{(0)}(\hat r) + T_{ij}^{(1)}(\hat r,\,\theta,\,\varphi)$. The perturbations of this tensor can be decomposed as \begin{equation} T^{(1)}_{ij} = t_1(\hat r)\,f^1_{ij} + \hat r\,t_2(\hat r)\,f^2_{ij} + \hat r^2\,t_3(\hat r)\,f^3_{ij} + \hat r^2\,t_4(\hat r)\,f^4_{ij} + t_5(\hat r)\,f^5_{ij} + \hat t_6(\hat r)\,f^6_{ij}, \label{eq:even} \end{equation} where \begin{equation} \mbox{$f^1_{ij}$} = \left(\matrix{ Y_{lm} & 0 & 0\cr 0 & 0 & 0 \cr 0 & 0 & 0 \cr }\right), \label{eq:f1} \end{equation} \begin{equation} \mbox{$f^2_{ij}$} = \left(\matrix{ 0 & \partial_\theta Y_{lm} & \partial_\varphi Y_{lm} \cr \hbox{symm} & 0 & 0 \cr \hbox{symm} & 0 & 0 \cr }\right), \label{eq:f2} \end{equation} \begin{equation} \mbox{$f^3_{ij}$} = \left(\matrix{ 0 & 0 & 0\cr 0 & Y_{lm} & 0 \cr 0 & 0 & \sin^2\theta\,Y_{lm} \cr }\right), \label{eq:f3} \end{equation} \begin{equation} \mbox{$f^4_{ij}$} = \left(\matrix{ 0 & 0 & 0\cr 0 & \partial^2_\theta\,Y_{lm} & (\partial_\theta\partial_\varphi - \cot\theta\partial_\varphi)\,Y_{lm} \cr 0 & \hbox{symm} & (\partial^2_\varphi + \sin\theta\,\cos\theta\, \partial_\theta)\,Y_{lm} \cr }\right) \label{eq:f4} \end{equation} \begin{equation} \mbox{$f^5_{ij}$} = \left(\matrix{ 0 & -\frac{1}{\sin\theta}\partial_\varphi Y_{lm} & \sin\theta\partial_\theta Y_{lm} \cr \hbox{symm} & 0 & 0 \cr \hbox{symm} & 0 & 0 \cr }\right), \label{eq:f5} \end{equation} \begin{equation} \mbox{$f^6_{ij}$} = \left(\matrix{ 0 & 0 & 0\cr 0 & \frac{1}{\sin\theta} (\partial_\theta\partial_\varphi-\cot\theta)\,Y_{lm} & \frac{\sin\theta}{2} \left(\frac{1}{\sin^2\theta} \partial^2_\varphi + \cot\theta\partial_\theta -\partial^2_\theta \right)\,Y_{lm} \cr 0 & \hbox{symm} & -\sin\theta(\partial_\theta\partial_\varphi - \cot\theta \partial_\varphi)\,Y_{lm} \cr }\right). \label{eq:f6} \end{equation} Above, $f^1_{ij}$, $f^2_{ij}$, $f^3_{ij}$, and $f^4_{ij}$ represent the even-parity tensor spherical harmonics and $f^5_{ij}$ and $f^6_{ij}$ the odd-parity counterparts. Using the Regge-Wheeler notation, the perturbations of the spatial metric (in physical space) read \begin{equation} \hat h_{ij}^{(1)} = \sum_{lm} e^{2\lambda}\,H_2(\hat r)\,f^1_{ij} + h_1^{even}(\hat r)\,f^2_{ij} + \hat r^2\,K(\hat r)\,f^3_{ij} + \hat r^2\,G(\hat r)\,f^4_{ij} + h_1^{odd}(\hat r)\,f^5_{ij} + h_2(\hat r)\,f^6_{ij}, \label{eq:rw} \end{equation} where $K$ must not be confused with the trace of the extrinsic curvature. From the previous section, we have that the spatial metric can be constructed from \begin{eqnarray} \hat h_{ij} & = & \hat h_{ij}^{(0)} + \hat h_{ij}^{(1)} = \phi^4\, h_{ij} = (\phi_{(0)} + \phi_{(1)})^4\,\, h_{ij}^{(0)}\nonumber \\ & = & \phi_{(0)}^4\,h_{ij}^{(0)} + 4\,\phi_{(0)}^3\phi_{(1)}\, h_{ij}^{(0)} \\ & = & \hat h_{ij}^{(0)} + 4\,\frac{\phi_{(1)}}{\phi_{(0)}}\,\hat h_{ij}^{(0)}, \end{eqnarray} where $\phi_{(0)} = (\hat r/r)^{1/2}$, $\hat h^{(0)}_{ij} = \hbox{diag}( e^{2\lambda},\,\hat r^2,\, \hat r^2\,\sin^2\theta)$ and $\phi_{(1)} = \sum_{lm} \phi_{(1)}(r)\,Y_{lm}$ with $\phi_{(1)}(r)$ a solution of the radial equation (\ref{eq:hamr}). Thus, our approach to construct initial data yields spatial metric perturbations of the form \begin{equation} \hat h_{ij}^{(1)} = \sum_{lm} 4\,\phi_{(1)}(r)\left(\frac{\hat r}{r}\right)^{1/2} (e^{2\lambda}\,f^1_{ij} + \hat r^2\,f^3_{ij}). \label{eq:hrw} \end{equation} Comparison of the metric perturbations (\ref{eq:hrw}) with (\ref{eq:rw}) shows that our procedure for constructing initial data yields in terms of the Regge-Wheeler notation $h_1^{odd} = h_1^{even} = h_2 = G = 0$ and \begin{equation} H_2 = K = 4\,\phi_{(1)}(r)\left(\frac{\hat r}{r}\right)^{1/2}. \end{equation} Consider now the extrinsic curvature. As mentioned before, the extrinsic curvature vanishes to zero-order. In addition, we made the choice of having vanishing first-order trace $K_{(1)}$ and transverse-traceless parts $A^{ij}_{*(1)}$. Thus, the extrinsic curvature is completely determined by the vector $W^i$ and the conformal factor of the background space $\phi_{(0)}$ from \begin{equation} \hat K_{ij} = \phi^{-2}_{(0)}\,(\l W)_{ij}. \end{equation} In terms of tensor spherical harmonics, the extrinsic curvature reads \begin{equation} \hat K_{ij} = \sum_{lm} \left(\frac{r}{\hat r}\right) \left[ k_1(r)\,f^1_{ij} + r\,k_2(r)\,f^2_{ij} + r^2\,k_3(r)\,f^3_{ij} + r^2\,k_4(r)\,f^4_{ij} +k_5(r)\,f^5_{ij} + k_6(r)\,f^6_{ij} \right], \label{eq:krw} \end{equation} where \begin{eqnarray} k_1 & = & \frac{2}{3\,r}\left[2\,r\,\frac{d}{dr} W_1 + \l(\l+1)\,W_2 - 2\,W_1\right] \nonumber \\ k_2 & = & \frac{1}{r}\left[r\,\frac{d}{dr}W_2 + W_1 - W_2\right] \nonumber \\ k_3 & = & \frac{2}{3\,r}\left[-r\,\frac{d}{dr}W_1 + W_1 + l(l+1)\,W_2\right] \\ k_4 & = & \frac{2}{r}\,W_2 \nonumber \\ k_5 & = & r\,\frac{d}{dr}W_3 - W_3 \nonumber \\ k_6 & = & -2\,r\,W_3\nonumber . \\ \end{eqnarray} Before proceeding, it is appropriate to discuss the implications of the assumptions we imposed for the presence of even and odd parity perturbations in the initial data. Consider first the perturbations of the 3-metric: The assumption of vanishing perturbations of the conformal 3-metric implies (as seen from Eq.~(\ref{eq:hrw})), that $h_1^{odd} = h_1^{even} = h_2 = G = 0$. This means that all odd parity perturbations of the 3-metric must vanish on the initial surface. This does not however exclude the presence of odd-parity perturbations. Such perturbations may enter via the Regge-Wheeler variable $h_0$, which is freely specifiable. Similarly for even-parity perturbations, the Regge-Wheeler quantities $H_0$ and $H_1$ are not part of the constraints and can be chosen freely. A choice of these three variables correspond to choosing a slicing for the spacetime, i.e. specifying the lapse and the shift vector. Specifically, Regge-Wheeler gauge corresponds to $h_1^{even}=G=h_2=0$ above, as well as perturbed lapse \begin{equation} \delta \alpha = e^{2\nu}H_0 \ , \end{equation} and shift vector \begin{equation} \delta \beta^i = H_1 e_1^i + h_0 e_3^i \ . \end{equation} Furthermore, there are no constraints on the odd or even parity nature of the extrinsic curvature initial data. This follows immediately from Eq.~(\ref{eq:krw}) since the coefficients $k_1$ through $k_6$ are in general non-vanishing. As a consequence, even if the initial 3-metric has vanishing odd-parity perturbations, their time evolution will in general include such perturbations. \section{Sample Initial data sets: Colliding neutron stars} We will now apply the method for constructing initial data, that was presented in Sec.~III, to a case of astrophysical relevance. We consider collisions of neutron stars under the close-limit approximation. Related results have already been discussed by Allen et al \cite{nsmergers}. That study was specialized to the case of head-on collisions. Here, we present a general discussion of neutron-star close-limit initial data, and give explicit results for both boosted head-on and inspiralling collisions. As stated at the end of Sec.~III, initial data are obtained by solving the system of equations equations~(\ref{eq:hamr}-\ref{eq:v3}). The input for these equations are the background conformal factor $\phi_{(0)}$ and density $\sigma_{(0)}$ and, in addition, the fluid perturbations $\sigma_{(1)}$ and $J_{i}$. For close-limit collisions, one specifies the background from the outcome of the collisions. The fluid perturbations $\sigma_{(1)}$ and $J_{i}$, on the other hand, are obtained by ``subtracting" the background from a suitable superposition of stars that represent the initial configuration. A suitable way of relating the two initial stars to the final configuration was presented in \cite{nsmergers}, and we refer the reader to that paper for further details. Let us first consider the perturbation to the background density. Neglecting complicating factors, such as the effects from tidal deformations, we approximate the total density of the binary system with the following superposition of density profiles of isolated neutron stars \cite{nsmergers}: \begin{equation} \sigma( r^i) = \sigma_*( r^i-\xi^i)+\sigma_*( r^i+\xi^i) -\left[\sigma_*( r^i-\xi^i)\,\sigma_*( r^i+\xi^i)\right]^{1/2}\,, \label{eq:rhotot} \end{equation} where $\sigma_*$ is the conformally transformed density profile of the colliding neutron stars in isolation located a distant $\xi^i$ in conformal space. This functional form (\ref{eq:rhotot}) for the total density is chosen since it leads to the correct zero-separation and infinite-separation limits. That is, for zero-separation $\sigma \rightarrow \sigma_*$, and for large separations $\sigma(r^i) \rightarrow \sigma_*( r^i-\xi^i)+\sigma_*( r^i+\xi^i)$. We now introduce the close-limit approximation, $\xi^i \ll 1$, and write \begin{equation} \sigma_*( r^i\pm\xi^i) = \sigma_*( r^i) \pm \xi^i\,\nabla_i\,\sigma_*( r^i) +\frac{1}{2}\,\xi^i\,\xi^j\,\nabla_i\, \nabla_j\,\sigma_*( r^i)\,. \label{eq:rhotaylor} \end{equation} Therefore, Eq.~(\ref{eq:rhotot}) takes the form \begin{equation} \sigma = \sigma_* + \frac{1}{2}\,(\xi^i\,\nabla_i\,\sigma_*)^2 + \frac{1}{2}\,\xi^i\,\xi^j\,\nabla_i\,\nabla_j\,\sigma_* \,. \end{equation} Given this total energy, the density perturbation is obtained from \begin{eqnarray} \sigma_{(1)} & = &\sigma-\sigma_{(0)} \nonumber \\ & = & \sigma_* - \sigma_{(0)} + \frac{1}{2}\,(\xi^i\,\nabla_i\,\sigma_*)^2 + \frac{1}{2}\,\xi^i\,\xi^j\,\nabla_i\,\nabla_j\,\sigma_*\,. \label{eq:rho11} \end{eqnarray} Similarly, the momentum density is given by the superposition of momentum densities of boosted isolated stars: \begin{equation} j^i(r^k) = j^i_+(r^k+\xi^k) + j^i_-(r^k-\xi^k) \,. \label{eq:jtot} \end{equation} In this case, there is no need for a counterpart of the last term in (\ref{eq:rhotot}). For both, head-on and inspiral collisions, $j^i_+ = - j^i_- = j^i_*$; thus, \begin{equation} j^i(r^k) = j^i_*(r^k+\xi^k) - j^i_*(r^k-\xi^k) \,, \label{eq:jtot2} \end{equation} which obviously has the appropriate zero-separation limit, namely $j^i(r^k) \rightarrow 0$ as $\xi^i \rightarrow 0$. Once again, we apply the close-limit condition and approximate \begin{equation} j^i_*( r^k\pm\xi^k) = j^i_{*}( r^k) \pm \xi^j\,\nabla_j\,j^i_{*}(r^k) +\frac{1}{2}\,\xi^l\,\xi^m\,\nabla_l\,\nabla_m\,j^i_{*}( r^k)\,. \end{equation} Therefore, \begin{equation} j^i = 2\,\xi^j\,\nabla_j\,j^i_{*}\,. \label{eq:j11} \end{equation} Notice that Eq.~(\ref{eq:j11}) directly gives the momentum density perturbation because, by construction, the background is static. \subsection{Head-on Collision Initial Data} For simplicity, we assume that the collision takes place along the $z$-axis. Therefore, \begin{eqnarray} \xi^i &=& \xi\,z^i \label{eq:vecxi} \\ j^i_* &=& -J_*\,z^i \label{eq:vecj}\,, \end{eqnarray} where \begin{equation} z^i = \left(\cos{\theta},\,-\frac{1}{ r}\,\sin{\theta},\, 0\right)\, \label{eq:vecz} \end{equation} is a unit vector along the $z$-axis. Substitution of (\ref{eq:vecxi}) into (\ref{eq:rho11}) yields \begin{eqnarray} \sigma_{(1)} &=& \sigma_* - \sigma_{(0)} + \frac{1}{2}\,\xi^2\,\left[\cos^2{\theta}\,\left(\frac{d}{d r}\sigma_*\right)^2 +\cos^2{\theta}\,\frac{d^2}{d r^2}\sigma_* +\sin^2{\theta}\,\frac{1}{ r}\frac{d}{d r}\sigma_*\right] \,. \nonumber \\ &=& \sqrt{4\,\pi}\,Y_{00}\, \bigg[ \sigma_* - \sigma_{(0)} + \frac{1}{6}\,\xi^2\,\biggl\lbrace\frac{d^2}{d r^2}\sigma_* + \left(\frac{d}{d r}\sigma_*\right)^2 + \frac{2}{ r}\frac{d}{d r}\sigma_*\biggr\rbrace\bigg] \nonumber \\ &+& \frac{2}{15}\,\sqrt{5\,\pi}\,Y_{20}\,\xi^2\, \left[\frac{d^2}{d r^2}\sigma_* + \left(\frac{d}{d r}\sigma_*\right)^2 -\frac{1}{ r}\frac{d}{d r}\sigma_*\right] \, . \label{eq:rho_b} \end{eqnarray} It is clear from (\ref{eq:rho_b}) that the conformal density perturbation has a monopole contribution ($m=0,\, \l= 0$). Since our main interest is associated with the gravitational waves generated during the merger, we will ignore this contribution; it obviously does not lead to any gravitational waves. The remaining part is the radiative quadrupole ($m=0,\, \l= 2$) perturbation: \begin{equation} \sigma_{(1)} = \frac{2}{15}\,\sqrt{5\,\pi}\,\xi^2\, \left[\frac{d^2}{d r^2}\sigma_* + \left(\frac{d}{d r}\sigma_*\right)^2 -\frac{1}{ r}\frac{d}{d r}\sigma_*\right]\,. \label{eq:rho_final} \end{equation} This initial data set is discussed in considerable detail in \cite{nsmergers}. Let us now consider the momentum density perturbation. Substitution of (\ref{eq:vecj}) into (\ref{eq:j11}) yields \begin{eqnarray} j^i &=& -2\,\xi\,z^j\,\nabla_j(\,J_*\,z^i)\nonumber \\ &=& -2\,\xi\,z^i\,z^j\,\nabla_j\,J_* \nonumber \\ &=& -2\,\xi\,\frac{d}{dr}J_*\,\left(\cos^2{\theta}, -\frac{1}{r}\sin{\theta}\,\cos{\theta},0 \right)\nonumber \\ &=& -\frac{4}{3}\,\sqrt{\pi}\,\xi\,\frac{d}{dr}J_*\,\left(Y_{00},0,0\right) -\frac{8}{15}\,\sqrt{5\,\pi}\,\xi\,\frac{d}{dr}J_*\,\left( Y_{20}, \frac{2}{r}\partial_\theta Y_{20}, 0\right)\,. \label{eq:jhead} \end{eqnarray} As expected and in agreement with the density perturbation, we again have a monopole momentum density perturbation that we shall ignore. Comparing the quadrupole terms in (\ref{eq:jhead}) with (\ref{eq:jvec}), we have $J_3 = 0$ and \begin{equation} J_1 = 2\,J_2 = -\frac{8}{15}\,\sqrt{5\,\pi}\,\xi\,\frac{d}{dr}J_*\,. \label{eq:jhfinal} \end{equation} From Eq.~(\ref{eq:sigma}) and (\ref{eq:j}) and the conformal transformations (\ref{eq:csigma}) and (\ref{eq:cj}), the expressions for $\sigma_*$ and $J_*$ in Eqs.~(\ref{eq:rho_final}) and (\ref{eq:jhfinal}) are given by: \begin{eqnarray} \sigma_* &=& \hat\rho_{*}\,\phi^{8}_{(0)} \label{eq:sigmamag}\\ J_* &=& (\hat\rho_{*} + \hat p_{*})\,v\,\phi^{10}_{(0)}\,, \label{eq:jmag} \end{eqnarray} with $v$ the magnitude of the collision velocity. In writing the above expressions, we used that to zero-order $\hat u^i = 0$ and $\hat\gamma = 1 + O(\hat u^2)$. Notice that the conformal factor in connection with the background star, $\phi_{(0)}$, was the one used in Eqs.~(\ref{eq:sigmamag}) and (\ref{eq:jmag}) to transform the physical TOV solutions of the colliding stars since it is the background star that provides the conformal space where our calculations are performed. Also important is to notice that, once the background and colliding neutron stars models have been completely determined, there are only two parameters that characterize the initial data: the separation $\xi$ and the velocity $v$. \subsection{Inspiral Collision Initial Data} For this case, we assume that the initial configuration is such that the neutron stars are along the $x$-axis and their momentum pointing along the $y$-axis. Then \begin{eqnarray} \xi^i &=& \xi\,x^i \label{eq:vecxi_r} \\ j^i_* &=& J_*\,y^i \label{eq:vecj_r}\,, \end{eqnarray} where \begin{eqnarray} x^i &=& \left(\sin{\theta}\,\cos{\varphi}, \frac{1}{r}\cos{\theta}\,\cos{\varphi}, -\frac{\sin{\varphi}}{r\,\sin{\theta}}\right) \label{eq:vecx} \\ y^i &=& \left(\sin{\theta}\,\sin{\varphi}, \frac{1}{r}\cos{\theta}\,\sin{\varphi}, \frac{\cos{\varphi}}{r\,\sin{\theta}}\right)\,. \label{eq:vecy} \end{eqnarray} are unit vectors along the $x$-axis and $y$-axis, respectively. Substitution of (\ref{eq:vecxi_r}) and (\ref{eq:vecj_r}) into (\ref{eq:rho11}) yields \begin{eqnarray} \sigma_{(1)} &=& \sigma_* - \sigma_{(0)} + \frac{1}{2}\,\xi^2\,\bigg[\sin^2{\theta}\,\cos^2{\varphi}\,\biggl\lbrace\left(\frac{d}{d r}\sigma_*\right)^2 +\frac{d^2}{d r^2}\sigma_*\biggr\rbrace \nonumber \\ &+&(\cos^2{\theta}\,\cos^2{\varphi}+\sin^2{\varphi})\frac{1}{ r}\frac{d}{d r}\sigma_*\bigg] \nonumber \\ &=& \sqrt{4\,\pi}\,Y_{00}\, \bigg[ \sigma_* - \sigma_{(0)} + \frac{1}{6}\,\xi^2\,\biggl\lbrace\frac{d^2}{d r^2}\sigma_* + \left(\frac{d}{d r}\sigma_*\right)^2 + \frac{2}{ r}\frac{d}{d r}\sigma_*\biggr\rbrace\bigg] \nonumber \\ &+& \left(-\sqrt{\frac{\pi}{5}}\,Y_{20}+ \sqrt{\frac{2\,\pi}{15}}\hbox{Re}Y_{22}\right)\,\xi^2\, \left[\frac{d^2}{d r^2}\sigma_* + \left(\frac{d}{d r}\sigma_*\right)^2 -\frac{1}{ r}\frac{d}{d r}\sigma_*\right] \, . \label{eq:rho_i} \end{eqnarray} As with the head-on collisions, we concentrate on the radiative quadrupole ($m=2,\,l=2$) term, so the density perturbation $\sigma_{(1)}$ source of Eq.~(\ref{eq:hamr}) is given by: \begin{equation} \sigma_{(1)} = \sqrt{\frac{2\,\pi}{15}}\,\xi^2\, \left[\frac{d^2}{d r^2}\sigma_* + \left(\frac{d}{d r}\sigma_*\right)^2 -\frac{1}{ r}\frac{d}{d r}\sigma_*\right]\,. \label{eq:rho_ins} \end{equation} Notice that the only difference between (\ref{eq:rho_ins}) and the corresponding source term in the head-on collision case, i.e. Eq.~(\ref{eq:rho_final}), is a numerical factor. For the momentum density perturbation, substitution of (\ref{eq:vecj_r}) into (\ref{eq:j11}) yields \begin{eqnarray} j^i &=& 2\,\xi\,x^j\,\nabla_j(\,J_*\,y^i)\nonumber \\ &=& 2\,\xi\,y^i\,x^j\,\nabla_j\,J_* \nonumber \\ &=& 2\,\xi\,\frac{d}{dr}J_*\,\left(\sin^2{\theta}\,\sin{\varphi}\,\cos{\varphi}, \frac{1}{r}\sin{\theta}\,\cos{\theta}\,\sin{\varphi}\,\cos{\varphi}, \frac{1}{r}\cos^2{\varphi}\right)\nonumber \\ &=& -\sqrt{\frac{4\,\pi}{3}}\,\xi\,\frac{d}{dr}J_*\, \left(0,0,\frac{1}{r\,\sin{\theta}}\partial_\theta\,Y_{10}\right) + 4\sqrt{\frac{2\,\pi}{15}}\,\xi\,\frac{d}{dr}J_*\,\hbox{Im}\left(Y_{22}, \frac{1}{r}\partial_\theta\,Y_{22}, \frac{1}{r\,\sin^2{\theta}} \partial_\varphi\,Y_{22}\right)\,. \label{eq:jrot1} \end{eqnarray} Comparing (\ref{eq:jrot1}) with (\ref{eq:jvec}), we deduce for the quadrupole term ($m=2,\,l=2$) $J_3 = 0$ and \begin{equation} J_1 = 2\,J_2 = 4\sqrt{\frac{2\,\pi}{15}}\,\xi\,\frac{d}{dr}J_*\,, \label{eq:jquadra} \end{equation} The dipole term ($m=0,\,l=1$) does not contribute to the emerging gravitational radiation and can be ignored. Once again, the momentum density perturbation for the inspiral case only differs from the head-on case by a numerical factor. The only non-trivial differences in the initial data will then arise from the $Y_{lm}$'s since in one case $m=0$ (head-on) and for the other $m=2$ (inspiral). The quantities $\sigma_*$ and $J_*$ in (\ref{eq:rho_ins}) and (\ref{eq:jquadra}) are obtained as in the head-on collision case, namely from Eqs.~(\ref{eq:sigmamag}) and (\ref{eq:jmag}) respectively. Figure~\ref{fig1} shows profiles of the density perturbation $\sigma_{(1)}$ and the momentum density perturbation $J_1$ for the close-limit, boosted, head-on collision. Recall that for inspiral and head-on collisions $J_1 = 2\,J_2$ and $J_3 = 0$. The perturbations $\sigma_{(1)}$ and $J_1$ were calculated from neutron stars with initial separation $\xi = 0.1\,R_{(0)}$ and velocity $v = 0.1\,c$. The corresponding perturbations for the inspiral case only differ from the perturbations shown in Fig.~\ref{fig1} by a constant numerical factor (compare Eqs.~(\ref{eq:rho_final}) and (\ref{eq:jhfinal}) with Eqs.~(\ref{eq:rho_ins}) and (\ref{eq:jquadra}). The TOV parameters for the background and colliding stars are: $\rho^{(0)}_c = 2.69\times 10^{15}\,\hbox{g/cm}^3$ and $\kappa_{(0)} = 100\,\hbox{km}^2 $. For these parameters, the mass and radius of the background star are $M_{(0)} = 1.24 \,M_\odot$ and $R_{(0)} = 9.0\,\hbox{km}$, respectively. The initial colliding stars, which are displaced a distance $0.1R_0$ from the center of mass, follow from $\rho^*_c = 2.98\times 10^{15}\,\hbox{g/cm}^3$ and $K_* = 90.25\,\hbox{km}^2 $. With these parameters, the colliding stars have a mass and radius of $M_* = 1.17 \,M_\odot$ and $R_0* = 8.58\,\hbox{km}$, respectively. Figure~\ref{fig2} shows the solutions to the conformal perturbation $\phi_{(1)}$ and the harmonic components $W_1$ and $W_2$ (see Eq.~\ref{eq:ww}) of the vector $W^i$ for the close-limit collision of neutron stars corresponding to the perturbations in Fig.\ref{fig1}. \section{Concluding remarks} In this paper, we have presented a framework for constructing initial data relevant for perturbative studies of neutron stars. Our approach was to ``linearize'' Lichnerowicz-York's standard procedure for the initial-value problem in General Relativity, and it facilitates (to a certain extent) setting astrophysical initial data for perturbation evolutions, cf. \cite{allen}. It is straightforward to compare our method (as well as the results) to the fully nonlinear one, which is important since a main motivation for perturbation studies is to provide benchmark tests for nonlinear numerical relativity. As examples of interesting initial data that can be constructed from our equations, we constructed data for merging neutron stars in the close-limit approximation. The simplest case of these data sets, that describe head-on collision of two initially static stars, has already been extensively discussed in \cite{nsmergers}. No studies of the more general data with initial momentum and for inspiralling collisions have yet been performed. Such simulations should obviously be carried out, and we hope to be able to discuss the relevant results, as well as possible extensions of the framework developed in this paper to, for example, rotating configurations, in the near future. \section*{Acknowledgments} We thank Johannes Ruoff for his assistance with some of the tensor harmonics manipulations. This work was partially supported by NATO grants CRG960260 and CRG971092, as well as NSF grants PHY9800973 and PHY9357219.
\section{Introduction} Understanding single hole dynamics in quantum antiferromagnets is a decisive step towards a comprehensive description of elementary excitations in strongly correlated systems. Experimental realizations are found in compounds such as $SrCuO_2$ \cite{kim96}, $Na_{0.96}V_2O_5$ \cite{kobayashi99} for chains, $Sr_{14}Cu_{24}O_{41}$ \cite{takahashi97} for ladders and $Sr_2 Cu O_2 Cl_2$ \cite{wells95} for planes. In particular chain compounds attract at present an increasing amount of interest in order to elucidate, whether signals of charge-spin separation as predicted from Luttinger-liquid theory can be observed experimentally. On the other hand, theoretical treatments based on Bethe-{\em Ansatz} (BA) results lead recently to a complete description of the spectral function of the Hubbard model at $U=\infty$ \cite{sorella92} and the low energy sector in the nearest-neighbour (NN) $t$-$J$ model, where explicit results are obtained at the supersymmetric (SuSy) point \cite{sorella96}. Further exact results - apart from exact diagonalizations which suffer from strong finite-size effects - are available only for the inverse-square exchange (ISE) \cite{kato98} $t$-$J$ model at the SuSy point. In order to be able to compare with experiments, it is crucial to extend such studies to realistic values of the parameters and possibly beyond the asymptotic low energy limit. In this letter, we present a simple finite-temperature quantum Monte Carlo (QMC) algorithm capable of dealing with this issue for the NN $t$-$J$ model. For single-hole excitations and in the absence of frustration, the method is free of the notorious sign problem, and applicable to chains, n-leg ladders and planes. Here, we concentrate on chains. Our simulations lead to the conclusion that the overall features of the spectral functions are well described by a charge-spin separation {\em Ansatz} (CSSA) based on a mean-field slave-boson picture \cite{suzuura97}, where the hole spectral function is given by the convolution of the spectral functions of free holons and spinons. The agreement with the simulations is obtained over all energy scales and values of $J/t$ ranging from $1/3$ to $4$. At the SuSy point a more detailed understanding of the spectrum is achieved by supplementing the simple model with BA results. A finite-size scaling on chains up to $L=96$ sites shows that the quasi-particle weight $Z_k$ vanishes as $1/\sqrt{L}$, a result which was beyond numerical capabilities up to now. Our starting point is the NN $t$-$J$ model, \begin{eqnarray} H_{t-J}= -t \sum\limits_{<i,j>,\sigma} \tilde c^{\dagger}_{i,\sigma} \tilde c^{}_{j,\sigma} +J \sum\limits_{<i,j>} \left( \vec S_i\cdot \vec S_j -\frac 1 4 \tilde n_i \tilde n_j \right). \end{eqnarray} Here $\tilde c^{\dagger}_{i,\sigma}$ are projected fermion operators $\tilde c^{\dagger}_{i,\sigma}=(1-c^{\dagger}_{i,-\sigma} c^{}_{i,-\sigma})c^{\dagger}_{i,\sigma}$ , $\tilde n_i=\sum\limits_{\alpha} \tilde c^{\dagger}_{i,\alpha}\tilde c^{}_{i,\alpha}$, $\vec S_i=(1/2)\sum\limits_{\alpha,\beta}c^{\dagger}_{i,\alpha} \vec{\sigma}_{\alpha,\beta}c^{}_{i,\beta}$, and the sum runs over nearest neighbours. After a canonical transformation this model is cast into the form \cite{khaliullin90} \begin{eqnarray} \tilde H_{t-J}= +t \sum\limits_{<i,j>} P_{ij}f^{\dagger}_if^{}_j + \frac J 2 \sum\limits_{<i,j>}\Delta_{ij}(P_{ij}-1) , \end{eqnarray} where $P_{ij}=(1+\vec\sigma_i\cdot\vec\sigma_j)/2$, $\Delta_{ij}=(1-n_i-n_j)$ and $n_i=f^{\dagger}_if^{}_i$. In this mapping, one uses the following identities for the standard creation ($c^{\dagger}_{i,\sigma}$) and annihilation ($c^{}_{i,\sigma}$) operators $ c^\dagger_{i\uparrow} = \gamma_{i,+} f_i - \gamma_{i,-} f_i^\dagger \, , \; \; \; c^\dagger_{i\downarrow} = \sigma_{i,-} (f_i + f_i^\dagger) \, , $ where $\gamma_{i,\pm} = (1 \pm \sigma_{i,z})/2$ and $\sigma_{i,\pm} = (\sigma_{i,x} \pm i \sigma_{i,y})/2$. The spinless fermion operators fulfill the canonical anticommutation relations $\{f_i^\dagger,f_j\} = \delta_{i,j}$, and $\sigma_{i,a}\, , \; a = x,y,$ or $z$ are the Pauli matrices. The constraint to avoid doubly occupied states transforms to the conserved and holonomic constraint $ \sum_i \gamma_{i,-} f^\dagger_i f_i = 0 $. The Green function in the spin up sector may be written as \begin{eqnarray} G_{\uparrow}(i-j, \tau)= \langle T \tilde c^{}_{i,\uparrow}(\tau) \tilde c^{\dagger}_{j,\uparrow} \rangle = \langle T f^{\dagger}_i(\tau) f^{}_j \rangle \end{eqnarray} where $T$ corresponds to the time ordering operator. Inserting complete sets of spin states the quantity above transforms as \begin{eqnarray} -G(i-j,-\tau) = & & \frac{\sum \limits_{\sigma_1} \langle v | \otimes \langle \sigma_1 | e^{-(\beta-\tau)\tilde H_{t-J}}f^{}_j e^{-\tau \tilde H_{t-J}}f^{\dagger}_i | \sigma_1 \rangle \otimes | v \rangle} {\sum \limits_{\sigma_1}\langle \sigma_1 | e^{-\beta\tilde H_{t-J}} | \sigma_1 \rangle} = \nonumber \\ & & \sum\limits_{\vec\sigma} P(\vec\sigma) \times \frac{ \langle v| f^{}_j e^{-\Delta\tau\tilde H(\sigma_n,\sigma_{n-1})} e^{-\Delta\tau\tilde H(\sigma_{n-1},\sigma_{n-2})} \ldots e^{-\Delta\tau\tilde H(\sigma_{2},\sigma_{1})} f^{\dagger}_i |v \rangle } { \langle\sigma_n | e^{-\Delta\tau\tilde H_{t-J}} | \sigma_{n-1}\rangle \ldots \langle \sigma_{2} e^{-\Delta\tau\tilde H_{t-J}} | \sigma_{1}\rangle } +{\cal O}(\Delta\tau^2) \nonumber \\ & & = \sum\limits_{\vec\sigma} P(\vec\sigma) G(i,j,\tau,\vec\sigma) +{\cal O}(\Delta\tau^2) \end{eqnarray} Here $m \Delta\tau=\beta$, $ n \Delta\tau=\tau$, $\Delta\tau t \ll 1$ and $\exp({-\Delta\tau\tilde H(\sigma_1,\sigma_2)})$ is the evolution operator for the holes, given the spin configuration $(\sigma_1,\sigma_2)$. In the case of single hole dynamics $|v\rangle$ is the vacuum state for holes, and $P(\vec\sigma)$ is the probability distribution of a Heisenberg antiferromagnet for the configuration $\vec\sigma$, where $\vec \sigma$ is a vector containing all intermediate states $(\sigma_1,\ldots\sigma_n,\ldots\sigma_{m},\sigma_1)$. The sum over spins is performed in a very efficient way by using a world-line cluster-algorithm \cite{evertz93}. As the evolution operator for the holes is a bilinear form in the fermion operators, $G(x,\tau,\vec\sigma)$ can be calculated exactly. $G(x,\tau,\vec\sigma)$ contains a sum over all possible fermion paths between $(i,0)$ and $(j,\tau)$, where $i-j=x$. This stands in contrast to the worm approach \cite{prokofev98}, where fermion paths are sampled stochastically. The numerical effort to calculate $G(x,\tau,\vec\sigma) \, \forall \, x,\tau$ scales as $L \tau$. Spectral properties are obtained by inverting the spectral theorem \begin{eqnarray} G (k,\tau)=\int\limits_{-\infty}^{\infty} {\rm d}\omega A(k,\omega)\frac{ \exp(-\tau\omega)}{\pi(1+\exp(-\beta\omega))} \end{eqnarray} with the Maximum Entropy method (MEM) \cite{jarrell96}. Since $P(\vec\sigma)$ is the probability distribution for the quantum antiferromagnet, the algorithm does not suffer from sign problems on bipartite lattices and next neighbour interactions in any dimension. However, when the spin and charge dynamics evolve according to very different time scales ($J \ltsim 0.2 t$), $G(x,\tau,\vec\sigma)$ shows an increasing variance. Best results are obtained at the SuSy point and an appreciable range of $J/t$ may be considered as shown below. We now concentrate on the one-dimensional $t$-$J$ model. The simulations were performed at temperatures $T \le {\rm min} (J,t) /15$, such that no appreciable changes with a further decrease in temperature can be seen: the results correspond to the zero temperature limit, a limit which is in general difficult to reach in other finite-temperature fermionic algorithms. We compare our results with the predictions of the CSSA, where free holons and spinons are described by \cite{suzuura97,anderson87} \begin{eqnarray} \label{sbos} H=-\frac {t_h} 2 \sum\limits_{<i,j>}h^{\dagger}_ih^{}_j - \frac {J_s} 2 \sum\limits_{<i,j>}s^{\dagger}_{i,\sigma}s^{}_{j,\sigma}. \end{eqnarray} Here the electron operator $c^{}_{i,\sigma}$ is given by the product of a holon ($h_i$) and a spinon ($s_{i,\sigma}$) operator, $c^{}_{i\sigma}=s_{i,\sigma}h^{\dagger}_i$, the holon being a boson and the spinon a spin-1/2 fermion. As a consequence of the above {\em Ansatz} , the dispersion relations of the free holons and spinons are given by $\epsilon_h = - t_h \cos q_h$ and $\epsilon_s = - J_s \cos q_s$ respectively, whereas the energy of the hole is $E(k) = \epsilon_h - \epsilon_s$ and by momentum conservation $k = q_h - q_s$. We take $t_h$ and $J_s$ as two free parameters in contrast to a mean-field approximation, where they have to be calculated self-consistently. The spectral function is then given by a convolution of the spinon and holon Green functions. The lowest attainable energy ($-t_h$) and highest one ($t_h + J_s$) define the bandwidth of the hole, $2 t_h + J_s$. Since the full band-width obtained by considering the compact support of the spectral function at $J=0$ is known to be exactly $4t$ \cite{sorella92}, we take $t_h = 2t$. In order to determine $J_s$, we consider the overall bandwidth, as obtained from the simulation. As can be seen in Fig.\ 1, for all values of $J$, \begin{figure}[bth] \centering \epsfig{file=Nomega.ps,width=11cm} \caption{Density of states $N(\omega)$ for different values of $J/t$. The vertical line indicates $4t+J$.} \end{figure} the width of the density of states $N(\omega)$ scales approximately as $4t+J$ in the parameter range considered, leading to $J_s = J$. Beyond predicting bandwidths, the CSSA describes accurately the support of the spectral function in the case $J=0$, when compared with exact results \cite{sorella92,suzuura97}. If furthermore phase string effects \cite{suzuura97} are taken into account, the singularities of $A(k,\omega)$ related to holons and spinons can be reproduced. For finite $J$, the minimal (maximal) possible energy of a hole in CSSA is given by $E(k) = - F_k$ ($E(k) = F_k$) for $k < k_0$ ($k> k_0$), where $F_k \equiv \sqrt{J^2+ 4t^2-4tJ\cos\left(k\right)}$ contains both holon and spinon contributions, and $k_0$ is determined by $\cos(k_0)=J/(2t)$. The remaining parts of the compact support are given by $E(k) = \mp 2t\sin (k)$ for $k > k_0$ (lower edge) and $k<k_0$ (upper edge) respectively. Such dispersions correspond to holons with momentum $k + q_s$, and a spinon with $q_s = \mp \pi/2$ \cite{sorella96,suzuura97}. As $J \rightarrow 2t$, $k_0 \rightarrow 0$ and the lower edge of the compact support is entirely determined by the dispersion of the holon. \begin{figure}[bth] \centering \epsfig{file=J04.ps,width=8.1cm} \epsfig{file=J12.ps,width=8.1cm} \epsfig{file=J2.ps,width=8.1cm} \caption{Spectral function $A(k,\omega)$ for $J=0.4t$ (a), $J=1.2t$ (b) and $J=2t$ (c). Here the wave vector $k\in [0,\pi]$ is given on the y-axis. For clarity, the data is rescaled by the number given at the right hand side of the plot. Further details are discussed in the text. } \end{figure} We now compare the above predictions with our QMC data. Figure 2 shows $A(k,\omega)$ for $J/t=0.4$ (a), $1.2$ (b) and $2$ (c). In all cases the compact support is reproduced very well by the CSSA. The {\em Ansatz} also predicts singularities at the lower (upper) edge for $k < k_0$ ($k> k_0$), and when phase strings are considered \cite{suzuura97} along the edges and the holon lines ($\pm 2t\sin (k)$) for all momenta. The singularities along the lower holon line are also supported by a recent low energy theory \cite{sorella96}. For all parameter values we observe dominant weight along the above mentioned lines. For $J/t=0.4$, we have checked that the results are consistent within the uncertainties of MEM with a peak along the edges and a further peak along the holon lines, signaled by a broad structure between the edges and the holon lines (Fig.\ 2.a). We observed such a behaviour for $0.33 \leq J/t \leq 0.6$. For $J/t \geq 1.2$ (Fig.\ 2.b and 2.c), the structure at the lower edge narrows considerably and the data are not any more consistent with an additional structure along the lower holon line for $k < k_0$, but only with a singularity for $k > k_0$. At $J/t=2$ the exact holon and spinon dispersions can be obtained by BA \cite{bares90}. Figure 2.c shows the comparison with the CSSA, where on the one side the original dispersions are used (full line) and on the other side, with the dispersions as given by BA (crosses). Whereas the BA holon dispersion reproduces very well the lower edge, showing that as anticipated by the CSSA, at the SuSy point that edge is completely determined by the holon dispersion, the full bandwidth is better described with the original dispersions. We assign the additional weight in the region $k > \pi/2$ to processes involving one holon and more than one BA spinon. In fact, that portion resembles the difference between the supports for one-holon/one-spinon and one-holon/three-spinon processes in the ISE model \cite{kato98}. In our case, no limitation on the possible number of spinons exists, such that in principle all odd number of them are allowed. It is interesting to notice that using a fermionic spinon one is able to describe both the case $J=0$ and $J=2t$. In the first case, the spinon in the exact solution is a fermion. At the SuSy point it is expected to be a semion \cite{kato98,haldane94} and on the basis of our results, we conclude that the fermionic spinon contains all possible states with an odd number of semionic spinons. \par Finally, we consider the quasiparticle residue $ Z_k = \left| \langle \Psi_0^{L-1} | \tilde c^{}_{k\sigma} | \Psi_0^{L} \rangle\right|^2$ at $k = \pi/2$ for $J=2t$. \begin{figure}[bth] \centering \epsfig{file=Z.ps,width=11cm} \caption{Quasiparticle weight at $k=\pi/2$. \hfill $\mbox{ }$ (a) $\tilde G(\pi/2,-\tau)\equiv G(\pi/2,-\tau) \exp\left[-\tau\left( E_0^L-E_0^{L-1}\left(\pi/2\right)\right)\right] $ versus $\tau/t$. At $\tau/t \gg 1$ this quantity converges to the quasiparticle weight $Z(\pi/2)$. \hfill $\mbox{ }$ (b) Finite size scaling of $Z(\pi/2)$ as obtained from (a). The solid line is a least square fit to the form $L^{-1/2}$. We consider $\beta J =30$ for $L\le 48$ and $\beta J=60$ for $L>48$, to guarantee convergence in $\tau$.} \end{figure} As Fig.\ 2.c shows, the lower edge is very sharp and without prior knowledge, the question may arise whether we are dealing with a quasiparticle. $ Z_k$ is related to the imaginary time Green function through: \begin{eqnarray} \lim_{\tau \rightarrow \infty} G(k, -\tau) \propto Z_k \exp \left[ \tau \left(E_0^{L} - E_0^{L-1}\left(k\right) \right) \right] . \end{eqnarray} Fig.\ 3.a shows $G(\pi/2, -\tau) \exp \left( - \tau (E_0^{L} - E_0^{L-1}(\pi/2) ) \right)$ versus $\tau$, where the energy difference is obtained by fitting the tail of $G(\pi/2, -\tau)$ to a single exponential form, for several sizes. The thus estimated $ Z(\pi/2) $ is plotted versus system size in Fig.\ 3.b. Our results are consistent with a vanishing quasiparticle weight $Z(\pi/2) \propto L^{-1/2}$ which is the scaling obtained by a combination of bosonization and conformal field theory \cite{sorella96}. Since the CPU-times scales as $V\beta$ ($V$ is the volume) the determination of the $Z$-factor may be efficiently extended to higher dimensions, in contrast to determinantal algorithms for the Hubbard model that scale as $V^3\beta$. In summary, we have developed a new QMC algorithm which allows the determination of single-hole dynamics in quantum antiferromagnets. This algorithm is extremely powerful in the sense, that the required CPU time scales as $V\beta$. For the one dimensional case, we showed that the spectral function is well described by a simple model with free spinons and holons with dispersions given by $J$ and $2t$ respectively. The comparison of our results at the supersymmetric point lead to a characterization of the excitation content of the spectra for this particular parameter, where additional information is available from the Bethe {\em Ansatz} solution. Finally we computed the quasiparticle weight and showed that it vanishes as $L^{-1/2}$. This work was supported by Sonderforschungsbereich 341 in T\"ubingen-Stuttgart. The numerical calculations were performed at HLRS Stuttgart and HLRZ J\"ulich. We thank the above institutions for their support.
\section{Introduction} Nonobservation of low-energy supersymmetry may be due to the fact that we live on a non-BPS topological defect, or brane, embedded in higher-dimensional space-time~\cite{DS1}. While the external space may or may not be supersymmetric, the effective low-energy theory on the non-BPS brane is not supersymmetric. The idea of the brane Universe is especially motivated by the solution of the hierarchy problem through lowering of the fundamental scale of quantum gravity down to TeV~\cite{add,AADD} (see also \cite{lw,ddg}). In the model suggested in Ref.~\cite{add} the standard-model particles are localized on a topological defect, or $3$-brane, embedded in space with large extra compact dimensions of size $R$, in which gravity can propagate freely. In other words, the original $(4 + N)$-dimensional space-time is assumed to be split into ${\cal M}_4\times {\cal M}^\prime_N$, where ${\cal M}_4$ is the four-dimensional Minkowskian space while ${\cal M}^\prime_N$ is a compact manifold. We will refer to it as to the external space. Within the scenario~\cite{add} one can lower the fundamental scale of gravity $M_{{\rm P}_{\rm f}}$ down to TeV, or so. The observed weakness of gravity at large distances is then due to a large volume of the extra space $R\gg M_{{\rm P}_{\rm f}}^{-1}$. The relation between the fundamental ($M_{{\rm P}_{\rm f}}$) and four- dimensional ($M_{\rm P}$) Planck scales can be derived~\cite{add} by virtue of the Gauss law, \begin{equation} M_{\rm P}^2 = M_{{\rm P}_{\rm f}}^{N + 2}\, V_N \, , \label{planckscale} \end{equation} where $V_N$ is the volume of the external space with radius $R$. Although this relation is valid for any values of $M_{{\rm P}_{\rm f}}$ and $R$, the case most interesting phenomenologically is $M_{{\rm P}_{\rm f}} \sim $ TeV. (For $R\sim $ 1 mm, this implies that the extra space is two-dimensional.) The most important ``technical" question to be addressed is dynamical localization of the standard-model particles on the brane. In the field-theoretic context the fermions can be localized due to an index theorem, as was suggested in \cite{localization1/2}, whereas the localization of the gauge fields requires the outside medium to be confining \cite{localization1}. In particular, this implies that free charges cannot exist in the bulk. On the other hand, in the string-theoretic context, the most natural framework for the brane-world picture is through the $D$-brane construction (for a review, see e.g. \cite{D-branes}). In this context the standard-model particles can be identified with the open string modes stuck on the brane, whereas gravity comes from the closed string sector propagating in the bulk \cite{AADD,ST,ignatios}. For the purposes of the present paper the precise nature of localization will be unimportant, since we will exploit the low-energy effective field-theory approach for which the high-energy nature of the brane is beyond ``resolution". In this paper we will present new theoretical observations regarding the branes on the manifolds ${\cal M}_4\times {\cal M}^\prime_N$ where ${\cal M}^\prime_N$ is compact. Then we discuss some possible cosmological consequences of the brane Universe with the low-scale quantum gravity. We point out that these theories admit stable solutions which could manifest themselves through a tiny spontaneous breaking of the Lorentz and rotational invariances. In four-dimensional field theory the solutions above can be interpreted as a recently discovered new class of {\em stable} vacuum configurations in supersymmetric theories, with the constant gradient energy, which may or may not break supersymmetry~\cite{DS1,DS2}. In the latter case these solutions generalize the notion of the BPS saturation to infinite values~\footnote{By the infinite central charges we do not mean trivial infinities associated with the area of the wall or the length of the string.} of the central charge~\cite{DS2}. Below we will show how these solutions can naturally emerge in the brane Universe picture. Before proceeding, let us note that other possible cosmological implications of branes can be due to ``brane inflation" driven by a displaced set of branes \cite{braneinflation} or due to nonconservation of global quantum numbers in the brane Universe \cite{gigad}. In a very different context a class of time-dependent cosmological solutions was discussed \cite{ovrut} within the Ho\v{r}ava-Witten approach \cite{HW}. Some aspects of thermal cosmology were considered in \cite{mmtr}. These issues are not directly related, however, to the present work. For definiteness assume $(4 + N)$-dimensional space-time to be ${\cal M}_4\times {\cal T}_N$, where ${\cal M}_4$ is four-dimensional Minkowskian space while ${\cal T}_N$ is an $N$-torus with the radius $R$. The standard-model particles are the modes living on a $3$-brane embedded in this space-time. Let the brane tension (the energy per unit three-volume) be $T$. Throughout the paper we will assume that $T \sim M_{{\rm P}_{\rm f}}^4$ (the only fundamental mass scale in the theory). Obviously, $T > 0$ if the brane is to be regarded as a dynamical object (the wall), since otherwise it will be unstable. Then, in the absence of matter and radiation, the effective four-dimensional energy density $\rho_{\rm eff}$ has two contributions: a positive one from the brane tension, and a (must-be) negative one from the bulk \begin{equation} \rho_{\rm eff} = T + \Lambda_{\rm bulk}\, V_N \, , \label{lambda} \end{equation} as seen in experiments at distances $\gg R$. Here $\Lambda_{\rm bulk}$ is the negative bulk cosmological constant. In the lowest-energy state these two must conspire to cancel each other. This is a usual fine-tuning of the cosmological constant, about which we have nothing new to say. The lowest-energy state is achieved in the above picture when the brane is straight, its ``surface" area is minimal per unit three-volume of the space (Fig. 1). However, since the brane is a dynamical object it can bend or curve in the external space and, in general, it will do so. It can be tilted with respect to ${\cal M}_4\times {\cal T}_N$. Note that this curvature {\it a priori} has nothing to do with the gravitational curvature of the whole (brane + bulk) effective four-dimensional space in the Friedmann cosmological equations emerging at distances $\gg R$. The latter will be assumed to vanish. Bending the brane in the external space, or tilting, it will produce an energy excess, or an effective energy density of the Universe, which can be estimated as \begin{equation} \rho_{\rm brane} \sim M_{{\rm P}_{\rm f}}^4\, (R/r)^2\, , \qquad r\gg R\, , \label{braneenergy} \end{equation} where $r$ is a typical curvature radius of the brane, {\em not} to be confused with the Friedmannian curvature radius, which we take infinite. (In the case of the tilted brane $r$ is its longitudinal dimension, $r\sim L$. In this case Eq.~(\ref{braneenergy}) can be obtained as follows: \begin{equation}} \def\eeq{\end{equation} \rho_{\rm brane} \sim M_{{\rm P}_{\rm f}}^4\, (\alpha)^2\, , \label{atwo} \eeq provided that the tilt angle $\alpha\sim R/L\ll 1$, see below.) This energy provides an effective force resisting to bending; the force tends to straighten out the brane. What would be the cosmological significance of this excess energy? \begin{figure} \epsfxsize=8cm \centerline{\epsfbox{dvali8fig.eps}} \caption{Domain walls (branes) in ${\cal M}_4\times {\cal T}_N$. $A$ -- the straight untilted brane, $B$ -- tilted brane, $C$ -- curved brane. The length of the cylinder in the $x$ direction is $L\sim H^{- 1}$.} \end{figure} To estimate its impact we have to know $r$. It is natural to assume that in the scale smaller than the Hubble size the brane is straightened out (no excessive crumpling). Whatever mechanism solves the horizon and isotropy problems, it would also help to this straightening. So, it seems reasonable to assume that \begin{equation}} \def\eeq{\end{equation} r^2 \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} H^{-2} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} \rho_c^{-1}\, M_{\rm P}^2\, , \label{aone} \eeq where $\rho_c$ is the critical density of the Universe today. This estimate obviously refers also to the longitudinal dimension of the tilted brane, $L\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} H^{-1}$. Now, substituting this in Eq. (\ref{braneenergy}) and using Eq. (\ref{planckscale}) with $N\geq 2$ we get \begin{equation} \rho_{\rm brane} \mathrel{\rlap{\lower3pt\hbox{\hskip0pt$\sim$} \rho_c \, . \label{energy2} \end{equation} The upper bound which can only, but not necessarily, appear for $N = 2$ is intriguing since it suggests that the brane can serve as some sort of dark matter in Universe. We will study the nature of its energy below. The potential importance of the domain walls in the cosmological considerations was recognized long ago~\cite{ZKO}. In Ref.~\cite{ZKO} it was shown that were the domain walls within our universe, serious (potentially terminal) cosmological problems might arise in the theory. Situation dramatically changes if our Universe itself is a wall. \section{What is the tilted wall?} To illustrate the idea we will consider, for simplicity, the four-dimensional space ${\cal M}_4$ compactified in ${\cal M}_3\times {\cal T}$. The situation is general and does not depend on particular details. Let us start from ${\cal M}_4$ and assume that the dynamical theory under consideration has multiple discrete degenerate vacua. The simplest example is the theory of the real scalar field with the potential of the double- well type. The simplest supersymmetric example is the minimal Wess-Zumino model with the cubic superpotential. The field configuration depending only on one coordinate (call it $z$) that interpolates between one vacuum at $z\to-\infty$ and another at $z\to\infty$ is the domain wall. The width of the wall $\delta$ in the $z$ direction is of order $\delta\sim M^{- 1}$ where $M$ is the mass scale of the field(s) of which the wall is made. It is assumed that $\delta$ is much smaller than any other relevant scale of dimension of length. On ${\cal M}_4$ it is meaningless to speak of the tilted wall -- the direction $z$ can be chosen arbitrarily. Now, we compactify $z$ and consider the theory on the cylinder ${\cal M}_3\times {\cal T}$. The underlying dynamical theory must be modified accordingly, in order to allow for the existence of the walls. In the case at hand it is sufficient to assume that the wall-forming field $\Phi$ lives on a circle, i.e. one can consider the model of the sine-Gordon type or its supergeneralizations. Both, the superpotential and the K\"{a}hler potential must be periodic in $\Phi$, with commensurate periods. For simplicity we assume these periods to be $2\pi$. Following an old tradition, we rename the compact coordinate, $z\to X$. The non-compact coordinates (including time) will be denoted by $x_\mu$. We look for the topologically nontrivial solutions of the soliton type on the cylinder, depending on one coordinate only. The solution of the type $\Phi_0 (X)$, which is independent of $x_\mu$, is the straight wall, see $A$ in Fig. 1. It satisfies the condition $\Phi_0 (X+2\pi R)=\Phi_0 (X) +2\pi$. This wall is aligned ``parallel" to the cylinder. The tilted wall ($B$ in Fig. 1) is a solution of the type $\Phi_\alpha ( X\cos\alpha + x\sin\alpha)$ where $\alpha$ is the tilt angle. Note that the function $\Phi_\alpha$ does not coincide with $\Phi_0$, generally speaking. In the limit of small $\alpha$ the difference between $\Phi_0$ and $\Phi_\alpha$ is $O(\alpha^2)$. The condition $\Phi_\alpha ( (X+2\pi R)\cos\alpha + x\sin\alpha)= \Phi_\alpha ( X\cos\alpha + x\sin\alpha )+2\pi$ must be satisfied. It is not difficult to see that the tilted wall solution exists, if so does the straight wall solution. The tilted wall is stable provided the solution is ``nailed" at the points $1\,,\,\, 2$ at the boundaries (see Fig. 1). Alternatively, one can glue the boundaries of the cylinder converting it into a two-torus. The tilted wall solution is automatically stable then if the wall winds around the two-torus. It is not difficult to prove that the domain walls on the cylinder cannot be BPS-saturated, strictly speaking. However, if $\delta/R\ll 1$ the straight walls may be very close to the BPS saturation, achieving the BPS saturation in the limit $\delta/R\to 0$, when the wall becomes the ``genuine brane". The tilted walls are never BPS-saturated, their tension being larger than that of the straight wall. This effect -- the increase of the internal tension $T$ of the tilted wall compared to that of the straight wall -- is proportional both to $\alpha$ and to $(\delta / R)^2$. It can be made arbitrarily small in the limit $\delta / R \to 0$. We will neglect it in what follows, using one and the same tension $T$ for the tilted and straight walls. The tilting does produce an impact on $\rho$ since the area of the tilted wall per unit length of the cylinder in the $x$ direction is larger. This effect is most conveniently described by the effective low-energy theory of the zero modes on the wall. The corresponding discussion is presented in the next section. \section{Calculating $\rho_{\rm brane}$ for non-vanishing tilt angles} To deal with the long wave-length ($\lambda\gg \delta$) deformations of the straight wall we will use an effective four-dimensional field theory emerging for the zero modes. The wall solution spontaneously break the translational invariance in one direction. Correspondingly, in the simplest case there arises one zero mode which is the Goldstone boson of the spontaneously broken symmetry. In more complicated models (see below) there may arise several zero modes $\phi_A$. In the absence of gravity the effective Lagrangian is \begin{equation}} \def\eeq{\end{equation} {\cal L} = M_{P_{\rm f}}^2 \,\int d^4x \, \frac{\partial \phi_A}{\partial x_\mu} \, \frac{\partial \phi_B}{\partial x^\mu} \, \eta^{AB}(\phi )\, , \label{dopone} \eeq where $\eta^{AB}$ is the external metrics depending on the structure of the manifold on which the fields $\phi$ live. For instance, in the case of the domain wall in Minkowski space $\eta^{AB}=\delta^{AB}$. All dimensional constants in the low- energy theory of the zero modes are related to the order parameter, which is the brane tension $T\sim M_{{\rm P}_{\rm f}}^4$. The dynamics of the Goldstone bosons on the brane is described~\cite{DS1} by $(3+1)$-dimensional field theory which may or may not be supersymmetric (in the latter case the brane must be BPS saturated). If we deal with the supersymmetric theory $\eta^{AB}$ is obtained from the K\"ahler potential. Let us assume for the time being that we have only one Goldstone boson and $\eta^{AB} =1$. Then, the solution $\phi = \alpha x$ goes through the equations of motion of the theory (\ref{dopone}). This is the constant energy density vacuum, discussed in Refs. \cite{DS1,DS2}. To make contact with the discussion above, we note that the vacuum $\phi =\alpha x$ represents the tilted wall described by the solution $\Phi_\alpha$ in the full theory. The additional contribution to $\rho$ compared to the straight brane is obviously \begin{equation}} \def\eeq{\end{equation} \Delta\rho_{\rm brane} = \frac{T\alpha^2}{2}\, . \eeq This result has a very transparent interpretation in the full theory. It exactly reproduces the increase of the brane surface per unit length of the cylinder for a non-vanishing tilt angle $\alpha$, see Fig. 2. Given this interpretation, one might ask why one needs to consider the effective low-energy theory at all. The point is that at the next step we want to switch on gravity in the bulk. Having the low-energy theory of the zero modes, describing matter on the brane, helps analyze the impact of gravity. \begin{figure} \epsfxsize=8cm \epsfysize=7cm \centerline{\epsfbox{dvali8fig2.eps}} \caption{The map of the cylinder of Fig. 1. The ratio of the surfaces of the branes $B$ to $A$ per unit length in the $x$ direction is $1+\alpha^2/2$.} \end{figure} In the language of the effective field theory~(\ref{dopone}), the bending of the brane is equivalent to $\phi^A$ acquiring some $x_{\mu}$ dependence in the vacuum. The generic $\phi^A(x)$ configuration is unstable and will decay producing $\phi$ waves (the sound waves on the brane). This is the mechanism of eliminating foldings on the brane. Since these waves travel with the speed of light, the brane will iron itself out in the horizon scale. Eventually it will evolve to a tilted brane. In the state with an arbitrarily bent brane we will distinguish two components. One can be viewed as a collection of all possible Goldstone waves traveling with the speed of light, which ``redshift" away like ordinary matter. Another component is the vacuum solution (more precisely, a family of solutions) that would be stable if it were not for the expansion of the Universe. This configuration ``redshifts" away slower than matter and can be called the tilted brane configuration (or wrapped, if there are several windings on the length of the cylinder $L$). It can only ``redshift" through the ``stretching" triggered by the Universe expansion. In the case of the generic massless fields $\phi_A$ with the flat $\eta^{AB}$, the solution $\phi = \alpha x$ is stable under any localized deformations. If $\phi_A$'s are the Goldstone bosons arising due to the spontaneous breaking of some compact symmetry, $\phi_A$'s are periodic (in fact, they are phases defined modulo shifts). This is exactly what happens in our case since the extra dimensions are assumed to be compact. Then the solution $\phi = \alpha x$ must be modified appropriately, see Sec. 2. Here we will add a comment regarding the issue of stability. To explain the point we will use a simple example \footnote{The analogy is much more precise than one might naively think, since the translation in the external space can be regarded as an internal U(1) rotation. In the presence of gravity, this is a gauge rotation gauged by the external components of the graviton (graviphoton(s)). We can neglect the coupling to gravity for the time being, due to the large sizes of extra dimension(s).} of the Goldstone field produced as a result of breaking of some global $U(1)$ symmetry (a similar model was treated in Sec. 6 of~\cite{DS1}). Start from the Lagrangian \begin{equation} {\cal L} = |\partial_{\mu}\Phi|^2 - {\lambda^2 \over 2}\left(|\Phi|^2 - v^2\right)^2\, . \label{othergold} \end{equation} The equations of motion have a solution \begin{equation} \Phi = c\,e^{i\mu x}\, , \qquad c = \sqrt{v^2 - {\mu^2 \over \lambda^2}}\, , \label{solution} \end{equation} which corresponds to winding of the phase with the period $2\pi / \mu$ as one moves along $x$. This solution has both the gradient and potential energies (cf. the solutions discussed in~\cite{DS2}). The potential term scales as $\sim \mu^4 / \lambda^2$. It is seen that at $\lambda \rightarrow \infty$ the solution (\ref{solution}) becomes pure gradient energy ($v$ must scale accordingly, of course, i.e. $v\sim\lambda^{-1}$). This is because in this limit, the $x$-dependence can not affect the order parameter and, thus, the potential energy. In the case of the tilted brane this would mean that the brane with a constant tilt carries purely gradient energy in the limit when bending can not affect its tension. This is true for any brane in the limit when one can ignore its thickness. The corresponding configuration can not decay into the $\phi$ waves. Its energy is reduced only through the Friedmann expansion of the Universe and, thus, scales as $\sim a^{-2}$, the scale factor in the Friedmann Universe. This means, in turn, that the state at hand will sooner or later dominate over both, the radiation and the matter densities. We will discuss the observational implications of this fact in Section 5. Before, however, let us discuss the effect of higher dimensional bulk gravity. \section{Switching on gravity in the bulk} The effective Lagrangian for the zero modes localized on the brane becomes \begin{equation} {\cal L}_{\rm brane} = M_{{\rm P}_{\rm f}}^2\, g^{\mu\nu}\left(\partial_{\mu} \phi_A\partial_{\nu}\phi_B \right) \eta^{AB} +\mbox{fermions}\, , \label{Goldstoneslag} \end{equation} where $\eta_{AB}$ and $g$ are the external and induced metrics, respectively. The fermion terms are relevant in the supersymmetric case. So far, we have discussed the brane dynamics on the cylinder neglecting gravity. Now we want to take it into account. In the scenario under consideration gravity is not confined to the wall, but, rather, propagates in the bulk. The components of the graviton belonging to ${\cal M}_4$ and ${\cal T}^N$ require separate treatment. The effect which is most important for us is due to the higher-dimensional components of the graviton (the so-called graviphotons). For simplicity we will consider only a single extra dimension compactified on a circle parametrized by the coordinate $X$. The zero mode component of $g_{\mu5}$ is the graviphoton $ A_{\mu} (x)$. Viewed as a four-dimensional gauge field, $A_{\mu} (x)$ gauges the translation in $X$, i.e. $X \rightarrow X + f(x_{\mu})$, which, from the four-dimensional standpoint, is an internal U(1) gauge symmetry. Since the brane spontaneously breaks the translational invariance, the corresponding gauge symmetry is realized nonlinearly. The Goldstone mode $\phi$ is eaten up by the gauge field $A_{\mu}$ (which gets a mass of order (1 mm)$^{-1}$~\cite{higgseffect}). As a result, the tilted wall solution $\phi =\alpha x$ we have considered previously is pure gauge. It can be compensated by the gauge field $A_{x} = \alpha$ and presents no physically observable effect. Thus, if we have only one zero mode on the wall, the tilted wall is indistinguishable from the straight one in the presence of gravity. To make the idea work we must have two or more zero boson modes on the wall. Then the tilted wall (the constant gradient energy solution) will lead to a physically observable excess in $\rho$. Let us explain this in more detail. Our solution can be made physical by ``projecting out" the graviphoton, provided the compact manifold breaks translational invariance in the extra space. The simplest dynamical realization is as follows. Consider topologically stable winding configurations~\cite{DS1}. (Such configurations may anyway be needed for supersymmetry breaking \cite{DS1} or for stabilizing radii at large distances~\cite{radius}). Consider a five-dimensional scalar field $\xi(x_{\mu}, X)$ transforming under an internal U(1)$_I$ symmetry as $\xi \rightarrow { e}^{i\alpha} \xi$. Assume that a potential forces the condensate $\langle \xi \rangle \neq 0$ to develop. The simplest choice can be $V = (|\xi|^2 - v)^2/m$, or any other similar function. Then the vacuum manifold is a circle, and there are topologically stable winding configurations \begin{equation} \xi = w\,e^{in X/R}\, , \qquad w = \sqrt{v - n^2m/R^2} \, , \end{equation} with integer $n$. They correspond to giving a vacuum expectation value to different Kaluza-Klein modes and, therefore, are {\it topologically} stable due to the mapping of the vacuum circle on the external compact space. Thus we are free to choose any of these states as the ground state. \footnote{Strictly speaking, we have to restrict ourselves to $n^2 < vR^2 / m$ since larger winding numbers force the field to vanish, and the solution will ``unwind". For our purposes, however, it is sufficient to consider small enough $n$.} Configurations with nonzero $n$ break spontaneously the $X$ translations and, thus, give mass to the graviphoton $A_{\mu}$. As a result, the graviphoton field can be integrated out in the low-energy effective theory. Then the brane Goldstone $\phi$ remains a physical field. More precisely, the picture is as follows. The theory at hand has two U(1) symmetries from the very beginning: the internal U(1)$_I$ and ``external" U(1)$_E$ gauge symmetry under translations in the extra dimension. However, from the point of view of a four-dimensional observer living on the brane, both of them are internal symmetries, one global and another gauged by $A_{\mu}$. The condensate $\langle \xi \rangle \neq 0$, with $n \neq 0$, breaks both U(1)'s down to a global U(1) describing the change of a relative position of the brane in the extra space. The brane breaks the latter down to nothing, but since there is no gauge degree of freedom left, the corresponding Goldstone is physical. A related mechanism of ``projecting out" the graviphoton while still keeping at least one Goldstone mode may be provided by multicomponent walls suggested in Ref.~\cite{SSV}. It was noted that in supersymmetric theories with multiple degenerate vacua, under certain conditions there exists a variety of the BPS-saturated walls. Some of them may coexist together; then, the tension of such ``multilayer" configuration is exactly equal to the sum of the tensions of the individual walls, independently of the distance between the layers (individual walls). A similar effect takes place for the $D$-branes which are the BPS-states in the limit of unbroken supersymmetry, and the net force between two straight parallel branes vanishes (see~\cite{D-branes} for a review). This means that, apart from the zero mode corresponding to the overall translation, there are extra zero modes corresponding to shifting the layers with respect to each other, without changing the position of the center of mass. As was mentioned, on the cylinder the exact BPS saturation can be achieved only in the limit $\delta / R\to 0$. Hence, only the first zero mode -- the one related to the overall translations -- is exactly zero, others become quasi-zero. The zero mode is eaten up by the graviphoton, eliminating it from the game. The quasi-zero modes remain physical, their dynamics is described by a low-energy Lagrangian. The supersymmetry breaking in general will give mass to this pseudo-Goldstones since the net force between branes (or walls) is non-zero. If there are no other massless states in the bulk, at large distances the only force between the branes is due to gravity which gives a small $r^N$-suppressed mass to the Goldstones ($r$ is the interbrane separation). This mass can be made arbitrarily small if the separation is large and, practically, it can be neglected. All dynamical consequences are the same as discussed in Sec. 3. \section{Implications for cosmology} Now let us discuss cosmology in more detail. We will treat the problem from the standpoint of the effective four-dimensional field theory at scales $\gg\gg R$. As was said above, in this picture the fluctuations of the brane in the external space $\phi_A(x_{\mu})$ can be described by the Goldstone modes which, at energies $\ll T$ (almost flat brane), behave as free particles. Canonically normalized fields are $\chi_A \sim (T)^{1/4} \phi_A$. For simplicity, consider a fluctuation in one transverse direction only. Then, very roughly, the issue is reduced to the behavior of a free field $\chi$ in an expanding Universe. The question is what are the initial conditions for such a field? The general solution of its equation of motion (the comoving coordinates are assumed) \begin{equation} \partial^2 \chi = 0 \label{freeequation} \end{equation} has the form \begin{equation} \chi = \sqrt{T} c x + \chi_p { e}^{ipx} \label{waves} \end{equation} (modulo the Lorentz transformations), where $c$ is a constant and the second term is some collection of massless plane waves. The energy stored in the second term will just ``redshift" away, like massless matter. However, the first term produces a rather different contribution. In the present context this term describes the tilted brane with the tilt angle given by $c \sim R/r$. In other words, when we move along $x$, the brane is wrapping around the extra dimension with the period $\sim r$. In the absence of gravity, in the limit of an infinitely thin brane, this is a stable configuration for any $c$. In the presence of gravity, however, its energy ``redshifts" away as $\sim a^{-2}$, due to the Friedmann expansion. Therefore, eventually this energy will dominate over matter. When this happens actually depends on the initial condition for $r (c)$ and on the subsequent evolution of the scale factor. Assume that initially $c \sim RH_{\rm in}$, where $H_{\rm in}$ is the Hubble parameter at that time. In other words, we assume that, when the brane was ``formed", it wrapped around the extra dimension once ( on average) per the causally connected region $\sim H_{\rm in}^{-1}$. Now, at present time, this region must have had evolved into a volume comparable to the present Hubble size (or larger). This is required by whatever mechanism solving the horizon problem. This means that \begin{equation} r_{\rm today} \sim \frac{a_{\rm today}}{H_{\rm in}a_{\rm in} }\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} H^{-1}_{\rm today}\, . \end{equation} Thus, the energy density of a tilted brane would be $ \mathrel{\rlap{\lower3pt\hbox{\hskip0pt$\sim$} \rho_c$. In particular, it is sufficient to have a period of inflation with the number of $e$-foldings \begin{equation} N \simeq {\rm ln}{H_{\rm in}M_P \over T_{\rm today} T_R} \, , \label{ef} \end{equation} and the subsequent reheating temperature $T_R$, where $T_{\rm today} \sim 3 K^o$. For instance, a brief period of the ``brane inflation" \cite{braneinflation} can do the job.\footnote{The wrapped brane can provide an additional (time-dependent) force stabilizing $R$. This may have implications for the early cosmology \cite{radiuscosmology}.} In reality, however, we expect inflation to have more $e$-foldings and, thus, $\rho_{\rm brane} \mathrel{\rlap{\lower3pt\hbox{\hskip0pt$\sim$} \rho_c$ is rather natural. \section{The Lorentz Symmetry Breaking} The important fact is that the solution at hand spontaneously breaks the Lorentz and rotational invariances in four dimensions. This would result in a global anisotropy in the expansion if $\rho_{\rm brane}$ were to dominate the Universe. Thus, we must require $\rho_{\rm brane}$ to be subdominant today, but it can become dominant at any time in the future. In this respect any observational evidence of a global anisotropy would be extremely important for constraining $\rho_{\rm brane}$. The tilted brane would induce rotational (or Lorentz) noninvariant terms in the effective four-dimensional standard-model Lagrangian. The important fact is that the brane Goldstones (or pseudo-Goldstones, which are similar in this respect) necessarily couple to all particles living on the brane through an induced metric on the brane. Thus, the operators of the form \begin{equation}} \def\eeq{\end{equation} \frac{\partial \phi}{\partial x_\mu} \, \frac{\partial \phi}{\partial x^\nu}\, \psi\gamma_{\mu}\partial x^\nu\psi + \frac{\partial \phi}{\partial x_\mu} \, \frac{\partial \phi}{\partial x^\nu} \, F_{\mu\alpha}F^{\nu\alpha} +... \eeq will appear (accompanied by the appropriate powers of $T$). Here, $\psi$ are the matter fermions and $F_{\mu\alpha}$ stands for the matter gauge fields. In the background of a titled brane $\phi = \alpha x$, these terms will induce effective Lorentz-violating interactions among the standard-model fields. Can one have tilted branes and still avoid the anisotropy and violations of the rotational invariance? In principle, the answer is positive. To this end one must deal with more complicated manifolds, such that the topology of the manifold ${\cal M}^\prime_N$ is the same as that of our $3$-dimensional space ${\cal M}_3$ (of course, in any case, locally ${\cal M}_3$ must be Minkowskian). In other words, ${\cal M}^\prime_N$ must include unshrinkable surfaces $C_3$ that are isomorphic to ${\cal M}_3$. In this case the brane that corresponds to a point on $C_3$ can move around when one moves along the brane in ${\cal M}_3$. The topologically nontrivial configuration of interest emerges when this motion maps $C_3$ onto ${\cal M}_3$. For such a configuration, one can find a preferred reference frame for which rotations on ${\cal M}_3$ will not be broken, but the Lorentz invariance will be broken in the arbitrary reference frame. To be more specific, let us give an example. Imagine that we have an external space, with three extra dimensions, which forms a manifold $K$ of radius $R$ (say, one can think of $K$ as of a three-sphere). Assume that ``our" three-dimensional space is also a very large manifold $K$, its radius is much larger than $H^{-1}$. Let us call these two manifolds ``small" and ``big", respectively. Clearly, at human scales the ``large" manifold (plus time) is identical to ${\cal M}_4$, but not globally. Now, our brane corresponds to a point on the ``small" $K$. Imagine a configuration such that when one moves in ``large" $K$, the point in ``small" $K$ also moves in the same way. Thus, we get a mapping $K \to K$, which is certainly topologically stable. Roughly speaking, the brane wraps isotropically in all directions. Such a wrapping is isotropic and stable. What is the low-energy picture corresponding to this construction? Since $K$ has tree dimensions, the brane breaks three translational invariances. Thus, there are three massless Goldstone modes in the effective low-energy theory, $\Phi^A\, , \,\,\, A=1,2,3$. Let $x^A$ be three coordinates on our large $ S_3$ which locally look as the Cartesian coordinates in our Minkowski space. Then, the solution is $\Phi^A = \alpha x^A$. This solution is isotropic because of the spherical symmetry of the problem. Its energy density will still scale $\propto 1/a^2$ because this is essentially the same gradient energy solution we have discussed previously. \section{Conclusions} The idea of confining our Universe to a wall which ensures an appropriate supersymmetry breaking~\cite{DS1} seems to be promising. At the very least, it deserves further investigation. Being combined with the idea of compactification of the extra dimensions and allowing gravity to propagate in the bulk~\cite{add} it leads to potentially realistic and reach phenomenology. In this paper we have shown that the walls on ${\cal M}_4\times {\cal M}^\prime_N$ where ${\cal M}^\prime_N$ is compact generate peculiar theoretical effects due to tilting. The situation becomes especially interesting when gravity is switched on. The appropriate theoretical framework for its analysis is provided by the effective low-energy theories of the Goldstone modes on the brane. After gravity is switched on one of these modes is eaten up by the graviphoton (making it massive and eliminating it from the massless particle spectrum). We presented models where there are residual physical Goldstone (or pseudo-Goldstone) modes. These solutions produce a framework for the spontaneous breaking of the Lorentz and rotational invariance and may have observable consequences. \vspace{0.5cm} {\bf Acknowledgments}: \hspace{0.2cm} We would like to thank A. Dolgov, I. Kogan, K. Olive, T. Piran, and A. Vilenkin for useful discussions and comments. This work was supported in part by DOE under the grant number DE-FG02-94ER40823.
\section{Introduction} The cosmic X--ray background (XRB) above $\sim 1$ keV is the result of the integrated emission of discrete sources, since the contribution of any intergalactic hot medium must be negligible (Wright et al. 1994). In the soft X--ray band from 0.5 to 2 keV the largest fraction of the XRB has already been resolved into sources (Hasinger et al. 1998), most of which turned out to be broad line active nuclei (Schmidt et al. 1998), i.e. Quasi Stellar Objects (QSOs) and Seyfert galaxies of type~1. The spectra of these sources are however too steep to reproduce also the hard XRB at several tens of keV, where the bulk of the energy resides, and a population of objects with flatter spectra is therefore required. The most popular synthesis models of the XRB are based on the so--called unification schemes for Active Galactic Nuclei (AGNs), where the orientation of a molecular torus surrounding the nucleus determines the classification of the source. At a zeroth--order approximation level, sources observed along lines of sight free from the torus obscuration should have unabsorbed X--ray spectra and optical broad lines (type 1 AGNs), while sources seen through the torus should have absorbed X--ray spectra and appear as narrow line objects in the optical (type 2 AGNs, e.g. Seyfert 2 galaxies). In this framework type 2 AGNs provide a natural class of sources with X--ray spectra flattened by absorption. The intrinsic X--ray luminosity function (XLF) of type 2 objects is unknown and has been usually assumed to be the same as the one derived for type 1s (e.g. Boyle et al. 1993), apart from a normalization factor. The cosmological evolution has also been taken identical for type 1s and type 2s. Under these assumptions it has been shown that the broad band 3--100 keV spectrum of the XRB can be reproduced by an appropriate mix of unabsorbed and absorbed AGNs (Matt \& Fabian 1994; Madau, Ghisellini \& Fabian 1994; Comastri et al. 1995, hereafter Co95). The number ratio $R$ of type 2 to type 1 objects, as well as the distribution of the absorbing column densities $N_{\rm H}$, are key parameters of the models; these have been assumed to be independent of redshift and of intrinsic source luminosity, and have been treated as free parameters in the fitting procedure. Since the overall parameter space of the models is quite large and a good fit to the XRB can be obtained with different set of values, it is important to compare the model predictions with the largest number of observational constraints. Indeed, Co95 showed that the source counts in the 0.5--2 keV and 2--10 keV energy bands, as well as the redshift distributions, could successfully be reproduced by their model. Very recently an additional set of observational constraints has become available. Deep surveys from ROSAT have extended our knowledge to the low luminosity part of the AGN XLF (Miyaji, Hasinger \& Schmidt 1999a, hereafter Mi99a). Contrary to previous results (Boyle et al. 1993; Page et al. 1996; Jones et al. 1997) a pure luminosity evolution (PLE) of AGNs with redshift is no longer consistent with the data, and a luminosity dependent density evolution (LDDE) is required. From the X--ray data of an optically selected sample of Seyfert galaxies Risaliti, Maiolino \& Salvati (1999) have determined the $N_{\rm H}$ distribution for local Seyfert 2 galaxies, pointing out that a significant fraction of sources have columns exceeding $N_{\rm H}=10^{25}$ cm$^{-2}$ and are therefore completely thick to Compton scattering. The $R$ ratio between type 2s and type 1s has been determined in the local Universe for low luminosity AGNs, i.e. Seyfert galaxies (Maiolino \& Rieke 1995), while the existence of a relevant number of high luminosity absorbed sources, the so--called QSO~2s, which is a basic assumption of previous models, is still uncertain (Akiyama et al. 1998). An observational constraint to the QSO~2 number density can be obtained from the infrared source counts. Indeed, QSO~2s are expected to have strong infrared counterparts, since the dust present in the torus should re--emit in the IR band the nuclear radiation absorbed by the gas. The ultraluminous infrared galaxies (ULIRGs) discovered by IRAS are the only local objects with QSO--like bolometric luminosities (Soifer et al. 1986; Kim \& Sanders 1998). Thus, even if all ULIRGs were powered by a hidden AGN, the local QSO~2s could not be more numerous than ULIRGs. Finally, source counts in the 5--10 keV band have been derived for the first time by the BeppoSAX satellite with the HELLAS survey (Giommi et al. 1998; Comastri et al. 1999). In the present paper we test the standard synthesis model to verify if it remains compatible with the new data. These data leave still some latitude to important parameters of the model, and various choices are possible to fit the XRB equally well. However, in all cases we find moderate but consistent evidence that at least some of the standard assumptions have to be relaxed: extra hard spectrum AGNs are needed at intermediate or high redshifts, in addition to those expected in the usual scenario. The additional sources could be analogous to local Seyfert 2s, if they evolve faster than type 1s, or they could be other astrophysical sources not yet enlisted among the contributors to the XRB. We discuss the observations which could distinguish between the alternatives. Throughout this paper the deceleration parameter and the Hubble constant are given the values $q_{0}=0.5$ and $H_{0}=50$ km s$^{-1}$ Mpc$^{-1}$. \section{AGN X--ray properties} \subsection{The spectra} After the observations of X--ray satellites like GINGA, ASCA and BeppoSAX, different components have been recognized in the X--ray spectra of AGNs. Starting with Sey 1 galaxies, the basic component is a power law with energy spectral index $\alpha\sim0.9$ (Nandra et al. 1997a) and an exponential cut off at high energies. A mean value for the $e$--folding energy can probably be set at $\sim 300$ keV, although the observed dispersion is very high (Matt 1998). Some of the primary radiation is reprocessed by an accretion disc and/or the torus around the nucleus, producing a flattening of the spectral slope above $\sim 10$ keV, and a strong iron line at 6.4 keV (Nandra \& Pounds 1994). Below 1--1.5 keV a radiation excess with respect to the power law emission is detected in a large fraction of Sey 1s (sometimes resulting from a misfit of the ``warm absorber'' component). The spectrum of QSO~1s is similar to that of Sey 1s, but there is no evidence for the iron line and the reflection hump to be as common (Lawson \& Turner 1997). Assuming that the accretion disc produces most of the line and hump, Nandra et al. (1997b) ascribe these differences to a higher ionization state of the disc in higher accretion rate sources, so that in QSO~1s the spectral features due to photoelectric processes are quenched. Recently, Vignali et al. (1999) have derived a mean spectral slope of $\langle \alpha \rangle =0.67\pm 0.11$ from a sample of 5 QSO~1s at redshifts above 2. Although the statistics is poor, this result seems to suggest that the spectra of high redshift QSOs are flatter than those of local ones. In Sey 2 galaxies the power law is cut off by photoelectric absorption at energies increasing with the column density of the intercepted torus. For highly absorbed objects the X--ray luminosity may be dominated by that fraction of the nuclear radiation which is reflected off the torus surface towards the observer. When $N_{\rm H}>10^{25}$ cm$^{-2}$ the obscuring medium is completely thick to Compton scattering and the spectrum is a pure reflection continuum as described by Lightman \& White (1988), with a 2--10 keV luminosity about two orders of magnitude lower than that of Sey 1s (Maiolino et al. 1998). On the contrary, when $N_{\rm H}<10^{24}$ cm$^{-2}$ the medium is Compton--thin and the spectrum is dominated by the component transmitted through the torus. In the range $10^{24}<N_{\rm H}<10^{25}$ cm$^{-2}$ both a transmitted and a reflected component contribute to the observed luminosity, the Circinus galaxy (Matt et al. 1999) being a typical example. Also Sey 2 galaxies often have soft emission in excess of the absorbed power law (Turner et al. 1997). These soft excesses are however two orders of magnitude weaker than those of Sey 1s of the same intrinsic luminosity and their nature is still unclear (probably scattered or starburst radiation). \subsection{The XLF and cosmological evolution} The most recent results about the AGN XLF and cosmological evolution have been obtained by Mi99a by combining data from several ROSAT surveys. Down to a limiting flux of $10^{-15}$ erg s$^{-1}$ cm$^{-2}$, reached by the deep survey in the Lockman Hole, they collected a sample of about 670 sources, which is the largest X--ray selected sample of AGNs presently available. The local XLF is described with a smoothed double power law of the following form: \begin{displaymath} \phi(L_{\rm x})=\frac{{\rm d}\,\Phi(L_{\rm x})}{{\rm d\,log}L_{\rm x}}={A}\,\left[(L_{\rm x}/{L_*}) ^{{\gamma_1}} +(L_{\rm x}/{L_*})^{{\gamma_2}} \right]^{-1}, \end{displaymath} \noindent where $L_{\rm x}$ is the observed 0.5--2 keV X--ray luminosity, ranging from $10^{41.7}$ to $10^{47}$ erg s$^{-1}$. The best fit values for the cosmology adopted here are: $A=(1.57 \pm 0.11)\times 10^{-6}$ Mpc$^{-3}$, $L_*=0.57^{+0.33}_{-0.19} \times 10^{44}$ erg s$^{-1}$, $\gamma_1=0.68\pm 0.18$ and $\gamma_2=2.26\pm 0.95$. The XLF has been found to evolve from redshift 0 up to $z_{cut}=1.51\pm 0.15$, with an evolution rate which drops at low luminosities according to the factor: \begin{eqnarray} e(z,L_{\rm x})= \left\{ \begin{array}{ll} (1+z)^{\max(0,{p1}-{\alpha}({\rm log}\; {L_{\rm a}} - {\rm log}\;L_{\rm x}))} & L_{\rm x}<L_{\rm a} \\ (1+z)^{{p1}} & L_{\rm x}\ge L_{\rm a}\;;\\ \end{array} \right. \nonumber \end{eqnarray} \noindent here $p1=5.4\pm 0.4$, $\alpha = 2.3\pm 0.8$ and ${\rm log} L_{\rm a}=44.2$ (fixed). The X--ray AGNs have been observed at redshifts up to $z=4.6$ and there is no evidence for a decline in their space density beyond $z\sim 3$, unlike what is found in optical (Schmidt, Schneider \& Gunn 1995) and radio surveys (Shaver et al. 1997). We note that the XLF parametrization of Mi99a is a preliminary result and is not a unique representation of the ROSAT data. The extrapolation of the high redshift XLF into the low luminosity range, where few data are available, is not well constrained. Indeed, the number of low luminosity, high redshift AGNs could be higher than the Mi99a representation (Hasinger et al. 1999). Another cause of uncertainty is the possible presence of type 2 AGNs in the Mi99a sample: unlike previous works, where only (optical) type 1 AGNs were included, Mi99a do not discriminate between type 1s and type 2s; some of the latter could then appear in the ROSAT bandpass because of their soft excesses and, for sources at high redshifts, because of the $K$--correction. Objects with type 2 optical spectra are relatively rare in the ROSAT sample (Hasinger et al. 1999). As for the X--ray spectral type, within any given model the raw counts can be corrected for the contribution of the absorbed sources, and these can be subtracted from the XLF: in the following we consider also this approach, and investigate the robustness of our conclusions with respect to the correction. In general, the correspondence between optical and X--ray spectral classification is broadly verified in the local Universe, albeit with some blurring (see Section 2.3); at high redshifts the question is still unsettled. Previous works about the XLF of AGNs used a PLE model both in the soft X rays, by combining observations from ROSAT and Einstein (Boyle et al. 1993, 1994; Jones et al. 1997), and in the hard X rays from ASCA data (Boyle et al. 1998a). In the Mi99a data the fit with a PLE model is rejected with a high significance. A pure density evolution model is marginally rejected, and LDDE models are preferred, even if several variants are still being discussed. \subsection{The number and column densities of local type 2 AGNs} In the local Universe 5--10\% of the galaxies show Seyfert activity (Maiolino \& Rieke 1995; Ho, Filippenko \& Sargent 1997). From a sample of $\sim 90$ nearby Seyfert galaxies limited in the B magnitude of the host galaxy, Maiolino \& Rieke (1995) derived an estimate for the local ratio $R$ of type 2 to type 1 Seyferts. From our point of view Seyfert types 1.8, 1.9 and 2, which have flat X--ray spectra due to absorption by cold gas, can be grouped as type 2 objects, while types 1, 1.2 and 1.5, which have steep X--ray spectra without significant cold absorption ($N_{\rm H}<10^{21}$ cm$^{-2}$), can be grouped as type 1s. Here it is noted that the relation between Seyfert type and X--ray absorption is not univocal. By observing with ROSAT the complete sample of Piccinotti et al. (1982), Schartel et al. (1997) showed that at least a fraction of type 1 AGNs suffer from X--ray absorption by more than $N_{\rm H}=10^{21}$ cm$^{-2}$. However this fraction (grouping Seyfert 1, 1.2 and 1.5) is only 20\%, and on average their $N_{\rm H}$ does not exceed $10^{22}$ cm$^{-2}$. The inclusion of some moderate--absorption type 1s should not change significantly our results. By considering Seyfert types 1.8, 1.9 and 2 as type 2s, and Seyfert types 1, 1.2 and 1.5 as type 1s, Maiolino \& Rieke found $R$=4.0$\pm$0.9, in agreement with the results of Osterbrock \& Martel (1993) and, more recently, Ho et al. (1997). From the Maiolino \& Rieke sample Risaliti et al. (1999) have derived a distribution of X--ray column densities for Sey 2s. The selection of the sample by means of optical narrow emission lines, rather than in the X--rays, should avoid biases against X--ray absorbed sources. It turned out that most of the sources are affected by strong absorption, $\sim 75\%$ of the objects having $N_{\rm H}>10^{23}$ cm$^{-2}$. Furthermore, a significant fraction of sources ($>25\%$) are absorbed by $N_{\rm H}>10^{25}$ cm$^{-2}$. Their results are shown in Fig.~1 and compared with the $N_{\rm H}$ distribution assumed by Co95. \begin{figure} \epsfig{file=8484_fig1.ps, width=9.cm, height=9.cm, angle=0} \caption{The normalized $N_{\rm H}$ distribution in Sey 2 galaxies derived by Risaliti et al. (1999) [Panel a)], compared with the distribution assumed by Co95 [Panel b)]. The shaded area represents the fraction of sources for which only a lower limit of log$N_{\rm H}>24$ is available.} \end{figure} In the limited luminosity range of Seyfert galaxies Risaliti et al. (1999) did not find evidence of a correlation between absorption and luminosity. In the wider luminosity domain which includes the QSOs the evidence is contradictory. Recent results from the IRAS 1--Jy survey (Kim \& Sanders 1998) show that the space density of ULIRGs at $z\stackrel{<}{_{\sim}} 0.1$ is similar to that of optically selected QSOs of comparable bolometric luminosity. By considering that the number of obscured QSOs cannot exceed that of ULIRGs, and assuming that every ULIRG is powered only by nuclear activity, we can set a conservative upper limit of 2 to the $R$ ratio at high luminosities. The actual value of $R$ might be significantly lower. Indeed, there is evidence that the fraction of AGN--powered ULIRGs decreases from 50\% for IR luminosities $L_{\rm IR}\stackrel{>}{_{\sim}} 1.7\times 10^{46}$ erg s$^{-1}$ to 15\% below this value (Lutz et al. 1998), the remaining ones being dominated by starburst activity. \section{The model} \subsection{The XRB spectrum} Our models are completely analogous to those of the canonical lineage, the only differences arising from updated input parameters. A key set of such parameters is the one referring to the XLF and its cosmological evolution, and for model A1 we adopt the results of Mi99a. Strictly speaking, the XLF of Mi99a refers to the observed 0.5--2 keV luminosities and could be considered as the 0.5--2 keV XLF in the rest frame only by assuming simple power law spectra with energy index $\alpha=1$ (i.e. zero $K$--correction). Indeed, in the rest frame energy range seen by ROSAT at different redshifts, the spectra of type 1 AGNs assumed in our models do not differ significantly from a power law with $\alpha=1$. Since we refer the Mi99a XLF to unabsorbed AGNs, without correcting for the contribution of absorbed AGNs to the ROSAT counts, model A1 might be biased in favor of a soft XRB. We discuss the strength of this bias in connection with models A2 and B in the following. The absorption distribution in type 2 AGNs is no longer derived from best fitting, instead it is taken equal to the local one, as measured by Risaliti et al. (1999). The objects for which only a lower limit is available, $N_{\rm H}>10^{24}$ cm$^{-2}$, have been assigned to the bin $10^{24}<N_{\rm H}<10^{25}$ cm$^{-2}$. Because of the evidence that the $R$ ratio decreases with the intrinsic luminosity of the AGNs, at least locally, we introduce a change with respect to the canonical scenario: the XLF is divided in two luminosity regions as follows: \begin{displaymath} \phi(L_{\rm x})=\phi(L_{\rm x})e^{-\frac{L_{\rm x}}{L_s}}+\phi(L_{\rm x})(1-e^{-\frac{L_{\rm x}}{L_s}})\;, \end{displaymath} \noindent with the 0.5--2 keV $e$--folding luminosity set equal to $L_s$=10$^{44.3}$ erg s$^{-1}$, following Miyaji, Hasinger \& Schmidt (1999b; hereafter Mi99b). The first and second term represent the XLF of Sey 1s and QSO 1s, respectively, and, apart from the exponential factors, are equal to the Mi99a functions as given in Section 2.2. The XLF of Sey 2s and QSO 2s are $R_{\rm S}$ and $R_{\rm Q}$ times the XLF of the corresponding type~1 objects. In this parametrization we can explore various hypotheses, including for instance the effects of eliminating altogether the QSO 2s. Following Co95, and the experimental evidence referred to in Section 2.1, we assume that the basic spectrum for type 1 AGNs is a power law with energy index $\alpha=0.9$ and exponential cut off with $e$--folding energy $E_c=320$ keV. Below 1.5 keV the soft excess is modeled with a power law of index 1.3. A reflection component from the accretion disc has been included for Sey 1s with relative normalization $f_d=1.29$ (Co95). Beside the disc, we have also included for Sey 1s a torus reflection component which is normalized in accordance with the prescriptions of Ghisellini, Haardt \& Matt (1994). In type 1 AGNs the relative contribution of the torus at 30 keV is 29\% and 55\% for $N_{\rm H}=10^{24}$ and $10^{25}$ cm$^{-2}$, respectively. If we assume that the column density of the torus is approximately the same for all obscured lines of sight, from the measured $N_{\rm H}$ distribution we find that the torus contributes on the average 28\% at 30 keV. The same disc and torus reflection components of Sey 1s have been included also in the QSO spectra: this is against the evidence at low redshifts, but mimics the harder power law seen at high redshifts (Vignali et al. 1999), where most of the XRB is produced. If anything, this assumption tends to reduce the need for additional hard spectrum sources, thus strengthening our results. \begin{figure} \epsfig{file=8484_fig2.ps, width=9.cm, height=9.cm, angle=0} \caption{The fit to the XRB. The data of higher intensity below 6 keV are from ASCA (Gendreau et al. 1995), while those above 3 keV are a compendium of various experiments including HEAO--1 A2 (Gruber 1992). The solid line is the overall fit while the other labeled curves represent the contributions of the different classes of sources. The radiation excess due to the iron line is also shown.} \end{figure} Sey 2 spectra have been computed for different amounts of intrinsic absorption (log$N_{\rm H}$=21.5, 22.5, 23.5, 24.5, 25.5) to cover all the observed column densities. In the Compton thin regime the adopted spectrum is that of Sey 1s with a photoelectric cut off and a lower amount of disc reflection ($f_d=0.88$, Co95). In this regime the component reflected by the torus does not contribute significantly to the observed radiation (5\% at 30 keV for log$N_{\rm H}$=23.5, inclusive of orientation effects). For the sources with log$N_{\rm H}$=25.5 we have adopted a pure reflection continuum. The normalization of the spectrum is determined so as to reproduce the contribution of thick tori to the flux of Sey 1s (55\% at 30 keV) after correcting for orientation effects (Ghisellini, Haardt \& Matt, 1994). This approach predicts that the 2--10 keV continuum luminosity of completely Compton thick sources is about 2\% of the typical luminosity of Sey 1s, in agreement with the results of Maiolino et al. (1998). A composite reflected/transmitted spectrum has been considered for Circinus--like sources with log$N_{\rm H}$=24.5, where the reflected and transmitted components have been normalized in analogy with the previous cases. \footnote{For log $N_{\rm H}$=24.5 the effects of Compton scattering begin to be important. We have checked the error introduced by our approximation with respect to the MonteCarlo simulations of Matt, Pompilio \& La Franca (1999). The counts remain unaffected, while the model XRB at 30 keV should be reduced by 10--15\%, and even more type 2s should be included in order to maintain the agreement with the data.} We have modeled the soft excess of Sey 2s with a power law of index 1.3 and a normalization at 1 keV which is 3\% of the primary de--absorbed power law. In analogy with type 1s, the spectra of QSO 2s --if at all present-- are assumed to be identical to those of Sey 2s. Finally, we have added to the input spectra an iron emission line at 6.4 keV. Following Gilli et al. (1999) we have considered lines with different equivalent widths according to the spectral absorption, and have not included the iron line in the spectra of QSOs. \begin{figure} \epsfig{file=8484_fig3.ps, width=9.cm, height=9.cm, angle=0} \caption{The predictions of model A1 compared with the source counts observed in the 0.5--2 keV band by ROSAT. In this and in the following Figures, source counts are plotted as $S_{14}^{1.5}\times N(>S)$, where $S_{14}$ is the flux in units of $10^{-14}$ erg s$^{-1}$ cm$^{-2}$. The AGN data (squares) are adapted from Mi99a. Cluster counts (stars) at $1\times 10^{-14}$ erg s$^{-1}$ cm$^{-2}$ and $6\times 10^{-14}$ erg s$^{-1}$ cm$^{-2}$ are from Rosati et al. (1995) and Jones et al. (1998), respectively. The cluster surface density at $1.2\times 10^{-11}$ erg s$^{-1}$ cm$^{-2}$ is adapted from De Grandi et al. (1999). The fluctuation limits to the total counts (dotted area) are adapted from Hasinger (1998).} \end{figure} In our model A1 we assume $R_{\rm Q}=0$, i.e. we do not include QSO~2s. The Mi99a XLF are integrated in the range $10^{41}<L_{\rm x}<10^{49}$ erg s$^{-1}$ up to $z_{max}=4.6$. The contribution of clusters of galaxies has been included by considering thermal bremsstrahlung spectra with a distribution of temperatures. We have adopted the 2--10 keV luminosity vs temperature relation of David et al. (1993), and the 2--10 keV X--ray luminosity function of Ebeling et al. (1997). The cluster XLF is assumed not to evolve, and is integrated in the range $10^{42}<L_{\rm x}<10^{47}$ erg s$^{-1}$ up to $z_{max}=2$. The overall XRB spectrum resulting from the model is shown in Fig.~2 as a solid line, which is the sum of the contibutions of the other labeled curves. In order to fit the observed XRB spectrum we need a ratio $R_{\rm S}=4.2$, in good agreement with the local value. Above $\sim 1$ keV, where the XRB is completely extragalactic, the model provides a good fit to the data from ASCA (Gendreau et al. 1995) and the compilation of Gruber (1992) based on HEAO--1 A2 measurements. The contribution of the AGN iron line to the model XRB is found to be less than 7\% at $\sim 6.4/(1+z_{cut})$ in agreement with the results of Gilli et al. (1999) obtained in a different framework (PLE). Clusters of galaxies are found to contribute to the model XRB by $\sim 12\%$ at 1 keV, in agreement with the results of Oukbir, Bartlett \& Blanchard (1997). \subsection{The X--ray source counts} We now compare the predictions of model A1 with the observed source counts in different X--ray bands. The results in the soft 0.5--2 keV band are shown in Fig.~3. The expected AGN counts, which are dominated by unabsorbed sources, agree with the data of Mi99a; the expected cluster counts agree with the data of Jones et al. (1998), Rosati et al. (1995), and De Grandi et al. (1999). Since the XLF and its evolution are derived from the ROSAT counts, this is no more than a self--consistency check; the slight overprediction at low fluxes ($\sim$30\% at $\sim 2\times 10^{-15}$ erg cm$^{-2}$ s$^{-1}$) is due to the $K$--displaced type~2 objects, as anticipated previously. \begin{figure} \epsfig{file=8484_fig4.ps, width=9.cm, height=9.cm, angle=0} \caption{The predictions of model A1 (lines are as in Fig.~3) compared with the hard counts between 2 and 10 keV. The open square at $\sim 3\times 10^{-11}$ erg s$^{-1}$ cm$^{-2}$ represents the AGN surface density in the Piccinotti et al. (1982) sample corrected by 20\% after Co95. ASCA data are represented by the filled squares (Cagnoni et al. 1998), the open circle at $\sim 2\times 10^{-13}$ erg s$^{-1}$ cm$^{-2}$ (Ueda et al. 1998) and the open square at $\sim 4\times 10^{-14}$ erg s$^{-1}$ cm$^{-2}$ (Ogasaka et al. 1998). The open triangle at $5\times 10^{-14}$ erg s$^{-1}$ cm$^{-2}$ is from BeppoSAX (Giommi et al. 1998).} \end{figure} In the hard 2--10 keV band the predictions of the model have to be compared with the results of HEAO--1 A2 (Piccinotti et al. 1982), ASCA (Cagnoni, Della Ceca \& Maccacaro 1998; Ogasaka et al. 1998; Ueda et al. 1998) and BeppoSAX (Giommi et al. 1998). At the flux limit of $S\simeq 3\times 10^{-11}$ erg s$^{-1}$ cm$^{-2}$ Piccinotti et al. (1982) found that AGNs and clusters of galaxies have the same surface density of $1.1\times 10^{-3}$ deg$^{-2}$. However, the AGN density found by these authors is likely to be overestimated by $\sim 20\%$ due to the local supercluster (Co95). As shown in Fig.~4, after the Piccinotti et al. point is corrected by 20\%, the model is in agreement with the data within $1\sigma$. On the contrary the disagreement cannot be solved at fainter fluxes. At $S\sim 2\times 10^{-13}$ erg s$^{-1}$ cm$^{-2}$ the AGN surface density expected in our model is about a factor of 2 lower than the measurement of Cagnoni et al. (1998). This corresponds to a $\sim 2\sigma$ discrepancy. When comparing the model with the data of the ASCA Large Sky Survey (Ueda et al. 1998) the discrepancy is even larger. The situation is worse still in the 5--10 keV band. The only available counts in this band are from the HELLAS survey performed by BeppoSAX (Giommi et al. 1998; Comastri et al. 1999). At the flux of $\sim 2\times 10^{-13}$ erg s$^{-1}$ cm$^{-2}$ the observed surface density is $2.7\pm0.7$ deg$^{-2}$, which is a factor of 4 ($\sim 3\sigma$) above the predictions (Fig.~5). \begin{figure} \epsfig{file=8484_fig5.ps, width=9.cm, height=9.cm, angle=0} \caption{The expected counts in model A1 (lines are as in Fig.~3) compared with the 5--10 keV BeppoSAX data (Giommi et al. 1998; Comastri et al. 1999). The open square represents the 20\% corrected AGN surface density found by Piccinotti et al. (1982), converted into the 5--10 keV band. The conversion has been performed by assuming a mean of the input spectra of our model, weighted for the $N_{\rm H}$ distribution of the Piccinotti et al. sample (Schartel et al. 1997).} \end{figure} \subsection{Correction for absorbed sources} One of the main assumptions of model A1 is that the XLF and evolution derived by Mi99a refer to unabsorbed AGNs. However, as discussed in Sect. 2.2, this is likely not the case. In order to evaluate the effects of our assumption, we have computed a different variant, A2, which adopts the XLF and LDDE of Mi99b. These authors allow self--consistently within their model for the $K$--correction, and for the absorbed sources which, especially at faint fluxes and high redshifts, appear in the ROSAT counts; thus the parameters they provide refer to unabsorbed sources only. Of course, our model is different from theirs, and the self--consistency is lost; however, this is likely to be a higher order effect, and for a first order estimate we can include their parameters in our computation. The results are shown in Fig.~6. Note that around 1 keV the unabsorbed sources produce 30\% of the XRB, exactly as in Mi99b. Note also that in order to fit the XRB spectrum a ratio $R_{\rm S}=13$ is now required, which is much higher than the local value and implies additional hard spectrum sources. Since the contribution of type 2s has increased with respect to A1, in order to make up for the reduced contribution of type 1s, the mean spectrum of the population producing the XRB is harder, and the discrepancies between the model predictions and the hard counts are somewhat reduced (though not completely eliminated). \begin{figure} \epsfig{file=8484_fig6a.ps, width=9.cm, height=7.cm, angle=0} \epsfig{file=8484_fig6b.ps, width=9.cm, height=7.cm, angle=0} \epsfig{file=8484_fig6c.ps, width=9.cm, height=7.cm, angle=0} \caption{From top to bottom, the same as in Figs.~2, 4, 5 but for model A2. Curves are as in the previous Figures. Note the very hard spectrum of the model XRB due to the high contribution of absorbed sources.} \end{figure} \section{Discussion} The main difference between models A1 and A2 is the fractional contribution of type 1 AGNs to the XRB; this contribution is dominated by objects close to the XLF break at redshifts close to $z_{cut}$, and is not well constrained by the data. In the former model 60\% of the 1--keV XRB is due to type 1s, so the local value of the type ratio $R_{\rm S}$ is sufficient to account for the entire XRB; the average spectrum, though, is too soft, and the softness shows up in a marginal discrepancy with the XRB spectrum at $>40$~keV (Fig.~2), and unacceptable discrepancies with the hard counts (Figs.~4 and 5). In the latter model the type 1s account for only 30\% of the 1--keV XRB, and making up the entire XRB requires an $R_{\rm S}$ much larger than the local value; now the average spectrum is harder, the shape discrepancy disappears and the count discrepancies are reduced (Fig. 6). By extrapolating from these two models we can make qualitative predictions on still different parametrizations of the Mi99a sample: for instance, models adopting density evolution with a dependence on luminosity weaker than Mi99a would predict a type 1 soft X--ray contribution higher than 60\%, and would miss the hard counts by factors larger than Figs.~4 and 5. Density evolution with a dependence on luminosity stronger than Mi99b (if at all acceptable) would require $R_{\rm S}>13$, which is already three times the local value. The main result of our analysis is precisely this one: no matter which variant is adopted for the XLF and the evolution, the models which incorporate the most recent observations within the standard prescriptions always produce some discrepancy. The discrepancy may appear as an underprediction of the observed hard counts, or a type 2 to type 1 ratio higher than the observed local value, but in all cases it points to additional hard spectrum sources at intermediate or high redshifts. For the sake of completeness, we present in the Appendix a PLE model with $R_{\rm Q}=R_{\rm S}$ (model B): this region of the parameter space is not favored by the most recent data, but was adopted in practically all previous works on the XRB. Note that model B has the same type 1 soft X--ray contribution as model A2 (30\%), but the type 2 contribution here is due to higher luminosity sources, which show up in higher flux bins: indeed, the XRB spectrum is well fitted, the counts in the ASCA band are matched, and the discrepancy with the HELLAS counts is reduced to $2\sigma$. The cost to be paid is a number density $R_{\rm Q}=7.7$, which --again-- is definitely higher than the local upper limit $R_{\rm Q}<2$. Irrespective of the plausibility of PLE and QSO~2s, we stress that even model B results in a discrepancy, and the discrepancy is concordant with the results of the other, less controversial variants. A population of absorbed or hard spectrum AGNs evolving more rapidly than the type~1s could accomodate all the problems discussed above. In this context one should be reminded that the hard counts already resolve $\sim$30\% of the XRB at fluxes $\sim~5\times 10^{-14}$ erg cm$^{-2}$ s$^{-1}$, so they must converge rapidly just below these values. The optical identifications of the counts in the 2--10 keV and 5--10 keV bands are still largely incomplete. Up to now 34 X--ray sources detected in the ASCA LSS survey (Ueda et al. 1998) have been identified (Akiyama et al. 1998), and 28 objects turned out to be AGNs. They are 22 broad line AGNs (type 1--1.5) with $0<z<1.7$, and 6 type 2 AGNs with $0<z<0.7$. The number of identified sources of the BeppoSAX HELLAS survey is lower, but the distribution of the AGNs seems similar to the ASCA one: 7 broad line QSOs with $0.2<z<1.3$ and 5 Seyferts 1.8--1.9 with $0.04<z<0.34$ (Fiore et al. 1999). If one accepts these low redshift type 2 identifications, one has to find a physical reason for a convergence so recent in comparison with all other AGNs (BL--Lacs excepted) and star forming galaxies. Alternatively, one could rely on the poor statistics to maintain that the hard counts are mostly due to optically empty fields, containing very absorbed, very powerful sources at redshifts $>1$. There are prospective candidates in both scenarios. In the low--$z$ hypothesis one could assume that ``normal'' Seyfert 2 galaxies evolve more rapidly than type 1s, so that $R_{\rm S}$ increases with redshift up to the required value. Not only the number ratio, but also the $N_{ \rm H}$ distribution could change with cosmic time (Franceschini et al. 1993). Local Sey 2s are associated with a star formation activity higher than Sey 1 and normal galaxies ( Maiolino et al. 1995; Rodr\'{\i}guez-Espinoza et al. 1986), so this assumption would have interesting implications on the star formation history. One could also invoke Advection Dominated Accretion Flows (ADAFs, Di Matteo et al. 1998) whose luminosity is proportional to $\dot M^2$, where $\dot M$ is the mass accretion rate, and which become normal QSOs at large $\dot M$: thus, they should evolve more rapidly than normal AGNs at intermediate redshifts, and should undergo a change of class at high redshifts. In the high--$z$ hypothesis one should resort to ULIRGs, which indeed are absorbed and powerful, and appear to evolve as fast as required [$(1+z)^{7.6\pm3.2}$, Kim \& Sanders 1998]. As mentioned before there is evidence that the IR emission of ULIRGs is powered both by starburst and AGN processes; Kim, Veilleux \& Sanders (1998) and Lutz et al. (1998) find that the fraction of AGN--powered infrared luminous galaxies increases with the bolometric luminosity, and reaches 30--50\% in the ULIRG range. While normal starbursts are inefficient emitters in the hard X--rays, obscured AGNs in ULIRGs could easily explain the hard counts. Finally, it should be noted that the optical identifications of the hard X--ray counts, scanty as they are, suggest that at the given X--ray flux the type 1s are more numerous than the type 2s. Concordant evidence is provided by recent BeppoSAX observations of the Marano field and the Lockman hole (Hasinger et al. 1999): most of the BeppoSAX sources have ROSAT counterparts, which in most cases are optically identified with type 1 AGNs. This type composition is in agreement with the predictions of, for instance, model A1 (Figs.~4 and 5). But if the hard counts in excess of the model are attributed entirely to obscured AGNs, then the predicted type ratio is reversed, with the type 2s more numerous than the type 1s. The numbers involved are too small to draw any conclusion, however they seem to suggest that some of the hard counts are due to flat X--ray spectrum sources with type 1 optical spectra; a few similar sources might have been found already in the ASCA LSS (Akiyama et al. 1998). Clearly, a decisive progress in this area will require more numerous and more secure identifications of hard X--ray counts; given the various hypotheses, counterparts should be looked for not only at optical wavelengths, but also in the infrared and submillimeter domains, where AGN--dominated ULIRGs should be conspicuous. \section{Summary and conclusions} In this paper we have shown that the standard prescriptions for synthesizing the XRB from the integrated emission of AGNs are not consistent with a number of recent observational constraints, and some of them must be relaxed. We have worked out models (A1 and A2) which take into account detailed input spectra of AGNs, the $N_{\rm H}$ distribution observed in local Seyfert 2s, and the XLF and evolution newly determined from the largest ROSAT sample. The latter data do not define a unique parametrization, and the two models explore different variants. As prescribed by the standard model, the XLF and evolution of type 2 AGNs are taken from type 1s, and the spectra of both types are taken independent of redshift; the only fitting parameter is the number ratio $R$ of type 2s to type 1s. We find that model A1 reproduces the XRB and the soft counts with a ratio $R$ compatible with the local value, but underestimates the hard counts. Model A2 is less discrepant as far as the counts are concerned, but requires a ratio $R$ definitely larger than observed locally. We have also computed a model adopting a canonical pure luminosity evolution (model B). In agreement with the results of Co95, model B can reproduce the XRB, the soft X--ray counts and the ASCA hard counts in the 2--10 keV band. It is also consistent within 2$\sigma$ (or discrepant at 2$\sigma$) with the preliminary BeppoSAX counts in the 5--10 keV band. Nevertheless, it requires a number of type 2 QSOs much higher than the local upper limit, and perhaps already ruled out by the deep X--ray surveys. The discrepancies found in all models are to some extent model dependent, but all of them point in the same direction, and suggest that hard spectrum sources at intermediate or high redshifts are needed in addition to the predictions of the standard scenario. The X--ray spectrum of these additional sources could be flattened by absorption, or could be intrinsically hard. In the former hypothesis reasonable candidate counterparts could be rapidly evolving, ``normal'' Seyfert 2s. One should also note that a fraction of ULIRGs seem to be powered by AGNs, and their cosmological evolution seems faster than that of unabsorbed QSOs. The alternate hypothesis could instead require the presence of ADAFs. Optical identifications of the hard X--ray sources are still largely incomplete and do not allow yet to decide between the various possibilities. \begin{acknowledgements} We are grateful to A. Comastri and G. Zamorani for a careful reading of the manuscript, and to T. Miyaji G. Hasinger and M. Schmidt for permission to use their LDDE model in advance of publication. Our presentation was greatly improved by the comments of the referee, Prof. G. Hasinger. This work was partly supported by the Italian Space Agency (ASI) under grant ARS--98--116/22 and by the Italian Ministry for University and Research (MURST) under grant Cofin98--02--32. \end{acknowledgements}
\section{Introduction} The CPT theorem \cite{cpt} and Lorentz symmetry have both been tested to very high accuracy in a variety of physical systems.\cite{pdg} Papers presented at this meeting have described experiments in astrophysical, nuclear, particle, and atomic systems, all of which provide very stringent bounds on possible CPT or Lorentz breaking. To date, the best bound on CPT has been obtained in particle-physics experiments involving neutral kaons. Since different particle sectors are largely independent, it is important to consider possible CPT and Lorentz breaking in all particle sectors, including mesons, leptons, baryons, and gauge bosons. While kaon experiments clearly provide the best test of CPT in the meson sector, it is interesting to note that the sharpest tests of CPT breaking in both the lepton and baryon systems have not been obtained in high-energy particle experiments. Instead, low-energy experiments on single isolated particles in Penning traps have yielded the best bounds on CPT in the lepton and baryon sectors. These experiments involve comparisons of electrons and positrons, protons and antiprotons, and hydrogen ions and antiprotons. One consequence of CPT invariance is that particles and antiparticles have equal charge-to-mass ratios and gyromagnetic ratios. Experiments in Penning traps are ideally suited for making very precise comparisons of these quantities. A Penning trap \cite{penning} captures a single charged particle in the cavity between two cap electrodes and a ring electrode. The electrodes are charged and create a quadrupole electric field. A static magnetic field is created using external current coils. A charged particle in the trap is bound due to the combination of the static electric and magnetic fields. By nesting two Penning traps, particles and antiparticles can be probed and quickly switched in the same magnetic field, but with the electric field reversed. The dominant structure of the energy levels for spin-${\textstyle{1\over 2}}$ particles at low temperature is that of relativistic Landau levels, with two ladders of energies for the two spin states. Transition frequencies between these levels can be measured with very high precision. Typically, two types of frequency comparisons of particles and antiparticles are possible in Penning traps. They involve making accurate measurements of the cyclotron frequency $\omega_c$ (for transitions between Landua levels with no spin flip) and the anomaly frequency $\omega_a$ (for transitions between Landua levels accompanied by a spin flip) of single isolated particles confined in the trap. The first type of experiment is an anomalous magnetic moment or $g-2$ experiment. These compare the ratio $2 \omega_a / \omega_c$ for particles and antiparticles. In the context of conventional quantum electrodynamics, this ratio equals $g-2$ for the particle or antiparticle. The second type of experiment compares values of $\omega_c \sim q/m$, where $q>0$ is the magnitude of the charge and $m$ is the mass. These therefore involve comparisons of the charge-to-mass ratios for the particle and antiparticle. Both $g-2$ and charge-to-mass ratio experiments have been performed with electrons and positrons. With protons and antiprotons, however, only charge-to-mass ratio comparisons have been performed. Because the magnetic moments of protons and antiprotons are much weaker than those of electrons and positrons, $g-2$ experiments with protons and antiprotons require much lower temperatures and greater sensitivity for detecting spin-flip transitions. Although these $g-2$ experiments with protons and antiprotons have not been performed to date some suggestions for making these experiments feasible in the future exist in the literature.\cite{hw,qg} To compare the sensitivities of CPT tests in Penning traps with those in the meson system, we list some of the relevant figures of merit. The conventional figure of merit in the neutral kaon system is given by \begin{equation} r_K \equiv \fr {|m_K - m_{\overline{K}}|} {m_K} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 2 \times 10^{-18} \quad , \label{rK} \end{equation} whereas for $g-2$ experiments with electrons and positrons,\cite{vd} electron-positron charge-to-mass ratio experiments,\cite{schwin81} and charge-to-mass ratio experiments with protons and antiprotons,\cite{gg1,gg2} respectively, the conventional figures of merit are given as \begin{equation} r_g^e \equiv \fr {|g_{e^-} - g_{e^+}|} {g_{\rm avg}} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 2 \times 10^{-12} \quad , \label{rg} \end{equation} \begin{equation} r_{q/m}^e \equiv \fr {\left|({q_{e^-}}/{m_{e^-}}) - ({q_{e^+}}/{m_{e^+}})\right|} {|q/m|_{\rm avg}} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 1.3 \times 10^{-7} \quad , \label{rqme} \end{equation} \begin{equation} r_{q/m}^p \equiv \fr {|(q_p/m_p) - (q_{\overline{p}}/m_{\overline{p}})|} {|q/m|_{\rm avg}} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 1.5 \times 10^{-9} \quad . \label{rqmp} \end{equation} Recently, an experiment comparing the cyclotron frequencies of hydrogen ions $H^-$ and antiprotons has been performed in a Penning trap.\cite{gg2} This experiment has the advantage that both particles in the trap have the same electric charge, thereby reducing systematic errors associated with reversing the sign of the electric field. An improved charge-to-mass ratio comparison for protons and antiprotons has been obtained from these results, which is given by \begin{equation} r_{q/m}^{H^-} \equiv \fr {|(q_p/m_p) - (q_{\overline{p}}/m_{\overline{p}})|} {|q/m|_{\rm avg}} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 9 \times 10^{-11} \quad . \label{rqmH} \end{equation} Measurements of frequencies in Penning traps typically have parts-per-billion (ppb) accuracies, which are four or five orders of magnitude better than the measurements made in kaon experiments. This raises some interesting questions concerning the sensitivity of these experiments to different possible types of CPT breaking. One goal of this work is to understand the Penning-trap experiments better and to address the question of why they do not provide sharper tests of CPT. To accomplish this, we must work in the context of a theoretical framework that permits CPT breaking. Such a framework has been developed by Colladay and Kosteleck\'y.\cite{ck} In the following sections, the parts of the framework providing an extension of quantum electrodynamics are described, and the results of our theoretical analysis of CPT and Lorentz tests in Penning traps are presented. In particular, we analyze $g-2$ experiments on electrons and positrons,\cite{bkr1} charge-to-mass ratio experiments on protons and antiprotons, and comparisons of cyclotron frequencies for $H^-$ and antiprotons.\cite{bkr2} Since the framework we use includes both a CPT-violating sector and a CPT-preserving sector (both of which violate Lorentz symmetry) in addition to investigating the sensitivity of Penning-trap experiments to CPT, we also examine how these experiments test CPT-preserving Lorentz symmetry. \section{Theoretical Framework} The theoretical framework of Colladay and Kosteleck\'y \cite{ck} is an extension of the $SU(3) \times SU(2) \times U(1)$ standard model. It originates from the idea of spontaneous CPT and Lorentz breaking in a more fundamental theory.\cite{kps} This type of CPT violation is a possibility in string theory because the usual axioms of the CPT theorem do not apply to extended objects like strings. In a theory with spontaneous symmetry breaking, the dynamics of the action remains CPT invariant, which means the framework can preserve desirable features of quantum field theory such as gauge invariance, power-counting renormalizability, and microcausality. CPT and Lorentz violation occurs only in the solutions of the equations of motion. This mechanism is similar to the spontaneous breaking of the electroweak theory in the standard model. In our analysis of Penning-trap experiments, we use a restriction of the full particle-physics framework to quantum electrodynamics. The effects of possible CPT and Lorentz violation in this context lead to a modification of the Dirac equation. The modified form (in units with $\hbar = c = 1$) is given by $$ ( i \gamma^\mu D_\mu - m - a_\mu \gamma^\mu - b_\mu \gamma_5 \gamma^\mu - {\textstyle{1\over 2}} H_{\mu \nu} \sigma^{\mu \nu} \quad\quad\quad\quad\quad $$ \begin{equation} \quad\quad\quad\quad\quad + i c_{\mu \nu} \gamma^\mu D^\nu + i d_{\mu \nu} \gamma_5 \gamma^\mu D^\nu) \psi = 0 \quad . \label{dirac} \end{equation} Here, $\psi$ is a four-component spinor, $A_\mu$ is the electromagnetic field, $i D_\mu \equiv i \partial_\mu - q A_\mu$ is the covariant derivative, and $a_\mu$, $b_\mu$, $H_{\mu \nu}$, $c_{\mu \nu}$, $d_{\mu \nu}$ are the parameters describing possible violations of CPT and Lorentz symmetry. The terms involving $a_\mu$, $b_\mu$ break CPT and those involving $H_{\mu \nu}$, $c_{\mu \nu}$, $d_{\mu \nu}$ preserve CPT, while all five terms break Lorentz symmetry. Since no CPT or Lorentz breaking has been observed in experiments to date, the quantities $a_\mu$, $b_\mu$, $H_{\mu \nu}$, $c_{\mu \nu}$, $d_{\mu \nu}$ must all be small. We can estimate the suppression scale for these quantities by taking the scale governing the fundamental theory as the Planck mass $m_{\rm Pl}$ and the low-energy scale as the electroweak mass scale $m_{\rm ew}$. The natural suppression scale for Planck-scale effects in the standard model would then be of order $m_{\rm ew}/m_{\rm Pl} \simeq 3 \times 10^{-17}$. If instead, we consider the electron mass scale as the low-energy scale, we obtain $m_{\rm e}/m_{\rm Pl} \simeq 5 \times 10^{-23}$. Since a more fundamental theory (which would determine these parameters more precisely) remains unknown, these ratios give only an approximate indication of the suppression scale. We use this theoretical framework to analyze comparative tests of CPT and Lorentz symmetry on particles and antiparticles in Penning traps. Some technical issues include the following. First, the time-derivative couplings in Eq.\ \ref{dirac} alter the standard procedure for obtaining a hermitian quantum-mechanical hamiltonian operator. To overcome this, we perform a field redefinition at the lagrangian level that eliminates the additional time derivatives. Second, to obtain a hamiltonian for the antiparticle, we use charge conjugation to find the Dirac equation describing the antiparticle. Perturbative calculations can then be carried out for both the particle and antiparticle, and the leading-order effects of CPT and Lorentz breaking can be obtained. \section{ Electron-Positron Experiments} Experiments testing CPT in the electron-positron system compare cyclotron frequencies $\omega_c$ and anomaly frequencies $\omega_a$ of particles and antiparticles in a Penning trap. A result of the CPT theorem is that electrons and positrons of opposite spin in a Penning trap with the same magnetic fields but opposite electric fields should have equal energies. The experimental relations $g-2 = 2 \omega_a / \omega_c$ and $\omega_c = qB/m$ provide connections to the quantities $g$ and $q/m$ that appear in the figures of merit $r_g^e$ and $r_{q/m}^e$. Calculations are performed using Eq.\ \ref{dirac} to obtain possible shifts in the energy levels due to either CPT-breaking or CPT-preserving Lorentz violation. The effectiveness of Penning-trap experiments on electrons and positrons as tests of both CPT-breaking and CPT-preserving Lorentz violation can then be analyzed. From the calculated energy shifts we determine how the frequencies $\omega_c$ and $\omega_a$ are affected and whether the conventional figures of merit are appropriate. The dominant contributions to the energy of an electron or positron in a Penning trap come from interactions with the constant magnetic field of the trap. Interactions with the quadrupole electric field generate smaller effects. In a perturbative treatment, the dominant CPT- and Lorentz-breaking effects can therefore be obtained by working with the relativistic Landau levels as unperturbed states. Conventional perturbations, such as the usual corrections to the anomalous magnetic moment, do not break CPT or Lorentz symmetry and are the same for electrons and positrons. Any violations of CPT or Lorentz symmetry result in either differences between electrons and positrons or in unconventional effects such as diurnal variations in measured frequencies. Our calculations \cite{bkr1} show that the leading-order corrections to the energies $E_{n,s}^{e^-}$ for the electron and $E_{n,s}^{e^+}$ for the positron due to the effects of CPT and Lorentz violation are $$ \quad \delta E_{n,\pm 1}^{e^-} \approx a_0^e \mp b_3^e - c_{00}^e m_e \pm d_{30}^e m_e \pm H_{12}^e $$ \begin{equation} \quad\quad\quad\quad\quad\quad - {\textstyle{1\over 2}} (c_{00}^e + c_{11}^e +c_{22}^e) (2n + 1 \pm 1) \omega_c \quad . \label{Eelec} \end{equation} $$ \quad \delta E_{n,\pm 1}^{e^+} \approx - a_0^e \mp b_3^e - c_{00}^e m_e \mp d_{30}^e m_e \mp H_{12}^e $$ \begin{equation} \quad\quad\quad\quad\quad\quad - {\textstyle{1\over 2}} (c_{00}^e + c_{11}^e + c_{22}^e) (2n + 1 \mp 1) \omega_c \quad . \label{Epos} \end{equation} From these we find the modified transition frequencies including the leading-order effects of CPT and Lorentz breaking. These are given by \begin{equation} \omega_c^{e^-} \approx \omega_c^{e^+} \approx (1 - c_{00}^e - c_{11}^e - c_{22}^e) \omega_c \quad , \label{wcelec} \end{equation} \begin{equation} \omega_a^{e^\mp} \approx \omega_a \mp 2 b_3^e + 2 d_{30}^e m_e + 2 H_{12}^e \quad . \label{waelec} \end{equation} Here, $\omega_c$ and $\omega_a$ represent the unperturbed electron or positron frequencies, while $\omega_c^{e^\mp}$ and $\omega_a^{e^\mp}$ denote the frequencies including corrections. Superscripts have been added to the parameters $b_\mu$, etc.\ to denote that they describe the electron-positron system. From these relations we find the electron-positron differences for the cyclotron and anomaly frequencies to be \begin{equation} \Delta \omega_c^e \equiv \omega_c^{e^-} - \omega_c^{e^+} \approx 0 \quad , \label{delwc} \end{equation} \begin{equation} \Delta \omega_a^e \equiv \omega_a^{e^-} - \omega_a^{e^+} \approx - 4 b_3^e \quad . \label{delwa} \end{equation} In the context of this framework, comparisons of cyclotron frequencies to leading order do not provide a signal for CPT or Lorentz breaking, since the corrections to $\omega_c$ for electrons and positrons are equal. On the other hand, comparisons of $\omega_a$ provide unambiguous tests of CPT. We also find that there are no leading-order corrections due to CPT or Lorentz violation to the $g$ factors for either electrons or positrons. This leads to some unexpected results concerning the figure of merit $r_g$ in Eq.\ \ref{rg}. With $g_{e^-}$ and $g_{e^+}$ equal to leading order, we find that $r_g$ vanishes, which would seem to indicate the absence of CPT breaking. However, this conclusion would be incorrect because the framework we are working in contains explicit CPT violation. In addition, calculations in the context of our framework show that with $\vec b \ne 0$ the experimental ratio $2 \omega_a / \omega_c$ depends on the magnetic field and is undefined in the limit of a vanishing $B$ field. Because of this, the usual relation $g-2 = 2 \omega_a / \omega_c$ does not hold in the presence of CPT violation. For these reasons, we conclude that in the context of our framework the figure of merit $r_g$ in Eq.\ \ref{rg} is inappropriate, and an alternative is suggested next. Since it is a prediction of the CPT theorem that electron and positron states of opposite spin in the same magnetic field have equal energies, we propose as a model-independent figure of merit \begin{equation} r^e_{\omega_a} \equiv \fr{|{E}_{n,s}^{e^-} - {E}_{n,-s}^{e^+}|} {{E}_{n,s}^{e^-} } \quad . \label{re} \end{equation} Here, ${E}_{n,s}^{e^\mp}$ are the Landau-level energies, with $n$ denoting the Landau level, and $s=\pm 1$ the spin. In the context of our framework, we find $r^e_{\omega_a} \approx |2 b_3^e | / m_e$, which can be bounded by experiments. Assuming ppb frequency resolutions, we estimate as a bound, \begin{equation} r^e_{\omega_a} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10^{-21} \quad . \label{relim} \end{equation} The figure of merit $r^e_{\omega_a}$ is compatible with the corresponding figure of merit $r_K$ which describes the neutral-kaon system. This is because both figures of merit involve energy ratios, which makes comparisons across experiments more meaningful. In contrast, the figures of merit $r_g^e$ and $r_K$ involve ratios of different physical quantities. Moreover, our estimated bound for $r^e_{\omega_a}$ is more in line with the high precision that is experimentally accessible in frequency measurements in a Penning trap and appears to improve on the bound given in terms of $r_K$. It is important to stress, however, that performing CPT tests in the meson sector remains essential because CPT violation in this sector is controlled by distinct CPT-violating parameters that appear only in the quark sector.\cite{k98} Alternative signatures of CPT and Lorentz violation can be considered as well.\cite{bkr2} These include possible diurnal variations in the anomaly and cyclotron frequencies. We estimate bounds for these quantities based on ppb accuracies in $\omega_a$ and $\omega_c$. They are \begin{equation} r^e_{\omega_{a}^{\mp},\rm diurnal} \approx \fr {2|\mp b_3^e + d_{30}^e m_e + H_{12}^e|}{m_e} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10^{-21} \quad , \label{rwadiurnal} \end{equation} \begin{equation} r^e_{\omega_c, \rm diurnal} \approx \fr {|c_{11}^e + c_{22}^e| \omega_c} {m_e} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10^{-18} \quad . \label{rwcdiurnal} \end{equation} Tests for these effects would provide bounds on some of the components of the CPT-preserving but Lorentz-violating parameters $c_{\mu \nu}^e$, $d_{\mu \nu}^e$, and $H_{\mu \nu}^e$. One type of experiment searching for diurnal variations would involve the electron alone or the positron alone in a Penning trap. Diurnal variations in the cyclotron and anomaly frequencies would occur because the spatial components of the parameters in Eq.\ref{waelec} would change as the Earth rotates. A figure of merit can be defined which is based on the relative size of the diurnal energy variations. First, consider the following quantities for the electron and positron: \begin{equation} \Delta^{e}_{\omega_a^{e^-}} \equiv \fr {|{E}_{0,+1}^{e^-} - {E}_{1,- 1}^{e^-}|} {{E}_{0,-1}^{e^-}} \quad , \quad \quad \Delta^{e}_{\omega_a^{e^+}} \equiv \fr {|{E}_{0,-1}^{e^+} - {E}_{1,+1}^{e^+}|} {{E}_{0,+1}^{e^+}} \quad . \label{Deleorpdnlom} \end{equation} Suitable figures of merit $r^e_{\omega_{a}^-,\rm diurnal}$ and $r^e_{\omega_{a}^+,\rm diurnal}$ can then be defined as the amplitude of the diurnal variations in $\Delta^{e}_{\omega_a^{e^-}}$ and $\Delta^{e}_{\omega_a^{e^+}}$, respectively. In the context of the framework we are using, we compute that \begin{equation} r^e_{\omega_{a}^{\mp},\rm diurnal} \approx \fr {2|\mp b_3^e + d_{30}^e m_e + H_{12}^e|}{m_e} \quad . \end{equation} Among the experimental issues involved in obtaining a bound on $r^e_{\omega_{a}^{\mp},\rm diurnal}$ is maintaining stability in the magnetic field. For example, drifts in the magnetic field at a level of about 5 parts in $10^9$ over the duration of the experiment would correspond to a 1 Hz frequency resolution. The data would then need to be plotted and fitted as a function of the orientation of the magnetic field with respect to a celestial coordinate system. Bounds obtained in an experiment on electrons alone or positrons alone would involve the combination $\mp b_3^e + d_{30}^e m_e + H_{12}^e$ of parameters in the standard-model extension. The dominant signal would therefore involve corrections to the anomaly and cyclotron frequencies which exhibit periodicities of approximately 24 hours. Subleading order corrections might exhibit 12-hour periodicities. However, these effects would be suppressed relative to the leading-order effects. All three of the quantities $b_3^e$, $d_{30}^e$, and $H_{12}^e$ break Lorentz symmetry, but only the coupling $b_3^e$ breaks CPT. If a signal were detected, it would indicate Lorentz breaking but not necessarily CPT violation. A subsequent experiment comparing anomaly frequencies of electrons and positrons which would bound the CPT-breaking parameter $b_3^e$ in isolation would then need to be performed. A preliminary analysis of this type of experiment on electrons alone has recently been performed.\cite{mit} With a precision of approximately 1 Hz in detecting diurnal variations, an estimated bound on Lorentz breaking is given as \begin{equation} r^e_{\omega_{a}^{\mp},\rm diurnal} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10^{-20} \quad . \end{equation} \section{Proton-Antiproton Experiments} We also investigate the sensitivity to CPT and Lorentz violations of charge-to-mass-ratio experiments and possible future $g-2$ experiments involving protons and antiprotons in Penning traps. In this analysis, it suffices to work at the level of an effective theory in which the protons and antiprotons are regarded as basic objects described by a Dirac equation. The coefficients $a_\mu^p$, $b_\mu^p$, $H_{\mu \nu}^p$, $c_{\mu \nu}^p$, $d_{\mu \nu}^p$ represent effective parameters, which at a more fundamental level depend on underlying quark interactions. To leading order, we find the proton-antiproton differences for the cyclotron and anomaly frequencies are \begin{equation} \Delta \omega_c^p \equiv \omega_c^{p} - \omega_c^{\bar p} \simeq 0 \quad , \label{omcp} \end{equation} \begin{equation} \Delta \omega_a^p \equiv \omega_a^{p} - \omega_a^{\bar p} \simeq 4 b_3^p \quad . \label{omap} \end{equation} Assuming $\omega_a^p$ and $\omega_a^{\bar p}$ can be measured with ppb accuracies, and defining an appropriate figure of merit, we estimate for $g-2$ experiments \begin{equation} r^p_{\omega_a} \equiv \fr{|{ E}_{n,s}^{p} - { E}_{n,-s}^{\bar p}|} {{ E}_{n,s}^{p}} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10^{-24} \quad , \label{romap} \end{equation} whereas in experiments searching for diurnal variations we estimate \begin{equation} r^p_{\omega_{a}^{\mp},\rm diurnal} \approx \fr {2|\mp b_3^p + d_{30}^p m_p + H_{12}^p|}{m_p} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10^{-24} \quad , \label{pwadiurnal} \end{equation} \begin{equation} r^p_{\omega_c, \rm diurnal} \approx \fr {|c_{11}^p + c_{22}^p| \omega_c} {m_p} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10^{-24} \quad . \label{pwcdiurnal} \end{equation} A recent experiment\cite{gg2} compares antiproton cyclotron frequencies with those of an $H^-$ ion instead of a proton. This comparison provides a sharp test of CPT-preserving Lorentz symmetry. In the context of our frameowrk, the difference between the cyclotron frequencies of the $H^-$ hydrogen ion and the antiproton can be computed and is given by $$ \Delta \omega_{c,\rm th}^{H^-} \approx (c_{00}^{p} + c_{11}^{p} +c_{22}^{p}) ( \omega_c - \omega_c^{H^-} ) \quad\quad\quad\quad $$ \begin{equation} \quad\quad\quad\quad - \fr{2m_e}{m_p} (c_{00}^{e} + c_{11}^{e} + c_{22}^{e} -c_{00}^{p} - c_{11}^{p} -c_{22}^{p}) \omega_{c}^{H^-} \quad . \label{romacH} \end{equation} The estimated bound that follows from this is\cite{bkr2} \begin{equation} r^{H^-}_{\omega_c} \approx |\Delta \omega_{c,\rm th}^{H^-}| / {m_p} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10^{-25} \quad . \label{rH} \end{equation} \section{Conclusions} In summary, we find that the use of a general theoretical framework incorporating CPT and Lorentz violation permits a detailed investigation of possible experimental signatures in Penning-trap experiments. In the electron-positron system, our results indicate that the sharpest tests of CPT in Penning-trap experiments emerge from comparisons of anomaly frequencies in $g-2$ experiments and that bounds of order $10^{-21}$ are attainable. In the context of our theoretical framework, we find that the conventional figure of merit $r_g^e$ does not provides an appropriate bound on CPT, and we have suggested an alternative. We find that comparisons of cyclotron frequencies are not sensitive to leading-order CPT or Lorentz violation, whereas diurnal variations in $\omega_a$ and $\omega_c$ can provide new signals for Lorentz violation with bounds of order $10^{-21}$ and $10^{-18}$, respectively. Experiments searching for diurnal variations in electrons alone can provide a bound on Lorentz breaking at a level of approximately $10^{-20}$. In the proton-antiproton system, our results show that future $g-2$ experiments on protons and antiprotons could provide stringent test of CPT, with bounds of order $10^{-24}$. Experiments searching for diurnal variations in the proton-antiproton system can also provide bounds on Lorentz and CPT breaking at a level of approximately $10^{-24}$. A recent comparison of $H^-$ and antiproton cyclotron frequencies have provided a new test of CPT-preserving Lorentz invariance at a level of $10^{-25}$. Table I contains a summary of the estimated bounds attainable in Penning-trap experiments in the three systems considered here. \begin{table}[h] \caption{Estimated CPT- and Lorentz-violating bounds for electron-postron, proton-antiproton, and $H^-$-antiproton experiments in Penning traps. The estimated bounds are based on ppb accuracies in $\omega_a$ and $\omega_c$ (except in the $H^-$ $\bar p$ experiments which have $\sim 10^{-10}$ accuracy). The first two columns specify the type of experiment. The third column lists the figures of merit, while the fourth gives the corresponding bounds estimated from current or future experiments. The fifth column shows the affected parameters, and the last shows which symmetry is tested, CPT-violating Lorentz symmetry (CPT) or CPT-preserving Lorentz symmetry (Lorentz).\label{tab:exp}} \vspace{0.2cm} \begin{center} \begin{tabular}{|l|c|c|c|l|l|} \hline System & Expt.\ & Fig.\ Merit & Est.\ Bound & Parms.\ &Test \\ \hline\hline $e^- \, e^+$ & $\Delta \omega_a$ & $r^e_{\omega_a}$ & $10^{-21}$ & $b_j^e$ & CPT \\[2mm] \cline{2-6} & $\omega_a$ \, diurnal & $r^e_{\omega_a, \rm diurnal}$ & $10^{-21}$ & $d_{j0}^e,\ H_{jk}^e$ & Lorentz \\[2mm] \cline{2-6} & $\omega_c$ \, diurnal & $r^e_{\omega_c, \rm diurnal}$ & $10^{-18}$ & $c_{jj}^e$ & Lorentz \\[2mm] \cline{1-6} $p \, \bar p$ & $\Delta \omega_a$ & $r^p_{\omega_a}$ & $10^{-24}$ & $b_j^p$ & CPT \\[2mm] \cline{2-6} & $\omega_a$ \, diurnal & $r^e_{\omega_a, \rm diurnal}$ & $10^{-24}$ & $d_{j0}^p,\ H_{jk}^p$ & Lorentz \\[2mm] \cline{2-6} & $\omega_c$ \, diurnal & $r^e_{\omega_c, \rm diurnal}$ & $10^{-24}$ & $c_{jj}^p$ & Lorentz \\[2mm] \cline{1-6} $H^- \, \bar p$ & $\Delta \omega_c$ & $r^{H^-}_{\omega_c}$ & $10^{-25}$ & $c_{jj}^e$, $c_{jj}^p$ & Lorentz \\[2mm] \hline \end{tabular} \end{center} \end{table} \section*{Acknowledgments} This work is supported in part by the National Science Foundation under grant number PHY-9801869. \section*{References}
\section{Introduction} {\sc Discrepancies} are measures of non-uniformity of point sets that play an important r\^ole in the Quasi-Monte Carlo method of numerical integration \cite{nieder1}. A certain class of these discrepancies, the so called {\em quadratic discrepancies}, can be defined as an average-case complexity over a class of functions, the {\em problem class} \cite{woz,kleiss,hak1}. In a number of publications \cite{hk1,hk2,hk3} the problem of calculating the probability distribution of discrepancies for sets of $N$ truly random points has been tackled. One of the results was the calculation of the asymptotic distributions in the limit of infinite $N$. In \cite{hak1}, we introduced techniques from quantum field theory (QFT) to calculate the moment generating function $G$ of the probability distribution, suitable to calculate it as a series expansion in $1/N$. In this paper, we extend the formalism by the introduction of fermions, and give the explicit diagrammatic expansion of $\log G$ up to and including ${\mathcal O}(1/N^2)$. For the Lego discrepancy, the $L_2^*$-discrepancy in one dimension and the Fourier diaphony in one dimension, we give the explicit $1/N$ correction. \section{The formalism} We start with a short repetition of the formalism of \cite{hak1} and continue with the introduction of some new tools. \subsection{Quadratic discrepancies and path integrals} We shall always take the integration region to be the $s$-dimensional unit hypercube ${\bf K}=[0,1)^s$. The point set $X_N$ consists of $N$ points $x_k\in{\bf K}$, $k=1,2,\ldots,N$. Defined as an average-case complexity on the problem class of functions $\phi:{\bf K}\mapsto{\bf R}$ with measure $\mu$, a discrepancy $D_N$ of the point set $X_N$ is given by \begin{equation} D_N \;=\; N\int\eta^2_N[\phi]\,d\mu[\phi] \quad,\quad \eta_N[\phi] \;=\; \frac{1}{N}\sum_{k=1}^N\phi(x_k) - \int_{\bf K} \phi(x)\,dx \;\;. \label{CorEq001} \end{equation} It is the quadratic integration error, averaged over the problem class. The probability density $H$ of the discrepancy as a random variable of random point sets is calculated as the inverse Laplace transform of the moment generating function $G$: \begin{equation} H(D_N=t) \;=\; \frac{1}{2\pi i}\int\limits_{-i\infty}^{+i\infty}e^{-zt}G(z)\,dz \quad,\quad G(z) \;=\; \Exp{e^{zD_N}} \;\;. \label{CorEq026} \end{equation} Here, $\mathsf{E}$ denotes the expectation value of a random variable. The measure $\mu$ is assumed to be Gaussian and in \cite{hak1} it is shown that the generating function is therefore given by \begin{equation} G(z) \;=\; \int \left(\int_\Kube e^{g[\phi(x)-\int_\Kube \phi(y)\,dy]}\,dx\right)^N d\mu[\phi] \quad,\quad g \;=\; \sqrt{\frac{2z}{N}} \;\;. \label{CorEq002} \end{equation} In \cite{hak1} we suggested to put the $N^{\textrm{th}}$ power in the exponential and interpret $G(z)$ as an Euclidean path integral \begin{equation} G(z) \;=\; \int \mathcal{D}\phi\,\exp(-S[\phi]) \end{equation} with an action \begin{equation} S[\phi] \;=\; \frac{1}{2}\int_{\Kube^2} \phi(x)\varLambda(x,y)\phi(y)\,dxdy - N\log\left(\int_\Kube e^{g[\phi(x)-\int_\Kube \phi(y)\,dy]}\,dx\right) \;\;, \label{CorEq016} \end{equation} where $\varLambda$ is the symmetric linear operator with boundary conditions which is the inverse of the two-point Green function under the measure $\mu$: \begin{equation} \int_\Kube \varLambda(x_1,y)\mathcal{C}(y,x_2)\,dy \;=\; \delta(x_1-x_2) \label{CorEq003} \quad,\quad \mathcal{C}(y,x_2) \;=\; \int\phi(y)\phi(x_2)\,d\mu[\phi] \;\;. \end{equation} The ``infinitesimal volume element'' $\mathcal{D}\phi$ can be seen as defined by the rule that $d\mu[\phi]=\mathcal{D}\phi\,\exp(-S_0)$, where $S_0$ is the action with $z=0$. For notational convenience we put $\mathcal{D}\phi$ to the left of the exponential. One of the features of this formalism is that the action has a {\em gauge freedom}; a {\em global translation} $\varTheta_c:\phi(x)\mapsto\phi(x)+c$, $c\in{\bf R}$ leaves $\eta_N[\phi]$ and {\em a fortiori} $G(z)$ invariant, and results in a change of the action at most linear in $\phi$: \begin{equation} S[\varTheta_c\phi] \;=\; S[\phi]+\alpha c\chi[\phi]+{\textstyle\frac{1}{2}}\alpha c^2 \quad \textrm{with} \quad \chi[\phi_1+\phi_2] \;=\; \chi[\phi_1]+\chi[\phi_2] \;\;. \end{equation} As a result of this, the path integral can be generally written as \begin{equation} G(z) \;=\; \frac{1}{I[F]}\sqrt{\frac{2\pi}{\alpha}}\int\mathcal{D}\phi\, \exp(-F(\xi[\phi]) - S_\varTheta[\phi]) \;\;,\quad S_\varTheta[\phi]=S[\phi]-{\textstyle\frac{1}{2}}\alpha\chi[\phi]^2 \;\;, \end{equation} where $F$ is restricted such that \begin{equation} I[F] \;\equiv\; \int_{-\infty}^{\infty} \exp(-F(c))\,dc \end{equation} exists, and $\xi$ is only restricted such that $\xi[\varTheta_c\phi]=\xi[\phi]+c$. It is, for example, possible to take $\chi[\phi]=\int_{\bf K}\phi(x)\,dx$ and $F(\chi)=M\chi^2$ with $M\rightarrow\infty$. In this gauge, $\chi[\phi]$ will vanish from the action, which is reflected in a two-point function that integrates to zero with respect to each of its variables. We shall refer to this gauge as the {\em Landau} gauge. Clearly, the first equation of (\ref{CorEq003}) cannot be satisfied in the Landau gauge, because then $\mathcal{C}$ integrates to zero~\footnote{There is a misprint in Eq.~(73) of \cite{hak1} with respect to this, where $\delta(x-y)$ should be replaced by $\delta(x-y)-1$.}. To see what happens, we assume that the problem class is a vector space with a basis of at least square integrable functions $\{u_n\}$, so that a member $\phi$ of the problem class can be written as \begin{equation} \phi(x) \;=\; \sum_n \phi_nu_n(x) \;\;,\quad \phi_n\in{\bf R} \;\;, \end{equation} and that the Gaussian measure on this function space is defined by \begin{equation} d\mu[\phi] \;=\; \prod_{n}\frac{\exp(-\phi_n^2/2\sigma_n^2)}{\sqrt{2\pi\sigma_n^2}}\,d\phi_n \;\;,\quad \sigma_n\in{\bf R} \;\;. \label{CorEq004} \end{equation} For the measure to be suitably defined, the strengths $\sigma_n$ have to satisfy certain restrictions which can be translated into the requirement that $\Exp{D_N}$ exists. They are the inverse of the eigenvalues of $\mathcal{C}$ and the basis consists of the eigenfunctions. Therefore, $\mathcal{C}$ and $\varLambda$ can be expressed in terms of the basis: \begin{equation} \mathcal{C}(x_1,x_2) \;=\; \sum_n\sigma_n^2u_n(x_1)u_n(x_2) \;\;,\quad \varLambda(x_1,x_2) \;=\; \sum_n\frac{1}{\sigma_n^2}\,u_n(x_1)u_n(x_2) \;\;, \label{CorEq005} \end{equation} where the last equation holds if the basis is orthonormal. The boundary conditions satisfied by $\mathcal{C}$ and $\varLambda$ are those satisfied by the basis functions. With different gauges come different bases and strengths. We call a gauge in which the basis is orthonormal a {\em Feynman} gauge. If the Landau gauge is used, in which $\int_\Kube\phi(x)\,dx=0$, then the basis functions have to integrate to zero: \begin{equation} \int_\Kube u^{(\textrm{L})}_n(x)\,dx \;=\; 0 \quad \forall\,n \;\;, \end{equation} where the label $\textrm{L}$ indicates the Landau gauge. This means that the basis cannot be ``complete'' in the sense that $\sum_nu_n(x)u_n(y)=\delta(x-y)$, but that we have \begin{equation} \sum_n u^{(\textrm{L})}_n(x)u^{(\textrm{L})}_n(y) \;=\; \delta(x-y) - 1 \;\;. \end{equation} The zero mode is isolated and integrated out of the path integral. We want to stress that the gauge freedom is something that comes from the original measure $\mu$, and that the Landau gauge exists for every quadratic discrepancy. It is a result of the fact that an integration error is the same for integrands that differ only by a constant. From now on we will denote the two-point function in the Landau gauge by $\mathcal{B}$. It satisfies \begin{equation} \int_\Kube\varLambda^{(\textrm{L})}(x_1,y)\mathcal{B}(y,x_2)\,dy \;=\; \delta(x_1-x_2)-1\;\;, \end{equation} and the discrepancy can then be written as \begin{equation} D_N \;=\; \frac{1}{N}\sum_{k,l=1}^N\mathcal{B}(x_k,x_l) \;\;. \end{equation} \subsection{Fermions as tools to calculate the $\boldsymbol{1}\boldsymbol{/}\boldsymbol{N}$ corrections} In \cite{hak1} we suggested to make a straight forward expansion in $1/N$ of $\exp(-S)$ to calculate $G$ perturbatively. This way, however, the calculation of the perturbation series becomes very cumbersome, and the reason for this is the following. We want to use the fact that an expansion in $1/N$ corresponds to an expansion around $\phi=0$ of the part of the action that is non-quadratic in $\phi$. The subsequent terms in the expansions are therefore proportional to moments of a Gaussian measure, and can be calculated using diagrams (cf. \cite{Rivers}). These diagrams, the {\em Feynman diagrams}, consist of lines representing two-point functions and vertices representing convolutions of two-point functions. Because the action is non-local, i.e. it cannot be written as a single integral over a Lagrangian density because of the logarithm in \eqn{CorEq016}, the total path integral, thus the total sum of all diagrams, cannot be seen as the exponential of all {\em connected} diagrams, and it is this that makes the calculations difficult. In order to circumvent this obstacle, we introduce $2N$ Grassmann variables $\bar{\psi}_i$ and $\psi_i$, $i=1,2,\ldots,N$. They all anti-commute with each other and commute with complex numbers: \begin{align} \bar{\psi}_i\bar{\psi}_j + \bar{\psi}_j\bar{\psi}_i = 0 \quad&,\quad \bar{\psi}_i\psi_j + \psi_j\bar{\psi}_i = 0 \quad,\quad \psi_i\psi_j + \psi_j\psi_i = 0 \quad\quad i,j=1,2,\ldots,N \\ c\bar{\psi}_i - \bar{\psi}_i c = 0 \quad&,\quad c\psi_i - \psi_i c = 0 \quad\quad i=1,2,\ldots,N \;\;,\quad c\in{\bf C} \;\;. \end{align} Now we use the well known Gaussian integration rules for Grassmann variables to write the $N^{\textrm{th}}$ power in \eqn{CorEq002} as an exponent and get \begin{align} & G(z) \;=\; \int \mathcal{D}\phi\,D\bar{\psi} D\psi\, \exp(-S[\phi,\bar{\psi},\psi]) \;\;, \\ & S[\phi,\bar{\psi},\psi] \;=\; \frac{1}{2}\int_{\Kube^2} \phi(x)\varLambda(x,y)\phi(y)\,dxdy + \int_\Kube e^{g[\phi(x)-\int_\Kube \phi(y)\,dy]}\,dx\sum_{i=1}^N\bar{\psi}_i\psi_i \;\;, \label{CorEq006} \end{align} where we introduced the notational shorthand \begin{equation} D\bar{\psi} = d\bar{\psi}_1d\bar{\psi}_2\cdots d\bar{\psi}_N \quad,\quad D\psi = d\psi_1d\psi_2\cdots d\psi_N \;\;. \end{equation} Notice that this action contains the same gauge freedom, so that the action becomes completely local if the Landau gauge is used. We have to introduce the ``fermion fields'' to achieve this, but for calculating the perturbation expansion they are much easier to handle than the logarithmic potential. From the action (\ref{CorEq006}) we obtain the Feynman rules. To calculate a term in the $1/N$-expansion of $G$, the contribution of all diagrams that can be drawn using the Feynman rules and carry the right power of $1/N$ has to be calculated. Because in this paper we want to calculate the path integral itself, and no correlation functions, we only have to consider {\em vacuum} diagrams, i.e., diagrams without external legs. Furthermore, we will always use the Landau gauge, because then action is local, so that we only need to calculate {\em connected} diagrams. The whole generating function is the exponential of the sum of these diagrams. The contribution of the connected diagrams we denote by $W$, so that \begin{equation} G(z) \;=\; e^{W(z)} \;\;. \end{equation} The diagrammatic expansion is an expansion in $1/N$, the subsequent terms of which we denote by \begin{equation} W(z) \;=\; W_0(z) + \frac{1}{N}W_1(z) + \frac{1}{N^2}W_2(z) + \cdots \;\;. \end{equation} In the Landau gauge the rules can be summarized by~\footnote{In \cite{hak1} we called another two-point function the {\em propagator}, namely the inverse of $\varLambda(x,y)-2z\delta(x-y)+2z$, denoted by $\mathcal{G}_z(x,y)$. This one is obtained from $\mathcal{B}(x,y)$ by the re-summation of certain diagrams, as we will show in the following.} \begin{align} &\textrm{boson propagator:}\hspace{6pt}\quad x\,\diagram{dcP2}{50}{-1}\,y \;=\; \mathcal{B}(x,y) \;\;; \\ &\textrm{fermion propagator:}\quad i\,\diagram{dcP1}{50}{2}\,j \;=\; \delta_{i,j} \;\;;\\ &\textrm{vertices:}\hspace{60pt}\quad \diagram{dcV1}{65}{34} \;=\; -g^p\times\textrm{convolution} \;\;,\quad p\geq 2 \label{CorEq007}\;\;. \end{align} In the vertices, boson propagators are convoluted as $\int_\Kube\mathcal{B}(y,x_1)\mathcal{B}(y,x_2)\cdots\mathcal{B}(y,x_p)\,dy$, fermion propagators as $\sum_{j=1}^{\raisebox{-1pt}{${\scriptstyle N}$}}\delta_{i_1,j}\delta_{j,i_2}$, and then these convolutions are multiplied. As a result of this, the bosonic part of each diagram decouples completely from the fermionic part. The contribution of the fermionic part can easily be determined, for every fermion loop only gives a factor $-N$. The main problem is now to calculate the remaining bosonic part. Finally notice that, as a result of the Landau gauge, vertices with {\em only one} bosonic leg do not exist. \section{The diagrammatic expansion} We want to stress again that we only need to calculate the connected diagrams. The sum of the contributions of all these diagrams gives $W=\log G$. Usually, a Feynman diagram is a mnemonic representing a certain contribution to a term in a series expansion, i.e. a label. We will use the same drawing for the contribution itself, apart of the symmetry factor of the diagram. For example, the contribution of the diagram $\diagram{dcZ1}{30}{4.5}$ is equal to ${\textstyle\frac{1}{2}} Ng^2\int_\Kube\mathcal{B}(x,x)\,dx$ and its symmetry factor is equal to ${\textstyle\frac{1}{2}}$, so that we write \begin{equation} \frac{1}{2}\,\diagram{dcZ1}{40}{8} \;=\; \frac{Ng^2}{2}\int_\Kube\mathcal{B}(x,x)\,dx \;\;. \end{equation} \subsection{The zeroth order} The contribution to the zeroth order in $1/N$ can only come from diagrams in which the power of $1/N$ coming from the vertices cancels the power of $N$, coming from the fermion loops. This only happens in diagrams with vertices with two bosonic legs only, and in which the fermion lines begin and end on the same vertex. To write down their contribution, we introduce the two-point functions $\mathcal{B}_p$, $p=1,2,\ldots$, defined by \begin{equation} \mathcal{B}_1(x_1,x_2) = \mathcal{B}(x_1,x_2) \;\;,\quad \mathcal{B}_{p+1}(x_1,x_2) = \int_\Kube\mathcal{B}_p(x_1,y,)\mathcal{B}(y,x_2)\,dy \;\;. \end{equation} The zeroth order term is given by \begin{align} \frac{1}{2}\,\diagram{dcZ1}{40}{8} \;+\; \frac{1}{4}\,\diagram{dcZ2}{60}{8} \;+\; \frac{1}{6}\,\diagram{dcZ3}{50}{16} \;+\; \cdots \;=\; \sum_{p=1}^{\infty}\frac{(Ng^2)^p}{2p}\int_\Kube\mathcal{B}_p(x,x)\,dx \;\;. \end{align} The factor $1/2p$ is the symmetry factor of this type of diagram with $p$ fermion ``leaves''. If we substitute $g=\sqrt{2z/N}$ in this expression, we find exactly the result of Eq.~(21) in \cite{hkh1}. If we use the spectral representation of $\mathcal{B}$ and assume that the basis functions are orthonormal in the Landau gauge, we get \begin{equation} W_0(z) \;=\; -\frac{1}{2}\sum_n\log(1-2z\sigma_{\textrm{L},n}^2) \;\;, \end{equation} where $\textrm{L}$ indicates the Landau gauge. This expression is the same as Eq.~(68) in \cite{hk1}. \subsection{The first order} As we have seen before, bosonic two-point vertices with a closed single fermion line contribute with a factor $2z$, and without any dependence on $N$. Therefore, it is useful to introduce the following effective vertex \begin{equation} \diagram{dcV3}{65}{34} \;=\; \diagram{dcV2}{65}{34} \;=\; Ng^p\times\textrm{convolution}\;\;, \end{equation} and the following {\em dressed} boson propagator \begin{align} x\,\diagram{dcP6}{50}{-1}\,y \;&=\; x\,\diagram{dcP2}{50}{-1}\,y \;+\; x\,\diagram{dcP3}{50}{0}\,y \;+\; x\,\diagram{dcP4}{75}{0}\,y \notag\\ &\hspace{174pt} \;+\; x\,\diagram{dcP5}{100}{0}\,y \;+\; \cdots \notag\\ \;&=\; \sum_{p=1}^\infty(2z)^{p-1}\mathcal{B}_p(x,y) \;\;. \end{align} If we assume that the basis is orthonormal in the Landau gauge, we can write \begin{equation} x\,\diagram{dcP6}{50}{-1}\,y \;=\; \sum_n\frac{\sigma_{\textrm{L},n}^2}{1-2z\sigma_{\textrm{L},n}^2}\, u^{(\textrm{L})}_n(x)u^{(\textrm{L})}_n(y)\;\;, \end{equation} which is, apart of a factor $2z$, the same expression as in Eq.~(67) in \cite{hk1}. The dressed propagator is equal to the propagator in the Landau gauge as we defined it in Eq.~(24) in \cite{hak1}: \begin{equation} x\,\diagram{dcP6}{50}{-1}\,y \;=\; \mathcal{G}^{(\textrm{L})}_z(x,y) \;\;. \end{equation} This is easy to see, because it satisfies \begin{equation} \int_\Kube\left[\varLambda^{(\textrm{L})}(x_1,y)-2z\delta(x_1-y)+2z\right] \sum_{p=1}^\infty(2z)^{p-1}\mathcal{B}_p(x,y)\;dy \;=\; \delta(x_1-x_2) - 1 \;\;, \end{equation} just like $\mathcal{G}^{(\textrm{L})}_z$ by definition. Notice that $\mathcal{G}^{(\textrm{L})}_z$ and $\mathcal{B}$ satisfy the relation \begin{equation} \lim_{z\ra0}\mathcal{G}^{(\textrm{L})}_z(x,y) \;=\; \mathcal{G}^{(\textrm{L})}_{z=0}(x,y) \;=\; \mathcal{B}(x,y) \;\quad \forall\,x,y,\in{\bf K} \;\;. \end{equation} Furthermore, notice that $\mathcal{G}^{(\textrm{L})}_z$ and $W_0$ satisfy \begin{equation} \frac{\partial}{\partial z}\,W_0(z) \;=\; \int_\Kube\mathcal{G}^{(\textrm{L})}_z(x,x)\,dx \;\;, \label{CorEq008} \end{equation} and that this relation determines $W_0$ uniquely, because we know that $W_0(0)$ has to be equal to $0$ in order for the asymptotic probability distribution to be normalized to $1$. From now on, we will omit the label $\textrm{L}$. The first order term in the expansion of $W(z)$ is \begin{equation} \frac{1}{N}W_1(z) \;=\; \frac{1}{8}\,\diagram{dcA11}{40}{5} + \frac{1}{8}\,\diagram{dcA22}{60}{7} + \frac{1}{4}\,\diagram{dcA21}{30}{12} + \frac{1}{8}\,\diagram{dcA13}{58}{5} + \frac{1}{12}\,\diagram{dcA12}{30}{11} \;\;, \label{CorEq009} \end{equation} or, more explicitly, \begin{align} W_1(z) &\;=\; \frac{z^2}{2}\int_\Kube\mathcal{G}_z(x,x)^2\,dx \;-\; \frac{z^2}{2}\left(\int_\Kube\mathcal{G}_z(x,x)\,dx\right)^2 \;-\; z^2\int_{\Kube^2}\mathcal{G}_z(x,y)^2\,dxdy \notag\\ &\hspace{22pt}+ z^3\int_{\Kube^2}\mathcal{G}_z(x,x)\mathcal{G}_z(x,y)\mathcal{G}_z(y,y)\,dxdy \;+\; \frac{2z^3}{3}\int_{\Kube^2}\mathcal{G}_z(x,y)^3\,dxdy \;\;. \end{align} \subsection{The second order} The second order term in the expansion of $W(z)$ is denoted by $\sfrac{1}{N^2}W_2(z)$ and is given by \begin{align} & \frac{1}{48}\,\diagram{dcB11}{35}{11} \;+\;\frac{1}{48}\,\diagram{dcB12}{30}{11} \;+\;\frac{1}{16}\,\diagram{dcB13}{60}{6} \;+\;\frac{1}{12}\,\diagram{dcB14}{50}{9} \;+\;\frac{1}{24}\,\diagram{dcB111}{30}{12} \;+\;\frac{1}{16}\,\diagram{dcB15}{45}{19} \notag\\ &\;+\;\frac{1}{8}\,\diagram{dcB16}{75}{5} \;+\;\frac{1}{8}\,\diagram{dcB17}{50}{11} \;+\;\frac{1}{8}\,\diagram{dcB18}{30}{12} \;+\;\frac{1}{16}\,\diagram{dcB19}{60}{16} \;+\;\frac{1}{12}\,\diagram{dcB110}{65}{9} \notag \end{align} \begin{align} &\;+\;\frac{1}{48}\,\diagram{dcB112}{60}{20} \;+\;\frac{1}{8}\,\diagram{dcB113}{65}{11} \;+\;\frac{1}{16}\,\diagram{dcB114}{90}{5} \;+\;\frac{1}{16}\,\diagram{dcB115}{40}{13} \notag\\ &\;+\;\frac{1}{4}\,\diagram{dcB23}{50}{13} \;+\;\frac{1}{8}\,\diagram{dcB24}{30}{16} \;+\;\frac{1}{4}\,\diagram{dcB25}{50}{13} \;+\;\frac{1}{4}\,\diagram{dcB28}{30}{16} \;+\;\frac{1}{4}\,\diagram{dcB29}{30}{12}\notag\\ &\;+\;\frac{1}{2}\,\diagram{dcB210}{30}{12} \;+\;\frac{1}{16}\,\diagram{dcB36}{40}{15} \;+\;\frac{1}{8}\,\diagram{dcB37}{35}{18} \;+\;\frac{1}{4}\,\diagram{dcB38}{35}{15} \;+\;\frac{1}{12}\,\diagram{dcB211}{30}{12} \notag\\ &\;+\;\frac{1}{8}\,\diagram{dcB21}{80}{7} \;+\;\frac{1}{16}\,\diagram{dcB22}{55}{17} \;+\;\frac{1}{8}\,\diagram{dcB26}{65}{11} \;+\;\frac{1}{12}\,\diagram{dcB27}{70}{10} \notag\\ &\;+\;\frac{1}{8}\,\diagram{dcB212}{95}{7} \;+\;\frac{1}{16}\,\diagram{dcB213}{60}{27} \;+\;\frac{1}{8}\,\diagram{dcB214}{60}{14} \;+\;\frac{1}{4}\,\diagram{dcB215}{65}{11}\notag\\ &\;+\;\frac{1}{4}\,\diagram{dcB216}{65}{18} \;+\;\frac{1}{8}\,\diagram{dcB217}{50}{20} \;+\;\frac{1}{8}\,\diagram{dcB218}{85}{11} \;+\;\frac{1}{4}\,\diagram{dcB219}{65}{12}\notag \\ &\;+\;\frac{1}{4}\,\diagram{dcB220}{75}{12} \;+\;\frac{1}{8}\,\diagram{dcB221}{60}{12} \;+\;\frac{1}{4}\,\diagram{dcB222}{55}{18} \;+\;\frac{1}{3}\,\diagram{dcB31}{30}{13}\notag \\ &\;+\;\frac{1}{24}\,\diagram{dcB32}{54}{18} \;+\;\frac{1}{4}\,\diagram{dcB33}{50}{12} \;+\;\frac{1}{16}\,\diagram{dcB34}{90}{7} \;+\;\frac{1}{4}\,\diagram{dcB35}{75}{12}\notag \end{align} \subsection{One-vertex decomposability} For some discrepancies, the contribution of a bosonic part of a diagram that consists of two pieces connected by {\em only one} vertex, is equal to the product of the contribution of those pieces. Such diagrams we call {\em one-vertex reducible}, and discrepancies with this property we call {\em one-vertex decomposable}. Examples of such discrepancies are those for which $\mathcal{B}$ is translation invariant, i.e., $\mathcal{B}(x,y)=\mathcal{B}(x+a,y+a)$ $\forall\,x,y,a\in{\bf K}$, such as the Fourier diaphony. Also the Lego discrepancy with equal bins is one-vertex decomposable. In contrast, the $L_2^*$-discrepancy is not one-vertex decomposable. As a result of the one-vertex decomposability, many diagrams cancel or give zero. For example, the first and the second diagram in (\ref{CorEq009}) cancel, and the fourth gives zero, so that \begin{equation} \frac{1}{N}W_1(z) \;=\; \frac{1}{4}\,\diagram{dcA21}{30}{12} \;+\; \frac{1}{12}\,\diagram{dcA12}{30}{11} \;\;. \label{CorEq010} \end{equation} To second order, only the following remains: \begin{align} \frac{1}{N^2}W_2(z) \;=\; & \frac{1}{48}\,\diagram{dcB12}{30}{11} \;+\;\frac{1}{24}\,\diagram{dcB111}{30}{12} \;+\;\frac{1}{8}\,\diagram{dcB18}{30}{12} \;+\;\frac{1}{16}\,\diagram{dcB115}{40}{13} \notag\\ &\;+\;\frac{1}{8}\,\diagram{dcB24}{30}{16} \;+\;\frac{1}{4}\,\diagram{dcB28}{30}{16} \;+\;\frac{1}{4}\,\diagram{dcB29}{30}{12} \;+\;\frac{1}{2}\,\diagram{dcB210}{30}{12}\notag\\ &\;+\;\frac{1}{16}\,\diagram{dcB36}{40}{15} \;+\;\frac{1}{8}\,\diagram{dcB37}{35}{18} \;+\;\frac{1}{4}\,\diagram{dcB38}{35}{15} \;+\;\frac{1}{12}\,\diagram{dcB211}{30}{12} \;+\;\frac{1}{3}\,\diagram{dcB31}{30}{13} \;\;. \label{CorEq015} \end{align} We now derive a general rule of diagram cancellation. First, we extend the notion of one-vertex reducibility to complete diagrams, including the fermionic part, with the rule that the two pieces both must contain a bosonic part. Consider the following diagram \begin{equation} \diagram{lcLA}{65}{0} \;\;. \label{CorEq011} \end{equation} The only restriction we put one the ``leave'' $A$ is that it must be one-vertex irreducible with respect to the vertex that connects it to the fermion loop. For the rest, it may be anything. We define the contribution of the leave by the contribution of the whole diagram divided by $-N$, and denote it with $C(A)$. This contribution includes internal symmetry factors. Now consider a diagram consisting of a fermion loop as in diagram (\ref{CorEq011}) with attached to the one vertex $n_1$ leaves of type $A_1$, $n_2$ leaves of type $A_2$, and so on, up to $n_p$ leaves of type $A_p$. The extra symmetry factor of such a diagram is $(n_1!n_2!\cdots n_p!)^{-1}$, and, for one-vertex decomposable discrepancies, the contribution is equal to the product of the contributions of the leaves, so that the total contribution is given by \begin{equation} -N\prod_{q=1}^p\frac{C(A_q)^{n_q}}{n_q!} \;\;. \end{equation} Now we sum the contribution of all possible diagrams of this kind that can made with the $p$ leaves, and denote the result by \begin{align} \diagram{lcFl1}{30}{17} \;=\; -N\sum_{n_1,n_2,\ldots\geq1}\prod_{q=1}^p \frac{C(A_q)^{n_q}}{n_q!} \;=\; -N\left(e^{\sum_{q=1}^pC(A_q)} - 1\right) \;\;. \label{CorEq012} \end{align} Because the black square in l.h.s. of \eqn{CorEq012} represents all possible ways to put the leaves together onto one vertex, the sum of all possible ways to put the leaves onto one fermion loop is given by \begin{equation} \diagram{lcFl1}{30}{17} \;+\; \diagram{lcFl2}{36}{17} \;+\; \diagram{lcFl3}{35}{17} \;+\; \cdots \;=\; -N\sum_{n=1}^{\infty}\frac{(-1)^{n-1}}{n} \left(e^{\sum_{q=1}^pC(A_q)} - 1\right)^n \;\;. \label{CorEq013} \end{equation} The $(-1)^{n-1}$ in the sum comes from the vertices and $1/n$ is the extra symmetry factor of such diagram with $n$ vertices. The sum can be evaluated further and is equal to \begin{equation} -N\log e^{\sum_{q=1}^pC(A_q)} \;=\; -N\sum_{q=1}^pC(A_q) \;\;, \end{equation} i.e., the sum of all possible ways to put $p$ different leaves onto one fermion loop is equal to the sum of all leaves, each of them put onto its own fermion loop. This means that diagrams, consisting of two or more leaves put onto one fermion loop, cancel. Now consider the following equation, which holds for every one-vertex decomposable discrepancy: \begin{equation} \diagram{lcD2}{50}{15} \;=\; -\; \diagram{lcD1}{80}{15} \;\;, \label{CorEq014} \end{equation} where we only assume that $B$ is not of the type on the l.h.s. of \eqn{CorEq013}. The minus sign comes from the fact that the first diagram has one vertex less. Because the number of fermion lines, a fermion loop consists of is equal to the number of vertices it contains, we can always pair the diagrams into one diagram of the l.h.s. type and one of the r.h.s. type so that they cancel. We can summarize the result with the rule that {\em for one-vertex decomposable discrepancies, only the one-vertex irreducible diagrams contribute}. \section{Applications} We apply the general formulae given above to the Lego discrepancy, the $L_2^*$-discrepancy in one dimension and the Fourier diaphony in one dimension. \subsection{The Lego discrepancy} For the Lego problem class, for example defined in \cite{hak1}, the basis consists of a set of characteristic functions $\vartheta_n$, $n=1,2,\ldots,M$ of $M$ disjunct subspaces of ${\bf K}$. We will denote the measure $\int_\Kube\vartheta_n(x)\,dx$ of subspace $n$ by $w_n$ and we have $\sum_{n=1}^Mw_n=1$. The strengths $\sigma_n$ are equal to $1/\sqrt{w_n}$\,, and the propagator is given by \begin{equation} \mathcal{B}(x,y) \;=\; \sum_{n=1}^M\frac{\vartheta_n(x)\vartheta_n(y)}{w_n}-1 \;\;. \end{equation} With this choice of $\sigma_n$, the discrepancy is just the $\chi^2$ statistic that determines how well the points are distributed over the bins. It is easy to see that $\mathcal{B}_p(x,y)=\mathcal{B}(x,y)$, $p=2,3,\ldots$, so that the dressed propagator is given by \begin{equation} \mathcal{G}_z(x,y) \;=\; \frac{1}{1-2z}\,\mathcal{B}(x,y) \;\;. \end{equation} In \cite{hak1}, the propagator is given as an $M\times M$-matrix with matrix elements \begin{equation} \mathcal{G}^{(z)}_{n,m} \;=\; \frac{1}{1-2z}\left[\frac{\delta_{n,m}}{w_n}-1\right] \end{equation} and acting in the $M$-dimensional function space, rather than as a two-point function. This follows naturally from the path integral, which is an $M$-dimensional integral. The obvious and correct relation between the two is that \begin{equation} \mathcal{G}_z(x,y) \;=\; \sum_{n,m=1}^M\vartheta_n(x)\mathcal{G}^{(z)}_{n,m}\vartheta_m(y) \;\;. \end{equation} The zeroth order term can be found with the relation of \eqn{CorEq008}, which results in the following expression \begin{equation} W_0(z) \;=\; -\frac{1}{2}\log(1-2z)\int_\Kube\mathcal{B}(x,x)\,dx \;=\; -\frac{M-1}{2}\log(1-2z) \;\;, \label{CorEq017} \end{equation} in agreement with Eq.~(44) in \cite{hak1}. To write down the first order term, we introduce \begin{equation} M_2 = \sum_{n=1}^M\frac{1}{w_n} \;\;, \quad\textrm{and}\quad \eta(z) = \frac{2z}{1-2z} \;\;, \end{equation} so that \begin{equation} W_1(z) \;=\; \frac{1}{8}\left(M_2-M^2-2M+2\right)\eta(z)^2 +\frac{1}{24}\left(5M_2-3M^2-6M+4\right)\eta(z)^3 \;\;. \label{CorEq018} \end{equation} If the bins are equal, so that $w_n=1/M$ $n=1,2,\ldots,M$, then only the contribution of the diagrams of \eqn{CorEq011} remains, and the result is \begin{equation} W_1(z) \;=\; -\frac{1}{4}\,E\eta(z)^2 + \frac{1}{12}(E^2-E)\eta(z)^3 \;\;, \label{CorEq019} \end{equation} where we denote \begin{equation} E = M-1 \;\;. \end{equation} To second order in $1/N$, the contribution comes from the diagrams in \eqn{CorEq015}, and is given by \begin{align} W_2(z) \;=\; &(5E^3-12E^2+7E)\frac{\eta(z)^6}{48} \;+\; (E^3-6E^2+5E)\frac{\eta(z)^5}{8} \notag\\ \;&+\; (E^3-28E^2+43)\frac{\eta(z)^4}{48} \;+\; (-E^2-5E)\frac{\eta(z)^3}{12} \;\;. \label{CorEq020} \end{align} In Appendix A, we present the expansion of $G(z)$ in the case of equal bins, up to and including the $1/N^4$ term. It is calculated using the path integral expression (\ref{CorEq002}) of $G(z)$ and computer algebra. The reader may check that this expression for $G(z)$ and the above terms of $W(z)$ satisfy $G(z)=e^{W(z)}$ up to the order of $1/N^2$. \subsection{The $\boldsymbol{L}_{\boldsymbol{2}}^{\boldsymbol{*}}$-discrepancy} For the $L_2^*$-discrepancy in one dimension, for example defined in \cite{hak1}, the gauge freedom is a freedom in the boundary conditions which the members of the problem class have to satisfy. $\varLambda$ acts on the members as \begin{equation} (\varLambda\phi)(x) \;=\; -\frac{d^2\phi}{dx^2}(x) \;\;, \label{CorEq111} \end{equation} and in the Landau gauge, the boundary conditions are given by \begin{equation} \int_\Kube\phi(x)\,dx = 0 \;\;,\quad \frac{d\phi}{dx}(0)=\frac{d\phi}{dx}(1)=0\;\;. \end{equation} The basis in the Landau gauge is given by the set of eigenfunctions of $\varLambda$ with the boundary conditions above, which is $\{\sqrt{2}\cos(n\pi x),\; n=1,2,\ldots\}$, so that the propagator is given by \begin{align} \mathcal{B}(x,y) \;=\; \sum_{n=1}^\infty\frac{2\cos(n\pi x)\cos(n\pi y)}{n^2\pi^2} \;=\; \min(x,y) - x + {\textstyle\frac{1}{2}} x^2 - y + {\textstyle\frac{1}{2}} y^2 + \sfrac{1}{3}\;\;. \end{align} The dressed propagator is given by \begin{align} \mathcal{G}_z(x,y) \;&=\; \sum_{n=1}^\infty\frac{2\cos(n\pi x)\cos(n\pi y)}{n^2\pi^2-2z} \\ \;&=\; \frac{1}{u^2} - \frac{1}{2u\sin u} \{\cos[u(1-|x+y|)] \;+\; \cos[u(1-|x-y|)]\} \;\;, \end{align} with \begin{equation} u=\sqrt{2z} \;\;. \end{equation} The zeroth order term can be obtained using \eqn{CorEq008}: \begin{equation} W_0(z) \;=\; -\frac{1}{2}\log\left(\frac{\sin u}{u}\right) \;\;, \label{CorEq021} \end{equation} which is the well-known result. After some algebra, also the first order term follows: \begin{equation} W_1(z) \;=\; \frac{1}{288}\left(24-8\frac{u}{\sin u}-7\frac{u^2}{\sin^2u} -7\frac{u}{\tan u}-2\frac{u^2}{\tan^2u}\right) \;\;. \label{CorEq022} \end{equation} \subsection{The Fourier diaphony} Usually, the Fourier diaphony is defined in terms of a basis that is in the Landau gauge already. It is, for example, given in \cite{hkh1}, and in one dimension, the basis is given by the functions \begin{equation} u_{2n-1}(x)=\sqrt{2}\sin(2\pi nx) \;\;,\quad u_{2n}(x) =\sqrt{2}\cos(2\pi nx) \;\;,\quad n=1,2,\ldots \;\;, \end{equation} and for the strengths we take \begin{equation} \sigma_{2n-1} = \sigma_{2n} = \frac{1}{n} \;\;,\quad n=1,2,\ldots \;\;. \end{equation} Notice that the basis functions satisfy $-\frac{d^2}{dx^2}u_n(x)=\frac{4\pi^2}{\sigma_n^2}u_n(x)$, so that, from this point of view, the only relevant difference between the $L_2^*$-discrepancy and the Fourier diaphony are the boundary conditions on the members of the problem class. The propagator is given by \begin{align} \mathcal{B}(x,y) \;=\; \sum_{n=1}^\infty\frac{2\cos(2n\pi\{x-y\})}{n^2} \;=\; \frac{\pi^2}{3}\left[1-6\{x-y\}(1-\{x-y\})\right] \;\;, \end{align} where we use the notation $\{x\}=x\mod1$. The dressed propagator is given by \begin{equation} \mathcal{G}_z(x,y) \;=\; \sum_{n=1}^\infty\frac{2\cos(2n\pi\{x-y\})}{n^2-2z} \;=\; \frac{\pi^2}{v^2}\left(1-\frac{v\cos[v(2\{x-y\}-1)]}{2\sin v}\right) \;\;, \end{equation} where \begin{equation} v=\sqrt{2\pi^2 z} \;\;. \end{equation} This two-point function is, apart of a factor $\pi^2/v^2$, the same as the one in Eq.~(26) in \cite{hk2}. The zeroth order term can easily be obtained from the dressed propagator and is given by \begin{equation} W_0(z) \;=\; -\log\left(\frac{\sin v}{v}\right) \;\;, \label{CorEq023} \end{equation} which is in correspondence with Eq.~(21) in \cite{hk2}. Because the propagator is translation invariant, i.e., $\mathcal{B}(x+a,y+a)=\mathcal{B}(x,y)$ $\forall\,x,y,a\in{\bf K}$, the contributions of the first two diagrams in \eqn{CorEq009} cancel, and the contribution of the fourth diagram is zero. The contribution of the remaining diagrams gives \begin{equation} W_1(z) \;=\; \frac{1}{36}\left(3 + v^2 - 3\frac{v^2}{\sin^2v}\right) \;\;. \label{CorEq024} \end{equation} \section{Conclusions} In addition to the machinery of QFT introduced in \cite{hak1}, we introduced fermions to calculate the moment generating function $G$ of the probability distribution under sets of random points of a quadratic discrepancy $D_N$. They allow for an expansion in the inverse of the number of points $N$ of the logarithm $W$ of $G$, where the contribution to each term in the expansion can be represented by a finite number of connected Feynman diagrams. We have presented the diagrams up to the order of $1/N^2$ for the general case, and derived a rule of diagram cancellation in the case of special discrepancies, which we call one-vertex decomposable. We have applied the formalism to the Lego discrepancy, the $L_2^*$-discrepancy in one dimension and the Fourier diaphony in one dimension, and calculated the first two terms $W_0$ and $W_1/N$ in the expansion. For the Lego discrepancy, this resulted in \eqn{CorEq017} and \eqn{CorEq018}, for the $L_2^*$-discrepancy in \eqn{CorEq021} and \eqn{CorEq022}, and for the Fourier diaphony in \eqn{CorEq023} and \eqn{CorEq024}. The Fourier diaphony and the Lego discrepancy with equal binning are one-vertex decomposable. For the latter, we also calculated the term $W_2/N^2$, which is in correspondence with the result of an alternative calculation up to the order of $1/N^4$, given in Appendix A. Calculations become very cumbersome for high orders because of the number of diagrams involved. A situation in which the formalism can still be powerful is when another parameter in the definition of the discrepancy, such as the dimension of the integration region or the number of bins in case of the Lego discrepancy, becomes large. This parameter can then serve as an extra order parameter in the determination of the importance of the contribution of the diagrams, and can lead to a substantial reduction in the number of relevant diagrams. In \cite{hak4}, we will present our results with respect to this for the Lego discrepancy. \section*{Appendix A} If we define, for the Lego-discrepancy with equal bins, $E=M-1$, $\eta(z)=2z/(1-2z)$ and \begin{equation} (1-2z)^{E/2}G(z) \;=\; \sum_{n,p\geq0}\frac{\eta(z)^p}{N^n}\,C^{(p)}_n(E) \;\;, \end{equation} then the only non-zero $C^{(p)}_n(E)$ up to $n=4$ are given by \begin{align} C_1^{(2)}(E) &= -\frac{1}{4}E\notag\\ C_1^{(3)}(E) &= E\bigg(\frac{1}{12}E-\frac{1}{12}\bigg)\notag\\ C_2^{(3)}(E) &= E\bigg(-\frac{1}{12}E+\frac{5}{12}\bigg)\notag\\ C_2^{(4)}(E) &= E\bigg(\frac{1}{48}E^2-\frac{53}{96}E +\frac{43}{48}\bigg)\notag\\ C_2^{(5)}(E) &= E\bigg(\frac{5}{48}E^2-\frac{35}{48}E +\frac{5}{8}\bigg)\notag\\ C_2^{(6)}(E) &= E\bigg(\frac{1}{288}E^3+\frac{7}{72}E^2 -\frac{71}{288}E+\frac{7}{48}\bigg)\notag\\ C_3^{(4)}(E) &= E\bigg(-\frac{1}{48}E^2+\frac{7}{12}E -\frac{61}{48}\bigg)\notag \end{align} \begin{align} C_3^{(5)}(E) &= E\bigg(\frac{1}{240}E^3-\frac{17}{30}E^2 +\frac{583}{120}E-\frac{1451}{240}\bigg)\notag\\ C_3^{(6)}(E) &= E\bigg(\frac{53}{576}E^3-\frac{1153}{384}E^2 +\frac{7423}{576}E-\frac{527}{48}\bigg)\notag\\ C_3^{(7)}(E) &= E\bigg(\frac{1}{576}E^4+\frac{461}{1152}E^3 -\frac{6581}{1152}E^2+\frac{8663}{576}E -\frac{467}{48}\bigg)\notag\\ C_3^{(8)}(E) &= E\bigg(\frac{11}{1152}E^4+\frac{85}{144}E^3 -\frac{5125}{1152}E^2+\frac{1555}{192}E -\frac{17}{4}\bigg)\notag\\ C_3^{(9)}(E) &= E\bigg(\frac{1}{10368}E^5+\frac{29}{3456}E^4 +\frac{955}{3456}E^3-\frac{12475}{10368}E^2 +\frac{953}{576}E-\frac{53}{72}\bigg)\notag\\ C_4^{(5)}(E) &= E\bigg(-\frac{1}{240}E^3+\frac{37}{80}E^2 -\frac{337}{80}E+\frac{1397}{240}\bigg)\notag\\ C_4^{(6)}(E) &= E\bigg(\frac{1}{1440}E^4-\frac{349}{960}E^3 +\frac{7193}{720}E^2-\frac{15283}{320}E +\frac{67021}{1440}\bigg)\notag\\ C_4^{(7)}(E) &= E\bigg(\frac{49}{960}E^4-\frac{29069}{5760}E^3 +\frac{372169}{5760}E^2-\frac{571727}{2880}E +\frac{21503}{144}\bigg)\notag\\ C_4^{(8)}(E) &= E\bigg(\frac{13}{23040}E^5+\frac{13979}{23040}E^4 -\frac{2290601}{92160}E^3+\frac{1446743}{7680}E^2\notag\\ &\phantom{= E\bigg(\frac{13}{23040}E^5} -\frac{9583187}{23040}E+\frac{294773}{1152}\bigg)\notag\\ C_4^{(9)}(E) &= E\bigg(\frac{73}{6912}E^5+\frac{35077}{13824}E^4 -\frac{781079}{13824}E^3+\frac{993515}{3456}E^2 -\frac{564301}{1152}E+\frac{24607}{96}\bigg)\notag\\ C_4^{(10)}(E) &= E\bigg(\frac{1}{13824}E^6+\frac{139}{3072}E^5 +\frac{162721}{34560}E^4-\frac{596467}{9216}E^3\notag\\ &\phantom{= E\bigg(\frac{1}{13824}E^6} +\frac{1653251}{6912}E^2-\frac{253799}{768}E +\frac{145199}{960}\bigg)\notag\\ C_4^{(11)}(E) &= E\bigg(\frac{17}{41472}E^6+\frac{895}{13824}E^5 +\frac{55025}{13824}E^4-\frac{1505645}{41472}E^3\notag\\ &\phantom{= E\bigg(\frac{17}{41472}E^6} +\frac{19783}{192}E^2-\frac{137875}{1152}E +\frac{1565}{32}\bigg)\notag\\ C_4^{(12)}(E) &= E\bigg(\frac{1}{497664}E^7+\frac{11}{31104}E^6 +\frac{2431}{82944}E^5+\frac{155735}{124416}E^4\notag\\ &\phantom{= E\bigg(\frac{1}{497664}E^7} -\frac{3942431}{497664}E^3+\frac{249239}{13824}E^2 -\frac{250141}{13824}E+\frac{2575}{384}\bigg)\notag \end{align}
\subsection*{Full title:} Recent Trends in Algebraic Geometry -- EuroConference on Algebraic Geo\-metry (Warwick, July 1996), Editors: Klaus Hulek (chief editor), Fabrizio Catanese, Chris Peters and Miles Reid, CUP, May 1999 \subsection*{Contents:} Victor V. Batyrev: Birational Calabi--Yau $n$-folds have equal Betti numbers, 1--11 (alg-geom/9710020) {\parindent0pt Arnaud Beauville: A Calabi--Yau threefold with non-Abelian fundamental group, 13--17 (alg-geom/9502003) K. Behrend: Algebraic Gromov--Witten invariants, 19--70 (compare alg-geom/9601011) Philippe Eyssidieux: K\"ahler hyperbolicity and variations of Hodge structures, 71--92 Carel Faber: Algorithms for computing intersection numbers on moduli spaces of curves, with an application to the class of the locus of Jacobians, 93--109 (alg-geom/9706006) Marat Gizatullin: On some tensor representations of the Cremona group of the projective plane, 111--150 Y. Ito and I. Nakamura: Hilbert schemes and simple singularities, 151--233 Oliver K\"uchle and Andreas Steffens: Bounds for Seshadri constants, 235--254 (alg-geom/9601018) Marco Manetti: Degenerate double covers of the projective plane, 255--181 (see further math.AG/9802088) David R. Morrison: The geometry underlying mirror symmetry, 283--310 (alg-geom/9608006) Shigeru Mukai: Duality of polarized K3 surfaces, 311--326 Roberto Paoletti: On symplectic invariants of algebraic varieties coming from crepant contractions, 327--346 Kapil H. Paranjape: The Bogomolov--Pantev resolution, an expository account, 347--358 (math.AG/9806084) Tetsuji Shioda: Mordell--Weil lattices for higher genus fibration over a curve, 359--373 Bernd Siebert: Symplectic Gromov--Witten invariants, 375--424 Claire Voisin: A generic Torelli theorem for the quintic threefold, 425--463 P.M.H. Wilson: Flops, Type~III contractions and Gromov--Witten invariants on Calabi--Yau threefolds, 465--483 } \section*{Foreword} \markboth{\qquad Foreword\hfill}{\hfill Foreword\qquad} The volume contains a selection of seventeen survey and research articles from the July 1996 Warwick European algebraic geometry conference. These papers give a lively picture of current research trends in algebraic geo\-metry, and between them cover many of the outstanding hot topics in the modern subject. Several of the papers are expository accounts of substantial new areas of advance in mathematics, carefully written to be accessible to the general reader. The book will be of interest to a wide range of students and nonexperts in different areas of mathematics, geometry and physics, and is required reading for all specialists in algebraic geometry. The European algebraic geometry conference was one of the climactic events of the 1995--96 \hbox{EPSRC} Warwick algebraic geometry symposium, and turned out to be one of the major algebraic geometry events of the 1990s. The scientific committee consisted of A.~Beauville (Paris), F.~Catanese (Pisa), K.~Hulek (Hannover) and C.~Peters (Grenoble) representing AGE (\hbox{Algebraic} Geometry in Europe, an EU HCM--TMR network) and N.J.~Hitchin (Oxford), J.D.S.~Jones and M.~Reid (Warwick) representing Warwick and British mathematics. The conference attracted 178 participants from 22 countries and featured 33 lectures from a star-studded cast of speakers, including most of the authors represented in this volume. \paragraph{The expository papers} Five of the articles are expository in intention: among these a beautiful short exposition by Paranjape of the new and very simple approach to the resolution of singularities; a detailed essay by Ito and Nakamura on the ubiquitous ADE classification, centred around simple surface singularities; a discussion by Morrison of the new special Lagrangian approach giving geometric foundations to mirror symmetry; and two deep and informative survey articles by Behrend and Siebert on Gromov--Witten invariants, treated from the contrasting viewpoints of algebraic and symplectic geometry. \paragraph{Some main overall topics} Many of the papers in this volume group around a small number of main research topics. Gromov--Witten invariants was one of the main new breakthroughs in geometry in the 1990s; they can be developed from several different starting points in symplectic or algebraic geometry. The survey of Siebert covers the analytic background to the symplectic point of view, and outlines the proof that the two approaches define the same invariants. Behrend's paper explains the approach in algebraic geo\-metry to the invariants via moduli stacks and the virtual fundamental class, which essentially amounts to a very sophisticated way of doing intersection theory calculations. The papers by Paoletti and Wilson give parallel applications of Gromov--Witten invariants to higher dimensional varieties: Wilson's paper determines the Gromov--Witten invariants that arise from extremal rays of the Mori cone of Calabi--Yau 3-folds, whereas Paoletti proves that Mori extremal rays have nonzero associated Gromov--Witten invariants in many higher dimensional cases. The upshot is that extremal rays arising in algebraic geometry are in fact in many cases invariant in the wider symplectic and topological setting. Another area of recent spectacular progress in geometry and theoretical physics is Calabi--Yau 3-folds and mirror symmetry. This was another major theme of the EuroConference that is well represented in this volume. The paper by Voisin, which is an extraordinary computational tour-de-force, proves the generic Torelli theorem for the most classical of all Calabi--Yau 3-folds, the quintic hypersurface in $\mathbb P^4$. The survey by Morrison explains, among other things, the Strominger--Yau--Zaslow special Lagrangian interpretation of mirror symmetry. Beauville's paper gives the first known construction of a Calabi--Yau 3-fold having the quaternion group of order 8 as its fundamental group. The paper by Batyrev proves that the Betti numbers of a Calabi--Yau 3-fold are birationally invariant, using the methods of $p$-adic integration and the Weil conjectures; the idea of the paper is quite startling at first sight (and not much less so at second sight), but it is an early precursor of Kontsevich's idea of motivic integration, as worked out in papers of Denef and Loeser. Several other papers in this volume (those of Ito and Nakamura, Mukai, \hbox{Shioda} and Wilson) are implicitly or explicitly related to Calabi--Yau 3-folds in one way or another. \paragraph{Other topics} The remaining papers, while not necessarily strictly related in subject matter, include some remarkable achievements that illustrate the breadth and depth of current research in algebraic geometry. Shioda extends his well-known results on the Mordell--Weil lattices of elliptic surfaces to higher genus fibrations, in a paper that will undoubtedly have substantial repercussions in areas as diverse as number theory, classification of surfaces, lattice theory and singularity theory. Faber continues his study of tauto\-logical classes on the moduli space of curves and Abelian varieties, and gives an algorithmic treatment of their intersection numbers, that parallels in many respects the Schubert calculus; he obtains the best currently known partial results determining the class of the Schottky locus. Gizatullin initiates a fascinating study of representations of the Cremona group of the plane by birational transformations of spaces of plane curves. Eyssidieux gives a study, in terms of Gromov's K\"ahler hyperbolicity, of universal inequalities holding between the Chern classes of vector bundles over Hermitian symmetric spaces of noncompact type admitting a variation of Hodge structures. K\"uchle and Steffens' paper contains new twists on the idea of Seshadri constants, a notion of local ampleness arising in recent attempts on the Fujita conjecture; they use in particular an ingenious scaling trick to provide improved criteria for the very ampleness of adjoint line bundles. Manetti's paper continues his long-term study of surfaces of general type constructed as iterated double covers of $\mathbb P^2$. He obtains many constructions of families of surfaces, and proves that these give complete connected components of their moduli spaces, provided that certain naturally occuring degenerations of the double covers are included. This idea is used here to establish a bigger-than-polynomial lower bound on the growth of the number of connected components of moduli spaces. In more recent work, he has extended these ideas in a spectacular way to exhibit the first examples of algebraic surfaces that are proved to be diffeomorphic but not deformation equivalent. The Fourier--Mukai transform is now firmly established as one of the most important new devices in algebraic geometry. The idea, roughly speaking, is that a sufficiently good moduli family of vector bundles (say) on a variety $X$ induces a correspondence between $X$ and the moduli space $\widehat X$. In favourable cases, this correspondence gives an equivalence of categories between coherent sheaves on $X$ and on $\widehat X$ (more precisely, between their derived categories). The model for this theory is provided by the case originally treated by Mukai, when $X$ is an Abelian variety and $\widehat X$ its dual; Mukai named the transform by analogy with the classical Fourier transform, which takes functions on a real vector space to functions on its dual. It is believed that, in addition to its many fruitful applications in algebraic geometry proper, this correspondence and its generalisations to other categories of geometry will eventually provide the language for mathematical interpretations of the various ``dualities'' invented by the physicists, for example, between special Lagrangian geometry on a Calabi--Yau 3-fold and coherent algebraic geometry on its mirror partner (which, as described in Morrison's article, is conjecturally a fine moduli space for special Lagrangian tori). Mukai's magic paper in this volume presents a Fourier--Mukai transform for K3 surfaces, in terms of moduli of semi-rigid sheaves; under some minor numerical assumptions, he establishes the existence of a dual K3 surface, the fact that the Fourier--Mukai transform is an equivalence of derived categories, and the biduality result in appropriate cases. The paper of Ito and Nakamura is the longest in the volume; it combines a detailed and wide-ranging expository essay on the ADE classification with an algebraic treatment of the McKay correspondence for the Kleinian quotient singularities $\mathbb C^2/G$ in terms of the $G$-orbit Hilbert scheme. The contents of their expository section will probably come as a surprise to algebraic geo\-meters, since alongside traditional aspects of simple singularities and their ADE homologues in algebraic groups and representation of quivers, they lay particular emphasis on partition functions in conformal field theory with modular invariance under $\operatorname{SL}(2,\mathbb Z)$ and on II$_1$ factors in von Neumann algebras. Their study of the $G$-Hilbert scheme makes explicit for the first time many aspects of the McKay correspondence relating the exceptional locus of the Kleinian quotient singularities $\mathbb C^2/G$ with the irreducible representations of $G$; for example, the way in which the points of the minimal resolution can be viewed as defined by polynomial equations in the character spaces of the corresponding irreducible representations, or the significance in algebraic terms of tensoring with the given representation of $G$. Ito and Nakamura and their coworkers are currently involved in generalising many aspects of the $G$-orbit Hilbert scheme approach to the resolution of Gorenstein quotient singularities and the McKay correspondence to finite subgroups of $\operatorname{SL}(3,\mathbb C)$, and this paper serves as a model for what one hopes to achieve. \paragraph{Thanks to all our sponsors} The principal financial support for the Euro\-Conference was a grant of ECU40,000 from EU TMR (Transfer and Mobility of Researchers), contract number ERBFMMACT 950029; we are very grateful for this support, without which the conference could not have taken place. The main funding for the 1995--96 Warwick algebraic geometry symposium was provided by British EPSRC (Engineering and physical sciences research council). Naturally enough, the symposium was one of the principal activities of the Warwick group of AGE (European Union HCM project Algebraic Geometry in Europe, Contract number ERBCHRXCT 940557), and financial support from Warwick AGE and the other groups of AGE was a crucial element in the success of the symposium and the EuroConference. We also benefitted from two visiting fellowships for Nakamura and Klyachko from the Royal Society (the UK Academy of Science). Many other participants were covered by their own research grants. The University of Warwick, and the Warwick Mathematics Institute also provided substantial financial backing. All aspects of the conference were enhanced by the expert logistic and organisational help provided by the \hbox{Warwick} Math Research Centre's incomparable staff, Elaine Greaves Coelho, Peta McAllister and Hazel Graley. \bigskip \centerline{Klaus Hulek and Miles Reid, November 1998} \end{document}
\section{introduction} Compact quantum groups were introduced by Woronowicz \cite{woron1} and studied by several other authors \cite{andr1,ed-ru1,daele1}, as non-commutative analogues of the function algebras on compact groups. The aim of this work is to construct some quantum analogues of the compact group $C^*$-algebras, which turn out to be non-cocommutative Hopf $C^*$-algebras with approximate unit. More precisely, starting from a compact Hopf $*$-algebra, we construct a Hopf $C^*$-algebra with approximate unit in such a way that when the original compact Hopf $*$-algebra is the function algebra on certain compact group $G$, the resulting Hopf $C^*$-algebra reduces to the group $C^*$-algebra $C^*(G)$. Recall that for a compact group $G$, the group $C^*$-algebra $C^*(G)$ is the completion by operator norm of the algebra $L^1(G)$ of absolutely integrable complex-valued functions on $G$. The product on $L^1(G)$ is the convolution-product, given by $$(g*f)(x)=\displaystyle\testint_{G}f(y)g(y^{-1}x)dy.$$ With respect to this product, $C^*(G)$ is a non-unital $C^*$-algebra. In that case, there is an important notion of $\delta$-type sequences in $C^*(G)$, that approximate the unity. Such a sequence is called an approximate unit in $C^*(G)$. Moreover, the action of the group $G$ on any representation can be recovered from the corresponding action of the algebra $C^*(G)$, by using $\delta$-type sequences. In our case, however, we do not know what the quantum group defining our $C^*$-algebra is, hence our defining a Hopf algebra structure to exhibit the ``group property'' of our algebra. Our construction is done in several steps. First we construct a convolution product on an arbitrary Hopf algebras with integrals, being motivated by the classical convolution product for $L_1(G)$, the completion of which (under the operator norm) defines $C^*(G)$ (see \ref{chH}). Denoting by \mbox{$\check{H}$}\ the vector space $H$ with the new convolution product, we show that \mbox{$\check{H}$}\ is a non-unital algebra that is an ideal of $H^*$ (the dual of $H$); that the category of completely reducible $H$-comodules is equivalent to the category of completely reducible \mbox{$\check{H}$}-comodules (\ref{thmact-coact}, \ref{chh-hsq}) and that when $H$ is co-semi-simple then \mbox{$\check{H}$}\ is isomorphic to a direct sum of full endomorphism rings of simple $H$-comodules \rref{thmchH}. In section \ref{sec3}, we now focus on the case when $H$ is a compact Hopf $*$-algebra. Through the works of Woronowicz, Koorwinder and Djikhuizen, \cite{woron1,dk}, compact Hopf $*$-algebra are known to play a role analogous to that of the algebra of functions on compact groups in the classical theory. Since by definition, $H$ has a complex scalar product, we complete it to a Hilbert space \mbox{$H_{L^2}$}, which is the quantum analogue of the algebra $L^2(G)$. We show in \ref{extend-pr} that the convolution product on \mbox{$\check{H}$}\ extends to \mbox{$H_{L^2}$}\ and that \mbox{$H_{L^2}$}\ possesses a (topological) coproduct and an approximate antipode and hence is a Hopf algebra with approximate unit (\ref{thm-hlh}). Since the antipode on $H$ is not involutive, the $*$-structre defined on $H$ by $h^\star:=S(h^*)$ cannot be extended to \mbox{$H_{L^2}$}\ because it is not continous with respect to the given norm. To overcome this problem, we pass to the $C^*$-enveloping of \mbox{$H_{L^2}$}\ to obtain \mbox{$H_{C^*}$}\ that is the required Hopf $C^*$-algebra with approximate unit (see 4.5). We show that there is a one-one correspondence between the irreducible unitary representations of \mbox{$H_{C^*}$}\ and those of \mbox{$\check{H}$}\ (see 4.4). \section{The Convolution Product and Co-semi-simplicity}\label{sec1} We work over an algebraically closed field ${\Bbb K}$ of characteristic zero. Let $(H,m,\eta)$ be an algebra over a field ${\Bbb K}$, where $m$ denotes the product, $\eta$ denotes the map ${\Bbb K}\longrightarrow H$, $1_{{\Bbb K}}\longrightarrow 1_H$. A bialgebra structure on $H$ is a pair of linear maps $\Delta:H\longrightarrow H\otimes H,\ \varepsilon:H\longrightarrow {\Bbb K}$, satisfying \begin{itemize}\item $\Delta$ and $\varepsilon$ are homomorphism of algebras. \item $(\varepsilon\otimes {\mathchoice{\mbox{\rm id}_H)\Delta= ({\mathchoice{\mbox{\rm id}_H\otimes\varepsilon)\Delta={\mathchoice{\mbox{\rm id}_H.$ \item $(\Delta\otimes{\mathchoice{\mbox{\rm id}_H)\Delta=({\mathchoice{\mbox{\rm id}_H\otimes \Delta)\Delta.$ \end{itemize} An antipode on $H$ is a linear map $H\longrightarrow H$, satisfying \begin{itemize} \item $m\circ(S\otimes{\mathchoice{\mbox{\rm id}_H)\Delta=m\circ({\mathchoice{\mbox{\rm id}_H\otimes S)\Delta=\eta\varepsilon.$ \end{itemize} A bialgebra equipped with an antipode is called Hopf algebra. The antipode is then uniquely determined. Let $(H,m,\eta,\Delta,\varepsilon)$ be a Hopf algebra. A right coaction of $H$ on a vector space $V$ is a linear map $\delta:V\longrightarrow V\otimes H$, satisfying \begin{equation}\label{coaction} \begin{array}{c}({\mathchoice{\mbox{\rm id}_V\otimes\varepsilon)\delta={\mathchoice{\mbox{\rm id}_V,\\ (\delta\otimes{\mathchoice{\mbox{\rm id}_H)\delta=({\mathchoice{\mbox{\rm id}_V\otimes \Delta)\delta,\end{array}\end{equation} in the first identity we identify $V$ with $V \otimes \mathbb{C}$. $V$ is then called a right $H$-co-module. We shall frequently use Sweedler's notation, in particular $\Delta(x):=\sum_{(x)}x_{(1)}\otimes x_{(2)}=x_1\otimes x_2, \delta(v)=\sum_{(v)}v_{(0)}\otimes v_{(1)}=v_0\otimes v_1.$ The elements $v_1$'s in the presentation $\delta(v)=v_0\otimes v_1$ are called coefficients of the coaction $\delta$, and the space they span is called coefficient space. This space is a subcoalgebra of $H$. To see this, fix a basis $\daysub[d]{x}$. Then, the coaction is given by $\delta(x_i)=x_j\otimes a_i^j$ and the $\{a_i^j\}$ span the coefficient space. On the other hand, from \rref{coaction} it follows $\Delta(a_i^j)=a^j_k\otimes a^k_i$, $\varepsilon(a_i^k)=\delta^k_i$. The comodule $V$ is simple iff $\{a_i^j\}$ is a basis for the coefficient space, see \cite{schmuedgen1}. A left (right) integral on a Hopf algebra is a (non-trivial) linear functional $\displaystyle\testint:H\longrightarrow {\Bbb K}$, which is a left (right) $H$-comodule homomorphism, where $H$ is a left (right) $H$-comodule by means of the coproduct $\Delta$ and ${\Bbb K}$ is a left (right) $H$-comodule by means of the unity map $\eta$. Explicitly, a left (resp. right) integral $\displaystyle\testint_l$ $\left(\mbox{resp.,} \displaystyle\testint_r\right)$ on $H$ satisfies \begin{eqnarray}\label{axiom-int}\displaystyle\testint_l(x)=x_1\displaystyle\testint_l(x_2)\quad \left(\mbox{resp.} \displaystyle\testint_r(x)=\displaystyle\testint_r(x_1)\cdot x_2\right).\end{eqnarray} It was shown by Sullivan \cite{sul} that the integral on a Hopf algebra, if it exists, is defined uniquely up to a constant. Further, we have \begin{lem}\label{lemnon-degenerate}\cite{sweedler,lin1,stefan1} Let $\displaystyle\testint $ be a left integral on $H$. Then the bilinear form $b(g,h)=\displaystyle\testint(gS(h))$ is non-degenerate on $H$, that is \begin{eqnarray}\label{non-degenerate}\begin{array}{l} \displaystyle\testint(gS(h))=0, \forall h\quad \Longrightarrow\quad g=0.\\ \displaystyle\testint(gS(h))=0, \forall g\quad \Longrightarrow\quad h=0.\end{array}\end{eqnarray}\end{lem} In the rest of this work, we fix a left integral $\displaystyle\testint$ on $H$. We mention an identity, due originally to Sweedler \cite{sul}, which plays a crucial role in our computations: \begin{equation}\label{sweedler-id} \displaystyle\testint(gS(h_1))\cdot h_2=g_1\displaystyle\testint(g_2S(h)),\end{equation} if $\Delta(g) = g_1 \otimes g_2$ and $\Delta(h) = h_1 \otimes h_2$ in the notations of Sweedler. The coalgebra structure on $H$ induces an algebra structure on its dual $H^*$ -- the space of linear forms on $H$. The product of $\phi,\psi$ in $H^*$ is given by $$\phi*\psi(h)=\phi(h_1)\psi(h_2), \forall h\in H.$$ The unit element in $H^*$ is the counit $\varepsilon$ of $H$. \subsection*{Convolution product on $H$ We first recall the classical structure. Let $G$ be a compact group. Then there exists a unique normalized Haar measure on $G$ which induces the Haar integral on $L^1(G)$. The $*$-product is defined on $L^1(G)$ as follows: \begin{equation}\label{conv-pr} (g*f)(x):=\displaystyle\testint_{G}f(y)g(y^{-1}x)dy.\end{equation} Now, let $H$ be a Hopf algebra with an integral. Being motivated by \rref{conv-pr} we define the convolution product on $H$ by: \begin{equation}\label{conv-pr1} g*f:=\displaystyle\testint (fS(g_1))\cdot g_2.\end{equation} According to \rref{sweedler-id}, we also have $g*f=f_1\displaystyle\testint(f_2S(g)).$ \begin{lem}\label{chH} $H$ equipped with the $*$-product defined above is a (non-unital) algebra.\end{lem} \proof We only have to check the associativity: \begin{eqnarray*} (h*g)*f&=& \displaystyle\testint\left(fS((h*g)_1)\right)\cdot (h*g)_2\\ &=& \displaystyle\testint\left(fS\left(\displaystyle\testint(gS(h_1)\cdot h_2)\right)\right)\cdot h_2\\ &=& \displaystyle\testint\left(fS(h_2)\cdot \displaystyle\testint(gS(h_1)\right)\cdot h_3\\ &=& \displaystyle\testint\left(f\displaystyle\testint(gS(h_1))\cdot S(h_2)\right)\cdot h_3 \mbox{(using \rref{sweedler-id})}\\ &=&\displaystyle\testint\left(fS(g_1)\displaystyle\testint(g_2S(h_1))\right)\cdot h_2\\ &=&\displaystyle\testint\left(fS(g_1)\cdot \displaystyle\testint(g_2S(h_1))\right)\cdot h_2\\ &=& \displaystyle\testint\left(\displaystyle\testint(fS(g_1))\cdot g_2 S(h_1)\right)\cdot h_2\\ &=& \displaystyle\testint((g*f)S(h_1))\cdot h_2\\ &=& h*(g*f).\end{eqnarray*} Lemma \ref{chH} is proved. \rule{.75ex}{1.5ex}\vskip1ex We denote by $\check H$ the vector space $H$, equipped with the convolution product $*$. \subsection*{$H$-comodules and \mbox{$\check{H}$}-modules We now study the correspondence between $H$-comodules and \mbox{$\check{H}$}-modules. Let $V$ be an $H$-comodule. We define an action of $\mbox{$\check{H}$}$ on $V$ as follows: \begin{equation}\label{coact-act}\textstyle h*v:=v_0\displaystyle\testint( v_1S(h))\end{equation} We check the associativity: \begin{eqnarray*} g*(h*v)&=& v_0\displaystyle\testint(v_1S(g))\cdot \displaystyle\testint(v_2S(h))\\ (\mbox{using }\rref{sweedler-id}) &=& v_0\displaystyle\testint\left(v_1S(h_1)\displaystyle\testint(h_2S(g))\right)\\ &=& v_0\displaystyle\testint(v_1S(g*h))\\ &=& (g*h)*v.\end{eqnarray*} Let $\phi:V\longrightarrow W$ be a morphism of $H$-comodules, i.e., $\phi(v)_0\otimes \phi(v)_1=\phi(v_0)\otimes v_1$. Then, for $h\in H, v\in v$, \begin{eqnarray*} \phi(h*v)&=& \phi(v_0\displaystyle\testint(v_1S(h))\\ &=& \phi(v_0)\displaystyle\testint(v_1S(h))\\ &=& \phi(v)_0\displaystyle\testint(\phi(v)_1S(h))\\ &=&h*\phi(v).\end{eqnarray*} Thus, $\phi$ is a morphism of \mbox{$\check{H}$}\ modules. We therefore have a functor $\mathcal F$ from the category of $H$-comodules into the category of \mbox{$\check{H}$}-modules, which is the identity functor on the underlying category of vector spaces. \begin{pro}\label{funF} The functor $\mathcal F$ defined above is full, faithful and exact.\end{pro} \proof From the definition of $\mathcal F$, we see that, as vector spaces, $\mbox{$\mathcal F$}(V)=V$ and $\mbox{$\mathcal F$}(\phi)=\phi$. Thus, the functor $\mbox{$\mathcal F$}$ is faithful and exact. It remains to show that $\mbox{$\mathcal F$}$ is full, which amounts to showing that if $\phi:V\longrightarrow W$ is a morphism of $\mbox{$\check{H}$}$-module then it is a morphism of $H$-comodule. By assumption, $\phi$ satisfies $$ h*(\phi(v))=\phi(h*v),\quad \forall h\in\mbox{$\check{H}$}$$ or, equivalently, $$\phi(v)_0\displaystyle\testint\left(\phi(v_1)S(h)\right)=\phi(v_0)\displaystyle\testint(\left(v_1S(h)\right), \quad \forall h\in\mbox{$\check{H}$}.$$ Since $\displaystyle\testint$ is faithful (see \rref{non-degenerate}), we conclude that $$\phi(v_0)\otimes \phi(v_1)=\phi(v_0)\otimes v_1.$$ In other words, $\phi$ is a homomorphism of $H$-comodules.\rule{.75ex}{1.5ex}\vskip1ex Let $V$ now be a cyclic \mbox{$\check{H}$}-module, i.e., there exists an element $\bar{v}\in V$, such that all $v\in V$ is obtained from $\bar{v}$ by the action of some $f\in H$. We then define the coaction of $H$ on $V$ by \begin{equation}\label{act-coact} \delta(v):=f_1* \bar{v}\otimes S(f_2),\end{equation} where $f$ is such that $f*\bar{v}=v$. First, we have to show that this coaction is well defined, which means that it does not depend on the choice of the representative elements $\bar{v}$ and $f$. To show that the definition does not depend on $\bar{v}$ we let $\bar{v}=g*\tilde{v}$ and show that the definition does not change when $\bar v$ is replaced by $\tilde{v}$, which means \begin{displaymath} f_1*\bar{v}\otimes S(f_2)=(f*g)_1*\tilde{v}\otimes S((f*g)_2).\end{displaymath} Replacing $\bar v$ on the left-hand side of this equation by $g*\tilde{v}$ and canceling $\tilde{v}$ in both sides, one is led to the following equation \begin{displaymath} f_1*g\otimes f_2=(f*g)_1\otimes (f*g)_2,\end{displaymath} which follows immediately from the definition. To show the independence on the choice of $f$ we assume $f*\bar{v}=0$ and show that $f_1*\bar{v}\otimes S(f_2)=0$. Indeed, we have $$\displaystyle\testint(f_2S(g))\cdot f_1*\bar{v}=(g*f)*\bar{v}=g*(f*v)=0,$$ for all $g$. Hence, according to Lemma \ref{lemnon-degenerate} $f_1*\bar{v}\otimes f_2=0$. We now proceed to check the co-associativity and co-unitary: \begin{eqnarray*} ({\mathchoice{\mbox{\rm id}_H\otimes\delta)\delta(v)&=&\delta(f_1*\bar{v})\otimes S(f_2)\\ &=& f_1*\bar{v}\otimes f_2\otimes f_3)\\ &=& f_1*\bar{v}\otimes \Delta(f_2)\\ &=&({\mathchoice{\mbox{\rm id}\otimes\Delta_H)\delta(v).\end{eqnarray*} \begin{eqnarray*}({\mathchoice{\mbox{\rm id}_V\otimes \varepsilon)\delta f(v)&=&f_1*\bar{v}\otimes\varepsilon(f_2)\\ &=&f*\bar{v}=v. \end{eqnarray*} Since simple modules are cyclic, we obtain a functor $\mathcal G$ from the category of completely reducible \mbox{$\check{H}$}-modules to the category of $H$-comodules. \begin{thm}\label{thmact-coact} The category of completely reducible $H$-comodules is equivalent to the category of completely reducible \mbox{$\check{H}$}-modules. \end{thm} \proof We show that the functors $\mathcal F\circ \mathcal G$ and $\mathcal G\circ \mathcal F$ are identity-functors on the categories of completely reducible $H$-comodules and \mbox{$\check{H}$}-comodules, respectively. Let $V$ be a simple $H$-comodule. Then $\mbox{$\check{V}$}={\mathcal F}(V)$ is a simple \mbox{$\check{H}$}-module. Let us fix $\bar{v}\in\mbox{$\check{V}$}$ and for $v\in\mbox{$\check{V}$}$, let $f\in\mbox{$\check{H}$}$ be such that $f*\bar{v}=v$. By definition, we have \begin{displaymath} v=f*\bar{v}=\bar{v}_0\displaystyle\testint(\bar{v}_1S(f)).\end{displaymath} The coaction of $H$ on $\mbox{$\mathcal G$}\circ \mbox{$\mathcal F$}(V)$ is \begin{eqnarray*}\delta_{\mathcal{G}\circ\mathcal{F}(V)}(v)&=&f_1*\bar{v}\otimes f_2\\ &=&\bar{v}_0\displaystyle\testint(\bar{v}_1S(f_1))\otimes f_2\\ &=&\bar v_0\otimes \bar v_1\displaystyle\testint(v_2S(f))\\ &=&\delta_V(v).\end{eqnarray*} Thus, $\mbox{$\mathcal G$}\circ\mbox{$\mathcal F$}$ is the identity functor. The assertion for $\mbox{$\mathcal F$}\circ\mbox{$\mathcal G$}$ is proved analogously. \rule{.75ex}{1.5ex}\vskip1ex \subsection*{\mbox{$\check{H}$}\ is an ideal of $H^*$ In the previous section we have seen that there exists a correspondence between $H$-comodules and \mbox{$\check{H}$}-modules. On the other hand, there exits a one-to-one correspondence between $H$-comodules and rational $H^*$-modules. It is then natural to ask about the relationship between $H^*$ and \mbox{$\check{H}$}. We now show that \mbox{$\check{H}$}\ is isomorphic to the left ideal generated by the integrals in $H^*$. It is well-known, that the rational submodule \mbox{$H^{\square} $}\ of $H^*$, considered as a left module on itself, i.e. the sum of all left ideals of $H^*$, which are finite dimensional (over ${\Bbb K}$), is an $H$-Hopf module (with an appropriate $H$-action) and hence isomorphic to the tensor product of $H$ with the space spanned by the integrals \cite[Thm 5.1.3]{sweedler}. Since the space of integrals is one-dimensional, we have an isomorphism between the two vector spaces $H$ and \mbox{$H^{\square} $}. This isomorphism can be given explicitly as follows. By means of the integral, every element of $H$ can by considered as a linear functional on $H$ itself: $H\ni h\longmapsto \displaystyle\testint^h\in H^*: \displaystyle\testint^h(g):=\displaystyle\testint(gS(h)).$ \begin{pro}\label{chh-hsq} The map $H\ni h\longmapsto \displaystyle\testint^h\in H^*$ defined above is an isomorphism of algebras $\mbox{$\check{H}$}\longrightarrow H^\square\subset H^*$.\end{pro} \proof We have \begin{eqnarray*} \displaystyle\testint^{f*h}(g)&=& \displaystyle\testint\left(gS(h_1\displaystyle\testint (h_2Sf))\right)\\ &=&\displaystyle\testint(gS(h_1))\displaystyle\testint(h_2S(f))\\ &=&\displaystyle\testint(g_1S(f))\displaystyle\testint(g_2S(h))\\ &=&(\displaystyle\testint^f*\displaystyle\testint^h)(g).\nonumber \mbox{ \rule{.75ex}{1.5ex}}\end{eqnarray*} In the strictly algebraic sense, \mbox{$\check{H}$}\ is not a Hopf algebra, unless $H$ is finite-dimensional. In the next section we will show that for compact Hopf $*$-algebra, there exists a natural topology in $H$ such that the completion of $H$ with respect to this topology is a topological Hopf algebra with approximate unit. \subsection*{Co-semi-simple Hopf algebras A Hopf algebra $H$ is called co-semisimple if any finite dimensional $H$-comodule decomposes into a direct sum of simple comodules. A Hopf algebra is co-semisimple if and only if it possesses an integral whose value at its unit element is nonzero. In this case, left and right integrals are equal \cite{sweedler}. Let $H$ be co-semi-simple Hopf algebra. Then it decomposes into a direct sum of simple sub-coalgebras, each of which is the coefficient space of a simple $H$-comodule \cite{sweedler} \begin{eqnarray}\label{coalg-decom}H\cong\bigoplus_{\lambda\in\Lambda}H_\lambda.\end{eqnarray} The set $\Lambda$ contains an element $0$ for which $H_0\cong{\Bbb K}$. The integral computed on $H_\lambda$ is zero for $\lambda\neq 0$. We consider the discrete topology on $\Lambda$. Let $C_0(\Lambda)$ denote the set of all compact subsets in $\Lambda$ containing 0. For any compact $K$, let \begin{eqnarray}\label{hk} H_K:=\bigoplus_{\lambda\in K}H_\lambda.\end{eqnarray} Then $H_K$ are subcoalgebra of $H$. For any $f\in H$, $f=\sum_{\lambda\in K} f_\lambda, f_\lambda\in H_\lambda,$ for some compact $K$. For each $\lambda\in\Lambda$, let $V_\lambda$ be the corresponding simple $H$-comodule. Note that the $V_\lambda$ are finite dimensional for all $\lambda\in \Lambda$. The isomorphism \rref{coalg-decom} becomes now an isomorphism of algebras between \mbox{$\check{H}$}\ and the direct sum of endomorphism ring of $V_\lambda$. \begin{eqnarray}\label{alg-decom} \mbox{$\check{H}$}\cong\bigoplus_{\lambda\in\Lambda}\mbox{$\check{H}$}_\lambda\cong\bigoplus_{\lambda\in\Lambda}\mbox{\rm End}_{\Bbb K}(V_\lambda).\end{eqnarray} Thus we have proved \begin{thm}\label{thmchH} Let $H$ be a co-semisimple Hopf algebra. Then \mbox{$\check{H}$}\ is isomorphic to the direct sum of full endomorphism rings of simple $H$-comodules. \end{thm} The algebra \mbox{$\check{H}$}\ does not have a unit element. Adding a unit to this algebra is problematic when we are dealing with the norm -- the unit element is some thing like `` the Dirac delta function " which never has finite norm. Instead we have a notion of $\delta$-type sequences, which approximate the unit. \begin{dn}\label{dnau}\rm \begin{itemize}\item[1)] Let $A$ be an algebra without unit. A system $\{e_i,i\in I\}$ of idempotents in $A$ is an {\it approximate unit} if \begin{itemize}\item[(i)] $I$ is a partially ordered set, \item[(ii)] For any $a\in A$, there exists $i=i(a)$, such that $e_ia=ae_j=a$, for all $j\geq i$.\end{itemize} For an approximate unit in a bialgebra we require further that \begin{itemize} \item[(iii)] $\varepsilon(e_i)=1, \forall i\in I$.\end{itemize} For an algebra with involution we require that \begin{itemize} \item[(iv)] There exist $f_i$, such that $e_i=f_if_i^*$, for all $i\in I$. \end{itemize} \item[2)]For a topological algebra, the condition (ii) above is replaced by \begin{itemize}\item[(ii')] the nets $\{ e_ia|i\in I\}$ and $\{ ae_i|i\in I\}$ converge to $a$.\end{itemize} \item[3)]A Hopf algebra with approximate unit is a bialgebra with approximate unit together with a system of endomorphism $\{ S_i|i\in I\}$, called an {\it approximate antipode}, satisfying $$m(S_i\otimes {\mathchoice{\mbox{\rm id})\Delta=m({\mathchoice{\mbox{\rm id}\otimes S_i)\Delta=e_i\varepsilon.$$ \end{itemize} \end{dn} The existence of such a sequence in our \mbox{$\check{H}$}\ is obvious. Indeed, let $e_K$ be the unit element in \mbox{$\check{H}$}$_K$. Then $\{e_K,K\in C_0(\Lambda)\}$ is an approximate unit in \mbox{$\check{H}$}. Thus \mbox{$\check{H}$}\ is an algebra with approximate unit. \section{Compact Hopf $*$-algebras}\label{sec2} In this section we construct from a compact Hopf $*$-algebra a bialgebra which is a Hilbert space. Thus, $\Bbb K=\Bbb C$. This is the first step toward the construction of our Hopf $C^*$-algebra. A good reference on compact Hopf $*$-algebra, where the algebra is referred to as CQG-algebra, is Dijkhuizen and Koornwinder \cite{dk}. By definition, a compact Hopf $*$-algebra over $\Bbb C$ is a co-semisimple Hopf algebra with an involutive $\Bbb C$-anti-linear anti-homomorphism $^*$ such that every simple comodule is unitarizable, i.e. we can define a scalar product on this comodule such that \begin{equation}\label{unitary-mod} <v_0,w>S(v_1)=<v,w_0>w_1^*,\mbox{ for }v,w\in H.\end{equation} For any orthonormal basis of this comodule, the corresponding coefficient matrix satisfy the orthogonality condition. More precisely, let ${\bf x}_1,{\bf x}_2,\ldots, {\bf x}_d$ be an orthonormal basis of the comodule and ${\bf U}=({\bf u}_i^j)$ be the corresponding coefficient matrix, i.e., $\delta({\bf x}_i)={\bf x}_k\otimes {\bf u}_i^k$. Then ${\bf U}$ satisfies ${\bf U} {\bf U}^*={\bf U}^*{\bf U}=I$, where ${{\bf U}^*}^j_i:={\bf u}_j^{i*}$. \begin{lem}\label{lemssquare}Let $H$ be a co-semisimple Hopf algebra. Then the square of the antipode on $H$ is co-inner, i.e., it can be given in terms of an invertible element $q$ of $H^*$ (i.e. a linear form of $H$): $$S^2(h)=q(h_1)h_2q^{-1}(h_2),$$ where the linear form $q$ is given by $q(h)=\displaystyle\testint(S^2(h_1)S(h_2)).$ \end{lem} \proof First we show that $S^2(h_1)q(h_2)=q(h_1)h_2,$ or equivalently \begin{eqnarray}\label{sq}S^2(h_1)\displaystyle\testint(S^2(h_2)S(h_3))=\displaystyle\testint(S^2(h_1)S(h_2))\cdot h_3.\end{eqnarray} We have \begin{eqnarray*} \displaystyle\testint(S^2(h_2)S(h_3))\cdot h_4S(h_1)&=&S^2(h_2)\displaystyle\testint(S^2(h_3)S(h_4))\cdot S(h_1) (\mbox{by } \rref{sweedler-id})\\ &=&\displaystyle\testint(S^2(h_1)S(h_2)).\end{eqnarray*} Thus, the lemma will be proved if we can show that $q$ is invertible as an element of $H^*$. Let $V$ be a finite dimensional $H$-comodule and $V^{**}$ be its double-dual. As vector space, $V^{**}$ is isomorphic to $V$ and the coaction of $H$ on $V^{**}$ is given by $\delta_{V^{**}}(v)=v_0\otimes S^2(v_1)$. Equation \rref{sq} shows that the map $v\longmapsto v_0\otimes q(v_1):V\longrightarrow V^{**}$ is a morphism of $H$-comodules. If $V$ is simple then $V^*$ and hence $V^{**}$ are simple. Therefore, the map above should be zero or invertible. To see that it cannot be zero we fix a basis ${\bf x}_1,{\bf x}_2,\ldots,{\bf x}_d$ of $V$ and let ${\bf U}=({\bf u}_i^j)$ be the coefficient matrix. Since have $\Delta({\bf u}_i^j)=\sum_k{\bf u}^j_k\otimes {\bf u}_i^k, \varepsilon({\bf u}_i^j)=\delta_i^j$, $$q(\sum_i{\bf u}_i^i)=\displaystyle\testint({\bf u}^k_iS({\bf u}_k^i))=d.$$ Thus, the map $v\longmapsto v_0\otimes q(v_1):V\longrightarrow V^{**}$ is an isomorphism of $H$-comodules. Therefore the form $q$ is invertible. \rule{.75ex}{1.5ex}\vskip1ex Set $Q_i^j=q({\bf u}_i^j)$. Then, according to Lemma \ref{lemssquare} the matrix $Q$ is the matrix of the isomorphism $V\longrightarrow V^{**}$ with respect to the basis ${\bf x}_1,{\bf x}_2,\ldots,{\bf x}_d$ of $V$ and we have \begin{equation}\label{ssquare}S^2({\bf u}_i^j)=Q^j_k{\bf u}^k_l{Q^{-1}}^l_j.\end{equation} The matrix $Q$ is very important in the study of $H$ and will be called {\it reflection matrix}. If $V$ is irreducible then the integral on ${\bf u}_i^jS({\bf u}_k^{l})$ can be given in terms of $Q$. In fact, from the left-invariance of $\displaystyle\testint$ we have \begin{equation}\label{compute-int} \displaystyle\testint({\bf u}_i^jS({\bf u}_k^l))={\bf u}_m^jS({\bf u}_k^n)\displaystyle\testint({\bf u}_i^mS({\bf u}_n^l)) \end{equation} Or equivalently \begin{displaymath} \displaystyle\testint({\bf u}_i^jS({\bf u}_k^l))\cdot {\bf u}_n^k={\bf u}_m^j\displaystyle\testint({\bf u}_i^mS({\bf u}_n^l)).\end{displaymath} Since $\{{\bf u}_i^j\}$ are linearly independent, $\displaystyle\testint({\bf u}_i^jS({\bf u}_k^l))=\delta_k^jC_i^l$ for some matrix $C=(C_i^l)$. On the other hand, according to \rref{ssquare}, we have ${\bf u}^m_iQ_m^nS({\bf u}_n^j)=Q^j_i$. Substituting this into \rref{compute-int} we get $\delta_n^mC_i^jQ_m^n=Q^i_j$. Thus $\mbox{\rm tr}(Q)\neq 0$ and $C_i^j=Q^j_i/\mbox{\rm tr}(Q)$, and we get \begin{equation}\label{int} \displaystyle\testint({\bf u}_i^jS({\bf u}^{l}_k))=\delta_k^jQ_i^l/\mbox{\rm tr}(Q).\end{equation} Analogously we have \begin{equation}\label{int2} \displaystyle\testint(S({\bf u}_i^{j}){\bf u}^l_k)=\delta_k^j(Q^{-1})_i^l/\mbox{\rm tr}(Q^{-1}).\end{equation} From Equations \rref{int} and \rref{int2} we get the rule for the convolution product for coefficients of simple comodules. Let $V_\lambda$ and $V_\mu$ be simple comodules and ${{\bf U}_\lambda},{{\bf U}_\mu}$ be the coefficient matrix with respect to some (orthogonal) bases of $V_\lambda$ and $V_\mu$, and $Q_\lambda:=q({\bf U}_\lambda), Q_\mu:=q({\bf U}_\mu)$. An element $f$ of $H_\lambda$ can be represented by a matrix $F_\lambda$: $f={F_\lambda}_j^i{{\bf u}_\lambda}^j_i=\mbox{trace}(F_\lambda {\bf U}_\lambda).$ Let $c_\lambda=\mbox{trace}(C_\lambda {\bf U}_\lambda), d_\mu=\mbox{trace}(D_\mu {\bf U}_\mu)$. The convolution product has the form \begin{eqnarray}\label{product-rule} c_\lambda * d_\lambda&=&{C_\lambda}^i_j{D_\lambda}^k_l({\bf u}_\lambda)_i^j*({\bf u}_\mu)_k^l\nonumber\\ &=&{C_\lambda}^i_j{D_\lambda}^k_l\delta^\lambda_\mu \frac{{Q_\lambda}^j_k}{\mbox{\rm tr}(Q_\lambda)}({\bf u}_\lambda)^l_i\nonumber\quad(\mbox{from \rref{conv-pr1} and \rref{int}})\\ &=&\frac{C_\lambda Q_\lambda D_\lambda}{\mbox{\rm tr}(Q)},\label{conv-pr2}\end{eqnarray} where in the left-hand side is the matrix product: $(CD)_i^j=C^i_kD^k_j$. Thus, with the $*$-product, $H_\lambda$ becomes a unital algebra, denoted by $\mbox{$\check{H}$}_\lambda$, and $\mbox{$\check{H}$}_\lambda\cong\mbox{\rm End}_{\Bbb C}(V_\lambda).$ Moreover, for any compact $K\subset\Lambda$, $\mbox{$\check{H}$}_K:=\bigoplus_{\lambda\in K}\mbox{$\check{H}$}_\lambda$ is also a unital algebra. Let $V_\lambda$ be a simple $H$-comodule, $\mbox{$\check{V}$}_\lambda$ be the corresponding \mbox{$\check{H}$}-module. The representation of \mbox{$\check{H}$}\ on $\mbox{$\check{V}$}_\lambda$ will be denoted by $\pi_\lambda$. Then according to \rref{coact-act}, the action of $f=\mbox{trace}(F_\lambda {\bf U}_\lambda)$ on ${\bf x}_i$ is given by \begin{equation}\label{f-action}f*{\bf x}_i={\bf x}_j\displaystyle\testint({\bf u}^j_iS(f))={\bf x}_j{F_\lambda}^j_l{Q_\lambda}^l_i.\end{equation} In other words, the matrix of $\pi_\lambda(f)$ with respect to the basis $\{{\bf x}_i\}$ is $F_\lambda Q_\lambda$. For compact Hopf $*$-algebra, we can show that the matrix $Q$, with respect to an orthonormal basis, is positive definite \cite{dk,woron1}. Let $V^*$ be the dual to the comodule $V$. Let ${\mathbf \xi}_1,{\mathbf \xi}_2,\ldots,{\mathbf \xi}_d$ be the basis of $V^*$, dual to the basis ${\bf x}_1,{\bf x}_2,\ldots,{\bf x}_d$. The corresponding coefficient matrix is then $S({\bf U}^t)$. Since $V^*$ is unitarizable, there exists a basis ${\mathbf \eta}_1,{\mathbf \eta}_2,\ldots, {\mathbf \eta}_d$ of $V^*$ which is orthonormal with respect to a scalar product, such that the corresponding coefficient matrix ${\bf W}=({\bf w}_i^j)$ satisfies $S({\bf W})={\bf W}^{*}$. Let $T$ be a matrix such that $\xi^i={\mathbf \eta}_jT_j^i$ then \begin{equation}\label{WtoU}{\bf w}^i_j={T^{-1}}^i_kS({\bf u}_l^k)T^l_j={T^{-1}}^i_k{\bf u}^{l*}_kT^l_j,\end{equation} since $({\bf u}^i_j)$ is also unitary. Therefore, by the involutivity of $*$ \begin{displaymath} S({\bf w}_i^j)={\bf w}^{i*}_j={\bar{T}}^{-1i}_{\:k}{\bf u}_k^l\bar{T}^l_j.\end{displaymath} Substituting $S(w_i^j)$ into the preceding equation and using \rref{ssquare} we, get $Q=\mbox{const}\cdot T^*T.$ A direct consequence of this fact is that $\displaystyle\testint$ is positive definite on $H$. For any $\lambda\in\Lambda$, let $\lambda^*$ be such that $(V_\lambda)^*=V_{\lambda^*}$. According to Lemma \ref{lemssquare}, $*$ is an involutive map on $\Lambda$. From above we see that the involution $*$ on $H$ maps $H_\lambda$ onto $H_{\lambda^*}$. Since the integral is zero on $H_\lambda$ except for $\lambda=0$, we see that $\displaystyle\testint (a)=\overline{\displaystyle\testint (a^*).}$ Consequently $\displaystyle\testint$ defines a scalar product $<,>$ on $H$, $<a,b>:=\displaystyle\testint(ab^*)$ giving rise to a norm on $H$ called $L^2$-norm. Let \mbox{$H_{L^2}$}\ ddenote the completion of $H$ with respect to the $L^2$-norm. Then \mbox{$H_{L^2}$}\ is a Hilbert space. \subsection*{The Hopf Algebra \mbox{$H_{L^2}$} Our aim is to define a Hopf algebra structure on \mbox{$H_{L^2}$}, where the product is an extension of the convolution product on \mbox{$\check{H}$}. First we have to extend the convolution product to \mbox{$H_{L^2}$}. This can be done provided that we showed that this product is continuous with respect to the $L^2$-norm. \begin{lem}\label{extend-pr} The convolution product on \mbox{$\check{H}$}\ ssatisfies $$\|f*g\|_{L^2}\leq \|f\|_{L^2}\cdot \|g\|_{L^2}.$$\end{lem} \proof It is sufficient to show this inequality for $f$ and $g$ belonging to the same coefficient space of a simple comodule. Then, in that case, $f$ and $g$ can be represented in the form $f = \mbox{\rm tr}(CU), \quad g = \mbox{\rm tr}(DU),$ for a unitary multiplicative matrix $U$ and some matrices $C$, $D$ with complex scalar entries. Using \rref{int}, \rref{product-rule} the indicated inequality is equivalent to the following $$\frac{\mbox{\rm tr}(CQC^*)\mbox{\rm tr}(DQD^*)}{tr(Q)^2} \geq \frac{\mbox{\rm tr}(DQC Q (DQC)^*)}{\mbox{\rm tr}(Q)^3}.$$ Let $T$ be such that $Q = TT^*$. Set $C_1 = CT,$ $D_1 = DT$ the above inequality has the form $$\mbox{\rm tr}(C_1C_1^*)\mbox{\rm tr}(D_1D_1^*)\mbox{\rm tr}(TT^*) \geq \mbox{\rm tr}(C_1T^*D_1D_1^*TC_1^*).$$ The last inequality follows immediately from the Minkowski inequality since $$\mbox{\rm tr}(CC^*)=\sum|c_i^j|^2. \mbox{ \rule{.75ex}{1.5ex}}$$ Therefore, the convolution product on \mbox{$\check{H}$}\ can be extended to \mbox{$H_{L^2}$}. \begin{lem} The family $\{ e_K \vert K\in C_0(\Lambda) \}$, where $e_K$ is the unit element of $\mbox{$\check{H}$}_K$, is an approximate unit in $H_{L^2}$. \end{lem} \proof What we need to show is that for any $h\in H_{L^2}$, there exists a composition sequence $K_1 \subset K_2 \subset \dots $ such that $$\lim_{n\to \infty} \Vert e_{K_n}*h - h\Vert = \lim_{n\to \infty} \Vert h*e_{K_n} - h\Vert = 0$$ Now, for any $h\in H_{L^2}$, there exists at most a countable set of $\lambda$, $\lambda \in \Lambda$ such that $h_\lambda = h*e_\lambda \ne 0$ and $h = \sum_\lambda h_\lambda$. The last series converges absolutely whence the assertion follows. \rule{.75ex}{1.5ex}\vskip1ex Then we have to define the coproduct. Notice that, the coproduct on \mbox{$H_{L^2}$}, if it exists, should be dual to the original product on $H$ by means of the integral. Thus, we consider, for an element $h\in H$, a linear functional $\phi_h:H\otimes H\longrightarrow \Bbb C$, $\phi_h(g\otimes f)=\displaystyle\testint(hgf)$. This is obviously continuous hence is extendable on $\mbox{$H_{L^2}$}\hat{\otimes}\mbox{$H_{L^2}$}$ ($\hat{\otimes}$ denotes the tensor product of Hilbert spaces). Hence, by Riesz theorem, there exists an element $\Delta_*(f)$ of $\mbox{$H_{L^2}$}\hat{\otimes}\mbox{$H_{L^2}$}$, such that $\phi_h(g\otimes f)=\displaystyle\testint\int(\Delta_*(h),g\otimes f)$. Thus, we get a map $H\longrightarrow \mbox{$H_{L^2}$}$, $h\longmapsto\Delta_*(h)$. Again this is continuous and hence induces a map $\mbox{$H_{L^2}$}\longrightarrow \mbox{$H_{L^2}$}\hat{\otimes}\mbox{$H_{L^2}$}$ which is the coproduct on \mbox{$H_{L^2}$}. The counit is given by $\varepsilon_*(f)=\displaystyle\testint(f)$. We check the axioms for the coproduct and counit. \begin{thm}\label{thm-hlh} $(H_{L^2}, *, e_K,\Delta_*,\varepsilon_*)$ is a Hopf algebra with approximate unit.\end{thm} \proof The proof consists of some lemmas. For convenience we shall use the notation $$\Delta_*(f) = \sum_{(f)} f^{(1)} \otimes f^{(2)} = f^{1} \otimes f^{2},$$ where the sum on the right-hand side is an absolute convergent series in $H_{L^2} \hat{\otimes} H_{L^2}.$ \rule{.75ex}{1.5ex}\vskip1ex \begin{lem} \label{coact-axiom} The coproduct and the counit satisfy $$(\Delta_* \otimes id_{H_{L^2}}) \Delta_* = (id_{H_{L^2}} \otimes \Delta_*) \Delta_*,$$ $$(\varepsilon_* \otimes id_{H_{L^2}})\Delta_* = (id_{H_{L^2}} \otimes \varepsilon_*)\Delta_* = id_{H_{L^2}},$$ in the second equation we identify ${\mathbb C} \hat{\otimes} H_{L^2}$ and $\mbox{$H_{L^2}$}\hat{\otimes} {\mathbb C}$ with $ H_{L^2}$. \end{lem} \proof By the faithfulness of $\displaystyle\testint$, the first equation is equivalent to $$\iiint (\Delta_* \otimes id)\Delta_*(f)\cdot g\otimes h \otimes k = \iiint((id \otimes \Delta_*)\Delta_* f\cdot g \otimes h\cdot k), \forall g,h,k \in H, f\in H_{L^2}$$ By definition, the left hand side is equal to $$\iiint (f^{11} \otimes f^{12} \otimes f^2\cdot g\otimes h \otimes k) = \displaystyle\testint(f^1\cdot gh)\displaystyle\testint(f^2k) = \displaystyle\testint fghk\cdot $$ So is also the right hand side, whence we obtain the first equation of the lemma. For the second equation of the lemma, we have, for all $g\in H$, $$\displaystyle\testint(\varepsilon(f^1) f^2\cdot g = \displaystyle\testint(f^1)\displaystyle\testint(f^2g) = \displaystyle\testint fg\cdot $$ The lemma \ref{coact-axiom} is therefore proved. \rule{.75ex}{1.5ex}\vskip1ex \begin{lem} \label{copr-pr} The map $\Delta_* : H_{L^2} \to H_{L^2} \hat{\otimes} H_{L^2}$ is a homomorphism of algebras. \end{lem} \proof This is equivalent to the fact that for any $h\cdot k\in H$, $f,g\in H_{L^2}$, $$\displaystyle\testint(f*g)hk = \displaystyle\testint((f^1 * g^1)h)\cdot \displaystyle\testint((f^2*g^2)k).$$ The left hand side is equal to $$\displaystyle\testint (fh_1k_1)\cdot \displaystyle\testint(gh_2k_2) = \displaystyle\testint(f^1h_1)\displaystyle\testint(f^2k_1)\displaystyle\testint(g^1h_2)\displaystyle\testint(g^2k_2) = \displaystyle\testint((f^1*g^1)h)\cdot \displaystyle\testint((f^2*g^2)k).$$ The Lemma \ref{copr-pr} is proved. \rule{.75ex}{1.5ex}\vskip1ex The above two lemmas imply that $H_{L^2}$ is a bialgebra with approximate unity. It remains to find an approximate antipode. By definition, we have to find a system $\{ S_{*K} \vert K \in C_0(\Lambda) \}$ of linear endomorphism on $H_{L^2}$, such that $m_*(S_{*K} \otimes {\mathchoice{\mbox{\rm id}) \Delta_*$ and $m_*({\mathchoice{\mbox{\rm id} \otimes S_{*K})\Delta_*$ are approximate units. For any $h\in H_\lambda$, we define $S_*(h)\in H_{\lambda *}$ to be such that $\forall g \in H_\lambda$, $$\displaystyle\testint (hS(g)) = \displaystyle\testint (S_*(h) g).$$ In fact, $S_*(h)$ can be computed explicitly $$S_*(h) = S(h_1)p(h_2)q(h_3),$$ where $q$ is the linear functional defined in Lemma \ref{lemssquare}, $p$ is given by $$p(h) = \displaystyle\testint S(h_2) h_1, \forall h\in H.$$ Set $S_{*\lambda}\vert_{H_\mu} = 0,\forall \mu \ne \lambda$. For any $K \subset C_0(\Lambda)$ set $S_{* K} = \bigoplus_{\lambda \in K} S_{* \lambda}$. \begin{lem} $\{ S_{* K} \vert K\in C_0(\Lambda) \}$ is an approximate antipode. \end{lem} \proof We show that $$S_{* K}(f^1) * f^2 =\varepsilon_*(f)e_K= \displaystyle\testint( f) e_K, \forall f \in H, K \in C_0(\Lambda).$$ Since $\displaystyle\testint S_{* K}(f) g = \displaystyle\testint fS(g_K),$ we have $$\left.\begin{matrix} \displaystyle\testint\left((S_{* K}(f^1) * f^2) g\right) & = & \displaystyle\testint(S_{* K}(f^1)g_1) . \displaystyle\testint (f^2 g_2)\\ & = & \displaystyle\testint(f^1S(g_{K1})).\displaystyle\testint (f^2 g_2)\\ & = & \displaystyle\testint (f) \displaystyle\testint (e_K g), \end{matrix}\right.$$ whence the assertion follows. The lemma is proved. \rule{.75ex}{1.5ex}\vskip1ex We have therefore finished the proof of Theorem \ref{thm-hlh}. \rule{.75ex}{1.5ex}\vskip1ex \noindent{\bf Remark.} Since the antipode on $H$ is not involutive, the $*$-structure on $\check{H}$ defined by $h^\star = S(h^*)$ cannot be extended to $H_{L^2}$, because it is not continuous with respect to the given norm. To overcome this obstruction we have to pass to the $C^*$ envelope of $H_{L^2}$ which is the subject of our next section. \section{The $C^*$-algebra $H_{C^*}$}\label{sec3} Let $H$ be a compact Hopf $*$-algebra. Let $\star$ be the involutive map on \mbox{$\check{H}$}\ defined in the previous section $f^\star=S(f^*)$. \begin{lem}\label{lemunitary-mod} All simple \mbox{$\check{H}$}-modules have the structure of $\star$-modules.\end{lem} \proof According to Theorem \ref{thmact-coact}, a simple \mbox{$\check{H}$}-comodule is equivalent to some module $\mbox{$\check{V}$}_\lambda$ induced from a simple unitary $H$-comodule $V_\lambda$. We check that, with respect to the given scalar product on $V_\lambda$, $$<h*v,w>=<v,h^\star*w>, \forall h\in \mbox{$\check{H}$}, v,w\in \mbox{$\check{V}$}_\lambda.$$ Indeed, \begin{eqnarray*} <h*v,w>&=&<v_0,w>\overline{\displaystyle\testint(hS(v_1))}\\ &=&<v,w_0>\overline{\displaystyle\testint(hw_1*)}\quad\mbox{ according to \rref{unitary-mod} }\\ &=& <v,w_0>\displaystyle\testint(w_1h^*)\\ &=&<v,w_0>\displaystyle\testint(S(h^*)S(w_1))\\ &=&<v,h^\star*w>. \mbox{ \rule{.75ex}{1.5ex}}\end{eqnarray*} We introduce the following semi-norm on \mbox{$\check{H}$}\ \begin{equation}\label{c*norm} \|f\|_{C^*}:=\sup_{\lambda\in\Lambda}\|\pi_\lambda(f)\|.\end{equation}\ The lemma below will show that this is a norm, i.e., it is bounded. Let ${\bf U}={\bf U}_\lambda$ be a unitary coefficient matrix corresponding to the simple unitary comodule $V_\lambda$ and $Q=Q_\lambda$ be the corresponding reflection matrix. If $0\neq f\in\mbox{$\check{H}$}_\lambda$, then $f=F_j^i{\bf u}_i^j$ for a complex matrix $F$. Since $f$ act on $\mbox{$\check{V}$}_\mu$ as zero if $\mu\neq \lambda$, then according to \rref{f-action}, \begin{equation}\label{norm-f} \|f\|_{C^*}=\|\pi_\lambda(f)\|=\|FQ\|>0.\end{equation} \begin{lem}\label{leml2c*-norm}The $L^2$- and $C^*$-norms on \mbox{$\check{H}$}\ satisfy $$\|f\|_{L^2}\geq \|f\|_{C^*}.$$ \end{lem} \proof Since $\|f\|_{L^2}^2=\sum_\lambda\|f_\lambda\|_{L^2}^2$ and $\|f\|_{C^*}=\sup_\lambda\|f_\lambda\|_{C^*}$, it is sufficient to check the above equation for $f\in\mbox{$\check{H}$}_\lambda$. Thus $f=F^i_j{\bf u}_i^j$, for $({\bf u}_i^j)=({\bf u}_\lambda)_i^j$ a unitary multiplicative matrix. The desired inequality is then equivalent to $$\frac{\mbox{\rm tr}(FQF^*)}{\mbox{\rm tr}(Q)}\geq \frac{\|FQ\|^2}{\mbox{\rm tr}(Q)^2}.$$ Since $Q$ is positive definite, $Q=T^*T$. Hence $$\|FQ\|^2=\|FTT^*\|\leq \|FT\|^2\|T^*\|^2\leq\mbox{\rm tr}(FQF^*)\mbox{\rm tr}(Q).$$ Here we use the inequality $$\|A\|^2\leq \mbox{\rm tr}(AA^*). \mbox{ \rule{.75ex}{1.5ex}}$$ We define the algebra \mbox{$H_{C^*}$}\ to be the completion of \mbox{$\check{H}$}\ with respect to the norm $\|\cdot\|_{C^*}$. Since $\|\cdot\|_{C^*}$ is an operator norm, \mbox{$H_{C^*}$}\ is a $C^*$-algebra. By virtue of Lemma \ref{leml2c*-norm} we have \begin{eqnarray}\label{hl2hcs}\mbox{$H_{L^2}$}\subset\mbox{$H_{C^*}$}.\end{eqnarray} From the decomposition \rref{alg-decom} of \mbox{$\check{H}$}\ and the definition of the norm $\|\cdot\|$, we can easily obtain an isomorphism of the two $C^*$-algebras \begin{eqnarray}\label{gelfand} \mbox{$H_{C^*}$}\cong {\prod_{\lambda\in\Lambda}}'\mbox{\rm End}_{\Bbb C}(V_\lambda),\end{eqnarray} where the product on the right-hand side of the equation is over all families $\{x_\lambda\in \mbox{\rm End}_{\Bbb C}(V_\lambda)\}$ with $\|x_\lambda\|\longrightarrow 0$ as $\lambda\longrightarrow \infty$. \begin{lem}\label{uni-repr} All irreducible unitary representations of \mbox{$H_{C^*}$}\ are finite-dimensional and irreducible over \mbox{$\check{H}$}.\end{lem} \proof Let $\pi:\mbox{$H_{C^*}$}\longrightarrow {\mathcal B}(\H)$ be an irreducible unitary representation of \mbox{$H_{C^*}$}\ in a Hilbert space $\H$. Then $\H$ is a \mbox{$\check{H}$}-module. Let $M\subset\H$ be a simple \mbox{$\check{H}$}-submodule. Then it has a structure of a simple $H$-comodule, hence is finite dimensional over ${\Bbb C}$. Thus $M$ is closed in $\H$. Since \mbox{$\check{H}$}\ is dense in \mbox{$H_{C^*}$}\ and $M$ is closed in $\H$, $M$ is a representation of \mbox{$H_{C^*}$}. By the irreduciblity of $\pi$, we conclude that $M=\H$.\rule{.75ex}{1.5ex}\vskip1ex As a corollary of Lemmas \ref{lemunitary-mod} and \ref{uni-repr}, we have \begin{pro} There exists a 1-1 correspondence between irreducible unitary representations of \mbox{$H_{C^*}$}\ and those of \mbox{$\check{H}$}.\end{pro} Thus \mbox{$H_{C^*}$}\ is a $C^*$-algebra of type I. From the construction of \mbox{$H_{C^*}$}\, it is easy to see, that if $G$ is a compact group and $H={\BBb C}[G]$ -- the Hopf algebra of representative functions on $G$, which is a dense subalgebra of the algebra $C^\infty(G)$ of all continous complex valued functions on $G$, then $\mbox{$H_{C^*}$}\cong C^*(G)$ -- the group $C^*$-algebra of $G$. In the case of a compact group $G$, the action of $G$ on any representation can be recovered by the action of $C^*(G)$, by using $\delta$-type sequences. In our case, we do not know, what the quantum group defining $H$ is. Therefore we have to introduce a Hopf algebra structure on \mbox{$H_{C^*}$}\ to exhibit the ``group'' property of this algebra. \begin{thm}\label{c*withau} \mbox{$H_{C^*}$}\ is a Hopf $C^*$-algebra with approximate unit.\end{thm} \proof Recall that the topological tensor product $\mbox{$H_{C^*}$}\hat{\otimes}\mbox{$H_{C^*}$}$ is the completion of $\mbox{$\check{H}$}\otimes\mbox{$\check{H}$}$ with respect to the norm $$\|f\otimes g\|_{C^*}:=\sup_{\lambda,\mu\in\Lambda}\|\pi_\lambda(f)\otimes \pi_\mu(g)\|.$$ For any $f\in\mbox{$\check{H}$}\subset\mbox{$H_{L^2}$}$, $\Delta(f)\in\mbox{$H_{C^*}$}\hat{\otimes}\mbox{$H_{C^*}$}.$ Thus, it is sufficient to show that $$\|\Delta(f)\|_{C^*}\leq\|f\|_{C^*}.$$ But this inequality is obvious, as for any $\lambda,\mu\in\Lambda$, $V_\lambda\otimes V_\mu$ is a representation of \mbox{$H_{L^2}$}\ by means of the map $(\pi_\lambda\otimes\pi_\mu)\Delta$, hence decomposes into a direct sum of irreducible representations, namely $\pi_\lambda\otimes\pi_\mu\cong \oplus_\gamma c_{\lambda\mu}^\gamma \pi_\gamma$, consequently $$\|(\pi_\lambda\otimes\pi_\mu)\Delta(f)\|\leq\sup_{\gamma\in\Lambda}\|\pi_\gamma(f)\|=\|f\|_{C^*}.$$ To show that $\varepsilon_*$ extends onto \mbox{$H_{C^*}$}\ we remark that $\varepsilon_*:\mbox{$\check{H}$}\longrightarrow {\Bbb C}$ is multiplicative, i.e., induces a representation, for $$\varepsilon_*(f*g)=\displaystyle\testint(f_1\displaystyle\testint(f_2S(g)))=\displaystyle\testint(f)\displaystyle\testint(g)=\varepsilon_*(f)\varepsilon_*(g).$$ Hence $$|\varepsilon_*(f)\leq \|f\|_{C^*}$$ thus $\varepsilon_*$ extends onto \mbox{$H_{C^*}$}. According to Lemma \ref{lemunitary-mod}, for $f\in\mbox{$\check{H}$}$, $$\pi_\lambda(f^\star)=\pi_\lambda(f)^\star, \forall \lambda\in\Lambda.$$ Therefore the involutive map $\star$ extends on $H_{C^*}$. Using \rref{ssquare} and \rref{conv-pr2}, we can show that the unit element $e_\lambda$ in $H_\lambda$ satisfies $e_\lambda^\star=e_\lambda$. Since $\{e_K|K\in C_0(\Lambda)\}$ is an approximate unit of \mbox{$\check{H}$}\ and since \mbox{$\check{H}$}\ is dense in \mbox{$H_{C^*}$}, $\{e_K|K\in C_0(\Lambda)\}$ is an approximate unit on \mbox{$H_{C^*}$}. Consequently, $\{S_K|K\in C_0(\Lambda)\}$ is an approximate antipode on \mbox{$H_{C^*}$}. The proof of Theorem \ref{c*withau} is completed. \rule{.75ex}{1.5ex}\vskip1ex
\section{Introduction} If theoretical cosmologists are the flyboys of astrophysics, they were flying on fumes in the 1990s. Since the early 1980s inflation and cold dark matter (CDM) have been the dominant theoretical ideas in cosmology. However, a key prediction of inflation, a flat Universe (i.e., $\Omega_0 \equiv \rho_{\rm total}/\rho_{\rm crit} = 1$), was beginning to look untenable. By the late 1990s it was becoming increasingly clear that matter only accounted for 30\% to 40\% of the critical density (see e.g., Turner, 1999). Further, the $\Omega_M =1$, COBE-normalized CDM model was not a very good fit to the data without some embellishment (15\% or so of the dark matter in neutrinos, significant deviation from from scale invariance -- called tilt -- or a very low value for the Hubble constant; see e.g., Dodelson \mbox{\it et al.}, 1996). Because of this and their strong belief in inflation, a number of inflationists (see e.g., Turner, Steigman \& Krauss, 1984 and Peebles, 1984) were led to consider seriously the possibility that the missing 60\% or so of the critical density exists in the form of vacuum energy (cosmological constant) or something even more interesting with similar properties (see Sec. 3 below). Since determinations of the matter density take advantage of its enhanced gravity when it clumps (in galaxies, clusters or superclusters), vacuum energy, which is by definition spatially smooth, would not have shown up in the matter inventory. Not only did a cosmological constant solve the ``$\Omega$ problem,'' but $\Lambda$CDM, the flat CDM model with $\Omega_M\sim 0.4$ and $\Omega_\Lambda\sim 0.6$, became the best fit universe model (Turner, 1991 and 1997b; Krauss \& Turner, 1995; Ostriker \& Steinhardt, 1995; Liddle \mbox{\it et al.}, 1996). In June 1996, at the Critical Dialogues in Cosmology Meeting at Princeton University, the only strike recorded against $\Lambda$CDM was the early SN Ia results of Perlmutter's group (Perlmutter \mbox{\it et al.}, 1997) which excluded $\Omega_\Lambda > 0.5$ with 95\% confidence. The first indirect experimental hint for something like a cosmological constant came in 1997. Measurements of the anisotropy of the cosmic background radiation (CBR) began to show evidence for the signature of a flat Universe, a peak in the multipole power spectrum at $l=200$. Unless the estimates of the matter density were wildly wrong, this was evidence for a smooth, dark energy component. A universe with $\Omega_\Lambda \sim 0.6$ has a smoking gun signature: it is speeding up rather than slowing down. In 1998 came the SN Ia evidence that our Universe is speeding up; for some cosmologists this was a great surprise. For many theoretical cosmologists this was the missing piece of the grand puzzle and the confirmation of a prediction. \section{The theoretical case for accelerated expansion} The case for accelerated expansion that existed in January 1998 had three legs: growing evidence that $\Omega_M \sim 0.4$ and not 1; the inflationary prediction of a flat Universe and hints from CBR anisotropy that this was indeed true; and the failure of simple $\Omega_M =1$ CDM model and the success of $\Lambda$CDM. The tension between measurements of the Hubble constant and age determinations for the oldest stars was also suggestive, though because of the uncertainties, not as compelling. Taken together, they foreshadowed the presence of a cosmological constant (or something similar) and the discovery of accelerated expansion. To be more precise, Sandage's deceleration parameter is given by \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} where 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 $w_X = -1$. For $\Omega_0 = 1$, $\Omega_M =0.4$ and $w_X < -{5\over 9}$, the deceleration parameter is negative. The kind of dark component needed to pull cosmology together implies accelerated expansion. \subsection{Matter/energy inventory: $\Omega_0 =1\pm 0.2$, $\Omega_M=0.4\pm 0.1$} There is a growing consensus that the anisotropy of the CBR offers the best means of determining the curvature of the Universe and thereby $\Omega_0$. 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 \mbox{\it et al.}, 1994). It works like this. 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 and are from 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, CBI, 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 and Wilkinson, 1999). \begin{figure} \centerline{\psfig{figure=knox_better.eps,width=3.5in}} \caption{Current CBR anisotropy data, averaged and binned to reduce error bars and visual confusion. The theoretical curve is for the $\Lambda$CDM model with $H_0=65\,{\rm km\, s^{-1}\,Mpc^{-1}}$ and $\Omega_M =0.4$; note the goodness of fit (Figure courtesy of L. Knox). } \label{fig:cbr_knox} \end{figure} 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 \mbox{\it 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. Mohr \mbox{\it et al.}\ (1998) have compiled the gas to total mass ratios determined from x-ray measurements for a sample of 45 clusters; they find $f_{\rm gas} = (0.075\pm 0.002)h^{-3/2}$ (see Fig.~\ref{fig:gas}). 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, most of which have now been tested. For example, the agreement of S-Z and x-ray gas masses implies that gas clumping is not significant; the agreement of x-ray and lensing estimates for the total mass implies that hydrostatic equilibrium is a good assumption; the gas fraction does not vary significantly with cluster mass. \begin{figure} \centerline{\psfig{figure=cbf2.eps,width=3.5in}} \caption{Cluster gas fraction as a function of cluster gas temperature for a sample of 45 galaxy clusters (Mohr \mbox{\it 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} \subsection{Dark energy} The apparently contradictory results, $\Omega_0 = 1\pm 0.2$ and $\Omega_M = 0.4\pm 0.1$, can be reconciled by the presence of a dark-energy component that is nearly smoothly distributed. The cosmological constant is the simplest possibility and it has $p_X=-\rho_X$. There are other possibilities for the smooth, dark energy. As I now discuss, other constraints imply that such a component must have very negative pressure ($w_X \mathrel{\mathpalette\fun <} -{1\over 2}$) leading to the prediction of accelerated expansion. To begin, parameterize the bulk equation of state of this unknown component: $w \equiv p_X/\rho_X$ (Turner \& White, 1997). 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 \mbox{\it et al.}, 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=3.5in}} \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=3.5in}} \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 $q_0 = 0.5 + 1.5w_X\Omega_X \simeq 0.5 + 0.9w < 0$ for $w<-{5\over 9}$. \subsection{$\Lambda$CDM} The cold dark matter scenario for structure formation is the most quantitative and most successful model ever proposed. Two of its key features are inspired by inflation: almost scale invariant, adiabatic density perturbations with Gaussian statistical properties and a critical density Universe. The third, nonbaryonic dark matter is a logical consequence of the inflationary prediction of a flat universe and the BBN-determination of the baryon density at 5\% of the critical density. There is a very large body of data that is consistent with it: the formation epoch of galaxies and distribution of galaxy masses, galaxy correlation function and its evolution, abundance of clusters and its evolution, large-scale structure, and on and on. In the early 1980s attention was focused on a ``standard CDM model'': $\Omega_0=\Omega_M =1$, $\Omega_B = 0.05$, $h=0.50$, and exactly scale invariant density perturbations (the cosmological equivalent of DOS 1.0). The detection of CBR anisotropy by COBE DMR in 1992 changed everything. First and most importantly, the COBE DMR detection validated the gravitational instability picture for the growth of large-scale structure: The level of matter inhomogeneity implied at last scattering, after 14 billion years of gravitational amplification, was consistent with the structure seen in the Universe today. Second, the anisotropy, which was detected on the $10^\circ$ angular scale, permitted an accurate normalization of the CDM power spectrum. For ``standard cold dark matter'', this meant that the level of inhomogeneity on all scales could be accurately predicted. It turned out to be about a factor of two too large on galactic scales. Not bad for an ab initio theory. With the COBE detection came the realization that the quantity and quality of data that bear on CDM was increasing and that the theoretical predictions would have to match their precision. Almost overnight, CDM became a ten (or so) parameter theory. For astrophysicists, and especially cosmologists, this is daunting, as it may seem that a ten-parameter theory can be made to fit any set of observations. This is not the case when one has the quality and quantity of data that will soon be available. In fact, the ten parameters of CDM + Inflation are an opportunity rather than a curse: Because the parameters depend upon the underlying inflationary model and fundamental aspects of the Universe, we have the very real possibility of learning much about the Universe and inflation. The ten parameters can be organized into two groups: cosmological and dark-matter (Dodelson \mbox{\it et al.}, 1996). \smallskip \centerline{\it Cosmological Parameters} \vspace{3pt} \begin{enumerate} \item $h$, the Hubble constant in units of $100\,{\rm km\,s^{-1}}\,{\rm Mpc}^{-1}$. \item $\Omega_Bh^2$, the baryon density. Primeval deuterium measurements and together with the theory of BBN imply: $\Omega_Bh^2 = 0.02 \pm 0.002$. \item $n$, the power-law index of the scalar density perturbations. CBR measurements indicate $n=1.1\pm 0.2$; $n=1$ corresponds to scale-invariant density perturbations. Many inflationary models predict $n\simeq 0.95$; range of predictions runs from $0.7$ to $1.2$. \item $dn/d\ln k$, ``running'' of the scalar index with comoving scale ($k=$ wavenumber). Inflationary models predict a value of ${\cal O}(\pm 10^{-3})$ or smaller. \item $S$, the overall amplitude squared of density perturbations, quantified by their contribution to the variance of the CBR quadrupole anisotropy. \item $T$, the overall amplitude squared of gravity waves, quantified by their contribution to the variance of the CBR quadrupole anisotropy. Note, the COBE normalization determines $T+S$ (see below). \item $n_T$, the power-law index of the gravity wave spectrum. Scale-invariance corresponds to $n_T=0$; for inflation, $n_T$ is given by $-{1\over 7}{T\over S}$. \end{enumerate} \smallskip \centerline{\it Dark-matter Parameters} \vspace{3pt} \begin{enumerate} \item $\Omega_\nu$, the fraction of critical density in neutrinos ($=\sum_i m_{\nu_i}/90h^2$). While the hot dark matter theory of structure formation is not viable, we now know that neutrinos contribute at least 0.3\% of the critical density (Fukuda \mbox{\it et al.}, 1998). \item $\Omega_X$ and $w_X$, the fraction of critical density in a smooth dark-energy component and its equation of state. The simplest example is a cosmological constant ($w_X = -1$). \item $g_*$, the quantity that counts the number of ultra-relativistic degrees of freedom. The standard cosmology/standard model of particle physics predicts $g_* = 3.3626$. The amount of radiation controls when the Universe became matter dominated and thus affects the present spectrum of density inhomogeneity. \end{enumerate} A useful way to organize the different CDM models is by their dark-matter content; within each CDM family, the cosmological parameters vary. One list of models is: \begin{enumerate} \item sCDM (for simple): Only CDM and baryons; no additional radiation ($g_*=3.36$). The original standard CDM is a member of this family ($h=0.50$, $n=1.00$, $\Omega_B=0.05$), but is now ruled out (see Fig.~\ref{fig:cdm_sum}). \item $\tau$CDM: This model has extra radiation, e.g., produced by the decay of an unstable massive tau neutrino (hence the name); here we take $g_* = 7.45$. \item $\nu$CDM (for neutrinos): This model has a dash of hot dark matter; here we take $\Omega_\nu = 0.2$ (about 5\,eV worth of neutrinos). \item $\Lambda$CDM (for cosmological constant) or more generally xCDM: This model has a smooth dark-energy component; here, we take $\Omega_X = \Omega_\Lambda = 0.6$. \end{enumerate} \begin{figure} \centerline{\psfig{figure=cdm_sum.eps,width=3in}} \caption{Summary of viable CDM models, based upon CBR anisotropy and determinations of the present power spectrum of inhomogeneity (Dodelson \mbox{\it et al.}, 1996).} \label{fig:cdm_sum} \end{figure} Figure \ref{fig:cdm_sum} summarizes the viability of these different CDM models, based upon CBR measurements and current determinations of the present power spectrum of inhomogeneity (derived from redshift surveys). sCDM is only viable for low values of the Hubble constant (less than $55\,{\rm km\,s^{-1}}\,{\rm Mpc}^{-1}$) and/or significant tilt (deviation from scale invariance); the region of viability for $\tau$CDM is similar to sCDM, but shifted to larger values of the Hubble constant (as large as $65\,{\rm km\,s^{-1}}\,{\rm Mpc}^{-1}$). $\nu$CDM has an island of viability around $H_0\sim 60\,{\rm km\,s^{-1}}\,{\rm Mpc}^{-1}$ and $n\sim 0.95$. $\Lambda$CDM can tolerate the largest values of the Hubble constant. While the COBE DMR detection ruled out ``standard CDM,'' a host of attractive variants were still viable. However, when other very relevant data are considered too -- e.g., age of the Universe, determinations of the cluster baryon fraction, measurements of the Hubble constant, and limits to $\Omega_\Lambda$ -- $\Lambda$CDM emerges as the hands-down-winner of ``best-fit CDM model'' (Krauss \& Turner, 1995; Ostriker \& Steinhardt, 1995; Liddle \mbox{\it et al.}, 1996; Turner, 1997b). At the time of the Critical Dialogues in Cosmology meeting in 1996, the only strike against $\Lambda$CDM was the absence of evidence for its smoking gun signature, accelerated expansion. \begin{figure} \centerline{\psfig{figure=kraus.eps,width=3.5in}} \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} \subsection{Missing energy found!} In 1998 evidence for the accelerated expansion anticipated by theorists 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 \mbox{\it et al.}, 1998) and the High-z Supernova Search Team (Riess \mbox{\it et al.}, 1998), working independently and using different methods of analysis, each found evidence for accelerated expansion. Perlmutter \mbox{\it 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! As I have tried to explain, cosmologists were quick than most to believe, as accelerated expansion was the missing piece of the puzzle found. Recently, two other studies, one based upon the x-ray properties of rich clusters of galaxies (Mohr \mbox{\it et al.}, 1999) and the other based upon the properties of double-lobe radio galaxies (Guerra \mbox{\it 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 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; 2) Because of the modeling uncertainties and lack of sensitivity for $\Omega_\Lambda <0.55$, lensing has little power in strictly constraining $\Lambda$ or a dark component; and 3) 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. \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. 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. Further, these numbers point to $\Lambda$CDM (or something similar) as the cold dark matter model. Another body of observations already support this as the best fit model. Cosmic concordance indeed! \begin{figure} \centerline{\psfig{figure=omegalambda.eps,width=3.5in}} \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} \section{What is the 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 gravitational 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. The combination of SN Ia, CBR and large-scale structure data are already beginning to significantly constrain models (Perlmutter, Turner \& White, 1999), and interestingly enough, the cosmological constant is still the best fit (see Fig.~\ref{fig:composite}). The rolling scalar field scenario (aka mini-inflation or quintessence) has received a lot of attention over the past decade (Freese \mbox{\it et al.}, 1987; Ozer \& Taha, 1987; Ratra \& Peebles, 1988; Frieman \mbox{\it et al.}, 1995; Coble \mbox{\it et al.}, 1996; Turner \& White, 1997; Caldwell \mbox{\it et al.}, 1998; Steinhardt, 1999). 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 \mbox{\it 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). \begin{figure} \centerline{\epsfxsize=12cm \epsfbox{fig1.ps}} \caption{Contours of likelihood, from $0.5\sigma$ to $2\sigma$, in the $\Omega_M$--$w_{\rm eff}$ plane. Left: The thin solid lines are the constraints from LSS and the CMB. The heavy lines are the SN Ia constraints for constant $w$ models (solid curves) and for a scalar-field model with an exponential potential (broken curves). Right: The likelihood contours from all of our cosmological constraints for constant $w$ models (solid) and dynamical scalar-field models (broken). Note: at 95\% cl $w_{\rm eff}$ must be less $-0.6$, and the cosmological constant is the most likely solution (from Perlmutter, Turner \& White, 1999).} \label{fig:composite} \end{figure} \section{Looking ahead} Theorists often require new results to pass Eddington's test: No experimental result should be believed until confirmed by theory. While provocative (as Eddington had apparently intended it to be), it embodies the wisdom of mature science. Results that bring down the entire conceptual framework are very rare indeed. \begin{figure} \centerline{\psfig{figure=cos.pol4.eps,width=3.5in}} \caption{The 95\% confidence interval for the reconstructed potential assuming luminosity distance errors of 5\% and 2\% (shaded areas) and the original potential (heavy line). For this reconstruction, $\Omega_M = 0.3$ and $V(\phi ) = V_0[1+\cos (\phi /f) ]$ (from Huterer \& Turner, 1998). } \label{fig:ht_recon} \end{figure} Both cosmologists and supernova theorists seem to use Eddington's test to some degree. It seems to me that the summary of the SN Ia part of the meeting goes like this: We don't know what SN Ia are; we don't know how they work; but we believe SN Ia are very good standardizeable candles. I think what they mean is they have a general framework for understanding a SN Ia, the thermonuclear detonation of a Chandrasekhar mass white dwarf, and have failed in their models to find a second (significant) parameter that is consistent with the data at hand. Cosmologists are persuaded that the Universe is accelerating both because of the SN Ia results and because this was the missing piece to a grander puzzle. Not only have SN Ia led us to the acceleration of the Universe, but also I believe they will play a major role in unraveling the mystery of the dark energy. The reason is simple: we can be confident that the dark energy was an insignificant component in the past; it has just recently become important. While, the anisotropy of the CBR is indeed a cosmic Rosetta Stone, it is most sensitive to physics around the time of decoupling. (To be very specific, the CBR power spectrum is almost identical for all flat cosmological models with the same conformal age today.) SNe Ia probe the Universe just around the time dark energy was becoming dominant (redshifts of a few). My student Dragan Huterer and I (Huterer \& Turner, 1998) have been so bold as to suggest that with 500 or so SN Ia with redshifts between 0 and 1, one might be able to discriminate between the different possibilities and even reconstruct the scalar potential for the quintessence field (see Fig.~\ref{fig:ht_recon}). \begin{acknowledgments} My work is supported by the US Department of Energy and the US NASA through grants at Chicago and Fermilab. \end{acknowledgments}
\section{Introduction} The panorama of optical observations of Isolated Neutron Stars (INS) is limited by the faintness of the vast majority of them. Only one object, the Crab pulsar, has good, medium resolution optical (Nasuti et al. 1996) and near-UV spectral data (Gull et al. 1998), while for PSR0540-69 the synchrotron continuum has been measured by HST (Hill et al. 1997). For a few more cases acceptable multicolour photometry exists, while the rest of the data base (a grand total of less than 10 objects currently) consists of one/two - wavelengths detections (see Caraveo 1998 for a summary of the observational panorama). \\ Apart from the very young objects, characterized by flat, synchrotron-like spectra arising from energetic electron interactions in their magnetosphere, of particular interest are the middle-aged ones ($\sim 10^{5}~ yrs$ old). The non-thermal, magnetospheric emission should have faded enough (in the X-ray waveband at least) to render visible the thermal emission from the hot INS surface. Standard cooling calculations predict a surface temperature in the range $10^{5}-10^{6}~^{\circ}K$, in excellent agreement with recent X-ray observations of INS with thermal spectra (e.g. Becker \& Tr\"umper 1997). It is easy to predict the IR-optical-UV fluxes generated along the $\sim E^{2}$ Rayleigh-Jeans slope of the Planck curve best fitting the X-ray data, and to compare predictions to observations, where available. \\ In what follows, we shall concentrate on Geminga, which is certainly the most studied object of its class (Bignami \& Caraveo 1996 and refs. therein), and possibly of all INSs (except for the Crab). Its IR-optical-UV data (Bignami et al. 1996; Mignani et al. 1998) show the presence of a well defined emission feature. Such a feature is superimposed on the thermal continuum expected from the extrapolation of the black-body X-ray emission detected by ROSAT (e.g. Halpern \& Ruderman 1993). The presence of a clear maximum centered on V has been recently questioned by Martin et al. (1998) on the basis of a spectrum which is at the limit of the capability of the Keck telescope. The spectrum of Geminga, detected at a level of just 0.5\% of the dark sky, seems fairly flat and, although broadly consistent with earlier measurements, it is definitely above the flux measured in the B-band by the HST/FOC (Mignani et al. 1998). Therefore, in view of the faintness of the target, we shall concentrate on the photometric measurements, which have been repeated using different instrumental set-ups and appear more reliable than the available spectral data. \\ Recently, pulsations in the B-band have been tentatively detected by Shearer et al. (1998). Also in this case, the faintness of the source limits quite severely the S/N ratio and thus the statistical significance of the result. Indeed, a pulsed signal at just the 3.5 $\sigma$ level was found during only one of the three nights devoted to the project. If confirmed, these measurements would have deep implications on the mechanisms responsible for the optical emission of Geminga. However, in view of the rather low statistical significance of these results, we shall stand by the interpretation of Mignani et al. (1998) and propose a phenomenological model interpreting the feature as an atmospheric cyclotron line emission from Geminga's polar caps. \section{The Spectral Distribution} \begin{figure} \centerline{\hbox{\psfig{figure=8267_fig1.ps,height=8cm,clip=}}} \caption{The plot summarizes the complete multiband HST and ground-based photometry of Geminga. Three digits identify WFPC2/FOC imaging filters. The dashed line represents the extrapolation in the optical of the blackbody fit to the ROSAT data (Halpern \& Ruderman 1993), which, for the best fitting temperature of $T=5.77~10^{5}~^{\circ}K$ and the measured distance $d \simeq 157~ pc$ (Caraveo et al. 1996), yield an emitting radius $R=10~km$.} \end{figure} Fig.1 shows the complete I-to-near UV colours of Geminga. These were obtained both from the ground and from HST during ten years of continuing observational effort (from Mignani et al. 1998). Comparing the repeteadly confirmed V-band magnitude with the relative fluxes of the three FOC points (430W, 342W and 195W) in the B/UV and the upper limit in I, it is easy to recognize an emission feature superimposed to a $\sim E^{2}$ RJ-like continuum. It is thus obvious to directly compare the measured optical fluxes to the extrapolation of the soft X-ray blackbody spectrum. However, this apparently trivial step is easier said than done, since different X-ray observations have yielded for Geminga slightly different best fitting temperatures, with significantly different bolometric fluxes. For example, two independent ROSAT observations yielded best fitting temperatures of $5.77~10^{5}~^{\circ}K$ (Halpern \& Ruderman 1993) and $4.5 ~10^{5}~^{\circ}K$ (Halpern \& Wang 1997), respectively, implying, for the same Geminga distance, a factor of 3 difference in the emitting area. As a reference, we show in Fig. 1 the Rayleigh-Jeans extrapolation of the ROSAT X-ray spectrum obtained using the best fitting temperature derived by Halpern \& Ruderman (1993) which, at the Geminga distance (Caraveo et al. 1996), yields an emission radius of 10 km. Thus, while the optical, RJ-like, continuum is largely consistent with thermal emission from the neutron star surface, the feature around $\sim 6,000 \AA$, requires a different interpretation. \\ In the following, we propose a phenomenological model interpreting the feature seen in Fig.1 as an atmospheric cyclotron line emission from Geminga's polar caps. \section{The Ion Cyclotron Emission Model.} Comparisons between theory and observed cooling NS spectra have been performed, e.g., by Romani (1987) and Meyer et al. (1994). Basically, the star surface is thought to be surrounded by a colder, partially ionized, atmosphere. Such an atmosphere behaves as a broad-band, absorbing-emitting medium. The observed spectrum is due to a rather complicate process, involving transfer of radiating energy between regions of different depths, temperatures and chemical compositions. These models predict a deviation from the blackbody emission law marginally observed in cooling neutron star spectra. Models that account for the effects of the star magnetic field $B$ foresee absorption lines at the cyclotron frequencies. This is easily explained by the presence of resonant frequencies, created by the magnetic field, at which the radiation emitted from deeper regions is quite efficiently absorbed. On the other hand, a cyclotron emission "line", superimposed to the blackbody continuum, requires a thermal inversion of the stellar atmosphere. We will see in the following that a thin hot plasma layer is optically thick at the cyclotron frequency, so that the cyclotron emission is far more efficient than Bremsstrahlung in the same spectral range. For reasonable value of the plasma density, therefore, Bremsstrahlung can be neglected. Geminga's emission "line" might then be explained as an (ion) cyclotron emission from a magnetoplasma, consisting of a mixture of Hydrogen and Helium, surrounding the star surface. Any feature of the observed radiation spectrum of Geminga in the region of the proton cyclotron frequency is ultimately due to the accelerated motion of charged particles in the magnetic and electric fields of the rotating star. We assume, however, the upper part of the star atmosphere to be a fully ionized "classical" plasma layer, with a Maxwellian ion distribution function. The plasma layer is immersed in the inhomogeneous star dipole magnetic field. Accordingly, we develop the theory of the emission under the condition of applicability of the Kirchhoff's law. To this end it is necessary to consider the full dielectric response of the magnetized plasma, and address two key questions: (i) identification of an electromagnetic (e.m.) wave which can propagate in vacuo toward the observer and assessment of its thermal (black body) absorption/emission properties in a single or multiple species plasma, (ii) assessment of the emission line broadening mechanism as a consequence of Doppler effect, collisions, temperature and density inhomogeneity and averaging over the magnetic field. \begin{figure} {\bf (a)} \\ \centerline{\hbox{ \psfig{figure=8267_fig2a.ps,height=6cm,clip=}}} {\bf (b)} \\ \centerline{\hbox{ \psfig{figure=8267_fig2b.ps,height=6cm,clip=}}} \caption{ (a) Plot of the real part of the square of refractive index $n^{2}(\omega,\theta)$ of the fast wave vs $\omega/\omega_{ci}-1$ (where $\omega_{ci}$ is the ion cyclotron frequency), for different values of the angle $\theta$ between the direction of propagation and the magnetic field. Propagation toward the observer, on the "low field side" $\omega/\omega_{ci} \ge 1$ is guaranteed by the fact that $n^{2}$ is positive. (b) Plot of the absorption coefficient $\alpha(\omega,\theta) [cm^{-1}]$ of the fast wave vs $\omega/\omega_{ci} - 1$, for different values of the angle $\theta$ between the direction of propagation and the magnetic field. Calculation is done for a Hydrogen Maxwellian plasma with $B = 3.8~ 10^{11}~ G$, $n =10^{19}[cm^{-3}]$, $T_{i}(\beta=0^{\circ})= 9~10^{7}~^{\circ}K$ (where $T_{i}$ is the ion temperature) and it shows that optical thickness is obtained for plasma layer of the order of 1 cm at all propagation angles.} \end{figure} \subsection{The Electromagnetic Wave and its Absorption/Emission Properties.} If a medium emits radiation at a given frequency, it also absorbs it at the same frequency. The quantity $I(\omega)$, i.e. the radiated power per unit area per unit solid angle per unit angular frequency, is given by the solution of the radiative transfer equation. For a single propagating e.m. mode in a medium of refractive index $n$ this is written as: $$ n^{2} \frac{d}{ds} \Bigl( \frac{I(\omega)}{n^2} \Bigr) = j(\omega) - \alpha(\omega)I(\omega) \eqno (1) $$ where $j(\omega)$ and $\alpha(\omega)$ are the emissivity and the absorption coefficient respectively and $s$ is the axis along the radiation path. In the present problem only one normal e.m. mode can propagate in vacuo and be observed. The emissivity, $j(\omega)$, is computed from the plasma dispersion relation and the Kirchhoff's law: $$ j(\omega) = \alpha(\omega)K(\omega,T) \eqno (2) $$ where $K (\omega,T)$ is the Planck function at the temperature $T$, where $T$ is here the plasma temperature. Near the cyclotron frequency $\omega_{ci}$ ($2\pi~ \nu_{ci} = \omega_{ci} = ZeB/A c$; $Z$ is the atomic charge A the atomic mass), two independent e.m. modes can propagate in the magnetized plasma atmosphere. They are identified by an index of refraction given by the complex roots $n_{\pm}= n'+i~n''$ of the complex dispersion relation written for a real frequency $\omega \simeq \omega_{ci}$ in the biquadratic form (Akhiezer et al. 1975): $$ An^{4}_{\pm} + Bn^{2}_{\pm} + C = 0 \eqno (3) $$ \begin{figure} \centerline{\hbox{\psfig{figure=8267_fig3.ps,height=8cm,clip=}}} \caption{Main geometrical parameters used in our model: $\gamma$ is the angle between the magnetic and the rotation axis. $\beta$ defines the latitude of the polar caps.} \end{figure} \begin{figure} \centerline{\hbox{\psfig{figure=8267_fig4.ps,height=6cm,clip=}}} \caption{Apart from $\gamma$ and $\beta$, our model call for a third angle, $\alpha$, between the wave vector, pointing toward the observer, and the rotation axis. The $\gamma$ angle between the magnetic and the rotation axis is also shown.} \end{figure} The coefficients $A, B$ and $C$ are function of the full dielectric tensor and depend on the plasma frequency $\omega_{p}$, the ion cyclotron frequency $\omega_{ci}$, the ion temperature $T_i$ and the angle $\theta$ between the propagation wave vector $\vec k = \frac{\omega}{c} \vec n$ and the local magnetic field $\vec B$. Of these two waves, the slow $n_+$ (ordinary) wave diverges at the cyclotron resonance with a finite bandwidth (roughly given by the ratio of the thermal velocity to the phase velocity) of anomalous dispersion around the cyclotron frequency. The ordinary wave in the region where $\omega > \omega_{ci}$ has an evanescence gap, which even if bridging by thermal and collisional effects is considered, produces substantial attenuation of wave propagating in tenuous plasma toward the observer (Shafranov 1058, Akhiezer et. al. 1975). The Geminga plasma atmosphere is characterized by $\omega_{p}/\omega_{ci} < 1$. This prevents the use of the customary ion cyclotron approximation of the dielectric tensor and a numerical solution of the dispersion relation for both the ordinary and extraordinary mode has been performed . The numerical evaluation of absorption and propagation properties of the mode propagating in vacuo are shown in Fig. 2. as function of the angle between the direction of propagation and the magnetic field . The extraordinary mode, which can propagate in vacuo with frequency $\omega \ge \omega_{ci}$, is the so called magnetosonic, compressional Alfven or fast wave. It has a regular index of refraction and a very narrow band of frequency of anomalous dispersion with a correspondingly narrow frequency range of absorption due to thermal effects (Akhiezer et al, l975). The fast wave is electromagnetic in nature and in the cold plasma limit it is righthanded polarized. It becomes lefthanded (i.e. in the direction of the ion motion), thus allowing absorption at the fundamental harmonic, owing to small thermal effects or to the presence in the plasma of small traces of isotopes of the minority species. The Geminga plasma is characterized by $\omega_{p}/\omega_{ci} < 1$. This prevents the use of the customary ion cyclotron approximation of the dielectric tensor and requires a numerical solution of the dispersion relation. Assuming an appropriate temperature and density, even in a pure $H^+$ plasma the fast wave reaches a sufficiently large absorption coefficient $\alpha(\omega,\theta) = 2 ~(\omega/c)~ n'' (\omega,\theta)$ . The optical depth, $\tau$, can be estimated as $$ \tau(\omega,\theta) = \int_0^L {\alpha(\omega,\theta) ds'} \simeq \alpha(\omega,\theta) L \gg 1 ~ \eqno (4) $$ where $L$ is the plasma layer thickness, of the order of few cm. The refraction index of the fast wave and its absorption coefficient $\alpha$ are given in Fig. 2a,b for different $\theta$ angles, a plasma density of $10^{19}cm^{3}$ and a plasma temperature of $9~10^{7}~^{\circ}K$. It is worth noticing that, owing to the weak dependence of $\alpha$ on $\theta$ (see Fig.2b), a layer of thickness less than 1 cm is sufficient to match the blackbody emission, independently of the angle of propagation with respect to the magnetic field at any point on the star surface. We will use this condition to greatly simplify the computation of the intensity and of the broadening of the emitted line. The intensity, $i(\omega,\omega_{c})$, radiated from any point of the star surface, is then simply given by $$ i(\omega,\omega_{ci})= K(T,\omega) [1- e^{- \tau(\omega,\omega_{ci})}] \eqno (5) $$ A density limit of about $10^{19}$ particles $cm^{-3}$ represents an upper limit to the plasma density. Above such limit, bremsstrahlung, with its absorption coefficient $\alpha \simeq 10^{-2} (cm^{-1})$ (Bekefi 1966) in the range of temperature frequency of interest here, would introduce severe spectral distortions, not observed in the data. \begin{figure} \centerline{\hbox{\psfig{figure=8267_fig5.ps,height=6cm,clip=}}} \caption{Radiation flux ($erg~cm^{2} s^{-1} Hz^{-1}$) versus frequency (Hz) in the region near the cyclotron resonance frequency. The data of Fig.1 (close symbols) are compared with the the ion cyclotron emission computed on the basis of our model (solid line). In order to make the curve immediately comparable to the data points, we have filtered the curve through the passbands actually used in the photometric observations. The resulting fluxes are shown as open symbols.} \end{figure} Blackbody emission conditions could be met in a lower density region of the outer part of the atmosphere, with a slightly larger plasma thickness. This, however, would not change the main issue of the present model: the observed Geminga spectral feature comes from a plasma which belongs to the star atmosphere. One may alternatively think that the cyclotron line was emitted by the low density matter surrounding the star. This would imply that what is observed depends upon a process of photon collection over a wide plasma volume and, consequently, over a wide range of cyclotron frequencies. However, this is contradicted by the comparatively well defined peak now observed. \subsection{Emission Line Broadening Mechanism.} The real part of $n^{2}(\omega,\theta)$ shown in Fig.2 exhibits a very narrow frequency band of anomalous dispersion. This indicates that the width of the observed emission line cannot be explained in terms of the thermal Doppler broadening ($\Delta \omega \simeq \omega | n cos \theta | \frac{v_{thi}}{c}$, where $v_{thi}$ is the ion thermal velocity). This is by far insufficient to fit the observations in the range of temperatures required to explain the intensity of the line. Another possible source of broadening are collisional effects. These, however, have been shown to be insufficient to account for the width of the cyclotron feature. The only mechanism left to account for such an observed width is the structure of a dipole-like magnetic field associated with the star. In the case of a perfect magnetic dipole located at the centre of the star, the intensity of the magnetic field increases by a factor two when moving on the star surface from the (magnetic) equator to the poles. As a consequence, the ion cyclotron frequency should also change by a factor two, if the emitting plasma were to cover the whole star surface. The shape of the emission line is then determined by the superposition of radiation emitted by regions of the star with different magnetic field values and possibly different plasma temperature. The intensity of the line in our model is obtained as the integral of the emission from each point of the star surface: $$ I(\omega) = \frac{1}{D^2} \int_\Sigma i(\omega,\omega_{ci}) \frac{\vec n \bullet \vec k}{|\vec k|} d^{2}\Sigma \eqno (6) $$ where $D$ is the distance of the star from the observer and the integral is extended to the whole surface, $\Sigma$, taking into account the usual geometry effects ($\vec n$ represents the unit vector perpendicular to $d^{2}\Sigma$, while $\vec k$ is the propagation vector pointing toward the observer). The non-homogeneous emission also depends on the relative position of the observer with respect to the star rotation axis and, to properly account for the shape and the intensity of the feature, on the angle $\gamma$ between the rotation and magnetic axis. The computed spectrum, of course, will be averaged over the star rotation period. These model assumptions also allow us to determine the modulation depth expected for the observed spectral feature, yielding a clear observational test. \section{Model and Interpretation.} In the context of the model described above, the intensity of the emission line depends upon 1) temperature of the cyclotron-emitting plasma 2) emitting fraction of the star surface. The range of the dipole field spanned by the emitting plasma surface determines the frequency width of the observed feature. Clearly, the ion gyrofrequency formula allows us to determine the value of the star' s magnetic field as a function of the frequency of the observed line. Since the ratio A/Z (which defines the chemical composition and the ionization level of the emitting atmosphere) is unknown, the magnetic field value is, in principle, determined only within a factor of $\sim$ 2. On the other hand, our model yields a clear prediction of a sharp intensity decrease at the frequency value corresponding to the B field maximum, located close to the magnetic poles. To compute a B field value, the emitting medium is assumed to be either H or He, yielding a well-defined A/Z ratio. This is consistent with the strong stratification of the elements induced by the huge gravitational field of the neutron star. Hydrogen and Helium differ by a factor two in their ratio A/Z, so that the cyclotron second harmonic of the heavier element overlaps exactly the fundamental one of the lighter. The position of the polar caps with respect to the star magnetic axis is shown in Fig.3. The polar cap extension, given by the angle $\beta$, determines the frequency width of the cyclotron emission. The magnetic axis forms an angle $\gamma$ with the rotation axis. The model will be fully determined once the angle, $\alpha$, between the observer and the rotation axis, is defined (see Fig.4). Fig.5 compares the data (filled circles) to the cyclotron emission model (open circles). The thin continous line shows the detailed shape of the feature as determined by the model. The open circles are obtained by integrating the model data over the filter passbands. The computed magnetic field ranges from $3.8~ 10^{11}~ G$ for the case of a pure Hydrogen plasma to $7.6~ 10^{11} G$ in the case of Helium. The model data shown in Fig. 5 have been obtained with the following assumptions: magnetic pole plasma temperature $T_{0} = 9~ 10^{7}~^{\circ}K$; temperature profile along the polar caps assumed to be gaussian-like $T = T_{0} exp[-(\beta/\beta_{0})^{4}]$ with $\beta_{0} =57^{\circ}$. With this choice of parameters the plasma temperature drops to 1/10 of its maximum value in about $60^{\circ} $. Such an extension of the plasma polar caps is required to explain the width of the line. Of course, the observed line intensity and the plasma temperature are also determined by Geminga's radius and by its distance from the observer. We assumed a radius $r_{0} =10~km$ and the parallax distance of 157 pc. Reasonable geometry uncertainties, however, do not change the order of magnitude of the plasma temperature required to emit such an intense cyclotron line. Fig.6 gives a prediction of our model under the assumption of the oblique rotation geometry ($\gamma = 90^{\circ}$) and for $\alpha$ close to $20^{\circ}$, both the line intensities and profiles are seen to vary with the rotation phase. For this choice of parameters the modulation factor is about 15\% and the profile is seen to sharpen close to the emission maximum (i.e. when the $\vec B$ axis sweeps over the observer direction). Obviously, the region around $5500 \AA$ should be the ideal one for observing the feature modulation and profile. \begin{figure} \centerline{\hbox{\psfig{figure=8267_fig6.ps,height=6cm,clip=}}} \caption{Prediction of the model under the assumption of the oblique rotation geometry ($\gamma=90$) and for $\alpha$ close to $20^{\circ}$. Both the line intensities (triangles and circles refer to the maximum and minimum emission respectively) and profiles (solid and dotted lines) are seen to vary with the rotation phase.} \end{figure} \section{Electron Cyclotron Emission and Absorption.} Geminga's polar cap atmosphere has been modelled as a fully ionized gas with a density $\simeq 10^{19} cm^{-3}$ and temperature $\simeq 9~10^{7}~^{\circ}K$. The associated electron-ion energy equipartion time, under these model assumptions, keeps ions and electrons close in temperature. The electron cyclotron emission process is thus quite efficient and the electron cyclotron line, associated to the companion ion cyclotron line, is expected to be an observable feature. This is in contradiction with the experimental data (see the combined ROSAT/ASCA X-ray spectrum) that show no line feature at E=4 or 8 keV i.e. in the X-rays spectral range where the electron spectral line should fall. A possible explanation based on the observation that, within a blackbody emission approximation, the power emitted at the electron cyclotron frequency is so high that the rate of relaxation between parallel and perpendicular (to the magnetic field) temperature is not sufficient to keep the electron distribution function isotropic (Ichimaru 1973; Trubnikov 1965). The anisotropy of the electron distribution function, following this line of thought, would then reach a steady state when the cyclotron emitted power, which is proportional to the perpendicular temperature (Bornatici et al. 1983), is lower than the one predicted by the isotropic case. In order to give a quantitative estimate of the steady state anisotropy and emission a kinetic calculation is required by means of a relativistic Fokker-Planck equation which includes a quasilinear term of interaction with electron cyclotron radiation. This is, at present, beyond the target of this work. A possible explanation of the fact that the residual emission is not observed is given by the polar cap heating model proposed for Geminga by Halpern \& Ruderman (1993). Following this model a flux of $e^{+} e^{-}$ pairs is created on closed field lines lying outside the star and channelled into the polar caps of Geminga with a residual energy of about 6.5 erg each (Halpern \& Ruderman 1993). The $e^{\pm}$ cloud embedded in the dipole magnetic field, which decreases by increasing the distance from the star surface, can thus act as a "second harmonic" resonant absorber of cyclotron radiation emitted from regions closer to the star surface. The resonant radial position is located at $r_{2nd}=2^{1/3}r_{0}$, and, owing to the electron cyclotron line width $\Delta \omega/\omega_{ce} = v_{the}/c$ (where $\omega_{ce}$ is the angular frequency of the electron cyclotron emission and $v_{the}$ the thermal velocity of the electrons) extends over about 300 m. It can be shown (Bornatici et al. 1983) that the E=4 keV line is efficiently absorbed ("optical depth" $\tau \simeq 16$) by the $e^{\pm}$ cloud, providing that ($n_{\pm}T_{\pm}$) $\simeq 100$ where $n_{\pm}$ and $T_{\pm}$ are density and temperature of $e^{\pm}$ respectively ($n_{\pm}$ in $10^{15}$ units and $T_{\pm}$ in keV). Typical densities $n_{\pm} \simeq 10^{17} cm^{-3}$, corresponding to column densities $\simeq 10^{22}cm^{-2}$, with a temperature of $\simeq 2~ 10^{6}~^{\circ}K$ (0.2 keV), for instance, will attenuate the electron cyclotron line by a factor $\simeq 10^{7}$. The electron cyclotron emission, following these model assumptions, can no longer be observed as a spectral feature. \section{Conclusions.} The data shown in Fig.1 leave no room for doubt that a wide emission feature exists in the optical region of Geminga's thermal continuum. The feature falls in the wavelength region where the atmospheric ion-cyclotron emission will be located close to the surface of a magnetic neutron star. Since Geminga is a magnetic neutron star, to wit its periodic $\gamma$-ray emission, and most probably has an atmosphere, to wit its soft X-ray emission, we have provided here a semi-quantitative interpretation for such feature. It is based on the reasonable assumption that the polar cap regions of the NS are covered by a thin plasma layer heated to a temperature higher that the global surface atmosphere by, e.g., infalling particles. This is not a new scenario per se. It was foreseen both in the case of INS accretion of ionized matter funnelled towards the poles by the B-field configuration and of magnetospheric particles drawn back to the polar surface by the strong E field induced by the oblique rotator. \\ The plausibility of this emission model in the visible range frequencies is also supported by estimate of power balance performed along the line proposed by Halpern \& Ruderman. In the case of Geminga a pair flux in excess of $\dot N = 10^{38} s^{-1}$ can release in the emitting plasma layer a linear power density $\dot N dE/dr \simeq 10^{28} erg~cm^{-1} s^{-1}$ (Jackson 1975). This power is sufficient to compensate plasma losses mainly due to the ion cyclotron emission and Bremsstrahlung over the whole star surface.\\ What is new here is the excellent fit obtained to the multiple experimental data by our physical model using a minimum of assumption. In particular, we have shown that the feature could not originate over the whole star surface, because global B-field variations would induce a feature wider than observed. The only free parameter is the geometry of the emission with respect to the observer; note, however, that our geometry is fully compatible with the oblique rotator proposed for Geminga by Halpern \& Ruderman (1993). The assumption that the composition of the outer emitting layer is either H or a light, fully ionized element mixture is supported by the estimated value for the magnetic field. Such a value is in good agreement with the standard pulsar magnetic field prediction. It represents, in fact, the first independent measurement of the surface magnetic field of an INS. \begin{acknowledgements} Useful discussions with Prof. Bruno Bertotti are gratefully acknowledged. \end{acknowledgements}
\section{Introduction} Upon deriving black hole radiance, Hawking\cite{Hawk} found that $\om$, the frequency of the in-modes involved in the Bogoliubov coefficients, grows exponentially according to \be \om = \la \; e^{\kappa (u - u_0)} \label{dopef} \ee where $\kappa$ is the surface gravity of the hole and where $u$ is the retarded time around which the out-particle of energy $\la$ is centered. This point was emphasized by Gerlach\cite{gerlach} (and subsequently in \cite{bp}) who showed that the constant emission rate arises from a steady conversion of vacuum configurations of frequency $\om$ into red-shifted on shell photons of energy $\la$. This observation questions the validity of the settings used to derive Hawking radiation, namely free field propagation in a given geometry. Indeed, knowing that gravitational interactions grow with the energy, what is the validity of describing these high frequency configurations by free field theory\cite{THooft,unruh,Verl3,MP2}. This question becomes particularly important when one considers the interactions between the collapsing matter and Hawking quanta. In this respect, what is well understood is that if one first solves Einstein's equations to determine the collapsing metric and then study free field propagation in this geometry, one obtains Hawking radiation through the frequency mixing described in eq. (\ref{dopef}). In this derivation, the motion (and the quantum state) of infalling matter is unaffected by the emission of Hawking quanta. Indeed, the configurations giving rise to these quanta freely propagate through the collapsing matter\cite{Hawk,MP2}. The aim of this article is to provide a more dynamical description of the interactions between infalling matter and the radiation field $\phi$. To this end we analyze a simple model composed of $\psi$, a field of mass $m$, coupled to $\phi$ by current-current interactions and propagating in Schwarzschild geometry. Thus the quanta of $\psi$ represent additional particles falling into an already formed black hole. Since the interactions are described by Feynman diagrams, the infalling particles are now properly scattered according to energy-momentum conservation. This is crucial since it reveals the dynamical role played by $\om$, the `transplanckian' energy of the `partner' of an asymptotic quantum of energy $\la$. Indeed, we now obtain two different regimes. First, as long as $\om$ is much smaller than $m$, the mass of the infalling dust particle, the S-matrix elements of our model identically reproduces the Bogoliubov coefficients obtained by attributing a given inertial trajectory to the infalling particle. In this regime one thus recovers the infalling mirror description of Hawking radiation\cite{Grov,Carl,Wilc,CV,TV} with one important improvement: the residual energy crossing the future horizon is equal to the energy of the infalling particle minus the energy carried away by the Hawking quanta. Therefore, one does not need to appeal to Einstein's equations in order to obtain the notion of black hole evaporation. The second regime occurs when $\om$ becomes comparable to $m$. Then the scattering amplitudes no longer agree with those found by Hawking. In fact, because of energy conservation, we shall see that the production of thermal photons (induced by scattering on the infalling particle) stops when $\om$ reaches $m$. With this result, we reach the heart of the problem: how to obtain a steady conversion of vacuum configurations giving rise to a constant thermal flux once recoil effects are no longer neglected. The questions raised by our negative result are addressed at the end of the letter. It should also be mentioned that other derivations of Hawking radiation do not confront the transplanckian problem in those radical terms. First, superspring theory succeeded in deriving Hawking radiation in completely different settings since it results from the degeneracy of black hole microstates\cite{calmad}. Moreover, the derivation is performed in flat space. Thus one confronts neither exponentially growing Doppler effects, which are the all mark of regular horizons, nor therefore the high frequencies. Secondly, phenomenological models based on analogies with condensed matter physics have been proposed in order to question the relevance of these high frequencies\cite{dumbUnruh,dumbus}. At present however, one does not know how to justify dynamically these rather ad hoc models. Perhaps the present approach can provide the roots for such a justification. \section{The Model} In this letter, for simplicity, we shall consider only radial motion. Therefore, we can work with 2 dimensional fields. It is then convenient to use the conformally flat coordinates $t, z$ in which the 2D line element reads \ba ds^2 &=& (1 - {2M \over r}) \left( - dt^2 + dz^{2} \right) \nonumber\\ &=& (1 - {2M \over r})\; \eta_{\mu \nu} dx^\mu dx^\nu \label{conffl} \ea where $\eta^{\mu \nu}$ is the 2D Minkowski unit matrix and $z$ the tortoise coordinate ($= r + 2M \ln(r/2M -1)$). In this coordinate system, the action of the system is \ba S &=& \int \!dtdz \left( - \eta^{\mu \nu} \partial_\mu \psi^* \partial_\nu \psi -(1 - {2M \over r}) m^2\vert \psi \vert^2 \right) \nonumber \\ &&+ \int \!dtdz \left( \ - \eta^{\mu \nu} \partial_\mu \phi^* \partial_\nu \phi - g \eta^{\mu \nu} J^\psi_\mu J^\phi_\nu \right) \label{stot} \ea where $J^\psi_\mu = \psi^*\! \lr{i\partial_\mu} \psi $, $J^\phi_\nu =\phi^*\! \lr{i\partial_\nu} \phi $ are the currents carried by the complex fields $\psi$ and $\phi$. $g$ is the coupling constant. We have chosen to work with two complex fields in order to have well defined current operators before splitting positive and negative frequencies, see \cite{recmir} for more details concerning this model. In the absence of interactions ($g=0$), the modes of the fields freely propagate in Schwarzschild geometry. Since the metric is static, they can be labeled by their constant energy. The massless modes describing infalling and outgoing photons are respectively \ba \phi_\om &=& {e^{- i \om (t + z)} \over \sqrt{4 \pi \om}} = {e^{- i \om_\mu x^\mu} \over \sqrt{4 \pi \om}} \nonumber\\ \phi_\la &=& {e^{- i \la (t - z)} \over \sqrt{4 \pi \la}} = {e^{- i \la_\mu x^\mu} \over \sqrt{4 \pi \la}} \label{inout} \ea In the WKB approximation, the infalling mode of the massive field with energy $\e$ is \be \psi_\e= {e^{- i (\e t + \int^{z} \!dz' p_\e) } \over \sqrt{4 \pi p_\e(z)}} \label{inf} \ee where $p_\e(z)$ is the classical momentum, the positive solution of the mass-shell condition \be \eta^{\mu \nu} p_\mu p_\nu = -\e^2 + p^2_\e = - m^2 (1 - 2M/r) \label{psquare} \ee For $m \gg \kappa$ the WKB approximation is valid if one does not approach the turning point $p_\e(z) =0$ which exists when $\e < m$. In the interacting picture, to first order in $g$, the transition amplitudes are given by the matrix elements of $g\int\!dtdz J^\psi_\mu J_{\phi}^{\mu}$. Denoting $A(\e; \om, \la)$ the amplitude for an infalling photon of energy $\om$ to be scattered by a dust particle of energy $\e$ and converted into an outgoing photon of energy $\la$, one gets \ba A(\e; \om, \la) &=& -i g \int\!dz \left( p_\mu(\e) + p_\mu(\e + \om - \la) \right) \left( \om^\mu + \la^\mu \right) \nonumber\\ && \quad \times {e^{-i z(\la + \om)} \over 4 \pi \sqrt{\la \om}} { e^{-i \int^{z}\! dz' (p_\e -p_{\e + \om - \la})} \over 2 \sqrt{ p_\e \; p_{\e + \om - \la}}} \label{A} \ea where the final energy of the particle is $\e+\om -\la$ since energy conservation is implemented by the integration over $t$. The prefactor in the first line of eq. (\ref{A}) comes from the matrix elements of the two current operators. By crossing symmetry, the amplitude to spontaneously produce these two photons of energy $\om$ and $\la$ from scattering by an infalling particle of energy $\e$ is given by \be B(\e; \om, \la) = A(\e; -\om, \la) \label{B} \ee As we shall see, it is through these Bremsstrahlung-like amplitudes that Hawking radiation will be recovered in the low energy regime. \section{Low Energy Regime} In this regime, i.e. for $\om, \la \ll m$, one can develop the phase and the prefactor of the integrand of $A(\e; \om, \la)$ in powers of $\om$ and $\la$. To first order, we get \ba A(\e; \om, \la) &\simeq& -i g \int\!dz \left[ (\om + \la) dt_{cl} /dz - (\om - \la ) \right] \nonumber\\ &&\quad \times {e^{-i z (\la + \om)} e^{-i t_{cl}(z) (\om - \la)} \over 4 \pi \sqrt{\la \om} } \label{A2} \ea where \be t_{cl}(z) = - \partial_\e \int^{z} \!dz' p_\e(z') \label{hj} \ee is the time lapse evaluated along the infalling particle trajectory. We also used $\e/p_\e(z)=dt_{cl}/dz$. To first order in $\om, \la$, we find that $A(\e; \om, \la)$ is proportional to $\alpha(\om, \la)$, the overlap of the photons wave functions evaluated along the classical trajectory of the infalling particle characterized by $\e$ and parametrized by $z$ through $t_{cl}(z)$. This is easily verified by direct computation. Similarly, one finds that the amplitude $B(\e; \om, \la)$ is proportional, with the {\it same} factor, to $\beta(\om, \la)= \alpha(-\om, \la)$, the Bogoliubov coefficient encoding pair creation. It should be stressed that these agreements to first order in the energy-momentum transfers, here given by $\om$ and $\la$, are generic in character, see \cite{bfa}. We now focus on the late time regime, for $r/2M-1 \simeq e^{2 \kappa z} \ll 1$. In this case one has\cite{gerlach,bp} \be {dt_{cl} \over dz} = - 1 - e^{2\kappa (z - z_0) } \label{traj} \ee Then upon applying the stationary phase condition to the integrand of eq. (\ref{A2}) one recovers eq. (\ref{dopef}) since in the late time regime $z = - u/2 + constant$ along the infalling trajectory. This confirms that \be \vert{ B(\e; \om, \la) \over A(\e; \om, \la) }\vert^2 = \vert{ \beta(\om, \la) \over \alpha(\om, \la) }\vert^2 = e^{- 2\pi \la / \kappa} \label{bovera} \ee Eq. (\ref{bovera}) establishes the thermal character of the outgoing photons spontaneously emitted by the scattering on the infalling particle. As usual `spontaneously' means that one starts from vacuum configurations on ${\cal{I}}^-$. It is now appropriate to show how the notion of a constant rate occurs. To order $g^2$, the mean number of photons of energy $\la$ found on ${\cal{I}}^+$, is given by \be \langle n_\la \rangle = \int\!d\om \; \vert B(\e; \om, \la) \vert^2 \label{w} \ee To first order in $\om, \la$ it is proportional to $\int\! d\om /\om$, as in \cite{bp,GO}. To give meaning to this integral, one uses the fact that at time $u$, the $\la$ photons arise from frequencies $\om$ centered according to eq. (\ref{dopef}). This yields $d\om /\om = \kappa du$, i.e. that the production rate is constant. Similarly, the mean number of photons received before a certain time $u$, is obtained by integration over $\om$ up to $\la e^{\kappa(u - u_0)}$. Thus it increases linearly with $u - u_0$. This establishes that the thermal flux originates from a steady conversion of vacuum configurations into pairs of photons whose energy are related by eq. (\ref{dopef}). Thus, in this linear approximation in $\om, \la$, one recovers Hawking radiation as obtained from free field theory. Moreover, in our description based on matrix elements, the {\it residual} energy which crosses the future horizon is equal to $\e$ minus the energy of the outgoing $\la$ quanta since energy is conserved and since the $\om$ quanta fall into the hole and do not enter into this global energy balance. Therefore we obtain the notion of evaporation without having used Einstein's equations, but simply by having followed the standard rules of quantum field theory. The most remarquable feature of these results is the steadiness of the production rate. However, as we shall see, it is a consequence of having performed a first order expansion in $\om$. Moreover, because of eq. (\ref{dopef}), $\om$ reaches $m$ after a few e-folds in the units of $1/\kappa$. Therefore, since the low energy regime is brief, it is mandatory to take into account higher order effects in $\om/\e$. \section{High Energy Regime} To characterize the high energy regime we first analyze the classical channel, i.e. the scattering of an infalling photon of frequency $\om$ which is sent from ${\cal{I}}^-$. The simplest way to proceed consists in applying the stationary phase condition to the phase of the integrand of $A(\e; \om, \la)$. Then, the dominant contribution arises from values of $z$ centered around the saddle point value $z_{sp}$, the solution of \be \om + \la = p_{\e +\om -\la}(z_{sp}) - p_\e(z_{sp}) \label{stpc} \ee In the late time regime and for $\om \gg \la \simeq \kappa$, it obeys \be e^{2 \kappa z_{sp}} = { 4 \la \over \om} {\e^2 \over m^2} ( 1 + {\om \over \e} ) \label{stpc2} \ee As long as $\om /\e \ll 1$ one recovers the low energy regime since eq. (\ref{stpc2}) is equivalent to eq. (\ref{dopef}). Instead, when $\om$ aproaches $\e$, one reaches a new regime: the maximal value of $\la$, the energy of the scattered photon found on ${\cal{I}}^+$, is bounded by $m e^{-\kappa (u- u_0)}/4$ no matter how $\om$ big is. Thus one looses the illusionary possibility offered by eq. (\ref{dopef}), of receiving, on ${\cal{I}}^+$ and at arbitrarily large times, photons with a given frequency $\la$ by sending from ${\cal{I}}^-$ quanta with sufficiently high frequencies $\om$. In loosing this possibility, we recover conventional physics: a particle of mass $m$ cannot properly reflect photons of energy (measured in its rest frame) higher than its mass, see eq. (62) in \cite{recmir}. Then, because of the red shift from the scattering locus $z_{sp}$ to ${\cal{I}}^+$, one obtains the announced bound for $\la$. We believe that a similar conclusion should also emerge from taking into account the deformation of the geometry induced by a infalling wave of energy $\hbar \om$. In both cases, eq. (\ref{dopef}) would be invalidated after time lapses of the order of a few $1/\kappa$. Having clarified the direct channel, we now turn to spontaneous pair creation amplitudes. Due to crossing symmetry, the stationary phase conditions applied to $B(\e; \la, \om)$ are given by eqs. (\ref{stpc}) and (\ref{stpc2}) with $\om$ replaced by $-\om$. As for the amplitude $A$, the new regime arises when $\om$ approaches $\e$. To characterize the onset of this regime, one should determine the first order corrections in $\om/\e$. This is easily achieved by exploiting the following fact. Besides the exponential $e^{i(\om - \la ) z}$, the phase of the integrand of $B$ is an anti-symmetric function in $\om + \la$ when developed around the `mean' energy $\bar \e = \e /2 - (\e - \om -\la)/2$. Therefore, there is no quadratic terms in $\om$ when one develops around $\bar \e$. (Note in passing that this result directly follows from quantum mechanics which dictates that the wave functions of the `heavy' system enter transition amplitudes always in products of the form $\psi^*_{\e_{f\!i\!n}} \psi_{\e_{i\!n}}$, see \cite{bfa,area}.) Thus the improved value of $B(\e; \om, \la)$ is given by the first order expression with $\e$ replaced by $\bar \e$. From this one deduces the following. Firstly, the first order correction to Hawking temperature vanishes. Indeed, eq. (\ref{bovera}) still obtains since the value of $\e$ plays no role to first order in $\om$. Secondly, the emission rate at which these thermal quanta are emitted is modified. Indeed, upon computing their mean number, eq. (\ref{w}), one faces a new phase space volume. We remind the reader that when dealing with eq. (\ref{dopef}), one had $d\om/\om = \kappa du$ which guaranteed a constant flux. Instead, upon using eq. (\ref{stpc2}), one gets \be { d\om \over \om } \simeq \kappa du \left( 1 - { \om \over \e} \right) \label{newr} \ee which implies a decreasing rate when $\om/\e $ approaches 1. It is of little interest to further characterize this decrease because one encounters an unavoidable barrier: the final energy of the scattered particle ($=\e -\om -\la$) cannot be negative. Indeed, if one follows the conventional rules of QFT, wave functions with negative energy do not correspond to on shell particles and therefore do not appear in matrix elements at the tree approximation. Therefore, the high energy regime (whose unbounded character was at the origin of the constant rate in the usual derivation) is simply not available since the upper bound in the integral of eq. (\ref{w}) is given by $\e -\la$. Thus the steadiness of the emission rate is lost and the total number of quanta emitted is bounded. (Notice that a different result has been obtained in \cite{TV} from a somehow more kinematical model.) \section{Conclusions} Upon analyzing the scattering amplitudes $A(\e; \om, \la)$ and $B(\e; \om, \la)$ we found two regimes. As long as $\om/\e \ll 1$, it is legitimate to perform a first order expansion in $\om/\e$ and one recovers the usual Bogoliubov coefficients. In fact this agreement defines the physical interpretation of these coefficients: they are the correct amplitudes as long as backreaction effects can be neglected. This is why they only depend on $\om, \la$ and $\kappa$. When $\om$ approaches $\e$, after a few $1/\kappa$ time lapses, the thermal production stops. Indeed, because of energy conservation, one can no longer appeal to unbounded frequencies $\om$, as one does in the absence of backreaction. In this we have reached our aim: to show when how and why a simple approach based on QFT fails to reproduce Hawking radiation at large times. Needless to say that it is over hasty to deduce from this that Hawking radiation actually stops. Rather it poses with greater accuracy the question: which fundamental theory is Hawking's derivation an approximation of ? A first radical option consists in postulating that Hawking radiation {\it cannot} be recovered from QFT based on General Relativity once backreaction effects are included. Instead it should emerge from other settings such as string theory\cite{calmad}. Then, our negative result can be considered as an indication in favour of this belief. A less radical approach consists in hoping that one shall recover Hawking radiation by improving the present analysis. (Notice that this conservative approach does work for the Unruh effect\cite{Unr}. In that case, the backreaction effects obtained from QFT are bounded and the corrections to the Unruh effect stay finite\cite{rec}.) The improvements can be pursued in two directions. Firstly, by including scattering on many dust particles, one might get collective effects which significantly differ from what we obtained. Secondly, by including loop corrections one could recover access to the infinite reservoir of high frequencies and find that Hawking radiation does not emerge from on-shell diagrams. Indeed, one should explore the consequences of modifying Feynman rules for the external legs which correspond to quanta falling into the hole since there is no a priori reason to treat asymptotic and infalling quanta on the same footings. In this respect one can already notice that, due to the $r$-dependence in eq. (\ref{psquare}), the final energy of the scattered particle can be smaller than $m$ even though the corresponding particle cannot be found asymptotically.
\section{Introduction} \label{subsec:diffusion} The transport of a solute in an incompressible fluid mixture of \emph{density} $\varrho$ is commonly modeled by the convection-diffusion equation\footnote{$\,$ Here, the operations of material and spatial time-differentiation are denoted with a superposed dot and a superposed grave accent, respectively; thus, for instance, $\dot{c}=\grave{c}+\boldsymbol{v}\!\zcdot\!\xgc$, with $\boldsymbol{v}$ being the \emph{mixture velocity}.} \begin{equation} \dot{c} =\xd\bigl(D(c)\,\xgc\bigr), \label{eq:diffusion} \end{equation} with $c$ the \emph{solute mass fraction} and $D\ge0$ the mass fraction-dependent \emph{diffusivity}. This equation is often derived (see, for example, Landau \& Lifshitz \cite{Landau59}, de~Groot \& Mazur \cite{Degroot83}, and Cussler \cite{Cussler84}) by adjoining to the solute mass balance \begin{equation} \varrho\mskip1mu\dot{c}=-\xd\boldsymbol{\jmath}, \label{eq:mass} \end{equation} a constitutive relation \begin{equation} \boldsymbol{\jmath}=-\varrho\mskip1mu D(c)\, \xgc \label{eq:constitutive} \end{equation} of the sort proposed by Fick \cite{Fick55} for the \emph{diffusive solute flux} $\boldsymbol{\jmath}$. Here, we propose and systematically develop an alternative approach to solute transport. Our approach is based on the viewpoint that to each independent kinematic process there should correspond a system of power-conjugate forces subject to a distinct force balance.\footnote{A foundation for this viewpoint can be devised using the {\em principle of virtual power}, see, for example, Germain \cite{Germain}, Antman \& Osborn \cite{Antman}, Maugin \cite{Maugin}, and Fr\'emond \cite{Fremond}. We follow here an alternative, but essentially equivalent, approach taken by Ericksen \cite{E61}, Goodman \& Cowin \cite{GC72}, Capriz \& Podio-Guidugli \cite{CP-G}, Capriz \cite{C89}, Fried \& Gurtin \cite{FG}, Fried \cite{F}, and Gurtin \cite{G}.} Specifically, in the context of solute transport, \emph{we identify mixture convection and solute diffusion as independent kinematic processes}: mixture convection is a macroscopic phenomenon described kinematically by the convective mixture velocity $\boldsymbol{v}$, whereas solute diffusion is a microscopic phenomenon described kinematically at a continuum level by the diffusive solute flux $\boldsymbol{\jmath}$, an object that measures the motion of solute particles relative to the macroscopic motion of the mixture comprised of solute and solvent. Further, to account for power expenditures associated with the diffusive motion of atoms, we introduce at a continuum level a system of forces conjugate to the diffusive solute flux $\boldsymbol{\jmath}$. Since they are associated with the microscopic processes governing the diffusion of the solute, we use the expression ``microforce'' to identify these forces. Additionally, we require that these microforces be consistent with a \emph{solute microforce balance}. This balance is a postulate ancillary to the conventional momentum balance, which involves the forces that act power conjugate to the convective mixture velocity $\boldsymbol{v}$. Importantly, we do not assume that the diffusive solute flux $\boldsymbol{\jmath}$ is determined by a constitutive relation---Fickian or otherwise. Rather, \emph{we treat $\boldsymbol{\jmath}$ as an independent kinematic variable whose evolution is governed by the microforce balance}. There is ample historical precedence for basing solute transport on additional, independent force balances. Nernst \cite{Nernst88}, following upon van't Hoff's \cite{Hoff87} analogy between the \emph{osmotic pressure} $\varpi$ of the molecules of a sufficiently dilute solute and the pressure of an ideal gas, viewed diffusion as driven by osmotic pressure gradients and developed a theory of diffusion based on a balance of forces acting on the solute molecules. Nernst proposed such force balances to account for the influences of osmotic, viscous, electrostatic, gravitational, centrifugal, and magnetic forces. For example, for electrolyte solutions, Nernst assumed separate force balances for the positive and negative ions. Subsequently Planck \cite{Planck90} also exploited these ideas in his work on electrolyte solutions. Further, in developing his theory of Brownian motion, Einstein \cite{einstein05,einstein06,einstein08} employed force balances on the Brownian particles. Neglecting the inertia of the diffusing solute, we may write Nernst's solute force balance as \begin{equation} \boldsymbol{h}+{\boldsymbol{\ell}}=\frac{1}{n}\,\xg\varpi = \frac{M}{\varrho\mskip1muc}\,\xg\varpi , \label{eq:nernst} \end{equation} wherein $\boldsymbol{h}$ is the \emph{hydrodynamic drag force} acting on the solute particles during motion through the solvent, $\boldsymbol{\ell}$ represents the \emph{external forces} acting on the solute particles, $M$ is the \emph{molecular weight} of the solute, and $n$ is the number density---so that the right-hand side of (\ref{eq:nernst}) can be interpreted as the \emph{osmotic force} per solute particle. Typically the hydrodynamic drag force is taken to be linear in the solute velocity relative to the mixture, whereby \begin{equation} \boldsymbol{h}=-\zeta\mskip1mu\boldsymbol{\jmath}, \label{eq:frictional_force} \end{equation} with $\zeta$ a non-negative coefficient called the \emph{frictional resistance} and $\zeta^{-1}$ called the \emph{mobility}.\footnote{Since we take the solute flux $\boldsymbol{\jmath}$ as a generalized velocity, our definitions of frictional resistance and mobility differ from the usual ones by a factor of $\varrhoc$.} For an ideal solution, the \emph{osmotic pressure} $\varpi$ is linear in the number density $n$, that is \begin{equation} \varpi=\kBT n=\frac{\varrho\mskip1mu\kBTc}{M}, \label{eq:osmotic} \end{equation} with $k_{\rm B}$ being \emph{Boltzmann's constant} and $\theta$ being the \emph{absolute temperature}. Thus, in an isothermal, constant density mixture, \begin{equation} \xg \varpi={\kBT}\,\xg{n}=\frac{\varrho\mskip1mu{\kBT}}{M}\xgc, \label{eq:osmotic_grad} \end{equation} which, in conjunction with (\ref{eq:nernst}) leads to the flux \begin{equation} \boldsymbol{\jmath}=-\frac{\kBT}{c\zeta}\,\xgc+\frac{\boldsymbol{\ell}}{\zeta}. \end{equation} Hence, provided that external forces are absent, the balance (\ref{eq:nernst}) and constitutive relations (\ref{eq:frictional_force}) and (\ref{eq:osmotic_grad}) yield an expression similar to the Fickian constitutive equation (\ref{eq:constitutive}) with the classical relation (see Nernst \cite[Page 615]{Nernst88}) \begin{equation} D=\frac{\kBT}{\varrhoc\mskip1mu\zeta}, \label{eq:diffusion_coef} \end{equation} between diffusivity and frictional resistance (or mobility). There is, however, disagreement over the form and interpretation of the solute force balance. For example, Einstein \cite[\S~4]{einstein06b} asserted that the osmotic pressure is a ``virtual'' force, so that it cannot be treated as a force acting on the individual molecules.\footnote{The exact meaning of this statement is unclear to the authors.} In fact, Einstein \cite{einstein06}, in his original treatment of Brownian motion, proposed instead a fictitious force $\boldsymbol{k}$ that acted on the individual particles but that balanced the osmotic pressure gradient: \begin{equation}\boldsymbol{k}=\frac{M}{\varrho\mskip1muc}\,\xg \varpi. \end{equation} Einstein \cite{einstein06} then took the solute force balance as \begin{equation}\boldsymbol{h}+\boldsymbol{l}=\boldsymbol{k} \end{equation} and thereby claimed (see Einstein \cite{einstein06b}) to have resolved the apparent difficulty of having a virtual force act on the individual particles. Additionally, Debye \cite[Page~106]{Debye62} asserted that osmotic forces are not ``true'' mechanical forces, so that the solute force balance should be instead \begin{equation} \boldsymbol{h}+{\boldsymbol{\ell}}={\bf0}. \label{eq:debye} \end{equation} In this alternative point of view, the flux $\boldsymbol{\jmath}$ is decomposed as a sum \begin{equation} \boldsymbol{\jmath}=\boldsymbol{\jmath}_D+\boldsymbol{\jmath}_F \label{eq:debye_flux} \end{equation} involving a purely diffusive part $\boldsymbol{\jmath}_D$ (usually called the ``ordinary diffusion'') and a part $\boldsymbol{\jmath}_F$ due to external forces (usually called the ``forced diffusion''). Commonly, Fick's first law is invoked for the ordinary diffusive flux, so that \begin{equation} \boldsymbol{\jmath}_D=-D\xg c, \label{eq:debye_flux1.5} \end{equation} and the flux due to the external forces is assumed to have the form \begin{equation} \boldsymbol{\jmath}_F=\frac{\boldsymbol{\ell}}{\zeta}, \label{eq:debye_flux2} \end{equation} which follows from (\ref{eq:frictional_force}) and (\ref{eq:debye}). Combining (\ref{eq:debye_flux2}) with the balance (\ref{eq:mass}) leads to the \emph{generalized convection-diffusion equation} \begin{equation} \dot{c}=\xd \bigl(D\xgc-\frac{\boldsymbol{\ell}}{\varrho \zeta}\bigr), \label{eq:debye_mass} \end{equation} typically referred to as a Smoluchowski \cite{smoluchowski15} equation. The diffusivity $D$ can be determined by considering the special case where the external forces are given by a potential, viz., \begin{equation} \boldsymbol{\ell}=-\,\xg{V}, \label{eq:potential} \end{equation} and requiring the steady-state solution to correspond to that of a Boltzmann \cite{boltzmann72} distribution, so that \begin{equation} c=c_0\exp\big(\frac{-V}{\kBT}\bigr). \label{eq:boltzmann} \end{equation} This yields the relation (\ref{eq:diffusion_coef}), which is usually enforced even when (\ref{eq:boltzmann}) no longer holds. This latter format was used by Debye \cite{debye13,debye29} in his theory of rotary diffusion in dielectrics to include the effect of the external electric field on the diffusing dipoles. It is still commonly used to obtain extensions of the Fickian constitutive relation to include, for instance, the effects of external electric fields on electrolytes. Additionally it is used to model the interaction of diffusion and external fields on dilute suspensions (see, for example, Giesekus \cite{Giesekus58} and, more recently, Doi \& Edwards \cite{Doi86}). However, Bird, Armstrong, \& Hassager \cite{Bird77} based their approach to polymeric suspensions on a force balance such as (\ref{eq:nernst}) that explicitly includes a contribution attributed to the Brownian forces on the polymer molecule (see, for example, \cite[Equation~13.2-1]{Bird77}). Interestingly, in the case of linear constitutive relations, the approaches described above lead to the same resulting equations---namely (\ref{eq:debye_mass}) and (\ref{eq:diffusion_coef}). However, the assumptions involved are rather different. In particular, it is not clear how to generalize either approach to allow for constitutive nonlinearities. For example, in (\ref{eq:debye_flux}) the ordinary diffusive flux $\boldsymbol{\jmath}_D$ is determined by a constitutive relation, whereas the forced diffusive part $\boldsymbol{\jmath}_F$ is determined by a balance law---a procedure highly atypical for a continuum theory. Furthermore, the validity of (\ref{eq:diffusion_coef}), derived from an assumed steady-state solution, is not clear in unsteady situations. In our opinion, these issues can be resolved by a more general treatment of the notion of forces in a binary mixture. Consequently, in the following, we develop a theoretical framework for diffusion based on a \emph{generic} microforce balance without a priori assumptions on the nature of the microforces or of steady-state concentration profiles. Furthermore, we allow for constitutive nonlinearities and external fields. We follow the procedure of modern thermodynamics, wherein consistency with the second law restricts the form of the constitutive relations. Our proposed format provides a systematic procedure for the generalization of convection-diffusion equation models for solute transport. In particular, it leads to a force balance that explicitly includes the diffusive or Brownian forces acting on the solute. Only in special cases, such as linear constitutive relations, can this force balance be expressed in the oft used form (\ref{eq:debye}) with the corresponding decomposition (\ref{eq:debye_flux}) of the flux proposed by Debye. For clarity, we restrict our attention to purely mechanical processes in a binary mixture of solute and solvent, ignoring thermal and other effects. In this context, the first and second laws of thermodynamics are replaced by a \emph{free-energy imbalance} that accounts for power expenditures associated not only with mixture convection but also with solute diffusion. That diffusion does, indeed, lead to the expenditure of power was clearly recognized and utilized by van't Hoff \cite{Hoff87} as an experimental means of measuring the osmotic pressure. First, we consider simple diffusion without mixture convection. This special case allows us to develop the theory without the complications of coupling with the macroscopic mixture motion. We then treat both convection and diffusion. \section{Simple theory accounting only for diffusion} We consider a binary mixture of solute and solvent that occupies a fixed region $\zcR$ of three-dimensional space. In the present context, where we ignore the macroscopic motion of the mixture, the mixture density is constant. Without loss of generality, we take that constant to be unity. \subsection{Basic laws} \label{sect:balance} The simple theory is based upon the following basic laws: \begin{itemize} \item[$\bullet$] solute mass balance; \item[$\bullet$] solute microforce balance; \item[$\bullet$] mixture free-energy imbalance. \end{itemize} We formulate these laws in global form for an arbitrary subregion $\mathcal{P}$ of $\zcR$. The outward unit-normal field on the boundary $\partial\mathcal{P}$ of $\mathcal{P}$ is denoted by $\bfnu$. \subsubsection{Solute mass balance} \label{subsect:solute_mass} We introduce fields\footnote{Since the mixture density is assumed to be unity, there is no distinction between solute mass and solute mass fraction, or between solute mass flux and solute mass fraction flux.} \begin{center} \begin{tabular}{cl} $c$ & \emph{solute mass fraction}, \\ [0.5ex] $\boldsymbol{\jmath}$ & \emph{diffusive solute flux}, \\ [0.5ex] $m$ & \emph{external solute supply}, \\ [0.5ex] \end{tabular} \end{center} in which case the integrals \begin{equation} \int\limits_{\mathcal{P} c\:\textsl{dv}, \qquad\qquad \int\limits_{\partial\mathcal{P}}\boldsymbol{\jmath}\!\cdot\!\bfnu\:\textsl{da}, \qquad\text{and}\qquad \int\limits_{\mathcal{P}}m\:\textsl{dv} \end{equation} represent, respectively, the solute mass in $\mathcal{P}$, the solute mass added to $\mathcal{P}$, per unit time, by diffusion across $\partial\mathcal{P}$, and the solute mass added to $\mathcal{P}$, per unit time, by external agencies. \emph{Solute mass balance} is the postulate that, for each subregion $\mathcal{P}$ of $\zcR$ and each instant, \begin{equation} \grave{\overline{\int\limits_{\mathcal{P} c\:\textsl{dv}}} =-\int\limits_{\partial\mathcal{P}}\boldsymbol{\jmath}\!\cdot\!\bfnu\:\textsl{da} +\int\limits_{\mathcal{P}}m\:\textsl{dv}. \label{eq:solute_mass_balance} \end{equation} Equivalently, since $\mathcal{P}$ is arbitrary, we may enforce solute mass balance via the local field equation \begin{equation} \grave{c}=-\xd\boldsymbol{\jmath}+m. \label{eq:local_solute_mass_balance} \end{equation} \subsubsection{Solute microforce balance} \label{subsect:micromoment} Associated with the evolution of the diffusive solute flux $\boldsymbol{\jmath}$, we introduce a system of power-conjugate \emph{microforces}, consisting of \begin{center} \begin{tabular}{cl} $\zSigma$ & \emph{solute microstress tensor}, \\[0.5ex] $\boldsymbol{h}$ & \emph{internal solute body microforce}, \\[0.5ex] $\boldsymbol{\ell}$ & \emph{external solute body microforce}, \\[0.5ex] \end{tabular} \end{center} in which case the integrals \begin{equation} \int\limits_{\mathcal{P}}\zSigma\bfnu\:\textsl{da}, \qquad\qquad \int\limits_{\mathcal{P}} \boldsymbol{h}\:\textsl{dv}, \qquad\text{and}\qquad \int\limits_{\mathcal{P}}\boldsymbol{\ell}\:\textsl{dv} \end{equation} represent, respectively, the solute microforces exerted on the region $\mathcal{P}$ by the solute microtraction distributed over $\partial\mathcal{P}$, by agencies within $\mathcal{P}$, and by agencies external to $\mathcal{P}$, respectively. In the present mixture context, internal forces arise from the action of the solvent on the solute, or from solute-solute interactions. External forces are those arising from sources remote from the mixture, such as those arising from applied electromagnetic or gravitational fields. The microstress tensor $\zSigma$ allows for surface tractions that are not parallel to the surface normal. We assume that the inertia associated with solute diffusion is negligible and, hence, impose a \emph{solute microforce balance}, which is the postulate that, for each subregion $\mathcal{P}$ of $\zcR$ and each instant, the resultant microforce acting on $\mathcal{P}$ must vanish: \begin{equation} \int\limits_{\partial\mathcal{P}}\zSigma\bfnu\:\textsl{da} +\int\limits_{\mathcal{P}}(\boldsymbol{h}+\boldsymbol{\ell})\:\textsl{dv}={\bf0}. \label{eq:orientational_balance} \end{equation} Equivalently, since $\mathcal{P}$ is arbitrary, we may enforce solute microforce balance via the local field equation \begin{equation} \xd\zSigma+\boldsymbol{h}+\boldsymbol{\ell}={\bf0}. \label{eq:local_microforce_balance} \end{equation} A comparison to the Nernst force balance (\ref{eq:nernst}) suggests an interpretation of the quantity $\zSigma$ as a generalized stress tensor that allows for the possibility of tangential osmotic forces on a semipermeable membrane. At this point, we impose no constitutive assumptions concerning the microforces. In particular, we allow for the possibility that $\zSigma$ vanish identically, which would correspond to the alternative balance (\ref{eq:debye}) proposed by Debye. \subsubsection{Free-energy imbalance} \label{subsect:energy} In this purely mechanical theory, the first and second laws of thermodynamics are replaced by a \emph{free-energy imbalance}. To formulate this imbalance, we introduce fields \begin{center} \begin{tabular}{cl} $\psi$ & \emph{mixture free-energy density}, \\[0.5ex] $\mu$ & \emph{solute diffusion potential}, \\[0.5ex] \end{tabular} \end{center} in which case that the integrals \begin{equation} \int\limits_{\mathcal{P} \psi\:\textsl{dv} \qquad\text{and}\qquad \int\limits_{\mathcal{P}}\mu\mskip1mu{m}\:\textsl{dv} \end{equation} represent, respectively, the free-energy in $\mathcal{P}$ and the rate at which free-energy is added to $\mathcal{P}$ by the external supply of solute to $\mathcal{P}$. Further, the integrals \begin{equation} \int\limits_{\partial\mathcal{P}}\zSigma\bfnu\!\zcdot\!\boldsymbol{\jmath}\:\textsl{da} \qquad\text{and}\qquad \int\limits_{\mathcal{P}}\boldsymbol{\ell}\!\zcdot\!\boldsymbol{\jmath}\:\textsl{dv} \end{equation} provide an accounting of the power expended on $\mathcal{P}$ by the solute microtraction distributed over $\partial\mathcal{P}$ and by the agencies acting external to that region. These expressions represent the power expended by the diffusion of the solute particles. Consistent with our assumption that solute inertia is negligible, we neglect the kinetic energy of solute diffusion. \emph{Free-energy imbalance} is, then, the postulate that, for each subregion $\mathcal{P}$ of $\zcR$ and each instant: \begin{equation} {\grave{\overline{\int\limits_{\mathcal{P} \psi\:\textsl{dv}}}} \ \le\ \int\limits_{\partial\mathcal{P}}\zSigma\bfnu\!\zcdot\!\boldsymbol{\jmath}\:\textsl{da} +\int\limits_{\mathcal{P}}(\boldsymbol{\ell}\!\zcdot\!\boldsymbol{\jmath}+\mu\mskip1mu{m})\:\textsl{dv}. \label{eq:difference} \end{equation} Equivalently, since $\mathcal{P}$ is arbitrary, we may enforce free-energy imbalance via the local field inequality \begin{equation} \grave{\psi} \mu\mskip1mu\grave{c} -\boldsymbol{h}\!\zcdot\!\boldsymbol{\jmath}+(\zSigma+\mu\idem)\!\zcdot\!\xg\boldsymbol{\jmath}\ge0. \label{eq:local_strong_energy_imbalance_2} \end{equation} \subsection{Constitutive equations} \label{sect:const} The preceding balance laws do not form a determinate system of equations. We now develop constitutive equations that will provide closure relations. We assume that the free-energy density $\psi$, the solute diffusion potential $\mu$, the solute microstress tensor $\zSigma$, and the internal solute body microforce $\boldsymbol{h}$ are determined constitutively as functions of the solute mass fraction $c$ and the diffusive solute flux $\boldsymbol{\jmath}$, so that \begin{equation} (\psi,\mskip2mu\mu,\mskip2mu\zSigma,\mskip2mu\boldsymbol{h}) =\bigl(\hat{\psi}(c,\boldsymbol{\jmath}),\mskip2mu\hat{\mu}(c,\boldsymbol{\jmath}), \mskip2mu\phat{\zSigma}(c,\boldsymbol{\jmath}),\mskip2mu\hat{\boldsymbol{h}}(c,\boldsymbol{\jmath})\bigr). \label{eq:cr} \end{equation} We emphasize that, in contrast to the standard approach to the theory of mass diffusion, the diffusive flux $\boldsymbol{\jmath}$ is not given by a constitutive relation. Inserting the constitutive relations (\ref{eq:cr}) into the local free-energy imbalance (\ref{eq:local_strong_energy_imbalance_2}), we arrive at the functional inequality \begin{multline} \bigl(\frac{\partial\hat{\psi}}{\partialc}(c,\boldsymbol{\jmath}) -\hat{\mu}(c,\boldsymbol{\jmath})\bigr)\grave{c} \frac{\partial\hat{\psi}}{\partial\boldsymbol{\jmath}} (c,\boldsymbol{\jmath})\!\zcdot\!\skew3\grave{\boldsymbol{\jmath}} +\hat{\boldsymbol{h}}(c,\boldsymbol{\jmath})\!\zcdot\!\boldsymbol{\jmath} -\bigl(\phat{\zSigma}(c,\boldsymbol{\jmath})+\hat{\mu}(c,\boldsymbol{\jmath})\idem\bigr) \!\zcdot\!\xg\boldsymbol{\jmath} \le0, \label{eq:dissipation_imbalance_with_constitutive_relations} \end{multline} and, following the procedure founded by Coleman \& Noll \cite{CN} in their incorporation of the second law into continuum thermomechanics, conclude that: \begin{itemize} \item[({\em i\/})] the constitutive response functions $\hat{\psi}$, $\phat{\mu}$, and $\phat{\zSigma}$ delivering the free-energy density $\psi$, the diffusion potential $\mu$, and the stress $\zSigma$ must be independent of the diffusive mass flux $\boldsymbol{\jmath}$, with\footnote{Here, a prime indicates with respect to the argument of a function depending on a scalar field. It is noteworthy that the solute microstress reduces to an isotropic tensor because we do not include the gradient $\xg\boldsymbol{\jmath}$ of the diffusive solute flux $\boldsymbol{\jmath}$ in the list of independent constitutive variables.} \begin{equation} \psi=\hat{\psi}(c), \qquad \mu=\hat{\psi}{}^{\prime}(c), \qquad\text{and}\qquad \zSigma=-\hat{\psi}{}^{\prime}(c)\idem; \label{eq:basic_model_restrictions} \end{equation} \item[({\em ii\/})] the constitutive response function $\hat{\boldsymbol{h}}$ for the internal solute body microforce $\boldsymbol{h}$ must be consistent with the \emph{residual inequality} \begin{equation} \hat{\boldsymbol{h}}(c,\boldsymbol{\jmath})\!\zcdot\!\boldsymbol{\jmath}\le0. \label{eq:residual_inequality} \end{equation} \end{itemize} Granted smoothness of the response function $\hat{\boldsymbol{h}}$, a result appearing in Gurtin \& Voorhees \cite{gurtin93} yields the general solution \begin{equation} \hat{\boldsymbol{h}}(c,\boldsymbol{\jmath})=-\boldsymbol{Z}(c,\boldsymbol{\jmath})\mskip1mu\boldsymbol{\jmath} \label{eq:solution_of_residual_inequality} \end{equation} of (\ref{eq:residual_inequality}), where the symmetric \emph{generalized frictional resistance tensor} $\boldsymbol{Z}$ must obey \begin{equation} \boldsymbol{\jmath}\!\zcdot\!\boldsymbol{Z}(c,\boldsymbol{\jmath})\mskip1mu\boldsymbol{\jmath}\ge0. \end{equation} Objectivity further requires that under the transformation \begin{equation} c\mapsto c, \qquad \boldsymbol{\jmath}\mapsto\boldsymbol{Q}\boldsymbol{\jmath}, \end{equation} with $\boldsymbol{Q}$ an arbitrary rotation, the quantities $\psi$, $\mu$, $\zSigma$, $\boldsymbol{h}$, and $\boldsymbol{\ell}$ transform as \begin{equation} \psi\mapsto\psi, \qquad \mu\mapsto\mu, \qquad \zSigma\mapsto\boldsymbol{Q}\zSigma\boldsymbol{Q}^{\trans}, \qquad \boldsymbol{h}\mapsto\boldsymbol{Q}\boldsymbol{h}, \qquad\boldsymbol{\ell}\mapsto\boldsymbol{Q}\boldsymbol{\ell}, \end{equation} which, in particular, implies that \begin{equation} \boldsymbol{Z}(c,\boldsymbol{\jmath})=\zeta(c,\jmath)\idem, \qquad\jmath=|\boldsymbol{\jmath}|, \label{eq:Z=zeta_idem} \end{equation} where the \emph{scalar frictional resistance} $\zeta$ must obey \begin{equation} \zeta(c,\jmath)\ge0 \end{equation} for all $(c,\jmath)$. Combining (\ref{eq:solution_of_residual_inequality}) and (\ref{eq:Z=zeta_idem}), we see that \begin{equation} \boldsymbol{h}=-\zeta(c,\jmath)\mskip1mu\boldsymbol{\jmath}, \end{equation} so that the internal solute body microforce $\boldsymbol{h}$ must be collinear with the diffusive solute flux $\boldsymbol{\jmath}$. The foregoing results show that the behavior of a medium of the sort considered here is completely determined by the provision of two constitutive response functions: (\emph{i}) $\hat{\psi}$ determining the mixture free-energy density as a function of the solute mass fraction $c$; and, (\emph{ii}) the scalar frictional resistance $\zeta$, which determines the internal solute body microforce $\boldsymbol{h}$ and, in general, may depend on both the solute mass fraction $c$ and the magnitude $\jmath$ of the diffusive solute flux $\boldsymbol{\jmath}$. The governing equations that arise on substituting the foregoing thermodynamically consistent constitutive equations in the local field equations (\ref{eq:local_solute_mass_balance}) and (\ref{eq:local_microforce_balance}) expressing solute mass balance and solute microforce balance are \begin{equation} \left\delimiter0 \begin{split} \grave{c}+\xd\boldsymbol{\jmath}&=m, \\[4pt] \zeta(c,\jmath)\mskip1mu\boldsymbol{\jmath} +\xg\bigl(\hat{\psi}{}^{\prime}(c)\bigr)&=\boldsymbol{\ell}. \quad \end{split} \right\} \label{eq:basic_model_final_equations} \end{equation} Provided that $\zeta$ is nonvanishing, (\ref{eq:basic_model_final_equations})$_2$ yields an expression \begin{equation} \boldsymbol{\jmath}= -\frac{\hat{\psi}{}^{\prime\prime}(c)}{\zeta(c,\jmath)}\, \xgc +\frac{\boldsymbol{\ell}}{\zeta(c,\jmath)}, \label{eq:expression_for_flux} \end{equation} for the diffusive solute flux $\boldsymbol{\jmath}$. This leads, in conjunction with (\ref{eq:basic_model_final_equations})$_1$, to the \emph{generalized Smoluchowski equation} \begin{equation} \grave{c}=\xd\big( \frac{\hat{\psi}{}^{\prime\prime}(c)}{\zeta(c,\jmath)}\,\xgc -\frac{\boldsymbol{\ell}}{\zeta(c,\jmath)}\big). \label{eq:smoluchowski} \end{equation} Note that, when the frictional resistance $\zeta$ depends on $\jmath$, the evolution equation (\ref{eq:smoluchowski}) \emph{does not} decouple from the microforce balance (\ref{eq:basic_model_final_equations})$_2$. In this case, the flux $\boldsymbol{\jmath}$ does not separate into distinct ordinary and forced diffusive contributions. Comparing (\ref{eq:expression_for_flux}) with the conventional relation (\ref{eq:constitutive}), we obtain a generalized expression \begin{equation} D(c,\jmath)=\frac{\hat{\psi}{}^{\prime\prime}(c)}{\zeta(c,\jmath)} \label{eq:generalized_stokes_einstein} \end{equation} for the diffusivity, which is not predicated on a particular steady-state solution and, moreover, is not restricted to the linear regime. Hence, within our framework, the Fickian constitutive relation (\ref{eq:constitutive}) can be interpreted as an expression of the solute microforce balance, granted that the internal body microforce is linear in the diffusive solute flux and that the external solute body microforce vanishes. In the special case governed by (\ref{eq:potential}), the steady-state solution is \begin{equation} \mu=\hat{\psi}^\prime(c)=-V+\text{constant}, \label{eq:solution} \end{equation} which need not correspond to the Boltzmann distribution (\ref{eq:boltzmann}). Our results for the microforce balance are consistent with the Nernst balance (\ref{eq:nernst}). In fact, a comparison of (\ref{eq:nernst}) to (\ref{eq:basic_model_final_equations})$_2$ shows that they are identical in form if \begin{equation} \varpi^\prime(c)=\frac{c}{M}\hat{\psi}^{\prime\prime}(c) =\frac{c}{M}\hat{\mu}^\prime(c). \label{eq:osmotic_eq} \end{equation} Thus, (\ref{eq:basic_model_restrictions}) shows that the osmotic traction acts only normal to a semipermeable membrane and is determined by the mixture free-energy density. Note that (\ref{eq:basic_model_final_equations})$_2$ is consistent with the Debye (\ref{eq:debye}) balance only when the frictional resistance $\zeta$ is independent of the flux $\boldsymbol{\jmath}$. \subsection{Special case: diffusion of non-interacting spheres} \label{subsubsect:ideal} We now specialize the preceding results to the case of diffusing, non-interacting spherical particles, which corresponds to the dilute Brownian spheres originally treated by Einstein \cite{einstein05}. Our framework requires only the specification of constitutive relations for the mixture free-energy density and the frictional resistance. We stipulate that $\psi$ be determined by a response function $\hat{\psi}$ of the form \begin{equation} \hat{\psi}(c)=\kBT\mskip1m c\log{c}, \label{eq:assumption1b} \end{equation} which corresponds to an \emph{ideal mixture} in the sense that it consists (to within an arbitrary additive constant) only of a classical \emph{entropic} contribution, a contribution that drives diffusion. This choice does not encompass particle-particle interactions. Insertion of (\ref{eq:assumption1b}) into (\ref{eq:osmotic_eq}) leads to \begin{equation} \varpi(c)=\frac{\kBT \mskip1mu c}{M}. \end{equation} Further, we require that the response function $\hat{\boldsymbol{h}}$ be linear in the diffusive solute flux $\boldsymbol{\jmath}$ with the specific form \begin{equation} \boldsymbol{h}=\hat{\boldsymbol{h}}(c,\boldsymbol{\jmath})=-\zeta(c)\mskip1mu\boldsymbol{\jmath}, \qquad\text{with}\qquad\zeta(c)= \frac{6\pi\eta_s r} c }, \label{eq:assumption2b} \end{equation} so that the internal body microforce is consistent with Stokes' \cite{Stokes} law for the drag on a rigid spherical particle slowly translating in an incompressible Newtonian fluid. Here, $\eta_s$ is the shear viscosity of the pure fluid solvent and $r$ is the particle radius. Finally, we suppose that the external body microforce is generated by the gradient of a potential according to (\ref{eq:potential}). Granted these assumptions, we may use (\ref{eq:expression_for_flux}) to obtain \begin{equation} \boldsymbol{\jmath}= -\frac{\kBT }}{6\pi\eta_s r}\,\xg c -\frac c }{6\pi\eta_s r}\,\xg V, \label{eq:debye_eq} \end{equation} which, when inserted into the mass balance (\ref{eq:basic_model_final_equations})$_2$ yields the generalized diffusion equation \begin{align} \grave{c} &=D\mskip1mu\xd\bigl(\xgc+\frac{c}{\kBT}\,\xg{V}\bigr), \label{eq:debye_eq1} \end{align} with the diffusivity $D$ having the explicit form \begin{equation} D=\frac \kBT}{6\pi\eta_s r} \label{eq:stokes-einstein} \end{equation} consistent with the diffusivity expression originally derived by Einstein \cite{einstein05} for Brownian spheres. Thus, the contribution of $\psi$ in (\ref{eq:debye_eq}) leads to a classical, linear second-order diffusion term that tends to randomize concentrations, which justifies our description of it as \emph{entropic}. Additionally the steady-state mass fraction has the classical Boltzmann form (\ref{eq:boltzmann}). Importantly, our derivation of (\ref{eq:debye_eq1}) and (\ref{eq:stokes-einstein}) relies neither on a Fickian constitutive relation for the diffusive mass flux nor on an assumed form for the mass fraction distribution in equilibrium. Rather, the relation (\ref{eq:stokes-einstein}) is a \emph{consequence} of the thermodynamic theory and the specialized ideal constitutive expressions (\ref{eq:assumption1b}) and (\ref{eq:assumption2b}). \section{Theory accounting for both mass diffusion and convection} \label{subsect:generalization} We now account for motion of the mixture as described by a convective velocity $\boldsymbol{v}$. Further, we allow the region $\zcR$ of three-dimensional space occupied by the mixture to evolve with time. \subsection{Basic laws} \label{sect:balance2} The theory is based upon the following laws: \begin{itemize} \item[$\bullet$] mixture mass balance; \item[$\bullet$] solute mass balance; \item[$\bullet$] mixture momentum balance; \item[$\bullet$] moment of mixture-momentum balance; \item[$\bullet$] solute microforce balance; \item[$\bullet$] mixture free-energy imbalance. \end{itemize} As in our treatment of the theory without mixture convection, we formulate the basic laws in global form for an arbitrary \emph{material} subregion $\mathcal{P}$ of $\zcR$. \subsubsection{Mixture mass balance} \label{subsect:mixture_mass} We now allow for convection and variations in the mixture mass density, so that we introduce fields \begin{center} \begin{tabular}{cl} $\varrho$ & \emph{mixture mass density}, \\ [0.5ex] $\boldsymbol{v}$ & \emph{mixture velocity}. \\ [0.5ex] \end{tabular} \end{center} \emph{Mixture mass balance} is the postulate that, for each material subregion $\mathcal{P}$ of $\zcR$ and each instant, \begin{equation} \grave{\overline{\int\limits_{\mathcal{P}}\varrho\:\textsl{dv}}}=0. \label{eq:mixture_mass_balance} \end{equation} Equivalently, since $\mathcal{P}$ is arbitrary, we may enforce mixture mass balance via the local field equation \begin{equation} \dot{\varrho}+\varrho\,\xd\boldsymbol{v}=0. \label{eq:local_mixture_mass_balance} \end{equation} \subsubsection{Solute mass balance} \label{subsect:solute_mass2} With mixture convection taken into account, the global statement of solute mass balance for a material subregion $\mathcal{P}$ of $\zcR$ has the form (\ref{eq:solute_mass_balance}) considered in the absence of convection, provided that $c$ is replaced by $\varrho\mskip1muc$. Thus, bearing in mind the mixture mass balance (\ref{eq:local_mixture_mass_balance}), the local equivalent of solute mass balance now reads \begin{equation} \varrho\mskip1mu\zdot{c}=-\xd\boldsymbol{\jmath}+m. \label{eq:local_solute_mass_balance2} \end{equation} \subsubsection{Solute microforce balance} \label{subsect:micromoment2} With mixture convection taken into account, the global and local statements, (\ref{eq:orientational_balance}) and (\ref{eq:local_microforce_balance}), of solute microforce balance remain unchanged from those introduced earlier. \subsubsection{Mixture momentum balance and moment of mixture-momentum balance} \label{sect:momentum} Associated with the convective mixture velocity, we introduce fields \begin{center} \begin{tabular}{cl} $\boldsymbol{T}$ & \emph{mixture stress}, \\[0.5ex] $\boldsymbol{b}$ & \emph{external mixture body force density}, \\[0.5ex] \end{tabular} \end{center} in which case the integrals \begin{equation} \int\limits_{\mathcal{P}}\boldsymbol{T}\bfnu\:\textsl{da} \qquad\text{and}\qquad \int\limits_{\mathcal{P}}\boldsymbol{b}\:\textsl{dv} \end{equation} represent the mixture forces exerted on the region $\mathcal{P}$ by the mixture traction distributed over $\partial\mathcal{P}$, by agencies within $\mathcal{P}$, and by agencies external to $\mathcal{P}$, respectively. In addition to the mixture mass balance (\ref{eq:local_mixture_mass_balance}), the solute mass balance (\ref{eq:local_solute_mass_balance2}), and the solute microforce balance (\ref{eq:local_microforce_balance}), we enforce balances of mixture momentum and momentum of mixture momentum. \emph{Mixture momentum balance} is the postulate that, for each material subregion $\mathcal{P}$ of $\zcR$ and each instant, the time-rate at which the convective momentum within $\mathcal{P}$ changes be equal to the resultant of the convective forces acting on $\mathcal{P}$: \begin{equation} \grave{\overline{\int\limits_{\mathcal{P}}\varrho\mskip1mu\boldsymbol{v}\:\textsl{dv}}} =\int\limits_{\partial\mathcal{P}}\boldsymbol{T}\bfnu\:\textsl{da} +\int\limits_{\mathcal{P}}\boldsymbol{b}\:\textsl{dv}. \label{eq:cauchy_balance} \end{equation} Further, \emph{moment of mixture momentum balance} is the postulate that, for each material subregion $\mathcal{P}$ of $\zcR$ and each instant, the time-rate at which the moment of the convective momentum within $\mathcal{P}$ changes be equal to the resultant of the convective torques acting on $\mathcal{P}$: \begin{equation} \grave{\overline{\int\limits_{\mathcal{P}}\boldsymbol{x}\!\wedge\!\varrho\mskip1mu\boldsymbol{v}\:\textsl{dv}}} =\int\limits_{\partial\mathcal{P}}\boldsymbol{x}\!\wedge\!\boldsymbol{T}\bfnu\:\textsl{da} +\int\limits_{\mathcal{P}}\boldsymbol{x}\!\wedge\!\boldsymbol{b}\:\textsl{dv}. \label{eq:moment_balance} \end{equation} Equivalent to these balances, we have the local field equations \begin{equation} \varrho\mskip1mu\skew2\dot{\boldsymbol{v}}=\xd\boldsymbol{T}+\boldsymbol{b} \qquad\text{and}\qquad \boldsymbol{T}=\boldsymbol{T}^{\trans}. \label{eq:local_cauchy_balance} \end{equation} \subsubsection{Free-energy imbalance} \label{sect:energy} Our previous statement of free-energy imbalance must be modified to account for variations in the mixture mass density, the kinetic energy of the mixture, and the power expended by the mixture stress and external mixture body force. Specifically, we now require that for each material subregion $\mathcal{P}$ of $\zcR$ and each instant: \begin{equation} {\grave{\overline{\int\limits_{\mathcal{P}}\varrho(\psi+\half|\boldsymbol{v}|^2)\:\textsl{dv}}}} \ \le\ \int\limits_{\partial\mathcal{P}} \bigl(\zSigma^{\trans}\boldsymbol{\jmath}+\boldsymbol{T}^{\trans}\boldsymbol{v}\bigr)\!\zcdot\!\bfnu\:\textsl{da} +\int\limits_{\mathcal{P}}(\boldsymbol{\ell}\!\zcdot\!\boldsymbol{\jmath}+\mu\mskip1mu{m}+\boldsymbol{b}\!\zcdot\!\boldsymbol{v})\:\textsl{dv}. \label{eq:difference2} \end{equation} Equivalently, since $\mathcal{P}$ is arbitrary, we may enforce mixture energy imbalance via the local inequality \begin{equation} -\varrho\mskip1mu\dot{\psi}+\varrho\mskip1mu\mu\mskip1mu\dot{c} -\boldsymbol{h}\!\zcdot\!\boldsymbol{\jmath}+\boldsymbol{T}\!\zcdot\!\boldsymbol{D}+(\zSigma+\mu\idem)\!\zcdot\!\xg\boldsymbol{\jmath}\ge0, \label{eq:delta2} \end{equation} with \begin{equation} \boldsymbol{D}=\half(\xg\boldsymbol{v}+(\xg\boldsymbol{v})^{\trans}) \end{equation} the symmetric \emph{strain-rate}. \subsection{Constitutive equations} We consider the cases of compressible and incompressible mixtures separately. \subsubsection{Compressible mixture} To the lists of dependent and independent constitutive variables we add the convective stress $\boldsymbol{T}$ and the mixture density $\varrho$ and strain-rate $\boldsymbol{D}$, respectively, so that \begin{equation} (\psi,\mskip2mu\mu,\mskip2mu\zSigma,\mskip2mu\boldsymbol{h},\mskip2mu\boldsymbol{T}) =\bigl(\hat{\psi}(z),\mskip2mu\hat{\mu}(z),\mskip2mu\hat{\zSigma}(z), \mskip2mu\hat{\boldsymbol{h}}(z),\mskip2mu\hat{\boldsymbol{T}}(z)\bigr), \label{eq:cr2} \end{equation} where, for brevity, we have introduced \begin{equation} z=(\varrho, c, \boldsymbol{\jmath}, \boldsymbol{D}). \end{equation} Inserting the constitutive relations (\ref{eq:cr2}) into the local energy imbalance (\ref{eq:delta2}), we arrive at the functional inequality \begin{multline} \varrho\bigl(\frac{\partial\hat{\psi}}{\partialc}(z) -\hat{\mu}(z)\bigr)\dot{c} +\varrho\frac{\partial\hat{\psi}}{\partial\boldsymbol{\jmath}}(z)\!\zcdot\!\skew3\dot{\boldsymbol{\jmath}} +\varrho\frac{\partial\hat{\psi}}{\partial\boldsymbol{D}}(z) \!\zcdot\!\skew3\dot{\boldsymbol{D}} +\hat{\boldsymbol{h}}(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})\!\zcdot\!\boldsymbol{\jmath} \\[4pt]-\bigl(\hat{\boldsymbol{T}}(z) -\varrho^2\frac{\partial\hat{\psi}}{\partial\varrho} (z)\idem\bigr)\!\cdot\!\boldsymbol{D} -\bigl(\hat{\zSigma}(z) +\hat{\mu}(z)\idem\bigr) \!\zcdot\!\xg\boldsymbol{\jmath} \le 0. \label{eq:dissipation_imbalance_with_constitutive_relations2} \end{multline} Hence, proceding as in the our treatment of the theory without mixture convection, we find that: \begin{itemize} \item[({\em i\/})] the constitutive response functions $\hat{\psi}$, $\hat{\mu}$, and $\hat{\zSigma}$ delivering the mixture energy density $\psi$, the mixture diffusion potential $\mu$, and the solute microstress $\zSigma$ must be independent of $\boldsymbol{\jmath}$, with \begin{equation} \psi=\hat{\psi} (\varrho,c),\qquad \mu=\frac{ \partial}{\partial c} \hat{\psi} (\varrho,c), \qquad\text{and}\qquad \zSigma=-\hat{\mu}(\varrho,c) \idem; \label{eq:basic_model_restrictions2} \end{equation} \item[({\em ii\/})] the constitutive response function $\hat{\boldsymbol{h}}$ for the internal solute body force density $\boldsymbol{h}$ and the mixture stress $\boldsymbol{T}$ must be consistent with the \emph{residual inequality} \begin{equation} \hat{\boldsymbol{h}}(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})\!\zcdot\!\boldsymbol{\jmath} -{\boldsymbol{T}}_ {\!{\rm vis}} (\varrho,c,\boldsymbol{\jmath},\boldsymbol{D}) \!\zcdot\!\boldsymbol{D} \le0, \label{eq:residual_inequality2} \end{equation} in which \begin{equation} \hat{\boldsymbol{T}}_ {\!{\rm vis}} (\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})=\varrho^2\frac{\partial}{\partial\varrho} \hat{\psi}(\varrho,c)\idem+\boldsymbol{T} \end{equation} determines the dissipative contribution ${\boldsymbol{T}}_ {\!{\rm vis}} $ to the mixture stress. \end{itemize} Further, granted smoothness of the response functions $\hat{\boldsymbol{h}}$ and $\hat{\boldsymbol{T}}_ {\!{\rm vis}} $, the previously employed result of Gurtin \& Voorhees \cite{gurtin93} yields a general solution \begin{equation} \left\delimiter0 \begin{split} \boldsymbol{h}&=-\boldsymbol{Z}(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{\jmath} -\boldsymbol{Y}(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{D}, \\[4pt] \boldsymbol{T}_ {\!{\rm vis}} &=\phantom{-} \boldsymbol{U}(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{\jmath} +\boldsymbol{V}(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{D}, \label{eq:solution_of_residual_inequality2} \end{split} \right\} \end{equation} of (\ref{eq:residual_inequality2}), where $\boldsymbol{Z}$, $\boldsymbol{Y}$, $\boldsymbol{U}$, and $\boldsymbol{V}$ are isotropic due to objectivity and must obey \begin{equation} \boldsymbol{\jmath}\!\cdot\! \boldsymbol{Z}(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{\jmath} +\boldsymbol{\jmath}\!\cdot\! \boldsymbol{Y}(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{D} +\boldsymbol{D}\!\cdot\! \boldsymbol{U}(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{\jmath} +\boldsymbol{D}\!\cdot\! \boldsymbol{V}(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{D} \ge0 \end{equation} for all $(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})$. Despite their objectivity, these coefficients need not reduce to scalars. In this case, the full system of governing equations is \begin{equation} \left\delimiter0 \begin{split} \dot{\varrho}&=-\varrho\mskip1mu\xd\boldsymbol{v}, \phantom{\frac{\partial}{\partial}}\\ \varrho\mskip1mu\dot{c}&=-\xd\boldsymbol{\jmath}+m, \phantom{\frac{\partial}{\partial}}\\ \varrho\mskip1mu\skew2\dot{\boldsymbol{v}}&= -\xg\bigl(\varrho^2\frac{\partial\hat{\psi}}{\partial\varrho} (\varrho,c)\bigr +\xd\bigl(\boldsymbol{U}(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})\boldsymbol{\jmath}+\boldsymbol{V}(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})\boldsymbol{D} \bigr) +\boldsymbol{b},\\ {\boldsymbol{Z}}(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{\jmath} &= -\xg \bigl(\frac{\partial\hat{\psi}}{\partialc} (\varrho,c)\bigr) -\boldsymbol{Y}(\varrho,c,\boldsymbol{\jmath},\boldsymbol{D}) \boldsymbol{D} +\boldsymbol{\ell}, \label{eq:} \end{split} \right\} \end{equation} These equations are susbstantially more complicated than those for the case of simple diffusion without convection. However, if we require that $\hat{\boldsymbol{h}}$ and $\hat{\boldsymbol{T}}_ {\!{\rm vis}} $ be linear functions of $(\boldsymbol{\jmath},\boldsymbol{D})$, the general result (\ref{eq:solution_of_residual_inequality2}) specializes to yield \begin{equation} \left\delimiter0 \begin{split} \boldsymbol{h}&=-{\zeta}(\varrho,c)\mskip1mu\boldsymbol{\jmath}, \\[4pt] \boldsymbol{T}_ {\!{\rm vis}} &=\phantom{}2{\eta}_1(\varrho,c)\mskip1mu\boldsymbol{D} +{\eta}_2(\varrho,c)\mskip1mu(\textsl{tr}\:\boldsymbol{D})\idem, \label{eq:solution_of_residual_inequality3} \end{split} \right\} \end{equation} with \begin{equation} \zeta(\varrho,c)\ge0, \qquad \eta_1(\varrho,c)\ge0,\qquad \text{and}\qquad 2\eta_1(\varrho,c)+3{\eta}_2(\varrho,c)\ge0 \end{equation} for all $(\varrho,c)$. In this case, the full system of governing equations has the form \begin{equation}\left\delimiter0\begin{split} \dot{\varrho}&=-\varrho\,\xd\boldsymbol{v}, \phantom{\frac{\partial}{\partial}}\\ \varrho\mskip1mu\dot{c}&=-\xd\boldsymbol{\jmath}+m, \phantom{\frac{\partial}{\partial}}\\ \varrho\mskip1mu\skew2\dot{\boldsymbol{v}}&= -\xg\bigl(\varrho^2\frac{\partial\hat{\psi}}{\partial\varrho} (\varrho,c)\bigr) +\xd\bigl(2{\eta_1}(\varrho,c)\boldsymbol{D}\bigr) +\xg\bigl(\eta_2(\varrho,c)\xd\boldsymbol{v}\bigr) +\boldsymbol{b}, \\ {\zeta}(\varrho,c)\mskip1mu\boldsymbol{\jmath} &= -\xg \bigl(\frac{\partial\hat{\psi}}{\partialc}(\varrho,c)\bigr) +\boldsymbol{\ell}, \label{eq:compressible_model_final_equations}\end{split}\ \right\} \end{equation} which constitutes a system of coupled equations for the fields $(\varrho,c,\boldsymbol{v},\boldsymbol{\jmath})$. \subsubsection{Incompressible mixture} If the convective mixture velocity obeys \begin{equation} \xd\boldsymbol{v}=\textsl{tr}\:\boldsymbol{D}=0, \label{eq:div_v=0} \end{equation} so that the motion of the mixture is \emph{isochoric}, we may without loss of generality assume that $\varrho=1$ and consequently delete the mixture density from the list of independent constitutive variables. For convenience, we introduce the fields \begin{center} \begin{tabular}{cl} $p$ & \emph{mixture pressure}, \\ [0.5ex] $\boldsymbol{S}$ & \emph{mixture extra stress}, \\ [0.5ex] \end{tabular} \end{center} defined by the decomposition \begin{equation} \boldsymbol{T}=-p\idem+\boldsymbol{S}, \qquad p=-\textstyle{\frac{1}{3}}\textsl{tr}\:\boldsymbol{T}, \qquad \textsl{tr}\:\boldsymbol{S}=0, \end{equation} of the mixture stress into a constitutively indeterminate \emph{pressure} $p$ that reacts to the constraint (\ref{eq:div_v=0}) and a traceless but constitutively determinate \emph{extra stress} $\boldsymbol{S}$. We then assume that \begin{equation} (\psi,\mskip2mu\mu,\mskip2mu\zSigma,\mskip2mu\boldsymbol{h},\mskip2mu\boldsymbol{S}) =\bigl(\hat{\psi}(z),\mskip2mu\hat{\mu}(z),\mskip2mu\hat{\zSigma}(z), \mskip2mu\hat{\boldsymbol{h}}(z),\mskip2mu\hat{\boldsymbol{S}}(z)\bigr), \label{eq:cr3} \end{equation} with \begin{equation} z=(c,\boldsymbol{\jmath},\boldsymbol{D}), \end{equation} and, requiring that these relations satisfy the energy imbalance (\ref{eq:delta2}) in all processes, find that the constitutive response functions $\hat{\psi}$, $\hat{\mu}$, and $\hat{\zSigma}$ delivering the mixture energy density $\psi$, the solute diffusion potential $\mu$, and the solute microstress $\zSigma$ must, as in (\ref{eq:basic_model_restrictions}), be independent of the diffusive mass flux $\boldsymbol{\jmath}$ and the strain-rate $\boldsymbol{D}$, so that \begin{equation} \psi=\hat{\psi}(c), \qquad \mu=\hat{\psi}{}^{\prime}(c), \qquad\text{and}\qquad \zSigma=-\hat{\psi}{}^{\prime}(c)\idem. \end{equation} Further, granted that the constitutive response functions $\hat{\boldsymbol{h}}$ and $\hat{\boldsymbol{S}}$ for the internal solute body microforce $\boldsymbol{h}$ and the traceless component $\boldsymbol{S}$ of the mixture stress are smooth functions of $(c,\boldsymbol{\jmath},\boldsymbol{D})$, we find that \begin{equation} \left\delimiter0 \begin{split} \boldsymbol{h}&=-\boldsymbol{Z}(c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{\jmath}-\boldsymbol{Y}(c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{D}, \\[4pt] \boldsymbol{S}&=\phantom{-} \boldsymbol{U}(c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{\jmath}+\boldsymbol{V}(c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{D}, \label{eq:solution_of_residual_inequality4} \end{split} \right\} \end{equation} where $\boldsymbol{Z}$, $\boldsymbol{Y}$, $\boldsymbol{U}$, and $\boldsymbol{V}$ must obey \begin{equation} \boldsymbol{\jmath}\!\cdot\!\boldsymbol{Z}(c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{\jmath} +\boldsymbol{\jmath}\!\cdot\!\boldsymbol{Y}(c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{D} +\boldsymbol{D}\!\cdot\!\boldsymbol{U}(c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{\jmath} +\boldsymbol{D}\!\cdot\!\boldsymbol{V}(c,\boldsymbol{\jmath},\boldsymbol{D})\mskip1mu\boldsymbol{D}\ge0 \end{equation} for all $(c,\boldsymbol{\jmath},\boldsymbol{D})$. If, as in the compressible case considered above, we require that $\hat{\boldsymbol{h}}$ and $\hat{\boldsymbol{S}}$ be linear functions of $(\boldsymbol{\jmath},\boldsymbol{D})$, the general result (\ref{eq:solution_of_residual_inequality4}) reduces to \begin{equation} \left\delimiter0 \begin{split} \boldsymbol{h}&=-{\zeta}(c)\mskip1mu\boldsymbol{\jmath}, \\[4pt] \boldsymbol{S}&=\mskip3mu\phantom{}2\mskip1mu{\eta}_1(c)\mskip1mu\boldsymbol{D}, \label{eq:solution_of_residual_inequality5} \end{split} \right\} \end{equation} with \begin{equation} \zeta(c)\ge0 \qquad\text{and}\qquad \eta_1(c)\ge0 \end{equation} for all $c$. In this case, the full system of governing equations is: \begin{equation}\left\delimiter0\begin{split} 0 &=\xd\boldsymbol{v}, \\ \dot{c}&=-\xd\boldsymbol{\jmath}+m,\\ \skew2\dot{\boldsymbol{v}}&= -\xg p +\xd\bigl(2{\eta}_1(c)\boldsymbol{D}\bigr) +\boldsymbol{b},\\ {\zeta}(c)\mskip1mu\boldsymbol{\jmath} &= -\xg \bigl( \hat{\psi}^\prime(c)\bigr) +\boldsymbol{\ell}, \label{eq:incompressible_model_final_equations}\end{split}\ \right\} \end{equation} which constitutes a system of coupled equations for the fields $(p,c,\boldsymbol{v},\boldsymbol{\jmath})$. \section{Discussion} \label{sect:discussion} Our theory is based upon the interpretation of the diffusive flux as a generalized velocity and the introduction of corresponding power-conjugate forces, which we call microforces, into an otherwise conventional continuum-mechanical description. A systematic, thermodynamically consistent derivation generates a final system of governing equations---(\ref{eq:basic_model_final_equations}) when convection is ignored, and (\ref{eq:compressible_model_final_equations}) or (\ref{eq:incompressible_model_final_equations}) when convection is taken into account---that constitutes a generalization of common convection-diffusion equations, a generalization that accounts for constitutive nonlinearities and external forces acting on the diffusing constituents. In the absence of mixture convection, we find that the constitutive response of the material is determined by the provision of two scalar functions, one for the energy density and the other for the frictional resistance. This result generalizes naturally to account for dissipation associated with the convective motion of the mixture. Importantly, we arrive at our theory without introducing a constitutive relation such as (\ref{eq:constitutive}) for the diffusive flux. Additionally, our proposed format also leads to a generalization, (\ref{eq:generalized_stokes_einstein}), of the classical relation between diffusion coefficient and frictional resistance. We find that the classical form for the relation (\ref{eq:diffusion_coef}) holds when the free-energy density is given by the special constitutive relation (\ref{eq:assumption1b}). A distinction between the final governing equations of our theory and those arising from more standard derivations of diffusion equations is that, in our approach, a vector-valued microforce balance provides a generalization between diffusive solute flux and gradient of the solute diffusion potential. Furthermore, consistent with Nernst's treatment of osmotic forces, the gradient of solute diffusion potential emerges naturally as the diffusive force in this balance. This result for the diffusive driving force is consistent with Batchelor's \cite{Batchelor76} ``thermodynamic force'' on particles in suspension. Typically, however, this diffusive force is not formally included in force balances, but, rather, a diffusive flux is added \emph{ad hoc} to the total flux. We point out that Bird, Armstrong \& Hassager \cite{Bird77} include such a diffusive force in their force balance for polymer suspensions. \section*{Acknowledgements} This work was partially supported by the U.\ S.\ Department of Energy and the EPSRC (UK).
\section{Introduction} The embedding of supersymmetric gauge theories into the framework of string theory and the subsequent application of symmetries and dualities of the latter has shown to be an extremely fruitful approach to the study of non-perturbative properties of gauge systems. One way to proceed is to exploit the fact that type IIA string theory gives rise to non-abelian gauge symmetries when compactified on certain singular Calabi-Yau manifolds. These are naturally $K3$-fibrations and a gauge theory of ADE-type arises when the $K3$ fiber develops a corresponding singularity. ``Geometric engineering'' \cite{kkv} furthermore exploits the fact that in the field theory limit the relevant part of the compactification geometry is the local singularity structure. This allows us, for the purpose of extracting field theoretical properties, to just model manifolds that exhibit the correct configuration of exceptional divisors (${\mbox{I}\!\mbox{P}}^1$'s that occur in resolving the singularities of the $K3$ fiber and whose intersection matrix equals the negative of the Cartan matrix of the respective gauge group) fibered over the appropriate base. The literature on geometric engineering of $N=2$ supersymmetric gauge theories in four dimensions is already vast. In this paper we follow \cite{fi} in applying this method to Calabi-Yau 4-folds, which leads to $N=(2,2)$ supersymmetric gauge theories in two dimensions. The specific example we will investigate is an $SU(3)$ gauge theory. This implies that the local compactification geometry on the type IIA side contains two intersecting ${\mbox{I}\!\mbox{P}}^1$'s fibered over a common compact complex two dimensional base, which we take to be a $\mbox{I}\!\mbox{P}^2$. The rigid limit of the local mirror to this geometry is a complex surface, which plays the r\^ole of the Seiberg-Witten curve. In analogy to the 3-fold case, this surface is no longer a Calabi-Yau manifold as it was in the $SU(2)$ example of \cite{fi}. Instead, it has two holomorphic 2-forms that stem from the same meromorphic 2-form as the derivatives w.r.t.\ the two moduli, respectively. A major novelty of 4-folds, as compared to 3-folds, is that 4-forms are no longer dual to 2-forms but represent an independent part of the cohomology. In particular the primitive subspace of $\dol{2}{2}(X)$, i.e.\ the one generated by forms in $\dol{1}{1}(X)$, will play a predominant r\^ole. It is related to the occurrence of 4-fluxes $\nu_a$, which as well as part of the intersection form $\eta^{(2)}$ on this subspace appear in the twisted chiral potential $\widetilde{W}$. It is given by \begin{equation} \widetilde{W}=\nu \cdot \eta^{(2)} \cdot \sigma_{D1} \ , \label{intro:W} \end{equation} where $\sigma_{D1}$ are the middle periods of the meromorphic 2-form on the rigid surface. We will explicitly compute these periods and the four-point Yukawa couplings and then perform a non-trivial test of the generalization of rigid special geometry to 4-folds with several moduli. Recall that for 3-folds, $X_3$, special geometry implies that the prepotential $F$ of the complex structure moduli space is given in terms of the periods $Z^i,F_j$ of the holomorphic 3-form w.r.t.\ a symplectic basis of $H_3(X_3,{\mathbb Z})$ as $F=\frac{1}{2}Z^i F_i(Z)$. The three-point Yukawa couplings are then given by $C_{(3)}=\partial^3 F$, where derivatives are w.r.t.\ the special projective coordinates $Z^i$. We will verify that for 4-folds the analogous structure in the rigid limit is given by $\widetilde{W}$ as in (\ref{intro:W}) and by \[ C_{(4)}\sim (\partial^2 \sigma_{D1})\cdot \eta^{(2)} \cdot (\partial^2 \sigma_{D1}) \ . \] We will make this more precise in the following. The computation of the four-point Yukawa couplings is done in a global model of the compact Calabi-Yau 4-fold, a realization of which is given by the resolution of the Fermat hypersurface of degree 36 in the weighted projective space ${\mbox{I}\!\mbox{P}}^{5}(18,12,3,1,1,1)$, which according to common convention we will call $X_{36}(18,12,3,1,1,1)$. This manifold does not only represent a $K3$ fibration over ${\mbox{I}\!\mbox{P}}^2$, but the $K3$ is itself a fibration of an algebraic 2-torus over ${\mbox{I}\!\mbox{P}}^1$. Such elliptically fibered 4-folds are interesting in themselves as they give rise to phenomenologically more interesting $N=1$ supersymmetric gauge theories in four dimensions when used as compactification manifolds for $F$-theory \cite{fth}. \section{The holomorphic Fayet-Iliopoulos potentials} It is well known that for generic points in the moduli space spanned by the gauge multiplets the non-abelian gauge group of $N=2$ supersymmetric gauge theories is broken down to the maximal abelian torus. Hence we will have to consider an effective gauge theory with $N=(2,2)$ supersymmetry and abelian gauge group $U(1)^k$ in two dimensions, where in our specific example $k=2$. The superfields appearing in such theories comprise $k$ real vector superfields $V_a$, $a=1,\ldots,k$, with component expansion \begin{eqnarray*} V_a &=& -\sqrt{2} \left( \theta^- \bar{\theta}^- v_{\bar{z},a}+ \theta^+ \bar{\theta}^+ v_{z,a}-\theta^- \bar{\theta}^+ \sigma_a -\theta^+ \bar{\theta}^- \bar{\sigma}_a \right) + \\ & & +\, i \left( \theta^2 \bar{\theta}^{\dot{\alpha}} \bar{\lambda}_{\dot{\alpha},a} -\bar{\theta}^2 \theta^{\alpha} \lambda_{\alpha,a} \right) +\frac{1}{2} \theta^2 \bar{\theta}^2 D_a \ , \end{eqnarray*} $r$ chiral superfields $\Phi_i$, $i=1,\ldots,r$, obeying $\bar{D}_{+} \Phi_i=\bar{D}_{-} \Phi_i=0$, with component expansion \[ \Phi_i = \phi_{i}+\sqrt{2} \left(\theta^+ \psi_{+,i} + \theta^- \psi_{-,i} \right) + \theta^2 F_{i} + \ldots \ , \] where $\ldots$ are total derivative terms, as well as their complex conjugates $\bar{\Phi}_i$. As a novelty of two dimensions there are in addition twisted chiral superfields $\Sigma_{a}$, $a=1,\ldots,k$, that satisfy $\bar{D}_{+} \Sigma_a=D_{-} \Sigma_a=0$. Their component expansion reads \begin{eqnarray} \Sigma_a &=& \frac{1}{\sqrt{2}} \bar{D}_{+} D_{-} V_a \nonumber \\ &=& \sigma_a-i\sqrt{2} \left( \theta^+ \bar{\lambda}_{+,a} + \bar{\theta}^{-} \lambda_{-,a} \right) + \sqrt{2} \theta^+ \bar{\theta}^- \left( D_a-i f_a \right) + \ldots \label{expand:sig} \end{eqnarray} The most general Lagrangian involving these superfields consists of a generalized K\"{a}hler potential $K(\Sigma,\bar{\Sigma},\Phi,\bar{\Phi})$ as well as holomorphic chiral and twisted chiral potentials, $W(\Phi)$ and $\widetilde{W}(\Sigma)$. For generic points in the moduli space the chiral matter fields will be massive, so that in the infrared, after having them integrated out, we are left with only the twisted chiral fields as the light degrees of freedom and an effective action involving $K(\Sigma,\bar{\Sigma})$ and $\widetilde{W}(\Sigma)$. Taking the scaling dimension of $\Sigma$ to equal $1$, the K\"ahler potential has to be multiplied by the squared inverse of a dimensionful gauge coupling and therefore becomes irrelevant in the infrared. The twisted chiral potential on the other hand generalizes the Fayet-Iliopoulos term \begin{equation} \frac{i}{2\sqrt{2}} \int d\theta^+ d\bar{\theta}^- \left. \widetilde{W}(\Sigma) \right|_{\theta^-=\bar{\theta}^+=0} + \mbox{c.c.} = \sum_{a=1}^k \left(-\xi_{a}(\sigma) D_a+\frac{\theta_{a}(\sigma)}{2\pi} f_a \right) \end{equation} and gives rise to effective, field dependent, complex FI couplings \begin{equation} \tau_{a}(\sigma) \equiv i \xi_{a}(\sigma)+\frac{\theta_{a}(\sigma)}{2\pi} = \left. \frac{\partial \widetilde{W}(\Sigma)}{\partial \Sigma_{a}} \right|_{\theta=\bar{\theta}=0} \ , \qquad a=1,\ldots,k \ . \label{def:tau} \end{equation} These dimensionless couplings are known \cite{phases,sum} to receive logarithmic perturbative corrections to one-loop order \begin{equation} \tau_{a}(\sigma)=\tau_{a,0}-\frac{1}{2\pi i}\sum_{i=1}^{r} Q_{i}^{a} \log\left(\frac{\sqrt{2}}{\mu}\sum_{b=1}^{k} Q_{i}^{b} \sigma_{b}\right) +\ldots \ , \label{tau:RG} \end{equation} where $\mu$ is the RG scale and $Q_{i}^{a}$ is the charge of the $i^{\rm th}$ massive chiral matter field under the $a^{\rm th}$ $U(1)$ factor. In addition we expect non-perturbative corrections. A further difference from four dimensional gauge theories is the appearance of a non-trivial scalar potential \begin{equation} V(\sigma) \sim \sum_{a=1}^{k} \left| \tau_{a}(\sigma) \right|^2 \ . \label{spot} \end{equation} This potential makes the vacuum energy depend on the theta-angle \cite{col,phases} and implies that supersymmetry is broken unless $\tau_{a}(\sigma)=0$ for all $a=1,\ldots,k$. \section{The twisted chiral potential via string theory} In compactifications of type IIA string theory on a Calabi-Yau 4-fold $X$, the twisted chiral superfields $\Sigma_a$ of the resulting two dimensional gauge theory are in one-to-one correspondence with K\"ahler classes in $\dol{1}{1}(X)$, while the chiral matter fields $\Phi_{i}$ correspond to the complex structure moduli belonging to $\dol{3}{1}(X)$. The twisted chiral potential $\widetilde{W}(\Sigma)$ is holomorphic in the $\Sigma$'s and is thus a holomorphic section of a line bundle over the moduli space of K\"ahler deformations of $X$. Since the type II dilaton resides in a different multiplet, $\widetilde{W}(\Sigma)$ does not receive any quantum space-time corrections and we can restrict ourselves to string tree level. Nevertheless $\widetilde{W}(\Sigma)$ suffers from non-perturbative corrections due to embeddings of the worldsheet ($\mbox{I}\!\mbox{P}^1$, since we are in string tree-level) into $X$, i.e.\ from worldsheet instantons. These lead to quantum corrections of the K\"ahler moduli space of $X$ that show up in $\widetilde{W}(\Sigma)$. The way to compute these corrections is to use mirror symmetry, which allows us to consider instead type IIA string theory on the mirror manifold $X^*$, while the K\"ahler moduli space of $X$ is mapped to the complex structure moduli space of $X^*$ (isomorphic to $\dol{3}{1}(X^*)$) and vice versa. Here the twisted chiral potential is a holomorphic section of a line bundle over the moduli space of complex structure deformations of $X^*$; as such, it receives no quantum corrections at all. As was done in \cite{fi}, we can identify the tree-level correlator we have to compute in order to obtain $\widetilde{W}(\Sigma)$ by considering the following tree-level Chern-Simons term in ten dimensional type IIA string theory \[ {\cal L}_{\rm CS} = B \wedge F_4 \wedge F_4 \ , \] where $F_4$ is the field strength of the RR 3-form field of type IIA theory. We expand the above forms in topological bases $\{{\cal O}^{(i)}_{a_{i}}\}$ of $\dol{i}{i}(X)$ as \begin{equation} B=\sum_{a=1}^{h^{1,1}} \sigma_a {\cal O}^{(1)}_a \ , \qquad F_4=\sum_{b=1}^{\tilde{h}^{2,2}} \nu_b {\cal O}^{(2)}_b + \sum_{a=1}^{h^{1,1}} F_a \wedge {\cal O}_a^{(1)} \ , \end{equation} where $F_a$ is the field strength of the twisted chiral superfield, related to the component $f_a$ in the expansion (\ref{expand:sig}) of $\Sigma_a$ by $F_a=f_a d^2x$, with $d^2x$ the volume form on the complement of $X$. Actually the expansion in the space $\dol{2}{2}(X)$ is restricted to its primitive subspace, whose dimension we denote by $\tilde{h}^{2,2}$. Novel to 4-folds as compared to 3-folds is the r\^ole played by elements ${\cal O}^{(2)}_{b}$ and their corresponding coefficients $\nu_{b}$. Denoting the dual cycles in $H_{4}(X,{\mathbb Z})$ by $\{\gamma_{b}\}$, we have \begin{equation} \int_{\gamma_{b}} F_{4} = \nu_{b} \ , \label{4fluxes} \end{equation} and it is known \cite{4flux} that at the quantum level these 4-fluxes have to be integers (or possibly half-integers if $\frac{p_1}{4}$ is not an integral class). Using the above expansion we obtain \begin{equation} \langle {\cal L}_{\rm CS} \rangle_{\rm 10 d} \, \sim \sigma_{a} n_{b} \nu_{c} \langle {\cal O}^{(1)}_{a} {\cal O}^{(1)}_{b} {\cal O}^{(2)}_{c} \rangle_{X} \ , \label{Chern:Simons} \end{equation} where $n_b=1/(2 \pi) \int F_b$ is the first Chern class of the $b^{\rm th}$ $U(1)$ bundle, i.e.\ the instanton number of the $b^{\rm th}$ gauge factor. Note that the correlator on the right hand side is a Yukawa coupling in a topological sigma model, known as the A model \cite{top}, which is obtained by twisting the superconformal sigma model on the worldsheet with target space $X$. The algebra of observables of this model is identified with the quantum deformation of the classical intersection algebra on ${\cal A}=\oplus_{p=0}^{d} H^p(X,\wedge^p T^*X)$, where $d$ is the complex dimension of $X$. Another twist leads to the B model, whose algebra of observables is the algebra on ${\cal B}=\oplus_{p=0}^{d} H^p(X,\wedge^p TX)\cong \oplus_{p=0}^{d} H^p(X,\wedge^{d-p} T^*X)$. Mirror symmetry relates the A and B models to each other on a pair of mirror manifolds . It was shown in \cite{mird,klemm4,mayr4} how to compute the Frobenius subalgebra of ${\cal A}$ corresponding to the primitive part of the vertical cohomology of $X$ around the large radius point, which is entirely determined in terms of the two- and three-point functions. For the case of several moduli, the authors of \cite{klemm4,mayr4} found in particular that the couplings $C^{(1,1,d-2)}_{a,b,c} : \dol{1}{1}(X)\times \dol{1}{1}(X)\times \dol{d-2}{d-2}(X) \rightarrow {\mathbb C}$ are given by the period integrals of the holomorphic $d$-form on the mirror $X^*$ as \[ C^{(1,1,d-2)}_{a,b,c}=\partial_{t_a} \partial_{t_b} \Pi^{(2)}_d \eta^{(2)}_{dc} \ . \] Here $t_a$ are the periods linear in logarithms\footnote{These are the flat coordinates on the complex structure moduli space of $X^*$, i.e.\ the ones in which the Gauss-Manin connection on the bundle ${\cal B} \cong H^d(X^*,{\mathbb C})$ over this moduli space is flat. Under the mirror map they become the flat coordinates on the moduli space of the complexified K\"ahler structure of $X$.} and $\Pi^{(2)}_d$ are the quadratic ones, both in a gauge such that the unique series solution for the periods around the large complex structure point is equal to $1$. On the other hand, $\eta^{(2)}_{dc}$ are purely topological two-point functions. They define a metric on the primitive subspace of $\dol{2}{2}(X)$ by the cup product pairing on a fixed topological basis $\{{\cal O}^{(2)}_a\}$ of this subspace as \begin{equation} \eta^{(2)}_{ab}=\langle {\cal O}^{(2)}_a , {\cal O}^{(2)}_b \rangle =\int_X {\cal O}^{(2)}_a \wedge {\cal O}^{(2)}_b \ . \label{2pt:fun} \end{equation} Via mirror symmetry it coincides with the analogous cup product pairing between B model observables on the mirror $X^*$, which are related to the above basis elements ${\cal O}^{(2)}_a$ by the mirror map. Hence $\eta^{(2)}$ is a symmetric, invertible $\tilde{h}^{2,2}\times \tilde{h}^{2,2}$ matrix with integer entries. We denote its inverse by $\eta_{(2)}$. Our situation differs from that above in that we are interested in the point of the string moduli space where an enhanced gauge symmetry arises, which is not the large radius point. Furthermore we will take the field theory limit in which gravitational and stringy modes decouple. In this limit the mirror $X^*$ turns into a complex surface $W_{\rm rig}$, which we will call the local rigid surface. The attribute local refers to the fact that it is determined by the local singularity structure of $X$, i.e.\ the fibration of exceptional divisors of the resolved $K3$ over the base $\mbox{I}\!\mbox{P}^2$. In particular the number of twisted chiral superfields that do not decouple in this limit is given by the number $k$ of such exceptional divisors. Moreover, as was established in \cite{selfd} for 3-folds, there exists a map $f: H_4(X^*,{\mathbb Z})\rightarrow H_2(W_{\rm rig},{\mathbb Z})$, such that the subset of periods of the holomorphic 4-form $\Omega_{(4,0)}$ on $X^*$ that form a closed monodromy problem by themselves is given in the rigid limit by periods of a meromorphic 2-form $\lambda$ on $W_{\rm rig}$ as $\left. \int_{\gamma} \Omega_{(4,0)}\right|_{\rm rigid}=\int_{f(\gamma)} \lambda$. These periods give the scalar components of the twisted chiral superfields not decoupling in the rigid limit and of their magnetic duals. In \cite{fi} it turned out that in the resulting $U(1)^k$ gauge theory the r\^ole of the flat coordinates $t_a$ is played by the $k$ series solutions $\sigma_a$ of the periods of $\lambda$ and that the $\Pi^{(2)}_d$ are replaced by the $k$ logarithmic solutions $\sigma_{D1,d}$. The latter arise as periods over cycles that are the image under the map $f$ of 4-cycles in $X^*$, which are dual to observables of the B model on $X^*$ that correspond, via the mirror map, to elements of the primitive subspace of $\dol{2}{2}(X)$. Identifying the field theory limit of the exponential of (\ref{Chern:Simons}) with $\sim \exp(2 \pi i\tau_b(\sigma) n_b)$ we are hence led to \begin{equation} \partial_{\sigma_a}\partial_{\sigma_b}\widetilde{W}(\sigma)\sim \sum_{c} \nu_c \left. \langle {\cal O}^{(1)}_{a} {\cal O}^{(1)}_{b} {\cal O}^{(2)}_{c} \rangle_{X}\right|_{\rm rigid} \sim \sum_{c,d} \nu_c \partial_{\sigma_a}\partial_{\sigma_b}\sigma_{D1,d}(\sigma)\: \eta^{(2)}_{dc} \label{dW} \end{equation} or (modulo an additive constant and linear terms in $\sigma$) \begin{equation} \widetilde{W}(\sigma)\sim \sum_{c,d} \nu_c\: \eta^{(2)}_{cd} \, \sigma_{D1,d}(\sigma) \ . \label{W} \end{equation} The indices $a,b,c,d=1,\ldots,k$ correspond to the $k$ indices in $\{1,\ldots,h^{1,1}(X)\}$ and $\{1,\ldots,\tilde{h}^{2,2}(X)\}$, respectively, which are related to periods surviving the field theory limit. Furthermore we kept the symbol $\eta^{(2)}$ for the resulting non-degenerate $k\times k$ submatrix of (\ref{2pt:fun}). In view of (\ref{def:tau}) and (\ref{W}) the Fayet-Iliopoulos couplings are given (modulo a possible additive constant) by \begin{equation} \tau_a(\sigma) \sim \sum_{c,d=1}^{k} \nu_c\: \eta^{(2)}_{cd} \, \frac{\partial \sigma_{D1,d}}{\partial \sigma_{a}} \equiv \sum_{c,d=1}^{k} \nu_c\: \eta^{(2)}_{cd}\, \hat{\tau}_{ad}(\sigma) \ , \qquad a=1,\ldots,k\ . \label{tau} \end{equation} The objects $\hat{\tau}_{ad}=\partial_{\sigma_{a}} \sigma_{D1,d}$ are very reminiscent of the gauge couplings $\tau_{ij}=\partial_{a_i} a_{D,j}$ of the $N=2$ supersymmetric $SU(k+1)$ gauge theory in four dimensions. In that case the rigid limit of the mirror Calabi-Yau 3-fold is a genus $k$ Riemann surface $S$, whose $k$ holomorphic 1-forms $\omega_i$ stem from a meromorphic 1-form $\lambda_{\rm SW}$ as derivatives w.r.t.\ the $k$ moduli of the Coulomb branch. Its $2k$ independent 1-cycles can be chosen to form a symplectic basis $\{\alpha_j,\beta_j\}_{j=1}^{k}$ of $H_1(S,{\mathbb Z})$ such that the gauge couplings are given in terms of the $k\times k$ parts $A_{ij}=\int_{\alpha_j}\omega_i=\partial_{u_{i+1}}\int_{\alpha_j} \lambda_{\rm SW}=\partial_{u_{i+1}} a_j$ and $B_{ij}=\int_{\beta_j}\omega_i=\partial_{u_{i+1}}\int_{\beta_j} \lambda_{\rm SW}=\partial_{u_{i+1}} a_{D,j}$ of the period matrix as $\tau_{ij}=(A^{-1}B)_{ij}$. In complete analogy, the local rigid surface $W_{\rm rig}$, which arises from the Calabi-Yau 4-fold we consider, is furnished with $k$ holomorphic 2-forms $\omega_a$ stemming from a meromorphic 2-form $\lambda$ as derivatives w.r.t.\ the $k$ moduli. However, for even complex dimensional manifolds the vanishing of $\int \Omega \wedge \Omega$ implies quadratic algebraic dependences between the periods of the holomorphic form $\Omega$. We therefore expect to find $3k$ independent 2-cycles $\{\alpha_j,\beta_j,\gamma_j\}_{j=1}^{k}$ in $H_2(W_{\rm rig},{\mathbb Z})$, such that the $\alpha$s intersect with $\gamma$s and the $\beta$s only among themselves. Their $k\times 3k$ period matrix then comprises the $k\times k$ parts \begin{eqnarray} A_{ab} &=& \int_{\alpha_b}\omega_a = \partial_{u_{a+1}}\int_{\alpha_b} \lambda = \partial_{u_{a+1}} \sigma_b\ , \nonumber \\ B_{ab} &=& \int_{\beta_b}\omega_a = \partial_{u_{a+1}}\int_{\beta_b} \lambda = \partial_{u_{a+1}} \sigma_{D1,b}\ , \nonumber \\ C_{ab} &=& \int_{\gamma_b}\omega_a = \partial_{u_{a+1}}\int_{\gamma_b} \lambda = \partial_{u_{a+1}} \sigma_{D2,b} \ , \label{period:matrix} \end{eqnarray} such that the couplings $\hat{\tau}_{ab}$, as defined in (\ref{tau}), are given by $\hat{\tau}_{ab}=(A^{-1}B)_{ab}$. Whereas the $\sigma_b$ are series in the moduli, $\sigma_{D1,b}$ and $\sigma_{D2,b}$ are logarithmic and double logarithmic, respectively, and this is precisely what accounts for the logarithmic one-loop correction in (\ref{tau:RG}). In the following we will compute the periods $\sigma_a$, $\sigma_{D1,a}$ and $\sigma_{D2,a}$ using geometric engineering and in a subsequent subsection (\ref{sec:yuk}) the four-point Yukawa couplings. This will allow us to exhibit the analogue of rigid special geometry for 4-folds. \subsection{Geometric engineering} The relevant part of the 4-fold geometry is most efficiently described with the help of toric geometry. Doing so, the Mori vectors describing the two intersecting $\mbox{I}\!\mbox{P}^1$'s fibered over a base $\mbox{I}\!\mbox{P}^2$ (together with the canonical line bundle that will be irrelevant for the purpose of this section) are \begin{equation} \begin{array}{c@{\:=\: \left( \, \right.}*{6}{r@{\, , \,}} r@{ \left. \,\right)}l} l^{(1)} & 1 &-2 & 1 & 0 & 0 & 0 & 0 & \ , \\ l^{(2)} & 0 & 1 &-2 & 1 & 0 & 0 & 0 & \ , \\ l^{(3)} &-3 & 0 & 0 & 0 & 1 & 1 & 1 & \ . \end{array} \label{mori:loc} \end{equation} The local mirror $X_{\rm loc}^*$ is then the complex surface given by \begin{equation} X_{\rm loc}^* = \left\{ 0=P(y)=\sum_{i=1}^7 a_i y_i \right\} \ , \end{equation} where the variables $\{y_i\}$ are projective and subject to the constraints $1=\prod_{i=1}^7 y_i^{l_{i}^{(j)}}$, $j=1,2,3$. A solution to these is given by \[ (y_1,\ldots,y_7)=\left(s^3,s^2 t,s t^2,t^3,s^2 z,s^2 w,\frac{s^5}{z w}\right) \] with $[s,t,z,w]\in \mbox{I}\!\mbox{P}^3$, such that $P$ is a homogeneous Laurent polynomial of degree 3 \begin{equation} P=a_1 s^3+a_2 s^2 t+a_3 s t^2+a_4 t^3+a_5 s^2 z+a_6 s^2 w+a_7 \frac{s^5}{z w} \ . \label{mir:loc} \end{equation} In terms of the algebraic coordinates $a,b,c$ on the moduli space of complex structure deformations of $X_{\rm loc}^*$, where \begin{equation} a=\frac{a_1 a_3}{a_2^2} \ , \qquad b=\frac{a_2 a_4}{a_3^2} \ , \qquad c=\frac{a_5 a_6 a_7}{a_1^3} \ , \label{alg:var} \end{equation} the discriminant of $X_{\rm loc}^*$ reads \begin{equation} \tilde{\Delta}=a b c \Delta=a b c ((\Delta_{\rm cl})^3+27 a^3 c\, q(a,b,c)) \ , \label{dis:loc} \end{equation} with $q(a,b,c)$ a polynomial and \[ \Delta_{\rm cl}=-1+4 a +4 b-18 a b+27 a^2 b^2 \ . \] This discriminant is itself singular at the point $(a,b,c)=(1/3,1/3,0)$ and expanding around this singularity as \begin{equation} a = \frac{1}{3} - \left(\frac{1}{3}\right)^{2/3} \epsilon^2 u \ , \qquad b = \frac{1}{3} - \left(\frac{1}{3}\right)^{2/3} \epsilon^2 u +3 \epsilon^3 v \ , \qquad c = \epsilon^9 \Lambda^9 \ , \label{exp:locx} \end{equation} we find to lowest order in $\epsilon$ \begin{equation} \frac{1}{\epsilon^{18}}\Delta= -(4 u^3-27 v^2)^3-162 \Lambda^9 v (4 u^3+9v^2)+27 \Lambda^{18} +{\cal O}(\epsilon) \ . \label{dis:field} \end{equation} Note that whereas for 3-folds the classical discriminant splits quadratically, for 4-folds we have a cubic splitting \cite{fi}. In order to find the correct variables on the field theory moduli space, we blow up the singular point $(a,b,c)=(1/3,1/3,0)$ until we get divisors with only normal crossings \cite{pointpart}. For one particular choice of coordinate patch this leads to the following variables \begin{eqnarray} z_1 &=& b-a = 3 \epsilon^3 v \ , \nonumber \\ z_2 &=& \frac{(a-\frac{1}{3})^3}{(b-a)^2} = -\frac{1}{81} \frac{u^3}{v^2} \ , \nonumber \\ z_3 &=& \frac{c}{(b-a)^3} = \frac{1}{27} \frac{\Lambda^9}{v^3} \ . \label{var:floc} \end{eqnarray} The sections $\sigma,\sigma_{D1},\sigma_{D2}$ obey a Picard-Fuchs system of regular singular differential equations. In terms of the algebraic coordinates $(a,b,c)$ this system takes the form \begin{eqnarray} {\cal L}_1 &=& (\theta_a -3\theta_c)(\theta_a -2\theta_b)- a(-2\theta_a +\theta_b)(-2\theta_a+\theta_b-1) \ , \nonumber \\ {\cal L}_2 &=& \theta_b (-2\theta_a+\theta_b)- b(\theta_a-2\theta_b) (\theta_a-2\theta_b-1) \ , \nonumber \\ {\cal L}_3 &=& \theta_c^3-c(\theta_a-3\theta_c)(\theta_a-3\theta_c-1) (\theta_a-3\theta_c-2) \ , \label{pic:lstr} \end{eqnarray} where $\theta_a=a \partial_a$, etc. After transforming to the variables $z_1,z_2,z_3$, rescaling the periods $\pi_{\rm old}=\epsilon\pi_{\rm new}=\sqrt{\alpha^{\prime}}\pi_{\rm new}$, where $\epsilon\sim z_1^{1/3} z_3^{1/9}$, and taking the field theory limit $\epsilon \rightarrow 0$, we are left with two independent differential operators \begin{eqnarray} {\cal D}_1 &=& \frac{1}{27} \theta_2 \left(\theta_2-\frac{1}{3}\right)+z_2 \left(\frac{2}{3} \theta_2+\theta_3\right)\left(\frac{2}{3} \theta_2+\theta_3+\frac{1}{3}\right) \ , \nonumber \\ {\cal D}_2 &=& -\left(\theta_3+\frac{1}{9}\right)^3+ z_3\left(\frac{2}{3}\theta_2+\theta_3\right) \left(\frac{2}{3}\theta_2+\theta_3+\frac{1}{3}\right) \left(\frac{2}{3}\theta_2+\theta_3+\frac{2}{3}\right) \ , \label{pic:field} \end{eqnarray} where $\theta_i=z_i \partial_{z_i}$. Solutions to these are easily found, using the Frobenius method, by making the ansatz \begin{equation} \tilde{\sigma}(s,t;z_2,z_3)=\sum_{n,m\geq 0} c(n,m;s,t)z_2^{n+s}z_3^{m+t} \ . \label{ansatz} \end{equation} This determines the indices to be \begin{equation} (s,t) \in \left\{ \left(0,-\frac{1}{9}\right),\left( \frac{1}{3},-\frac{1}{9} \right) \right\} \label{ind} \end{equation} and, remembering that we chose the twisted chiral fields to have scaling dimension $1$, we fix the first coefficient to be $c(0,0;s,t)=\Lambda$. Recursion relations then imply that the general coefficient is given by \begin{eqnarray} c(n,m;s,t) &=& \Lambda \left\{ \prod_{i=1}^n \left[ (-27)\left(\frac{2}{3}(i-1+s)+t\right) \left(\frac{2}{3}(i-1+s)+t+\frac{1}{3}\right) \right] \right\} \times \nonumber \\ & & \times \frac{(\frac{2}{3}(n+s)+t)_m (\frac{2}{3}(n+s)+t+\frac{1}{3})_m (\frac{2}{3}(n+s)+t+\frac{2}{3})_m}{(s+1)_n (s+\frac{2}{3})_n [(t+\frac{10}{9})_m ]^3 } \ , \label{coeff} \end{eqnarray} where we have used the Pochhammer symbol $(a)_m = \Gamma(a+m)/\Gamma(a)=\prod_{i=0}^{m-1} (a+i)$. Since the derivatives w.r.t.\ $s$ and $t$ commute with ${\cal D}_1$ and ${\cal D}_2$ and since the first two derivatives w.r.t.\ $t$ of the indicial equations vanish at our given pairs of indices, we find the following set of solutions in a neighbourhood of $(z_2,z_3)\sim (0,0)$: two series solutions \begin{equation} \tilde{\sigma}_i(z_2,z_3)=\left. \tilde{\sigma}(s,t;z_2,z_3)\right|_{(s,t)_i} \ , \label{series} \end{equation} two logarithmic solutions \begin{eqnarray} \tilde{\sigma}_{D1,i}(z_2,z_3) &=& \partial_t \left. \tilde{\sigma}(s,t;z_2,z_3)\right|_{(s,t)_i} \label{log} \\ &=& \log(z_3) \tilde{\sigma}_i(z_2,z_3)+\sum_{n,m\geq 0}\left(\partial_t c(n,m;s,t) \right) \left. z_2^{n+s} z_3^{m+t}\right|_{(s,t)_i} \nonumber \end{eqnarray} and two double-logarithmic solutions \begin{equation} \tilde{\sigma}_{D2,i}(z_2,z_3) = \partial_t^2 \left. \tilde{\sigma}(s,t;z_2,z_3)\right|_{(s,t)_i} \ . \label{log2} \end{equation} The expansions for the series solutions read \begin{eqnarray} \tilde{\sigma}_1(z_2,z_3) &=& \frac{\Lambda}{z_3^{1/9}}\left(1+z_2-\frac{10}{729}z_3 -4 z_2^2+\frac{440}{729}z_2 z_3 -\frac{1540}{531441}z_3^2 + \frac{77}{3}z_2^3 - \right. \nonumber \\ & & \hspace{10mm} \left. - \, \frac{10472}{729}z_2^2 z_3 +\frac{261800}{531441}z_2 z_3^2-\frac{12042800}{10460353203}z_3^3 +\ldots \right) \nonumber \\ \tilde{\sigma}_2(z_2,z_3) &=& \frac{\Lambda z_2^{1/3}}{z_3^{1/9}} \left(1-z_2+\frac{28}{729}z_3 +5 z_2^2-\frac{910}{729}z_2 z_3 +\frac{7280}{531441}z_3^2 - \frac{104}{3}z_2^3 + \right. \nonumber \\ & & \hspace{10mm} \left. + \, \frac{19760}{729}z_2^2 z_3 -\frac{760760}{531441}z_2 z_3^2+\frac{76076000}{10460353203}z_3^3 +\ldots \right) \label{exp:series} \end{eqnarray} and for the logarithmic solutions as functions of the series solutions \begin{eqnarray} \tilde{\sigma}_{D1,1}(\tilde{\sigma}_1,\tilde{\sigma}_2) &=& 9\; \tilde{\sigma}_1 \log\left(\frac{\Lambda}{\tilde{\sigma}_1}\right) +\frac{27 \tilde{\sigma}_2^9}{8 \tilde{\sigma}_1^8} -\frac{27 \tilde{\sigma}_2^6}{10 \tilde{\sigma}_1^5} +\frac{9 \tilde{\sigma}_2^3}{2 \tilde{\sigma}_1^2}+\frac{55 \Lambda^9 \tilde{\sigma}_2^3}{27 \tilde{\sigma}_1^{11}} - \nonumber \\ & & -\, \frac{11 \Lambda^9}{243 \tilde{\sigma}_1^8}-\frac{2443 \Lambda^{18}}{354294 \tilde{\sigma}_1^{17}} +\ldots \nonumber \\ \tilde{\sigma}_{D1,2}(\tilde{\sigma}_1,\tilde{\sigma}_2) &=& 9\; \tilde{\sigma}_2 \log\left(\frac{\Lambda}{\tilde{\sigma}_1}\right) -\frac{27 \tilde{\sigma}_2^{10}}{10 \tilde{\sigma}_1^9} +\frac{27 \tilde{\sigma}_2^7}{14 \tilde{\sigma}_1^6} -\frac{9 \tilde{\sigma}_2^4}{4 \tilde{\sigma}_1^3}-\frac{5 \Lambda^9 \tilde{\sigma}_2^4}{\tilde{\sigma}_1^{12}} + \nonumber \\ & & +\, \frac{59 \Lambda^9 \tilde{\sigma}_2}{243 \tilde{\sigma}_1^9} +\frac{12157 \Lambda^{18}\tilde{\sigma}_2}{177147 \tilde{\sigma}_1^{18}} +\ldots \label{exp:logs} \end{eqnarray} These sections have a geometric interpretation as period integrals of a meromorphic 2-form $\lambda$ over 2-cycles in the rigid surface, which arises from the local mirror in the field theory limit and which generalizes the Seiberg-Witten curve arising from 3-folds. The physical set of periods $\sigma_{b}$, $\sigma_{D1,b}$, $\sigma_{D2,b}$ is one that corresponds to an integral basis of 2-cycles with Weyl-invariant intersection form such that the series solutions satisfy the Casimir relations \begin{eqnarray} \sigma_1^2(u,v)+\sigma_2^2(u,v)-\sigma_1(u,v) \sigma_2(u,v) &=& u+\ldots \ , \nonumber \\ \sigma_1(u,v) \sigma_2(u,v) [\sigma_1(u,v)-\sigma_2(u,v)] &=& v+\ldots \ , \label{Casimir} \end{eqnarray} where $\ldots$ indicate corrections that vanish in the classical limit (which in our coordinate patch (\ref{var:floc}) is at $v\rightarrow \infty$). Using the notation $\sigma_{(0,b)}\equiv \sigma_b$, $\sigma_{(1,b)}\equiv \sigma_{D1,b}$ and $\sigma_{(2,b)}\equiv \sigma_{D2,b}$, such a set of periods is obtained from the above solutions of the Picard-Fuchs equations by the linear transformation \begin{equation} \sigma_{(a,b)}=c_a \sum_{c=1}^{2} M_{bc} \tilde{\sigma}_{(a,c)} \ , \qquad a=0,1,2 \label{mon:trafo} \end{equation} with \begin{equation} M = \left( \begin{array}{cc} - (1-i\sqrt{3})/(2\; 3^{1/3}) & -1 \\ - 1/(3^{1/3}) & (-1)^{2/3} \end{array} \right) \label{trafo:M} \end{equation} and $c_0=1$, $c_1=i/(6 \pi)$ and $c_2=-1/(36 \pi^2)$. The monodromy around $v\sim \infty$ for constant $u$ then acts on the period vector $(\sigma_{1},\sigma_{2},\sigma_{D1,1},\sigma_{D1,2},\sigma_{D2,1},\sigma_{D2,2})^t$ as matrix multiplication by\footnote{ Actually the blocks below the diagonal could be altered by a change of homology basis, which adds series solutions to logarithmic ones and series and logarithmic solutions to double logarithmic ones and would still meet our requirements. It is however important to notice that such a change of basis leaves the relations (\ref{dW}) and (\ref{sg}) invariant and reflects the indeterminacy mentioned above (\ref{W}).} \begin{equation} M_v^{(\infty)}=\left( \begin{array}{ccc} N & 0 & 0 \\ N & N & 0 \\ N & 2N & N \end{array} \right) \qquad \mbox{where} \quad N=\left( \begin{array}{cc} -1 & 1 \\ -1 & 0 \end{array} \right) \ . \label{mon:N} \end{equation} Note that $N^3$ is the identity matrix. For our specific example we can easily write down the equation for the local rigid surface and the meromorphic 2-form $\lambda$. After shifting the $t$-variable such that the quadratic term in $t$ disappears from (\ref{mir:loc}), and using (\ref{alg:var}) and (\ref{exp:locx}), the local rigid surface takes the form \[ W_{\rm rig}=\left\{ z+w+\frac{\Lambda^9}{z w}+P_{A_2}(x;u,v) =0 \right\} \ , \] with $P_{A_2}(x;u,v)=x^3-u x-v$ the simple singularity of type $A_2$. Equivalently we can take the polynomial form \begin{equation} f_{\rm rig}=z^2 w+z w^2+\Lambda^9+z w P_{A_2}(x;u,v) \ . \label{rig:surlo} \end{equation} As for the Seiberg-Witten curve, this rigid surface is no longer a Calabi-Yau space. Rather, we expect to find two holomorphic 2-forms that are derivatives of $\lambda$ w.r.t.\ $u$ and $v$, respectively. There are curves in the field theory moduli space (spanned by $u$ and $v$) above which $\{f_{\rm rig}(w,x,z)=0\}$ is a singular space. Away from such subvarieties, i.e.\ where $f_{\rm rig}$ defines a smooth surface, we can use the Poincar\'e residue map to construct a holomorphic 2-form. Indeed, whenever we have $\left. \partial_w f_{\rm rig}\right|_{\{f_{\rm rig}=0\}} \equiv \sqrt{-4\Lambda^9 z+z^2(z+P_{A_2})^2}\neq 0$ a suitable meromorphic 2-form is given by \begin{equation} \lambda=-\frac{1}{z}\log\left[ 2z(z+P_{A_2})+2\sqrt{-4\Lambda^9 z+z^2(z+P_{A_2})^2}\, \right] dz\wedge dx \ , \label{lambda} \end{equation} such that \[ \omega_1 = \partial_v \lambda = \frac{dz\wedge dx}{\sqrt{-4\Lambda^9 z+z^2(z+P_{A_2})^2}}= \frac{dz\wedge dx}{\partial_w f_{\rm rig}} \] and \[ \omega_2 = \partial_u \lambda = \frac{x\; dz\wedge dx}{\sqrt{-4\Lambda^9 z+z^2(z+P_{A_2})^2}} = \frac{x\; dz\wedge dx}{\partial_w f_{\rm rig}} \] are two holomorphic 2-forms, as we had expected. In the case $\left. \partial_w f_{\rm rig}\right|_{\{f_{\rm rig}=0\}}=0$, but $\left. \partial_z f_{\rm rig}\right|_{\{f_{\rm rig}=0\}} \equiv \sqrt{-4\Lambda^9 w+w^2(w+P_{A_2})^2}\neq 0$, we just have to trade $z$ for $w$ in the above forms. \subsection{The Yukawa couplings and rigid special geometry} \label{sec:yuk} In (\ref{dW}) we had found that the second derivatives of the twisted chiral potential are proportional to linear combinations of three-point Yukawa couplings of the A model with target space the Calabi-Yau 4-fold $X$. These couplings in turn were identified with second derivatives of the logarithmic solutions to the period integrals of a meromorphic 2-form over 2-cycles in the rigid surface. In the previous subsection we used geometric engineering and local mirror symmetry to compute these periods. The purpose of this subsection is to compute the four-point Yukawa couplings directly from the Picard-Fuchs system. This allows us to exhibit the structure that generalizes rigid special geometry to 4-folds in an example with two moduli. The 4-fold we used above is a non-compact toric Calabi-Yau manifold, more precisely the total space of a canonical line bundle. This non-compactness did not bother us as long as we were only interested in properties of the rigid local mirror since, for these, only the compact base space of the bundle was relevant. On the other hand, since the ring structure on the cohomology of a non-compact manifold is not well defined, we need to use a global model of the compact Calabi-Yau 4-fold $X$ in order to compute the Yukawa couplings (taking the field theory limit not until the end). As already alluded to in the introduction, such a model is furnished, for example, by the Fermat hypersurface of degree 36 in the resolution of the weighted projective space $\mbox{I}\!\mbox{P}^5(18,12,3,1,1,1)$, which we called $X_{36}(18,12,3,1,1,1)$. This space is a fibration over $\mbox{I}\!\mbox{P}^2$ with fiber a $K3$ surface given as $X_{12}(6,4,1,1)$ that is itself a fibration over $\mbox{I}\!\mbox{P}^1$ with fiber this time a 2-torus $X_{6}(3,2,1)$. Note that this is the same geometry as that of $X_{24}(12,8,2,1,1)$ \cite{k3fib,kava}, except that the base $\mbox{I}\!\mbox{P}^1$ of that $K3$-fibration has been traded for a $\mbox{I}\!\mbox{P}^2$ base. Since it is the family of $K3$-fibers that determines the gauge symmetries, most of the analysis is similar to the 3-fold case \cite{hoso1}. The weighted projective space $\mbox{I}\!\mbox{P}^5(18,12,3,1,1,1)$ is a toric variety whose fan comprises the one-dimensional cones \[ \begin{array}{c@{\:=\:}l@{\ , \qquad}c@{\:=\:}l} \nu_1 & (1,0,0,0,0) & \nu_2 & (0,1,0,0,0) \ , \\ \nu_3 & (0,0,1,0,0) & \nu_4 & (0,0,0,1,0) \ , \\ \nu_5 & (0,0,0,0,1) & \nu_6 & (-18,-12,-3,-1,-1) \ . \end{array} \] They represent vertices of a reflexive polyhedron $\Delta$, which contains in addition the origin $\nu_0$ as the only interior point, as well as the two integral vertices $\nu_7=(\nu_4+\nu_5+\nu_6)/3$ and $\nu_8=(3\nu_3+\nu_4+\nu_5+\nu_6)/6$. The latter represent exceptional divisors that are introduced in the process of resolving the quotient singularities of the weighted projective space. The enlarged vertices, $\bar{\nu}_i=(1,\nu_i) \in {\mathbb Z}^{6}$, satisfy three independent relations $0=\sum_{i=0}^8 l_i^{(k)} \bar{\nu}_i$, $k=1,2,3$, which define the Mori vectors \begin{equation} \begin{array}{c@{\:=\: \left( \, \right.}*{8}{r@{\, , \,}} r@{ \left. \,\right)}l} l^{(1)} &-6 & 3 & 2 & 0 & 0 & 0 & 0 & 0 & 1 & \ , \\ l^{(2)} & 0 & 0 & 0 & 0 & 1 & 1 & 1 &-3 & 0 & \ , \\ l^{(3)} & 0 & 0 & 0 & 1 & 0 & 0 & 0 & 1 &-2 & \ . \end{array} \label{mori:glob} \end{equation} Applying standard techniques \cite{hoso1,baty,mondiv,hoso2} it follows that the mirror $X^*$ is again a hypersurface in the same weighted projective space $\mbox{I}\!\mbox{P}^5(18,12,3,1,1,1)$, given by \begin{eqnarray} p &=& a_0 z_1 z_2 z_3 z_4 z_5 z_6+a_1 z_1^2+a_2 z_2^3+a_3 z_3^{12}+ a_4 z_4^{36}+a_5 z_5^{36}+a_6 z_6^{36} \nonumber \\ & & +\, a_7 (z_4 z_5 z_6)^{12}+a_8 (z_3 z_4 z_5 z_6)^{6} \ . \label{mir:glob} \end{eqnarray} In terms of the algebraic coordinates $a=a_1^3 a_2^2 a_8/a_0^6$, $b=a_4 a_5 a_6/a_7^3$, $c=a_3 a_7/a_8^2$ on the moduli space of complex structure deformations of $X^*$ the Picard-Fuchs system for the periods $\pi_i(a,b,c)$ of the holomorphic 4-form on $X^*$ reads \begin{eqnarray} {\cal D}_1 &=& \theta_a (\theta_a-2\theta_c)-12 a (6\theta_a+1) (6\theta_a+5) \ , \nonumber \\ {\cal D}_2 &=& \theta_b^3-b (-3\theta_b+\theta_c) (-3\theta_b+\theta_c-1)(-3\theta_b+\theta_c-2) \ , \nonumber \\ {\cal D}_3 &=& \theta_c (-3\theta_b+\theta_c)- c (\theta_a-2\theta_c)(\theta_a-2\theta_c-1) \ . \label{pic:globstr} \end{eqnarray} From the analysis of the $X_{24}(12,8,2,1,1)$ model in \cite{k3fib} we can furthermore infer that the point of $SU(3)$ gauge symmetry enhancement is located in the string moduli space in rescaled variables $(x,y,z)=(432a,27b,4c)$ at $(x,y,z)=(\infty,0,1)$. Expanding around this point as \begin{equation} x = \frac{1}{2\epsilon^3 u^{3/2}} \ , \qquad y = 3 \sqrt{3}\: \epsilon^9 \Lambda^9 \ , \qquad z = 1-2\epsilon^3 u^{3/2} -3\sqrt{3}\: \epsilon^3 v \ , \label{exp:globx} \end{equation} we indeed find the discriminant of (\ref{pic:globstr}) to coincide with (\ref{dis:field}) at lowest order and up to an irrelevant factor. Next, we compute the four-point Yukawa couplings following \cite{hoso2}. This means that we first compute the four-point functions of the B model on $X^*$ directly from the Picard-Fuchs system (\ref{pic:globstr}), transform them to the field theory variables and to the correct gauge and take the rigid limit. Whereas in \cite{hoso2} this gauge was chosen such that the unique fundamental period around the large complex structure point was scaled to $1$, in our situation it is fixed by the requirement that the Picard-Fuchs system (\ref{pic:globstr}), when transformed to this gauge, reduces in the field theory limit to the rigid system (\ref{pic:field}). The rigid limit of the A model four-point Yukawa couplings is then obtained\footnote{These computations are straightforward but somehow unhandy. Therefore we refer to the appendix for details and present the verification of (\ref{sg}) only for one particular four-point coupling.} by going to the flat coordinates $\sigma_i$. As has already been mentioned, the four-point couplings are entirely determined by the two- and three-point functions as \begin{equation} C_{\sigma_i \sigma_j \sigma_k \sigma_l} = \frac{1}{4!} \sum_{\pi \in S_4} \sum_{e,f=1}^{\tilde{h}^{2,2}} C_{\sigma_{\pi(i)},\sigma_{\pi(j)},e}^{ (1,1,2)} \; \eta_{(2)}^{ef} \; C_{\sigma_{\pi(k)},\sigma_{\pi(l)},f}^{ (1,1,2)} \ . \label{yuk:fact} \end{equation} Our identification of the A model couplings $\left. C_{i,j,k}^{(1,1,2)}\right|_{\rm rigid} \sim \partial_{\sigma_i}\partial_{\sigma_j} \sigma_{D1,h} \eta^{(2)}_{hk}$ hence leads to the following relations \begin{eqnarray} \left. C_{\sigma_i \sigma_j \sigma_k \sigma_l} \right|_{\rm rigid} &=& \frac{\mbox{const}}{4!} \sum_{\pi \in S_4} \sum_{e,f,g,h} \left(\partial_{\sigma_{\pi(i)}}\partial_{\sigma_{\pi(j)}} \sigma_{D1,g} \: \eta^{(2)}_{ge} \right) \eta_{(2)}^{ef} \left(\partial_{\sigma_{\pi(k)}}\partial_{\sigma_{\pi(l)}} \sigma_{D1,h} \: \eta^{(2)}_{hf} \right) \nonumber \\ &=& \frac{\mbox{const}}{4!} \sum_{\pi \in S_4} \sum_{e,f} \left(\partial_{\sigma_{\pi(i)}}\partial_{\sigma_{\pi(j)}} \sigma_{D1,e} \right) \eta^{(2)}_{ef} \left(\partial_{\sigma_{\pi(k)}}\partial_{\sigma_{\pi(l)}} \sigma_{D1,f} \right) \ . \label{sg} \end{eqnarray} Here $\eta^{(2)}$ is the non-degenerate submatrix of the intersection form (\ref{2pt:fun}) on the primitive subspace of $\dol{2}{2}(X)$ corresponding to periods that survive the rigid limit. It is a symmetric, invertible $2\times 2$ matrix with integer coefficients, whose inverse we denote by $\eta_{(2)}$. The constant of proportionality appearing in (\ref{sg}) is of course the same for all couplings. Having calculated the four-point functions and the periods, we can indeed match the left and right sides of (\ref{sg}), which fixes the constant of proportionality and moreover determines the intersection form $\eta^{(2)}$ to be the Cartan matrix of $SU(3)$ \begin{equation} \eta^{(2)}= \left( \begin{array}{cc} 2 & -1 \\ -1 & 2 \end{array} \right) \ . \label{eta} \end{equation} Relations (\ref{W}) and (\ref{sg}) represent the analogue of rigid special geometry for 4-folds with several moduli, in the same sense that special geometry for 3-folds manifests itself in the relations $F=\frac{1}{2}Z^i F_i(Z)$ for the prepotential $F$, where $Z^i$ and $F_i$ are periods of the holomorphic 3-form w.r.t.\ a symplectic basis of $H_3(X_3,{\mathbb Z})$ and $C_{ijk}=\partial_i \partial_j \partial_k F$ for the three-point Yukawa couplings, where the derivatives are w.r.t.\ the special projective coordinates $Z^i$. They are the rigid limit of the structure found in \cite{mayr4} for the non-rigid case. \section{Conclusion} We have investigated the field theory limit of type IIA string compactification on a Calabi-Yau 4-fold whose relevant part for the purpose of extracting field theoretic properties consists of two intersecting $\mbox{I}\!\mbox{P}^1$'s fibered over a common base $\mbox{I}\!\mbox{P}^2$. The rigid limit of the local mirror is a complex surface that generalizes the Seiberg-Witten curve and on which there exist two holomorphic 2-forms that stem from the same meromorphic 2-form as derivatives w.r.t.\ the two moduli. The effective field theory that is the appropriate description in the infrared is an $N=(2,2)$ supersymmetric gauge theory in two dimensions with abelian gauge group $U(1)^2$. Its twisted chiral potential is of the form $\widetilde{W}=\nu\cdot\eta^{(2)}\cdot\sigma_{D1}$, where $\nu$ is a vector of 4-fluxes, $\eta^{(2)}$ an intersection form and $\sigma_{D1}$ a period vector. By explicit computation of the period integrals as solutions of the Picard-Fuchs equations and of the four-point Yukawa couplings we were able to exhibit the generalization of rigid special geometry to 4-folds in a non-trivial example with two moduli. This structure manifests itself in the relation $C_{(4)}=(\partial^2 \sigma_{D1}) \cdot \eta^{(2)} \cdot (\partial^2 \sigma_{D1})$ between these four-point functions and derivatives of the middle periods w.r.t.\ flat coordinates. We briefly mention a number of conclusions about two-dimensional gauge theories as derived from type IIA compactifications on 4-folds, which have already been discussed in \cite{fi} but apply to our example as well. The major novelty of 4-folds is the r\^ole played by the primitive subspace of $\dol{2}{2}(X)$ and its dual 4-cycles, respectively. They lead to new discrete moduli of the gauge theories in two dimensions, the 4-fluxes. If they all vanish, the theory just exhibits a non-trivial K\"ahler potential. But once the 4-fluxes are switched on, the structure of the theory becomes richer, as a twisted chiral potential, FI couplings and a scalar potential are generated. Generically the last seems to break supersymmetry, as was the case in the one-modulus example of \cite{fi}. The coefficient of the logarithmic term in the FI couplings can furthermore be interpreted as indicating the presence of massive chiral matter we had not accounted for in the geometrical set-up. Finally, the choice of base is not unambiguous, in contrast to the case of 3-folds, but the instanton series depends on this choice. We do not know how to interpret or resolve this ambiguity. \begin{center} {\bf Acknowledgements} \end{center} I would like to thank Ulrike Feichtinger, Peter Mayr and especially Wolfgang Lerche for discussions and sharing their insight.
\section{Introduction} Notably with the advent of the 'Dirichlet brane', which is nowadays referred to as the D-brane, discovered by Polchinski \cite{Polchinski}, the D-brane theory in various world-volume dimensions has been widely recognized as an essential ingredient in which to discuss a variety of non-perturbative aspects of superstring theory and M-theory, especially the ones concerning the intricate web of string dualities \cite{Polchinski2}. It is also remakable that D-branes not only explain the origin of a black hole entropy \cite{Strominger, Horowitz} but also provide the elementary building blocks of matrix models \cite{BFSS, IKKT}. Thus it is desirable to achieve as thorough an understanding of physical properties of D-branes as possible. In a previous paper \cite{KO} we have shown that the supersymmetric and $\kappa$-symmetric Dp-brane actions \cite{Cederwall1, Cederwall2, Bergshoeff} in a type II supergravity background have the same duality transformation properties as those in a flat Minkowski background \cite{Aganagic2}. Specially, it is shown that super D-string transforms in a covariant way while super D3-brane is self-dual under the $SL(2,Z)$ duality. Also, D2-brane and D4-brane transform in ways expected from the relation between type IIA superstring theory and M-theory. However, in the study \cite{KO} we have confined ourselves to a restricted background geometry where the dilaton and the axion are set to zero or a constant although the other fields are considered to be general superfields. This restriction is obviously unsatisfactory from the following three reasons. First, the dilaton and the axion are ordinary interacting particles in the low energy theory of superstring and thus should be treated on same footing as the other low energy fields such as the graviton, the antisymmetric second rank tensor and fermoins. Second, in a type IIB superstring or supergravity it is well known that the dilaton and the axion as well as the two 2-form antisymmetric tensor fields, those are, the NS-NS 2-form and the R-R 2-form, are doublets of the $SL(2,R)$ Mobius group whereas the graviton and the 4-form gauge field are singlets in the Einstein metric. Thus, if we wish to show that they transform covariantly under the $SL(2,R)$ transformation these local fields must be treated as not constants but fields. In this context, notice that the vanishing or constant dilaton and axion imply the vanishing 3-form field strengths, so in order to understand the transformation rules under the $SL(2,R)$ we are led to consider the non-constant dilaton and axion. \footnote{One of authors (I. O.) would like to thank M. Tonin for pointing out this issue.} Finally, a study of general dilaton and axion may shed light on F-theory \cite{F}, a conjectured 12-dimensional quantum field theory underlying a type IIB superstring. F-theory is expected to give a systematic description of the IIB superstring with $\it{non-trivial}$ dilaton and axion background to which the complex structure of a two-dimensional torus would correspond. According to these reasons, the extension of our previous work \cite{KO} to the non-constant dilaton and axion system is not only valuable but also quite non-trivial. In this paper, we will remove our previous restriction on the dilaton and the axion completely and prove that various duality transformations found in a type II supergravity background with the constant or vanishing dilaton and axion \cite{KO} can be extended to a more general situation where the dilaton and the axion are superfields depending on the supercoordinates $Z^M = (X^m, \theta^\mu)$. Therefore, the present study offers a complete proof that various duality relations in the super D-brane actions, which were constructed in \cite{Cederwall1, Cederwall2, Bergshoeff}, exist in the general on-shell type II supergravity background geometry. Here let us summarize the results that will be obtained in this paper. As in the cases of a flat Minkowski \cite{Aganagic2} and a type II supergravity backgrounds \cite{KO} with the constant dilaton and axion, we can show that the super D-string action is the $(m,1)$ string action with the NS-NS charge $m$ and the R-R charge 1. We prove that the super D3-brane action in the general background is self-dual under the $SL(2,Z)$ duality transformation in both the classical and the quantum-mechanical exact approaches. Furthermore, it is shown that under a duality transformation the super D2-brane and D4-brane actions transform in ways expected from the relation between type IIA superstring theory and M-theory. Namely, the dual action of the super D2-brane action is identified as the M2-brane action with a circular 11th dimension. As for D4-brane, the double-dimensional reduction of the M5-brane action gives rise to the dual action of the super D4-brane action. It is worthwhile to point out the technical differences between the case of the constant (or zero) dilaton and axion and that of the non-constant dilaton and axion. Moving the former on to the latter, we would encounter at least two complications. The one is that we have to introduce a new superfield $\Lambda$, i.e., the dilatino superfield, associated with the dilaton superfield in a theory, which gives rise to rather complicated constraint structures \cite{Cederwall1, Cederwall2}. Indeed, due to this complication we have confined our consideration to the constant or zero dilaton and axion in our previous work \cite{KO}. The second complication that arises is that the non-constant axion makes it impossible to use a convenient technique; in the case of the constant axion we are free to add a theta term to the dual action at the end of calculations by hand to derive the $SL(2,Z)$ covariance of the super D-string tension and the $SL(2,Z)$ self-duality of the super D3-brane \cite{Aganagic2} while in the case of the non-constant axion we have to deal with the corresponding Wess-Zumino term from the beginning. As mentioned in the above, this convenient prescription is far from complete from the point of view of proof of the $SL(2,R)$ symmetry. At this point let us comment briefly several previous articles relating to the study at hand. The duality transformations of the bosonic D-brane action have been investigated in Refs.\cite{Schmid, Tseytlin, de Alwis, Lozano}. Afterwards, supersymmetric D-brane actions with local kappa symmetry were constructed in a flat background \cite{Aganagic1} and in a type II supergravity background \cite{Cederwall1, Cederwall2, Bergshoeff}. Various duality transformations in a flat background have been clarified in the context of the super D-brane actions in Ref.\cite{Aganagic2}. (See also an important paper \cite{Townsend}.) Recently, motivated by AdS/CFT correspondence \cite{Maldacena} the duality transformations of super D-string action \cite{Oda1} and super D3-brane action \cite{Oda2, Kimura, Park} were clarified in the $AdS_5 \times S^5$ background. More recently, we have generalized these works to a type II supergravity background \cite{Oda0, KO}. However, these works are not completely general in that we have limited our consideration to the constant or zero dilaton and axion except super D-string \cite{Oda0}. The main motivation in this paper is to present a full detail of the duality transformations of super D-brane actions constructed in Refs.\cite{Cederwall1, Cederwall2, Bergshoeff} in a general type II supergravity background. Incidentally, there are several interesting papers which attempt to construct actions of super D-branes with the manifest $SL(2,Z)$-duality and a dynamical tension \cite{TBC, CW}. It would be of interest to ask possible relations between these works and the present study in future. This article is organized as follows. Section 2 reviews super D-brane actions in a general II supergravity background \cite{Cederwall1, Cederwall2, Bergshoeff}. In Section 3 it is then proved in both classically and quantum-mechanically exact manner that the super D-string action in a type IIB on-shell supergravity background is transformed to the type IIB Green-Schwarz superstring action \cite{GS} in a background with the NS-NS charge $-\lambda$ and the RR-charge 1. Section 4 deals with the super D2-brane in a type IIA on-shell supergravity background and presents that the super D2-brane action can be transformed to the super M2-brane action with a circular eleventh dimension by a duality transformation. In Section 5 we show that the super D3-brane action in a type IIB on-shell supergravity background is mapped into itself by an S-duality transformation, thereby verifying the $SL(2,Z)$ self-duality of the action. We shall offer both classical and quantum-mechanical proofs here. In Section 6 it is shown that the super D4-brane action becomes identical to the supersymmetric action which is obtained in terms of double-dimensional reduction of the super M5-brane action in the eleven dimensional space-time through a duality transformation. The final section will be devoted to discussions. In Appendix a brief review of the $SL(2,R)/SO(2)$ coset description of the dilaton and the axion in the type IIB supergravity is given and the type IIB supergravity constraint equations are shown to be invariant under the $SL(2,R)$ duality transformation. \section{Super D-brane actions} Before describing our results for various duality transformations of super D-brane actions \cite{Cederwall1, Cederwall2, Bergshoeff} in a general type II supergravity background, we shall review the salient points of the superspace formulation of the super Dp-brane theories. It is well known that super Dp-brane actions are divided into two pieces, namely, the Dirac-Born-Infeld action and the Wess-Zumino action in (p+1)-dimensional world-volume. The former includes the NS-NS two-form, dilaton and world-volume metric in addition to Abelian gauge field while the latter action contains the coupling of the D-brane to the R-R fields. The existence of this Abelian gauge field is a peculiar feature of D-brane theories. The two terms are separately invariant under type II superspace reparametrizations as well as $(p+1)$-dimensional general coordinate transformations. However, local $\kappa$ symmetry is achieved by a suitable conspiracy between the two pieces. Then, the super Dp-brane actions in a general type II on-shell supergravity background which we consider in this paper are given by \begin{eqnarray} S = S_{DBI} + S_{WZ}, \label{2.1} \end{eqnarray} with \begin{eqnarray} &{}& S_{DBI} = - \int_{M_{p+1}} d^{p+1} \sigma \ e^{\frac{p-3}{4} \phi} {\sqrt{- \det ( G_{ij} + e^{-\frac{1}{2} \phi} {\cal F}_{ij} )}}, \nn\\ &{}& S_{WZ} = \int_{M_{p+1}} e^{\cal{F}} \wedge C = \int_{M_{p+1}} \Omega_{p+1} = \int_{M_{p+2}} I_{p+2}, \label{2.2} \end{eqnarray} where $\sigma^i \ (i = 0, 1, \ldots, p)$ are the world-volume coordinates, $\phi$ the dilaton superfield, and $G_{ij}$ the induced Einstein metric of the world-volume. We have defined various quantities as follows: \begin{eqnarray} {\cal F} &=& F - b_2, \nn\\ F &=& dA, \nn\\ C &=& \displaystyle \bigoplus_{n=0}^{9} C_{(n)}, \nn\\ I_{p+2} &=& d \Omega_{p+1} = d ( e^{\cal{F}} \wedge C ), \nn\\ M_{p+1} &=& \partial M_{p+2}, \label{2.3} \end{eqnarray} where $F$ is the Maxwell field strength 2-form, and the 2-form $b_2$ is introduced in ${\cal F}$ such that ${\cal{F}}$ is invariant under supersymmetry. And the R-R $n$-form fields $C_{(n)}$ are collected in $C$ with $n$ taking odd integers for type IIA and even integers for type IIB. The integration of the integrand (\ref{2.2}) involves forms of various rank; the integral picks out precisely the terms that are proportional to the volume form of the p-brane world-volume. In addition, in order to describe the curved target superspace geometry we have to introduce the superspace vielbein 1-form $E^A$ defined by \begin{eqnarray} E^A = dZ^M E_M \ ^A, \label{2.4} \end{eqnarray} with $dZ^M$ denoting the superspace differential $(dX^m, d\theta^\mu)$, and the torsion 2-form $T^A = DE^A$ as well as the curvature 2-form defined in terms of the spin connection $\omega_A \ ^B$ as \begin{eqnarray} R_A \ ^B = d\omega_A \ ^B + \omega_A \ ^C \wedge \omega_C \ ^B. \label{2.5} \end{eqnarray} Note that we have also defined as $M = (m, \mu)$ in curved superspace while $A = (a, \alpha)$ in flat superspace as usual. The superspace vielbein 1-form $E^A$ in the tangent space decomposes under the action of the Lorentz group into a vector $E^a$ and a spinor $E^{\alpha}$. We shall follow the conventions that the Lorentz spinor $E^{\alpha}$ is a 32-component Majorana spinor for the type IIA superspace, on the other hand, a pair of 16-component Majorana-Weyl spinors for the type IIB superspace so that the latter may be written as $E^{I \alpha}$ with $I$ being the $N=2$ index ($I=1, 2$). Then the world-volume metric $G_{ij}$ is represented by \begin{eqnarray} G_{ij} = E_i \ ^a E_j \ ^b \eta_{ab}, \label{2.6} \end{eqnarray} where $E_i \ ^A = \partial_i Z^M E_M \ ^A$ and $\eta_{ab}$ = diag$(-,+, \ldots, +)$. Throughout this paper we use following conventions for superspace forms. Firstly, a general $n$-form superfield $\Omega_{(n)}$ is expanded as \begin{eqnarray} \Omega_{(n)} &=& \frac{1}{n!} dZ^{M_n} \wedge \ldots \wedge dZ^{M_1} \Omega_{M_1 \ldots M_n}, \nn\\ &=& \frac{1}{n!} E^{A_n} \wedge \ldots \wedge E^{A_1} \Omega_{A_1 \ldots A_n}, \label{2.7} \end{eqnarray} where the superspace differential $dZ^M$ and the superspace vielbein $E^A$ are antisymmetric with respect to bosonic coordinates while they are symmetric with respect to fermionic coordinates. Secondly, we define the exterior derivative as an operator acting from the right \begin{eqnarray} d (\Omega_{(m)} \wedge \Omega_{(n)}) = \Omega_{(m)} \wedge d\Omega_{(n)} + (-)^n d\Omega_{(m)} \wedge \Omega_{(n)}. \label{2.8} \end{eqnarray} Now, following the paper \cite {Cederwall2}, let us define the NS-NS 3-form superfield $H_3$ and the R-R $n$-form superfield $R$ as \footnote{See the ref.\cite{Bergshoeff} for type IIA massive supergravity, i.e., $R_{0}=m$} \begin{eqnarray} H_{3} &=& db_2, \nn\\ R &=& e^{b_2} \wedge d( e^{-b_2} \wedge C ) = \displaystyle \bigoplus_{n=1}^{10} R_{(n)}. \label{2.9} \end{eqnarray} It is obvious that from these definitions the field strengths obey the following Bianchi identities \begin{eqnarray} dH_{3} &=& 0, \nn\\ e^{b_2} \wedge d( e^{-b_2} \wedge R ) = dR - R \wedge H_{3} &=& 0. \label{2.10} \end{eqnarray} In order to reduce the enormous unconstrained field content included in the superfields to the field content of the on-shell type II supergravity theory, one has to impose the constraints on the torsion and the field strengths by hand, which make various Bianchi identities coincide with the equations of motion of type II supergravity. The nontrivial constraints imposed on the torsion and field strength components \cite{Cederwall2} take the following forms for type IIA: \begin{eqnarray} T_{\alpha\beta} \ ^c &=& 2i \gamma_{\alpha\beta} \ ^c, \nn\\ T_{\alpha\beta} \ ^{\gamma} &=& \frac{3}{2} \delta_{(\alpha} \ ^{\gamma} \Lambda_{\beta)} + 2 (\gamma_{11})_{(\alpha} \ ^{\gamma} (\gamma_{11} \Lambda)_{\beta)} - \frac{1}{2} (\gamma_a)_{\alpha\beta} (\gamma^a \Lambda)^{\gamma} \nn\\ &{}& + (\gamma_a \gamma_{11})_{\alpha\beta} (\gamma^a \gamma_{11} \Lambda)^{\gamma} + \frac{1}{4} (\gamma_{ab})_{(\alpha} \ ^{\gamma} (\gamma^{ab} \Lambda)_{\beta)}, \nn\\ H_{a \alpha\beta} &=& - 2i e^{\frac{1}{2} \phi} (\gamma_{11} \gamma_a)_{\alpha\beta}, \nn\\ H_{ab \alpha} &=& e^{\frac{1}{2} \phi} (\gamma_{ab} \gamma_{11} \Lambda)_{\alpha}, \nn\\ R_{(n) a_1 \ldots a_{n-2} \alpha\beta} &=& 2i e^{\frac{n-5}{4} \phi} (\gamma_{a_1 \ldots a_{n-2}} (\gamma_{11})^{\frac{n}{2}})_{\alpha\beta}, \nn\\ R_{(n) a_1 \ldots a_{n-1} \alpha} &=& - \frac{n-5}{2} e^{\frac{n-5}{4} \phi} (\gamma_{a_1 \ldots a_{n-1}} (-\gamma_{11})^{\frac{n}{2}} \Lambda)_{\alpha}, \label{2.11} \end{eqnarray} and for IIB: \begin{eqnarray} T_{\alpha\beta} \ ^c &=& 2i \gamma_{\alpha\beta} \ ^c, \nn\\ T_{\alpha\beta} \ ^{\gamma} &=& - ({\cal I})_{(\alpha} \ ^{\gamma} ({\cal I} \Lambda)_{\beta)} + ({\cal K})_{(\alpha} \ ^{\gamma} ({\cal K} \Lambda)_{\beta)} \nn\\ &{}& + \frac{1}{2} (\gamma_a {\cal I})_{\alpha\beta} (\gamma^a {\cal I} \Lambda)^{\gamma} - \frac{1}{2} (\gamma_a {\cal K})_{\alpha\beta} (\gamma^a {\cal K} \Lambda)^{\gamma}, \nn\\ H_{a \alpha\beta} &=& - 2i e^{\frac{1}{2} \phi} ({\cal K} \gamma_a)_{\alpha\beta}, \nn\\ H_{ab \alpha} &=& e^{\frac{1}{2} \phi} (\gamma_{ab} {\cal K} \Lambda)_{\alpha}, \nn\\ R_{(n) a_1 \ldots a_{n-2} \alpha\beta} &=& 2i e^{\frac{n-5}{4} \phi} (\gamma_{a_1 \ldots a_{n-2}} {\cal K}^{\frac{n-1}{2}} {\cal E})_{\alpha\beta}, \nn\\ R_{(n) a_1 \ldots a_{n-1} \alpha} &=& - \frac{n-5}{2} e^{\frac{n-5}{4} \phi} (\gamma_{a_1 \ldots a_{n-1}} {\cal K}^{\frac{n-1}{2}} {\cal E} \Lambda)_{\alpha}, \label{2.12} \end{eqnarray} where the round brackets enclosing indices denote symmetrization with 'strength one', and then $\Lambda_{\alpha} \equiv \frac{1}{2} \partial_{\alpha} \phi$, and ${\cal E}$, ${\cal I}$, and ${\cal K}$ describing the $SL(2,R)$ matrices are defined in terms of the conventional Pauli matrices $\sigma_i$ as follows: \begin{eqnarray} {\cal E} = i \sigma_2 = \pmatrix{ 0 & 1 \cr -1 & 0 \cr }, \ {\cal I} = \sigma_1 = \pmatrix{ 0 & 1 \cr 1 & 0 \cr }, \ {\cal K} = \sigma_3 = \pmatrix{ 1 & 0 \cr 0 & -1 \cr }. \label{2.13} \end{eqnarray} Based on this formulation of the super Dp-brane actions in a type II on-shell supergravity background, we shall explore various duality symmetries in subsequent sections. \section{The super D-string} In this section we would like to consider the super D-string (i.e. the super D1-brane) first. The super D2, D3 and D4-branes will be treated in order in subsequent sections. In these sections we shall prove various duality symmetries of the super D-brane actions in a type II on-shell supergravity background. The corresponding proofs have been already done in a flat Minkowskian background \cite{Aganagic2} and a type II supergravity background with the constant or zero dilaton and axion \cite{KO}. A partial result was also presented in a type IIB supergravity background with the non-constant dilaton and axion \cite{Oda0}. The purpose of this section is to offer the full detail and address the issue of the $SL(2,Z)$ symmetry in such a most general background. Our study appears to be quite important for future development in string theory and M-theory since the global discrete symmetries such as the $SL(2,Z)$ S-duality are nowadays believed to be exact symmetries in still mysterious underlying theory \cite{Hull, Witten2} so that these symmetries should be valid even in a curved background geometry. In the case at hand, by making use of the superspace convention (\ref{2.7}) the action (\ref{2.1}), (\ref{2.2}) and the relevant constraints (\ref{2.12}) for type IIB reduce to \begin{eqnarray} S &=& S_{DBI} + S_{WZ}, \nn\\ S_{DBI} &=& - \int_{M_2} d^2 \sigma e^{-\frac{1}{2} \phi} \sqrt{- \det ( G_{ij} + e^{-\frac{1}{2} \phi} {\cal F}_{ij} )}, \nn\\ S_{WZ} &=& \int_{M_2 = \partial M_3} (C_2 + C_0 {\cal F}) = \int_{M_3} I_3, \label{3.1} \end{eqnarray} \begin{eqnarray} H_{3} &=& db_2 = i e^{\frac{1}{2} \phi} \bar{E} \wedge \hat{E} \wedge {\cal K} E + \frac{1}{2} e^{\frac{1}{2} \phi} \bar{E} \wedge \gamma_{ab} {\cal K} \Lambda E^b \wedge E^a , \nn\\ R_{(1)} &=& dC_0 = 2 e^{-\phi} \bar{E} {\cal E} \Lambda, \nn\\ R_{(3)} &=& dC_2 - H_3 C_0 \nn\\ &=& - i e^{-\frac{1}{2} \phi} \bar{E} \wedge \hat{E} \wedge {\cal I} E + \frac{1}{2} e^{-\frac{1}{2} \phi} \bar{E} \wedge \gamma_{ab} {\cal I} \Lambda E^b \wedge E^a, \label{3.d} \end{eqnarray} where $\bar{E}$, $\hat{E}$ and $E$ represent the Dirac conjugate of $E^{I\alpha}$, $E^a \gamma_a$ and $E^{I\alpha}$, respectively. The equations in (\ref{3.1}) and (\ref{3.d}) will serve as our basis for all subsequent considerations in this section. \subsection{The classical analysis} In this subsection, we turn our attention to a duality transformation of the super D-string action in a classical approach. Next subsection will provide a quantum-mechanical exact treatment of the same model. Having exhibited the explicit forms of the classical action and the constraints of the super D-string, the next task is to construct a dual action out of them. The strategy for this purpose is now standard, but for definiteness let us present the detailed exposition. As a first step, one needs to incorporate a Lagrange multiplier field $\tilde{H}^{ij} = - \tilde{H}^{ji}$ into the classical action as follows \cite{Tseytlin}: \begin{eqnarray} S &=& - \int_{M_2} d^2 \sigma e^{-\frac{1}{2} \phi} \sqrt{- \det ( G_{ij} + e^{-\frac{1}{2} \phi} {\cal F}_{ij} )} + \int_{M_2} (C_2 + C_0 {\cal F}) \nn\\ &{}& + \int_{M_2} d^2 \sigma \frac{1}{2} \tilde{H}^{ij} (F_{ij} - 2 \partial_i A_j), \label{3.2} \end{eqnarray} and then regard $F_{ij}$ as an independent superfield. The equation of motion of $A_i$ gives rise to $\partial_i \tilde{H}^{ij} = 0$ whose classical solution is found to be $\tilde{H}^{ij} = \varepsilon^{ij} \lambda$ with a constant scalar superfield $\lambda$. Next, substituting this classical solution into the original action (\ref{3.2}) gives us an action $S' = S_1 + S_{2D}$, where \begin{eqnarray} S_1 &=& \int_{M_2} d^2 \sigma \left[ - e^{-\frac{1}{2} \phi} \sqrt{- \det ( G_{ij} + e^{-\frac{1}{2} \phi} {\cal F}_{ij} )} + \frac{1}{2} (\lambda + C_0) \varepsilon^{ij} {\cal F}_{ij} \right], \nn\\ S_{2D} &=& \int_{M_2} (C_2 + \lambda b_2). \label{3.3} \end{eqnarray} The important procedure to get a dual action is to solve the equation of motion obtained by varying $F_{ij}$ to express the action $S'$ in terms of $\lambda$ instead of $F_{ij}$. Note that since $S_{2D}$ does not include $F_{ij}$ in it, this part of the action is manifestly unaffected by the duality transformation. A simple calculation leads to \begin{eqnarray} {\cal F}_{01} = - e^{\frac{1}{2} \phi} \frac{\lambda + C_0}{\sqrt{e^{-2 \phi}+(\lambda + C_0)^2}} \sqrt{- \det{G_{ij}}}, \label{3.4} \end{eqnarray} where we have used the formula holding for $2 \times 2$ matrices \begin{eqnarray} \det ( G_{ij} + e^{-\frac{1}{2} \phi} {\cal F}_{ij} ) &=& \det{G_{ij}} + e^{- \phi} {\cal F}_{01}^2 \nn\\ &=& \det{G_{ij}} + \frac{1}{4} e^{- \phi} (\varepsilon^{ij} {\cal F}_{ij})^2. \label{3.5} \end{eqnarray} It follows that inserting Eq.(\ref{3.4}) to $S_1$ together with $S_{2D}$ yields the dual action of the super D-string action \begin{eqnarray} S_D = - \int_{M_2} d^2 \sigma e^{\frac{1}{2} \phi} \sqrt{e^{-2 \phi} + (\lambda + C_0)^2} \sqrt{- \det G_{ij}} + \int_{M_2} (C_2 + \lambda b_2). \label{3.6} \end{eqnarray} We may interpret the dual action (\ref{3.6}) in the following way. If we regard the Wess-Zumino term in (\ref{3.6}), $\int_{M_2} (C_2 + \lambda b_2)$, as the source term for the NS-NS and the R-R 2-form potentials, it turns out that this dual action carries the NS-NS charge $-\lambda$ and the R-R charge 1 \footnote{Here we have used the fact that $(b_2, -C_2)$ is the $SL(2,R)$ doublet as shown around the end of this section.}. Then provided that we identify the constant scalar superfield $- \lambda = m$ with the NS-NS charge corresponding to the $(m,1)$ string \footnote{See the next subsection about the reason why the scalar $\lambda$ becomes an integer $m$.}, the result obtained above, $\sqrt{e^{-2 \phi} + (m - C_0)^2}$, agrees with the tension formula for the $SL(2,Z)$ S-duality spectrum of strings in the type IIB superstring \cite{Schwarz} \footnote{For comparison of the string tension, we need to consider the string metric defined as $G_{ij}^{(s)} = e^{\frac{1}{2} \phi} G_{ij}$ where $G_{ij}$ is the Einstein metric.}. This identification means that the D-string action may be actually the action for an arbitrary number of 'fundamental' IIB strings bound to a single D-string. This singleness of D-string of course reflects the $U(1)$ character of the Abelian gauge field since a system of $N$ coincident Dp-branes would be described by the non-abelian version of the Dirac-Born-Infeld action \cite{DBI}. Now let us examine the implication of Eq.(\ref{3.6}) in more detail and consider the exact connection with the type IIB Green-Schwarz superstring action. For that purpose we rewrite the Wess-Zumino term in (\ref{3.6}) in the form \begin{eqnarray} d\Omega_D &\equiv& d (C_2 + \lambda b_2) \nn\\ &=& R_{(3)} + (\lambda + C_0) H_3 \nn\\ &=& i \bar{E} \wedge \gamma_a \left[e^{-\frac{1}{2} \phi} {\cal I} - (\lambda + C_0) e^{\frac{1}{2} \phi} {\cal K} \right] E \wedge E^a \nn\\ &{}& + \frac{1}{2} \bar{E} \wedge \gamma_{ab} \left[e^{-\frac{1}{2} \phi}{\cal I} + (\lambda + C_0) e^{\frac{1}{2} \phi} {\cal K} \right] \Lambda E^b \wedge E^a, \label{3.e} \end{eqnarray} where Eq.(\ref{3.d}) was used. Although the two matrices in the square brackets which are $2 \times 2$ hermitian matrices, have the same eigenvalues $\pm e^{-\frac{1}{2} \phi} \sqrt{1 + (\lambda + C_0)^2 e^{2 \phi}}$, these matrices are not mutually commutable so that they cannot be diagonalized simultaneously. In the case of constant dilaton and axion the last term in (\ref{3.e}) vanishes and the first term can be diagonalized by a suitable $SO(2)$ spinor-rotation. As a result, (\ref{3.e}) can be written as \begin{eqnarray} d\Omega_{D} = e^{\frac{1}{2} \phi} \sqrt{e^{-2 \phi} + (\lambda + C_0)^2} i\bar{E}^{\prime}\wedge \gamma_{a}E^{a}\wedge{\cal K}E^{\prime}. \nonumber \end{eqnarray} {}From this form of the dual action, it has been concluded previously \cite{Aganagic2,Oda0,KO} that the dual action of D-string is equivalent to the type IIB Green-Schwarz superstring with the $SL(2,Z)$ covariant string tension $\sqrt{e^{-2 \phi} + (\lambda + C_0)^2}$ \cite{Schwarz}. But this argument is a little naive because in this case we can change the string tension at will through the field redefinitions. In the present formulation, however, we have to take notice of at least two problems in this argument. Firstly in the process of the diagonalization the $SO(2)$ spinor-rotation must be accompanied by some associated $SL(2,R)$ transformation of the background fields in order to preserve the type IIB supergravity constraint equations; namely type IIB supergravity equations of motion. Secondly since in the general case of non-constant dilaton and axion the last term in (\ref{3.e}) does not vanish, two terms in (\ref{3.e}) can not be simultaneously diagonalized by some $SO(2)$ spinor-rotation. It is remakable to see that these problems are closely related and resolved in the following way. To do so let us rewrite (\ref{3.6}) in the form \begin{eqnarray} S_D^{\prime} = - \int_{M_2} d^2 \sigma e^{\frac{1}{2}\phi^{\prime}} \sqrt{- \det G_{ij}} + \int_{M_2} b_{2}^{\prime}, \label{3.a} \end{eqnarray} where \begin{eqnarray} e^{\frac{1}{2}\phi^{\prime}} &=&e^{\frac{1}{2} \phi} \sqrt{e^{-2\phi}+ (\lambda + C_0)^2}, \nonumber \\ b_{2}^{\prime} &=& C_2 + \lambda b_2. \label{3.b} \end{eqnarray} According to the $SL(2,R)/SO(2)$ coset description of dilaton and axion in type IIB supergravity reviewed in Appendix, Eq.(\ref{3.b}) is just the $SL(2,Z)$ transformation $\tau\rightarrow \tau^{\prime}=\frac{-1}{\tau + \lambda}$ associated with the $SL(2,Z)$ matrix \begin{equation} S= \left( \begin{array}{cc} 0 & -1 \\ 1 & \lambda \end{array} \right), \ (S^{T})^{-1}= \left( \begin{array}{cc} \lambda & -1 \\ 1 & 0 \end{array} \right), \label{3.c} \end{equation} where $(b_{2},-C_{2})$ belongs to an $SL(2,R)$ doublet which transforms like $B$ in (\ref{A-11}). As discussed in Appendix the type IIB supergravity constraints are invariant under the $SL(2,R)$ transformation combined with some suitable $SO(2)$ spinor-rotation and therefore $b_{2}^{\prime}$ satisfies the type IIB supergravity constraint equation for the NS-NS 2-form. To compare with the Wess-Zumino term of the type IIB Green-Schwarz superstring action, let us take an exterior derivative of the integrand in the Wess-Zumino term, which is calculated as follows: \begin{eqnarray} d \Omega_D^{\prime} &\equiv& db_2^{\prime} \nn\\ &=&e^{\frac{1}{2}\phi^{\prime}} \left[i\bar{E}^{\prime} \wedge \gamma_{a}E^{a} \wedge{\cal K}E^{\prime} + \frac{1}{2} \bar{E}^{\prime} \wedge \gamma_{ab}{\cal K} \Lambda^{\prime} E^b \wedge E^a \right]. \label{3.7} \end{eqnarray} where $E^{\prime}$ and $\Lambda^{\prime}$ are the $SO(2)$ spinor-rotated supervielbein and dilatino superfield, respectively. The spinor-rotation angle is given by (\ref{A-17}) and (\ref{3.c}). Thus it follows that the dual action (\ref{3.a}) is the type IIB Green-Schwarz superstring action with the unit string tension, $T=1$ in the $SL(2,R)$ transformed type IIB supergravity background \cite{Grisaru}. (Note that this $SL(2,Z)$ transformed background has the NS-NS charge $-\lambda$ and the RR-charge 1.) In this way the dual action (\ref{3.6}) is precisely transformed into the type IIB Green-Schwarz superstring action by a $SL(2,Z)$ duality transformation (combined with a suitable $SO(2)$ spinor-rotation). Of course, this does not imply that the dual action (\ref{3.6}) of the super D-string is equivalent to the Green-Schwarz superstring action with the unit NS-NS charge, for the $SL(2,Z)$ transformation is not a symmetry in the case of string theory, as opposed to D3-brane case. Instead, the above demonstration means that the dual action (\ref{3.6}) is indeed nothing but the type IIB superstring action in a background with the NS-NS charge $-\lambda$ and the RR-charge 1. As alluded in Section 1, in order to obtain this result, we have not adopted the convenient technique for which the Wess-Zumino action containing the constant axion, which is a topological theta term, is taken into account at the end of calculations in order to produce the $SL(2,Z)$ covariant tension, instead treated its non-trivial Wess-Zumino action due to the non-constant character of the axion field from the beginning in a direct manner. \subsection{The quantum analysis} The results we have obtained in the previous subsection unfortunately rely on the classical analysis in the sense that we have made use of the field equations at least in the two stages, namely, we have used the field equations of $A_i$ and $F_{ij}$. Now we are ready to present a quantum-mechanical exact proof of $SL(2,Z)$ S-duality of the super D-string action in a general ten dimensional IIB supergravity background. (The results in this subsection were briefly reported in a short article \cite{Oda0}.) To this aim, let us utilize the first-order Hamiltonian form of path integral following the techniques developed in \cite{de Alwis, Oda4}. As a first step of the Hamiltonian formalism, let us introduce the canonical conjugate momenta $\pi^i$ corresponding to the gauge field $A_i$ defined as \begin{eqnarray} \pi^i \equiv \frac{\partial S} {\partial \dot{A}_i}, \label{3.9} \end{eqnarray} where the dot denotes the time derivative. Then the canonical conjugate momenta $\pi^i$ are calculated to be \begin{eqnarray} \pi^0 = 0, \ \pi^1 = e^{-\frac{3}{2} \phi} \frac{{\cal F}_{01}} {\sqrt{-\det ( G_{ij} + e^{-\frac{1}{2} \phi} {\cal F}_{ij} )}} + C_0, \label{3.10} \end{eqnarray} where the former equation implies the existence of the $U(1)$ gauge invariance. From the latter equation $\dot{A}_1$ can be expressed in terms of $\pi^1$ whose result is given by \begin{eqnarray} \dot{A}_1 = e^{\frac{1}{2} \phi} \frac{\pi^1 - C_0}{\sqrt{(\pi^1 - C_0)^2 + e^{-2 \phi}}} \sqrt{- \det{G_{ij}}} + b_{01} + \partial_1 A_0. \label{3.11} \end{eqnarray} Then we will see that the Hamiltonian density takes the form \begin{eqnarray} {\cal H} = e^{\frac{1}{2} \phi} \sqrt{e^{-2 \phi}+(\pi^1 - C_0)^2} \sqrt{- \det G_{ij}} - A_0 \partial_1 \pi^1 + \partial_1 (A_0 \pi^1) + \pi^1 b_{01} - C_{01}. \label{3.12} \end{eqnarray} Now the partition function is defined by the first-order Hamiltonian form with respect to only the gauge field as follows: \begin{eqnarray} Z &=& \frac{1}{\int {\cal D}\pi^0} \int {\cal D}\pi^0 {\cal D}\pi^1 {\cal D}A_0 {\cal D}A_1 \exp{ i \int d^2 \sigma ( \pi^1 \partial_0 A_1 - {\cal H} ) } \nn\\ &=& \int {\cal D}\pi^1 {\cal D}A_0 {\cal D}A_1 \exp{ i \int d^2 \sigma} \nn\\ &{}& \times \left[ - A_1 \partial_0 \pi^1 + A_0 \partial_1 \pi^1 - e^{\frac{1}{2} \phi} \sqrt{e^{-2 \phi}+ (\pi^1 - C_0)^2} \sqrt{- \det G_{ij}} - \pi^1 b_{01} + C_{01} \right], \label{3.13} \end{eqnarray} where we have taken the boundary condition for $A_0$ such that the surface term identically vanishes. Then we can carry out the integrations over $A_i$ explicitly, which gives rise to $\delta$ functions \begin{eqnarray} Z &=& \int {\cal D}\pi^1 \delta(\partial_0 \pi^1) \delta(\partial_1 \pi^1) \exp{ i \int d^2 \sigma} \nn\\ &{}& \times \left[ -e^{\frac{1}{2} \phi} \sqrt{e^{-2\phi}+(\pi^1 - C_0)^2}\sqrt{- \det G_{ij}} + C_{01} - \pi^1 b_{01} \right]. \label{3.14} \end{eqnarray} The existence of the $\delta$ functions reduces the integral over $\pi^1$ to the one over only its zero-modes. If we require that one space component is compactified on a circle, these zero-modes are quantized to be integers \cite{Witten}. As a consequence, the partition function becomes \begin{eqnarray} Z &=& \displaystyle{ \sum_{m \in {\tenbf} \def\it{\tenit} \def\sl{\tensl Z}} } \exp{ i \int d^2 \sigma \left[ - e^{\frac{1}{2} \phi} \sqrt{-e^{2 \phi}+ (m - C_0)^2} \sqrt{- \det G_{ij}} + C_{01} - m b_{01} \right] }, \label{3.15} \end{eqnarray} from which we can read off the effective action \begin{eqnarray} S = \int d^2 \sigma \left( -e^{\frac{1}{2} \phi} \sqrt{e^{-2 \phi} + (m - C_0)^2} \sqrt{- \det G_{ij}} + C_{01} - m b_{01} \right). \label{3.16} \end{eqnarray} Note that if we replace $m$ with $-\lambda$ this action precisely becomes equivalent to (\ref{3.6}). In this way, we have derived the dual action of the super D-string action in a type IIB on-shell supergravity background geometry in quantum-mechanical exact manner. To close this section let us present an interesting observation which will play an important role especially in proving the self-duality of the super D3-brane action in Sec.5. Notice that the original action (\ref{3.1}) possesses the following two symmetries; \begin{eqnarray} C_{0}&\rightarrow& C_{0}^{\prime}=C_{0}+1, \nonumber \\ b_{2}&\rightarrow& b_{2}^{\prime}=b_{2}, \ C_{2}\rightarrow C_{2}^{\prime}=C_{2}-{\cal F}=C_{2}+b_{2}-F, \label{3-17} \end{eqnarray} and \begin{eqnarray} C_{0}&\rightarrow& C_{0}^{\prime}=C_{0}+1, \nonumber \\ b_{2}&\rightarrow& b_{2}^{\prime}=b_{2}, \ C_{2}\rightarrow C_{2}^{\prime}=C_{2}+b_{2}, \label{3-18} \end{eqnarray} which stem from the structure of the Wess-Zumino action. The super D-string action (\ref{3.1}) is exactly invariant under (\ref{3-17}) and invariant only up to the topological term F under (\ref{3-18}). These symmetries do not affect the type IIB supergravity constraints (\ref{3.d}), which implies that these are real symmetries of the theory. Of course, it is obvious that the partition function (\ref{3.15}) is also invariant under these symmetries. Noticing $F=dA$ the transformation (\ref{3-17}) may be interpreted as the transformation (\ref{3-18}) combined with a gauge transformation of $C_{2}$ whose gauge parameter is the world-volume Abelian gauge field itself. Then it is natural to ask what the meaning of these symmetries is. In the following we shall follow the notations in Appendix. First, when we introduce the complex variable $\tau$ by $\tau = C_{0}+ie^{-\phi}$, the shift of $C_{0}$ corresponds to $\tau \rightarrow \tau +1$, being an element of $SL(2,R)$ associated with $SL(2,R)$ matrix \begin{equation} S= \left( \begin{array}{cc} 1 & 1 \\ 0 & 1 \end{array} \right), (S^{T})^{-1}= \left( \begin{array}{cc} 1 & 0\\ -1 & 1 \end{array} \right), \label{3-20} \end{equation} Next we should also investigate how this symmetry acts on the other superfields in the theory. Comparing the transformation (\ref{3-18}) with (\ref{A-11}), $(b_{2},-C_{2})$ is expected to form an $SL(2,R)$ doublet which transforms like $B$ in (\ref{A-11}). This statement is consistent with the result obtained in the previous subsection and also supported in Sec.5 by the fact that under the weak-strong duality $\tau\rightarrow -\frac{1}{\tau}$, $(b_{2},-C_{2})$ correctly transforms as an $SL(2,R)$ doublet. It is illuminating that super D-brane actions in the general type IIB supergravity background has a shift symmetry $\tau\rightarrow \tau+1$ as expected from the fact that this is a symmetry of pertubation theory. This fact has been thus far understood in the target space approach where the R-R scalar $C_{0}$ appears only through its field strength. But in the present world-volume approach this comes from the non-trivial symmetry of the super D-brane action. We wish to stress that it is possible to prove this interesting observation only in the present formalism within the context of the world-volume theory as mentioned in Sec.1. \section{The super D2-brane} Next we turn to the $\it{classical}$ derivation of a duality transformation between the super D2-brane (i.e., the super D-membrane) in a type IIA supergravity background and the super M2-brane in eleven dimensional supergravity. The authors in a paper \cite{Bergshoeff} have already dealt with this problem from a different viewpoint. The method adopted there is to start with the super M2-brane in eleven dimensions, achieve the dimensional reduction to ten dimensions a la Kaluza-Klein ansatz, then perform a duality transformation for the purpose of getting the super D2-brane action and its $\kappa$-symmetry. Our method is similar to that of Aganagic et al. \cite{Aganagic2} where the above arguments were reversed, namely, the super M2-brane action was obtained from starting with the super D2-brane action through a duality transformation of the world-volume gauge field. The case of the constant or zero dilaton was already considered in our previous work \cite{KO}. Making use of the formulation explained in Sec.2, we focus on the problem of how our previous work should be extended to the case of the non-constant dilaton superfield. {}From Eqs.(\ref{2.1}) and (\ref{2.2}), the super D2-brane action is of the form \begin{eqnarray} S &=& S_{DBI} + S_{WZ} + S_{\tilde{H}}, \nn\\ S_{DBI} &=& - \int_{M_3} d^3 \sigma e^{-\frac{1}{4} \phi} \sqrt{- \det ( G_{ij} + e^{-\frac{1}{2} \phi} {\cal F}_{ij} )}, \nn\\ S_{WZ} &=& \int_{M_3} ( C_3 + C_1 \wedge {\cal F} ) = \int_{M_4} I_4, \nn\\ S_{\tilde{H}} &=& \int_{M_3} d^3 \sigma \frac{1}{2} \tilde{H}^{ij} ( F_{ij} - 2 \partial_i A_j ), \label{4.1} \end{eqnarray} where we have added $S_{\tilde{H}}$ to the original action in order to perform a duality transformation. Moreover, in this case the constraints (\ref{2.11}) on the field strengths reduce to \begin{eqnarray} H_3 &=& db_2 = i e^{\frac{1}{2} \phi} \bar{E} \wedge \gamma_{11} \hat{E} \wedge E + \frac{1}{2} e^{\frac{1}{2} \phi} \bar{E} \wedge \gamma_{ab} \gamma_{11} \Lambda E^b \wedge E^a , \nn\\ R_{(2)} &=& dC_1 = i e^{-\frac{3}{4} \phi} \bar{E} \wedge \gamma_{11} E - \frac{3}{2} e^{-\frac{3}{4} \phi} \bar{E} \wedge \gamma_a \gamma_{11} \Lambda E^a, \nn\\ R_{(4)} &=& dC_3 + H_3 \wedge C_1 \nn\\ &=& \frac{i}{2} e^{-\frac{1}{4} \phi} \bar{E} \wedge \gamma_{ab} E \wedge E^b \wedge E^a + \frac{1}{12} e^{-\frac{1}{4} \phi} \bar{E} \wedge \gamma_{abc} \Lambda E^c \wedge E^b \wedge E^a, \label{4.2} \end{eqnarray} where (\ref{2.7}), (\ref{2.8}) and (\ref{2.9}) were also utilized. Of course, $C_1$ and $C_3$ are determined by solving Eq.(\ref{4.2}). Varying $A_i$ in the action (\ref{4.1}) gives the equation of motion, $\partial_i \tilde{H}^{ij} = 0$ like the super D-string case. Then the solution is obviously given by $\tilde{H}^{ij} = \epsilon^{ijk} \partial_k B$ with $B$ being a scalar superfield. Next, substituting this classical solution into the original action (\ref{4.1}) leads to an action $S' = S_1 + S_{2D}$, where \begin{eqnarray} S_1 &=& \int_{M_3} d^3 \sigma \left[ - e^{-\frac{1}{4} \phi} \sqrt{- \det ( G_{ij} + e^{-\frac{1}{2} \phi} {\cal F}_{ij} )} + \frac{1}{2} \varepsilon^{ijk} (\partial_k B + C_k) {\cal F}_{ij} \right], \nn\\ S_{2D} &=& \int_{M_3} (C_3 + b_2 \wedge dB). \label{4.3} \end{eqnarray} Since $S_{2D}$ does not include $F_{ij}$, it is manifestly duality invariant. Solving the equation of motion for $F_{ij}$ in order to rewrite the action $S_1$ in terms of $B$ instead of $F_{ij}$, we arrive at the dual action $S_D$ of (\ref{4.1}) \begin{eqnarray} S_D = - \int_{M_3} d^3 \sigma e^{-\frac{1}{4} \phi} \sqrt{- \det G'_{ij}} + \int_{M_3} ( C_3 + b_2 \wedge dB ), \label{4.4} \end{eqnarray} where we have defined as \begin{eqnarray} G'_{ij} = G_{ij} + e^{\frac{3}{2} \phi} (\partial_i B + C_i) (\partial_j B + C_j). \label{4.5} \end{eqnarray} Incidentally, in order to derive the dual action we have used the mathematical formulas holding for $3 \times 3$ matrices \begin{eqnarray} \det ( G_{ij} + A_i A_j ) &=& (\det G_{ij}) \times ( 1 + G^{ij} A_i A_j ), \nn\\ \det ( G_{ij} + {\cal F}_{ij} ) &=& (\det G_{ij}) \times ( 1 + \frac{1}{2} G^{ij} G^{kl} {\cal F}_{ik} {\cal F}_{jl} ), \label{4.6} \end{eqnarray} where ${\cal F}_{ij} = -{\cal F}_{ji}$. Eq.(\ref{4.5}) implies the identification $dB + C_1 = e^{-\frac{3}{4} \phi} E^{11}$, in other words, identifying the world-volume scalar with the coordinate of a compact extra target-space dimension. Consequently, the Dirac-Born-Infeld action in Eq.(\ref{4.4}) takes the standard form for the induced metric of the M2-brane. The remaining work is to show that the second term in the right hand side of Eq.(\ref{4.4}) equals to the expression for the Wess-Zumino term of the super M2-brane. Taking the exterior derivative and using the relation (\ref{4.2}) we can evaluate this term as follows: \begin{eqnarray} d \Omega_D &\equiv& d ( C_3 + b_2 \wedge dB ) \nn\\ &=& R_{(4)} + (dB + C_1) \wedge H_3 \nn\\ &=& \frac{i}{2} e^{-\frac{1}{4} \phi} \bar{E} \wedge \gamma_{ab} E \wedge E^b \wedge E^a + \frac{1}{12} e^{-\frac{1}{4} \phi} \bar{E} \wedge \gamma_{abc} \Lambda E^c \wedge E^b \wedge E^a \nn\\ &{}& + (i e^{-\frac{1}{4} \phi} \bar{E} \wedge \gamma_{11} \gamma_a E \wedge E^a - \frac{1}{2} e^{-\frac{1}{4} \phi} \bar{E} \wedge \gamma_{ab} \gamma_{11} \Lambda E^b \wedge E^a) \wedge E^{11} \nn\\ &=& \frac{i}{2} e^{-\frac{1}{4} \phi} \bar{E} \wedge \gamma_{\hat{a}\hat{b}} E \wedge E^{\hat{b}} \wedge E^{\hat{a}} \nn\\ &{}& + \frac{1}{12} e^{-\frac{1}{4} \phi} \bar{E} \wedge \gamma_{abc} \Lambda E^c \wedge E^b \wedge E^a - \frac{1}{2} e^{-\frac{1}{4} \phi} \bar{E} \wedge \gamma_{ab} \gamma_{11} \Lambda E^b \wedge E^a \wedge E^{11} \nn\\ &\equiv& d \Omega_1 + d \Omega_2 , \label{4.7} \end{eqnarray} where $\hat{a} \equiv (a, 11)$ denotes 11 dimensional index, and $d \Omega_1$ and $d \Omega_2$ are respectively defined as $d \Omega_1 = \frac{i}{2} e^{-\frac{1}{4} \phi} \bar{E} \wedge \gamma_{\hat{a}\hat{b}} E \wedge E^{\hat{b}} \wedge E^{\hat{a}}$ and $d \Omega_2 = \frac{1}{12} e^{-\frac{1}{4} \phi} \bar{E} \wedge \gamma_{abc} \Lambda E^c \wedge E^b \wedge E^a - \frac{1}{2} e^{-\frac{1}{4} \phi} \bar{E} \wedge \gamma_{ab} \gamma_{11} \Lambda E^b \wedge E^a \wedge E^{11}$. Accordingly, the dual action (\ref{4.4}) of the super D2-brane can be written as \begin{eqnarray} S_D = - \int_{M_3} d^3 \sigma e^{-\frac{1}{4} \phi} \sqrt{- \det G'_{ij}} + \int_{M_3} \Omega_1 + \int_{M_3} \Omega_2, \label{4.8} \end{eqnarray} where $G'_{ij} = E_i^{\hat{a}} E_j^{\hat{b}} \eta_{\hat{a}\hat{b}}$ \footnote{$\eta_{\hat{a}\hat{b}}$ is the 11 dimensional flat metric defined as diag$(-,+, \ldots,+,+)$.}. In order to make (\ref{4.8}) coincide with the super M2-brane action where there is no dilaton, we should remove the dilaton superfield except $\Omega_2$ by rescaling the other superfields. It then turns out that the following rescaling makes a job \begin{eqnarray} E^{\hat{a}} \rightarrow e^{\frac{1}{12} \phi} E^{\hat{a}}, \ E^{\alpha} \rightarrow e^{\frac{1}{24} \phi} E^{\alpha}, \ \Lambda \rightarrow e^{\frac{1}{24} \phi} \Lambda. \label{4.9} \end{eqnarray} After this rescaling, the dual action takes the form \begin{eqnarray} S_D = - \int_{M_3} d^3 \sigma \sqrt{- \det G^{11}_{ij}} + \int_{M_3} \Omega^{11} + \int_{M_3} \Omega', \label{4.10} \end{eqnarray} where $d \Omega^{11} = \frac{i}{2} \bar{E} \wedge \gamma_{\hat{a}\hat{b}} E \wedge E^{\hat{b}} \wedge E^{\hat{a}}$ and $d \Omega' = \frac{1}{12} e^{\frac{1}{12} \phi} \bar{E} \wedge \gamma_{abc} \Lambda E^c \wedge E^b \wedge E^a - \frac{1}{2} e^{\frac{1}{12} \phi} \bar{E} \wedge \gamma_{ab} \gamma_{11} \Lambda E^b \wedge E^a \wedge E^{11}$. Provided that we neglect the last term $\int_{M_3} \Omega'$ in the Wess-Zumino action for the time being, this dual action is nothing but the standard form of the M2-brane action \cite{Sezgin}. Thus, we have proved that the super D2-brane action in a type IIA supergravity background is transformed to the super M2-brane action with a circular compactified 11th dimension in eleven dimensional supergravity background through a duality transformation of the world-volume gauge field as expected from IIA/M-duality. In fact, we can show the rescaling (\ref{4.9}) that was required arises the well-known relation between the dilaton and the radius of a compactified circle in the 11th direction as follows. {}From Eq.(\ref{4.5}) and the rescaling, we have a relation \begin{eqnarray} G^{11}_{ij} = e^{-\frac{1}{6} \phi} G_{ij} + e^{\frac{4}{3} \phi} (\partial_i B + C_i) (\partial_j B + C_j). \label{4.11} \end{eqnarray} Next in order to get the desired relation we have to rewrite the Einstein metric $G_{ij}$ in terms of the string metric $G^{(s)}_{ij} \equiv e^{\frac{1}{2} \phi} G_{ij}$ whose result is given by \begin{eqnarray} G^{11}_{ij} = e^{-\frac{2}{3} \phi} G^{(s)}_{ij} + e^{\frac{4}{3} \phi} (\partial_i B + C_i) (\partial_j B + C_j). \label{4.12} \end{eqnarray} This correctly yields the relation between the 11 dimensional metric and the 10 dimensional string metric \cite{Witten2}. Particularly, the coefficient in front of $(\partial B)^2$ gives us the relation that $R_{11} = e^{\frac{2}{3} \phi}$ \cite{Witten2}, where $R_{11}$ is the radius of a compactified circle in the 11th direction. {}Finally, let us comment the term $\int_{M_3} \Omega'$ in the Wess-Zumino action. Since $\Lambda_{\alpha} \equiv \frac{1}{2} \partial_{\alpha} \phi$ as defined in Sec.2, this term never decouples from the theory unless the dilaton is a constant. This fact is physically reasonable from the fact that $\Lambda_{\alpha}$ is the dilatino which is a superpartner of the dilaton. Thus, at first sight, since the dilaton and the dilatino are contained, the dual action (\ref{4.10}) appears more general than the standard M2-brane action formulated in an 11-dimensional supergravity background whose bosonic fields are the metric and a 3-form gauge potential. But this is illusory since it has been already shown \cite{Duff} that by suitable redefinitions of the superconnection and parts of the supervielbein the terms involving the dilatino can be set to zero in 11 dimensions. Accordingly, we have precisely shown that the dual of the super D2-brane action can be identified as the M2-brane action with a circular 11th dimension. \section{The super D3-brane} In this section let us show the $SL(2, Z)$ self-duality of the super D3-brane action in a general type IIB supergravity background with non-constant dilaton and axion in both classically and quantum mechanically exact manner. In the case of non-constant dilaton and axion there appear several new features as we will discuss in this section. {}From Eqs.(1), (2) and (3), the super D3-brane action is given by \begin{eqnarray} S &=& S_{DBI} + S_{WZ}, \nonumber \\ S_{DBI} &=& - \int_{M_4}d^4\sigma \sqrt{-\det(G_{ij}+ e^{-\frac{\phi}{2}} {\cal F}_{ij})}, \nonumber \\ S_{WZ} &=& \int_{M_4}(C_4 + C_2\wedge {\cal F} + \frac{1}{2}C_0 {\cal F}\wedge {\cal F}) = \int_{M_{4}}\Omega_{4}. \label{5-1} \end{eqnarray} The constraints (12) on the field strengths of NS-NS and R-R form-potentials are given by \begin{eqnarray} H_3 &= & db_2 = ie^{\frac{\phi}{2}}\bar{E}\wedge\hat{E} \wedge{\cal K}E + \frac{1}{2}e^{\frac{\phi}{2}} \bar{E} \wedge \gamma_{ab}{\cal K}\Lambda E^b\wedge E^a,\nonumber \\ R_{(1)} &=& dC_0 = 2e^{-\phi}\bar{E}{\cal E}\Lambda, \nonumber\\ R_{(3)} &=& dC_2 - H_3 C_0 = -ie^{-\frac{\phi}{2}}\bar{E}\wedge\hat{E}\wedge{\cal I}E + \frac{1}{2}e^{-\frac{\phi}{2}}\bar{E} \wedge \gamma_{ab} {\cal I}\Lambda E^b\wedge E^a, \nonumber \\ R_{(5)} &=& dC_4 - H_3 \wedge C_2 = \frac{i}{6}\bar{E}\wedge \gamma_{abc}{\cal E}E\wedge E^c\wedge E^b\wedge E^a. \label{5-2} \end{eqnarray} Note that in the case of the non-constant dilaton and axion the Abelian worldvolume gauge field enters into the axion term in the Wess-Zumino term only through the gauge invariant and supersymmetric ${\cal F}$. Therefore this term is no longer a topological term and the symmetry under the constant shift of the axion seems to be lost. However, as we will see in the next subsection, it is this form of the Wess-Zumino action that ensures the NS-NS and the R-R 2-form potentials transform as a doublet under the $SL(2,R)$ duality transformation. \subsection{The classical analysis} In this subsection we show that the super D3-brane action in a general type IIB supergravity background is classically self-dual. For this we first add a Lagrangian multiplier term \cite{Tseytlin} \begin{eqnarray} S_{\tilde{H}} = \int_{M_4}d^4\sigma\frac{1}{2}\tilde{H}^{ij} (F_{ij}-2\partial_{i} A_{j} ), \label{5-3} \end{eqnarray} to the above action (\ref{5-1}) and solve the equation of motion for $A_i$ by $\tilde{H}^{ij} = \epsilon^{ijkl}\partial _k B_l$ with a dual vector potential $B_i$. Then substitution of this solution into the action (\ref{5-1}) leads to an action $S'=S_{1}+S_{2D}$, where \begin{eqnarray} S_1 &=& \int_{M_4} \left[-d^4\sigma \sqrt{-\det(G_{ij}+ e^{-\frac{\phi}{2}} {\cal F}_{ij})} + (C_2 + \tilde{F})\wedge {\cal F} + \frac{1}{2}C_0 {\cal F}\wedge {\cal F} \right], \nonumber \\ S_{2D} &=& \int_{M_4}(C_4 + \tilde{F}\wedge b_{2}), \label{5-4} \end{eqnarray} where $\tilde{F}=dB$. Next we solve the equation of motion for $F_{ij}$ and substitute its solution into the action $S'$. Following \cite{Tseytlin}, we perform the calculation in the specific local Lorentz frame where $G_{ij}=\eta_{ij}=$ diag$(-1,1,1,1)$ and ${\cal F}$ has the following block diagonal form \begin{equation} {\cal F}_{ij} = \left( \begin{array}{cccc} 0 & {\cal F}_{01} & 0 & 0 \\ -{\cal F}_{01} &0 & 0 & 0 \\ 0 & 0 & 0 & {\cal F}_{23} \\ 0 & 0 & -{\cal F}_{23} & 0 \end{array} \right). \label{5-5} \end{equation} Then the action (\ref{5-4}) is written in this local Lorentz frame as \begin{eqnarray} S^{\prime}&=&\int_{M_{4}}d^{4}\sigma [-\sqrt{(1-e^{-\phi}{\cal F}_{01}^{2}) (1+e^{-\phi}{\cal F}_{23}^{2})} + (C_2 +\tilde{F})_{23} {\cal F}_{01} + (C_2 +\tilde{F})_{01}{\cal F}_{23} \nonumber \\ &+& C_{0}{\cal F}_{01}{\cal F}_{23} + C_{0123} + \tilde{F}_{01}b_{23} +\tilde{F}_{23}b_{01} ]. \label{5-6} \end{eqnarray} Solving the equation of motion for $F_{01}$ and $F_{23}$ yields \begin{eqnarray} {\cal F}_{01} = -\frac{1}{A} \left(e^{\phi}C_{0}b + a \sqrt{\frac {A-b^{2}}{A+a^{2}}} \right), \ {\cal F}_{23} = -\frac{1}{A} \left(e^{\phi}a - b\sqrt{\frac{A+a^{2}} {A-b^{2}}} \right), \label{5-7} \end{eqnarray} where \begin{eqnarray} A\equiv e^{-\phi}+e^{\phi}C_{0}^{2}, \ a\equiv (C_2 +\tilde{F})_{23}, \ b\equiv (C_2 +\tilde{F})_{01}. \label{5-8} \end{eqnarray} Inserting these solutions into the action (\ref{5-6}) and returning to the general coordinate reference frame, we arrive at the dual action $S_{D}$ in a simple form \begin{eqnarray} S_D &=& -\int_{M_4}\sqrt{-\det \left[G_{ij} + \frac{1}{\sqrt{e^{-\phi}+e^{\phi}C_{0}^{2}}}(\tilde{F}_{ij}+C_{2ij})\right]} + \int_{M_4}\Omega_D, \nonumber \\ \Omega_D &=& C_4 + b_2\wedge \tilde{F} -\frac{1}{2}\frac{e^{\phi}C_{0}}{e^{-\phi}+e^{\phi}C_{0}^{2}} (\tilde{F}+C_{2})\wedge (\tilde{F}+C_{2}). \label{5-9} \end{eqnarray} Now if we define a new primed dilaton, axion and form-potentials which are of the form, \begin{eqnarray} e^{-\phi^{\prime}}=\frac{1}{e^{-\phi}+e^{\phi}C_{0}^{2}}, \nonumber \\ C_{0}^{\prime}=\frac{e^{\phi}C_0}{e^{-\phi}+e^{\phi}C_{0}^{2}}, \label{5-10} \end{eqnarray} and \begin{eqnarray} b_2^{\prime} = -C_2, \ C_2^{\prime} = b_2, \ C_4^{\prime} = C_4 - b_2 \wedge C_2, \label{5-11} \end{eqnarray} then the dual action can be written as \begin{eqnarray} S_D &=& -\int_{M_4}d^{4}\sigma\sqrt{-\det(G_{ij} + e^{-\frac{\phi^{\prime}}{2}}\tilde{{\cal F}}_{ij}^{\prime})} \nonumber \\ &{}& + \int_{M_4}(C_4^{\prime} + C_2^{\prime} \wedge \tilde{{\cal F}}^{\prime} +\frac{1}{2}C_{0}^{\prime} \tilde{{\cal F}}^{\prime}\wedge\tilde{{\cal F}}^{\prime}), \label{5-12} \end{eqnarray} where $\tilde{{\cal F}}^{\prime}=\tilde{F}-b_{2}^{\prime}$. The resulting action is completely the same form as the original action (\ref{5-1}) with which we have started. This means that under the transformations (\ref{5-10}), (\ref{5-11}) and \begin{eqnarray} \tilde{F}\rightarrow F, \ F\rightarrow \tilde{F}, \label{5-13} \end{eqnarray} the action (\ref{5-1}) and (\ref{5-12}) are classically equivalent. In order to establish the $SL(2,R)$ self-duality of the super D3-brane action in a general type IIB supergravity background, we have to answer the following two problems. The first problem is to clarify the transformation properties of form-potentials under the $SL(2,R)$ duality transformation. The second one is whether the type IIB constraint equations (\ref{5-2}) are preserved under the $SL(2,R)$ duality transformation. First of all, let us introduce complex variable $\tau$ as in Sec.3, \begin{eqnarray} \tau = C_{0} + ie^{-\phi}, \label{5-14} \end{eqnarray} then the transformation (\ref{5-10}) is written as $\tau \rightarrow \tau^{\prime}=-\frac{1}{\tau}$, which is an element of the $SL(2,R)$ with the transformation matrix $S= \left( \begin{array}{cc} 0 & 1 \\ -1 & 0 \end{array} \right)$. According to the transformation rules (\ref{A-10}) and (\ref{A-11}) the transformations (\ref{5-11}) and (\ref{5-13}) imply that $(b_{2}, -C_{2})$ and $(\tilde{F}, -F)$ are expected to form $SL(2,R)$ doublets which transform like B in (\ref{A-11}) and A in (\ref{A-10}), respectively. Next we must find the $SL(2,R)$ element $\tau \rightarrow \tau^{\prime}=\tau +1$ which corresponds to the shift of the axion $C_{0}\rightarrow C_{0}+1$. It is easily shown that as in the super D-string action the super D3-brane action (\ref{5-1}) is invariant under the following two transformations \begin{eqnarray} C_{0} &\rightarrow& C_{0}+1, \nonumber \\ b_{2} &\rightarrow& b_{2} , \ C_{2}\rightarrow C_{2} -{\cal F} =C_{2}+b_{2}-F, \nonumber \\ C_{4} &\rightarrow& C_{4} + \frac{1}{2}{\cal F}\wedge {\cal F}, \nonumber \\ F &\rightarrow& F, \label{5-15} \end{eqnarray} and \begin{eqnarray} C_{0} &\rightarrow& C_{0}+1, \nonumber \\ b_{2} &\rightarrow& b_{2} , \ C_{2}\rightarrow C_{2}+b_{2}, \nonumber \\ C_{4} &\rightarrow& C_{4} + \frac{1}{2}b_{2}\wedge b_{2}, \nonumber \\ F &\rightarrow& F. \label{5-16} \end{eqnarray} The super D3-brane action (\ref{5-1}) is exactly invariant under (\ref{5-15}) and invariant only up to the topological term $\frac{1}{2}F\wedge F$ under (\ref{5-16}). It is also easily shown that the type IIB constraint equations (\ref{5-2}) are invariant under these transformation. Thus the transformations (\ref{5-15}) and (\ref{5-16}) are real symmetries of the theory. These transformations are also consistent with the postulate that $(b_{2}, -C_{2})$ and $(\tilde{F}, -F)$ form $SL(2,R)$ doublets which transform like B in (\ref{A-11}) and A in (\ref{A-10}), respectively. The transformation (\ref{5-15}) is an exact symmetry of the theory and seems to be more attractive than another one (\ref{5-16}). It may be interpreted in such a way that there may exist a manifestly self-dual formulation of the super D3-brane which are broken by some gauge fixing to the present formulation of super D3-brane action. Unfortunately, however, as discussed in the next subsection, the transformation (\ref{5-15}) does not satisfy the consistency condition of the constructive relation (\ref{5-25}) in the next subsection. Therefore we will not consider this transformation any more in this paper. At this point we should stress the following fact. In the case of constant or vanishing dilaton and axion the following form is sometimes taken as the Wess-Zumino action; \begin{eqnarray} S_{WZ}=\int(C_{4} + C_{2}\wedge{\cal F} +\frac{1}{2} C_0 F\wedge F), \nonumber \end{eqnarray} Although this action is trivially invariant up to the topological term $\frac{1}{2}F\wedge F$ under the shift of the axion $C_{0}$, the R-R 2-form potential $C_{2}$ can not transform in such a way to form $SL(2,R)$ doublet with the NS-NS 2-form potential $b_{2}$ and preserve the constraint equation on the R-R 5-form strength $R_{(5)}$. Therefore with this Wess-Zumino action the $SL(2,R)$ duality symmetry would be lost even in the case of constant or zero dilaton and axion. The general $SL(2,R)$ transformations are generated by $\tau\rightarrow -\frac{1}{\tau}$ and $\tau\rightarrow \tau +1$. Here we recaptulate the $SL(2,R)$ transformation rule for various fields in the theory for later use in the next subsection. Corresponding to the $SL(2,R)$ transformation of the complex variable $\tau$ \begin{eqnarray} \tau \rightarrow \tau^{\prime}=\frac{a\tau + b}{c\tau + d}, \ ad-bc=1, \label{5-17} \end{eqnarray} which in terms of $C_{0}$ and $\phi$ is equivalent to \begin{eqnarray} C_{0}^{\prime}&=&\frac{(aC_{0}+b)(cC_{0}+d)+ace^{-2\phi}} {(cC_{0}+d)^2 +c^2 e^{-2\phi}}, \nonumber \\ e^{-\phi^{\prime}}&=&\frac{e^{-\phi}} {(cC_{0}+d)^2 +c^2 e^{-2\phi}}, \label{5-18} \end{eqnarray} we define the matrix $S$ as \begin{equation} S = \left( \begin{array}{cc} a & b \\ c & d \end{array} \right), \ (S^{T})^{-1} = \left( \begin{array}{cc} d & -c \\ -b & a \end{array} \right). \label{5-19} \end{equation} Then it is expected that $\tilde{F}, F, b_{2}, C_{2}$ and $C_{4}$ transform in the following manner; \begin{eqnarray} \tilde{F}\rightarrow \tilde{F}^{\prime}=a\tilde{F}-bF, \ F\rightarrow -c\tilde{F} + dF, \label{5-20} \end{eqnarray} \begin{eqnarray} b_{2}\rightarrow b_{2}^{\prime}=db_{2}+cC_{2}, \ C_{2}\rightarrow C_{2}^{\prime}=bb_{2}+aC_{2}, \label{5-21} \end{eqnarray} and \begin{eqnarray} C_{4}\rightarrow C_{4}^{\prime}=C_{4}+\frac{bd}{2} b_{2}\wedge b_{2} + bcb_{2}\wedge C_{2} +\frac{ac}{2}C_{2}\wedge C_{2}. \label{5-22} \end{eqnarray} The transformation of $C_{4}$ is determined by the requirement that $R_{(5)}$ is invariant under the $SL(2,R)$ transformations (\ref{5-16}) and (\ref{5-20}). The $SL(2,R)$ transformations (\ref{5-17})-(\ref{5-22}) are just the same as those in \cite{GG}, but the authors of \cite{GG} gave neither the transformation of $C_{4}$ nor $SO(2)$ spinor-rotation (\ref{A-16}) since they considered only the bosonic D3-brane action. The $SL(2,R)$ transformation (\ref{5-22}) of $C_{4}$ and the $SO(2)$ spinor-rotaition (\ref{A-16}) are indispensable ingredients for the type IIB supergravity constraints to be invariant under the $SL(2,R)$ duality transformation. Next we turn to the problem of the invariance (or covariance) of the type IIB supergravity constraints under the $SL(2,R)$ duality transformation. In Appendix it is shown that all the type IIB supergravity constraint equations (of course including the torsion constraints) are invariant under general $SL(2,R)$ duality transformations (\ref{5-17}), (\ref{5-21}), (\ref{5-22}) and the $SO(2)$ spinor-rotation (\ref{A-16}). Finally we note that, since the invariance under $\tau\rightarrow \tau + 1$ holds only up to the topological term $\frac{1}{2}F\wedge F$, the $SL(2,R)$ invariance of the theory would be broken to the $SL(2,Z)$ invariance in a quantum theory. \subsection{The exact analysis} Let us turn to the other analysis of duality condition which was initiated by Gaillard and Zumino \cite{GZ1} and extended by several authors \cite{GZ2,GR,GG,IIK1}. Let us define the dual field strength $K_{ij}$ by the following constructive relation; \begin{eqnarray} \ast K^{ij} = \frac{\partial L}{\partial F_{ij}}, \ \frac{\partial F_{kl}}{\partial F_{ij}}= \delta_{k}^{i} \delta_{l}^{j}-\delta_{l}^{i} \delta_{k}^{j}, \label{5-25} \end{eqnarray} where the Hodge dual components $\ast K_{ij}$ for the antisymmetric tensor $K_{ij}$ are defined by \begin{eqnarray} \ast K_{ij} \equiv \frac{1}{2}\epsilon_{ij} ^{kl} K_{kl}, \ \ast\ast K_{ij}=-K_{ij}, \label{5-26} \end{eqnarray} where $\epsilon_{ijkl}$ is the Levi-Civita symbol in 4 dimensions, $\epsilon^{0123}=1$ and the signature of $G_{ij}$ is $(-,+,+,+)$. At the classical level $K_{ij}$ introduced here is equivalent to the dual variable $-\tilde{F}_{ij}$ in the last subsection. Therefore it is natural to assume that $(K_{ij}, F_{ij})$ transforms as Eq.(\ref{5-20}); for the infinitesimal transformation with $S= \left( \begin{array}{cc} 1+\alpha & \beta \\ \gamma & 1-\alpha \end{array} \right)$ it is given by \begin{eqnarray} \delta K_{ij}= +\alpha K_{ij} + \beta F_{ij}, \ \delta F_{ij}= -\alpha F_{ij} + \gamma K_{ij}, \label{5-27} \end{eqnarray} The sufficient condition for the consistency of the constructive relation (\ref{5-25}), the invariance of the field equation and the invariance of the (world-volume) energy-momentum tensor under the $SL(2,R)$ duality transformation is that the following equation is satisfied for arbitrary $SL(2,R)$ parameters $\alpha, \beta$ and $\gamma$; \begin{eqnarray} \frac{\gamma}{4}*K^{ij}K_{ij} -\frac{\beta}{4}*F^{ij}F_{ij} -\frac{\alpha}{2}*K^{ij}F_{ij} + \delta_{\Phi} L =0, \label{5-28} \end{eqnarray} where $\delta_{\Phi}$ means to take the $SL(2,R)$ variation for all fields except the world volume Abelian gauge field. This equation is derived by the similar argument in \cite{IIK1} where a similar equation was derived for the special case of $\alpha=0$ and $\beta=-\gamma$ ($SO(2)$ duality transformation). We call Eq.(\ref{5-28}) the Gaillard-Zumino (G-Z) duality condition. In \cite{IIK1} it has been shown that the G-Z condition is actually the necessary and sufficient condition in order that one can define off-shell (non-local) duality transformation for the $U(1)$ gauge potential itself under which the action is invariant up to some surface terms. Therefore if one can show for an action to satisfiy the G-Z condition, then one establishes the exact self-duality of the theory described by this action without resort to any classical approximation. It has been shown that the super D3-brane action on a flat \cite{IIK2}, an $AdS_5 \times S^5$ \cite{Kimura} and a general type IIB supergravity backgrounds with constant dilaton and axion \cite{KO} indeed satisfies the G-Z condition. In this subsection we show that the D3-brane action on the most general type IIB supergravity background described by the action (\ref{5-1}) and constraints (\ref{5-2}) satisfies the G-Z duality condition under the $SL(2,R)$ duality transformations (\ref{5-17})-(\ref{5-22}) and the $SO(2)$ spinor-rotation (\ref{A-16}). Before we enter the detailed discussions we should note that the transformation (\ref{5-15}) does not satisfiy the consistency condition for the constructive relation (\ref{5-25}). The consistency condition of the constructive relation demands the equation \begin{eqnarray} \delta *K^{ij}= \frac{1}{2}\frac{\partial^{2}L} {\partial F_{kl}\partial F_{ij}}\delta F_{kl} + \frac{\partial^{2}L}{\partial \Phi \partial F_{ij}}\delta \Phi, \nonumber \end{eqnarray} to be satisfied. A simple calculation leads us to a contradictory result; $*F^{ij}=0$. It seems to be curious that the exact symmetry of Lagrangian does not satisfy the consistency condition. The super D3-brane Lagrangian (\ref{5-1}) in terms of component fields is written as \begin{eqnarray} L = L_{DBI} + \epsilon^{ijkl}(\frac{1}{24}C_{ijkl} + \frac{1}{4}C_{ij}{\cal F}_{kl} + \frac{1}{8}C_{0} {\cal F}_{ij}{\cal F}_{kl}), \label{5-29} \end{eqnarray} \begin{eqnarray} L_{DBI}&=& -\sqrt{-\det(G_{ij} + e^{-\frac{\phi}{2}}{\cal F}_{ij})} \nonumber \\ &=& -\sqrt{-G}\sqrt{1 + \frac{e^{-\phi}}{2}{\cal F}_{ij}{\cal F}^{ij} - \frac{e^{-2\phi}}{16}({\cal F}_{ij}\ast {\cal F}^{ij})^{2}}, \label{5-30} \end{eqnarray} where $G = \det G_{ij}$. Then the dual field strength defined by (\ref{5-25}) is given by \begin{eqnarray} *K^{ij}&=& \frac{\partial L_{DBI}}{\partial F_{ij}} + *C^{ij} + C_{0}*{\cal F}^{ij}, \nonumber \\ K_{ij}&=&-(*\frac{\partial L_{DBI}}{\partial F})_{ij} + C_{ij} + C_{0}{\cal F}_{ij}. \label{5-31} \end{eqnarray} The infinitesimal $SL(2,R)$ transformations with $S= \left( \begin{array}{cc} 1+\alpha & \beta \\ \gamma & 1-\alpha \end{array} \right)$ of various fields in our theory are given by \begin{eqnarray} \delta C_{0}&=&2\alpha C_{0} + \beta - \gamma(C_{0}^2 - e^{-2\phi}), \nonumber \\ \delta \phi &=& 2\gamma C_{0} -2\alpha, \label{5-32} \end{eqnarray} \begin{eqnarray} \delta K_{ij}= +\alpha K_{ij} + \beta F_{ij}, \ \delta F_{ij}= -\alpha F_{ij} + \gamma K_{ij}, \label{5-33} \end{eqnarray} \begin{eqnarray} \delta b_{ij}= -\alpha b_{ij} + \gamma C_{ij}, \ \delta C_{ij}= +\alpha C_{ij} + \beta b_{ij}, \label{5-34} \end{eqnarray} and \begin{eqnarray} \delta{C_{4}}=\frac{\beta}{2}b_{2}\wedge b_{2} + \frac{\gamma}{2} C_{2}\wedge C_{2}. \label{5-35} \end{eqnarray} The infinitesimal $SO(2)$ spinor-rotation is given by \begin{eqnarray} \delta \theta&=&-\frac{\gamma e^{-\phi}}{2}{\cal E}\theta, \ \delta \partial_{\alpha}=\frac{\gamma e^{-\phi}}{2} ({\cal E}\partial)_{\alpha}, \nonumber \\ \delta E &=&\frac{\gamma e^{-\phi}}{2}{\cal E}E, \ \delta \bar{E} =-\frac{\gamma e^{-\phi}}{2}\bar{E}{\cal E}. \label{5-36} \end{eqnarray} The infinitesimal $SO(2)$ spinor-rotation (\ref{5-36}) ensures the invariance of the type IIB supergravity constraints. Under these infinitesimal $SL(2,R)$ transformations $\delta_{\Phi}L$ is given by \begin{eqnarray} \delta_{\Phi}L &=& \frac{1}{2}\frac{\partial L}{\partial b_{ij}}\delta b_{ij} + \frac{1}{2}\frac{\partial L}{\partial C_{ij}}\delta C_{ij} + \frac{\partial L}{\partial\phi}\delta \phi + \frac{\partial L}{\partial C_{0}}\delta C_{0} + \frac{1}{24}\frac{\partial L}{\partial C_{ijkl}}\delta C_{ijkl}. \label{5-37} \end{eqnarray} The G-Z duality condition demands that Eq.(\ref{5-28}) must hold for arbitrary variations $\alpha, \beta$ and $\gamma$. Therefore coefficients of $\alpha, \beta$ and $\gamma$ should identically vanish, respectively. Substituting (\ref{5-37}) with (\ref{5-32})-(\ref{5-35}) into (\ref{5-27}) we obtain following three equatios for $\alpha, \beta, \gamma$ coefficients; \begin{eqnarray} -\frac{1}{4}*K^{ij}K_{ij}-\frac{1}{2}\frac{\partial L} {\partial b_{ij}}b_{ij}+\frac{1}{2}\frac{\partial L} {\partial C_{ij}}C_{ij} -2\frac{\partial L}{\partial \phi} +2C_{0}\frac{\partial L}{\partial C_{0}}=0, \label{5-38} \end{eqnarray} \begin{eqnarray} -\frac{1}{4}*F^{ij}F_{ij} +\frac{1}{2}\frac{\partial L} {\partial C_{ij}}b_{ij}+\frac{\partial L} {\partial C_{0}} +\frac{1}{8}\epsilon^{ijkl}b_{ij}b_{kl}=0, \label{5-39} \end{eqnarray} and \begin{eqnarray} \frac{1}{4}*K^{ij}K_{ij} +\frac{1}{2}\frac{\partial L} {\partial b_{ij}}C_{ij} +2C_{0}\frac{\partial L} {\partial \phi} -(C_{0}^{2}-e^{-2\phi})\frac{\partial L} {\partial C_{0}} +\frac{1}{8}\epsilon^{ijkl}C_{ij}C_{kl}=0, \label{5-40} \end{eqnarray} respectively. Eq.(\ref{5-39}) ($\beta$ coefficient) is almost trivially satisfied. Eq.(\ref{5-38}) ($\alpha$ coefficient) is also easily shown to be satisfied, using the following identity \begin{eqnarray} \frac{\partial L_{DBI}}{\partial \phi}=-\frac{1}{4} \frac{\partial L_{DBI}}{\partial F_{ij}}{\cal F}_{ij}, \label{5-41} \end{eqnarray} which is derived by the fact that the dilaton $\phi$ is contained in $L_{DBI}$ in the form of $e^{-\frac{\phi}{2}}{\cal F}$. The $\gamma$ coefficient (\ref{5-40}) is reduced after straightforward but somewhat lengthy calculations to the following equation \begin{eqnarray} \frac{1}{2}\epsilon^{ijkl}(\frac{\partial L_{DBI}}{\partial F_{ij}} \frac{\partial L_{DBI}}{\partial F_{kl}} + e^{-2\phi}{\cal F}_{ij}{\cal F}_{kl}) = 0. \label{5-42} \end{eqnarray} This is the only equation which explicitly depends on the specific form of $L_{DBI}$. Using the explicit expression (\ref{5-30}) for $L_{DBI}$, it is easily shown that the equation (\ref{5-41}) is satisfied. Therefore we have shown that the super D3-brane action (\ref{5-1}) in the most general type IIB supergravity background indeed satisfies the Gaillard-Zumino duality condition. In \cite{IIK2} it has been shown that the super D3-brane action in a flat background with constant dilaton and axion which satisfies the Gaillard-Zumino condition is pseudo-invariant under the $SL(2,R)$ (non-local) duality transformation of the world-volume Abelian gauge field in both the Lagrangian and the Hamiltonian formalism. According to the theorem proved in \cite{IIK1} our above result strongly suggests that the super D3-brane action in the most general type IIB supergravity background is also exactly self-dual without resort to any semi-classical approximation. It would be an interesting problem to establish it in both the Lagrangian and the Hamiltonian formalism. \section{The super D4-brane} In this section let us start with the super D4-brane action and perform a duality transformation of the world-volume gauge field to reach the action obtained by the double-dimensional reduction of the super M5-brane \cite{Aganagic3, Mario}. The method we consider is analogous to the one adopted in Section 4, so we shall follow a similar path of argument as in the super D2-brane. Like the super D2-brane, the analysis in this section is purely $\it{classical}$. This time, from (\ref{2.1}) and (\ref{2.2}) the super D4-brane action with a Lagrange multiplier term becomes \begin{eqnarray} S &=& S_{DBI} + S_{WZ} + S_{\tilde{H}}, \nn\\ S_{DBI} &=& - \int_{M_5} d^5 \sigma e^{\frac{1}{2} \phi} \sqrt{- \det ( G_{ij} + e^{-\frac{1}{2} \phi}{\cal F}_{ij} )}, \nn\\ S_{WZ} &=& \int_{M_5 = \partial M_6} ( C_5 + C_3 \wedge {\cal F} + \frac{1}{2} C_1 \wedge {\cal F} \wedge {\cal F}) = \int_{M_6} I_6, \nn\\ S_{\tilde{H}} &=& \int_{M_5} d^5 \sigma \frac{1}{2} \tilde{H}^{ij} ( F_{ij} - 2 \partial_i A_j ). \label{6.1} \end{eqnarray} And the constraints (\ref{2.11}) for type IIA on the field strengths are given by \begin{eqnarray} H_3 &=& db_2 = i e^{\frac{1}{2} \phi} \bar{E} \wedge \gamma_{11} \hat{E} \wedge E + \frac{1}{2} e^{\frac{1}{2} \phi} \bar{E} \wedge \gamma_{ab} \gamma_{11} \Lambda E^b \wedge E^a, \nn\\ R_{(2)} &=& dC_1 = i e^{-\frac{3}{4} \phi} \bar{E} \wedge \gamma_{11} E - \frac{3}{2} e^{-\frac{3}{4} \phi} \bar{E} \wedge \gamma_a \gamma_{11} \Lambda E^a, \nn\\ R_{(4)} &=& dC_3 + H_3 \wedge C_1 \nn\\ &=& \frac{i}{2} e^{-\frac{1}{4} \phi} \bar{E} \wedge \gamma_{ab} E \wedge E^b \wedge E^a + \frac{1}{12} e^{-\frac{1}{4} \phi} \bar{E} \wedge \gamma_{abc} \Lambda E^c \wedge E^b \wedge E^a, \nn\\ R_{(6)} &=& dC_5 + H_3 \wedge C_3 \nn\\ &=& \frac{i}{24} e^{\frac{1}{4} \phi} \bar{E} \wedge \gamma_{abcd} \gamma_{11} E \wedge E^d \wedge E^c \wedge E^b \wedge E^a \nn\\ &{}& + \frac{1}{240} e^{\frac{1}{4} \phi} \bar{E} \wedge \gamma_{a_1 \ldots a_5} \gamma_{11} \Lambda E^{a_5} \wedge \ldots \wedge E^{a_1}, \label{6.2} \end{eqnarray} from which $C_5$, $C_3$ and $C_1$ are determined. As usual, we take the variation with respect to $A_i$, which gives rise to the solution $\tilde{H}^{ij} = \frac{1}{6} \epsilon^{ijklm} K_{klm}$ with $K = d B$ with $B$ being a second rank tensor superfield. After substituting this solution into the action, we obtain the action $S' = S_1 + S_{2D}$ where $S_1$ and $S_{2D}$ are defined as \begin{eqnarray} S_1 &=& - \int_{M_5} d^5 \sigma e^{\frac{1}{2} \phi} \sqrt{- \det ( G_{ij} + e^{-\frac{1}{2} \phi} {\cal F}_{ij} )} + \int_{M_5} ( {\cal H} \wedge {\cal F} + \frac{1}{2} C_1 \wedge {\cal F} \wedge {\cal F}), \nn\\ S_{2D} &=& \int_{M_5} ( C_5 + K \wedge b_2 ), \label{6.3} \end{eqnarray} with ${\cal H} = K + C_3$. As before, $S_{2D}$ is unaffected under a duality transformation. Therefore a duality transformation amounts to solving the equation of motion for $F_{ij}$ in $S_1$ in order to rewrite the action in terms of $B$ (or its field strength $K$) instead of $F_{ij}$. Following the formula in ref.\cite{Aganagic2}, it is tedious but straightforward to derive the dual action $S_D = S_{1D} + S_{2D}$ where $S_{1D}$ is given by \begin{eqnarray} S_{1D} &=& - \int_{M_5} d^5 \sigma \left[ e^{\frac{1}{2} \phi} \sqrt{-G} \sqrt{1 + z_1 + \frac{z_1^2}{2} - z_2} - \frac{e^{\phi}}{8(1 + e^{\phi} C_1^2)} \epsilon_{ijklm} C^i \tilde{{\cal H}}^{jk} \tilde{{\cal H}}^{lm} \right], \label{6.4} \end{eqnarray} where \begin{eqnarray} z_1 &=& \frac{1}{2(-G)(1 + C_1^2)} tr (\tilde{G}\tilde{{\cal H}} \tilde{G}\tilde{{\cal H}}), \nn\\ z_2 &=& \frac{1}{4(-G)^2 (1 + C_1^2)^2} tr (\tilde{G}\tilde{{\cal H}} \tilde{G}\tilde{{\cal H}}\tilde{G}\tilde{{\cal H}} \tilde{G}\tilde{{\cal H}}), \nn\\ G &=& \det{G_{ij}}, \nn\\ \tilde{G}_{ij} &=& G_{ij} +C_i C_j, \nn\\ \tilde{{\cal H}}^{ij} &=& \frac{1}{6} \epsilon^{ijklm} {\cal H}_{klm}. \label{6.5} \end{eqnarray} Now let us turn our attention to $S_{2D}$. The conditions (\ref{6.2}) yield the equation \begin{eqnarray} d\Omega_D &\equiv& d (C_5 + K \wedge b_2) \nn\\ &=& R_{(6)} + (K + C_3) \wedge H_3 \nn\\ &=& \frac{i}{24} e^{\frac{1}{4} \phi} \bar{E} \wedge \gamma_{abcd} \gamma_{11} E \wedge E^d \wedge E^c \wedge E^b \wedge E^a - i e^{\frac{1}{2} \phi} \bar{E} \wedge \gamma_a \gamma_{11} E \wedge E^a \wedge {\cal H} \nn\\ &{}& + \frac{1}{240} e^{\frac{1}{4} \phi} \bar{E} \wedge \gamma_{a_1 \ldots a_5} \gamma_{11} \Lambda E^{a_5} \wedge \ldots \wedge E^{a_1} - \frac{1}{2} e^{\frac{1}{2} \phi} \bar{E} \wedge \gamma_{ab} \gamma_{11} \Lambda E^b \wedge E^a \wedge {\cal H} \nn\\ &\equiv& d \Omega_1 + d \Omega_2, \label{6.6} \end{eqnarray} where $d \Omega_1$ and $d \Omega_2$ are respectively defined as the first plus second terms independent of the dilatino $\Lambda$ and the third plus forth terms dependent on the dilatino in the third line of the above equation. As a result, we have the dual action of the super D4-brane in type IIA supergravity background \begin{eqnarray} S_D &=& - \int_{M_5} d^5 \sigma \left[ e^{\frac{1}{2}\phi} \sqrt{-G} \sqrt{1 + z_1 + \frac{z_1^2}{2} - z_2} - \frac{e^{\phi}}{8(1 + e^{\phi} C_1^2)} \epsilon_{ijklm} C^i \tilde{{\cal H}}^{jk} \tilde{{\cal H}}^{lm} \right] \nn\\ &{}& + \int_{M_5} (\Omega_1 + \Omega_2). \label{6.7} \end{eqnarray} As in the case of the super D2-brane in Sec.4, we can also set $\int_{M_5} \Omega_2$ to be zero in terms of redefinitions of the superconnection and the supervielbein. With this situation, if the dilaton is vanishing, this dual action of (\ref{6.7}) exactly reduces to our previous action \cite{KO}, while if the dilaton is vanishing and the background is in a flat Minkowski spacetime, it becomes the action considered in Ref.\cite{Aganagic2}. Then it is straightforward to show that each action is identical to its corresponding action which is obtained by the double-dimensional reduction of the super M5-brane \cite{Aganagic3, Mario} after rearranging the constant dilaton factor in a suitable way. Following the similar line of argument, it turns out that the double-dimensional reduction of the super M5-brane action coincides with the dual super D4-brane action even in a general type IIA supergravity background under consideration as suggested by the duality between M-theory and IIA superstring theory. \section{Discussions} In this paper, we have studied the properties of a duality transformation of super Dp-brane actions ($p = 1, 2, 3, 4$) in a general type II on-shell supergravity background, which have been constructed in Refs.\cite{Cederwall1, Cederwall2, Bergshoeff}. They have already been investigated in the case of a flat background with the zero or constant dilaton and axion \cite{Aganagic2} and in a type II on-shell supergravity background with the zero or constant dilaton and axion \cite{KO}. Our presentation in this paper is most general compared to the previous approaches so we believe that we have succeeded in showing that various duality symmetries in the super D-brane actions are indeed valid in a general type II supergravity background geometry. The main motivation of the present paper was to take account of the non-constant dilaton and axion superfields in carrying out a duality transformation of the super D-brane actions. To the best of our knowledge, this is the first attempt and consequently gives us several new fruitful results and insights as shown thus far. In particular, from the viewpoint of the world-volume field theory we have succeeded in showing that the dilaton and the axion as well as the two 2-form gauge fields, those are, the NS-NS 2-form and the R-R 2-form, are doublets of the $SL(2,R)$ Mobius group whereas the graviton and the 4-form gauge potential are singlets in the Einstein metric. These facts are of course familiar to us but have been so far proved only in the target space formulation while our presentation is purely from the world-volume field theory. The main new issue in the theory at hand is the existence of a non-trivial symmetry corresponding to a shift of the modulus field $\tau$. The discovery of this symmetry enables us to prove the expected duality relations of the super D-brane actions under the $SL(2,R)$ transformation, in particular, the self-duality of the super D3-brane action. Moreover, it was shown that the dual actions of the super D2-brane and D4-brane actions on a IIA supergravity background with the constant dilaton, respectively, have the standard forms which are expected from the dimensional reduction of the M2-brane and the M5-brane actions on the 11 dimensional supergravity background. Note that the dilaton and the dilatino appear in 10 dimensional spacetime through the Kaluza-Klein compacification on a circle from the M2- and M5-brane actions in an 11 dimensional supergravity background. Of course, in this paper our considerations have been focused on only the super D-brane actions \cite{Cederwall1, Cederwall2, Bergshoeff}, thus a detailed comparison between the formulation at present and the other approaches treating different forms of the p-brane actions \cite{TBC, CW} certainly merits further investigation in trying to clarify a number of non-perturbative aspects of D-brane theory. In the last section in a paper \cite{Aganagic2}, it is stated that "...... For the most part, our analysis has been classical and limited to flat backgrounds. The results should not depend on these restrictions, however." In our pevious paper \cite{KO}, we have relaxed these restrictions to some extent, while in this paper, we have removed such restrictions completely for the super D1-brane and D3-brane. On the other hand, for the super D2-brane and D4-brane we have removed the restriction of 'flat background', but we have presented only the classical analysis. This restriction should be also removed in future. But as emphasized in the previous paper \cite{KO}, it is not always clear for us whether it is necessary to remove this restriction or not since the Dp-brane actions with $p > 1$ are in essence unrenormalizable so these actions might describe the low energy effective theory of some underlying renormalizable theory. This problem still deserves further investigation. \vs 1 \begin{flushleft} {\tenbf} \def\it{\tenit} \def\sl{\tensl Acknowledgement} \end{flushleft} One of authors (I.O.) is grateful to M. Tonin for valuable discussions. His work was supported in part by Grant-Aid for Scientific Research from Ministry of Education, Science and Culture No.09740212.
\section{The nuclear mean field} The cross sections involving neutrons and protons impinging on nuclei display a marked energy variation generally interpreted as an interference between the incident and the transmitted waves. This occurrence in turn implies a mean free path for collisions between the nuclear constituents large compared not only to the internucleonic distance, but even, sometimes, to the dimensions of the nucleus itself. This finding is strongly suggestive of a mean field approach to the structure of the nucleus, especially as far as its ground state is concerned. The nuclear mean field is implemented at the empirical level with the shell model. At the theoretical level a natural treatment resorts to the Hartree--Fock (HF) self--consistent theory, in a frame viewing nuclei as self--bound composite systems of nucleons interacting via a static two--body potential and governed by non--relativistic quantum mechanics. Care is however required in handling the HF theory for the atomic nuclei. Indeed any realistic nucleon--nucleon (NN) force $V_{NN}$ embodies so much repulsion that the expectation value of the nuclear Hamiltonian in the HF ground state is far from being enough attractive. In other words the HF wave function does not prevent two nucleons to come close to each other, where they experience a violent repulsion. The Brueckner theory provides a remedy to this flaw. Indeed it yields an effective interaction $G$ between two nucleons in a nucleus such that the interaction of two ``uncorrelated'' nucleons through $G$ equals the interaction of two ``correlated'' nucleons through $V_{NN}$. Formally this result is achieved in the framework of perturbation theory by summing up (with the Bethe--Goldstone equation) the infinite set of ladder diagrams representing the scattering of two nucleons in the medium out of the Fermi sea, the other nucleons remaining passive. Is the Brueckner--Hartree--Fock (BHF) accounting satisfactorily for the nuclear ground state properties? While the answer to this question is negative (the BHF mean field yielding a too large nuclear central density), yet an important lesson has been learned from BHF. It amounts to recognize that the perturbative corrections to the BHF mean field, necessary to reconcile theory and experiment, should not be computed by expanding in $G$, but rather by grouping together diagrams involving three, four, etc. nucleons repeatedly interacting among themselves through $G$ ({\it hole line} expansion). In spite of the fact that a proof of the convergence of the hole line expansion has never been provided, three-- and four--hole line contributions have been actually computed in nuclear matter; however to achieve the same goal in finite nuclei has proved to be an almost impossible task. Accordingly short cuts have been seeked to incorporate the effects going beyong BHF still in a self--consistent mean field framework for the nuclear ground state.\footnote{We do not mention, for brevity, relativistic mean field approaches which also look promising in reproducing the nuclear ground state properties.} In this connection a successfull approach allows for a density dependence of the interaction $G$, in addition to the one naturally induced by the Pauli operator in the Bethe--Goldstone equation. Indeed this procedure conveniently simulates the energy dependence and non--locality of $G$. Moreover it explains quite successfully the experimentally well established fact that in a nucleus the single particle orbits below the Fermi level are not 100\% occupied as the HF (but not the BHF) approach would imply. Clearly the prize to be payed for the simplicity (and the success) is the introduction of a certain amount of phenomenology empirically fixing the local density dependence of $G$. In the above outlined scheme, Negele\cite{Negele} has been able to impressively reproduce over a wide range of momentum transfer the data of elastic electron scattering on several nuclei (like $^{16}$O, $^{40}$Ca and $^{208}$Pb). This findings strongly support the view that most of the physics of the nuclear ground state lends indeed itself to be embodied in a ``mean field''. Should this be the case, then we would actually know and understand the nucleons' distribution in nuclei. However, before drawing this conclusion, one should remind that electrons only probe the {\it proton} distribution. It would be therefore desirable to test the mean field scheme on other observables, the most natural one being the neutron distribution. Although experimental information on the latter has been gathered in the past, e.g. with pion scattering on nuclei, still our knowledge of it remains rather poor. Even the recurring question of whether or not the nuclear surface is neutron rich cannot be presently answered with certainty. Furthermore the present workshop stresses the urgence of reaching an accurate knowledge of the neutron distribution in order to achieve a precise interpretation of the atomic parity--violating (PV) experiments. Thus in the following we propose a method to measure the neutron distribution in the nuclear ground state, which is based on PV polarized electron--nucleus scattering. \section{The formalism of PV electron scattering} The helicity asymmetry, as measured in the scattering of rigth-- and left-- handed electrons off nuclei, is defined as follows \begin{eqnarray} {\cal A} &&= \frac{d^2\sigma^+ - d^2\sigma^-} {d^2\sigma^+ + d^2\sigma^-} \nonumber\\ &&={\cal A}_0\frac{v_L R^L_{AV}(q,\omega) + v_T R^T_{AV}(q,\omega) + v_{T'} R{T'}_{VA}(q,\omega)} {v_L R^L(q,\omega) + v_T R^T(q,\omega)} \label{asym} \end{eqnarray} where \begin{equation} v_L = \left(\frac{Q^2}{{\vec q}^2}\right)^2, \label{vL} \end{equation} \begin{equation} v_T = \frac{1}{2}\left\vert \frac{Q^2}{{\vec q}^2}\right\vert + \tan^2\frac{\theta}{2} \label{vT} \end{equation} and \begin{equation} v_{T'} = \sqrt{\left\vert \frac{Q^2}{{\vec q}^2}\right\vert + \tan^2\frac{\theta}{2}} \tan\frac{\theta}{2} \label{vTp} \end{equation} are the usual lepton factors, $\theta$ is the electron scattering angle and $Q^2=\omega^2-{\vec q}^2 < 0$ is the space--like four momentum transfer of the vector boson carrying the electromagnetic ($\gamma$) or the weak neutral ($Z_0$) interaction. In (\ref{asym}) the nuclear and nucleon's structure are embedded in both the parity conserving (electromagnetic) longitudinal and transverse ($R^L$ and $R^T$) response functions and in the parity violating (weak neutral) ones. These classify as well in vector longitudinal and transverse ($R^L_{AV}$, $R^T_{AV}$) and axial transverse $R^{T'}_{VA}$ responses, the first (second) index in the subscript referring to the vector/axial nature of the weak neutral leptonic (hadronic) current. Finally the scale of the asymmetry is set by \begin{equation} {\cal A}_0 = \frac{\sqrt{2}G m^2_{\scriptscriptstyle N}}{\pi\alpha} \frac{|Q^2|}{4m_{\scriptscriptstyle N}^2} \approx 6.5\times10^{-4}\tau \qquad \left(\tau=\frac{|Q^2|}{4m_{\scriptscriptstyle N}^2} \right) \label{A0} \end{equation} in terms of the electromagnetic ($\alpha$) and Fermi ($G$) coupling constants ($m_{\scriptscriptstyle N}$ is the nucleon mass). To gain informations on the structure of nuclei and nucleons the standard Coulomb, electric and magnetic multipole decomposition \begin{equation} R^L(q,\omega) = \sum_{J\ge 0} F_{CJ}^2 (q) \label{RL} \end{equation} and \begin{equation} R^T(q,\omega) = \sum_{J\ge 1}\left\{ F_{EJ}^2(q) + F_{MJ}^2(q) \right\} \label{RT} \end{equation} for the parity conserving (electromagnetic) and \begin{eqnarray} R^L_{AV}(q,\omega) &=& a_A\sum_{J\ge 0} F_{CJ}(q){\widetilde F}_{CJ}(q) \label{RLAV}\\ R^T_{AV}(q,\omega) &=& a_A\sum_{J\ge 1} \left\{ F_{EJ}(q){\widetilde F}_{EJ}(q) + F_{MJ}(q){\widetilde F}_{MJ}(q)\right\} \label{RTAV}\\ R^{T'}_{VA}(q,\omega) &=& -a_V\sum_{J\ge 1} \left\{ F_{EJ}(q){\widetilde F}_{MJ_5}(q) + F_{MJ}(q){\widetilde F}_{EJ_5}(q)\right\} \label{RTPAV} \end{eqnarray} for the parity violating (weak neutral) responses are performed. In the above formulas $\omega$ is supposed to be given so the responses actually become functions of $q$ only. Furthermore the Standard Model for the electroweak interaction is assumed: thus the vector and axial--vector leptonic coupling at the tree level read \begin{eqnarray} a_V &=& 4\sin^2\theta_W - 1 \label{av}\\ a_A &=& -1 \label{aa} \end{eqnarray} in terms of the Weinberg's angle. Finally the form factors, chosen to be real, with (without) a tilde are the weak--neutral (electromagnetic) ones. They, of course, split into isoscalar and isovector components according to the isospin decomposition \begin{equation} J_{\mu}= \xi^{(0)}\left(J_\mu\right)_{00} + \xi^{(1)}\left(J_\mu\right)_{10} \label{current} \end{equation} of the hadronic current. Again, on the basis of the Standard Model at tree level, one has \begin{equation} \xi^{(0)}=\xi^{(1)} = 1 \label{csiem} \end{equation} in the electromagnetic sector. In the weak neutral sector, instead, one has \begin{eqnarray} \xi^{(0)}&=&\beta_V^{(0)}= -2\sin^2\theta_W \simeq -0.461, \label{csiw1}\\ \xi^{(1)}&=&\beta_V^{(1)}= 1-2\sin^2\theta_W \simeq 0.538 \label{csiw2} \end{eqnarray} for the vector coupling and \begin{equation} \xi^{(0)}=\beta_A^{(0)}=0, \qquad \xi^{(1)}=\beta_A^{(1)}=1 \label{csiwax} \end{equation} for the axial one. All the above formulas hold valid in the approximation of exchanging only one vector boson between the lepton and the hadron and up to a possible parity admixture in the nucleon itself (anapole moment of the nucleon) or in the nuclear states.The latter would stem from parity--violating components in the nucleon--nucleon interaction. These items will not be dealt with here. For a comprehensive treatment we refer the interested reader to the specialized literature on the subject. \section{Elastic polarized electron scattering from spin zero, isospin zero nuclei} As pointed out long time ago by Feinberg and Walecka\cite{Feen} formula (\ref{asym}) applied to the elastic scattering of polarized electrons from a spin--zero, isospin--zero nuclear target leads to the simple expression \begin{equation} {\cal A} = {\cal A}_0 2 \sin^2\theta_W \label{asymel} \end{equation} for the asymmetry. It thus seemed that the opportunity was there to address the physics of the Standard Model in the low energy regime, being (\ref{asymel}) a ``model--independent'' expression. However things are not so simple because the isospin purity of the nuclear states, on which (\ref{asymel}) relies, is not realized in nature: indeed the proton and neutron quantum states are different in a nucleus. Accordingly, to account for the isospin breaking, (\ref{asymel}) should be recast as follows \begin{eqnarray} {\cal A}&&={\cal A}_0 a_A\beta_V^{(0)} \left\{ {\displaystyle \frac {1+ {\displaystyle\frac{\beta_V^{(1)}}{\beta_V^{(0)}}} {\displaystyle\frac{<J_i |{\hat M}_{0;10}| J_i>} {<J_i |{\hat M}_{0;00}| J_i> }} } {1+ {\displaystyle\frac{<J_i |{\hat M}_{0;10}| J_i>} {<J_i |{\hat M}_{0;00}| J_i> }} }} \right\} \label{nuclasym}\\ &&\simeq {\cal A}_0 2\sin^2\theta_W \left\{ 1 + \left(\frac{\beta_V^{(1)}}{\beta_V^{(0)}} -1\right) {\displaystyle \frac {<J_i |{\hat M}_{0;10}| J_i> }{<J_i |{\hat M}_{0;00}| J_i>} } \right\} \nonumber \\ &&= {\cal A}_0 2\sin^2\theta_W \left\{ 1 - \frac{1}{2\sin^2\theta_W} {\displaystyle \frac {<J_i |{\hat M}_{0;10}| J_i> }{<J_i |{\hat M}_{0;00}| J_i>} } \right\}\, . \nonumber \end{eqnarray} The above formula should of course be used ``cum grano salis''. Indeed the expansion in the arguably small ground state matrix element of the {\it isovector} monopole operator ${\hat M}_{0;10}$ is warranted except where the ground state matrix element of the {\it isoscalar} monopole operator ${\hat M}_{0;00}$ is also or vanishing or very small, which happens, of course, at or close to the diffraction minima of the elastic cross--sections. Thus, barring for these small domains, one would like to estimate, in the formula for the asymmetry \begin{equation} {\cal A}={\cal A}_0 2\sin^2\theta_W \left[ 1 + \Gamma(q)\right]\ , \label{asym2} \end{equation} the impact of the nuclear dependent term $\Gamma(q)$, whose definition follows by comparing (\ref{nuclasym}) with (\ref{asym2}). Before specifically addressing this issue, let us first answer the question: given that $\Gamma(q)$ is not zero, is it still small enough to render worth trying a measurement of $\sin^2\theta_W$ with parity violating electron scattering? Possibly this was the case a decade ago. Indeed around 1990 the value of the Weinberg's angle quoted in the literature read \begin{equation} \sin^2 \theta_W = 0.227\pm 0.005 \, , \label{Wein} \end{equation} namely it was given with a precision of 2.2\%. Accordingly in such a condition a meaningful determination of $\sin^2\theta_W$ would have required on the one hand a measurement of the asymmetry to an accuracy of $1\div 2\%$ and on the other to assume the isospin impurity in a nucleus to be below such a level. Let us now see how these figures translate into a kinematical constraint. Given that an accuracy of $10^{-7}$ could be reached in measuring ${\cal A}$, a test of the Standard Model, as expressed by (\ref{Wein}), would have had a good chance to be performed providing ${\cal A}\ge 10^{-5}$, in turn implying [from (\ref{A0})] $q \ge 1.75$~fm$^{-1}$. Now, always from (\ref{A0}), it follows that ${\cal A}$ grows with $q$, but, at the same time, the elastic form factor falls off with $q$: one would thus choose the range \begin{equation} 1.75 \le q \le 3.5 \,{\mathrm fm}^{-1} . \label{range} \end{equation} as a reasonnable compromise between these two opposite requirements. The question to be addressed was (and is) then: how large is $\Gamma (q)$ in the range (\ref{range})~? Is there $\Gamma(q) \le 10^{-2}$~? Whatever the answer to these questions might be, today the above arguing is of course untenable since now \begin{equation} \sin^2\theta_W = 0.23055 \pm 0.00041 \, , \label{Weinlast} \end{equation} i.e, the Weinberg's angle is known with an accuracy of $0.18\%$. Yet for others relevant observables, for example the strange parity content of the nucleon, although a generalization of the expression (20) for the asymmetry is required, a reliable theoretical handling of $\Gamma (q)$ is still crucial. This issue will be addressed in the next Section. \section{A simple model for isospin breaking} The isospin symmetry is broken in atomic nuclei by the Coulomb force which pushes the protons orbits outward with respect to the neutrons ones. On the other hand the strong proton--neutron interaction tries to equalize the proton's and neutron's Fermi energies, thus acting in the direction of restoring the isospin symmetry: the balance between these two effects is believed to leave a generally modest symmetry breaking in nuclei,which thus lends itself to a perturbative treatment. To explore this physics Donnelly {\it et al.}\cite{Donn} worked out a simple model characterized by two monopole states ($J^\pi =0^+$), one isoscalar and one isovector, mixed by the isospin breaking interaction, in particular the Coulomb one. In this scheme the dominantly isospin $T = T_0$ ground state is actually represented by the superposition \begin{equation} |``T_0''> = \cos\chi\,|T_0> + \sin\chi\, |T_0+1>\, , \label{T0} \end{equation} and the dominantly $T_1 = T_0 + 1$ excited state by the orthogonal combination \begin{equation} |``T_1''> = -\sin\chi\,|T_0> + \cos\chi\, |T_0+1>\, . \label{T1} \end{equation} The associated ground state matrix elements of the isoscalar and isovector monopole Coulomb operators read then, respectively, \begin{eqnarray} &&<``T_0''|{\hat M}_{0;00}(q)|``T_0''> \nonumber \\ &&\qquad = \cos^2\chi\,<T_0|{\hat M}_{0;00}|T_0> + \sin^2\chi\, <T_0+1|{\hat M}_{0;00}|T_0+1> \label{M00me} \end{eqnarray} and \begin{eqnarray} &&<``T_0''|{\hat M}_{0;10}(q)|``T_0''> = \cos^2\chi\,<T_0|{\hat M}_{0;10}|T_0> \label{M10me} \\ && + \sin^2\chi\, <T_0+1|{\hat M}_{0;10}|T_0+1> + 2\sin\chi\cos\chi\, <T_0+1|{\hat M}_{0;10}|T_0>\, . \nonumber \end{eqnarray} Let us then consider $N = Z$ nuclei. The Clebsch--Gordon (CG) coefficients entering into the reduction of the above matrix elements in isospace are 1 and $1/\sqrt{3}$ , respectively, in formula (\ref{M00me}), whereas in (\ref{M10me}) only the third term survives and the associated CG is $1/\sqrt{3}$. One thus obtains for the asymmetry in leading order of the mixing angle $\chi$ the expression \begin{eqnarray} {\cal A} &&= {\cal A}_0 a_A {\displaystyle\frac {\beta_V^{(0)}<0^+;0\|{\hat M}_{0;00}\|0^+;0> + \beta_V^{(1)}2\chi\frac{1}{\sqrt{3}}<0^+;1\|{\hat M}_{0;10}\|0^+;0>} { <0^+;0\|{\hat M}_{0;00}\|0^+;0> + 2\chi\frac{1}{\sqrt{3}}<0^+;1\|{\hat M}_{0;10}\|0^+;0>} } \nonumber \\ &&\quad \simeq {\cal A}_0 2\sin^2\theta_W \{1 +\Gamma(q)\} \label{asym3} \end{eqnarray} where \begin{equation} \Gamma(q) = 2\left(\frac{\beta_V^{(1)}}{\beta_V^{(0)}} - 1\right) \chi {\cal R}(q) = -\frac{1}{\sin^2\theta_W}\chi {\cal R}(q) \label{Gamma} \end{equation} and \begin{equation} {\cal R}(q) = \frac{1}{\sqrt{3}} \frac{<0^+;1\|{\hat M}_{0;10}\|0^+;0>}{<0^+;0\|{\hat M}_{0;00}\|0^+;0>} = \frac{F_{C0}(0^+;``1'';``0''\to 0^+)} {F_{C0}(0^+;``0'';``0''\to 0^+)}\, . \label{Erre} \end{equation} In the above the double bar matrix elements are meant to be reduced in isospace. Note that ${\cal R}(q)$ simply represents the ratio between the inelastic and the elastic form factors associated with the two monopoles states of our $N = Z$ nucleus. Explicit calculations of $\Gamma(q)$ for a few nuclei have been performed by Donnelly {\it et al.}\cite{Donn} in a Wood--Saxon single particle wave functions basis with various effective interactions and in different configurations spaces. We display in Fig.~1a and 1b their results for $\Gamma(q)$ in C$^{12}$ and Si$^{28}$. The singularities in the curves should be disregarded since they don't have, as previously discussed, any physical significance. Although the results of ref.\cite{Donn} are model dependent (they do change significantly according to the effective interaction employed), yet they convey the message that for ligth $N = Z$ nuclei like C$^{12}$ the isospin breaking remains tiny indeed, below 1\% over the whole range (\ref{range}) of $q$. \begin{figure}[t] \mbox{\epsfig{file=fig1.ps,width=0.9\textwidth}} \vskip 0.1cm \caption[Fig.~\ref{fig1}]{\label{fig1} The nuclear--structure--dependent part of the parity--violating asymmetry as defined by eqs,~(\ref{Gamma}) and (\ref{Erre}). {\it Left:} calculations for elastic scattering from $^{12}$C using Woods--Saxon single--particle wave functions. The dotted line at $|\Gamma(q)|=10^{-2}$ indicates the level above which the structure--dependent effects would have confused the interpretation of the asymmetry as a test of electroweak theories, when these were known with the precision given by (\ref{Wein}). {\it Right:} the same as in the left panel, but for $^{28}$Si. For this nucleus Harmonic Oscillator single--particle wave functions and the shell model amplitudes as given by W.C. Haxton (unpublished) have been used. } \end{figure} Note also that in reaching this result the mixing angle $\chi$ has been extracted, with a quite conservative attitude, from the perturbative formula \begin{equation} \sin^2\chi = \frac{\left|<T_0+1|H_{CSV}|T_0>\right|^2} {\left(E_{T_0+1}-E_{T_0}\right)^2} \label{sinchi} \end{equation} ( $H_{CSV}$ is the charge symmetry violating part of the nuclear hamiltonian). In fact, while the experiments indicate for the matrix element appearing in the numerator of (\ref{sinchi}) a value ranging, in C$^{12}$, between 150 and 300~keV, the latter value has been adopted in obtaining Fig.~1a and 1b (moreover $\Delta E = E_{T_0+1}-E_{T_0}\simeq 17$~MeV according to the simple two levels model). These findings are supported by the results of a quite sophisticated calculation of Ramavataram {\it et al.}\cite{Ramava}. These authors, using a state of the art variational wave function originally due to Pandharipande {\it et al.}\cite{Pandha}, obtain in He$^4$ a $\Gamma(q)$ always well below 1\% in the whole range (\ref{range}). Of course this outcome also reflects the particularly rigid structure of He$^4$, whose first excited state lies at 20.1~MeV (the first excited state of C$^{12}$ is at 4.44~MeV). It should however be pointed out that, for $N = Z$ but heavier nuclei like Si$^{28}$, the isospin breaking, while small, grows and reaches a few \% in the range of q given by (\ref{range}). \section{The case of $N \ne Z$ nuclei} For nuclei with $N \ne Z$ a novel (and important) feature appears. Although it will be illustrated in the specific case of nuclei with $N = Z + 2$, and therefore with third isospin component $M_T = -1$, it remains valid in all cases. Sticking always to the two levels model\cite{Donn} we consider then a nucleus with a dominant isospin $T_0 = 1$ component in the ground state. Let in addition the nucleus have an excited state with a dominant isospin $T_1 = T_0 + 1 = 2$. Both states are further characterized by having $J^\pi = 0^+$. Proceeding as in the case of a $Z = N$ nucleus, one arrives to an asymmetry still given by an expression like (\ref{asym3}), but with the model dependent nuclear term $\Gamma(q)$ reading now as follows \begin{eqnarray} &&\Gamma(q) = \frac{1}{2}\beta_V^{(1)} \left\{ {-\frac{1}{\sqrt{6}}<0^+;1\|{\hat M}_{0;1}\|0^+;1> + 2\sqrt{\frac{1}{10}}\chi <0^+;2\|{\hat M}_{0;1}\|0^+;1> } \right\} \nonumber\\ &&\qquad\quad\times\left\{ \sqrt{\frac{1}{3}} <0^+;1\|{\hat M}_{0;0}\|0^+;1> -\frac{1}{\sqrt{6}} <0^+;1\|{\hat M}_{0;1}\|0^+;1> + \right. \nonumber\\ &&\qquad\qquad\qquad\qquad \left. + 2\sqrt{\frac{1}{10}}\chi <0^+;2\|{\hat M}_{0;1}\|0^+;1> \right\}^{-1} \label{Gamma1}\\ \end{eqnarray} which can again be recast according to \begin{equation} \Gamma(q) = \frac{1}{2}\beta_V^{(1)} {\displaystyle\frac {F_{C0}(0^+;``1'';``1''\to 0^+)_{\mathrm isovector}} {F_{C0}(0^+;``1'';``1''\to 0^+)_{\mathrm total}} }\, . \label{Gamma2} \end{equation} Namely $\Gamma(q)$ turns out to be, like before, proportional to the ratio between the inelastic (isovector) and the elastic (this time both isoscalar and isovector) form factors. Now from (\ref{Gamma1}) a new feature is immediately apparent: unlike the $N = Z$ case, where $\Gamma(q)$ was found to be proportional to the mixing parameter $\chi$, here $\Gamma(q)$, in addition to terms proportional to $\chi$, also embodies terms {\it independent} from it. As a consequence, $\Gamma(q)$ turns out to be now much larger, as it is clearly observed in Fig.~2a and 2b, where the results of ref.\cite{Donn} are displayed for C$^{14}$ and Si$^{30}$. \begin{figure}[t] \mbox{\epsfig{file=fig2.ps,width=0.9\textwidth}} \vskip 0.1cm \caption[Fig.~\ref{fig2}]{\label{fig2} The structure--dependent part of the parity--violating asymmetry for elastic scattering as defined in eqs.~(\ref{tildegam}) and (\ref{betap}). The results for $^{14}$C using a 1p--shell model space (solid line) and a $2\hbar\omega$--space (dashed line) are displayed. Also displayed are the results for $^{30}$Si in the extreme single--particle shell model (solid line) and in a full 2s1d shell--model calculation (dashed line). No isospin mixing effects are included. } \end{figure} Actually what is shown in the figures 2a and 2b is not $\Gamma(q)$, but ${\widetilde\Gamma} (q)$ since the $N \ne Z$ nuclei are more appropriately discussed in terms of protons and neutrons rather than in term of isospin and, accordingly, the expression (\ref{asym2}) for the asymmetry is now more conveniently recast into the form: \begin{equation} {\cal A}={\cal A}_0 a_A\left(\beta_V^p+\frac{N}{Z}\beta_V^n\right) \left[ 1 + {\widetilde\Gamma}(q)\right] \label{asym4} \end{equation} where \begin{eqnarray} \beta_V^p &=& \frac{1}{2}\left(\beta_V^{(0)} +\beta_V^{(1)}\right) =0.038\, , \label{betavp}\\ \beta_V^n &=& \frac{1}{2}\left(\beta_V^{(0)} -\beta_V^{(1)}\right) = -\frac{1}{2} \label{betavn} \end{eqnarray} and \begin{equation} {\widetilde\Gamma}(q) = \frac{1}{2}{\tilde\beta}'_V {\displaystyle\frac {\left\langle 0^+\left|\frac{1}{2}\left(1-\frac{Z}{N}\right){\hat M}_{0;00} +\frac{1}{2}\left(1+\frac{Z}{N}\right){\hat M}_{0;10}\right|0^+ \right\rangle} {<0^+\left| {\hat M}_{0;00} + {\hat M}_{1;10}\right|0^+>} }\, , \label{tildegam} \end{equation} being \begin{equation} {\tilde\beta}'_V = 4\frac{N}{Z}\frac{\beta_V^n}{\beta_V^p + {\displaystyle\frac{N}{Z}}\beta_V^n}\, . \label{betap} \end{equation} It is immediately checked that by setting $N = Z$ in the above formulas one gets ${\widetilde\Gamma}(q)\to \Gamma(q)$ and ${\tilde\beta}'_V = 1/\sin^2\theta_W$. Let us further observe that, although in the figures 2a and 2b ${\widetilde\Gamma}(q)$ appears indeed to be markedly model dependent, yet the basic message previously anticipated clearly stands out: in the range of momentum transfers (\ref{range}) ${\widetilde\Gamma}(q)$ assumes values typically ranging between 10 and 50\%. In conclusion, from the previous two Sections it follows that for light $N = Z$ nuclei, like He$^4$ and C$^{12}$, \begin{itemize} \item{} a test of the electroweak theory with parity--violating polarized elastic electron scattering experiments is out of question or, perhaps, only marginally, if at all, possible at very low momentum transfer in He$^4$; \item{} however many investigations show that elastic (and, also, quasielastic, not addressed here) parity violating polarized electron scattering experiments can be usefully exploited to unravel the strange form factor of the nucleon. \end{itemize} For medium--heavy and heavy nuclei $\Gamma(q)$ grows with the mass number $A$, especially when $N \ne Z$ (large terms not proportional to $\chi$ appear!). Therefore parity--violating polarized electron scattering can be advantageously used to measure the amount of isospin breaking in a nucleus or, better yet, the neutron distribution. \section{The neutron distribution} To appreciate how the neutron distribution can be measured with polarized elastic electron scattering it helps to revisit the previous concepts with a somewhat different language. Let us thus observe that for $T = 0$ nuclei ($N = Z$) if isospin is a ``good symmetry'' then the standard unpolarized electron scattering measures the {\it isoscalar} nuclear density. In these conditions the latter also fixes the polarized elastic electron scattering which is thus independent from any further nuclear structure information. If, however, isospin is sligthly broken, then the {\it isovector} nuclear density also enters, as a small perturbation, in the elastic polarized electron scattering thus introducing an additional model dependence. Quite on the contrary, for $T \ne 0$ ($N \ne Z$) nuclei both the {\it isoscalar} and the {\it isovector} densities enter into the scattering process, no matter if isospin is or is not a perfect symmetry. In this instance, as already pointed out, it is preaferable to use the neutron--proton language. Accordingly the ground state matrix element of the Coulomb monopole operator \begin{equation} <0^+\left|{\hat M}_0(q)\right|0^+> = \frac{1}{\sqrt{4\pi}}\int d{\vec x} j_0(qx)\rho({\vec x})\, , \label{monopole} \end{equation} where $j_0$ is the zeroth order spherical Bessel function and $\rho({\vec x})$ is the matter density of the nucleus, will display both an isoscalar and an isovector components, according to the expressions \begin{equation} <0^+\left|{\hat M}_{0;00}(q)\right|0^+> = \frac{1}{\sqrt{4\pi}}\int d{\vec x} j_0(qx) \frac{\rho_p({\vec x})+\rho_n({\vec x})}{2} \label{isosca} \end{equation} and \begin{equation} <0^+\left|{\hat M}_{1;00}(q)\right|0^+> = \frac{1}{\sqrt{4\pi}}\int d{\vec x} j_0(qx) \frac{\rho_p({\vec x})- \rho_n({\vec x})}{2} \label{isovec} \end{equation} In the above $\rho_p$ and $\rho_n$ are the protons and neutrons densities, respectively. When (\ref{isosca}) and (\ref{isovec}) are inserted into the formula (\ref{asym4}) it is then an easy matter to obtain for the asymmetry the expression\footnote{Note that, by comparison, (\ref{Gamma3}) yields \begin{equation} {\widetilde\Gamma}(q) = \frac{N}{Z} \frac{\beta_V^n}{\beta_V^p + {\displaystyle\frac{N}{Z}}\beta_V^n} \left[ 1 - {\displaystyle\frac {Z \int d{\vec x} j_0(qx)\rho_n({\vec x})} {N \int d{\vec x} j_0(qx)\rho_p({\vec x})} } \right] \, . \label{Gamma3} \end{equation} } \begin{equation} {\cal A} = {\cal A}_0 a_A \left\{ \beta_V^p + \beta_V^n {\displaystyle\frac {\int d{\vec x} j_0(qx)\rho_n({\vec x})} {\int d{\vec x} j_0(qx)\rho_p({\vec x})} } \right\} \label{asym5} \end{equation} Since, according to the Standard model, \begin{equation} \beta_V^p = 0.038 \qquad{\mathrm and}\quad \beta_V^n = - 0.5 \,, \label{boh} \end{equation} (\ref{Gamma3}) shows that {\it the asymmetry in the parity--violating elastic polarized electron scattering represents an almost direct measurement of the Fourier transform of the neutron density}, the analogous quantity for the protons being fixed by the elastic unpolarized electron scattering. In particular a rigorous $q$--independence of ${\cal A}/{\cal A}_0$ would imply \begin{equation} Z \rho_n({\vec x}) = N \rho_p({\vec x}), \end{equation} i.e. pure isospin symmetry! Of course, as previously discussed, in nuclei the distribution of neutrons differs from the one of protons. To gain a first insight on how this difference would be perceived in a parity violating elastic electron scattering experiment Donnelly {\it et al.} have computed the asymmetry (\ref{asym5}) and the ${\widetilde\Gamma}(q)$ of eq.~(\ref{Gamma3}) for Ca$^{40}$, Ca$^{48}$ and Pb$^{208}$ using phenomenological proton densities which well accomplish for the elastic unpolarized electron scattering. For example, for Ca$^{40}$ they employ the well--known 3 parameters Fermi distribution: \begin{equation} \rho({\vec r})=\rho_0{\displaystyle\frac{1+\omega{\displaystyle\frac{r^2}{R^2}}} {1+e^{{\displaystyle (r-R)/a}}} }\, . \label{Fermi3} \end{equation} For the neutrons they use the same densities, euristically enlarging however, for explorative purposes, the radius parameter by 0.2~fm. Their results are displayed in Fig.~3. To grasp the significance of this figure it helps to expand ${\widetilde\Gamma}(q)$ as follows: \begin{equation} {\widetilde\Gamma}(q) = \frac{\beta_V^n}{\beta_V^p + {\displaystyle\frac{N}{Z}}\beta_V^n} \left[ \frac{N}{Z} - {\displaystyle\frac {N - \frac{1}{6}q^2 R^2_n} { \int d{\vec x} j_0(qx)\rho_p({\vec x})} } + \dots \right] \, . \label{Gamma4} \end{equation} It is thus clear that the region around the first dip/peak carries information on the neutron radius whereas the fourth moment of the neutron distribution will be observed nearby the second dip/peak and so on. On the basis of this expansion one finds that in the case of Pb$^{208}$ at $q\simeq 0.5$~fm$^{-1}$, with ${\cal A}\simeq 8\times 10^{-8}$, a 1\% change in the neutron radius is reflected in a change of about 6\% in the asymmetry. \begin{figure}[t] \mbox{\epsfig{file=fig3.ps,width=0.9\textwidth}} \vskip 0.1cm \caption[Fig.~\ref{fig3}]{\label{fig3} The parity--violating asymmetry (upper row) and the structure--dependent part of the asymmetry as defined in eq.~(\ref{Gamma4}) (lower row) for elastic electron scattering from $^{40}$Ca, $^{48}$Ca and $^{208}$Pb. Displayed are the calculations with $\rho_n(r)/N=\rho_p(r)/Z$ [solid line, corresponding to ${\widetilde\Gamma}(q)=0$] and for $\rho_n(r)/N\ne \rho_p(r)/Z$ (dashed line). The density parameterizations are discussed in the text and specified in ref.\cite{Donn}. } \end{figure} It might be interesting to observe that formula (\ref{Gamma4}) can be generalized in the sense that the Fourier transform of the pure Fermi distribution can be, although not exactly, quite accurately analytically expressed. Indeed for this density the form factor turns out to be\cite{noi1} \begin{equation} F_{\mathrm Fermi}(q) =\rho_0 (2\pi)^2R^2 a \left\{j_1(qR){\tilde y}_0(\pi qa) - {\tilde j}_1(\pi qa)y_0(qR)\right\} \frac{\pi qa}{{\tilde j}_0(\pi qa)} \label{FForm} \end{equation} where the $j(y)$ and the ${\tilde j}({\tilde y})$ are the ``spherical'' and the ``modified spherical'' Bessel functions of first (second) kind, respectively \cite{Abram}. The form factor $F_3$ of the three--parameters Fermi distribution is then expressed in terms of $F_{\mathrm Fermi}$ according to: \begin{equation} F_3(q) \simeq F_{\mathrm Fermi}(q) -\frac{\omega}{R^2} \frac{d^2}{dq^2}\left[ q F_{\mathrm Fermi}(q)\right]\, . \label{F3} \end{equation} When inserted into (\ref{asym5}) and (\ref{Gamma3}), the above formula allows to analytically recover the results of Donnelly for Ca$^{40}$. \section{Flaws and hopes} The results of the previous section are flawed essentially by two shortcomings: one relates to the factorization of the single nucleon physics, which has been assumed in deducing (\ref{asym5}) from (\ref{asym4}). Especially when $q$ is large, and thus relativistic effects become substantial, such a procedure almost certainly becomes unwarranted. It is clear that this point must be more carefully addressed in future research. The second problem relates to the approximation of considering just the Fourier transform of the charge and neutron distributions, which corresponds to consider a single vector boson exchange between the impinging field and the target, leaving the electron to be described by plane waves. This is clearly insufficient, especially in heavy nuclei, where, in fact, the electron wave is quite distorted by the nuclear Coulomb field. To account for this effect a heavy computational effort is presently carried out by an MIT--Indiana University collaboration. A more modest, but perhaps also useful approach, resorts to a kind of eikonal approximation to describe the distortion of the electron wave. We feel confident that these difficulties will in the end be overcome. It will thus become possible to measure the neutron distribution in the ground state of atomic nuclei with parity violating {\it elastic} scattering experiments. This most remarkable occurrence goes in parallel with the finding of Alberico {\it et al.}\cite{noi2} for the {\it quasi--elastic} polarized electron--nucleus scattering. Indeed these authors show that, being the $Z_0$ almost ``blind'' to protons, the PV longitudinal response of an uncorrelated system of protons and neutrons, like the relativistic Fermi gas, to a polarized beam of electrons is almost vanishing. Departure from this expectation will thus signal the effect of neutron--proton correlations in nuclei, which are expected to be especially relevant in the isoscalar channel. Thus polarized electron scattering experiments appear to open a window on crucial, and till now insufficiently explored, aspects of the nuclear structure. A final remark is in order: parity--violating {\it nuclear} electron scattering experiments, initially conceived as a tool for exploring the Standard Model at the nuclear level, will turn out, in the end, to represent a tool for providing the information the {\it atomic} parity violating experiments need to accurately test the Standard Model at the atomic level: namely the neutron distribution. Indeed the precision on the measured energy levels of the Cesium atoms, which is required to test the Standard Model, is so high that it cannot be reached without controlling also the neutron distribution in nuclei.
\section{Introduction} The lumpy distribution of matter resulting in Toomre \& Kalnajs' (1991, hereafter TK) local shearing-sheet experiments, reminds us of the ubiquitous inhomogeneous state of the ISM as well as the flocculent structures of many spirals. Therefore we decided to take up their 2D-model and to extend it on 3 dimensions order to investigate more precisely the clumpy structure of galactic matter. It is important to include the third dimension in the model, because as soon as clumping develops dynamical coupling with vertical motion must be strong, contrary to the weak coupling existing in a smoothly distributed thin disk. Our model considers scales of the order of ${\cal O}(1 \rm{kpc})$. Thus we are able to investigate the transition regime between the molecular cloud scales and the galactic disk scale. Furthermore we can investigate whether gravitation and dissipation can maintain the power laws observed in molecular clouds on scales at which shearing is dominant. This paper is divided up in two main parts. The first part presents the model (Sects. 2.1-2.4) and the second part shows the results (Sect. 3.1) and their analysis, i.e. the determination of the fractal dimension (Sect. 3.2) and the verification of Larson's law (Sect. 3.3). \section{Model} \subsection{Principle} In local models only a small part (e.g., everything inside a box with a given size) of the system is simulated and more distant regions are represented by replicas of the local box. In such a model the orbital motion of the particles is determined by Hill's approximation of Newton's equations of motion (Hill 1878) \begin{equation} \label{eq1} \begin{array}{ccccccc} \ddot{x}&-&2\Omega_0\dot{y}&=&4\Omega_0 A_0 x& + &F_x\\ \ddot{y}&+&2\Omega_0\dot{x}&=& & &F_y\\ \ddot{z}& & &=&-\nu^2 z&+&F_z \end{array} \end{equation} where $A_0=-\frac{1}{2} r_0 \left(\frac{d\Omega}{dr}\right)_{r_0}$ is the Oort constant of differential rotation and $\nu$ is the vertical epicycle frequency. $F_x, F_y$ and $F_z$ are local forces due to the self-gravitating particles. Because of the shearing flow the relative positions of the rectangular boxes (local box and replicas) change with time, so that a initially periodic arrangement of the boxes relative to a fixed Cartesian coordinate system can not be maintained. As a consequence the forces of the self-gravitating particles must be determined by direct summation, requiring ${\cal O}(N^2)$ operations for $N$ particles. To increase the performance in our models we calculate the forces with the FFT-convolution method, requiring ${\cal O}(N_c{\rm log} N_c)$ operations, where $N_c$ is the number of cells, which should have the same order of magnitude as the number of particles. This requires a system spatially periodic at each time. Therefore we use a pair of time-dependent affine coordinate systems, whose affinity angle change with time and are determined by the shear flow (see Fig. 1). The angle difference between the two grids is constant and given by ${\rm atan}(L_y/L_x)$. The reason to evaluate twice the forces in two different affine grids is for avoiding force discontinuities in time when the affine angle would have reached a maximum value beyond which a smaller angle would exist. After the calculation of the forces for the two affined grids, they are transformed into Cartesian coordinates, where they are weighted and added. The weighting factors are proportional to the grid inclination and normalized. \begin{figure}[htb] \centerline{ \psfig{file=grid8.ps,width=\hsize}} \caption{Schematic diagram of the inclined grids seen from above. {\it a.)} The initial state of the two grids $(t=0)$. The dashed grid is the forward grid and the solid one is the backward grid. Below them, the affinity angles of the grids are indicated, $\alpha_f$ resp. $\alpha_b$. The angle between the two grids remains the same for all times. {\it b.)} The grids at $t=L_y/(2 L_x \Omega_0)$, where $\alpha_f=-\alpha_b$. {\it c.)} When the grids attain these inclinations, they jump back to the positions shown in {\it (a)} and the process starts again without discontinuity in the dynamics.} \end{figure} \subsection{Cooling} \label{cooling} To avoid an increase of the random epicyclic motion due to gravitational heating, TK proposed to add artificial cooling forces. Following them we include the damping terms $-C_x \dot{x}$ and $-C_z \dot{z}$ in the radial resp. vertical forces $(F_x, F_z)$ controlling the particle motions via Eq. (\ref{eq1}). Two different damping terms are necessary to reflect the different directional collision rates induced by the anisotropic velocity ellipsoid. The cooling coefficients are chosen in such a way that the velocity dispersions of the random motions reach a stable level and don't differ much from their initial values. The same stability conditions holds for the disk scale height $z_0$. \section{Results} \subsection{Shearing boxes} Fig. \ref{xy}\footnote{Full resolution paper available at http://obswww.unige.ch/Preprints/cgi-bin/Preprintshtml.cgi?\#DYNAMIC.} and \ref{xz} reveals the spatial distribution of matter in a box, representing a section of a galactic disk. The size of the box sides $(L_x\times L_y\times L_z)$ is given by the critical wavelength $\lambda_{\rm crit}$, defining the scale for which the theory of swing amplification predicts the strongest response (Toomre \citeyear{Toomre81}). Because we are interested in long time behavior, the simulations were performed for $t=20$ galactic rotations. In the initial state $t=0$ the particles are distributed uniformly in the x-y-plane. The particle distribution in $z$-direction obeys $\rho\propto{\rm sech}^2(z/z_0)$, where $\rho$ is the density and $z_0$ is the disk scale height. The velocities at $t=0$ are determined by the shear-flow \begin{equation} \begin{array}{ccccccccccc} \dot{x}&=&0,& &\dot{y}&=&-2A_0 x,& &\dot{z}&=&0 \end{array} \end{equation} and the Schwarzschild velocity ellipsoid. \begin{figure*} \psfig{file=xy1.ps,angle=90,width=\hsize} \caption{ The evolution of the particle positions seen from above. The number of rotations of the shearing box around the galactic center are indicated above each panel. The number density at the beginning of the simulation is $n=10100 \lambda_{\rm crit}^{-2}$, corresponding to $32724$ particles.} \label{xy} \end{figure*} \begin{figure*} \psfig{file=xz.eps,angle=90,width=\hsize} \caption{The positions of the particles inside the slice, $-0.1\; \lambda_{\rm crit}< y < 0.1\; \lambda_{\rm crit}$, seen in the direction of orbital motion.} \label{xz} \end{figure*} Like the 2D-model of TK, our 3D-model reveals a fast ``structure formation''. After the first rotations the striations are already developed and are more or less maintained during the rest of the simulation. In our model $t_{\rm osc} < t_{{\rm cool},x} < t_{{\rm cool},z}$ is valid, where $t_{\rm osc}$ is the period of the unforced epicyclic motion. $t_{{\rm cool},x}$ and $t_{{\rm cool},z}$ are the cooling times for the radial resp. vertical damping in fact $t_{cool,x} \propto C_x^{-1}$ and $t_{cool,z} \propto C_y^{-1}$. Furthermore, the cooling times depend on the particle number density $n$. For the simulation in Fig. \ref{xy} the following relation is valid: $t_{\rm osc} : t_{{\rm cool},x} : t_{{\rm cool},z} \approx 1 : 40 : 300$. The random velocity dispersions and the disk scale height were calculated after every half rotation and are plotted in Fig. \ref{dispz0}. Gravitational instability and shearing, which are responsible for an important part of the striations, have an effect particularly in the x-y-plane and leads to a fast increase of $\sigma_x$ and $\sigma_y$. \begin{figure}[htb] \centerline{ \psfig{file=dispz0.eps,angle=90,width=\hsize}} \caption{The left panel shows the velocity dispersions $\sigma_x$, $\sigma_y$ and $\sigma_z$ as a function of time $t$, indicated in galactic rotation unit. The right panel reveals the evolution of the disk scale height $z_0$.} \label{dispz0} \end{figure} \subsection{Fractal dimension} During this studies our interest was particularly focused on the following question: Can self-gravitation, shearing and dissipation be dominantly responsible for maintaining a fractal distribution of the matter in a disk and can they account for Larson's law (Larson 1981)? To answer this question we calculated the appropriate scaling laws and determined the fractal dimension for the structures shown in Fig. \ref{xy} and \ref{xz}. We determined the fractal dimension, quantifying how much space our system fills, as follows: We choose a representative set of particles and count for each particle the number of neighboring particles $N(R)$ inside a certain radius $R$. If we repeat this for other values of $R$ we can find the fractal dimension $D$ via \begin{equation} N(R)\propto R^D. \end{equation} Because the structures we examine are not an idealized mathematical set, but model a physical system, we have to take into account upper and lower cutoffs in the analysis. An upper limit due to the numerical model is given by the size of the simulation box in the x-y-plane. On this scale the system becomes periodic, meaning that it can not be fractal on scales ${\cal O}(L_x)$. A lower limit is due to the finite resolution of the box grid. If the grid cells have the size $l_x\times l_y\times l_z$ and $l=l_x=l_y>l_z$ then we can't expect fractal structures below $2 l$. Fig. \ref{fractal3} reveals the fractal dimension $D=d {\rm ln}(N)/d{\rm ln}(R)$ as a function of the radius $R$, i.e. of the scale. The solid line corresponds to the initial distribution of matter: $\rho(x,y)={\rm const.},\; \rho(z)\propto{\rm sech}^2(z/z_0)$. The other lines give the mean dimensions of the structures for the simulation terminal phase. The intensity of the structures and thus the structure dimension $D$ depend on the cooling. To show this clearly we performed simulations with varying cooling forces. The figure reveals clearly that the stronger the cooling, the more intensive the structures and the lower the structure dimension $D$ are. The structures in Fig. \ref{xy} and \ref{xz} were performed with a ``strong'' cooling. For this structures $D$ runs in a thin band between $0.2\lambda_{\rm crit}$ and the upper model limit with a minimum of $D\approx 1.83$ at $R=0.35 \lambda_{\rm crit}$. Pfenniger (\citeyear{Pfenniger96}) showed, that a fractal-like, hierarchical gravitating system in statistical equilibrium can only exist with $D<2$. In the range, in which $D$ runs in a thin band this condition is fulfilled, but the dynamical range is too small to call the structures fractals. \begin{figure} \centerline{ \psfig{file=fractal3.eps,width=7cm}} \caption{ The structure dimension $D$ as a function of the scale $R$. The solid line corresponds to the initial state. On large scales this state represents a 2D matter distribution, whereas on small scales $R<z_0$ it tends to $D>2.6$. The other curves represent the simulation terminal phase for different cooling forces (weak, middle, strong). The relative magnitudes of the cooling coefficients are: $C_{x,{\rm weak}} : C_{x,{\rm middle}} : C_{x,{\rm strong}}\approx 1:1.5:2$. Above the figure the natural logarithm of the lower model limit is indicated.} \label{fractal3} \end{figure} \subsection{Larson's law} If the fractal organization of the matter extends to scales at which the shear flow becomes important and the system is still virialized, then the validity range of Larson's law \begin{equation} \sigma \propto R^{\delta} \quad \hbox{\rm or} \quad \delta=\frac{d\ln \sigma}{d\ln R} \end{equation} should exceed the size of Giant Molecular Clouds. We checked Larson's law for scales $>100$ pc by analyzing the phase-space of the particles in Fig. \ref{xy}. Fig. \ref{larson} shows $\delta$ as a function of the scale $R$. There is no plateau between $0.3<\delta<0.5$, but $\delta$ reaches this range for the bigger scales, where also $2\delta+1= D$ is roughly fulfilled. \begin{figure} \centerline{ \psfig{file=larson.eps,width=7cm}} \caption{$\delta$ of Larson's law as a function of the scale $R$. To calculate $\delta$ we used the simulation with the particle density $n=40450$ and the resolution $l=0.03 \lambda_{\rm crit}$ The solid line corresponds to the initial state, whereas the dashed line indicate $\delta$ for the simulation terminal phase.} \label{larson} \end{figure} \section{Conclusion} Our local 3D-simulations show, that self-gravitation and dissipation ensure a statistical equilibrium at scales at which the shear flow is important. Moreover they reveal a fractal-like distribution of matter on kpc-scales. The fractal dimension found in the simulations can be lower than 2, as observed in molecular clouds. The specific value found depends on the relative strengths of the competing gravitational and dissipation processes. Because the dynamical range of our simulations is still small, it would be premature to call the found structures strict fractals. The anisotropy of the velocity-dispersion ellipsoid, resulting from our simulations, has systematically the same ordering $(\sigma_R>\sigma_\Phi>\sigma_z)$ as observed in the Galaxy and in N-body simulations of spirals. \begin{acknowledgements} This work has been supported by the Swiss Science Foundation. \end{acknowledgements} \input{references} \end{article} \end{document}
\section{Introduction} Before the famous experiment by Michelson and Morley in 1887, physicists believed that there should exist an ether in space, in order to explain the propagation of electromagnetic waves, by means of the mechanical theory of waves. However, that experiment showed that there is no relative motion of our planet with respect to a physical medium usually refered to as the ether. Now there is the belief that there should exist space-time as a medium which allows to order physical events. The `new' ether is a collection of physical properties of a continuum space-time (Einstein, 1991). In this paper we propose a discrete picture for the dynamics in classical particle mechanics, where the continuum has no physical meaning at all. We work on the possibility that the world is, in some sense, atomistic. And space-time, as one of the constituents of the world, is also atomistic. We show that we may have causality without any time interval, in the usual sense, between two events. Some decades ago the japanese physicist T. Tati began a researche program about a description for physical theories where the concepts of space and time are not {\em primitive\/} notions (Tati, 1964, 1983, 1986, 1987). For a quick reference on his work see (Tati, 1986). For a detailed presentation of the non-space-time picture of classical mechanics, electromagnetism, quantum mechanics and quantum electrodynamics (QED) see (Tati, 1964). Tati's objective is to solve a specific problem, namely, the divergences in quantum field theory. Tati argues that it is possible {\em to define} space and time in some physical theories, by using the description proposed by him. Space and time are fundamental concepts that should remain in classical physics. But in quantum field theories, the classical approach is meaningless. So, space and time do not exist at microscopic levels. Tati's work supports a theory of finite degree of freedom in QED, which allows to eliminate the divergences. Although Tati's motivation was a {\em physical\/} problem, we consider that a non-space-time description for physical theories is also interesting from the {\em philosophical\/} point of view. The investigation of all logically possible physical theories may conduct to a better understanding of the real role of fundamental concepts and principles in physical theories. For example, we have recently showed how to define a set-theoretical predicate (in the sense of P. Suppes' (1967) program of axiomatization) for classical particle mechanics without the concept of force (Sant'Anna, 1995; 1996), based on Hertz's mechanics (Hertz, 1894; 1956). Now, we are defining an axiomatic framework for non-relativistic classical particles mechanics without space-time. Thus, our work is in agreement with P. Suppes' words about the role for philosophy in the sciences: \begin{quote} We are no longer Sunday's preachers for Monday's scientific workers, but we can participate in the scientific enterprise in a variety of constructive ways. Certain foundational problems will be solved better by philosophers than by anyone else. Other problems of great conceptual interest will really depend for their solution upon scientists deeply immersed in the discipline itself, but illumination of the conceptual significance of the solutions can be a proper philosophical role. (Suppes, 1990). \end{quote} Tati does not assume space-time as a primitive notion. Rather than space-time, he considers {\em causality\/} as a primitive concept, whatever it does really mean. We do not present in this paper all details of Tati's theory because we consider his formulation a bit confuse from the logico-mathematical standpoint. So, our starting point is the intuition presented in Tati's work which, we believe, is somehow preserved in our paper (at least in principle). We venture to interpret Tati's work at our own risk, as it follows. Physical observations may be associated to elements of a discrete set which corresponds, intuitivelly speaking, to {\em state\/} measurement values. Each observation is related to another one by a causal relation. This causal relation may be expressed by equations which are very similar to numerical solutions of differential equations commonly used in physics. The notion of a {\em continuum\/} space-time in theoretical physics allows the use of standard differential and integral calculus on those equations. Space-time intervals, which are associated to elements of a {\em continuum}, should be regarded as `unknowables', if we use Tati's terminology. Such unknowables behave like hidden variables in the sense that we cannot actually measure {\em real\/} space-time intervals. Measurements do not have arbitrary precision. It should also be emphasized that even measurements of mass, position, momentum, etc., should be associated to elements of a discrete set of numbers, if we are interested to eliminate anthropomorphical notions like the {\em continuum} and real numbers. We recall the well known words by L. Kronecker: ``God made the integers, all the rest is the work of man.'' In the next Section we present the axiomatic framework for classical particle mechanics by McKinsey, Sugar and Suppes. In Section 3, a non-space-time description for classical particle mechanics is presented, based on the formulation described in Section 2. In Section 4 we show that our description is consistent. In Section 5 we discuss some possible applications of our picture, as well as other related lines for future works. \section{McKinsey-Sugar-Suppes Predicate for Classical Particle Mechanics} This section is essentially based on the axiomatization for classical particle mechanics presented in (Suppes, 1957), which is a variant of the formulation in (McKinsey et al., 1953). We abbreviate McKinsey-Sugar-Suppes System for Classical Particle Mechanics as M.S.S. system. Our intention is to apply Tati's ideas on the M.S.S. system in order to illustrate our non-space-time description of physics. We have chosen M.S.S. system because it represents the simplest case of a physical theory ever axiomatized. M.S.S. system has six primitive notions: $P$, $T$, $m$, $s$, $f$, and $g$. $P$ and $T$ are sets; $m$ is a vector-valued unary function defined on $P$; $s$ and $g$ are vector-valued binary functions defined on the Cartesian product $P\times T$; and $f$ is a vector-valued ternary function defined on the Cartesian product $P\times P\times T$. Intuitivelly, $P$ corresponds to the set of particles and $T$ is to be physically interpreted as a set of real numbers measuring elapsed times (in terms of some unit of time, and measured from some origin of time). $m(p)$ is to be interpreted as the numerical value of the mass of $p\in P$. $s_{p}(t)$, where $t\in T$, is a $3$-dimensional vector which is supposed to be physically interpreted as the position of particle $p$ at instant $t$. $f(p,q,t)$, where $p$, $q\in P$, corresponds to the internal force that particle $q$ exerts over $p$, at instant $t$. Finally, the function $g(p,t)$ is to be understood as the external force acting on particle $p$ at instant $t$. Now, we can give the axioms for M.S.S. system. \begin{definicao} ${\cal P} = \langle P,T,s,m,f,g\rangle$ is a M.S.S. system if and only if the following axioms are satisfied: \begin{description} \item [P1] $P$ is a non-empty, finite set. \item [P2] $T$ is an interval of real numbers. \item [P3] If $p\in P$ and $t\in T$, then $s_{p}(t)$ is a $3$-dimensional vector such that $\frac{d^{2}s_{p}(t)}{dt^{2}}$ exists. \item [P4] If $p\in P$, then $m(p)$ is a positive real number. \item [P5] If $p,q\in P$ and $t\in T$, then $f(p,q,t) = -f(q,p,t)$. \item [P6] If $p,q\in P$ and $t\in T$, then $[s_{p}(t), f(p,q,t)] = -[s_{q}(t), f(q,p,t)]$. \item [P7] If $p,q\in P$ and $t\in T$, then $m(p)\frac{d^{2}s_{p}(t)}{dt^{2}} = \sum_{q\in P}f(p,q,t) + g(p,t).$ \end{description} \end{definicao} The brackets [,] in axiom {\bf P6} denote vector product. Axiom {\bf P5} corresponds to a weak version of Newton's Third Law: corresponding to every force there is always a counterforce. Axioms {\bf P6} and {\bf P5}, correspond to the strong version of Newton's Third Law, since axiom {\bf P6} establishes that the direction of force and counterforce is the direction of the line between particles $p$ and $q$. Axiom {\bf P7} corresponds to Newton's Second Law. \begin{definicao} Let ${\cal P} = \langle P,T,s,m,f,g\rangle$ be a M.S.S. system, let $P'$ be a non-empty subset of $P$, let $s'$, $g'$, and $m'$ be the functions $s$, $g$, and $m$ with their first arguments restricted to $P'$, and let $f'$ be the function $f$ with its domain $P\times P\times T$ restricted to $P'\times P'\times T$. Then ${\cal P'} = \langle P',T,s',m',f',g'\rangle$ is a subsystem of ${\cal P}$ if the following condition is satisfied: \begin{equation}\label{P7p} m'(p)\frac{d^{2}s'_{p}(t)}{dt^{2}} = \sum_{q\in P'}f'(p,q,t) + g'(p,t). \end{equation} \label{P7} \end{definicao} Actually, in (Suppes, 1957), definition (\ref{P7}) does not have equation (\ref{P7p}). On the other hand, such an equation is really necessary to prove the following theorem: \begin{teorema} Every subsystem of a M.S.S. system is again a M.S.S. system. \end{teorema} \begin{definicao} Two M.S.S. systems \[{\cal P} = \langle P,T,s,m,f,g\rangle\] and \[{\cal P'} = \langle P',T',s',m',f',g'\rangle\] are equivalent if and only if $P=P'$, $T=T'$, $s=s'$, and $m=m'$. \end{definicao} \begin{definicao} A M.S.S. system is isolated if and only if for every $p\in P$ and $t\in T$, $g(p,t) = {\bf 0}$, where ${\bf 0}$ is the null vector. \end{definicao} \begin{teorema} If \[{\cal P} = \langle P,T,s,m,f,g\rangle\] and \[{\cal P'} = \langle P',T',s',m',f',g'\rangle\] are two equivalent systems of particle mechanics, then for every $p\in P$ and $t\in T$ \[\sum_{q\in P}f(p,q,t)+g(p,t) = \sum_{q\in P'}f'(p,q,t)+g'(p,t).\]\label{somaforcas} \end{teorema} The imbedding theorem is the following: \begin{teorema} Every M.S.S. system is equivalent to a subsystem of an isolated system of particle mechanics.\label{Her} \end{teorema} \section{Classical Particle Mechanics Without Space-Time} In this section we define a set-theoretical predicate for a classical particle mechanics system, inspired on Tati's ideas. By set-theoretical predicate we mean Suppes predicate (Suppes-1967). We are aware about the limitations of particle mechanics, but that is not the point. The issue is Tati's picture for physical theories. Obviously we intend to apply these same ideas on other physical theories. But that is a task for future papers. Our main goal in the present work is to give an axiomatic framework for a simple case of physical theory, where space-time is not stated as one of the primitive notions. The simplest case of a physical theory, in our opinion, is M.S.S. system. In this paragraph we settle some notational features to be used in the paper from now on. We denote the set of real numbers by $\Re$, the set of integer numbers by ${\bf Z}$, the set of positive integers by ${\bf Z}^+$ and the cartesian products $\Re\times\Re\times\Re$ and ${\bf Z}\times{\bf Z}\times{\bf Z}$, respectively by $\Re^3$ and ${\bf Z}^3$. When there is no risk of confusion, we say that $\Re^3$ is a real vector space. We say that $h$ is a $(C^0,\tau)$-function iff $h = h(\tau)$ and $h$ is continuous with respect to $\tau$. Moreover, we say that $h$ is a $(C^k,\tau)$-function iff $h = h(\tau)$ and it is continuous and continuously differentiable (with respect to $\tau$) $k$ times. Our system has sixteen primitive concepts: $P$, $I$, $T$, $m$, $\bar{s}$, $\bar{f}$, $\bar{g}$, $s$, $v$, $f$, $g$, $c_s$, $c_v$, $c_f$, $c_g$, and $c_t$. \begin{definicao} ${\cal F} = \langle P, I, T, m, \bar{s}, \bar{f}, \bar{g}, s, v, f, g, c_s, c_v, c_f, c_g, c_t\rangle$ is a {\em non-relativistic classical particle mechanics system without space-time}, which we abbreviate as CM-Tati's system, if and only if the following axioms are satisfied: \begin{description} \item[CM-1] $P$ is a non-empty finite set; \item[CM-2] $I\subset {\bf Z^+}$; \item[CM-3] $T$ is an interval of real numbers; \item[CM-4] $m:P\to{\bf Z}^+$ is a function whose images are denoted by $m^p$; \item[CM-5] $s:P\times I\to{\bf Z}^3$ is a function whose images are denoted by $s_i^p$, where $i\in I$ and $p\in P$. Yet, if $p\neq p'$ then $s_i^p\neq s_i^{p'}$; \item[CM-6] $v:P\times I\to{\bf Z}^3$ is a function whose images are denoted by $v_i^p$, where $i\in I$ and $p\in P$; \item[CM-7] $f:P\times P\times I\to{\bf Z}^3$ is a function whose images are denoted by $f_i^{pq}$, where $i\in I$ and $p, q \in P$; \item[CM-8] $g:P\times I\to{\bf Z}^3$ is a function whose images are denoted by $g_i^p$, where $i\in I$ and $p\in P$; \item[CM-9] $c_t:I\to T$ is a function whose images are denoted by $c_t(i)$, where $i\in I$; \item[CM-10] $\bar{s}:P\times T\to\Re^3$ is a $(C^2,\tau)$-function whose images are denoted by $\bar{s}^p(\tau)$, satisfying the following property: if there exists $i\in I$ such that $\tau = c_t(i)$, then $\bar{s}^p(\tau) = s_i^p$; \item[CM-11] If there exists $i\in I$ such that $\tau = c_t(i)$, then $\frac{d}{d\tau}\bar{s}^p(\tau) = v_i^p$; \item[CM-12] $\bar{f}:P\times P\times T\to\Re^3$ is a $(C^0,\tau)$-function whose images are denoted by $\bar{f}^{pq}(\tau)$, satisfying the following property: if there exists $i\in I$ such that $\tau = c_t(i)$, then $\bar{f}^{pq}(\tau) = f_i^{pq}$; \item[CM-13] $\bar{g}:P\times T\to\Re^3$ is a $(C^0,\tau)$-funcion whose images are denoted by $\bar{g}^p(\tau)$, satisfying the following property: if there exists $i\in I$ such that $\tau = c_t(i)$, then $\bar{g}^p(\tau) = g_i^p$; \item[CM-14] For all $p,q\in P$ and $\tau\in T$ we have $\bar{f}^{pq}(\tau) = -\bar{f}^{qp}(\tau)$; \item[CM-15] For all $p,q\in P$ and $\tau\in T$ we have $[\bar{s}^p(\tau), \bar{f}^{pq}(\tau)] = -[\bar{s}^p(\tau), \bar{f}^{qp}(\tau)]$; \item[CM-16] For all $p,q\in P$ and $\tau\in T$ we have \[m^p\frac{d^{2}\bar{s}^p(\tau)}{d\tau^2} = \sum_{q\in P}\bar{f}^{pq}(\tau) + \bar{g}^p(\tau);\] \item[CM-17] $c_s$ is a recursive function such that $s_{i+1}^p = c_s(s_i^p,v_i^p,f_i^{pq},g_i^p)$; \item[CM-18] $c_v$ is a recursive function such that $v_{i+1}^p = c_v(s_i^p,v_i^p,f_i^{pq},g_i^p)$; \item[CM-19] $c_f$ is a recursive function such that $f_{i+1}^{pq} = c_f(s_i^p,v_i^p,f_i^{pq},g_i^p)$; \item[CM-20] $c_g$ is a recursive function such that $g_{i+1}^p = c_g(s_i^p,v_i^p,f_i^{pq},g_i^p)$; \item[CM-21] The diagram \[P\times P\times I\stackrel{\mbox{$\gamma$}}{\longrightarrow}{\bf Z}^3\times{\bf Z}^3\times{\bf Z}^3\times{\bf Z}^3\] \[\varphi\downarrow\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\downarrow c\;\;\;\] \[P\times P\times I\stackrel{\mbox{$\gamma$}}{\longrightarrow}{\bf Z}^3\times{\bf Z}^3\times{\bf Z}^3\times{\bf Z}^3\] \noindent commutes, where $\varphi(p,q,i) = (p,q,i+1)$, $\gamma(p,q,i) = (s_i^p,v_i^p,f_i^{pq},g_i^p)$ and $c(s_i^p,v_i^p,f_i^{pq},g_i^p) =$\\ $(c_s(s_i^p,v_i^p,f_i^{pq},g_i^p),c_v(s_i^p,v_i^p,f_i^{pq},g_i^p),c_f(s_i^p,v_i^p,f_i^{pq},g_i^p),c_g(s_i^p,v_i^p,f_i^{pq},g_i^p))$; \end{description} \end{definicao} This paragraph follows with some aditional intuitive hints about our axioms. $P$ corresponds to the set of particles. Axiom {\bf CM-1} says that we are dealing only with a finite number of particles. Axiom {\bf CM-2} says that $I$ is a set of positive integers; intuitivelly, each $i\in I$ corresponds to an {\em observation\/}. When we refer to observations we talk about either {\em performed\/} observations or {\em potentially performable\/}(in a sense) observations. Axiom {\bf CM-3} says that $T$ is an interval of real numbers. The intuitive meaning of $T$ is clarified when we discuss about the function $c_t$. In M.S.S. system, $T$ is interpreted as time. We say that $m^p$, in axiom {\bf CM-4}, corresponds to the (inertial) mass of particle $p$, which is a positive integer number. Such a condition demands an adequate measurement unit for mass, obviously different from the usual units. $s_i^p$ in axiom {\bf CM-5} corresponds to the position of particle $p$ at the $i$-th observation, while $v_i^p$ in axiom {\bf CM-6} is the speed of particle $p$ at observation $i$. In axiom {\bf CM-7} $f_i^{pq}$ corresponds to the internal force that particle $q$ exerts over $p$, at the $i$-th observation. In the next axiom $g_i^p$ is interpreted as the external force over $p$ at observation $i$. Function $c_t$ in axiom {\bf CM-9} is the correspondence that physicists make between their observations and the working of an ideal (in some sense) chronometer, represented by $T$. It should be emphasized that we are not imposing that $c_t$ is an injective function. That means that we may have causal relations without any passage of a `time' interval $[0,\tau]$, which is a very common situation in quantum mechanics as well as in the usual descriptions for classical mechanics. See, for instance, the problem of instantaneous actions-at-a-distance in newtonian mechanics. Functions $\bar{s}$, $\frac{d}{d\tau}\bar{s}$, $\bar{f}$ and $\bar{g}$ in axioms {\bf CM-10}, {\bf CM-11}, {\bf CM-12} and {\bf CM-13} are extensions of, respectively, $s$, $v$, $f$ and $g$. Such extensions allow us to use differential and integral calculus according to axiom {\bf CM-16}, although there is no physical interpretation for $\bar{s}$, $\frac{d}{d\tau}\bar{s}$, $\bar{f}$, and $\bar{g}$ when there is no $i\in I$ such that $\tau = c_t(i)$. Axioms {\bf CM-14}, {\bf CM-15} and {\bf CM-16} are analogous to axioms {\bf P5}, {\bf P6} and {\bf P7} in M.S.S. system. Functions $c_s$, $c_v$, $c_f$ and $c_g$ are called `causal relations'. The commutative diagram in axiom {\bf CM-21} is necessary in order to state a sort of compatibility between the causal relations and positions, speeds and forces. For those who are not familiar with the language of category theory, we say that the diagram in axiom {\bf CM-21} commutes when $\gamma\circ c = \varphi\circ\gamma$, where $\circ$ referes to functions composition. In the following theorems we denote the imbedding function by $1$. \begin{teorema}\label{1} The diagram \[P\times I\stackrel{\mbox{\it s}}{\longrightarrow}{\bf Z}^3\] \[\;\;\;\alpha\downarrow\;\;\;\;\;\;\;\;\;\;\;\;\downarrow 1\] \[P\times T\stackrel{\bar{\mbox{\it s}}}{\longrightarrow}\Re^3\] \noindent commutes, where $\alpha(p,i) = (p,c_t(i))$. \end{teorema} \noindent {\bf Proof:} direct from axiom {\bf CM-10}. \begin{teorema}\label{2} The diagram \[P\times I\stackrel{\mbox{\it v}}{\longrightarrow}{\bf Z}^3\] \[\;\;\;\alpha\downarrow\;\;\;\;\;\;\;\;\;\;\;\;\downarrow 1\] \[P\times T\stackrel{\frac{d}{d\tau}\bar{\mbox{\it s}}}{\longrightarrow}\Re^3\] \noindent commutes, where $\alpha(p,i) = (p,c_t(i))$. \end{teorema} \noindent {\bf Proof:} direct from axiom {\bf CM-11}. \begin{teorema}\label{3} The diagram \[P\times P\times I\stackrel{\mbox{\it f}}{\longrightarrow}{\bf Z}^3\] \[\;\;\;\;\;\;\beta\downarrow\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\downarrow 1\] \[P\times P\times T\stackrel{\bar{\mbox{\it f}}}{\longrightarrow}\Re^3\] \noindent commutes, where $\beta(p,q,i) = (p,q,c_t(i))$. \end{teorema} \noindent {\bf Proof:} direct from axiom {\bf CM-12}. \begin{teorema}\label{4} The diagram \[P\times I\stackrel{\mbox{\it g}}{\longrightarrow}{\bf Z}^3\] \[\;\;\;\alpha\downarrow\;\;\;\;\;\;\;\;\;\;\;\;\downarrow 1\] \[P\times T\stackrel{\bar{\mbox{\it g}}}{\longrightarrow}\Re^3\] \noindent commutes, where $\alpha(p,i) = (p,c_t(i))$. \end{teorema} \noindent {\bf Proof:} direct from axiom {\bf CM-13}.\\ The commuting diagram in axiom {\bf CM-21} describes the dynamics of our discrete picture for physical phenomena. Obviously, such a dynamics does not depend on time $T$. Nevertheless, the diagrams of theorems (\ref{1}), (\ref{2}), (\ref{3}) and (\ref{4}) show the compatibility of the discrete picture and the continuous description given in axioms {\bf CM-10} $\sim$ {\bf CM-16}, which are in correspondence, in a certain sense, with some of the axioms of M.S.S. system. Those diagrams suggest that the natural language for CM-Tati's system is category theory. It seems reasonable to describe the causal relations $c_s$, $c_v$, $c_f$ and $c_g$ as morphisms of a given category, whose objects are in correspondence with the observations of $s_i^p$, $v_i^p$, $f_i^{pq}$ and $g_i^p$. The definitions of subsystem, isolated system and equivalent systems, as well as the corresponding theorems (see Section 2), can be easily stated. \section{CM-Tati's System is Consistent} In this section we present a model for CM-Tati's system. Axioms {\bf CM-1}, {\bf CM-3} and {\bf CM-10}$\sim${\bf CM-16} are coincident with M.S.S. system assumptions up to function $c_t$ and to the fact that mass is a positive integer. There is no trouble with respect to integer mass. We may assume M.S.S. system with integer masses as our model. We also consider that $c_t$ is the imbedding function, if we make an appropriate choice for $I$ and $T$ ($I = \{1,2,3,4\}$ and $T = [1,4]$, for example). Now, the question is: how to interpret functions $s$, $v$, $f$, $g$, $c_s$, $c_v$, $c_f$, and $c_g$ and the diagram given in axiom {\bf CM-21}. Let us consider $f$ and $g$ as constants with respect to $i$, i.e., $\forall i \forall i' ((i\in I, i'\in I) \to (f_i^{pq} = f_{i'}^{pq} \wedge g_i^p = g_{i'}^p))$, where $\wedge$ and $\forall$ stand, respectivelly, for the logical connective `and' and the universal quantifier. Consider also that $\bar{f}$ and $\bar{g}$ are constants with respect to $\tau$ and that $\forall i( i\in I \to c_f(s_i^p,v_i^p,f_i^{pq},g_i^p) = f_i^{pq} \wedge c_g(s_i^p,v_i^p,f_i^{pq},g_i^p) = g_i^p)$, i.e., $f_{i+1}^{pq} = f_i^{pq}$ and $g_{i+1}^p = g_i^p$. Under these assumptions, the solution of the differential equation given in axiom {\bf CM-16} is: \begin{equation} \bar{s}^p(\tau) = \bar{s}^p(\tau_0) + \frac{d\bar{s}^p(\tau_0)}{d\tau}(\tau-\tau_0) + \frac{1}{2}\frac{\sum_{q\in P}\bar{f}^{pq}(\tau_0) + \bar{g}^p(\tau_0)}{m^p}(\tau-\tau_0)^2.\label{horas} \end{equation} This solution is valid, in particular, for integer values $\tau_0 = i$ and $\tau = i+1$ according to our statement that $c_t$ is the imbedding function. In this case it is easy to verify the commutativity of the diagram in axiom {\bf CM-21} if we consider that $s_{i+1}^p = c_s(s_i^p,v_i^p,f_i^{pq},g_i^p)$ is equivalent to equation (\ref{horas}) when $\tau = i+1$ and $\tau_0 = i$. And adequate choice for the values of forces $f$ and $g$, as well as to the mass $m$, will conduct to integer solutions $s$ of equation \ref{horas}. That means that we should choose adequate measurement units for mass, force, speed and position. \section{Final Remarks} There are some open problems related to our discrete picture for classical particle mechanics: \begin{enumerate} \item How to describe transformations of coordinates systems in ${\bf Z}^3$? What is a coordinate system in ${\bf Z}^3$? Concepts like invariance and covariance should be revised. We know that ${\bf Z}^3$ may be taken as a free ${\bf Z}$-module, where ${\bf Z}$ is a ring, i.e., ${\bf Z}^3$ may have a basis and a respective dimension. We may also define rotations and translations with respect to a given basis. Nevertheless, a group of transformations in this discrete space is a task for future works. \item How to extend such ideas to continuum mechanics? Is there any continuum mechanics in this picture? \item How to extend our picture to quantum mechanics or quantum field theory? That is an interesting point, since in EPR (Einstein-Podolsky-Rosen experiment) we have an instantaneous and non-local interaction between two elementary particles. If we extend our picture to quantum physics we may be able to explain EPR as a causal relation with no elapsed time interval, by considering that $c_t$ is not injective. That does not violate special relativity if we consider that relativity holds only when $c_t$ is injective. \item How to make a discrete picture for physical theories, without any reference to continuum spaces? It is well known that Turing machines do not provide an adequate notion for computability in the set of real numbers. If a discrete picture for physics is possible, some undecidable problems which are very common in theoretical physics should disappear. \end{enumerate} These questions cannot be answered in this paper. But we certainly intend to do it in future works. \section{Acknowledgments} We gratefully acknowledge the important criticisms and suggestions made by D\'ecio Krause (Universidade Federal do Paran\'a, Brazil), Heinz-J\"urgen Schmidt (Universit\"at Osnabr\"uck, Germany), and Newton da Costa (Universidade de S\~ao Paulo, Brazil). We were also benefitted by criticisms of one anonymous referee of this journal.\\\\\\ \noindent {\em References} \small \begin{enumerate} \item Einstein, A., 1991, `On the ether', in S. Saunders and H.R. Brown (eds.) {\em The Philosophy of Vacuum}, Oxford Un. Press, New York, 13-20. Originally published as `\"Uber den \"Ather', 1924, {\em Schweizerische naturforschende Gesellschaft, Verhanflungen\/}, {\bf 105}, 85-93. \item Hertz, H.R., 1894, {\em Die Prinzipien der Mechanik in Neuem Zusammenhange Dargestellt}, Barth, Leipzig. \item Hertz, H.R., 1956, {\em The Principles of Mechanics}, English translation by D.E. Jones and J.T. Walley, Dover Publications, New York. \item McKinsey, J.C.C., A.C. Sugar and P. Suppes, 1953, `Axiomatic foundations of classical particle mechanics', {\em J. Rational Mechanics and Analysis}, {\bf 2} 253-272. \item{Sant'Anna, A. S., 1995, `Set-theoretical structure for Hertz's mechanics', in {\em Volume of Abstracts of the 10th International Congress of Logic, Methodology and Philosophy of Science} (LMPS95), International Union of History and Philosophy of Science, 491-491.} \item Sant'Anna, A.S., 1996, `An axiomatic framework for classical particle mechanics without force', {\em Philosophia Naturalis} {\bf 33} 187-203. \item Suppes, P., 1957, {\em Introduction to Logic}, Van Nostrand, Princeton. \item Suppes, P., 1967, {\em Set-Theoretical Structures in Science}, mimeo. Stanford University, Stanford. \item Suppes, P., 1990, `Philosophy and the sciences', in W. Sieg (editor) {\em Acting and Reflecting}, Kluwer Academic, Dordrecht, 3-30. \item Tati, T., 1964, `Concepts of space-time in physical theories', {\em Prog. Theor. Phys. Suppl.} {\bf 29} 1-96. \item Tati, T., 1983, `The theory of finite degree of freedom', {\em Prog. Theor. Phys. Suppl.} {\bf 76} 186-223. \item Tati, T., 1986, `Macroscopic world and microscopic world: a physical picture of space-time', {\em Annals of the Japan Association for Philosophy of Science} $\bf{7}$ 15-31. \item Tati, T., 1987, `Local quantum mechanics'', {\em Prog. Theor. Phys.} $\bf{78}$ 996-1008. \end{enumerate} \end{document}
\section*{Acknowledgments} This work was supported in part by the DOE (at Chicago, Fermilab and Case Western Reserve) and by the NASA (at Fermilab through grant NAG 5-7092).
\section{Introduction} \label{sec_intro} A central concept in the theory of critical phenomena is the idea of {\em universality}\/, which states that phase-transition systems can be divided into a relatively small number of ``universality classes'' (determined primarily by the system's spatial dimensionality and the symmetries of its order parameter) within which certain features of critical behavior are universal. In the 1950s and 1960s it came to be understood that critical exponents are universal in this sense \cite{Domb_96}. Later, in the 1970s, it was learned that certain dimensionless ratios of critical amplitudes are also universal \cite{Privman_91}. The past quarter-century has seen enormous progress in the determination of critical exponents for a wide variety of universality classes, including exact analytical results for two-dimensional (2D) models \cite{Nienhuis_87,Cardy_87,Itzykson_88,DiFrancesco_97} and increasingly precise numerical determinations for three-dimensional models by a variety of techniques (field-theoretic renormalization group \cite{Zinn_Justin_book,Guida_98}, series extrapolation \cite{Bhanot_92,Janke_93,Guttmann_93,Guttmann_94,% Arisue_95,Butera_97a,Butera_97b,Butera_99}, Monte Carlo \cite{Holm_93,Gottlob_93,Li_95,Bloete_95,Nightingale_96,Bloete_96,% Caracciolo_98,Ballesteros_98,Ballesteros_99,Tsypin_99}). As a result, attention has turned quite naturally to universal amplitude ratios: these include amplitude ratios in infinite volume and those in finite-size scaling (FSS). Though much numerical work has been done, few exact results are known.\footnote{ Among the models studied are the 2D Ising model \cite{Zinn_96b,Pelissetto_98,Pelissetto_98a,Sokolov_98,Kim_99}, 2D nonlinear $\sigma$-models \cite{CEPS_o3_prl,CEMPS_LAT95_4loop,mgsu3,Pelissetto_98,Pelissetto_98a,% Balog_97,Balog_99}, 2D Potts models \cite{Caselle_99,Delfino_99}, the Baxter 8-vertex model \cite{Naon_89}, 2D and 3D self-avoiding walks \cite{Li_95}, the 3D Ising model \cite{Liu_Fisher_89,Kim-Patrascioiu_93,Bloete_95,Gutsfeld_96,% Zinn_Fisher_96,Zinn_96b,Guida_Zinn-Justin_97,Caselle_Hasenbusch_97,% Pelissetto_98,Pelissetto_98a,Pelissetto_98b,Ballesteros_98,Ballesteros_99,% Kim_99,Tsypin_99}, 3D $O(N)$ spin models \cite{Gottlob_93,Janke_98,Butera_98,Pelissetto_98,Pelissetto_98a,Butera_99,% Sokolov_99,Petkou_99}, 3D site percolation \cite{Ballesteros_99}, and the 5D Ising model \cite{Luitjen_99}. This list of references is far from exhaustive. } The critical behavior of many 2D models can be studied analytically using conformal field theory (CFT) \cite{Cardy_87,Itzykson_88,DiFrancesco_97}. Many critical exponents have been determined exactly, along with a few universal amplitude ratios \cite{Wu_76,Di_Francesco_87,Di_Francesco_88,Cardy_88a,Cardy_88b,% Cardy_Saleur_89,Caracciolo_90,Cardy_Mussardo_93,Cardy_Guttmann_93,Smilga_97,% Delfino_98,Delfino_Cardy_98,Delfino_99}. The main goal of the present paper is to compute, using CFT, a few more universal amplitude ratios for the 2D Ising model and to test these predictions by a high-precision Monte Carlo study. The amplitude ratios considered here arise in finite-size scaling; they can be computed starting from the correlation functions of the critical 2D Ising model on a torus. The first class of quantities we study concern the shape of the magnetization distribution $\rho({\cal M})$ at criticality on a symmetric torus ($L_x = L_y$).\footnote{ One of the insights of conformal field theory is that universal finite-size-scaling properties (such as the universal amplitude ratios considered here) ought to be studied as analytic functions of the modular parameter $\tau$ of the torus. Nevertheless, we think that the case of a symmetric torus ($\tau = i$) is of sufficient practical importance in Monte Carlo simulations to warrant special attention. } We study the rescaled shape of this distribution (i.e.\ normalizing by its width $\< {\cal M}^2 \>^{1/2}$) as well as the dimensionless ratios of its moments, \begin{equation} V_{2n} \;=\; { \<{\cal M}^{2n}\> \over \<{\cal M}^2\>^n } \;. \label{def_ratios} \end{equation} We can also define the dimensionless cumulants \begin{equation} U_{2n} \equiv { \<{\cal M}^{2n}\>_{\rm conn} \over \<{\cal M}^2\>^n } \;. \label{def_ratios_bis} \end{equation} For any symmetric distribution $\rho({\cal M}) = \rho(-{\cal M})$ these satisfy\footnote{ These relations can be computed from the generating functions $$ \sum\limits_{n=1}^\infty {U_{2n} \over (2n)!} z^{2n} \;=\; \log\!\left( \sum\limits_{n=0}^\infty {V_{2n} \over (2n)!} z^{2n} \right) $$ with $V_0 = V_2 = 1$ and $V_{2n+1} = 0$. } \begin{subeqnarray} \slabel{def_U4} U_4 &=& V_4 - 3 \\ U_6 &=& V_6 - 15 V_4 + 30 \\ U_8 &=& V_8 - 28 V_6 - 35 V_4^2 + 420 V_4 - 630 \\ U_{10} &=& V_{10} - 45 V_8 - 210 V_4 V_6 + 1260 V_6 + 3150 V_4^2 - 18900 V_4 + 22680 \\ & \vdots & \nonumber \end{subeqnarray} Note that $V_4$ and $U_4$ are closely related to the so-called Binder cumulant \cite{Binder_81} \begin{equation} U_{\rm 4,Binder} \equiv 1 - {\<{\cal M}^{4}\>\over 3 \<{\cal M}^2\>^2} = 1 - {V_4 \over 3} = - {U_4 \over 3} \;. \label{def_Binder_cumulant} \end{equation} For $\beta<\beta_c$ the ratios $V_{2n}$ tend in the infinite-volume limit to those characteristic of a Gaussian distribution, \begin{subeqnarray} V_{2n}({\rm Gaussian}) & = & (2n-1)!! \slabel{V2n_Gaussian} \\ U_{2n}({\rm Gaussian}) & = & 0 \label{ratios_Gaussian} \end{subeqnarray} while for $\beta>\beta_c$ they tend to those characteristic of a sum of two delta functions, \begin{subeqnarray} V_{2n}(\mbox{\rm two deltas}) & = & 1 \slabel{V2n_2deltas} \\ U_{2n}(\mbox{\rm two deltas}) & = & {2^{2n-1} (2^{2n} - 1) \over n} \, B_{2n} \label{ratios_2deltas} \end{subeqnarray} where $B_{2n} = (-1)^{n-1} (2n)! \, \zeta(2n) / 2^{2n-1} \pi^{2n}$ is a Bernoulli number. At $\beta=\beta_c$, however, these ratios acquire non-trivial values in-between \reff{V2n_Gaussian} and \reff{V2n_2deltas}.\footnote{ The Schwarz inequality implies that $V_{2n} \ge 1$ for any model, and the Gaussian inequality \cite{Newman_75a,Bricmont_77} implies that $V_{2n} \le (2n-1)!!$ for ferromagnetic Ising models. In particular, we have $-2 \le U_4 \le 0$. Moreover, Newman \cite{Newman_75b} and Shlosman \cite{Shlosman_86} have proven, for ferromagnetic Ising models, that $(-1)^{n-1} U_{2n} \ge 0$ for all $n$; and Newman \cite{Newman_75b} has proven some additional inequalities on the $U_{2n}$. } These values, which are universal, can in principle be computed by integrating the spin correlators for the critical 2D Ising model on a torus, which were determined by Di Francesco {\em et al.}\/ \cite{Di_Francesco_87,Di_Francesco_88} using CFT. In practice, however, the formula for $V_{2n}$ rapidly gets more complicated as $n$ grows. Di Francesco {\em et al.}\/ \cite{Di_Francesco_87,Di_Francesco_88} computed $V_4$ to roughly three decimal places by Monte Carlo integration. Here we shall improve this result by three orders of magnitude, and shall also compute $V_6$ to five decimal places, $V_8$ to almost four decimal places, and $V_{10}$ to three decimal places: \begin{eqnarray} V_4 & = & 1.1679229 \pm 0.0000047 \\ V_6 & = & 1.4556491 \pm 0.0000072 \\ V_8 & = & 1.89252 \pm 0.00018 \\ V_{10} & = & 2.53956 \pm 0.00034 \end{eqnarray} Finally, we shall measure the ratios $V_{2n}$ for $2 \le n \le 10$ by Monte Carlo simulation, with an accuracy that gradually deteriorates as $n$ grows. For $n=2,3,4,5$ our Monte Carlo estimates agree well with the theoretical predictions (but are of course less precise). Another interesting quantity is the second-moment correlation length $\xi$. It has a sensible definition in finite volume (see Section \ref{sec_simul}), and its expected FSS behavior is \begin{equation} \xi \;\sim\; L \left[ x^\star + A L^{-\Delta} + \cdots \right] \;, \label{def_FSS_xi} \end{equation} where the leading coefficient \begin{equation} x^\star \;=\; \lim_{L\rightarrow\infty} \xi/L \end{equation} is universal.\footnote{ The quantity $x^\star$ also plays an important role in a recently developed method for extrapolating finite-volume Monte Carlo data to infinite volume \cite{Luscher_91,Kim_93,fss_greedy,mgsu3,fss_greedy_fullpaper}. }${}^,$\footnote{ An analogous situation holds in a cylindrical ($L \times \infty$) geometry for the exponential correlation length in the longitudinal direction, $\xi_{exp}(L)$ [which can be defined in terms of the logarithm of the ratio of the two largest eigenvalues of the transfer matrix]. Privman and Fisher \protect\cite{Privman_Fisher_84} showed that $\lim\limits_{L \to\infty} \xi_{exp}(L)/L$ at criticality is universal, and Cardy \protect\cite{Cardy_84} showed that for 2D conformal-invariant systems it is equal to $1/(\pi \eta)$. } Here we shall compute $x^\star$ for the 2D Ising model at criticality by numerically integrating the known spin correlators \cite{Di_Francesco_87,Di_Francesco_88}; we find \begin{equation} x^\star \;=\; 0.90504 88292 \pm 0.00000 00004 \;. \end{equation} Our Monte Carlo data confirm this prediction. To our knowledge, this is the first exact determination of $x^\star$ for any universality class. Monte Carlo estimates of $x^\star$ are available for many other 2D models --- including the 3-state Potts model \cite{Salas_Sokal_Potts3}, the 4-state Potts model \cite{Salas_Sokal_AT,Salas_Sokal_FSS}, the 3-state square-lattice Potts antiferromagnet \cite{Salas_Sokal_swaf3,Ferreira_Sokal_prep}, the XY model \cite{Kim_private}, and several points on the self-dual curve of the symmetric Ashkin--Teller model \cite{Salas_Sokal_AT} --- and it would be very interesting to compute $x^\star$ analytically for some of these models. Numerical estimates of $x^\star$ are also available for some three-dimensional spin models \cite{Gottlob_93,Kim-Patrascioiu_93,Janke_98,Ballesteros_98}. Consider, finally, the dimensionless renormalized four-point coupling constant \begin{equation} g_4 \;=\; - \, {\bar{u}_4 \over \chi^2 \xi^d} \;=\; - \, {U_4 \over (\xi/L)^d} \;, \end{equation} where $\bar{u}_4$ is the connected four-point function at zero momentum. In the FSS limit $L \to \infty$, $\beta \to \beta_c$ with $\xi/L$ fixed, $g_4$ is a nontrivial function of the FSS variable $\xi/L$: \begin{equation} g_4 \;=\; F_{g_4}(\xi/L) \;. \end{equation} Therefore, the function $g_4(\beta,L)$ fails to be jointly continuous at $(\beta,L) = (\beta_c,\infty)$; many limiting values are possible depending on the mode of approach, and the massive and massless scaling limits \begin{eqnarray} g_4^* & = & \lim\limits_{\beta \uparrow \beta_c} \lim\limits_{L \to \infty} g_4(\beta,L) \\[1mm] G_4^* & = & \lim\limits_{L \to \infty} \lim\limits_{\beta \uparrow \beta_c} g_4(\beta,L) \;=\; \lim\limits_{L \to \infty} g_4(\beta_c,L) \end{eqnarray} correspond to the two extreme cases $g_4^* = F_{g_4}(0)$, $G_4^* = F_{g_4}(x^\star)$. As a corollary of our computation of $V_4$ and $x^\star$, we obtain the value of $g_4$ at criticality on a symmetric torus: \begin{equation} G_4^* \;=\; - \, {U_4 \over x^{\star 2}} \;=\; 2.2366587 \pm 0.0000057 \;. \end{equation} More generally, consider the dimensionless renormalized $2n$-point coupling constant \begin{equation} g_{2n} \;=\; {\chi^n \, \bar{\Gamma}_{2n} \over \xi^{(n-1)d}} \;, \end{equation} where $\bar{\Gamma}_{2n}$ is the amputated one-particle-irreducible $2n$-point function at zero mo\-men\-tum.\footnote{ The $\bar{\Gamma}_{2n}$ are defined by the generating-function relation $$ \sum\limits_{n=1}^\infty {\bar{\Gamma}_{2n} \over (2n)!} \phi^{2n} \;=\; \sup_J \left[ J\phi \,-\, \sum\limits_{n=1}^\infty {\bar{u}_{2n} \over (2n)!} J^{2n} \right] \;. $$ Recall that $\bar{u}_2 = \chi$ and that $\bar{u}_{2n} = L^{-d} \<{\cal M}^{2n}\>_{\rm conn}$. % % % } We can predict the next three renormalized coupling constants at criticality on a symmetric torus: \begin{subeqnarray} G_6^* &=& - \, {U_6 - 10 U_4^2 \over x^{\star 4}} \;=\; 29.25457 \pm 0.00015 \\[1mm] G_8^* &=& - \, {U_8 - 56 U_4 U_6 + 280 U_4^3 \over x^{\star 6}} \;=\; 942.6095 \pm 0.0072\\[1mm] G_{10}^* &=& - \, {U_{10} - 120 U_4 U_8 - 126 U_6^2 + 4620 U_4^2 U_6 - 15400 U_4^4 \over x^{\star 8}} \nonumber \\ & & \hspace{4.5cm} =\; 56110.24 \pm 0.56 \end{subeqnarray} In addition, we shall provide Monte Carlo estimates of $G_{12}^*$ through $G_{20}^*$ [cf.\ \reff{G_2n_MC_results} below]. This paper is organized as follows: In Section~\ref{sec_theor} we review the relevant exact results available for the 2D Ising model at criticality on a torus, and we compute (by numerical integration) the CFT prediction for the quantities $x^\star$, $V_4$, $V_6$, $V_8$ and $V_{10}$. In Section~\ref{sec_simul} we explain the Monte Carlo algorithm we have used to simulate this model. In Section~\ref{sec_res_static} we analyze our numerical results for the static observables and compare them against the available exact results. Finally, in Section~\ref{sec_conclusions} we present our final conclusions and discuss prospects for future work. In Appendix~\ref{app_integrals} we explain how we carried out the numerical integrations involved in computing $x^\star$, and in Appendix~\ref{app_theta} we summarize the definitions and principal properties of the Jacobi theta functions. \section{Theoretical Results} \label{sec_theor} The universal amplitudes we consider in this paper ($x^\star$ and $V_{2n}$) can be written in terms of integrals of the $2n$-point spin correlation functions of the the critical 2D Ising continuum field theory on a torus. These correlators were obtained by Di Francesco {\em et al}.\ \cite{Di_Francesco_87,Di_Francesco_88} using an approach based on conformal field theory. The result is\footnote{ There is a misprint in the normalization of the 4-spin correlator in equation (9) of \protect\cite{Di_Francesco_88}, and in the normalization of the $2n$-spin correlator in equation (6.6) of \protect\cite{Di_Francesco_87}. We have rederived both correlators using the chiral bosonization prescription presented in \protect\cite{Di_Francesco_87}. With the correct normalization, shown in (\protect\ref{def_ising_spin_correlators})--% (\protect\ref{def_ising_spin_correlators_full}) below, we are able to reproduce the numerical value of $V_4$ reported in \protect\cite{Di_Francesco_88}, as well as the numerical estimates of $V_4$, $V_6$, $V_8$ and $V_{10}$ obtained in our simulation. } \begin{equation} \<\sigma_{z_1}\cdots\sigma_{z_{2n}}\> \;=\; {\sum_{\nu=1}^4 Z_\nu \<\sigma_{z_1}\cdots\sigma_{z_{2n}}\>_\nu \over \sum_{\nu=1}^4 Z_\nu } \label{def_ising_spin_correlators} \end{equation} where \begin{equation} Z_\nu \;=\; {|\theta_\nu(0)| \over 2 |\eta|} \label{def_ising_partition} \end{equation} and \begin{subeqnarray} Z^2_\nu \<\sigma_{z_1}\cdots\sigma_{z_{2n}}\>^2_\nu &=& {1 \over 2^{n+2} |\eta|^2} \sum_{ \stackrel{\epsilon_j=\pm1}{\sum_j \epsilon_j=0}} \left| \theta_\nu \!\left({\sum_j \epsilon_j z_j \over 2}\right)\right|^2 \prod_{i<j} \left| {\theta_1(z_i-z_j)\over \theta_1^\prime(0)}\right|^{\epsilon_i \epsilon_j / 2} \qquad \\[1mm] &=& \label{def_ising_spin_correlators_full} {\theta_1^\prime(0)^{n/2} \over 2^{n+2} |\eta|^2} \sum_{ \stackrel{\epsilon_j=\pm1}{\sum_j \epsilon_j=0}} \left| \theta_\nu \!\left({\sum_j \epsilon_j z_j \over 2}\right)\right|^2 \prod_{i<j} \left| \theta_1(z_i-z_j) \right|^{\epsilon_i \epsilon_j / 2} \qquad \end{subeqnarray} Here we have used the complex-number notation $z = x_1+i x_2$; $\theta'_1(0) \approx 2.8486946040$ is the derivative of $\theta_1(z,\tau)$ with respect to $z$ evaluated at $z=0$ and $\tau=i$; and $\eta \approx 0.7682254223$ is the usual Dedekind function $\eta(\tau)$ evaluated at $\tau=i$. Please note that $\theta'_1(0) = 2\pi \eta^3$ [cf.\ (\protect\ref{eq_A.13})]. Note also that the contribution of $\{\epsilon_j\}$ to \reff{def_ising_spin_correlators_full} is equal to that of $\{-\epsilon_j\}$, so in the numerical evaluation of this expression we need only take half the terms. The expression \reff{def_ising_spin_correlators} gives the FSS limit for the Ising-model correlation functions at criticality: here $z_i$ denotes the position in lattice units divided by the lattice linear size $L$. \bigskip \noindent {\bf Remark}. Although the sector $\nu=1$ does not contribute to the partition function [since $Z_1 \sim \theta_1(0)=0$], it does contribute to the correlation functions [since $Z_1 \<\sigma_{z_1}\cdots\sigma_{z_{2n}}\>_1 \neq 0$]. So this sector cannot simply be discarded. See ref.~\cite{Di_Francesco_87} for details. \bigskip The two correlators that are needed in the evaluation of the Binder cumulant are \begin{eqnarray} \label{def_ising_2point} Z_\nu \<\sigma_{z_1}\sigma_{z_2}\>_\nu &=& {|\theta'_1(0)|^{1/4} \over 2 |\eta|} {\left|\theta_\nu \!\left({z_1-z_2\over2}\right)\right| \over |\theta_1(z_1 - z_2)|^{1/4}} \\[4mm] \label{def_ising_4point} Z_\nu \<\sigma_{z_1}\sigma_{z_2} \sigma_{z_3}\sigma_{z_4} \>_\nu &=& {|\theta'_1(0)|^{1/2}\over 2\sqrt{2} |\eta|} \left\{ \left|\theta_\nu \!\left({z_1+z_2-z_3-z_4\over2}\right)\right|^2 \right. \nonumber \\ & & \qquad \times \left| {\theta_1(z_1-z_2)\theta_1(z_3-z_4) \over \theta_1(z_1-z_3)\theta_1(z_1-z_4) \theta_1(z_2-z_3)\theta_1(z_2-z_4)} \right|^{1/2} \nonumber \\ & & \left. \phantom{1\over1} + (2\leftrightarrow3) + (2\leftrightarrow4) \right\}^{1/2} \end{eqnarray} We also need the 6-point correlator to compute $V_6$. Its exact expression can be deduced easily from the general equation \reff{def_ising_spin_correlators_full}: \begin{eqnarray} & & Z_\nu \<\sigma_{z_1}\sigma_{z_2} \sigma_{z_3}\sigma_{z_4} \sigma_{z_5}\sigma_{z_6} \>_\nu = {|\theta'_1(0)|^{3/4}\over 4 |\eta|} \left\{ \left|\theta_\nu \!\left({z_1+z_2+z_3-z_4-z_5-z_6\over2}\right)\right|^2 \right. \nonumber \\ & & \qquad \qquad \phantom{1\over1} \times \Psi(z_1,z_2,z_3,z_4,z_5,z_6) + (2\leftrightarrow4) + (2\leftrightarrow5) + (2\leftrightarrow6) \nonumber \\ & & \qquad \qquad \phantom{1\over1} + (3\leftrightarrow4) + (3\leftrightarrow5) + (3\leftrightarrow6) + (2\leftrightarrow4;3\leftrightarrow5) \nonumber \\ & & \left. \qquad \qquad \phantom{1\over1} + (2\leftrightarrow4;3\leftrightarrow6) + (2\leftrightarrow5;3\leftrightarrow6) \right\}^{1/2} \label{G6} \end{eqnarray} where the function $\Psi$ is defined as \begin{equation} \Psi(z_1,z_2,z_3,z_4,z_5,z_6) = \left( {\theta_1(z_{12})\theta_1(z_{13}) \theta_1(z_{23}) \theta_1(z_{45})\theta_1(z_{46}) \theta_1(z_{56}) \over \prod_{i=1,2,3;j=4,5,6} \theta_1(z_{ij}) } \right)^{1/2} \end{equation} and we have used the shorthand notation $z_{ij} \equiv z_i -z_j$. {}From these equations we can obtain the values of $x^\star = \lim_{L\rightarrow\infty} \xi/L$ and $V_{2n}$ by numerical integration. In particular, the correlation length on a periodic lattice of size $L$ is defined to be \begin{equation} \xi \;=\; {1 \over 2 \sin(\pi/L)} \left( {\chi \over F} - 1 \right)^{1/2} \;, \label{def_xi} \end{equation} where $\chi$ is the susceptibility (i.e., the Fourier-transformed two-point correlation function at zero momentum) and $F$ is the corresponding quantity at the smallest nonzero momentum $(2\pi/L,0)$ [see \reff{def_msquare}/\reff{def_f} and \reff{def_susceptibility}/\reff{def_f_density} below for details]. This is a finite-lattice generalization of the second-moment correlation length. Then, the universal amplitude $x^\star$ is given by \begin{equation} x^\star \;=\; {1 \over 2\pi} \left( {\chi \over F} - 1 \right) ^{\! 1/2} \end{equation} where \begin{eqnarray} \chi &\sim& \int d^2z \, \<\sigma_0 \sigma_z\> \\ F &\sim& \int d^2z \, \<\sigma_0 \sigma_z\> \, \cos (2 \pi x_1) \end{eqnarray} and $\int d^2z = \int_0^1 \int_0^1 dx_1 \; dx_2$. The details of this computation are given in Appendix~\ref{app_integrals}. We obtain \begin{eqnarray} \int d^2z \, \<\sigma_0 \sigma_z\> & = & 1.55243 29546 5 \pm 0.00000 00000 4 \label{chi_numerical} \\[2mm] \int d^2z \, \<\sigma_0 \sigma_z\> \, \cos (2 \pi x_1) & = & 0.04656 74468 2 \pm 0.00000 00000 4 \label{F_numerical} \end{eqnarray} As a result, we obtain $x^\star$ with 10 digits of precision: \begin{equation} x^\star = 0.90504 88292 \pm 0.00000 00004 \;. \label{xstar_theor} \end{equation} We repeated the computation requiring 11 digits of precision in the integrals, and the result was the same. The universal moment ratio $V_4$ is given by \begin{equation} V_4 \;=\; { \int d^2z_2 \, d^2z_3 \, d^2z_4 \, \<\sigma_{0}\sigma_{z_2} \sigma_{z_3}\sigma_{z_4} \> \over \left[ \int d^2z \, \<\sigma_0 \sigma_z\> \right]^2 } \;. \label{ratio_V4} \end{equation} Di Francesco {\em et al.}\/ \cite{Di_Francesco_87,Di_Francesco_88} performed the integrals in numerator and denominator by Monte Carlo and obtained \begin{equation} V_4 \;=\; 1.168 \pm 0.005 \;. \label{V4_DiFrancesco} \end{equation} We have improved this value, as follows: For the denominator of \reff{ratio_V4}, we use the very precise estimate \reff{chi_numerical} coming from deterministic numerical integration. For the numerator, we performed a Monte Carlo integration using $10^9$ measurements. Our result is \begin{equation} \label{V4_exact} V_4 \;=\; 1.1679 229 \pm 0.0000 047 \;, \end{equation} which is compatible with \reff{V4_DiFrancesco} but three orders of magnitude more precise. This value also agrees closely with the estimate of Kamieniarz and Bl\"ote \cite{Kamieniarz_93} based on extrapolation of the exact results (computed by transfer-matrix methods) for $L \le 17$: \begin{equation} V_4 \;=\; 1.1679 296 \pm 0.0000 014 \;, \label{v4_Kamieniarz} \end{equation} where the error bar is of course somewhat subjective.\footnote{ Unfortunately, Kamieniarz and Bl\"ote \cite{Kamieniarz_93} reported only meager details of the fits that led to this extraordinarily precise estimate. That is a shame, as information on the presence or absence of particular correction-to-scaling terms could be of considerable theoretical interest. } More generally, the universal moment ratio $V_{2n}$ is given by \begin{equation} V_{2n} \;=\; { \int d^2z_2 \, \cdots \, d^2z_{2n} \, \<\sigma_{0}\sigma_{z_2} \cdots \sigma_{z_{2n}} \> \over \left[ \int d^2z \, \<\sigma_0 \sigma_z\> \right]^n } \;. \end{equation} We have been able to compute the (exact except for the numerical integration) values of the ratios $V_6$, $V_8$ and $V_{10}$. We performed the integrals in the numerator by Monte Carlo, using $10^9$ measurements for $V_6$, $4\times 10^6$ measurements for $V_8$ and $2.5\times 10^6$ measurements for $V_{10}$. We obtain \begin{eqnarray} V_6 & = & 1.4556491 \pm 0.0000072 \label{V6_exact} \\ V_8 & = & 1.89252 \pm 0.00018 \label{V8_exact} \\ V_{10} & = & 2.53956 \pm 0.00034 \label{V10_exact} \end{eqnarray} In general, the formula for the $2n$-point function contains $(2n)! /[ 2 (n!)^2 ]$ terms [this takes into account the $\{\epsilon_j\} \leftrightarrow \{-\epsilon_j\}$ symmetry], and this grows asymptotically like $4^n$. Thus, in computing $V_4$ (resp.\ $V_6$, $V_8$, $V_{10}$) we had to include 3 (resp.\ 10, 35, 126) terms, and the computation of $V_{12}$ would require handling 462 terms. Moreover, the numerator has to be integrated over a $(4n-2)$-dimensional torus. These facts make the high-precision numerical integration of $V_{2n}$ extremely time-consuming as soon as $n$ becomes moderately large. Let us consider, finally, the dimensionless renormalized four-point coupling constant $g_4$ defined by \begin{equation} g_4 \;=\; - \, {\bar{u}_4 \over \chi^2 \xi^d} \;=\; - \, {U_4 \over (\xi/L)^d} \;, \end{equation} where $\bar{u}_4$ is the connected four-point function at zero momentum, and more generally the dimensionless renormalized $2n$-point coupling constant $g_{2n}$ defined by \begin{equation} g_{2n} \;=\; {\chi^n \, \bar{\Gamma}_{2n} \over \xi^{(n-1)d}} \;, \end{equation} where $\bar{\Gamma}_{2n}$ is the amputated one-particle-irreducible $2n$-point function at zero momentum. In the FSS limit $L \to \infty$, $\beta \to \beta_c$ with $\xi/L$ fixed, $g_{2n}$ becomes a nontrivial function of the FSS variable $\xi/L$, \begin{equation} g_{2n} \;=\; F_{g_{2n}}(\xi/L) \;. \end{equation} (There is some evidence that $F_{g_4}$ is a {\em decreasing}\/ function of $\xi/L$.\footnote{ Baker and Kawashima \cite{Baker_96} conjecture that $g_4(\beta,L)$ [resp.\ $\xi(\beta,L)$] is a decreasing [resp.\ increasing] function of $\beta$ for each fixed $L < \infty$; these two facts, if true, would immediately imply that $F_{g_4}(\xi/L)$ is a decreasing function of its argument $\xi/L$. Numerical data for the 2D \cite{Kim-Patrascioiu_93,Baker_94} and three-dimensional \cite{Kim-Patrascioiu_93,Baker_96} Ising models clearly support the Baker--Kawashima conjecture. }) In particular, the massive and massless scaling limits \begin{eqnarray} g_{2n}^* & = & \lim\limits_{\beta \uparrow \beta_c} \lim\limits_{L \to \infty} g_{2n}(\beta,L) \\[1mm] G_{2n}^* & = & \lim\limits_{L \to \infty} \lim\limits_{\beta \uparrow \beta_c} g_{2n}(\beta,L) \;=\; \lim\limits_{L \to \infty} g_{2n}(\beta_c,L) \end{eqnarray} correspond to the two extreme cases $g_{2n}^* = F_{g_{2n}}(0)$, $G_{2n}^* = F_{g_{2n}}(x^\star)$. The best currently available estimates for the 2D Ising model are \begin{eqnarray} g_4^* & = & \cases{14.694 \pm 0.002 & by high-temperature expansion \protect\cite{Butera_96,Pelissetto_98} \cr \noalign{\vskip 2mm} 14.66 \pm 0.42 & by $\epsilon$-expansion \protect\cite{Pelissetto_98} \cr \noalign{\vskip 2mm} 15.50 \pm 0.84 & by $g$-expansion \protect\cite{LeGuillou_80} \cr \noalign{\vskip 2mm} 14.66 \pm 0.06 & by expansion around $d=0$ \protect\cite{Bender_93,Bender_95} \cr \noalign{\vskip 2mm} 14.7 \pm 0.2 & by Monte Carlo \protect\cite{Kim_99} \cr } \\[4mm] G_4^* & = & 2.239 \pm 0.007 \qquad\hbox{by Monte Carlo \protect\cite{Kim_99}} \label{Gstar4_MC_Kim} \\[4mm] g_6^* & = & \cases{794.1 \pm 0.6 & by high-temperature expansion \protect\cite{Zinn_96b,Pelissetto_98,Pelissetto_98a} \cr \noalign{\vskip 2mm} 797 \pm 9 & by $\epsilon$-expansion \protect\cite{Pelissetto_98,Pelissetto_98a} \cr \noalign{\vskip 2mm} 792 \pm 40 & by $g$-expansion with constrained $g_4^*$ \protect\cite{Sokolov_98} \cr \noalign{\vskip 2mm} 691 \pm 29 & by expansion around $d=0$ \protect\cite{Bender_93,Bender_95} \cr \noalign{\vskip 2mm} 850 \pm 25 & by Monte Carlo \protect\cite{Kim_99} \cr } \\[4mm] G_6^* & = & 29.34 \pm 0.20 \qquad\hbox{by Monte Carlo \protect\cite{Kim_99}} \label{Gstar6_MC_Kim} \\[4mm] g_8^* & = & \cases{(82.5 \pm 0.6) \times 10^3 & by high-temperature expansion \protect\cite{Pelissetto_98,Pelissetto_98a} \cr \noalign{\vskip 2mm} (83.8 \pm 3.2) \times 10^3 & by $\epsilon$-expansion \protect\cite{Pelissetto_98,Pelissetto_98a} \cr \noalign{\vskip 2mm} (89 \pm 5) \times 10^3 & by Monte Carlo \protect\cite{Kim_99} \cr } \\[4mm] G_8^* & = & 947 \pm 10 \qquad\hbox{by Monte Carlo \protect\cite{Kim_99}} \label{Gstar8_MC_Kim} \\[4mm] g_{10}^* & = & \cases{(12.8 \pm 0.7) \times 10^6 & by high-temperature expansion \protect\cite{Pelissetto_98,Pelissetto_98a} \cr \noalign{\vskip 2mm} (8.0 \pm 1.4) \times 10^6 & by $\epsilon$-expansion \protect\cite{Pelissetto_98,Pelissetto_98a} \cr } \end{eqnarray} Our own Monte Carlo data, reported in Section \ref{sec4.Vn}, combined with the theoretical value \reff{xstar_theor} for $x^\star$, improve \reff{Gstar4_MC_Kim}/\reff{Gstar6_MC_Kim}/\reff{Gstar8_MC_Kim} to $G_4^* = 2.23685 \pm 0.00016$, $G_6^* = 29.2602 \pm 0.0047$ and $G_8^* = 942.91 \pm 0.25$, respectively, and also give values for $G_{10}^*$ through $G_{20}^*$ [see \reff{G_2n_MC_results} and the footnote following it]. {}From \reff{xstar_theor} and \reff{V4_exact}--\reff{V10_exact} we obtain the theoretical predictions \begin{subeqnarray} G_4^* &=& - \, {U_4 \over x^{\star 2}} \;=\; 2.2366587 \pm 0.0000057 \\[1mm] G_6^* &=& - \, {U_6 - 10 U_4^2 \over x^{\star 4}} \;=\; 29.25457 \pm 0.00015 \\[1mm] G_8^* &=& - \, {U_8 - 56 U_4 U_6 + 280 U_4^3 \over x^{\star 6}} \;=\; 942.6095 \pm 0.0072 \\[1mm] G_{10}^* &=& - \, {U_{10} - 120 U_4 U_8 - 126 U_6^2 + 4620 U_4^2 U_6 - 15400 U_4^4 \over x^{\star 8}} \nonumber \\ & & \hspace{4.5cm} =\; 56110.24 \pm 0.56 \label{G_2n_exact} \end{subeqnarray} The error bars in \reff{G_2n_exact} are obtained by carefully propagating the statistical errors from the $V_{2n}$ [in which the errors are independent except for an extremely small effect arising from the value of the denominator \reff{chi_numerical}] to the $U_{2n}$ (in which the errors are correlated) and thence to the $G_{2n}^*$. \section{Numerical Simulations} \label{sec_simul} We consider the two-dimensional nearest-neighbor Ising model on a $L \times L$ square lattice with periodic boundary conditions, given by the Hamiltonian \begin{subeqnarray} \slabel{Hamiltonian_Ising} {\cal H}_{\rm Ising} &=& -{\beta\over 2} \sum_{\< ij \>} \sigma_i\sigma_j \\ \slabel{Hamiltonian_Potts} &=& -\beta \sum_{\< ij \>} \delta_{\sigma_i,\sigma_j} \,+\, \hbox{const} \;. \end{subeqnarray} Note that we use throughout this paper a non-standard normalization of $\beta$, which is motivated by considering the Ising model as a special case of the $q$-state Potts model; it differs by a factor of 2 from the usual Ising normalization. In our normalization, the critical point is at \begin{equation} \beta_c \;=\; \log (1 + \sqrt{2}) \;\approx\; 0.881373587 \;. \end{equation} \subsection{Observables to be measured} We have performed simulations of this system using the Swendsen-Wang algorithm \cite{Swendsen_87,Edwards_Sokal,Sokal_Cargese}. In particular, we have measured the following basic observables: \begin{itemize} \item the energy density (i.e., the number of unsatisfied bonds) \begin{equation} {\cal E} \;\equiv\; \sum_{\<xy\>} (1 - \delta_{\sigma_x,\sigma_y} ) \label{def_energy} \end{equation} \item the bond occupation \begin{equation} {\cal N} \;\equiv\; \sum_{\<xy\>} n_{xy} \label{def_n} \end{equation} \item the nearest-neighbor connectivity (which is an energy-like observable \cite{Salas_Sokal_Potts3}) \begin{equation} {\cal E}' \;\equiv\; \sum_{\<xy\>} \gamma_{xy} \;, \label{def_energyp} \end{equation} where $\gamma_{xy}$ equals 1 if both ends of the bond $\<xy\>$ belong to the same cluster, and 0 otherwise. More generally, the connectivity $\gamma_{ij}$ can be defined for an arbitrary pair $i,j$ of sites: \begin{equation} \gamma_{ij}(\{n\}) \;=\; \left\{ \begin{array}{ll} 1 & \quad \hbox{\rm if $i$ is connected to $j$} \\ 0 & \quad \hbox{\rm if $i$ is not connected to $j$} \end{array} \right. \end{equation} \item the squared magnetization \begin{subeqnarray} {\cal M}^2 &=& \left( \sum_x \mbox{\protect\boldmath $\sigma$}_x \right)^2 \\ &=& {q \over q-1} \sum_{\alpha=1}^q \left( \sum_x \delta_{\sigma_x,\alpha} \right)^2 - {V^2 \over q-1} \label{def_msquare} \end{subeqnarray} where $\mbox{\protect\boldmath $\sigma$}_x \equiv {\bf e}^{(\sigma_x)} \in \hbox{{\bf R}}^{q-1}$ is the Potts spin in the hypertetrahedral representation\footnote{ Let $\{ {\bf e}^{(\alpha)} \} _{\alpha=1}^q$ be unit vectors in $\hbox{{\bf R}}^{q-1}$ satisfying ${\bf e}^{(\alpha)} \cdot {\bf e}^{(\beta)} = (q \delta^{\alpha\beta} - 1)/(q-1)$, and let $\mbox{\protect\boldmath $\sigma$}_x \equiv {\bf e}^{(\sigma_x)}$. For $q=2$ this means $\mbox{\protect\boldmath $\sigma$}_x = \cos (\pi \sigma_x) = \pm 1$. } and $V=L^2$ is the number of lattice sites \item powers of the squared magnetization \begin{equation} {\cal M}^{2n} \;=\; ({\cal M}^2)^n \end{equation} \item the square of the Fourier transform of the spin variable at the smallest allowed non-zero momentum \begin{subeqnarray} {\cal F} &=& {1 \over 2} \left( \left| \sum_x \mbox{\protect\boldmath $\sigma$}_x \, e^{2\pi i x_1/L} \right|^2 + \left| \sum_x \mbox{\protect\boldmath $\sigma$}_x \, e^{2\pi i x_2/L} \right|^2 \right) \\ &=& {q \over q-1} \times {1 \over 2} \sum_{\alpha=1}^q \left( \left| \sum_x \delta_{\sigma_x,\alpha} \, e^{2\pi i x_1/L} \right|^2 + \left| \sum_x \delta_{\sigma_x,\alpha} \, e^{2\pi i x_2/L} \right|^2 \right) \label{def_f} \end{subeqnarray} where $(x_1,x_2)$ are the Cartesian coordinates of point $x$. Note that ${\cal F}$ is normalized to be comparable to its zero-momentum analogue ${\cal M}^2$. \item the mean-square and mean-fourth-power size of the clusters \begin{eqnarray} {\cal S}_2 &=& \sum_{\cal C} \#({\cal C})^2 \label{def_s2} \\ {\cal S}_4 &=& \sum_{\cal C} \#({\cal C})^4 \label{def_s4} \end{eqnarray} where the sum is over all the clusters ${\cal C}$ of activated bonds and $\#({\cal C})$ is the number of sites in the cluster ${\cal C}$. \end{itemize} \noindent {}From these observables we compute the following expectation values: \begin{itemize} \item the energy density $E$ per spin \begin{equation} E \;=\; {1 \over V} \< {\cal E} \> \label{def_energy_density} \end{equation} \item the specific heat \begin{equation} C_H \;=\; {1 \over V} \hbox{var} ({\cal E}) \equiv {1\over V} \left[ \< {\cal E}^2 \> - \< {\cal E} \>^2 \right] \label{def_cv} \end{equation} \item the magnetic susceptibility \begin{equation} \chi \;=\; {1 \over V} \< {\cal M}^2 \> \label{def_susceptibility} \end{equation} \item the higher magnetization cumulants \begin{equation} \bar{u}_{2n} \;=\; {1 \over V} \< {\cal M}^{2n} \>_{\rm conn} \label{def_ubar_2n} \end{equation} \item the magnetization moment ratios \begin{equation} V_{2n} \; = \; {\< {\cal M}^{2n} \> \over \< {\cal M}^2 \>^n} \end{equation} \item the correlation function at momentum $(2\pi/L,0)$ \begin{equation} F \;=\; {1 \over V} \< {\cal F} \> \label{def_f_density} \end{equation} \item the second-moment correlation length \begin{equation} \xi \;=\; {1 \over 2 \sin(\pi/L)} \left( {\chi \over F} - 1 \right)^{1/2} \label{def_xi_sec3} \end{equation} \item the variant second-moment correlation length \begin{equation} \xi' \;=\; {L \over 2\pi} \left( {\chi \over F} - 1 \right)^{1/2} \;, \label{def_xi'_sec3} \end{equation} which differs from $\xi$ only by correction-to-scaling terms of order $L^{-2}$ \end{itemize} \bigskip \noindent {\bf Remarks.} 1. Using the Fortuin--Kasteleyn identities \cite{Kasteleyn_69,Fortuin_Kasteleyn_72,Fortuin_72,Sokal_Cargese}, it is not difficult to show that \begin{eqnarray} \< {\cal N} \> & = & p (B- \< {\cal E} \>) \label{check_NE} \\[1mm] \< {\cal E} \> & = & {q \over q-1} (B- \< {\cal E}' \>) \label{check_EE'} \\[1mm] \< {\cal M}^2 \> & = & \< {\cal S}_2 \> \label{check_stwo} \\[1mm] \< {\cal M}^4 \> & = & {q+1\over q-1} \< {\cal S}_2^2 \> - {2 \over q-1} \< {\cal S}_4 \> \label{check_sfour} \end{eqnarray} where $p = 1 - e^{-\beta}$ and $B=2V$ is the number of bonds in the lattice. As a check on the correctness of our simulations, we have tested these identities to high precision, in the following way: Instead of comparing directly the left and right sides of each equation, which are strongly positively correlated in the Monte Carlo simulation, a more sensitive test is to define new observables corresponding to the differences (i.e., ${\cal N} - p(B - {\cal E})$ and so forth). Each such observable should have mean zero, and the error bars on the sample mean can be estimated using the standard error analysis outlined below. The comparison to zero yields the following $\chi^2$ values: \begin{eqnarray} \hbox{For \reff{check_NE}:} & \; & \chi^2 = 10.23 \hbox{ (14 DF, level = 75\%)} \\ \hbox{For \reff{check_EE'}:} & \; & \chi^2 = 17.00 \hbox{ (14 DF, level = 26\%)} \\ \hbox{For \reff{check_stwo}:} & \; & \chi^2 = \phantom{1}9.93 \hbox{ (14 DF, level = 77\%)} \\ \hbox{For \reff{check_sfour}:} & \; & \chi^2 = \phantom{1}9.05 \hbox{ (14 DF, level = 83\%)} \end{eqnarray} Here DF means the number of degrees of freedom, and ``level'' means the confidence level of the fit (defined at the beginning of Section \ref{sec_res_static} below). The agreement is excellent. 2. We also compared our data for $V_4$ for $L=4,6,8,12,16$ with the exact values computed by Kamieniarz and Bl\"ote \cite{Kamieniarz_93} using transfer-matrix methods. We get $\chi^2 = 5.14$ (5 DF, level = 40\%), indicating good agreement. \bigskip \medskip For each observable ${\cal O}$ discussed above we have measured its autocorrelation functions in the Swendsen-Wang dynamics, \begin{eqnarray} C_{\cal OO}(t) &=& \<{\cal O}_s {\cal O}_{s+t}\> - \<{\cal O}\>^2 \\[2mm] \rho_{\cal OO}(t) &=& {C_{\cal OO}(t) \over C_{\cal OO}(0) } \end{eqnarray} where the expectations are taken in equilibrium. {}From these functions we have estimated the corresponding integrated autocorrelation time \begin{subeqnarray} \tau_{{\rm int},{\cal O}} & = & {1 \over 2} \sum\limits_{t=-\infty}^\infty \rho_{\cal OO}(t) \\[2mm] & = & {1 \over 2} \,+\, \sum\limits_{t=1}^\infty \rho_{\cal OO}(t) \end{subeqnarray} by the methods of Ref.~\cite[Appendix C]{Madras_Sokal}, using a self-consistent truncation window of width $6 \tau_{{\rm int},{\cal O}}$. This autocorrelation time is needed to compute the correct error bar on the sample mean $\overline{\cal O}$. \bigskip \noindent {\bf Remarks.} 1. The error bar of the second-moment correlation length is computed by considering the random variable \begin{equation} {\cal O}' \;=\; { {\cal M}^2 \over \mu_{{\cal M}^2} } - { {\cal F} \over \mu_{{\cal F}} } \;, \label{def_O'} \end{equation} which automatically has zero mean. Then, \begin{equation} \hbox{var} (\widehat{\xi})^{1/2} \;=\; {1 \over 4\sin(\pi/L)} {\chi \over F} \left({\chi\over F} - 1\right)^{-1/2} \hbox{var} ({\cal O}')^{1/2} \;, \end{equation} where $\widehat{\xi}$ denotes our Monte Carlo estimate of $\xi$. In practice, the values of $\mu_{{\cal M}^2}$ and $\mu_{{\cal F}}$ are replaced by their corresponding sample means (which should be computed first). 2. The error bar on the ratio $V_{2n}$ is computed in a similar fashion: \begin{equation} \hbox{var} (\widehat{V}_{2n})^{1/2} \;=\; { \< {\cal M}^{2n} \> \over \< {\cal M}^2 \>^n } \hbox{var} ({\cal O}''_{2n})^{1/2} \;, \end{equation} where $\widehat{V}_{2n}$ denotes our Monte Carlo estimate of $V_{2n}$, and the observable ${\cal O}''_{2n}$ is defined as \begin{equation} {\cal O}''_{2n} \;=\; { {\cal M}^{2n} \over \mu_{{\cal M}^{2n}} } - n{ {\cal M}^2 \over \mu_{{\cal M}^{2}} } + n-1 \end{equation} and has mean zero. Again, the mean values $\mu_{{\cal M}^{2n}}$ and $\mu_{{\cal M}^2}$ are replaced in practice by their sample means. 3. As a further check on the correctness of our simulations, we have computed both sides of the identity \begin{equation} \rho_{\cal NN}(1) \;=\; 1 - { (1-p)(2 - E) \over p C_H + (1-p)(2 - E) } \end{equation} proven in \cite[equation 7]{Li_Sokal} (see also \cite{Salas_Sokal_Potts3}).\footnote{ Please note that \protect\cite{Li_Sokal} used a definition of energy that is slightly different from the one used here: $E({\rm Ref.~\protect\cite{Li_Sokal}}) = (1/V)\<\sum_{\<xy\>} \delta_{\sigma_x,\sigma_y}\> = 2 - E$. } This is a highly nontrivial test, as it relates static quantities (energy and specific heat) to a dynamic quantity (autocorrelation function of the bond occupation at time lag 1). We have also checked with great accuracy the identities \cite{Salas_Sokal_Potts3} \begin{eqnarray} C_{\cal EE}(t) &=& {1 \over p^2} C_{\cal NN}(t+1) \\ \rho_{\cal EE}(t) &=& {\rho_{\cal NN}(t+1) \over \rho_{\cal NN}(1)} \\ C_{\cal E'E'}(t) &=& \left({q \over q-1}\right)^2 C_{\cal EE}(t+1) \\ \rho_{\cal E'E'}(t) &=& {\rho_{\cal EE}(t+1) \over \rho_{\cal EE}(1)} \end{eqnarray} \subsection{Summary of the simulations} We have run our Monte Carlo program on lattices with $L$ ranging from 4 to 512 (see Table~\ref{table_statics}). In all cases the initial configuration was random, and for $L\leq 64$ (resp.\ $L\geq 96$) we discarded the first $5 \times 10^4$ (resp.\ $10^5$) iterations to allow the system to reach equilibrium; this discard interval is in all cases greater than $10^4 \, \tau_{{\rm int},{\cal E}}$.\footnote{ Such a discard interval might seem to be much larger than necessary: $10^2 \tau_{\rm int}$ would usually be more than enough. However, there is always the danger that the longest autocorrelation time in the system may be much larger than the longest autocorrelation time that one has {\em measured}\/, because one has failed to measure an observable having sufficiently strong overlap with the slowest mode. As an undoubtedly overly conservative precaution against the possible (but unlikely) existence of such a (vastly) slower mode, we decided to discard up to 2\% of the entire run. This discard amounts to reducing the accuracy on our final estimates by a mere 1\%. } The total number of iterations ranges from $2.15 \times 10^6$ ($L=4$) to $8.2 \times 10^6$ ($L=512$), and is selected to be approximately $10^6 \, \tau_{{\rm int},{\cal E}}$. These statistics allow us to obtain a high accuracy in our estimates of the static and dynamic quantities (error $\mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$}$ 0.17\% and $\mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$}$ 0.51\%, respectively). The static data are displayed in Table~\ref{table_statics} ($\chi,F,\xi$) and Table~\ref{table_ratios} (the ratios $V_{2n}$). The dynamic data will be reported elsewhere. The CPU time required by our program is approximately 6.3 $L^2$ $\mu$s per iteration on a Linux Pentium machine running at 166 MHz. The total CPU time used in the project was approximately 7.5 months on this machine. We have improved the precision of our analysis of the correlation length $\xi$ by supplementing our own Monte Carlo data with comparable data from Ballesteros {\em et al.}\/ \cite{complutense}. They performed single-cluster \cite{Wolff_89} simulations of the 2D site-diluted Ising model at various concentrations ${\sf p}$. Their data for ${\sf p}=1$ (i.e., the usual Ising model) correspond to anywhere from $4 \times 10^5$ to $7 \times 10^5$ statistically independent measurements at each lattice size from $L=12$ to $L=512$ (see Table~\ref{table_xi_JJ}). The statistical independence of two consecutive measurements was achieved by allowing 100 single-cluster moves between them. Their error bars are slightly larger than ours. As a matter of fact, their error bars $\sigma'(\xi)$ and our error bars $\sigma(\xi)$ satisfy approximately the relation \begin{equation} {\sigma(\xi) \over \sigma'(\xi)} \;=\; \sqrt{2 \tau_{{\rm int},{\cal O}'} N' \over N} \;\,, \end{equation} where $N$ (resp. $N'$) is the number of measurements of our (resp.\ their) work, and ${\cal O}'$ is the observable \reff{def_O'} we used to compute the correct correlation-length error bar. This supports the belief that their measurements are indeed essentially independent and that their error bars are correctly computed. Comparison of their raw data to ours at the eleven overlapping $L$ values yields $\chi^2 = 10.28$ (11 DF, level = 51\%). The two data sets are therefore compatible. The corresponding merged data are shown in Table~\ref{table_xi_merged}. \section{Data Analysis} \label{sec_res_static} For each quantity ${\cal O}$, we carry out a variety of fits using the standard weighted least-squares method. As a precaution against corrections to scaling, we impose a lower cutoff $L \ge L_{min}$ on the data points admitted in the fit, and we study systematically the effects of varying $L_{min}$ on the estimated parameters and on the $\chi^2$ value. In general, our preferred fit corresponds to the smallest $L_{min}$ for which the goodness of fit is reasonable (e.g., the confidence level\footnote{ ``Confidence level'' is the probability that $\chi^2$ would exceed the observed value, assuming that the underlying statistical model is correct. An unusually low confidence level (e.g., less than 5\%) thus suggests that the underlying statistical model is {\em incorrect}\/ --- the most likely cause of which would be corrections to scaling. } is $\mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$}$ 10--20\%) and for which subsequent increases in $L_{min}$ do not cause the $\chi^2$ to drop vastly more than one unit per degree of freedom (DF). \subsection{Corrections to scaling} \label{sec4.fss} In the data analysis we should take into account the effect of corrections to scaling in order to get reliable estimates of the physical quantities. In particular, the value at criticality of any observable $O(L)$ is typically given for large $L$ by \begin{equation} O(L) = A L^{p_O} \left( 1 + A' L^{-\Delta} + \cdots \right) \end{equation} where $p_{O}$ is the critical exponent associated to the observable $O$, $\Delta$ is the leading correction-to-scaling exponent, and the dots indicate higher-order corrections. In finite-size-scaling (FSS) theory \cite{Privman} for systems with periodic boundary conditions, three simplifying assumptions have frequently been made: \begin{itemize} \item[(a)] The regular part of the free energy, $f_{\rm reg}$, is independent of $L$ \cite{Privman} (except possibly for terms that are exponentially small in $L$). \item[(b)] The relations connecting the nonlinear scaling fields $g_t$ and $g_h$ to the conventional thermodynamic parameters $t \equiv \beta_c-\beta$ and $h$ are independent of $L$ \cite{Guo_87}. \item[(c)] The scaling field $g_L$ associated to the lattice size equals $L^{-1}$ exactly, with no corrections $L^{-2}$, $L^{-3}$, \ldots\ \cite{Privman}. \end{itemize} Moreover, in the nearest-neighbor spin-1/2 2D Ising model, it has further been assumed that \begin{itemize} \item[(d)] There are no irrelevant operators \cite{Aharony_83,Gartenhaus_88}. \end{itemize} This latter assumption has been confirmed numerically (in the infinite-volume theory) through order $t^3$, at least as regards the bulk behavior of the susceptibility \cite{Gartenhaus_88}. However, both numerical \cite{Nickel_99a,Nickel_99b} and theoretical \cite{Pelissetto_private} evidence has recently emerged suggesting that irrelevant operators {\em do}\/ contribute to the susceptibility at order $t^4$. The absence of irrelevant operators implies that the corrections to scaling in this model are due to the smooth but in general nonlinear connection between the conventional thermodynamic parameters $t$ and $h$ and the renormalization-group nonlinear scaling fields \cite{Wegner_72,Wegner_D+G_vol6,Aharony_83,Gartenhaus_88}. Starting from the FSS Ansatz for the Ising-model free energy and using the above assumptions, it is possible to obtain a FSS expression for the usual observables at criticality as functions of the lattice size $L$ \cite{Salas_Sokal_FSS_paper2}. In particular, the leading correction term in the expansion of the susceptibility is the $L$-independent term coming from the regular part of the free energy. This implies that for this observable \begin{equation} \Delta \;=\; {7\over 4} \;\,. \label{Delta_prediction} \end{equation} The same result is plausible for the observable $F$ defined in \reff{def_f_density}; thus, we expect $\Delta=7/4$ for the second-moment correlation lengths $\xi$ and $\xi'$ and the corresponding amplitude $x^\star$. The expansion for the magnetization cumulant $\bar{u}_{2n}$ gives an exponent $\Delta = 1 + \gamma/\nu = 11/4$ (perhaps with a multiplicative logarithmic correction). Thus, we expect that the ratios $V_{2n}$ also have a correction-to-scaling exponent given by \reff{Delta_prediction} [due to the power of the susceptibility appearing in its definition \reff{def_ratios}]. For a more detailed theoretical and numerical analysis of the corrections to scaling in this model, see \cite{Salas_Sokal_FSS_paper2}. \subsection{Second-moment correlation length} \label{sec4.xi} The second-moment correlation length $\xi$ and its variant $\xi'$ [cf.\ \reff{def_xi_sec3}/\reff{def_xi'_sec3}] are expected to behave as \begin{equation} \left\{ {\xi \atop \xi'} \right\} \;=\; L^p \left[ x^\star + A L^{-\Delta} + \cdots \right] \label{ansatz_xi} \end{equation} with $p=1$. We can estimate $p$ by ignoring correction-to-scaling terms and performing a simple power-law fit. We get \begin{eqnarray} \hbox{For $\xi$:} & & p = 0.99974 \pm 0.00036 \qquad \hbox{ ($L_{min} = 32$, $\chi^2=1.48$, 6 DF, level = 96\%)} \nonumber \\ \\ \hbox{For $\xi'$:} & & p = 1.00018 \pm 0.00036 \qquad \hbox{ ($L_{min} = 32$, $\chi^2=1.37$, 6 DF, level = 97\%)} \nonumber \\ \end{eqnarray} The agreement with the theoretical prediction is excellent. The value of the constant $x^\star$ can be estimated most simply by fitting the ratio $\xi/L$ or $\xi'/L$ to a constant, ignoring corrections to scaling. We get \begin{eqnarray} \hbox{For $\xi$:} & & x^\star = 0.90577 \pm 0.00028 \qquad \hbox{ ($L_{min} = 32$, $\chi^2=2.02$, 7 DF, level = 96\%)} \nonumber \\ \label{xstar_xi_simple} \\ \hbox{For $\xi'$:} & & x^\star = 0.90557 \pm 0.00026 \qquad \hbox{ ($L_{min} = 16$, $\chi^2=2.73$, 9 DF, level = 97\%)} \nonumber \\ \label{xstar_xi'_simple} \end{eqnarray} The estimate based on $\xi$ lies 2.6 standard deviations away from the value $x^\star \approx 0.90505$ predicted by CFT [cf.\ \reff{xstar_theor}]. The estimate based on $\xi'$ is slightly better: it lies two standard deviations away from the theoretical prediction, and works also for smaller $L_{min}$. Indeed, the corrections to scaling in $\xi'/L$ are negligible (compared to our statistical error) already for $L \ge 16$ (see the last column of Table~\ref{table_xi_merged}).\footnote{ Table~\ref{table_xi_merged} is based on merging our data with that of Ballesteros {\em et al.}\ \cite{complutense}; but virtually identical results are obtained using our data alone. } This fact makes it almost hopeless to study corrections to FSS in $\xi'$. If we fit $\xi$ to \reff{ansatz_xi}, keeping the first correction-to-scaling term and trying to estimate simultaneously the three parameters $x^\star$, $A$ and $\Delta$, a good fit is obtained for $L_{min} = 8$: \begin{subeqnarray} x^\star &=& 0.90546 \pm 0.00033 \slabel{fit_xi_deltavar_a} \\ A &=& 0.75 \pm 0.30 \\ \Delta &=& 1.76 \pm 0.18 \slabel{fit_xi_deltavar_Delta} \label{fit_xi_deltavar} \end{subeqnarray} with $\chi^2=2.97$ (9 DF, level = 97\%). The value of $x^\star$ is again two standard deviations away from the theoretical prediction \reff{xstar_theor}. The estimate of $\Delta$ is very close to $7/4$, and is only 1.4 standard deviations away from $2$; but perhaps this estimate ought not be taken too seriously, as the correction-to-scaling amplitude is only 2.5 standard deviations away from zero (a deviation that is, moreover, comparable to the discrepancy in $x^\star$). The analogous fit for $\xi'$ is even more hopeless (the amplitude $A$ is compatible with zero within 0.7 standard deviations), so we omit the details. This correction-to-scaling exponent $\Delta \approx 2$ can be understood as arising simply from the ratio $\xi/\xi' \equiv [(L/\pi) \sin (\pi/L)]^{-1} = 1 + (\pi^2/6) L^{-2} + \ldots\;$. Indeed, if we fit the data to the Ansatz $\xi/L = x^\star + A L^{-2}$ we get for $L_{min}=8$: \begin{subeqnarray} x^\star &=& 0.90569 \pm 0.00027 \\ A &=& 1.269 \pm 0.064 \label{xstar_xi_Lminus2} \end{subeqnarray} with $\chi^2=4.56$ (10 DF, level = 92\%). Then, $A/x^\star \approx 1.40$, which is not far from $\pi^2/6 \approx 1.64$. We can improve the precision of our numerical estimates by using the merged data of Table~\ref{table_xi_merged} (our data plus that of Ballesteros {\em et al.}\ \cite{complutense}). The simple power-law fit yields \begin{eqnarray} \hbox{For $\xi$:} & & p = 1.00047 \pm 0.00033 \qquad \hbox{ ($L_{min} = 48$, $\chi^2=0.86$, 5 DF, level = 97\%)} \nonumber \\ \\ \hbox{For $\xi'$:} & & p = 1.00074 \pm 0.00033 \qquad \hbox{ ($L_{min} = 48$, $\chi^2=0.82$, 5 DF, level = 98\%)} \nonumber \\ \end{eqnarray} The fits to a constant give \begin{eqnarray} \hbox{For $\xi$:} & & x^\star = 0.90565 \pm 0.00023 \qquad \hbox{ ($L_{min} = 48$, $\chi^2=2.92$, 6 DF, level = 82\%)} \nonumber \\ \\ \hbox{For $\xi'$:} & & x^\star = 0.90555 \pm 0.00020 \qquad \hbox{ ($L_{min} = 16$, $\chi^2=6.99$, 9 DF, level = 64\%)} \nonumber \\ \end{eqnarray} The three-parameter fit $\xi/L = x^\star + A L^{-\Delta}$ is good for $L_{min} = 8$: \begin{subeqnarray} x^\star &=& 0.90552 \pm 0.00026 \slabel{fit_xi_merged_deltavar_xstar} \\ A &=& 0.86 \pm 0.29 \\ \Delta &=& 1.82 \pm 0.15 \label{fit_xi_merged_deltavar} \end{subeqnarray} with $\chi^2 = 7.57$ (9 DF, level = 58\%). The analogous fit with $\xi'$ yields a correction-to-scaling amplitude compatible with zero within errors. In conclusion, one can extract accurate estimates of the critical exponent $p$ and the amplitude $x^\star$ using our Monte Carlo data; the results agree with the theoretical prediction \reff{xstar_theor} within two standard deviations. However, it is very difficult to estimate from our numerical data the correction-to-scaling exponent (or the corresponding amplitude). Indeed, for $\xi'$ the corrections to scaling are negligible (compared to our statistical error) for $L \ge 16$. \subsection{Magnetization moment ratios} \label{sec4.Vn} If we study the magnetization distribution $\rho({\cal M})$ as $L \to \infty$ at fixed $\beta$, we expect three distinct behaviors depending on the value of $\beta$: \begin{itemize} \item[(a)] At $\beta < \beta_c$, we are in the high-temperature regime, where correlations decay exponentially. A variant of the central limit theorem \cite{Newman_private} guarantees that the finite-$L$ distributions will converge, after rescaling by the factor $\sqrt{V\chi}$, to a Gaussian distribution of mean zero and unit variance.\footnote{ Previously this had been proven for finite subvolumes of an infinite system \cite{Newman_80,Newman_83}, by a technique using the FKG inequalities. It has recently been proven by Newman \cite{Newman_private} also for finite systems with periodic or free boundary conditions, by a different (but simple and elegant) method using the GKS, GHS and Simon-Lieb inequalities. We thank Professor Newman for communicating to us this unpublished result, which we hope he will someday publish. } \item[(b)] At $\beta>\beta_c$ we are in the low-temperature regime, and the finite-$L$ distributions should converge, after rescaling by the factor $V M_0$ (where $M_0$ is the spontaneous magnetization), to the sum of two delta functions. There are Gaussian fluctuations around these two delta functions, but their width is much smaller, namely $\sqrt{V \chi_0}$, where $\chi_0$ is the susceptibility in a pure phase. \item[(c)] At $\beta=\beta_c$ [or more generally, at fixed value of the FSS variable $L^{1/\nu}(\beta-\beta_c)$], the finite-$L$ distributions will converge, after rescaling by the factor $\sqrt{V \chi}$, to some non-Gaussian distribution characteristic of the critical Ising model in a finite box. This distribution is not, to our knowledge, known exactly. \end{itemize} We have computed the magnetization histograms at $\beta=\beta_c$ for $L=4,\ldots,512$. The sequence of histograms is expected to converge to a limiting distribution when we normalize the magnetization by $\sqrt{V \chi}$ and normalize the height of the bins so that the area enclosed by the histogram is 1. For $L\gtapprox64$ the histograms converge well to a limiting histogram (Figure~\ref{figure_histo_mag_256}). For $L\ltapprox48$ small corrections to scaling are observed: the peaks of the histogram are slightly taller than in the limiting histogram The limiting distribution is symmetric and very strongly two-peaked (with maxima at ${\cal M}/\sqrt{V \chi} \approx \pm 1.11$); clearly the 2D Ising model at criticality in a finite symmetric torus is very far from Gaussian (e.g., we will find $V_4$ much closer to 1 than to 3). In order to characterize quantitatively this limiting distribution, we have measured its moments $\<{\cal M}^{2n}\>$ for $n=1,\ldots,10$ and have computed the corresponding ratios $V_{2n} \equiv \<{\cal M}^{2n}\> / \<{\cal M}^2\>^n$. We expect a behavior \begin{equation} V_{2n} \;=\; V_{2n}^\infty + B_{2n} L^{-\Delta} + \cdots \;. \label{ansatz_V2n} \end{equation} For each $n$, we have fitted our numerical data (Table~\ref{table_ratios}) in two ways: a one-parameter fit to a constant $V_{2n} = V_{2n}^\infty$ (fits marked with a C on the second column of Table~\ref{table_fit_ratios}) and a three-parameter fit to $V_{2n} = V_{2n}^\infty + B_{2n} L^{-\Delta}$ (fits marked P in Table~\ref{table_fit_ratios}). As expected, the estimates of $V_{2n}^\infty$ lie in-between the values associated to a Gaussian distribution \reff{ratios_Gaussian} and those associated to a two-delta-function distribution \reff{ratios_2deltas}. However, they are much closer to the latter values, reflecting the strongly two-peaked shape of the magnetization distribution. The fits to a constant are excellent for $L_{min} \mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$} 32$--64; for $2n = 4,6,8,10$ the estimates of $V_{2n}^\infty$ agree with the theoretical predictions within about 2.5 standard deviations. The three-parameter fits are excellent already for $L_{min}=8$: the correction-to-scaling amplitude $B_{2n}$ grows in magnitude with $n$, while the values of the correction-to-scaling exponent $\Delta$ are quite stable and are consistent with the theoretical prediction $\Delta = \gamma/\nu = 7/4$ [cf.\ \reff{Delta_prediction}] within two standard deviations. Let us now look more closely at $V_4$. With the three-parameter fit, we obtain for $L_{min}=8$: \begin{subeqnarray} V_4^\infty &=& 1.16777 \pm 0.00013 \slabel{fit_v4_deltavar_v4} \\ \Delta &=& 2.01 \pm 0.22 \\ B_4 &=& -0.48 \pm 0.23 \label{fit_v4_deltavar} \end{subeqnarray} with $\chi^2=1.53$ (9 DF, level = 100\%). The estimate of $V_4^\infty$ is about one standard deviation away from the theoretical prediction $V_4^\infty \approx 1.16792$ [cf.\ \reff{V4_exact}]. The estimate of $\Delta$ is reasonably close to $\Delta=7/4$; but since the estimated correction amplitude $B_4$ is only 2 standard deviations away from zero, this estimate of $\Delta$ should perhaps not be taken too seriously. Similarly, the estimates \begin{eqnarray} V_6^\infty & = & 1.45517 \pm 0.00037 \\ V_8^\infty & = & 1.89163 \pm 0.00079 \\ V_{10}^\infty & = & 2.5377 \pm 0.0015 \end{eqnarray} from the three-parameter fit are compatible with the predicted exact values \reff{V6_exact}/\newline\reff{V8_exact}/\reff{V10_exact} within 1.5, 1.3 and 1.2 standard deviations, respectively. The estimates of the exponent $\Delta$ ($1.90 \pm 0.16$, $1.83 \pm 0.13$ and $1.78 \pm 0.11$, respectively) are compatible with 7/4. The values of the first nine dimensionless renormalized $2n$-point coupling constants at criticality on a symmetric torus can be obtained from the results contained in Table~\ref{table_fit_ratios}: \begin{subeqnarray} G_{4\phantom{0}}^* &=& 2.23685 \pm 0.00016 \\[1mm] G_{6\phantom{0}}^* &=& 29.2602 \pm 0.0047 \\[1mm] G_{8\phantom{0}}^* &=& 942.91 \pm 0.25 \\[1mm] G_{10}^* &=& (5.6135 \pm 0.0021)\times 10^4 \\[1mm] G_{12}^* &=& (5.3281 \pm 0.0026)\times 10^6 \\[1mm] G_{14}^* &=& (7.3681 \pm 0.0046)\times 10^8 \\[1mm] G_{16}^* &=& (1.3969 \pm 0.0010)\times 10^{11} \\[1mm] G_{18}^* &=& (3.4746 \pm 0.0031)\times 10^{13} \\[1mm] G_{20}^* &=& (1.0969 \pm 0.0011)\times 10^{16} \label{G_2n_MC_results} \end{subeqnarray} The central values in \reff{G_2n_MC_results}, which are intended as our ``best estimates'', are computed using the {\em theoretical}\/ value \reff{xstar_theor} for $x^\star$.\footnote{ A ``pure'' Monte Carlo value, using the estimate \reff{fit_xi_merged_deltavar_xstar} for $x^\star$ in place of \reff{xstar_theor}, can be obtained by dividing the central value in \reff{G_2n_MC_results} by $1.005206^{2n-2}$ and adding $(2n-2) \times 0.000287$ to the fractional error bar. Note that the {\em dominant}\/ contribution to the error bar on $G_{2n}^*$ would then come from the uncertainty on $x^\star$: for example, we would have $G_4^* = 2.2345 \pm 0.0014$, $G_6^* = 29.199 \pm 0.038$, $G_8^* = 939.97 \pm 1.87$, etc.\ in place of \reff{G_2n_MC_results}. } The error bars quoted in \reff{G_2n_MC_results} are upper bounds computed using the triangle inequality, since we did not bother to compute the covariances among our Monte Carlo estimates of $V_{2n}^\infty$. Of course, for $G_4^*, G_6^*, G_8^*, G_{10}^*$ the estimates (\ref{G_2n_MC_results}a--d) are supplanted by the much more precise theoretical values (\ref{G_2n_exact}a--d). In summary, we have been able to estimate the limiting values $V_{2n}^\infty$ with great accuracy; and in the cases where the exact values are known, our numerical estimates agree with the theoretical predictions within less than two standard deviations. Our numerical estimates for $\Delta$ are compatible (within less than two standard deviations) with $\Delta=7/4$. \section{Conclusions} \label{sec_conclusions} We have computed, using results from conformal field theory (CFT), the exact (except for numerical integration) values of five universal amplitude ratios characterizing the 2D Ising model at criticality on a symmetric torus: the correlation-length ratio $x^\star$ and the magnetization moment ratios $V_4$, $V_6$, $V_8$ and $V_{10}$. All except for $V_4$ are new, and we have improved previous CFT determinations of $V_4$ by three orders of magnitude (reaching precision similar to that obtained by transfer-matrix approaches). As a corollary, we have computed the exact values $G_4^*$, $G_6^*$, $G_8^*$ and $G_{10}^*$ of the first four dimensionless renormalized $2n$-point coupling constants at criticality on a symmetric torus. We have checked all these theoretical predictions by means of a high-precision Monte Carlo simulation. Using finite-size-scaling (FSS) techniques, we have tried to determine the leading term as well as the correction-to-scaling terms. We confirm to high precision the theoretically predicted universal amplitude ratios $x^\star$, $V_4$, $V_6$, $V_8$ and $V_{10}$ (error bars $\mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$} 0.06\%$). The determination of the leading correction-to-scaling exponent $\Delta$ has proved to be difficult. For the modified correlation length $\xi'$, the corrections to FSS are so weak that they are essentially invisible for $L \ge 16$; and no reliable conclusions can be obtained from our data for $L=4,6,8,12$. For the standard correlation length $\xi$, the leading correction to scaling might be $\Delta=7/4$, or it might be $\Delta=2$ arising from $\xi/\xi' \equiv [(L/\pi) \sin (\pi/L)]^{-1} = 1 + (\pi^2/6) L^{-2} + \ldots\;$. For the magnetization moment ratios $V_{2n}$ we obtain stable results compatible with $\Delta=7/4$ within two standard deviations, in agreement with the theoretical prediction \reff{Delta_prediction}. It would be interesting to extend the analytic computation of $x^\star$ to other two-dimensional models, in particular those that can be mapped onto Gaussian models via height representations (see e.g.\ \cite{Nijs_82,Burton_Henley_97,Salas_Sokal_swaf3}). This work is currently in progress \cite{CPSS_in_prog}.
\section{Introduction} One modern approach to therapeutics is to identify and to block the molecular interactions responsible for disease. Within this context, combinatorial peptide screening experiments play an important role in the discovery of inhibitors. Libraries of peptides are used to discover the crucial molecular structure, or pharmacophore, that can block effectively the aberrant binding event. Cyclic peptides are preferred for this purpose since they display higher binding affinities due to their reduced conformational entropy in solution.\cite{Jackson_94,Bisang_98} The cyclic scaffolds and templates derived from combinatorial cyclic peptide studies have been used to assemble various spatially-defined functional groups, and highly active analogs have been synthesized in this way.\cite{Bisang_98,Haubner_96,Huang_93} A classic example is the blocking of the molecular event responsible for blood platelet aggregation by peptides of the form CRGDxxxC\put(-5,15){\line(-1,0){53}} \put(-5,10){\line(0,1){5}} \put(-58,10){\line(0,1){5}}, CxxxRGDC\put(-5,15){\line(-1,0){53}} \put(-5,10){\line(0,1){5}} \put(-58,10){\line(0,1){5}}, and CxxxKGDC\put(-5,15){\line(-1,0){53}} \put(-5,10){\line(0,1){5}} \put(-58,10){\line(0,1){5}}.\cite{II_ONeil} Organic analogs of these RGD peptides have been synthesized, and several companies are pursuing therapeutic applications in clinical trials. The ability of a peptide molecule to bind selectively to a receptor depends on its structural and conformational properties, which are in turn determined by its constituent amino acids. As the only natural secondary amino acid, proline plays a particular role in peptide and protein structural biology as a turn-promoting unit.\cite{Stryer_88} Although most amide bonds in native peptides and proteins are in the {\em trans} state, a significant fraction (6\%) of the X-Proline amide bonds are in the {\em cis} state.\cite{Stryer_88} In some proteins~\cite{Higgins_88} and cyclic peptides,\cite{Weisshoff_96} prolyl {\em cis/trans} isomerization has been detected in solution. This isomerization event is frequently a rate-limiting step in protein folding. The free energy barrier for the {\em cis/trans} isomerization of typical X-Pro amide bonds is about 19 Kcal/mol.\cite{Stein_93} While isomerization of non-prolyl amide bonds is rare, the energy barrier to isomerization is similar.\cite{Scherer_98} Conventional molecular simulations cannot follow the natural {\em cis/trans} isomerization dynamics, since the typical equilibration time is in the range of 10 to 100 s.\cite{Schmid_93} Monte Carlo simulations have been successful only in the case where isomerization is experimentally known to occur. In such cases, an isomerization reaction coordinate can be defined, and special techniques such as umbrella sampling can be applied. Conventional molecular dynamics for cyclic peptides, even in the absence of {\em cis/trans} isomerization, turns out to be non-ergodic as well, failing to sample the multiple solution conformations during the accessible simulation time. The difficulties arise from the cyclic constraints that, along with the intrinsic high energy barriers, isolate the accessible conformations to several separated regions in phase space. Monte Carlo methods, however, do not require the system to follow the natural trajectory. Therefore, larger and unphysical moves can be performed to overcome these energy barriers. A numerical peptide rebridging method, inspired by the alkane rebridging method~\cite{Dodd_93} and the configurational bias Monte Carlo (CBMC) method,\cite{Frenkel_92,dePablo_92} has been successfully applied to the simulation of non-proline-containing, cyclic peptides.\cite{Wu_99} Parallel tempering was shown to be a key factor in the efficiency of equilibration. Proline poses some geometrical complexity in the rebridging approach, however, and has not yet been treated. Here we present a new Monte Carlo method, the analytical rebridging scheme, that is suitable for equilibration of proline-containing, cyclic peptides. Our analytical method was inspired by the solution for a related inverse kinematics problem in robotic control.\cite{Manocha_94} The method can accommodate any rigid unit geometry. The rebridging method is not restricted to peptides and can be readily applied to other molecules. The {\em cis/trans} isomerization is naturally incorporated in the method by allowing for two states of the amide bond. As an added benefit, we find the analytical approach to be at least ten times more computationally efficient than the previous numerical method,\cite{Wu_99} even in the simplest, non-proline-containing cases. The analytical rebridging method and other components of our simulation methodology are described in Section~\ref{sec:methods}. We apply our method to five cyclic peptides that have previously been characterized by NMR methods in Section~\ref{sec:results}. We show that our method can effectively equilibrate these molecules, yielding conformations consistent with the NMR analyses. We discuss the results in Section~\ref{sec:discussion} and conclude in Section~\ref{sec:conclusion}. \section{Simulation Methods} \label{sec:methods} \subsection{Analytical Peptide Rebridging} \label{sec:methods-apr} Our equilibration scheme involves three types of moves: rebridging moves, semi-look-ahead (SLA) moves for side chains, and swapping moves. The rebridging move is described below and in Sec.~\ref{sec:methods-cti}. We describe the SLA move in Sec.~\ref{sec:methods-sg} and the swapping move in Sec.~\ref{sec:methods-pt}. In our system, bond lengths and bond angles are kept at their equilibrium value. We focus on sampling the biologically-relevant, torsional degrees of freedom. With this simplification, a molecule is comprised of a set of rigid units.\cite{I_Deem1} The rebridging scheme can easily be generalized to accommodate flexible bond angles and bond lengths. A peptide rebridging move causes a local conformational change within the molecule, leaving the rest of the molecule fixed. Consider the segment of a peptide backbone shown in Fig.~\ref{fig:reb_demo}. The angles $\phi_{0}$ and $\phi_{7}$ are rotated, causing the rigid units between 0 and 6 to change. The range of rotation is within $\pm\Delta\phi_{\mathrm{max}}$. These two rotations break the connectivity of the molecule. We then find all the solutions that re-insert the backbone units in a valid way between rigid units 1 to 6. The rebridging move is based upon enforcing six geometrical constraints, and this is why we choose to modify six rigid units. Modification of more than six units in a single move is possible by the rotation of more than two angles. Such a move is likely, however, to lead to an infeasible geometry most of the time and so to result in ineffective equilibration. Therefore, we choose to rotate $\phi_{0}$ and $\phi_{7}$ only. Our peptide rebridging scheme features an analytical solution of the geometrical problem arising from the reconnection of the backbone units, a problem previously solved in a numerical way.\cite{Wu_99,I_Deem1} The side chains are rigidly rotated for each of the solutions. For rebridging moves, the solutions for both the new and the old configurations are needed so as to satisfy detailed balance. The analytical solution involves the reduction of twenty linear equations to an 8$\times$8 determinant equation of one torsional angle. The determinant equation is equivalent to a polynomial of degree sixteen. Therefore, the maximum number of new geometrical solutions is strictly limited to 16, a bound that is obeyed in previous simulations using the numerical rebridging method.~\cite{crankshaft} The determinant equation is reformulated as an eigenvalue problem~\cite{Manocha_94} and solved using the QR algorithm.\cite{C_recipes} The details of the analytical rebridging method are described in Appendix~\ref{AppA}. Following the ``with Jacobian'' (WJ) biasing,\cite{Wu_99} one of the solutions is picked with a probability proportional to ${\mathrm J^{(n)}}\exp (-{\mathrm \beta U^{(n)}})$, the product of the Jacobian and the Boltzmann factor of the solution. The Jacobian $\mathrm{J}$ accounts for the correction to the non-uniform distribution generated by rebridging moves, and can be expressed in several ways. Here we present one form, involving the determinant of a $5\times 5$ matrix ${\mathbf B}$: \begin{eqnarray} \label{eqn:jac_RB} {\mathrm J} & = & \frac{ {\hat{\mathbf u}}_{6}\cdot{\hat{\mathbf e}}_3 } { \det \vert{\mathrm B} \vert }\nonumber \\ {\mathbf B}_{ij} & = & [{\hat{\mathbf u}}_{j}\times({\mathbf r}_{6}-{\mathbf r}_{\mathit{j}})]_i \mbox{, if $i\leq3$}, \nonumber \\ & & [{\hat{\mathbf u}}_{j}\times{\hat{\mathbf u}}_{6}]_{i-3} \mbox{, if $i=4,\ 5$} . \end{eqnarray} Here ${\hat{\mathbf u}}_{i}$ is the unit vector about which the torsional angle $\phi_{i}$ is measured, ${\mathbf r}_{i}$ is the position of the atom in unit $i-1$ that is bonded by a sigma bond to unit $i$, and ${\hat{\mathbf e}}_3$ is a unit vector along the laboratory $z$-axis. The subscript outside the brackets refers to the component of the bracketed vector in the laboratory frame. The Jacobian can also be written in terms of a determinant of a $4\times 4$ matrix, although the definition of the components of the $4\times 4$ matrix is more involved.~\cite{Wu_99} The attempted move is accepted with the probability \begin{equation} \label{eqn:acc-WJ} \mathrm{acc}(\mathrm{o\rightarrow n}) = \min \left(1, \frac{W^{\mathrm{(n)}}}{W^{\mathrm{(o)}}}\right)\ , \end{equation} where ${\mathrm W^{(n)}}$ and ${\mathrm W^{(o)}}$ are the normalization (Rosenbluth) factors for the new and old solutions, respectively.\cite{Wu_99,Frenkel_96} \subsection{The {\em cis/trans} Isomerization} \label{sec:methods-cti} For each amide bond that we wish to allow to isomerize, we assign two discrete states to the corresponding rigid unit. As shown in Fig.~\ref{fig:cistrans}, the amide unit takes the torsional values of $\omega=0^\circ$ or $\omega=180^\circ$ in the {\em cis} and {\em trans} conformations, respectively. The partition function includes a summation over both states. Because of the sum over states, solutions corresponding to all the possible {\em cis/trans} states between unit 0 and unit 6 are included in the calculation of the Rosenbluth factor. The same approach can be applied to both non-prolyl and prolyl amide bonds. \subsection{Side Groups} \label{sec:methods-sg} The chemical functionality of peptides lies mostly in the side chains. Efficient equilibration for the side chains, therefore, is very important. The semi-look-ahead (SLA) move, based on CBMC methods, has been shown to equilibrate effectively long and bulky chains.\cite{Wu_99} We use the SLA method to equilibrate side chains and end groups in this work. A SLA move proceeds by regrowing a randomly selected side chain, unit by unit, beginning from the bond that connects the backbone to the side chain. The reverse move is performed so as to satisfy detailed balance. The Jacobian for each solution is unity. The attempted move is accepted with the probability \begin{equation} \label{eqn:acc-SLA} \mathrm{acc}(\mathrm{o\rightarrow n}) = \min \left(1, \frac{W^{\mathrm{(n)}}}{W^{\mathrm{(o)}}}\right)\ , \end{equation} where ${\mathrm W^{(n)}}$ and ${\mathrm W^{(o)}}$ are the normalization (Rosenbluth) factors for the new and old geometries, respectively.\cite{Frenkel_96} \subsection{Parallel Tempering} \label{sec:methods-pt} Parallel tempering was first proposed for the study of glassy systems with large free energy barriers.\cite{geyer_91} It has since been successfully applied to a variety of systems.\cite{hukushima_96,marinari_98,tesi_96,Boyd_98,Falcioni_99} This method achieves rigorously correct canonical sampling, and it significantly reduces the equilibration time in a simulation. Instead of a single system, we consider in parallel tempering a larger ensemble with $n$ systems, each equilibrated at a distinct temperature $T_i$, $i=1,\ \ldots,\ n$. The system with the lowest temperature is the one of our interest; the higher temperature systems are added to aid in the equilibration of the system of interest. In addition to the normal Monte Carlo moves performed in each system, swapping moves are proposed that exchange the configurations between two systems $i$ and $j=i+1$, $1\leq i< n$. A swapping move is accepted with the probability \begin{eqnarray} \label{eqn:acc-pt} {\mathrm acc}[(i,\ j)\rightarrow (j,\ i)] & = & \min[1,\exp(-\Delta\beta\Delta\mathrm{U})]\ , \end{eqnarray} where $\Delta\beta$ and $\Delta\mathrm{U}$ are the difference of the reciprocal temperatures and energies, respectively. The higher temperature systems are included solely to help the lowest temperature system to escape from local energy minima via the swapping moves. To achieve efficient sampling, the highest temperature should be such that no significant free energy barriers are observed. So that the swapping moves are accepted with a reasonable probability, the energy histograms of systems adjacent in the temperature ladder should overlap. \section{Comparison with Experimental Results} \label{sec:results} We perform simulations on five distinct cyclic peptides that have been previously characterized by NMR methods. We focus on the backbone structure of proline-containing, cyclic peptides that were observed experimentally to undergo {\em cis/trans} isomerization. Molecular interactions are described by the AMBER force field with explicit atoms.\cite{Weiner_86} Aqueous solvent effects are estimated by simple dielectric theory.\cite{Smith_93} Five or six systems were used in the parallel tempering for each simulation, with the highest temperatures ranging from 10$^5$ K to 10$^7$ K. The lowest temperature system in each case is 298 K. For the first three peptides, {\em cis/trans} isomerization is allowed in prolyl amide bonds only. For the last two peptides, isomerization of all amide bonds is allowed. The simulations take 5-8 CPU hours for the first three peptides and 15-20 CPU hours for the last two peptides. All the simulations were performed on an Intel Pentium II 450 MHz Linux workstation. Rapid equilibration is achieved for all peptides. The results for each of the peptides are presented below: \begin{enumerate} \item c(Pro-Phe-D-Trp-Lys-Thr-Phe). This analog of somatostatin displays high activity in inhibiting the release of a growth hormone.\cite{Huang_93,Veber_81} Analysis of the NMR spectrum in D$_2$O solution indicated a unique backbone conformation.\cite{Veber_81} The prolyl amide bond at Phe-Pro adopted a {\em cis} conformation. In our simulation, we find an essentially unique conformation, possessing the same amide bond {\em cis/trans} sequence. A representative conformation for this peptide is shown in Fig.~\ref{fig:ex1}. \item c(Phe-Phe-Aib-Leu-Pro). This pentapeptide contains the Pro-Phe-Phe sequence that has been proposed to be responsible for the cytoprotective ability of antamanide and cyclolinopeptides.\cite{Kessler_86,Zanotti_98} NMR analysis indicated that the peptide is conformationally non-homogeneous at room temperature. Two predominant {\em cis/trans} isomers for the Leu-Pro amide bond were identified in acetonitrile at 240 K.\cite{Zanotti_98} Our simulation led to two inter-converting conformers in the simulation, as shown in Fig.~\ref{fig:ex2}. The {\em cis} and {\em trans} conformers occur with probability 58\% and 42\%, respectively. \item c(Gly-Glu(OBzl)-Pro-Phe-Leu-Pro). This cyclic hexapeptide was synthesized for use as a possible chiral site for enantiomeric separation.\cite{McEwen_93} NMR studies in dimethyl sulfoxide (DMSO) reported two isomers, one having two {\em cis} prolyl bonds, and the other having all-{\em trans} bonds. We find only the 2-{\em cis} conformer in simulation, as shown in Fig.~\ref{fig:ex4}. All the torsional angles fluctuate around mean values, except that the amide group between Pro and Gly flips between two opposite orientations. The all-{\em trans} conformer was found at higher temperatures. \item c(Pro-Ala-Pro-Ala-Ala). This cyclic peptide has been designed to serve as a rigid structural template. An unique conformation with two {\em cis} prolyl amide bonds in DMSO solution was found, according to NMR analysis.\cite{Mueller_93} The backbone consists of two intertwined type-VIb $\beta$ turns, centered about the two prolyl amide bonds respectively. We find the same unique conformation. The torsional angles are close to the values derived from restrained molecular dynamics.\cite{Mueller_93} Figure~\ref{fig:ex7} depicts the geometry of this peptide. \item Tentoxin, c(MeAla-Leu-MePhe[(Z)$\triangle$]-Gly). This tetrapeptide selectively induces chlorosis in the seedlings of plants. Although tentoxin lacks proline, its two methylated amide bonds were found to adopt the {\em cis} conformation in a nearly saturated aqueous solution.\cite{Pinet_95} The other two non-methylated amide bonds adopt the {\em trans} conformation. The observation of this {\em cis}-{\em trans}-{\em cis}-{\em trans} sequence of the backbone, along with other experimental data, led to a proposed boat-like conformation, with the two {\em cis} bonds located on the same side of the mean plane. In our simulation, we find the same amide bond sequence and boat-like conformation that was found in the experimental structure. The conformation is shown in Fig.~\ref{fig:ex8}. All the carbonyl groups lie in the same side of the mean plane, which implies that we have found the the third major (conformer C, 8\% abundance) of the four conformers found at 268 K in ref.~\onlinecite{Pinet_95}. The four conformers differ only in the orientation of the two non-methylated amide groups. The other three conformers were found at higher temperatures in the simulation. \end{enumerate} \section{Discussion} \label{sec:discussion} By analyzing the energy trajectories and conformational data, we find that all the peptides are effectively equilibrated with relatively few Monte Carlo steps. For example, we find that the {\em cis/trans} equilibrium for peptide 2 was attained within the first 10\% of the simulation time. Parallel tempering is crucial for this inter-conversion, since the peptide essentially does not undergo \emph{cis/trans} isomerization in a single, room-temperature, canonical simulation. For peptide 1, the NMR-based conformational study indicated a type-II'$\beta$ turn in the Phe-D-Trp-Lys-Thr region. This turn is characterized by the hydrogen bond between the C=O of Phe and the N-H of Thr. In our simulation, The C=O of Phe and the N-H of Thr are close to each other, but are not in precise alignment. Such disorder is expected at finite temperature. Bond angle inflexibility may influence the predicted equilibrium properties of these highly-strained molecules. We have investigated the dependence on bond angles by changing the angle between C$_\alpha$-C and N-C$_\alpha$ in the amide bond on peptide 2 from 6$^\circ$ to 0$^\circ$. The predicted {\em cis/trans} equilibrium shifts from 42\% {\em trans} to 14\% {\em trans}, which is a non-trivial, although energetically small, effect. We suspect, therefore, that the absence of the all-{\em trans} conformer of peptide 3 in our simulation may be due partly to inflexibility of the bond angles. We used a rigid proline ring, with a $\phi_{\mathrm{Pro}}=-75^\circ$. This constraint suppresses the small fluctuations that occur in the proline ring. In exceptionally-constrained, cyclic peptides, $\phi_{\mathrm{Pro}}$ can take on other values. For example, $\phi_{\mathrm{Pro}}\simeq -50^\circ$ has been observed for some unusual prolines in the {\em trans} state.\cite{Mueller_93} Fluctuations in the proline ring may, therefore, be important in some highly-strained systems. Although experimental conformational analysis of peptide 3 in water was not performed, NMR analysis in CHCl$_3$ indicated a minor conformer in addition to the two major ones found in DMSO solution.\cite{McEwen_93} It is clear, then, that solvent effects play a role in the equilibrium structure of peptide 3. Note that the dielectric constants are $\varepsilon_{\mathrm{CHCl_3}}=4.8$,\cite{CRC_87} $\varepsilon_{\mathrm{DMSO}}=45$,\cite{Merck_83} and $\varepsilon_{\mathrm{H_2O}}= 78$.\cite{CRC_87} The low-dielectric CHCl$_3$ solution favored the all-{\em trans} conformer. Water, with a high dielectric constant, may favor the 2-{\em cis} conformer, which is what we observe in simulation. Indeed, upon reducing the dielectric constant in our implicit solvent model, we find a small amount of the all-{\em trans} state. However, the detailed structure of the solvent molecules around the peptide is likely to be important, and a more accurate description of solvent is likely necessary to account fully for the solvent effects. Tentoxin (peptide 5) in water at 268 K was experimentally found to aggregate in a way that suggested micellar organization. The critical micelle concentration (CMC) was estimated to be roughly 35 $\mu$M.\cite{Pinet_95} Conformations observed at concentrations above the CMC may differ from that of the dilute-limit, monomeric form. In fact, of all four conformers found in ref.~\onlinecite{Pinet_95}, only conformer C yielded chemical shifts for the $\delta$ and $\gamma$ protons of Leucine close to that expected for a monomeric form. The other two major conformers, A and B, displayed strong shielding effects. The shielding constants for the minor conformer D were not reported. These shielding effects were explained by an aggregated structure of these two conformers. Since the concentration of conformer C (250$\mu$M) is still well above the critical micelle concentration, it is not entirely clear why the chemical shifts for conformer C were relatively unaffected by potential aggregation.\cite{Pinet_95} Nonetheless, the experimental data suggests that conformer C may be either in monomeric form or in an environment similar to aqueous solution. This gives one possible explanation for the absence of conformers A and B in our simulation, since these conformers are certainly not in monomeric form. \section{Conclusion} \label{sec:conclusion} We presented a new method, the analytical rebridging scheme, for the simulation of chemically-diverse, chain-like molecules. The method naturally accommodates {\em cis/trans} isomerizations. Our rebridging scheme, combined with parallel tempering and biased Monte Carlo, is very successful at equilibration of proline-containing, cyclic peptides. Our method is not limited to cyclic peptides and can simulate any chain-like molecule. We compared our simulations with experimental data on five cyclic peptides and found the predicted conformations to be reasonably accurate. We were able to sample multiple, relevant conformations separated by high energy barriers, a feat not possible with conventional molecular dynamics or Monte Carlo. The numerical quality of our predictions, while not limited by sampling issues, may be limited by our choice of a simple forcefield. Nonetheless, our method can be easily extended to accommodate flexible bond angles and bond lengths. In addition, solvent effects may be represented more accurately by better implicit solvent models.\cite{Ashbaugh_98,Paulaitis_97} The methods described here are powerful enough and general enough to influence the preferred approach to simulating biological systems. For example, our method should be a valuable tool for the fitting of new potential parameters for biological systems. New simulation methods for long alkanes have made possible the optimization of force fields that significantly reduce the discrepancies between simulation and experiment.\cite{Smit_95,Nath_98} The same approach should lead to improved forcefields for biological systems. In addition, we expect that our peptide rebridging scheme, combined with parallel tempering, should replace high temperature molecular dynamics and is readily suitable for use in NMR-based conformational analyses of biomolecules. \section*{Acknowledgments} This research was supported by the National Science Foundation through grant no. CHE-9705165. \newpage
\section{Introduction} \label{sec:intro} A working hypothesis in most approaches to cosmological structure formation is that the density fluctuation field $\delta({\bf x})$ is Gaussian at early times. This assumption follows naturally from the idea that the primordial fluctuations may be described as superpositions of statistically independent Fourier modes, \begin{equation} \delta({\bf x}) = \sum \delta_{\bf k} e^{i{\bf k}\cdot{\bf x}}\,. \label{eq:fourier} \end{equation} Such modes are generic predictions of inflation, in which quantum fluctuations are exponentially enlarged to become the classical density fluctuations. Whatever the form of the power spectrum $P(k) = |\delta_{\bf k}|^2,$ the independence of the Fourier modes guarantees, by the central limit theorem, that the real-space density fluctuations are Gaussian. Although the Gaussian hypothesis is well-motivated, several lines of argument suggest that it merits further scrutiny. On the theoretical side, Peebles (1999a,b) has presented a model in which nongaussianity emerges naturally in a modified inflationary scenario. The key ingredient is that the scalar field which drives inflation, $\psi,$ couples quadratically to a massive scalar field $\phi$ which, at the end of inflation, produces the cold dark matter (CDM). The CDM density field is then given by $\rho_{CDM}({\bf x}) =\mu^2 \phi^2({\bf x})/2,$ where $\phi({\bf x})$ is a Gaussian random field. The fluctuations in the CDM component, $\delta({\bf x}) \propto \rho_{CDM}({\bf x})-\vev{\rho_{CDM}},$ are thus distributed like a $\chi^2$ function shifted to have vanishing mean. Such a distribution is strongly nongaussian, with significant skew and excess kurtosis. (Other inflationary models which lead to nongaussian density fluctuations have recently been described by Salopek [1999] and by Martin, Riazuelo, \& Sakellariadou [1999].) Peebles' model has been extended by White (1998) and by Koyama, Soda, \& Taruya (1999), who have considered models in which the CDM is produced by a finite number, $m,$ of quadratically coupled fields. The resulting density fluctuations would then have a $\chi^2_m$ distribution (again, shifted to have mean zero), which approaches Gaussianity in the limit of large $m.$ In this way one can generate primordial fluctuations with any desired amount of nongaussianity. However, the required number of quadratically coupled fields must be quite large ($\simgt100$) if the wanted amount of nongaussianity is small. Such a large number of inflationary fields arguably makes this model contrived. On the observational side, hints of nongaussianity have emerged, first, from Cosmic Background Radiation (CBR) anisotropy maps. Ferreira {\sl et al.}\ (1998, 1999; see also Magueijo 1999) analyzed the full-sky COBE-DMR maps (Bennett {\sl et al.}\ 1996), finding evidence of nongaussianity on large ($\lower.5ex\hbox{\gtsima} 10\ifmmode^\circ\else$^\circ$\fi$) angular scales using a statistic known as the normalized bispectrum estimator. This finding was confirmed by two independent analyses using alternative statistical techniques (Novikov, Feldman, \& Shandarin, 1998; Pando, Valls-Gabaud, \& Fang, 1998). These studies appear to rule out Gaussianity of the CBR anisotropies at about the 95\% confidence level. A curious feature of these COBE-DMR results is that the nongaussian signal goes away if the north galactic cap is excluded from the analysis; this is a surprising property, one that argues for caution in ascribing reality to these findings (see Bromley \& Tegmark 1999 for a detailed treatment of this issue). In an unrelated study of CBR anisotropy detections on degree scales, Gazta\~naga, Fosalba, \& Elizalde (1998) found evidence for nongaussianity based on the large variance among the reported anisotropy amplitudes. This finding, like that of Ferriera {\sl et al.}, should be considered preliminary, as it is based on an intercomparison of very different data sets and is strongly dependent on the accuracy of the reported errors. Nonetheless, these results from CBR anisotropy data call into question the hypothesis of primordial gaussianity. The second line of observational evidence comes from the abundance and clustering of rich clusters of galaxies. As the most massive virialized systems in the universe, rich clusters are diagnostic of the background cosmological parameters in a number of ways (cf.\ Bahcall 1999 for a recent review). In particular, their number density and correlation length, as a function of mass and redshift, can be predicted from the Press-Schechter (1974, hereafter PS) formalism, summarized in \S~\ref{sec:PS} below. The original PS formalism assumes primordial Gaussianity, but this assumption is readily relaxed, leading to a generalized PS approach (\S~2.3). Robinson, Gawiser, \& Silk (1998) applied the generalized PS formalism to low-redshift ($z\lower.5ex\hbox{\ltsima} 0.1$) cluster abundance and correlation data. They found the data to be consistent with the Gaussian hypothesis in an $\Omega_{\rm M}=1$ universe.\footnote{Here and throughout this paper, the present value of the mass density parameter is denoted $\Omega_{\rm M},$ while the density parameter associated with a cosmological constant $\Lambda$ (or, equivalently, vacuum energy density) is denoted $\Omega_\Lambda.$ The Hubble constant is written as $H_0=100\,h\ \kms\ {{\rm Mpc}}^{-1}.$} For lower values of the density parameter, however, Robinson {\sl et al.}\ found that a substantial degree of primordial nongaussianity is required to reconcile the number density of massive clusters with their observed correlation length. Koyama, Soda, \& Taruya (1999) further narrowed these constraints by applying the generalized PS formalism to intermediate-redshift ($z\simeq 0.5$--0.6) cluster abundance data, as well as to the low redshift data sets studied by Robinson {\sl et al.}\ They found that $\Omega_{\rm M}>0.5$ was ruled out by the combined data sets regardless of the nature of the initial density fluctuations, and that Gaussian fluctuations were ruled out for $\Omega_{\rm M}\leq 0.5.$ We defer a fuller discussion of their results to the final section of this paper, after introducing the necessary terminology in \S~\ref{sec:nongauss}. The studies cited above provide evidence that the hypothesis of Gaussian density fluctuations is violated at some level. The purpose of this paper is to extend the analyses of Robinson {\sl et al.}\ (1998) and Koyama {\sl et al.}\ (1999) by applying the generalized PS approach to a single high-redshift cluster, MS1054--03, which at $z=0.83$ is one of the highest redshift clusters known. It is, moreover, the most extensively studied of the known high-redshift clusters; its mass has been accurately estimated by the techniques of weak gravitational lensing (Luppino \& Kaiser 1997), X-ray temperature analysis (Donahue {\sl et al.}\ 1998), and galaxy velocity dispersion (Tran {\sl et al.}\ 1999). These studies all found that MS1054 is an unusually massive cluster, with a virial mass $\lower.5ex\hbox{\gtsima} 10^{15}\,\h1M_\odot.$ Its high redshift, large and accurately determined mass, and its having been selected from the Einstein Medium Sensitivity Survey (Henry {\sl et al.}\ 1992; EMSS), whose selection criteria are very well understood, combine to make MS1054 especially well suited for the present study. A previous analysis of MS1054 (Bahcall \& Fan 1998) noted these features, and derived important cosmological constraints, but did so under the assumption of primordial Gaussianity. We will extend their results by including nongaussianity as an additional degree of freedom. This paper is also meant to serve two additional purposes. First, we will introduce (\S~\ref{sec:nongauss}) a new mathematical description of nongaussian density fluctuations, in the form of a probability distribution function (PDF) we refer to as the $\psi_\lambda$-distribution. We will discuss the relationship between the $\psi_\lambda$ and $\chi^2_m$ distributions, and argue that the former is perferable to the extent that a fully generic, model-independent parameterization of nongaussianity is desired. Second, we will present (\S~\ref{sec:clmass}) a detailed discussion of the proper determination of cluster virial masses from X-ray temperature, galaxy velocity dispersion, and weak lensing data sets, and apply the results to MS1054. This discussion will, it is hoped, be useful in clarifying the ways in which virial masses are dependent on cosmological parameters and, more subtly, on cluster mass models, and thus inform future studies based on larger data sets. The outline of this paper is as follows. In \S~\ref{sec:PS}, we describe the Gaussian and generalized PS formalisms. In \S~\ref{sec:nongauss}, we introduce the $\psi_\lambda$ distribution. In \S~\ref{sec:clmass} we discuss the determination of cluster virial masses. In \S~\ref{sec:ms1054} we use published data to estimate the virial mass of MS1054. In \S~\ref{sec:results} we apply the generalized PS formalism to MS1054 and derive constraints on nongaussianity as a function of $\Omega_{\rm M}.$ Finally, in \S~\ref{sec:disc}, we further discuss and summarize the main results of the paper. \section{The Press-Schechter Formalism and its Extension to the Nongaussian Case} \label{sec:PS} Under the assumption of primordial Gaussianity, the PS formula for the comoving number density of virialized objects of mass $M$ is \begin{equation} n(M,z) = \sqrt{\frac{2}{\pi}}\,\frac{\overline\rho}{M^2}\, \frac{\delta_c(z)}{\sigma_M}\left|\frac{d\ln \sigma_M}{d\ln M}\right| \,e^{-\delta_c(z)^2/2\sigma_M^2}\,, \label{eq:PS} \end{equation} where $\sigma_M$ is the rms mass fluctuation on a mass scale $M,$ $\overline\rho\equiv\Omega_{\rm M}\rho_{crit}$ is the comoving mean mass density, and $\delta_c(z)$ is the critical density for collapse at redshift $z.$ Because $\sigma_M$ is by definition the rms density fluctuation linearly extrapolated to the present, $\delta_c(z)$ is similarly normalized to the present, \begin{equation} \delta_c(z;\Omega_{\rm M},\Omega_\Lambda) = \delta_0(z) \frac{D(z=0;\Omega_{\rm M},\Omega_\Lambda)} {D(z;\Omega_{\rm M},\Omega_\Lambda)}\,. \label{eq:deltacz} \end{equation} The linear fluctuation growth factor is given by \begin{equation} D(z;\Omega_{\rm M},\Omega_\Lambda) = \frac{5}{2} \Omega_{\rm M} E(z) \int_z^{\infty} \frac{1+z'}{E(z')^3} dz'\,, \label{eq:dz} \end{equation} where $E(z)=\sqrt{\Omega_{\rm M}(1+z)^3+\Omega_R(1+z)^2+\Omega_\Lambda},$ $\Omega_{\rm M}+\Omega_R+\Omega_\Lambda=1,$ and we are following Peebles' (1993) notation. Throughout this paper we assume a flat universe, $\Omega_R=0.$ The quantity $\delta_0(z)$ in equation~\ref{eq:deltacz} is the famous factor $3(12\pi)^{2/3}/20=1.686$ for an Einstein de-Sitter universe. It has a weak dependence on cosmological parameters and redshift. In this paper, we use the forms of $\delta_0(z)$ derived by Kitayama \& Suto (1996). \subsection{Calculation of $\sigma_M$} The rms mass fluctuation $\sigma_M$ in equation~\ref{eq:PS} is computed from the linear power spectrum $P(k)$ as follows: \begin{equation} \sigma_M^2 = \int_0^\infty \frac{dk}{k} \Delta^2(k)\,W^2(kR)\,, \label{eq:sigmr} \end{equation} where $\Delta^2(k)\equiv k^3 P(k)/2\pi^2$ is the mass variance per logarithmic wavenumber interval, $W(kR)$ is the Fourier transform of the window function defining the mass scale $M,$ which we take to be the usual top-hat, and $R$ is the radius of a sphere which contains mass $M$ in an unperturbed universe. In this paper we assume a CDM-dominated universe, in which the power spectrum is well-approximated by \begin{equation} \Delta^2(k) = \delta_H^2(\Omega_{\rm M},\Omega_\Lambda) \left(\frac{ck}{H_0}\right)^{3+n}\,T^2(k/\Gamma) \label{eq:defdelta} \end{equation} (Bunn \& White 1997), where $\delta_H$ is the rms overdensity at horizon-crossing, $n$ is the primordial spectral index, and $T(q)$ is the CDM transfer function, for which we adopt the analytic approximation of Bardeen {\sl et al.}\ (1986). The parameter $\Gamma$ determines the position of the power-spectrum ``turnover'' in $k$-space. For CDM it is given by $\Gamma \approx \Omega_{\rm M} h.$ For the calculations of this paper, the precise value of $\Gamma$ is relatively unimportant, and we adopt the value $\Gamma = 0.20,$ consistent with observations of large-scale structure data (e.g., Liddle {\sl et al.}\ 1996) and with the CDM expectation for reasonable values of the cosmological parameters. We also assume $n=1,$ consistent with estimates derived from the large-scale CBR anisotropies observed by COBE (Gorski {\sl et al.}\ 1996). Changing our adopted values of $\Gamma$ and $n$ by $\lower.5ex\hbox{\ltsima} 20\%,$ roughly their allowed ranges, would have little effect on the main conclusions of this paper. \subsection{Conversion to useful units} To make practical calculations, we reexpress the PS formula in terms of suitably scaled variables. We first define a dimensionless mass $m$ by \begin{equation} M = \frac{4\pi}{3} r_8^3 \rho_{crit}\, m = 5.95\times 10^{14} m\, \ifmmode h^{-1}\else$h^{-1}$\fi\ M_\odot\,, \label{eq:defm} \end{equation} where $r_8\equiv 8\,\ifmmode h^{-1}\else$h^{-1}$\fi$ Mpc and $\rho_{crit}=3H_0^2/8\pi G$ is the critical density. For rich clusters, $m$ is of order unity. Let $n(m) dm$ be the comoving number density of clusters with dimensionless masses in the range $(m,m+dm).$ Then \[ n(m,z) = n(M,z) \frac{dM}{dm} \] \begin{equation} = \sqrt{\frac{2}{\pi}} \frac{3\,\Omega_{\rm M}} {4\pi\,r_8^3\,m^2} \frac{\delta_c(z)}{\sigma_M} \left|\frac{d\ln \sigma_M}{d\ln M}\right| e^{-\delta_c(z)^2/2\sigma_M^2}\,, \label{eq:ps1} \end{equation} where we have substituted the definition of $m$ into equation~1. Furthermore, as noted above, we compute $\sigma_M$ not as a function of mass but of a length scale $R_m$ defined by $M=\frac{4}{3}\pi R_m^3 \Omega_{\rm M} \rho_{crit},$ or equivalently, \begin{equation} R_m = r_8\left(\frac{m}{\Omega_{\rm M}}\right)^{1/3}\,. \label{eq:defrm} \end{equation} Thus, $d\ln\sigma_M/d\ln M=1/3\times d\ln\sigma_M/d\ln R.$ Substituting this into equation~\ref{eq:ps1} and evaluating numerical factors yields \begin{equation} n(m,z) = n_0\,\frac{\Omega_{\rm M}}{m^2} \left|\frac{d\ln\sigma_M} {d\ln R}\right|_{R_m}\!\nu_m \phi(\nu_m)\,, \label{eq:ps2} \end{equation} where $n_0=3.11\times 10^{-4}(\ifmmode h^{-1}\else$h^{-1}$\fi{\rm Mpc})^{-3},$ $\phi(x)$ is a Gaussian of zero mean and unit variance, and $\nu_m(z) \equiv \delta_c(z;\Omega_{\rm M})/\sigma_M(R_m).$ Equation~\ref{eq:ps2} gives the comoving number density per unit dimensionless mass; note that most of the $\Omega_{\rm M}$-dependence of this expression resides in the factor $\nu_m,$ not in the $\Omega_{\rm M}$ out in front. Finally, the directly observed quantity is $N(\ge\!m,z),$ the comoving number density of all clusters of mass $\ge m$ at redshift $z,$ given by \begin{equation} N(\ge\!m,z) = n_0\, \Omega_{\rm M}\!\int_m^\infty \frac{dm'}{m'^2} \left|\frac{d\ln\sigma_M}{d\ln R}\right|_{R_{m'}}\!\nu_{m'} \phi(\nu_{m'})\,. \label{eq:numden} \end{equation} \subsection{Extension to Nongaussian Fluctuations} The above formulae assumed a Gaussian PDF. We now relax this assumption, and assume only that the PDF, $P(\delta|M),$ is such that $\vev{\delta}=0$ and that $\vev{\delta^2}=\sigma_M^2.$ It is useful to express the PDF in terms of a dimensionless function $\psi(x)$ as follows: \begin{equation} P(\delta|M) = \sigma_M^{-1}\,\psi\left(\frac{\delta}{\sigma_M}\right)\,, \label{eq:pgen} \end{equation} where $\psi(x)$ satisfies the conditions $\int \psi(x)\,dx=1,$ $\int x\, \psi(x)\,dx =0,$ and $\int x^2\,\psi(x)\,dx =1\,.$ If one now retraces the usual steps leading to the PS abundance formula, equation~\ref{eq:PS}, but using the PDF given by equation~\ref{eq:pgen} in place of the usual Gaussian, one arrives at the expression \begin{equation} n(M,z) = \frac{2 f_\psi\,\overline\rho}{M^2}\, \frac{\delta_c(z)}{\sigma_M}\left|\frac{d\ln \sigma_M}{d\ln M}\right| \psi\left(\frac{\delta_c(z)}{\sigma_M}\right)\,. \label{eq:psgen} \end{equation} The quantity $f_\psi$ is given by $\left(2 \int_0^\infty \psi(x)\,dx\right)^{-1}$; it corrects for underdense regions that are incorporated into virialized structures according to the standard PS ansatz. Equation~\ref{eq:psgen} is our generalization of the PS formalism for non-Gaussian density perturbations. Setting $\psi(x)$ to a Gaussian yields equation~1, as may readily be verified. Adopting dimensionless mass units and repeating the steps of \S~2.2, we obtain for $N(\ge\!m,z)$ the result \begin{equation} N(\ge\!m,z) = f_\psi n_0\, \Omega_{\rm M}\!\int_m^\infty \frac{dm'}{m'^2} \left|\frac{d\ln\sigma_M}{d\ln R}\right|_{R_{m'}}\!\nu_{m'} \psi(\nu_{m'})\,, \label{eq:numden_gen} \end{equation} where the meaning of $n_0$ and $\nu_m$ are the same as above. The difference between the standard and generalized PS abundance predictions thus consists simply in replacing the Gaussian PDF $\phi$ by the function $\psi,$ apart for the factor $f_\psi,$ which as shown below is $\sim 1$ for cases of interest. \section{Description of NonGaussianity} \label{sec:nongauss} We wish to consider non-Gaussian fluctuations generically, i.e., to construct a PDF meeting the above conditions but which is not tied to a particular model. Moreover, we want our distribution to contain a parameter which quantifies the degree of nongaussianity. As noted above, a possible choice is the $\chi^2_m$ distribution, a generalization of the Peebles (1999a,b) model proposed by White (1998) and by Koyama {\sl et al.}\ (1999). Here we suggest an alternative parameterization. Our reasons for doing so are twofold: first, the $\chi^2_m$ distribution is associated with a particular physical model, whereas our preference is to avoid such association at this preliminary stage in our understanding of the primordial fluctuations; and second, the $\chi^2_m$ model approaches gaussianity only in the limit of very large $m,$ a fact which makes the physical interpretation difficult to sustain if the observationally required degree of nongaussianity is slight. Our model is based on a modified form of the Poisson distribution. Consider, first, an integer random variable $n$ that is Poisson-distributed with expectation value $\lambda.$ The probability that $n$ takes on a particular integer value $m$ is \begin{equation} P(n=m) = \frac{\lambda^m}{m!} e^{-\lambda}\,. \label{eq:poisson} \end{equation} The rms deviation of $n$ is $\sqrt{\lambda}.$ We now imagine that $n$ is continuous rather than discrete, and define a related random variable $x\equiv (n-\lambda)/\sqrt{\lambda}.$ The probability distribution of $x,$ which we refer to as $\psi_\lambda(x),$ is a modified form of the Poisson distribution, but shifted and scaled so that it has mean zero and unit variance. We obtain an explicit representation of $\psi_\lambda(x)$ by letting $m=\sqrt{\lambda} x + \lambda$ in equation~\ref{eq:poisson}, and then multiplying by $\sqrt{\lambda}$ to renormalize the distribution. In addition, the factorial function in the denominator, which is defined only for integer values of its argument, must be replaced by its appropriate generalization for continuous variables, the $\Gamma$-function. This yields \begin{equation} \psi_\lambda (x) = \frac{\lambda^{\sqrt{\lambda}x+\lambda+\frac{1}{2}}\, e^{-\lambda}} {\Gamma(\sqrt{\lambda}x+\lambda+1)} \,. \label{eq:defpsi} \end{equation} This function is defined for $x>-(\sqrt{\lambda}+1/\sqrt{\lambda}).$ As we now show, $\psi_\lambda(x)$ is a suitable representation of quantifiable, generic departures from Gaussianity.\footnote{We note that equation~\ref{eq:defpsi} does not guarantee that $\psi_\lambda(x)$ will have the key properties we seek: normalization, vanishing mean, and unit variance. For small $\lambda,$ the transition from integer to continuous arguments does not preserve these properties, which were inherent in the parent Poisson distribution. By direct integration we have found that departures from these properties are completely negligible for $\lambda > 10.$ For $\lambda < 10,$ we correct $\psi_\lambda(x)$ using formulae derived by fitting the deviations from normalization, vanishing mean, and unit variance. However, these corrections are extremely small for the cases of interest, so that equation~\ref{eq:defpsi} is fully adequate for our purposes.} \vbox{% \begin{center} \leavevmode \hbox{% \epsfxsize=8.9cm \epsffile{figure1.ps}} \begin{small} \figcaption{% Comparison of a Gaussian, $\phi$ (thin red curve) with the $\psi_\lambda$ distribution (heavy blue curve), equation~\ref{eq:defpsi}, for four values of the parameter $\lambda.$ Note the progression of $\psi_\lambda$ toward Gaussianity for $\lambda\gg 1.$ \label{fig:phi_psi_1}} \end{small} \end{center}} Figure~\ref{fig:phi_psi_1} compares the $\psi_\lambda$ distribution with a Gaussian for four values of the parameter $\lambda.$ For $\lambda=3,$ the differences between the two are readily apparent; in particular, one sees the significant skewness of the $\psi_\lambda$ distribution. For $\lambda=5,$ analogous differences can be seen, but they are noticeably smaller, and for $\lambda=10$ the differences are very small. For $\lambda=100,$ the $\psi_\lambda$ is indistinguishable, on this plot, from a Gaussian. The differences between $\psi_\lambda$ and a Gaussian, $\phi,$ become more apparent, even for $\lambda\lower.5ex\hbox{\gtsima} 10,$ if we examine their behavior for large values of $\delta/\sigma.$ This is shown in Figure~\ref{fig:phi_psi_2} for the same four values of $\lambda$ as in the previous figure. We see that for $\lambda \lower.5ex\hbox{\ltsima} 10,$ $\psi_\lambda(x)$ can exceed $\phi(x)$ by 2--3 or more orders of magnitude for $\delta/\sigma \lower.5ex\hbox{\gtsima} 4.$ As noted in \S~\ref{sec:PS}, this same factor enters directly into the PS-predicted cluster abundance. Thus, we expect that if the PDF is described by the $\psi_\lambda$ distribution, massive clusters, representing rare peaks in the initial density field, will be far more abundant than they will in the Gaussian case, provided $\lambda$ is not too large. \vbox{% \begin{center} \leavevmode \hbox{% \epsfxsize=8.9cm \epsffile{figure2.ps}} \begin{small} \figcaption{% The ratio of $\psi_\lambda$ to a Gaussian, $\phi,$ plotted for values of the argument $x \ge 1.$ The plotted ratio shows that ``rare events,'' $\delta/\sigma \lower.5ex\hbox{\gtsima} 3,$ are considerably more likely if the distribution of overdensities is described by the $\psi_\lambda$ function for $\lambda \lower.5ex\hbox{\ltsima} 10,$ than they are in the Gaussian case. \label{fig:phi_psi_2}} \end{small} \end{center}} \vbox{% \begin{center} \leavevmode \hbox{% \epsfxsize=8.9cm \epsffile{figure3.ps}} \begin{small} \figcaption{% The T statistics of the $\psi_\lambda$ (solid line) and $\chi^2_m$ distributions (open circles) plotted versus their respective parameters, $\lambda$ and $m.$ The axes are scaled such that $m=9\lambda.$ The close agreement of the T statistics for the two distributions reveals their strong similarity in terms of the likelihood of rare ($\lower.5ex\hbox{\gtsima} 3\sigma$) events, if one chooses $m=9\lambda.$ \label{fig:Tpsi}} \end{small} \end{center}} The factor $f_\psi$ that appears in equation~\ref{eq:psgen} may be calculated numerically for the $\psi_\lambda$ distribution. We have done so for $2 \leq \lambda \leq 300.$ Within this range the numerical results are very well approximated by the expression \begin{equation} f_\psi(\lambda) = 1 + 0.149 \lambda^{-0.528}\,. \label{eq:fcorr} \end{equation} Note that $f_\psi$ differs from unity by less than 10\% for $\lambda\lower.5ex\hbox{\gtsima} 2.$ \subsection{Comparison with the $\chi^2_m$ distribution} The model introduced by White (1998) and by Koyama, Soda, \& Taruya (1999) postulates that the CDM density field is given by \begin{equation} \rho_{CDM}({\bf x}) = \frac{\mu}{2}\sum^m_{i=1} \phi_i^2({\bf x}) \label{eq:cdmchi} \end{equation} where the $\phi_i({\bf x})$ are statistically independent Gaussian fields. The overdensity $\nu=\delta/\sigma$ in this model is distributed like a $\chi^2$ variable with $m$ degrees of freedom, shifted to have vanishing mean and unit variance. Explicitly, \begin{equation} P(\nu) d\nu = \frac{\left(1+\sqrt{\frac{2}{m}}\nu\right)^{ m/2-1}}{\left(\frac{2}{m}\right)^{(m-1)/2} \Gamma\left(\frac{m}{2}\right)} \exp\left(-\frac{m}{2}-\sqrt{\frac{m}{2}}\nu\right) d\nu\,, \label{eq:chi2m} \end{equation} a distribution we henceforth label $\chi^2_m.$ Koyama, Soda, \& Taruya (1999) (see also Robinson, Gawiser, \& Silk 1998) have advocated quantifying ``rare event'' nongaussianity in terms of a ``T-statistic'' defined by' \begin{equation} T = \frac{\sqrt{2\pi} \int_3^\infty P(\nu) d\nu}{\int_3^\infty e^{-\nu^2/2} d\nu}\,, \label{eq:defT} \end{equation} the likelihood relative to Gaussian of $3\,\sigma$ or rarer events. In terms of this statistic the relationship between the $\psi_\lambda$ and $\chi^2_m$ distributions becomes particularly clear. In Figure~\ref{fig:Tpsi}, the T statistics for the two distributions are plotted versus their respective parameters, with the axes scaled such that $m=9\lambda.$ With this scaling, the two distributions have T statistics which agree to within a few percent for all $\lambda.$ This shows that in terms of the likelihood of rare events, the $\psi_\lambda$ and $\chi^2_m$ distributions are extremely similar for $m=9\lambda.$ It is also interesting to note that, despite the near-indistinguishability of $\psi_\lambda$ from a Gaussian for $\lambda\lower.5ex\hbox{\gtsima} 100$ (Figure 1), the two distributions approach the Gaussian value $T=1$ extremely slowly as $\lambda\rightarrow\infty.$ Indeed, one requires $m \simeq 200$ even to achieve $T=2,$ twice the Gaussian value. It is for this reason that the $\chi^2_m$ model may not be suitable for describing mild Gaussianity---at least if one were to take its physical basis seriously---because the required number of primordial Gaussian fields is excessively large. \section{On the Determination of Cluster Virial Masses} \label{sec:clmass} It is essential when applying PS abundance estimates that one use the rigorous definition of virial mass. Specifically, the PS formulae apply to the mass $M_V$ interior to a radius $r_V$ such that $3 M_V/4\pi r_V^3 = \Delta_V(z,\Omega_{\rm M},\Omega_\Lambda) \overline\rho (z),$ where $\Delta_V(z,\Omega_{\rm M},\Omega_\Lambda)$ is a cosmology- and redshift-dependent overdensity factor. In an Einstein-de Sitter universe, $\Delta_V\simeq 178$ at all redshifts; for $\Omega_{\rm M}<1,$ $\Delta_V$ is larger than this value, and increases with decreasing redshift. Kitayama \& Suto (1996) have derived analytic approximations for $\Delta_V(z)$ as a function of $\Omega_{\rm M}$ for flat and open cosmologies, and we use their formulae in this paper. The definition of virial mass above means that one cannot compute $M_V$ from observational data, such as velocity dispersion, X-ray temperature, or weak lensing, without specifying both a cosmology ($\Omega_{\rm M}$ and $\Omega_\Lambda$) and a model for the radial mass distribution of the cluster. The X-ray or velocity data are usually derived from the dense, central parts of the cluster ($r\lower.5ex\hbox{\ltsima} 500\ifmmode h^{-1}\else$h^{-1}$\fi$ kpc), whereas the virial radius is generally in the range 1--$1.5\ifmmode h^{-1}\else$h^{-1}$\fi$ Mpc for massive clusters. Consequently, an extrapolation is entailed. Weak lensing data do yield mass as a function of aperture that, in the case of MS1054, extend nearly to the virial radius, but the mass in question is a projected mass density, so that, again, a model is required to obtain the virial mass. These issues are not always well appreciated; in this section we derive $M_V$ in the context of a specific mass model in order to make clear the assumptions involved. We adopt the profile of Navarro, Frenk, \& White (1997; NFW), which has been shown to be a good fit to cluster-scale dark matter halos in N-body distributions. The NFW profile has a density distribution given by \begin{equation} \rho(r) = \frac{\rho_{crit}\, \delta_c}{x (1+x)^2}\,, \label{eq:rhonfw} \end{equation} where $x\equiv r/r_s$ and $r_s$ is a scale radius for the halo. The central density is determined by the parameter $\delta_c,$ which one expects to be of order $10^4$ for massive clusters. The mass interior to radius $r$ is then given by \begin{equation} M(r) = 4\pi\rho_{crit}\,\delta_c\,r_s^3 \left[\ln(1+x)-\frac{x}{1+x}\right]\,. \label{eq:mnfw} \end{equation} Substituting equation~\ref{eq:mnfw} into the expression for virial mass yields \begin{equation} \frac{\ln(1+x_V) - \frac{x_V}{1+x_V}}{x_V^3} = \frac{\Delta_V (1+z)^3\,\Omega_{\rm M}}{3\,\delta_c}\,, \label{eq:xv} \end{equation} where $x_V\equiv r_V/r_s.$ \subsection{Obtaining $M_V$ from X-ray temperature or velocity dispersion} Now let us suppose that the observational data consist either of a measurement of the X-ray temperature $T_X$ of the intracluster gas, or of the rms line-of-sight velocity dispersion of cluster galaxies, $\sigma_v.$ Applying the equation of hydrostatic equilibrium to the former, or the Jeans equation to the latter, leads to the equation \begin{equation} \frac{G M(r)}{r^2}\,\rho_{X,g}\, = \, -\frac{d}{dr}\left(\rho_{X,g} u^2\right)\,, \label{eq:hydro} \end{equation} where $\rho_{X,g}$ is the density of the X-ray emitting gas or the galaxy population at radius $r,$ and \begin{equation} u^2 = \left\{ \begin{array}{ll} \frac{kT_X}{\mu\,m_p} \,, & {\rm X-ray\ gas\/} \,;\\ & \\ \sigma_v^2\,, & {\rm galaxy\ velocities}\,. \end{array} \right. \end{equation} We can solve equation~\ref{eq:hydro} if we make the simplifying assumption that the tracer (gas or galaxies) distribution is roughly isothermal at the radii from which most of the X-ray emission (galaxy velocities) are derived, $r \approx r_s.$ I.e., we assume that $d\ln\rho_{X,gas}/d\ln r \approx -2,$ $u^2 \approx$ constant. This assumption is reasonable if the gas (galaxies) approximately trace the dark matter potential, because the NFW profile itself is nearly isothermal at such radii. With the isothermal assumption we obtain from equation~\ref{eq:hydro} the relation \begin{equation} \delta_c \approx 7\left(\frac{u}{H_0 r_s}\right)^2\,. \label{eq:rsdc} \end{equation} \vbox{% \begin{center} \leavevmode \hbox{% \epsfxsize=8.9cm \epsffile{figure4.ps}} \begin{small} \figcaption{% The virial mass (upper left panel), virial radius (lower left panel), NFW concentration parameter (upper right panel), and dimensionless central density $\delta_c$ (lower right panel), for a cluster with an X-ray temperature of 12.3 keV, plotted as a function of the adopted value of the NFW scale radius $r_s.$ See text for details. \label{fig:nfw_X}} \end{small} \end{center}} Equation~\ref{eq:rsdc} shows that $\delta_c$ is determined from observations (i.e., $T_X$ or $\sigma_v$) only if the scale radius $r_s$ is known. In other words, the two NFW parameters cannot be determined from a single piece of information. Nevertheless, one might hope that the virial mass itself is insensitive to the particular values of $\delta_c$ and $r_s$ provided they are related by equation~\ref{eq:rsdc}. This indeed turns out to be the case, as we show by adopting the following procedure. First, we pick a value of $r_s$ in the range 250--1100\ifmmode h^{-1}\else$h^{-1}$\fi\ kpc. From this we derive $\delta_c$ from equation~\ref{eq:rsdc} and insert the result into equation~\ref{eq:xv} to obtain $x_V.$ Substituting into equation~\ref{eq:mnfw} then yields, after some algebra, the result \begin{equation} M_V = \frac{21}{2}\,\frac{u^2}{G}\,r_s \left[\ln(1+x_V)- \frac{x_V}{1+x_V}\right]\,. \label{eq:mvir} \end{equation} Several aspects of equation~\ref{eq:mvir} merit further comment. First, $x_V$ is a function of the observable $u$ and the adopted value of $r_s.$ Thus, $M_V$ is not quadratic in $u$ and linear in $r_s.$ Indeed, the virial mass is virtually independent of the adopted value of $r_s,$ as was argued above. This is shown in the upper left panel of Figure~\ref{fig:nfw_X}, in which the $M_V$ is computed from an X-ray temperature of $T_X=12.3$ keV, the value obtained by Donahue {\sl et al.}\ (1998) for MS1054 (cf.\ \S~\ref{sec:ms1054}). (The figure assumes a flat, $\Omega_{\rm M}=0.3$ cosmology at $z=0.83,$ the redshift of MS1054.) For $300 \lower.5ex\hbox{\ltsima} r_s \lower.5ex\hbox{\ltsima} 1000\ifmmode h^{-1}\else$h^{-1}$\fi$ kpc, $M_V$ varies by less than $\sim\,$ 10\% from its mean value of $1.3\times 10^{15}\, \h1M_\odot$ for the adopted $T_X.$ The virial radius $r_V$ changes by an even smaller percentage over this range. In contrast, the inferred central density $\delta_c$ is quite sensitive to the adopted $r_s.$ To constrain $r_s$ further, the NFW concentration index $c$ is plotted in the upper right hand panel. (This quantity is not simply equal to $r_V/r_s,$ as it is in Navarro {\sl et al.}, because the virial overdensity is not taken to be exactly 200.) For massive clusters, the simulations of Navarro {\sl et al.}\ suggest that $c\simeq 3$--5. Thus, the plausible values of $r_s$ are in the range $300 \lower.5ex\hbox{\ltsima} r_s \lower.5ex\hbox{\ltsima} 700\ifmmode h^{-1}\else$h^{-1}$\fi$ kpc. As we have seen, $M_V$ varies little for $r_s$ in this range. Consequently, we can be assured that our procedure introduces less than about 10\% error in the virial mass. Second, the virial mass as determined from equation~\ref{eq:mvir} differs from what would be obtained had we assumed (as is usually done) a purely isothermal structure for the cluster. Specifically, the virial mass obtained from the NFW model is 10--20\% {\em larger\/} than an isothermal mass estimate, with the precise factor depending on the typical concentration index. In a sense it is reassuring that the difference is relatively small; on the other hand, as we shall see below, even small mass differences can be crucial in terms of inferences about nongaussianity. Constraining the mass profiles of clusters is thus an important ongoing task for cluster physics. Last, it is important to bear in mind that $M_V$ as derived from equation~\ref{eq:mvir} is dependent on the adopted cosmology and redshift, via equation~\ref{eq:xv}, from which $x_V$ is derived. Thus, {\em the virial mass of a cluster, derived from temperature or velocity observations, cannot be specified independently of adopted cosmology.} This property of the virial mass, as defined by the PS formalism, will be important in the analysis presented below. \subsection{Obtaining $M_V$ from weak lensing data} Weak lensing data constitute the third and, in principle, most rigorous means of deriving cluster masses, as there is no need to assume hydrostatic or dynamical equilibrium. Deriving a virial mass given aperture mass measurements nonetheless requires care, as we now discuss. \def{\overline \kappa}{{\overline \kappa}} The quantity measured directly by a weak lensing shear analysis (e.g., Luppino \& Kaiser 1996) is the mean dimensionless surface density ${\overline \kappa}$ within an angular radius $\theta$ of the cluster center. The projected mass within $\theta$ is then given by \begin{equation} M(\theta) = {\overline \kappa} \frac{c^2}{4\pi G} \frac{D_S D_L}{D_{LS}} \pi\theta^2\,, \label{eq:mtheta} \end{equation} where $D_L,$ $D_S,$ and $D_{LS}$ are the lens, source, and lens-source angular diameter distances respectively. These angular diameter distances are all cosmology-dependent; moreover, the typical redshift of the lensed sources must be assumed. These factors all influence the derived virial mass. Evaluating numerical factors in equation~\ref{eq:mtheta} yields \begin{equation} M(\theta) = 2.13\times 10^{15} \frac{{\overline \kappa}}{0.1} \left(\frac{\theta} {4'}\right)^2 \frac{y_L y_S}{(1+z_L) y_{LS}} \h1M_\odot\,, \label{eq:mtheta1} \end{equation} where the $y$'s are the dimensionless comoving angular diameter distances as defined by Peebles (1993). To make the connection with the NFW halo parameters, we must equate the observationally derived aperture mass above with the projected mass calculated from the NFW formulae; we show elsewhere (Willick \& Padmanabhan 1999) that this is given by \begin{equation} M(\theta;\delta_c,r_s) = 4\pi\rho_{crit}\delta_c r_s^3 \beta(D_L\theta/r_s)\,, \label{eq:mthetanfw} \end{equation} where \begin{equation} \beta(x) = \int_0^x y\,dy \int_0^\infty (y^2+w^2)^{-1/2} (1+y^2+w^2)^{-1}\,dw \end{equation} is the projected (dimensionless) mass density for the NFW profile. (Explicit expressions for $\beta$ are given by Willick \& Padmanabhan 1999.) Equating the lensing-inferred aperture mass (equation~\ref{eq:mtheta1}) with the NFW model aperture mass (equation~\ref{eq:mthetanfw}) yields the following expression for $\delta_c$ as a function of $r_s,$ ${\overline \kappa},$ and $\theta:$ \begin{equation} \delta_c= 4886 \frac{{\overline \kappa}}{0.1} \left(\frac{\theta} {4'}\right)^2 \beta^{-1} \left(\frac{r_s}{500\ {\rm kpc}}\right)^{-3} \frac{y_L y_S}{(1+z_L)y_{LS}}\,. \label{eq:dclens} \end{equation} \vbox{% \begin{center} \leavevmode \hbox{% \epsfxsize=8.9cm \epsffile{figure5.ps}} \begin{small} \figcaption{% Same as the previous figure, except that now the virial mass and related quantities are derived from lensing data. A mean convergence of ${\overline \kappa}=0.08$ within $\theta=4',$ the values for MS1054, have been used in the calculation. \label{fig:nfw_L}} \end{small} \end{center}} Equation~\ref{eq:dclens} is the analogue of equation~\ref{eq:rsdc} for the case of a lensing measurement; it gives $\delta_c$ from the data for any adopted value of $r_s.$ This suggests that we proceed as above: pick a value of $r_s$ within a sensible range; calculate the corresponding $\delta_c$ from equation~\ref{eq:dclens}, and thus the virial radius in units of $r_s,$ $x_V,$ from equation~\ref{eq:xv}; obtain the virial mass $M_V=4\pi r_V^3 \Delta_V\overline\rho/3.$ We expect that, as found in the temperature/velocity dispersion case, the virial mass will be insensitive to the choice of $r_s.$ Figure~\ref{fig:nfw_L}, in which virial mass, radius, NFW concentration index, and central density are plotted for the lensing parameters of MS1054 (see next section), shows that this is once again the case. \section{The Virial Mass of MS1054} \label{sec:ms1054} We now apply the methods of the previous section to estimate the virial mass of MS1054 as a function of $\Omega_{\rm M}.$ We assume a flat universe, $\Omega_{\rm M}+\Omega_\Lambda=1,$ so that the quantity $\Delta_V(z)$ discussed in the previous section is fully determined by $\Omega_{\rm M}.$ MS1054 lies at a redshift of $z=0.833$ (Tran {\sl et al.}\ 1999, T99), making it one of the most distant confirmed rich clusters. Its mass can be estimated via each of the three principal methods of the previous section. T99 found it to have a velocity dispersion of $1170\pm 150\ \ifmmode{\rm km}\,{\rm s}^{-1}\else km$\,$s$^{-1}$\fi$ based on 24 galaxies. An earlier determination of the velocity dispersion by Donahue {\sl et al.}\ (1998; D98) found $\sigma_v=1360\ \ifmmode{\rm km}\,{\rm s}^{-1}\else km$\,$s$^{-1}$\fi$ based on 12 galaxies. D98 also obtained ASCA data for MS1054, from which they deduced a temperature of $T_X=12.3^{+3.1}_{-2.2}$ keV for the intracluster gas, equivalent to a velocity dispersion $kT_X/\mu m_p=(1413\ \ifmmode{\rm km}\,{\rm s}^{-1}\else km$\,$s$^{-1}$\fi)^2$ for $\mu=0.59,$ the value adopted by Borgani {\sl et al.}\ (1999). Luppino \& Kaiser (1997, LK97) obtained $V$ and $I$ band images of MS1054 at the 2.2 m telescope on Mauna Kea in good ($\sim 1''$) seeing, and used these data to estimate ${\overline \kappa}(\theta).$ Their mass estimates (for which they assumed $\Omega_{\rm M}=1$) indicate a ${\overline \kappa}=0.1$ for $\theta=4$ arcmin. However, a more conservative estimate, obtained from their Figure 7, is ${\overline \kappa}=0.08$ for $\theta=4',$ and we use this more conservative estimate here. Using these data we can determine the virial mass of MS1054 for an adopted cosmology. Because of the $\sim 10\%$ variations in the deduced mass with the adopted value of $r_s,$ we must impose an additional constraint. The most reasonable one to adopt is to require the NFW concentration parameter $c$ to take on a particular value. Here we choose $c=4.$ For the lensing mass, an additional assumption is required, namely, the mean redshift, $z_S,$ of the lensed galaxy population. Here we choose $z_S=1.5,$ which LK97 considered the most suitable value, and which yields mass estimates midway between the minimum ($z_S\approx 1$) and maximum ($z_S\approx 3$) allowed values (LK97). Figure~\ref{fig:ms1054_mass} shows the virial masses obtained via each of the three methods as a function of $\Omega_{\rm M}.$ The cosmological dependence of the virial mass is clear. We have also calculated, for use in the next section, a weighted average mass and the $1\,\sigma$\ errors on this average. Mass errors were estimated by considering the principal sources of error for each method: velocity dispersion uncertainty for the dynamical mass, temperature uncertainty for the X-ray mass, and source redshift uncertainty for the weak lensing mass. The corresponding mass errors were estimated to be 42\%, 35\%, and 40\% respectively. The individual masses were weighted inversely with the squares of the fractional mass errors to obtain the average mass, which is shown as a heavy solid curve in the plot; the shaded region around the average shows its $\pm 1\,\sigma$ uncertainty, which is $\sim 22\%.$ As the figure shows, while the X-ray and lensing masses agree to within $\sim\,$ 10\% for $\Omega_{\rm M}\lower.5ex\hbox{\ltsima} 0.5,$ the dynamical mass is 40--50\% smaller than the other two. This difference is within the observational errors, and thus is not indicative of systematic differences. Nonetheless, the sensitivity of the PS-predicted cluster abundance to mass makes this difference potentially significant. We discuss this important issue further in \S~7.2 \vbox{% \begin{center} \leavevmode \hbox{% \epsfxsize=8.9cm \epsffile{figure6.ps}} \begin{small} \figcaption{% The virial mass of MS1054 computed from X-ray, galaxy velocity, and lensing data, plotted as a function of $\Omega_{\rm M}.$ The weighted average is plotted as a solid black line, with the light shading showing the $\pm 1\,\sigma$ errors. \label{fig:ms1054_mass}} \end{small} \end{center}} \section{Method of Analysis and Basic Results} \label{sec:results} \def\sigma_8{\sigma_8} In order to apply the PS formalism to high-redshift clusters, one must first normalize the mass fluctuations on cluster scales by requring the PS prediction to match observations at {\em low\/} redshift. The result is generally expressed as a relationship between $\sigma_8,$ the value of $\sigma_M$ within a top-hat sphere of radius $8\ifmmode h^{-1}\else$h^{-1}$\fi$ Mpc, and $\Omega_{\rm M}.$ Recent results obtained from X-ray temperature and luminosity data include those of Borgani {\sl et al.}\ (1999): \begin{equation} \sigma_8 = (0.58\pm 0.06) \times \Omega_{\rm M}^{-0.47+0.16\Omega_{\rm M}}\,; \label{eq:B99} \end{equation} Wang \& Steinhardt (1998): \begin{equation} \sigma_8 = (0.50\pm 0.10) \times \Omega_{\rm M}^{-0.43-0.33\Omega_{\rm M}}\,; \label{eq:BWS98} \end{equation} and Pen (1998): \begin{equation} \sigma_8 = (0.53 \pm 0.05) \times \Omega_{\rm M}^{-0.53}\,. \label{eq:P98} \end{equation} (In each case, results for a flat universe, $\Omega_{\rm M}+\Omega_\Lambda=1,$ are given; in the case of Wang \& Steinhardt (1998) nominal values of other cosmological parameters have been adopted. See the original papers for further details.) Similar expressions have been given by Eke, Cole, \& Frenk (1996), Girardi {\sl et al.}\ (1998), and Suto {\sl et al.}\ (1999). \vbox{% \begin{center} \leavevmode \hbox{% \epsfxsize=8.9cm \epsffile{figure7.ps}} \begin{small} \figcaption{% Three recently published $\sigma_8$-$\Omega_{\rm M}$ relations derived from low-redshift cluster X-ray temperature and luminosity data. \label{fig:sig8omegam}} \end{small} \end{center}} Figure~\ref{fig:sig8omegam} shows $\sigma_8$ as a function of $\Omega_{\rm M}$ as given by equations~\ref{eq:B99}, ~\ref{eq:BWS98}, and~\ref{eq:P98}. The agreement is within the reported errors, and is especially good for $0.4\lower.5ex\hbox{\ltsima} \Omega_{\rm M} \lower.5ex\hbox{\ltsima} 0.7.$ However, because of the exponential sensitivity of the PS abundance predictions to $\sigma_M,$ the small differences in figure~\ref{fig:sig8omegam} have a nonnegligible effect on the inferred degree of nongaussianity. In what follows, we adopt the Borgani {\sl et al.}\ (1999) calibration, which is intermediate between the other two for $\Omega_{\rm M}$ in the observationally preferred range ($\sim\,$ 0.2--0.5). Furthermore, it is based on the most extensive and recent compilation of X-ray data, on which the conclusions of the present paper heavily depend. \subsection{Effect of nongaussianity on the $\sigma_8$-$\Omega_{\rm M}$ relation} The above low-redshift $\sigma_8$-$\Omega_{\rm M}$ relations were, of course, derived under the assumption of Gaussian primordial density perturbations. If we relax this assumption---e.g., assume the fluctuations are described by the $\psi_\lambda$ distribution for finite $\lambda$---we expect the $\sigma_8$-$\Omega_{\rm M}$ relation to change as well. To obtain the relation for finite $\lambda$ we first note that the relations essentially represent the requirement that the $z=0$ PS abundance prediction, for any $\Omega_{\rm M},$ match the observed abundances at a characteristic rich cluster mass. Specifically, the Borgani {\sl et al.}\ (1999) calibration yields $M n(M) = 4.2 \times 10^{-6}\ \ifmmode h^{-1}\else$h^{-1}$\fi\ {\rm Mpc}^{-3}$ for $M = 5.95 \times 10^{14}\,\h1M_\odot$ (i.e., for $m=1$). By requiring that this condition hold for finite $\lambda$ as well, we obtain modified $\sigma_8$-$\Omega_{\rm M}$ relations for any desired degree of nongaussianity. \vbox{% \begin{center} \leavevmode \hbox{% \epsfxsize=8.9cm \epsffile{figure8.ps}} \begin{small} \figcaption{% The ratio of $\sigma_8$ for the $\psi_\lambda$ distribution to the Borgani {\sl et al.}\ value of $\sigma_8,$ plotted as a function of $\Omega_{\rm M}.$ The values of $\sigma_8(\lambda)$ were obtained by requiring the generalized PS abundance of $M=6\times 10^{14}\,\ifmmode h^{-1}\else$h^{-1}$\fi\ M_\odot$ clusters to match that obtained for the Gaussian case, at each value of $\Omega_{\rm M}.$ \label{fig:s8sl}} \end{small} \end{center}} Figure~\ref{fig:s8sl} shows the effect on the $\sigma_8$-$\Omega_{\rm M}$ relation of using the generalized PS abundance formula, equation~\ref{eq:psgen}, with the $\psi_\lambda$ distribution for $\lambda=3,$ $10,$ and $100.$ It can be seen that the changes relative to the Gaussian case are rather small, especially for $\Omega_{\rm M}\leq 0.2.$ This is because low-redshift clusters of moderate mass do not represent very high peaks in the initial density field, for low $\Omega_{\rm M}.$ For significant nongaussianity $\lambda\lower.5ex\hbox{\ltsima} 3$ and $\Omega_{\rm M}\lower.5ex\hbox{\gtsima} 0.4,$ the fluctuation normalization is more subtantially modified. \subsection{Calculating the expected number of MS1043-like clusters} Once we have modified the normalization of the density fluctuations $\sigma_M$ for a given value of $\lambda$ as described above, we may calculate the predicted comoving number density of clusters above a given mass threshold using equation~\ref{eq:numden_gen} with $\psi=\psi_\lambda.$ When Gaussian fluctuations are assumed, we use equation~\ref{eq:numden}. To determine the number of clusters, of mass and redshift as large or larger than that of MS1054, expected in the EMSS sample, we integrate this number density over redshift, multiplying by the appriate comoving volume element: \begin{equation} {\cal N}_{exp} = \omega_{1054} \int_{z_1}^{z_{max}} N(\ge\!m_{1054},z) \frac{dV}{dz}\,dz\,, \label{eq:nexp} \end{equation} where: \begin{enumerate} \item $m_{1054}$ is the virial mass, in dimensionless units (cf.\ \S~2.2), of MS1054 as determined by the methods of \S~\ref{sec:clmass}; \item $\omega_{1054}$ is the solid angle covered by the Einstein satellite to a limiting X-ray flux fainter than the observed flux of MS1054. From Tables 1 and 2 of Henrey {\sl et al.}\ (1992) we find $\omega_{1054}=0.041$ sterradian; \item $z_1=0.833$ is the redshift of MS1054 (T99); \item $z_{max}$ is the maximum redshift out to which a cluster whose X-ray luminosity is equal to that of MS1054 could have been detected by the EMSS. In practice, $N(\ge\!m,z)$ is dropping so rapidly with increasing redshift that the result is insensitive to $z_{max}$ for $z_{max}\lower.5ex\hbox{\gtsima} 1,$ and we conservatively set $z_{max}=1.3$ in the calculations to follow; \item $dV/dz$ is the cosmology-dependent comoving volume per unit redshift (e.g., equation~13.61 of Peebles 1993). \end{enumerate} \vbox{% \begin{center} \leavevmode \hbox{% \epsfxsize=8.9cm \epsffile{figure9.ps}} \begin{small} \figcaption{% Expected number of MS1054-like clusters in the EMSS survey, plotted as a function of mass, under the assumption of Gaussian initial density fluctuations. The $\sigma_8$-$\Omega_{\rm M}$ relation is that of Borgani {\sl et al.}\ (1999), and a flat universe is assumed. The different curves are for four different values of $\Omega_{\rm M}.$ The heavy dashed line sloping down from right to left indicates the variation of the virial mass of MS1054 with $\Omega_{\rm M}$ (see Figure~\ref{fig:ms1054_mass}). The horizontal dotted line indicates a one in ten chance that a cluster as massive as MS1054 would be present in the EMSS survey. Note that only this occurs only for $\Omega_{\rm M}=0.15.$ \label{fig:nexp_gauss}} \end{small} \end{center}} Figures~\ref{fig:nexp_gauss} and~\ref{fig:nexp_lambda} show the results of calculating ${\cal N}_{exp}$ for a Gaussian PDF and a $\lambda=1.5$ $\psi_\lambda$-PDF respectively. Each figure plots ${\cal N}_{exp}$ for four representative values of $\Omega_{\rm M}.$ The heavy dashed line sloping down and to the left indicates how the virial mass of MS1054 changes with $\Omega_{\rm M}$ (cf.\ \S~\ref{sec:ms1054}). Where that line intersects the curve for a given $\Omega_{\rm M}$ yields the predicted number of MS1054-like clusters in the EMSS. As Figure~\ref{fig:nexp_gauss} shows, this number is quite small for all values of $\Omega_{\rm M}\ge 0.25$ in the Gaussian case. A cluster at $z=0.83,$ and as massive as the lensing, X-ray, and velocity data for MS1054 indicate that it is, is unlikely to have been found if the fluctuations are Gaussian, unless $\Omega_{\rm M}\lower.5ex\hbox{\ltsima} 0.2.$ The situation is significantly changed for a $\psi_\lambda$-PDF with $\lambda=1.5.$ Figure~\ref{fig:nexp_lambda} shows that an MS1054-like cluster now has a better than 10\% chance of being found for $\Omega_{\rm M}\lower.5ex\hbox{\ltsima} 0.4.$ Thus, nongaussian fluctuations allow a much larger value of $\Omega_{\rm M}$ to be consistent with the MS1054 data. \vbox{% \begin{center} \leavevmode \hbox{% \epsfxsize=8.9cm \epsffile{figure10.ps}} \begin{small} \figcaption{% Same as the previous figure, except that now the expected number of clusters is calculated using a PDF described by the $\psi_\lambda$ distribution, with $\lambda=1.5.$ \label{fig:nexp_lambda}} \end{small} \end{center}} Because we are dealing with a sample of one, we cannot hope to estimate a value of $\lambda$ per se. It is, however, reasonable to derive an upper limit on $\lambda$ (i.e., a lower limit on the required degree of nongaussianity) as a function of $\Omega_{\rm M}.$ To do so, we calculate, for each $\Omega_{\rm M},$ the value of $\lambda$ required to yield ${\cal N}_{exp}=0.1,$ i.e., a one in ten chance that an MS1054-like cluster would be found in the EMSS for that value of $\Omega_{\rm M}.$ The resultant value of $\lambda$ may be thought of as a 90\% confidence level upper limit on $\lambda.$ The corresponding value of the $T$-statistic (see Figure~\ref{fig:Tpsi}) would then be a 90\% confidence level lower limit. Figure~\ref{fig:lambda_T} shows the results of carrying out this calculation, and thus summarizes the main results of this paper. The upper limit on $\lambda$ decreases from $\infty$ (Gaussian fluctuations) for $\Omega_{\rm M}=0.17,$ to $\sim\,$ 100 for $\Omega_{\rm M}=0.20,$ to much smaller values for $\lambda\lower.5ex\hbox{\gtsima} 0.25.$ As the upper limit on $\lambda$ decreases with increasing $\Omega_{\rm M},$ signifying increasing nongaussianity, the lower limit on $T$ increases, from 1 (Gaussian fluctuations) at $\Omega_{\rm M}=0.17,$ to $\sim\,$ 2 at $\Omega_{\rm M}=0.25,$ to $\lower.5ex\hbox{\gtsima} 6$ for $\Omega_{\rm M}\ge 0.4.$ The required amount of nongaussianity rapidly increases with increasing $\Omega_{\rm M}.$ This is ultimately a reflection of of the later ``freeze-out'' time for fluctuation growth in higher density universes. \vbox{% \begin{center} \leavevmode \hbox{% \epsfxsize=8.9cm \epsffile{figure11.ps}} \begin{small} \figcaption{% The upper panel shows the 90\% confidence level upper bound on $\lambda$ as a function of $\Omega_{\rm M}.$ The lower panel shows the corresponding {\em lower\/} bound on $T.$ These 90\% bounds correspond to the amount of nongaussianity required for the expected number of MS1054-like clusters in the EMSS to be $\ge 0.1.$ See main text for further details. \label{fig:lambda_T}} \end{small} \end{center}} \section{Discussion} \label{sec:disc} We have shown that the X-ray selected cluster MS1054--03, at $z=0.83,$ adds to a small but growing body of evidence that the primordial density fluctuation field may be nongaussian. Specifically, the high mass of MS1054, $\lower.5ex\hbox{\gtsima} 10^{15}\,\h1M_\odot,$ indicates that rare fluctuations, $\delta/\sigma_M\geq 3,$ are more probable at early times than they would be in the Gaussian case, if $\Omega_{\rm M}\lower.5ex\hbox{\gtsima} 0.2.$ In this concluding section, we further discuss several aspects of this result. \subsection{Comparison with previous work} \subsubsection{Previous analyses based on MS1054} Two other recent papers have considered the cosmological implications of MS1054: D98 (cf.\ \S~\ref{sec:ms1054}) and Bahcall \& Fan (1998; BF98). Both papers reached the conclusion that, by virtue of its high mass, MS1054 by itself rules out an Einstein-de Sitter universe at a high confidence level. D98 stated that MS1054 is consistent with a flat $\Omega_{\rm M}=0.3$ cosmology, while BF98 found that MS1054 is most consistent with $\Omega_{\rm M}=0.2.$ Our results are in rough accord with theirs, although for Gaussian fluctuations we find $\Omega_{\rm M}\le 0.17,$ somewhat smaller than BF98 and markedly lower than D98. The latter difference may result from D98's conservative estimate of the virial mass of MS1054, $\sim 7\times 10^{14}\,\h1M_\odot,$ which assumed $\Omega_{\rm M}=1$ and an isothermal mass model. In any case, comparison with D98 and BF98 is imprecise because they did not consider nongaussianity as an additional degree of freedom in the abundance analysis; their constraints on $\Omega_{\rm M}$ are valid only if the initial density field is Gaussian. \subsubsection{Other cluster analyses allowing nongaussianity} A more direct comparison of our results may be made with the recent papers by Robinson, Gawiser, \& Silk (1998; RGS98) and Koyama, Soda, and Taruya (1999; KST99). The main points of those papers were summarized in \S~\ref{sec:intro}; here we discuss them further in light of our findings. RGS98 and KST99 considered not only the cluster abundance, but also the cluster correlation length, using observational constraints on the latter obtained from the APM survey (Croft {\sl et al.}\ 1997), and compared with the predictions of the generalized PS formalism. While increasingly nongaussian PDFs {\em increase\/} the predicted cluster abundance (for given $\Omega_{\rm M}$ and $\sigma_8$), they {\em decrease\/} the predicted correlation length. This helps break the degeneracy between $\Omega_{\rm M}$ and nongaussianity (e.g., $T$) present in any analysis which, like the present one, considers only abundance data. In particular, while our analysis allows Gaussian fluctuations for sufficiently low $\Omega_{\rm M},$ RGS98 and KST99 found that large $T$ is in fact required for low ($\lower.5ex\hbox{\ltsima} 0.2$) $\Omega_{\rm M}$ in order to accommodate both abundance and correlation length data. The analyses of RGS98 and KST99 differed in that only the latter used intermediate-redshift cluster abundance as an observational constraint (each considers $z\lower.5ex\hbox{\ltsima} 0.1$ abundances and correlations). This makes KST99 the more powerful probe of parameter space. While RGS98 found an Einstein-de Sitter universe with a Gaussian PDF to be a good fit to the data, KST99 ruled out $\Omega_{\rm M}\ge 0.5.$ They found that the required degree of nongaussianity is a minimum for $\Omega_{\rm M}\simeq 0.3,$ at which $T=3.8\pm 1.0.$ From Figure~\ref{fig:lambda_T} we see that for $\Omega_{\rm M}\simeq 0.3$ our study indicates $T \ge 3,$ consistent with the KST99 result. For a flat $\Omega_{\rm M}=0.3$ universe with $\Gamma=0.20,$ RGS98 obtained $T=4.0^{+3.6}_{-2.0},$ consistent with both KST99 and this paper. \subsection{On the form of nongaussian PDF} KST99 assumed $P(\delta|M) \propto \chi^2_m(\delta/\sigma_M),$ whereas we have taken $P(\delta|M) \propto \psi_\lambda(\delta/\sigma_M).$ The two PDFs differ more in spirit than in practice. The $\chi^2_m$ distribution has its roots in a physical model, in which the CDM is produced at the end of an inflationary epoch from $m$ scalar fields which couple quadratically to a Gaussian inflaton field. The $\psi_\lambda$ distribution, on the other hand, has no particular physical motivation, being simply a mathematical transformation of the familiar Poisson distribution (\S~\ref{sec:nongauss}). As we showed in \S~\ref{sec:nongauss}, the two distributions are remarkably similar in terms of their $T$-statistic, provided one makes the identification $m=9\lambda.$ In fact, $\chi^2_{m=9\lambda}(x)$ and $\psi_\lambda(x)$ are similar (though not identical) at all values of $x,$ not only in their $T$ values, for $\lambda \lower.5ex\hbox{\gtsima} 3.$ Our reason for introducing the $\psi_\lambda$ distribution is primarily to make the philosophical point that nongaussianity can be modeled phenomenologically, and that as a result we should be cautious in interpreting our findings as evidence of any particular physical model. Thus, although the cluster data point toward a PDF with $T\approx 4,$ this does not imply that the quadratically coupled inflationary model with $m\approx 40$ is correct. Indeed, any theory which invokes such a large number of identical but independent scalar fields is suspect on Occam's Razor grounds alone. More to the point, any PDF with the indicated level of nongaussianity---such as the $\psi_\lambda$ distribution with $\lambda\simeq 3$---can can account for the cluster data. Of course, even if the $\psi_\lambda$ distribution proves able to describe the cluster abundance data, this will not necessarily constitute evidence that it is a good description of the PDF. The cluster data test the positive tail of the PDF, not its shape near the peak, which can only be probed via statistics sensitive to regions of average density. The clustering of galaxies in the mildly nonlinear regime, in which perturbation theory (PT) is valid, may allow such a probe. Frieman \& Gazta\~naga (1999) have applied PT to the angular 3-point correlation function, and have compared their predictions to the APM data. They are able to rule out a strongly nongaussian PDF, the Peebles $\chi^2$ density field (cf. \S~\ref{sec:intro}), which has $T=16.3,$ via this approach. They did not, however, consider the milder levels of nongaussianity ($T \approx 3$--4) indicated by the cluster data for $\Omega_{\rm M}\simeq 0.3.$ Inspection of Figure~\ref{fig:phi_psi_1} shows that at such levels ($\lambda \sim 5$) the PDF is close to Gaussian near its peak; it may be quite difficult for the approach of Frieman \& Gazta\~naga to constrain nongaussianity at this level. Another test of the PDF will be provided by CMB anisotropy data. As mentioned at the outset of the paper, analyses of the COBE data have unearthed indications of nongaussianity. However, those results correspond to a comoving scale of $\sim\,$ 1000\ifmmode h^{-1}\else$h^{-1}$\fi\ Mpc, much larger than the $\sim\,$ 10\ifmmode h^{-1}\else$h^{-1}$\fi\ Mpc cluster scales probed by clusters. Future CMB measurements sensitive to $\lower.5ex\hbox{\ltsima} 10'$ scale anisotropies will probe cluster mass scales directly. It is not yet clear, however, that the signal-to-noise ratio per resolution element will be sufficient to detect mild nongaussianity. It may well be that for the forseeable future the cluster data sets will provide the most sensitive tests of PDF nongaussianity. \subsection{What does it mean?} If future tests confirm that the initial density field is moderately nongaussian, what would this tell us about the early universe? It was argued above that this would not in and of itself lend credence to any particular inflationary model. Rather, the basic significance of such a finding would be that the fundamental mechanism leading to real-space Gaussianity of the initial fluctuations---their origin as superpositions of numerous, statistically indpendent individual Fourier modes (equation~1)---is not fully operative. This could be because either (i) only a small number of Fourier modes contribute effectively to $\delta({\bf x}),$ and the mode amplitudes are themselves nongaussian; (ii) the various Fourier modes are not statistically independent; or (iii) some combination of (i) and (ii). It is the second of these effects which is responsible for nongaussianity in Peebles' (1999a,b) model, but the first cannot be excluded. Only as observational constraints on nongaussianity improve will be able to construct realistic models of how it arises. This may lead us to a deeper understanding of the fluctuations themselves. Indeed, perhaps the greatest lesson to be learned from this paper and the literature cited herein is that we still know relatively little about the true origin and character of the primordial density fluctuations. \subsection{Reasons for Caution} As seen in Figures~9 and~10, the PS-predicted comoving density $N(\ge\!M,z)$ is a rapidly decreasing function of mass for $M\lower.5ex\hbox{\gtsima} 10^{15}\,\h1M_\odot.$ Consequently, the predicted number of MS1054-like clusters is extraordinarily sensitive to the mass we assign to MS1054. We have selected this cluster because three high-quality data sets, LK97, D98, and T99, enable us to estimate its mass by three independent methods. However, Figure~\ref{fig:ms1054_mass} shows that the velocity dispersion mass estimate for MS1054 is about 50\% below the X-ray and lensing estimates. This discrepancy is within the observational uncertainties and thus does not suggest the presence of major systematic errors. On the other hand, if the velocity dispersion mass estimate of T99 is correct, our conclusions would be markedly changed. We would in that case find, using the same criteria as above, that Gaussian fluctuations are allowed for $\Omega_{\rm M}\le 0.33$ as compared with $\Omega_{\rm M}\leq 0.17,$ and that for $\Omega_{\rm M}=0.45$ the 90\% constraint is $\lambda\le 25$ ($T\ge 2$) as compared with $\lambda \le 1$ ($T\ge 7$). In short, our conclusion that substantial nongaussianity is indicated by MS1054 is correct only if the X-ray and lensing mass estimates are closer to the truth than the dynamical one. This extreme sensitivity to mass is both the strength and weakness of the PS approach. It is a strength because it enables even one high-redshift, high-mass cluster such as MS1054 to powerfully constrain cosmology. It is a weakness because it means that even modest random errors in mass estimation drastically affect our quantitative conclusions. To best utilize the PS approach as a cosmological probe, we will need much larger high-redshift ($z\lower.5ex\hbox{\gtsima} 0.6$) cluster samples than are presently available. Each sample cluster will need to have its mass as accurately and robustly measured as MS1054. Only then will the effect of mass errors be reduced, by $\sqrt{N}$ statistics, to levels at which nongaussianity can be constrained (as a function of $\Omega_{\rm M}$) with high confidence. The EMSS cluster sample has been the major source of known intermediate and high redshift clusters to date, but its sky coverage is too small at the faintest X-ray flux levels to suffice for future work. Future distant cluster samples will most likely be derived from deep optical surveys to which automated cluster-finding algorithms (e.g., Postman {\sl et al.}\ 1996; Kepner {\sl et al.}\ 1998) are applied, with cluster candidates followed up with spectroscopy at 8--10 m class telescopes, X-ray satellite observations, and ground- or space-based deep imaging, to obtain dynamical, X-ray temperature, and weak lensing mass estimates respectively. The second reason for caution is our reliance on the PS formalism, whose validity at high masses is not yet confirmed. It has long been known from N-body simulations that the PS formula overestimates the number of collapsed objects at low masses, $M \ll M_*,$ where $M_*$ is the ``nonlinear mass'' at a given epoch defined by $\sigma_M(M_*)=\delta_c(z).$ At masses $M\lower.5ex\hbox{\gtsima} M_*,$ however, N-body studies have found PS abundance predictions to be remarkably accurate (see, e.g., Borgani {\sl et al.}\ 1999 and references therein). The nonlinear mass at the present time is $\sim 5\times 10^{13}\,\ifmmode h^{-1}\else$h^{-1}$\fi M_\odot,$ so that rich clusters are safely above $M_*$ and thus expected to be well-described by the PS formalism. However, the PS formula has not been exhaustively tested in the very high-mass ($M\lower.5ex\hbox{\gtsima} 100 M_*$) regime, for the simple reason that most N-body simulations contain very few objects in this mass range. With the recent advent of extremely large simulations, such tests have become possible. In a set of simulations of cubic volumes $\sim\,$ 500\ifmmode h^{-1}\else$h^{-1}$\fi\ Mpc on a side, each containing hundreds of rich clusters, Governato {\sl et al.}\ (1998) were able to test the accuracy of the PS abundance formula for virial masses up to $\sim 3\times 10^{15}\,\h1M_\odot.$ They found excellent agreement between the predicted and observed number of clusters in simulations of open, $\Omega_{\rm M}=0.3$--0.4 universes. However, in their $\Omega_{\rm M}=1$ simulation with a low present-day normalization ($\sigma_8\lower.5ex\hbox{\ltsima} 0.5$), they found that the PS formula underpredicted the number of clusters by a factor of $\sim\,$ 3--10 for masses greater than $\sim 10^{15}\,\h1M_\odot.$ The case for which Governato {\sl et al.}\ found PS to be inaccurate, $\Omega_{\rm M}=1,$ is not one that is relevant to this paper (we considered only $\Omega_{\rm M}\leq 0.5$), and for low-density universes (the case of interest here) Governato {\sl et al.}\ confirmed the PS abundance predictions. Still, these results are cause for concern, because if PS underpredicts abundances, we may be led to spurious evidence for nongaussianty. On balance, then, while the present evidence from N-body simulations favors continued use of the PS formalism for cluster analysis, the Governato {\sl et al.}\ findings suggest that continued testing with larger simulations is needed. \subsection{Conclusion} We have used the generalized PS formalism to calculate the likelihood that a cluster as massive as MS1054--03, at a redshift of $z=0.83,$ would have been found in the EMSS sample. The calculations assumed a flat universe, $\Omega_{\rm M}+\Omega_\Lambda=1,$ and mass fluctuations $\sigma_M$ normalized to the observed abundance of low-redshift ($z\leq 0.1$) clusters. The expected number of MS1054-like clusters then depends on $\Omega_{\rm M}$ and the deviations from Gaussianity of the PDF, $P(\delta|M),$ on cluster mass scales. We characterized departures from Gaussianity by assuming $P(\delta|M)=\sigma_M^{-1}\psi_\lambda(\delta/M),$ where $\psi_\lambda(x)$ is defined by equation~16. The parameter $\lambda$ quantifies the degree of nongaussianity; $\psi_\lambda$ approaches Gaussianity for $\lambda \gg 1.$ Special attention has been given to the problem of estimating cluster virial masses from galaxy velocity, X-ray temperature, and weak lensing data. Such estimates are dependent on the assumed density profile of the cluster, for which we have adopted the NFW form, which has been shown in N-body simulations to be more realistic than any pure power law. We elected to work with MS1054 alone because quality X-ray temperature, galaxy velocity, and lensing data have been obtained for it (D98, T99, LK97), making it a uniquely well-studied high-redshift cluster at this time. The X-ray and lensing mass estimates are in excellent agreement; the dynamical mass is $\sim\,$ 50\% smaller than the other two, but this is within the observational errors. If the initial density fluctuations are Gaussian, it is improbable that a cluster of the mass, redshift, and X-ray flux of MS1054 would be found in the EMSS if $\Omega_{\rm M}\ge 0.2.$ For example, the chances of finding an MS1054-like cluster, for a Gaussian PDF, in the EMSS search volume is less than about one in fifty if $\Omega_{\rm M}=0.25,$ and less than one in three hundred if $\Omega_{\rm M}=0.35,$ as shown in Figure~9. If the PDF is nongaussian, however, the likelihood can be greatly enhanced. To constrain the required amount of nongaussianity, we have determined the value of $\lambda$ required for an MS1054-like cluster to have a one in ten chance of being found in the EMSS. The results are shown in Figure~\ref{fig:lambda_T} as a function of $\Omega_{\rm M}.$ The maximum allowed value of $\lambda$ estimated using this criterion decreases from $\sim 25$ for $\Omega_{\rm M}=0.25$ to $\sim 1$ for $\Omega_{\rm M}=0.45.$ These results may also be expressed in terms of the parameter $T$ (equation~20), which measures the likelihood of $\ge 3\sigma$ peaks in the density field relative to the Gaussian case. We find $T\ge 2$ for $\Omega_{\rm M}=0.25,$ increasing to $T\lower.5ex\hbox{\gtsima} 6$ for $\Omega_{\rm M}\lower.5ex\hbox{\gtsima} 0.4.$ In short, for any value of $\Omega_{\rm M}\lower.5ex\hbox{\gtsima} 0.25$ the initial density field must be significantly nongaussian, with the extent of nongaussianity increasing rapidly with increasing $\Omega_{\rm M}.$ Because our analysis has considered only the predicted cluster {\em abundance,} it does not exclude Gaussian fluctuations if we are willing to accept $\Omega_{\rm M} < 0.2.$ However, the case for a nongaussian PDF irrespective of $\Omega_{\rm M}$ is strengthened if our results are taken in conjunction with those of RGS98 and KST99, who analyzed cluster correlation as well as abundance data. In particular, KST99 found that nongaussianity was required, at the level $T\simeq 2$--6, for {\em all\/} values of $\Omega_{\rm M}$ less than 0.5, and ruled out larger values of $\Omega_{\rm M}.$ Our estimate $T\ge 3$ for $\Omega_{\rm M}\ge 0.3$ is consistent with their results. If one takes the findings of RGS98, KST99, and this paper at face value, the case for nongaussian initial density fluctuations is strong indeed. We have, however, identified two potential weaknesses in our anlaysis, and by extension with any attempt to constrain cosmological parameters from cluster abundance data. First, the predicted abundances drop precipitously with increasing cluster virial mass at the high masses ($\lower.5ex\hbox{\gtsima} 10^{15}\,\h1M_\odot$) and redshifts ($\lower.5ex\hbox{\gtsima} 0.5$) of interest. This sensitivity translates into large errors in the derived cosmological parameters for even modest ($\sim 30\%$) errors in virial mass. This fundamental problem can only be remedied by much larger catalogs of high-redshift massive clusters; at present, MS1054 is one of but a handful of such objects known. These clusters, once identified, will need to be followed up with X-ray, galaxy velocity, and weak lensing measurements to ensure reliable mass estimates. Efforts are presently under way by this author and collaborators, as well as other groups, to obtain such data sets. In 5--10 years we will undoubtedly know much more than we do now about the evolution of the cluster abundance at $z \sim 1.$ The second problem concerns the validity of the Press-Schechter formalism in the high-mass regime. It has become conventional wisdom in recent years that the PS abundance formula ``works much better than it should,'' based as it is on a simple, spherical collapse model, and as a result it has been widely and fruitfully used. However, if the PS formula underpredicts the actual abundance of high-mass objects, as at least one study suggests (Governato {\sl et al.}\ 1998), one would underestimate $\Omega_{\rm M}$ (if Gaussian fluctuations are assumed) or overestimate departures from Gaussianity by applying the PS formula to cluster abundance data. A challenge for theory and numerical simulations in the coming years is to rigorously test the PS formula at high masses and, if necessary, replace it with a more accurate semianalytical framework. Such theoretical groundwork will be crucially important if the high-quality cluster data sets of the coming decade are to be fully exploited. \acknowledgements The author gratefully acknowledges the support of NSF grant AST-9617188, the Research Corporation, and a Frederick Terman Fellowship from Stanford University. The author thanks Nikhil Padmanabhan and Puneet Batra for assistance in developing the Press-Schechter computer codes, and Nikhil Padmanabhan, Puneet Batra, Keith Thompson, and Sarah Church for useful discussions. Finally, the author would like to thank Joanne Cohn for bringing a number of papers on primordial nongaussianity to his attention.
\subsection*{Abstract} We have detected extraplanar cold dust at distances out to $> 10\,$kpc, situated in the halo of the interacting galaxy NGC\,4631. The dust emission disk is much thinner than the warped H{\sc i} disk and new structures emerge. In particular, a giant arc has been found that is linked to anomalies in the kinematical structure of the atomic gas. Most of the extraplanar dust is closely associated with H{\sc i} spurs that have been found earlier [Weliachew, L., Sancisi, R. \& Gu\'elin, M. (1978) {\it Astron. Astrophys.} 65, 37-45; Rand, R.J. (1994) {\it Astron. Astrophys.} 285, 833-856]. These spurs obviously are traces of the interaction [Combes, F. (1978) {\it Astron. Astrophys.} 65, 47-55]. The dust emission within the plane reaches the border of the optical disk. The activity of the disk of NGC\,4631 is moderately enhanced by the interaction, but no gas moving in the $z$-direction could be found [Rand, R.J., Kulkarni, S.R. \& Hester, J.J. (1992) {\it Astrophys. J.} 396, 97-103; Golla, G., Dettmar, R.-J. \& Domg\"orgen, H. (1996) {\it Astron. Astrophys.} 313, 439-447]. Hence, it seems unlikely that strong winds have deposited the high-$z$ dust. Instead, the coincidence with the H{\sc i} features suggests that we see a track left behind by the interaction. In addition, the H{\sc i} shows a supershell formed by an impact [Rand, R.J. \& Stone, J.M. (1996) {\it Astron. J.} 111, 190-196] in the zone where the dust trail crosses the disk. This region is also characterized by disturbances in the distribution of the H$\alpha$ light. The masses associated with the dust can be estimated only very roughly on the basis of the existing data; they are of the order of a few $10^9$ M$_{\odot}$ of gas. \section{Introduction} Cold dust has come into focus only recently because it had to await the development of sensitive millimeter/sub-millimeter bolometer arrays to be detectable unambiguously. The IRAS survey could provide only hints at its existence because it was blinded by the strong emission from the small percentage of warmer dust that is radiating far more brightly. The large amount of cold dust ($T_{d}\le 25$ K) can be detected only at (sub)mm wavelengths where the radiation of the warmer components has vanished. To give an example: The peak brightness of a blackbody at 30\,K is 30$\times$ higher than that of a 15-K object, all other parameters being equal. On the other hand, the radiation of a blackbody at 15\,K peaks at $\sim200\,\mu$m and remains more than one order of magnitude brighter at $\lambda$ 1.2\,mm than that of a blackbody at 30\,K with the same peak brightness. Now, the emissivity of interstellar dust is roughly proportional to $T_{d}^{6}$ -- for a blackbody $B(T)=\sigma\,T^4$ -- so even a very large amount of cold dust emits only weakly. Because of this $T^{6}$ dependence of the emission, a very large energy input is needed to heat dust, and the majority of it remains at lower temperatures. This cold component thus is an important tracer, and, indeed, it may represent $> 90$\% of the interstellar dust (cf.\ refs.\ 1 and 2). In itself, the contribution of the dust to the total mass of a galaxy is $< 1$\% of the gas mass, but there are indications that the dust-to-gas ratio is relatively constant, independent of the type of gas (atomic or molecular). Indeed, the dust grains are believed to play a crucial r\^ole in the formation of molecules. The study of their properties thus also offers an independent means of studying the molecular gas content of galaxies. This is important because the standard practice of observing the CO molecule and deriving, thereby, the properties of the H$_{2}$ has substantial uncertainties particularly concerning the derived masses. On the other hand, investigating the cold dust is technically difficult and cannot provide any information about the kinematics because it is based on broadband continuum observations. In several runs, the IRAM (Institute for Radio Astronomy in the Millimeter domain) 30-m telescope equipped with MPIfR (Max-Planck-Institute for Radioastronomy) bolometer arrays has been used to map nearby galaxies in the $\lambda$ 1.2-mm continuum emission. The first maps led to the impression that it is well correlated with the CO emission, and drops off similarly steeply with increasing distance from the center. This behavior was shown, for example, for galaxies NGC\,891 (3), M\,51 (4), and NGC\,4631 (5). It soon became evident, however, that this is not generally the case. The Sb galaxy NGC\,4565 is significantly more extended in the emission of $\lambda$ 1.2-mm continuum than in that of the CO line (6). The cold dust is even detected in the warped outermost rim of the disk. As an intermediate case, NGC\,5907 also shows an extended dust disk (7). The sensitivity needed to detect this extended emission has been achieved only recently, however, and the sample is still small. So it is not yet clear what determines the extent of the cold dust -- the profile of the cold dust along the major axis in NGC\,891 remains very close to the rapidly vanishing CO, even when studied with much higher sensitivity than previously published (R. Zylka, personal communication). \section{The observations and the object: NGC 4631} \subsection{Observational details} All recent maps were obtained with bolometer arrays consisting of 19 elements whose sensitivity is in practice about a factor of 2 better than that of the 7-element detector used before. The 19-element bolometer array has a bandwidth of $\sim 80$\,GHz centered at $\sim$ 230\,GHz. The individual elements are arranged in a closely packed hexagonal pattern. The beam size at the 30-m telescope is $11''$ and the spacing between the beams $20''$. The observations were made in March 1997 during a period of stable weather with zenith opacities typically $<0.2$. We monitored the sky opacity before and after each subimage and mapped Mars every night to determine the absolute flux scale. To obtain a map, the object is scanned in azimuthal direction including parts of blank sky on both sides to define a proper zero level. In addition, the sub-reflector of the telescope is oscillating at a frequency of 2\,Hz which makes the beam switch between two positions separated by $45''$ in the orientation of the scanning. This yields an ``on-off'' measurement that cancels atmospheric variations at short time scales. The whole area of NGC\,4631 was covered with a mosaic of 19 individual fields. Their distribution is rather uniform along the whole disk and, hence, the sensitivity drops only at the outer edges of the final map. Furthermore, we kept the scanning orientation close to the minor axis of the galaxy by carefully chosing the hour angles of the individual observations. This minimizes spurious contributions and the noise in the map. The field presented here is about $15'\times 8'$ after the cutoff of the edges with lower sensitivity. In the central part, the noise is $\sim 2$ mJy/beam for the data smoothed to an angular resolution of $20''$ and rises to $\sim 3.5$ mJy/beam at the edges. \subsection{NGC\,4631} It is obviously best to choose edge-on galaxies for studies of weak phenomena, because the lines of sight are long -- remember that the dust emission in the mm regime is optically thin and thus the whole disk contributes to the detectable flux. The 19-element bolometer made it possible to map large areas, so we decided to re-observe NGC\,4631. This moderately active galaxy has long been a favorite candidate for an interacting system. It is relatively nearby, at 7.5 megaparsecs (Mpc) (ref. 8; $1'$ corresponds to $\sim 2$\,kpc), and two obvious companions are close by. The dwarf elliptical NGC\,4627 is situated $3'$ northwest of the nucleus and $30'$ to the southeast the distorted spiral NGC\,4656 can be found. The whole group has been extensively studied in H{\sc i} (8,9) in order to understand the traces of the interaction (cf.\ Fig.~\ref{fig:synops}). According to a modeling of the encounter (10) the prominent streamers of atomic gas can be explained as being pulled out of the members of this group during the interaction. Presumably also as a result of the interaction, the disk of NGC\,4631 has a disturbed appearance in the optical continuum and H$\alpha$ line emission. At two positions in the disk highly energetic ``supershells'' have been found (11) of which one is described as being caused by the impact of a high velocity object (12). Almost simultaneously with the early high-resolution H{\sc i} observations a large radio halo was detected (13). It is of nonthermal origin and one of the most prominent radio halos known. Detailed investigations (14) have subsequently shown that the magnetic field lines of the galactic disk open into the halo -- in stark contrast to most spiral galaxies, where the field is more or less confined in the disk (15). This magnetic field structure allows electrons, cosmic ray particles and hot gas to escape from the active disk into the halo. ROSAT (Roentgen Satellite) detections (16,17) show a large X-ray envelope that is a natural consequence of this configuration. The investigation of the molecular gas indicates rather normal conditions, however. The central region of $\sim 2'\times 1'$ was completely mapped in the (1-0) and (2-1) transitions of CO (18). In addition, a major axis strip of $7'$ length was obtained with a higher sensitivity of $\sim 20$\,mK. Some additional spectra at other locations did not reveal significant emission. \section{The distribution of the cold dust} \subsection{The dusty disk} At a first glance, the $\lambda$ 1.2-mm emission of NGC\,4631 (Fig.\,\ref{fig:map}) is characterized by a narrow, extended disk, with a double-peaked central region. About three times weaker than the brightest peaks, at a level of $\simeq 30$\,mJy/beam, the disk is stretched out to a distance of $\sim 13$\,kpc on either side of the nucleus, gently decreasing in the west, and at a relatively constant level for some 10\,kpc in the east -- see Fig.\,\ref{fig:macut}. This distribution is similar to that found in NGC\,4565 (6): The correlation between dust and CO is restricted to the nuclear region, whereas the dust in the outer parts of the disk seems to follow the H{\sc i}. The CO emission drops in a similarly steep manner as the centimetric radio continuum: At a radius of $2.5'$ ($=5.5\,$kpc) it has decreased to a tenth of the peak value, and a bit further out, no CO emission could be detected at all. At this radius, the dust emission is still at about one-third of the peak level, however, and it actually stretches out at least to the edge of the optical disk. In fact, the limits are set by the border of the map, not by the vanishing emission. The old bolometer map (5) is limited to the innermost part due to its restricted coverage and sensitivity. The narrow main emission ridge is somewhat surprising given the fact that NGC\,4631 does not even show a dust lane. Its thickness is similar to that of the undisturbed edge-on galaxies NGC\,4565 (6) or NGC\,891 (3). This suggests that the dust is concentrated in the midplane of a galaxy in any case (cf. ref.\ 19), and the absence of an optical dust lane may just reflect a high clumpiness with large ``holes''. A closer inspection shows some radial variations in comparison with the H$\alpha$ map (Fig.~\ref{fig:hamap}). Near the center, the optically bright regions surround the dust emission peaks. Here, the bulk of the optical emission is produced in the central region and later absorbed on its way through the disk. In contrast, in the outer parts of the disk several dust maxima coincide with optically bright spots. It is unlikely that all of these places are just accidental line-of-sight coincidences. The source of the radiation could be young stars in their dusty birthplaces (e.g.\ in a spiral arm tangent). In any case, the material along the light path should be rather transparent. Such large variations of the opacity are not surprising for spiral galaxies, however (see ref.\ 20 for a compilation). \subsection{The extraplanar dust} Completely new is the detection of significant dust emission out to $z$-distances of at least 10\,kpc. The distribution of this intergalactic dust strongly suggests that it has been brought there by the same mechanism that formed the four H{\sc i} streamers (8,9). All three H{\sc i} spurs that are touched by our map are connected with the thin optical disk by corresponding dust features (see Fig.~\ref{fig:synops}). The disk formed by the atomic gas is much thicker than the optical or dust emission disk, however, so that most of the $\lambda$ 1.2-]mm emission still lies within their boundaries. On the other hand, CO emission could only be detected out to $z\sim 1\,$kpc. So either there is hidden molecular gas in those outer regions or the dust is associated with atomic gas. Although these streamers follow the already known H{\sc i} features rather closely, north of the center of NGC\,4631 a structure has been unveiled that was invisible in the thick atomic gas disk: A giant arc spans over the central region with its footpoints about 4\,kpc east and west of the central region. The eastern footpoint is situated opposite the onset of the southern streamer 2, and the western part of the arc is blending into spur 4. The thickness of the disk emission seems to be reduced in the central region, but this might be an artifact of the data-reducing technique. The question arises regarding whether there is a common origin for these extraplanar structures. \section{Origin of the extraplanar dust} Unfortunately, we do not have any velocity information from the $\lambda$ 1.2-mm continuum observations, and hence have to rely on indirect arguments for the determination of the history of this dust distribution. To some degree, it seems reasonable to assume that the H{\sc i} velocities are a good indicator also for the dust kinematics. Indeed, there are at least two clear coincidences between H{\sc i} velocity anomalies and the arc. The velocity gradient along the major axis is much steeper in the northern part of the disk than in the southern. A possible interpretation suggests that the disk is somewhat inclined and warped along the line of sight, so we would be looking at different portions of the galaxy (8) -- north of the center at the inner disk with a steeper gradient and south of it at the near part of the outer disk. If we compare the velocity field with the dust emission (see Fig.~\ref{fig:rotbol}), a different explanation arises. In addition to the steeper gradient, the northern disk shows anomalous velocity components at two places -- and they coincide perfectly with the footpoints of the arc. These anomalous components are visible as regions with almost closed isovelocity contours in Fig.~\ref{fig:rotbol} (marked as A1 and A2). Although we are not able to determine the precise location of the extragalactic dust with respect to the disk, it seems clear that there is a local interaction. The additional velocity components point towards us in the west and away from us in the east. In addition there are two structural indications, both of them suggesting that material has followed the arc in an counterclockwise direction and hit the disk in the east: ({\it i}) The appearance of the disk in the light of the H$\alpha$ line is much more disturbed at the eastern footpoint of the arc (cf.\ Fig.~\ref{fig:hamap}: In broadband red light the eastern part seems truncated. The disturbed region lies between the arc and the streamer 2, whereas the western side seems unaffected by the arc or streamers 3 and 4. ({\it ii}) In the H{\sc i} emission of NGC\,4631 two supershells have been found near the midplane and subsequently have been modeled. Shell~2 (west of the nucleus at Right Ascension 12$^h$39$^m$30$^s$) seems to be an expanding bubble, whereas shell~1 (at 12$^h$39$^m$56$^s$) is most probably caused by an impact of a cloud with a mass of $\sim 10^{7} M_{\odot}$ at a speed of 200 km/s coming {\it from the north} (12). The geometry of the collision is very well constrained by its kinematical signature in the H{\sc i} data and indicates that he material not only came from the north but, moreover, must have had a velocity component away from us. The location of this shell is between arc and spur 2. This fits to the other indications of the trajectory of the arc (the shells and the anomalous components mentioned above are distinct features). If we try to describe all extraplanar dust features as a single trail, its path could be as follows: starting from spur 3, it runs through the western disk, follows it to the western footpoint of the arc, sweeps along the arc, penetrates the eastern disk and leaves it via spur 2. In the way, some material is forced away: for example, to follow streamer 4. Of course, such a concatenation is purely speculative and, at this point, merely meant to summarize the structure of the dust distribution. In principle, the concept of a continuous trail is a natural explanation within the framework of an interaction, however. In any case, we have to answer the question regarding which of the three involved galaxies has left the dust traces. The interaction has been modeled rather early on the basis of the first high-resolution H{\sc i} data (10). This description suggests that the bridge 1 and the streamer 4 had been parts of the disk of NGC\,4631. The two perpendicular features 2 and 3 are made from material pulled out of the now dwarf elliptical NGC\,4627 which might have been of a different type before. However, the H{\sc i} emission of structures 2 and 3 is only weakly connected with the disk of NGC\,4631 and not at all with the dwarf. It has to be stressed that this model is by no means unique -- three involved galaxies open a vast parameter space for an interaction scenario. Today, the location of part of the H{\sc i} gas can be determined even in the third dimension with the help of X-ray data: The soft band is very sensitive to absorption by atomic hydrogen, and it can be safely concluded from the distributions that the southern part of the disk is at the near side (16,17). Looking a bit more in detail, the onset of spur 4 seems to be located in front of the X-ray halo (cf.\ Figure 6 of ref.\ 17) but, higher up, the X-ray emission becomes stronger again, suggesting that this streamer is pointing away from us. Such additional information will help to constrain the parameters of the interaction much better than it was possible before. \section{Discussion and outlook} \subsection{Origin of the mm emission} NGC\,4631 is characterized not only by the interaction with two other galaxies but also by a unusually large radio halo. In particular the enhanced star formation might be responsible for the uncommon magnetic field configuration and thus for the radio and X-ray halo -- could this possibly produce extraordinary continuum emission at $\lambda$ 1.2\,mm as well? In addition to radiation from dust, there are three candidates for the source of radiation: free-free emission in ionized clouds, nonthermal emission and line radiation within the bandpass of the detector. The free-free emission should be correlated with H$\alpha$ -- so a contribution is only expected in the disk since the sparse high-$z$ H$\alpha$ emission (21) is not correlated with dust features. Within the disk, even in the brightest spots the emission measure reaches only a relatively low value of 1000\,pc$\cdot$cm$^{-6}$ (22). So the {\it total} flux due to free-free radiation is of the order of 10\,mJy only. The synchrotron emission is very extended around NGC\,4631 at cm wavelengths (14). From the fluxes and the determined spectral index (23) we can estimate its contribution in our mm band. In the center, the spectral index is rather flat at about $\alpha\sim -0.65$. The peak flux here is 60\,mJy/$84''$\,beam at 10.55\,GHz, this translates into 0.3 mJy/$20''$\,beam at 230\,GHz. Outside the disk the spectral index is even steeper, so this contribution is smaller than a tenth of the noise level. The only significant addition thus comes from the molecular line emission: The eastern peak reaches 70 K$\cdot$km\,s$^{-1}$ of $^{12}$CO(2-1) emission. Within our $20''$ beam, this contributes to a flux density of $\sim 20$\,mJy/beam (cf.\ ref.\ 4), about a fifth of the total value. In the outer disk, however, the line emission adds $<3$\,mJy/beam -- just the figure of the noise level in the map. So, the observed mm continuum radiation is pure thermal dust emission almost everywhere, but how did the dust reach such enormous $z$-heights? Could it simply be blown out of the active plane by large-scale winds? Outflows are known for many galaxies with strong starbursts like, for example, M\,82 (see, e.g., refs.\ 24 and 25). But the disk of NGC\,4631 is forming stars at an only moderately enhanced rate (18). H$\alpha$ kinematic data as a more direct tool do not show signs of a gas outflow from the disk (22) -- in contrast to M\,82 which has a similar orientation (24). Peculiar velocities have been found, but they are more likely explained by the direct influence of the interaction. \subsection{Temperatures and Masses} The dust around the disk of NGC\,4631 seems to be cold -- the spatially resolved IRAS CPC data as an indicator for warmer dust show at most a somewhat thicker disk near the center (26). Its width is of the order of $90''$ at $100\mu$m and $45''$ at $50\mu$m wavelength using the published beam sizes of $95''$ and $80''$, respectively. No significant emission can be seen further out. Unfortunately, the CPC instrument could not be calibrated properly, and because of the uncertainties of the order of $\pm60\% $ we could not derive dust temperatures from these data. In view of the moderate star forming activity the presence of very warm material does not seem very likely. Presumably the temperature of the coldest dust component is rather low even in the disk -- similarly to other spirals in which it could be measured so far (2,6). We expect, therefore, temperatures in the range of 15 to 20 K for the disk, and even lower values for the extraplanar dust. To determine such temperatures, observations in the sub-mm range between 1\,mm and 100\,$\mu$m are needed, but, here, the opaque atmosphere renders them very difficult. This is why only few data of spiral galaxies have been published in this range (e.g.\ refs.\ 2 and 27). We will do more observations at $\lambda\lambda$ 450\,$\mu$m and 850\,$\mu$m to complement the existing data. Together, they should provide the spectral information needed for the determination of the temperatures. Until then, we have to postpone the proper determination of masses and energies from the dust emission. In principle, the dust mass is directly proportional to the detected flux $S_{\lambda}$. With some additional assumptions about the dust and gas properties, we can derive the hydrogen column density. In the millimeter/sub-millimeter regime, the observed flux density per beam (of half power beam width $\theta$) produced by dust of temperature $T_{d}$ is given by $S_{\lambda}= 1.13\,\theta^{2}\,(1-e^{\tau_{\lambda}})\,B(\lambda,T_{d})$. Here, $\tau_{\lambda}=\sigma_{\lambda}^{\rm H}\cdot N_{\rm H}$ gives the dust absorption cross section per hydrogen atom, and $B(\lambda,T_{d})$ is the Planck law for the radiation of a black body (see ref.\ 1 for a derivation, and ref.\ 6 for a determination of a possibly typical value). Such a calculation yields about $2\cdot10^{9}$ M$_{\odot}$ of hydrogen for the gas associated with a dust component of 21.5\,K in the central region (5). As already stated above, the warmer dust radiates much more efficiently, and, hence, a second dust component at 55\,K is associated with $<1$\% of this mass (5). These values are poorly constrained, however, and we face large uncertainties in the temperatures of the coldest component. If we nevertheless assume 15\,K for the extraplanar dust the total gas mass in the arc should be of the order of $1.7\cdot10^{9}$ M$_{\odot}$, somewhat less than the mass in the central region. A similar value is found for its presumed continuation, spur 2, so that the total gas mass in this structure would be about half of the atomic hydrogen mass of the whole galaxy (8). \subsection{Conclusion} Obviously, the interaction is the dominant event in the history of NGC\,4631. It has caused a partial disruption of the disk, triggered the star formation activity, provoked an upturned magnetic field configuration, and, last but not least, left trails of atomic gas and dust all around the disk. It is therefore crucial to know exactly how the interaction took place. The existing model (10) was obtained by using a ``trial-and-error'' method for the parameters. It is virtually impossible to find the proper solution that way; the number of possible parameter sets for an interaction in such a group with three members is simply too large. Moreover, the now existing data (including the distribution of the cold dust) give new and more detailed information to check the model against. The problem is how to explore the parameter space to find the best set. Now there are new, powerful techniques available like so-called `genetic codes' (28), so we hope to obtain a much better view of the interaction, which in turn should help to fix the roots of the related phenomena. In any case, the detection of cold dust outside of a galactic disk has unveiled the existence of hitherto invisible baryonic matter in the halo region. This will not solve the dark matter problem, even using a favorable estimation of the associated mass, but it is a step towards a more ``normal'' neighborhood of galaxies -- containing less exotic material, probably simpler to understand, but in any case easier to investigate. \vspace{2em}\noindent {\small {\bf Acknowledgments} We thank G. Golla for the H$\alpha$ picture and the CO data set, R. Rand for several sets of H{\sc i} data, A. Vogler for the ROSAT data and J. Kerp for help with their interpretation. F. Combes supplied the trajectories for her model so that we got an idea of the time evolution. Part of this work was supported by the Deutsche Forschungsgemeinschaft within the frame of SFB301.} \small \noindent \subsubsection*{References} 1. Cox, P. \& Mezger, P.G. (1989) {\it Astron.\ Astrophys. Rev} {\bf 1}, 49-83 \\ 2. Kr\"ugel, E., Siebenmorgen, R., Zota, V. \& Chini, R. (1998), {\it Astron.\ Astrophys.} {\bf 331}, L9-L12 \\ 3. Gu\'elin, M., Zylka, R., Mezger, P.G., Haslam, C.G.T., Kreysa, E., Lemke, R. \& Sievers, A. (1993) {\it Astron.\ Astrophys.} {\bf 279}, L37-L40 \\ 4. Gu\'elin, M., Zylka, R., Mezger, P.G., Haslam, C.G.T. \& Kreysa, E. (1995) {\it Astron.\ Astrophys.} {\bf 298}, L29-L32 \\ 5. Braine, J., Kr\"ugel, E., Sievers, A. \& Wielebinski, R. (1995) {\it Astron.\ Astrophys.} {\bf 295}, L55-L58 \\ 6. Neininger, N., Gu\'elin, M., Garc\'{\i}a-Burillo, S., Zylka, R. \& Wielebinski, R. (1996) {\it Astron.\ Astrophys.} {\bf 310}, 725-736 \\ 7. Dumke, M., Braine, J., Krause, M., Zylka, R., Wielebinski, R. \& Gu\'elin, M. (1997) {\it Astron.\ Astrophys.} {\bf 325}, 124-134 \\ 8. Rand, R.J. (1994) {\it Astron.\ Astrophys.} {\bf 285}, 833-856 \\ 9. Weliachew, L., Sancisi, R. \& Gu\'elin, M. (1978) {\it Astron.\ Astrophys.} {\bf 65}, 37-45 \\ 10. Combes, F. (1978) {\it Astron.\ Astrophys.} {\bf 65}, 47-55 \\ 11. Rand, R.J. \& van der Hulst, J.M. (1993) {\it Astron.\ J.} {\bf 105}, 2098-2106 \\ 12. Rand, R.J. \& Stone, J.M. (1996) {\it Astron.\ J.} {\bf 111}, 190-196 \\ 13. Ekers, R.D. \& Sancisi, R. (1977) {\it Astron.\ Astrophys.} {\bf 54}, 973-974 \\ 14. Golla, G. \& Hummel, E. (1994) {\it Astron.\ Astrophys.} {\bf 284}, 777-792 \\ 15. Dumke, M., Krause, M., Wielebinski, R. \& Klein, U. (1995) {\it Astron.\ Astrophys.} {\bf 302}, 691-703 \\ 16. Wang, Q.D., Walterbos, R.A.M., Steakley, M.F., Norman, C.A. \& Braun, R. (1995) {\it Astrophys.\ J.} {\bf 439}, 176-184 \\ 17. Vogler, A. \& Pietsch, W. (1996) {\it Astron.\ Astrophys.} {\bf 311}, 35-48 \\ 18. Golla, G. \& Wielebinski, R. (1994) {\it Astron.\ Astrophys.} {\bf 286}, 733-747 \\ 19. Xilouris, E.M., Byun, Y.I., Kylafis, N.D., Paleologou, E.V. \& Papamastorakis, J. (1999) {\it Astron.\ Astrophys.}, {\bf 344}, 868-878 \\ 20. Davies, J.I. \& Burstein, D. (eds.), (1995) {\it The Opacity of Spiral Disks} (Kluwer, Dordrecht), Vol.\ 469 \\ 21. Rand, R.J., Kulkarni, S.R. \& Hester, J.J. (1992) {\it Astrophys.\ J.} {\bf 396}, 97-103 \\ 22. Golla, G., Dettmar, R.-J. \& Domg\"orgen H. (1996) {\it Astron.\ Astrophys.} {\bf 313}, 439-447 \\ 23. Golla, G. (1993) {PhD Thesis}, University of Bonn \\ 24. McKeith, C.D., Greve, A., Downes, D. \& Prada, F. (1995) {\it Astron.\ Astrophys.} {\bf 293}, 703-709 \\ 25. Bregman, J.N., Schulman, E. \& Tomisaka K. (1995) {\it Astrophys.\ J.} {\bf 439}, 155-162 \\ 26. van Driel, W., de Graauw, Th., de Jong, T. \& Wesselius, P.R. (1993) {\it Astron.\ Astrophys.\ Sup.} {\bf 101}, 207-252 \\ 27. Alton, P.B., Bianchi, S., Rand, R.J., Xilouris, E.M., Davies, J.I. \& Trewhella, M. {\it Astrophys.\ J.} {\bf 507}, L125-L129 \\ 28. Theis, Ch. (1999) {\it Reviews in Modern Astronomy} {\bf 12} (in press) \normalsize \begin{figure}[ht] \epsfig{bbllx=135,bblly=41,bburx=544,bbury=707,% figure=f_map.ps,height=8.0cm,angle=270} \caption{Map of the $\lambda$ 1.2-mm emission of NGC\,4631, overlaid on an image taken from the Digital Sky Survey. The levels are -6 (dotted), 6, 11, 21, 41, and 81 mJy/beam. Only significant emission is shown and the outer parts of the map with higher noise have been cut off. The small object north of the disk is the dwarf elliptical galaxy NGC\,4627; the other companion, NGC\,4656, is situated about half a degree away in the south-east. } \label{fig:map} \end{figure} \begin{figure}[ht] \psfig{file=f_macut.ps,height=8cm,angle=270} \vspace{2mm} \caption{Cuts along the major axis; the CO and the 2.8\,cm data are provided by G. Golla (14,18) and the H{\sc i} curve is extracted from a map supplied by R. Rand (8). The left axis gives the units for the dust and CO emission, the right axis for H{\sc i} and cm continuum. Similar to the galaxies NGC 4565 (6) and 5907 (7), the profiles of the CO and the cm continuum drop off more steeply than that of the dust and the H{\sc i}. Note, however, that such cuts are one-dimensional and miss emission that is close to, but not on the major axis. } \label{fig:macut} \end{figure} \begin{figure}[ht] \epsfig{bbllx=281,bblly=41,bburx=544,bbury=707,% figure=f_hamap.ps,height=16.0cm,angle=270} \caption{Comparison of the $\lambda$ 1.2-mm emission of NGC\,4631 with the H$\alpha$ emission which has been smoothed to $6''$. The dwarf elliptical galaxy NGC\,4627 is invisible here. Note the large disturbed area in the west with its subsequent `cutoff'. Near the center, the dust and H$\alpha$ tend to be anti-coincident whereas in the outer disk there are clear correspondences. } \label{fig:hamap} \end{figure} \begin{figure}[ht] \psfig{file=f_synops.ps,height=16cm,angle=270} \caption{An overview of the distribution of the extraplanar H{\sc i} (purple contours) and cold dust (in color); the H{\sc i} is smoothed to $22''$ spatial resolution. For comparison, optical isophotes are added in red. The H{\sc i} spurs are numbered as introduced in ref.\ 9. The axis labels are given in arc minutes from the center of NGC\,4631 on the lower and left sides; the other sides gives the projected length scale for an assumed distance of 7.5\,Mpc. } \label{fig:synops} \end{figure} \begin{figure}[ht] \epsfig{bbllx=143,bblly=41,bburx=544,bbury=753,% file=f_rotbol.ps,height=16cm,angle=270} \caption{A comparison between the cold dust emission (at $20''$ resolution) and the velocity field of the atomic gas (at $12''\times 22''$). Velocity contours are plotted every 20\,km/s (cf.\ also the right-hand scale). Note the velocity anomalies at the foot points of the arc (arrows). The insert shows position-velocity curves parallel to the major axis; the box is labelled in velocity vs.\ minutes of arc. The red line goes through A1, the green line through A2 as indicated by short lines in the main plot. In the southern disk the gradient is flatter (purple line). The stars mark the position of the two H{\sc i} supershells. } \label{fig:rotbol} \end{figure} \end{document}
\section{Introduction} Absolute distance measurements in astronomy are extremely valuable and are very difficult to make. The extragalactic distance scale in particular has profound implications for cosmology, but the distance-ladder methods have generated long-standing controversies. Direct methods of measuring cosmological distances, for example via gravitational lensing effects, offer alternatives to the distance ladder. One such method exists through combining X-ray observations of clusters of galaxies with the Sunyaev--Zel'dovich (SZ) effect \cite{sandz}. This method takes advantage of the fact that the SZ effect depends on a different combination of the intrinsic properties of the cluster gas (density, temperature, physical size) than the observed X-ray flux: hence the physical size of the cluster, and via the angular size, its distance, can be measured. To see how the distance depends on the measurable quantities, consider a blob of gas of size $l$, electron number density $n_e$, at temperature $T_e$. The observed X-ray flux density is \[ F_X \propto n_e^2 f(T_e) l^3D_l^{-2}\label{Xbrem}\] where $D_l$ is the luminosity distance and $f(T_e)$ is the temperature dependence of the X-ray emissivity (including Gaunt factor). The SZ effect from the same gas is \[ \Delta T_{\rm SZ} \propto n_e T_e l. \label{SZeffect}\] If the blob subtends an angle $\theta$ then \[ \theta = l/D_a,\] where the angular size distance $D_a = D_l (1+z)^{-2}$. Rearranging, \[ H_0 \propto D_a^{-1} \propto \frac{T_e^2 F_X}{\Delta T^2 f(T_e) \theta}, \label{calcH0} \] where we have dropped the explicit $z$ dependence. $H_0$ can therefore be determined for a cluster of known redshift provided we can measure the gas temperature, the X-ray flux and the SZ effect, and provided we can model the distribution of gas in the cluster. This method (e.g. Cavaliere et al \shortcite{cavaliere}) was suggested shortly after the SZ itself was proposed, and since the first successful detections of the SZ effect in the mid-1980s several authors have published determinations of $H_0$ based on it, using both single-dish (e.g. \cite{birk94,myers}) and interferometric (e.g. \cite{jones-capri,reese}) SZ measurements. Despite the advantages of the SZ method in by-passing the entire distance ladder, it does have its own sources of error, which must be carefully addressed. These include: \begin{enumerate} \item{ Beam switching that does not go fully outside the cluster; contamination by radio sources in the cluster or (for single-dish SZ measurements) in the reference beams; resolving out most of the total flux of the cluster (for interferometric measurements); poor spatial sampling of the cluster.} \item{Errors in the temperature measurement of the cluster, due to poor photon statistics, temperature substructure and clumping of the cluster gas} \item{Errors due to fitting a simplified smooth model of the gas distribution to a real complicated cluster} \item{Biases due to selection effects in the chosen cluster sample.} \end{enumerate} In this paper we describe our analysis techniques for determining $H_0$ by combining SZ observations from the Ryle Telescope~(RT) with X-ray data from {\sc ROSAT} and {\sc ASCA}. We apply this to data for the cluster A1413, for which we have previously published SZ results \cite{grainge96}. Finally, we consider the various sources of error, both random and systematic, which affect our estimate. \section{Cluster modelling} \label{model} We use our data to measure the Hubble constant by creating a numerical model of the temperature and density distributions in a cluster deduced from X-ray data, and then using these to generate mock RT data that can then be compared with real observations. The gas density is modelled as a King profile in one or three dimensions (depending on whether the cluster shows significant asphericity), and arbitrary temperature profiles can be specified. The cosmological parameters $H_0$, $\Omega_M$ and $\Omega_\Lambda$, upon which the relative scaling of the X-ray and SZ data depend, can be adjusted until both sets of mock data agree with the observations. The method is essentially similar to those used in other SZ measurements of $H_0$, especially other interferometric measurements \cite{reese}. Unlike Reese et al we do not use the SZ data to constrain the shape parameters, as the X-ray image contains far more information and would dominate any joint fit; we do however allow for a non-spherical cluster. \subsection{Physical model} Our modelling is based on the analytical King approximation~\cite{CF76} for the potential due to a self-gravitating isothermal sphere, which gives results consistent with X-ray images, at least in the inner parts of the gas distribution. The cluster total density $\rho$ at radius $r$ from the cluster centre is given by \[ \rho(r)=\rho_0 \left(1+\left({r/r_c}\right)^2 \right)^{-3/2} , \] where $r_c$is the core radius of the cluster. Assuming that the gas density,~$\rho_g$ is related to the total mass density by $\rho_g \propto \rho^{\beta}$, then \begin{equation} \rho_g(r)=\rho_{g0} \left(1+\left({r/r_c}\right)^2 \right)^{-3\beta/2} . \label{kingden} \end{equation} The average value of $\beta$ determined by fits to X-ray surface brightness of a large number of clusters is found to be $<\beta_{fit}>=0.67$ \cite{JF84}, but we fit for $\beta$ separately for each cluster. Although we typically use isothermal models for fitting X-ray and SZ data, we can vary the temperature to investigate specific effects. We do not vary the X-ray emissivity constant in these models; however, for typical rich cluster temperatures (7--10 keV) this constant is a weak function of temperature when observing in the {\sl ROSAT} 0.5--2~keV band. \subsection{Simulating observations} \label{simmodel} We generate two-dimensional images, of typically $256 \times 256$ pixels, of the model skies in SZ and in X-rays. This results in a spatial resolution in the model of $\sim 50$~kpc and a total simulation size of $\sim 10$~Mpc -- numerical experiments show the results to be insensitive to these parameters. We can use either a spherical King model (equation \ref{kingden}) for the density, or a 3-d modification of it in which core radii ($r_{cx}, r_{cy}, r_{cz}$) are specified in three perpendicular directions, the directions relative to the plane of the sky being specified by three Euler angles. The density at a position ($x, y, z$) in these coordinates is thus given by \[ \rho = \rho_0(1+x^2/r_{cx}^2+y^2/r_{cy}^2+z^2/r_{cz}^2)^{-3\beta/2}\] We integrate the functions shown below along the rows of a three-dimensional array to generate the SZ brightness temperature and X-ray images: \begin{equation} \Delta T(i,j) = \sum_k n(i,j,k) T(i,j,k){{g(x)}\over{x^2}} {{k_B T_0 \sigma_T}\over{m_e c^2}} L , \label{szeffect} \end{equation} \begin{equation*} g(x) = {{x^4 e^x}\over{(e^x-1)^2}} \left( x \coth{x\over2}-4 \right), \end{equation*} \begin{equation*} x = {{h \nu}\over{k_B T_0}} , \end{equation*} \begin{equation} {\rm X}(i,j) = \sum_k K n^2(i,j,k) t_{exp} L^3 , \end{equation} where $T_0$ is the CMB temperature, $t_{exp}$ is the exposure time of the X-ray image and $K$ is a telescope, redshift and cluster-temperature dependent constant determining the received count rate, incorporating the bremsstrahlung and line emissivities, hydrogen absorption, detector bandpass and K-correction, expressed as {\sl ROSAT} PSPC counts per second due to a volume of $1 \rm m^3$ of gas, of electron density $1 \rm m^{-3}$, at a luminosity distance of $1$ Mpc. $L$ is the size in metres corresponding to the angular extent of each pixel $\theta_{cell}$; this is calculated using the cluster redshift, $z$, and assumed values of $H_0$, $\Omega_M$ and $\Omega_\Lambda$. We have checked that we obtain results consistent with the analytic expressions for both the X-ray and SZ profiles for the spherical isothermal King model. We choose not to use the analytic expressions in our code since the approach described above can more easily be modified to allow different distributions of gas density and temperature. For example, we have been able to investigate the effect of non-isothermal temperature profiles, and also central cooling flow regions. The latter have no effect on the SZ signal, however, since cooling flows are considered to be in pressure equilibrium with the rest of the cluster. The brightness temperature decrement map is converted to flux density, multiplied by the RT primary beam response, and Fast Fourier Transformed to obtain the simulated aperture-plane response of the RT to the SZ decrement. The best-fit cluster parameters found for the cluster A1413 (see section~\ref{a1413xfit}) give a profile in the aperture plane for different observing baselines shown in Figure~\ref{pmax}. Mock data that correspond to the real RT observations can then be generated by sampling this array at the observed coordinates in the aperture plane. We also simulate the process of observing the sky with the {\sl ROSAT} PSPC or HRI. The model X-ray sky is convolved with the X-ray telescope point-spread function, and a suitable background level added. For making model fits a smooth background is used; we can also make images in which each pixel value (signal plus background) is replaced with a sample from a Poisson distribution of the same mean, thus generating realistic X-ray images. We tested the program to ensure that it gave the expected dependence upon central number density, $\beta$, core radius, gas temperature and value for $H_0$ and to check that the results for the simulations were not dependent upon pixel size or overall array size. \subsection{Predictions of SZ observations}\label{predict} We have predicted the effect that changes in various cluster parameters would have on our RT observations. In particular we were concerned about the behaviour of the flux density measured on the shortest RT baselines, since these are the most sensitive to the SZ effect. (The shortest physical RT baseline is $870~\lambda$, but at for lower declinations this can be projected down to a minimum of the antenna diameter of $640~\lambda$.) The results here were obtained using the model parameters for Abell~1413, but are typical of all the rich clusters we observe. We simulated a cluster with using the best-fit parameters that we have found for A1413 (see section~\ref{a1413xfit}) but varied the core radius. We found that the central decrement is directly proportional to the core radius, as expected, and that the flux density measured on the 870~$\lambda$ baseline quickly began to flatten off (see Figure \ref{flxvrad}) as the RT starts to resolve out the SZ signal. This resolving out starts to become significant for core radii of greater than about 70~arcseconds for the 870~$\lambda$ baselines. In a further test we varied the central number density, $n_0$, for various core radii to give the same observed flux density using a $870 \lambda$ baseline, and recorded the central temperature decrement that these models would give. The results are also shown in Figure~\ref{flxvrad}. The central decrement does not vary greatly with core radius, especially around 60~arcseconds which is the expected size for a cluster with core radius of about 200~kpc at moderate redshift. We have also assessed the effects of redshift and cosmology on the SZ signal seen. We simulated observations of clusters with the same intrinsic parameters lying at different redshifts. First we calculated the clusters' angular core radii, corresponding to a constant physical radius of 250~kpc, assuming Einstein-de Sitter ($\Omega = 1.0$) and Milne ($\Omega = 0.0$) cosmologies. The results for observations with baselines of 870 and 1740~$\lambda$ are shown in Figure~\ref{flx900}. It can be seen that the RT has similar sensitivity to a cluster at redshifts between 0.2 and 10 on the shortest baseline in both cosmologies. On the 1740~$\lambda$ baseline the signal peaks for clusters at redshifts between 0.3 and 4 in an Einstein-de Sitter cosmology, but tends to a maximum value in a Milne universe beyond a redshift of 1. We therefore conclude that observations of the SZ effect with the RT should be possible for clusters at any redshift from $z=0.1$ to $z=10$. Although the SZ effect predicted to be detected by the RT from a cluster of given physical size is largely independent of cosmology, the value of $H_0$ that we will calculate from observations will be affected by the values of $\Omega_M$ and $\Omega_\Lambda$ that we adopt. The true value of $H_0$ and the calculated value $H_0^{\rm calc}$ will be related by \[ H_0 = {{H_0^{\rm calc} D_a^{\rm calc}}\over{D_{a}}} \] where $D_a^{ \rm calc}$ and $D_{a}$ are the calculated and true angular distances to the cluster respectively. The correction factor $H_0^{\rm calc} / H_0$ is shown in Figure~\ref{q0corr}. \subsection{Fitting to X-ray images} \label{x-fitting} Since {\sl ROSAT} X-ray images contain far more information about the gas distribution in a cluster than RT SZ data, we use the X-ray image to model the density distribution. We assume the cluster is a triaxial ellipsoid with two principal axes in the plane of the sky, and with the core radius along the line of sight equal to the geometric mean of the other two core radii. We then use a downhill simplex fitting method, implemented by the routine {\sc amoeba}~\cite{numrep}, to optimise the central density, core radii, ellipsoid centre, position angle and $\beta$ parameter. The likelihoods of the data given the model are calculated as the product of the Poisson likelihoods of obtaining the observed count at each pixel in the X-ray image, given the parameters. It is necessary to use the Poisson statistic rather than $\chi^2$ in order to avoid biases in the regions of low count rate where the two statistics differ significantly. This method also allows arbitrary regions to be excluded from the fit; this is used for example to exclude point sources in the X-ray image. We use vignetting-corrected X-ray images (i.e. divided by the image of effective exposure time) for the fitting. Although this results in a distortion of the photon statistics across the map (partly due to the presence of the non-vignetted instrumental background), this effect is small in the central part of the image where all the cluster signal lies. We also use images that are binned so as to under-sample the PSF (pixel size equal to FWHM of PSF) so that the position dependence of the PSF is negligible over the region of interest. There is a degeneracy between $\beta$ and $\theta_c$, and $n_0$, in that almost equally good fits can be obtained by changing the value of $\beta$, and re-fitting for $\theta_c$ and $n_0$. However, although the resulting value of the central temperature decrement is changed for these different fits, the flux density observed by the interferometer is almost unchanged. This is because the observed flux density depends on both the central decrement and the shape of the cluster, and on our most sensitive baselines the two effects almost cancel. The effect of this can be seen in Figure~\ref{degenerate}, which shows contours of likelihood of the X-ray fit in the $\beta$--$\theta_c$ plane overlaid with contours of the predicted flux density observed by the RT. The flux density contours (which are used to calculate $H_0$) lie parallel to the $\beta$--$\theta_c$ degeneracy, which thus has little effect on the value of $H_0$ found. Reese et al.~\shortcite{reese} find a similar result for the OVRO/BIMA arrays. \subsection{Fitting to SZ data} Interferometers measure the Fourier transform of the sky brightness distribution at discrete points defined by the orientation of the baselines, with independent noise values at each point. Inverting the data to form an image results in both signal and noise being convolved with the Fourier transform of the sampling function (the synthesised beam), resulting in long-range correlations across the map. With well-filled apertures and high signal-to-noise, this is not a problem. Since the RT has few baselines on which the SZ effect is detectable, these correlations are significant, ie the synthesised beam has large far-out sidelobes. It is therefore very difficult to do meaningful fitting to the data in the image plane. Accordingly, we fit our data to models in the aperture plane, although we also make maps, particularly in order to identify positive sources in the field. Once a fit has been made to the X-ray image using an assumed value of $H_0$, an SZ sky model and simulated RT aperture are generated as described above. Variations in $H_0$ result in a simple scaling of the amplitude of the SZ signal. The calibrated, source-subtracted RT visibility data are read in and a $\chi^2$ statistic calculated between the real and mock data as a function of amplitude scaling; we multiply by a prior that is uniform in log space (since $H_0$ is a scale parameter) to find the peak and extent of the posterior probability distribution. \section{Calculating $H_0$ from A1413} \label{calchub1413} A1413, at a redshift $z=0.143$, is a massive and highly luminous cluster with a {\sl ROSAT}-band luminosity of $2.0 \times 10^{38}$ W \cite{allen95}. We have previously published SZ data on the cluster Abell~1413 \cite{grainge96} but did not derive an $H_0$ value there due to the significant ellipticity of the cluster and the presence of a moderate cooling flow, both of which complicate the process. Here we calculate $H_0$ from A1413 taking both these effects into consideration. As well as our RT SZ data, we use an observation made with the {\sl ROSAT} PSPC, and a temperature derived from {\sl ROSAT} HRI and {\sl ASCA} data. \subsection{X-ray image fitting} \label{a1413xfit} We fitted an ellipsoidal model (as described in section \ref{x-fitting}) to the {\sl ROSAT} PSPC image observed on 1991 November 27, which has an effective exposure of 7696~s. We use only the hard (0.5--2~keV) data in order to minimise the effect of Galactic absorption; the image is shown in Figure~\ref{A1413-0-1}, along with the RT SZ image. Allen et al.~\shortcite{allen95} find that A1413 has a cooling flow with a mass deposition rate of 200 M$_{\odot} \rm yr^{-1}$ within a radius of 200 kpc of the centre, and that the temperature in this region is depressed by some 20\% from the value in the outer regions. Therefore we excluded the X-ray data within the cooling flow radius from our analysis, along with several X-ray point sources which appear in the image. Keeping the cooling flow region in the fit would result in a more compact model (higher $\beta$, smaller core radius) that would not reflect the true distribution of the gas, most of which is outside the cooling flow. Allen and Fabian \shortcite{allen98} use {\sl ASCA} and {\sl ROSAT} HRI data to determine the temperature of A1413 in two models; one assuming the gas is isothermal, and one also fitting for a cool component, with the fraction of the total emission from the cool gas normalised by the fitted mass deposition rate of the cooling flow. These methods yield $7.5^{+0.3}_{-0.3}$ keV and $8.5^{+1.3}_{-0.8}$~keV respectively (90\% confidence errors). The latter value is a better estimate of the temperature of the bulk of the cluster gas, outside the cooling flow region. Since we have excluded the cooling flow region from our fit to the surface brightness image, we use this higher temperature as the more consistent model. The value of $H_0$ that would be obtained from using the cooler temperature can be found from the scaling given in section \ref{theanswer}. The X-ray emissivity constant was calculated assuming an absorbing H column of $1.62 \times 10^{24}$ m$^{-2}$, a metallicity of 0.3 solar and a temperature of 8.5~keV, giving a value of $3.41\pm0.24 \times 10^{-69}$ $\rm counts~s^{-1}$ from 1~m$^3$ of gas of electron density 1~m$^{-3}$ at a luminosity distance of 1~Mpc. The best-fitting model parameters were $\beta = 0.57$, core radii of~$53''$ and $35''$ with a position angle of the major axis of $0.1^{\circ}$ and a central density $n_0 = 1.51 \times 10^4 \, h^{1/2} \rm m^{-3}$ (assuming a core radius in the line of sight of $43=(53\times35)^{1/2}$ arcsec), where $H_0 = 100h \, \rm km^{-1}Mpc^{-1}$. These parameters agree well with the fit by Allen et al \shortcite{allen95}. The {\sl ROSAT} image, our X-ray model and a map of the residuals are shown in Figure~\ref{residuals}. The residual map has zero mean in the regions included in the fitting, although the excess flux in the cooling-flow region and a point source stand out. The Poisson noise also rises towards the centre of the cluster as the square root of the signal level. The reduced value of $\chi^2$ for this fit is 1.02; the mean log likelihood over the region of the fit is $L=-1.585$, and the expected distribution of $L$, as discussed in Grainge et al.~\shortcite{G01}, is $L= -1.585\pm0.021$. We therefore conclude that the fit is a good one. There is a degeneracy between $\beta$ and the core radii (see section~\ref{x-fitting}) which lies parallel to lines of constant predicted flux density observed by the RT. Figure~\ref{degenerate} shows that acceptable fits to the X-ray map correspond to uncertainties in the predicted SZ flux density of $\pm 14 \, \mu$Jy and $\pm 24 \, \mu$Jy at 67\% and 95\% confidence. Therefore we estimate that the error in $H_0$ that arises solely from our X-ray fitting procedure is $2\times {{14}\over{650}} = 4\%$. \subsection{Relativistic Corrections} The formula for the SZ effect quoted above (eqn~\ref{szeffect}) uses non-relativistic approximations, which become more significant at higher gas temperatures and higher observing frequency. Challinor and Lasenby \shortcite{challinor98} show that a better approximation in the Rayleigh-Jeans region (correct to second order in $T_e$) is \[ \frac{\Delta T}{T_0} = -2y\left(1-\frac{17}{10}\frac{k_BT_e}{m_e c^2}+\frac{123}{40}\left(\frac{k_BT_e}{m_e c^2}\right)^2\right). \] For A1413 this correction amounts to some 2.5\%, and we use the corrected value in the results below. \subsection{Fitting for \bf {$H_0$}} The SZ data used were those described by Grainge et al \shortcite{grainge96}, who also describe the procedure for subtracting the effects of radio sources within the field. We used only the visibilities from the RT baselines shorter than 2 k$\lambda$ for the fitting, since the SZ effect is completely resolved out on longer baselines. The mock data based on our X-ray model were compared with the source-subtracted data from our observations to find best-fit values of $H_0$ and the corresponding value of $n_0$; these were $H_0$ = $57 ^{+16}_{-13}$~$\rm km s^{-1} Mpc^{-1}$ and $n_0 = 9.0 \times 10^3$~$\rm m^{-3}$ (for an Einstein-de Sitter universe and apply the correction for primary flux calibration discussed in section~\ref{prim}). A plot of the posterior probability distribution is shown in Figure~\ref{likelihood}. The central SZ temperature decrement in this model is~$825~\mu$K. The quoted (1-$\sigma$) error on $H_0$ is just that due to the noise on the SZ data and the fitting to the X-ray image; other sources of error are discussed below. \section{Other errors in the calculation of $H_0$} There are other sources of error which may affect our estimate of $H_0$. These arise in the radio observations, in the X-ray data, and in the modelling process. \subsection{Primary flux calibration of the RT}\label{prim} The RT is primary flux-calibrated every day on either 3C~48 or 3C~286, assumed to have flux densities in the observed Stokes' parameter {\it I}+{\it Q} of 1.70 and 3.50~Jy respectively. These fluxes are derived from Baars et al. \shortcite{B77}. We estimate that the daily random error in primary calibration is $5\%$. Since the total integration time on A1413 is 65~days, the overall error in primary calibration due to variation during each day is $\pm 0.6 \%$. Since the calculated value of $H_0$ is proportional to the square of the measured SZ flux, this leads to an error of $\pm 1.2 \%$ in $H_0$. Measurements by the VLA in `D' configuration have shown that the flux densities of both 3C~48 and 3C~286 show slight variations with time \cite{vlacalib}. This time variability is small, less than $5\%$. Measurements in 1995, roughly contemporary with our observations, showed that the true flux density of 3C48 was 3.4\% lower than the Baars et al value, while that of 3C286 was 1.2\% higher. Since roughly equal numbers of the individual RT observations of A1413 were calibrated from each source, we average the errors and take the likely systematic error as $1.1\pm2.5$\%. This results in a reduction in $H_0$ of 2.2\%, and a further error term of $\pm 5$\%. \subsection{Source subtraction}\label{soursub} We removed the effect of contaminating radio sources using higher resolution data from both the longer baselines of the RT and from VLA imaging of the field. However, the accuracy to which the sources can be removed depends on the level of uncalibrated amplitude and phase errors in the RT. From maps of secondary calibrator sources we have found that the dynamic range of the RT is typically 100:1. Since the total flux density of the six sources removed from the A1413 data is 3.1~mJy, the maximum spurious flux remaining after source subtraction will be $30~\mu$Jy. However, taking into account the distribution of the sources on the sky and the contribution each makes at the position of the SZ decrement, we find that the maximum unsubtracted contribution of the detected sources is only $3~\mu$Jy. However, a more significant contribution is from the sources which were not subtracted due to their being too faint to detect. To calculate the confusion noise contribution from unsubtracted sources, we take the $8.4~$GHz $\mu$Jy source counts of Windhorst et al \shortcite{windhorst}, and extrapolate to 15~GHz using an effective spectral index calculated from their 5--8.4~GHz spectral measurements. Integrating the source counts over the beam area corresponding to our shortest baseline, from zero flux density up to our source detection limit of $120~\mu$Jy, and taking into account the increased source population in clusters compared to the field \cite{cooray}, we find a residual confusion noise of $60~\mu$Jy, which adds in quadrature with the thermal noise. However, since we estimate our thermal noise from the scatter in the visibilities, the confusion noise is already included in our noise estimate, and does not need to be added in again. \subsection{X-ray emission constant} The X-ray emission constant, $K$, used in simulating the X-ray data is dependent upon the telescope detector response in the appropriate frequency band, the K-correction for redshift, the cluster temperature and the absorption due to Galactic hydrogen. We estimate that this constant could be in error by $\pm 7\%$ leading to a $\pm 7\%$ error in $H_0$. \subsection{Extended radio emission} Some clusters of galaxies are known to have diffuse (`halo') emission on arcminute scales. Any such emission at 15~GHz would be resolved out on the longer RT baselines used for source subtraction, and would therefore contaminate the SZ data. Observations at 1.4~GHz from the NVSS and FIRST surveys show that there is 1.9~mJy of flux associated with the central cluster galaxy which is extended on scales of $\approx 45''$, but none on larger scales. This source was detected (and indeed slightly resolved) in the RT longer baselines and was removed from the data (source 3 in Grainge et al \shortcite{grainge96}). The spectral index of this source $\alpha^{15}_{1.4}$ is 0.8, typical for a cluster galaxy but quite atypical for a diffuse halo source, where one would expect $\alpha^{15}_{1.4} > 1.4$ \cite{H82}. Also, halo emission is unknown in clusters with strong cooling flows such as A1413. We therefore conclude that there is no evidence for any unsubtracted diffuse emission in A1413. \subsection{X-ray estimate of gas temperature} The temperature we use has a quoted 90\% confidence error of $^{+15}_{-9}\%$. Converting this to a 1-$\sigma$ errors for consistency gives $^{+10}_{-6}\%$, which will lead to a $^{+21}_{-12}\%$ error in $H_0$. \subsection{Cluster ellipticity} X-ray images of clusters show that clusters are typically elliptical with ellipticities of up to 1.2:1 being common. Therefore one cluster is not sufficient to provide a definitive value of $H_0$. Making the necessary corrections to equation \ref{calcH0}, one finds that the calculated value $ H_0^{\rm calc}$ is related to the true value by: \[ {H_0^{\rm calc}}\propto {H_0}^{\rm true} {{l_{\perp}}\over{l_{\parallel}}},\] for a cluster with line-of-sight depth $l_{\perp}$ and observed diameter $l_{\parallel}$. Therefore the derived $H_0$ from a sample where there is no bias in the orientation of clusters will form a distribution with a geometric mean close to the true value. It is thus necessary to observe such a sample of clusters in order to reduce the uncertainty due to this effect. An X-ray surface-brightness limited sample will be biased towards selecting clusters elongated towards us. However, an X-ray sample selected on X-ray flux will be independent of cluster orientation. In a case such as the present where one is estimating $H_0$ from a {\em single} cluster, it is appropriate to make some allowance for orientation uncertainty. For A1413 we have assumed that the line-of-sight depth through the cluster is the mean of the two elliptical axes in the plane of the sky. This is an unbiased estimator of the true depth, assuming the cluster is drawn from an unbiased population~\cite{grainger2001}. (Sulkanen \shortcite{sulkanen} finds a similar result taking the arithmetic mean.) In order to estimate the error introduced by this assumption we compare model distributions of elliptical clusters with those observed in a sample of ROSAT-selected clusters over a redshift range of $z=0.1$--0.5. We find that the effect of cluster ellipticity adds an error of $14\%$ for each cluster to the calculated value of $H_0$. \subsection{Effect of temperature structure}\label{coldhalos} Considerations of hydrostatic equilibrium, as well as observations on nearby clusters \cite{markevitch} show that the isothermal assumption breaks down at radii much bigger than $r_c$, the temperature falling by a factor of about 2 at $6r_c$. This can in principle have a significant effect on the derived value of $H_0$, tending to bias it downwards \cite{inagaki,roettiger}. For A1413, one core diameter corresponds to 2 arcminutes, which is approximately the scale to which the RT is sensitive. Gas on larger scales than this does not affect the flux density seen by the RT, despite the effect it has on the SZ central decrement. {\sl ASCA} observations do encompass larger scales, but the temperature measurement is necessarily emission-weighted, and hence is a good measurement of the temperature of the gas that the RT sees. We therefore conclude that SZ observations by the RT are not going to be significantly affected by the presence of a cold halo of gas, and so we can calculate $H_0$ without requiring knowledge of the gas temperature at large radii. A future paper will deal with this effect in more detail. \subsection{Clumping of the intracluster gas} X-ray imaging on nearby, large angular size clusters has found no evidence for clumping, and has placed constraints on the degree of clumping which may be present~\cite{FCEM94}. However it has been suggested that the scatter in the temperature--luminosity correlation for clusters is due to the existence of clumping below this level and further that the degree of clumping is correlated with the strength of the cooling flow \cite{KTK91}. If the clumps are close to being in pressure equilibrium with their surroundings the clumping will have no effect on the SZ signal. The X-ray luminosity will increase but the effect of this in the $H_0$ estimate will largely be cancelled due to the decreased emission-weighted temperature. The net result would be a modest underestimate of $H_0$. Initial simulations indicate that fractionation leading to an underestimate of $2\%$ in $H_0$ would be consistent with the temperatures calculated by Allen and Fabian~\shortcite{allen98}. \subsection{The kinetic SZ effect} Bulk motion of the cluster gas along the line of sight will result in a further distortion of the CMB spectrum due to the Doppler effect. Watkins \shortcite{w97} finds that the 1-d rms peculiar velocity of clusters is $\sigma_{v_z} = 265^{+106}_{-75}~\rm{km~s^{-1}}$ (at $90\%$ confidence). At 15~GHz, a cluster with a peculiar velocity of 265~km~$\rm s^{-1}$ will have a kinetic SZ effect with a magnitude $2.5\%$ of that due to the thermal SZ effect and so introduce a $5\%$ error in the determination of $H_0$ from a single cluster. Since randomly selected clusters will have random directions of peculiar motion, the error introduced into $H_0$ calculations can be reduced by averaging over a large sample of clusters. \subsection{Radio bremsstrahlung} It is possible that the electron cluster gas could emit sufficient thermal bremsstrahlung radiation at 15~GHz to significantly affect the SZ decrement which we observe. We calculated the emissivity of the hot thermal component of the gas using our cluster model, taking the Gaunt factor to be 10 at a frequency of 15~GHz. We find that the total integrated flux density from the entire cluster is $15~\mu$Jy. This emission will be distributed like the X-ray emission, and the resolving-out of our shortest baseline means that we will observe only about 20\% of it. Since the emissivity due to brem\-sstrah\-lung is proportional to $n^2$, it is possible that the effect of thermal bremsstrahlung will be much greater in a cluster which has a strong cooling flow and thus a high gas density at the cluster centre \cite{tarter,schlickaiser}. We investigated this possibility by using a good empirical fit to the form of a cooling flow, determined from X-ray observations, allowing the density to increase as $1/r$ for $r$ within the cooling flow radius of 100--200~kpc until $r = 10$~kpc. We find that this cooling flow contributes a further $4~\mu$Jy of flux density. Therefore we conclude that even for this worst case example, and assuming that none of the cooling-flow flux is resolved out by the RT, the effect of thermal bremsstrahlung will be to overestimate $H_0$ by up to $2\%$. \subsection{Effects of gravitational lensing} The strong gravitational potential of the cluster will inevitably lead to lensing of background radio sources in the field. Refregier and Loeb \shortcite{RL97} have calculated the systematic bias this introduces into $H_0$ calculations. Assuming $\beta = 0.6$, using a $2'$ beam and subtracting sources above $100~\mu$Jy, we find that approximately $0.5\%$ of the SZ temperature decrement may be due to this effect. This leads to an underestimate in our value to $H_0$ of $1\%$. \subsection{Combining the errors} \label{theanswer} All the errors listed above are summarised in table \ref{table}. Writing the result as a function of the systematic quantities concerned, we find \[ H_0 = 57^{+16}_{-13} \times \left( {{{1.64~\rm{Jy}}\over{S_{3C48}}} + {{3.54~\rm{Jy}}\over{S_{3C286}}}\over{2}} \right)^2 \times \left({{3.41\times10^{-69}}\over{K}} \right) \cr \times \left ( {{k_BT_e}\over{8.5~\rm{keV}}} \right)^2 \times \left({{\theta_z}\over{57''}}\right) \times \left(1+0.085{{v_{\rm z}}\over{1000~\rm{km~s^{-1}}}} \right)^2 \cr \times \left(1-{{S_{\rm{brems}}}\over{640~\mu\rm{Jy}}} \right)^2 \times \left(1+{{S_{\rm{lens}}}\over{640~\mu\rm{Jy}}} \right)^2 \rm \, km \, s^{-1} \, Mpc^{-1} \] where $S_{3C48}$ and $S_{3C286}$ are the true primary calibrator flux densities (I+Q), $K$ is the X-ray emissivity constant as defined in section \ref{a1413xfit}, $T_e$ is the gas temperature, $\theta_z$ the line-of-sight core radius, $v_z$ the line-of-sight peculiar velocity, and $S_{\rm brems}$ and $S_{\rm lens}$ are the flux densities due to bremsstrahlung emission and lensed background sources respectively. Inserting our best estimates for the errors in these quantities, we find in the worst case, where the errors conspire to add together linearly, $H_0 = 57^{+58}_{-27}$~$\rm km s^{-1} Mpc^{-1}$. However, since these errors are all independent of each other, we can justifiably combine them in quadrature with the random error. Doing this we obtain $H_0$ = $57^{+23}_{-16}$~$\rm km s^{-1} Mpc^{-1}$. The largest errors are due to the uncertainty in X-ray temperature, the noise on the SZ measurement and uncertainty in cluster line-of-sight depth. \section{Conclusions} We have described our methods of determining the Hubble constant from SZ and X-ray observations of clusters of galaxies, giving careful consideration to sources of both random and systematic error. We have used observations of the cluster A1413 to estimate $H_0$. We conclude that: \begin{enumerate} \item The error due to unsubtracted radiosources is comparable to the thermal noise, and adds in quadrature with it; our estimate of the noise from the scatter in the visibilities includes both effects. \item Fitting the X-ray emission using a King model gives a degeneracy in the $\beta, \ \theta$ plane. The predicted flux density that the RT will observe based on any of these good-fit models is almost constant, with the result that the error in $H_0$ that is introduced solely by our X-ray fitting is small, approximately~$\pm 4\%$. \item There is a random uncertainty in the calculation of the X-ray emissivity of the cluster gas which we estimate to be 7\%. \item The orientation of any non-spherical cluster causes a random uncertainty in the value of $H_0$ of some 14\%. This error may be reduced by observing a complete sample of randomly-oriented clusters. \item A colder cluster atmosphere at large radius will not affect our determination of $H_0$ much, but clumps of cool gas within the cluster may result in an underestimate the value of $H_0$ of up to 2\%. \item The kinetic SZ effect will cause a random uncertainty of approximately 2.5\% the value of the thermal SZ effect. This error may be reduced by observing a complete sample of randomly-oriented clusters. \item We estimate that the contribution of thermal bremsstrahlung at radio wavelengths will cause an overestimate of $H_0$ of 2\% in the worst case. \item Gravitational lensing will cause the {\em over}-subtraction of flux from contaminating background radiosources; this is a small effect causing an underestimate in $H_0$ of about 1\%. \item The dominant sources of error are thus the thermal/confusion noise in the SZ measurement, the temperature of the gas, and the unknown line-of -sight depth. Combining these sources of random and systematic error in quadrature, we determine from our observations of A1413 a value of $H_0 = 57^{+23} _{-16}$~$\rm km s^{-1} Mpc^{-1}$. \end{enumerate} \section{Acknowledgements} We thank the staff of the Cavendish Astrophysics group who ensure the continued operation of the Ryle Telescope. Operation of the RT is funded by PPARC. AE acknowledges support from the Royal Society; WFG acknowledges support from a PPARC studentship. \clearpage
\section{Introduction} The 3D XY universality class is unique in the respect that experimental estimates for critical exponents are more precise than any theoretical estimate. These experiments are performed in the neighbourhood of the super-fluid transition of $^4$He. The specific heat or the super-fluid density is measured as a function of the temperature \cite{helium1,helium2,helium3}. In the present study we try to close the gap between theory and experiment by a high statistics Monte Carlo simulation of the two-component $\phi^4$ (or Landau-Ginzburg) model on a three dimensional simple cubic lattice. The action is given by \begin{equation} S = \sum_x \{- 2 \kappa \; \sum_{\mu} \vec{\phi}_x \vec{\phi}_{x+\hat \mu} +\vec{\phi}_x^2 + \lambda \; (\vec{\phi}_x^2 -1)^2 \} \;\; , \label{action} \end{equation} where the field variable $\vec{\phi}_x$ is a vector with two real components and $x=(x_1,x_2,x_3)$, where $x_i$ is integer, labels the lattice sites. $\mu$ labels the directions and $\hat\mu$ is a unit-vector in $\mu$-direction. The Boltzmann factor is $\exp(-S)$. For $\lambda=0$ we get the Gaussian model on the lattice. In the limit $\lambda=\infty$ the XY-model is recovered. In addition to statistical errors Monte Carlo estimates of critical exponents are affected by systematical errors that result from corrections to scaling. These systematical errors can be reduced (in a finite size scaling study) by increasing the linear size $L$ of the lattices that are simulated. A more elegant approach is to remove corrections by a suitable choice of the action. Recently it was demonstrated that leading order corrections to scaling can be removed by a suitable tuning of the coupling constant $\lambda$ in the one-component $\phi^4$ theory on the lattice \cite{spain,us,MH}. Leading order corrections to scaling are proportional to $\xi^{-\omega}$ ($L^{-\omega}$ in finite size scaling), where $\xi$ is the correlation length and $\omega\approx 0.8 $. The paper is organised as follows: In section 2 we discuss the observables that are measured. In section 3 we explain the algorithm that has been used for the simulation and we summarise the simulation parameters. In section 4 the data are analysed. In section 5 our results for exponents are compared with experimental and theoretical estimates given in the literature. In section 6 we give our conclusions and an outlook. \section{The measured quantities} In the case of the one-component model the Binder cumulant turned out to be a good indicator for corrections to scaling \cite{MH}. The Binder cumulant is defined by \begin{equation} U = \frac{<(\vec{m}^2)^2>}{<\vec{m}^2>^2} \;\; , \end{equation} where \begin{equation} \vec{m} = \frac{1}{V} \sum_x \vec{\phi}_x \;\; \end{equation} is the magnetisation per lattice site of a given configuration. The volume is $V=L^3$. In the following we will always consider systems with periodic boundary conditions. In ref. \cite{MH} the Binder cumulant was computed at a fixed value of the ratio of partition functions $Z_a/Z_p$. $Z_a$ is the partition function for anti-periodic boundary conditions and $Z_p$ for periodic boundary conditions. This ratio can also be computed for an arbitrary number of components. For a simulation of the XY model see ref. \cite{GoHa}. However in the present paper we have replaced $Z_a/Z_p$ by the dimension-less ratio $\xi_{2nd}/L$ because the second moment correlation length $\xi_{2nd}$ is easier to implement as $Z_a/Z_p$. Note that $\xi/L$, where $\xi$ is the exponential correlation length on a strip of width $L$, was used in the pioneering work of Nightingale \cite{Night} on the phenomenological renormalization group approach. The second moment correlation length is defined by \begin{equation} \xi_{2nd} = \left(\frac{\chi/F -1}{4 \sin^2(\pi/L)} \right)^{1/2} \;\;, \end{equation} where the magnetic susceptibility is given by \begin{equation} \chi= {V} \; \langle \vec{m}^2 \rangle \end{equation} and \begin{equation} F = \frac{1}{V} \; \langle|\sum_x \exp\left(\;i \; \frac{ 2 \pi x_1}{L} \right) \; \vec{\phi}_x|^2 \rangle \end{equation} is the Fourier transform of the correlation function at minimal momentum. In the simulation we averaged over all three directions to reduce the statistical error. Note that in the following $\xi_{2nd}$ is always evaluated at a finite value of $L$ and not for the thermodynamic limit. We performed some simulations of the one-component model to compare $\xi_{2nd}/L$ and $Z_a/Z_p$. We found that the physical as well as statistical properties of $\xi_{2nd}/L$ and $Z_a/Z_p$ are similar. In order to compute observables in the neighbourhood of the simulation parameter $\kappa_s$ we computed the first two coefficients of the Taylor expansion in $\kappa-\kappa_s$. We always checked that the errors made by the truncation of the Taylor series are much smaller than the statistical errors of the quantities that were computed. \section{The Simulations} \subsection{The Monte Carlo algorithm} We generalise the idea of Brower and Tamayo \cite{BrTa} to simulate the one-component $\phi^4$ theory. They use the Swendsen-Wang cluster algorithm \cite{SwWa} to update the sign of the field $\phi$. In order to obtain an ergodic update they supplement the cluster-update with a Metropolis update that also allows to update the modulus of the field. In our case we use the single cluster algorithm \cite{Wolff} only to update the direction of the field. The modulus is updated with the Metropolis algorithm. Let us briefly recall the steps of the single cluster algorithm applied to the two-component $\phi^4$ theory. First a direction $\vec{n}$ is chosen \begin{equation} n_1 = \sin(2\pi \theta) \;\;\; , \;\; n_2 = \cos(2\pi \theta) \;\; , \end{equation} where $\theta$ is a random-number that is uniformly distributed in $[0,1)$. Next randomly a site of the lattice is picked as seed of the cluster. The cluster is build recursively. New sites enter the cluster when they freeze onto their neighbours that are already members of the cluster. The freezing probability is $p_f=1-p_d$ with \begin{equation} p_d = \mbox{min}\left[1,\; \exp(- 4 \kappa \; (\vec{n} \vec{\phi}_x) \; (\vec{n} \vec{\phi}_y) ) \right] \;\; . \end{equation} The fields of all sites in the cluster are reflected \begin{equation} \vec{\phi}_x' = \vec{\phi}_x - 2 \; (\vec{n} \vec{\phi}_x) \; \vec{n} \;\;. \end{equation} The modulus of $\vec{\phi}$ is changed with a local Metropolis update. A proposal for the field is generated by \begin{equation} \phi_{i,x}' = \phi_{i,x} + s (r_i - 0.5) \end{equation} for $i=1,2$, where $r_i$ is a random-number that is uniformly distributed in $[0,1)$. The acceptance probability is given by \begin{equation} A = \mbox{min}[1,\; \exp(S-S')] \;\;, \end{equation} where $S$ and $S'$ are the action for the original field and the proposal, respectively. We found that a step-size $s = 2$ yields an acceptance rate of about $50 \%$. In one sweep we go trough the lattice in lexicographic order. \subsection{The simulation parameters} We performed simulations at a large range of $\lambda$ values and linear lattice sizes $L$. In table \ref{statist} we give an overview of the simulation parameters and the number of measurements for each set of simulation parameters. Most of our simulations were performed on 200MHz Pentium Pro PCs running under Linux. The program is written in C. As random number generator we used our own implementation of G05CAF of the NAG-library. The total amount of CPU-time used for the simulations was about 3 years on the 200MHz Pentium Pro PCs. \begin{table}[h] \caption[bx1] {\label{statist} \sl Summary of simulation parameters. In the first row we give the value of $\lambda$, in the second row the linear lattice size $L$ and in the third row the number of measurements divided by $3 \times 10^{6}$. } \vskip 0.2cm \begin{center} \begin{tabular}{|l|l|l|} \hline $\lambda$ & $L$ & stat/$3 \times 10^{6} $ \\ \hline 0.5 & 8,16 & 17,5 \\ 1.0 & 6,8,10,12,14,16,18,20,22,24 & 50,10,10,10,10,10,10,10,10,11 \\ 1.5 & 8,16 &13,6 \\ 1.7 & 8,12,24 & 15,10,2.5 \\ 1.8 & 3,4,5,6,7,8,9,10,12,16 &20,67,67,20,40,15,45,30,15,8\\ 1.9 & 3,4,5,6,7,8,12,16,20,24 &33,27,20,20,15,20,10,10,11,10\\ 1.98 & 8,12,16,20,24 &20,15,10,11,15\\ 2.0 &3,4,5,6,7,8,9,10,11,12,13,14& 133,67,67,50,25,20,24,20,20,20,20,30\\ &15,16,18,20,22,24,26,28,32,40,48& 20,20,25,25,25,20,16,15,22,10,10 \\ 2.2 & 3,4,5,6,7,8,9,10,12,16,24& 133,67,67,50,40,15,45,30,15,8,5 \\ 4.0 & 6,7,8,9,10,11,12,14,16,18,20,22,24 & 50,20,10,10,10,9,10,10,10,10,10,10,10\\ \hline \end{tabular} \end{center} \end{table} Per measurement we performed one sweep with the Metropolis algorithm and $m$ single cluster updates. The number of cluster updates was chosen roughly proportional to the linear lattice size $L$. For some lattice sizes we searched for the $m$ that gives the optimal performance of the algorithm. For $L=48$ we found $m=40$ as optimal. \section{Analysing the data} \subsection{The Binder cumulant and corrections to scaling} We analysed the Binder cumulant at $\xi_{2nd}/L=0.5927$ fixed. This means that first (at fixed $\lambda$) $\kappa_f$ is computed for that $\xi_{2nd}/L=0.5927$. Then the Binder cumulant is computed at $\kappa_f$. In the following we denote the Binder cumulant at $\xi_{2nd}/L=0.5927$ by $\bar{U}$. From preliminary simulations we know that $\xi_{2nd}/L=0.5927$ is a good approximation of \begin{equation} \xi_{2nd}/L^* = \lim_{L \rightarrow \infty} \;\; \xi_{2nd}/L|_{\kappa_c} \;\;\; . \end{equation} The advantage of this approach is that we need not to search for $\kappa_c$ and that due to cross-correlations the statistical error of the Binder cumulant at $\xi_{2nd}/L=0.5927$ fixed is smaller than at a given value of $\kappa$. (See e.g. ref. \cite{mc2}.) For large $L$ $\bar{U}$ approaches a universal constant $\bar{U}^*$. Leading order corrections are given by \begin{equation} \label{simple1} \bar{U}(L,\lambda) = \bar{U}^* + c_1(\lambda) \;\; L^{-\omega} \;\;. \end{equation} We fitted the data for all values of $\lambda$ simultaneously with this ansatz. The free parameters of this fit are $\bar{U}^*$, $\omega$ and $c_1(\lambda)$ for each value of $\lambda$. The results for various minimal lattice sizes $L_{min}$ that have been included in the fit are summarised in table \ref{omega1}. \begin{table}[h] \caption[bx2] {\label{omega1} \sl Fit results for the Binder cumulant evaluated at $\xi_{2nd}/L=0.5927$ fixed. The ansatz is given in eq. (\ref{simple1}). We give results for various minimal lattice sizes $L_{min}$, $\omega$ is the correction to scaling exponent.} \vskip 0.2cm \begin{center} \begin{tabular}{|c|c|c|c|} \hline \rule[0mm]{0mm}{4mm} $L_{min}$ & $\chi^2$/d.o.f. & $\bar{U}^*$ & $\omega$ \\ \hline \phantom{0}6 & 5.42 & 1.24357(3) & 0.786(6)\phantom{0} \\ \phantom{0}8 & 2.34 & 1.24324(4) & 0.775(6)\phantom{0} \\ 10 & 2.15 & 1.24311(5) & 0.788(10) \\ 12 & 1.80 & 1.24297(6) & 0.782(14) \\ 14 & 1.75 & 1.24279(8) & 0.790(20) \\ 16 & 1.86 & 1.24274(9) & 0.819(31) \\ \hline \end{tabular} \end{center} \end{table} The values for $\chi^2$/d.o.f. stay rather large as $L_{min}$ is increased. We could not pin-point a particular problem that caused this effect. On the other hand the result for the exponent $\omega$ is quite stable as $L_{min}$ is varied. As our final result for the correction to scaling exponent we quote $\omega=0.79(2)$. It is hard to give reliable estimates for the systematical errors. At least the fact that the result for $\omega$ stays almost constant starting from $L_{min}=6$ indicates that these errors should be small. For $L_{min}=12$, $14$ and $16$ we give the results for $c_1(\lambda)$ in table \ref{constant1}. Linear interpolation of the result for $c_1$ at $\lambda=2.0$ and $\lambda=2.2$ yields $\lambda_{opt} = 2.046(9)$, $2.086(9)$ and $2.101(10)$ for $L_{min}=12$, $14$ and $16$, respectively. Where $\lambda_{opt}$ is defined by $c(\lambda_{opt})=0$. There is still an increase in $\lambda_{opt}$ visible as $L_{min}$ increases. We quote $\lambda_{opt}=2.10(1)[5]$ as our final result. As a rough estimate of systematical errors we give (in the square brackets) the difference of the result for $L_{min}=12$ and $L_{min}=16$. \begin{table}[h] \caption[bx3] { \label{constant1} \sl The correction to scaling amplitude $c_1(\lambda)$ as a function of $\lambda$ from fits with the ansatz (\ref{simple1}). We give the results for three values of $L_{min}=12$, $14$ and $16$. } \vskip 0.2cm \begin{center} \begin{tabular}{|c|c|c|c|} \hline $\lambda$ & $L_{min}=12$ & $L_{min}=14$ & $L_{min}=16$ \\ \hline 0.5\phantom{0}&\phantom{--}0.2152(83)& \phantom{--0}0.2220(122) & \phantom{--}0.2408(207) \\ 1.0\phantom{0}&\phantom{--}0.0956(37)& \phantom{--}0.0999(59)& \phantom{--}0.1094(101)\\ 1.5\phantom{0}&\phantom{--}0.0398(21)& \phantom{--}0.0424(28)& \phantom{--}0.0464(43)\phantom{0}\\ 1.7\phantom{0}&\phantom{--}0.0229(12) & \phantom{--}0.0280(32) & \phantom{--}0.0314(42)\phantom{0} \\ 1.8\phantom{0}&\phantom{--}0.0153(8)\phantom{0} & \phantom{--}0.0186(17) & \phantom{--}0.0207(23)\phantom{0} \\ 1.9\phantom{0}&\phantom{--}0.0077(8)\phantom{0} & \phantom{--}0.0099(11) & \phantom{--}0.0114(15)\phantom{0} \\ 1.98 &\phantom{--}0.0038(7)\phantom{0} & \phantom{--}0.0067(11) & \phantom{--}0.0079(14)\phantom{0} \\ 2.0\phantom{0}&\phantom{--}0.0022(6)\phantom{0} & \phantom{--}0.0043(9)\phantom{0} & \phantom{--}0.0057(13)\phantom{0} \\ 2.2\phantom{0}& --0.0074(7)\phantom{0} & --0.0057(13) & --0.0056(16)\phantom{0} \\ 4.0\phantom{0}& --0.0604(24) & --0.0601(37) & --0.0649(63)\phantom{0} \\ \hline \end{tabular} \end{center} \end{table} Following ref. \cite{MH} we tried to fit our data with the extended ansatz \begin{equation} \label{simple2} \bar{U}(L,\lambda) = \bar{U}^* + c_1(\lambda) \;\; L^{-\omega} + c_2 \; c_1(\lambda)^2 \;\; L^{-2 \omega} \;\;. \end{equation} However it turned out that we have too few data with a large difference $\bar{U}-\bar{U}^*$ to resolve $c_2$. Finally we fitted the difference of the Binder cumulant at $\lambda=2.0$ and $\lambda=2.2$ with the ansatz \begin{equation} \label{delta} \bar{U}(L,\lambda=2.0) -\bar{U}(L,\lambda=2.2) = c \;\; L^{-\omega} \;\;. \end{equation} Results are given in the table \ref{constant}. It turns out that the $\chi^2$/d.o.f. becomes order $1$ already for the very small $L_{min}=3$. Also the value obtained for $\omega$ with this small $L_{min}$ is consistent with the result obtained above. Hence corrections beyond $L^{-\omega}$ do depend very little on $\lambda$ and are cancelled in $\bar{U}(L,\lambda=2.0) -\bar{U}(L,\lambda=2.2)$. The same observation holds in the case of the one-component model \cite{MH}. \begin{table}[h] \caption[bx4] { \label{constant} \sl Fitting the difference of $\bar{U}$ at $\lambda=2.0$ and $\lambda=2.2$ with the ansatz (\ref{delta}). } \vskip 0.2cm \begin{center} \begin{tabular}{|c|c|c|c|} \hline $L_{min}$ & $\omega$ & $c_1(2.0)-c_2(2.2)$ & $\chi^2$/d.o.f. \\ \hline 3 & 0.787(18) & 0.0106(3) & 0.98 \\ 4 & 0.780(31) & 0.0104(5) & 1.09 \\ 5 & 0.794(43) & 0.0107(9) & 1.21 \\ \hline \end{tabular} \end{center} \end{table} \subsection{The critical line $\kappa_c(\lambda)$} As approximation of the critical $\kappa_c$ we take $\kappa_f$ where $\xi_{2nd}/L=0.5927$. In table \ref{kritisch} we give the result for the largest lattice size available for each value of $\lambda$ that has been studied. Leading corrections are given by \begin{equation} \kappa_f - \kappa_c = a \; L^{-1/\nu} + b \; L^{-1/\nu-\omega} \; + \;... \;\;\; . \end{equation} The constant $a$ should be very small since we have chosen $\xi_{2nd}/L=0.5927$ as a good approximation of $\xi_{2nd}/L^*$. The value of $b$ depends on $\lambda$ and vanishes at $\lambda_{opt}$. Nevertheless we assume pessimistically that errors decay with $L^{-1/\nu}$. Systematical errors are then computed by comparing $\kappa_f$ at $L$ with $\kappa_f$ at $L/2$. These errors are given in square brackets. Whenever statistical errors reach a similar size as the systematical ones they are quoted in addition in round brackets. \begin{table}[h] \caption[bx5] { \label{kritisch} \sl Estimates of the critical $\kappa_c$ for all values of $\lambda$ that have been simulated. The value for $\lambda=\infty$ has been taken from ref. \cite{mc2}. The systematical errors are given in square brackets and statistical errors in round brackets. } \vskip 0.2cm \begin{center} \begin{tabular}{|l|c|l|} \hline \multicolumn{1}{|c|}{$\lambda$} & $L$ & \multicolumn{1}{|c|}{2 $\kappa_{c}$} \\ \hline 0 & & 0.33... \\ 0.5 & 16 & 0.4828[6] \\ 1.0 & 24 & 0.50754[7] \\ 1.5 & 16 & 0.51197[7] \\ 1.7 & 24 & 0.51160[2] \\ 1.8 & 16 & 0.51115[2] \\ 1.9 & 24 & 0.510576(2)[7] \\ 1.98 & 24 & 0.510049(1)[7] \\ 2.0 & 48 & 0.5099049(6)[9]\\ 2.2 & 24 & 0.508344(2)[4] \\ 4.0 & 24 & 0.49243[5] \\ $\infty$ & & 0.454165(4) \\ \hline \end{tabular} \end{center} \end{table} \subsection{The exponent $\eta$} We computed the exponent $\eta$ from the finite size behaviour of the magnetic susceptibility $\chi$ at either $\xi_{2nd}/L = 0.5927$ or $U=1.243$ fixed. We denote the magnetic susceptibility at $\xi_{2nd}/L$ or $U$ fixed by $\bar{\chi}$. It scales as \begin{equation} \label{chifit1} \bar{\chi} = d \, L^{2-\eta} \;\; . \end{equation} First we analysed our data for $\lambda=2.0$ which is close to $\lambda_{opt}$ and where we have accumulated most data. Results for $\xi_{2nd}/L$ fixed are given in table \ref{chixi1} and for $U$ fixed in table \ref{chiu1}. In both cases rather large $L_{min}$ are needed to reach a $\chi^2/$d.o.f. close to 1. Since $\bar{\chi}$ at fixed $\xi_{2nd}/L$ has a smaller statistical error than $\bar{\chi}$ at fixed $U$ also the statistical error of $\eta$ is smaller for $\xi_{2nd}/L$ fixed than for $U$ fixed. \begin{table}[h] \caption[bx6] {\label{chixi1} \sl Fits of the magnetic susceptibility at $\xi_{2nd}/L=0.5927$ fixed with the ansatz (\ref{chifit1}). } \vskip 0.2cm \begin{center} \begin{tabular} {|c|c|c|c|} \hline $L_{min}$ & $d$ & $\eta$ & $\chi^{2}/$d.o.f. \\ \hline 14 & 1.25629(20) & 0.03667(5)\phantom{0} & 10.28 \\ 24 & 1.25957(50) & 0.03742(11) & \phantom{0}2.91 \\ 26 & 1.26067(61) & 0.03766(14) & \phantom{0}0.70 \\ 28 & 1.26117(75) & 0.03777(17) & \phantom{0}0.40 \\ \hline \end{tabular} \end{center} \end{table} \begin{table}[h] \caption[bb27] {\label{chiu1} \sl Fits of the magnetic susceptibility at $U=1.243$ fixed with the ansatz (\ref{chifit1}). } \vskip 0.2cm \begin{center} \begin{tabular} {|c|c|c|c|} \hline $L_{min}$ & $d$ & $\eta$ & $\chi^2/$ d.o.f. \\ \hline 14 & 1.2598(6)\phantom{0} & 0.03740(16) & 2.27 \\ 22 & 1.2628(13) & 0.03811(31) & 1.20 \\ 24 & 1.2644(16) & 0.03845(38) & 0.83 \\ 26 & 1.2625(20) & 0.03804(46) & 0.29 \\ \hline \end{tabular} \end{center} \end{table} Because we had to go to large $L_{min}$ with the simple ansatz (\ref{chifit1}) we added an analytic correction \begin{equation} \label{chifit2} \bar{\chi} =c + d \, L^{2-\eta} \;\;\; . \end{equation} Note that also corrections that decay like $L^{-x}$ with $x\approx 2$ are effectively parametrised by this ansatz. Results for fits with this ansatz are given in table \ref{chixi2} for $\xi_{2nd}/L$ fixed and for $U$ fixed in table \ref{chiu2}. We see that a small $\chi^2/$d.o.f. is already reached for $L_{min}=7$ and $L_{min}=6$ respectively. Despite the fact that $\chi^2/$d.o.f. of order 1 is reached the results for $\eta$ do not match within statistical errors. This is a reminder that a small $\chi^2/$d.o.f. does not imply that systematical errors are of the same size as the statistical ones. Since the statistical error with $\xi_{2nd}/L$ fixed is smaller we take our final result from these fits. In order to estimate systematical errors we compare results of fits with the range $L_{min},L_{max}$ and $L_{min}'=2 L_{min}$, $L_{max}'=2 L_{max}$. Then the error due to $L^{-2}$ (which we assume to be the leading corrections beyond $L^{-\omega})$ corrections in the second interval should be $1/3$ of the difference of the two results (up to a difference in the distribution of the data with the interval). As our final estimate we take the fit result from $L_{min}=14$ and $L_{max}=48$. For comparison we fitted with $L_{min}=7$ and $L_{max}=24$. For this interval we get $\eta=0.03800(13)$. Hence the systematical error from $L^{-2}$ corrections should be smaller than $0.00012$ (taking statistical errors into account). \begin{table}[h] \caption[bbb11] {\label{chixi2} \sl Fits of the magnetic susceptibility at $\xi_{2nd}/L = 0.5927$ fixed with the extended ansatz (\ref{chifit2}). } \vskip 0.2cm \begin{center} \begin{tabular}{|c|c|c|c|c|} \hline $L_{min}$ & $c$ & $d$ & $\eta$ & $\chi^{2} \mbox{/d.o.f.}$ \\ \hline \phantom{0}6 & -0.3602(40) & 1.26187(21) & 0.03784(5)\phantom{0}& 2.07 \\ \phantom{0}7 & -0.3809(68) & 1.26246(26) & 0.03798(6)\phantom{0}& 1.33 \\ \phantom{0}8 & -0.395(11)\phantom{0}& 1.26280(34)&0.03805(8)\phantom{0} & 1.17 \\ 10 & -0.381(18)\phantom{0} & 1.26254(43) & 0.03800(10) & 1.06 \\ 12 & -0.393(32)\phantom{0} & 1.26275(62) & 0.03804(14) & 1.21 \\ 14 & -0.405(43)\phantom{0} & 1.26289(73) & 0.03807(16) & 1.40 \\ 16 & -0.436(72)\phantom{0} & 1.26330(99) & 0.03815(21) & 1.32 \\ \hline \end{tabular} \end{center} \end{table} \begin{table} \caption[bb1r] {\label{chiu2} \sl Fits of the magnetic susceptibility at $U = 1.243$ fixed with the extended ansatz (\ref{chifit2}). } \vskip 0.2cm \begin{center} \begin{tabular} {|c|c|c|c|c|} \hline $L_{min}$ & $c$ & $d$ & $\eta$ & $\chi^2/\mbox{d.o.f.}$ \\ \hline \phantom{0}4 & -0.464(4)\phantom{0}& 1.2651(4)\phantom{0}& 0.03845(11) & 3.27 \\ \phantom{0}6 & -0.525(12) & 1.2681(6)\phantom{0}& 0.03917(16) & 0.76 \\ \phantom{0}8 & -0.526(30) & 1.2682(10) & 0.03918(23) & 0.85 \\ 10 & -0.553(55) & 1.2688(14) & 0.03931(31) & 0.96 \\ 12 & -0.574(90) & 1.2691(18) & 0.03937(40) & 1.05 \\ 14 & -0.47(13)\phantom{0} & 1.2676(22) & 0.03907(49) & 1.17 \\ 16 & -0.24(23)\phantom{0} & 1.2649(32) & 0.03851(67) & 1.27 \\ \hline \end{tabular} \end{center} \end{table} Finally we checked for systematical errors due to residual leading order corrections to scaling at $\lambda=2.0$. For this purpose we fitted our data for $\lambda=1.0$ and $\lambda=4.0$ also with $L_{min}=14$ and ansatz (\ref{chifit2}). We get $\eta=0.0375(13)$ and $\eta=0.0373(13)$ respectively. Taking into account the statistical errors we find that \begin{equation} \left| \frac{\Delta \eta_{eff}}{\Delta c_1(\lambda)} \right| < 0.018 \;\;\; . \end{equation} From the previous section we know that the coefficient $c_1(2.0)$ should be smaller than $0.007$. (Taking the fit result for $L_{min}=16$ plus the statistical error). Therefore the systematical error in our final estimate of $\eta$ due to residual leading order corrections should be smaller than 0.00013. As a check we repeated the error-analysis along the lines of ref. \cite{us} and came up with a similar estimate. As final estimate for $\eta$ we take the result from fitting the magnetic susceptibility at $\xi_{2nd}/L$ fixed with the ansatz (\ref{chifit2}) and $L_{min}=14$ \begin{equation} \eta=0.0381(2)[2] \;\;. \end{equation} The estimate of the systematical error is given in the second bracket. It covers residual $L^{-\omega}$ corrections and higher order corrections. \subsection{The exponent $\nu$} We computed the derivate of the Binder cumulant $U$ with respect to $\kappa$ at the fixed value of the Binder cumulant $U=1.243$ and at the fixed value of $\xi_{2nd}/L=0.5927$. These quantities behave as \begin{equation} \label{simplenu} \overline{\frac{\partial U}{\partial \kappa}} = c \;\; L^{1/\nu} \;\; . \end{equation} Results of the fits are summarised in table \ref{abx} and \ref{abu} for fixed $\xi_{2nd}/L$ and for fixed $U$, respectively. The $\chi^2/$d.o.f. becomes order 1 starting from $L_{min}=8$ and $L_{min}=7$ respectively. The statistical errors are slightly smaller in the case of fixed $U$. \begin{table}[h] \caption[bbrx] {\label{abx} \sl Fits of $\frac{\partial U}{\partial \kappa}$ at $\xi_{2nd}/L$ fixed with the ansatz (\ref{simplenu}). } \vskip 0.2cm \begin{center} \begin{tabular} {|c|c|c|c|} \hline $L_{min}$ & $c/2$ & $\nu$ & $\chi^2/\mbox{d.o.f.}$ \\ \hline \phantom{0}6 & -0.5542(5)\phantom{0} & 0.6709(1) & 3.82 \\ \phantom{0}7 & -0.5565(6)\phantom{0} & 0.6715(2) & 1.76 \\ \phantom{0}8 & -0.5578(7)\phantom{0} & 0.6719(2) & 1.02 \\ 10 & -0.5586(9)\phantom{0} & 0.6721(2) & 0.96 \\ 12 & -0.5595(11) & 0.6723(3) & 0.80 \\ 14 & -0.5608(13) & 0.6727(4) & 0.65 \\ 16 & -0.5620(18) & 0.6729(5) & 0.69 \\ 20 & -0.5610(24) & 0.6727(6) & 0.63 \\ 24 & -0.5632(36) & 0.6732(9) & 0.50 \\ \hline \end{tabular} \end{center} \end{table} \begin{table} \caption[brxz] {\label{abu} \sl Fits of $\frac{\partial U}{\partial \kappa}$ at $U$ fixed with the ansatz (\ref{simplenu}).} \vskip 0.2cm \begin{center} \begin{tabular} {|c|c|c|c|} \hline $L_{min}$ & $c/2$ & $\nu$ & $\chi^2/\mbox{d.o.f.}$ \\ \hline \phantom{0}6 & -0.5551(4)\phantom{0} & 0.6712(1) & 2.07 \\ \phantom{0}7 & -0.5564(5)\phantom{0} & 0.6716(1) & 1.19 \\ \phantom{0}8 & -0.5572(6)\phantom{0} & 0.6718(2) & 0.80 \\ 10 & -0.5577(7)\phantom{0} & 0.6719(2) & 0.73 \\ 12 & -0.5583(9)\phantom{0} & 0.6721(3) & 0.62 \\ 14 & -0.5593(11) & 0.6723(3) & 0.51 \\ 16 & -0.5601(15) & 0.6725(4) & 0.55 \\ 20 & -0.5592(21) & 0.6723(5) & 0.48 \\ 24 & -0.5610(31) & 0.6727(7) & 0.30 \\ \hline \end{tabular} \end{center} \end{table} As in the case of the exponent $\eta$ we expect in addition to the statistical error systematical errors due to the fact that the coefficient of $L^{-\omega}$ corrections does not vanish exactly and due to sub-leading $L^{-2}$ corrections. In order to estimate these errors we proceed as in the previous section. As our final result we take the fit with $L_{min}=14$ and $L_{max}=48$ of $\frac{\partial U}{\partial \kappa}$ at fixed $U$. In order to estimate $L^{-2}$ corrections we fitted the data in the interval $L_{min}=7$ and $L_{max}=24$. For these lattices sizes we obtain $\nu=0.6712(2)$. Hence the estimate for a $L^{-2}$ error is $0.0011(5)/3 \approx 0.0005$. In order to estimate the error due to residual $L^{-\omega}$ corrections we fitted our data for $\lambda=1.0$ and $\lambda=4.0$. From $L_{min}=14$ we obtain $\nu=0.6706(11)$ for $\lambda=1.0$ and $\nu=0.6758(10)$ for $\lambda=4.0$. Hence \begin{equation} \left| \frac{\Delta \nu_{eff}}{\Delta c_1(\lambda) } \right| < 0.04 \;\;\; . \end{equation} From the previous section we know that $ c_1(2.0) \approx 0.007 \;$. Therefore the estimate of the systematical error in $\nu$ is $0.04 \times 0.007 \approx 0.0003 \;$ . We arrive at our final estimate \begin{equation} \nu=0.6723(3)[8] \;\; , \end{equation} where the statistical error is given in the first bracket and the systematical error that is given in the second bracket covers $L^{-2}$ and residual $L^{-\omega}$ corrections. \section{Comparison with the literature} In table \ref{literature} we give for comparison recent results for critical exponents. Critical exponents for the XY-universality class have been calculated using the high temperature series expansions, the $\epsilon$-expansion, perturbation theory in three dimension and Monte Carlo simulations. Our result for $\nu$ is consistent within error-bars with (almost) all other theoretical results given in table \ref{literature}. The result of the MC study \cite{mc1} seems to be a little too small. Our error-bar is smaller than that of all previous estimates. Our estimate for $\eta$ is consistent with the other theoretical estimates except with the Monte Carlo results. The values of refs. \cite{mc0,mc1} are too small compared with our present estimate. Note that in these studies no careful check of systematical errors due to corrections to scaling was performed. On the other hand the result of ref. \cite{mc2}, which takes into account $L^{-\omega}$ corrections, is by two standard deviations larger than our result. In contrast to the one-component case \cite{MH} our result for the correction to scaling exponent $\omega$ is consistent with that obtained with field theoretic methods. \begin{table}[h] \caption[bx10] { \label{literature} \sl Recent results for critical exponents obtained with Monte Carlo simulations (MC), $\epsilon$-expansion, Perturbation-Theory in three dimensions (3D,PT) and High temperature series expansions. When only $\nu$ and $\gamma$ are given in the reference we computed $\eta$ with the scaling law. These cases are indicated by $^*$. In ref. \cite{helium1} a result for $\alpha$ is given. In the table it is converted to $\nu$ using the scaling relation $\alpha=2-d \nu$. For a discussion see the text. } \vskip 0.2cm \begin{center} \begin{tabular}{|c|c|l|l|l|} \hline Ref. & Method &\multicolumn{1}{|c|}{$\nu$} & \multicolumn{1}{|c|}{$\eta$} & \multicolumn{1}{|c|}{$\omega$} \\ \hline present work & MC & 0.6723(3)[8] & 0.0381(2)[2] & 0.79(2) \\ \cite{mc0} & MC & 0.670(2)& $0.025(7)^*$ & \\ \cite{mc1} & MC & 0.662(7) & 0.026(6) & \\ \cite{mc2} & MC & 0.6721(13) & 0.042(2) & \\ \cite{guzi} & 3D,PT & 0.6703(15) & 0.0354(25) & 0.789(11) \\ \cite{guzi} & $\epsilon$,bc &0.6680(35) & 0.0380(50)& 0.802(18) \\ \cite{guzi} & $\epsilon$,free &0.671 & 0.0370 & 0.802(18) \\ \cite{buco} & HT & 0.674(2) & $0.039(7)^*$ & \\ \cite{helium1} & $^4$He & 0.67095(13) & & \\ \cite{helium2} & $^4$He & 0.6705(6) & & \\ \cite{helium3} & $^4$He & 0.6708(4) & & \\ \hline \end{tabular} \end{center} \end{table} Experimental results for the exponent $\nu$ have been obtained for the $\lambda$-transition of $^4$He. These results have smaller error-bars than our Monte Carlo result. The experimental results are all smaller then our value but still the error-bars touch. \section{Conclusion and outlook} In this paper we have improved the accuracy of the theoretical estimate of $\nu$ of the 3D XY universality class considerably. In particular we give in addition to the statistical error a careful estimate of systematical errors that are caused by corrections to scaling. Our value $\nu=0.6723(3)[8]$ is consistent with other theoretical estimates. However it is larger than the experimental results obtained from the $\lambda$-transition of $^4$He \cite{helium1,helium2,helium3} that give values from $0.6704$ up to $0.6709$ with an error in the last digit. It would be interesting to further improve the theoretical estimate to the claimed accuracy of the experimental results. This could be achieved by simulating at our best estimate for $\lambda_{opt}=2.1$ and going to linear lattice sizes roughly twice as large as in the present study to reduce the effect of sub-leading corrections. At a sustained statistical accuracy this would require about 10 years of CPU-time on a modern PC. In addition to critical exponents amplitude ratios are universal and have been experimentally determined for the $\lambda$-transition of $^4$He. For example the specific heat behaves in the neighbourhood of the phase transition as \begin{equation} C = A_{\pm} \; |t|^{-\alpha} \; (1 + D_{\pm} \; |t|^{\theta} + E_{\pm} \; t ) \;+\; B \;\;, \end{equation} where $t=(T-T_c)/T_c$ is the reduced temperature. The constants $A_{\pm}$, $D_{\pm}$, $E_{\pm}$ and $B$ depend on the system that is considered. The subscript $\pm$ indicates the low and high temperature phase. However renormalization group predicts the ratio $A_+/A_-$ to be universal. Setting $\lambda=\lambda_{opt}$ leads to $D_{\pm}=0$ which greatly simplifies the determination of $A_+/A_-$ in a Monte Carlo simulation.
\section{Introduction} In our earlier papers~\cite{hkn97josa,hkn97pre}, we have formulated the Jones vectors and Stokes parameters in terms of the two-by-two and four-by-four matrix representations of the six-parameter Lorentz group~\cite{brown95}. It was seen there that, to every two-by-two transformation matrix for the Jones vector, there is a corresponding four-by-four matrix for the Stokes parameters. It was found also that the Stokes parameters are like the components of Minkowskian four-vectors, and two-component Jones vectors are like two-component spinors in the relativistic world. This enhances our capacity to approach polarization optics in terms of the kinematics of special relativity. Indeed, we can now design specific experiments which will test some of the consequences derivable from the principles of special relativity. The most widely known example is the Wigner rotation. This has been extensively discussed in the literature in connection with the Thomas effect~\cite{hks87cqg}, Berry's phase~\cite{chiao88,kitano89}, and squeezed states of light~\cite{knp91}. In our earlier papers, we discussed an optical filter which will exhibit the matrix form of \begin{equation}\label{shear1} \pmatrix{1 & u \cr 0 & 1} \end{equation} applicable to two transverse components of the light wave, where u is a controllable parameter. When applied to a two-component system, this matrix performs a superposition in the upper channel while leaving the low channel invariant. The question is whether it is possible to produce optical filters with this property. In Ref.~\cite{hkn97josa}, we approached this problem in terms of the generators of the Lorentz group. It is very difficult, if not impossible, to manufacture optical devices performing the function of group generators. In the case of optical filters, this means an infinite number of layers of zero thickness. In the present paper, we deal with the same problem from the experimental point of view. We will present a specific design for optical filters performing this function. We will of course present our case in terms of a combination of three filters of finite thickness. In order to achieve this goal, we use the fact that polarization optics and special relativity shares the same mathematics. This aspect was already noted in the literature for the case of the Wigner rotation~\cite{kitano89}. The concept of the Wigner rotation comes from the kinematics of special relativity, in which two successive non-collinear Lorentz boosts do not end up with a boost. The result is a boost followed or preceded by a rotation. Thus we can achieve a rotation from three non-collinear boosts starting from a particle at rest. Since each boost corresponds to an attenuation filter, it requires three attenuation filters to achieve a Wigner rotation in polarization optics. While the Wigner rotation is based on Lorentz transformations of massive particles, there are similar transformations for massless particles. Here, two non-collinear Lorentz boosts do not result in one boost. They become one boost preceded or followed by a transformation which corresponds to a gauge transformation. In two-by-two formalism, the transformation takes the form of Eq.(\ref{shear1}). We shall show in this paper that the filter possessing the property of Eq.(\ref{shear1}) can be constructed from one rotation filter and one attenuation filter. In mathematics, this type of decomposition is called the Iwasawa decomposition~\cite{iwa49,simon98}. While the primary purpose of this paper is to discuss filters and their combinations in polarization optics, we provide also concrete illustrative examples of Wigner's little group~\cite{wig39}. The little group is the maximal subgroup of the Lorentz group whose transformations leave the four-momentum of a given particle invariant, and has a long history~\cite{knp86}. The Wigner rotation and the Iwasawa decomposition are transformations of the little groups for massive and massless particles respectively. It is interesting to note that these transformations can be also achieved in optics laboratories. In Sec.~\ref{formul}, we review the formalism for optical filters based on the Lorentz group and explain why filters are like Lorentz transformations. It is shown in Sec. \ref{wigrot}, that a rotation can be achieved by three non-collinear Lorentz boosts. In Sec.~\ref{iwasa}, we spell out in detail how the Iwasawa decomposition can be achieved from the combination of two optical filters. \section{Formulation of the Problem}\label{formul} In studying polarized light propagating along the $z$ direction, the traditional approach is to consider the $x$ and $y$ components of the electric fields. Their amplitude ratio and the phase difference determine the degree of polarization. Thus, we can change the polarization either by adjusting the amplitudes, by changing the relative phases, or both. For convenience, we call the optical device which changes amplitudes an ``attenuator'' and the device which changes the relative phase a ``phase shifter.'' Let us write these electric fields as \begin{equation}\label{expo1} \pmatrix{E_{x} \cr E_{y}} = \pmatrix{A \exp{\left\{i(kz - \omega t + \phi_{1})\right\}} \cr B \exp{\left\{i(kz - \omega t + \phi_{2})\right\}}} . \end{equation} where $A$ and $B$ are the amplitudes which are real and positive numbers, and $\phi_{1}$ and $\phi_{2}$ are the phases of the $x$ and $y$ components respectively. This column matrix is called the Jones vector. In dealing with light waves, we have to realize that the intensity is the quantity we measure. Then there arises the question of coherence and time average. We are thus led to consider the following parameters. \begin{eqnarray}\label{sii} S_{11} &=& <E_{x}^{*}E_{x}> , \qquad S_{22} = <E_{y}^{*}E_{y}> , \cr S_{12} &=& <E_{x}^{*}E_{y}> , \qquad S_{21} = <E_{y}^{*}E_{x}> . \end{eqnarray} Then, we are naturally invited to write down the two-by-two matrix: \begin{equation}\label{cohm1} C = \pmatrix{<E^{*}_{x}E_{x}> & <E^{*}_{y} E_{x}> \cr <E^{*}_{x} E_{y}> & <E^{*}_{y} E_{y}>} , \end{equation} where $<E^{*}_{i}E_{j}>$ is the time average of $E^{*}_{i}E_{j}$. The above form is called the coherency matrix~\cite{born80}. It is sometimes more convenient to use the following combinations of parameters. \begin{eqnarray}\label{stokes} &{}& S_{0} = S_{11} + S_{22}, \cr &{}& S_{1} = S_{11} - S_{22}, \cr &{}& S_{2} = S_{12} + S_{21}, \cr &{}& S_{3} = -i\left(S_{12} - S_{21}\right). \end{eqnarray} These four parameters are called the Stokes parameters in the literature~\cite{born80}. We have shown in our earlier papers that the Jones vectors and the Stokes parameters can be formulated in terms of the two-by-two spinor and four-by-four vector representations of the Lorentz group. This group theoretical formalism allows to discuss three different sets of physical quantities using one mathematical device. In our earlier publications, we used the concept of Lie groups extensively and used their generators based on infinitesimal generators. In this paper, we avoid the Lie groups and work only with explicit transformation matrices. For this purpose, we start with the following two matrices. \begin{eqnarray} &{}& B = \pmatrix{\cosh\chi & \sinh\chi & 0 & 0 \cr \sinh\chi & \cosh\chi & 0 & 0 \cr 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 1}, \nonumber \\[2ex] &{}& R = \pmatrix{1 & 0 & 0 & 0 \cr 0 & \cos\phi & -\sin\phi & 0 \cr 0 & \sin\phi & \cos\phi & 0 \cr 0 & 0 & 0 & 1} . \end{eqnarray} If the above matrices are applied to the Minkowskian space of $(ct, z, x, y)$, the matrix $B$ performs a Lorentz boost: \begin{eqnarray} &{}& t' = (\cosh\chi) t + (\sinh\chi) z , \nonumber \\[2ex] &{}& z' = (\sinh\chi) t + (\cosh\chi) z, \end{eqnarray} while $R$ leads to a rotation: \begin{eqnarray} &{}& z' = (\cos\phi) z - (\sin\phi) x , \nonumber \\[2ex] &{}& x' = (\sin\phi) z + (\cos\phi) x . \end{eqnarray} In our previous paper, we discussed in detail what these matrices do when they are applied to the Stokes four-vectors. In the two-component spinor space, the above transformation matrices take the form \begin{equation} \pmatrix{e^{\chi/2} & 0 \cr 0 & e^{-\chi/2}}, \qquad \pmatrix{\cos(\phi/2) & -\sin(\phi/2) \cr \sin(\phi/2) & \cos(\phi/2)} . \end{equation} We discussed the effect of these matrices on the Jones spinors in our earlier publications. In this paper, we discuss some of nontrivial consequences derivable from the algebra generated by these two sets of matrices. We shall study Wigner rotations and Iwasawa decompositions. The Wigner rotation has been discussed in optical science in connection with Berry's phase, but the Iwasawa decomposition is a relatively new word in optics. We would like to emphasize here that both the Wigner rotation and Iwasawa decomposition come from the concept of subgroup of the Lorentz groups whose transformations leave the momentum of a given particle invariant. \section{Wigner Rotations}\label{wigrot} There are many different versions of the Wigner rotation in the literature. Basically, this rotation is a product of two non-collinear Lorentz boosts. The result of these two boosts is not a boost, but a boost preceded or followed by a rotation. This rotation is called the Wigner rotation. \begin{figure}[thb] \centerline{\psfig{figure=triangle.ps,angle=0,height=60mm}} \caption{Closed Lorentz boosts. Initially, a massive particle is at rest with its four momentum $P_{a}$. The first boost $B_{1}$ brings $P_{a}$ to $P_{b}$. The second boost $B_{2}$ transforms $P_{b}$ to $P_{c}$. The third boost $B_{3}$ brings $P_{c}$ back to $P_{a}$. The particle is again at rest. The net effect is a rotation around the axis perpendicular to the plane containing these three transformations. We may assume for convenience that $P_{b}$ is along the $z$ axis, and $P_{c}$ in the $zx$ plane. The rotation is then made around the $y$ axis.} \end{figure} In this paper, we approach the problem by using three boosts described in Fig.~1. Let us start with a particle at rest, with its four momentum \begin{equation}\label{4a} P_{a} = (m, 0, 0, 0), \end{equation} where we use the metric convention $(ct, z, x, y)$. Let us next boost this four-momentum along the $z$ direction using the matrix \begin{equation}\label{boost1} B_{1} = \pmatrix{\cosh\eta & \sinh\eta & 0 & 0 \cr \sinh\eta & \cosh\eta & 0& 0 \cr 0 & 0 & 1 & \cr 0 & 0 & 0 & 1} , \end{equation} resulting in the four-momentum \begin{equation}\label{4b} P_{b} = m (\cosh\eta, \sinh\eta, 0, 0) . \end{equation} Let us rotate this vector around the $y$ axis by an angle $\theta$. Then the resulting four-momentum is \begin{equation}\label{4c} P_{c} = m \left(\cosh\eta, (\sinh\eta)\cos\theta, (\sinh\eta)\sin\theta, 0 \right) . \end{equation} Instead of this rotation, we propose to obtain this four-vector by boosting the four-momentum of Eq.(\ref{4b}). The boost matrix in this case is \widetext \begin{equation}\label{boost2} B_{2} = \pmatrix{1 & 0 & 0 & 0 \cr 0 & \cos\psi & -\sin\psi & 0 \cr 0 & \sin\psi & \cos\psi & 0 \cr 0 & 0 & 0 & 1} \pmatrix{\cosh\lambda & \sinh\lambda & 0 & 0 \cr \sinh\lambda & \cosh\lambda & 0 & 0 \cr 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 1} \pmatrix{1 & 0 & 0 & 0 \cr 0 & \cos\psi & \sin\psi & 0 \cr 0 & -\sin\psi & \cos\psi & 0 \cr 0 & 0 & 0 & 1} , \end{equation} with \begin{equation}\label{psi} \lambda = 2 \tanh^{-1}\left\{[\sin(\theta/2)] \tanh\eta\right\} , \qquad \psi = {\theta \over 2} + {\pi \over 2} . \end{equation} If we carry out the matrix multiplication, \begin{equation}\label{boost22} B_{2} = \pmatrix{\cosh\lambda & -\sin(\theta/2)\sinh\lambda & \cos(\theta/2)\sinh\lambda & 0\cr -\sin(\theta/2)\sinh\lambda & 1 + \sin^{2}(\theta/2)(\cosh\lambda - 1) & -\sin\theta \sinh^{2}(\lambda/2) & 0 \cr \cos(\theta/2) \sinh\lambda & -\sin\theta \sinh^{2}(\lambda/2) & 1 + \cos^{2}(\theta/2)(\cosh\lambda - 1) & 0 \cr 0 & 0 & 0 & 1} . \end{equation} Next, we boost the four-momentum of Eq.(\ref{4c}) to that of Eq.(\ref{4a}). The particle is again at rest. The boost matrix is \begin{equation}\label{boost3} B_{3} = \pmatrix{1 & 0 & 0 & 0 \cr 0 & \cos\theta & -\sin\theta & 0 \cr 0 & \sin\theta & \cos\theta & 0 \cr 0 & 0 & 0 & 1} \pmatrix{\cosh\eta & -\sinh\eta & 0 & 0 \cr -\sinh\eta & \cosh\eta & 0 & 0 \cr 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 1} \pmatrix{1 & 0 & 0 & 0 \cr 0 & \cos\theta & \sin\theta & 0 \cr 0 & -\sin\theta & \cos\theta & 0 \cr 0 & 0 & 0 & 1} . \end{equation} After the matrix multiplication, \begin{equation}\label{boost33} B_{3} = \pmatrix{\cosh\eta & -\cos\theta \sinh\eta & -\sin\theta \sinh\eta & 0 \cr -\cos\theta \sinh\eta & 1 + \cos^{2}\theta (\cosh\eta - 1) & \sin\theta \cos\theta (\cosh\eta - 1) & 0 \cr -\sin\theta \sinh\eta & \sin\theta \cos\theta (\cosh\eta - 1) & 1 + \sin^{2}\theta(\cosh\eta - 1) & 0 \cr 0 & 0 & 0 & 1} . \end{equation} \narrowtext The net result of these transformations is $B_{3}B_{2}B_{1}$. This leaves the initial four-momentum of Eq.(\ref{4a}) invariant. Is it going to be an identity matrix? The answer is No. The result of the matrix multiplications is \begin{equation} W = \pmatrix{1 & 0 & 0 & 0 \cr 0 & \cos\Omega & -\sin\Omega & 0 \cr 0 & \sin\Omega & \cos\Omega & 0 \cr 0 & 0 & 0 & 1} , \end{equation} with \begin{equation}\label{omeg} \Omega = 2\, \sin^{-1}\left\{{(\sin\theta)\sinh^{2}(\eta/2) \over \sqrt{\cosh^{2}\eta - \sinh^{2}\eta \sin^{2}(\theta/2)} }\right\} . \end{equation} This matrix performs a rotation around the $y$ axis and leaves the four-momentum of Eq.(\ref{4a}) invariant. This rotation is an element of Wigner's little group whose transformations leave the four-momentum invariant. This is precisely the Wigner rotation. This relativistic effect manifests itself in atomic spectra as the Thomas precession. Otherwise, the experiments on Wigner rotation in special relativity is largely academic. On the other hand, as was noted in the literature, this effect could be tested in optics laboratories. As for the Stokes parameters, the above four-by-four matrices are directly applicable. Indeed, each four-by-four matrix corresponds to one optical filter applicable to polarized light. In order to see this effect more clearly, let us use the Jones matrix formalism. The two-by-two squeeze matrix corresponding to the boost matrix $B_{1}$ of Eq.(\ref{boost1}) is \begin{equation} S_{1} = \pmatrix{e^{\eta/2} & 0 \cr 0 & e^{-\eta/2} } . \end{equation} The two-by-two squeeze matrix corresponding to the boost matrix of Eq.(\ref{boost2}) is now \widetext \begin{equation} S_{2} = \pmatrix{\cos(\psi/2) & -\sin(\psi/2) \cr \sin(\psi/2) & \cos(\psi/2) } \pmatrix{e^{\lambda/2} & 0 \cr 0 & e^{-\lambda/2}} \pmatrix{\cos(\psi/2) & \sin(\psi/2) \cr -\sin(\psi/2) & \cos(\psi/2)} , \end{equation} where the parameters $\psi$ and $\lambda$ are given in Eq.(\ref{psi}). After the matrix multiplication, $S_{2}$ becomes \begin{equation} S_{2} = \pmatrix{\cosh(\lambda/2) - \sin(\theta/2) \sinh(\lambda/2) & \cos(\theta/2) \sinh(\lambda/2) \cr \cos(\theta/2) \sinh(\lambda/2) & \cosh(\lambda/2) + \sin(\theta/2) \sinh(\lambda/2)} . \end{equation} This is a matrix which squeezes along the direction which makes the angle $(\pi + \theta)/2$ with the $z$ axis. The two-by-two squeeze matrix corresponding to $B_{3}$ of Eq.(\ref{boost3}) is \begin{equation} S_{3} = \pmatrix{\cosh(\eta/2) - \cos\theta \sinh(\eta/2) & - \sin\theta \sinh(\eta/2) \cr - \sin\theta \sinh(\eta/2) & \cosh(\eta/2) + \cos\theta \sinh(\eta/2) } . \end{equation} \narrowtext Now the matrix multiplication $S_{3} S_{2} S_{1}$ corresponds to the closure of the kinematical triangle given in Fig. 1. The result is \begin{equation} S_{3}S_{2}S_{1} = \pmatrix{\cos(\Omega/2) & -\sin(\Omega/2) \cr \sin(\Omega/2) & \cos(\Omega/2) } , \end{equation} where $\Omega$ is given in Eq.(\ref{omeg}). \section{Iwasawa Decompositions}\label{iwasa} In Sec.~\ref{wigrot}, the Lorentz kinematics was based on a massive particle at rest. If the particle is massless, there are no Lorentz frames in which the particle is at rest. Thus, we start with a massless particle whose momentum is in the $z$ direction: \begin{equation}\label{ka} K_{a} = (k, k, 0, 0) , \end{equation} where $k$ is the magnitude of the momentum. We can rotate this four-vector to \begin{equation}\label{kb} K_{b} = (k, -k\sin\alpha, k\cos\alpha, 0) \end{equation} by applying to $K_{a}$ the rotation matrix \begin{equation}\label{r+} R_{+} = \pmatrix{1 & 0 & 0 & 0 \cr 0 & \cos\alpha_{+} & -\sin\alpha_{+} & 0 \cr 0 & \sin\alpha_{+} & \cos\alpha_{+} & 0 \cr 0 & 0 & 0 & 1} , \end{equation} with $\alpha_{+} = \alpha + \pi/2$. \begin{figure}[thb] \centerline{\psfig{figure=fan.ps,angle=-90,height=50mm}} \vspace{3mm} \caption{Two rotations and one Lorentz boost which preserve the four-momentum of a massless particle invariant. The four-momentum $K_{a}$ is rotated to $K_{b}$ by $R_{+}$. It is then boosted to $K_{c}$ by the boost matrix $B$. The rotation matrix $R_{-}$ brings back the four-momentum to $K_{a}$. The initial momentum is along the $z$ direction, and the boost $B$ is made also along the same direction. The rotations are performed around the $y$ axis.} \end{figure} If we rotate $K_{b}$ around the $y$ axis by $-2\alpha$, the resulting four-momentum will be \begin{equation}\label{kc} K_{c} = (k, k\sin\alpha, k\cos\alpha, 0) . \end{equation} It is possible to transform $K_{b}$ to $K_{c}$ by applying to $K_{b}$ the boost matrix \begin{equation}\label{boost4} B = \pmatrix{\cosh\gamma & \sinh\gamma & 0 & 0 \cr \cosh\gamma & \sinh\gamma & 0 & 0 \cr 0 & 0 & 1 & 0 \cr 0 & 0 & 0 & 1} . \end{equation} with \begin{equation}\label{gamal} \sinh\gamma = {2 \sin\alpha \over \cos^{2}\alpha} , \qquad \cosh\gamma = {1 + \sin^{2}\alpha \over \cos^{2}\alpha } . \end{equation} We can transform $K_{c}$ to $K_{a}$ by rotating it around the $y$ axis by $(\alpha - \pi/2)$. The rotation matrix takes the form \begin{equation}\label{r-} R_{-} = \pmatrix{1 & 0 & 0 & 0 \cr 0 & \cos\alpha_{-} & -\sin\alpha_{-} & 0 \cr 0 & \sin\alpha_{-} & \cos\alpha_{-} & 0 \cr 0 & 0 & 0 & 1} , \end{equation} with $\alpha_{-} = \alpha - \pi/2$. Thus, the multiplication of the three matrices, $R_{-} B R_{+}$, gives \begin{equation} T = \pmatrix{ 1 + u^{2}/2 & - u^{2}/2 & u & 0 \cr u^{2}/2 & 1 - u^{2}/2 & u & 0 \cr u & -u & 1 & 0 \cr 0 & 0 & 0 & 1 } , \end{equation} with $$ u = - 2 \tan\alpha . $$ This $T$ matrix plays an important role in studying space-time symmetries of massless particles. If this matrix is applied to the four-momentum $K_{a}$ given in Eq. (\ref{ka}), the four-momentum remains invariant. If this matrix is applied to the electromagnetic four-potential for the plane wave propagating along the $z$ direction with the frequency $k$, the result is a gauge transformation. Again, the above four-by-four matrices are directly applicable to the Stokes parameters. On the other hand, if we are interested in designing optical filters, we need two-by-two representations corresponding to the four-by-four matrices given so far. The two-by-two squeeze matrix corresponding to the boost matrix $B$ of Eq.(\ref{boost4}) is \begin{equation} S = \pmatrix{e^{\gamma/2} & 0 \cr 0 & e^{-\gamma/2}} , \end{equation} while the two-by-two matrices corresponding to $R_{+}$ of Eq.(\ref{r+}) and $R_{-}$ of (\ref{r-}) are \begin{equation} R_{\pm} = \pmatrix{\cos(\alpha_{\pm}/2) & -\sin(\alpha_{\pm}/2) \cr \sin(\alpha_{\pm}/2) & \cos(\alpha_{\pm}/2) } , \end{equation} where $\alpha_{+}$ and $\alpha_{-}$ are given in Eq.(\ref{r+}) and Eq.(\ref{r-}) respectively. They satisfy the equations $$ \alpha_{+} + \alpha_{-} = 2\alpha, \qquad \alpha_{+} - \alpha_{-} = \pi . $$ The relation between $\gamma$ and $\alpha$ given in Eq.(\ref{gamal}) can also be written as $\cosh(\gamma/2) = 1/\cos\alpha$, which is more useful for carrying out the two-by-two matrix algebra. The matrix multiplication $R_{-} S R_{+} $ leads to \begin{equation}\label{iwa1} T = R_{-}SR_{+} = \pmatrix{1 & -2\,\tan\alpha \cr 0 & 1} . \end{equation} Conversely, we can write the \begin{equation} \pmatrix{1 & -2\,\tan\alpha \cr 0 & 1} = R_{-}SR_{+} . \end{equation} The $T$ matrix can be decomposed into rotation and squeeze matrices. This possibility is called the Iwasawa decomposition. In the present case, $T$ of Eq.(\ref{iwa1}) can also be written as \begin{eqnarray} &{}& T = R_{-}S \left\{\left(R_{-}\right)^{-1} R_{-1}\right\} R_{+} \nonumber \\[2ex] &{}& \hspace{2ex} = \left\{R_{-}S\left(R_{-}\right)^{-1}\right\} \left(R_{-1} R_{+}\right) \end{eqnarray} The matrix chain $R_{-}S\left(R_{-}\right)^{-1}$ is one squeeze matrix whose squeeze axis is rotated by $\alpha_{-}/2$, and the matrix product $R_{-1}R_{+}$ becomes one rotation matrix. The result is \begin{equation} T = S(\alpha_{-}) R(2\alpha) , \end{equation} with \widetext \begin{eqnarray} &{}& S(\alpha_{-}) = \pmatrix{\cosh(\gamma/2) + \cos\alpha_{-}\,\sinh(\gamma/2) & \sin\alpha_{-}\,\sinh(\gamma/2) \cr \sin\alpha_{-}\,\sinh(\gamma/2) & \cosh(\gamma/2) - \cos\alpha_{-}\,\sinh(\gamma/2) } , \nonumber \\[2ex] &{}& R(2\alpha) = \pmatrix{\cos\alpha & -\sin\alpha \cr \sin\alpha & \cos\alpha}. \end{eqnarray} \narrowtext It is indeed gratifying to note that the $T$ matrix can be decomposed into one rotation and one squeeze matrix. The squeeze is made along the direction which makes an angle of $\alpha_{-}/2$ or $-(\pi/2 - \alpha)/2$ with the $z$ axis. The angle $\alpha$ is smaller than $\pi/2$. We have discussed in our earlier papers~\cite{hkn97josa,hkn97pre} optical filters with the property given in Eq.(\ref{iwa1}). We said there that the filters with this property can be produced from an infinite number of infinitely thin filters. This argument was based on the theory of Lie groups where transformations are generated by infinitesimal generators. This may be possible these days, but the method presented in this paper is far more practical. We need only two filters~\cite{hks86jm}. We are able to achieve this improvement because we used here the analogy between polarization optics and Lorentz transformations who share the same mathematical framework. \section*{Concluding Remarks} In this paper, we noted first that both the Wigner rotation and the Iwasawa decomposition come from Wigner's little group whose transformations leave the four-momentum of a given particle invariant. Since the Lorentz group is applicable also to the Jones vector and the Stokes parameters, it is possible to construct corresponding transformations in polarization optics. We have shown that both the Wigner rotation and the Iwasawa decomposition can be realized in optics laboratories. The matrix of Eq.(\ref{shear1}) performs a shear transformation when applied to a two-dimensional object, and has a long history in physics and engineering. It also has a history in mathematics. The fact that a shear can be decomposed into a squeeze and rotations is known as the Iwasawa decomposition~\cite{iwa49}. Among the many interesting applications of shear transformations, there is a special class of squeezed states of photons or phonons having the symmetry of shear~\cite{ky92}. The wave-packet spread can be formulated in terms of shear transformations~\cite{kiwi90aj}. As we can see from this paper, a set of shear transformations can be formulated as a subset of Lorentz transformations. This set plays an important role in understanding internal space-time symmetry of massless particles, such as gauge transformation and neutrino polarizations~\cite{knp86,wein64,hks82}. \section*{Acknowledgments} We would like to thank A. E. Bak for bringing to our attention to his early works with C. S. Brown on applications of the Lorentz group to polarization optics. We are also grateful to S. Baskal for telling us about the recent paper by Simon and Mukunda on the Iwasawa decomposition~\cite{simon98}.
\section{Introduction} Three gauge coupling constants of the standard-model gauge interactions determined at the weak scale $\mu \simeq m_Z$ strongly suggest supersymmetric (SUSY) grand unification (GUT) of all gauge groups SU$(3)_C$, SU$(2)_L$ and U$(1)_Y$ at a very high energy scale, $M_{GUT} \simeq 2\times10^{16}$GeV. If the gravitational scale $M_* \simeq 2.4\times 10^{18}$GeV is the fundamental cut-off scale of field-theory description of nature, it implies the presence of a mild hierarchy $M_{GUT}/M_* \simeq 10^{-2}$. It is very natural to consider such a hierarchy is a low-energy manifestation of more fundamental theory. There is a mechanism which generates such a mild hierarchy. It is an anomalous U$(1)_A$ gauge symmetry whose anomalies are cancelled by Green-Schwarz mechanism~\cite{GS}. The anomalous U$(1)_A$ gauge symmetry often appears in low energy effective theories of string theories~\cite{anom}. Since the sum of U$(1)_A$ charges is non-zero, $\mathop{\rm Tr}\nolimits Q_A \neq 0$, in the anomalous U$(1)_A$ gauge symmetry, Fayet-Iliopoulos (FI) term is induced at one-loop level~\cite{FI}. In pertubative heterotic string theories it is given by \begin{eqnarray} {\cal L_{\rm FI}}= g_A^2\frac{\mathop{\rm Tr}\nolimits Q_A}{192\pi^2} M_*^2 D_A \equiv \xi^2 D_A. \end{eqnarray} If all fields have no VEVs, D-term is non-zero so that SUSY is broken. Instead, one usually assumes that some field $\phi$ has VEV to restore SUSY and define the normalization of U$(1)_A$ charge so that $\phi$ has a U$(1)_A$ charge $-1$. In this normalization $\VEV{\phi}=\xi$ and a mild hierarchy is generated as $\VEV{\phi}/M_* \equiv \epsilon \sim {\cal O}(10^{-1})$. In this paper, we discuss the anomalous U$(1)_A$ gauge symmetry as a mechanism of generating the mild hierarchy $M_{GUT}/M_*$.\footnote{ There are several attempts to generate the hierarchy $M_{GUT}/M_*$ \cite{GUTscale}.} We first show that the standard SUSY GUT based on a simple group cannot be realized unless some parameters are ``tuned'' and that models with product gauge groups are preferred instead. Here, we construct the ``$R$-invariant natural unification'' model~\cite{R_inv} based on gauge groups SU$(5)_{GUT}\times$U$(3)_H$, in which the symmetry breaking scale $\simeq 2\times 10^{16}$GeV is induced through the anomalous U$(1)_A$ breaking. We also show that this model has various phenomenologically desirable features that the doublet-triplet splitting problem is solved, proton decays induced by dimension four and five operators~\cite{P_decay,dim_five} are suppressed by $R$-symmetry and $m_s \neq m_\mu$ without violating the GUT relation, $m_b = m_\tau$. \section{Standard SUSY GUT models} \label{sec:2} In this section, we show that it is difficult to generate the GUT scale through anomalous U$(1)_A$ breaking in the standard SUSY SU$(5)$ GUT models. In order to break SU$(5)$ down to SU$(3)_C\times$SU$(2)_L\times$U$(1)_Y$, a Higgs field $\Sigma$ in adjoint representation must have an appropriate vacuum expectation value (VEV), $\VEV{\Sigma}/M_* \sim {\cal O}(10^{-2})$. Since there is no scale but the breaking scale of the anomalous U$(1)_A$ gauge symmetry, VEV of a field with a negative U$(1)_A$ charge $-q$ is controlled by its U$(1)_A$ charge and is of order $\epsilon^q M_*$ in general. That is, the field $\Sigma$ is required to have a negative U$(1)_A$ charge $-2$. Now, we give a simple example for generating the GUT scale. Since $\Sigma$ has U$(1)_A$ charge $-2$, we have to introduce another field $\Sigma'$ in adjoint representation with a positive U$(1)_A$ charge $4$ to have a nontrivial superpotential. Consider the following superpotential,\footnote { The last term is necessary to give masses for $(\bf 3,2^\star)$ and $(\bf 3^\star,2)$ components in $\Sigma'$, since $(\bf 3,2^\star)$ and $(\bf 3^\star,2)$ in $\Sigma$ are absorbed into broken gauge bosons. } \begin{eqnarray} W= \kappa \Sigma' \left\{ \left(\frac{\phi}{M_*}\right)^2 M_* \Sigma + \kappa' \Sigma^2\right\} + \kappa'' \left(\frac{\phi}{M_*}\right)^8 M_* {\Sigma'}^2, \label{standard} \end{eqnarray} where parameters, $\kappa$'s, are assumed to be of order unity. There is a desirable SUSY vacuum which breaks SU$(5)$ down to SU$(3)_C\times$SU$(2)_L\times$U$(1)_Y$, \begin{eqnarray} \VEV{\Sigma}/M_* \simeq (\epsilon^2/\kappa'){\rm diag}(2,2,2,-3,-3). \end{eqnarray} The generated GUT scale is indeed ${\cal O}(\epsilon^2 M_*)$ and the masses for broken gauge bosons, $M_V$, and those for $(\bf 8,1)$, $(\bf 1,3)$ components in $\Sigma$ and $\Sigma'$, $M_{\Sigma}$, are of order $\epsilon^2 M_*$, while masses for $(\bf 3,2^\star)$ and $(\bf 3^\star,2)$ components in $\Sigma'$, $M_X$, of order $\epsilon^{8} M_* \ll M_{GUT}$. However, the disparity of the masses for various fields which belong to the same representation under SU$(5)$ disturbs the unification of gauge coupling constants. We show it explicitly by considering the following combination of standard-model gauge coupling constants. From renormalization group equations (RGEs) at one-loop level, we get~\cite{HMY} \begin{eqnarray} \left(\frac{5}{\alpha_1}-\frac{3}{\alpha_2}-\frac{2}{\alpha_3} \right)(m_Z) &=& \frac{8}{2\pi} \ln \left(\frac{m_{SUSY}}{m_Z}\right) +\frac{12}{2\pi}\ln \left(\frac{M_V^2 M^2_\Sigma}{m_Z^3 M_X}\right). \end{eqnarray} The observed values for standard-model gauge couplings constrain the combination of masses in the last term as \begin{eqnarray} \left(\frac{M_V^2 M_\Sigma^2}{M_X}\right)^{1/3} \simeq M_{GUT} \simeq 2\times 10^{16}{\rm GeV}. \label{mass_relation} \end{eqnarray} On the other hand, $(M_V^2 M_\Sigma^2/M_X)^{1/3}$ is almost equal to $M_*$ in the present model, since factors $\epsilon$ in $M_V$, $M_\Sigma$ and $M_X$ are cancelled out. It implies that the standard-model gauge coupling constants do not unify at the high-energy scale due to the mass splitting among various fields in the SU(5) multiplet $\Sigma'$. We show, in Appendix, that such a situation cannot be avoided by adding fields in various representations. One possibility for avoiding it is to ``tune'' some parameters. For example, if we assume that $\kappa$ is not of order unity but ${\cal O}(10^{-3})$ in the superpotential Eq.~(\ref{standard}), $M_\Sigma \simeq 10^{-3} \epsilon^2 M_*$ and the combination $(M_V^2 M_\Sigma^2/M_X)^{1/3}$ is $10^{-2} M_* \simeq 10^{16}$GeV. In this paper we pursue another possibility in which no ``tuning'' is required. It implies that we discard a ``simple'' idea that the standard-model gauge groups unify to one simple gauge symmetry. That is, we need GUT models with product gauge groups. As such models, we consider ``$R$-invariant natural unification'' model with a gauge group SU$(5)\times$U$(3)_H$~\cite{R_inv}, since the anomalous U$(1)_A$ gauge symmetry is naturally embedded in this model. Moreover, it has various desirable features in view of phenomenology. \section{$R$-invariant unification model} Let us first discuss briefly the ``$R$-invariant natural unification'' model which is based on a SUSY SU$(5)_{GUT}\times$U$(3)_H$ gauge theory with an unbroken $R$ symmetry~\cite{R_inv}. The SU$(5)_{GUT}$ is the usual GUT gauge group and its coupling constant is in a perturbative regime, while the U$(3)_H$ is a hypercolor gauge group whose coupling is strong at the GUT scale. Here, GUT means unification of all gauge groups, SU$(3)_C$, SU$(2)_L$ and U$(1)_Y$, in the standard model. The quarks and leptons obey the usual transformation law under the GUT group SU$(5)_{GUT}$ and they are all singlets of the hypercolor group U$(3)_H$. A pair of Higgs multiplets $H_r$ and $\bar H^r$ ($r=1,\cdots,5$) transform as ${\bf 5}+{\bf 5^\star}$ under the SU$(5)_{GUT}$ and as singlets under the U$(3)_H$. All the matter multiplets introduced so far are the same as in the minimal SUSY SU$(5)$ model. We now introduce six pairs of hyperquarks $Q^\rho_\alpha$ and $\bar Q^\alpha_\rho$ ($\alpha=1,\cdots,3$; $\rho=1,\cdots,6$) which transform as ${\bf 3}$ and ${\bf 3^\star}$ under the hypercolor SU$(3)_H$ and have U$(1)_H$ charges $1$ and $-1$, respectively (SU$(3)_H\times$U$(1)_H\equiv$U$(3)_H$). The first five pairs of $Q_\alpha^r$ and $\bar Q^\alpha_r$ ($r=1,\cdots,5$) belong to ${\bf 5^\star}$ and ${\bf 5}$ of SU$(5)_{GUT}$ and the last pair of $Q_\alpha^6$ and $\bar Q^\alpha_6$ are singlets of SU$(5)_{GUT}$. To cause a breaking of the total gauge group SU$(5)_{GUT}\times$U$(3)_H$ down to the standard-model gauge groups, SU$(3)_C\times$SU$(2)_L\times$U$(1)_Y$, we introduce another chiral multiplet $X^\alpha_\beta$ coupled to $Q_\alpha^r$ and $\bar Q^\alpha_r$, which is an adjoint representation of the SU$(3)_H$. Since $Q_\alpha^r$ and $\bar Q^\alpha_r$ are supposed to have VEV of order of the GUT scale, they must be trivial representations of the $R$ symmetry U$(1)_R$ \cite{R_inv}, and hence the $X^\alpha_\beta$ carry $R$-charge two. $R$ charges for all matter multiplets besides quark and lepton multiplets are given in Table~\ref{charge}. To suppress unwanted nonrenormalizable interactions in superpotential we further impose an axial U$(1)_A$ symmetry, under which the hyperquarks $Q_\alpha^r$ and $\bar Q^\alpha_r$ and the adjoint $X^\alpha_\beta$ transform as \begin{eqnarray} Q_\alpha^r \to e^{i\theta} Q_\alpha^r, \hspace*{5mm} \bar Q^\alpha_r \to e^{i\theta} \bar Q^\alpha_r, \hspace*{5mm} X^\alpha_\beta \to e^{-2i\theta} X^\alpha_\beta. \end{eqnarray} U$(1)_A$ charges for all matter multiplets besides quark and lepton multiplets are also given in Table~\ref{charge}. \begin{table}[h] \begin{eqnarray} \begin{array}{|c|cc|cc|c|cc||cc|} \hline & Q_\alpha^r & \bar Q^\alpha_r & Q^6_\alpha & \bar Q^\alpha_6 & X^\alpha_\beta & H_r & \bar H^r & \phi & \chi\\ \hline {\rm U}(1)_R & 0 & 0 & 2 & 2 & 2 & 0 & 0 & 0 & 2\\ {\rm U}(1)_A & -2 & -2 & 2 & 2 & 4 & 0 & 0 & -1 & 4\\ \hline \end{array} \nonumber \end{eqnarray} \caption{Charge assignments in the Higgs sector.} \label{charge} \end{table} \\ Then, we have a superpotential\footnote { The imposition of U$(1)_A$ and U$(1)_R$ alone allows nonrenormalizable interactions which include arbitrary powers of $H_r \bar H^r /M_*^2$. } \begin{eqnarray} W=\lambda Q_\alpha^r \bar Q^\beta_r X^\alpha_\beta +h Q^r_\alpha \bar Q^\alpha_6 H_r +\bar h Q_\alpha^6 \bar Q^\alpha_r \bar H^r. \label{Rinvariant} \end{eqnarray} Notice that the $R$ charges for $H_r$ and $\bar H^r$ are vanishing and those for $Q_\alpha^6$ and $\bar Q^\alpha_6$ are two. As shown in Ref.~\cite{R_inv} we have the desirable vacua; \begin{eqnarray} \BigVEV{X^\alpha_\beta}=\BigVEV{Q_\alpha^6}=\BigVEV{\bar Q^\alpha_6}=0, \hspace*{3mm} \BigVEV{H_r}=\BigVEV{\bar H^r}=0, \hspace*{3mm} \BigVEV{Q_\alpha^r} = v \delta_\alpha^r, \hspace*{3mm} \BigVEV{\bar Q^\alpha_r} = v \delta^\alpha_r. \label{vacuum} \end{eqnarray} For $v\neq 0$, the total gauge group, SU$(5)_{GUT}\times$U$(3)_H$, is broken down to SU$(3)_C$ $\times$SU$(2)_L\times$U$(1)_Y$ and hence the $v$ corresponds to the GUT scale. In these vacua, the color SU$(3)_C$ is an unbroken linear combination of an SU$(3)$ subgroup of the SU$(5)_{GUT}$ and the hypercolor SU$(3)_H$, and the hypercharge U$(1)_Y$ is that of a U$(1)$ subgroup of the SU$(5)_{GUT}$ and the strong U$(1)_H$. Thus, the gauge coupling constants $\alpha_3$, $\alpha_2$ and $\alpha_1$ of SU$(3)_C\times$SU$(2)_L\times$U$(1)_Y$ are given by \begin{eqnarray} \alpha_3 \simeq \frac{\alpha_{\rm GUT}} {1+\alpha_{\rm GUT}/\alpha_{3H}}, \hspace*{3mm} \alpha_2 = \alpha_{\rm GUT}, \hspace*{3mm} \alpha_1 \simeq \frac{\alpha_{\rm GUT}} {1+\frac{1}{15}\alpha_{\rm GUT}/\alpha_{1H}}, \end{eqnarray} where $\alpha_{3H}$ and $\alpha_{1H}$ are gauge coupling constants for the hypercolor SU$(3)_H$ and U$(1)_H$, respectively. We see that the unification of three gauge coupling constants, $\alpha_3$, $\alpha_2$ and $\alpha_1$, is practically achieved in a strong coupling regime of the hypercolor gauge interactions, that is, $\alpha_{3H}$ and $\alpha_{1H} \gg {\cal O}(1)$. Interesting is that the color triplets $H_a$ and $\bar H^a$ ($a=1,\cdots,3$) acquire masses of order $v$ together with the sixth hyperquarks $\bar Q^\alpha_6$ and $Q^6_\alpha$ in the vacua Eq.~(\ref{vacuum}), while the weak doublets $H_l$ and $\bar H^l$ ($l=4,5$) remain massless since there are no partners for them to form $R$-invariant masses. The masslessness for these doublets is guaranteed by the unbroken U$(1)_R$ symmetry.\footnote { The U$(1)_R$ symmetry is broken down to a discrete subgroup $Z_{4R}$ by the hypercolor SU$(3)_H$ anomaly. However, the unbroken subgroup $Z_{4R}$ is sufficient to keep the Higgs doublets massless. } So far, the GUT scale $v$ is undetermined because of the presence of a flat direction in the present vacua. We identify the axial U$(1)_A$ with the anomalous U$(1)_A$ gauge symmetry and show that the GUT scale $v$ is determined by the breaking scale of the anomalous U$(1)_A$. To obtain the desirable value for $v\simeq 2\times10^{16}$GeV we introduce a singlet $\chi$ with a U$(1)_A$ charge $+4$ and the following superpotential \begin{eqnarray} W = k Q_\alpha^r \bar Q^\alpha_r \chi + \frac{k'}{M_*^2} \phi^4 \chi. \end{eqnarray} Here, we have assumed $R$ charges for $\chi$ and $\phi$ are two and zero, respectively. Then, we get the GUT scale with a correct magnitude \begin{eqnarray} v \simeq \sqrt{\frac{k'}{k}} \frac{\VEV{\phi}^2}{M_*} \sim {\cal O}(10^{16}{\rm GeV}), \end{eqnarray} for $k,k' \sim {\cal O}(1)$. \section{Quark and lepton mass matrices} Let us turn to discuss quark and lepton mass matrices. It is very attractive to use the above $\phi$ field generating hierarchies in quark and lepton mass matrices~\cite{FN,Yukawa}. We tentatively assume U$(1)_A$ charges for quark and lepton multiplets, ${\bf 5}^\star_i$ and ${\bf 10}_i$ ($i=1,\cdots,3$) as shown in Table~\ref{FN_charge}~\cite{large} where U$(1)_R$ charges are also given. There are two cases for the assignment of $U(1)_A$ charges corresponding that $\tan\beta$ is small or large, where $\tan\beta$ is the ratio of the VEVs of the Higgs ($\tan\beta=\VEV{H}/\langle\bar H\rangle$). That is, $\tau=1$ for small $\tan\beta \sim {\cal O}(1)$ and $\tau= 0$ for large $\tan\beta\sim {\cal O}(1/\epsilon)$. \begin{table}[h] \begin{eqnarray} \begin{array}{|c|ccc|ccc||cc|} \hline & {\bf 10}_1 & {\bf 10}_2 & {\bf 10}_3 & {\bf 5}^\star_1 & {\bf 5}^\star_2 & {\bf 5}^\star_3 & H'({\bf 45}) & \bar H'({\bf 45^\star})\\ \hline {\rm U}(1)_R & 1 & 1 & 1 & 1 & 1 & 1 & 0 & 2\\ {\rm U}(1)_A & 2 & 1 & 0 & \tau+1 & \tau & \tau & -\tau-1 & 4\\ \hline \end{array} \nonumber \end{eqnarray} \caption{Charge assignments for the matter fields.} \label{FN_charge} \end{table} Because of the U$(1)_A$ invariance and $\VEV{\phi}/M_* = \epsilon$, we obtain the Yukawa matrix for up-type quarks as \begin{eqnarray} \widehat{\lambda}_u \simeq \left( \begin{array}{ccc} \epsilon^4 & \epsilon^3 & \epsilon^2 \\ \epsilon^3 & \epsilon^2 & \epsilon \\ \epsilon^2 & \epsilon & 1 \end{array} \right), \end{eqnarray} and those for down-type quarks and charged leptons as \begin{eqnarray} \widehat{\lambda}_d =\widehat{\lambda}_l \simeq \epsilon^{\tau} \left( \begin{array}{ccc} \epsilon^3 & \epsilon^2 & \epsilon^2\\ \epsilon^2 & \epsilon & \epsilon\\ \epsilon & 1 & 1 \end{array} \right). \label{wrong_relation} \end{eqnarray} Each element in the matrices has an undetermined coefficient of order unity. If $\epsilon \simeq 1/20$, the above mass matrices are roughly consistent with observations except for the unwanted GUT relations \begin{eqnarray} m_s=m_\mu \;\;\mbox{and}\;\; m_d=m_e. \end{eqnarray} It is only the GUT condensation, $\BigVEV{Q^r_\alpha}=v \delta^r_\alpha$ and $\BigVEV{\bar Q^\alpha_r}=v \delta^\alpha_r$, that can break these unwanted GUT relations. In the above model, however, a possible operator is only \begin{eqnarray} \frac{1}{M_*^2} {\bf 5}^\star_1 {\bf 10}_1 \bar H \VEV{Q_\alpha \bar Q^\alpha} \end{eqnarray} for $\tau=1$ and nothing for $\tau=0$. Thus, it cannot lead to sufficient contributions to the quark and lepton mass matrices. A way to solve this problem is to introduce another pair of Higgs multiplets $H'$ and $\bar H'$ transforming as ${\bf 45}$ and ${\bf 45^\star}$ under the SU$(5)_{\rm GUT}$.\footnote{ Another way to solve the problem is to lower the cut-off scale, $M_*$, and change the U$(1)_A$ charge assignments in Table~\ref{charge}. } We assign U$(1)_A$ charge for ${\bf 45}$ so that the following Yukawa coupling is possible, \begin{eqnarray} {\bf 5}^\star_2 {\bf 10}_2 H'({\bf 45}). \end{eqnarray} On the other hand, since the existence of ${\bf 5}^\star_3 {\bf 10}_3 H'({\bf 45})$ would break the successful GUT relation $m_b=m_\tau$, U$(1)_A$ and U$(1)_R$ charges for $H'({\bf 45})$ are determined as in Table~\ref{FN_charge}. Now, massless Higgs doublet $\bar H_f$ is a linear combination of doublets, $\bar H_f({\bf 5}^\star)$ in $\bar H({\bf 5}^\star)$ and $H'_f({\bf 45})$ in $H'({\bf 45})$. Mixing angle $\theta$ is determined by the following superpotential \begin{eqnarray} W= H'({\bf 45}) \bar H'({\bf 45}^\star) \frac{\phi^2}{M_*} \;(\;{\rm or}\;\frac{\phi^3}{M_*^2}\;)\; + \frac{1}{M_*} \bar H({\bf 5}^\star) \bar H'({\bf 45}^\star) \VEV{Q_\alpha \bar Q^\alpha}, \end{eqnarray} for $\tau =1$ (or $\tau =0$). Here, we have taken U$(1)_A$ and U$(1)_R$ charges for $\bar H'({\bf 45^\star})$ as in Table~\ref{FN_charge}. We get $\theta \simeq \VEV{\phi}^2/M_*^2$ ( or $\VEV{\phi}/M_*$ ) where $\theta$ is defined as \begin{eqnarray} \bar H_f=\bar H_f({\bf 5}^\star)\cos\theta +H'_f({\bf 45}) \sin\theta. \end{eqnarray} This gives \begin{eqnarray} \VEV{H'_f({\bf 45})} & \simeq & \epsilon^2 \;({\rm or}\; \epsilon) \VEV{\bar H_f},\\ \VEV{\bar H_f({\bf 5}^\star)}& \simeq & \VEV{\bar H_f}, \end{eqnarray} which leads to a required magnitude of the violation of the unwanted GUT relations, $m_s = m_\mu$ and $m_d = m_e$, as follows: \begin{eqnarray} \delta \widehat{\lambda}_d \simeq (-2)\times\epsilon^{\tau}\left( \begin{array}{ccc} \epsilon^3 & \epsilon^2 & \epsilon^2\\ \epsilon^2 & \epsilon & \epsilon \\ \epsilon & 0 & 0 \end{array} \right), \hspace*{1cm} \delta \widehat{\lambda}_l \simeq 3\times\epsilon^{\tau}\left( \begin{array}{ccc} \epsilon^3 & \epsilon^2 & \epsilon^2\\ \epsilon^2 & \epsilon & \epsilon \\ \epsilon & 0 & 0 \end{array} \right). \end{eqnarray} Note that the contribution from the GUT condensation for down-type quarks has an opposite sign to that for charged leptons. It shows that we can get the relation $m_s < m_\mu$ when the signs of $(2,2)$ elements in $\widehat{\lambda}_d$ and $\delta\widehat{\lambda}_d$ are opposite. If the GUT condensation is not taken into account, Cabbibo angle is given by $\epsilon$, which is smaller than the observed value $0.22$. However, due to the cancellation in (2,2) element in the mass matrix for down-type quarks, we can get the desirable value. \section{Anomaly cancellations} In this section, we discuss the cancellation of the U$(1)_A$ mixed anomalies, $C_G$, with the gauge groups $G$ ($G = $SU$(5)_{GUT}$, SU$(3)_H$ and U$(1)_H$). In the above model, we obtain $C_G$ as \begin{eqnarray} \begin{array}{lcc} C_{{\rm SU}(5)_{GUT}}&= & 35-\frac{21}{2}\tau, \\ C_{{\rm SU}(3)_H} &= & 4, \\ C_{{\rm U}(1)_H} &= & -48. \end{array} \label{anomaly} \end{eqnarray} We find that U$(1)_A$ has the negative anomaly coefficient $C_{{\rm U}(1)_H}$, which may not be expected in usual heterotic string theories.\footnote { Apparently, it seems possible that we add U$(1)_H$-charged fields with positive U$(1)_A$ charges so that the anomaly coefficient $C_{{\rm U}(1)_H}$ becomes positive. In that case, however, the gauge coupling constant for U$(1)_H$ blows up below $M_*$. } In the heterotic string case there is the {\it universal} relation to the mixed anomaly coefficients as follows: \begin{eqnarray} \frac{C_G}{k_G} = {\rm const.}, \end{eqnarray} where $k_G$ is the Kac-Moody level, since only the dilaton field $S$ plays a role in anomaly cancellations or, in other words, there is only one antisymmetric tensor field, $B_{\mu\nu}$ (Green-Schwarz mechanism)~\cite{GS}. However, we can consider the case where various moduli fields other than the dilaton cancel these anomalies (``generalized'' Green-Schwarz mechanism). Indeed, such a situation is realized in Type I and Type IIB string theories with orientifold compactifications~\cite{Type12}. The anomalous U(1) gauge symmetries in these theories have been studied recently and it has been revealed that several anomalous U$(1)$ gauge symmetries exist corresponding to the existence of several antisymmetric tensors in twisted sectors. The ``generalized'' Green-Schwarz mechanism is illustrated as follows. In this mechanism there are many moduli fields $M_k$ which are coupled to field strength superfields as \begin{eqnarray} {\cal L}_{gauge} = \int d^2\theta \left( k_G S + \sum_k c_G^k M_k \right) W_G^\alpha W_{G\alpha}, \end{eqnarray} and these moduli fields transform under the U$(1)_A$ transformation, $V_A \rightarrow V_A + \frac i2 (\Lambda - \Lambda^{\dagger})$, as \begin{eqnarray} M_k \rightarrow M_k + i\delta_k \Lambda, \end{eqnarray} where $c_G^k$ and $\delta_k$ are model-dependent constants. This gives the {\it nonuniversal} relation to the mixed anomaly coefficients \begin{eqnarray} C_G = 4\pi^2 \sum_k c_G^k \delta_k. \end{eqnarray} It shows that the three anomalies in Eqs.~(\ref{anomaly}) can be cancelled by the non-linear transformation of the moduli fields $M_k$. The origin of FI term is also different from that in the case of perturbative heterotic string theories. FI term is not generated at one-loop level but is given by the VEVs of moduli fields $\VEV{M_k}$ at tree level as \begin{eqnarray} {\cal L}_{FI} \simeq -\sum_k \delta_k \VEV{M_k} M_*^2 D_A \equiv \xi^2 D_A. \end{eqnarray} Although the VEVs of the moduli fields $\VEV{M_k}$ are undetermined in the present framework, it seems natural to consider that the resulting FI term is one order smaller than $M_*$, i.e. $\epsilon = \xi/M_* \sim 10^{-1}$. Then, the conclusions in the previous sections are not affected. \section{Summary and Discussions} We have found that the idea that the GUT scale is generated through the anomalous U$(1)_A$ breaking is compatible with the ``$R$-invariant natural unification'' model with gauge groups SU$(5)_{GUT}\times$U$(3)_H$. The Higgs field which breaks the GUT group has an anomalous U$(1)_A$ charge and the unwanted GUT relations, $m_s = m_\mu$ and $m_d = m_e$, can be avoided in a simple manner without violating the successful GUT relation, $m_b=m_\tau$. Also, the doublet-triplet splitting problem is solved naturally in the present model. As for proton decay, the vacuum in this model preserves the U$(1)_R$ invariance and $R$-parity is included in the U$(1)_R$ symmetry. Therefore, the dimension four and five operators contributing to the proton decay~\cite{P_decay,dim_five} are suppressed. Instead, the process induced by the dimension six operators is dominant. It is a crucial difference from the usual GUT models that the SU$(3)_C$ is an unbroken linear combination of an SU$(3)$ subgroup of the SU$(5)_{GUT}$ and the hypercolor SU$(3)_H$. It predicts a smaller value of the strong coupling constant $\alpha_3$. The current experimental values of $\alpha_3$~\cite{PDG} are consistent with the present model~\cite{gaugino}. We conclude this paper with a comment on neutrino masses~\cite{neutrino}. Since neutrino masses are written by the effective operators, $({\bf 5}^\star H)^2$, they are independent of U$(1)_A$ charges of right handed neutrinos and are given by,~\cite{large} \begin{eqnarray} \widehat{m}_\nu \simeq \epsilon^{2\tau} \left( \begin{array}{ccc} \epsilon^2& \epsilon &\epsilon\\ \epsilon & 1 & 1\\ \epsilon & 1 & 1 \end{array} \right) \frac{m^2_t}{M_*}. \label{neutrino_mass} \end{eqnarray} This mass matrix implies the large angle between the second and third generations and either small or large angle between the first and second ones. However, the mass of tau neutrino is of order of $10^{-5}$eV for $\tan\beta \sim {\cal O}(\epsilon^{-1})$ or $10^{-7}$eV for $\tan\beta \sim {\cal O}(1)$. It is smaller by several orders than the scale indicated by the atmospheric neutrino oscillation~\cite{SuperK}. There are two ways to generate a desirable mass scale. One way is to introduce the U$(1)_{B-L}$ breaking scale as well as the anomalous U$(1)_A$ breaking scale. The other is to extend the assignment of U$(1)_A$ charges given in Table~\ref{charge} and Table~\ref{FN_charge}. This can be done because the assignment of U$(1)_A$ charges is not uniquely determined by the following conditions: $Q\bar Q$ has charge $-4$, the mixed anomaly U$(1)_H$-U$(1)_A^2$ vanishes, $H \bar H$ has charge $0$ in order to use the Giudice-Masiero mechanism for $\mu$ and $B\mu$ terms~\cite{GM}, and the superpotential (\ref{Rinvariant}) is U$(1)_A$ invariant. The U$(1)_A$ charges for various fields are given in Table~\ref{neutrino}. \begin{table}[h] \begin{eqnarray} \begin{array}{|c|cc|cc|c|cc||cc|} \hline & \makebox[14mm]{$Q_\alpha^r$} & \makebox[14mm]{$\bar Q^\alpha_r$} & \makebox[14mm]{$Q^6_\alpha$} & \makebox[14mm]{$\bar Q^\alpha_6$} & \makebox[9.6mm]{$X^\alpha_\beta$} & \makebox[10mm]{$H_r$} & \makebox[10mm]{$\bar H^r$} & \makebox[9mm]{$\phi$} & \makebox[9mm]{$\chi$}\\ \hline {\rm U}(1)_A & -2-p & -2+p & 2-5p & 2+5p & 4 & -4p & 4p & -1 & 4\\ \hline \end{array} \nonumber\\ \begin{array}{|c|ccc|ccc||cc|} \hline & {\bf 10}_1 & {\bf 10}_2 & {\bf 10}_3 & {\bf 5}^\star_1 & {\bf 5}^\star_2 & {\bf 5}^\star_3 & H'({\bf 45}) & \bar H'({\bf 45^\star})\\ \hline {\rm U}(1)_A & 2+2p & 1+2p & 2p & \tau+1-6p & \tau-6p & \tau-6p & -\tau-1+4p & 4-4p\\ \hline \end{array} \nonumber \end{eqnarray} \caption{Assignment of U$(1)_A$ charges. $p$ is an arbitrary parameter. Taking $p=0$ corresponds to the assignment given in Table~\ref{charge} and \ref{FN_charge}.} \label{neutrino} \end{table} For this charge assignment we are to replace the coefficient $\epsilon^{2\tau}$ with $\epsilon^{-20p+2\tau}$ in Eq.~(\ref{neutrino_mass}). If we take $p=1/4$ and $\tau=1$, we can get desirable tau neutrino mass, $m_{\nu_\tau} \simeq \epsilon^{-3} m_t^2/M_* \sim 0.1$eV. With this charge assignment, dimension four baryon-number-violating operators are completely suppressed by an unbroken discrete gauge symmetry. It is also desirable that no other scales are required to understand the neutrino masses. \section*{Acknowledgments} Y.N. thanks the Japan Society for the Promotion of Science for financial support. This work is supported in part by the Grant-in-Aid, Priority Area ``Supersymmetry and Unified Theory of Elementary Particles''(\#707).
\section{INTRODUCTION} One requirement for measuring the Hubble constant using SNe Ia as standard candles is a sample of well-measured distant supernovae. They should be distant enough that their measured redshifts are dominated by the Hubble flow, but not so distant that the still-uncertain dynamics associated with deceleration due to the mass density of the universe and possible acceleration due to conjectured repulsive forces are important. A broad range of redshift $z$ from 0.01 to 0.2 is suitable, with $z$ about 0.05 being optimum. For this purpose the carefully measured and uniformly analyzed Cal\'an-Tololo collection of 29 SNe Ia (Hamuy et al. 1996) covering the range $0.01 < z < 0.1$ is very well-suited. The collection has a spread in relative luminosity covering more than 1.5 magnitudes, but it has been shown that a two-parameter luminosity correction using the color and decline rate of each supernova is both necessary and sufficient to standardize them to a common luminosity (Tripp 1998). A second requirement for measuring $H_0$ is to establish the absolute luminosity of SNe Ia that have been standardized in the above way. At the present time there are seven nearby (typically an order of magnitude closer) galaxies that have hosted SNe~Ia whose galaxy distances have been determined using Cepheid variables measured by the Hubble Space Telescope. This includes the recently discovered SN 1998bu in the Leo I group galaxy NGC 3368 whose Cepheid distance had already been measured. The Cepheid-determined distances each have a typical accuracy of about 5\%, plus an additional overall uncertainty of about 7\% associated with the absolute distance scale for Cepheid variables. These seven galaxies have hosted eight recorded SNe Ia. The oldest, SN 1895B, was observed photographically in only one color so cannot be used here. Three of the others preceded the use of CCDs in astronomy and, despite being well-measured by the techniques of the day, are not of the quality that can now be achieved. Thus we also consider an expanded list containing three more recent SNe Ia which, although not found in Cepheid-calibrated galaxies, are very near on the sky to such a galaxy or are generally considered to be within groups of galaxies that have at least one member that has been so calibrated. Results are presented both with and without these additional supernovae. An extensive bibliograpy on the use of Type Ia supernovae for measuring the Hubble constant can be found in the recent review by Branch (1998). \section{PROCEDURE} We treat the 29 distant SNe of Hamuy et al. (1996) following the method of Tripp (1998), except that we allow the absolute magnitude to vary in a simultaneous fit with the Cepheid-calibrated supernovae. We calculate the Hubble constant $H_0$ for each distant supernova using \begin{equation} \log H_0 = { {M_B - B + 52.38} \over 5 } + \log {{1 - q_0 + q_0 z - (1-q_0) \sqrt{ 1 + 2 q_0 z } } \over { q_0^2}} \end{equation} Here $B$ and $M_B$ are the apparent and absolute blue magnitudes at maximum light, while $q_0$ is the deceleration parameter and $z$ is the measured red shift. For the 29 distant SNe, $B$ and $z$ are tabulated by Hamuy et al. (1996). These supernovae are not so remote $(z \le 0.1)$ that the still uncertain value of $q_0$ leads to a significant uncertainty. Conforming to the current evidence of Riess et al. (1998) and Perlmutter et al. (1998) for a negative $q_0$, we fix it at -0.45 (see Section 4).As shown previously (Tripp 1998), in order to fit the data of Hamuy et al. (1996) within their quoted errors, the above absolute magnitude $M_B$ for each supernova must be adjusted according to both its rate of decline $\Delta m_{15}$ and its color $(B - V)$. Here $B$ and $V$ are the maximum apparent blue and visual magnitudes. Empirically, linear dependences prove to be more than adequate for both corrections. We therefore write for $M_B$ in equation (1) \begin{equation} M_B = <M_B^0> + b ( \Delta m_{15} - 1.05) + R (B - V) \end{equation} \noindent The parameter $R$ incorporates both intrinsic color differences between the SNe and any reddening caused by dust in the host galaxies. These two effects are often difficult to distinguish, particularly for the more distant supernovae. The reddening parameter due to dust alone, usually denoted by $R_B$, should be about 4 if the dust surrounding extragalactic supernovae is similar to dust in the Milky Way.\footnote{ However, recent measurements of extinction from diffuse Galactic cirrus clouds (Szomoru \& Guhathakurta 1999) find $R_B$ values $\lesssim 3$ due perhaps to a smaller average grain size. Since SNe Ia are not closely associated with star formation it may well be that cirrus values are relevant for them. In any event, all SNe with evidence of strong dust obscuration are removed from the sample for the case where we apply a color selection to the data.} The weighted average value $<M_B^0>$ in equation (2) is obtained from the Cepheid-calibrated SNe in the following manner. We use a similar dependence to standardize each of them to the value it would have if $\Delta m_{15} = 1.05$ and $B - V = 0$ for that supernova. Thus the corrected absolute magnitude becomes \begin{equation M_B^0 = M_B - b (\Delta m_{15} - 1.05) - R (B - V) \end{equation} A least-squares fit of the Cepheid-calibrated SNe gives the weighted average $<M_B^0>$ appearing in equation (2), along with the $\chi ^2$ for the fit. Both are functions of $b$ and $R$. The value 1.05 in equations (2) and (3) is the average $\Delta m_{15}$ for 13 SNe Ia compiled by Branch et al. (1996). Chosen for convenience, the choice is arbitrary and has no effect on $H_0$, although it will directly affect $<M_B^0>$. The same remarks apply to the arbitrary choice of a $<B-V> = 0 $ subtraction in the color term. For each Cal\'an-Tololo distant supernova, we use equation. (1) to evaluate $H_0$ using the value of $M_B$ from equation (2) with $b$ and $R$ as parameters. The uncertainty in $H_0$ for each supernova is obtained by combining in quadrature the quoted errors $\delta B$, $\delta m_{15}$, and $\delta (B - V)$ with an uncertainty in luminosity distance due to possible peculiar motion $\delta v = 400$ km/s of the host galaxy with respect to the Hubble flow. Neglecting correlations between errors (since they are not reported), we have: \begin{equation \delta H_0 = H_0 \sqrt{ \left ( {\ln 10 \over 5 } \right )^2 \left [ \delta B^2 + ( b \delta \Delta m_{15} )^2 + ( R \delta (B-V))^2 \right ] + \left [ \left ( {1 \over z} + { {1- q_0} \over 2 } \right ) \delta z \right ] ^2 } \end{equation} \noindent A weighted average of the 29 values of $H_0 \pm \delta H_0$ then gives a least-squares value for $H_0$ along with its uncertainty. The $\chi^2$ for this fit of $H_0$ is added to the $\chi^2$ for the Cepheid- calibrated SNe fit of $<M_B^0>$, both with $b$ and $R$ as parameters. These are then varied to minimize the overall $\chi^2$. When the parameters to be varied enter into the evaluation of the uncertainty $\delta H_0$ and thus into the $\chi^2$ of the fit, as they do here, there is a bias favoring larger values of $b$ and $R$ leading to a larger $\delta H_0$ and thus to a lower $\chi^2$. We try to eliminate this by temporarily fixing $b$ and $R$ for the evaluation of $\delta H_0$ in equation (4) at the first solution values, then reminimizing and again fixing them at the new values. After a few such iterations, stable values of $b$ and $R$ are found which are also $\chi ^2$ minima. In the end, this results in a small decrease in $b$, a substantial decrease in $R$ (by about 0.6 unit), and a consequent increase in $H_0$ by about 1.3 units. \section{RESULTS} In Table 1, we list the seven SNe from Cepheid-calibrated galaxies and the three from Cepheid-calibrated galaxy groups, along with the measured values used in our fits; these are the distance modulus $\mu$, the measured apparent magnitude $B$ and the resulting absolute magnitude $M_B = B - \mu$ appearing in equation (3), the decline in $B$ magnitude, $\Delta m_{15}$, during the first 15 days after maximum, and the color, taken to be $B_{\rm max} - V_{\rm max}$. Where appropriate, the distance modulus incorporates the HST long-exposure correction of 0.05 as well as an estimate of the effect of host-galaxy absorption on the Cepheid distance determination. Conforming to the procedure used by Hamuy et al. (1996) for their more distant SNe, both $B$ and $B-V$ are corrected for Galactic absorption using the estimates of Burstein \& Heiles (1984), but no correction is made for absorption within the host galaxy. Our method accomodates this host-galaxy absorption as well as any intrinsic reddening with the parameter $R$ used in the fits. Thus the generally uncertain dust absorption is effectively corrected for by this procedure. For the 29 Cal\'an-Tololo SNe, we use values of red-shift, $B$, $B - V$, and $\Delta m_{15}$ given in Table 1 of Hamuy et al.(1996). Table 2 presents the results of our joint fitting of the Cepheid-calibrated (CC) SNe + Cal\'an-Tololo (CT) SNe for two data selections. Shown are the number of fitted supernovae in each category, the best-fit values of $H_0$, $<M_B^0>$, $b$, and $R$, and the individual confidence levels for the best joint fit of the two subsets of data. In the first row we limit the sample to the six directly calibrated CC SNe that also satisfy the color selection $(B - V) < 0.2 $ (Vaughan et al., 1998) and to the 26 color-selected CT SNe. In the second row, the full nearby sample of 10 CC SNe are jointly fitted with all 29 of the CT SNe. Both yield, fortuitously, identical values for $H_0 = 62.8$. It can be seen from Table 2 that the CT data alone fit extremely well in both cases even though the $b$ and $R$ parameters are optimized to fit the combined (CT and CC) data. The confidence levels for the CT data are in all cases considerably higher than the most likely value of 0.5. This reflects the presumed overestimate of the errors in the CT data set described in Tripp (1998) where these data alone lead for the 29 (26) SNe to an unrealistically high confidence level of 0.98 (0.97) for $b = 0.52\ (0.53)$ and $R = 2.09\ (2.44)$. In both cases the joint confidence levels are acceptable due in large measure to these overly good fit of the CT data. For the case of six CC SNe, good fits are also realized for the CC subsample. But this is not the case for the 10 CC SNe where their confidence level falls below 1\%. This is due to a conflict between the two most recent CC SNe: 1991T and 1998bu. They are both somewhat reddened slow decliners, with the first being superluminous and the second being subluminous according to this analysis, so that no $b$ and $R$ corrections can adequately reconcile the two and at the same time fit the CT data.\footnote{However, if we fit just the 10 CC SNe alone, then a satisfactory confidence level CL = 0.15 is found, with $b = 1.56$ and $R = 2.80$. Alternatively, if one or the other of the two conflicting SNe is eliminated, then (CC+CT) fits can be found that are also good CC fits. Thus, making the color cut, thereby discarding 98bu and retaining 91T, yields $H_0 = 60.8$ with a CC confidence level of 0.36, while selecting just the directly calibrated SNe and making no color cut discards 91T and retains 98bu, yielding $H_0 = 64.9$ with a CC confidence level of 0.30. It is to be expected from the extreme nature of these two SNe, lying, as they do, far out on either side of the $\chi^2$ distribution, that when one or the other is eliminated it will substantially impact $H_0$. It has been suggested (Fisher et al. 1999) that SN 1991T forms a separate class of relatively rare superluminous objects not seen among the CT sample. It is best explained as the result of a merger of two white dwarfs leading to a super-Chandrasekhar explosion. SN 1998bu is a normal supernova, but with considerable dust obscuration. Making the conventional color cut $(B - V) < 0.2$ eliminates it. Since there is no evidence coming from interstellar absorption lines for strong dust obscuration among the 29 CT SNe, whereas all three of the 10 CC SNe falling outside this cut show clear evidence for this, in order to minimize bias between the samples the safest procedure is to apply the color cut to both groups as we do in the CC6/CT26 fit.} We list in Table 3 the values of $M_B^0$ for each of the Cepheid-calibrated SNe along with their residuals, i.e., $\chi = [M_B^0- <M_B^0>] \slash \delta M_B^0 $, for both fits of Table 2. In Figure 1, we show the corrected values (filled circles) and the uncorrected values $M_B$ (open circles) for the full sample (CC10/CT29). These are displayed as a function of $\Delta m_{15}$ in Figure 1a and as a function of B-V in Figure 1b. The strong dependences of the uncorrected open circles on $\Delta m_{15}$ and $B-V$ are mostly removed by the best-fit parameters $b = 0.55$ and $R = 2.40$, as seen in the corrected full circles. However, as noted above, this fit to a common value of $<M_B^0> = -\!19.44$ has an unsatisfactory $\chi^2$ with CL = 0.009 due to the conflicting demands of SN 1991T and SN 1998bu. \section{DISCUSSION} >From the confidence levels in Table 2, it is evident that the combined >data from the Cal\'an-Tololo and the Cepheid-calibrated supernovae can be simultaneously fit in a satisfactory manner. However, the good confidence level, primarily the result of exceedingly good fits of the CT SNe by themselves, disguises a significant difference in the data sets. This is apparent in Figure 2, which shows plots of $B - V$ vs. $\Delta m_{15}$ for the 29 CT SNe (Figure 2a) and the 10 CC SNe (filled circles in Figure 2b). The striking difference between the two groups is the complete absence of unreddened $(B - V < 0.2)$ events beyond $\Delta m_{15} = 1.2$ among the 10 CC SNe and their abundance among the 29 CT SNe. Since $B - V$ and $\Delta m_{15}$ are both independent of distance, we may include in Figure 2b six additional nearby SNe in the Virgo and Fornax clusters (open circles) whose distances are still uncertain. This inclusion makes the two samples more compatible, suggesting that the void in the nearby sample of 10 may be, in part, due to a statistical fluctuation. Thus, while the 29 CT SNe in Figure 2a display very little correlation between $B - V$ and $\Delta m_{15}$, within the limited statistics of the 10 CC SNe there is a strong correlation, but one which is much reduced by the inclusion of the six distance-uncalibrated SNe. A possible bias affecting any measurement of the Hubble constant, both in our two parameter fit and previous one and zero parameter fits, may arise because the nearby and distant samples come from galaxies of somewhat different morphologies. Since a galaxy must contain a population of young stars in order to produce Cepheid variables, the CC SNe are generally found in spiral galaxies. (An exception to this is NGC5253, the parent of SNe 1972E and 1895B, which is often classified as an E/SO peculiar galaxy. A putative explanation for its star forming regions is that they are the result of a galactic merger.). Thus, apart from SN 1972E, our CC sample comes from spiral galaxies, whereas the CT sample is about equally divided between spirals and E, E/SO, and SO galaxies. Eliminating from CT26 all but the 13 spirals gives for CC6/CT13, after reminimizing $\chi^2$ for this smaller sample, a value of $61.9 \pm 2.0$ for $H_0$ compared to $62.8 \pm1.6$ for the full sample CT6/CT26. Since the former is presumably freer of potential bias we take $H_0 = 62$ as our best estimate of the Hubble constant and the associated value of -19.46 for $<M^0_B>$. We now discuss the uncertainty in $H_0$ arising from a variety of sources. The 1-$\sigma$ statistical error in $\delta H_0$ is found to be $\pm 1.0$. In addition, there is uncertainty in $b$ and $R$ which we display in Figure 3 where $H_0$, as a function of $b$ and $R$, is superposed on the 1-, 2-, 3-, and 4-$\sigma$ contours of the CC6/CT26 fit. >From this we obtain the 1-$\sigma$ error from the $b$ and $R$ uncertainty >of $\delta H_0 = \pm1.2$. Throughout this analysis we have fixed the insensitive deceleration parameter at $q_0 = -0.45$, found by making a three parameter $(b,\ R,\ q_0)$ fit of the 26 CT SNe jointly with the 9 cosmological SNe of Riess et al. (1998) for which $B-V$ colors were available. If we assign an uncertainty of $\pm 0.5$ to this value of $q_0$, this leads to an uncertainty $\delta H_0 = \mp 0.6$. Combining these three sources in quadrature yields $\delta H_0 = \pm 1.7$. To these uncertainties arising from the SNe Ia analysis must be added a larger uncertainty coming from the calibration of the Cepheid variables and their possible metallicity dependence. The Madore and Freedman (1991) distance scale in current use fixes the distance modulus to the Large Magellanic Cloud (LMC) to be 18.50 mag (50.1 kpc) and scales more distant Cepheid-based galaxy distances to this value. Because of the importance of this number for the determination of the Hubble constant, much attention has been devoted to methods for obtaining a more accurate value. A recent review from a post-Hipparcos perspective (Walker 1998) recommends a mean modulus of $18.55 \pm 0.10$, found using a variety of distance indicators. However, even more recent analyses of an eclipsing binary in the LMC (Guinan et al. 1998), found during the OGLE microlensing search, yield a value as low as $18.22 \pm 0.13$ mag (Udalski et al. 1998). For this and other reasons evident in the Walker(1998) summary, we use without change the distance moduli of Cepheid-calibrated SNe Ia found by the various observers which are all based on the 18.50 mag LMC value. We assign to this value a 1-$\sigma$ error of $\pm 0.15$ mag, with the usual statistical view that a true distance modulus to the LMC differing by twice that value would be surprising and one differing by 3-$\sigma$ would be very unlikely. Using $\delta H_0 = 29(H_0 \slash 62) \delta M_B$ obtained from equation (1), this uncertainty leads to a 1-$\sigma$ error of $\delta H_0 =\pm 4.34$. For some years there has been concern that metallicity differences between the LMC and the Cepheid- calibrated galaxies can alter the derived distances to these galaxies. As the most easily measured proxy for metallicity, Kochanek (1997) has collected the logarithmic abundance ratio $[O/H]$ for a number of the Cepheid-calibrated galaxies relative to that of the LMC. For galaxies hosting SNe Ia, these cover a range of $[O/H]$ from -0.35 to +0.69. We have used these to alter the measured distance moduli $\mu$ by means of the expression $\mu^\prime = \mu + \gamma [O/H]$, where the parameter $\gamma$ is varied to obtain a best fit . For CC6/CT26, the data set for which five of the Cepheid-calibrated galaxies have measured values of $[O/H]$, we find $\gamma= +0.23 \pm 0.77$, in agreement with other recent findings which fall between 0.14 and 0.31 (Kochanek (1997), Kennicutt et al. (1998), Nevalainen and Roos (1998)), but showing that this data set has little sensitivity to $\gamma$. If we fix $\gamma = 0.3$ and follow the previous procedure of varying $b$ and $R$ to find a $\chi^2$ minimum, then $H_0$ increases by 0.8 km s$^{-1}$ Mpc$^{-1}$. Since the question of a metallicity dependence is still in dispute (see Beaulieu et al. 1997, Saha et al. 1997, and Saio and Gautschy 1998 for recent differing views), we retain the value of $H_0$ without correction but assign an uncertainty in $\gamma$ of $ {+0.3 \atop -0}$ resulting in $\delta H_0 = {+0.8 \atop -0}$ due to this source. Combining all these errors in quadrature yields $\delta H_0 = \pm 4.7$ for the overall uncertainty. Despite significant differences with other analyses, our value of $H_0 \approx 62$ falls squarely in the middle of the range spanned by recent determinations using SNe Ia. These values of $H_0$, all using the distance modulus of 18.50 to the LMC, range between 55 and 69 km s$^{-1}$ Mpc$^{-1}$. One difference between our analysis and others is that they all seem to use the nearby approximation to equation (1), which is tantamount to setting $q_0 = 1$. If instead they were to use equation (1) with a negative $q_0$, as is now apparently required by the cosmological data, then their values would each increase by about 1.7 units in cases when the CT SNe are used for the Hubble flow sample. The other major difference is that only our analysis uses two independent parameters ($b$ and $R$) to standardize the luminosity of the supernovae; this is required in order to obtain a fit to the CT data with an acceptable $\chi^2$ (Tripp 1998). Among the recent analyses using a collection of nearby calibrating supernovae, Saha et al. (1997), employing seven Cepheid-calibrated SNe (1895B, 1937C, 1960F, 1972E, 1981B, 1989B, 1990N) and no standardizing parameters, find $H_0 = 58$. Suntzeff et al. (1998), using five CC SNe (1937C, 1972E, 1981B, 1990N, 1998bu) along with the Hubble flow CT SNe and with a one parameter correction for $\Delta m_{15}$, obtain $H_0 = 64$. The difference in $H_0$ found in these two analyses lies primarily in the introduction of the $b$ parameter. Our analysis, involving both $b$ and $R$, reduces $b$ and leads to an intermediate result. This is immediately apparent from Figure 3. As can be discerned from the figure, the Saha et al.(1997) $b = R = 0$ fit should yield about 60 for $H_0$. This is equivalent to their 58 after correcting for the above effect of $q_0$. Likewise the Suntzeff et al. (1998) fit with only $b$ as a free parameter should, after the $q_0$ correction, yield about 66, as can be seen from the figure for $R = 0$. Apparently, different selections of Cepheid-calibrated SNe have only a minor effect on $H_0$. Thus future augmentations to the present small number of Cepheid-calibrated SNe will probably have little impact on $H_0$. Assuming that there are no significant observational selection differences between the CT and CC samples, only a revision of the Cepheid distance scale will be capable of altering $H_0$ by more than a few units. This work was supported in part by the Director, Office of Energy Research, Office of High Energy and Nuclear Physics, Division of High Energy Physics of the U.S. Department of Energy under contract AC03- 76SF00098 and (for D.B.) an NSF grant AST 9417102. \clearpage \centerline{\bf REFERENCES} \begin{description} \item Beaulieu J. P., Sasselov D.D., Renault C. et al. 1997, A\&A, 318, L47 \item Branch D., Romanishin W., Baron E. 1996, ApJ, 465, 73 \item Branch D. 1998, ARA\&A, 36, 17 \item Burstein D. \& Heiles, C. 1984, ApJS, 54, 33 \item Fisher, A., Branch, D., Hatano, K. \& Baron, E. astro-ph/980732, MNRS (in press). \item Guinan, E. F., Fitzpatrick, L.,E., DeWarf, F.,P. et al. 1999, ApJ, 509, L21 \item Hamuy, M., Phillips, M. M., Schommer, R. A. et al. 1996, AJ, 112, 2398 \item Kennicutt, R. C., Stetson, P. B., Saha, A. et al. 1998, ApJ, 498, 181 \item Kochanek, C. S. 1997, ApJ, 491, 13 \item Lira, P., Suntzeff, N. B., Phillips, M. M. et al. 1998, AJ, 115, 234 \item Madore, B. F., Freedman, W. L. 1991, PASP, 103, 933 \item Nevalainen, J. \& Roos, M. 1998, A\&A, 339, 7 \item Perlmutter, S. et al. 1998, ApJ, in press \item Phillips, M. M. 1993, ApJ, 413, L105 \item Phillips, M. M., Phillips, A. C., Heathcote, S. R. et al. 1987, PASP, 99, 592 \item Riess, A. G., Filippenko, A. V., Challis, P. et al. 1998, AJ, 116, 1009 \item Saha, A., Sandage, A., Labhardt, L. et al. 1996, ApJ, 466, 55 \item Saha, A., Sandage, A., Labhardt, L. et al. 1997, ApJ, 486, 1 \item Saio, H. \& Gautschy, A. 1998, ApJ, 498, 360 \item Schaefer, B. E. 1996, ApJ, 460, L19 \item Schaefer, B. E. 1998, ApJ, 509, 80 \item Suntzeff, N. B., Phillips, M.M., Covarrubias, R. et al. astro-ph/9811205 \item Szomoru, A. \& Guhathakurta, P., astro-ph/9901422 \item Tanvir, N. R., Shanks T., Ferguson, H. C., Robinson, D. R. T. 1995, Nat, 377, 27 \item Turner, A., Ferrarese, L., Saha, A. et al. 1998, ApJ, 505, 207 \item Tripp, R. 1998, A\&A, 331, 815 \item Vaughan, T. E., Branch, D., Miller, D. L. \& Perlmutter, S. 1995, ApJ, 439, 558 \item Udalski, A., Pietrzynski, G., Wozniak, P. et al. 1999, ApJ, 509, L25 \item Walker A., astro-ph/9808336, 1998, in Post Hipparcos Candles, eds. F. Caputo \& A. Heck (Dordrecht, Kluwer Academic Publ.) \end{description} \clearpage \figcaption{The corrected values $M_B^0$ (filled circles) for the CC10/CT29 fit and the uncorrected values $M_B$ (open circles) (top) as a function of $\Delta m_{15}$ and (bottom) these quantities as a function of $B_{\rm max}-V_{\rm max}$. The best-fit corrected value of -19.44 is shown by the horizontal lines.\label{Fig.1}} \figcaption{Plot of the distance-independent quantities $B_{\rm max} - V_{\rm max}$ color vs. $\Delta m_{15}$ (top) for the 29 SNe of the Cal\'an-Tololo distant sample and (bottom) for the 10 Cepheid- calibrated SNe (filled circles) and six other nearby SNe (open circles) from the Virgo and Fornax clusters. The latter tend to populate the void of unreddened SNe beyond $\Delta m_{15} = 1.2$, thereby making the two distributions more compatible.\label{Fig.2}} \figcaption{1-, 2-, 3-, and 4-$\sigma$ contours (corresponding to an increase in $\chi^2$ of 1, 4, 9, 16) for the CC6/CT26 data set as a function of $b$ and $R$. Also shown are $H_0 = 60$ and 65 curves as a function of $b$ and $R$. From these curves it is apparent why the $b = 0$, $R = 0$ fit of Saha et al. (1997) yields a lower value than ours and the $R = 0$ fit of Suntzeff et al. (1998) yields a larger value. The $\sigma$ contours show only a weak correlation between $b$ and $R$ for this data set, as would be expected, following the imposition of the color cut from Figure 2. The $1-\sigma$ uncertainties for $b$ and $R$ are $\pm 0.13$ and $\pm 0.25$ respectively.\label{Fig.3}} \clearpage \begin{deluxetable}{cllllll} \footnotesize \tablecaption{Cepheid-Calibrated Supernovae\tablenotemark{a} \label{tbl-1}} \tablewidth{0pt} \tablehead{ \colhead{SN} & \colhead{Galaxy} & \colhead{$\mu(\delta \mu)$} & \colhead{$B(\delta B)$} & \colhead{$M_B(\delta M_B)$} & \colhead{$\Delta m_{15}(\delta\Delta m_{15})$} & \colhead{$B-V[\delta(B-V)]$} } \startdata 1937C&IC4182&28.36(9) (1)&8.80(9) (6)&-19.56(13)&0.87(10) (6)&-0.02(6) (6)\nl 1960F&N4496A &31.13(10) (1)&11.60(10) (1)&-19.53(14)&1.06(8) (11)&0.09(8) (1)\nl 1972E&N5253&27.94(8) (1,2)&8.30(14) (1)&-19.64(16)&0.87(10) (6)&-0.05(8) (6)\nl 1974G&N4414&31.41(23) (3)&12.48(5) (7)&-18.93(24)&1.11(6) (7)&0.18(5) (7)\nl 1981B&N4536 &31.10(13) (4)&12.03(3) (6)&-19.07(13)&1.10(7) (2)&0.10(4) (6)\nl 1986G\tablenotemark{*}&N5128 &28.13(46) (1)&11.96(7) (8)&-16.17(46)&1.73(7) (12)&0.83(10) (8)\nl 1989B\tablenotemark{*}&N3627 &30.28(26) (1)&12.29(6) (9)&-17.99(27)&1.31(7) (12)&0.34(5) (9)\nl 1990N&N4639 &32.03(22) (1)&12.71(3) (10)&-19.32(22)&1.07(5) (2)&0.05(4) (10)\nl 1991T\tablenotemark{*}&N4527 &31.11(10) (1) &11.70(2) (10)&-19.41(10)&0.94(5) (10)&0.19(3) (10)\nl 1998bu&N3368&30.37(16) (5,2)&12.10(3) (2)&-18.27(16)&1.01(5) (2)&0.30(4) (2)\nl \enddata \tablenotetext{a}{A list of the 10 Cepheid-calibrated supernovae along with the input data used in the fits. The uncertainty for each quantity is shown in parentheses, followed by the reference for the measurement.} \tablenotetext{*} {means the distance is inferred from association with other galaxies with the distance uncertainty increased by the projected distance.} \tablenotetext{}{References: (1) Saha et al. 1997; (2) Suntzeff et al. 1998; (3) Turner et al. 1998; (4) Saha et al. 1996; (5) Tanvir et al. 1995; (6) Hamuy et al. 1996; (7) Schaefer 1998; (8) Phillips et al. 1997; (9) Wells et al. 1994; (10) Lira et al. 1998; (11) Schaefer 1996; (12) Phillips 1993.} \end{deluxetable} \clearpage \begin{deluxetable}{crrrrrrrr} \footnotesize \tablecaption{Best-fit Solutions\tablenotemark{a} \label{tbl-2}} \tablewidth{0pt} \tablehead{ SNe& $H_0$& $<M_B^0>$& b& R&\quad&\multispan{3}{\hfil Confidence level (CL)\hfil}\nl &&&&&&CT+CC&CT&CC} \startdata CC6/CT26 &62.8&-19.42&0.483&1.883&&0.955&0.933&0.719\nl CC10/CT29 &62.8&-19.44&0.553&2.397&&0.385&0.962&0.009\nl \enddata \tablenotetext{a}{The best-fit solutions for the two data sets discussed in the text.} \end{deluxetable} \clearpage \begin{deluxetable}{lrrrrr} \footnotesize \tablecaption{$M_B^0$ and residuals.\tablenotemark{a} \label{tbl-3}} \tablewidth{0pt} \tablehead{ Supernova&\multispan{2}{\hfil CC6/CT26\hfil}&\quad&\multispan{2}{\hfil CC10/CT29\hfil}\nl &$M_B^0\quad$&Residual&&$M_B^0\quad$&Residual } \startdata 1937C&-19.44 (18)&-0.10&&-19.41 (20)&0.15\nl 1960F&-19.70 (21)&-1.36&&-19.75 (24)&-1.27\nl 1972E&-19.46 (23)&-0.18&&-19.42 (26)&0.90\nl 1974G&-19.30 (26)&0.47&&-19.39 (27)&0.18\nl 1981B&-19.28 (16)&0.87&&-19.34 (17)&0.63\nl 1986G*&&&&-18.54 (52)&1.73\nl 1989B*&&&&-18.95 (30)&1.67\nl 1990N&-19.39 (24)&0.11&&-19.45 (24)&-0.03\nl 1991T*&&&&-19.80 (13)&-2.83\nl 1998bu&&&&-18.97 (19)&2.49\nl \enddata \tablenotetext{a}{Note: Values of $M_B^0$ with its error $\delta M_B^0$ and the residuals $\chi = [M_B^0 -<M_B^0>] \slash \delta M_B^0$ for each Cepheid-calibrated supernova for the two fits discussed in the text. } \end{deluxetable} \end{document}
\section{Introduction} Recent developments in string theory are understandings of the correspondence between the brane dynamics and the physics of gauge theories in various dimensions. Hanany and Witten have first pointed out \cite{HW} that $N{=}4$ supersymmetric gauge theory in three-dimensions is realized as worldvolume effective theory on D3-branes stretched between two NS5-branes and the ``mirror symmetry'' can be understand as a consequence of $SL(2,{\bf Z})$ duality of Type IIB superstring theory. Their construction is extended to different numbers of supersymmetries \cite{dBHOY,dBHO} and gauge theory in various dimensions (for a review see \cite{GK,Karch}). The advantage of this approach is that we can geometrically analyze the structure of the moduli spaces of vacua as the configuration space of branes. The geometrical construction of moduli space make easy to classify possible supersymmetric vacuum. From the point of view of brane solutions in Type IIB superstring theory, the three-dimensional Hanany-Witten type configuration is generalized to the case including rotated 5-branes at arbitrary angles \cite{GGPT}. It contain the interesting 5-brane solution preserving $3/16$ of supersymmetry. This means that in three-dimensions we have $N{=}3$ supersymmetric theory. One of the realization of the three-dimensional $N{=}3$ supersymmetry is obtained by adding the Chern-Simons term \cite{ZK,KL,Kao}. So the brane configuration with $3/16$ supersymmetry can be associated with three-dimensional Chern-Simons theory. Moreover, the systematic classification of possible Hanany-Witten type configuration with various supersymmetries has been studied in ref. \cite{KOO}. Since the Type IIB brane configuration with $3/16$ supersymmetry contains a $(p,q)$5-brane which is a bound state of NS5 and D5-brane, it is reasonable to consider that the replacement of one of the NS5-branes with the $(p,q)$5-brane is related to adding the Chern-Simons interaction. Under this identification, it is found that the vacuum structure of Maxwell Chern-Simons theory is consistent with the brane picture and a ``mirror symmetry'' is similarly realized by Type IIB S-duality. This ``mirror symmetry'' in supersymmetric Chern-Simons theory is recently confirmed by using a generalized Fourier transformation in ref. \cite{KS}. In this paper, we investigate the correspondence between the $(p,q)$5-brane configuration introduced by ref. \cite{KOO} and Yang-Mills (Maxwell) Chern-Simons theories in detail. More specifically, we construct a explicit Lagrangian for the supersymmetric Chern-Simons theories and identify vevs of emergent fields with moduli parameters of Type IIB branes. This identification gives us the knowledge to understand the complicated dynamics of three-dimensional massive gauge theories and the dynamics of Type IIB branes. This paper is organized as follows: In section 2 we briefly review the construction of the brane configurations with $(p.q)$5-brane. These configurations in Type IIB theory are reduced from M5-brane configurations in M-theory with various supersymmetries. We give all possible Hanany-Witten type configuration with $N{\geq}2$ supersymmetry. In section 3, we see how to describe the Coulomb branch by vevs of adjoint scalars in supersymmetric Chern-Simons theory. The number of the adjoint scalar fields corresponds to the number of freely moving directions of D3-branes. Finally, we consider the vacuum of Maxwell Chern-Simons theory with matter superfields. In this model, there are two distinct vacua. One of vacua, which is called as a symmetric phase, exists on in the Maxwell Chern-Simons Higgs system. We can find the brane configuration of the symmetric phase. We also identify the topological vortex in the asymmetric phase as a bound state of strings. \section{Supersymmetric Configuration with $(p,q)$5-brane} In this section we briefly review the construction of the brane configuration with $(p,q)$5-brane. The brane configurations in Type IIB string theory which are given by Hanany and Witten \cite{HW} describe three dimensional supersymmetric gauge theories. In this configuration D3-branes are suspended between two parallel NS5-branes and the $3+1$ dimensional worldvolume of D3-brane is compactified on a finite line segment. We denote the extended directions of two NS5-branes and $N_c$ D3-branes as the following symbols: \begin{eqnarray*} 2\times {\rm NS5} & : & (012345),\\ N_c\times {\rm D3} & : & (012|6|), \end{eqnarray*} where the numbers in the parenthesis are the worldvolume directions and the vertical lines on both side of $6$ mean that the worldvolume of D3-branes are restricted to finite interval in the $x^6$-direction. The ``left'' NS5-brane is located at $x^6=-L/2$ and the ``right'' is at $x^6=L/2$. The low-energy effective theory on $N_c$ parallel D3-branes is $3+1$ dimensional $SU(N_c)$ supersymmetric Yang-Mills theory in the limit that string scale $l_s$ goes to zero, but this compactification on the line reduces the D3-brane worldvolume theory to $2+1$ dimensions. Upon this Kaluza-Klein reduction on the line segment we find the gauge coupling constant $g$ of this worldvolume theory is given by \begin{equation} \frac{1}{g^2}=\frac{L}{g_s}, \label{coupling} \end{equation} where $g_s$ is a string coupling constant of Type IIB theory and $L$ is a length of the line segment, namely, a distance between two NS5-branes in the $x^6$-direction. In this configuration the presence of these two kinds of branes preserves 8 of 32 supercharges of Type IIB string theory since each kind of brane preserves half of supercharges. Therefore we have $N{=}4$ supersymmetry in $2+1$ dimensions. To add matter in the fundamental representation of the gauge group we can introduce $N_f$ D5-branes, which have the worldvolume: \begin{eqnarray*} N_f\times{\rm D5} & : & (012789), \end{eqnarray*} without breaking any supersymmetries. The matter hypermultiplets come from strings stretched between the $N_c$ suspended D3-branes and the $N_f$ D5-branes. So $N_f$ represents the number of flavors. In addition to these types of branes there also exists the $(p,q)$5-brane which is a bound state of NS5- and D5-branes. One of the extension of the above setup is replacement of the right NS5-brane in the Hanany-Witten configuration with this $(p,q)$5-brane. However, if we simply substitute the $(p,q)$5-brane for the right NS5-brane and the left NS5-brane and the right $(p,q)$5-brane is parallel then all supersymmetry is broken. To preserve some of the supercharges we must rotate the $(p,q)$5-brane by some angles. In order to count the residual supercharges we now lift the above Type IIB configuration to M5- and M2-brane configurations in M-theory. T-duality in the $x^2$-direction maps the NS5-brane and the D5-brane into the NS5-brane and the D4-brane in Type IIA theory, respectively. By lifting to M-theory these objects become M5-branes. The NS5-brane corresponds to the M5-brane wrapping on the $x^2$-direction and the D5-brane corresponds to the M5-brane wrapping on the 11th-direction $x^{10}$. So two cycles of the torus $T^2$ with coordinates $(x^2,x^{10})$ are related to NSNS-charge and RR-charge in Type IIB theory. Similarly, $(p,q)$5-brane in Type IIB theory is also unified as a M5-brane which is obliquely wrapping on the torus $T^2$ at some angle $\theta$. This angle and charges of the $(p,q)$5-brane are related each other as follows \begin{equation} \tan\theta=\frac{2\pi R_{10}p}{2\pi R_{2}q}=g_s\frac{p}{q}, \label{tan} \end{equation} where $R_{2}$ and $R_{10}$ are radii of the torus $T^2$. Using the above procedure D3-branes in Type IIB theory become M2-branes stretched between two M5-branes in M-theory. In this situation the right M5-brane tilted by the angle $\theta$ can be rotated by more three angles in the configuration space of M-theory. So, we have a general configuration in M-theory: \begin{eqnarray*} {\rm M5} & : & (012345)\\ N_c\times {\rm M2} & : & (01|6|)\\ {\rm M5}' & : & (01 \rot{2}{10}{\theta} \rot{3}{7}{\psi} \rot{4}{8}{\varphi} \rot{5}{9}{\rho} ), \end{eqnarray*} where the symbol $\rot{\mu}{\nu}{\theta}$ means that the extended direction of the ${\rm M5}'$-brane are rotated by the angle $\theta$ in the $(x^\mu,x^\nu)$-space. If the angle $\theta$ is zero ${\rm M5}'$-brane become a rotated NS5-brane in Type IIB configuration space. On the other hand, the case of $\theta=\pi/2$ corresponds to a D5-brane. The presence of these branes imposes the constrains on the 11-dimensional Killing spinors $\epsilon$ \cite{OT} \begin{eqnarray} {\rm M5} &:& \Gamma_{012345}\epsilon = \epsilon, \label{M5}\\ {\rm M2} &:& \Gamma_{016}\epsilon = \epsilon, \label{M2}\\ {\rm M5}'&:& R\Gamma_{012345}R^{-1}\epsilon = \epsilon, \label{M5'} \end{eqnarray} where $R$ is the rotation matrix in the spinor representation and give by \[ R=\exp\left\{ \frac{\theta}{2}\Gamma_{2,10} + \frac{\psi}{2}\Gamma_{37} + \frac{\varphi}{2}\Gamma_{48} + \frac{\rho}{2}\Gamma_{59} \right\}. \] The numbers of the remaining supersymmetry are obtained by solving eqs. (\ref{M5})-(\ref{M5'}) simultaneously. All possible solutions are completely classified in ref. \cite{KOO} by setting some relations between four angles and we have $1/16$, $1/8$, $3/16$ and $1/4$ of the original 32 supercharges. These fractions of supersymmetry correspond to $N{=}1,2,3,4$ supersymmetries, respectively, in $2+1$ dimensions. In present paper we treat only the configuration for $N{\geq}2$ supersymmetry. This is because quantum corrections of higher supersymmetric is suppressed rather than the $N{=}1$ case and it is not so complicated to find the correspondence to the branelogy. $N{\geq}2$ supersymmetric configurations are divided into the following two cases. One is the NS5 configuration with $N{=}2,4$ supersymmetry where the right 5-brane is the NS5-brane. The other is the configuration including $(p,q)$5-brane with $N{=}2,3$. We first consider the NS5 configuration. For $N{=}4$ configuration we must set all angles to zero, $\theta=\psi=\varphi=\rho=0$. The right ${\rm M5}'$-brane is a NS5-brane in Type IIB theory, which is parallel to the left NS5-brane. This is nothing but the Hanany-Witten configuration. As we mentioned above, this configuration describes $N{=}4$ $SU(N_c)$ supersymmetric Yang-Mills theory. This theory includes three real scalar fields in adjoint representation. Vevs of these adjoint scalars and scalars dual to the gauge fields parametrize the Coulomb branch. On the other hand, D3-branes in the brane configuration can freely move along the $(x^3,x^4,x^5)$-space. These positions of D3-branes on NS5-brane correspond to the vevs of real adjoint scalars. In addition the $SO(3)$ rotation group in the $(x^3,x^4,x^5)$ can be identified with a part of the $R$-symmetry $SU(2)_V$ of $N{=}4$ supersymmetry algebra in three dimensions. The other part of the $R$-symmetry $SU(2)_H$ corresponds to rotations in the $(x^7,x^8,x^9)$-space. The $N{=}2$ NS5 configuration is obtained by a rotation of the right NS5-brane. In this rotation we must set same values on two pairs of angles. Since $\theta$ is zero in NS5 configuration one of our choice for angles is $\theta=\rho=0$ and $\varphi=\psi\neq0$. The special case of $\varphi=\psi=\pi/2$ is investigated in ref. \cite{dBHOY} and this brane configuration space well describes the moduli space of vacua of $N{=}2$ supersymmetric Yang-Mills theory. The Coulomb branch of this theory is parametrized by vevs of one real massless adjoint scalar and dual gauge fields. In this configuration D3-branes can not move in the $(x^3,x^4)$-space any longer since two NS5-branes are completely twisted in that space. Then the vev of real adjoint scalar corresponds to the positions of D3-branes in $x^5$-direction. For the general value of $\psi$ this rotation angle is related to a mass of two real adjoint scalars \cite{EGK,Barbon}, which originally belong to the $N{=}4$ vector multiplet. Therefore we can identify the rotation of NS5-brane with the supplement of the mass term of two adjoint scalars into $N{=}4$ theory. Actually, this brane rotation breaks the $SU(2)_V \times SU(2)_H$ rotation symmetry to $SO(2)\simeq U(1)_R$, which coincide with the $N{=}2$ $R$-symmetry group. Let us next mention about the configuration with $(p,q)$5-brane. This configuration is realized by setting $\theta\neq0$. A maximal supersymmetric configuration with $(p,q)$5-brane is the case of $\theta=\psi=\varphi=-\rho$. This configuration preserves $N{=}3$ supersymmetry in $2+1$ dimensions. Since the all rotation angles are same except for the signature, the rotation symmetry $SU(2)_V\times SU(2)_H$ in the $(x^3,x^4,x^5)$- and $(x^7,x^8,x^9)$-spaces breaks to a diagonal $SU(2)_D$. This rotational symmetry $SU(2)_D$ can be considered as the $R$-symmetry of the three-dimensional $N{=}3$ supersymmetry algebra. $N{=}2$ configuration is given by setting $\rho=-\theta\neq\varphi=\psi$, that is, two different sets of same angles. If we take $\theta=\rho=0$, then we have the $N{=}2$ NS5 configuration as mentioned above. Similarly, we can choose as $\varphi=\psi=0$. In this case the right 5-brane is $(p,q)$5-brane but still $N{=}2$ supersymmetric. It is field theoretically shown that the three-dimensional $N{=}3$ maximal supersymmetric system can be constructed by adding Chern-Simons terms \cite{ZK,KL,Kao}. So one expect \cite{GGPT} that the worldvolume effective theory on this $2+1$-dimensional intersection of rotated 5-branes is gauge theories with Chern-Simons terms. In fact, since there exists a non-trivial vev of axion (RR-scalar) field $C_0$ around $(p,q)$5-brane in Type IIB theory, the coupling $\int C_0 F\wedge F$ on the D3-brane induces the Chern-Simons term with coupling $\kappa=p/q$ \cite{KOO}. In the following section, we investigate the correspondence between these supersymmetric configurations with the $(p,q)$5-brane and the moduli space of vacua of Chern-Simons theories. \section{Coulomb Branches of The Non-Abelian Theories} \subsection{$N{=}3$ supersymmetric Yang-Mills Chern-Simons theory} We first begin with the $N{=}3$ configuration. The configuration in Type IIB theory is written as follows in the notation of the previous section: \begin{eqnarray*} {\rm NS5} & : & (012345)\\ N_c\times {\rm D3} & : & (012|6|)\\ (p,q)5 & : & (01 \rot{3}{7}{\theta} \rot{4}{8}{\theta} \rot{5}{9}{-\theta} ), \end{eqnarray*} where $\theta=\tan^{-1}\left(g_s\frac{p}{q}\right)$. There originally exists $SU(N_c)$ Yang-Mills gauge theory on the D3-branes. As we have explained, the right $(p,q)$5-brane provides the Chern-Simons term into the Yang-Mills theory. So we have the Yang-Mills Chern-Simons theory as the effective theory on this configuration. The coupling constant of Chern-Simons term is given by $\kappa=p/q$. However, the coupling constant of the non-Abelian Chern-Simons theory should be a integer in order to preserve gauge invariance under a large transformation. Therefore we set the charges of $(p,q)$5-brane as $(p,q)=(n,1)$. $N{=}3$ supersymmetric field theory in three dimensions has the $SU(2)_D$ $R$-symmetry. The vector multiplet contains one spin 1 massive vector field $A_\mu$, three spin $1/2$ spinors $\lambda_i$, three spin 0 real adjoint fields $X_i$ and one spin $-1/2$ spinor $\chi$. Three spinors and adjoint scalars form a triplet under $SU(2)_D$. The bosonic part of the Lagrangian of this $N{=}3$ supersymmetric Yang-Mills Chern-Simons theory is given by \cite{Kao} \begin{eqnarray} {\cal L}_{\rm B}&=& -\frac{1}{g^2}{\rm Tr}\left\{ \frac{1}{2}{F_{\mu\nu}}^2 +\left(\nabla_\mu X_i\right)^2 +\frac{1}{2}\left[X_i,X_j\right]^2 -\left(\frac{\kappa g^2}{4\pi}\right)^2 {X_i}^2 \right\}\\ &&+\frac{\kappa}{4\pi}{\rm Tr}\left\{ \epsilon^{\mu\nu\rho}\left(A_\mu\partial_\nu A_\rho+\frac{2}{3}iA_\mu A_\nu A_\rho\right) -\frac{i}{3}\epsilon^{ijk}X_i\left[X_j,X_k\right] \right\}. \end{eqnarray} Due to the Chern-Simons term the vector field topologically gains mass $M=\left|\frac{\kappa g^2}{4\pi}\right|$ which is the same mass as the adjoint scalars have. Note that in the limit of $\kappa\rightarrow0$, which corresponds to the limit $\theta\rightarrow0$ in the brane configuration, one obtains the $N{=}4$ supersymmetric pure Yang-Mills Lagrangian. The Coulomb branch of supersymmetric gauge theory is described by solutions of the following flatness condition for the adjoint scalar fields \begin{equation} M^2 X_i -\frac{M}{2} i\epsilon_{ijk}\left[X_j,X_k\right] -\left[X_j,\left[X_i,X_j\right]\right]=0. \label{flat} \end{equation} For real $X_i$, the general solution of this condition is $\left<X_i\right>=0$. So the Coulomb branch of this theory is completely lifted because of the mass term coming from the Chern-Simons interaction. From the point of view of the brane configuration, it is easy to understand this Coulomb branch moduli as the following. In the $N{=}4$ brane configuration, $N_c$ D3-branes can freely moved on the $(x^3,x^4,x^5)$-space since two NS5-branes are parallel. So we can identify positions of D3-branes in the $(x^3,x^4,x^5)$-space with vevs of the three real adjoint scalar fields, which describe a part of the Coulomb branch. However, in the $N{=}3$ configuration, since the left NS5-brane and the right $(p,q)$5-brane is completely twisted in $(x^3,x^4,x^5)$- and $(x^7,x^8,x^9)$-space, D3-branes can not move anywhere. This means that the corresponding vevs of the adjoint scalars must be at origin, namely, $\left<X_i\right>=0$, which coincide with the solution of the flatness condition (\ref{flat}). The mass of the adjoint scalars can be written in terms of the rotation angle of the $(p,q)$5-brane as \begin{equation} \left|\frac{\kappa g^2}{4\pi}\right|=\frac{1}{4\pi L}|\tan\theta|, \end{equation} where we use the relation (\ref{coupling}) and (\ref{tan}). This relation also agree with the results in ref. \cite{EGK,Barbon}. In the $\theta\rightarrow 0$ limit the mass of the adjoint scalar fields vanishes and at the same time the Chern-Simons interaction is dropped. In the brane configuration the right $(p,q)$5-brane become NS5-brane and to be parallel to the left NS5. So we have $N{=}4$ supersymmetric pure Yang-Mills theory. On the other hand, in the limit $\theta\rightarrow \pi/2$ the mass of the vector multiplet goes to infinity and all gauge degrees of freedom are decoupled. In this limit the right $(p,q)$5-brane become D5-brane which extends in the $(x^0,x^1,x^2,x^7,x^8,x^9)$-space. This configuration also recover the $N{=}4$ supersymmetry again and appears in the Higgs branch of the Hanany-Witten configuration \cite{HW}. \subsection{Breaking to $N{=}2$} We next consider the supersymmetry breaking of $N{=}3$ theory down to $N{=}2$. In the brane language this is done by making two of the angles different from the others. Namely, we have the configuration \begin{eqnarray} {\rm NS5} & : & (012345)\nonumber\\ N_c\times {\rm D3} & : & (012|6|)\\ (p,q)5 & : & \left(012 \rot{3}{7}{\psi} \rot{4}{8}{\psi} \rot{5}{9}{-\theta} \right).\nonumber \end{eqnarray} As the result, the corresponding mass of the adjoint scalars become different. We can field theoretically understand this supersymmetry breaking. Originally $N{=}3$ vector multiplet contains the fields $(A_\mu,\lambda_i,X_i,\chi)$. If we give the different mass $\mu$ to the $X_1,X_2$ and $\lambda_3,\chi$, these fields form the $N{=}2$ neutral chiral supermultiplet $\Phi$. The rests of the fields $(A_\mu,\lambda_a,X_3)$ are contents of the $N{=}2$ vector multiplet, which still have the mass $\left|\frac{\kappa g^2}{4\pi}\right|$. In this $N{=}2$ theory three adjoint scalars are massive as long as $\psi\neq0$. So the Coulomb branch of this theory is still lifted and there is no moduli. This corresponds to the twisting configuration of the NS5-brane and the rotated $(p,q)$5-brane. However, if we set $\psi=0$ the NS5-brane and $(p,q)$5-brane become parallel in the $(x^3,x^4)$-space. Therefore D3-branes can now freely move in the $(x^3,x^4)$-space. In this configuration, two adjoint scalars $X_1$ and $X_2$ become massless and the Coulomb branch appears. \section{Broken and Unbroken Phases of Abelian Theory} In this section we consider the Higgs branches of the $N{=}2$ supersymmetric Chern-Simons theory. We first briefly review the description of the Higgs branch in the Hanany-Witten configuration. The inclusion of the matter hypermultiplets corresponds to the addition of the D5-branes, which extend in the $(x^0,x^1,x^2,x^7,x^8,x^9)$-space, between two NS5-branes. The matter hypermultiplets come from the open string stretched between D5-branes and D3-branes\footnote{There is a couple of chiral fields with opposite chirality with respect to orientation of strings.}. For simplicity, we consider only Abelian theory with one matter hypermultiplet, which includes two complex scalar fields $(q,\widetilde{q})$ from now on. If we denote the distance between the D5-brane and the D3-brane in the $(x^3,x^4,x^5)$-space as $d$, the mass of the hypermultiplet is given by \begin{equation} m = T_{\rm F1}d = \frac{d}{2\pi {l_s}^2}, \end{equation} where $T_{\rm F1}$ is the string tension and $l_s$ is the string length. In the decoupling limit of massive excitation of strings $l_s\rightarrow 0$, the quantity of $m$ must be fixed. The Higgs branch of this theory emanates from the point where the hypermultiplet is massless, namely, $d=0$ in the brane configuration. When $d=0$, the D5-brane can separate the D3-brane into two parts. As the result, two NS5-branes can shift each other's positions in the $(x^7,x^8,x^9)$-space. We denote this difference as $\widetilde{d}$. The gauge symmetry in this branch is completely broken since the D3-brane is divided into two parts by the D5-brane and all gauge degrees of freedom are decoupled. In the broken phase of the Abelian gauge theory, there exists a vortex state, which is an electric-magnetic dual object of the fundamental matter. Since the electric-magnetic duality of the brane effective theory is achieved by the Type IIB S-duality, the stringy counterpart of the vortex state is a D-string stretched between two separated parts of D3-brane. The mass of this state is \begin{equation} \widetilde{m} = T_{\rm D1}\widetilde{d} =\frac{\widetilde{d}}{2\pi {\widetilde{l_s}}^2}, \end{equation} where $\widetilde{l_s}={g_s}^{1/2} l_s$ is a string length of the dual theory. Filed theoretically, the mass of the vortex state is given by the Fayet-Iliopoulos (FI) parameter $\zeta$. If we define as \begin{equation} \zeta=\frac{\widetilde{m}}{4\pi}=\frac{\widetilde{d}}{8\pi^2{\widetilde{l_s}}^2}, \label{vortex mass} \end{equation} the FI parameter is proportional to the difference of two NS5-brane positions in the $x^9$-direction. Next, we consider a similar situation in the $N{=}2$ brane configuration with $(p,q)$5-brane. We add only one set of massless matter superfields which plays a role of Higgs field and consider the Abelian case, that is, the Maxwell Chern-Simons Higgs system. Moreover, we set the angles $\psi$ to $\pi/2$ in the $N{=}2$ brane configuration. So the brane configuration is as follows \begin{eqnarray*} {\rm NS5} & : & (012345)\\ {\rm D3} & : & (012|6|)\\ {\rm D5} & : & (012789)\\ (p,q)5 & : & \left(01278\rot{5}{9}{-\theta}\right). \end{eqnarray*} The two adjoint scalars $X_1$ and $X_2$ are decoupled from the theory and so we can easily pay attention to the true nature of the Higgs phase of the Abelian theory. The squark Higgs fields $q,\widetilde{q}$ in the matter superfields couple to a non-dynamical D-field in the off-shell vector multiplet. If we integrate out the auxiliary D-field we have a bosonic scalar potential \begin{equation} V=-\frac{g^2}{2}\left( |q|^2-|\widetilde{q}|^2-\zeta+\frac{\kappa}{4\pi} X_3 \right)^2 -{X_3}^2\left(|q|^2+|\widetilde{q}|^2\right), \label{potential} \end{equation} where $\zeta$ is the FI parameter. The potential of the aforementioned Maxwell Higgs theory is obtained by setting $\kappa=X_3=0$. In this case there is only the asymmetric phase $|q|^2-|\widetilde{q}|^2=\zeta$. However, the potential (\ref{potential}) admits two distinct vacua, an asymmetric phase and a symmetric phase. We examine the correspondence between each branch and the brane configuration from now on.\\\\ \underline{\it asymmetric phase} The asymmetric phase is described by $|q|^2-|\widetilde{q}|^2=\zeta$ and $X_3=0$. This phase is the same as the asymmetric phase in the previous case. So the corresponding configuration is also very similar to the Maxwell Higgs one. Namely, the D5-brane divide the D3-brane into two segments. One of the segments of the D3-brane stretch between the left NS5-brane and the D5-brane and the other is between the D5-brane and the right $(p,q)$5-brane. In this phase the relative position of the 5-branes along the $x^9$-direction can differ because of this split of the D3-brane and the gauge symmetry on the D3-brane is broken simultaneously. We denote the difference of 5-branes in the $x^9$-direction as $\widetilde{d}$. (See Fig.\ref{asymmetric}.) \begin{figure} \centerline{ \includegraphics[scale=0.9]{asymmetric.eps} } \caption{The brane configuration in the asymmetric phase.} \label{asymmetric} \end{figure} There also exists the topological vortex state in this broken phase of the Maxwell Chern-Simons Higgs system. However, in the $(p,q)$5-brane configuration we can not simply identify this vortex state with the D-string stretch between two fragments of the D3-brane. From the SUGRA solution of the $(p,q)$5-brane, there is a non-trivial vev of the axion field like as \cite{KOO} \begin{equation} \chi(x^6,x^7,x^8,x^9)= \frac{1}{g_s} \frac{\sin\theta\cos\theta(H^{1/3}-H^{2/3})} {H^{-1/3}\sin^2\theta+H^{2/3}\cos^2\theta}, \end{equation} where $H=1+\frac{{l_s}^2}{r^2}$ is a harmonic function on the $(x^6,x^7,x^8,x^9)$-space and $r$ is a distance from the $(p,q)$5-brane. We now assume that the distance $L$ between two 5-branes is satisfied $L\ll l_s$\footnote{This limit is connected with the pure Chern-Simons limit ($\kappa$ is fixed and $g\rightarrow\infty$).}. Since the D5-brane sits between two 5-branes, the distance from $(p,q)$5-brane $r$ is also satisfied $r\ll l_s$. In this limit, that is, the axion field take a constant value near the D5-brane \begin{equation} \chi=-\frac{1}{g_s}\tan\theta=-\kappa. \end{equation} In this background, D-string must appear as a dyon with electric charge proportional to $\chi$ by the Witten effect. In fact, since the tension of $(a,b)$-string in this background is given by \cite{Schwarz} \begin{equation} T_{(a,b)} =\frac{1}{2\pi{l_s}^2} \sqrt{\left(\frac{b}{g_s}\right)^2+\left(a+b\chi\right)^2}, \end{equation} the D-string, {\it i.e.} $(0,1)$-string, is not lightest state. A lighter state is a $(\kappa,1)$-string with the tension \begin{equation} T_{(\kappa,1)}=\frac{1}{2\pi{l_s}^2 g_s}, \end{equation} which is the same tension as the D-string in the Maxwell theory configuration. Therefore the vortex state near the $(p,q)$5-brane configuration is not only magnetically charged but also electrically charged. Field theoretically, the Chern-Simons interaction gives the relation between electric charge $Q$ and magnetic flux $\Phi$ \begin{equation} Q=\kappa \Phi \end{equation} from the Gauss law constraint. So the vortex solution carry both magnetic charge and electric charge in the ratio of 1 to $\kappa$. This consequence exactly agree with the $(\kappa,1)$-string state around the $(p,q)$5-brane background.\\\\ \underline{\it symmetric phase} For the symmetric phase the vevs of the scaler fields are \begin{eqnarray} &&q=\widetilde{q}=0,\\ &&X_3=4\pi\zeta/\kappa. \label{sym vev} \end{eqnarray} This phase does not exist in the Maxwell Higgs system. And also there is no corresponding configuration in the $N{=}4$ Hanany-Witten setup. However, in our case there exists another possible vacuum configuration for $\zeta\neq0$ due to the rotation of the 5-brane. If we want to set $\zeta\neq0$ we must shift the positions of the 5-branes each other along the $x^9$-direction. In general, the D3-brane suspended between the 5-branes prevents this motion. One of settlements is the division of the D3-brane into two parts as we mentioned before. Another solution is that the D3-brane slide up along the $x^5$-direction and is attached again on the intersecting point of the NS5-brane and the rotated $(p,q)$5-brane in the $(x^5,x^9)$-plane. (See Fig. \ref{symmetric}.) \begin{figure} \centerline{ \includegraphics[scale=0.8]{symmetric.eps} } \caption{The brane configuration in the symmetric phase} \label{symmetric} \end{figure} In this symmetric phase configuration, the position of the D3-brane in the $x^5$ coordinate is $x^5=\widetilde{d}/\tan\theta$ where $\widetilde{d}$ is a gap of the 5-brane positions. The gap $\widetilde{d}$ corresponds to the FI parameter $\zeta$ as in eq. (\ref{vortex mass}) and the $x^5$ position of the D3-brane corresponds to the vev of the adjoint scalar $X_3$. So we find that the adjoint scalar vev is given by \begin{eqnarray} X_3&=&\frac{\widetilde{d}}{2\pi {l_s}^2\tan\theta}\\ &=&\frac{4\pi\zeta}{p/q}. \end{eqnarray} This result is exactly agree with (\ref{sym vev}). Moreover, the distance between the D3 and D5-brane is $\widetilde{d}/\tan\theta$. This means that the mass of the matter superfields which is coming from the fundamental string between D3 and D5-brane is $4\pi\zeta/\kappa$. Correspondingly, when the $X_3$ field has the vev (\ref{sym vev}) the scalar potential (\ref{potential}) gives the mass $4\pi\zeta/\kappa$ to the $q$ and $\widetilde{q}$. This is also consistent with the brane picture. Finally, we would like to comment on vortex states in this phase. In this symmetric phase, it is known that there exists a non-topological vortex \cite{JW,JLW}. However, since charges of the non-topological vortex contain a continuous parameter, it seems to be difficult to find the corresponding state in string theory. \section*{Acknowledgments} I would like to thank T. Kitao and N. Ohta for arguments during the early stage of this work. I also grateful to B.-H. Lee for useful discussions and comments at Niseko Winter School. \input refs.tex \end{document}
\section*{Acknowledgements} I would like to thank Prof. J. Katriel for a critical reading of the manuscript and useful comments.
\section{Introduction} \label{secintro} Billiard systems with flux lines are simple dynamical systems that can model the typical behaviour of low dimensional quantum systems. One main application has been in the investigation of universal properties of quantum systems that are chaotic in the classical limit \cite{BR86}. The high lying states of these systems have statistical properties that depend only on the symmetries of the system and can be described by random matrix theory \cite{Boh91,Haa92,Meh91}. Ordinary billiards are used to model systems that are invariant under time inversion and are described by the Gaussian orthogonal ensemble (GOE). The introduction of an additional flux line allows to study a further universality class since the flux line breaks the time-reversal symmetry. If the flux strength is sufficiently large (when considering levels in a fixed energy range) the corresponding universality class is that of the Gaussian unitary ensemble (GUE), but by varying the flux strength one can investigate also the GOE-GUE transition or parametric correlations in the GUE regime. Another application is for integrable billiard systems where the introduction of a flux line can break the integrability of the quantum billiard \cite{DJM95}, or to use billiards with a flux line to model the properties of quantum dots (see e.g.\ \cite{BLM96}). A major advantage in all these applications is that the classical trajectories are not changed by the presence of a flux line, except for the set of measure zero that hit the flux line. In a semiclassical analysis the same set of periodic orbits appears independent of the flux strength and only the semiclassical contributions of these orbits are changed. For example, in approximations to the density of states the periodic orbit contributions have an additional phase $2 \pi m \alpha$ where $\alpha$ is related to the flux strength and $m$ is the number of times the orbit winds around the flux line. However, this is not the only semiclassical effect of a flux line. Its presence leads also to wave diffraction that can be semiclassically described by an additional set of trajectories that are closed but not periodic and which start and end on the flux line. In semiclassical arguments these so-called diffractive orbits are often neglected since their semiclassical contribution is estimated to be of order $\sqrt{\hbar}$ smaller than the contributions of periodic orbits. Such an estimate is, however, only valid if the scattering angle is not too close to the forward direction. In the forward direction diffractive orbits contribute in the same order as periodic orbits. In this article we investigate the semiclassical contributions of diffractive orbits to the density of states. We derive an approximation for orbits that are scattered once on a flux line. This approximation is valid for all scattering angles, also in the forward scattering direction. We show that these contributions are necessary in order to cancel discontinuities in periodic orbit contributions that occur when the position of the flux line is changed and crosses a periodic orbit. We also discuss the importance of diffractive orbits in the semiclassical limit. The article is organised as follows. In section 2 we consider the scattering on a flux line in a plane and derive a uniform approximation for the Green function. With this input we derive in section 3 a uniform approximation for semiclassical contributions to the density of states from isolated diffractive orbits that are scattered once on a flux line. As an example of a system with non-isolated diffractive orbits we consider in section 4 the integrable circular billiard with a flux line in its center. The results are discussed in section 5. \section{A flux line in a plane} \label{secplane} The scattering of a wave function on a flux line in a plane without boundaries has been studied in great detail since the problem was first treated by Aharonov and Bohm \cite{AB59}. There exists, for example, an exact integral representation for the propagator. A review on the quantum effects of electromagnetic fluxes is given in \cite{OI85}. In this section we review some of the results and derive a uniform approximation for the Green function that will be used in the following section. The Schr\"odinger equation for a particle with mass $M$ and charge $q$ in a magnetic field in two dimensions is given in Gaussian units by \begin{equation} \label{diffpla1} \frac{1}{2 M} \left[ \frac{\hbar}{i} \nabla - \frac{q}{c} \boldsymbol A \right]^2 \Psi(\r) = E \Psi(\r) \; , \end{equation} and for a flux line the vector potential can be chosen as \begin{equation} \label{vecpot} \boldsymbol A = \frac{\Phi}{2 \pi r} \hat{\phi} \; , \qquad \nabla \times \boldsymbol A = \Phi \delta(\r) \hat{z} \; , \qquad \nabla \cdot \boldsymbol A = 0 \; , \end{equation} which describes a magnetic field with flux $\Phi$ that is concentrated in form of a delta-function in the origin. In polar coordinates the Schr\"odinger equation with vector potential (\ref{vecpot}) has the form \begin{equation} \label{diffpla2} - \left[ \frac{\partial^2}{\partial r^2} + \frac{1}{r} \frac{\partial}{\partial r} + \frac{1}{r^2} \left( \frac{\partial}{\partial \phi} - i \alpha \right)^2 \right] \Psi(r, \phi) = k^2 \Psi(r,\phi) \; , \end{equation} where $\alpha = q \Phi/(2 \pi \hbar c)$ and $k = \sqrt{2 M E}/\hbar$. The solutions that are regular in the origin are given by \begin{equation} \label{solpla} \Psi_{m,k}(r,\phi) = \sqrt{\frac{k}{2 \pi}} \, J_{|m-\alpha|}(k r) \, \exp\{ i m \phi \} \; , \end{equation} where the normalisation is chosen such that \begin{equation} \int_0^{2 \pi} \! d \phi \; \int_0^\infty \! dr \; r \, \Psi_{m,k}(r,\phi) \, \Psi_{m',k'}^*(r,\phi) = \delta_{m,m'} \, \delta(k - k') \; , \end{equation} as follows from the orthogonality relations of the exponential and the Bessel functions \cite{Wat66}. It is sufficient to consider only values $0 \leq \alpha \leq 0.5$ since the solutions for other values of $\alpha$ can be obtained by a multiplication of a ($\phi$-dependent) phase factor and possibly a complex conjugation. With the normalised solutions (\ref{solpla}) the propagator can be written down directly \begin{align} \label{propa} K_\alpha(\r,\boldsymbol r_0,t) = \, & \sum_{m=-\infty}^\infty \int_0^\infty \! dk \; \Psi_{m,k}(r,\phi) \, \Psi_{m,k}^*(r_0,\phi_0) \, \exp\left\{-\frac{i}{\hbar} \frac{\hbar^2 k^2}{2 M} t \right\} \notag \\ & = \sum_{m=-\infty}^\infty \frac{M}{2 \pi i \hbar t} \, J_{|m-\alpha|} \left( \frac{r r_0 M}{\hbar t} \right) \, \exp \left\{ i m (\phi - \phi_0) - \frac{M(r^2 + r_0^2)}{2 i \hbar t} - i \frac{\pi}{2} |m- \alpha| \right\} \; , \end{align} and in the same way one obtains an exact representation for the Green function \begin{align} \label{greenex} G_\alpha(\r,\boldsymbol r_0,E) = \, & \lim_{\varepsilon \rightarrow 0} \frac{1}{i \hbar} \int_0^\infty \! dt \; K_\alpha(\r,\boldsymbol r_0,t) \, \exp \left\{ \frac{i}{\hbar} t (E + i \varepsilon) \right\} \notag \\ & = \lim_{\varepsilon \rightarrow 0} \sum_{m=-\infty}^\infty \int_0^\infty \! dk' \; \Psi_{m,k'}(r,\phi) \, \Psi_{m,k'}^*(r_0,\phi_0) \, \frac{1}{E + i \varepsilon - \frac{\hbar^2 {k'}^2}{2 M} } \notag \\ & = \sum_{m=-\infty}^\infty \frac{M}{2 i \hbar^2} \, e^{i m (\phi - \phi_0)} J_{|m-\alpha|} (k r_<) \, H_{|m - \alpha|}^{(1)}(k r_>) \; , \end{align} where $r_<$ and $r_>$ are the smaller and larger values of $r$ and $r_0$, respectively. For further evaluations the representation in the basis of the angular momentum eigenstates is not convenient. One can derive, however, from the sum in (\ref{propa}) an exact integral representation for the propagator (see \cite{OI85}). It involves an integral over Hankel functions. In the following we use an approximation to this integral representation that is obtained after replacing the Hankel functions by their leading asymptotic form. This gives a semiclassical approximation for the propagator that is valid if $0 < \alpha < 1$ and $M r r_0/t \gg \hbar$, i.\,e.\ for not too long times or too close distances to the flux line. \begin{align} \label{propuni} K_\alpha(\r,\boldsymbol r_0,t) \approx \, & \frac{M}{2 \pi i \hbar t} \exp \left\{ \frac{i M}{2 \hbar t} (\r - \boldsymbol r_0)^2 + i \alpha (\phi - \phi_0) \right\} \notag \\ & - \frac{M \sin(\alpha \pi)}{\pi i \hbar t} \exp \left\{ \frac{iM}{2 \hbar t} (r + r_0)^2 + \frac{i}{2} (\phi - \phi_0) \right\} \, K \left[ \sqrt{\frac{2 M r r_0}{\hbar t}} \, \cos \left( \frac{\phi - \phi_0}{2} \right) \right] \; . \end{align} The angles in (\ref{propuni}) have to be chosen such that $|\phi - \phi_0| \leq \pi$, and the function $K(z)$ is described below. The semiclassical propagator in (\ref{propuni}) consists of two parts. The first part is almost identical to the propagator in a free plane with the difference of an additional phase proportional to $\alpha$. Semiclassically it can be interpreted as the contribution from the direct path from $\r_0$ to $\r$. The additional phase results from the dependence of the Lagrangian on the vector potential. The second term in (\ref{propuni}) describes the scattering on the flux line and is discussed in more detail in connection with the Green function. Both terms are discontinuous in the forward direction $|\phi - \phi_0| = \pi$ but the sum is continuous and uniformly valid for all values of $\phi$ and $\phi_0$. The function $K(z)$ in (\ref{propuni}) is a modified Fresnel function that is defined by \begin{equation} K(z) = \frac{1}{\sqrt{\pi}} \exp\left\{ - i z^2 - i \frac{\pi}{4} \right\} \int_z^\infty \! dy \; e^{i y^2} = \frac{1}{2} e^{-i z^2} \operatorname{erfc} \left( e^{- i \pi/4} z \right) \; , \end{equation} and it has the following limiting properties \begin{equation} \label{kasym} K(0) = \frac{1}{2} \; ; \qquad \qquad K(z) \sim \frac{e^{i \pi/4} }{2 z \sqrt{\pi}} \; , \; \; |z| \rightarrow \infty \; , \; \; -\frac{\pi}{4} < \arg(z) < \frac{3 \pi}{4} \; . \end{equation} An important alternative representation of $K(z)$ is given by the integral ($\beta$, $z > 0$) \begin{equation} \label{k2} K(\sqrt{\beta} z) = \frac{1}{2 \pi i} \int_{-i \infty}^{i \infty} \! dx \; \frac{e^{-i \beta x^2}}{x + z} \; . \end{equation} Due to this form the function $K$ arises in semiclassical evaluations of oscillatory integrals in which a pole of the integrand is close to a stationary point. In figure \ref{figk} we show the real and imaginary parts of the function $K(z)$ for positive $z$. The function has its largest absolute value at $z=0$, and from approximately $z=3$ on it agrees well with its asymptotic approximation (\ref{kasym}). \begin{figure}[thb] \label{figk} \begin{center} \mbox{\epsfxsize8cm\epsfbox{kr.eps}} \mbox{\epsfxsize8cm\epsfbox{ki.eps}} \end{center} \caption{Real and imaginary parts of the modified Fresnel function $K(z)$ (full line) and its asymptotic approximation (dotted line).} \end{figure} With the relation between the propagator and the Green function in (\ref{greenex}) one can obtain a uniform approximation for the Green function. We consider the contributions from the two parts of the propagator in (\ref{propuni}) separately. These parts are called the geometrical and the diffractive part in the following \begin{equation} \label{greentot} G_\alpha(\r,\boldsymbol r_0,E) \approx G_g(\r,\boldsymbol r_0,E) + G_d(\r,\boldsymbol r_0,E) \; . \end{equation} For the first part one can evaluate the integral in (\ref{greenex}) by stationary phase evaluation and one obtains, analogously to the propagator, the free semiclassical Green function modified by a phase \begin{equation} \label{greengeo} G_g(\r,\boldsymbol r_0,E) = \frac{M}{\hbar^2 \sqrt{2 \pi k |\r - \boldsymbol r_0|}} \; \exp \left\{ i k |\r - \boldsymbol r_0| + i \alpha (\phi -\phi_0) - i \frac{3 \pi}{4} \right\} \; . \end{equation} For the second part we express the Fresnel function by the integral (\ref{k2}) and arrive at \begin{align} \label{laptra} G_d(\r,\boldsymbol r_0,E) & \approx - \frac{i M \sin(\alpha \pi)}{2 \pi^2 \hbar^2} \lim_{\varepsilon \rightarrow 0} \int_0^\infty \! dt \; \int_{-i \infty}^{i \infty} \! dz \; \frac{1}{t} \, \frac{1}{z + \cos \frac{\phi - \phi}{2}} \notag \\ & \qquad \qquad \times \exp \left\{ \frac{iM}{2 \hbar t} (r + r_0)^2 + \frac{i}{2} (\phi - \phi_0) - i \frac{2 M r r_0 z^2}{\hbar t} +\frac{i}{\hbar} t (E + i \varepsilon) \right\} \; . \end{align} This double integral has a stationary point at $(z,t)=(0,t_{cl})$ where $t_{cl} = \sqrt{M/(2E)}(r + r_0)$ is the classical time for the path from $\r$ to $\r_0$ via the origin at energy $E$. In order to evaluate the integral we expand the exponent in (\ref{laptra}) up to second order around the stationary point. Then the integral over $t$ can be evaluated by stationary phase approximation, since the stationary point at $t=t_{cl}$ is well separated from the pole at $t=0$. In the integral over $z$, however, the $z$-dependence of the denominator has to be taken into account and this yields again a modified Fresnel function. In this way a uniform approximation for the diffractive part of the Green function is obtained \begin{equation} \label{greenuni} G_d(\r,\boldsymbol r_0,E) = \frac{M \sin(\alpha \pi) \sqrt{2 \pi i}}{ \pi \hbar^2 \sqrt{k (r+r_0)}} \, \exp \left\{ i k (r + r_0) + \frac{i}{2} (\phi -\phi_0) \right\} \; K \left[ \sqrt{\frac{2 k r r_0}{r + r_0}} \, \cos \left( \frac{\phi - \phi_0}{2} \right) \right] \; , \end{equation} where again the angular coordinates have to be chosen such that $|\phi - \phi_0| \leq \pi$. This approximation for the Green function will be the ingredient for the derivation of semiclassical contributions of diffractive orbits in trace formulas in the next section. It is valid as long as $k r$, $k r_0 \gg 1$. We consider now an approximation to this expression that is valid if $|\phi - \phi_0|$ is not too close to $\pi$. Then the argument of the Fresnel function can be replaced by its leading asymptotic term (\ref{kasym}) and yields \begin{equation} \label{greengtd} G_d(\r,\boldsymbol r_0,E) \sim \frac{M \sin(\alpha \pi)}{2 \pi k \hbar^2 \sqrt{r r_0} \cos \left( \frac{\phi - \phi_0}{2} \right) } \exp \left\{ i k (r + r_0) + \frac{i}{2} (\phi - \phi_0) + i \frac{\pi}{2} \right\} \end{equation} This approximation is of the general form that is obtained within the Geometrical Theory of Diffraction (GTD) (see e.\,g.\ \cite{Kel62}) \begin{equation} \label{gtd1} G_d(\r,\boldsymbol r_0,E) \approx \frac{\hbar^2}{2 M} \, G_0(\r,0,E) \, {\cal D}(\phi,\phi_0) \, G_0(0,\boldsymbol r_0,E) \; . \end{equation} In this theory the scattering is described by a free Green function from $\r_0$ to the scattering source at the origin, multiplied by a diffraction coefficient ${\cal D}(\phi,\phi_0)$ that contains the information about the particular scattering process and a further free Green function from the origin to the point $\r$. A comparison with (\ref{greengtd}) shows that the diffraction coefficient for the present case is given by \begin{equation} \label{gtd2} {\cal D}(\phi,\phi_0) = \frac{2 \sin(\alpha \pi)}{ \cos \left(\frac{\phi - \phi_0}{2} \right) } \exp\left\{ i \frac{\phi - \phi_0}{2} \right\} \; . \end{equation} The term (\ref{greengtd}) can be interpreted as the contribution of a trajectory that runs from $\r_0$ to the origin and then to $\r$. It is of order $k^{-1/2}$ smaller than the contribution of the direct trajectory from $\r_0$ to $\r$ in (\ref{greengeo}). However, the GTD-approximation breaks down in the forward direction $|\phi-\phi_0|=\pi$ where the diffraction coefficient (\ref{gtd2}) diverges. This reflects the fact that the diffractive part of the Green function has a different leading asymptotic term in the forward direction. Here the diffractive trajectory contributes in the same order in $k$ as the direct trajectory. The uniform approximation (\ref{greenuni}) interpolates between these two different asymptotic regimes. In figure \ref{figgreen} we compare the different approximations to the Green function with the exact result (\ref{greenex}). The dotted line is the geometrical part (\ref{greengeo}). The approximation is already reasonably good, but one can still see a clear deviation from the exact curve. By adding the diffractive part in the GTD-approximation (dashed line) the difference becomes much smaller for most values of $\phi - \phi_0$, except near the endpoints $\phi - \phi_0 = \pm \pi$ where it diverges. This divergence is removed by the uniform approximation in (\ref{greenuni}) (dashed-dotted line) for which the difference with the exact line can hardly be seen, even though we chose relatively small values of $k r$ and $k r_0$. \begin{figure}[htb] \label{figgreen} \begin{center} \mbox{\epsfxsize8cm\epsfbox{gr.eps}} \mbox{\epsfxsize8cm\epsfbox{gi.eps}} \end{center} \caption{Real and imaginary parts for different approximations of the Green function $G(\r,\r_0,E)$ with $r=1$, $r_0=2$, $k=5$ and $\alpha = 0.4$. Full line: exact result, dotted line: $G_g(\r,\r_0,E)$, dashed line: GTD-approximation, and dashed-dotted line: uniform approximation.} \end{figure} \section{Semiclassical Contributions of Isolated Diffractive Orbits} \label{secbound} For the derivation of semiclassical approximations in billiard systems it is convenient to apply the boundary integral method. It provides an alternative formulation of the quantum mechanical eigenvalue problem in terms of an integral equation along the billiard boundary. All semiclassical contributions due to diffraction in billiard systems that go beyond GTD have so far been derived by this method \cite{PSSU96,SPS97}. The boundary integral method has been developed for billiard systems without internal fields (see e.g.\ \cite{KR74,Rid79}), but it can be extended to more general situations. In \cite{TCA97} Tiago et al.\ derived an integral equation for billiard systems with magnetic field which is described by a vector potential in the Coulomb gauge. The resulting integral equation has the same form as for billiards without field. For a billiard system with a flux line and Dirichlet boundary conditions one obtains \begin{equation} \label{inteq} - 2 \int_{\partial {\cal D}} \! d s' \, \partial_{\hat{n}} \, G_\alpha(\r\,',\r,E) \, u(\r\,') = u(\r) \end{equation} where from here on we use units in which $\hbar = 2M = 1$. The difference to the field free case is that the Green function in (\ref{inteq}) is the one for a flux line in a plane and the solutions $u(\r)$ are the complex conjugate of the normal derivatives of the wave functions. The form of the integral equation in (\ref{inteq}) is different from that in \cite{TCA97} since we applied an additional normal derivative to the equation in order to obtain a non-divergent integral kernel. Also the Green function differs by a factor $(-8\pi)$ from that in \cite{TCA97}. In (\ref{inteq}) the points $\r$ and $\r\,'$ lie on the boundary, the position of $\r\,'$ is parametrised by $s'$, and $\partial_{\hat{n}}$ denotes the outward normal derivative at $\r$. The integral is evaluated along the boundary $\partial {\cal D}$ of the billiard system. The left-hand-side of (\ref{inteq}) can be abbreviated by $\hat{Q} u(\r)$ where $\hat{Q}$ is the integral operator acting on $u$. Equation (\ref{inteq}) is a Fredholm equation of the second kind and it has non-trivial solutions only if the determinant $\Delta(E) := \det(\boldsymbol{1} - \hat{Q}(E))$ vanishes. This condition determines the quantum energies of the problem. For a summary of the properties of the Fredholm determinant $\Delta(E)$ for two-dimensional billiards without fields with corresponding references see \cite{Sie98}. From the condition of the vanishing of the Fredholm determinant one can obtain an expression for the density of states in terms of the integral operator $\hat{Q}$. It is given by \begin{equation} \label{secc1} d(k) = d_{\text{sm}}(k) + \frac{1}{\pi} \frac{d}{d k} \Im \sum_{n=1}^\infty \frac{1}{n} \operatorname{Tr} \hat{Q}^n (k)\; , \end{equation} where \begin{equation} \label{secc2} \operatorname{Tr} \, \hat{Q}^n (k) = (-2)^n \int_{\partial {\cal D}} \! d s_1 \dots d s_n \; \partial_{\hat{n}_1} G_\alpha(\r_2,\r_1,E) \; \partial_{\hat{n}_2} G_\alpha(\r_3,\r_2,E) \dots \partial_{\hat{n}_n} G_\alpha(\r_1,\r_n,E) \; . \end{equation} The sum in (\ref{secc1}) is not convergent for real $k$, and for the derivation we assume that $k$ has a sufficiently large imaginary part and that $\Re k > 0$. The density of states in (\ref{secc1}) has a smooth part and a sum over oscillatory integrals. A semiclassical evaluation of the oscillatory integrals yields the leading semiclassical contributions to the density of states from periodic orbits and diffractive orbits. These contributions can be separated by writing $G_\alpha$ as sum of its geometrical and its diffractive part. Then (\ref{secc2}) consists of $2^n$ terms. The periodic orbit contributions are contained in the term that contains only geometrical Green functions. By evaluating this term in stationary phase approximation one obtains the usual Gutzwiller expression for isolated periodic orbits \cite{Gut71}, modified by an additional phase $2 \pi m \alpha$ where $m$ is the winding number of the orbit around the flux line. The terms with $l$ diffractive Green functions on the other hand contain the contributions of diffractive orbits that are scattered $l$ times on the flux line. In this article we consider only diffractive orbits that are scattered once on a flux line. We thus have to evaluate an integral with a product of $n-1$ geometrical Green functions and one diffractive Green function. This integral can be considerably simplified by applying a composition law for the geometrical part of the Green functions. \begin{equation} \label{secc5} G_g^{(n)}(\r,\r\,',E) \approx (-2)^n \int_{\partial {\cal D}} \! d s_1 \dots d s_n \; G_g(\r_1,\r\,',E) \; \partial_{\hat{n}_1} G_g(\r_2,\r_1,E) \dots \partial_{\hat{n}_n} G_g(\r,\r_n,E) \; . \end{equation} where the approximate sign denotes that the integrals are evaluated in stationary phase approximation. The composition law was proved for ordinary billiard systems in \cite{SPS97} and it is a semiclassical version of the multiple reflection expansion of the Green function of Balian and Bloch. Since the presence of a flux line adds only a phase to the Green function which doesn't effect the stationary points in leading order, the composition law holds also in the present case and the phase is simply additive. The semiclassical expression for $G_g^{(n)}$ is then (see \cite{SPS97}) \begin{equation} \label{secc7} G_g^{(n)} (\r_2,\r_1,E) = \sum_{\xi_n} \frac{1}{\sqrt{8 \pi k |\tilde{M}_{12}|}} \exp \left\{ i k \tilde{L} - i \frac{\pi}{2} \tilde{\nu} - i \frac{3 \pi}{4} + i \alpha \phi_{21} \right\} \; . \end{equation} Here $\phi_{21}$ is the total winding angle of the trajectory around the flux line. It can be written as $\phi_{21} = 2 \pi m + \Delta \phi$ where $m$ is an integer and $|\Delta \phi| \leq \pi$. Furthermore, $\tilde{M}_{12}$ is the $12$-element of the stability matrix, $\tilde{L}$ is the length of the trajectory, and $\tilde{\nu}$ is the number of conjugate points from $\r_1$ to $\r_2$. The stability matrix $\tilde{M}$ is evaluated at unit energy and is energy independent. We use a tilde here in order to distinguish the quantities from the corresponding ones for the diffractive orbit. With (\ref{secc5}) one obtains the following expression for the partial contribution to the level density from diffractive orbits with one scattering event \begin{equation} \label{startap} d_{part} = \frac{4}{\pi} \Im \frac{d}{d k} \int_{\partial {\cal D}} \! d s_1 \, d s_2 \; \partial_{\hat{n}_1} G_d(\r_2,\r_1,E) \; \partial_{\hat{n}_2} G_g^{(n-1)}(\r_1,\r_2,E) \; . \end{equation} We insert now (\ref{secc7}) and (\ref{greenuni}) with (\ref{k2}) for the semiclassical and the diffractive Green function, respectively. The normal derivative gives in both cases in leading order a factor $i k \cos \alpha$ where $\alpha$ is the angle between the normal and the outgoing trajectory. We obtain \begin{equation} d_{part} = \Im \frac{d}{dk} \frac{k \cos \alpha_1 \cos \alpha_2 \sin(\alpha \pi)}{ 2 \pi^3 \sqrt{|\tilde{M}_{12}| (r_1 + r_2)}} \exp \left\{ - i \frac{\pi}{2} \tilde{\nu} + i \alpha \phi_{21} - i \frac{\Delta \phi}{2} \right\} \, I \; , \end{equation} where \begin{equation} \label{diffint} I = \int_{\partial {\cal D}} \! d s_1 \, d s_2 \, \int_{-i \infty}^{i \infty} \! dz \; \frac{\displaystyle \exp \left\{ i k (\tilde{L} + r_1 + r_2) - i \frac{2 k r_1 r_2}{r_1 + r_2} z^2 \right\} }{\displaystyle z + \cos \frac{\Delta \phi}{2}} \; . \end{equation} The main part in the derivation consists in the evaluation of the diffraction integral $I$. If all three integrals are evaluated by a stationary phase approximation one obtains the contribution of the diffractive orbit in the GTD approximation. This expression diverges when $|\Delta \phi|=\pi$. In order to obtain a semiclassical approximation that is uniformly valid for all angles one has to take the dependence of the denominator on the integration variables into account. One can do this by expanding the exponent in (\ref{diffint}) up to second order in $s_1$, $s_2$ and $z$, and the denominator up to first order in these variables. In the denominator one obtains in this way \begin{equation} \label{denom} \cos \frac{\Delta \phi}{2} \approx \cos \frac{\Delta \phi_0}{2} + \frac{1}{2} \left( \frac{s_1 \cos \alpha_1}{r_1} - \frac{s_2 \cos \alpha_2}{r_2} \right) \sin \frac{\Delta \phi_0}{2} \; . \end{equation} This method corresponds to a particular choice of a uniform approximation. It yields an approximation that is correct at the `optical boundary' $|\Delta \phi|=\pi$ and in the GTD limit, and it interpolates between them. There are other possibilities to obtain a uniform approximation, for example by mapping the exponent onto a quadratic function in the three variables and then choosing an appropriate approximation for the amplitude function. This reflects the fact that uniform approximations are in general not unique. For example, a different interpolating approximation can be obtained by dropping the sine-term in (\ref{denom}) which corresponds to replacing it by its value at the optical boundary. The evaluation of the integral $I$ is rather lengthy. It consists in carrying out the expansions and performing linear transformations of the variables such that, finally, the denominator depends only on one of the variables and the integral can be expressed in terms of the modified Fresnel function. A comparison with \cite{SPS97} shows, however, that the same diffraction integral occurs for the diffraction on billiard corners. For that reason, the calculations don't have to be done again and the result can be inferred from this paper. One has ($c>0$) \begin{align} \label{intres} & \int_{-\infty}^\infty \! d s_1 \! \int_{-\infty}^\infty \! d s_2 \! \int_{-i \infty}^{i \infty} \! dz \; \frac{\displaystyle \exp \left\{ i k (\tilde{L} + r_1 + r_2 - c\,z^2) \right\} }{\displaystyle a + z + b \left( \frac{s_1 \cos \alpha_1}{2 r_1} - \frac{s_2 \cos \alpha_2}{2 r_2} \right) } \notag \\[5pt] = & - \frac{4 \pi^2 \operatorname{sign}(a) \tau}{k |b| \cos \alpha_1 \cos \alpha_2} \sqrt{\frac{2 r_1 r_2 |\tilde{M}_{12}|}{c |\operatorname{Tr} M - 2|}} \, K \left[ i^{\displaystyle \kappa} \left| \frac{a}{b} \right| \sqrt{\frac{2 k |M_{12}|}{|\operatorname{Tr} M - 2|}} \right] \text{\Large e}^{\displaystyle i k L - i \pi(\nu + \kappa - \tilde{\nu})/2} + b.c. \; , \end{align} where b.c.\ denotes a boundary contribution whose origin is the discontinuity of the diffractive part of the Green function. It is cancelled by a corresponding boundary contribution from the geometrical part of the Green function. In (\ref{intres}) $L$ and $M$ denote the length and stability matrix of the orbit, respectively, and $\nu$ is the number of conjugate points along the orbit. $\kappa$ is given by \begin{equation} \kappa = \begin{cases} 0 & \text{if} \frac{M_{12}}{\operatorname{Tr} M - 2} > 0 \, , \\ 1 & \text{if} \frac{M_{12}}{\operatorname{Tr} M - 2} < 0 \, , \end{cases} \end{equation} and $\tau = 1 - 2 \kappa$. With (\ref{intres}) and noting that the derivative gives in leading order a factor $(i L)$ one obtains the final result for the contribution of a diffractive orbit $\xi$ \begin{equation} \label{final} d_\xi(k) = - \Re \left( \frac{\displaystyle 2 \tau L \sin(\alpha \pi) \, \text{\Large e}^{\displaystyle i k L - i \pi \mu /2 + i \phi_{21} \alpha - i \Delta \phi/2} }{\displaystyle \pi \, |\sin(\frac{\Delta \phi}{2})| \, \sqrt{|\operatorname{Tr} M - 2|}} \, K \left[ i^{\displaystyle \kappa} \left| \cot \frac{\Delta \phi}{2} \right| \sqrt{\frac{2 k |M_{12}|}{|\operatorname{Tr} M - 2|}} \right] \, \right) \end{equation} where $\mu = \nu + \kappa$ agrees with the usual definition of the Maslov index for periodic orbits. Formula (\ref{final}) is the main result of this article. It describes the contribution of a diffractive orbit to the density of states and is valid for all scattering angles $\Delta \phi$, and for $0 < \alpha < 1$. A slightly simpler approximation can be obtained by dropping the sine-term in (\ref{final}) and replacing the cotangent by a cosine, corresponding to the discussion after (\ref{denom}). The use of the full Fresnel function in (ref{final}) is necessary in an angular region around the forward direction with width of order $k^{-1/2}$. For angles outside this range the Fresnel function can be replaced by its leading asymptotic form which results in \begin{equation} \label{finlim} d_\xi(k) \approx \Re \left( \frac{L \sin(\alpha \pi)}{\pi \cos(\Delta \phi/2)} \frac{1}{\sqrt{2 \pi k |M_{12}|}} \exp \left\{ i k L - i \frac{\pi}{2} \nu + i \alpha \phi_{21} - \frac{i}{2} \Delta \phi - i \frac{3 \pi}{4} \right\} \right) \; . \end{equation} This agrees with the GTD approximation for the diffractive orbit \cite{GTD} \begin{equation} \label{fingtd} d_\xi(k) = \Re \left( \frac{L}{\pi} G^\xi_g(\r_2,\r_1) {\cal D}(\phi_1,\phi_2) \right) \, . \end{equation} It is a contribution that is of order $\sqrt{k}$ smaller than the contribution of a isolated periodic orbit. In (\ref{fingtd}) $G_g^\xi$ denotes the contribution from the trajectory $\xi$ to the geometrical part of the Green function where $\r_1=\r_2$ is the position of the flux line. In the opposite limiting case at $|\Delta \phi| = \pi$ the uniform approximation is of the same order as that of a periodic orbit and it is discontinuous. As $|\phi|$ goes through $\pi$ the contribution makes a step of size \begin{equation} \label{discon} \left. d_\xi(k) \right|_{\Delta \phi = \pi-0} - \left. d_\xi(k) \right|_{\Delta \phi = -\pi+0} = - \frac{2 \tau L \sin(\alpha \pi)}{\pi \sqrt{|\operatorname{Tr} M - 2|}} \sin \left(k L - \frac{\pi}{2} \mu + (2m+1) \pi \alpha \right) \, . \end{equation} In order to understand this discontinuity we consider an angle $\Delta \phi = \pi - \varepsilon$ which is infinitesimally different from $\pi$. By examining the linearised motion around the diffractive orbit one finds that there is, in general, a periodic orbit in an infinitesimal neighbourhood of the diffractive orbit. This orbit is obtained from the conditions that $dq=dq'$ and $dp=dp'-k\varepsilon$ where the primed and the unprimed quantities are infinitesimal perpendicular deviations from the starting point and endpoint of the diffractive orbit, respectively. From these conditions one finds that the position of the periodic orbit is given by $dq=-k \varepsilon M_{12}/(\operatorname{Tr} M - 2)$. Depending on the sign of the left-hand-side, i.e.\ on the value of $\kappa$, there can be two possible cases that are shown in figure \ref{figtraj} for positive $\Delta \phi$ near $\pi$. \begin{figure}[htb] \label{figtraj} \begin{center} \mbox{\epsfxsize8cm\epsfbox{traj.eps}} \end{center} \caption{Parts of a diffractive orbit and a nearby periodic orbit for $\kappa=0$ and $\kappa=1$.} \end{figure} If one moves the flux line in a way that $\Delta \phi$ goes through $\pi$ then the flux line crosses this periodic orbit at the same instant. For the periodic orbit the semiclassical contribution is also discontinuous since the winding number of the periodic orbit changes. For a primitive orbit the winding number changes by $+1$ if $\kappa=0$ and by $-1$ if $\kappa=1$. By writing the product of sines in (\ref{discon}) as a sum of two cosines one can show that the two discontinuities cancel exactly and the sum of both contributions is continuous. This shows that the uniform approximation for diffractive contributions is necessary in order to make semiclassical approximations continuous if the position of a flux line is changed. \section{Diffractive orbits in the circular billiard} \label{seccirc} A simple example in which diffractive orbits are not isolated is a circular billiard with a flux line in its centre. This system is integrable since the energy and the angular momentum around the centre are conserved. All diffractive orbits run from the flux line to the boundary and are reflected back directly onto the flux line. They appear in one- or more-parameter families, and their lengths are multiples of $2R$ where $R$ is the radius of the circle. Contributions of these diffractive orbits to the density of states have been observed in \cite{RBMBM96}. We are interested in the leading order semiclassical contributions of the diffractive orbits. For this purpose one cannot apply formula (\ref{final}) since the orbits have a stability matrix with trace two. Although one can in principle obtain the diffractive contributions from the boundary element method, it is now much more convenient to start from the torus-quantisation conditions for integrable systems (EBK-conditions). In this way one can obtain all leading order diffractive contributions to the density of states at once, even those for multiple diffraction. The exact solutions of the Schr\"odinger equation are given by the solutions of the flux line in a plane (\ref{solpla}) with a different normalisation constant and the additional condition that the wave functions have to vanish at $r=R$. From this condition it follows that the energies are determined by the zeros of the Bessel functions $E_{m,n} = \hbar^2 k_{m,n}^2/(2M)$, where $k_{m,n} = j_{|m-\alpha|,n+1}/R$, and $j_{\nu,n}$ is the $n$-th zero of the Bessel function with index $\nu$. The EBK-quantisation conditions yield semiclassical approximations to these energy levels. They have the form \begin{equation} \label{ebk1} \oint \! d\phi \; p_\phi = 2 \pi \hbar m \; , \; \; m \in \mathbb{Z} \; , \qquad \quad \oint \! dr \; p_r = 2 \pi \hbar \left( n + \frac{3}{4} \right) \; , \; \; n = 0, 1, 2, \dots \end{equation} Due to the conservation of the angular momentum the first condition gives $p_\phi = \hbar m$ and the second condition is evaluated with the energy conservation law $ E = ( p_r^2 + (p_\phi - \hbar \alpha)^2/r^2)/(2M)$ and yields \begin{equation} \label{ebk2} \sqrt{k_{m,n}^2 R^2 - (m-\alpha)^2} - |m-\alpha| \, \arccos \frac{|m-\alpha|}{k_{m,n} R} = \pi \left( n + \frac{3}{4} \right) \end{equation} which determines $k_{m,n}$ as a function of the two quantum number $m$ and $n$. From (\ref{ebk2}) the periodic orbit contributions to the density of states are obtained by applying the Poisson summation formula \cite{BT76} \begin{align} \label{poisson} d(k) & = \sum_{M,N=-\infty}^\infty \int_{-\infty}^\infty \! dm \; \int_{-3/4}^\infty \! dn \; e^{2 \pi i (Mm + Nn)} \, \delta ( k - k_{m,n}) \notag \\ & = \sum_{M,N=-\infty}^\infty \int_{\alpha-kR}^{\alpha+kR} \! dm \; \frac{1}{\pi k} \sqrt{k^2 R^2 - (m - \alpha)^2} \, \exp\{2 \pi i (Mm + Nn(m,k))\} \notag \\ & = \sum_{M,N=-\infty}^\infty \int_0^{kR} \! dm \; \frac{1}{\pi k} \sqrt{k^2 R^2 - m^2} \times \notag \\ & \qquad \exp \left\{ 2 \pi i M (m+\alpha) + 2 i N \left[ \sqrt{k^2 R^2 - m^2} - m \arccos \frac{m}{kR} \right] - \frac{3 \pi}{2} i N \right\} + [\alpha \rightarrow -\alpha] \; . \end{align} We consider now the main semiclassical contributions to the integrals in (\ref{poisson}). Altogether these are four contribution. One is the only non-oscillatory term $M=N=0$, which yields the leading area-term for the mean density of states \begin{equation} d_A(k) = \frac{2}{\pi k} \int_0^{kR} \! dm \; \sqrt{k^2 R^2 - m^2} = \frac{1}{2} k R^2 \; . \end{equation} This agrees with the leading term in Weyl's law $Ak/(2 \pi)$ with area $A=\pi R^2$. The second main contributions arises from the stationary points of the integrals. A stationary phase approximation gives the contributions of the periodic orbits that have been obtained also in \cite{RBMBM96}. \begin{equation} \label{circpo} d_{po}(k) = \sum_{N=2}^\infty \sum_{M+1}^{[N/2]} g_{M,N} \sqrt{\frac{4k}{\pi N} R^3 \sin^3 \frac{\pi M}{N} } \cos \left( 2 N k R \sin \frac{\pi M}{N} - \frac{3 \pi}{2} N + \frac{\pi}{4} \right) \, \cos (2 \pi M \alpha) \end{equation} where $g_{M,N} = 1$ if $M = N/2$ and $g_{M,N} = 2$ if $M \neq N/2$. This factor arises from the fact that the stationary points for $M=N/2$ lie on the endpoint $m=0$ of the integrals in (\ref{poisson}) and give only half the contribution. The remaining two contributions follow from the boundaries of the oscillatory integrals. They can be obtained by an integration by part which yields \begin{equation} \label{bc} \text{b.\,c.\ of} \; \left\{ \int_{-\infty}^z \! dx \; g(x) \, e^{if(x)} \right\} = - i \frac{g(z)}{f'(z)} \, e^{if(z)} \; , \end{equation} for an upper end of an integration range. If $z$ is on the lower end of an integration range the end point contribution is given by minus the right hand side of (\ref{bc}). With (\ref{bc}) one obtains for the contribution from $m=kr$ \begin{align} \label{peri} d_L(k) & = \sum_{N,M=-\infty}^\infty \!\!\!\!\!\!{}' \, \, \frac{1}{\pi k} \left. \frac{\sqrt{k^2 R^2 - m^2} \exp \left\{ 2 \pi i M (m+\alpha) - \frac{3 \pi}{2} i N \right\} }{ 2 \pi i M - 2 i N \arccos \frac{m}{k R} } \right|_{ m \rightarrow k R} + [\alpha \rightarrow -\alpha] \notag \\ & = \frac{2 R}{\pi} \sum_{N=1}^\infty \frac{\sin \left( \frac{3 \pi}{2} N \right)}{N} \notag \\ & = - \frac{R}{2} \; . \end{align} where the prime denotes that the term $(N,M)=(0,0)$ is excluded from the sum since it corresponds to a non-oscillatory integral. The result in (\ref{peri}) can be identified with the perimeter term in the asymptotic expansion of the mean density of states. From the other end points of the integrals at $m=0$ one obtains \begin{align} d_d(k) & = \frac{iR}{2 \pi^2} \sum_{\substack{M,N=-\infty \\ M \neq N/2}}^\infty \frac{1}{M - \frac{N}{2}} \exp \left\{ 2 \pi i M \alpha + 2 i N k R - \frac{3 \pi}{2} i N \right\} + [\alpha \rightarrow -\alpha] \notag \\ & = d_d^e(k) + d_d^o(k) \; . \end{align} Here the terms $M=N/2$ have to be excluded from the sum since they have already been taken into account by the stationary phase evaluation. For convenience the sum is split into two parts, corresponding to a summation over even and odd values of $N$, respectively. For the first part $N$ is replaced by $2N$, the summation index $M$ is shifted by $N$, and terms for positive and negative values of $M$ are combined \begin{align} \label{de} d_d^e(k) & =- \frac{R}{\pi^2} \sum_{N=-\infty}^\infty \sum_{M=1}^\infty \frac{\sin [ 2 \pi \alpha M]}{M} \exp \left\{ i 2 N (\pi \alpha + 2 k R - \frac{3 \pi}{2})\right\} + [\alpha \rightarrow -\alpha] \notag \\ & = \frac{2 R}{\pi} (1 - 2 \alpha) \sum_{N=1}^\infty \sin \left( 2N(2kR - \frac{3 \pi}{2}) \right) \sin(\pi \alpha 2N) \; , \end{align} and for the second part $N$ is replaced by $2N+1$, the summation index $M$ is shifted by $N$, and terms for $M \geq 1$ and $M<1$ are combined \begin{align} \label{do} d_d^o(k) & = - \frac{2 R}{\pi^2} \sum_{N=-\infty}^\infty \sum_{M=1}^\infty \frac{\sin [ \pi \alpha ( 2 M - 1) ]}{2 M - 1} \exp \left\{ i (2 N + 1)(\pi \alpha + 2 k R - \frac{3 \pi}{2})\right\} + [\alpha \rightarrow -\alpha] \notag \\ & = \frac{2 R}{\pi} \sum_{N=0}^\infty \sin \left( (2N+1)(2kR - \frac{3 \pi}{2}) \right) \sin(\pi \alpha (2N+1)) \; . \end{align} The formulas for evaluating the summations in (\ref{peri}), (\ref{de}) and (\ref{do}) can be found in \cite{GR80} and are valid for $0 < \alpha < 1$. Collecting all results, the complete trace formula is given by \begin{align} \label{trace} d(k) \approx \; & \bar{d}(k) + \sum_{N=2}^\infty \sum_{M=1}^{[N/2]} g_{M,N} \sqrt{\frac{4k}{\pi N} R^3 \sin^3 \frac{\pi M}{N} } \cos \left( 2 N k R \sin \frac{\pi M}{N} - \frac{3 \pi}{2} N + \frac{\pi}{4} \right) \, \cos (2 \pi M \alpha) \notag \\ & + \sum_{N=1}^\infty \frac{2R}{\pi} \sin \left( 2 N k R - \frac{3 \pi}{2} N \right) \sin(\pi N \alpha) - 2 \alpha \sum_{N=1}^\infty \frac{2R}{\pi} \sin \left( 4 N k R - 3 \pi N \right) \sin(2 \pi N \alpha) \; . \end{align} It consists of the mean level density, the contributions of periodic orbits, and the contributions of diffractive orbits which are by an order $k^{-1/2}$ smaller than those of the periodic orbits. The formula is similar to the corresponding one for a harmonic oscillator with a flux line \cite{BBLMM95}. The diffractive orbit term in (\ref{trace}) can be given a more direct interpretation since it can be transformed into a sum of delta functions by using the Poisson summation formula. One obtains \begin{align} \label{ddiff} d_d(k) = & - \frac{R}{2 \pi} \sum_{n=-\infty}^\infty \left[ \delta \left( \frac{kR}{\pi} - \frac{3}{4} + \frac{\alpha}{2} + n \right) - \delta \left( \frac{kR}{\pi} - \frac{3}{4} - \frac{\alpha}{2} + n \right) \right] \notag \\ & + \frac{\alpha R}{\pi} \sum_{n=-\infty}^\infty \left[ \delta \left( \frac{2kR}{\pi} - \frac{3}{2} + \alpha + n \right) - \delta \left( \frac{2kR}{\pi} - \frac{3}{2} - \alpha + n \right) \right] \end{align} The argument of one of the delta functions can be identified with the semiclassical quantisation condition for eigenvalues with vanishing angular momentum ($m=0$): $kR \approx \pi \alpha / 2 + 3 \pi/4 + n \pi$. Since the full trace formula produces delta peaks at the eigenvalues given by the EBK conditions (\ref{ebk2}), the diffractive orbits have the following role: they contribute to peaks at eigenvalues with vanishing angular momentum and they cancel wrong peaks that are produced by the periodic orbit sum. The fact that the periodic orbits produce wrong peaks can be understood by noting that the periodic orbit contributions (\ref{circpo}) are invariant under $\alpha \rightarrow -\alpha$. Due to this symmetry the periodic orbits give also rise to wrong peaks at $kR \approx - \pi \alpha / 2 + 3 \pi/4 + n \pi$. The first two terms in (\ref{ddiff}) provide a correction to this failure in the approximation for states with zero angular momentum. The other two terms in (\ref{ddiff}) are necessary in order that the approximation is invariant under $\alpha \rightarrow 1 - \alpha$, since the spectrum is invariant under this replacement. A test of the trace formula (\ref{trace}) is presented in figure \ref{figcirc} which shows a comparison between quantum mechanical and semiclassical results. The value of $\alpha$ is chosen to be $\alpha=0.25$, since for this value the periodic orbit contributions start at $L=8R$. The difference between the two curves cannot be seen on this scale. \begin{figure}[htb] \label{figcirc} \begin{center} \mbox{\epsfxsize8cm\epsfbox{d1.eps}} \end{center} \caption{Fourier transform of the oscillatory part of the density of states for the circular billiard with $\alpha=0.25$ and $R=1$ (full line) in comparison with the semiclassical approximation (dotted line). Inset: Magnification of one peak.} \end{figure} \section{Conclusions} \label{seccon} One general property of uniform approximations for isolated diffractive orbits is that they make semiclassical approximations continuous as a system parameter is changed. They cancel discontinuities in semiclassical contributions of periodic orbits. In billiards with concave boundaries or with corners those discontinuities are connected with the appearance or disappearance of periodic orbits. In the present case the discontinuity is due to a phase change in periodic orbit contributions. Besides that, the diffraction on a flux line is very similar to the diffraction on corners. The final formula is expressed in terms of the same interpolating function, but it is also simpler since it consists only of one term instead of four. This suggests, for example, that a semiclassical study of two-dimensional systems which are non-integrable but in which the classical motion is still restricted to a two-dimensional surface in phase space might be performed more easily on billiards with a flux line instead of pseudo-integrable polygonal billiards. In this article we considered only diffractive orbits that are scattered once on a flux line. The same method can be used to obtain semiclassical contributions of diffractive orbits with multiple scattering, but the formulas become increasingly more complex and have to be expressed in terms of multiple Fresnel integrals. The treatment of an integrable billiard with a flux line like the circular billiard is much simpler. There all leading order diffractive contributions including those from multiple diffraction can be obtained in one step from the EBK conditions. Moreover, they can be summed up and shown to contribute only to states with zero angular momentum. One remaining question is whether in semiclassical arguments concerning spectral statistics in chaotic systems it is sufficient to consider only periodic orbits. One argument for this is that the semiclassical contributions of most diffractive orbits (GTD region) are by an order $1/\sqrt{k}$ smaller than those of periodic orbits where $k$ is the wavenumber. For scattering angles near the forward direction they contribute more strongly, up to the order of a periodic orbit, but the corresponding angular regime of this transitional region decreases proportional to $1/\sqrt{k}$. So one might argue that diffractive orbits become less and less important in the semiclassical regime. Let us discuss this point in more detail. By applying the trace formula in order to resolve adjacent energy levels with wave numbers of order $k$ one has to take into account all orbits up to the Heisenberg length $L_H \propto k$. In a chaotic system the number of diffractive orbits increases exponentially with the orbit length. If we assume that the scattering angle is uniformly distributed then the relative number of diffractive orbits in the transitional region decreases like $1/\sqrt{k}$. However, the total number of diffractive orbits in the transitional region is still increasing exponentially due to the exponential increase of all diffractive orbits. A similar argument can be applied to orbits with multiple scattering. It is not obvious that these orbits can be neglected. Moreover one can show that even in the GTD approximation diffractive orbits have a non-vanishing influence on spectral statistics in the semiclassical limit \cite{Sie99}. \bigskip \bigskip \bigskip \noindent {\large \bf Acknowledgements} \bigskip \noindent Financial support by the Deutsche Forschungsgemeinschaft in form of a `Habilitandenstipendium' (SI 380/2-1) is gratefully acknowledged.
\section{Introduction} One of the counterintuitive features of quantum mechanics is nonlocality, whereby operations on a system can instantaneously affect measurements on a distant system with which it had interacted in the past but is now no longer in dynamical interaction with. It was first considered by Einstein, Podolsky and Rosen (EPR), who showed that quantum mechanics is either incomplete or nonlocal \cite{einstein}. They rejected the nonlocal possibility because it implied a superluminal influence in apparent contradiction with Special Relativity. However, today quantum nonlocality is widely accepted as a fact thanks to experiments since the mid-1980's performed both on spin entangled systems \cite{astiw}, based on the Bohm version \cite{bohm} of the EPR thought-experiment, which are shown to violate Bell's inequality \cite{bella} (apart from technical caveats concerning detection loopholes \cite{thompson}), and also on systems entangled in continuous variables (Refs. \cite{ghosh,kwi93,strekalov,zei2000} and references therein), where nonlocality is manifested directly in multi-particle interferences. It is generally accepted that although quantum nonlocality enforces distant superluminal correlations at the level of individual events, the inherent randomness of the correlated processes prevents the transmission of controllable signals \cite{eberhard2}. One sees this by noting that the expectation value of an observable of one of a pair of entangled systems remains unaffected by the interaction of a measuring apparatus with the other \cite{bohi,ghirardi,shi93}. Bussey proved that the probability for outcomes of a measurement on an entangled particle is not affected by a measurement on the other entangled particle \cite{bussey}. Jordan argued that the particles of an entangled pair belongs to different subsystems and hence operations on them should commute \cite{jordan}. On this basis, he concluded that expectation value for the given observable at one of the particles remains unaffected by operations at the other. In the language of Bayesian probability, Garrett pointed out that the {\it a priori} and {\it a posteriori} probabilities of results of measurments on a particle are the same given a measurement on its entangled counterpart \cite{garrett}. Using the density matrix formulation to resolve the EPR `paradox', Cantrell and Scully proved that the reduced density matrix remains the same for both particles in an EPR pair before and after measurement on one of them \cite{cantrell}. On this basis , it is conventionally accepted that that nonlocality and Relativistic causality can ``coexist peacefully" \cite{shi89}. However Peacock and Hepburn have argued that most existing no-signalling proofs are to an extent tautological since they begin by assuming an explicitly local interaction Hamiltonian between the measuring instrument and the entangled system \cite{peacock}. Here we prove by means of a thought experiment that the superluminal nonlocal influence is in fact noncausal. The assumption of instantaneous collapse of the wavefunction throughout configuration space is thus shown to be not of purely epistemological significance as held in the orthodox Copenhagen interpretation but as leading to controllable superluminal physical effects. \section{A Thought Experiment}\label{thought} For the purpose of superluminal communcation, the advantage of entanglement in a spatial variable like momentum over spin is that one can modify the wavefunction by using mirrors, lenses, filters and double (or multiple) slits before making a measurement. Such a modification is crucial to avoid the consequences of the no-signalling theorem, as we later see. Mirrors and lenses can be used to rotate momentum states, while slits can be used for getting the wavefunction to self-interfere-- a feature central to the argument below. \subsection{The experimental set-up} The proposed thought experiment is a modification of the original Einstein-Podolsky-Rosen experiment \cite{einstein}. It involves source S of light consisting of entangled photon-pairs. Two spatially seperated observers, Alice and Bob analyze the light. Alice is equipped with a device, labelled K, to measure momentum or position of the photons in $x$ and $y$ direction. Bob is equipped with a Young's double-slit interferometer, labelled L, with the slits seperated along the $y$-axis. A directional filter assembly is mounted just before the double slit diaphragm, as shown in Figure \ref{bm}, to ensure that only rays (almost) orthogonal to its surface will fall on it. It consists of two convex lenses in tandem such that their foci coincide. Entry of light towards lens 2 is restricted to a small hole in the focal plane surrounding the focus. All other light passing from the lens 1 is reflected backwards by the reflector. A spectral filter behind lens 2 is used to restrict the radiation to a narrow bandwidth about wavelength $\lambda$. The double slit is shielded from stray radiation by an enclosure with a reflecting exterior. Thus, the only radiation entering the experiment is via lens 1. Together the filters and the enclosure pass only a unidirectional almost monochromatic light onto the double slit. The use of reflectors rather than absorbers is to avoid collapsing the wavefunction. \begin{figure} \centerline{\psfig{file=bm.eps}} \caption{Bob's equipment: the tandem lens system acts as a directional filter that permits only horizontal rays to fall on the double slit. The spectral filter restricts the radiation from lens 2 to a narrow bandwidth. The entire system is placed within an enclosure with a reflecting exterior, so that the only radiation entering the experiment is via lens 1.} \label{bm} \end{figure} All measurements and detectors are considered to be ideal. Part of the light (denoted A) from S is analyzed by Alice at time $t_1$ while its EPR counterpart (denoted B) is analyzed by Bob at $t_2 \equiv t_1 + \delta t$, where $\delta t$ includes the time taken by Bob to complete his observation. By prior agreement, at $t_1$ Alice will measure either the transverse position coordinate $y$ and longitudinal momentum componant $p_x$ of the photons in ensemble A, or the momentum componants $p_x, p_y$, while Bob will observe the resulting interference pattern on the screen of L. Just before her measurement, only Alice knows what she will measure. We shall prove that quantum nonlocality implies that at $t_2$ Bob can know what measurement Alice has made. \subsection{Nonlocal communication} At time $t_1$ let the pure state of the entangled EPR pairs be described by the wavefunction: \begin{equation} \label{wavefunc1} \Psi ({\bf x}_a, {\bf x}_b) = \int\int_{-\infty}^{\infty} e^{(2\pi i/\hbar )({\bf x}_a - {\bf x}_b + {\bf x_0})\cdot{\bf p}}dp_xdp_y, \end{equation} where ${\bf x} \equiv (x, y)$, the subscripts $a$ and $b$ refer to the particles in ensembles A and B, respectively, ${\bf x}_0$ is the Alice-Bob spatial seperation, and ${\bf p} \equiv (p_x, p_y)$ is momentum. At time $t_1$, Alice has the choice of making either a transverse position or momentum measurement on the particles in A. The corresponding operators are $y$ and $(\hbar /2\pi i)\nabla$. If Alice measures the momentum, then a photon in A is collapsed to a momentum eigenstate: \begin{equation} \label{eigenmom1} \pi_a({\bf x}_a) = e^{(2\pi i/\hbar){\bf x_a}\cdot{\bf p}} \end{equation} with the eigenvalue of ${\bf p} = (p_x, p_y)$. Simultaneously, the photon B is left in the momentum eigenstate: \begin{equation} \label{eigenmom2} \pi_b({\bf x}_b) = e^{(2\pi i/\hbar)({\bf x}_0 - {\bf x}_b)\cdot{\bf p}}, \end{equation} with eigenvalue $-{\bf p}$. Eq. (\ref{wavefunc1}) can now be written as:- \begin{equation} \Psi ({\bf x}_a, {\bf x}_b) = \int\int_{-\infty}^{\infty}\pi_a({\bf x}_a)\pi_b({\bf x}_b)d^2p. \end{equation} Eq. (\ref{eigenmom2}) corresponds to a monochromatic plane wave moving in the $-{\bf p}$ direction. Since $t_2 > t_1$ and also assuming that Alice's measuring apparatus is arbitrarily large, we can ensure that the A twins of all photons observed by Bob have already been observed by Alice. Now after Alice's measurement B no longer necessarily exists in a pure state. Each particle in it may have collapsed to different momenta in the $xy$ plane. However the directional and spectral filters in front of Bob's double slit as shown in Figure 1 ensure that only horizontal momentum eigenstates within a narrow bandwidth centered on wavelength $\lambda$ are incident upon the double slit. As a result, $all$ of the ``hits" on Bob's screen are coincidences corresponding to a narrow spectral and angular ranges in A measurements. The wavefronts have transverse positional uncertainty larger than the slitwidth $d$ of the double slit. Hence, they pass through both slits to interfere and form a Young's double slit pattern on the screen in the single counts. Beyond the double-slit, the particle is represented by $\Psi(x_b,y_b) = \Psi_1(x_b, y_b) + \Psi_2$, where $\Psi_i$ is the monochromatic wave spreading from slit~$i$. The observed intensity pattern at the screen seen in the single counts is: \begin{subeqnarray} \slabel{p1} I(y) &\propto& |\Psi_1 + \Psi_2|^2, \\ \slabel{p2} &=& {\rm sinc}^2\left(\frac{\pi apy}{\hbar D}\right) \cos^2\left(\frac{\pi dpy}{\hbar D}\right), \end{subeqnarray} which is the Young's double-slit diffraction-interference pattern. Here $a$ is the slit-width, $d$ the slit seperation, and $D$ the distance between the slit plane and the detection screen. We note that non-measurement by Alice also produces the double slit pattern because of the presence of the filters. On the other hand, if Alice measures the transverse position and longitudinal momentum of the photons in ensemble A, these photons are collapsed to eigenfunctions of the form: \begin{equation} \label{eigenpos1} \xi_a({\bf x}_a) = \delta (y_a - y)e^{(2\pi i/\hbar )(x_a p_x)} \end{equation} which represents a monochromatic photon wave in A moving with momentum ${\bf p}^{\prime} = (p_x, 0)$ being localized at $y_a = y$. If $\xi_b({\bf x}_b)$ is the eigenfunction of the counterpart photon in ensemble B, Eq. (\ref{wavefunc1}) is written as:- \begin{equation} \Psi ({\bf x}_a, {\bf x}_b) = \int_{-\infty}^{\infty}\xi_a({\bf x}_a)\xi_b({\bf x}_b)dydp_x, \end{equation} so that a photon in B is left in a state given by the eigenfunction: \begin{equation} \label{eigenpos2} \xi_b({\bf x}_b) = \int^{\infty}_{-\infty} e^{(2\pi i/\hbar )([y - y_b + y_0]p_y + [x_0 - x_b]p_x)}dp_y = \hbar\delta (y - y_b + y_0)e^{(2\pi i/\hbar )(x_0 - x_b)p_x} \end{equation} This represents a monochromatic photon wave in B moving with momentum $-{\bf p}^{\prime}$ being transversely localized at $y_b = y + y_0$. Given that the wave represented by Eq. (\ref{eigenpos2}) moves perpendicular to the double slit plane, it cannot be deselected by Bob's directional filter. Provided $y + y_0$ falls within the area of lens 1 in Figure \ref{bm}, it passes to the spectral filter. If its frequency falls within the allowed narrowband, it is incident upon the double slit diaphragm. Since spectral filters restrict only longitudinal frequency, the localization of the eigenstate (\ref{eigenpos2}) in the transverse direction is not affected by the filter. Being localized in the transverse direction, photon B can pass through at most one slit. Each photon B diffracts through one slit or the other in L and does not self-interfere. Because of the low spectral and angular uncertainty of the incident rays, the resulting single slit pattern is observable in the single counts. Furthermore, according to the chronology of the Alice-Bob measurement and assuming that K is arbitrarily large, all of Bob's hits are coincidences corresponding to a restricted transverse localization of photons in K. This ensures high visibility in the pattern observed by Bob. The wavefunction behind the double slit subsequent to Alice's position measurement is:- $\Psi(x_b,y_b) = \Psi_1(x_b, y_b)$ or $\Psi_2(x_b, y_b)$, where $\Psi_i$ is now the wavefunction of a particle localized at slit~$i$ to spread out therefrom. The intensity pattern on the screen will be determined by the classical addition of probabilities: \begin{subeqnarray} \slabel{x1} I(y) &\propto& |\Psi_1|^2 + |\Psi_2|^2 \\ \slabel{x2} & \approx & {\rm sinc}^2\left(\frac{\pi apy}{\hbar D}\right). \end{subeqnarray} This single slit diffraction pattern in contrast to the diffraction-interference pattern of Eq. (\ref{p2}). We note that the inequality of the right-hand sides of Eqs. (\ref{p1}) and (\ref{x1}) lies at the heart of the difference in Bob's observed patterns. The ensembles A and B are assumed to be sufficiently large that a finite number of photons will be found in L after Bob's measurement. Depending on whether Alice chooses to measure $(p_x, y)_a$ or $(p_x, p_y)_a$, she induces a single-slit diffraction pattern or a Young's double-slit pattern on Bob's screen, respectively. Conversely, based on his observation of the interference pattern, Bob instantaneously can know what quantity Alice has just measured. Thus she is able to transmit a classical binary signal which Bob unambiguously interprets. The effective speed of the signal is ${\bf v}_{\rm eff} = {\bf x}_0/\delta t$, which can be made arbitrarily large by increasing ${\bf x}_0$ and decreasing $\delta t$. Nonlocality allows ${\bf v}_{\rm eff}$ to exceed $c$, since the wavefunction is postulated to collapse instantaneously at all points in space. A possible practical realization of this idealized superluminal quantum telegraph is discussed in Section \ref{sqt}. \section{On circumventing no-signalling proofs} The no-signalling proofs mentioned in the Introduction would prevent Bob from extracting a classical signal if he makes a momentum or position measurement {\it before} processing the B ensemble through the directional and spectral filters, and the double slit. The processing is such that a transverse position state is transformed to a particle spreading out from one slit or the other and does not self-interfere, while the momentum state is transformed to a monochromatic wave spreading from both slits that eventually self-interferes. As a result, the subsequent evolution of the position and momentum states beyond the double slit cannot be represented by the same evolutionary operator. But for this, the indistinguishability of the two mixed states resulting from Alice's position or momentum measurment would be merely carried over to a new eigenbasis representation. Let us consider a simplified situation in which position and momentum are two-valued in the entangled state. The particles A and B can exist in two possible transverse momentum componant eigenstates $|p_j\rangle$ ($j$ = {\rm q, r}). Alternatively the two position eigenstates are $|y_i\rangle$ ($i = 1, 2)$, corresponding to a particle being found at slit $i$. In a momentum eigenstate the particle passes through both slits. The entangled state in which the particles are assumed to exist before Alice's measurement is: \begin{subeqnarray} \Psi_{1,2} &=& \frac{1}{\sqrt{2}}(|p_{\rm q}\rangle|p_{\rm r}\rangle - |p_{\rm r}\rangle|p_{\rm q}\rangle )\\ &=& \frac{1}{\sqrt{2}}(|y_1\rangle|y_2\rangle - |y_2\rangle|y_1\rangle ), \end{subeqnarray} wherein the first eigenket in each product pair refers to an A photon, the second to a B one. The transformation from momentum to position basis is assumed to be given by: \begin{subeqnarray} \label{transform} |y_1\rangle &=& \frac{1}{\sqrt{2}}(|p_{\rm q}\rangle + |p_{\rm r}\rangle ) \\ |y_2\rangle &=& \frac{1}{\sqrt{2}}(|p_{\rm q}\rangle - |p_{\rm r}\rangle ). \end{subeqnarray} The no-signaling argument can be expressed by the fact that the reduced density operator ${\cal D}_p$ for Bob's particles subsequent to Alice making a momentum measurement is indistinguishable from the reduced density operator ${\cal D}_y$ for Bob's particles subsequent to her making a position measurement, as seen from: \begin{subeqnarray} {\cal D}_p &\equiv& \frac{1}{2}(|p_{\rm q}\rangle\langle p_{\rm q}| + |p_{\rm r}\rangle\langle p_{\rm r}|)\\ &=& \frac{1}{2}(|y_1\rangle\langle y_1| + |y_2\rangle\langle y_2|)\\ &\equiv& {\cal D}_y \end{subeqnarray} This shows that Bob cannot know whether Alice made a position or momentum measurement by observing the B photons directly (i.e, before they pass filters and double slit). To simplifiy the post-double-slit scenario, we once again associate a two-valuedness for the wavefunction spreading from the slits. We designate by $|p_{i{\rm q}}\rangle$ and $|p_{i{\rm r}}\rangle$ the two possible momentum eigenstates spreading from slit $i$. The double slit modifies a momentum eigenstate as: \begin{equation} \label{ptransform} |p_j\rangle \longrightarrow |p_j^{\prime}\rangle \equiv \frac{1}{\sqrt{2}}(|p_{1j}\rangle + |p_{2j}\rangle); ~~~j = {\rm q,~r}. \end{equation} This depicts that a momentum eigenstate impinging on the double slit diaphragm becomes two monochromatic wavefronts spreading from the two slits. A position eigenstate is modified to: \begin{equation} \label{xtransform} |y_i\rangle \longrightarrow |y_i^{\prime}\rangle \equiv \frac{1}{\sqrt{2}}(|p_{i{\rm q}}\rangle + |p_{i{\rm r}}\rangle); ~~~i = 1,~2. \end{equation} This depicts that a position eigenstate localized at slit $i$ spreads out from the same slit as a nonmonochromatic wave. Note that the $|p^{\prime}\rangle$'s and $|y^{\prime}\rangle$'s are not mutually orthogonal, and in specific not related according to Eq. (\ref{transform}). To understand this, we must picture what happens to a wavefront impinging on a double slit. It is collapsed to a wavepacket with the momentum distribution such that it peaks at the two slits. As a result, the transformations (\ref{ptransform}) and (\ref{xtransform}) need not be unitary. This situation is forced on us by the fundamental way in which slits affect the wavefunction. For example, if we form $|y^{\prime}_1\rangle \equiv (1/\sqrt{2})(|p^{\prime}_{\rm q}\rangle + |p^{\prime}_{\rm r}\rangle)$ in analogy with Eq. (\ref{transform}), then the ket would contain spherical momentum eigenstates originating from the other slit, which is disallowed by the complementarity principle. If Alice made a momentum measurement, Bob's particles behind the double slit are described by the density operator:- \begin{equation} \label{pdtransform} {\cal D}^{\prime}_p = \frac{1}{2}(|p_{\rm q}^{\prime}\rangle\langle p_{\rm q}^{\prime}| + |p_{\rm r}^{\prime}\rangle\langle p_{\rm r}^{\prime}|) \end{equation} Alternatively, in case of position measurement by Alice:- \begin{equation} \label{xdtransform} {\cal D}^{\prime}_y = \frac{1}{2}(|y_1^{\prime}\rangle\langle y_1^{\prime}| + |y_2^{\prime}\rangle\langle y_2^{\prime}|) \end{equation} Substituting Eqs. (\ref{ptransform}) and (\ref{xtransform}) into Eqs. (\ref{pdtransform}) and (\ref{xdtransform}), we find:- \begin{equation} {\cal D}^{\prime}_p - {\cal D}^{\prime}_y = \frac{1}{2\sqrt{2}}\left[(|p_{1{\rm q}}\rangle - |p_{2{\rm r}}\rangle) (\langle p_{2{\rm q}}| - \langle p_{1{\rm r}}) \right]. \end{equation} The fact that ${\cal D}^{\prime}_p \ne {\cal D}^{\prime}_y$ in general implies that Bob can know from the modified mixed states whether Alice made a position or momentum measurement. \section{A possible realization of the thought experiment} \label{sqt} Recent technological advances permit the performance of experiments involving two-photon wavefunctions entangled in continuous variables like momentum (Refs. \cite{strekalov,zei2000} and references cited therein). Typically, in such experiments a UV beam from a pump entering a nonlinear optical crystal spontaneously fissions into two momentum entangled photons by the process known as spontaneous parametric down-conversion (SPDC). We describe a particular experiment closest to the thought-experiment described above. An `unfolded' diagram of it is given in (Figure \ref{downconvert}). One of the down-converted photons (say A) is observed by Alice scanning the focal plane $y_2$ or imaging plane $y_3$ behind a Heisenberg lens using a detector. The other photon (say B) from the down-conversion pair enters a (say Bob's) double slit assembly, at distance $g$ from the crystal. Registration of a photon A by Alice's detector at the focal plane constistutes a momentum measurement on A. Because of the strong correlation between the two photons, this will project photon B to a momentum in the opposite direction. A diffraction-interference pattern is observed at plane $y_1$ behind the double slit in the coincidences. Registration of a photon A by Alice's detector at the imaging plane constistutes a position measurement on A. This will project photon B to a corresponding position states As this can give path information of photon B, it cannot pass through both slits simultaneously and hence does not self-interfere. No double slit interference pattern is observed in the coincidences. In our variant of this experiment, the distance $g$ is large enough that the crystal subtends a very small angle at the double slit, and wavefronts intercepted by the latter can be approximated by a single mode. Thus, the large distance acts as a directional filter. Alternatively, the tandem lens system described in Section \ref{thought} could also be used. Alice has the choice between making a momentum measurement on photons A by intercepting them at the focal plane, or making a position measurement by intercepting them at the imaging plane. \begin{figure} \centerline{\psfig{file=downconvert.ps}} \caption{`Unfolded' view of a two-photon experiment: A nonlinear crystal creates a pair of entangled optical photons by down conversion of a UV photon from the pump. Photon A is registered by means of a ``Heisenberg" detector system by Alice. Photon B passes into a double slit assembly. If the detector registers a photon at the focal plane $y_2$ of the lens, photon B is projected into a momentum eigenstate. Bob observes a double slit pattern in the coincidences. If the detector registers a photon at the imaging plane $y_3$, photon B is projected into a position eigenstate. It can pass through at most one slit or the other and hence no interference pattern is seen in the coincidences.} \label{downconvert} \end{figure} If Alice registers photon A in the $y_2$ plane, photon B is left with a well defined momentum but the position of origin at the crystal is uncertain (Figure \ref{mom}). The corresponding photon B is instantaneously collapsed to the modes $j$ and $k$ representing a coherent momentum state proceeding from points $J$ and $K$ in the crystal. In practice, the situation is more complicated since the entanglement angle of the down-converted light is usually finite \cite{joo96}, but we can ignore it for our purpose. The state $|\Psi\rangle$ of the SPDC light is assumed to be given by the superposition of the vacuum state and a twin state with a single photon in the two modes $\mu$ and $\nu$ of the signal and idler beams: \begin{equation} \label{spdcfield} |\Psi\rangle = |{\rm vac}\rangle + \epsilon(|s_\mu i_\mu\rangle + |s_\nu i_\nu\rangle ) . \end{equation} where $\epsilon \ll 1$ is proportional to the classical pump field and the crystal's nonlinearity. Here $s_X$ stands for the signal beam on ray $X$ from the source, received by Alice; $i_X$ stands for the idler beam on ray $X$, received by Bob. \begin{figure} \centerline{\psfig{figure=mom.ps}} \caption{Simplified view of Alice's momentum measurement: If Alice detects a photon in the $y_2$ plane she collapses the photon B to a momentum eigenstate, and thus the direction in which the photon B leaves the crystal is well defined. However, the position is uncertain since the photon could have originated at points $J$ or $K$. The momentum eigenstate passes through Bob's double slit to self-interfere and form a double slit interference pattern. Because of Bob's low acceptance angle, the momentum eigenstate intercepted by him can be approximated to a unidirectional wave. Hence the interference pattern will be visible in the single counts.} \label{mom} \end{figure} In Figure \ref{mom}, the field operator $E^{(+)}_{J^{\prime}}$ ($E^{(+)}_{K^{\prime}}$) at the point $J^{\prime}$ ($K^{\prime}$) just in front of the slits plane, and $E^{(+)}_A$, the field at Alice's detector placed in the focal plane of the lens are assumed to be given in terms of annihilation operators: \begin{subeqnarray} \slabel{momfielda} E^{(+)}_{J^{\prime}} &=& \hat{j}_i {\rm exp}(ikr_{JJ^{\prime}}) \\ \slabel{momfieldb} E^{(+)}_{K^{\prime}} &=& \hat{k}_i {\rm exp}(ikr_{KK^{\prime}}) \\ \slabel{momfieldc} E^{(+)}_A &=& \hat{j}_s {\rm exp}(ikr_{Jy_2}) + \hat{k}_s {\rm exp}(ikr_{Ky_2}), \end{subeqnarray} where $\hat{j}, \hat{k}$ are the annihilation operators in the $j$ and $k$ modes, with suffixes $s$ and $i$ used to refer to signal and idler photons. The quantities $r_{XY}$ refer to the distance between points $X$ and $Y$ along the respective mode. The second-order cross-correlation function for the fields at the two points gives the two-photon wavefunction amplitude: \begin{equation} C_{AJ^{\prime}} \equiv \langle E^{(+)}_{J^{\prime}}E^{(+)}_A\rangle = \epsilon ~{\rm exp}(ikr_{J^{\prime}y_2}). \end{equation} The angles $\langle\cdot\cdot\cdot\rangle$ stand for expectation value with the SPDC field given by Eq. (\ref{spdcfield}) with $\mu = j$ and $\nu = k$. Similarly for the $k$ mode we find: \begin{equation} C_{AK^{\prime}} = \epsilon~{\rm exp}(ikr_{K^{\prime}y_2}). \end{equation} Since $r_{J^{\prime}y_2} \approx r_{K^{\prime}y_2}$, therefore $C_{AJ^{\prime}} \approx C_{AK^{\prime}}$. Thus the B photon that is incident upon the double slit is a coherent wavefront seemingly originating from Alice's detector on the $y_2$ plane, and has transverse positional uncertainty. Therefore, the wave passes through both slits and forms a double slit pattern. The twins of not all photons registered by Alice are registered by Bob. If the photon is registered by Bob behind the double slit, it should be a horizontal mode that has passed within his narrow acceptance angle. Because of low angular uncertainty, the photons registered by Bob form a double slit pattern at the $y_1$-plane in the single counts. We note that because the largeness of $g$ acts as a directional filter, the pattern is seen even if Alice does not measure momentum. We assume that the lens subtends approximately the same or larger angle at the double slit than the crystal. Thus, if Alice scans the focal plane, all registrations by Bob will have a coincident registration by Alice (though, as mentioned, the converse is not true), corresponding to a momentum making a small angle with the principal axis of the lens. Furthermore, Alice will find a double slit pattern in the coincidences even though her photons did not pass a double slit \cite{strekalov,zei2000}. If Alice registers a photon in the $y_3$ plane by positioning the detector there, then the twin photon B is instantaneously collapsed to an entangled state of the position eigenstates represented by the modes $j$ and $m$ (Figure \ref{pos}). In this case, the point of origin of the photons at the crystal is definite ($J$), but momentum is uncertain. If Bob registers a coincidence, then the mode $j$ is selected, which corresponds to a definite transverse position on the double slit. Photon B can at most pass through one of the slits and hence forms a plain diffraction pattern. \begin{figure} \centerline{\psfig{file=pos.ps}} \caption{Simplified view of Alice's position measurement: If Alice detects a photon in the $y_3$ plane, she collapses the photon B to modes $j$ and $m$. The point where the photons originated in the crystal is well defined ($J$). However, the momentum is uncertain since the paths of the modes $j$ and $m$ have different directions. If further the photon is detected by Bob, then mode $j$ is selected which corresponds to a well defined transverse position on Bob's slit plane. Hence, the photon will not self-interfere to form a double slit pattern. Because of Bob's low acceptance angle, the direction of these photons is almost horizontal. As a result, a single slit pattern emerges behind Bob's slit in the single counts.} \label{pos} \end{figure} If Alice makes a position measurement as given in Figure \ref{pos}, then the field $E^{(+)}_A$ at the detector placed at a point $y_3$ on the imaging plane, and the fields at $J^{\prime}$ and $K^{\prime}$ are given by: \begin{equation} E^{(+)}_A = \hat{j}_s {\rm exp}(ikr_{j}) + \hat{m}_s {\rm exp}(ikr_{m}) \end{equation} where $r_j$ ($r_{m}$) refers to the distance between point $J$ and point $y_3$ along mode $j$ ($m$). Using this and the detector fields given by Eqs. (\ref{momfielda}) and (\ref{momfieldb}) we find the correlation functions: \begin{eqnarray} \langle E^{(+)}_{J^{\prime}}E^{(+)}_A\rangle &=& \epsilon~{\rm exp} \left(ikr_{J^{\prime}y_2}\right), \nonumber \\ \langle E^{(+)}_{K^{\prime}}E^{(+)}_A\rangle &= &0, \end{eqnarray} where the averaging is done in the state given by Eq. (\ref{spdcfield}) with $\mu = j$ and $\nu = m$. Thus the position measurement by Alice selects a single longitudinal mode and localizes the photon B transversely. Since this represents path information, photon B does not self-interfere but instead forms a single slit diffraction pattern. Because of Bob's low acceptance angle, only almost horizontal modes $j$ are incident upon the double slit. Thus, Bob's time-intergrated registrations will produce the single slit diffraction pattern in the single counts. Because of the angular containment of the source within the lens as seen by Bob, all registrations by Bob will have a coincident registration by Alice (as before, the converse not being true). Thus the time-integrated pattern observed by Bob will have a high visibility. By choosing to measure position or momentum, Alice can nonlocally control the single particle interference pattern seen by Bob at the $y_1$-plane thereby transmitting a classical binary signal. In practice, she can be located midway between the focal and imaging plane of the lens, with a mechanism to send light signals to activate either the detector at the focal plane or that at the imaging plane. By making the number flux of correlated pairs emanating from the crystal very high the time required by Bob to integrate the received pattern can be made sufficiently low. From Figure \ref{downconvert} the effective speed with which Alice transmits the signal is: \begin{equation} v_{\rm eff} \approx \frac{3.5f + g}{f/2c} = \left(7 + \frac{2g}{f}\right)c \end{equation} where the origin of the signal is taken to be Alice's position. \section{Conclusion} We find that nonlocality is intrinsically noncausal. This leads us to doubt that quantum nonlocality as we understand it now can peacefully coexist with Special Relativity. A possible test of the noncausal effect is suggested. \acknowledgements I am grateful to Dr. R. Tumulka, Dr. R. Plaga, Dr. M. Steiner and Dr. J. Finkelstein for criticism of an earlier version of this manuscript. I thank Dr. Caroline Thompson for discussions.
\section{INTRODUCTION} After about two decades of research, the superconductor-insulator (SI) transition in disordered films of metals remains a controversial subject, mainly due to contradictory results in both theoretical and experimental studies. This work aims to improve the understanding of this phenomenon, which might also be relevant for high-T$_{c}$ superconductors and possibly connected to novel metal-insulator transitions in 2D electron systems. The superconductor-insulator transition in ultrathin films of metals is believed to occur at the absolute zero of temperature when the quantum ground state of the system is changed by tuning disorder, film thickness, carrier concentration or magnetic field. Unlike finite temperature phase transitions in which thermal fluctuations are crucial, $T=0$ phase transitions are driven purely by quantum fluctuations. At finite temperatures, an underlying quantum phase transition manifests itself in the scaling behavior of the resistance with the appropriate tuning parameter and the temperature, along with the coherence length and dynamical critical exponents, $\nu $ and $z$ respectively\cite{Sondhi}. Various models of the superconductor-insulator transition in disordered films can be roughly divided in two groups: those in which the superconductivity is destroyed by fluctuations of the amplitude of the order parameter, and those which focus only on the phase fluctuations. If superconductivity is destroyed only by phase fluctuations, then Cooper pairs persist on the insulating side of the transition and the transition may be described by a model of interacting bosons in the presence of disorder. Based on this assumption, Fisher and co-workers \cite{Fisher g} suggested a scaling theory and a phase diagram for a two-dimensional system as a function of temperature, disorder, and magnetic field\cite{Fisher B,Fisher d}. The superconducting phase is considered to be a condensate of Cooper pairs with localized vortices, and the insulating phase is a condensate of vortices with localized Cooper pairs. At the transition, both vortices and Cooper pairs are mobile as they exchange their roles, which leads to a finite resistance. Some important predictions of the model are the universal value of this critical resistance and specific values of the critical exponents $\nu $ and $z$. This so-called ''dirty boson'' problem has been extensively studied using quantum Monte Carlo simulations \cite {Scalettar,Makivic,Cha1,Cha2,Wallin,Kisker,Wagenblast,Lidmar}, real-space renormalization group techniques \cite{Singh,Zhang}, strong-coupling expansions \cite{Freericks} and in other ways \cite {Gold,Ma,Fradkin,Herbut1,Ghosal}. Finite temperature behavior in the vicinity of a quantum critical point was also studied analytically\cite {Sachdev,Herbut2}. A transition from a superfluid to a Mott insulator was found in the pure case, and to a Bose glass insulator in the presence of disorder, but there is still considerable disagreement as to the universality class of the transition, as well as the value of the critical resistance. An alternative picture of interacting electrons \cite {Maekawa,Belitz,Finkelshtein} proposes a different mechanism: the density of states and the Cooper pairing are suppressed on the insulating side of the superconductor-insulator transition due to an enhanced Coulomb interaction. The SI transition occurs as a consequence of fluctuations in the amplitude, rather than the phase of the order parameter. In other words, Cooper pairs break up into single electrons at the transition. Therefore the superconducting gap would also vanish at the transition. The model of interacting electrons has also been studied numerically. Quantum Monte Carlo simulations of an attractive fermion Hubbard model with on-site interactions \cite{Trivedi} yielded a direct superconductor-to-insulator transition in two dimensions without an intervening metallic phase. The critical resistance was found to depend on the strength of the attractive interaction, as a function of which, a crossover from a fermionic to a bosonic regime occurs. The results of this theory qualitatively resemble the experimental data. A recent calculation of the effect of disorder on the gap in the density of states, using a similar model \cite{Huscroft}, showed that the existence of a gap on the insulating side of the transition depends on the coupling strength, allowing for a Fermi insulator at weak and a Bose insulator at strong coupling. Experimentally, the destruction of superconductivity by disorder has been studied in films of MoGe \cite{Graybeal,Yazdani}, InO$_{x}$ \cite {Hebard,Paalanen,Gantmakher1,Gantmakher2,Okuma} and Bi, Pb, Ga, Al \cite {Jaeger,Liu,Wu} among others. Evidence was found of $T_{c}$ going to zero with increasing disorder \cite{Graybeal} implying the destruction of Cooper pairs at the transition. Tunneling experiments also seem to support the fermionic picture. Valles {\it et al}. found that the superconducting gap and the mean field transition temperature are both suppressed as disorder is increased, and that the gap vanishes on the insulating side of the superconductor-insulator transition \cite{Valles}. Hsu {\it et al}. carried out tunneling studies of the superconductor-insulator transition in PbBi/Ge films, and found a large number of quasiparticle states near the Fermi energy \cite{Hsu}. They estimated the average number of Cooper pairs in a coherence volume to be on the order of one at the superconductor-insulator transition. This result, in combination with the disappearance of the energy gap, was interpreted as evidence of the superconductor-insulator transition being driven by fluctuations in the amplitude of the order parameter. Alternatively, it is possible for the superconducting energy gap to be reduced or the tunneling density of states to be broadened as a consequence of phase fluctuations \cite{Ferrel}. Thus, the absence of the gap in these tunneling studies does not necessarily mean that Cooper pairs are absent on the insulating side of the superconductor-insulator transition, but it may imply that a full picture might have to include fermionic degrees of freedom. Evidence of the importance of the bosonic picture can be found in the work of Paalanen{\it \ et al}.\cite{Paalanen}. These workers studied the magnetoresistance and the Hall effect in amorphous InO$_{x}$ films and observed two distinct transitions: one at a critical field $B_{xx}^{c}$ where the longitudinal resistance diverges and the system presumably undergoes a transition from a superconducting phase to a Bose glass insulator with localized Cooper pairs, and the other at a higher field $% B_{xy}^{c}$, where the transverse resistance diverges and the Cooper pairs of the Bose glass insulator presumably unbind. The transition in the transverse resistance occurred at the same magnetic field where the longitudinal resistance showed a maximum. Since a Bose insulator might be expected to have a higher resistance than an insulator with localized single electrons, and from the disorder dependence of $B_{xx}^{c}/B_{xy}^{c}$, this was interpreted as evidence of the bosonic nature of the insulating state close to the superconductor-insulator transition. Similar behavior was observed by other groups \cite{Okuma}. Magnetoresistance studies of amorphous $InO_{x}$ films by Gantmakher et al.\cite{Gantmakher1} also seem to support the bosonic picture. Furthermore, a linear component of the magnetoresistance observed in the insulating regime in amorphous Bi films can be interpreted as a signature of vortex motion \cite{Markovic1}. In the context of the scaling behavior, the thickness tuned transition of ultrathin films of amorphous $Bi$ has been studied in zero magnetic field% \cite{Liu}. A scaling analysis of the magnetic field tuned SI transition has been carried out for thin films of $In0_{x}$ \cite{Hebard} and $MoGe$. \cite {Yazdani} All of these investigations found $\nu \approx 1.3$ and $z\approx 1 $, consistent with the theoretical predictions of the boson Hubbard model. Yet another interpretation of the experimental data has recently been proposed by Shimshoni {\it et al}. \cite{Shimshoni} and expanded upon by Mason and Kapitulnik \cite{Mason}. In this picture, a film contains both insulating and superconducting puddles, and transport is dominated by tunneling or activated hopping between them. The SI transition then occurs as a consequence of the percolation of one phase or the other. Since the correlation length exponent in 2D classical percolation is 4/3, this is consistent with $\nu {\approx }1.3$ observed in most experiments. This model also predicts a saturation of the resistance at very low temperatures, which seems to be supported by the experimental data of Ephron{\it \ et al}. \cite {Ephron}, and Yazdani and Kapitulnik\cite{Yazdani}. Similar effects have been observed in the much earlier work of Wang {\it et al. }\cite{WangT}{\it % \ }on underdoped high-$T_{c}$ (cuprate) films. These ideas may be relevant to similar features of the results of Kravchenko {\it et al}. on two dimensional electron gas systems \cite{Kravchenko}. In all studies in which there is flattening in $R(T)$ at low temperatures, one must be concerned with the possibility of electrical noise being the source of the effect. Also in multi-component materials such as MoGe and underdoped cuprates there is always a possibility of second phases affecting the outcome. Furthermore, it has recently been proposed that the flattening in $R(T)$ at low temperatures may be a signature of Bose metal, a phase in which the Cooper pairs are mobile but do not condense \cite{Das1}. The quantitative results of the study of the magnetic field tuned superconductor-insulator transition presented here for disordered metal systems are in serious disagreement with previous measurements of this transition, adding yet another puzzle to this problem, and calling for a re-examination of existing models. The thickness-tuned transition has also been studied in a nonzero magnetic field. This allows for the construction of a phase diagram and a direct comparison of the two different ways of tuning the SI transition, by varying thickness or magnetic field. This paper is organized as follows: the finite-size scaling procedures used to determine the critical exponents are described in Section II. Experimental details are given in Section III. Section IV focuses on the magnetic field-tuned transition, while the analysis of the thickness-tuned transition in finite magnetic field, which has not been studied before, is presented in Section V. In Section VI, the phase diagram as a function of thickness and magnetic field is presented. The critical resistance and its apparent non-universality are discussed in Section VII. The results and their implications are summarized and further discussed in Section VIII. A brief account of a portion of this work has been previously reported \cite {Markovic2}. \bigskip \section{SCALING\ PROCEDURES} The scale of fluctuations on either side of a quantum phase transition is set by a diverging correlation length $\xi \propto \delta ^{-\nu }$ and a vanishing characteristic frequency $\Omega \propto \xi ^{-z}.$ Here $\delta $ is the deviation from the critical point $\delta =|K-K_{c}|,$ where $K$ is the control or tuning parameter, which drives the system through the transition (i.e. disorder, thickness, magnetic field, etc.), $K_{c}$ is the critical value of $K$ at the transition, $\nu $ is the correlation length exponent and $z$ is the dynamical critical exponent. The exponents $\nu $ and $z$ determine the universality class of the transition. They may not depend on the microscopic details of the physics of the system under study, but on its dimensionality, the symmetry group of its Hamiltonian and the range of interactions. The resistance of a two-dimensional system in the quantum critical regime follows the scaling relation \cite{Sondhi,Fisher B}: \begin{equation} R(\delta ,T)=R_{c}\ f(\delta T^{-1/\nu z}) \label{T scaling} \end{equation} Here $\delta =|d-d_{c}|$ in the case of the thickness-tuned transition and $% \delta =|B-B_{c}|$ in the case of the magnetic field-tuned transition. $% R_{c} $ is the critical resistance at $\delta =0$, and $f(x)$ is a universal scaling function such that $f(0)=1$. The first step in the analysis of the experimental data is to determine the critical value of the tuning parameter and plot the resistance as a function of $\delta $. The $\delta $-axis is then re-scaled by a factor $t$: \begin{equation} R(\delta ,t)=R_{c}\ f(\delta t) \label{rescaled} \end{equation} where the parameter $t(T)$ is determined at each temperature by performing a numerical minimization which yields the best collapse of the data. If the resistance really follows the scaling law (Eq. \ref{T scaling}), it is obvious that $t(T)$ has to be a power law in temperature, $t(T)\equiv T^{-1/\nu z}$. The exponent product $\nu z$ is then found by plotting $t(T)$ as a function of $T$ on a log-log scale, and determining the slope which is then equal to $-1/\nu z$. Similarly, at a constant temperature \cite{Yazdani}: \begin{equation} R(\delta ,E)=R_{c}\ f(\delta E^{-1/\nu (z+1)}) \label{E scaling} \end{equation} where E is the electric field across the sample. This time, the $\delta $% -axis is re-scaled by a field-dependent factor, $t(E)$, which should be a power law in electric field, $t(E)\equiv E^{-1/\nu (z+1)}$, and the exponent $\nu (z+1)$ can then be determined from the field dependence of the parameter $t(E)$. The main advantage of this scaling procedure is that it requires neither prior knowledge of the critical exponents, nor the temperature and thickness dependence of the resistance. The critical exponents are determined empirically from the data, with the critical exponent product as the only adjustable parameter, while the critical value of the tuning parameter is determined independently. The temperature scaling determines the product $% \nu z$, while the electric field scaling determines $\nu (z+1)$. Combining the two results, the correlation length exponent $\nu $ and the dynamical exponent $z$ can be determined separately. An alternative way to determine these critical exponent products is to evaluate a derivative of the resistance with respect to $K$ at its critical value $K_{c}$ \cite{Hebard}: \begin{equation} (\partial R/\partial K)_{K_{c}}\propto R_{c}T^{-1/\nu z}\ f^{^{\prime }}(0) \label{exponent} \end{equation} where $K\equiv d$ at the thickness-tuned transition and $K\equiv B$ at the magnetic field-tuned transition, and $f^{\prime }(0)$ is a constant. Plotting $(\partial R/\partial K)_{K_{c}}$ as a function of $T^{-1}$ on a log-log scale should yield a straight line, with a slope equal to $1/\nu z$. The same method can be applied to the electric field scaling to determine $% 1/\nu (z+1)$, and then $\nu $ and $z$ can be calculated from the results. In the work described below, both scaling procedures were used to obtain the critical exponents, in order to check their consistency. The exponents obtained using two different methods were found to be the same, within the experimental uncertainty. \bigskip \section{EXPERIMENTAL\ METHODS} Ultrathin Bi films were evaporated on top of a $10\AA $\ thick layer of amorphous Ge, which was pre-deposited onto either $SrTiO_{3}$ or glazed alumina substrates. The substrate temperature was kept well below $20K$ during all depositions and all the films were grown{\it \ in situ} under UHV conditions ($\sim $ $10^{-10}$ Torr). The film thickness was gradually increased through successive depositions in increments of $0.1-0.2\AA $. Resistance measurements were carried out between the depositions using a standard DC four-probe technique, with currents up to 50 nA. A detailed temperature dependence of the resistance in zero field and in magnetic field was recorded at each film thickness in the temperature range between 0.14K and 15K, where the lowest temperatures were achieved using a dilution refrigerator. As the film thickness increased from 7\AA\ to 15\AA , the temperature dependence of the resistance of the system changed from insulator-like to superconductor-like at low temperatures, with no sign of reentrant behavior typically observed in granular films \cite{Jaeger}. The films that were superconducting in zero field were driven insulating by applying a magnetic field of up to 12 kG perpendicular to the plane of the sample using a superconducting split-coil magnet. The scaling procedures described above were applied to the magnetic field-tuned transition, as well as to the thickness-tuned transition in both zero field and in a fixed magnetic field. \bigskip \section{MAGNETIC FIELD-TUNED SI TRANSITION} The resistance as a function of temperature for seven films with varying degrees of disorder was studied in magnetic fields up to 12 kG applied perpendicular to the plane of the sample. A typical temperature dependence of the resistance as the magnetic field changes is shown in Fig. \ref{R vs T in B}. In zero field, the resistance decreases with decreasing temperature suggesting the existence of superconducting fluctuations. A magnetic field destroys this downward curvature, and at some critical magnetic field, $% B_{c} $, the resistance is independent of temperature. In magnetic fields higher than $B_{c}$ the film is insulating, with $\partial R/\partial T<0$. Figure \ref{R vs B} shows the resistance as a function of magnetic field for different temperatures. If the sheet resistance is normalized by the value of the critical resistance at each thickness, $R/R_{c}(d),$ then all the data can be collapsed onto a single curve. The collapse of the normalized resistance data as a function of $\delta t$ for five samples is shown in Fig. \ref{B collapse}. The critical exponent product $\nu z$, determined from the temperature dependence of the parameter $t$ (inset of Fig.\ref{B collapse}. ), is found to be $\nu z=0.7\pm 0.2$, apparently independent of the film thickness. The same exponent products were obtained using the alternative method of plotting $(\partial R/\partial B)_{B_{c}}\ $vs. $T^{-1}$ on a log-log plot and determining the slope which is equal to $1/\nu z$, as shown on Fig. \ref{B exponent}. Electric field scaling was also carried out for one of the samples. Unfortunately, there was not enough data available for the insulating side of the transition to carry out a complete analysis, but the data on the superconducting side was sufficient to obtain the value of the critical exponent product $\nu (z+1)$. The magnetic field dependence of the sheet resistance for different values of electric field applied across the sample is shown on Fig. \ref{R vs B in E}. The resistance data were then plotted as a function of $(B-B_{c})$, and re-scaled by a parameter $t(E)$ to obtain the best collapse of the data, shown in Fig. \ref{E collapse}. For the electric field dependence of the parameter $t(E)$, shown in the inset of Fig. \ref{E collapse}., the best power law fit was obtained for $\nu (z+1){\approx }1.4.$% Combining this result with the result of the temperature scaling, it follows that $z\approx 1$ and $\nu {{}\approx 0.7}$ for the magnetic field tuned superconductor-insulator transition. In contrast with our findings, previous studies of thin films of amorphous $% InO_{x}$ \cite{Hebard} and $MoGe$ \cite{Yazdani} both showed $\nu \approx 1.3 $ and $z=1$ for the magnetic field tuned superconductor-insulator transition. Our surprising result is also in obvious disagreement with the prediction of the scaling theory (from which $\nu \geq 1$ \cite {Chayes,Fisher B} for a disordered system), as well as with the percolation-based models \cite{Shimshoni} (from which $\nu \approx 1.3$ would be expected). \bigskip \section{THICKNESS-TUNED SI TRANSITION} For very thin films, the resistance increases exponentially with decreasing temperature, while for the thicker films the resistance goes to zero as the films become superconducting. At the critical thickness $d_{c},$ the resistance is temperature independent, and the system is expected to stay metallic down to $T=0$. Using the same methods described above, the critical exponent product $\nu z$ was determined to be $1.2\pm 0.2$ when the superconductor-insulator transition was tuned by changing the film thickness in zero magnetic field \cite{Markovic2}. A similar scaling behavior has been found in ultrathin films of Bi by Liu et al.\cite{Liu}, with the critical exponent product $\nu z=2.8$ on the insulating side and $\nu z=1.4$ on the superconducting side of the transition. The fact that $\nu z$ was found to be different on the two sides of the transition raises the question of whether the measurements really probed the quantum critical regime. It is likely that the scaling was carried out too deep into the insulating phase, forcing the scaling form (Eq. \ref{T scaling}.) on films which were in a fundamentally different insulating regime. Such films should not be expected to scale together with the superconducting films, hence the discrepancy on the insulating side of the transition. In the present work, the measurements were carried out at lower temperatures than previously studied and with more detail in the range of thicknesses close to the transition. Both sides of the transition scaled with $\nu z{{}\approx }1.2$, which is close to the value obtained by Liu et al. on the superconducting side of the transition. This result is also in very good agreement with the predictions of the scaling theory \cite{Fisher d}, renormalization group calculations \cite{Singh,Zhang,Herbut1}, and Monte Carlo simulations \cite{Cha1,Wallin,Makivic}. All previous experiments which studied the thickness or disorder tuned superconductor-insulator transition were carried out in zero magnetic field. An applied magnetic field is generally expected to change the universality class of the transition since it breaks time reversal symmetry. One would therefore expect the critical exponent product $\nu z$ to be different in the presence of a finite magnetic field. Furthermore, the thickness-tuned transition in a finite magnetic field might be expected to be in the same universality class as the magnetic field-tuned transition at fixed thickness. The thickness-tuned superconductor-insulator transition in a finite magnetic field was probed by sorting the magnetoresistance data which were carefully taken as a function of temperature and magnetic field for each film. A detailed scaling analysis was carried out at fixed magnetic fields of: 0.5 kG, 1 kG, 2 kG, 3 kG, 4.5 kG and 7 kG for one set of films, and 12 kG for a different set of films. For each value of the magnetic field, the resistance was plotted as a function of the film thickness at different temperatures, ranging from 0.14 K to 0.5 K, in order to determine the critical thickness at that field. If the sheet resistance is normalized by the critical value at each field $R/R_{c}(B)$, then the normalized resistance data as a function of the scaling variable for all temperatures and all values of the magnetic field collapsed onto a single curve, as shown on Fig. \ref{d collapse}. The critical exponent product determined from the parameter $t(T)$ (Inset of Fig. \ref{d collapse}) was found to be $\nu z=1.4\pm 0.2$, apparently independent of the magnetic field. Once again, the alternative scaling procedure yielded very similar results, as shown in Fig. \ref{d exponent}. This value of the product $\nu z$ is a factor of two larger than that obtained for the magnetic field-tuned transition. It is, however, very close to that obtained from the analysis of the zero-field transition carried out using data from the same set of films, which was $\nu z=1.2\pm 0.2.$ Given the experimental uncertainties, it is hard to say whether this difference in value of the exponent products reflects a difference between the universality classes of the thickness driven transitions in zero and finite magnetic field. These exponent products are close to those found in Monte Carlo simulations of the (2+1)-dimensional classical XY model with disorder by Cha and Girvin \cite{Cha1}. \bigskip \section{THE\ PHASE\ DIAGRAM} Combining the data obtained from the thickness-tuned transitions in a fixed magnetic field and the field-tuned transitions at the fixed thickness, one can construct a phase diagram with thickness and magnetic field as independent variables. This is shown in Fig.\ref{phase diagram}. The films characterized by parameters which lie above the phase boundary are ''insulating'' ($\partial R/\partial T<0$ at finite temperatures), and the ones below it are ''superconducting'' ($\partial R/\partial T>0$ at finite temperatures). The phase boundary is a power law: \begin{equation} B_{c}\propto \left| d-d_{c}\right| ^{x} \label{phase boundary} \end{equation} The best fit to the data yields $x=0.7$. Near the critical thickness for the zero field transition, a simple dimensionality argument \cite{Fisher B} suggests that the critical magnetic field should scale as: \begin{equation} B_{c}\propto \frac{\Phi _{0}}{\xi ^{2}} \label{critical field} \end{equation} where $\Phi _{0}$ is the flux quantum. Since the correlation length is $\xi \propto \left| d-d_{c}\right| ^{-\nu }$, one might expect the critical field to be: \begin{equation} B_{c}\propto \left| d-d_{c}\right| ^{2\nu } \label{theory phase boundary} \end{equation} According to the phase boundary obtained in this experiment (see Eq. \ref {phase boundary}.), this would mean that $\nu =0.35$, a value not consistent with the results of the scaling analysis carried out on the same films. It also does not agree with $\nu =1.3$ obtained by Refs. \cite{Hebard} and \cite {Yazdani}. There is no obvious physical reason for such a small value of $% \nu $ and implied large values of $z$, so this discrepancy is a mystery at this time. It is possible that the simple argument expressed in Eqs. \ref {critical field} and \ref{theory phase boundary} is too naive. Another surprising feature of the experimental results is that the critical exponent product $\nu z$ evidently depends on whether the phase boundary is crossed vertically (changing the thickness at a constant magnetic field), in which case $\nu z\approx 1.4$, or horizontally (changing the magnetic field at a fixed thickness), in which case $\nu _{B}z_{B}{{}\approx }\ 0.7$. One might expect the critical exponents to not depend on the direction in which the boundary is crossed. If, however, the actual tuning parameter were not film thickness, but some other physical parameter which was a function of thickness, a factor of two in the critical exponent product determined from an analysis using thickness rather than the ''correct'' control parameter might result. The ''correct'' control parameter might be some measure of disorder, electron screening, damping, or Cooper pair density. The detailed functional form of the thickness dependence of these parameters for quench-condensed films is not known. Another possibility is that there are actually two phase boundaries, separating three different regimes, so that each exponent belongs to a different phase boundary. There has been some indication of a vortex liquid phase in between the superconducting (vortex glass) phase and the insulating (Bose glass) phase \cite{Chervenak,vanOtterlo}. Since there only appears to be one phase boundary, that is probably not the case. It is possible, however, that the two boundaries could be indistinguishable over the range of parameters explored in these studies, but would become apparent at higher fields, greater film thicknesses, or lower temperatures. These matters need to be investigated further. \bigskip \section{THE CRITICAL RESISTANCE} The critical resistance for the field-driven transition, contrary to the predictions of the dirty boson models, does not seem to be universal. Figure \ref{Rc vs B} shows that $R_{c}$ decreases as the critical field increases, roughly in a linear fashion. Since thicker films have lower normal state resistances and higher critical fields, this also means that $R_{c}$ decreases with increasing thickness and decreasing normal state resistance. Very similar behavior was observed by Yazdani and Kapitulnik \cite{Yazdani}. In order to explain the non-universal behavior of the critical resistance, these authors proposed a two-channel conduction model, in which the conductance due to the electron (fermion) channel adds to the conductance due to the boson channel. When the unpaired electrons are strongly localized, the conduction is mostly due to bosons, and the resistance is close to $R_{Q}=h/4e^{2},$ as predicted by the boson Hubbard model. In the opposite limit, unpaired electrons contribute significantly to the conduction at the transition. Films with lower normal state resistances would then have lower critical resistances due to the larger fraction of normal electrons. The critical resistances in our experiment, however, are all {\it greater} than $R_{Q}$ and their values could only be explained this way if the quantum resistance due to pairs was itself greater than $R_{Q}$. The conductance due to the electronic channel in a magnetic field might also depend on the strength of the spin-orbit interactions, which is another difference between our samples and those of Refs. \cite{Hebard} and \cite {Yazdani}. The strength of the spin-orbit interactions is typically proportional to $Z^{4}$, where $Z$ is the atomic number. Since $Bi$ is a heavy metal, spin-orbit interactions are stronger than in the lighter $% InO_{x}$ and $MoGe$. It is known that in the weakly localized systems with strong spin-orbit interactions the magnetoresistance is positive, while it is negative otherwise \cite{Bergmann,Lee}. If weakly localized unpaired electrons really contributed significantly to the conduction at the magnetic field-tuned superconductor-insulator transition at the experimentally accessible finite temperatures, the contribution to the magnetoresistance due to localization effects could have a positive or a negative sign, depending on the strength of the spin-orbit interactions. This would make $% R_{c}$ bigger in the case of Bi films, and smaller in the case of $InO_{x}$ and $MoGe$, consistent with experimental observations. There is, however, a striking similarity in the magnetic field and normal state resistance dependence of the critical resistance of the $Bi$ films and $MoGe$ films of Ref. \cite{Yazdani}: even though their critical resistances fall on the opposite sides of $R_{Q}$, they both decrease with magnetic field roughly linearly, with almost the same slope. Strictly speaking, the critical resistance is predicted to be universal only at $T=0$, while the finite temperature corrections are expected to be scaled by the with the Kosterlitz-Thouless transition temperature, $T_{c}$ \cite {Fisher B}: \begin{equation} R_{c}(B_{c},T)=R_{c}^{*}\ +O(\frac{T}{T_{c}})^{2} \label{Rc} \end{equation} where $R_{c}^{*}$ is the universal resistance at $T=0$, and $R_{c}$ is the critical resistance at some finite temperature as measured in the experiments. A closer look at the crossing plots such as that of Fig. \ref{R vs B}, reveals that the critical resistance is indeed slightly temperature dependent. A considerable amount of noise over the accessible temperature range made it hard to compare this temperature dependence with Eq. \ref{Rc}, but qualitative behavior is shown on Fig. \ref{Rc vs T}. Normal state resistances of the $MoGe$ films \cite{Yazdani} are a factor of 3-10 lower than the $Bi$ films considered here, which means that our samples are probing a different part of the phase diagram (normal state resistances are inversely proportional to the film thickness in our experiment), and the finite temperature corrections might be more important in one case then the other. Indeed, somewhat higher critical resistances were found in $InO_{x}$ films if the temperature dependence of $R_{c}$ is taken into account \cite {Gantmakher2}. A recent analytical calculation of the critical resistance of a two dimensional system at finite temperatures in the dirty boson model including Coulomb interactions\cite{Herbut2} yielded a critical resistance of ${% \approx }1.4R_{Q}$. The author suggested that the next order correction would bring the result closer to $R_{Q}$. This result is in excellent agreement with the critical resistance found in the present measurements, which was $1.1-1.2R_{Q}$. Monte Carlo simulations of the $(2+1)$-dimensional XY model without disorder \cite{Cha2} also find the critical resistance to be $R_{c}=7.7k\Omega $, again very close to the value found in this work. \bigskip \section{DISCUSSION} A lot of attention has been focused recently on the effects of dissipation on SI transitions \cite {Shimshoni,Mason,Wagenblast,Kramer,Chakravarty,Rimberg}. Within the picture proposed by Shimshoni {\it et al}.\cite{Shimshoni}, the transition between the superconducting and the insulating state is of a percolative nature. On the insulating side of the transition, electrical transport occurs through activation or tunneling of Cooper pairs between superconducting domains. Likewise, on the superconducting side, vortices tunnel from one insulating domain to another. Using incoherent Boltzmann transport theory, Shimshoni {\it et al}. derive resistivity laws in different temperature regimes and predict finite dissipation at $T=0$ for all values of the magnetic field. Their results seem to be supported by measurements on several different systems: thin films \cite{Ephron,Yazdani}, 2D Josephson junction arrays \cite {vdZant}, Si MOSFETs \cite{Kravchenko} and QH systems \cite{Shahar}, where a saturation of the resistance at low temperatures is observed and attributed to dissipation effects. The percolative nature of the transition can explain the value of $\nu \approx 1.3$ found in most of the field-tuned experiments on thin films \cite{Hebard,Yazdani}, as well as the apparent symmetry between insulating and the conducting phase observed in other experiments \cite{vdZant,Kravchenko,Simonian,Shahar}. In contrast with the above mentioned results, we do not observe any saturation in the temperature dependence of the resistance as the temperature decreases, or in other words, $\delta R/\delta T$ is non-zero down to the lowest temperatures, which were above 0.1K. Of course investigation down to even lower temperatures might lead to a different conclusion. However a satisfactory fit to the resistivity laws predicted by Shimshoni {\it et al.}\cite{Shimshoni} could not be obtained. Mason and Kapitulnik \cite{Mason} recently proposed a new phase diagram for the SI transition which takes into account the possibility of a coupling of the system to a dissipative bath. They argued that this coupling, which becomes important when the critical point is approached, can result in a new, metallic-like phase. In this picture, a direct SI transition is still possible for very weak coupling, while for a stronger coupling the system goes through a metallic phase and is truly superconducting only at the lowest magnetic fields. The fact that the typical sheet resistances of our samples are about a factor of five higher than those in which resistance leveling was observed \cite{Ephron} might just mean that our samples are in the weak coupling regime. However, the correlation length exponent determined in our experiment for the magnetic field-tuned transition, using two different methods, on different physical samples and at several levels of disorder was found to be $\nu \approx 0.7$,which is not consistent with the exponent expected from the classical 2D percolation theory, $\nu =4/3,$ even with much more generous error bars. A coherence length exponent of 0.7 is also inconsistent with what was believed to be an exact theorem, \cite{Chayes} which predicts $\nu \geq 1$ in two dimensions in a presence of disorder. It is interesting to note that our exponent agrees with the result of the classical 3D XY model which is suggested to be relevant in the absence of disorder \cite{Fisher g}. Numerical simulations of a (2+1)-dimensional XY model \cite{Cha2} and the Boson-Hubbard model at $T=0$ \cite{Kisker} without disorder also find $z=1$ and $\nu =0.7.$ However, recently it was suggested that the nature of disorder averaging may introduce a new correlation length, different from the intrinsic one, which might lead to $\nu <1$ even for a disordered system \cite{Pazmandi}. There is also a possibility that the local dissipation coupled to the phase of the superconducting order parameter due to gapless electronic excitations might change the universality class of the system and lead to a non-universal critical resistance \cite{Wagenblast}. The critical resistance would then be expected to increase with increasing damping due to dissipation. The latter would be expected to increase with decreasing normal state resistance. However, we observe that the critical resistance decreases as the normal state resistance decreases, which is exactly the {\it opposite }of the behavior predicted by Wagenblast {\it et al}. We should note that $\nu \approx 0.7$ was also found for the magnetic field-tuned insulator-conductor transition in Si MOSFET samples \cite {Phillips}, suggesting a possible connection between the two phenomena. Our results for the magnetic field-tuned SI transition seem to be consistent with the predictions of bosonic models, rather than percolation models. This is further supported by the transport studies in the insulating regime, where the magnetoresistance cannot be explained by the weak localization theory only \cite{Markovic1}, and the temperature dependence of the resistance fits the predictions of Das and Doniach \cite{Das2} for the bosonic conduction. These observations still need to be reconciled with the results of the tunneling experiments which find no superconducting gap in the insulating regime. The tunneling experiments might however be emphasizing regions of the samples containing quasi-localized single electron states below the gap, or those in which the amplitude fluctuations break the system into superconducting ''islands'' with finite spectral gaps in the density of states, as recently predicted \cite{Ghosal}. A highly non-uniform gap has also been predicted by Herbut \cite{Herbut3} for the case of large disorder. This problem might be clarified using spatially resolved scanning tunneling spectroscopy at low temperatures, which may be able to detect{\it \ local} variations in the density of states. Such studies might also help answer the question as to why $\nu $ different for the thickness- and magnetic field-tuned transitions on the same samples. In the case of the thickness-tuned transition, the correlation length exponent is close to what might be expected from the percolation theory. There is a major difference between the magnetic field-tuned and the thickness-tuned transitions: when the transition is tuned by the magnetic field, the microstructure of the sample stays fixed, while in the case of the thickness-tuned transitions it changes slightly with each film in the sequence. It may be that in this case the percolation effects become relevant, complicating the determination of the critical exponents \cite {Sheshadri}. Finally, the shape of the phase boundary poses a further challenge to theorists. We are currently investigating the role of the dissipation in this system in more detail, using a 2D electron gas as a substrate, similar to the experiment of Rimberg {\it et al}. \cite{Rimberg}. We gratefully acknowledge useful discussions with A. P. Young, S. Sachdev and P. Phillips. This work was supported in part by the National Science Foundation under Grant No. NSF/DMR-9623477.
\section{Introduction} The calculation of thermodynamic properties and excitation spectra of different magnetic materials is one of the most important motivations of the microscopic theory of magnetism. The main approach for such type of investigations is the local spin density functional (LSDF) scheme \cite{lsdf}% . However, this method has some serious shortcomings when applying to transition metal and rare-earth magnetic materials. The main defect is the absence of the ``Hubbard'' type correlations which are most important for real magnets (see, e.g., recent reviews \cite{pwa,vons,dmft}). This leads to, generally speaking, incorrect description of electronic structure for such important groups of magnetic materials as rare-earth metals and their compounds, metal-oxide compounds (including ``classical'' Mott insulators such as NiO and MnO as well as high-$T_c$ cuprates) and even for the iron-group metals \cite{pwa,dmft,lda++}. At the same time, the experience with the Hubbard model shows that the description of electronic structure and magnetic properties of highly correlated materials are closely connected. Recently we propose a rather general scheme (so-called ``LDA++ approach'') for first-principle calculations of the electronic structure with the local correlation effects being included \cite{lda++}. In this technique the full matrix of on-site Coulomb repulsion for the correlated states is taken into account in the local approximation for the electron self- energy. In such a way, we could provide a rather reasonable description of the electronic structure for different correlated systems such as Fe, NiO, and TmSe. It will be very useful to develop this approach for the description of different magnetic characteristics. The most rigorous way to consider properties of magnetic excitations is the calculation of frequency-dependent magnetic susceptibilities\cite{dmft}. However, for many important cases we can restrict ourselves to more simple problem of the calculation of {\it static }response functions. More explicitly, one can consider the variations of total energy $E_{tot}$ (or thermodynamic potential $\Omega $) with respect to magnetic moments rotations. This approach results in the magnetic interactions of different types: the variation of total energy of a ferromagnet over the rotation of all spins at the same angle determines the magnetic anisotropy energy, while the variation of $E_{tot}$ over the relative rotations of spins on two sites gives the parameters of pairwise exchange interactions, etc. This approach was proposed earlier in the the framework of the LSDF scheme \cite{LKG}. It is sufficient for the calculation of ``phenomenological'' exchange parameters which are important for the consideration of domain wall widths and other ``micromagnetic'' properties. In the adiabatic approximation when the energies of magnetic excitations are small in comparison with typical electronic energies this is also sufficient for the calculation of the spin-wave spectrum. In the mean-field approximation these ``exchange parameters'' can be used for the estimation of Curie or Neel temperature. In this work we derive general expressions for the parameters of magnetic interactions in LDA++ approach and calculate the exchange parameters for ferromagnetic iron. It is a first attempt to investigate magnetic interactions, taking into account correlation effects in the electronic structure for real materials. \section{General formalism} \subsection{Local force theorem in LDA++ approach} An important trick for the definition of exchange interactions in the LSDF approach is the use of so called ''local force theorem''. This reduces the calculation of the total energy change to the variations of one-particle density of states \cite{MA,LKG}. First of all, let us prove the analog of local force theorem in the LDA++ approach. In contrast with the standard density functional theory, it deals with the real dynamical quasiparticles defined via Green functions for the correlated electrons rather than with Kohn-Sham ``quasiparticles'' which are, strictly speaking, only auxiliary states to calculate the total energy. Therefore, instead of the working with the thermodynamic potential $\Omega $ as a {\it density} functional we have to start from its general expression in terms of an exact Green function \cite{luttinger} \begin{eqnarray} \Omega &=&\Omega _{sp}-\Omega _{dc} \label{first} \\ \Omega _{sp} &=&-Tr\left\{ \ln \left[ \Sigma -G_0^{-1}\right] \right\} \nonumber \\ \Omega _{dc} &=&Tr\Sigma G-\Phi \nonumber \end{eqnarray} where $G,G_0$ and $\Sigma $ are an exact Green function, its bare value and self-energy, correspondingly; $\Phi $ is the Luttinger generating functional (sum of the all connected skeleton diagrams without free legs), $% Tr=Tr_{\omega iL\sigma }$ is the sum over Matsubara frequencies $Tr_\omega ...=T\sum\limits_\omega ...,$ $\omega =\pi T\left( 2n+1\right) ,$ $n=0,\pm 1,...,$ $T$ is the temperature, and $iL\sigma $ are site numbers ($i$), orbital quantum numbers ($L={l,m}$) and spin projections $\sigma $ , correspondingly. We have to keep in mind also Dyson equation \begin{equation} G^{-1}=G_0^{-1}-\Sigma \label{DYSON} \end{equation} and the variational identity \begin{equation} \delta \Phi =Tr\Sigma \delta G. \label{var} \end{equation} We represent the expression (\ref{first}) as a difference of ''single particle'' ($sp$) and ''double counted'' ($dc$) terms as it is usual in the density functional theory. When neglecting the quasiparticle damping, $% \Omega _{sp}$ will be nothing but the thermodynamic potential of ''free'' fermions but with exact quasiparticle energies. Suppose we change the external potential, for example, by small spin rotations. Then the variation of the thermodynamic potential can be written as \begin{equation} \delta \Omega =\delta ^{*}\Omega _{sp}+\delta _1\Omega _{sp}-\delta \Omega _{dc} \label{var2} \end{equation} where $\delta ^{*}$ is the variation without taking into account the change of the ''self-consistent potential'' (i.e. self energy) and $\delta _1$ is the variation due to this change of $\Sigma $. Taking into account Eq. (\ref {var}) it can be easily shown (cf. \cite{luttinger}) that \begin{equation} \delta _1\Omega _{sp}=\delta \Omega _{dc}=TrG\delta \Sigma \label{var3} \end{equation} and hence \begin{equation} \delta \Omega =\delta ^{*}\Omega _{sp}=-\delta ^{*}Tr\ln \left[ \Sigma -G_0^{-1}\right] \label{var4} \end{equation} which is an analog of the ''local force theorem'' in the density functional theory \cite{LKG}. In the LSDF scheme all the computational results expressed in terms of the retarded Green function and not in the Matsubara one. The relations of ``real'' and ``complex'' Green-function formulae are based on the identity \begin{equation} Tr_\omega F(i\omega )=-\frac 1\pi \int\limits_{-\infty }^\infty dzf(z)ImF(z+i0) \label{id} \end{equation} where $f(z)=\left[ \exp z/T+1\right] ^{-1}$ is the Fermi function, $F(z)$ is a function regular in all the complex plane except real axis. Therefore Eq.(% \ref{var4}) takes the following form \begin{equation} \delta \Omega =\frac 1\pi \int\limits_{-\infty }^\infty dzf(z)ImTr_{iL\sigma }\ln G^{-1}(z+i0) \label{lloyd} \end{equation} which is the starting point for the calculations of magnetic interactions in LSDF approach \cite{LKG}. However note, that in the case of frequency-dependent self-energy (LDA++ approach) it is more convenient to work with the Matsubara Green functions \cite{lda++}. \subsection{Effective spin Hamiltonian} Further considerations are similar to the corresponding ones in LSDF approach. The most suitable way based on the sum rule is proposed in \cite {physica}. In the LDA++ scheme, the self energy is local, i.e. is diagonal in site indices. Let us write the spin-matrix structure of the self energy and Green function in the following form \begin{eqnarray} \Sigma _i &=&\Sigma _i^c+{\bf \Sigma}_i^s{\bf { \sigma }} \label{spin} \\ G_{ij} &=&G_{ij}^c+{\bf G}_{ij}^s{\bf { \sigma }} \nonumber \end{eqnarray} where $\Sigma _i^{\left( c,s\right) }=\frac 12\left( \Sigma _i^{\uparrow }\pm \Sigma _i^{\downarrow }\right)$, ${\bf \Sigma}_i^s=\Sigma _i^s{\bf e}_i, $ with ${\bf e}_i$ being the unit vector in the direction of effective spin-dependent potential on site $i$, ${\bf { \sigma }}=(\sigma_x,\sigma_y,% \sigma_z)$ are Pauli matrices, $G_{ij}^c=\frac 12Tr_\sigma (G_{ij})$ and $% {\bf G}_{ij}^s=\frac 12Tr_\sigma (G_{ij} {\bf {\sigma}})$. We assume that the bare Green function $G^0$ does not depend on spin directions and all the spin-dependent terms including the Hartree-Fock terms are incorporated in the self energy. Spin excitations with low energies are connected with the rotations of vectors ${\bf e}_i$: \begin{equation} \delta {\bf e}_i=\delta {\bf \varphi}_i\times {\bf e}_i \label{rot1} \end{equation} According to the ''local force theorem'' (\ref{var4}) the corresponding variation of the thermodynamic potential can be written as \begin{equation} \delta \Omega =\delta ^{*}\Omega _{sp}={\bf V}_i\delta {\bf \varphi}_i \label{rot2} \end{equation} where the torque is equal to \begin{equation} {\bf V}_i=2Tr_{\omega L}\left[ {\bf \Sigma }_i^s\times {\bf G}_{ii}^s\right] \label{torque} \end{equation} Further we have to use an important sum rule for the Green function which is the consequence of Dyson equation. Using Eq.(\ref{DYSON}) one has \begin{eqnarray} G &=&G^c\cdot \widehat{I}+{\bf G}^s\cdot {\bf {\sigma}}=GG^{-1}G= \label{sum1} \\ &&\ \ G\left( \left( G_0^{-1}-\Sigma ^c\right) \cdot \widehat{I}-{\bf \Sigma}% ^s\cdot {\bf {\sigma}}\right) \left( G^c\cdot \widehat{I}+{\bf G}^s\cdot {\bf {\sigma}}\right) . \nonumber \end{eqnarray} Separating the spin-dependent and the spin-independent parts in Eq. (\ref {sum1}) we have the following sum rules for $G^c$% \begin{equation} G^c=G^c\left( G_0^{-1}-\Sigma ^c\right) G^c-G^c{\bf \Sigma}^s{\bf G}% ^s=G^c\left( G_0^{-1}-\Sigma ^c\right) G^c-{\bf G}^s{\bf \Sigma}^sG^c \label{sum2} \end{equation} and similarly for ${\bf G}^s$% \begin{eqnarray} {\bf G}^s &=&-\left( {\bf G}^s{\bf \Sigma}^s\right) {\bf G}^s+G^c{\bf \Sigma}% ^sG^c+iG^c\left( {\bf \Sigma}^s\times{\bf G}^s\right) = \label{sum3} \\ \ &=&-{\bf G}^s\left( {\bf \Sigma}^s{\bf G}^s\right)+G^c{\bf \Sigma}% ^sG^c+i\left( {\bf G}^s\times{{\bf \Sigma} }^s\right) G^c \nonumber \end{eqnarray} Then for diagonal elements of the Green function one obtains \begin{equation} {\bf G}_{ii}^s=-\sum_j\left[ \left( {\bf G}_{ij}^s{\bf \Sigma }_j^s\right) {\bf G}_{ji}^s-G_{ij}^c{\bf \Sigma }_j^sG_{ji}^c-iG_{ij}^c\left( {\bf \Sigma }_j^s\times {\bf G}_{ji}^s\right) \right] \label{sum4} \end{equation} Substituting (\ref{sum4}) into (\ref{torque}) we have the following expression for the torque \begin{eqnarray} {\bf V}_i &=&2Tr_{\omega L}\left[ {\bf \Sigma }_i^s\times {\bf G}_{ii}\right] \label{torque2} \\ \ &=&-2\sum_jTr_{\omega L}\left\{ {\bf \Sigma }_i^s\times \left( {\bf G}% _{ij}^s{\bf \Sigma }_j^s\right) {\bf G}_{ji}^s-{\bf \Sigma }_i^s\times G_{ji}^c{\bf \Sigma }_j^sG_{ij}^c-i{\bf \Sigma}_i^s\times G_{ji}^c\left( {\bf \Sigma }_j^s\times {\bf G}_{ji}^s\right) \right\} \nonumber \end{eqnarray} If we represent the total thermodynamic potential of spin rotations or the effective Hamiltonian in the form \begin{equation} \Omega _{spin}=-\sum_{ij}Tr_{\omega L}\left\{ \left( {\bf G}_{ij}^s{\bf % \Sigma }_j^s\right) \left( {\bf G}_{ji}^s{\bf \Sigma }_i^s\right) -{\bf % \Sigma }_i^sG_{ij}^c{\bf \Sigma }_j^sG_{ji}^c-i\left( {\bf \Sigma }% _i^s\times G_{ij}^c{\bf \Sigma }_j^s\right) {\bf G}_{ji}^s\right\} \label{Hamilt} \end{equation} one can show by direct calculations that \begin{equation} \left[ \frac{\delta \Omega _{spin}}{\delta {\bf \varphi }_i}\right] _{G=const}={\bf V}_i \label{torque3} \end{equation} This means that $\Omega _{spin}\left\{ {\bf e}_i\right\} $ is the effective spin Hamiltonian. The last term in Eq.(\ref{Hamilt}) is nothing but Dzialoshinskii- Moriya interaction term. It is non-zero only in relativistic case where ${\bf \Sigma }_j^s$ and ${\bf G}_{ji}^s$ can be, generally speaking, ``non-parallel'' and $G_{ij}\neq G_{ji}$ for the crystals without inversion center. In the following we will not consider this term. \subsection{Exchange interactions} In the nonrelativistic case one can rewrite the spin Hamiltonian for small spin deviations near collinear magnetic structures in the following form \begin{equation} \Omega _{spin}=-\sum_{ij}J_{ij}{\bf e}_i{\bf e}_j \label{heis} \end{equation} where \begin{equation} J_{ij}=-Tr_{\omega L}\left( \Sigma _i^sG_{ij}^{\uparrow }\Sigma _j^sG_{ji}^{\downarrow }\right) \label{Jij} \end{equation} are the effective exchange parameters. This formula generalize the LSDA expressions of \cite{LKG} to the case of correlated systems. The sum rule (Eq. \ref{sum4}) for the collinear magnetic configuration takes the following form \begin{equation} G_{ii}^{\uparrow }-G_{ii}^{\downarrow }=2\sum\limits_jG_{ij}^{\uparrow }\Sigma _j^sG_{ji}^{\downarrow } \label{sum5} \end{equation} Using Eq. (\ref{sum5}) we obtain the following expression for the total exchange interaction of a given site with the all magnetic environment. \begin{equation} J_i=\sum\limits_{j(\neq i)}J_{ij}=Tr_{\omega L}\left[ \Sigma _i^sG_{ii}^{\uparrow }\Sigma _i^sG_{ii}^{\downarrow }-\frac 12\Sigma _i^s\left( G_{ii}^{\uparrow }-G_{ii}^{\downarrow }\right) \right] \label{Jo} \end{equation} Spin wave spectrum in ferromagnets can be considered both directly from the exchange parameters or by the consideration of the energy of corresponding spiral structure (cf \cite{LKG}). In nonrelativistic case when the anisotropy is absent one has \begin{equation} \omega _{{\bf q}}=\frac 4M\sum\limits_jJ_{0j}\left( 1-\cos {\bf qR}_j\right) \equiv \frac 4M[J(0)-J({\bf q})] \label{om} \end{equation} where $M$ is the magnetic moment (in Bohr magnetons) per magnetic ion. Corresponding expressions can be easily written in {\bf k-}space also. In the short notation $q=({\bf q},i\omega )$ it is easy to write the general expression for $J({{\bf q)}\equiv }J({\bf q},0)$: \begin{equation} J(q)=-\frac 1{N_k}\sum_kTr_L[\Sigma ^s(k)G^{\uparrow }(k)\Sigma ^s(k+q)G^{\downarrow }(k+q)] \label{Jq} \end{equation} where $N_k$ is the total number of k-points. It should be noted that the expression for spin stiffness tensor $D_{\alpha \beta }$ defined by the relation \begin{equation} \omega _{{\bf q}}=D_{\alpha \beta }q_\alpha q_\beta \label{D} \end{equation} (${\bf q\rightarrow 0}$) in terms of exchange parameters has to be exact since it is the consequence of phenomenological Landau- Lifshitz equations which are definitely correct in the long-wavelength limit (cf \cite{LKG}). One has from Eqs. (\ref{Jij},\ref{om}) \begin{equation} D_{\alpha \beta }=\frac 2M\sum\limits_jJ_{0j}R_j^\alpha R_j^\beta =-\frac 2M% Tr_{\omega L}\sum\limits_{{\bf k}}\left( \Sigma ^s\frac{\partial G^{\uparrow }\left( {\bf k}\right) }{\partial k\alpha }\Sigma ^s\frac{\partial G^{\downarrow }\left( {\bf k}\right) }{\partial k_\beta }\right) \label{D2} \end{equation} where ${\bf k}$ is the quasimomentum and the summation is over the Brillouin zone. In cubic crystals $D_{\alpha \beta }=D\delta _{\alpha \beta }$. For arbitrary ${\bf q}$, the expression of magnon spectrum in terms of $J_{ij}$ is valid only in the adiabatic approximation, i.e. provided that the magnon frequencies are small in comparison with characteristic electronic energies. Otherwise, collective magnetic excitations which are magnons cannot be separated from non-coherent particle-hole excitations (Stoner continuum) \cite{moriya} and magnon frequencies (\ref{om}) are not the exact poles of transverse magnetic susceptibility (which are even not real at large ${\bf q} $). Now we have to consider the accuracy of expressions for $J_{ij}$ (\ref{Jij}) themselves. Eqs. (\ref{torque}) and (\ref{torque2}) are exact in LDA++ approach (i.e. with the only assumption about the {\it local} self-energy). Hence, if one postulate the existence of effective spin Hamiltonian in the sense of Eq.(\ref{torque3}), Eq. (\ref{Jij}) is also exact. However, they do not have rigorous connection with the static transverse spin susceptibility. The latter is expressed in terms of the matrix \[ \frac{\delta ^2\Omega _{}}{\delta \varphi _i^\alpha \delta \varphi _j^\beta }% =\frac{\delta V_i^\alpha }{\delta \varphi _j^\beta } \] {\it without }the restriction $G=const.$ They differ from our exchange parameters by the terms containing $\frac{\delta \widehat{G}}{\delta {\bf % \varphi }_i}.$ From the diagrammatic point of view in the framework of DMFT \cite{dmft} they are nothing but the vertex corrections. We do not present the corresponding expressions since the benefits of the introducing of exchange parameters beyond adiabatic approximation which is equivalent to (% \ref{torque3}) is doubtful. In more rigorous consideration it is convenient to work directly with DMFT expressions for static and dynamic susceptibility \cite{dmft}. Thus one can see that, generally speaking,the exchange parameters differ from the exact response characteristics defined via static susceptibility since the latter contains vertex corrections. At the same time, our derivation of exchange parameters seems to be rigorous in the adiabatic approximation for spin dynamics when spin fluctuation frequency is much smaller than characteristic electron energy. The situation is similar to the case of electron-phonon interactions where according to the Migdal theorem vertex corrections are small in adiabatic parameter (ratio of characteristic phonon energy to electron one)\cite{migdal}. The derivation of the exchange parameters from the variations of thermodynamic potential, being approximate, can be useful nevertheless for the fast and accurate calculations of different magnetic systems. Note that in the LDA++ approach, as well as in LDA+U method and in contrast with usual LSDF, one can rotate separately spins of states with given orbital quantum numbers $L,L^{\prime }$ . For example, for non-relativistic case one can obtain \[ \Omega _{spin}=-\sum_{iL,jL^{\prime }}J_{iL,jL^{\prime }}{\bf e}_{iL}{\bf e}% _{jL^{\prime }} \] where \[ J_{iL,jL^{\prime }}=-Tr_\omega \left( \Sigma _{iL}^sG_{iL,jL^{\prime }}^{\uparrow }\Sigma _{jL^{\prime }}^sG_{jL^{\prime },iL}^{\downarrow }\right) \] are orbital dependent exchange parameters. \subsection{Magnetic anisotropy} Let us consider now the change of spin energy at the rotation of all the spins at the same angle. It is definitely zero in nonrelativistic case. In the presence of spin-orbit coupling, it is nothing but the energy of magnetic anisotropy. One can obtain from Eq.(\ref{Hamilt}) \begin{eqnarray} \Omega _{anis} &=&Tr_{\omega L}\sum\limits_{ij}\left\{ \left( {\bf G}% _{ij}^s\times {\bf \Sigma }_j^s\right) \left( {\bf G}_{ji}^s\times {\bf % \Sigma }_i^s\right) \right\} = \label{anis} \\ &&\ \ Tr_{\omega L}\sum\limits_{{\bf k}}\left\{ \left( {\bf G}^s\left( {\bf k% }\right) \times {\bf \Sigma }^s\right) \left( {\bf G}^s\left( {\bf -k}% \right) \times {\bf \Sigma }^s\right) \right\} \nonumber \end{eqnarray} where the last equality is valid for ferromagnets with one magnetic atom per unit cell, {\bf k} is the quasimomentum and the summation is over Brillouin zone. Finally, note that we use essentially three properties of LDA++ approach: (i) locality of self-energy (ii) spin-independence of bare Green function (i.e. spin-independence of bare LDA spectrum; all magnetic effects including Hartree-Fock ones are included in self- energy) and (iii) approximations for self energy have to be conserving, or ``$\Phi $- derivable'' since only in that case analog of ``local force theorem'' (\ref{var4}) takes place. \section{Exchange interactions in ferromagnetic iron} \subsection{Computational technique} As an example we calculate the magnetic properties of ferromagnetic iron using the most accurate method to take into account local correlations. For this purpose we use the local quantum Monte-Carlo approach\cite{dmft} with the generalization to the multiband case\cite{MQMC}. We start from LDA+U Hamiltonian in the diagonal density approximation: \begin{eqnarray} H &=&\sum_{\{im\sigma \}}t_{im,i^{\prime }m^{\prime }}^{LDA}c_{im\sigma }^{+}c_{i^{\prime }m^{\prime }\sigma }+ \label{calc1} \\ &&\frac 12\sum_{imm^{\prime }\sigma }U_{mm^{\prime }}^in_{im\sigma }n_{im^{\prime }-\sigma }+ \nonumber \\ &&\frac 12\sum_{im\neq m^{\prime }\sigma }(U_{mm^{\prime }}^i-J_{mm^{\prime }}^i)n_{im\sigma }n_{im^{\prime }\sigma } \nonumber \end{eqnarray} where $t^{LDA}$ is effective single-particle Hamiltonian obtained from the non-magnetic LDA with the corrections for double counting of the average interactions among d-electrons; $i$ is the site index and $m$ is the orbital quantum numbers; $\sigma =\uparrow ,\downarrow $ is the spin projection; $% c^{+},c$ are the Fermi creation and annihilation operators ($n=c^{+}c$). The screened Coulomb and exchange vertex for the d-electrons{\ \begin{eqnarray} U_{mm^{\prime }} &=&<mm^{\prime }|V_{scr}^{ee}({\bf r-r}^{\prime })|mm^{\prime }> \label{calc2} \\ J_{mm^{\prime }} &=&<mm^{\prime }|V_{scr}^{ee}({\bf r-r}^{\prime })|m^{\prime }m> \nonumber \end{eqnarray} are expressed via the effective Slater integrals and corresponds to the average }$U=2.3$ eV and $J=0.9$ eV\cite{lda++}. We use the minimal $spd$% -basis in the LMTO-TB formalism and numerical orthogonalization for $t^{LDA}(% {\bf k})$ matrix \cite{lda++}. The Matsubara frequencies summation in our calculations corresponds to the temperature of about T=850 K. Local Green-function matrix has the following form \begin{equation} G(i\omega )=\sum_{{\bf k}}\{i\omega +\mu -t^{LDA}({\bf k})-\Sigma (i\omega )\}^{-1} \label{Gk} \end{equation} Note that due to cubic crystal symmetry of ferromagnetic bcc-iron the local Green function is diagonal both in the orbital and the spin indices and the bath Green function is defined as \begin{equation} G_m^0(i\omega )=G_m(i\omega )+\Sigma _m(i\omega ) \label{Sig} \end{equation} The local Green functions for the imaginary time interval $\left[ 0,\beta \right] $ with the mesh $\tau _l=l\Delta \tau $, $l=0,...,L-1,$ and $\Delta \tau =\beta /L$, where $\beta =\frac 1T$ is calculated in the path-integral formalism\cite{dmft,MQMC}:{\ \begin{equation} G_m^{ll^{\prime }}=\frac 1Z\sum_{s_{mm^{\prime }}^l}\det [O(s)]*G_m^{ll^{\prime }}(s) \label{QMC} \end{equation} here we redefined for simplicity $m\equiv \{m,\sigma \},Z$ is the partition function and the so-called fermion-determinant }$\det [O(s)]$ and the Green function for arbitrary set of the auxiliary fields{\ }$G(s)=O^{-1}(s)$ {are obtained via the Dyson equation\cite{hirsch} for imaginary-time matrix} $(% {\bf G}_m(s)\equiv G_m^{ll^{\prime }}(s))$: \[ {\bf G}_m=[{\bf 1}-({\bf G}_m^0-{\bf 1})(e^{V_m}-{\bf 1})]^{-1}{\bf G}_m^0 \] {where the effective fluctuation potential from the Ising fields }$% s_{mm^{\prime }}^l=\pm 1$ is {\ \begin{eqnarray*} V_m^l &=&\sum_{m^{\prime }(\neq m)}\lambda _{mm^{\prime }}s_{mm^{\prime }}^l\sigma _{mm^{\prime }} \\ \sigma _{mm^{\prime }} &=&\{ \begin{array}{c} 1,m<m^{\prime } \\ -1,m>m^{\prime } \end{array} \end{eqnarray*} and the discrete Hubbard-Stratonovich parameters are $\lambda _{mm^{\prime }}={\rm arccosh}[\exp (\frac 12\Delta \tau U_{mm^{\prime }})] $ \cite{hirsch}% . The main problem of the multiband QMC formalism is the large number of the auxiliary fields }$s_{mm^{\prime }}^l.$ For each time slices $l $ it is equals to $M(2M-1)$ where$M$ is the total number of the orbitals which is equal to 45 for d-states. We compute the sum over this auxiliary fields in Eq.\ref{QMC} using important sampling QMC algorithm and performed a dozen of self-consistent iterations over the self-energy Eqs.(\ref{Gk},\ref{Sig},\ref {QMC}). The number of QMC sweeps was of the order of 10$^5$ on the CRAY-T3e. The final $G_m(\tau )$ has very little statistical noise. We use maximum entropy method\cite{MEM} for analytical continuations of the QMC Green functions to the real axis. Comparison of the total density of states with the results of LSDA calculations (Fig.\ref{DOS}) shows a reasonable agreement for single-particle properties of not ``highly correlated'' ferromagnetic iron. Using the self-consistent values for $\Sigma _m(i\omega )$ we calculate the exchange interactions (Eq.\ref{Jq}) and spin-wave spectrum (Eq.\ref{om}) using the four-dimensional fast Fourier transform (FFT) method\cite{FFT} for $({\bf k},i\omega )$ space with the mesh $20^3\times 320$. We compare the results for the exchange interactions with corresponding calculations for the LSDA method\cite{LKG}. \subsection{Computational results} The spin-wave spectrum for ferromagnetic iron is presented in Fig.\ref{SW} in comparison with the results of LSDA-exchange calculations and with different experimental data\cite{omexp1,omexp2,omexp3} . Note that for high-energy spin-waves the experimental data\cite{omexp3} has large error-bars due to Stoner damping (we show one experimental point with the uncertainties in the ``$q$'' space). On the other hand, the expression of magnon frequency in terms of exchange parameters itself becomes problematic in that region due to breakdown of adiabatic approximation, as it is discussed above. Therefore we do not believe that good agreement of LSDA data with the experimental ones for large wave vectors is too important since it may be a result of ''mutual cancellation'' of inaccuracies of adiabatic approximation and LSDA itself. At the same time, our lower-energy spin-waves spectrum (where the theory is reliable) agree better with the experiments then the result of LSDA calculations. Our LSDA spin-wave spectrum agree well with the results of frozen magnon calculations \cite{Sandratskii,Halilov}. Note that in the LSDA scheme one could use the linear-response formalism\cite{Savr} to calculate the spin-wave spectrum with the Stoner renormalizations, which should gives in principle the same spin-wave stiffness as our LSDA calculations. The corresponding exchange parameters and spin-waves stiffness (Eq.\ref{D}) are presented in the Table I. The general trend in the distance dependence of exchange interactions in ferromagnetic iron is similar in both schemes, but relative strength of various interactions is quite different. Experimental value of spin-wave stiffness D=280 meV/A$^2$\cite{omexp2} agrees well with the theoretical LDA++ estimations. \section{Conclusions} In conclusion, we present a general method for the investigation of magnetic interactions in the correlated electron systems. This scheme is not based on the perturbation theory in ``$U$'' and could be applied for rare-earth systems where both the effect of the band structure and the multiplet effects are crucial for a rich magnetic phase diagram. Our general expressions are valid in relativistic case and can be used for the calculation of both exchange and Dzialoshinskii- Moriya interactions, and magnetic anisotropy. An illustrative example of ferromagnetic iron shows that the correlation effects in exchange interactions may be noticeable even in such moderately correlated systems. For rare-earth metals and their compounds, colossal magnetoresistance materials or high-$T_c$ systems, this effect may be much more important. For example, the careful investigations of exchange interactions in $MnO$ within the LSDA, LDA+U and optimized potential methods for $MnO$ \cite{Sol} show the disagreement with experimental spin-wave spectrum (even for small ${\bf q}$) , and indicate a possible role of correlation effects. As for the formalism itself, this work demonstrates an essential difference between spin density functional approach and LDA++ method. The latter deals with the thermodynamic potential as a functional of Green function rather than electron density. Nevertheless, there is a deep formal correspondence between two techniques (self-energy corresponds to the exchange- correlation potential, etc). In particular, an analog of local force theorem can be proved for LDA++ approach. It may be useful not only for the calculation of magnetic interactions but also for elastic stresses, in particular, pressure, and other physical properties. \section{Acknowledgments} The calculations were performed on Cray T3E computers in the Forschungszentrum J\"ulich with grants of CPU time from the Forschungszentrum and John von Neumann Institute for Computing (NIC). This work was partially supported by Russian Basic Research Foundation, grant 98-02-16279
\section{Introduction} Most of our knowledge of the $YN$ interaction has been obtained from heavy hypernuclei and it remains rather qualitative. In contrast, recent theoretical analyses of few-baryon systems with strangeness such as the three-body calculations of $^3_\Lambda H$\cite{hyp1,hyp2,hypbnl} and the four-body study of $^4_\Lambda H$\cite{hiyama} yield more qualitative results, being based on modern baryon-baryon forces and on rigorous solutions of the few-body Schr\"odinger equations. Thus, they offer a great advantage to scrutinize $YN$ interaction models. As an example, Refs.~\cite{hyp1,hyp2,hypbnl} demonstrated that the Nijmegen soft core $YN$ interaction NSC89 \cite{nsc89} binds the hypertriton, but the J\"ulich $\tilde A$ potential\cite{juelich} does not. This is possibly caused by the sizable difference between the $^1S_0$ force components in the low-energy region\cite{hyp2}. This shows that even basic quantities such as scattering lengths are still undetermined. The few-body analysis of $^3_\Lambda H$ \cite{hyp1,hyp2,hypbnl} has also clarified the effect of the $\Lambda -\Sigma$ conversion which was exactly included in the coupled channel formalism\cite{hyp1}. Although in the case of the hypertriton the admixture of $\Sigma NN$ states is only 0.5\%, the expectation values of the sum of the transition potentials $V_{\Lambda N, \Sigma N}$ and $V_{\Sigma N, \Lambda N}$ are approximately 8\% of the total potential energy, which is crucial for the binding of the hypertriton \cite{hyp2,hypbnl}. This knowledge about the $\Sigma$ states coupling, however, has been obtained from the bound state lying below the $\Lambda NN$ threshold. This should be extended to analyses close to the $\Sigma$ threshold, where $\Lambda -\Sigma$ conversion effects will emerge sharply. In this region, so-called unstable bound states\cite{kok} have received attention and have been searched for experimentally. Only one state in the $\Sigma (\Lambda) NNN$ system is confirmed in the reaction $^4He(K^-, \pi^-)$\cite{nagae}. The existence of such an unstable bound state in the A=4 system was predicted by Harada {\it et~al.} where the coupling to continuum $\Lambda NNN$ states was approximated by a $\Sigma N$ optical potential\cite{harada}. However, for understanding the actual features of $\Lambda -\Sigma$ conversion, it is highly desirable to treat it directly using a realistic $\Lambda N-\Sigma N$ coupling interaction. At present, it is technically possible to incorporate precisely the coupling to $\Lambda$ continuum states for only the $\Sigma N$ and $\Sigma NN$ systems. Afnan and Gibson\cite{afnan} calculated the $\Lambda d$ elastic scattering fully incorporating this coupling but using simple phenomenological $YN$ interactions. They then found and analyzed enhancements just below the $\Sigma NN$ threshold. To examine more closely the $\Lambda N -\Sigma N$ coupling interaction, a similar study applying more sophisticated meson-theoretical interactions is necessary. It is also important to analyze electromagnetic hyperon-production processes \cite{tjlab}, which are experimentally accessible. For the $YN$ interaction, there exists a variety of strengths for the $\Lambda N-\Sigma N$ coupling among extensively used meson-theoretical potentials. The soft core model\cite{nsc89} and the hard core model D\cite{nd} of the Nijmegen group show cusps in the $\Lambda N$ elastic total cross section just at the $\Sigma N$ threshold with different magnitudes, while the Nijmegen hard core model F\cite{nf} and the J\"ulich models\cite{juelich} show round resonance peaks below the threshold. We stress here that these prominent cusps do not mean simple threshold effects, but suggest the existence of $t$-matrix poles in unphysical Riemann sheets. This is important because the poles might move and become unstable bound state poles if the coupling strengths varied. Some examples for separable potentials are given in Ref.~\cite{pearce}. In this paper, we shall locate these poles and follow their trajectories for the Nijmegen potentials. Knowledge of $YN$ $t$-matrix around the $\Sigma N$ threshold is crucial for the analysis of $\Lambda NN-\Sigma NN$ continuum states using meson-theoretical interactions. This paper accordingly investigates poles of the $t$-matrices for the Nijmegen F, D and the two soft core interactions\cite{nsc89,nsc97}. This is achieved in momentum space and hence the results are directly applicable to the three-body calculation. The behavior of $t$-matrices around an inelastic threshold in coupled channel problems and the effects of nearby poles have often been studied~ \cite{kok,pearce}. In such analyses it is important to understand the connection between various Riemann energy sheets and how far from the physical region the poles are located. In this analysis we adopt a so-called uniformization given by Newton\cite{newton}, by which the $t$-matrix for two-channel problems becomes single-valued after a suitable variable is introduced in place of energy. We thereby clearly describe the positions and the trajectories of the $t$-matrix poles in the Riemann sheets. Section II gives the expression for $\Lambda N-\Sigma N$ $t$-matrix which is analytically continued to the complex energy plane. Section III describes a method to treat a hard core potential in momentum space. This is for the purpose of treating the Nijmegen hard core potentials which are influential in hypernuclear physics. In Sec.~IV, the uniformization mentioned above is introduced, thereby we discuss how the shape of the $\Lambda N$ elastic total cross section around the $\Sigma N$ threshold is related to the positions of nearby poles. In Sec.~V, the positions of the $t$-matrix poles for the Nijmegen soft and hard core models are described. We also show the trajectories of the poles when the strengths of the potentials are increased. \section{Analytic continuation of the $t$-matrix} In this section, we give the expression of the off-shell $t$-matrix for the $\Lambda N-\Sigma N$ system and continue it analytically into the complex energy plane. The coupled $t$-matrices for the $\Lambda N-\Sigma N$ system are defined by the integral equations \begin{equation} T_{ij}(z)=V_{ij}+\sum_k V_{ik}\, G_0^{(k)} (z)\, T_{kj}(z), \hspace{10mm} i,j,k=1,2 \label{eq:2.1} \end{equation} with \begin{equation} G_0^{(k)} (z) = \left( z-H_0^{(k)} \right)^{-1}, \hspace{10mm} z=E+i\varepsilon \label{eq:2.2} \end{equation} where $i$, $j$, $k$ have integer values of 1 and 2 for the $\Lambda N$ and $\Sigma N$ channels, respectively. The free Hamiltonian $H_0^{(k)}$ for channel $k$ is defined as \begin{equation} H_0^{(k)}=\frac{p^2_k}{2\mu_k}+m_N+m_Y^{(k)} \, . \label{eq:2.3} \end{equation} This refers to the total momentum zero frame and we denote the relative momentum between the nucleon and the hyperon by ${\bf p_k}$ and the reduced mass of channel $k$ by $\mu_k$. The masses $m_Y^{(k)}$$(k=1,2)$ indicate $m_\Lambda$ and $m_\Sigma$ respectively. After performing the partial-wave decomposition in momentum space, we express the projected $t$-matrix elements for a given total angular momentum and parity again by $T$. Then Eq.~(\ref{eq:2.1}) reads \begin{eqnarray} < p \,|\, T_{ij}(z) \,|\, p^\prime >&=&< p \,|\, V_{ij} \,|\, p^\prime > \nonumber \\ &+&\sum_k \int_0^\infty dp^{\prime\prime} {p^{\prime\prime}}^2 < p \,|\, V_{ik} \,|\, p^{\prime\prime} > \frac{1}{e_k - \frac{{p^{\prime\prime}}^2}{2\mu_k} + i\varepsilon } < p^{\prime\prime} |\, T_{kj}(z) |\, p^{\prime} > \label{eq:2.4} \end{eqnarray} with \begin{equation} e_k\equiv \frac{q_k^2}{2\mu_k}= E-m_N-m_Y^{(k)} \,\, . \label{eq:2.5} \end{equation} To simplify the notation, the $p$ indices have been omitted, and the partial-wave elements are assumed to have no coupling between different orbital angular momenta or channel-spin states. The extension to the case with couplings is straightforward. Now consider the energy $E$ to be a complex number. Hence $e_k$ and \begin{equation} q_k =\sqrt{2\mu_k e_k} \label{eq:2.6} \end{equation} are complex numbers. We introduce the function \begin{equation} h_k(p^{\prime\prime})\equiv < p \,|\, V_{ik} \,|\, p^{\prime \prime} > < p^{\prime\prime} \,|\, T_{kj}(z) \,|\, p^{\prime} > \label{eq:2.7} \end{equation} and define each term in the $k-$summation of the right-hand side of Eq.~(\ref{eq:2.4}) by $I_k(e_k)$ as \begin{equation} I_k(e_k)= \int^\infty_0 dp^{\prime\prime}\,\, \frac{{p^{\prime\prime}}^2 h_k(p^{\prime\prime})} {e_k -\frac{{p^{\prime\prime}}^2}{2\mu_k}} \,\,\, . \label{eq:2.8} \end{equation} This function has a cut for $e_k \geq 0$, in other words, a cut along $m_N+m_Y^{(k)}\leq E <\infty$. Thus, there are two cuts in the $E$ plane starting at the $N+\Lambda$ and $N+\Sigma$ thresholds, respectively. The function values beyond the cuts are defined by analytic continuation. This is achieved by modifying Eq.~(\ref{eq:2.8}) as \begin{equation} I_k(e_k)= \int^{\infty}_0 dp^{\prime\prime}\ \frac{ {p^{\prime\prime}}^2 h_k( p^{\prime\prime} ) - 2\mu_k e_k h_k(q_k) } { e_k - \frac{{p^{\prime\prime}}^2}{2\mu_k} } + h_k (q_k) \int^{\infty}_0 dp^{\prime\prime} \frac{2\mu_k e_k }{e_k -\frac{{p^{\prime\prime}}^2}{2\mu_k} } \label{eq:2.9} \end{equation} where we assume that $h_k(p)$ can be continued analytically and has no singularity in the trajectory from real $p$ to the complex value $q_k$ given in Eq.~(\ref{eq:2.6}). This is true for the case here. The cut now appears explicitly in the second term of Eq.~(\ref{eq:2.9}). It is then easy to show \begin{equation} \int^{\infty}_0 dp^{\prime\prime} \frac{2\mu_k e_k}{e_k -\frac{{p^{\prime\prime}}^2}{2\mu_k} }\ = -i\pi \mu_k \sqrt{2\mu_k e_k} = -i\pi \mu_k q_k \label{eq:2.10} \end{equation} which defines the integral in both sheets of the Riemann $e_k$ surface, corresponding to positive and negative imaginary parts of $q_k$. From Eqs.~(\ref{eq:2.9}) and (\ref{eq:2.10}), we can rewrite Eq.~(\ref{eq:2.4}) as \begin{eqnarray} \lefteqn{< p \,|\, T_{ij} (z) \,|\, p^{\prime} > =< p \,|\, V_{ij} \,|\, p^\prime > } \nonumber \\ &+& \sum_k \left[ \, \int^\infty_0 dp^{\prime\prime} \left( \frac{{p^{\prime\prime}}^2 < p \,|\, V_{ik} \,|\, p^{\prime\prime} > <p^{\prime\prime}\,|\, T_{kj}(z) \,|\, p^\prime >} {e_k-\frac{{p^{\prime\prime}}^2}{2\mu_k}} \right. \right. \nonumber \\ &-&\left. \left. \frac{ 2\mu_k e_k < p \,|\, V_{ik} \,|\, q_k > < q_k \,|\, T_{kj} (z) \,|\, p^\prime >} {e_k-\frac{{p^{\prime\prime}}^2}{2\mu_k}} \right) -i\pi \, \mu_k q_k < p \,|\, V_{ik} \,|\, q_k > < q_k \,|\, T_{kj}(z) \,|\, p^{\prime} > \right] . \nonumber \\ && \label{eq:2.11} \end{eqnarray} This equation contains a new $t$-matrix element $<q_k \mid T_{kj} \mid p'>$ which requires the additional equation \samepage{ \begin{eqnarray} \lefteqn{< q_k \,|\, T_{ij} (z) \,|\, p^{\prime} > =< q_k \,|\, V_{ij} \,|\, p^\prime >} \nonumber \\ &+&\sum_k \left[ \, \int^\infty_0 dp^{\prime\prime} \left( \frac{{p^{\prime\prime}}^2 < q_k \,|\, V_{ik} \,|\, p^{\prime\prime} > <p^{\prime\prime}\,|\, T_{kj}(z) \,|\, p^\prime > } {e_k-\frac{{p^{\prime\prime}}^2}{2\mu_k}} \right. \right. \nonumber \\ &-&\left. \left. \frac{2\mu_k e_k < q_k \,|\, V_{ik} \,|\, q_k > < q_k \,|\, T_{kj} (z) \,|\, p^\prime > } {e_k-\frac{{p^{\prime\prime}}^2}{2\mu_k}} \right) -i\pi \, \mu_k q_k < q_k \,|\, V_{ik} \,|\, q_k > < q_k \,|\, T_{kj}(z) \,|\, p^\prime > \right] . \nonumber \\ && \label{eq:2.12} \end{eqnarray} } The Eqs.~(\ref{eq:2.11}) and (\ref{eq:2.12}) form a closed set of integral equations\cite{gtext} and define the $t$-matrix elements on the entire $q_k$ planes or the Riemann surface of the complex energy $E$. This set is solved in the following sections. \section{$t$-matrix for a hard core potential} Although it is now rare to represent the short-range repulsion of the NN interaction by a hard core, the Nijmegen D and F models of the $YN$ interaction with hard cores are still used frequently in hypernuclear physics. Therefore, in this section we explore a method to obtain the off-shell $t$-matrix for a hard core potential in momentum space. The off-shell $t$-matrix can be expressed as \begin{equation} < \vec{p} \,|\, T(z) \,|\, \vec{k} > =< \vec{p} \,|\, V \,|\, \Psi_{q,\, \vec{k}}^{(+)} > \label{eq:3.1} \end{equation} where $\Psi_{q, \, \vec k}^{(+)} $ is defined by \begin{equation} |\, \Psi_{ q,\, \vec{k}}^{(+)} > = |\, \vec{k} > + \, G_0(z) \, V \,|\, \Psi_{ q,\, \vec{k}} ^{(+)} > \label{eq:3.2} \end{equation} with \begin{equation} z=\frac{q^2}{2\mu}+i\varepsilon \,\,\, . \label{eq:3.3} \end{equation} To simplify the notation, the $\Lambda N-\Sigma N$ coupling has been omitted. First, we divide the interaction $V$ into the pure hard core part $U$ and the remainder $\hat V$ as \begin{equation} V=U+\hat{V} \label{eq:3.4} \end{equation} and use the two-potential formula$\cite{gw}$ to obtain \begin{equation} < \vec{p} \,|\, T(z) \,|\, \vec{k} > = < \vec{p} \,|\, U \,|\, \Phi^{(+)}_{ q,\, \vec{k}} > + < \Phi^{(-)}_{ q,\, \vec{p}} \,|\, \hat{V} \,|\, \Psi_{ q,\, \vec{k}}^{(+)} > \label{eq:3.5} \end{equation} with \begin{equation} |\, \Phi_{ q,\, \vec{k}}^{(+)} >\, = |\, \vec{k} > + \,G_0(z)\, U \,|\, \Phi_{ q,\, \vec{k}}^{(+)} > \,\, , \label{eq:3.6} \end{equation} \begin{equation} |\, \Phi_{ q,\, \vec{p}}^{(-)} >\, = |\, \vec{p} > + \, G_0(z^{\ast}) \,U \,|\, \Phi_{ q,\, \vec{p}}^{(-)} > \,\, . \label{eq:3.7} \end{equation} As we shall show later, the first term of the right-hand side of Eq.~(\ref{eq:3.5}) is expressed analytically, and the second term satisfies an integral equation similar to the Lippmann-Schwinger equation which can be solved using a standard method. Our method is thus a natural extension of a standard treatment without a hard core, and is therefore useful not only for the present purpose, but also for other few-body calculations in momentum space. The analytic expression of the first term in Eq.~(\ref{eq:3.5}) has already been given by Takemiya$\cite{takemiya}$, who proposed a method to evaluate the off-shell $t$-matrix for a hard-core potential in coordinate space. Here, we use this method only in the treatment of the pure hard-core part of the formula. $\Phi_{ q,\, \vec{k}}^{(+)}$ and $\Phi_{ q,\, \vec{p}}^{(-)}$ in Eqs.~(\ref{eq:3.6}) and (\ref{eq:3.7}) can be expanded into partial waves \begin{equation} \Phi^{(\pm)}_{ q,\, \vec{k}}(\vec{r}) =\sum_{ \stackrel{\scriptstyle l^\prime s^\prime}{\stackrel{\scriptstyle ls}{JM}} } {\mbox{\myfont y}}^{JM}_{l^\prime s^\prime} (\hat{r}) \,{\Phi^{J(\pm)}_{l^\prime s^\prime ls}}(q,k,r) \, {{\mbox{\myfont y}}_{ls}^{JM}}^{\dag} (\hat{k}) \label{eq:3.8} \end{equation} and similarly $\Psi_{ q,\, \vec{k}}^{(+)}$. Here, $ {\mbox{\myfont y}}^{JM}_{L S}$ is the simultaneous eigenfunction of $L^2$, $S^2$, $ J^2$ and $ J_z$. We denote the pure-hard core part of Eq.~(\ref{eq:3.5}) as \begin{equation} < \vec{p} \,|\, \tilde{t}(z) \,|\, \vec{k} > \equiv < \vec{p} \,| \,U\, |\, \Phi^{(+)}_{ q,\, \vec{k}} > \label{eq:3.9} \end{equation} and decompose it into partial waves \begin{equation} < \vec{p} \,|\, \tilde{t} (z) \,|\, \vec{k} > =\sum_{ \stackrel{\scriptstyle l^\prime s^\prime}{\stackrel{\scriptstyle ls}{JM}} } {\mbox{\myfont y}}^{JM}_{l^\prime s^\prime} (\hat{r}) \, \tilde{t}_{l^\prime s^\prime ls}^J (p,k;z) \, { {\mbox{\myfont y}}_{ls}^{JM}}^{\dag}(\hat{k}) \,\, . \label{eq:3.10} \end{equation} Reference~\cite{takemiya} proves that if we introduce a function $\chi$ defined by \begin{equation} \frac{ {\chi_{l^\prime s^\prime ls}^{J(\pm)}}(q,k,r) }{r} \sqrt{ \frac{2}{\pi} } \:i^l \equiv {\Phi^{J(\pm)}_{l^\prime s^\prime ls}}(q,k,r)- \sqrt{ \frac{2}{\pi} } \:i^l j_l(kr) \label{eq:3.11} \end{equation} it satisfies the equation \begin{equation} \left( q^2 + \frac{d^2}{d r^2} - \frac{l^\prime (l^\prime + 1)}{r^2} \right) {\chi^{J(\pm)}_{l^\prime s^\prime ls}}(q,k,r) -\sum_{ l^{\prime\prime} s^{\prime\prime} } {2\mu} \, U^{J}_{l^\prime s^\prime l^{\prime\prime} s^{\prime\prime}} (r) \, {\chi^{J(\pm)}_{l^{\prime\prime} s^{\prime\prime} ls}}(q,k,r) =r{2\mu} \, U^J_{l^{\prime} s^{\prime} ls}(r) \, j_l(kr) \label{eq:3.12} \end{equation} and the off-shell element of $\tilde t$ is given by $\chi$ as \begin{equation} \tilde{t}^J_{l^{\prime} s^{\prime} ls}(p,k;z)=\frac{1}{2\mu}\, \frac{2}{\pi} \, i^{-l^{\prime}+l} \int^{\infty}_0 dr \, r \, j_{l^{\prime}} (pr) \left( q^2+\frac{d^2}{dr^2} -\frac{l^{\prime}(l^{\prime}+1)}{r^2}\right) {\chi^{J(+)}_{l^{\prime} s^{\prime} ls}} (q,k,r) \,\, . \label{eq:3.13} \end{equation} Following the method described in Ref.~\cite{takemiya} one arrives at the analytic expression of the $\tilde t$ element in Eq.~(\ref{eq:3.13}). For a pure hard core potential with radius $c$, Eq.~(\ref{eq:3.12}) has the solution \begin{equation} {\chi^{J(\pm)}_{l^{\prime} s^{\prime} ls}}(q,k,r) = \delta_{l^{\prime} l} \delta_{s^{\prime} s} \times \left\{ \begin{array}{ll} \begin{minipage}{0.3\textwidth} \[-r\,j_l (kr)\] \end{minipage} &(r\le c) \\ \begin{minipage}{0.3\textwidth} \[ -\frac{j_l(kc)}{h_l^{(\pm)}(qc)} \,r\,h_l^{(\pm)} (qr) \] \end{minipage} &(r\ge c) \end{array} \right. \label{eq:3.14} \end{equation} Note, there is no coupling in $\ell$ and $s$, a result which can be found by generating the hard-core potential matrix $U$ as limits of square well potentials (see Ref.~\cite{takemiya}). Further, from Eq.~(\ref{eq:3.14}) the integration in Eq.~(\ref{eq:3.13}) is limited to $r\leq c$ and performing the integration we obtain the final expression for $\tilde t$ \begin{eqnarray} \tilde{t}^J_{l^{\prime} s^{\prime} ls}(p,k;z)&=&\delta_{l^{\prime} l} \delta_{s^{\prime} s} \, \frac{1}{2\mu} \, \frac{2}{\pi} \, i^{-l^{\prime} + l} \nonumber \\ &\times&\left[ \left. -\,c \, j_l (pc) \frac{d}{dr} \,r\, h_l^{(+)} (qr) \right|_{r=c} \, \frac{j_l(kc)}{h_l^{(+)}(qc)}+ \left. \frac{d}{dr} \, r\, j_l(pr) \right|_{r=c} \,c \:j_l (kc) \right. \nonumber \\ &-& \left. \left( q^2-p^2 \right) \int^c_0 dr\, r^2 \, j_l (pr) \,j_l (kr) \right] . \label{eq:3.15} \end{eqnarray} The last term on the right-hand side of this equation is shown in Ref.~\cite{takemiya} to be \begin{equation} \int^c_0 dr \,r^2 \,j_l(pr) \,j_l(kr) =\left\{ \begin{array}{ll} \begin{minipage}{0.6\textwidth} \[\frac{c^2}{k^2-p^2} \left( \,k\, j_l(pc) \,j_{l+1}(kc)-p \, j_l (kc) \,j_{l+1} (pc) \right)\] \end{minipage} & (p\neq k) \\ \begin{minipage}{0.6\textwidth} \[ \frac{c^2}{2k} \left( \,k\,c \left( \,j_l^2 (kc) + j_{l+1}^2 (kc) \,\right) - (2l+1)\,j_l(kc)\, j_{l+1}(kc) \right)\] \end{minipage} & (p=k) \end{array} \right. \label{eq:3.16} \end{equation} Next, let us consider the second part of the two-potential formula~(\ref{eq:3.5}). The state $\Psi_{ q,\, \vec k}^{(+)} $ given in Eq.~(\ref{eq:3.2}) satisfies another equation\cite{gw} \begin{equation} | \,\Psi_{ q,\, \vec{k}}^{(+)} > \,= |\, \Phi_{q,\, \vec{k}}^{(+)} > + \,G_U (z) \,\hat{V} \,|\, \Psi_{ q,\, \vec{k}}^{(+)} > \label{eq:3.17} \end{equation} with \begin{equation} G_U(z) = G_0 (z) + G_0 (z) \,U\, G_U (z) \,\, . \label{eq:3.18} \end{equation} Hence, if we define $\hat t$ by \begin{equation} < \vec{p} \,|\, \hat{t} (z) \,|\, \vec{k} >\, \equiv \,< \Phi_{ q,\, \vec{p}}^{(-)} \,|\, \hat{V} \,|\, \Psi_{ q,\, \vec{k}}^{(+)} > \label{eq:3.19} \end{equation} then the second part of the two-potential formula becomes \begin{equation} < \vec{p} \,|\, \hat{t}(z) \,|\, \vec{k} > \,=\, < \Phi_{ q,\, \vec{p}}^{(-)} \,|\, \hat{V} \,|\, \Phi_{ q,\, \vec{k}}^{(+)} > + < \Phi_{ q,\, \vec{p}}^{(-)} \,|\, \hat{V} \,G_U (z) \,\hat{V} \,|\, \Psi_{ q,\, \vec{k}}^{(+)} > \, . \label{eq:3.20} \end{equation} Observing that \begin{eqnarray} < \vec{p}^{\:\prime} \,|\, G_U &=& <\vec{p}^{\:\prime} \,|\, G_0 \left(\, 1+UG_U \,\right) \nonumber \\ &=& \frac{1}{z-\frac{{p^{\prime}}^2}{2\mu}} <\Phi_{q, \, \vec{p}^{\:\prime}}^{(-)} | \label{eq:3.21} \end{eqnarray} and applying it to the second term of the right-hand side of Eq.~(\ref{eq:3.20}), we arrive at the integral equation for $\hat t$ \begin{equation} <\vec{p}\,|\, \hat{t}(z) \,|\,\vec{k}>=< \Phi_{ q,\, \vec{p}}^{(-)}\,|\, \hat{V} \,| \, \Phi_{ q,\, \vec{k}}^{(+)} > +\int d\vec{p}^{\:\prime} < \Phi_{ q,\, \vec{p}}^{(-)} \,|\, \hat{V} \,|\, \vec{p}^{\:\prime} > \frac{1}{z-\frac{{p^{\prime}}^2}{2\mu}} < \vec{p}^{\:\prime} \,|\, \hat{t} (z) \,|\, \vec{k} > \, . \label{eq:3.22} \end{equation} The inputs to this integral equation, $<\Phi^{(-)}_{ q,\, \vec p} \mid \hat V \mid \Phi^{(+)}_{ q,\, \vec k}>$ and $<\Phi^{(-)}_{ q,\, \vec p} \mid \hat V \mid \vec p\, ' >$, are expressed by the scattering states from the pure hard-core part, $\Phi^{(\pm)}$, and the remainder of the potential, $\hat V$. From Eqs.~(\ref{eq:3.11}) and (\ref{eq:3.14}), the scattering states $\Phi^{(\pm )}$ can be expressed simply by spherical Bessel and Hankel functions as \begin{equation} {\Phi^{J(\pm)}_{l^{\prime} s^{\prime} ls}} (q,k,r)= \left\{ \begin{array}{ll} \begin{minipage}{0.5\textwidth} \[0\] \end{minipage} &(r<c)\\ \begin{minipage}{0.5\textwidth} \[ \delta_{l^{\prime} l} \delta_{s^{\prime} s} \,\sqrt{\frac{2}{\pi}} \:i^l \left( \,j_l(kr) - \frac{j_l(kc)}{h_l^{(\pm)}(qc)} \,h_l^{(\pm)}(qr) \,\right) \] \end{minipage} &(r>c) \end{array} \right. \label{eq:3.23} \end{equation} allowing the inputs to be easily calculated. Thus, the integral equation~(\ref{eq:3.23}) can be solved in a similar manner as described in Sec.~II. Consequently, combining the analytic expression given in Eqs.~(\ref{eq:3.15}) and (\ref{eq:3.16}) with the solution of this integral equation, we easily obtain the off-shell $t$-matrix for a hard-core potential. \section{Cusps and round peaks caused by nearby poles} As described in Sec.~I, the main aim of this paper is to search $t$-matrix poles for various $YN$ interactions around the $\Sigma N$ threshold. In Eq.~(\ref{eq:2.4}), the $t$-matrix elements are defined by the relative momenta between the hyperons and the nucleon, $q_1$ in the case of $\Lambda$-N and $q_2$ in the case of $\Sigma$-N. However, these momenta are not independent and are related to the energy $E$ through Eq.~(\ref{eq:2.5}). This can be rewritten as \begin{equation} \frac{q^2_1}{2\mu_1}+m_N+m_{\Lambda}=\frac{q^2_2}{2\mu_2} +m_N+m_{\Sigma}=E \,\, . \label{eq:4.1} \end{equation} Thus, each $t$-matrix element is a function of the energy $E$, and has branch points at the two thresholds $E=m_N+m_\Lambda$ and $E=m_N+m_\Sigma$. We therefore encounter a somewhat complicated Riemann energy surface with four sheets, and must specify how they are related to the upper and lower halves of the $q_1$ and $q_2$ planes \cite{kok,pearce}. In two-channel problems, a procedure called uniformization\cite{newton} is very convenient to map the 4 Riemann sheets into one plane. This is used in the present analysis. The uniformization procedure introduces a new variable in place of the energy, in terms of which the $t$-matrix becomes single-valued. Following Ref.~\cite{newton}, we introduce such a variable $\omega$ which satisfies \begin{equation} \frac{q_1}{\sqrt{2\mu_1}} + \frac{q_2}{\sqrt{2\mu_2}} = \Delta\ \omega \label{eq:4.2x} \end{equation} and \begin{equation} \frac{q_1}{\sqrt{2\mu_1}} - \frac{q_2}{\sqrt{2\mu_2}} = \Delta\ \omega^{-1} \label{eq:4.2} \end{equation} with \begin{equation} \Delta^2 \equiv m_{\Sigma}-m_{\Lambda} \,\, . \label{eq:4.3} \end{equation} By these relations~(\ref{eq:4.2}) and (\ref{eq:4.2x}) it is easy to realize Eq.~(\ref{eq:4.1}). These equations constitute a mapping of the Riemann energy surface to the complex $\omega$ plane which is shown in Fig.~1. Of course, there are 4 possible quadrants where $q_1$ can be located, and for each $q_1$ two different values of $q_2$ are allowed by Eq.~(\ref{eq:4.1}). Hence, there are 8 possible cases in specifying to which quadrants both $q_1$ and $q_2$ belong on their own complex planes. The complex $\omega$ plane in Fig.~1 is divided accordingly into 8 parts, each of which contains two numbers inside square brackets indicating the quadrants to which $q_1$ and $q_2$ belong. The bold line expresses the region where bound or scattering states exist if present. The $\Lambda N$ threshold is located at $\omega=i$, and the $\Sigma N$ threshold resides at $\omega=1$. If moving counter-clockwise around $\omega=i$, the quadrant to which $q_1$ belongs changes as $1\rightarrow 2\rightarrow 3\rightarrow 4$, and at the same time the quadrant of $q_2$ changes as $1\rightarrow 2\rightarrow 1\rightarrow 2$. On the other hand, if one moves around $\omega=1$ which corresponds to the $\Sigma N$ threshold, the quadrant to which $q_2$ belongs varies as $1\rightarrow 2\rightarrow 3\rightarrow 4$ and the quadrants of $q_1$ as $1\rightarrow 4\rightarrow 1\rightarrow 4$. Let us now consider the relation between the shapes of the $\Lambda N$ elastic total cross section and the positions of a pole near the $\Sigma N$ threshold. One important difference to single channel problems is that there exists the region [1,3] touching the $\Sigma N$ threshold. Suppose a pole exists in this region close to the threshold, then the $\Lambda N$ elastic total cross section takes the shape of a cusp just at the threshold. On the other hand, if a pole resides in the region [4,2] or [4,4] close to the bold line mentioned above, the cross section shows a round peak of the Breit-Wigner form. A pole lying in the region [4,2] is often called an unstable bound state (UBS) pole\cite{kok}. We shall now discuss the above mentioned behavior of the cross sections. Assuming that the $t$-matrix has a pole at the position $(q_1,q_2)=(\alpha_1,\alpha_2)$ and the corresponding energy is $E_0$, it follows that \begin{equation} E_0 = \frac{\alpha_1^2}{2\mu_1}+m_N+m_{\Lambda} = \frac{\alpha_2^2} {2\mu_2}+m_N+m_{\Sigma} \,\, . \label{eq:4.4} \end{equation} Then, the $t$-matrix elements around the pole can be approximated as \begin{equation} <p\,|\,T_{ij}(E)\,|\,p^{\prime}> \,\simeq\, \frac{R_{ij}(p,p^{\prime})}{E-E_0} \, . \label{eq:4.5} \end{equation} The approximation (\ref{eq:4.5}) of a first-order pole holds even in the case when the pole resides in Riemann sheets other than the first, which is proved for example in Ref.~\cite{elster} for a single channel case. The extension to the coupled $\Lambda N-\Sigma N$ system is straightforward. Notice, however, for all pairs of $(q_1,q_2)=(\pm\alpha_1,\pm\alpha_2)$ the energies defined by Eq.~(\ref{eq:4.1}) take the same value $E_0$ but reside in different places on the Riemann energy surface. The approximation (\ref{eq:4.5}) is therefore valid only around $(q_1,q_2)=(\alpha_1,\alpha_2)$. From Eqs.~(\ref{eq:4.1}) and (\ref{eq:4.4}), we can rewrite Eq.~(\ref{eq:4.5}) as \begin{equation} <p\,|\,T_{ij}(E)\,|\,p^{\prime}> \,\simeq\, \frac{ \tilde{R}_{ij} (p,p^{\prime}) }{q_1-\alpha_1} \label{eq:4.6} \end{equation} or \begin{equation} <p\,|\,T_{ij}(E)\,|\,p^{\prime}> \,\simeq\, \frac{ \bar{R}_{ij} (p,p^{\prime}) } {q_2-\alpha_2} \,\, . \label{eq:4.7} \end{equation} However, the expression (\ref{eq:4.6}) is not appropriate around the $\Sigma N$ threshold. As already mentioned, this is because if the pole moves around the $\Sigma N$ threshold, the quadrant to which $\alpha_1$ belongs changes as $1\rightarrow 4\rightarrow 1\rightarrow 4$, while the quadrant in which $\alpha_2$ is located changes as $1\rightarrow 2\rightarrow 3\rightarrow 4$. Therefore, the expression (\ref{eq:4.6}) can not distinguish whether the pole is situated in the regions [4,2] or [4,4], or in the regions [1,1] or [1,3]. On the other hand, the expression (\ref{eq:4.7}) can distinguish between the regions and so will be used here. Let us now infer from Eq.~(\ref{eq:4.7}) the shapes of the $\Lambda N$ elastic total cross section $\sigma$ around the $\Sigma N$ threshold. Since \begin{equation} \sigma \propto \left| \,<q_1\,|\,T_{11}(E)\,|\,q_1> \,\right|^2 \label{eq:4.8} \end{equation} the $q_2$ dependence of the cross section becomes roughly \begin{equation} \sigma \propto \left| \frac{1}{q_2-\alpha_2} \right|^2 \, . \label{eq:4.9} \end{equation} Writing $\alpha_2 =a+ib$, we have \begin{equation} \sigma \propto \left\{ \begin{array}{ll} \begin{minipage}{0.25\textwidth} \[ \frac{1}{\left( |q_2| - b \right)^2 + a^2} \] \end{minipage} &(\ q_2=i|q_2|:\ {\rm below\ the\ \Sigma N\ threshold}\ ) \\ \begin{minipage}{0.25\textwidth} \[ \frac{1}{\left( q_2 - a \right)^2 + b^2} \] \end{minipage} &(\ q_2>0:\ {\rm above\ the\ \Sigma N\ threshold}) \end{array} \right. \label{eq:4.10} \end{equation} For the three cases, when the pole is located in the regions [4,2], [1,3] and [4,4], we plot the cross sections $\sigma$ expressed by Eq.~(\ref{eq:4.10}) in Figs.~2, 3 and 4, respectively. Notice that Figs.~2(c), 3(c) and 4(c) show the cross section as a function of the energy $E$, hence its derivative at the threshold energy is infinite according to the relation~(\ref{eq:4.1}). If the pole is located in the regions [4,2] or [4,4], the cross sections show round peaks of the Breit-Wigner form (Figs.~2(c) or 4(c)), and are quite similar to the resonances in single channel problems. In contrast, if the pole sits in the region [1,3], the cross section forms a large cusp just at the threshold (Fig.~3(c)). In Ref.~\cite{kok}, these types of poles are named inelastic virtual state poles. We should recognize that such a large cusp is caused by the pole, and is not a simple threshold effect. Some such poles, as we shall show in the next section, can actually move into the region [4,2] and convert to unstable bound state poles when the potential strength is slightly increased. \section{Results and discussions} We searched $t$-matrix poles for various meson theoretical $YN$ interactions in the manner described in Secs.~II and III. We used two soft core models of the Nijmegen group, NSC89 \cite{nsc89} and the recently proposed new soft core model NSC97 \cite{nsc97}, which includes six different versions named a, b, c, d, e and f. In this study we analyzed NSC97f. Both soft core models NSC89 and NSC97f reproduce the correct binding energy of the hypertriton\cite{hyp2,hypbnl,miyagawa}. We also chose hard core models D\cite{nd} and F\cite{nf} of the Nijmegen group (abbreviated as ND and NF respectively) which are still used in hypernuclear physics studies. Fig.~5 shows the $\Lambda N$ elastic total cross sections around the $\Sigma N$ threshold for the force models above. The model NF yields a round peak, while ND and NSC89 form cusps just at the threshold. For NSC97f, the shape is unclear. All the enhancements are found to be caused by the $^3S_1-^3D_1$ force component. Unfortunately, there exist only sparse experimental data of the $\Lambda N$ cross sections, and so we can not determine its actual shape. However, a prominent peak around the $\Sigma N$ threshold has been observed in the $K^-+d\rightarrow p+\Lambda+\pi^-$ reaction\cite{kd}. For every potential used, we found a pole near the $\Sigma N$ threshold in the $^3S_1-^3D_1$ wave. These are shown in Table~I. In Fig.~6, the poles are also displayed in the complex $q_2$ ($\Sigma -N$ relative momentum) plane. For NSC97f and NF, the poles are located in the [4,2] region of the $\omega$ plane, and for NSC89 and ND, they lie in the region [1,3]. The relation between the position of the pole and the shape of the $\Lambda-N$ cross section described in Sec.~IV holds for all potentials except NSC97f. The pole for NSC97f is close to the boundary between the regions [4,2] and [1,3], and it is farther from the imaginary axis of the $q_2$ plane than for NF. This explains why the shape of the $\Lambda N$ cross sections for this potential is not a definite example of a cusp or a round peak type. For all the interaction models, the poles are close to the $\Sigma N$ threshold and cause some enhancements. For NSC97f, the unstable bound state exists in the two-body $YN$ system, and very likely in the $YNN$ system. We should emphasize that the poles in the region [1,3] which produce the cusps are as equally important as those in the region [4,2]. To demonstrate this, we calculated the trajectory of the pole for the potential ND, multiplying it by an overall strength parameter $\lambda$. The trajectory is shown in Fig~7. The pole moves from the region [1,3] into [4,2] as the potential strength increases, and becomes an unstable bound state pole. As for the location of poles in the complex energy sheets, we refer the readers to Ref.~\cite{kok} where they are nicely illustrated. We discovered that poles also exist near the $\Lambda N$ threshold. In Table~II, the antibound-state poles below the $\Lambda N$ threshold are shown for the $^1S_0$ and $^3S_1-^3D_1$ waves. The $^1S_0$ poles are relatively close to the threshold, and as expected correlate to the scattering length. As mentioned earlier, the $\Lambda N$ scattering lengths have yet to be determined because of scant cross section data. However, the analyses of the hypertriton\cite{hyp2,hypbnl,miyagawa} constrain the $S$-wave scattering lengths. The potentials NSC97f and NSC89 which reproduce both the hypertriton binding energy and the $\Lambda N$ cross section data have a $^1S_0$ scattering length within $-2.6$ to $-2.4$ fm, and a $^3S_1$ scattering length within $-1.7$ to $-1.3$ fm. The corresponding position of the $^1S_0$ pole is at about $-0.27 i$ fm$^{-1}$ in the $q_1$($\Lambda -N$ relative momentum) complex plane. Finally, we would like to point out that the analyses of the kaon photoproduction processes, $d(\gamma,\, K^+)YN$ or $^3$He$(\gamma,\, K^+)YNN$ offer a very promising way to clarify the effects caused by the $YN$ final-state interaction around the $\Lambda N$ and $\Sigma N$ thresholds. These processes are experimentally feasible at TJLAB and SPring-8. Further, the interactions of the photon and $K^+$ meson with the baryons are comparatively weak, which enables one to formulate and calculate these reactions rather well. All the techniques and insights gained in this article are immediately applicable to those reactions and we plan to perform such calculations in the near future. \acknowledgements We would like to thank W.~Gl\"ockle for discussions, and critical and helpful comments. We are also grateful to L.~Anthony for proofreading the final manuscript.
\section{Introduction} The interest devoted to the three-band Hubbard model\cite{Emery87} is due to the fact that it describes the charge properties of a CuO$_{2}$ layer, like those found in the high-$T_{C}$ superconducting cuprates, while still being comparatively simple.\cite{Dagotto94} The basic assumption which leads to the three-band Hubbard model is that the only relevant orbitals are the Cu $3d_{x^{2}-y^{2}}$ and the O $2p_{x}$ and $2p_{y}$ orbitals. This granted, the CuO$_{2}$ layer may be described by a lattice with one Cu and two O sites per unit cell with hybridization between nearest neighbor Cu-O pairs and O-O pairs. In the hole picture the three-band Hubbard Hamiltonian reads \begin{mathletters} \begin{eqnarray} H &=&H_{0}+H_{1}~\mbox{,}\label{3-band1}\\ H_{0} &=&\Delta\sum_{j\sigma}n^p_{j\sigma} +U_{d} \sum_{i}n^d_{i\uparrow}n^d_{i\downarrow}~\mbox{,}\label{3-band2}\\ H_{1} &=&t_{pd}\sum_{\langle ij\rangle \sigma}\phi^{ij}_{pd} (p^\dagger_{j\sigma}d_{i\sigma}+h.c.) \nonumber \\ &+&t_{pp}\sum_{\langle jj^\prime\rangle \sigma}\phi^{jj^\prime}_{pp} p^\dagger_{j\sigma}p_{j^\prime\sigma}~\mbox{,}\label{3-band3} \end{eqnarray} \end{mathletters} where $d^\dagger_{i\sigma}$ ($p^\dagger_{j\sigma}$) create a hole with spin $\sigma$ in the $i$-th Cu $3d$ orbital ($j$-th O $2p$ orbital), while $n^d_{i\sigma}$ ($n^p_{j\sigma}$) are the corresponding number operators. $H_{0}$ is the atomic part of the Hamiltonian with the charge-transfer energy $\Delta$ and the on-site Coulomb repulsion $U_{d}$ between Cu $3d$ holes. $H_{1}$ represents the hybridization of Cu $3d$ and O $2p$ orbitals (hopping strength $t_{pd}$) and of O $2p$ orbitals (hopping strength $t_{pp}$). The factors $\phi^{ij}_{pd}$ and $\phi^{jj^\prime}_{pp}$ give the correct sign for the hopping processes\cite{Dagotto94}, and $\langle ij\rangle$ denotes the summation over nearest neighbor pairs. In Eq.~(\ref{3-band1}) only the most important Coulomb repulsion $U_{d}$ is included, while O on-site and inter-site Coulomb repulsions have been neglected. For explicit calculations the following typical set of values for the parameters involved in Eq.~(\ref{3-band1}) will be used\cite{McMahan88} \begin{eqnarray} \Delta&=&3.5\text{ eV, }U_{d}=8.8\text{ eV,} \nonumber \\ t_{pd}&=&1.3\text{ eV, }t_{pp}=0.65\text{ eV.} \label{parameterset} \end{eqnarray} Strong correlations due to $U_{d}$ are the reason why ground state properties of Hamiltonian $H$ can be calculated only approximately and/or on finite clusters. Besides analytical approaches like, e.g., different approximations for dynamical Green's functions\cite{Oles89} mostly numerical simulations\cite{Stephan89} have been applied to the three-band Hubbard model. The aim of the present work is to derive an analytical approximation for the ground state of an infinite system at half-filling (i.e. one hole per Cu site) which correctly describes charge properties and is still comparatively easy to evaluate. The resulting approximation does not only allow for a calculation of ground-state properties of the three-band model (\ref{3-band1}). It also provides a framework for the investigation of excitations. Furthermore, the approach is sufficiently general to be applied not only to a CuO$_{2}$ plane but also to different geometries like, e.g., that of a CuO$_{3}$ corner-sharing chain.\cite{Waidacher99} Starting point of the approximation is a N\'{e}el-ordered ground state of the atomic Hamiltonian $H_{0}$, which is denoted by $\left| \psi _{0}\right\rangle $. Due to fluctuations induced by $H_{1}$ this atomic ground state $\left| \psi _{0}\right\rangle $ differs from the full ground state $\left| \Psi \right\rangle $ of Hamiltonian $H$. In Sec.~III it will be shown that a perturbative treatment of these fluctuations breaks down for parameter values which are in the physically relevant range. Therefore, ground-state fluctuations have to be treated in a non-perturbative way. In the following we will present a systematic and non-perturbative scheme to introduce these fluctuations on the background of $\left| \psi _{0}\right\rangle $. The paper is organized as follows. In Sec.~II the general formalism is presented. As an illustration in Sec.~III this formalism is applied to the (exactly solvable) problem of a single CuO$_{4}$ plaquette. The approximative ground state of an infinite, half-filled CuO$_{2}$ plane is developed in Sec.~IV, and ground-state expectation values are evaluated in Sec.~V. In Sec.~VI the results of the analytical approach are compared to Projector Quantum Monte Carlo simulations. The conclusions are presented in Sec.~VII. Finally, a more detailed justification of the approach together with a discussion of\ its relationship to the cumulant formalism \cite{Becker88}\ is given in the Appendix. \section{General formalism} The basic idea is to start with a state $\left| \psi _{0}\right\rangle $ which is a first approximation to the full ground state $\left| \Psi \right\rangle $. In the present case we will choose $\left| \psi _{0}\right\rangle $ to be a N\'{e}el-ordered ground state of the atomic Hamiltonian $H_{0}$, Eq.~(\ref{3-band2}). In $\left| \psi _{0}\right\rangle $ every Cu site is singly occupied and all O sites are empty. Next, fluctuations on the background of $\left| \psi _{0}\right\rangle $ are introduced. These ground-state fluctuations are described by fluctuation operators $F_{\alpha }$ (introduced below in more detail) which approximately transform state $\left| \psi _{0}\right\rangle $ into the full ground state $\left| \Psi\right\rangle $. Under quite general conditions it can be shown \cite{Schork92} that the transformation leading from $\left| \psi_{0}\right\rangle $ to $\left| \Psi \right\rangle $ has to be of exponential form \begin{equation} \left| \Psi \right\rangle =\exp \left( \sum_{\alpha }\lambda _{\alpha }F_{\alpha }\right) \left| \psi _{0}\right\rangle \text{ .} \label{groundstate1} \end{equation} The parameters $\lambda _{\alpha }$ are fluctuation strengths of the fluctuation operators $F_{\alpha }$. They are determined using the set of equations \begin{equation} 0=\left\langle \Psi \right| \left[ H,F_{\alpha }^{\dagger }\right] \left| \Psi \right\rangle \text{ , }\alpha =1,2,\ldots\quad\mbox{.} \label{lambda-equations1} \end{equation} Eq.~(\ref{lambda-equations1}) follows from the condition that $\left| \Psi \right\rangle $ is an eigenstate of the full Hamiltonian $H$, Eq.~(\ref{3-band1}). From Eq.~(\ref{groundstate1}) all ground-state properties can be evaluated using \begin{equation} \left\langle A\right\rangle =\frac{\left\langle \Psi \right| A\left| \Psi \right\rangle }{\left\langle \Psi \mid \Psi \right\rangle }~\mbox{.} \label{expectationvalue} \end{equation} The ground-state energy, for instance, is calculated from Eq.(\ref {expectationvalue}) with $A=H$. Equations~(\ref{groundstate1}) and (\ref {lambda-equations1}), together with an appropriate choice of fluctuation operators $F_{\alpha }$, constitute the formal framework to be used in the remainder of this work. The above formalism will be applied to an exactly solvable problem in Sec.~III. This serves both as an illustration of the approach, and as an indication to which extent a perturbative treatment of ground-state fluctuations is possible. We finally remark that Eqs.~(\ref{groundstate1}) and (\ref{lambda-equations1}) are closely related to the cumulant formalism (see Appendix). Therefore the approach presented in this work preserves size consistency. Thus, for instance, the approximated ground-state energy remains an extensive quantity. \section{Application to a single plaquette} As an example, the method presented in the last section is now used to find an approximate expression for the ground state of Hamiltonian~(\ref{3-band1}) for the case of a single CuO$_{4}$ plaquette occupied by a single hole. In this case the unperturbed ground state $\left| \psi _{0}\right\rangle $ of $H_{0}$ is a state in which the Cu site is singly occupied while the four O sites are empty. For reasons of symmetry we may use a single fluctuation operator $F_{1}$ which describes fluctuations of the hole from the Cu site to the four surrounding O sites $j$, i.e. \begin{equation} F_{1}=-\sum_{j\sigma }\phi _{pd}^{j}p_{j\sigma }^{\dagger }d_{\sigma }~\mbox{,} \label{plaqF1} \end{equation} where $\phi _{pd}^{j}$ are the phase factors introduced in Eq.~(\ref{3-band3}). According to Eq.~(\ref{groundstate1}) the full ground state of a hole on a single plaquette is expressed by \begin{equation} \left| \Psi \right\rangle =\exp \left( \lambda _{1}F_{1}\right) \left| \psi _{0}\right\rangle~\mbox{.} \label{plaqground-state} \end{equation} The norm of this state is \begin{equation} \left\langle \Psi \mid \Psi \right\rangle =1+4\lambda _{1}^{2}\text{ .} \label{plaqnorm} \end{equation} The fluctuation strength $\lambda _{1}$ in Eq.~(\ref{plaqground-state}) is determined from condition~(\ref{lambda-equations1}) \begin{equation} 0=\left\langle \Psi \right| \left[ H,F_{1}^{\dagger }\right] \left| \Psi \right\rangle \text{ .} \label{plaqlambdaequation1} \end{equation} Non-vanishing contributions in Eq.~(\ref{plaqlambdaequation1}) arise only from terms up to order $\lambda _{1}^{2}$. The following quadratic equation for $\lambda _{1}$ is obtained \begin{equation} 0=4t_{pd}-4\left( \Delta -2t_{pp}\right) \lambda _{1}-16t_{pd}\lambda _{1}^{2}\text{ .} \label{plaqlambdaequation2} \end{equation} When the positive solution for $\lambda _{1}$ is used in Eq.~(\ref{plaqground-state}) one obtains the exact ground state. The ground-state energy $E_{G}$ is calculated using Eq.~(\ref{expectationvalue}) \begin{eqnarray} E_{G}&=&-4t_{pd}\lambda _{1} \nonumber\\ &=&\frac{1}{2}\left[ \Delta -2t_{pp}-\sqrt{\left( \Delta -2t_{pp}\right) ^{2}+\left( 4t_{pd}\right) ^{2}}\,\right] \text{ .} \label{plaqenergy} \end{eqnarray} For parameter set~(\ref{parameterset}) a value of $\lambda _{1}=0.33$ results. Notice that Eq.~(\ref{plaqenergy}) contains only a reduced effective charge-transfer energy $\Delta -2t_{pp}$. The Cu-occupation number $\left\langle n_{{\rm Cu}}\right\rangle =\left\langle \Psi \right| n_{d}\left| \Psi \right\rangle \left\langle \Psi \mid \Psi \right\rangle ^{-1}$ is given by \begin{eqnarray} \left\langle n_{{\rm Cu}}\right\rangle &=&\frac{\left\langle \psi _{0}\right| \exp \left( \lambda _{1}F_{1}^{\dagger }\right) n_{d}\exp \left( \lambda _{1}F_{1}\right) \left| \psi _{0}\right\rangle }{\left\langle \Psi \mid \Psi \right\rangle } \nonumber \\ &=&\frac{\left\langle \psi _{0}\right| \left( 1+\lambda _{1}F_{1}^{\dagger }\right) n_{d}\left( 1+\lambda _{1}F_{1}\right) \left| \psi _{0}\right\rangle }{\left\langle \Psi \mid \Psi \right\rangle } \nonumber \\ &=&\frac{1}{1+4\lambda _{1}^{2}}~\mbox{.} \label{plaqncu} \end{eqnarray} From this result one may conclude that a perturbative treatment of $F_{1}$ fluctuations (i.e.\ an expansion in $\lambda _{1}$, see Appendix), also in infinite systems, is in general not possible. Typically, $\lambda _{1}$ is of the order $1/2$. An expansion of Eq.~(\ref{plaqncu}) in $\lambda _{1}$, however, diverges for $\lambda_{1}\geq 0.5$. Condition $\lambda _{1}=0.5$ is equivalent to a vanishing effective charge-transfer energy $\Delta -2t_{pp}=0$, and to a Cu-occupation number of $1/2$. At this point state $\left| \psi _{0}\right\rangle $ ceases to be a good approximation of the exact ground state $\left| \Psi \right\rangle $. This divergence has been observed previously\cite{Becker90}, although its origin was unclear at that time. \section{Application to an infinite CuO$_{2}$ plane} We now apply the formalism presented in Sec.~II to the geometry of an infinite CuO$_{2}$ plane. State $\left| \psi _{0}\right\rangle $ is again the N\'{e}el-ordered ground state of the atomic Hamiltonian $H_{0}$, Eq.~(\ref{3-band2}). Let us introduce appropriate fluctuation operators $F_{\alpha}$. First, operator $F_{1}$ from Eq.~(\ref{plaqF1}) is generalized to all $N$ Cu sites $i$ \[F_{i,1}=-\sum_{j\sigma }\phi _{pd}^{ij}\,p_{j\sigma }^{\dagger }d_{i\sigma } ~\mbox{,}\] where the sum is over the four O sites $j$ which surround Cu site $i$. The remaining operators are constructed in accordance with the following principles: (i) All operators describe delocalizations of a hole initially located at Cu site $i$. (ii) The final site in the process is reached via the shortest path accessible by Cu-O hopping processes. (iii) A summation over all equivalent final sites is taken. (iv) The signs of the hopping processes are chosen to be the negative of the phases $\phi _{pd}^{ij}$ in Hamiltonian~(\ref{3-band3})(which guarantees non-negative values for the fluctuation strengths $\lambda _{\alpha }$). \epsfxsize=0.2\textwidth \epsfbox{gfigure01.eps} \begin{figure} \narrowtext \caption{Final sites of the fluctuation operators $F_{i,\alpha }$. Cu and O sites are symbolized by squares and circles, respectively. Arrows show spin orientation and position of the holes in the atomic ground state. The fluctuations $F_{i,\alpha }$ start from Cu site $i$ and lead to the final sites sites labelled by $\alpha=1,\ldots 4$. The area shown is a quarter of the full area accessible to the fluctuations.} \end{figure} Figure~1 shows final sites reached by fluctuation operators $F_{i,\alpha }$, $\alpha =1,\ldots 4$. For reasons of symmetry only a quarter of the allowed fluctuation range is shown. Fluctuation $F_{i,2}$, for instance, describes the hopping of the hole from Cu site $i$, via O site $j$, to the four nearest-neighbor Cu sites $k$ \[ F_{i,2}=-\sum_{jk\sigma }\left( 1-n_{j \sigma }^{p}\right) d_{k\sigma}^{\dagger }d_{i\sigma }~\mbox{.} \] Final states with singly or doubly occupied Cu sites differ by the Coulomb energy $U_{d}$. Since $U_{d}$ is large we have to distinguish between these two cases. Therefore we split $F_{i,2}$ into two operators which describe a process leading to a singly or doubly occupied Cu site, respectively \begin{eqnarray*} F_{i,2s} &=&-\sum_{jk\sigma }\left( 1-n_{k \overline{\sigma }}^{d}\right) \left( 1-n_{j\sigma }^{p}\right) d_{k\sigma }^{\dagger } d_{i\sigma }~\mbox{,} \\ F_{i,2d} &=&-\sum_{jk\sigma }n_{k\overline{ \sigma }}^{d} \left( 1-n_{j\sigma }^{p}\right) d_{k\sigma }^{\dagger } d_{i\sigma }~\mbox{.} \end{eqnarray*} Note that it is not necessary to introduce a fluctuation operator which leads to the nearest neighbor Cu sites in diagonal direction (e.g. the Cu site without a label in Fig.~1). Due to the Pauli principle fluctuations to these sites are largely excluded because of antiferromagnetic order. The neglect of fluctuations leading beyond the range shown in Fig.~1 will be justified \textit{a posteriori}. It will be shown that the fluctuation strengths $\lambda _{\alpha }$ decrease rapidly with increasing length of the fluctuation processes. According to Eq.~(\ref{groundstate1}) the ground state has the following form \begin{equation} \left| \Psi \right\rangle =\exp \left( \sum_{i\alpha }\lambda _{\alpha }F_{i\alpha }\right) \left| \psi _{0}\right\rangle~\mbox{,} \label{groundstate2} \end{equation} where $\alpha $ denotes the $5$ fluctuation operators described above. Because of translational symmetry the parameters $\lambda _{\alpha }$ do not depend on the Cu-site index $i$. To simplify Eq.~(\ref{groundstate2}) we approximately factorize the exponential function with respect to the fluctuations $F_{i,1}$ \begin{equation} \left| \Psi \right\rangle =\exp \left( \sum_{i\alpha >1}\lambda _{\alpha }F_{i\alpha }\right) \exp \left( \sum_{i^{\prime }}\lambda _{1}F_{i^{\prime },1}\right) \left| \psi _{0}\right\rangle~\mbox{.} \label{groundstate3} \end{equation} This approximation amounts to the assumption that far-reaching fluctuations $ F_{i,\alpha >1}$ occur on the background of $F_{i,1}$-fluctuations which in turn are influenced only indirectly (i.e. via $\lambda _{1}$) by the former. The second exponential function in Eq.~(\ref{groundstate3}) exactly factorizes with respect to $i^{\prime }$ \begin{equation} \left| \Psi \right\rangle =\exp \left( \sum_{i\alpha >1}\lambda _{\alpha }F_{i\alpha }\right) \prod_{i^{\prime }}\left( 1+\lambda _{1}F_{i^{\prime },1}\right) \left| \psi _{0}\right\rangle~\mbox{.} \label{groundstate4} \end{equation} In Eq.~(\ref{groundstate4}) every hole may fluctuate over a total range of five plaquettes each. Notice that all holes fluctuate simultaneously. This leads to a multitude of many-body effects, i.e. the fluctuation of a hole depends on the configuration of other holes. Basically there are three types of many body effects which are exemplified in Fig.~2. First, due to the Pauli principle the fluctuation of a hole may be blocked by the presence of other holes with the same spin, as in the fluctuation process labelled (a) in Fig.~2. Second, there are processes in which holes with the same spin change place, cf. process (b). In the following we will call these processes \textit{site-changing processes}. Third, there are strong correlations due to the Hubbard $U_{d}$ on doubly occupied Cu sites, as in process (c). One common feature of all these many-body effects is that they suppress fluctuations. \epsfxsize=0.2\textwidth \epsfbox{gfigure02.eps} \begin{figure} \narrowtext \caption{Examples for many body effects. There are three types of effects: (a) processes which are excluded by the Pauli-principle, (b) site-changing processes, and (c) correlations due to the Hubbard $U_{d}$ on doubly occupied Cu sites. The approach presented in this work accounts for all effects shown here.} \end{figure} This multitude of many-body effects makes an exact evaluation of expectation values using Eq.~(\ref{groundstate4}) impossible. Further approximations are therefore necessary. Let us consider processes in which two or more holes simultaneously leave their original plaquette. The fluctuation strengths $\lambda _{\alpha}$ for such far-reaching fluctuations turn out to be small compared to $\lambda _{1}$. Therefore it should be possible to neglect the many-body effects arising in these processes (except for site-changing processes in diagonal direction, see below). In the case of $F_{i,2d}$ fluctuations, for example, we neglect the possibility that the O site $j$ between the Cu starting and final sites $i$ and $k$ may already be occupied by a hole with the same spin. This amounts to the following simplification \[ F_{i,2d} =-\sum_{jk\sigma }n_{k\overline{ \sigma }}^{d}~d_{k\sigma }^{\dagger }d_{i\sigma }~\mbox{.} \nonumber \] In this way all of the aforementioned processes are included, some of them however only in a simplified way (i.e.\ by neglecting many-body effects). In addition to all many-body effects which are due to processes where only one hole leaves its original plaquette, we furthermore take account of all site-changing processes in diagonal direction, see Fig.~2(b). The suppression of charge fluctuations due to the diagonal sites turns out to be of great importance. On the other hand, site-changing processes involving next-nearest Cu neighbors in horizontal or vertical direction can be neglected for the following reason: During these processes the paths of the holes have to cross at the intermediate O sites (i.e. sites 1 and 4 in Fig.~1). However, since the holes have the same spin, they have to avoid each other due to the Pauli principle. Site-changing processes in horizontal or vertical direction are therefore unlikely. \section{Evaluation of expectation values} We now evaluate expectation values with state~(\ref{groundstate4}). Using the above approximations the norm of this state is $\left\langle \Psi \mid \Psi \right\rangle =\nu ^{N}$, where $N$ is the number of Cu sites and, in generalization of Eq.~(\ref{plaqnorm}) \begin{equation} \nu =1+\sum_{\alpha }z_{\alpha }p_{\alpha }\lambda _{\alpha }^{2}~\mbox{.} \label{norm} \end{equation} $z_{\alpha }$ is the number of equivalent final sites of the given process (e.g.\ $z_{2s}=4$), and $p_{\alpha }$ is the probability that the configuration of the other holes makes the process possible. This probability is defined by \begin{equation} p_{\alpha }=\frac{\left\langle \Psi \right| P_{i,\alpha }\left| \Psi \right\rangle }{\left\langle \Psi \mid \Psi \right\rangle }~\mbox{,} \label{probabilities} \end{equation} where $P_{i,\alpha }$ is a projection operator on all configurations which allow for process $\alpha $. For example, $P_{i,2s}$ projects on states in which the target Cu site of fluctuation $F_{i,2s}$ is empty, whereas $P_{i,2d}$ is the projector on states with a singly occupied final site. Due to translational symmetry the probabilities $p_{\alpha }$ do not depend on the site index $i$. Obviously, Eqs.~(\ref{norm}) and (\ref{probabilities}) have to be solved self-consistently since $\left| \Psi \right\rangle$ in Eq.(\ref{probabilities}) depends on the parameters $\lambda _{\alpha}$. In the case of $p_{2s}$, for instance, we obtain by explicit calculation \begin{eqnarray} p_{2s} &=&\frac{\left\langle \Psi \right| P_{i,2s}\left| \Psi \right\rangle } {\nu ^{N}} \nonumber\\ &=&\frac{1}{\nu }\sum_{\alpha }z_{\alpha }p_{\alpha }\lambda _{\alpha }^{2} \nonumber\\ &=&1-1/\nu~\mbox{.} \nonumber \end{eqnarray} In an analogous way one finds \begin{eqnarray} p_{1} &=&1~\mbox{,}~~p_{2s}=1-1/\nu~\mbox{,}~~p_{2d}=1/\nu~\mbox{,} \nonumber \\ p_{3} &=&1-2\lambda_{1}^{2}/\nu~\mbox{,}~~ p_{4}=1-\lambda _{1}^{2}/\nu~\mbox{.} \label{explicitprobabilities} \end{eqnarray} The interpretation of Eq.~(\ref{explicitprobabilities}) is straightforward. $p_{1}=1$ holds since we assume that far-reaching fluctuations occur on the background of $F_{i,1}$-fluctuations. $1/\nu $ is the probability to find a given hole at its original Cu site. $p_{2d}$ is therefore the probability that a target Cu site is singly occupied. This is a necessary prerequisite for the fluctuation process $F_{2d}$ which leads to a double occupancy, cf. Fig.~2(c). $p_{2s}$, on the other hand, is the probability that a target Cu site is empty, as required for fluctuation process $F_{2s}$. The probability to find a given hole at a specific O site on its original plaquette is $\lambda _{1}^{2}/\nu $. Thus $p_{4}$ is the probability that the target O site of fluctuation $F_{4}$ is not blocked by the hole of same spin which resides on the neighboring Cu site, cf. Fig.~2(a). Analogously, $p_{3}$ is the probability that the target O site of fluctuation $F_{3}$ is not blocked. The additional factor $2$ in $p_{3}$ (as compared to $p_{4}$) is due to site-changing processes, cf. Fig.~2(b). The fluctuation strengths $\lambda _{\alpha }$ are calculated using Eq.~(\ref {lambda-equations1}) for an arbitrary site $i=0$ \begin{equation} 0=\left\langle \Psi \right| \left[ H,F_{0,\alpha }^{\dagger }\right] \left| \Psi \right\rangle~\mbox{.} \label{lambda-equations2} \end{equation} One obtains the following nonlinear system of equations \begin{eqnarray} 0&=& \left(E_{G}-\Delta+2t_{pp}\right)\lambda_{1}+t_{pd} +t_{pd}\lambda_{2s}p_{2s} \nonumber \\ &&+t_{pd}\lambda_{2d}p_{2d}+2t_{pp}\lambda_{3}p_{3} -2t_{pp}\lambda _{1}^{2}\lambda _{3}/\nu~\mbox{,} \label{system1} \\ 0&=&E_{G}\lambda _{2s}+4t_{pd}\lambda _{2d}\lambda _{1}/\nu +t_{pd}\lambda _{1}p_{2s} \nonumber \\ &&+2t_{pd}\lambda _{3}p_{3}p_{2s}+t_{pd}\lambda _{4}p_{4}p_{2s}~\mbox{,} \label{system2} \\ 0&=&\left( 2E_{G}-U_{dd}\right)\lambda _{2d} +4t_{pd}\lambda _{2s}\lambda_{1} \nonumber \\ &&+t_{pd}\lambda _{1}+2t_{pd}\lambda _{3}p_{3} +t_{pd}\lambda _{4}p_{4}~\mbox{,}\label{system3} \\ 0&=&\left( E_{G}-\Delta +t_{pp}\right)\lambda_{3}p_{3} +t_{pp}\lambda_{1}p_{3} \nonumber \\ &&+t_{pd}\lambda _{2s}p_{2s}p_{3} +t_{pd}\lambda _{2d}p_{2d}p_{3} \nonumber \\ &&+t_{pp}\lambda _{4}p_{3} -\left( t_{pd}+2t_{pp}\lambda _{1}\right) \lambda _{1}\lambda _{3}/\nu~\mbox{,} \label{system4} \\ 0&=&\left( E_{G}-\Delta \right)\lambda _{4}p_{4} +t_{pd}\lambda _{2s}p_{2s}p_{4}+t_{pd}\lambda _{2d}p_{2d}p_{4} \nonumber \\ &&+2t_{pp}\lambda_{3}p_{3}-\left( t_{pd}+2t_{pp}\lambda _{1}\right) \lambda _{1}\lambda _{4}/ \nu ~\mbox{,} \label{system5} \end{eqnarray} where $E_{G}=-4t_{pd}\lambda _{1}$ is the ground-state energy per Cu site (see Eq.~(\ref{physicalquantities1})). This system of equations, together with Eqs.~(\ref{norm}) and (\ref{explicitprobabilities}) can be solved self-consistently for all $\lambda _{\alpha }$, $p_{\alpha }$ and for $\nu$. The solution with the lowest value of $E_{G}$ is then used in Eq.~(\ref{groundstate4}). In the case of $\lambda _{\alpha }=0$ for all $\alpha >1$ Eq.~(\ref{system1}) reduces to Eq.~(\ref{plaqlambdaequation2}) for the single plaquette. \epsfxsize=0.4\textwidth \epsfbox{gfigure03.ps} \begin{figure} \narrowtext \caption{Delocalization probability as a function of fluctuation length. The graph shows the delocalization probability of a hole which originates from Cu site $i$. The probability is summed over the sites displayed beneath the bars, and over all equivalent final sites. The hole remains predominantly on its original plaquette. Delocalization beyond the nearest-neighbor plaquette is negligible small. The probability has been calculated using Eq.~(\ref{groundstate4}) for parameter set~(\ref{parameterset}).} \end{figure} Figure~3 shows the delocalization probability $p_{\alpha }\lambda_{\alpha }^{2}/ \nu$ of a given hole summed over equivalent final sites as a function of fluctuation length for parameter set (\ref{parameterset}). In order to demonstrate the convergence of the results additional fluctuations have been introduced which lead beyond the fluctuation range shown in Fig.~1. The contribution of these additional fluctuations to the ground-state energy amounts to less than $0.1$ percent. No significant delocalization beyond the nearest neighbor plaquette occurs. A similar observation is made when other model-parameter values are chosen within the range which is relevant for cuprate compounds. These results retrospectively justify the neglect of far-reaching fluctuations and many-body effects. Notice, however, that the neglect of many-body effects allows for unphysical fluctuations which may decrease the calculated ground-state energy below the exact value. Thus, in contrast to an exact evaluation of Eq.~(\ref{expectationvalue}), our approximate solution does not guarantee an upper limit to the exact ground-state energy. From ground state~(\ref{groundstate4}) all expectation values are easily evaluated using Eq.~(\ref{expectationvalue}). The ground-state energy per Cu site, occupation numbers, and double occupancies of Cu and O sites are \begin{eqnarray} E_{G} &=&-4t_{pd}\lambda _{1}~\mbox{,} \label{physicalquantities1} \\ \left\langle n_{{\rm Cu}}\right\rangle &=&\frac{1}{\nu }\left( 1+4\lambda _{2s}^{2}p_{2s}+4\lambda _{2d}^{2}p_{2d}\right)~\mbox{,} \label{physicalquantities2} \\ \left\langle d_{{\rm Cu}}\right\rangle &=&\frac{1}{\nu }\left( 4\lambda _{2d}^{2}p_{2d}\right)~\mbox{,} \label{physicalquantities3} \\ \left\langle n_{{\rm O}}\right\rangle &=&\frac{2}{\nu }\left( \lambda _{1}^{2}+2\lambda _{3}^{2}p_{3} +\lambda _{4}^{2}p_{4}\right)~\mbox{,}\label{physicalquantities4} \\ \left\langle d_{{\rm O}}\right\rangle &=&\frac{1}{4}\left\langle n_{{\rm O}}\right\rangle ^{2}~\mbox{.} \label{physicalquantities5} \end{eqnarray} The number of holes is conserved, i.e. $\left\langle n_{{\rm Cu}}\right\rangle +2\left\langle n_{{\rm O}}\right\rangle =1$. By comparison with Quantum Monte Carlo calculations it will be shown in the next section that these results correctly reproduce the charge properties of the ground state. However, magnetic properties like the reduction of sublattice magnetization due to fluctuations are only partly described. We have neglected many-body effects in processes in which two holes simultaneously leave their original plaquette. Therefore, no spin-flip effects are included. Thus, in the (Heisenberg-) limit of infinitely large $\Delta $ and $U_{d}$ ground state (\ref{groundstate4}) reduces to the N\'{e}el state. \section{Discussion of the results and comparison to quantum-Monte Carlo simulations} We have carried out numerical simulations of the three-band Hubbard model~(\ref{3-band1}), using the Projector Quantum Monte Carlo (PQMC) algorithm\cite{vonderLinden92}, in order to compare them with the analytical result, Eq.~(\ref{groundstate4}). In the PQMC approach the ground state of a finite cluster is projected out from a suitable trial state by applying the exponential operator $e^{-\beta H}$ onto the trial state in the limit $\beta \rightarrow \infty$. However, in numerical calculations only finite values of the parameter $\beta$ are accessible. Therefore one has to check convergence of the results with respect to $\beta$. Furthermore one has to account for possible finite size effects. \epsfxsize=0.4\textwidth \epsfbox{gfigure04.eps} \begin{figure} \narrowtext \caption{Convergence of the PQMC calculations with respect to $\beta $ and the system size. The ground-state energy $E_{G}$ per Cu site for clusters of $4\times 4$ and $6\times 6$ plaquettes is shown as a function of inverse linear system size $L$. The solid and broken error bars are the results for $\beta =4$ and $\beta =8$, respectively. The parameters are those of set~(\ref{parameterset}).} \end{figure} In the present study two different cluster sizes have been investigated. Firstly, we have used a system consisting of $4\times 4$ plaquettes (i.e.\ $48$ sites). This is the smallest cluster which allows for periodic boundary conditions while still being fully two-dimensional. Secondly, we have studied the next largest cluster, a system of $6\times 6$ plaquettes (i.e.\ $108$ sites). We have used a mean-field version of the analytical ground state~(\ref{groundstate4}) as trial state. No sign problem occurred. \epsfxsize=0.4\textwidth \epsfbox{gfigure05.eps} \begin{figure} \narrowtext \caption{Comparison of analytically calculated ground-state energies and Cu-occupation numbers (solid lines) with the results of PQMC simulations for a cluster of $4\times 4$ plaquettes (error bars connected by broken lines). As a function of $\Delta $ the plots show the ground-state energy $E_{G}$ per Cu site (upper graph), and the Cu-occupation number $\left\langle n_{{\rm Cu}}\right\rangle $ (lower graph). The other parameters are those of set~(\ref{parameterset}).} \end{figure} It turns out that the results obtained for the $4\times 4$ system with $\beta =4$ are already reasonably well converged with respect to both $\beta $ and system size. As shown in Fig.~4 for the case of the ground-state energy per Cu site the error bars for different values of $\beta $ overlap and the values of the $4\times 4$ system differ only slightly from those of the $6\times 6$ system. For this reason we restrict our simulations to a system of $4\times 4$ plaquettes with $\beta =4$ and compare the results with the analytical approach. In Figs.~5 and 6 the ground-state energies $E_{G}$ per Cu site and several occupation numbers calculated using Eqs.~(\ref{physicalquantities1})--(\ref{physicalquantities5}) are compared to the results of the PQMC simulations. The number of holes is conserved in both approaches. Thus the O-occupation number $\left\langle n_{{\rm O}}\right\rangle$ is a function of the Cu-occupation number $\left\langle n_{{\rm Cu}}\right\rangle $, and the former is therefore not shown. While the values of the model parameters are those of set (\ref{parameterset}), the charge-transfer energy $\Delta $ is varied covering the range of very large charge fluctuations $\left( \Delta =1.5\text{ eV}\right) $ to fairly small charge fluctuations $\left( \Delta =4.5\text{ eV}\right) $. In general there is a good agreement between the analytical and numerical results, especially for larger values of $\Delta $. \epsfxsize=0.4\textwidth \epsfbox{gfigure06.eps} \begin{figure} \narrowtext \caption{Comparison of analytically calculated double occupancies (solid lines) with the results of PQMC simulations for a cluster of $4\times 4$ plaquettes (error bars connected by broken lines). As a function of $\Delta $ the plots show the Cu-double occupancy $\left\langle d_{{\rm Cu}}\right\rangle $ (upper graph), and the O-double occupancy $\left\langle d_{{\rm O}}\right\rangle $ (lower graph). The other parameters are those of set (\ref{parameterset}).} \end{figure} With increasing $\Delta $ both $E_{G}$ and $\left\langle n_{{\rm Cu}} \right\rangle$ increase while the O-double occupancy $\left\langle d_{{\rm O}}\right\rangle $ decreases. This behavior is due to the suppression of fluctuations for larger values of the charge-transfer energy. For smaller $\Delta $ the analytical value for the ground-state energy lies below the PQMC result. This can be explained by the neglect of many-body effects in Eq.~(\ref{groundstate4}) which allows for more unphysical fluctuations when $\Delta $ becomes smaller. These fluctuations also lead to values for $\left\langle n_{{\rm Cu}}\right\rangle $ which are slightly larger than the PQMC result. However, even for $\Delta =1.5$ eV the relative deviation for both $E_{G}$ and $\left\langle n_{{\rm Cu}}\right\rangle$ amounts to less than $3\%$. For larger values of $\Delta $ the decrease of $\left\langle d_{{\rm O}}\right\rangle $ with increasing $\Delta $ is about two times as large as the decrease of $\left\langle d_{{\rm Cu}}\right\rangle$. The reason for this weaker dependence of $\left\langle d_{{\rm Cu}}\right\rangle$ on $\Delta$ is that an increase in the charge-transfer energy affects $\left\langle d_{{\rm Cu}}\right\rangle$ only indirectly by reducing the effective Cu-Cu hopping, while - in contrast to $\left\langle d_{{\rm O}}\right\rangle $ - the on-site energy of the final site is not changed. Both analytical and numerical results show a maximum in the Cu-double occupancy $\left\langle d_{{\rm Cu}}\right\rangle$ for intermediate values of $\Delta $. This may be interpreted as the point where $\Delta $ is already sufficiently large to force holes from O sites onto already occupied Cu sites but still not large enough to suppress the effective Cu-Cu hopping. \section{Conclusion} Summing up, we have derived an analytical approximation, Eq.~(\ref{groundstate4}), for the ground state of the three-band Hubbard model~(\ref{3-band1}) on an infinite, half filled CuO$_{2}$ plane. The approach uses fluctuation operators $F_{\alpha}$ and fluctuation strengths $\lambda_{\alpha}$ which have a clear physical interpretation. The parameters contained in Eq.~(\ref{groundstate4}) are determined self-consistently by solving a nonlinear system of equations. While the approach is non-perturbative and conserves size consistency, expectation values with the approximate ground state are still easy to evaluate. By comparison with Projector Quantum Monte Carlo simulations we have demonstrated that Eq.~(\ref{groundstate4}) gives a reliable description of charge properties covering the range from small to very large charge fluctuations. Equation~(\ref{groundstate4}) can be generalized for other geometries. Furthermore, due to the use of fluctuation operators our approach provides a natural framework for the calculation of charge excitations, for example by using projection technique. This will be demonstrated in a forthcoming publication.\cite{Waidacher99} \acknowledgements Discussions with A.~H\"ubsch and M.~Vojta are gratefully acknowledged. This work is supported by DFG through the research program of the SFB 463, Dresden.
\section{Introduction} In experiments of sufficiently high sensitivity, fluctuations in the thermal energy of signal-band degrees of freedom can be an important source of noise. These fluctuations are due to coupling between the signal-band degrees of freedom of the detector and the thermal bath of other degrees of freedom. Since this coupling is also the cause of dissipation, one can relate the thermal noise in a detector to the dissipation of signal-band excitations. The relationship is quantified, for a wide class of couplings, by the Fluctuation-Dissipation Theorem: \begin{equation} \label{FDT} x^2(f)=\frac{k_BT}{\pi^2 f^2} Re\left[\frac{1}{Z(f)}\right], \end{equation} where $x(f)$ is the spectral density of the fluctuations, $T$ is the temperature of the detector, and $Z^{-1}(f)$ is the admittance of detector excitations. \cite{Callen, Saulson:PhysRev} The real part of the admittance is proportional to the dissipation, so to minimize the thermal noise of a detector at a given temperature we must minimize its dissipation. In mechanical experiments, such as gravitational wave detectors, the relevant thermally excited degrees of freedom are actual mechanical vibrations of detector components. The level at which such a detector is limited by thermal noise can therefore depend on the internal friction of the material from which the detector components are fabricated. The lower the dissipation in the material, the lower the thermal noise. Fused silica (synthetic amorphous $\mathrm{SiO_2}$) has extremely low dissipation for audio frequency oscillations at room temperature. Therefore, fused silica is an excellent choice of material for the fabrication of critical components of a thermal noise limited detector, operating at room temperature in the audio frequency regime.\cite{BraginskiiBook} Many authors have reported measurements of the dissipation in fused silica. \cite{Fraser,Bill,Gillespie:Thesis,Lunin3,Lunin2,Litten} However, the measured quality factors, $Q$, are usually limited by dissipation mechanisms other than the intrinsic dissipation of the bulk material. \cite{PaperYinglei,Braginski,Logan} Particularly worrisome is the effect of a structurally defective and chemically impure surface, such as might result from normal handling, exposure to the atmosphere, or differential cooling during fabrication.\cite{Lunin1} Although the current investigations are focused on fused silica, the condition of the surface is also a concern for other high-$Q$ materials. \cite{juli} To investigate the amount of dissipation induced by the surface, we measured the dissipation in fused silica fibers of varying diameters, and hence varying surface-to-volume ratios. In addition to providing information about the effect of the surface, our results may be extrapolated to give an approximate value for the dissipation in the bulk material. \section{The Experiment} Our fibers were hand drawn in air, from Heraeus Co.\ ``Suprasil 2''-brand fused silica rods, using a natural gas flame. The rods had diameters between $5$ and $9~\mathrm{mm}$, and the fibers were drawn from the rods as supplied, with no additional surface preparation, either before or after drawing. The dissipation in the fibers was measured at room temperature using the ringdown method. \cite{Kovalik}$^,$\cite{PaperYinglei} Figure~\ref{setup} shows the experimental setup. The fibers, hanging freely, were excited with a comb capacitor providing an oscillating electric field with a large gradient. \cite{Cadez} Due to the dielectric properties of fused silica, an oscillating force is felt by the fiber. After a resonant mode was excited, the field was turned off, the capacitor grounded, and the fiber allowed to ring freely. The displacement of the fiber as a function of time was measured by a split photodiode shadow sensor. The envelope of the displacement was extracted and fit to the functional form \begin{equation} x_0e^{-\pi f_n \phi(f_n)t}+C, \label{dissipation angle definition} \end{equation} where $t$ is time, $x_0$ is the initial amplitude, $C$ is the level of sensor and amplifier noise, and $\phi(f_n)$ is the loss angle at the resonance frequency $f_n$. (The quality factor of a resonance is equal to the reciprocal of the loss angle; $Q=\phi(f_n)^{-1}$.) In this experiment the dominant source of noise was sensor and amplifier noise. Equation~\ref{dissipation angle definition} does not include a term describing seismic excitation of the fiber, which unlike sensor and amplifier noise, adds to the signal in quadrature.\cite{Rowan1} At the frequencies of interest, seismic excitation of fiber resonant modes was not detectable above the sensor and amplifier noise, and including such a term in Eq.~\ref{dissipation angle definition} did not measurably improve the quality of the fits. Great care was taken to eliminate all extrinsic sources of dissipation, or ``excess loss.'' In any measurement of fiber $Q$'s, there are several sources of excess loss that must be considered: residual gas damping, rubbing at the clamp-fiber interface, recoil damping, and eddy-current damping. \cite{PaperYinglei} The damping due to residual gas molecules in the vacuum chamber is given by \begin{equation} \phi_{gas} = \frac{\bar{v}}{\pi f d}\, \frac{\rho_{\scriptscriptstyle gas}}{\rho_{\scriptscriptstyle fiber}}, \end{equation} where $d$ is the fiber diameter, $f$ is the frequency of vibration, $\bar{v}$ is the average speed of the gas molecules, $\rho_{\scriptscriptstyle gas}$ is the mass density of the gas, and $\rho_{\scriptscriptstyle fiber}$ is the mass density of the fiber. Gas damping was made negligible by conducting our measurements at pressures around $10^{-6}$~torr. Taking typical values for the parameters, we estimate $\phi_{gas}\approx 10^{-9}$. To eliminate rubbing at the clamp-fiber interface we ensured that the fiber oscillation did not induce elastic deformation of the fiber material in the clamp. This was achieved by drawing the fiber monolithically from a much thicker rod (i.e. leaving the fiber attached to the rod from which it was drawn), then clamping the rod in a collet as shown in Fig.~\ref{setup}.\cite{Quinn} Recoil damping is due to coupling between the resonant modes of the fiber and low-Q resonances of the support structure. Resonant modes of the fiber having frequencies close to resonances of the supporting structure will be very strongly damped, while other modes will not be as strongly coupled and will have less recoil damping. The main diagnostic for recoil damping is therefore a strong frequency dependence of the $Q$.\cite{Logan} In this experiment, recoil damping was minimized by isolating the fiber resonances from the support resonances by a structure analogous to a double pendulum. This also serves to isolate the fiber from seismic excitation at the frequencies of interest. Figure~\ref{fiber designs} shows the three different double pendulum type structures used in this experiment. The structures were monolithic to prevent interfacial rubbing, and in each case we measured the dissipation in the lowest fiber only. In the design of type~3, the lowest ``fiber'' is simply the undrawn rod, as supplied. Eddy-current damping occurs when the oscillating fiber carries a charge, and the motion of the charges induces eddy currents in nearby conductors. Resistance in the conductors dissipates the mechanical energy stored in the currents, degrading the $Q$. The dissipation in any given fiber was not noticeably dependent on the arrangement of, or distance to, nearby conductors. We conclude that eddy-current damping was negligible in our measurements. Finally, we note that our fibers were drawn from fused silica rather than fused quartz rods. Fused quartz is fabricated from naturally occurring SiO$_2$, while fused silica is made of synthetic SiO$_2$. Fused quartz has been measured to have intrinsic room temperature $Q$'s of at most a few million. \cite{Fraser, Rowan1,Rowan2,Geppo:talk} These values are too low for our purposes and would obscure the dissipation induced by the surface layer. Figure~\ref{Q improvement} shows the variations in $Q$ measured with and without some of the precautions mentioned above. If the fiber is held directly in a clamp instead of being left attached to the rod from which it was drawn, friction due to rubbing in the clamp dominates. Without an isolation bob, recoil damping dominates and there is strong frequency dependence. Fibers clamped and hung by the rods from which they were drawn, and having a central isolation bob show the highest $Q$'s, with much less frequency dependence. In these fibers, the difference between fused quartz and fused silica becomes apparent, with fused silica exhibiting substantially higher $Q$. \section{Dissipation versus Frequency for a Typical Fiber} For each fiber, the dissipation was measured at a number of resonance frequencies. For fibers of the type~1 design, the resonance frequencies agree well with the resonance frequencies of a beam of circular cross section clamped at one end: \begin{equation} f_n=\left\{\begin{array}{ll} \frac{\pi}{8}\sqrt{\frac{Yd^2}{\rho L^4}}(0.597)^2 & n=1 \\ \frac{\pi}{8}\sqrt{\frac{Yd^2}{\rho L^4}}(n-\frac{1}{2})^2 & n \ge 2, \end{array} \right. \label{resfreqs} \end{equation} where $Y$ is Young's modulus, $d$ is the fiber diameter, $L$ is the fiber length, $\rho$ is the mass per unit volume, and $n=1,2\ldots$~is the mode number. \cite{Morse} Figure~\ref{cantilever} shows the agreement between the measured and predicted resonance frequencies for a typical fiber of type~1. The mode frequencies may be used to calculate the diameter of the fiber, and we find good agreement with the average diameter measured using a micrometer. Similarly, for the fibers of type~3, the resonance frequencies agree with those of a free beam of circular cross section. As expected, the resonance frequencies of the fiber of type~2 are not well modeled by either a free or a clamped beam. This is due to the large diameter of the excited fiber relative to the first (lowest) isolation bob which takes some part in the motion. Because of this, the fiber of type~2 was made with an extra isolation bob. Figure~\ref{qvsf} shows the dissipation versus frequency for the fiber whose resonance frequencies are shown in Fig~\ref{cantilever}. The graph shows a loss peak around $100$ Hz which is due to thermoelastic damping. \cite{Zener} Thermoelastic damping is given by \begin{equation} \phi_{therm}=\frac{Y \alpha^2 T_0}{C} \frac{\sigma f}{1+\sigma^2f^2}, \end{equation} where $Y$ is Young's modulus, $\alpha$ is the thermal expansion coefficient, $T_0$ is the fiber temperature (absolute scale), $C$ is the heat capacity per unit volume, and $f$ is the frequency of vibrations. The constant $\sigma$ is $$\sigma=\frac{2\pi}{13.55} \frac{C d^2}{\kappa},$$ where $d$ is the fiber diameter and $\kappa$ is the thermal conductivity. We found good agreement between the average measured diameter and the diameter calculated from the position of the thermoelastic damping peak. As can be seen from the graph, the dissipation is not purely thermoelastic and can be modeled by the form \begin{equation} \phi(f)=\phi_{therm}(f)+\phi_c, \label{damping versus frequency model} \end{equation} where $\phi_{therm}(f)$ is thermoelastic damping and $\phi_c$ is a frequency-independent damping term. The dissipation in most of the fibers is in good overall agreement with the form of Eq.~\ref{damping versus frequency model}. (Several fibers have a small number of modes showing anomalously high dissipation. This indicates an undiagnosed source of dissipation affecting these modes.) The appearance of a frequency-independent loss angle $\phi_c$ suggests the presence of a dissipation source whose microscopic components have a wide range of activation energies.\cite{duPre} Dissipation in the bulk material is likely to be of this type. However, a frequency-independent term might also arise from defects or impurities in the surface layer, or possibly from other sources. One way of obtaining further information as to the source of $\phi_c$ is to measure its dependence on the fiber diameter. In this way it is possible to distinguish between dissipation occurring in the bulk volume of the fiber and dissipation occurring in the surface layer. \section{Model of Diameter Dependence} The constant loss angle $\phi_c$ is modeled as consisting of two parts, one due to dissipation in the bulk and one due to dissipation in the surface layer: \begin{equation} \label{assumption} \phi_c= {(\Delta E_{bulk}+\Delta E_{surf})}/{E}, \end{equation} where $\Delta E_{bulk}$ is the energy lost per cycle in the bulk material, $\Delta E_{surf}$ is the energy lost per cycle in the surface layer, and $E$ is the total energy stored in the oscillating fiber. If we make the rather general assumption that $\Delta E_{surf}$ is proportional to the surface area $S$ while $\Delta E_{bulk}$ is proportional to the volume $V$, we may write \begin{equation} \label{prop} \frac{\Delta E_{surf}}{\Delta E_{bulk}} \propto \frac{S}{V}. \end{equation} The coefficient of proportionality depends on the ratio of the loss angle of the surface layer to the loss angle of the bulk material. It also depends on the ratio of energy stored in the surface layer to energy stored in the bulk. Some complication arises because both the loss angle of the surface layer and the density of energy stored in the surface layer are functions of depth. While we can normally calculate the energy density from the strain profile of the mode shape, we know little about dissipation in the surface layer. Since we are interested in characterizing the dissipation in the surface independently of the mode of oscillation or sample geometry, we write the coefficient of proportionality as a product of two factors \begin{equation} \label{surftobulk} \frac{\Delta E_{surf}}{\Delta E_{bulk}} = \mu \frac{d_s}{V/S}. \end{equation} The geometrical factor $\mu$ depends only on the geometry of the sample and on the mode of oscillation, while the ``dissipation depth'' $d_s$ depends only on the strength of dissipation in the surface layer relative to the bulk. The appropriate expressions for $\mu$ and $d_s$ are calculated in the Appendix. When the Young's modulus of the surface layer is the same as that of the bulk, we have \begin{equation} \label{dissipation depth} d_s =\frac{1}{\phi_{bulk}}\int_0^h \phi(n)dn, \end{equation} where $n$ is a coordinate measuring the distance inward from the surface, $\phi_{bulk} \equiv \Delta E_{bulk}/E_{bulk}$ is the loss angle of the bulk material, $\phi(n)$ is the loss angle of the surface layer as a function of depth, and $h$ is the thickness of the surface. For our fibers, having circular cross section and oscillating in transverse modes, we have $\mu = 2$. For samples with simple geometries, $\mu$ is of order unity and the volume-to-surface ratio has the same order of magnitude as the minimum thickness of the sample. When the dissipation depth is small compared to the minimum thickness of the sample, the effect of the surface on the dissipation is also small. When the dissipation depth is greater than or on the order of the minimum thickness of the sample, dissipation in the surface is likely to dominate. Since $\phi(n)$ is seldom known explicitly, a measurement of the dissipation depth provides a convenient way of comparing the surface condition of different samples made of the same material. Since $E \approx E_{bulk}$ we may rewrite Eq.~\ref{assumption} in terms of the dissipation depth, \begin{equation} \label{general result} \phi_c = \phi_{bulk}\bigl(1+ \mu \frac{d_s}{V/S} \bigr). \end{equation} In our case $V/S=d/4$, and the theory predicts \begin{equation} \label{fiber model} \phi_c = \phi_{bulk}\bigl(1+ 8\frac{d_s}{d} \bigr). \end{equation} Equation~\ref{fiber model} is the model to which we shall compare our data on dissipation versus fiber diameter. \section{Dissipation Versus Diameter} For all of the fibers, the constant loss angle $\phi_c$ was measured with the surface condition ``as drawn''. The surface of fibers drawn in a flame is largely free from microcracks \cite{Doremus,Uhlman} and we tried to avoid damaging this surface. Although care was taken during transport and during installation in the apparatus, some of the fibers did get lightly knocked against aluminum or glass components. This represents the only physical contact with the fiber surface. Measurements of $\phi_c$ were made in the following way. Since $\phi(f) \rightarrow \phi_c$ far from the thermoelastic damping peak, a measurement of the total loss angle $\phi$ at a frequency where thermoelastic damping is known to be negligible, constitutes a direct measurement of $\phi_c$. In each case $\phi_c$ was taken as the lowest measured value of the loss angle for a particular fiber. In most cases, measurements of $\phi$ could be taken at a sufficiently large range of frequencies so that those modes exhibiting the lowest dissipation gave a good approximation to the $\phi_c$ asymptote. The three thickest fibers however, posed some problems. We were only able to excite two or three modes in each, and no direct verification could be made of the existence of a $\phi_c$ asymptote. In addition, the dissipation in these fibers is very small, and correspondingly more sensitive to excess loss. Table~\ref{Qtab} lists the $Q$'s measured for the three thickest fibers. While Fiber G exhibits constant $Q$'s, the $Q$'s of Fiber J and Fiber K are quite frequency dependent. This is particularly striking for the split-frequency, third resonance mode of Fiber K, where the $Q$ changes by a factor of $3$ within $4$ Hz. The source of this excess loss is undiagnosed. Figure~\ref{diameter dependence} shows $\phi_c$ versus diameter for the 10 fibers measured. Systematic errors are likely to be far larger than the uncertainty shown (which represents the repeatability). The main source of systematic error is the upward bias of the measured dissipation due to undiagnosed sources of excess loss. We fit the data to Eq.~\ref{fiber model} and have tried to minimize the error induced by undiagnosed excess loss by including in the fit only those fibers whose graphs of dissipation versus resonance frequency do not have points deviating significantly from the form predicted by Eq.~\ref{damping versus frequency model}. (For example, the fiber whose dissipation versus frequency data is shown in Fig.~\ref{qvsf} satisfies this criterion well.) The fit determines $d_s$ and $\phi_{bulk}$, which have the values \begin{eqnarray} \label{phi_bulk value} d_s &=&180 \pm 20~\mathrm{\mu m},\\ \phi_{bulk} &=&3.3 \pm 0.3 \times 10^{-8} . \label{d_s value} \end{eqnarray} The relationship between $d_s$ and other measures of surface condition (fiber strength, surface roughness, impurities, etc.) is an interesting and open question. Further research is required to understand the dependence on surface preparation methods, storage times, manufacturing and handling, etc. The above value for $\phi_{bulk}$ is consistent with the lowest dissipation measured in fused silica.\cite{Bill,Lunin2} A quality factor of approximately $3 \times 10^7$ has been seen in hemispherical resonators of surface-treated, Russian brand KS4V fused silica at $3.7$~kHz.\cite{Lunin:Private} With knowledge of the dissipation depth, we can use Eq.~\ref{surftobulk}, to calculate the fiber diameter, $d_{eq}$, at which surface-induced dissipation becomes equal in importance to bulk-induced dissipation: $$d_{eq}=8d_s \approx 1300~\mathrm{\mu m}.$$ In order to obtain an estimate for the average loss angle in the surface layer, we model the surface layer as a homogeneous shell of thickness $h$ having a depth-independent loss angle, \hbox{$\phi(\vec{r})\equiv\phi_{surf}$}. Equation~\ref{dissipation depth} gives \begin{equation} d_s = h\frac{\phi_{surf}}{\phi_{bulk}}. \end{equation} The literature suggests several mechanisms for chemical surface damage penetrating to a depth of order 1 $\mathrm{\mu m}$.\cite{Doremus2} Taking $h=1~\mathrm{\mu m}$ and using the values given by Eqs.~\ref{phi_bulk value}~and~\ref{d_s value} we obtain \begin{equation} \phi_{surf} \approx 10^{-5}. \end{equation} \section*{Acknowledgements} We would like to thank Peter Saulson for advice, suggestions, support, and careful reading of the manuscript. We also thank William Startin and Steven Penn for useful discussions and reading the manuscript. Thanks to Yinglei Huang for teaching us the basics of fiber measurements and to Mark Beilby for helpful discussions. Additional thanks are due to Vinod Balachandran for contributing his time in the lab during the summer of 1998, and to John Schiller who is extending this work by performing surface treatments on fused silica fibers. We thank James Hough and the anonymous referee for pointing out the correct way of including seismic noise in the description of a ringdown envelope. We would especially like to thank the glassblower to Syracuse University, John Chabot, who drew all the fibers used in these measurements. This work was supported by Syracuse University and by National Science Foundation grant PHY-9602157.
\section*{Introduction} \addcontentsline{toc}{section}{Introduction} Let $d,m_1,\ldots,m_r$ be $(r+1)$ positive integers. Denote by $V(d;m_1,\ldots,m_r)$ the variety of irreducible (complex) plane curves of degree $d$ having exactly $r$ ordinary singularities of multiplicities $m_1,\ldots,m_r$. In most cases, it is still an open problem to know whether this variety is empty or not. In this paper, we will concentrate on the case where the $r$ singularities can be taken in a general position. Precisely, let $(P_1,\ldots,P_r)$ be a general $r$-tuple of point in $(\Bbb P^2)^r$. Denote by $E$ the linear system of plane curves of degree $d$ passing through the points $P_i$ $(1\leq i\leq r)$ with multiplicity at least $m_i$. The expected dimension of $E$ is $\max (-1 ; d(d+3)/2-\sum m_i(m_i+1)/2)$. \begin{theorem} \label{theorem} Given a positive integer $m$, there exists an integer $\bo d'(m)$ such that, if $m_i\leq m$ for $1\leq i\leq r$ and $d\geq \bo d'(m)$, then~: The system $E$ has the expected dimension $e$ and, if $e\geq 0$, then a general curve in $E$ is irreducible, smooth away from the $P_i$, and has an ordinary singularity of multiplicity $m_i$ at each point $P_i$. As a consequence, $V(d;m_1,\ldots,m_r)$ is not empty. \end{theorem} The importance of this result comes from the fact that it is still valid when the expected dimension is small (which happens when the number $r$ of points is high) ; even say, when $e$ is zero. In this case, the curve is isolated in $E$, and Bertini's theorem can not be used. Recent existence results have been proved by Greuel, Lossen and Shustin in the case of ordinary singularities, (\cite{gls.blowup}, section 3.3) ; or even for general singularities \cite{gls.plane}. But in all these statements the dimension of the system $E$ must be, at least, quadratic in the degree $d$. Notice, however, that the method of \cite{gls.plane} together with the vanishing result of Alexander-Hirschowitz cited below (see \cite{al-hi.asymptotic}) would easily give theorem \ref{theorem} as soon as $e\geq d+1$ (see also section \ref{sec.highdim} for such considerations). As for zero dimensional systems, a previous theorem had been proved by the author \cite{mig.mono} for $m_i\leq 3$ and $d\geq 317$. \medskip An explicit value for $\bo d'(m)$ has been computed by the author~: one may take $\bo d'(m)=2((m+2)38)^{2^{m-1}}$. According to a theorem of Alexander and Hirschowitz \cite{al-hi.asymptotic}, it is already known that there exists a bound $\bo d(m)$ for the degree, above which $E$ has the expected dimension. In theorem \ref{theorem} $\bo d'(m)$ is slightly greater than $\bo d(m)$ and is expressed in terms of it. It is now possible to follow the proof of \cite{al-hi.asymptotic} and give an explicit bound for $\bo d(m)$ (let us recall that \cite{al-hi.asymptotic} holds for any projective variety, since our bound only holds for $\Bbb P^2$). With this approach, it seems that the doubly exponential growth for the explicit value of $\bo d(m)$ is unavoidable. However this bound is far from being sharp. In fact, according to a conjecture of Hirschowitz, if the $m_i$ are in decreasing order and if $d$ is greater than $m_1+m_2+m_3$, then the system $E$ should have the expected dimension and contain an irreducible and smooth curve away from the $P_i$ (except in the well-known case $(d;m_1,\ldots,m_r)\neq (3n;n,\ldots,n)$ with $r=9$). Thus, the conjectural bound for $\bo d'(m)$ is $3m+1$. Due to its length, the computation of this explicit value is not described here. The author places it at the reader's disposal. \medskip Theorem \ref{theorem} is also interesting in view of recent results on the varieties $V(d;m_1,\ldots,m_r)$. Recall that the first variety of this type, $V=V(d;2,\ldots,2)$ was studied by Severi \cite{sev.anhangf}. He proved that $V$ is not empty and smooth if and only if $r\leq (d-1)(d-2)/2$. If in addition $r \leq d(d+3)/6$ we also know that the nodes can be taken in generic position except in the case $d=6$, $r=9$, $m_1=\cdots=m_9=2$ (case of an isolated double cubic) (see \cite{arb-cor.footnote} and \cite{treg.nodes}). In 1985, Harris \cite{ha.onseveri} completed this work, proving that $V(d;2,\ldots,2)$ is always irreducible. The questions of irreducibility and smoothness of general varieties of curves with prescribed singularities have been treated in many papers. Let us mention recent results for general singularities \cite{sh.equisingular} or for nodal curves on general surfaces in $\Bbb P^3$ \cite{ch-ci.sevvar}. However in the case considered here, i.e. plane curves with ordinary singularities, A. Bruno announced that, $V(d;m_1,\ldots,m_r)$ is irreducible, smooth and has the expected codimension assuming that it is not empty and that the singularities can be taken in generic position (conference in Toledo, September 98). This is exactly what is proved in theorem \ref{theorem}. \medskip \noindent \emph{\textbf {Strategy of the proof}} The proof of theorem \ref{theorem} is based on a lemma proved by the author in \cite{mig.horgeo} (see also \cite{bossini} for a first --not differential-- approach of this lemma). This result, which we called ``Geometric Horace Lemma'', is inspired by the Horace method of Hirschowitz (see, for example, \cite{al-hi.asymptotic}). But, while the usual Horace method can only be used to compute the \emph{dimension} of linear systems like $E$, the geometrical lemma also yields conclusions about the \emph{irreducibility} and \emph{smoothness} of the curves in $E$. The principle of the Geometric Horace Lemma is the following~: Let us choose an irreducible and smooth plane curve $C$. Let us specialize some of the $r$ points on $C$. Denote by $y=(Q_1,\ldots,Q_r)$ this special point of $(\Bbb P^2)^r$ and by $x$ the generic point of $(\Bbb P^2)^r$. Two linear systems may be considered~: $E_x=E$ when the points are in generic position and $E_y$ when they are in special position. The specialization from $x$ to $y$ is done in such a way, that $C$ is a base component of the system $E_y$. Thus a curve in $E_y$ is the union of $C$ and of a \emph{residual} curve. Under some assumptions, detailed in \ref{horgeo}, if the generic residual curve is geometrically irreducible, smooth, and has ordinary singularities, then the general curve in $E_x$ also satisfies these properties. An important point must be mentioned~: if we do not specialize enough points on $C$, then $C$ is not a base component of $E_y$ and the method fails. But, if we specialize too many points, then the dimension of the linear system grows~: $\dim E_y > \dim E_x$. This phenomenon is controlled with the help of differential conditions. It means that we have to consider some sub-systems of curves bound to pass through infinitely near points. \medskip Here is the main point of the proof~: by specializing too many points on the curve $C$, it is possible to make the dimension of $E$ grow considerably~; i.e. grow as high as the degree $d$. Then, assuming that some vanishing property holds true, the residual system is base point free, and Bertini's theorem can be used. As a consequence, a general residual curve is smooth, irreducible, and the intersection variety described above is irreducible. \medskip To make all this strategy work, we still have to check the vanishing property referred to above. Roughly speaking, it means that the residual system has the expected dimension. To prove this, we make use of the following vanishing result of Alexander and Hirschowitz \cite{al-hi.asymptotic}~: Given an integer $m$, there exists an integer $\bo a(m)$, and for $a\geq \bo a(m)$, there exists another bound $d_0(a,m)$ such that, if $C$ is the generic curve of degree $a$, if $d\geq d_0(a,m)$ and if the points $P_i$ are either generic in $\Bbb P^2$ or generic on $C$ (not too many of them) then the system $E$ has the expected dimension. In view of this result, the last choices are made~: As for the curve $C$, we choose the generic curve of degree $\bo a(m)$ ; and we only consider systems of curves of degree higher than $\bo d(m)=d_0(\bo a(m),m)$ (in fact, the final value $\bo d'(m)$ is greater than $\bo d(m)$, as appears in theorem \ref{theorem2}). \medskip \noindent \emph{\textbf {Contents}} The article is organized as follows~: In the first part, notations and definitions are set. In particular, we describe the universal variety which parameterizes the curves we are studying. In the second and third sections, we restate respectively the Geometric Horace Lemma, and the vanishing theorem of Alexander and Hirschowitz. These are the two main tools in the proof of theorem \ref{theorem}. The fourth section is devoted to the study of linear systems of ``high'' dimension, (precisely, a dimension greater than the degree $d$). In particular, when the $r$ points are in a good position, so that the vanishing lemma can be used, we show that theorem \ref{theorem} is true for these systems. In the last section , theorem \ref{theorem} is proved. \section{Curves on rational surfaces} \label{sec.prelim} In the introduction, the situation has been described on the plane. Actually, most of the proofs will be done on the plane blown-up along the $r$ points $P_1,\ldots,P_r$. In this section, we shall describe the family of rational surfaces obtained by blowing up a family of $r$ disjoint sections in $\Bbb P^2$, and the families of curves on these surfaces. We shall also set up most of the notations used in the article. \subsection{Families of rational surfaces and relative divisors} Let $r$ be a positive integer, and $X\subset (\Bbb P^2)^r$ be the open subset of $r$-tuples of distinct points. The morphism $\Bbb P^2_X=\Bbb P^2\times X\longrightarrow X$ is naturally endowed with $r$ sections~: $$ \begin{array}{cccc} \gamma_i : & X & \longrightarrow & \Bbb P^2\times X\\ &(P_1,\ldots,P_r)& \longrightarrow& (P_i,(P_1,\ldots,P_r)) \end{array} $$ Let $\Gamma_i$ be the image of $\gamma_i$ ; $\Gamma=\cup_{i=1}^r \Gamma_i$ is a nonsingular variety of $\Bbb P^2_X$. Blowing up $\Bbb P^2_X$ along $\Gamma$ produces a family of rational surfaces, parameterized by $X$~: $S_X\morph{b}\Bbb P^2_X\longrightarrow X$. Let $\pi$ denote the composed morphism $S_X\longrightarrow X$. At any point $x=(P_1,\ldots,P_r)$ of $X$, the fiber of $\pi$ will be denoted by $S_x$. This surface $S_x$ is simply the projective plane blown up along the $r$ points $P_1,\ldots,P_r$. \medskip Let us keep in mind that a relative effective Cartier divisor of $S_X$ on $X$ is simply an ideal sheaf $\f I_D$ on $S_X$, locally principal, and not a zero-divisor in any fiber of $\pi$ (see \cite{gro.pic}). These ideal sheaves are flat on $X$. \medskip \noindent \textbf{Examples}~: \hyp 1 Consider a line $L$ on $\Bbb P^2$, $L\times X\subset \Bbb P^2\times X$ the trivial family of lines above $X$, and $H_X=b^{-1}(L\times X)$ the total transform of $L\times X$ in $S_X$. The ideal sheaf $\f I_{H_X}$ is a relative effective Cartier divisor of $S_X$ on $X$. \hyp 2 Consider now $E_{i,X}$, the exceptional divisor obtained by blowing up the irreducible smooth variety $\Gamma_i$ ; the ideal sheaf $\f I_{E_{i,X}}$ is also a relative effective Cartier divisor of $S_X$ on $X$. \subsection{Intersection pairing and linear systems} For any $x\in X$, the Picard group of the surface $S_x$ is endowed with the usual base~: $[H_x],[-E_{1,x}],\ldots,[-E_{r,x}]$. The relative Cartier divisors being flat on $X$, these bases satisfy the following property~: \begin{proposition} Let $D$ be a relative Cartier divisor of $S_X\longrightarrow X$. For any point $x\in X$, let $[D_x]\in \Bbb Z^{r+1}$ be the class of the sheaf $\f O_{S_x}(D)$ expressed in the base defined above ; then $[D_x]$ does not depend on $x$. \end{proposition} The canonical divisor of $\ensuremath{\mathrm{Pic\ }} S_x$ is $\m \omega_x=(-3;(-1)^r)$ (the notation $(-1)^r$ means that the integer $(-1)$ is repeated $r$ times ; this convention will be kept in the sequel). The intersection pairing of $\ensuremath{\mathrm{Pic\ }} S_x$ is as follows~: $(d;m_1,\ldots,m_r).(c;n_1,\ldots,n_r)=dc-m_1n_1-\ldots -m_rn_r$. Let $\m d=(d;m_1,\ldots,m_r)\in \ensuremath{\mathrm{Pic\ }} S_x$. The sheaf $\f O(dH_x-\sum_{i=1}^r m_iE_{i,x})$ will be denoted by $\f O(\m d)$, and the complete linear system of $\f O(\m d)$ (i.e. the projective space $\Bbb P(H^0(S_x,\f O(\m d)))$ will be denoted by $\f L_x(\m d)$. The Riemann-Roch theorem for surfaces allows us to compute the Euler-Poincar\'e characteristic of $\f O(\m d)$~: $\chi(\m d)=\frac{(d+1)(d+2)}{2}- \sum_{i=1}^r\frac{m_i(m_i+1)}{2}.$ One may also compute the arithmetical genus of the eventual sections of $\f O(\m d)$~: $g(\m d)=\frac{(d-1)(d-2)}{2}- \sum_{i=1}^r\frac{m_i(m_i-1)}{2}.$ Suppose that $d\geq -2$. By Serre Duality Theorem, $h^2(S_x,\f O(\m d))=0$. The \emph{expected dimension} for $\f L_x(\m d)$ is then~: $\max(\chi(\m d)-1,-1)$. (An empty system is supposed to have dimension $-1$). A system $\f L_x(\m d)$ will be said to be \emph{regular} if it has the expected dimension. More generally, a sheaf $\f F$ on $S_x$ will be said to be regular if $h^0(S_x,\f F)= \max(\chi(\f F),0)$ and $h^1(S_x,\f F)= \max(-\chi(\f F),0)$. \subsection{Universal family of divisors} \label{family} Let $\m d=(d;m_1,\ldots,m_r)$ be an $r$-tuple of integers, and let us define $m'_i=\max(m_i,0)$ and $\m d'=(d;m'_1,\ldots,m'_r)$. Let $\Gamma_i$ be the image of the $i^{th}$ natural section $\gamma_i$ defined above, and let $Z\subset \Bbb P^2_X$ be the scheme defined by the ideal $\f I_{\Gamma_1}^{m'_1}\ldots \f I_{\Gamma_r}^{m'_r}$. The scheme $Z$ is a flat family above $X$ whose fibers $Z_x$ are unions of $r$ fat points of multiplicities $m'_1,\ldots,m'_r$ (see section \ref{sec.vanish} for a definition of fat points). Consider $x=(P_1,\ldots,P_r)\in X$. The linear system $|\f I_{Z_x}(d)|$ is the system of plane curves of degree $d$ passing through each point $P_i$ with multiplicity at least $m'_i$. One can easily see that this system is isomorphic to $\f L_x(\m d)$. \medskip Consider now the linear system $\Bbb P(H^0(\Bbb P^2,\f O(d)))\stackrel{\sim}{\longrightarrow} \Bbb P^{d(d+3)/2}$ of plane curves of degree $d$. One may define (here, only under a ``set-theoretic'' point of view, but it is endowed with a natural scheme structure) a subscheme $F$ of $\Bbb P^{d(d+3)/2}\times X$ in the following way~: \smallskip $\begin{array}{rcll} F& = & \{(D,(P_1,\ldots,P_r))\ | & \mbox{$D$ passes through each point} \\ & & &\mbox{$P_i$ with multiplicity at least $m'_i$}\}\\ \end{array}$ \smallskip This scheme $F$ parameterizes a canonical family of curves $\f D'\subset \Bbb P^2\times F$~: given $x=(D,(P_1,\ldots,P_r))$, the fiber $\f D'_x$ simply is the curve $D$. As above, it is possible to blow-up the variety $\Bbb P^2_F=\Bbb P^2\times F$ along the $r$ disjoint natural sections. Let $b_F:S_F\longrightarrow \Bbb P^2_F$ be this blowing-up. By assumption, the divisor $\f D=b_F^{-1}(\f D')-\Sigma_{i=1}^r m_iE_i$ is effective~: it is a relative effective Cartier divisor of the family $S_F\longrightarrow F$. Moreover, $\f D$ is a universal divisor~: \begin{proposition} Let $\m d\in \Bbb Z^{r+1}$. The functor $\bo F_{\m d}$ from $\mathbf{Schemes/X}$ to $\mathbf{Sets}$ such that $\bo F_{\m d}(Y)=\{D\subset S_Y, \mbox{relative effective divisor of class $\m d$ on $Y$}\}$ is represented by the couple $(F,\f D)$. \end{proposition} (This proposition is detailed in \cite{mig.these} ; see also \cite{gro.pic} and \cite{no.planecur}). \medskip Let $p:F\longrightarrow X$ be the natural projection from $F\subset \Bbb P^{d(d+3)/2} \times X$ to $X$. The fiber of $p$ over a point $x\in X$ is nothing but the linear system $\f L_x(\m d)$. Let $x'$ be the generic point of this fiber. By definition of the universal divisor $\f D$, the curve $\f D_{x'}$ is the generic curve of $\f L_x(\m d)$. It will be denoted by $\mathfrak{D}_x(\m d)$. Suppose now that a point $y\in X$ is a specialization of $x$. We will say that the curve $\mathfrak{D}_x(\m d)$ \emph{specializes to the curve} $\mathfrak{D}_y(\m d)$ if the generic point of $p^{-1}(y)=\f L_y(\m d)$ is a specialization of $x'$. This notion is of special importance if one expects to find properties of $\mathfrak{D}_x(\m d)$ from those of $\mathfrak{D}_y(\m d)$. In particular, if $\mathfrak{D}_y(\m d)$ is geometrically irreducible, smooth, or has ordinary singularities, then the same holds for $\mathfrak{D}_x(\m d)$. If the dimension of $\f L_y(\m d) $ equals the dimension of $\f L_x(\m d)$, one can easily see that $\mathfrak{D}_x(\m d)$ always specializes to $\mathfrak{D}_y(\m d)$. If the dimension grows after the specialization, extra conditions are needed. In fact, it is sufficient to prove that the strata of the cohomological stratification (associated to the sheaf $\f O_{S_X}(\m d)$) have sufficiently big codimension. This can be done with the help of differential methods (see \cite{mig.horgeo}, and the lemma \ref{horgeo}). \section{The Geometric Horace Lemma} \label{sec.horgeo} In this section, the Geometric Horace Lemma is restated and commented on. This lemma was proved by the author in \cite{mig.horgeo}. \medskip Let us first give some notations and conventions~: Let $y=(Q_1,\ldots,Q_r)$ be a point of $X$ (the notation $(Q_1,\ldots,Q_r)$ is slightly incorrect, since $y$ is generally not a closed point). Let $G$ be a closed integral subscheme of $\Bbb P^2$. We will say that the $a$ points $Q_1,\ldots, Q_a$ $(0\leq a\leq r)$ are \emph{generic and independent} on $G$ if $y$ is the generic point of a subvariety $Y\subset X$ such that $Y=G^a\times V$, where $V$ is an irreducible subscheme of $(\Bbb P^2)^{r-a}$. Suppose now that the $i$-th point $Q_i$ is a nonsingular point of a plane curve $C$. On the rational surface $S_y$, the intersection point of the exceptional divisor $E_i$ and the strict transform $\til C$ will be denoted by $Q_i^C$ (and its ideal sheaf, $\f I_{Q_i^C}$). Let $\f O(\m d=(d;m_1,\ldots,m_r))$ be an invertible sheaf on $S_x$. Global sections of the sheaf $\f I_{Q_i^C}(\m d)$ can be seen as plane curves of degree $d$ having multiplicity at least $m_j$ at each point $Q_j$ and, if the multiplicity at $Q_i$ is exactly $m_i$ having a branch tangent to $C$ at this point. \begin{lemma} \label{horgeo} Let $\m d=(d;m_1,\ldots,m_r)$ be an $r$-tuple of positive integers, and $x=(P_1,\ldots,P_r)$, $y=(Q_1,\ldots,Q_r)$ be two points of $X$ such that $x$ specializes to $y$. Let $C$ be a plane curve, and $\til C:=\til C_y$ its strict transform on $S_y$. Assume that $\til C$ is geometrically irreducible and smooth, of class $\m c\in \ensuremath{\mathrm{Pic\ }} S_y$ and of genus $g(\m c)$. \smallskip \noindent \textbf{Dimension and specialization} Suppose that~: \begin{theo} \item[$\hyp 1$] $-\alpha:=\chi(\f O_{\til C}(\m d))=\m d.\m c+1-g(\m c)\leq 0$\ \ ; \item[$2^\circ )$] At the point $y$, $g(\m c)$ points are generic and independent on $C$. \item[$3^\circ )$] If $\alpha \geq 1$, there exist $\alpha +1$ integers $i_1,\ldots,i_{\alpha +1}$ such that~: \\ $P_{i_1},\ldots,P_{i_{\alpha +1}}$ are generic and independent in the plane, and \\ $Q_{i_1},\ldots,Q_{i_{\alpha +1}}$ are generic and independent on $C$. \item[$4^\circ )$] If $\alpha =0$, $H^0(S,\f O(\m d-\m c))$ has the expected dimension~: $\chi(\m d-\m c)=\chi(\m d)$.\\ If $\alpha \geq 1$, $H^0(S,\f I_{Q_{i_1}^C\cup\cdots\cup Q_{i_{\alpha +1}}^C}(\m d-\m c))$ has the expected dimension~: $\chi(\m d-\m c)-\alpha -1 = \chi(\m d)-1$, \end{theo} \noindent Then $\f L_x(\m d)$ is regular, and, if $\chi(\m d)>0$, $\mathfrak{D}_x(\m d)$ specializes to $\mathfrak{D}_y(\m d)$. \noindent \textbf{Irreducibility} Suppose, moreover, that $\chi(\m d)>0$ and~: \begin{theo} \item [$\hyp 5$] the system $\f L_x(\m c)$ is empty. \item [$\hyp 6$] $\mathfrak{D}_y(\m d-\m c)$ is geometrically irreducible. \end{theo} \noindent Then $\mathfrak{D}_x(\m d)$ is geometrically irreducible. \noindent \textbf{Smoothness} Suppose finally that~: \begin{theo} \item [$\hyp 7$] If $\alpha =0$, $y$ is normal and the closure of $x$~; \item [$\hyp 8$] $\mathfrak{D}_y(\m d-\m c)$ meets $\til C$ transversally~; \item [$\hyp 9$] $\mathfrak{D}_y(\m d-\m c)\cap \til C$ is irreducible (not: geometrically irreducible)~; \item [$\hyp {10}$] $\mathfrak{D}_y(\m d-\m c)$ is a smooth curve, \end{theo} then $\mathfrak{D}_x(\m d)$ is smooth. \end{lemma} The system $\f L_y(\m d-\m c)$, and the curve $\mathfrak{D}_y(\m d-\m c)$ are respectively called \emph{residual system} and \emph{residual curve}. In fact, $\mathfrak{D}_y(\m d)$ is the union of $\til C$ and $\mathfrak{D}_y(\m d-\m c)$. The curve $\til C$ being well-known, this lemma allows one to get information on $\mathfrak{D}_x(\m d)$, from the properties of $\mathfrak{D}_y(\m d-\m c)$ and the relation between $\mathfrak{D}_y(\m d-\m c)$ and $\til C$. \medskip Condition \hyp 9 is certainly best described in an example~: assume that $\dim \f L_y(\m d)=0$, and denote by $Y$ the adherence of $y$ in $X$. There is only one curve in the system $\f L_z(\m d)$ for every closed point $z$ in an open subset of $Y$. Then the intersection of $\til C$ and the residual curve in $\f L_z(\m d-\m c)$ makes, as $z$ varies, a covering of degree $(\m d-\m c).\m c$ over the open subset of $Y$. The condition \hyp 9 means that this ``intersection variety'' is irreducible ; or in other words, that the monodromy group of this covering acts transitively on the intersection points of $\til C$ and the residual curve $\mathfrak{D}_y(\m d-\m c)$. \medskip A case of special interest is the case where the number $\alpha$ of \hyp 1 is positive. Roughly speaking, this situation arises when one specializes too many points on the curve $C$. Let us suppose that $\chi(\m d)>0$. Considering the exact sequence \begin{eqnarray} \label{exact} 0\longrightarrow \f O_{S_y}(\m d-\m c)\longrightarrow \f O_{S_y}(\m d)\longrightarrow \f O_{\til C}(\m d)\longrightarrow 0, \end{eqnarray} we find that $\chi(\m d-\m c)=\chi(\m d)+\alpha $. Since $\f I_{Q_{i_1}^C\cup\cdots\cup Q_{i_{\alpha +1}}^C}(\m d-\m c) $ is regular, $\f L_y(\m d-\m c)$ also is regular and $\dim \f L_y(\m d)=\dim \f L_y(\m d-\m c)=\dim \f L_x(\m d)+\alpha $. Thus, the dimension has grown by $\alpha$. \begin{remark} \label{ordinary} \textnormal{ As regards to the ordinary singularities of the plane projection of $\mathfrak{D}_x(\m d)$, the Geometric Horace Lemma yields no conclusion. But, if $\mathfrak{D}_y(\m d)=\mathfrak{D}_y(\m d-\m c)\cup \til C$ meets the exceptional divisor $E_i$ in $m_i$ distinct points, $\mathfrak{D}_x(\m d)$ also possesses this property. Hence, its projection has only ordinary singularities of the expected multiplicity. } \textnormal{ This is in particular the case if $C$ has ordinary singularities and $\f L_y(\m d-\m c)$ is base point free (Bertini's theorem applied to $E_i$ and the restricted system $\f L_y(\m d-\m c)_{|E_i}$). } \end{remark} \section{An asymptotic vanishing theorem} \label{sec.vanish} In order to use the Geometric Horace Lemma, we have to check that some linear system is regular (condition \hyp 4 of \ref{horgeo}). This will be done with the help of an asymptotic vanishing theorem of Alexander and Hirschowitz. In this section, we restate this result and the adequate definitions. As a corollary, we write down precisely the vanishing lemma used in the proof of theorem~\ref{theorem}. Here, opposed to the other sections, all the work is done on the plane, without blowing it up. It is a natural choice when dealing with the \emph{dimension} of a linear system, without consideration of the smoothness of its sections. \medskip Let us first recall some definitions~: As usual, a \emph{fat point} of support $P\in \Bbb P^2$ is a subscheme $P^m$ of $\Bbb P^2$ defined by the ideal $\f I_P^m$ ; the integer $m$ is called the \emph{multiplicity } of $P$. If $Z$ is a zero dimensional subscheme of $\Bbb P^2$, the \emph{degree} of $Z$, denoted by $\deg Z$, is the length of the ring $\f O_Z$. As an example, $\deg P^m=m(m+1)/2$. \begin{definition} Let $P\in \Bbb P^2$ be a nonsingular point of a plane curve $C$, and $i,m$ be two integers such that $0\leq i\leq m-1$. The \emph{$i$-th residue point} supported by $P$, of multiplicity $m$, with respect to $C$, is the scheme defined by the ideal $\f I_P^{m-1}\cap(\f I_C^i+\f I_P^m)$. It is denoted by $D_{C}^i(P^m)$ or $D^i(P^m)$ if no confusion can arise. A residue of type $D^{m-1}(P^m)$ is called a \emph{simple residue} (\cite{al-hi.asymptotic}, 2.2.). \end{definition} Condition \hyp 4 of lemma \ref{horgeo} is easily expressed with the help of residue points~: For example, if $P_1,\ldots,P_r$ are points of $\Bbb P^2$, such that $P_1$ is a nonsingular point of a curve $C$, and if $\m d=(d;m_1,\ldots,m_r)$ is an $(r+1)$-tuple of positive integers, we consider the zero dimensional scheme $Z=D^1(P_1^{m_1+1})\cup P_{2}^{m_{2}} \cup\cdots\cup P_r^{m_r}$. Then $|\f I_{P_1^C}(\m d)|$ is isomorphic to $|\f I_Z(d)|$. What we still have to show is that $h^1(\Bbb P^2,\f I_Z(d))$ or $h^0(\Bbb P^2,\f I_Z(d))$ is zero. The vanishing result of Alexander and Hirschowitz deals with some special types of systems $|\f I_Z(d)|$ defined below~: \begin{definitions}[\cite{al-hi.asymptotic}, 3.1] \label{defcand} Let $d,m$ and $a$ be three positive integers ; denote by $C$ the generic plane curve of degree $a$. An \emph{$(m,a)$-configuration} is a zero dimensional scheme $Z=\ensuremath{\mathrm{Const}}(Z)\cup\ensuremath{\mathrm{Free}}(Z)$, where~: $\ensuremath{\mathrm{Free}}(Z)$ is the \emph{free} part of $Z$ ; it is a union of fat points, of generic and independent support in $\Bbb P^2$. $\ensuremath{\mathrm{Const}}(Z)$ is the \emph{constrained} part of $Z$ ; it is a union of fat points or simple residue points, of generic and independent support in $C$. All these points are supposed to have a multiplicity less than or equal to $m$. A \emph{$(d,m,a)$-candidate} is an $(m,a)$-configuration $Z$ such that $\chi(\f I_Z(d))\leq 0$ and $h^0(C,\f O_{C}(d))\geq \deg(Z\cap C)$. A $(d,m,a)$-candidate such that $H^0(\Bbb P^2,\f I_Z(d))=0$ is said to be \emph{winning}. \end{definitions} \begin{remark} \label{remark} \textnormal{ If $Z$ is an $(m,a)$-configuration such that $\chi(\f I_Z(d))>0$ and $h^0(C,\f O_{C}(d))\geq \deg(Z\cap C)$, one also says that $Z$ is a $(d,m,a)$-candidate. But in this case, $Z$ is winning means that $h^1$ vanishes, whereas $h^0$ is positive. } \end{remark} The crucial vanishing result is the following~: \begin{proposition}[\cite{al-hi.asymptotic}, 7.1] \label{theoalhi} Given a positive integer $m$, there exists an integer $\bo{a}(m)$ and, for each $a\geq \bo{a}(m)$, an integer $d_0(a,m)$ such that : if $a\geq \bo{a}(m)$ and $d\geq d_0(a,m)$ then any $(d,m,a)$- candidate is winning. \end{proposition} Unfortunately, this vanishing result involves \emph{simple residues}, of type $D^{m-1}$, whereas the needed condition involves residues of type $D^1$. With our ``asymptotical'' point of view, this is essentially a technical problem. But to solve it, a little more Horace method is needed~: Let $Z$ be a closed subscheme of $\Bbb P^2$, and $C$ be an irreducible and reduced plane curve. The \emph{trace} of $Z$, denoted $Z\cap C$, is the scheme defined by the ideal $\f I_Z+\f I_C$. The residue of $Z$ with respect to $C$, denoted $Z'$, is the scheme defined by the conductor ideal $(\f I_Z:\f I_C)$. \begin{proposition}[\cite{al-hi.asymptotic}, 2.3] \label{hordiff} Let $C$ be an irreducible and smooth plane curve of degree $a$ and genus $g=(a-1)(a-2)/2$, and $d$ be an integer greater than $a$. Let $Z=Z_0\cup P_1^{m_1}\cup\cdots \cup P_\beta ^{m_\beta }$ be a zero dimensional subscheme of $\Bbb P^2$ such that $P_1,\ldots,P_\beta $ are generic and independent points of $\Bbb P^2$. Denote also by $Q_1,\ldots,Q_\beta $, $\beta $ generic and independent points of $C$. Suppose that $\chi(\f I_Z(d))\leq 0$. If~: $i)$ $H^1(\Bbb P^2,\f I_{Q_1^{m_1}\cup\cdots\cup Q_\beta ^{m_\beta }}(d-a))=0$ ; $ii)$ $\beta = da+1-g-\deg(Z\cap C)$ ; $iii)$ $\f I_{(Z_0\cap C)\cup Q_1 \cup\cdots\cup Q_\beta }(d)$ is a regular invertible sheaf of $C$ ; $iv)$ $H^0(\Bbb P^2,\f I_{Z'_0\cup D^{m_1-1}(Q_1^{m_1})\cup\cdots\cup D^{m_\beta -1}(Q_{\beta}^{m_\beta })}(d-a))=0$ ; \noindent then $H^0(\Bbb P^2,\f I_Z(d))=0$, as expected. \end{proposition} \begin{corollary} \label{vanish} Let $m,a$ be two positive integers, and $C$ be the generic plane curve of degree $a$. Denote by $x=(O_1,\ldots,O_t,P_1,\ldots,P_r)$ the generic point of $C^t\times (\Bbb P^2)^r$, and consider an integer $\alpha$ such that $0\leq \alpha \leq t$. Let $\m d=(d;n_1,\ldots,n_t,m_1,\ldots,m_r)\in \ensuremath{\mathrm{Pic\ }} S_x$ such that $n_i\leq m-1,\ m_i\leq m$, and suppose that $\chi(\m d)-\alpha= 0$. If $i)$ $a\geq \max(\bo a(m)\ ,\ 4m)$ ; $ii)$ $d\geq \max(d_0(a,m)+a\ ,\ 2am)$ ; $iii)$ $da+1-g-n_1-\ldots-n_t-\alpha \geq 0$ ; \noindent then $\f I_{O_1^C\cup\ldots\cup O_{\alpha}^C}(\m d)$ is regular. \end{corollary} \noindent \textsc{Proof : \ } Let $Y_0=D^1(O_1^{n_1+1})\cup\cdots\cup D^1(O_\alpha^{n_\alpha+1})\cup O_{\alpha+1}^{n_{\alpha+1}} \cup\cdots\cup O_t^{n_t}$, and $Y=Y_0\cup P_1^{m_1}\cup\cdots\cup P_r^{m_r}$. Clearly, $\chi(\f I_Y(d))=\chi(\m d)-\alpha=0$, therefore $\f I_{O_1^C\cup\ldots\cup O_{\alpha}^C}(\m d)$ is regular if and only if $H^0(\Bbb P^2,\f I_Y(d))=0$. \medskip \noindent $\bullet$ Let us first prove that there exists a non negative integer $s$ such that~: \vspace{-0.5cm} \begin{eqnarray} \label{ajust} \beta&:=& da+1-g-\alpha-\sum_{i=1}^tn_i-\sum_{j=1}^sm_j\in [0;m-1]\\ \label{derivable} s+\beta&\leq r \end{eqnarray} \vspace{-0.1cm} Since $0<m_i\leq m$, it will be enough to show that~: $$ \begin{array}{ll} \ &\!\! da+1-g-\alpha-\sum_{i=1}^t n_i-\sum_{j=1}^{r-m+1}m_j\leq m-1\\ \Longleftarrow_{(m_j\leq m)}&\!\! da+1-g-\alpha-\sum_{i=1}^t n_i-\sum_{j=1}^{r} m_j + (m-1)m\leq m-1\\ \Longleftarrow_{(m_j,n_i\leq m)}&\!\! da+1-g-\alpha\\ &\mbox{\qquad}-\frac{2}{m+1}\left(\sum \frac{n_i(n_i+1)}2 +\sum\frac{m_i(m_i+1)}2\right) +(m-1)^2 \leq 0\\ \Longleftrightarrow &\!\! da+1-g-\alpha-\frac{2}{m+1}\left(\frac{(d+1)(d+2)}2-\chi(\m d)\right) + (m-1)^2 \leq 0\\ \Longleftrightarrow_{(\chi(\m d)= \alpha)} &\!\! da+1-g-\frac{(d+1)(d+2)}{(m+1)}-\alpha \left(1-\frac{2}{m+1}\right) + (m-1)^2 \leq 0\\ \Longleftarrow_{(m>0)} &\!\! da+1-\frac{(a-1)(a-2)}{2} -\frac{(d+1)(d+2)}{(m+1)} + (m-1)^2 \leq 0\\ \Longleftarrow_{\scriptscriptstyle{(\!d+1\geq a(\!m+1\!)\!)}} &\!\! 1-\frac{(a-1)(a-2)}{2} -2a + (m-1)^2 \leq 0\\ \Longleftarrow_{(a\geq 4 m)} &\!\! 1- \frac{(4 m-1)(4 m-2)}{2} -8 m + (m-1)^2 \leq 0\\ \Longleftrightarrow &\!\! -7m^2-4m+1\leq 0, \mbox{\ \ which is true.} \end{array} $$ \medskip \noindent $\bullet$ Consider $Q_1,\ldots,Q_{s+\beta}$, $(s+\beta)$ generic and independent points on $C$. Denote by $Z_0$ and $Z$ the schemes \begin{eqnarray*} Z_0&:=& Y_0\cup Q_1^{m_1}\cup\cdots\cup Q_s^{m_s}\cup P_{s+\beta+1}^{m_{s+\beta+1}}\cup \cdots\cup P_r^{m_r}\\ Z &:=& Z_0\cup P_{s+1}^{m_{s+1}}\cup \cdots\cup P_{s+\beta}^{m_{s+\beta}}. \end{eqnarray*} By the Semicontinuity Theorem, if $H^0(\Bbb P^2,\f I_Z(d))=0$, then $H^0(\Bbb P^2,\f I_Y(d))=0$ as expected. To prove that $H^0(\f I_Z(d))$ is equal to zero, we make use of proposition \ref{hordiff}. Let us specialize the points $P_{s+1},\ldots,P_{s+\beta}$ to the points $Q_{s+1},\ldots,Q_{s+\beta}$. The following relation,which bound the number of generic points on $C$, will be useful~: \begin{equation} \label{bigts} t+s+\beta \geq 2a^2-\frac{a^2}{2m} \end{equation} This inequality comes from (\ref{ajust}), which yields $\sum_{i=1}^t n_i + \sum_{j=1}^{s}m_j +\beta \geq da+1-g$. Since $n_i,m_j\leq m$ and $d\geq 2am$ one gets $m(t+s+\beta)\geq 2a^2m-a^2/2 +3a/2$. Let us check conditions $i$ to $iv$ of \ref{hordiff}~: $i)$ $H^1(\Bbb P^2,\f I_{Q_{s+1}^{m_{s+1}}\cup\cdots\cup Q_{s+\beta}^{m_{s+\beta}}}(d-a))=0$ by the lemma \ref{xu} below. $ii)$ By definition of $Z$, $\deg Z\cap C= \alpha+\sum_{i=1}^t n_i+\sum_{j=1}^s m_j$. So that $\beta=da+1-g-\deg Z\cap C$. $iii)$ The divisor of $C$ defined by the ideal $\f J=\f I_{(Z_0\cap C)\cup Q_{s+1} \cup\cdots\cup Q_{s+\beta}}$ is supported on the $t+s+\beta$ points $O_1\ldots,O_t,Q_1,\ldots,Q_{s+\beta}$ which are generic and independent on $C$. Hence, if $t+s+\beta\geq g$, $\f J(d)$ is a nonspecial invertible sheaf. But, $t+s+\beta \geq 2a^2-a^2/(2m)$ (\ref{bigts}) and then $t+s+\beta \geq a^2/2\geq g$. $iv)$ Let $T=Z'_0\cup D^{m_{s+1}-1}(Q_{s+1}^{m_{s+1}}) \cup\cdots\cup D^{m_{s+\beta}-1}(Q_{s+\beta}^{m_{s+\beta}})$. Since $Z'_0$ is a union of fat points (the $D^1$ have disappeared), $T$ is an $(m,a)$-configuration. Let us prove that it is a $(d-a,m,a)$-candidate. From the definition of $\beta$, one easily sees that $\chi(\f I_T(d-a))=\chi(\f I_Z(d))=0$. The second condition is~: $h^0(C,\f O_C(d-a))-\deg (T\cap C) \geq 0$. If $s+1\leq j\leq s+\beta$, then $\deg (D^{m_j-1}(Q_j^{m_j}))$ equals $m_j$ if $m_j> 1$ and $0$ if $m_j=1$. So the inequality can be checked as follows~: $$ \begin{array}{ll} & h^0(C,\f O_C(d-a))-\deg (T\cap C)\\ \geq & (d-a)a+1-g-\sum_{i=1}^t (n_i\!-\!1)-\! \sum_{j=1}^s(m_j\!-\!1)\!-\sum_{j=s+1}^{s+\beta}m_j \\ =_{(\ref{ajust})}& -a^2+t+s+\alpha+\beta - \sum_{j=s+1}^{s+\beta}m_j \\ \geq_{(m_j\leq m)}& -a^2+t+s+\alpha+\beta -m(m-1)\\ \geq_{\mathrm{(\ref{bigts})}}& -a^2+(2a^2-a^2/2m) -m(m-1)\\ \geq_{(a\geq 4 m)}& 15m^2 -7m \geq 0. \end{array} $$ Thus $T$ is a $(d-a,m,a)$-candidate. By assumption, $a\geq \bo a(m)$, and $d-a\geq d_0(a,m)$ ; hence, by Proposition \ref{theoalhi}, $T$ is winning and $H^0(\f I_T(d-a))=0$. \smallskip It is now allowed to apply proposition \ref{hordiff} ; it gives $H^0(\f I_Z(d))=0$. \ensuremath{\square} \begin{lemma} \label{xu} Under the assumptions of corollary \ref{vanish}, consider $Q_1,\ldots,Q_\beta$, $\beta $ generic points of $C$, and $m_1,\ldots,m_\beta$, $\beta $ integers bounded by $m$. If $\beta \leq (m-1)$, then $H^1(\Bbb P^2,\f I_{Q_1^{m_1}\cup\cdots\cup Q_\beta^{m_\beta}}(d-a))=0$. \end{lemma} \noindent \textsc{Proof : \ } The only case we need to consider is $\beta=m-1$ and $m_1=\ldots=m_\beta=m$. By Xu's theorem (\cite{xu.ample}, theorem 3), $H^1(\Bbb P^2,\f I_{Q_1^{m}\cup\cdots\cup Q_\beta^{m}}(d-a))=0$, as soon as $d-a\geq 3m$ and $(d-a+3)^2>(10/9)\sum^{\beta}_{i=1} (m_i+1)^2=(10/9)(m-1)(m+1)^2$. By assumption, $d-a\geq 2am-a\geq am$ and $a\geq \sqrt 2 m$, hence $(d-a+3)^2\geq a^2m^2 \geq 2m^4 \geq (10/9)(m-1)(m+1)^2$ for any positive $m$. If $a\geq 3$ then $d-a \geq 3m$. If $a=2$ then $m$ necessary equals $1$, and the lemma is clearly true. \ensuremath{\square} \section{Systems of high dimension} \label{sec.highdim} Given a system $\f L(\m d)$ of a sufficiently high dimension, Theorem \ref{theorem} may be proved with the help of Bertini's theorem. The main point consists in showing that $\f L(\m d)$ is base point free ; this is done here with the vanishing theorem of the preceding section and a kind of Castelnuovo-Mumford's argument. \begin{proposition} \label{highdim} Let $m$ be a positive integer, $a\geq \bo a(m)$ (see prop. \ref{theoalhi}) and $C$ the generic curve of degree $a$. Let $x=(P_1,\ldots,P_r)$ be the generic point of $C^s\times (\Bbb P^2)^{r-s}$, $(0\leq s\leq r)$ and $\m d=(d;m_1,\ldots,m_r)\in \ensuremath{\mathrm{Pic\ }} S_x$ be such that $m_i\leq m$ ($1\leq i\leq r$). The class of $\til C$ in $\ensuremath{\mathrm{Pic\ }} S_x$ is denoted by $\m c$, and its genus $g$. Suppose that $d\geq d_0(a,m)+1$ (see \ref{theoalhi}), $\chi(\m d)\geq d+1$, and $\m d.\m c+1-g \geq a$. Then, $\f L_x(\m d)$ is base point free, $\mathfrak{D}_x(\m d)$ is geometrically irreducible and smooth, $\mathfrak{D}_x(\m d)$ meets $\til C$ transversally, and $\mathfrak{D}_x(\m d)\cap \til C$ is irreducible. \end{proposition} \noindent \textsc{Proof : \ } Let us first prove that $\f L_x(\m d)$ is base point free~: the characteristic $\chi(\m d)$ is greater than $1$, so we just have to show that, given a point $Q\in S_x$, $h^1(S_x, \f I_Q(\m d))=0$. Whatever the position of $Q$ is (even on an exceptional divisor), there exists a line $L$ on $\Bbb P^2$ such that $Q$ belongs to the strict transform $\til L$ of $L$. Let $\m l\in \ensuremath{\mathrm{Pic\ }} S_x$ be the class of $\til L$ ; one may write $\m l=(1;\varepsilon_1,\ldots,\varepsilon_r)$, where $\varepsilon_i=1$ if $P_i\in L$ and $0$ otherwise. Consider the exact sequence~: $$ H^1(S_x,\f O(\m d-\m l))\longrightarrow H^1(S_x,\f I_Q(\m d))\longrightarrow H^1(\til L,\f I_{Q,\til L}(\m d))$$ The scheme $Z=P_1^{m_1-\varepsilon_1}\cup\cdots\cup P_r^{m_r-\varepsilon_r}$ is clearly an $(m,a)$-configuration. Since $(\m d-\m l).\m c+1-g\geq 0$, $h^0(C,\f O(d-1))\geq (d-1)a+1-g\geq \deg (Z\cap C)$, hence $Z$ is a $(d-1,m,a)$-candidate (in the extended sense of remark \ref{remark}). Moreover, $(d-1)\geq d_0(a,m)$ ; Proposition \ref{theoalhi} shows that $\f L_x(\m d-\m l)$ is a regular system. But $\chi(\m d-\m l)\geq\chi(\m d)-(d+1)\geq 0$, therefore $h^1(\m d-\m l)=0$. Moreover, since $\til L$ is a rational curve, $|\f I_{Q,\til L}(\m d)|$ is a regular system of degree $\m d.\m l-1$. If $\m d.\m l\geq 0$, then $h^1(\f I_{Q,\til L}(\m d))=0$, as expected. Otherwise, the exact sequence $ 0\longrightarrow \f O_{S_x}(\m d-\m l)\longrightarrow \f O_{S_x}(\m d)\longrightarrow \f O_{\til L}(\m d)\longrightarrow 0$ shows that $\til L$ is a base component of $\f L_x(\m d)$. Consider another line $L'$ of $\Bbb P^2$ containing none of the $r$ points $P_1,\ldots,P_r$. The preceding argument is true, with $L'$ in place of $L$, showing that the point $Q'=\til L\cap\til L'$ can not be a base point of $\f L_x(\m d)$. This is a contradiction. Therefore, $h^1(\f O_{S_x}(\m d-\m l))=h^1(\f I_{Q,\til L}(\m d))=0$, and the first exact sequence yields $h^1(\f I_Q(\m d))=0$. \medskip \noindent $\bullet$ Thus $\f L_x(\m d)$ is base point free. Bertini's theorem shows that $\mathfrak{D}_x(\m d)$ is a smooth curve. Suppose it is not geometrically irreducible, and denote by $D_1,\ldots,D_l \ (l\geq 2)$ its geometrically irreducible components (over a bigger base field) . Let $\m d_i\in \ensuremath{\mathrm{Pic\ }} S_x$ be the class of $D_i$ $(1\leq i\leq l)$. Consider two integers $1\leq i\neq j \leq l$. The curve $\mathfrak{D}_x(\m d)$ being smooth, $D_i$ does not intersect $D_j$ and $\m d_i.\m d_j=0$. Let $P$ be a point of $D_i$. The dimension of $\f L_x(\m d_j)$ being positive (this system is base point free), there exists a curve $D'_j\in \f L_x(\m d_j)$ containing $P$. But $D'_j.D_i=0$, so $D_i$ is a component of $D'_j$, and $\f L_x(\m d_j-\m d_i)$ is effective. By the same argument $\f L_x(\m d_i-\m d_j)$ is effective too, hence $\m d_i=\m d_j$ for every $i\neq j$. Thus $\m d=l\m d_1$ and $\m d^2=0$. The equality $\chi(\m d)+g(\m d)=\m d^2+2$ yields $g(\m d)\leq 1-d$. Moreover, one can easily see that $g(l\m d_1)=lg(\m d_1)-(l-1)$. Therefore, $lg(\m d_1)\leq l-d$ and (since $g(\m d_1)\geq 0$), $l\geq d$. The only possibility is $l=d$. Then $D_1$ is the strict transform of a line such that $D_1^2=0$. One may suppose that $\m d_1=(1;1)$, and $\m d=(d;d)$. This situation never happens since, by assumption $d\geq d_0(a,m)>m$. \medskip \noindent $\bullet$ We still have to prove that $\mathfrak{D}_x(\m d)$ meets $\til C$ transversally and that $\mathfrak{D}_x(\m d)\cap \til C$ is irreducible. The first point comes from Bertini's theorem, applied to the curve $\til C$ and the restricted (base point free) linear system $\f L_x(\m d)_{|\til C}$. As for the second point, consider $\f G\subset S_x\times \f L(\m d)$, the universal divisor over $\f L(\m d)$. Let $I=\f G\cap (\til C\times \f L(\m d))$ be the intersection variety. Since $\f L(\m d)$ is base point free, the fibers of the natural projection $I\longrightarrow \til C$ are projective spaces of constant dimension. Therefore, $I$ is irreducible and $\mathfrak{D}_x(\m d)\cap \til C$, which is the generic fiber of $I\longrightarrow \f L(\m d)$ also. \ensuremath{\square} \section{Proof of Theorem 1} \label{sec.proof} In this section we gather the preceding results to prove the announced theorem. Actually, the work is not made directly on the projective plane but, rather, on the plane blown up at the $r$ points. However, the theorem proved below clearly implies the statement of the introduction. \begin{theorem} \label{theorem2} Let $m$ be a positive integer, $x$ the generic point of $(\Bbb P^2)^r$ and $\m d=(d;m_1,\ldots,m_r)\in \ensuremath{\mathrm{Pic\ }} S_x$ such that $0<m_i\leq m$ ($1\leq i\leq r$). With the notations of \ref{theoalhi}, let \smallskip $a=\bo a'(m)=\max (\bo a(m), 4 m) ; $ $\bo d'(m)= \max(d_0(a,m)+2a, a(2m+1)).$ \smallskip \noindent If $d\geq \bo d'(m)$ then $\f L_x(\m d)$ is regular, and if $\dim \f L_x(\m d)\geq 0$, the generic curve $\mathfrak{D}_x(\m d)$ is geometrically irreducible, smooth, and meets each exceptional divisor $E_i$ in $m_i$ distinct points ($1\leq i\leq r$). As a consequence, the image of $\mathfrak{D}_x(\m d)$ on the plane has an ordinary singularity of the prescribed multiplicity $m_i$ at every $P_i$. \end{theorem} \noindent \textsc{Proof : \ } Let $C$ be the generic curve of degree $a$ and genus $g$. If $\chi(\m d)\leq 0$, then the scheme $Z=P_1^{m_1}\cup\cdots\cup P_r^{m_r}$ is a $(d,m,a)$-candidate with an empty constraint part. By Proposition \ref{theoalhi}, $Z$ is winning, $\f L_x(\m d)$ is empty and the theorem is true. If $\chi(\m d)>1$ one may consider $\chi(\m d)-1$ more general points $P_{r+1},\ldots,P_{r+\chi(\m d)-1}$ of $\Bbb P^2$ and study the sub-system of curves in $\f L_x(\m d)$ passing through these supplementary points with multiplicity at least $1$. It is equivalent to study the curves in $\f L(d;m_1,\ldots,m_r)$ or in $\f L(d;m_1,\ldots,m_r,1^{\chi(\m d)-1})$. As a consequence, we can make the assumption that $\chi(\m d)=1$. \medskip \noindent $\bullet$ There exists a positive integer $s\leq r$ such that~: \begin{eqnarray} \label{ajuste} -\alpha &= & da-m_1-\cdots -m_s +1 -g \in [-d+a-m,-d+a-1]\\ \label{bigs} s&\geq& (2da-a^2)/(2m) \end{eqnarray} The second inequality follows from the first one~: (\ref{ajuste}) together with $m_i\leq m$ gives $da -ms +1-g\leq -d+a-1$ hence (since $g\leq a^2/2$), $ms\geq (2da-a^2)/2$. As for (\ref{ajuste}), since $0<m_i\leq m$, it is sufficient to show that $da-\sum_{i=1}^r m_i +1-g\leq -d+a-1$. This is a consequence of the assumption $\chi(\m d)=1$~: $$ \begin{array}{ll} & da-(m_1+\cdots +m_r)+1-g\leq -d+a-1\\ \Longleftarrow_{(m_i\leq m,g>0)}& da-\frac{2}{m+1} \left(\frac{m_1(m_1+1)}2 +\cdots +\frac{m_r(m_r+1)}2\right)\leq -d+a-1\\ \Longleftrightarrow & da-\frac{2}{m+1}\left(\frac{(d+1)(d+2)}2-\chi(\m d)\right)\leq -d+a-1\\ \Longleftrightarrow_{(\chi(\m d)=1)} & d(a+1)-\frac{d(d+3)}{(m+1)}-a+1\leq 0 \end{array} $$ which is true when $d\geq a(2m+1)$. \noindent $\bullet$ Let $x=(P_1,\ldots,P_r)$ denote the generic point of $(\Bbb P^2)^r$, and $y=(Q_1,\ldots,$ $Q_s,P_{s+1},\ldots,P_r)$ the generic point of $C^s\times (\Bbb P^2)^{r-s}$. The $r$-tuple $y$ is a specialization of $x$. The class of $\til C_y$ is $\m c=(a;1^s,0^{r-s})$. In order to apply the Geometric Horace Lemma we are going to check the points $\hyp 1$ to $\hyp {10}$ of lemma \ref{horgeo}. Condition \hyp 1 is nothing but the relation (\ref{ajuste}) above. The assumption $d\geq a(2m+1)$ and (\ref{bigs}) yield $s\geq a^2\geq g$, hence \hyp 2 is true ; moreover (\ref{bigs}) and $a\geq 4 m$ also give $s\geq 4 d -2a\geq d-a+m+1$, hence \hyp 3 is true. As for the regularity of $\f I_{Q_1^C\cup\cdots\cup Q_{\alpha+1}^C} (\m d-\m c))$, we make use of the corollary \ref{vanish}. Since $\chi(\m d)=1$, the exact sequence \ref{exact} (section \ref{sec.horgeo}) and \ref{ajuste} yield $\chi(\m d-\m c)-(\alpha+1)=0$. The choice of $\bo a'(m)$ and $\bo d'(m)$ gives $a\geq \max(\bo a(m),4 m)$ and $d-a\geq \max(d_0(a,m)+a, 2am)$. Therefore, the only remaining condition of \ref{vanish} is $$ \begin{array}{ll} & (\m d-\m c).\m c - \alpha -1 +1-g \geq 0\\ \Longleftrightarrow & (d-a).a-\sum_{i=1}^s(m_i-1)+1-g-(\alpha +1)\geq 0\\ \Longleftrightarrow_{(\ref{ajuste})}& -2\alpha -a^2+s-1\geq 0 \\ \Longleftarrow_{(\ref{bigs})}& -2\alpha - a^2 +(2da-a^2)/(2m)-1 \geq 0\\ \Longleftarrow_{(\alpha \leq d-a+m)}& d\bigl(\frac a m - 2\bigr) -a^2 \bigl( 1 +\frac 1 {2m}\bigr) +2a - 2m -1 \geq 0\\ \Longleftarrow_{(d \geq a(2m+1))}& (2a^2 +\frac {a^2} m -4am-2a)-a^2 - \frac{a^2}{2m}+2a-2m-1 \geq 0\\ \Longleftrightarrow & a^2 - 4am +\frac{a^2}{2m} -2m -1 \geq 0 \\ \Longleftarrow_{(a\geq 4m)}& 8m -2m -1 \geq 0 \end{array} $$ which is true since $m\geq 1$. \medskip \noindent $\bullet$ We are now left with the ``irreducibility'' and ``smoothness'' part of lemma \ref{horgeo}. The sheaf $\f O(\m c)$ is effective if and only if $s\leq a(a+3)/2$, which is not the case by (\ref{bigs}) ; so $\hyp 5$ is true. The $7$-th point is empty since $\alpha\geq d-a+1>0$. Now, the residual system $\f L_y(\m d-\m c)$ has a ``high'' dimension~: Precisely, the exact sequence \ref{exact} shows that $\dim \f L_x(\m d-\m c)=\alpha\in [d-a+1,d-a+m]$. Thus, the remaining condition of the Horace lemma can be proved with the proposition \ref{highdim}. By assumption $d-a\geq d_0(a,m)+1$. It is then sufficient to prove that $(\m d-\m c).\m c+1-g\geq a$. But the preceding computation shows that $(\m d-\m c).\m c+1-g\geq \alpha + 1\geq d-a+2 \geq _{(d\geq 3a)} a$. \noindent $\bullet$ Let us turn now to the question of ordinary singularities . Recalling the remark \ref{ordinary}, we only have to check that the system $\f L_y(\m d-\m c)$ is base point free, which is the case by Proposition \ref{highdim}. \ensuremath{\square}
\section{Introduction} \setcounter{footnote}{0} \renewcommand{\thefootnote}{\arabic{footnote}} The nature of the dark matter in the haloes of galaxies is an outstanding problem in astrophysics. Over the last several decades there has been great debate about whether this matter is baryonic or must be exotic. Many astronomers believed that a stellar or substellar solution to this problem might be the most simple and therefore most plausible explanation. However, in the last few years, these candidates have been ruled out as significant components of the Galactic Halo. I will discuss limits on these stellar candidates, and argue for my personal conviction that: {\bf Most of the dark matter in the Galactic Halo must be nonbaryonic.} Until recently, stellar candidates for the dark matter, including faint stars, brown dwarfs, white dwarfs, and neutron stars, were extremely popular. However, recent analysis of various data sets has shown that faint stars and brown dwarfs probably constitute no more than a few percent of the mass of our Galaxy (Bahcall, Flynn, Gould, and Kirhakos \cite {bfgk}); Graff and Freese \cite{gf96a}; Graff and Freese \cite{gf96b}; Mera, Chabrier and Schaeffer \cite{mcs}; Flynn, Gould, and Bahcall \cite{fgb}; and Freese, Fields, and Graff \cite{freese}). Specifically, using Hubble Space Telescope and parallax data (with some caveats mentioned in the text), we showed that faint stars and brown dwarfs contribute no more than 1\% of the mass density of the Galaxy. Microlensing experiments, which were designed to look for Massive Compact Halo Objects (MACHOs), also failed to find these light stellar objects and place strong limits on dark matter candidates in the $(10^{-7} - 10^{-2}) M_\odot$ mass range. Recently white dwarfs have received attention as possible dark matter candidates. Interest in white dwarfs has been motivated by microlensing events interpreted as being in the Halo, with a best fit mass of $\sim 0.5 M_\odot$. However, I will show that stellar remnants including white dwarfs and neutron stars are extremely problematic as dark matter candidates, due to a combination of mass budget issues and chemical abundances (Fields, Freese, and Graff 1998): A significant fraction of the baryons of the universe would have to be cycled through the white dwarfs (or neutron stars) and their main sequence progenitors; however, in the process, an overabundance of carbon and nitrogen is produced, far in excess of what is observed both inside the Galaxy and in the intergalactic medium. Agreement with measurements of these elements in the Ly$\alpha$ forest would require $\Omega_{\rm WD} h \leq 2 \times 10^{-4}$. Throughout, $h$ is the Hubble constant in units of 100 km s$^{-1}$ Mpc$^{-1}$. Some uncertainty in the yields of C and N from low metallicity stars motivated us (Fields, Freese, and Graff 1999) to look also at D and He$^4$, whose yields are far better understood. The abundances of D and He$^4$ can be kept in agreement with observations only for low mass white dwarf progenitors $(m_{prog} \sim 2 M_\odot)$ and $\Omega_{\rm WD} < 0.003$. In addition, another constraint arises from considering the contribution of white dwarf progenitors to the infrared background. If galactic halos contain stellar remnants, the infra-red flux from the remnant progenitors would contribute to the opacity of multi-TeV $\gamma$-rays. But the HEGRA experiment does see multi-TeV $\gamma$-rays from the blazar Mkn501 at z= 0.034. By requiring that the optical depth due to $\gamma \gamma \rightarrow e^+ e^-$ be less than 1 for a source at z=0.034, we limit the cosmological density of stellar remnants (Graff, Freese, Walker, and Pinsonneault 1999), $\Omega_{\rm WD} \leq (1-3) \times 10^{-3} h^{-1}$. Hence white dwarfs, brown dwarfs, faint stars, and neutron stars are either ruled out or extremely problematic as dark matter candidates. Then the puzzle remains: What are the 14 MACHO events that have been interpreted as being in the Halo of the Galaxy? Are some of them actually located elsewhere, such as in the LMC itself? These questions are currently unanswered. As regards the dark matter in the Halo of our Galaxy, one is driven to nonbaryonic constituents as the bulk of the matter. Possibilities include supersymmetric particles, axions, primordial black holes, or other exotic candidates. \subsection{\bf Microlensing Experiments} The MACHO (Alcock {\it et al. } \cite{macho:1yr}, \cite{macho:2yr}) and EROS (Ansari {\it et al. } \cite{ansari}) experiments have attempted to find the dark matter of our Galactic Halo by monitoring millions of stars in the neighboring Large Magellanic Cloud (LMC), which is approximately $(45-60)$ kpc away; they have monitored stars in the Small Magellenic Cloud (SMC) as well. When a Macho crosses the line of sight between a star in the LMC and us, the Macho's gravity magnifies the light of the background star. The background star gets temporarily brighter and then dims back down. The Macho acts as a lens for the background star. The duration of the event scales as $\Delta t \propto {\sqrt{m} \over v}$, where $m$ is the mass of the Macho and $v$ is the velocity perpendicular to the line of sight. Thus there is a degeneracy in the interpretation of the data between $m$ and $v$. To break the degeneracy, one has to assume a galactic model, e.g., one has to assume that the lenses are in the Halo of our Galaxy. The three events in the first year MACHO data had a typical timescale of 40 days, which corresponds (with the above assumption) to a best fit mass for the Machos of $\sim 0.1M_\odot$. With reanalysis and more data, four years of data yield 14 events of longer duration, 35-150 days (T. Axelrod \cite{axelrod}; this is the Einstein diameter crossing time). Thus the new best fit mass is roughly $$m \sim 0.5 M_\odot \, . $$ From the experiments, one can estimate what fraction of the Halo is made of Machos. Using isothermal sphere models for the Galaxy with the two year data, the Macho group estimated that 50\% (+30\%,-20\%) of the Halo could be made of Machos. However, this estimate depends sensitively on the model used for the Galaxy. Gates, Gyuk, and Turner \cite{ggt} ran millions of models and found that the number of models vs. Halo mass fraction peaks at Machos comprising (0-30)\% of the Halo, with virtually no models compatible with a 100\% Macho Halo. Hence there is evidence that a {\bf nonbaryonic} component to the Halo of our Galaxy is required. Microlensing experiments have ruled out a large class of possible baryonic dark matter components. Substellar objects in the mass range $10^{-7}M_\odot$ all the way up to $10^{-2}M_\odot$ are ruled out by the experiments. In this talk I will discuss the heavier possibilities in the range $10^{-2}M_\odot$ to few $M_\odot$. \section{\bf Baryonic Candidates} In this talk I will concentrate on baryonic candidates. Hegyi and Olive \cite{ho} ruled out large classes of baryonic candidates. See also the work of Carr \cite{carr}. Until recently the most plausible remaining possibilities for baryonic dark matter were \noindent --Red Dwarfs ($0.2 M_\odot >$ mass $> 0.09 M_\odot$). These are stars just massive enough to burn hydrogen; they shine due to fusion taking place in the core of the star. Thus these are very faint stars. \noindent --Brown Dwarfs (mass $< 0.09 M_\odot$). These are sub-stellar objects that cannot burn hydrogen. They are too light to have fusion take place in the interior. \noindent --White Dwarfs (mass $\sim 0.6 M_\odot$). These are the end-products of stellar evolution for stars of mass $< 8 M_\odot$. In this talk, I will present limits on red dwarfs (Graff and Freese 1996a), brown dwarfs (Graff and Freese 1996b), and white dwarfs (Graff, Laughlin, and Freese \cite{glf}; Fields, Freese, and Graff \cite{ffg}; Fields, Freese, and Graff \cite{ffg2}; Graff, Freese, Walker, and Pinsonneault \cite{gfwp}) as candidates for baryonic dark matter. \section{Faint Stars and Brown Dwarfs} The number of stellar objects grows with decreasing stellar mass;. Hence, until recently, there was speculation that there might be a large number of faint stars or brown dwarfs that are just too dim to have been seen. However, as I will argue these candidates (modulo caveats below) have now been ruled out as dark matter candidates. Faint stars and brown dwarfs constitute no more than a few percent of the mass of our Galactic Halo. \subsection{\bf Faint Stars} First we used Hubble Space Telescope data (Bahcall, Flynn, Gould, and Kirhakos 1994) to limit the mass density in red dwarfs to less than 1\% of the Halo (Graff and Freese 1996a). The data of Bahcall et al (1994) from HST examined a small deep field and measured the relative magnitudes of stars in the V and I bands. We used the six stars that were seen with $1.7 < V-I < 3$ to limit the density of red dwarfs in the Halo. First we obtained the distances to these stars, which are shown in Figure 1. One can see that the survey is sensitive out to at least 10 kpc. Note that the closest stars are likely disk contaminants and not included in our final analysis. We obtained estimates of the stellar masses of these objects from stellar models of Baraffe et al (1996); the masses are in the range 0.0875$M_\odot$ - 0.2$M_\odot$. \begin{figure} \centering \psfig{figure=fig1.ps, width=9cm} \caption{(taken from Graff and Freese 1996a): Distances to six stars in HST data with $1.7 < V-I < 3$ obtained by comparing apparent with absolute magnitudes of these stars.} \end{figure} For the 6 stars in the HST data with $1.7 < V-I < 3$, we thus obtained a Halo red dwarf mass density. We then compared this red dwarf mass density with virial estimates of the Halo density to see what fraction is composed of red dwarfs. We took a local Halo mass density of $\rho_o \sim 9 \times 10^{-3} M_\odot/pc^3$. Bahcall et al (1994) had made this comparison by assuming that the red dwarfs had properties of stars at the edge of the high metallicity main sequence; these authors found that red dwarfs contribute less than 6\% of the Halo density. However, Halo red dwarfs are low metallicity objects, and we were thus motivated to redo the analysis as outlined above. A ground-based search for halo red dwarfs by Boeshaar, Tyson, and Bernstein (1994) found a much smaller number. We felt that a careful reinterpretation of the Bahcall et al (1994) data was in order. Our result is that Red dwarfs with $1.7 < V-I < 3$ (i.e., mass 0.0875 $< M/M_\odot < 0.2)$, make up less than 1\% of the Halo; our best guess is that they make up 0.14\% - 0.37\% of the mass of the halo. Subsequent examination of the Hubble Deep Field by Flynn, Gould, and Bahcall \cite{fgb} and work by Mera, Chabrier, and Schaeffer \cite{mcs2} reiterated that low-mass stars represent a negligible fraction of the Halo dark matter. \subsection{\bf Brown Dwarfs} With these strong limits on the contribution of faint stars to the Galactic Halo, we then obtained a Mass Function of these same red dwarfs in order to be able to extrapolate to the brown dwarf regime; in this way we were able to limit the contribution of brown dwarfs as well. We obtained the mass function from the following relation: \begin{equation} \label{derivemassfunc} {\rm Mass Function} = {({\rm dM}_{\rm V} / {\rm dm})} \times {\rm Luminosity Function} \, . \end{equation} Here, the Mass Function (hereafter MF) is the number density of stars with mass between $m$ and $m+dm$, and the Luminosity Function (hereafter LF) is the number density of stars in a magnitude range $M \rightarrow M+dM$ (note that $M$ refers to magnitude while $m$ refers to mass). The luminosity function is what is observed; we used parallax data taken by the US Naval Observatory (Dahn et al 1995) who identified 114 halo stars. We went from this observed luminosity function to the desired mass function via stellar models of $M_V(m)$ obtained by Alexander et al. \cite{models96}. \begin{figure} \centering \psfig{figure=fig2.ps, width=9cm} \caption{(taken from Dahn et al 1995): H-R diagram of nearby stars with measured parallax. The filled circles are high metallicity disk stars; the velocity dispersion of these disk stars is $\sim 30$ km/sec. The open circles are low metallicity halo ``subdwarfs"; these stars have high proper motions $\sim 200$ km/sec. We have superimposed a solid line which indicates the theoretical model of Baraffe et al (1995) with log(Z/Z$_\odot) = -1.5$.} \end{figure} The parallax data (Dahn et al 1995) are shown in Figure 2. This is an H-R diagram of nearby stars with measured parallax. The filled circles are high metallicity disk stars. The open circles are red dwarfs which are known to be in the Halo because of their low metallicities and high velocities. It is these 114 Halo stars that we used to get a mass function. We always took the most ``conservative" case, i.e., the steepest MF towards low mass; this case would give the largest number of brown dwarfs and low mass red dwarfs. For this reason, we considered a number of metallicities and used the lowest realistic value of $Z = 3 \times 10^{-4}$. There is a potential complication in that some of the stars in the survey may actually be unresolved binaries. If so, the observed light is the sum of the light from two stars. Then one may overestimate the mass of the star if one assumes the light is from a single star. We considered three models for binaries. The most extreme of these is that all the stars are really in binaries, with equal masses for the two stars in the binary system. Then the luminosity of each star is really half as big as if it had been a single star, each star has a smaller mass, and one obtains a steeper mass function towards low mass. This model is unphysical but simple, and we used it to illustrate an extreme for the largest number of stars at low mass that can be obtained from this data set. Figure 3 shows the mass functions that we obtained, for the case of no binaries and the extreme case of 100\% binaries. In these plots we multiplied the vertical axis by $m^2$ for simplicity of interpretation. With this factor of $m^2$, a mass function (MF) that is decreasing to the left converges, an MF that is increasing to the left diverges, while an MF that is flat diverges only logarithmically. In figure 3a, the case of no binaries, we can see that the MF $\times m^2$ decreases to the left (convergent); in Figure 4c, the case of 100\% binaries, the MF $\times m^2$ is flat (diverges logarithmically). Hence Figure 3 summarizes our results for the mass function for faint stars heavier than $0.09M_\odot$. \begin{figure} \centering \psfig{figure=fig3.ps, width=9cm} \caption{(taken from Graff and Freese 1996b): The mass function of red dwarf halo stars (multiplied by m$^2$). Each of the four models is derived from the LF of Dahn ${\it et \, al}$ (1995) but assumes different metallicity and binary content. In all three panels, crosses without errorbars illustrate the mass function derived for stars with metallicity Z = $3 \times 10^{-4}$ and no binary companions. The other model presented in panel (a) has Z = $6 \times 10^{-4}$ (no binaries) for comparison. Panels (b) and (c) show binary models II and III for $Z = 3 \times 10^{-4}$. Binary model III has been designed to exaggerate the number of low mass stars compared to high mass ones and is unrealistic.} \end{figure} Now, in order to proceed with an extrapolation of this red dwarf mass function past the hydrogen burning limit into the red dwarf regime, we need a brief theoretical interlude. Star formation theory indicates that, as one goes to lower masses, the MF rises no faster than a power law. The theories of Adams and Fatuzzo (1996), Larson (1992), Zinnecker (1984), and Price and Podsiadlowski (1995), while based on different physical principles, all find this same upper limit. Hence we looked for the power law describing the red dwarf mass function at the lowest masses, and then use this same power law to extrapolate into the brown dwarf regime. We took the mass function to scale as \begin{equation} \label{massfunc} MF\propto m^{- \alpha} \, . \end{equation} Then the total mass in the Halo is \begin{equation} \label{mtot} m_{tot} = \int_0^{0.09M_\odot} m \times MF \times dm \, . \end{equation} If $\alpha > 2$, then the total mass diverges. If $\alpha = 2$, then the total mass diverges only logarithmically. If $\alpha <2$, then the total mass converges. We found \begin{equation} \label{alpha} \alpha \leq 2 \, , \end{equation} for all models. More specifically, for the extreme case of 100\% binaries, we found $\alpha = 2$, i.e., each order of magnitude of mass range contains an equal total mass. Even for a lower limit $\sim m_{moon}$, the total mass in brown dwarfs is less than 3\% of the Halo mass. For all other models, including the case of no binaries, we find $\alpha <2$, and brown dwarfs constitute less than a percent of the Halo mass. Similar results were found by Mera, Chabrier, and Schaeffer \cite{mcs}. How might one avoid these conclusions? First, star formation theory might be completely wrong. Alternatively, there might be a spatially varying initial mass function so that brown dwarfs exist only at large radii and not in our locality, so that they were missed in the data (Kerins and Evans \cite{kerev}). The two year MACHO microlensing data have also shown that, for standard Halo models as well as a wide range of alternate models, the timescales fo the events are not compatible with a population of stars lighter than 0.1$M_\odot$ (Gyuk, Evans, and Gates 1998). \subsection{\bf Punchline} The basic result of this work is that the total mass density of local Population II Red Dwarfs and Brown Dwarfs makes up less than 1\% of the local mass density of the Halo; in fact, these objects probably make up less than $0.3\%$ of the Halo. \section{Mass Budget Issues} This section (based on work by Fields, Freese, and Graff \cite{ffg}) is general to all Halo Machos, no matter what kind of objects they are. \subsection{\bf Contribution of Machos to the Mass Density of the Universe} Dalcanton {\it et al. } were able to place strong limits on the cosmological mass density of Machos even before the galactic microlensing experiments produced their first results. They looked for a reduction in apparent equivalent width of quasar emission lines; such a reduction would be caused by compact objects. They found that $\Omega_m < 0.1$. There is a potential problem in that too many baryons are tied up in Machos and their progenitors (Fields, Freese, and Graff \cite{ffg}). We begin by estimating the contribution of Machos to the mass density of the universe: Microlensing results (Alcock {\it et al. } 1997a) predict that the total mass of Machos in the Galactic Halo out to 50 kpc is \begin{equation} \label{outto} M_{\rm Macho} = (1.3 - 3.2) \times 10^{11} M_\odot \, . \end{equation} Now one can obtain a ``Macho-to-light" ratio for the Halo by dividing by the luminosity of the Milky Way (in the B-band), \begin{equation} \label{lummw} L_{MW} \sim (1.3-2.5) \times 10^{10} L_\odot \, . \end{equation} We obtain \begin{equation} \label{masstolight} (M/L)_{\rm Macho} = (5.2-25)M_\odot/L_\odot \, . \end{equation} From the ESO Slice Project Redshift survey (Zucca {\it et al. } \cite{zuc}), the luminosity density of the Universe in the $B$ band is \begin{equation} \label{phi} {\cal L}_B = 1.9\times 10^{8} h \ L_\odot \ {\rm Mpc}^{-3} \end{equation} where the Hubble parameter $h=H_0/(100 \, {\rm km} \, {\rm sec}^{-1} \, {\rm Mpc}^{-1})$. If we assume that the $M/L$ which we defined for the Milky Way is typical of the Universe as a whole, then the universal mass density of Machos is \begin{equation} \label{rho} \rho_{\rm Macho} = (M/L)_{\rm Macho} \, {\cal L}_B = (1-5) \times 10^{9} h \ M_\odot \, {\rm Mpc}^{-3} \, . \end{equation} The corresponding fraction of the critical density $\rho_c \equiv 3H_0^2/8 \pi G = 2.71 \times 10^{11} \, h^2 \, M_\odot \ {\rm Mpc}^{-3}$ is \begin{equation} \label{omega} \Omega_{\rm Macho} \equiv \rho_{\rm Macho}/ \rho_c = (0.0036-0.017) \, h^{-1} \, . \end{equation} Note: see also the discussion by Fukugita, Hogan, and Peebles (\cite{fhp}). We will now proceed to compare our $\Omega_{\rm Macho}$ derived in Eq.\ \pref{omega} with the baryonic density in the universe, $\Omega_{\rm B}$, as determined by primordial nucleosynthesis. Recently, the status of Big Bang nucleosynthesis has been the subject of intense discussion, prompted both by observations of deuterium in high-redshift quasar absorption systems, and also by a more careful examination of consistency and uncertainties in the theory. To conservatively allow for the full range of possibilities, we will therefore adopt \begin{equation} \label{omegab} \Omega_{\rm B}= (0.005-0.022) \ h^{-2} \, . \end{equation} We can see that $\Omega_{\rm Macho}$ and $\Omega_{\rm B}$ are roughly comparable within this na\"i ve calculation. Thus, if the Galactic halo Macho interpretation of the microlensing results is correct, Machos make up an important fraction of the baryonic matter of the Universe. Specifically, the central values in eqs.\ (\ref{omega}) and (\ref{omegab}) give \begin{equation} \label{central} \Omega_{\rm Macho}/\Omega_{\rm B} \sim 0.7 \, . \end{equation} However, the lower limit on this fraction is considerably smaller and hence less restrictive. Taking the lowest possible value for $\Omega_{\rm Macho}$ and the highest possible value for $\Omega_{\rm B}$, we see that \begin{equation} \label{comp} {\Omega_{\rm Macho} \over \Omega_{\rm B}} \geq {1 \over 6} h \geq \frac{1}{12} \, . \end{equation} The only way to avoid these conclusions is to argue that the luminosity density in eqn. (8) is dominated by galaxies without Machos, so that the Milky Way is atypically rich in Machos. However, this is extremely unlikely, because most of the light contributing to the luminosity density ${\cal{L}}$ comes from galaxies similar to ours. Even if Machos only exist in spiral galaxies (2/3 of the galaxies) within one magnitude of the Milky Way, the value of $\Omega_{\rm Macho}$ is lowered by at most a factor of 0.17. \subsection{\bf Comparison with the Lyman-${\bf \alpha}$ Forest} We can compare the Macho contribution to other components of the baryonic matter of the universe. In particular, measurements of the Lyman-$\alpha$ (Ly$\alpha$) forest absorption from intervening gas in the lines of sight to high-redshift QSOs indicate that many, if not most, of the baryons of the universe were in this forest at redshifts $z>$2. It is hard to reconcile the large baryonic abundance estimated for the Ly$\alpha$\ forest with $\Omega_{\rm Macho}$ obtained previously (Gates, Gyuk, Holder, \& Turner \cite{gght}). Although measurements of the Ly$\alpha$\ forest only obtain the neutral column density, careful estimates of the ionizing radiation can be made to obtain rough values for the total baryonic matter, i.e. the sum of the neutral and ionized components, in the Ly$\alpha$\ forest. For the sum of these two components, Weinberg et al. \pcite{wmhk} estimate \begin{equation} \label{lyman} \Omega_{\rm Ly\alpha} \sim 0.02 h^{-3/2} \, . \end{equation} This number is at present uncertain. For example, it assumes an understanding of the UV background responsible for ionizing the IGM, and accurate determination of the quasar flux decrement due to the neutral hydrogen absorbers. Despite these uncertainties, we will use Eq.\ ({\ref{lyman}}) below and examine the implications of this estimate. We can now require that the sum of the Macho energy density plus the Ly$\alpha$\ baryonic energy density do not add up to a value in excess of the baryonic density from nucleosynthesis: \begin{equation} \label{insist} \Omega_{\rm Macho}(z) + \Omega_{{\rm Ly}\alpha} (z) \leq \Omega_{\rm B} \, ; \end{equation} this expression holds for any epoch $z$. Unfortunately, the observations of Machos and Ly$\alpha$\ systems are available for different epochs. Thus, to compare the two one must assume that there has not been a tradeoff of gas into Machos between the era of the Lyman systems ($z \sim 2-3$) and the observation of the Machos at $z=0$. That is, we assume that the Machos were formed before the Ly$\alpha$\ systems. Although Eq.\ (\ref{insist}) offers a potentially strong constraint, in practice the uncertainties in both $\Omega_{{\rm Ly}\alpha}$ and in $\Omega_{\rm B}$ make a quantitative comparison difficult. Nevertheless, we will tentatively use the numbers indicated above. We then have \begin{equation} \label{onehalf} (\Omega_{\rm Macho} = 0.007-0.04) + (\Omega_{{\rm Ly}\alpha} = 0.06) \leq (\Omega_{\rm B} = 0.02 - 0.09) \ {\rm for} \, h=1/2 \, , \end{equation} and \begin{equation} \label{one} (\Omega_{\rm Macho} = 0.004 - 0.02) + (\Omega_{{\rm Ly}\alpha} = 0.02) \leq (\Omega_{\rm B} = 0.005 - 0.02) \ {\rm for} \, h=1 \, . \end{equation} These equations can be satisfied, but only if one uses the most favorable extremes in both $\Omega_{\rm Macho}$ and $\Omega_{\rm B}$, i.e., for the lowest possible values for $\Omega_{\rm Macho}$ and the highest possible values for $\Omega_{\rm B}$. Recent measurements of Kirkman and Tytler \pcite{kt} of the ionized component of a Lyman limit system at z=3.3816 towards QSO HS 1422+2309 estimate an even larger value for the mass density in hot and highly ionized gas in the intergalactic medium: $\Omega_{\rm hot} \sim 10^{-2} h^{-1}$. If this estimate is correct, then Eq.\ (\ref{insist}) becomes even more difficult to satisfy. One way to avoid this mass budget problem would be to argue that the Ly$\alpha$\ baryons later became Machos. Then it would be inappropriate to add the Ly$\alpha$\ plus Macho contributions in comparing with $\Omega_{\rm B}$, since the Machos would be just part of the Ly$\alpha$\ baryons. However, the only way to do this would be to make the Machos at a redshift after the Ly$\alpha$\ measurements were made. Since these measurements extend down to about $z\sim 2-3$, the Machos would have to be made at $z<2$. However, this would be difficult to maneuver. A large, previously unknown population of stellar remnants could not have formed after redshift 2; we would see the light from the stars in galaxy counts (Charlot and Silk \cite{cs}) and in the Hubble Deep Field (Loeb \cite{loeb}). Until now we have only considered the contribution to the baryonic abundance from the Machos themselves. Below we will consider the baryonic abundance of the progenitor stars as well, in the case where the Machos are stellar remnants. When the progenitor baryons are added to the left hand side of Eq.\ (\ref{insist}), this equation becomes harder to satisfy. However, we wish to reiterate that measurements of $\Omega_{{\rm Ly} \alpha}$ are at present uncertain, so that it is possibly premature to conclude that Machos are at odds with the amount of baryons in the Ly$\alpha$ forest. \section{Machos as Stellar Remnants: White Dwarfs or Neutron Stars} In the last section on the mass budget of Machos, we assumed merely that they were baryonic compact objects. In this section (based on work by Fields, Freese, and Graff \cite{ffg}, Fields, Freese, and Graff \cite{ffg2}, and Graff, Freese, Walker, and Pinsonneault \cite{gfwp}): we turn to the specific possibility that Machos are stellar remnants white dwarfs, neutron stars, or black holes. The most complete microlensing data indicate a best fit mass for the Machos of roughly $(0.1 - 1)M_\odot$. Hence there has been particular interest in the possibility that these objects are white dwarfs. I will discuss problems and issues with this interpretation: in particular I will discuss the baryonic mass budget and the pollution due to white dwarf progenitors. \subsection{\bf Mass Budget Constraints from the Macho Progenitors} In general, white dwarfs, neutron stars, or black holes all came from significantly heavier progenitors. Hence, the excess mass left over from the progenitors must be added to the calculation of $\Omega_{\rm Macho}$; the excess mass then leads to stronger constraints. Previously we found that any baryonic Machos that are responsible for the Halo microlensing events must constitute a significant fraction of all the baryons in the universe. Here we show that, if the Machos are white dwarfs or neutron stars, their progenitors, while on the main sequence, are an even larger fraction of the total baryonic content of the universe. The excess mass is then ejected in the form of gas when the progenitors leave the main sequence and become stellar remnants. This excess mass is quite problematic, as there is more of it than is allowed by big band nucleosynthesis and it is chemically enriched beyond what is allowed by observations of Halo stars and the intergalactic medium. If all the Machos formed in $< 1$Gyr (the burst model), then (for different choices of the initial mass function) we can determine the additional contribution of the excess gas to the mass density of the universe. Typically we find the contribution of Macho progenitors to the mass density of the universe to be \begin{equation} \label{omegaprog} \Omega_{{\rm prog}} = 4 \Omega_{{\rm Macho}} = (0.016-0.08)h^{-1} \, . \end{equation} (As an extreme minimum, we find an enhancement factor of 2 rather than 4). From comparison with $\Omega_{\rm B}$, we can see that a very large fraction of the baryons of the universe must be cycled through the Machos and their progenitors. In fact, the central values of all the numbers now imply \begin{equation} \label{toomuch} \Omega_{\rm prog} \sim 3 \Omega_B \, , \end{equation} which is obviously unacceptable. One is driven to the lowest values of $\Omega_{rm Macho}$ and highest value of $\Omega_B$ to avoid this problem. \subsection{\bf Galactic Winds} The white dwarf progenitor stars return most of their mass in their ejecta, i.e., planetary nebulae composed of processed material. Both the mass and the composition of the material are potential problems. As we have emphasized, the cosmic Macho mass budget is a serious issue. Here we see that it is significant even when one considers only the Milky Way. The amount of mass ejected by the progenitors is far in excess of what can be accommodated by the Galaxy. Given the $M_{\rm Macho}$ of Eq.(5), a burst model requires the total mass of progenitors in the Galactic Halo (out to 50 kpc) to have been at least twice the total mass in remnant white dwarfs, i.e., $M_{{\rm prog}} \ge 2 M_{\rm Macho} = (2.4 - 5.8) \times 10^{11} M_\odot$. The gas that is ejected by the Macho progenitors is collisional and tends to fall into the Disk of the Galaxy. But the mass of the ejected gas $M_{\rm gas} = M_{{\rm prog}} - M_{\rm Macho} \sim M_{\rm Macho}$ is at least as large as the mass ($\sim 10^{11} M_\odot$) of the Disk and Spheroid of the Milky Way combined. We see that the Galaxy's baryonic mass budget---including Machos---immediately demands that some of the ejecta be {\it removed} from the Galaxy. This requirement for outflow is intensified when one considers the composition of the stellar ejecta. It will be void of deuterium, and will include large amounts of the nucleosynthesis products of $(1-8) M_\odot$ white dwarf progenitors, notably: helium, carbon, and nitrogen (and possibly s-process material). A possible means of removing these excess baryons is a Galactic wind. Indeed, as pointed out by Fields, Mathews, \& Schramm \pcite{fms}, such a wind may be a virtue, as hot gas containing metals is ubiquitous in the universe, seen in galaxy clusters and groups, and present as an ionized intergalactic medium that dominates the observed neutral Ly$\alpha$\ forest. Thus, it seems mandatory that many galaxies do manage to shed hot, processed material. Such a wind may be driven by some of the white dwarfs themselves (Fields, Freese, and Graff \cite {ffg2}). Some of the white dwarfs may accrete from binary red giant companions and give rise to Type I Supernovae, which serve as an energy source for Galactic winds. However, excess heavy elements such as Fe are overproduced in the process (Canal, Isern, and Ruiz--Lapuente \cite{pilar}). \subsection{\bf On Carbon and Nitrogen} \label{sect:carbon} The issue of carbon (Gibson \& Mould \cite{gm}) and/or nitrogen produced by white dwarf progenitors is the greatest difficulty faced by a white dwarf dark matter scenario. Stellar carbon yields for zero metallicity stars are quite uncertain. Still, according to the Van den Hoek \& Groenewegen (1997) yields, a star of mass 2.5$M_\odot$ will produce about twice the solar enrichment of carbon. If a substantial fraction of all baryons pass through intermediate mass stars, the carbon abundance in this model will be near solar. Then overproduction of carbon can be a serious problem, as emphasized by Gibson \& Mould \pcite{gm}. They noted that stars in our galactic halo have carbon abundance in the range $10^{-4}-10^{-2}$ solar, and argued that the gas which formed these stars cannot have been polluted by the ejecta of a large population of white dwarfs. The galactic winds discussed in the previous section could remove carbon from the star forming regions and mix it throughout the universe. However, carbon abundances in intermediate redshift Ly$\alpha$\ forest lines have recently been measured to be quite low. Carbon is indeed present, but only at the $\sim 10^{-2}$ solar level, (Songaila \& Cowie \cite{sc}) for Ly$\alpha$\ systems at $z \sim 3$ with column densities $N \ge 3 \times 10^{15} \, {\rm cm}^{-2}$. Ly$\alpha$ forest abundances have also been recently measured at low redshifts with HST (Shull {\it et al. } \cite{shull}) to be less than $3 \times 10^{-2}$ solar. Furthermore, in an ensemble average of systems within the redshift interval $2.2 \le z \le 3.6$, with lower column densities ($10^{13.5} \, {\rm cm}^{-2} \le N \le 10^{14} \, {\rm cm}^{-2}$), the mean C/H drops to $\sim 10^{-3.5}$ solar (Lu, Sargent, Barlow, \& Rauch \cite{lsbr}). In order to maintain carbon abundances as low as $10^{-2}$ solar, only about $10^{-2}$ of all baryons can have passed through the intermediate mass stars that were the predecessors of Machos. Such a fraction can barely be accommodated by our results in section 4.1 for the remnant density predicted from our extrapolation of the Macho group results, and would be in conflict with $\Omega_{{\rm prog}}$ in the case of a single burst of star formation. We note that progenitor stars lighter than 4$M_\odot$ overproduce Carbon; whereas progenitor stars heavier than 4$M_\odot$ may replace the carbon overproduction problem with nitrogen overproduction (Fields, Freese, and Graff \cite{ffg2}). The heavier stars may have a process known as Hot Bottom Burning, in which the temperature at the bottom of the star's convective envelope is high enough for nucleosynthesis to take place, and carbon is processed to nitrogen (Lattanzio \cite{latt}, Renzini and Voli \cite {rv}, Van den Hoek and Groenewegen (1997), Lattanzio and Boothroyd \cite{lattbooth}). In this case one gets a ten times solar enrichment of nitrogen, which is far in excess of the observed nitrogen in damped Lyman systems. In conclusion, both C and N exceed what's observed. Using the yields described above, we calculated the C and N that would result from the stellar processing for a variety of initial mass functions for the white dwarf progenitors. We used a chemical evolution model based on a code described in Fields \& Olive \cite{fo98} to obtain our numerical results. The star formation rate is chosen as an exponential $\psi \propto e^{-t/\tau}$ with an $e$-folding time $\tau = 0.1$ Gyr, although we have found that the results are insensitive to details of the star formation rate up to $\tau = 1$ Gyr. Our results are presented in panels b) of figures 4 and 5. The CN abundances are presented relative to solar via the usual notation of the form \begin{equation} [{\rm C/H}]= \log_{10} \frac{{\rm C/H}}{({\rm C/H})_\odot} \, . \end{equation} For example, in this notation $[{\rm C/H}]=0$ represents a solar abundance of C, while $[{\rm C/H}]=-1$is 1/10 solar. Our C and N abundances were obtained without including HBB, which would exchange a C overproduction problem for a N overproduction problem. \begin{figure} \centering \psfig{figure=fig4.ps, width=9cm} \caption{(taken from Fields, Freese, and Graff 1999): {\bf (a)} The D/H abundances and helium mass fraction $Y$ for models with $\Omega_{\rm WD} h = 6.1 \times 10^{-4}$, $h=0.7$, and IMF peaked at $2M_\odot$. The red curves show the changes in primordial D and He and a result of white dwarf production. The solid red curve is for the full chemical evolution model, the dotted red curve is for instantaneous recycling, and the long-dashed red curve for the burst model. The short-dashed blue curve shows the initial abundances; the error bars show the range of D and He measurements. This is the absolute minimum $\Omega_{\rm WD}$ compatible with cosmic extrapolation of white dwarf Machos if Machos are contained only in spiral galaxies with luminosities similar to the Milky Way. \hfill\break {\bf (b)} CNO abundances produced in the same model as {\bf a}, here plotted as a function of $\Omega_B$. The C and N production in particular are greater than 1/10 solar.} \end{figure} In Figure 4, we make the parameter choices that are in agreement with D and He$^4$ measurements (see the discussion below) and are the least restrictive when comparing with the Ly$\alpha$ measurements. We take $\Omega_{\rm WD} h = 6.1 \times 10^{-4}$, the minimum amount of WD required to explain the microlensing results if only galaxies similar to ours produce WD Machos. We take h= 0.7 and an initial mass function (IMF) sharply peaked at 2$M_\odot$, so that there are very few progenitor stars heavier than 3$M_\odot$ (this IMF is required by D and He$^4$ measurements). In Figure 5, we have the same values for $\Omega_{\rm WD}$ and h but have taken an IMF peaked at 4$M_\odot$. In both cases, by comparing with the observations, we obtain the limit, \begin{equation} \Omega_{\rm WD} h \leq 2 \times 10^{-4} \, . \end{equation} We also considered a variety of other parameter choices, and obtained the same limit (see the figures in Fields, Freese, and Graff 1999). \begin{figure} \centering \psfig{figure=fig5.ps, width=9cm} \caption{As in Fig. 4, but with an IMF peaked at 4$M_\odot.$ We see that the processing drives D and He out of the measured range.} \end{figure} Alternatively, we require an actual abundance distribution that is quite heterogeneous: those regions in which the observations are made must be underprocessed. This implies segregation efficiency of 97\%. Note that it is possible (although not likely) that carbon never leaves the white dwarf progenitors, so that carbon overproduction is not a problem (Chabrier \cite{chabriernew}). Carbon is produced exclusively in the stellar core. In order to be ejected, carbon must convect to the outer layers in the ``dredge up'' process. Since convection is less efficient in a zero metallicity star, it is possible that no carbon would be ejected in a primordial star. In that case, it would be impossible to place limits on the density of white dwarfs using carbon abundances. We have here assumed that carbon does leave the white dwarf progenitor stars. \subsection{\bf Deuterium and Helium} Because of the uncertainty in the C and N yields from low-metallicity stars, we have also calculated the D and He$^4$ abundances that would be produced by white dwarf progenitors. These are far less uncertain as they are produced farther out from the center of the star and do not have to be dredged up from the core. We use both the numerical model discussed above, in which the stars have finite lifetimes, and also two extreme analytical models to bracket the possible results. We consider a burst model, in which the timescale for star formation is much shorter than the lifetimes of the stars. We also consider the opposite limit, the instantaneous recycling approximation, in which the stellar lifetime is short compared to the star formation timescale. In the figures we present results from both analytical approaches and from the numerical model; we can see that the numerical results are closely approximated by the burst model. Panels a) in Fig. 4 and 5 display our results. Also shown are the initial values from big bang nucleosynthesis and the (very generous) range of primordial values of D and He$^4$ from observations. One can see right away that Fig. 4 obtains abundances compatible with the measurements, while the model in Fig. 5 fails to match the D and He$^4$ measurements. Thus, from D and He alone, we can see that the white dwarf progenitor IMF must be peaked at low masses, $\sim 2M_\odot$. We obtain \begin{equation} \Omega_{\rm WD} \leq 0.003 \, . \end{equation} \subsection{\bf Background Light} If galactic halos contain stellar remnants, the infrared flux from the remnant progenitors would contribute to the opacity of multi-TeV $\gamma$-rays (Konopelko {\it et al. } \cite{konop}). The multi-TeV $\gamma$-ray horizon is established to be at a redshift $z>0.034$ by the observation of the blazar Mkn501. By requiring that the optical depth due to $\gamma \gamma \rightarrow e^+ e^-$ be less than one for a source at $z=0.034$ we limit the cosmological density of stellar remnants (Graff, Freese, Walker, and Pinsonneault \cite{gfwp}), \begin{equation} \Omega_{\rm WD} \leq (1-3) \times 10^{-3} h^{-1} \, . \end{equation} In other words, if the density of white dwarfs exceeds this value, the infrared radiation from the progenitors would have prevented TeV $\gamma$-rays from Mkn501 from ever reaching us. \subsection{\bf Neutron Stars} The first issue raised by neutron star Macho candidates is their compatibility with the microlensing results. Neutron stars ($\sim 1.5 M_\odot$) and stellar black holes ($\mathrel{\mathpalette\fun >} 1.5 M_\odot$) are more massive objects, so that one would typically expect longer lensing timescales than what is currently observed in the microlensing experiments (best fit to $\sim 0.5 M_\odot$). As discussed by Venkatesan, Olinto, \& Truran \pcite{vot}, one must posit that as the experiments continue to take measurements, longer timescale events should begin to be seen. In this regard, it is intriguing that the first SMC results (Palanque-Delabrouille et al.\ \cite{eros:smc}; Alcock et al.\ \cite{macho:smc}) suggest lensing masses of order $\sim 2 M_\odot$. Note that these long timescales could be explained if the SMC events are due to SMC self lensing (Palanque-Delabrouille {\it et al. } \cite{eros:smc}; Graff \& Gardiner \cite{gg}). However, the same issues of mass budget and chemical enrichment arise for neutron stars as did for white dwarfs, only the problems are worse. In particular, the higher mass progenitors of neutron stars eject even more mass, so that $\Omega_{\rm prog}$ is even bigger than for the case of white dwarfs. The ejecta are highly metal rich and would need a great deal of dilution (as much as for the case of white dwarfs) in order to avoid conflict with observations. However, most of the baryons in the universe have already been used to make the progenitors (even more than for the case of white dwarfs); there are no baryons left over to do the diluting. \subsection{\bf Mass Budget Summary} If Machos are indeed found in halos of galaxies like our own, we have found that the cosmological mass budget for Machos requires $\Omega_{\rm Macho}/ \Omega_{\rm B} \geq {1 \over 6} h f_{\rm gal}$, where $f_{\rm gal}$ is the fraction of galaxies that contain Machos, and quite possibly $\Omega_{\rm Macho} \approx \Omega_{\rm B}$. Specifically, the central values in eqs.\ (\ref{omega}) and (\ref{omegab}) give $\Omega_{\rm Macho}/\Omega_{\rm B} \sim 0.7$. Thus a stellar explanation of the microlensing events requires that a significant fraction of baryons cycled through Machos and their progenitors. If the Machos are white dwarfs that arose from a single burst of star formation, we have found that the contribution of the progenitors to the mass density of the universe is at least a factor of two higher, probably more like three or four. We have made a comparison of $\Omega_{\rm B}$ with the combined baryonic component of $\Omega_{\rm Macho}$ and the baryons in the Ly$\alpha$\ forest, and found that the values can be compatible only for the extreme values of the parameters. However, measurements of $\Omega_{{\rm Ly} \alpha}$ are at present uncertain, so that it is perhaps premature to imply that Machos are at odds with the amount of baryons in the Ly$\alpha$ forest. In addition, we have stressed the difficulty in reconciling the Macho mass budget with the accompanying carbon and/or nitrogen production in the case of white dwarfs. The overproduction of carbon or nitrogen by the white dwarf progenitors can be diluted in principle, but this dilution would require even more baryons that have not gone into stars. At least in the simplest scenario, in order not to conflict with the upper bounds on $\Omega_{\rm B}$, this would require an $\Omega_{\rm Macho}$ slightly smaller than our lower limits from extrapolating the Macho results. Only 10$^{-2}$ of all baryons can have passed through the white dwarf progenitors, a fraction that is in conflict with our results for $\Omega_{{\rm prog}}$. \section{Zero Macho Halo?} The possibility exists that the 14 microlensing events that have been interpreted as being in the Halo of the Galaxy are in fact due to some other lensing population. One of the most difficult aspects of microlensing is the degeneracy of the interpretation of the data, so that it is currently impossible to determine whether the lenses lie in the Galactic Halo, or in the Disk of the Milky Way, or in the LMC. Evans {\it et al. } (\cite{evans:diskflare}) proposed that the events could be due to lenses in our own Milky Way Disk. Gould (\cite{gouldy}) showed that the standard model of the LMC does not allow for significant microlensing. Zhao (1998) has proposed that debris lying in a tidal tail stripped from the progenitor of the LMC or SMC by the Milky Way or by an SMC-LMC tidal interaction may explain the observed microlensing rate towards the LMC. Within this general framework, he suggests that the debris thrown off by the tidal interaction could also lead to a high optical depth for the LMC. There have been several observational attempts to search for this debris. Zaritsky \& Lin (1997) report a possible detection of such debris in observations of red clump stars, but the results of further variable star searches by the {\sc macho} group (Alcock {\it et al. } 1997b), and examination of the surface brightness contours of the LMC (Gould 1998) showed that there is no evidence for such a population. A stellar evolutionary explanation for the observations of Zaritsky \& Lin (1997) was proposed by Beaulieu \& Sackett (1998). However, possible evidence for debris within a few kpc of the LMC along the line of sight is reported by the {\sc eros} group (Graff {\it et al. } in preparation). These issues are currently unclear and are under investigation by many groups. Note that a recent microlensing event towards the SMC, MACHO-98-SMC-1, was due to a binary lens. In this case it was possible to clearly identify that the lens is in the SMC and not in our Halo (Albrow \cite{planet}). Parallax analysis of event MACHO-97-SMC-1 shows that this event also is likely to be in the SMC (Palanque-Delabrouille {\it et al. } 1998). Analysis of the binary lensing event MACHO-LMC-9 shows that this event lies in the LMC. So far, all the events which can be located lie in the Magellanic Clouds. However, the cause of the remaining events of the LMC remains ambiguous and awaits further observations. \section{Conclusions} Microlensing experiments have ruled out a large class of possible baryonic dark matter components. Substellar objects in the mass range $10^{-7}M_\odot$ all the way up to $10^{-2}M_\odot$ are ruled out by the experiments. In this talk I discussed the heavier possibilities in the range $10^{-2}M_\odot$ to a few $M_\odot$. I showed that brown dwarfs and faint stars are ruled out as significant dark matter components; they contribute no more than 1\% of the Halo mass density. White dwarfs and neutron stars are also extremely problematic. The chemical abundance constraints are formidable. The D and He$^4$ production by the progenitors of the white dwarfs can be in agreement with observation for low $\Omega_{WD}$ and an IMF sharply peaked at low masses $\sim 2M_\odot$. Unless carbon is never dredged up from the stellar core (as has been suggested by Chabrier \cite{chabriernew}), overproduction of carbon and/or nitrogen is problematic. The relative amounts of these elements that is produced depends on Hot Bottom Burning, but both elements are produced at the level of at least solar enrichment, in conflict with what is seen in our Halo and in Ly$\alpha$ systems. One must either abandon stellar remnants as dark matter or argue that the debris have remained hot and segregated from cooler neutral matter. However, the observations of TeV $\gamma$-rays from Mkn501 at z=0.034 restrict the infrared background of the universe and hence the white dwarf progenitors that would have produced infrared light. In sum, we have a constraint on the remnant density, $\Omega_{\rm WD} \leq (1-3) \times 10^{-3} h^{-1}$. Hence, in conclusion, \begin{description} \item{1.} Nonbaryonic dark matter in our Galaxy seems to be required, and \item{2.} The nature of the Machos seen in microlensing experiments and interpreted as the dark matter in the Halo of our Galaxy remains a mystery. Are we driven to primordial black holes (Carr 1994; Jedamzik \cite{jedam}), nonbaryonic Machos (Machismos?), mirror Machos (Mohapatra and Teplitz \cite{mohap}) or perhaps a no-Macho Halo? \end{description} \section{Acknowledgments} We are grateful for the hospitality of the Aspen Center for Physics, where part of this work was done. DG acknowledges the financial support of the French Ministry of Foreign Affairs' Bourse Chateaubriand. KF acknowledges support from the DOE at the University of Michigan. The work of BDF was supported in part by DOE grant DE-FG02-94ER-40823. \section{References}
\section{Introduction} Chiral anomalies in quantum field theory appear in several different forms. Historically they were first observed in perturbative calculations of certain 1-loop scattering amplitudes, as a breaking of the (classically valid) chiral symmetry, \cite{ABJ}. Later nonperturbative methods were developed for understanding the chiral symmetry breaking in the euclidean path integral formalism, \cite{NRS},\cite{AW}. It was also understood that even the symmetry under coordinate transformations could be broken when quantizing massless fermions. In the hamiltonian approach to chiral anomalies one considers the equal time commutation relations for the infinitesimal generators of the classical symmetry group. First one constructs the bundle of fermionic Fock spaces parametrized by various external (classical) fields: gauge potentials, metrics, scalar potentials etc. The quantization of the algebra of currents in the Fock spaces requires some renormalization procedure. In $1+1$ space-time dimensions usually a normal ordering is sufficient but in higher dimensions certain additional subtractions are needed. Typically the renormalization modifies the algebra of classical symmetries by the so-called Schwinger terms, \cite{S}. In $1+1$ dimensions the Schwinger term is normally just a c-number, leading to an affine Lie algebra (gauge transformations) or to the Virasoro algebra (diffeomorphisms). In higher dimensions the algebra is more complicated; instead of a central (c-number) extension the Schwinger terms lead to an extension by an abelian ideal \cite{FM}. Direct analytic computations of the Schwinger terms in higher dimensions, although they can be carried out \cite{M1} in the Yang-Mills case, are very complicated in the case of an external gravitational field. However, there are topological and geometrical methods which give directly the structure of the quantized current algebra. As in the case of euclidean path integral anomalies, a central ingredient in this discussion is the families index theorem. In a previous paper \cite{CMM} (see also \cite{CMM2} for a review) it was shown how the Schwinger terms in the gauge current algebra are related, via Atiyah-Patodi-Singer index theory, to the structure of a system of local determinant line bundles in odd dimensions. This system provides an example of a mathematical structure called a bundle gerbe, \cite{Mu}. In the present paper we want to extend the methods of \cite{CMM} for constructing the Schwinger terms which arise in fermionic Fock space quantization of the algebra of vector fields on an odd dimensional manifold. In section 2 we set up the notation and recall some basic results in the families index theory in case of compact manifolds with boundaries. In section 3 we compute the curvature forms for a local system of complex line bundles over a parameter space $B$ which consists of gauge potentials and Riemannian metrics. In section 4 we explain how the Schwinger terms in the Fock space quantization of vector fields (and infinitesimal gauge transformations) are obtained from the local curvature formulas. Finally, in section 5 we give some results of explicit computations of the Schwinger terms. \section{The family index theorem} Let $\tilde{\pi }:\tilde{M}\rightarrow B$ be a smooth fibre bundle with fibres diffeomorphic to a compact oriented spin manifold $M$ of even dimension $2n$. Assume that each fibre $M_z=\tilde{\pi }^{-1}(z)$, $z\in B,$ is equipped with a Riemannian metric. Assume further that $\tilde{M}$ is equipped with a connection. This means that at each point $x\in \tilde{M}$, the tangent space splits in a horizontal and a vertical part: $T_x\tilde{M}= H_x \oplus V_x,$ where $V_x$ consists of vectors tangential to the fibers. Let $\tilde{{\cal E} }$ be a vector bundle over $\tilde M$ which along each fiber $M_z$ is the tensor product ${\cal E}_z$ of the Dirac spinor bundle over $M_z$ (with Clifford action of the vertical vectors $V_x$) and a finite dimensional vector bundle $W_z$ (with trivial Clifford multiplication). It has a $\bf Z_2$ structure provided by the chirality operator $\Gamma$ according to: $c(\Gamma )=\pm 1$ on $\tilde{{\cal E} } ^{\pm }$, where $c$ denotes the Clifford action. $\tilde{{\cal E} }$ is assumed to be equipped with a hermitean fiber metric. This naturally induces an $L^2$-metric in the space of sections $\Gamma (M_z,{\cal E}_z)$. Finally, let $D$ be an operator on $\Gamma (\tilde{M},\tilde{\cal E})$ which is fiberwise defined as a family of Dirac operators $D_z : \Gamma (M_z,{\cal E}_z)\rightarrow\Gamma (M_z,{\cal E}_z)$ with $z\in B.$ With a Dirac operator we mean an operator that can be written as the sum of compositions of a covariant derivative and a Clifford multiplication. We would now like to apply the family index theorem to the case described above. Before doing this, some additional assumptions is needed. These are of different nature depending on if $M$ has a non-empty boundary or not. When the boundary is empty, it will only be assumed that the set $\{ D_z\} _{z\in B}$ should consist of self-adjoint Dirac operators. The assumptions in the more difficult case of a non-empty boundary $\partial M$ will now be described. We make the common simplifying assumption that for all $z\in B$, there exists a collar neighbourhood of $\partial M_z$ such that all structures in ${\cal E }_z$ are of 'product type\rq $ $ (see ref. \cite{APS}). This implies that $D_z^{+}=D|_{\Gamma (M_z,{\cal E} ^+_z)}$ can be written as $c_t(\frac{\partial}{\partial t}+D_z^{\partial })$ near the boundary, where $c_t$ is the Clifford multiplication by an element corresponding to the coordinate vector field $\frac{\partial}{\partial t}$ (which is along the unit inward normal vector field at the boundary) and $D_z^{\partial}$ is a self-adjoint Dirac operator on ${\tilde{\cal E } ^+}|_{\partial M_z}$. Our conventions are such that $c_t$ is unitary. Let $\mbox{Ind}{D}_{\lambda }$, be the family $\{\mbox{Ind}D_{z,\lambda }\} _{z\in U_{\lambda }}$, where $\mbox{Ind}{D}=\mbox{ker} D^+\ominus \mbox{ker} D^-$ is the index bundle in the sense of K-theory and $U_{\lambda }=\{ z\in B;\lambda \notin \mbox{spec} \left( D_z^{\partial }\right)\}$. The notation means that every operator $D_z^{+}$ will be restricted to the domain $\{ \psi \in \Gamma ( M_z,{\cal E } ^+_z; P_{z,\lambda }\psi |_{\partial M_z}=0\}$, while $D_z^{-}$ will be restricted to $\{ \psi \in \Gamma ( M_z,{\cal E } ^-_z ; (1-P_{z,\lambda }) c_t \psi |_{\partial M_z}=0\}$, where $ P_{z,\lambda }$ is the spectral projection of $D_z^{\partial }$ corresponding to eigenvalues $\geq \lambda $. We will assume that the Dirac operators $D_{z,\lambda }$ are self-adjoint. The family index theorem has been proven in \cite{AS2} when $\partial M=\emptyset$ and in \cite{BC}, based on \cite{APS}, when $\partial M\neq\emptyset$. It reads: \begin{eqnarray} \label{eq:FIT} \mbox{ch}\left( \mbox{Ind}{D}\right) (z) & = & \int _{M_z} \hat{A}(M_z)\mbox{ch}(W_z), \quad\partial M=\emptyset\nonumber \\ \nopagebreak \mbox{ch}\left( \mbox{Ind}D_{\lambda }\right) (z) & = & \int _{M_z} \hat{A}(M_z)\mbox{ch}(W_z) - \frac{1}{2}\tilde{\eta }_{\lambda }(z),\,\,\, z\in U_{\lambda },\partial M\neq\emptyset, \nonumber \end{eqnarray} where \begin{eqnarray} \label{eq:AR} \hat{A}(M_z) & = & \mbox{det}^{1/2}\left( \frac{iR_z/4\pi } {\sinh (iR_z/4\pi)}\right) \nonumber \\ \nopagebreak \mbox{ch}(W_z) & = & \mbox{tr exp}\left( iF_z/2\pi\right) . \end{eqnarray} We choose to not write down the definitions of the form $\tilde{\eta }_{\lambda }$ or the curvature 2-forms $\tilde{R}$ and $\tilde{F}$ in $\tilde{M}$, where $R_z=\tilde{R}|_{M_z} $ and $F_z=\tilde{F}|_{M_z}$, since we will only need their explicit expression in a simple special case. The only thing about the $\tilde\eta$ form we need to know later is that it depends only on the boundary spectral data. The zero degree part of $\tilde \eta$ is just the $\eta$-invariant of the boundary Dirac operator. For $M=\emptyset$, the determinant line bundle DET is an object closely related with the index bundle $\mbox{Ind}{D}$. It is a line bundle over $B$, fibre-wise defined as $(\mbox{det ker} D_z^{+})^{\ast }\otimes\mbox{det ker }D_z^{-}$. To define it globally over $B$ we must also account for that the dimension of $\mbox{ker}D_z^{+}$ and $\mbox{ker}D_z^{-}$ can jump as $z$ varies. For a detailed construction, see \cite{BF}. For $M\neq\emptyset $ there exist a similar construction of a bundle $\mbox{DET}_{\lambda }$ over $U_{\lambda }$, closely related to $\mbox{Ind}D_{\lambda }$, see ref. \cite{PZ}. In \cite{BF} and \cite{PZ} it has been shown that for $\partial M=\emptyset$ and $\partial M\neq\emptyset$ there exists a connection on DET and $\mbox{DET}_{\lambda }$, respectively, naturally associated with the Quillen metric, \cite{Q1}, with curvature given by \begin{eqnarray} \label{eq:FD} \frac{i}{2\pi} F^{\mbox{\footnotesize{DET}}}(z) & = & \left(\int _{M_z} \hat{A}(M_z)\mbox{ch}(W_z) \right) _{[2]}, \quad \partial M=\emptyset \nonumber \\ \nopagebreak \frac{i}{2\pi}F^{\mbox{\footnotesize{DET}}_{\lambda }}(z) & = & \left(\int _{M_z} \hat{A}(M_z)\mbox{ch}(W_z) - \frac{1}{2}\tilde{\eta }_{\lambda }(z)\right) _{[2]},\nonumber \\ \nopagebreak && z\in U_{\lambda },\partial M\neq\emptyset , \end{eqnarray} where $\left[ 2\right]$ denotes the part that is a 2-form. Notice that the 2-form part of the right hand side of the family index theorem, eq. \Ref{eq:FIT}, is equal to the right hand side of eq. \Ref{eq:FD}. In the case of an odd dimensional manifold $N$ one can produce in a similar way an element $\Omega\in H^3(B, \bf Z),$ by an integration over the fibers $N_z,$ \begin{eqnarray} \label{eq:FDD} \Omega (z) & = & \left(\int _{N_z} \hat{A}(N_z)\mbox{ch}(W_z) \right) _{[3]}, \quad \partial N=\emptyset \end{eqnarray} where this time we pick up the component of the form which is of degree $3$ in the tangential directions on the parameter space $B.$ This form plays an important role in the hamiltonian quantization of external field problems, \cite{CMM}. It is the Dixmier-Douady class of a gerbe. A nonvanishing Dixmier-Douady class is an obstruction to quantizing chiral fermions in a gauge invariant manner. \section{Local line bundles over boundary \\ geometries} Let $P$ be a fixed principal $G$ bundle over $M$ with a projection $\pi:P \to M$ and $FM$ the oriented frame bundle of $M.$ The bundle $FM$ is a also a principal bundle, with the structure group $GL_+(2n,\bf R),$ the group of real $2n\times 2n$ matrices with positive determinant. Let $Q$ denote the product bundle $P\times FM.$ Let $B=\cal A \times \cal M$ where $\cal A$ is the affine space of connections on $P$ and $\cal M$ is the space of Riemannian metrics on $M.$ Locally, an element of $\cal A$ is written as a Lie$(G)$ valued 1-form on $M.$ We may view $Q$ as a principal bundle $\tilde Q$ over $\tilde{M}=M\times B$ in a natural way, as the pull-back under the projection $M\times B\to M.$ With notations as in the previous section we define $M_z$ as the manifold $M$ with metric given by $z\in B$. Along the model fiber $M,$ let $\cal E$ be the tensor product of the Dirac spinor bundle and a vector bundle $W$ over $M,$ the latter being an associated bundle to $P(M,G).$ We view $\cal E$ in a natural way as a vector bundle $\tilde{\cal E}$ over $M\times B.$ Finally, we let $D_z:\Gamma (M_z,{\cal E }_z)\rightarrow \Gamma (M_z,{\cal E }_z)$ be the Dirac operator constructed from $z\in B$ in the usual way; in terms of a local coordinates $A=A_{\mu} dx^{\mu}, \Gamma= \Gamma_{\mu} dx^{\mu},$ and with respect to a local orthonormal frame $\{e_{a}\}_{a =1}^{2n}$ of $TM_z$ we have $$ D_{(A,\Gamma)}=\sum _{a,\mu =1}^{2n}\gamma ^{a}{e_a}^{\mu} \left( \partial _{\mu }+ A_{\mu}+ \Gamma _{\mu }\right) , $$ where the ${e_a}^{\mu}$'s are the components of the basis vectors $e_a$ in the coordinate frame and $\gamma^a$ is the Clifford multiplication by $e_a,$ with $\gamma^a\gamma^b +\gamma^b\gamma^a = 2\delta^{ab}.$ Let $\cal D$ be the group of orientation preserving diffeomorphims of $M$ and $\cal G$ the group of gauge transformations in $P,$ i.e., the group of automorphims of $P(M,G)$ which projects to the identity map on the base. The groups $\cal G$ and $\cal D$ act through pull-backs on $\cal A$ and $\cal M.$ The group actions induce a fiber structure in $B=\cal A \times \cal M$ but in order to obtain smooth moduli spaces we restrict to the subgroups ${\cal G}_0\subset \cal G$ and ${\cal D}_0 \subset \cal D.$ The former is the group of based gauge transformations $\phi$ leaving invariant some fixed base point $p_0\in P,$ $\phi(p_0) =p_0.$ The group ${\cal D}_0$ is defined as $${\cal D}_0 =\{\phi\in{\cal D} | \phi(x_0) =x_0 \mbox{ and } T_{x_0}\phi= \mbox{id} \}$$ for some fixed $x_0\in M.$ With these choices, we obtain the smooth fiber bundles ${\cal M}\to {\cal M}/{\cal D}_0$ and ${\cal A} \to {\cal A}/ {\cal G}_0.$ This leads also to a fibering $B={\cal A}\times {\cal M} \to ({\cal A}\times {\cal M})/({\cal D}_0\ltimes {\cal G}_0).$ Note that the group of symmetries is the semidirect product of ${\cal D}_0$ and ${\cal G}_0$ since locally an element of ${\cal G}_0$ is a $G$-valued function on $M$ and the diffeomorphisms act on the argument of the function. Following \cite{AS3} we have a connection form $\omega$ on $\tilde Q=P\times FM\times B\to M\times B$ which can be pushed forward to a connection form on ${\tilde Q}/({\cal D}_0\ltimes {\cal G}_0).$ Along $P\times FM$ the form $\omega$ is given by a connection $A\in \cal A$ and by the Levi-Civita connection given by a metric $g\in \cal M.$ Restricted to the second factor $B=\cal A\times \cal M$ the form $\omega$ is called the BRST ghost and will be denoted by $v.$ Since the total form $\omega$ should vanish along gauge and diffeomorphism directions it follows that its value along these directions in $B$ is uniquely determined by the value of the corresponding vector field on $P\times FM.$ In the case of gauge potentials, $B=\cal A,$ an infinitesimal gauge transformation is given locally as a Lie$(G)$ valued function $Z$ on $M$ and then $v_{p,A}(Z_A)= Z(x)$ where $x=\pi(p)$ and $Z_A=\delta_Z A$ is the vector field on $\cal A$ defined by the infinitesimal gauge transformation $Z.$ In the case of diffeomorphims, $Z$ is the ${\bf gl}(2n,\bf R)$ valued function defined as the Jacobian (in local coordinates) of a vector field on $M.$ Again, $v(Z_{\Gamma})$ is the 'tautological 1-form', $Z$ evaluated at a point $x\in M.$ The ghost $v$ in other directions is a nonlocal expression involving the Green's function of a gauge covariant Laplacian, \cite{AS3, Do}, but we shall make explicit use of $v$ only in the gauge and diffeomorphims directions. Next let $N$ be an odd dimensional manifold without boundary and let $M=[0,1] \times N,$ dim$M=2n.$ Given a principal bundle $P$ over $N$ we can extend it to a principal bundle over $M$ (to be denoted by the same symbol $P$) by a pull-back defined by the projection $[0,1]\times N \to N.$ We choose a fixed connection $A_0$ in the principal bundle $P$ over $N$ and a fixed metric $g_0$ in $N.$ If $A$ is an arbitrary connection in $P$ we form a connection $A(t)$ (with $t\in [0,1]$) in the principal bundle $P$ over $M$ by $$A(t) = (1-f(t)) A_0 + f(t) A,$$ where $f$ is a fixed smooth real valued function such that $f(0)=0, f(1)=1,$ and all the derivatives of $f$ vanish at the end points $t=0,1.$ Similarly, any metric $g$ in $N$ defines a metric in $M$ such that along $N$ directions it is given by $$g_{ij}(t) = (1-f(t)) (g_0)_{ij} + f(t) g_{ij}$$ and such that $\partial_t$ is a normalized vector field in $t$ direction, orthogonal to the $N$ directions. However, for computations below it is more convenient to use directly a homotopy connecting the the Levi-Civita connection $\Gamma$ (constructed from the metric $g$) to the connection $\Gamma_0$ (constructed from $g_0$). The formula for this homotopy is the same as for the gauge potentials above. We use the 'Russian formula', which is just an expression of the fact that in a principal bundle with a connection the curvature form has \it no components along fiber directions. \rm The formula tells that when the total curvature on $P\times FM\times{\cal A}\times {\cal M}$ is evaluated along vertical directions in ${\cal A}\times {\cal M} \to ({\cal A}\times {\cal M})/ ({\cal G}_0 \times {\cal D}_0)$ and along vector fields on $P\times FM$ the result is $$ F^{\omega} = F^{A,\Gamma} = dA +\frac12 [A,A] + d\Gamma + \frac12 [\Gamma, \Gamma],$$ where $\Gamma$ is the Levi-Civita connection. Next we replace $\omega = A+\Gamma +v$ by the 'time' dependent connection $$\omega(t) = (1-f(t)) (A_0 +\Gamma_0) +f(t) (A+\Gamma +v).$$ An evaluation of the curvature of a Dirac determinant bundle over $\cal A \times \cal M$ involves an integration of a characteristic class $p_{n+1} (F^{\omega(t)})$ over $M=[0,1]\times N$ and an evaluation of the $\tilde\eta$ form on the boundary. Here $p_i$ denotes a generic homogeneous symmetric invariant polynomial of degree $i$ in the curvature. Actually we have to restrict the construction of the determinant bundle to subsets $U_{\lambda}$ in the parameter space $B$ on the boundary. Here $U_{\lambda}$ is again the set of those points in $B$ such that the associated Dirac operator on $\partial M$ does not have the eigenvalue $\lambda.$ These sets form an open cover of $B.$ In each of the sets $U_{\lambda}$ one can define the $\tilde\eta$ form associated to Dirac operators $D_z -\lambda$ as a continuous function of the parameters $z\in B.$ We shall restrict to the problem of determining the curvature along gauge and diffeomorphism directions on the boundary. The $\tilde\eta$ form is a spectral invariant and therefore the only term which contributes is the appropriate characteristic class in the bulk $M.$ The integration of the index density in the bulk can be performed in two steps. First one integrates over the time variable $t$ and then the resulting expression is integrated over $N$ to produce a 2-form on ${\cal A}(N) \times {\cal M}(N).$ All the computations involving the ghost $v$ are restricted to vertical directions. Restricting to the case of gauge potentials (calculations involving the Levi-Civita connections are performed in the same way) we have \begin{eqnarray} F_{ij}^{\omega(t)} &= &\partial_i A_j(t) -\partial_j A_i(t) +[A_i(t), A_j(t)]\nonumber \\ \nopagebreak F_{0i}^{\omega(t)} &= &f'(t) (A -A_0)_i,\nonumber \end{eqnarray} where we use the index $0$ for the $t$ component. For intermediate times $0<t<1$ the curvature has components also to the vertical directions, \begin{eqnarray} (F^{\omega(t)})_{[0,2]}&= &\frac12 f(f-1)[v,v] \nonumber \\ \nopagebreak (F^{\omega(t)})_{[1,1]}&=& v f'(t) dt + f(f -1) [A-A_0, v].\nonumber\end{eqnarray} Here we have denoted by $(F)_{[i,j]}$ the component of a form $F$ which is of degree $i$ in the tangential directions in $P$ and of degree $j$ in the ghost $v.$ If $p_k$ is any homogeneous symmetric function of degree $k$ of the curvature we set $p_k(A,F) = p_k(A,F,F,\dots,F)$ and then \begin{eqnarray} \label{eq:POM} \int_M p_k(F^{\omega (t)})&=& k\int_N \int_{0}^{1} p_k(f'(t)(A+v-A_0), F^{\omega(t)}) dt \nonumber \\ \nopagebreak &\equiv &\int_N \omega_{2k-1}(A+v,A_0).\end{eqnarray} The form on the right, when expanded in powers of the ghost $v$, gives forms of various degrees on the parameter space $B.$ We are interested in the curvature form which is of degree 2 in $v.$ The degree zero part gives just the Chern-Simons form $\omega_{2k-1}(A,A_0)$ and if $N$ were an even dimensional manifold the degree 1 term would be the nonabelian gauge anomaly. In low dimensions one gets familiar explicit formulas; as an example, consider the case of a trivial bundle and $A_0 =0.$ When $n=1$ the relevant characteristic class is $p_2(F)=\frac{1}{2!} (\frac{i}{2\pi})^2 \mbox{tr}\, F^2$ and the curvature $c_1$ along vector fields given by a pair $X,Y$ of infinitesimal gauge transformations on the one-dimensional manifold $N$ is $$\frac{i}{2\pi}c_1(X,Y)= \frac{1}{8\pi^2} \int_N \mbox{tr} \, A[X,Y]$$ and in the case $n=2,$ dim$ N=3,$ $p_3(F)=\frac{1}{3!} (\frac{i}{2\pi})^3 \mbox{tr}\, F^3$, one gets \begin{eqnarray} \frac{i}{2\pi} c_3(X,Y) & = & \frac{1}{48\pi^3} \int_N \mbox{tr} \, \Big( (AdA +dA \,A + A^3)[X,Y] \nonumber \\ \nopagebreak &&+XdA\, YA -YdA\,XA\Big) .\nonumber\end{eqnarray} The case of Levi-Civita connection $\Gamma$ needs some extra remarks. The reason is that we have actually two types of Chern-Simons forms (and associated anomaly forms) depending whether we write the connection with respect to a (local) orthonormal frame in the tangent bundle $TM$ or with respect to the holonomic frame given by coordinate vector fields. Formally, the two Chern-Simons forms (and associated polynomials in $v$) look the same; they are given exactly by the same differential polynomials in $\Gamma_{\mu\nu}^{\lambda}$ (coordinate basis) or in $\Gamma_{\mu a}^b$ (with respect to an anholonomic basis $e_a^{\mu}$). The difference is (locally) an exterior derivative of a form (in $N$) of lower degree. The difference $\Delta \omega$ of the Chern-Simons forms involves the matrix function $e_a^{\mu}(x)$ on $N.$ Since this function takes values in the group $GL(2n-1,\bf R),$ which topologically is equivalent to $SO(2n-1),$ there might be a topological obstruction for writing $\Delta \omega$ globally as $d\theta$ for some form $\theta.$ The potential obstruction is the winding number of the map $e: N\to GL(2n-1,\bf R),$ given by the (normalized) integral $$w(e) = \frac{1}{(2n-1)!} \left(\frac{i}{2\pi}\right) ^{n+1} \int_N \mbox{tr}\, (e^{-1} de)^{2n-1}.$$ The choice $\Gamma_{\mu b}^a$ in the anholonomic frame $e_a$ leads to diffeomorphism invariant integral $\int_N \omega_{2n-1}(A,A_0)$ and there are no anomalies or 2-forms along Diff$(N)$ orbits in $\cal M.$ On the other hand, there is a frame bundle anomaly related to local frame rotations; this takes exactly the same form as the pure gauge anomaly discussed above, \cite{BZ}. The choice $\Gamma_{\mu\nu}^{\lambda}$ in the coordinate frame is insensitive to the frame rotations $e_a \mapsto e'_a$ but it responds to a local change of coordinates. Explicit formulas for the forms $c_{2n-1}$ along Diff$N$ orbits are given in the appendix. \section{Schwinger terms from the local system of line bundles} As we saw in the previous section, APS index theorem gives us a system of local determinant bundles $\mbox{DET}_{\lambda}$ over certain open sets $U_{\lambda} \subset B.$ The infinite-dimensional group ${\cal K} = {\cal D}_0 \times {\cal G}_0$ acts in the parameter space $B$ mapping each of the subsets $U_{\lambda}$ onto itself. We denote $\bf k=$ Lie$(\cal K).$ In general, the determinant bundles are topologically nontrivial and one cannot lift directly the action of $\cal K$ to the total space of $\mbox{DET}_{\lambda}.$ Instead, there is a an extension $\hat {\cal K}$ which acts in the determinant bundles. The Lie algebra $\hat{\bf k}$ of $\hat{\cal K}$ is given in a standard way. It consists of pairs $(X,\alpha),$ where $X\in {\bf k}$ and $\alpha$ is a function on $B,$ with commutation relations $$[(X,\alpha), (Y,\beta)] = ([X,Y], {\cal L}_X \beta -{\cal L}_Y\alpha + c(X,Y; \cdot) ),$$ where the Schwinger term $c$ is a purely imaginary function on $B$ and antisymmetric bilinear function on $\bf k.$ It is defined as the value of the curvature of $\mbox{DET}_{\lambda}$ at the given point in $B$ along the vector fields $X,Y$ on $B.$ Here ${\cal L}_X$ denotes the Lie derivative on $B$ along the vector field $X.$ The Jacobi indentity in $\hat{\bf k}$ is an immediate consequence of the fact that the curvature is a closed 2-form on $U_{\lambda}\subset B.$ Let $U_{\lambda\lambda'}=U_{\lambda}\cap U_{\lambda'}.$ Over $U_{\lambda \lambda'}$ there is a natural complex line bundle $\mbox{DET}_{\lambda\lambda'}$ such that the fiber at a point $z$ is the top exterior power of the finite-dimensional vector space spanned by all eigenvectors of the Dirac operator $D_z$ on $N$ with eigenvalues in the range $\lambda < \mu < \lambda'.$ If $\lambda' < \lambda,$ we set $\mbox{DET}_{\lambda \lambda'} = \mbox{DET}_{\lambda'\lambda}^*.$ By construction, we have a natural isomorphism $$\mbox{DET}_{\lambda\lambda'}\otimes \mbox{DET}_{\lambda'\lambda''}\simeq \mbox{DET}_{\lambda \lambda''},$$ for all triples $\lambda,\lambda',\lambda''.$ \begin{theorem} (\cite{CMM}) For any pairs $\lambda,\lambda'$ of real numbers one has $$\mbox{\rm DET}_{\lambda\lambda'} \simeq \mbox{\rm DET}_{\lambda} \otimes \mbox{\rm DET}_{\lambda'}^*$$ over the set $U_{\lambda\lambda'}.$ \end{theorem} Note that even though in \cite{CMM} the discussion was mainly around the case of gauge potentials and gauge transformations, the proof of the theorem was abstract and very general, not depending on the particular type of parameter space for Dirac operators. In the gerbe terminology the content of this theorem is that the gerbe defined by the system of local line bundles $DET_{\lambda\lambda'}$ is trivial. The line bundles can be pushed forward to give a family of local line bundles on $B/\cal K$ since the spectral subspaces transform equivariantly under gauge transformations and changes of coordinates. However, over $B/\cal K$ the gerbe is no more trivial, i.e., it cannot be given as tensor products of local line bundles over the sets $pr(U_{\lambda}),$ where $pr: B \to B/{\cal K}$ is the canonical projection. The obstruction to the trivialization is an element of $H^3(B/{\cal K}, \bf Z),$ the Dixmier- Douady class of the gerbe. In \cite{CMM} the DD class was computed from the index theory in the case of Yang-Mills theory; the generalization to the case involving metrics and diffeomorphism is straight-forward and the free part of the cohomology class is given by the integral formula \Ref{eq:FDD}, with $B$ replaced by $B/\cal K.$ The importance of the above theorem comes from the following simple observation. Let $H_z=H_+(z,\lambda) \oplus H_-(z,\lambda)$ be the spectral decomposition of the fermionic '1-particle' Hilbert space with respect to a spectral cut at $\lambda\in {\bf R}$, not in the spectrum of ${D_z}$. This determines a representation of the CAR algebra in a Fock space ${\cal F}(z,\lambda),$ with a normalized vacuum vector $|z,\lambda>.$ The defining property of this representation is that $$a(u)|z,\lambda>=0= a^*(u')|z,\lambda>,$$ for $u\in H_+(z,\lambda)$ and $u'\in H_-(z,\lambda)$. All creation operators $a^*(u)$ and annihilation operators $a(u)$ are anticommuting except $$a^*(u) a(u') + a(u') a^*(u) = <u',u>,$$ where $<\cdot,\cdot>$ is the inner product in $H_z.$ If we change the vacuum level from $\lambda$ to $\lambda'> \lambda,$ we have an isomorphism ${\cal F}(z, \lambda) \to {\cal F}(z,\lambda')$ which is natural up to a multiplicative phase. The phase is fixed by a choice of normalized eigenvectors $u_1,u_2,\dots u_p$ in the energy range $\lambda < D_z < \lambda'$ and setting $|z,\lambda'> = a^*(u_1) \dots a^*(u_p)|z,\lambda'>.$ But this choice is exactly the same as choosing a (normalized) element in $\mbox{DET}_{\lambda\lambda'}$ over the point $z\in B.$ Thus, setting ${\cal F}_z ={\cal F}(z,\lambda) \otimes \mbox{DET}_{\lambda}(z)$ for any $\lambda$ not in the spectrum of $D_z$ we obtain, according to Theorem 1, a family of Fock spaces parametrized by points of $B$ but which do not depend on the choice of $\lambda,$ \cite{M2}. This gives us a smooth Fock bundle $\cal F$ over $B.$ The $\cal K$ action on the base lifts to a $\hat{\cal K}$ action in $\cal F,$ the extension part in $\hat{\cal K}$ coming entirely from the action in the determinant bundles $\mbox{DET}_{\lambda}.$ The Schr\"odinger wave functions for quantized fermions in background fields (parametrized by points of $B$) are sections of the Fock bundle. It follows that the Schwinger terms for the infinitesimal generators of $\hat{\cal K},$ acting on Schr\"odinger wave functions, are given by the formula for $c$ which describes the curvature of the determinant bundle in the $\cal K$ directions. In the case of $B=\cal A$ and ${\cal K}={\cal G},$ the elements in the Lie algebra are the Gauss law generators. This case was discussed in detail in \cite{CMM}. More generally, we give explicit formulas for the Schwinger terms in section \ref{sec:EC}. \section{Explicit computations} \label{sec:EC} The Schwinger term in $(2n-1)$-dimensional space will now be computed. This will be done by using notations for Yang-Mills, but it works for diffeomorphisms as well if a different symmetric invariant polynomial is used. Eq. \Ref{eq:POM} gives that \begin{eqnarray} &&\omega _{2n+1}(A+v,A_0) = (n+1)\int _0^1f^{\prime }p_{n+1}\Big( A+v-A_0 , \nonumber \\ \nopagebreak && fdA+f^2A^2+(1-f)dA_0+ (1-f)^2A_0^2 -f(f-1)[A_0,A]\nonumber \\ \nopagebreak && +f( f-1) [A-A_0,v]+\frac{1}{2}f(f-1)[v,v]\Big) dt,\nonumber \end{eqnarray} The Schwinger term can be calculated from this expression. However, since we are only interested in the Schwinger term up to a coboundary, an alternative is to use the \lq triangle formula\rq $ $ as in \cite{MS}: \[ \omega _{2n+1}(A+v,A_0) \sim \omega _{2n+1}(A_0+v,A_0)+\omega _{2n+1}(A+v,A_0+v), \] where \lq $\sim$\rq $ $ means equality up to a coboundary with respect to $d+\delta$, where $\delta$ is the BRST operator. This gives a simpler expression for the non-integrated Schwinger term and also for all other ghost degrees of $\omega _{2n+1}(A+v,A_0)$. Straight forward computations gives the result \begin{eqnarray} && \omega _{2n+1}(A+v,A_0)_{(2)} \sim \frac{(n+1)n}{2}p_{n+1}\left( v, dv+[A_0,v],dA_0+A_0^2\right) \nonumber \\ \nopagebreak && +\frac{(n+1)n(n-1)}{2}\int _0^1f^{\prime }(1-f)^2p_{n+1}\Big( A-A_0, dv+[A_0,v], \nonumber \\ \nopagebreak && dv+[A_0,v] , fdA+f^2A^2+(1-f)dA_0\nonumber \\ \nopagebreak && +(1-f)^2A_0^2- f(f-1)[A_0,A]\Big) dt\nonumber \end{eqnarray} where the index $(2)$ means the part of the form that is quadratic in the ghost. Inserting $n=1,2$ and $3$ gives: \begin{eqnarray} && \omega _3(A+v,A_0)_{(2)} \sim p_2(v, dv+[A_0,v])\nonumber \\ \nopagebreak && \omega _5(A+v,A_0)_{(2)} \sim 3p_3\left( v, dv+[A_0,v],dA_0+A_0^2\right) \nonumber \\ \nopagebreak && +p_3(A-A_0, dv+[A_0,v], dv+[A_0,v])\nonumber \\ \nopagebreak && \omega _7(A+v,A_0)_{(2)} \sim 6p_4\left( v, dv+[A_0,v],dA_0+A_0^2,dA_0+A_0^2\right) \nonumber \\ \nopagebreak && +p_4\big( A-A_0, dv+[A_0,v], dv+[A_0,v], \nonumber \\ \nopagebreak && dA +\frac{2}{5}A^2+3dA_0+\frac{12}{5}A_0^2+\frac{3}{5}[A_0,A]\big) .\nonumber \end{eqnarray} This gives expressions for the non-integrated Schwinger term in a pure Yang-Mills potential if $p_{n+1}$ is the symmetrized trace. The appropriate polynomial to use for the Levi-Civita connection is $p_{n+1}=\hat{A}(M)_{n+1}$, according to eq. \Ref{eq:AR}. Using \begin{eqnarray} \hat{A}(M) & = & 1+\left(\frac{1}{4\pi }\right) ^2\frac{1}{12}\mbox{tr}\left( R^2\right) \nonumber \\ \nopagebreak &&+ \left(\frac{1}{4\pi }\right) ^4\left( \frac{1}{288}\left( \mbox{tr}\left( R^2\right) \right) ^2 + \frac{1}{360}\mbox{tr}\left( R^4\right)\right) + ...\nonumber \end{eqnarray} gives for $n=1$ and 2: \begin{eqnarray} \omega _3(\Gamma +v,\Gamma _0)_{(2)} & \sim & \left(\frac{1}{4\pi }\right) ^2\frac{1}{12}\left(vdv+2\Gamma _0v^2\right)\nonumber \\ \nopagebreak \omega _5(\Gamma +v,\Gamma _0)_{(2)} & \sim & 0.\nonumber \end{eqnarray} Since the expression for $\omega _7$ is rather long we will omit to write it down. However, for the special case $\Gamma _0=0$ it becomes: \begin{eqnarray} && \omega _7(\Gamma +v,0)_{(2)} \sim \nonumber \\ \nopagebreak && \left(\frac{1}{4\pi }\right) ^4\left( \frac{1}{288}\cdot\frac{2}{3}\mbox{tr} \left( \Gamma dv\right)\mbox{tr}\left( dv\left( d\Gamma+\frac{2}{5}\Gamma ^2\right)\right) \right. \nonumber \\ \nopagebreak &&+ \frac{1}{288}\cdot\frac{1}{3}\mbox{tr}\left( (dv)^2\right) \mbox{tr}\left( \Gamma \left( d\Gamma +\frac{2}{5}\Gamma ^2\right)\right) \nonumber \\ \nopagebreak && \left. + \frac{1}{360}\cdot\frac{1}{3}\mbox{tr} \left(\left( R-\frac{3}{5}\Gamma ^2\right) \left( (dv)^2\Gamma +(dv)\Gamma dv +\Gamma (dv)^2\right)\right)\right) .\nonumber \end{eqnarray} This expression can be simplified if subtracting the coboundary \begin{eqnarray} &&\left(\frac{1}{4\pi }\right) ^4\frac{1}{288}\cdot\frac{2}{3}\left( \delta\left(\mbox{tr}\left( \Gamma dv\right)\mbox{tr}\left( \Gamma d\Gamma +\frac{4}{5}\Gamma ^3\right)\right)\right. \nonumber \\ \nopagebreak && +\left. d\left(\mbox{tr}\left( v dv\right)\mbox{tr}\left( \Gamma d\Gamma +\frac{2}{3}\Gamma ^3\right)\right)\right) .\nonumber \end{eqnarray} The result is \begin{eqnarray} && \omega _7 (\Gamma +v,0)_{(2)} \sim \left(\frac{1}{4\pi }\right) ^4\left( \frac{1}{288}\mbox{tr} \left( v dv\right)\mbox{tr}R^2\right. \nonumber \\ \nopagebreak && \left. + \frac{1}{360}\cdot\frac{1}{3}\mbox{tr} \left(\left( R-\frac{3}{5}\Gamma ^2\right)\left( (dv)^2\Gamma +(dv)\Gamma dv +\Gamma (dv)^2\right)\right)\right) .\nonumber \end{eqnarray} The gravitational Schwinger terms are obtained by multiplying with the normalization factor $(i/2\pi )^{-1}$, inserting the integration over $N$ and evaluating on vector fields $X$ and $Y$ on $M$ generating diffeomorphisms. The Levi-Civita connection and curvature have components $(\Gamma )^{i^{\prime }}_{\,\, j^{\prime }}=\Gamma ^{i^{\prime }}_{i j^{\prime }}dx^i$ and $(R )^{ i^{\prime }}_{\,\, j^{\prime }}=R ^{i^{\prime }}_{ij j^{\prime }}dx^i\wedge dx^j$. Recall that $(v(X))^{i^{\prime }}_{\,\, j^{\prime }}=\partial _{j^{ \prime }}X^{i^{\prime }}$, see, for instance, \cite{BZ}. To illustrate how the Schwinger terms can be computed, we give the result for 1 space dimension: \[ -2\pi i\int _N\omega _3(\Gamma +v,\Gamma _0)_{(2)}(X,Y) \sim -\frac{i}{48\pi }\int _N \left(\partial _xX\right) \partial _x^2Ydx, \] where \lq $\sim$\rq $ $ now means equality up to a coboundary with respect to the BRST operator. When both a Yang-Mills field and gravity are present, the relevant polynomial is a sum of polynomial of type \[ p_k\left( F^{\omega (t)}\right) \tilde{p}_l\left( R^{\omega (t)}\right) , \] where the curvatures $F^{\omega (t)}$ and $R^{\omega (t)}$ are with respect to pure Yang-Mills and pure gravity, respective. This gives \begin{eqnarray} && \int _Mp_k\left( F^{\omega (t)}\right) \tilde{p}_l\left( R^{\omega (t)}\right) \nonumber \\ \nopagebreak &=& \int _N\int _0^1\Big[ kp_k\left( f^{\prime }(t)(A+v_A-A_0), F^{\omega (f(t))}\right) \tilde{p}_l\left( R^{\omega (h(t))}\right) \nonumber \\ \nopagebreak && +lp_k\left( F^{\omega (f(t))}\right) \tilde{p}_l\left( h^{\prime }(t)(\Gamma +v_{\Gamma }-\Gamma _0),R^{\omega (h(t))}\right) \Big] dt\nonumber . \end{eqnarray} The expression is independent of $f$ and $h$ (see below). With a choice such that $f^{\prime }(t)=0$, $t\in [1/2,1]$ and $h^{\prime }(t)=0$, $t\in [0,1/2]$, this implies that \begin{eqnarray} \label{eq:REFD} \int _Mp_k\left( F^{\omega (t)}\right) \tilde{p}_l\left( R^{\omega (t)}\right) & = &\int _N\left( \omega _{2k-1}(A+v_A,A_0)\tilde{p}_l(R_0)\right.\nonumber \\ \nopagebreak &&\left. +p_k(F)\tilde{\omega }_{2l-1}(\Gamma +v_{\Gamma },\Gamma _0)\right) . \end{eqnarray} Thus, the Schwinger term in combined Yang-Mills and gravity is up to a coboundary equal to the part of the expansion of \Ref{eq:REFD} that is of second ghost degree. In particular, this implies that Schwinger terms which have one Yang-Mills ghost and on diffeomorphism ghost are in cohomology equal to the Schwinger term obtained from the form in \Ref{eq:REFD}. Thus, truly mixed Schwinger terms do not exist. Notice that if the background fields are vanishing then the Schwinger term is gravitational (although some parts of the form degrees are taken up by the Yang-Mills polynomial). This can give anomalies of Virasoro typ in higher dimensions. Observe that there is nothing special about gravity, a Yang-Mills Schwinger term is obtained by interchanging the role of $f$ and $h$. This does however not mean that the gravitational Schwinger term differ from the Yang-Mills Schwinger term by a coboundary. The terms with $k=0$ respective $l=0$ ruin this argument. It is easy to see that our method of computing the Schwinger term agrees with one of the most common approaches: The polynomial $p_k(F^n)-p_k(F_0^n)$ is written as $(d+\delta )$ on a form, the (non-integrated) Chern-Simons form. The Schwinger term is the given by the part of the Chern-Simons form that is quadratic in the ghost. For the case when both Yang-Mills and gravity are present, the relevant polynomial is a sum of polynomials \begin{equation} \label{eq:POL} p_k(F)\tilde{p}_l(R)-p_k(F_0)\tilde{p}_l(R_0).\nonumber \end{equation} There is an ambiguity in the definition of the Chern-Simons form; it is for instance possible to add forms of type $(d+\delta )\chi$ to it. However, an ambiguity of this type will only change the Schwinger term by a coboundary. It will now be shown that the ambiguity in the definition of the Chern-Simons form is only of this type. Thus, we must prove that closeness with respect to $(d+\delta )$ implies exactness. This can be done by introducing the degree 1 derivation $\triangle$ defined on the generators by: $\triangle (d+\delta )(A+v_A)=A+v_A, \triangle (d+\delta )(\Gamma +v_{\Gamma })=\Gamma +v_{\Gamma }, \triangle dA_0=A_0,\triangle d\Gamma _0=\Gamma _0$, and otherwise zero. Then $\triangle (d+\delta )+(d+\delta )\triangle $ is a degree 0 derivation which is equal to 1 on the generators. Therefore, if $\chi$ is closed with respect to $(d+\delta )$, then $\chi$ is proportional to $(d+\delta )\triangle \chi$. An example of a (non-integrated) Chern-Simons form for the polynomial in \Ref{eq:POL} is \[ \omega _{2k-1}(A+v_A,A_0)\tilde{p}_l(R_0)+p_k(F)\tilde{\omega }_{2l-1}(\Gamma +v_{\Gamma },\Gamma _0). \] This is in complete agreement with \Ref{eq:REFD}. \newpage
\section{Introduction} Any renormalizable theory only becomes predictive once it has been supplied with a sufficient number of experimental inputs so as to fix the free parameters that appear in its Lagrangian. In order for the subsequent theoretical predictions to be as accurate as possible, these inputs are chosen from the experimentally best-measured quantities available. In the case of the Standard Model of electroweak interactions these are the electromagnetic coupling constant, $\alpha$, the Fermi coupling constant, $G_F$, and the mass of the $Z^0$ boson, $M_Z$. Their recognized best values, along with their absolute and relative errors that they represent are \cite{PDG,LEPEWWG} \begin{xxalignat}{2} {}\qquad\qquad\qquad\qquad \alpha&=1/(137.0359895\pm0.0000061)& (0.045\,{\rm ppm})& \\ G_F&=(1.16639\pm0.00002)\times10^{-5}\,{\rm GeV^{-2}}& (17\,{\rm ppm})& \\ M_Z&=91.1867\pm0.0021\,{\rm GeV}& (23\,{\rm ppm})& \end{xxalignat} The first of these, $\alpha$, is measured at momentum $q=0$ and when used in the analysis high energy data is afflicted with a `hadronic uncertainty' entering already at the 1-loop level and arising because it must be `run up' from low energy crossing, on the way, the hadronic resonance region. Most processes, such as $e^+e^-$ annihilation, that allow $M_Z$ to be accurately determined, are also mediated by the photon and thus contain a pair of characteristic energy scales. Experiments performed at low-energy therefore do not provide useful insight into physics at the high scale, $M_Z$. The Fermi coupling constant is obtained from the muon lifetime, $\tau_\mu$, via a calculation in the Fermi model, in which the weak interactions are represented by a contact interaction. It suffers from neither of the disadvantages described above. There can be no hadronic effects at 1-loop in this model and even in the full Standard Model 1-loop hadronic effects in muon decay are suppressed by a factor $m_f^2/M_W^2$ where $m_f$ is a light fermion mass. Charged current processes are only mediated by the $W$ boson whose effects are unobscured no matter at what energy experiments are performed. The measurement of the muon lifetime can thus be considered a high-energy experiment that is actually performed at a low scale \cite{Roberts}. This point is highlighted by the fact that the next generation of muon lifetime measurements will be sensitive to $W$ propagator effects. In the mid-80's, just before the turn on of LEP, a CERN report concluded that the error on $M_Z$ would be $\pm50$\,MeV or 550\,ppm and that ``A factor of 2--3 improvement can be reached with a determined effort'' \cite{CERN}. It was thus generally thought that $M_Z$ would be the input parameter limiting the accuracy with which theoretical predictions could be made. As is often case, due to a variety of unforeseen effects, the uncertainty turned out to have been significantly overestimated and it now approaches that of $G_F$. The lesson that should be learned from this is that it is extremely difficult to predict, even in the relatively short term, the accuracy to which fundamental parameters will be determined and it is important that these be extracted to the limits that the current theoretical and experimental technology allows. A great effort and expense was brought to bear in order to reduce the uncertainty on $M_Z$ to the level given in the table above and whereas it is unlikely to diminish significantly in the near future a muon collider offers the prospect of a reduction by a factor of 10 \cite{Blondel}. The Fermi coupling constant is closely related to the $\rho$-parameter that was introduced by Ross and Veltman \cite{RossVeltman} as a way of interrogating the mass generation sector of the theory \begin{equation} \rho=\frac{M_W^2}{M_Z^2\cos^2\theta_W}=1+\delta\rho \end{equation} The detailed understanding of the nature of the Higgs mechanism will come about through the scrutiny of systems and processes that are most sensitive to the Yukawa sector of the theory. The Fermi coupling constant obtained from the muon lifetime provides such a probe but stands alone for the tremendous accuracy with which in can be measured. For this reason a considerable amount of theoretical effort, reviewed in section~\ref{sec:WeakCorr}, has been devoted to the study of higher order weak corrections to the muon lifetime. It can be shown that the 2-loop QED corrections to the muon lifetime affect the predicted value of the Higgs mass at the percent level which is precisely the level needed to probe the detailed structure of the Higgs system by its radiative corrections. In this paper we look closely at the theoretical and experimental considerations necessary for the extraction of $G_F$ from $\tau_\mu$ to an accuracy of 1\,ppm or better as is anticipated for new experiments planned at the Brookhaven National Laboratory, the Paul Scherrer Institute and the Rutherford-Appleton Laboratory. What is known about the relationship between the muon lifetime and the Fermi coupling constant is reviewed in section~\ref{sec:taumuGF}. In section~\ref{sec:Definition} inadequacies that appear at higher orders with the usual definition of $G_F$ are pointed out and the relative merits of the one adopted here are discussed. In section~\ref{sec:AlphaRenorm} it is shown how to perform the renormalization of the electromagnetic coupling constant in such a way as to incorporate the large logarithms, $\ln(m_e^2/m_\mu^2)$. Section~\ref{sec:calcul} describes the methods used for the calculation of the 2-loop quantum electrodynamic (QED) corrections to the muon lifetime in the Fermi model. The sources of the dominant theoretical and experimental uncertainties are examined in sections~\ref{sec:TheoUncert} and \ref{sec:ExpUncert} respectively. The incorporation of the value of $G_F$ obtained here into analyses using the full electroweak Standard Model is discussed in section~\ref{sec:WeakCorr}. Appendix~\ref{sec:multloop} contains expressions for certain loop integrals for which exact analytic forms are known in dimensional regularization. Appendix~\ref{sec:diagresults} gives the results for the individual Feynman diagrams that occur in the calculation of the 2-loop QED corrections to muon lifetime and Appendix~\ref{sec:elecspec} gives an expression for the electron spectrum to ${\cal O}(\alpha)$ keeping the full electron mass dependence. Finally Appendix~\ref{sec:BR3e} examines the branching ratio for the process $\mu^-\rightarrow e^-\nu_\mu\bar\nu_e e^+e^-$. \section{The Muon Lifetime and The Fermi Coupling Constant} \label{sec:taumuGF} In the Minkowskian metric in which time-like momenta squared are positive, the Lagrangian, relevant for the calculation of the muon lifetime in the Fermi theory is \begin{equation} {\cal L}_F={\cal L}_{{\rm QED}}^0+{\cal L}_{{\rm QCD}}^0+{\cal L}_W \label{eq:FullLagrangian} \end{equation} Here ${\cal L}_{{\rm QED}}^0$ is the usual bare Lagrangian of Quantum Electrodynamics (QED), \begin{equation} {\cal L}_{{\rm QED}}^0= \sum_f\bar\psi_f^0(\slash p-m_f)\psi_f^0 -\frac{1}{4}\left(\partial_\rho A_\sigma^0 -\partial_\sigma A_\rho^0\right)^2 -ie^0\sum_f Q_f\bar\psi_f^0\gamma_\rho\psi_f^0 A_\rho^0. \end{equation} The sum is over all fermion species, $f$, with wavefunction, $\psi_f$, mass, $m_f$, and electric charge, $Q_f$. $A_\rho$ is the photon field. The superscript ${}^0$ indicates bare, as opposed to renormalized, quantities. ${\cal L}_{{\rm QCD}}^0$ is the bare Quantum Chromodynamic (QCD) Lagrangian responsible for strong interactions. The Fermi contact interaction that mediates muon decay is \begin{equation} {\cal L}_W=-2\sqrt{2} G_F \big[\bar\psi_{\nu_\mu}^0\gamma_\lambda\gamma_L\psi_\mu^0\big]. \big[\bar\psi_e^0\gamma_\lambda\gamma_L\psi_{\nu_e}^0\big] \label{eq:FermiLagrange} \end{equation} in which $\psi_\mu$, $\psi_e$, $\psi_{\nu_\mu}$ and $\psi_{\nu_e}$ are the wavefunctions for the muon, the electron and their associated neutrinos respectively. and $\gamma_L$ denotes the usual Dirac left-hand projection operator. For the present purposes the Fermi coupling constant, $G_F$, goes unrenormalized. To leading order in $G_F$ and all orders in $\alpha$ the formula obtained for the muon lifetime, $\tau_\mu$, by means of the ${\cal L}_{{\rm F}}$ takes the general form \begin{equation} \frac{1}{\tau_\mu}\equiv\Gamma_\mu=\Gamma_0(1+\Delta q). \label{eq:QEDcorr} \end{equation} where \begin{equation} \Gamma_0=\frac{G_F^2 m_\mu^5}{192\pi^3} \label{eq:Gamma0} \end{equation} and $\Delta q$ encapsulates the higher order QED corrections and can be expressed as power series expansion in the renormalized electromagnetic coupling constant $\alpha_r=e_r^2/(4\pi)$. \begin{equation} \Delta q=\sum_{i=0}^\infty\Delta q^{(i)} \label{eq:DeltaqSeries} \end{equation} in which the index $i$ gives the power of $\alpha_r$ that appears in $\Delta q^{(i)}$. Although ${\cal L}_F$ is not renormalizable, the $\Delta q^{(i)}$ can be shown to be finite \cite{BermSirl}. This follows from the fact that when Fierz rearrangement is used to rewrite ${\cal L}_W$ in so-called charge retention order, \[ {\cal L}_W\rightarrow-2\sqrt{2} G_F \big[\bar\psi_e^0\gamma_\lambda\gamma_L\psi_\mu^0\big]. \big[\bar\psi_{\nu_\mu}^0\gamma_\lambda\gamma_L\psi_{\nu_e}^0\big] \] the currents remain purely V$-$A in form. By contrast, in the case of neutron decay, the analogous transformation generates scalar and pseudo-scalar terms and the following arguments break down. The radiative corrections in that case are not finite. Considering the vector part $\bar\psi_e\gamma_\mu\psi_\mu$ of this effective $\mu$-$e$ current, one sees that after fermion mass renormalization is performed the remaining divergences are independent of the masses and thus cancel, as for the case of pure QED. The QED corrections to the axial vector part may be shown to also be finite by noting that the transformations $\psi_e\rightarrow\gamma_5\psi_e$ and $m_e\rightarrow -m_e$ leave ${\cal L}_{{\rm QED}}$ and ${\cal L}_{{\rm QCD}}$ invariant but exchange $\bar\psi_e\gamma_\lambda\psi_\mu\leftrightarrow \bar\psi_e\gamma_\lambda\gamma_5\psi_\mu$. Thus the radiative corrections to the axial-vector contribution to the muon decay matrix element can be obtained from those of the vector contribution by setting $m_e\rightarrow-m_e$. This argument was used by Roos and Sirlin \cite{RoosSirlin} to show that terms odd in $m_e$ would cancel between vector and axial-vector contributions in the expression for the differential decay rate. They then went on to show, by direct examination of known analytic form of the differential decay rate \cite{BehrFinkSirl}, that the phase-space integration could not generate terms linear in $m_e$ and thus that the leading electron mass corrections at 1-loop are ${\cal O}\left(\alpha(m_e^2/m_\mu^2)\ln(m_e^2/m_\mu^2)\right)$, ${\cal O}\left(\alpha(m_e^2/m_\mu^2)\right)$ and higher but they did not exclude the possibility of ${\cal O}\left(\alpha(m_e^3/m_\mu^3)\right)$ terms which, in fact, do occur. It is possible to show that $\Delta q$ cannot contain any terms that are odd in $m_e$ at any order in $\alpha$. The total decay rate may be calculated directly as the imaginary part of muon self-energy diagrams and was done in Ref.\cite{muonprl}. In charge retention order terms in the numerators that are odd in the electron mass lead to a helicity flip along the internal electron line that causes the purely left-handed V$-$A vertices to annihilate. However the integrals themselves generate non-analytic terms such as $m_e^2\sqrt{m_e^2}=|m_e|^3$ which may appear to be odd in $m_e$ if the absolute value is dropped. The above considerations are true in any regularization scheme and have the importance consequence that in the limit $m_e\rightarrow0$ only the radiative corrections to the vector pieces in ${\cal L}_W$ need to be calculated as they are equal to those to the axial vector part. This avoids entirely the problems associated with $\gamma_5$ when dimensional regularization is used. Dropping the electron neutrino mass and keeping only the leading term in that of the muon neutrino gives \begin{equation} \Delta q^{(0)}=-8x-12x^2\ln x+8x^3-x^4-8y+{\cal O}(xy), \ \ \ \ \ \ x=\frac{m_e^2}{m_\mu^2}, \ \ \ \ \ \ y=\frac{m_{\nu_\mu}^2}{m_\mu^2} \label{eq:Deltaq0} \end{equation} that comes from phase space integrations. Notice that the coefficient of the leading terms in the electron and muon neutrino masses are identical which again follows from the fact that their wavefunctions may be interchanged by Fierz rearrangement of ${\cal L}_{{\rm W}}$. The first order corrections to $\Delta q$ are \begin{equation} \Delta q^{(1)}= \left(\frac{\alpha_r}{\pi}\right) \left(\frac{25}{8}-3\zeta(2) -(34+12\ln x)x+96\zeta(2)\,x^{\frac{3}{2}} +{\cal O}(x\ln^2 x) \right) \label{eq:Deltaq1} \end{equation} where $\zeta$ is the Riemann zeta function and $\zeta(2)=\pi^2/6$. The leading electron mass-independent term was calculated Kinoshita and Sirlin \cite{KinoSirl} and it was the observation that the result is well-behaved in the limit $m_e\rightarrow0$ that ultimately lead to the discovery of the Kinoshita-Lee-Nauenberg (KLN) theorem \cite{KinoLeeNaue}. An exact expression for the full electron mass dependence in $\Delta q^{(1)}$ has been given by Nir \cite{Nir}. Starting from the expressions for the 1-loop QED corrections to the electron spectrum given by Behrends {\it et al.}\cite{BehrFinkSirl} and, after taking into account the correction given in Appendix~C of Ref.\cite{KinoSirl}, we obtain complete agreement with Eq.(12) of Ref.\cite{Nir}. As a spinoff we have considerably simplified the expression for the electron spectrum of Ref.\cite{BehrFinkSirl}. The result appears in Appendix \ref{sec:elecspec}. An analogous expression has been given by Czarnecki {\it et al.} \cite{CzarJezaKuhn} for the QCD corrections to $b\rightarrow c\tau\bar\nu_\tau$. We have also checked the electron mass-dependent terms in Eq.(\ref{eq:Deltaq1}) by performing a large momentum expansion of the imaginary part of propagator-type diagrams. The second order corrections to $\Delta q$ have been presented recently \cite{muonprl,muonhad} for $m_e=0$. The result, ignoring the effects of tau loops, is \begin{multline} \Delta q^{(2)}= \left(\frac{\alpha_r}{\pi}\right)^2 \bigg(\frac{156815}{5184} -\frac{1036}{27}\zeta(2) -\frac{895}{36}\zeta(3) +\frac{67}{8}\zeta(4)\\ +53\zeta(2)\ln2 -(0.042\pm0.002) \bigg) \label{eq:Deltaq2} \end{multline} where $\zeta(3)=1.20206...$ and $\zeta(4)=\pi^4/90$. The numerical constant is the hadronic contribution with a conservative estimate of its error. The contribution from tau loops has been shown to be very small \cite{muonhad} as anticipated from the decoupling theorem \begin{equation} \Delta q^{(2)}_{{\rm tau}}=-\left(\frac{\alpha_r}{\pi}\right)^2\,0.00058 \label{eq:Deltaq2tau} \end{equation} and will be neglected unless otherwise stated. The methods used to obtain the result (\ref{eq:Deltaq2}) will be described in some detail in section~\ref{sec:calcul}. A key feature is that the calculation was performed by means of cutting relations \cite{Cutkosky} applied to the 25 1-particle irreducible (1PI) diagrams that appear in Fig.\ref{4loopdiagrams}. In these diagrams the thick line represents a muon with the external legs being on-shell. The thin lines represent the electron, electron neutrino or muon neutrino all of which are taken to be massless. The wiggly line represents the photon. Cuts passing only through massless internal lines give non-zero contribution and all others vanish. In particular, since the external, leg is on-shell, any cut through a muon line is identically zero. The imaginary part of the diagrams of Fig.\ref{4loopdiagrams} then generates precisely those combinations of the products of amplitudes that appear in the calculation of the 2-loop QED corrections to muon decay. Moreover, when the calculation is performed in this way, the intricate cancellation of infrared (IR) divergences, that would occur between these products of amplitudes, coming from distinct cuts applied to a given diagram, is largely taken care of automatically. \begin{figure} \begin{picture}(600,280)(10,-280) \SetScale{.25} \SetOffset(0,0) \SetWidth{8} \Line(50,50)(75,50) \Line(175,50)(300,50) \SetWidth{2} \Line(75,50)(175,50) \Oval(125,50)(40,50)(0) \PhotonArc(175,73)(73,238,340){4}{11.5} \PhotonArc(179,58)(35,246,343){4}{5.5} \Text(20,-5)[m]{\small A1} \SetOffset(90,0) \SetWidth{8} \Line(50,50)(75,50) \Line(175,50)(300,50) \SetWidth{2} \Line(75,50)(175,50) \Oval(125,50)(40,50)(0) \PhotonArc(215,49)(22,5,175){-4}{5.5} \PhotonArc(190,90)(85,242,330){4}{10.5} \Text(20,-5)[m]{\small A2} \SetOffset(180,0) \SetWidth{8} \Line(50,50)(108,50) \Line(192,50)(300,50) \SetWidth{2} \Line(95,50)(195,50) \Oval(150,50)(30,42)(0) \PhotonArc(230,49)(22,5,175){-4}{5.5} \PhotonArc(175,110)(122,210,-30){4}{18.5} \Text(20,-5)[m]{\small A3} \SetOffset(270,0) \SetWidth{8} \Line(50,50)(132,50) \Line(218,50)(300,50) \SetWidth{2} \Line(125,50)(225,50) \Oval(175,50)(30,42)(0) \PhotonArc(175,27)(28,180,0){-4}{7.5} \PhotonArc(175,-2)(108,28,152){4}{17.5} \Text(20,-5)[m]{\small A4} \SetOffset(360,0) \SetWidth{8} \Line(50,50)(132,50) \Line(218,50)(300,50) \SetWidth{2} \Line(125,50)(225,50) \Oval(175,50)(30,42)(0) \PhotonArc(175,75)(80,200,-20){4}{15.5} \PhotonArc(175,-22)(130,35,145){4}{19.5} \Text(20,-5)[m]{\small A5} \SetOffset(0,-65) \SetWidth{8} \Line(50,50)(108,50) \Line(193,50)(300,50) \SetWidth{2} \Line(100,50)(200,50) \Oval(150,50)(30,42)(0) \PhotonArc(175,64)(60,225,345){-4}{9.5} \PhotonArc(165,13)(92,90,156){4}{8.5} \PhotonArc(165,-25)(130,34,90){4}{10} \Text(20,-5)[m]{\small A6} \SetOffset(90,-65) \SetWidth{8} \Line(50,50)(100,50) \Line(200,50)(300,50) \SetWidth{2} \Line(100,50)(200,50) \Oval(150,50)(40,50)(0) \PhotonArc(172,12)(25,41,184){4}{5.5} \PhotonArc(160,48)(58,198,270){-4}{5.0} \PhotonArc(160,98)(108,270,332){-4}{9.0} \Text(20,-5)[m]{\small A7} \SetOffset(180,-65) \SetWidth{8} \Line(50,50)(75,50) \Line(175,50)(300,50) \SetWidth{2} \Line(75,50)(175,50) \Oval(125,50)(40,50)(0) \PhotonArc(154,60)(60,238,350){4}{9.5} \PhotonArc(195,78)(67,238,335){4}{9.5} \Text(20,-5)[m]{\small B1} \SetOffset(270,-65) \SetWidth{8} \Line(50,50)(75,50) \Line(175,50)(300,50) \SetWidth{2} \Line(75,50)(175,50) \Oval(125,50)(40,50)(0) \PhotonArc(237,45)(30,15,165){4}{6.5} \PhotonArc(165,90)(85,250,330){4}{10.5} \Text(20,-5)[m]{\small C1} \SetOffset(360,-65) \SetWidth{8} \Line(45,50)(108,50) \Line(192,50)(300,50) \SetWidth{2} \Line(95,50)(195,50) \Oval(150,50)(30,42)(0) \PhotonArc(247,45)(30,15,165){4}{6.5} \PhotonArc(155,90)(100,205,-25){4}{16.5} \Text(20,-5)[m]{\small C2} \SetOffset(0,-130) \SetWidth{8} \Line(50,50)(75,50) \Line(175,50)(300,50) \SetWidth{5} \Oval(208,8)(16,16)(0) \SetWidth{2} \Line(75,50)(175,50) \Oval(125,50)(40,50)(0) \PhotonArc(190,90)(85,242,272){4}{4.0} \PhotonArc(190,90)(85,292,330){4}{4.5} \Text(20,-5)[m]{\small C3} \SetOffset(90,-130) \SetWidth{8} \Line(50,50)(132,50) \Line(218,50)(300,50) \SetWidth{5} \Oval(175,-8)(16,16)(0) \SetWidth{2} \Line(125,50)(225,50) \Oval(175,50)(30,42)(0) \PhotonArc(160,65)(75,190,270){4}{9.5} \PhotonArc(190,65)(75,270,350){4}{9.5} \Text(20,-5)[m]{\small C4} \SetOffset(180,-130) \SetWidth{8} \Line(50,50)(132,50) \Line(218,50)(300,50) \SetWidth{5} \Oval(175,-7)(16,16)(0) \SetWidth{2} \Line(125,50)(225,50) \Oval(175,50)(30,42)(0) \PhotonArc(170,20)(30,160,245){4}{4.5} \PhotonArc(180,20)(30,295,20){4}{4.5} \Text(20,-5)[m]{\small C5} \SetOffset(270,-130) \SetWidth{8} \Line(50,50)(100,50) \Line(200,50)(300,50) \SetWidth{2} \Line(100,50)(200,50) \Oval(150,50)(40,50)(0) \PhotonArc(185,90)(85,250,330){4}{8.5} \PhotonArc(160,-5)(40,48,150){4}{5.75} \Text(20,-5)[m]{\small D1} \SetOffset(360,-130) \SetWidth{8} \Line(50,50)(110,50) \Line(210,50)(300,50) \SetWidth{2} \Line(110,50)(220,50) \Oval(160,50)(40,50)(0) \PhotonArc(200,52)(48,239,354){-4}{7.5} \PhotonArc(137,25)(25,158,323){-4}{6.5} \Text(20,-5)[m]{\small D2} \SetOffset(0,-195) \SetWidth{8} \Line(50,50)(115,50) \Line(235,50)(300,50) \SetWidth{2} \Line(110,50)(240,50) \Oval(175,50)(40,60)(0) \PhotonArc(200,-5)(40,55,155){4}{5.75} \PhotonArc(165,33)(40,190,330){4}{7.75} \Text(20,-5)[m]{\small D3} \SetOffset(90,-195) \SetWidth{8} \Line(50,50)(115,50) \Line(235,50)(300,50) \SetWidth{2} \Line(110,50)(240,50) \Oval(175,50)(40,60)(0) \PhotonArc(175,-72)(105,65,115){-4}{7.5} \PhotonArc(175,17)(23,190,-10){-4}{5.5} \Text(20,-5)[m]{\small D4} \SetOffset(180,-195) \SetWidth{8} \Line(50,50)(115,50) \Line(235,50)(300,50) \SetWidth{2} \Line(110,50)(240,50) \Oval(175,50)(40,60)(0) \PhotonArc(144,23)(23,160,327){-4}{5.5} \PhotonArc(206,23)(23,213,20){-4}{5.5} \Text(20,-5)[m]{\small D5} \SetOffset(270,-195) \SetWidth{8} \Line(50,50)(75,50) \Line(175,50)(300,50) \SetWidth{2} \Oval(208,8)(16,16)(0) \SetWidth{2} \Line(75,50)(175,50) \Oval(125,50)(40,50)(0) \PhotonArc(190,90)(85,242,272){4}{4.0} \PhotonArc(190,90)(85,292,330){4}{4.5} \Text(20,-5)[m]{\small D6} \SetOffset(360,-195) \SetWidth{8} \Line(50,50)(132,50) \Line(218,50)(300,50) \SetWidth{2} \Oval(175,-8)(16,16)(0) \SetWidth{2} \Line(125,50)(225,50) \Oval(175,50)(30,42)(0) \PhotonArc(160,65)(75,190,270){4}{9.5} \PhotonArc(190,65)(75,270,350){4}{9.5} \Text(20,-5)[m]{\small D7} \SetOffset(0,-260) \SetWidth{8} \Line(50,50)(132,50) \Line(218,50)(300,50) \SetWidth{2} \Oval(175,-7)(16,16)(0) \SetWidth{2} \Line(125,50)(225,50) \Oval(175,50)(30,42)(0) \PhotonArc(170,20)(30,160,245){4}{4.5} \PhotonArc(180,20)(30,295,20){4}{4.5} \Text(20,-5)[m]{\small D8} \SetOffset(90,-260) \SetWidth{8} \Line(50,50)(125,50) \Line(225,50)(300,50) \SetWidth{2} \Line(110,50)(240,50) \Oval(175,50)(40,50)(0) \PhotonArc(125,54)(45,188,300){4}{7.5} \PhotonArc(225,54)(45,240,352){4}{7.5} \Text(20,-5)[m]{\small E1} \SetOffset(180,-260) \SetWidth{8} \Line(50,50)(132,50) \Line(217,50)(300,50) \SetWidth{2} \Line(110,50)(240,50) \Oval(175,50)(30,42)(0) \PhotonArc(152,60)(68,190,270){4}{7.0} \PhotonArc(152,55)(63,270,340){4}{5.5} \PhotonArc(198,55)(63,200,270){4}{6.0} \PhotonArc(198,60)(68,270,350){4}{7.5} \Text(20,-5)[m]{\small F1} \SetOffset(270,-260) \SetWidth{8} \Line(50,50)(118,50) \Line(202,50)(300,50) \SetWidth{2} \Line(110,50)(210,50) \Oval(160,50)(30,42)(0) \PhotonArc(204,85)(77,238,332){4}{9.5} \PhotonArc(160,23)(80,20,160){4}{16.5} \Text(20,-5)[m]{\small G1} \SetOffset(360,-260) \SetWidth{8} \Line(45,50)(132,50) \Line(218,50)(305,50) \SetWidth{2} \Line(125,50)(225,50) \Oval(175,50)(30,42)(0) \PhotonArc(185,15)(90,90,156){4}{8.5} \PhotonArc(185,-25)(130,34,90){4}{10} \PhotonArc(165,125)(130,214,270){4}{10} \PhotonArc(165,85)(90,270,338){4}{8.5} \Text(20,-5)[m]{\small G2} \end{picture} \caption{ \label{4loopdiagrams} 4-loop diagrams whose cuts give contributions to the muon decay rate. The diagrams are grouped according to the main integration topologies A--G (see Figs. \ref{TopoAFigs}--\ref{TopoDEFGFigs}). The results for the individual diagrams are given in Appendix \ref{sec:diagresults} } \end{figure} The result for $\Delta q^{(2)}$ in Eq.~(\ref{eq:Deltaq2}) is composed of several independent pieces. The part coming from the purely photonic diagrams of Fig.\ref{4loopdiagrams} that contain no closed charged-fermion loop and related 1-particle reducible (1PR) external leg corrections is \begin{eqnarray} \Delta q_{\gamma\gamma}^{(2)}&=& \left(\frac{\alpha_r}{\pi}\right)^2 \bigg(\frac{11047}{2592}-\frac{1030}{27}\zeta(2) -\frac{223}{36}\zeta(3) +\frac{67}{8}\zeta(4) +53\zeta(2)\ln (2)\bigg) \label{eq:Deltaq2photon}\\ &=&\left(\frac{\alpha_r}{\pi}\right)^2 3.55877. \end{eqnarray} This was calculated in a general $R_\xi$ gauge and was found to be both gauge invariant and ultraviolet (UV) finite. Feynman diagrams containing an electron loop or $e^+e^-$ pair in the final state are obtained from cuts to diagrams D6--D8 of Fig.\ref{4loopdiagrams}. Addition of the related 1PR external leg corrections and diagrams containing coupling constant counterterm insertions gives \begin{eqnarray} \Delta q^{(2)}_{{\rm elec}}&=& -\left(\frac{\alpha_r}{\pi}\right)^2 \left(\frac{1009}{288}-\frac{77}{36}\zeta(2) -\frac{8}{3}\zeta(3)\right) \label{eq:Deltaq2elec}\\ &=&\left(\frac{\alpha_r}{\pi}\right)^2 3.22034 \label{eq:electronnum} \end{eqnarray} The value obtained in Eq.(\ref{eq:electronnum}) is consistent with a numerical study presented in Ref.\cite{LukeSavaWise} in the context of semi-leptonic decays of heavy quarks, $Q\rightarrow X_q e\bar\nu_e$. The contribution from diagrams containing muon loops, diagrams C3--C5 of Fig.\ref{4loopdiagrams} can be shown to be identical in the on-shell and $\overline{{\rm MS}}$ renormalization schemes for a 't~Hooft mass $\mu=m_\mu$ (see section\ref{sec:AlphaRenorm}). These plus related 1PR external leg corrections and diagrams containing coupling constant counterterm insertions and give \begin{eqnarray} \Delta q^{(2)}_{{\rm muon}}&=&\left(\frac{\alpha_r}{\pi}\right)^2 \left(\frac{16987}{576}-\frac{85}{36}\zeta(2) -\frac{64}{3}\zeta(3)\right)\\ &=&-\left(\frac{\alpha_r}{\pi}\right)^2 0.0364333. \end{eqnarray} This was calculated both by means of dispersion relations in the on-shell scheme \cite{muonhad} and by the perturbative methods described in section~\ref{sec:calcul} in the $\overline{{\rm MS}}$ renormalization scheme. Full agreement was found. The hadronic contribution to (\ref{eq:Deltaq2}) is given in Ref.\cite{muonhad}. Notice that, because it was calculated using dispersion relations, it naturally involves a subtraction of the photon vacuum polarization at $q=0$ which corresponds to the on-shell scheme being adopted for electric charge renormalization. The same is true for the stated result for tau loops of Eq.~(\ref{eq:Deltaq2tau}). On the other hand the photonic and electronic contributions, (\ref{eq:Deltaq2photon}) and (\ref{eq:Deltaq2elec}), were calculated perturbatively assuming, where necessary, the $\overline{{\rm MS}}$ renormalization scheme. It will be shown in section~\ref{sec:AlphaRenorm} that the two sets of results can be combined in a consistent manner by setting renormalized electromagnetic coupling constant, $\alpha_r=\alpha_e(m_\mu)$=1/135.90. In doing so all corrections up to ${\cal O}(\alpha^2)$, ${\cal O}\left(\alpha^3\ln(m_e^2/m_\mu^2)\right)$ and ${\cal O}\left(\alpha^i\ln^{i-1}(m_e^2/m_\mu^2)\right)$ for all $i\ge2$ are taken into account. Fermion mass renormalization is always performed in the on-shell renormalization scheme in which the renormalized mass of a stable fermion is set equal to its pole mass. $\Delta q^{(2)}$ of Eq.~(\ref{eq:Deltaq2}), when combined with the tiny contribution of tau loops gives Eq.~(\ref{eq:Deltaq2tau}), \begin{equation} \Delta q^{(2)} = \Gamma_0\left(\frac{\alpha_e(m_\mu)}{\pi}\right)^2 (6.700\pm 0.002). \end{equation} When this is used in Eq.~(\ref{eq:QEDcorr}) along with the current best value for the muon lifetime, $\tau_\mu=(2.19703\pm0.00004)\,\mu$s \cite{PDG}, it gives a new value for the Fermi coupling constant \begin{xxalignat}{2} {}\qquad\qquad\qquad\qquad G_F&=(1.16637\pm0.00001)\times10^{-5}\,{\rm GeV^{-2}}& (9\,{\rm ppm})& \end{xxalignat} This represents a reduction in the overall error on $G_F$ of about a factor of 2 and a downward shift in the central value of twice the experimental uncertainty. The error is now entirely experimental. \section{The Definition of the Fermi Coupling Constant} \label{sec:Definition} At high precision it is essential to have a clear and unambiguous definition for the Fermi coupling constant, $G_F$. For the present purposes $G_F$ will taken to be defined by Eq.(\ref{eq:QEDcorr}) \begin{equation} \tau^{-1}_\mu \equiv \frac{G_F^2 m_\mu^5}{192\pi^3} \left( 1+\Delta q \right) \end{equation} with $\Delta q$ being calculated using the Fermi Theory Lagrangian ${\cal L}_F$ of Eq.(\ref{eq:FullLagrangian}). This expresses $G_F$ to arbitrary accuracy in terms of the physically observable quantities, $\tau_\mu$, $\alpha$, $m_\mu$ and $m_e$. In this way the infrared structure, which involves both real bremsstrahlung and virtual corrections, is entirely relegated to a single constant $\Delta q$ and does not have to be dealt with when $G_F$ is eventually used in the calculation of other physical constants such as the $W$ boson mass, $M_W$. Thus the, presumably well-understood, contribution of QED to the muon lifetime is eliminated from $G_F$ which is left with an enriched contribution from weak physics. $G_F$ is {\bf sometimes } defined via the formula \cite{PDG} \begin{equation} \tau^{-1}_\mu=\frac{G_F^2 m_\mu^5}{192\pi^3}F\left(\frac{m_e^2}{m_\mu^2}\right) \left(1+\frac{3}{5}\frac{m_\mu^2}{M_W^2}\right) \left[1+\frac{\alpha(m_\mu)}{2\pi} \left(\frac{25}{4}-\pi^2\right)\right] \label{eq:taumudef} \end{equation} where \[ F(x)=1-8x-12x^2\ln x+8x^3-x^4 \] and \[ \alpha(m_\mu)^{-1}=\alpha^{-1} -\frac{2}{3\pi}\ln\left(\frac{m_e}{m_\mu}\right) +\frac{1}{6\pi}\sim136. \] In as far as this can be derived from ${\cal L}_F$ it provides an adequate definition but it contains a number of features that become out of place at higher orders. The function $F$, coming from phase space integration, does not factorize in this way at higher orders. The factor $\left(1+(3m_\mu^2)/(5M_W^2)\right)$ is the effect of the $W$ boson propagator and is not generated by the ${\cal L}_F$. It is more naturally included along with the weak corrections in $\Delta r$ as is described in section~\ref{sec:WeakCorr}. Its presence is an historical artifact of an attempt to reconcile the observed muon and neutron beta decay rates with universality of weak interactions before the advent of the Cabibbo angle \cite{BermSirl,Hofstadter}. In that scenario a light $W$ boson with a mass slightly heavier than the kaon was needed so as to forbid the unobserved decay mode, $K^\pm\rightarrow W^\pm\gamma$. The expression for $\alpha(m_\mu)$ used in the above definition contains a term $1/(6\pi)$ that comes from $W$ boson loops in the photon self-energy. As these do not come from ${\cal L}_F$ they should be omitted so as not to risk double counting when $G_F$ is used as input in electroweak calculations. It is sometimes suggested that Eq.(\ref{eq:taumudef}) should be viewed as an exact definition of $G_F$. This has significant drawbacks and hard to justify now that the 2-loop QED corrections are available \cite{muonprl,muonhad}. It is already the case that these 2-loop QED corrections are larger than the current experimental error and are roughly an order of magnitude greater than those anticipated in the next generation of experiments (see section~\ref{sec:ExpUncert}). If Eq.~(\ref{eq:taumudef}) is taken to be exact then these 2-loop corrections will ultimately have to be incorporated in formulas that relate $G_F$ to other electroweak observables such as the mass of the $W$ boson, $M_W$. This is extremely inconvenient as one would prefer to deal with quantities from which QED has been eliminated as far as possible. \section{Renormalization of the Electromagnetic Coupling Constant} \label{sec:AlphaRenorm} In this section it will be shown how to set up the $\overline{{\rm MS}}$ renormalization scheme so as to obtain the leading logarithmic corrections to the muon lifetime of ${\cal O}\left(\alpha^3\ln(m_\mu^2/m_e^2)\right)$ and of ${\cal O}\left(\alpha^i\ln^{i-1}(m_\mu^2/m_e^2)\right)$ for all $i>0$. It is also shown how to incorporate hadronic contributions \cite{muonhad}, calculated in the on-shell renormalization scheme, in a consistent manner. In any self-consistent calculation using the Lagrangian (\ref{eq:FullLagrangian}), the quantity $\alpha$ that appears in the $\Delta q^{(i)}$ is the renormalized electromagnetic coupling constant, to be denoted $\alpha_r$, in whatever renormalization scheme has been chosen. It is generally convenient to adopt the on-shell renormalization scheme for the fermion masses which will be assumed throughout unless otherwise stated. No such restriction is placed on the coupling constant renormalization. For problems containing widely disparate scales, renormalization of the coupling constant in the $\overline{{\rm MS}}$ scheme is to be preferred as it automatically absorbs the dominant logarithmic corrections into $\alpha_r$ at the outset and avoids the need for resummation of large logarithms coming from higher-order contributions as is required when the on-shell renormalization scheme is adopted. In QED the numerical value of $\alpha_r$ is by obtained solving the equation \begin{equation} \alpha=\frac{\alpha_r}{1-\widehat\Pi_{\gamma\gamma}^\prime(0)}. \label{eq:alphadef} \end{equation} where $\alpha$ on the right-hand side of Eq.(\ref{eq:alphadef}) is the experimentally-measured quantity, $\alpha=1/137.0359895(61)$ \cite{PDG}. $\widehat\Pi_{\gamma\gamma}^\prime(0)$ is the photon vacuum polarization function which itself may be written as expansion in $\alpha_r$, \begin{equation} \widehat\Pi_{\gamma\gamma}^\prime(0) =\sum_{i=1}^\infty\widehat\Pi_{\gamma\gamma}^{\prime(i)}(0) \end{equation} and each term receives contributions from all fermion species. The hat, \ $\widehat{}$\ , indicates that counterterm contributions in the chosen renormalization scheme have been included. In the on-shell renormalization scheme the counterterms are adjusted so that $\widehat\Pi_{\gamma\gamma}^\prime(0)=0$ and it follows from Eq.(\ref{eq:alphadef}) that the renormalized coupling constant in this scheme satisfies, $\alpha_r\equiv\alpha_{{\rm OS}}=\alpha$. In the $\overline{{\rm MS}}$ renormalization scheme the counterterms are chosen to contain only divergent pieces plus certain uninteresting constants. In particular 1-loop counterterms are just proportional to $\Delta= 1/\varepsilon -\gamma_E +\ln4\pi + O(\varepsilon)$ where $D=4-2\varepsilon $ is the dimensionality of spacetime and $\gamma_E=0.57721566...$ is Euler's constant. In this scheme an appropriate choice for the 't~Hooft mass is $\mu=m_\mu$ and we write $\alpha_r\equiv\alpha(m_\mu)$. Defining $\widehat\Pi_{\gamma\gamma}^{\prime(i)}(0)=\alpha_r^i P^{(i)}$ and solving Eq.(\ref{eq:alphadef}) for $\alpha_r$ yields \begin{equation} \alpha_r=\alpha-\alpha^2 \widehat P^{(1)} -\alpha^3 \widehat P^{(2)}+{\cal O}(\alpha^4) \label{eq:alphaRsoln} \end{equation} The contribution of leptons to $\widehat\Pi_{\gamma\gamma}^{\prime(i)}(0)$ can be exactly calculated in perturbation theory using dimensional regularization \cite{dimreg,dimreg2}. The 1-loop contribution is \begin{eqnarray} \widehat\Pi_{\gamma\gamma}^{\prime(1)}(0)&=& -\frac{\alpha_r}{3\pi}(4\pi)^{2-\frac{D}{2}} \Gamma\left(2-\frac{D}{2}\right) \sum_l Q_l^2\left(\frac{m_l^2}{\mu^2}\right)^{\frac{D}{2}-2} +2\frac{\delta e^{(1)}_l}{e}\\ &=&-\frac{\alpha_r}{3\pi}\sum_l Q_l^2\left(\Delta-\ln \frac{m_l^2}{\mu^2}\right) +2\frac{\delta e^{(1)}_l}{e}+{\cal O}(D-4) \label{eq:PiAA1} \end{eqnarray} where $\delta e^{(1)}_l$ is the leptonic contribution to the 1-loop electromagnetic charge counterterm. At this order the $\overline{{\rm MS}}$ and on-shell renormalization schemes differ only in the the finite parts of their coupling constant counterterms, $\delta e^{(1)}_{\overline{{\rm MS}}}$ and $\delta e^{(1)}_{{\rm OS}}$ respectively. Notice that the contribution from muon loops to $\delta e^{(1)}_{\overline{{\rm MS}}}$ and $\delta e^{(1)}_{{\rm OS}}$ is identical for $\mu=m_\mu$ and thus their overall contribution to $\widehat\Pi_{\gamma\gamma}^{\prime(1)}(0)$ vanishes in both schemes. The 2-loop correction in a general renormalization scheme can be obtained from Ref.\cite{MaldeStuart1} taking into account the difference in measure used to define dimensional regularization \begin{multline} \widehat\Pi_{\gamma\gamma}^{\prime(2)}(0)= \frac{\alpha_r^2}{12\pi^2} \frac{(5D^2-33D+34)}{D(D-5)}(4\pi)^{4-D} \Gamma\left(3-\frac{D}{2}\right)\Gamma\left(2-\frac{D}{2}\right) \sum_l Q_l^4 \left(\frac{m_l^2}{\mu^2}\right)^{D-4}\\ +\frac{2\alpha_r}{3\pi}(4\pi)^{2-\frac{D}{2}}\Gamma\left(3-\frac{D}{2}\right) \sum_l Q_l^2\frac{\delta m_l^{(1)}}{m_l} \left(\frac{m_l^2}{\mu^2}\right)^{\frac{D}{2}-2} +2\frac{\delta e^{(2)}_l}{e} \label{eq:GeneralPi2} \end{multline} where $\delta m_l^{(1)}$ is the 1-loop lepton mass counterterm and $\delta e^{(2)}_l$ is the leptonic part of the 2-loop electromagnetic charge counterterm. Adopting the on-shell renormalization scheme for fermion masses gives \begin{equation} \label{masscounterinsert} \frac{\delta m_l^{(1)}}{m_l}=-\frac{\alpha_r}{4\pi} Q_l^2\left(\frac{m_l^2}{\mu^2}\right)^{\frac{D}{2}-2} \frac{(D-1)}{(D-3)} (4\pi)^{2-\frac{D}{2}} \Gamma\left(2-\frac{D}{2}\right) \end{equation} and substituting this in Eq.(\ref{eq:GeneralPi2}) yields \begin{eqnarray} \widehat\Pi_{\gamma\gamma}^{\prime(2)}(0)&=& \frac{\alpha^2_r}{4\pi^2} \frac{(D^3-12D^2+41D-34)}{D(D-3)(D-5)}(4\pi)^{4-D} \Gamma\left(3-\frac{D}{2}\right)\Gamma\left(2-\frac{D}{2}\right) \nonumber\\ & &\qquad\qquad\qquad \times\sum_l Q_l^4 \left(\frac{m_l^2}{\mu^2}\right)^{D-4} +2\frac{\delta e^{(2)}_l}{e}\\ &=&-\frac{\alpha_r^2}{4\pi^2} \sum_lQ_l^4\left(\Delta-\ln\frac{m_l^2}{\mu^2}+\frac{15}{4}\right) +2\frac{\delta e^{(2)}_l}{e}+{\cal O}(D-4) \end{eqnarray} whose leading logarithms are in agreement with the well-known result of Jost and Luttinger \cite{JostLuttinger}. The leading logarithms of ${\cal O}(\alpha^3)$ were obtained by Rosner \cite{Rosner}. The contributions to muon lifetime from hadron, muon and tau loops have been calculated using dispersion relations \cite{muonhad}. This method naturally generates a subtraction at momentum $q=0$ which corresponds to on-shell renormalization of the coupling constant. The calculation of Ref.\cite{muonhad} was performed by effectively replacing the photon propagators \begin{equation} \frac{g_{\mu\nu}}{q^2+i\epsilon}\longrightarrow \frac{g_{\mu\sigma}}{q^2+i\epsilon} (q^2 g_{\sigma\tau}-q_\sigma q_\tau) \widehat\Pi_{\gamma\gamma}^{\prime(1)}(q^2) \frac{g_{\tau\nu}}{q^2+i\epsilon}. \end{equation} The on-shell results of Ref.\cite{muonhad} are converted to $\overline{{\rm MS}}$ by adding a correction obtained by replacing the photon propagator by \begin{equation} \frac{g_{\mu\nu}}{q^2+i\epsilon}\longrightarrow \frac{g_{\mu\sigma}}{q^2+i\epsilon} (q^2 g_{\sigma\tau}-q_\sigma q_\tau) \frac{2}{e}\left(\delta e^{(1)}_{\overline{{\rm MS}}} -\delta e^{(1)}_{{\rm OS}}\right) \frac{g_{\tau\nu}}{q^2+i\epsilon}. \label{eq:MStoOS} \end{equation} where $\delta e^{(1)}_{\overline{{\rm MS}}}-\delta e^{(1)}_{{\rm OS}}$ is a finite constant that includes the effects of loops of all fermion species except the electron. As usual $\epsilon$ is a positive infinitesimal. When (\ref{eq:MStoOS}) is applied to the calculation of the muon lifetime the terms on the right hand side proportional to $q_\mu q_\nu$ cancel between diagrams hence its effect amounts to an overall multiplicative factor for the photon propagator and leads to a correction to $\Gamma_\mu$ of \[ \Delta\Gamma_{{\rm OS}\rightarrow\overline{{\rm MS}}} =\Gamma_0 \Delta q^{(1)} \frac{2}{e}\left(\delta e^{(1)}_{\overline{{\rm MS}}} -\delta e^{(1)}_{{\rm OS}}\right) \] This can be accounted for by a redefinition of the renormalized coupling constant, $\alpha_r$ of Eq.(\ref{eq:alphaRsoln}), which has the effect of eliminating the contributions of all fermion loops except those of the electron. It follows that the results of Ref.\cite{muonhad} can be used directly in the calculation of the muon lifetime provided that the $\overline{{\rm MS}}$ value $\alpha(\mu)$ is replaced by \begin{equation} \alpha(\mu)\longrightarrow \alpha_e(\mu)=\alpha +\alpha^2\left(\frac{1}{3\pi}+\frac{\alpha}{4\pi^2}\right) \ln\frac{\mu^2}{m_e^2}+\frac{15\alpha^3}{16\pi^2}. \label{eq:alphae} \end{equation} The KLN theorem \cite{KinoLeeNaue} requires that the only electron mass singularities contributing the muon lifetime are those that appear in (\ref{eq:alphae}) which arise as a consequence of $\alpha$'s having been defined at the exceptional momentum, $q^2=0$. Moreover, careful examination of Eq.(\ref{eq:alphadef}) reveals that writing \begin{equation} \alpha_e(m_\mu)=\frac{\alpha} {1-\frac{\displaystyle \alpha}{\displaystyle 3\pi} \ln\frac{\displaystyle m_\mu^2}{\displaystyle m_e^2}} +\frac{\alpha^3}{4\pi^2}\ln\frac{m_\mu^2}{m_e^2} \label{eq:alphaefinal} \end{equation} correctly resums logarithms of the form $\alpha^n\ln^{n-1}(m_\mu^2/m_e^2)$ for all $n>0$ and incorporates those of ${\cal O}(\alpha^3\ln(m_\mu^2/m_e^2))$ without generating spurious higher-order logarithms as there would be if the second term on the right hand side of Eq.(\ref{eq:alphaefinal}) where moved into the denominator of the first. The constant term of ${\cal O}(\alpha^3)$ has been dropped since terms of this order have not been accounted for in the present calculation. Terms that are not logarithmically enhanced are correct up to ${\cal O}(\alpha^2)$. For the extraction of $G_F$ in the Fermi theory the appropriate choice is $\mu=m_\mu$ which gives \[ \alpha_e(m_\mu)=1/135.90=0.0073582 \] however when $G_F$ is used in the analysis of electroweak data obtained at the weak scale $\mu=M_Z$ is generally a more convenient choice. The expression for $\alpha_e(m_\mu)$ could also be obtained be adopting a hybrid renormalization scheme in which electron loops were renormalized in $\overline{{\rm MS}}$ and all other fermions loops in the on-shell scheme. This is allowable in QED since the individual fermions are not connected by any gauge symmetry and provides a simple, if unconventional, way of deriving Eq.(\ref{eq:alphaefinal}). \section{The Calculation of the 2-loop QED Contributions} \label{sec:calcul} In this section the evaluation of Feynman diagrams that appear in the calculation of $\Delta q^{(2)}$ Eqs.\,(\ref{eq:DeltaqSeries}), (\ref{eq:Deltaq2}) is discussed. For the calculation of the 25 muon decay diagrams of Fig.\ref{4loopdiagrams} it is necessary to deal with only 7 basic 4-loop topologies labeled A--G (see Figs.\ref{TopoAFigs}--\ref{TopoDEFGFigs}). These 7 basic topologies are assumed to have scalar propagators raised to arbitrary (positive and negative) integer powers. The integrals are evaluated on the mass shell $Q^2= m_\mu^2$ using dimensional regularization \cite{dimreg,dimreg2} for both the UV and IR divergences. Within dimensional regularization a few integrals are known in a simple closed form for arbitrary powers of the propagators as a ratio of Euler $\Gamma$ functions. These are listed in Appendix~\ref{sec:multloop}. The imaginary parts must be obtained for the 4-loop topologies. Those 4-loop integrals that, on general grounds, have no imaginary part will be discarded along the way. This includes 4-loop vacuum bubble integrals and 4-loop integrals with a through-going on-shell line, as these have no cuts that can generate an imaginary part. To reduce the burden of analytic integration as much as possible we use the well-known method of integration-by-parts \cite{parialint} in dimensional regularization to express all integrals of a given topology, having arbitrary integer powers of the propagators, in terms of a small set of primitive integrals. Using integration-by-parts has the advantage of requiring the explicit calculation of only a fixed number of primitive integrals instead of hundreds of scalar integrals that generally appear in a multiloop calculation. On-shell integration-by-parts relations were used at the two-loop level for the first time in Ref.\cite{msbarpolemass} (but see also \cite{shell2}) and at the three-loop level in Ref.\cite{gmin2at3loops}. A given set of primitive integrals is not unique as one can eliminate one set in favor of another set of integrals, as long as they are independent with respect to the integration-by-parts identities. Therefore only the total number of primitive integrals is a fixed quantity for a given topology, and what is used as a primitive integral is a matter of convenience. Integration-by-parts relations follow from calculating the derivative in the identity \begin{equation} \label{integratebyparts} \int {\rm d}^Dp\; \frac{\partial}{\partial p^\mu} \; \left[ k^\mu f(p,l_1,\cdots, l_n) \right]=0 \end{equation} within dimensional regularization. Here $p$ is any momentum that is integrated over, $k\in \{ p,l_1,\cdots, l_n \} $ and $f$ is a scalar function that may depend in addition to $p$ on a set of momenta $l_i$. \begin{figure} \begin{center} \begin{picture}(230,60)(0,0) \SetOffset(0,10) \SetWidth{.5} \Line(0,21.65063510)(25,21.65063510) \Line(0,21.65063510)(12.5,0) \Line(25,21.65063510)(12.5,0) \Line(12.5,0)(12.5,-15) \Line(25,21.65063510)(40,34.64101616) \Line(0,21.65063510)(-15,34.64101616) \Text(12.5,27)[m]{\scriptsize 3} \Text(23,9)[m]{\scriptsize 1} \Text(2,9)[m]{\scriptsize 2} \Text(18,-12)[m]{\scriptsize 5} \Text(40,28)[m]{\scriptsize 4} \Text(-15,28)[m]{\scriptsize 6} \Text(-14,-18)[m]{$T_{0,0,0}$} \LongArrow(52,27)(42,19) \Text(53,21)[m]{\footnotesize $l$} \LongArrow(35,13)(29,2) \Text(38,6)[m]{\footnotesize $p$} \LongArrow(28,-8)(28,-19) \Text(34,-12)[m]{\footnotesize $k$} \SetOffset(110,10) \SetWidth{2} \Line(0,21.65063510)(25,21.65063510) \SetWidth{.5} \Line(0,21.65063510)(12.5,0) \Line(25,21.65063510)(12.5,0) \Line(12.5,0)(12.5,-15) \SetWidth{2} \Line(25,21.65063510)(40,34.64101616) \SetWidth{.5} \Line(0,21.65063510)(-15,34.64101616) \Text(12.5,27)[m]{\scriptsize 3} \Text(23,9)[m]{\scriptsize 1} \Text(2,9)[m]{\scriptsize 2} \Text(18,-12)[m]{\scriptsize 5} \Text(40,28)[m]{\scriptsize 4} \Text(-15,28)[m]{\scriptsize 6} \Text(-14,-18)[m]{$T_{m,0,0}$} \SetOffset(220,10) \SetWidth{2} \Line(0,21.65063510)(25,21.65063510) \Line(0,21.65063510)(12.5,0) \SetWidth{.5} \Line(25,21.65063510)(12.5,0) \SetWidth{2} \Line(12.5,0)(12.5,-15) \Line(25,21.65063510)(40,34.64101616) \SetWidth{.5} \Line(0,21.65063510)(-15,34.64101616) \Text(12.5,27)[m]{\scriptsize 3} \Text(23,9)[m]{\scriptsize 1} \Text(2,9)[m]{\scriptsize 2} \Text(18,-12)[m]{\scriptsize 5} \Text(40,28)[m]{\scriptsize 4} \Text(-15,28)[m]{\scriptsize 6} \Text(-14,-18)[m]{$T_{m,m,0}$} \end{picture} \end{center} \caption{ \label{TriangleFigs} Examples of triangle diagrams with different mass configurations for which the triangle relation Eq.~(\ref{defTriangleRel}) holds. The lines are numbered according to Eq.~(\ref{defTriangle}). } \end{figure} An important example of an integration-by-parts identity is the triangle relation for one-loop 3-point integrals \cite{parialint}. Let us define for masses $a,b,c$ \begin{multline} T_{a,b,c} (\alpha_1,\alpha_2,\alpha_3,\alpha_4,\alpha_5,\alpha_6 ) = \int {\rm d}^Dp\; \Biggl( \frac{ 1}{(p^2)^{\alpha_1} [(p-k)^2-a^2]^{\alpha_2} [(p-l)^2-b^2]^{\alpha_3} } \Biggr)\\ \times \frac{ 1}{(l^2-b^2)^{\alpha_4} (k^2-a^2)^{\alpha_5} [(k-l)^2-c^2]^{\alpha_6} } \label{defTriangle} \end{multline} Here and below $+i\epsilon$ in the propagators is not written explicitly to keep expressions compact (e.g. $1/p^2$ is understood to be $1/(p^2+i\epsilon)$). By taking the derivative in the integral \[ \int {\rm d}^Dp\; \frac{\partial}{\partial p^\mu} \Biggl( \frac{ p^\mu }{(p^2)^{\alpha_1} [(p-k)^2-a^2]^{\alpha_2} [(p-l)^2-b^2]^{\alpha_3} } \Biggr)=0 \] one obtains the triangle relation \begin{equation} \label{defTriangleRel} \biggl[ ( D -2\alpha_1 - \alpha_2 - \alpha_3) - \alpha_2 {\bf 2^+} ( {\bf 1^-} - {\bf 5^-} ) - \alpha_3 {\bf 3^+} ( {\bf 1^-} - {\bf 4^-} ) \biggr] T_{a,b,c} (\alpha_1,\alpha_2,\alpha_3,\alpha_4, \alpha_5,\alpha_6) = 0 \end{equation} where the notation of Ref.\cite{parialint} has been adopted. Thus the operator ${\bf 2}^{+}$ raises the power $\alpha_2$ in $T$ by one, ${\bf 4}^{-}$ lowers $\alpha_4$ in $T$ by one, etc.. By repeated use of this relation on integrals $T$ with positive integer powers $\alpha_1,\cdots,\alpha_6$ one is able to lower at least one of the powers $\alpha_1$, $\alpha_4$ or $\alpha_5$ to zero. Instead of applying the triangle relations recursively, the result can be obtained directly and expressed as a 3-fold nested sum \cite{resolvedtriangle}. Note that the configuration of masses in Eq.~(\ref{defTriangle}) is not completely general, and if all propagators carry different masses additional terms appear in the triangle relation, leading to a less useful identity. However, when a triangle with the mass configuration of $T$ appears as a subgraph of a multiloop integral, one can use the triangle relation to simplify the multiloop integrals considerably. Some subgraphs relevant to the present calculation are given in Fig.\ref{TriangleFigs} To express multiloop integrals of a certain topology in terms of primitive integrals, one generally needs to consider all integration-by-parts identities obtained from Eq.~(\ref{integratebyparts}) and solve them. To solve these relations there are two possible ways to proceed\footnote{It should be mentioned that new approaches to solving integration-by-parts identities are being developed as well \cite{newibpsolution} }. Integration-by-parts relations may be explicitly applied to a large set integrals with specific integer powers $\{\cdots,-2,-1,0,1,2, \cdots \}$ and solved the resulting set of linear equations (see, for example, Ref.\cite{gmin2at3loops}). Alternatively the relations may be considered as symbolic equations with unspecified powers of the propagators and combined into a set of recurrence relations that, after repeated application, express integrals in terms of primitives \cite{parialint,msbarpolemass}. Both approaches can have advantages in particular circumstances. The second approach is more commonly used as one may encounter integrals with many different powers in applications that involve expansions of diagrams. For the present work the recursive approach was used and a recurrence scheme was implemented in a FORM \cite{form} program. As some steps in the recurrence relations involve hundreds of terms we limit ourselves to the main conclusions for each of the basic 4-loop topologies and give a list of the primitive integrals were chosen. The calculation of these primitive integrals is discussed in the next section. \begin{figure} \begin{center} \begin{picture}(350,70)(0,0) \SetOffset(0,0) \SetWidth{2} \Line(30,30)(45,30) \Line(105,30)(164,30) \SetWidth{.5} \CArc(45,57)(27,270,353) \Oval(75,30)(24,30)(0) \CArc(102,39.0)(32,248,343) \Text(48,48)[m]{\scriptsize 1} \Text(100,50)[m]{\scriptsize 2} \Text(69,34)[m]{\scriptsize 3} \Text(63,14)[m]{\scriptsize 4} \Text(98,22)[m]{\scriptsize 5} \Text(126,10)[m]{\scriptsize 6} \Text(117,35)[m]{\scriptsize 7} \Text(30,4)[m]{ A$^{(3)}$} \SetOffset(165,0) \SetWidth{2} \Line(30,30)(45,30) \Line(105,30)(164,30) \SetWidth{.5} \CArc(45,57)(27,270,353) \Oval(75,30)(24,30)(0) \CArc(105,43.8)(43.800,238,340) \CArc(107.4,34.8)(21,246,343) \Text(48,48)[m]{\scriptsize 1} \Text(100,50)[m]{\scriptsize 2} \Text(69,34)[m]{\scriptsize 3} \Text(63,14)[m]{\scriptsize 4} \Text(89,14.5)[m]{\scriptsize 5} \Text(99,24)[m]{\scriptsize 6} \Text(131,2)[m]{\scriptsize 7} \Text(120,12)[m]{\scriptsize 8} \Text(117,35)[m]{\scriptsize 9} \Text(137,35)[m]{\scriptsize 10} \Text(38,4)[m]{ A} \end{picture} \end{center} \caption{ \label{TopoAFigs} Integration topologies A$^{(3)}$ and A. The lines are numbered according to Eq.(\ref{defTopoA3}) and Eq.~(\ref{defTopoA}) respectively } \end{figure} For the 3-loop topology A$^{(3)}$ of Fig.\ref{TopoAFigs}, it is possible to lower the power of at least one propagator to zero through the use of triangle identities. It is assumed that a tensor reduction on the massless 1-loop sub-diagram formed by lines 1 and 3 is performed, such that for A$^{(3)}$, all invariants can be expressed in terms of propagators. Integration-by-parts recurrence relations express all integrals of this topology in terms of one primitive integral with non-zero cuts. Here and below we do not count integrals that can be calculated by repeated application of the closed expressions Eqs.(\ref{1loopbubble})--(\ref{bubblem00}). For the normalization of the primitive integrals we define \begin{equation} \label{defnormloop} N_\varepsilon = \frac{\pi^2}{ (\pi s)^{\varepsilon}} \frac{\Gamma^2 (1-\varepsilon) \Gamma(1+\varepsilon) } {\Gamma(1-2\varepsilon) } \end{equation} where $\varepsilon$ is related to the space-time dimension as $D=4-2\varepsilon$. \begin{multline} I_{A^{(3)}} (\alpha_1,\alpha_2,\alpha_3,\alpha_4,\alpha_5,\alpha_6,\alpha_7) = i\; \int\int\int {\rm d}^Dp\; {\rm d}^Dk\; {\rm d}^Dl\; \Biggl( \frac{ 1}{(p^2)^{\alpha_1} [(k+Q)^2]^{\alpha_2} }\\ \times \frac{1}{ [(p-k-Q)^2]^{\alpha_3} (k^2)^{\alpha_4} [(k-l)^2]^{\alpha_5} (l^2)^{\alpha_6} (l^2+2l\cdot Q)^{\alpha_{7}} } \Biggr) \label{defTopoA3} \end{multline} \begin{multline} \Im( I_{A^{(3)}}(1,0,1,1,1,0,1) ) = \pi\, s\, (N_\varepsilon)^3 \biggl[ \frac{1}{\varepsilon} \Bigl( -\frac{1}{2} \Bigr) +\Bigl( -\frac{11}{2} + \zeta_2 \Bigr) + \varepsilon \Bigl( -\frac{77}{2} + \frac{15}{2}\zeta_2 + 8 \zeta_3 \Bigr) \\ + \varepsilon^2 \Bigl( - \frac{439}{2} + \frac{81}{2}\zeta_2 + 50 \zeta_3 + \frac{69}{2}\zeta_4 \Bigr) + {\cal O}(\varepsilon^3) \biggr] \end{multline} where for compactness the notation $\zeta_n\equiv\zeta(n)$ has been adopted. For 4-loop topology A of Fig.\ref{TopoAFigs} one can lower the power of at least one propagator to zero through the use of triangle identities. For all 4-loop topologies A--G of Figs.\ref{TopoAFigs}--\ref{TopoDEFGFigs} we assume that a tensor reduction on the massless 1-loop subdiagram formed by lines 1 and 3 is performed, such that one has to deal with only one invariant that cannot be expressed in terms of propagators. For topology A this invariant is chosen to be $p\cdot Q$. Integration-by-parts recurrence relations express all integrals of this topology in terms of 9 primitive integrals with a non-zero imaginary part. Again we do not count integrals that can be calculated by repeated application of the closed expressions Eqs.(\ref{1loopbubble})--(\ref{bubblem00}). \begin{multline} I_A (\alpha_1,\alpha_2,\alpha_3,\alpha_4,\alpha_5,\alpha_6,\alpha_7,\alpha_8 ,\alpha_9,\alpha_{10},\alpha_{11}) =\\ \int\int\int\int {\rm d}^Dp\; {\rm d}^Dk\; {\rm d}^Dl\; {\rm d}^Dr \; \Biggl( \frac{ (p\cdot Q)^{\alpha_{11}}}{(r^2)^{\alpha_1} [(k+Q)^2]^{\alpha_2} [(r-k-Q)^2]^{\alpha_3} (k^2)^{\alpha_4} [(k-l)^2]^{\alpha_5} }\\ \times \frac{1}{ [(k-l-p)^2]^{\alpha_6} (l^2)^{\alpha_7} (p^2)^{\alpha_8} [(l+p)^2 +2(l+p)\cdot Q]^{\alpha_9} (l^2+2l\cdot Q)^{\alpha_{10}} } \Biggr) \label{defTopoA} \end{multline} \begin{eqnarray*} \Im( I_A(1,1,1,1,0,1,1,1,1,1,0) )&=&\textstyle \frac{ \pi}{s} (N_\varepsilon)^4 \left[ \frac{1}{\varepsilon} \Bigl( - 5 \zeta_4 \Bigr) + {\cal O}(\varepsilon^0) \right] \\ \Im( I_A(1,1,1,0,1,0,1,1,1,0,0) )&=&\textstyle \pi\, s\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon^2}\Bigl( - \frac{1}{2}\Bigr) + \frac{1}{\varepsilon} \Bigl( - \frac{31}{4} + \zeta_2 \Bigr) + \Bigl( - \frac{591}{8} + \frac{29}{2}\zeta_2 + 9 \zeta_3 \Bigr) \\ & &\textstyle + \varepsilon \Bigl( - \frac{8979}{16} + \frac{533}{4} \zeta_2 + \frac{209}{2} \zeta_3 + \frac{65}{2} \zeta_4 \Bigr) + {\cal O}(\varepsilon^2) \biggr] \\ \Im( I_A(1,0,1,1,0,1,0,1,0,1,0) )&=&\textstyle \pi\, s^2\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon^2}\Bigl( -\frac{1}{4} \Bigr) + \frac{1}{\varepsilon} \Bigl( -\frac{7}{3} \Bigr) + \Bigl( - \frac{1703}{144} - \zeta_2 \Bigr) \\ & &\hspace{-4cm}\textstyle +\varepsilon \Bigl( - \frac{3697}{108} - \frac{119}{12} \zeta_2 - \frac{17}{2} \zeta_3 \Bigr) +\varepsilon^2 \Bigl( \frac{159001}{5184} - \frac{2053}{36} \zeta_2 - \frac{511}{6} \zeta_3 - \frac{263}{4} \zeta_4 \Bigr) + {\cal O}(\varepsilon^3) \biggr] \\ \Im( I_A(1,1,1,1,0,1,0,1,0,1,0) )&=&\textstyle \pi\, s\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon^3}\Bigl( \frac{1}{2} \Bigr) + \frac{1}{\varepsilon^2} \Bigl( 4 \Bigr) + \frac{1}{\varepsilon} \Bigl( \frac{165}{8} \Bigr) + \Bigl(\frac{343}{4} + \frac{3}{2} \zeta_2 - 3 \zeta_3 \Bigr) \\ & &\textstyle + \varepsilon \Bigl( \frac{9749}{32} + \frac{25}{2} \zeta_2 - 9 \zeta_3 - \frac{17}{2} \zeta_4 \Bigr) + {\cal O}(\varepsilon^2) \biggr]\\ \Im( I_A(1,0,1,0,1,1,1,0,1,0,0) )&=&\textstyle \pi\, s^2\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon} \Bigl( \frac{1}{12} \Bigr) + \Bigl( \frac{49}{18} - \zeta_2 \Bigr) + \varepsilon \Bigl( \frac{17161}{432} - \frac{127}{12} \zeta_2 - \frac{25}{2}\zeta_3 \Bigr) \\ & &\textstyle + \varepsilon^2 \Bigl( \frac{260347}{648} - \frac{2837}{36} \zeta_2 - \frac{376}{3} \zeta_3 - \frac{319}{4} \zeta_4 \Bigr) + {\cal O}(\varepsilon^3) \biggr] \\ \Im( I_A(1,1,1,0,1,1,1,0,1,0,0) )&=&\textstyle \pi\, s\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon^2} \Bigl( -\frac{1}{2}\Bigr) + \frac{1}{\varepsilon} \Bigl( -6 \Bigr) +\Bigl( - \frac{385}{8} + \frac{11}{2}\zeta_2 + \zeta_3 \Bigr) \\ & &\textstyle + \varepsilon \Bigl( - \frac{651}{2} + \frac{121}{2}\zeta_2 + 36 \zeta_3 + \frac{19}{2} \zeta_4 \Bigr) + {\cal O}(\varepsilon^2) \biggr] \\ \Im( I_A(1,0,1,1,1,1,0,0,1,1,0) )&=&\textstyle \pi\, s\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon^2} \Bigl( -\frac{1}{2} \Bigr) + \frac{1}{\varepsilon} \Bigl( - \frac{31}{4} + 2 \zeta_2 \Bigr) + \Bigl( - \frac{591}{8} + 15 \zeta_2 + 22 \zeta_3 \Bigr) \\ & &\textstyle + \varepsilon \Bigl( - \frac{8979}{16} + \frac{155}{2} \zeta_2 + 146 \zeta_3 + 155 \zeta_4 \Bigr) + {\cal O}(\varepsilon^2) \biggr] \\ \Im( I_A(1,0,1,1,0,1,1,1,1,0,0) )&=&\textstyle \pi\, s\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon^2} \Bigl( -\frac{1}{4} \Bigr) + \frac{1}{\varepsilon} \Bigl( - \frac{33}{8} + \zeta_2 \Bigr) + \Bigl( - \frac{665}{16} + \frac{13}{2} \zeta_2 +13 \zeta_3 \Bigr) \\ & &\textstyle + \varepsilon \Bigl( -\frac{10605}{32} + \frac{61}{4} \zeta_2 +\frac{179}{2}\zeta_3 + \frac{233}{2}\zeta_4 \Bigr) + {\cal O}(\varepsilon^2) \biggr] \\ \Im( I_A(1,1,1,0,0,1,1,1,1,0,0) )&=&\textstyle \pi\, s\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon} \Bigl( - \frac{7}{4} + \zeta_2 \Bigr) + \Bigl( - \frac{119}{4} + \frac{15}{2}\zeta_2 + 13 \zeta_3 \Bigr) \\ & &\textstyle + \varepsilon \Bigl( - \frac{4843}{16} + \frac{249}{4}\zeta_2 + \frac{195}{2} \zeta_3 + \frac{125}{2} \zeta_4 \Bigr) + {\cal O}(\varepsilon^2) \biggr] \end{eqnarray*} \begin{figure} \begin{center} \begin{picture}(350,70)(0,0) \SetOffset(0,0) \SetWidth{2} \Line(30,30)(45,30) \Line(105,30)(164,30) \SetWidth{.5} \CArc(45,59)(29,270,350) \Oval(75,30)(24,30)(0) \CArc(92,36)(36,238,350) \CArc(117,46.8)(40.2,238,335) \Text(48,48)[m]{\scriptsize 1} \Text(100,50)[m]{\scriptsize 2} \Text(69,34)[m]{\scriptsize 3} \Text(63,14)[m]{\scriptsize 4} \Text(86,13)[m]{\scriptsize 5} \Text(99,24)[m]{\scriptsize 6} \Text(119,19)[m]{\scriptsize 7} \Text(146,12)[m]{\scriptsize 8} \Text(117,35)[m]{\scriptsize 9} \Text(140,35)[m]{\scriptsize 10} \Text(35,4)[m]{ B} \SetOffset(165,0) \SetWidth{2} \Line(33,30)(45,30) \Line(105,30)(172,30) \SetWidth{.5} \CArc(45,59)(29,270,350) \Oval(75,30)(24,30)(0) \CArc(142.2,27)(18,15,165) \CArc(99,54)(51,250,330) \Text(48,48)[m]{\scriptsize 1} \Text(100,50)[m]{\scriptsize 2} \Text(69,34)[m]{\scriptsize 3} \Text(63,14)[m]{\scriptsize 4} \Text(94,18)[m]{\scriptsize 5} \Text(114,10)[m]{\scriptsize 6} \Text(142,50)[m]{\scriptsize 7} \Text(116,35)[m]{\scriptsize 8} \Text(134,35)[m]{\scriptsize 9} \Text(150,35)[m]{\scriptsize 10} \Text(39,2)[m]{ C} \end{picture} \end{center} \caption{ \label{TopoBCFigs} Integration topologies B and C. The lines are numbered according to Eqs. (\ref{defTopoB},\ref{defTopoC}) } \end{figure} Topology B has 7 non-trivial primitive integrals with a non-zero imaginary part, not counting the integrals that can be considered cases of Topology A. In choosing primitive integrals one may prefer in one case to keep a power of $(p\cdot Q)$ in the numerator to avoid an infrared divergence when the three momenta $p$, $k$ and $l$ go simultaneously to zero. \begin{multline} I_B (\alpha_1,\alpha_2,\alpha_3,\alpha_4,\alpha_5,\alpha_6,\alpha_7,\alpha_8 ,\alpha_9,\alpha_{10},\alpha_{11}) =\\ \int\int\int\int {\rm d}^Dp\; {\rm d}^Dk\; {\rm d}^Dl\; {\rm d}^Dr \; \Biggl( \frac{ (p\cdot Q)^{\alpha_{11}}}{(r^2)^{\alpha_1} [(k+Q)^2]^{\alpha_2} [(r-k-Q)^2]^{\alpha_3} (k^2)^{\alpha_4} [(k-p)^2]^{\alpha_5} }\\ \times \frac{1}{ [(k-l-p)^2]^{\alpha_6} (p^2)^{\alpha_7} (l^2)^{\alpha_8} [(l+p)^2 +2(l+p)\cdot Q]^{\alpha_9} (l^2+2l\cdot Q)^{\alpha_{10}} } \Biggr) \label{defTopoB} \end{multline} \begin{eqnarray*} \Im( I_B(1,1,1,1,1,1,1,1,1,1,1) )&=& \textstyle \frac{ \pi}{s} (N_\varepsilon)^4 \left[ \frac{1}{\varepsilon} \Bigl( - \frac{17}{16} \zeta_4 \Bigr) + {\cal O}(\varepsilon^0) \right] \\ \Im( I_B(1,1,1,1,1,1,0,1,1,1,0) )&=& \textstyle \frac{ \pi}{s} (N_\varepsilon)^4 \left[ \frac{1}{\varepsilon} \Bigl( - \frac{51}{8} \zeta_4 \Bigr) + {\cal O}(\varepsilon^0) \right] \\ \Im( I_B(1,0,1,0,1,1,1,0,1,1,0) )&=& \textstyle \pi\, s\, (N_\varepsilon)^4 \biggl[ \Bigl( -\frac{7}{4}+\zeta_2 \Bigr) + \varepsilon \Bigl( -\frac{287}{8}+10 \zeta_2 +15\zeta_3 \Bigr) \\ & & \textstyle + \varepsilon^2 \Bigl( -\frac{6847}{16} +67\zeta_2 +159 \zeta_3 +108\zeta_4 \Bigr) + {\cal O}(\varepsilon^3) \biggr] \\ \Im( I_B(1,1,1,1,1,0,0,1,1,1,0) )&=& \textstyle \pi\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon^3} \Bigl( \frac{1}{2} \Bigr) + \frac{1}{\varepsilon^2} \Bigl( 5 \Bigr) + \frac{1}{\varepsilon} \Bigl(\frac{63}{2} -2 \zeta_3 -2\zeta_2 \Bigr) \\ & & \textstyle + \Bigl( 159-19\zeta_2 -15 \zeta_3 - \frac{11}{2} \zeta_4 \Bigr) + {\cal O}(\varepsilon) \biggr] \\ \Im( I_B(1,1,1,0,1,0,1,1,1,1,0) )&=& \textstyle \pi\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon} +\Bigl( 20 - \zeta_2 - 4\zeta_3 \Bigr) + \varepsilon \Bigl( 237 - 24 \zeta_2 \\ & & \textstyle - 53 \zeta_3 - \frac{67}{2}\zeta_4 \Bigr) + {\cal O}(\varepsilon^2) \biggr] \\ \Im( I_B(1,1,1,1,1,1,0,0,1,2,0) )&=& \textstyle \frac{ \pi}{s} (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon} \Bigl( \frac{5}{4} \zeta_3 - 3 \zeta_2 \ln(2) \Bigr) + \Bigl( 5\zeta_3 - 12 \zeta_2 \ln(2) \\ & & \textstyle - \frac{49}{8} \zeta_4 + 6{\rm Li_4}(1/2) + \frac{1}{4} \ln^4(2) \Bigr) + {\cal O}(\varepsilon) \biggr] \\ \Im( I_B(1,0,1,1,1,1,0,0,2,2,0) )&=& \textstyle \frac{ \pi}{s} (N_\varepsilon)^4 \biggl[ \Bigl( -3 \zeta_2 +\frac{3}{4}\zeta_3 +3\zeta_2\ln(2) \Bigr) + \varepsilon \Bigl( - 21\zeta_2 - \frac{25}{2}\zeta_3 \\ & & \textstyle + 6\zeta_2\ln(2) + \frac{291}{8}\zeta_4 - 6\, {\rm Li_4}(1/2) - \frac{1}{4}\ln^4(2) \Bigr) + {\cal O}(\varepsilon^2) \biggr] \end{eqnarray*} Here ${\rm Li_4}(x)$ is the fourth order Polylogarithm \cite{polylog}, ${\rm Li_4}(1/2) = 0.517479\cdots$. For topology C one can lower the power of at least one propagator to zero through the use of integration-by-parts recurrence relations. For this topology there are three additional primitive integrals with a non-zero imaginary part, not counting the integrals that are special cases of Topologies A and B. \begin{multline} I_C (\alpha_1,\alpha_2,\alpha_3,\alpha_4,\alpha_5,\alpha_6,\alpha_7,\alpha_8 ,\alpha_9,\alpha_{10},\alpha_{11}) =\\ \int\int\int\int {\rm d}^Dp\; {\rm d}^Dk\; {\rm d}^Dl\; {\rm d}^Dr \; \Biggl( \frac{ (kl)^{\alpha_{11}}}{(r^2)^{\alpha_1} [(k+Q)^2]^{\alpha_2} [(r-k-Q)^2]^{\alpha_3} (k^2)^{\alpha_4} [(k-p)^2]^{\alpha_5} }\\ \times \frac{1}{ (p^2)^{\alpha_6} (l^2)^{\alpha_7} (p^2+2p\cdot Q)^{\alpha_8} [(l+p)^2 +2(l+p)\cdot Q]^{\alpha_9} (l^2+2l\cdot Q)^{\alpha_{10}} } \Biggr) \label{defTopoC} \end{multline} \begin{eqnarray*}\textstyle \Im( I_C(1,0,1,1,1,0,0,1,1,1,0) )&=& \textstyle \pi\, s\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon^2} \Bigl( -\frac{1}{4} \Bigr) + \frac{1}{\varepsilon} \Bigl( -\frac{33}{8} +\zeta_2 \Bigr) \\ & & \textstyle + \Bigl( -\frac{657}{16} +11 \zeta_2 + 9 \zeta_3 \Bigr) + {\cal O}(\varepsilon) \biggr] \\ \Im( I_C(1,0,1,1,1,1,0,1,1,1,0) )&=& \textstyle \pi\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon} \Bigl( 3 \zeta_2 \Bigr) + \Bigl( 11 + 26\zeta_2 + \zeta_3 \Bigr) + {\cal O}(\varepsilon) \biggr] \\ \Im( I_C(1,1,1,1,0,0,0,1,1,1,0) )&=& \textstyle \pi\, s (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon^3} \Bigl( \frac{3}{2} \Bigr) + \frac{1}{\varepsilon^2} \Bigl( \frac{47}{4} \Bigr) + \frac{1}{\varepsilon} \Bigl( \frac{457}{8} - 3 \zeta_2 \Bigr) \\ & & \textstyle + \Bigl( \frac{3507}{16} - \frac{31}{2} \zeta_2 - 3\zeta_3 \Bigr) + \varepsilon \Bigl( \frac{22985}{32} - 48\zeta_2 \ln(2) \\ & & \textstyle - \frac{89}{4} \zeta_2 + \frac{9}{2} \zeta_3 - \frac{15}{2} \zeta_4 \Bigr) + {\cal O}(\varepsilon^2) \biggr] \end{eqnarray*} For topology D one can apply integration-by-parts relations in the form of the massless triangle rule to the triangle formed by lines 5, 6 and 7. Further use of integration-by-parts expresses integrals of this topology into integrals that can be calculated using the closed expressions Eqs.(\ref{1loopbubble})--(\ref{bubblem00}) or integrals with no imaginary part, or integrals that are of topology A. Therefore no new non-trivial primitives with a non-zero imaginary part appear for topology D. \begin{figure} \begin{center} \begin{picture}(350,160)(0,0) \SetOffset(-10,100) \SetWidth{2} \Line(38,30)(60,30) \Line(120,30)(172,30) \SetWidth{.5} \CArc(60,59)(29,270,350) \Oval(90,30)(24,30)(0) \CArc(111,54)(51,250,330) \CArc(96,-3)(24,48,150) \Text(62,48)[m]{\scriptsize 1} \Text(117,50)[m]{\scriptsize 2} \Text(84,34)[m]{\scriptsize 3} \Text(68,20)[m]{\scriptsize 4} \Text(104,25)[m]{\scriptsize 5} \Text(100,12)[m]{\scriptsize 6} \Text(85,1)[m]{\scriptsize 7} \Text(150,14)[m]{\scriptsize 8} \Text(123,20)[m]{\scriptsize 9} \Text(137,35)[m]{\scriptsize 10} \Text(45,0)[m]{ D} \SetOffset(165,100) \SetWidth{2} \Line(30,30)(75,30) \Line(135,30)(180,30) \SetWidth{.5} \CArc(75,59)(29,270,350) \Oval(105,30)(24,30)(0) \CArc(75,32.4)(27,188,300) \CArc(135,32.4)(27,240,352) \Text(78,48)[m]{\scriptsize 1} \Text(132,50)[m]{\scriptsize 2} \Text(100,35)[m]{\scriptsize 3} \Text(105,11)[m]{\scriptsize 4} \Text(128,21)[m]{\scriptsize 5} \Text(84,21)[m]{\scriptsize 6} \Text(63,14)[m]{\scriptsize 7} \Text(148,14)[m]{\scriptsize 8} \Text(62,35)[m]{\scriptsize 9} \Text(150,35)[m]{\scriptsize 10} \Text(42,0)[m]{ E} \SetOffset(-10,10) \SetWidth{2} \Line(40,30)(79.2,30) \Line(130.2,30)(170,30) \SetWidth{.5} \CArc(79,57)(27,270,340) \Oval(105,30)(18,25.2)(0) \CArc(91.2,36)(40.8,190,270) \CArc(91.2,27)(31.8,270,340) \CArc(118.8,27)(31.8,200,270) \CArc(118.8,36)(40.8,270,350) \Text(84,46)[m]{\scriptsize 1} \Text(126,48)[m]{\scriptsize 2} \Text(102,35)[m]{\scriptsize 3} \Text(105,16)[m]{\scriptsize 4} \Text(123,23)[m]{\scriptsize 5} \Text(88.5,23)[m]{\scriptsize 6} \Text(63,14)[m]{\scriptsize 7} \Text(148,14)[m]{\scriptsize 8} \Text(64,35)[m]{\scriptsize 9} \Text(148,35)[m]{\scriptsize 10} \Text(45,-7)[m]{ F} \SetOffset(165,10) \SetWidth{2} \Line(30,30)(70.8,30) \Line(121.2,30)(180,30) \SetWidth{.5} \CArc(71,57)(27,270,340) \Oval(96,30)(18,25.2)(0) \CArc(122.4,51)(46.2,238,332) \CArc(96,13.8)(50,20,160) \Text(76,47)[m]{\scriptsize 1} \Text(110,39)[m]{\scriptsize 2} \Text(94,33)[m]{\scriptsize 3} \Text(83.5,20)[m]{\scriptsize 4} \Text(110,20)[m]{\scriptsize 5} \Text(63,58)[m]{\scriptsize 6} \Text(130,10)[m]{\scriptsize 7} \Text(62,35)[m]{\scriptsize 8} \Text(130,35)[m]{\scriptsize 9} \Text(152,35)[m]{\scriptsize 10} \Text(42,0)[m]{ G} \end{picture} \end{center} \caption{ \label{TopoDEFGFigs} Integration topologies D -- G. The lines are numbered according to Eqs. (\ref{defTopoE}) and (\ref{defTopoF}) } \end{figure} For topology E one can lower the power of at least one propagator to zero through the use of integration-by-parts relations. There is one primitive integral with a non-zero imaginary part, not counting the integrals that can be considered as cases of topology A and B. \begin{multline} \label{defTopoE} I_E (\alpha_1,\alpha_2,\alpha_3,\alpha_4,\alpha_5,\alpha_6,\alpha_7,\alpha_8 ,\alpha_9,\alpha_{10},\alpha_{11}) =\\ \int\int\int\int {\rm d}^Dp\; {\rm d}^Dk\; {\rm d}^Dl\; {\rm d}^Dr \; \Biggl( \frac{ (p\cdot l)^{\alpha_{11}}}{(r^2)^{\alpha_1} [(k+Q)^2]^{\alpha_2} [(r-k-Q)^2]^{\alpha_3} (k^2)^{\alpha_4} [(k-l)^2]^{\alpha_5} }\\ \times \frac{1}{ [(k-p)^2]^{\alpha_6} (p^2)^{\alpha_7} (l^2)^{\alpha_8} (p^2 +2p\cdot Q)^{\alpha_9} (l^2+2l\cdot Q)^{\alpha_{10}} } \Biggr) \end{multline} \[\textstyle \Im( I_E(1,1,1,0,1,1,1,1,1,1,0) ) = \frac{ \pi}{s} (N_\varepsilon)^4 \left[ \frac{1}{\varepsilon} \Bigl( -6 \zeta_4 \Bigr) + {\cal O}(\varepsilon^0) \right] \] Non-planar topology F has 6 non-trivial primitive integrals with a non-zero imaginary part, not counting the integrals that can be considered cases of topologies A--E. In choosing primitive integrals one may prefer in one case to keep a power of $(p\cdot l)$ in the numerator to avoid an infrared divergence when the three momenta $p$, $k$ and $l$ go simultaneously to zero. \begin{multline} I_F (\alpha_1,\alpha_2,\alpha_3,\alpha_4,\alpha_5,\alpha_6,\alpha_7,\alpha_8 ,\alpha_9,\alpha_{10},\alpha_{11}) =\\ \int\int\int\int {\rm d}^Dp\; {\rm d}^Dk\; {\rm d}^Dl\; {\rm d}^Dr \; \Biggl( \frac{ (p\cdot l)^{\alpha_{11}}}{(r^2)^{\alpha_1} [(p+k+l+Q)^2]^{\alpha_2} [(r-p-k-l-Q)^2]^{\alpha_3} (k^2)^{\alpha_4} }\\ \times \frac{1}{ [(k+p)^2]^{\alpha_5} [(k+l)^2]^{\alpha_6} (p^2)^{\alpha_7} (l^2)^{\alpha_8} (p^2 +2p\cdot Q)^{\alpha_9} (l^2+2l\cdot Q)^{\alpha_{10}} } \Biggr) \label{defTopoF} \end{multline} \begin{eqnarray*} \Im( I_F(1,1,1,1,1,1,1,1,1,1,1) )&=&\textstyle \frac{ \pi}{s} (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon} \Bigl( \frac{3}{4} \zeta_4 \Bigr) + {\cal O}(\varepsilon^0) \bigg]\\ \Im( I_F(1,1,1,1,1,1,1,0,1,1,0) )&=&\textstyle \frac{ \pi}{s} (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon} \Bigl( -\frac{37}{8} \zeta_4 \Bigr) + {\cal O}(\varepsilon^0) \bigg]\\ \Im( I_F(1,0,1,1,1,1,1,0,1,2,0) )&=&\textstyle \frac{ \pi}{s} (N_\varepsilon)^4 \biggl[ \Bigl( -\frac{27}{8} \zeta_4 \Bigr) + {\cal O}(\varepsilon) \bigg]\\ \Im( I_F(1,1,1,1,0,1,1,0,1,1,0) )&=&\textstyle \pi\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon^2} \Bigl( 1 \Bigr) + \frac{1}{\varepsilon} \Bigl( 15 -\zeta_2 -\zeta_3 \Bigr)\\ & &\qquad\textstyle + \Bigl( 139 -19 \zeta_2 -17\zeta_3 - 9\zeta_4 \Bigr) + {\cal O}(\varepsilon) \bigg]\\ \Im( I_F(1,1,1,1,0,0,1,1,1,1,0) )&=&\textstyle \pi\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon} \Bigl( -1 + \zeta_2 \Bigr) + \Bigl( -20 + 11 \zeta_2 + 11 \zeta_3 \Bigr)\\ & &\qquad\textstyle + \varepsilon \Bigl( -237 + 93 \zeta_2 + 121 \zeta_3 + 37 \zeta_4 \Bigr) + {\cal O}(\varepsilon^2) \bigg]\\ \Im( I_F(1,1,1,1,1,1,0,0,1,1,0) )&=&\textstyle \pi\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon} \Bigl( -2 + \zeta_2 + \frac{13}{4} \zeta_3 -3\zeta_2\ln(2) \Bigr) + {\cal O}(\varepsilon^0) \bigg] \end{eqnarray*} For integrals of topology G one is able to remove enough propagators through the use of integration-by-parts relations to make these integrals cases of topologies A--F. Therefore no new primitives appear for topology G. \subsection{The evaluation of primitive integrals} In this section we will discuss the calculation of primitive integrals on the mass shell using a large mass expansion as an intermediate representation. The general theory of Euclidean asymptotic expansions that is invoked was developed in Refs. \cite{formalexpansions1,formalexpansions2} and in practice the techniques of Ref.\cite{largemass} were used. In many physical applications the large mass expansion is truncated at a finite order, however it has been used in recent works \cite{largemassUntrunc,largemassUntrunc2} also in an untruncated form. Before treating the more complicated 4-loop integrals we will illustrate some basic features of the approach using a simple massive one-loop integral as an example \begin{equation} \label{prototypeoneloop} I(\alpha_1,\alpha_2) = \int {\rm d}^Dk\; \frac{ 1}{(k^2)^{\alpha_1} [(k+Q)^2-M^2]^{\alpha_2} } = \int {\rm d}^Dp\; \frac{ 1}{[(p+Q)^2]^{\alpha_1} (p^2-M^2)^{\alpha_2} } \end{equation} Again we do not write explicitly $+i\epsilon$ in the denominators to keep expressions compact. For a large mass, $M^2 > Q^2$ one may expand this integral by making an ordinary Taylor expansion in the integrand: \begin{eqnarray} \frac{1}{[(p+Q)^2]^{\alpha_1}} & \rightarrow & \frac{1}{(p^2)^{\alpha_1}} \sum_{i=0}^\infty (-1)^i \left( \frac{ 2p\cdot Q+Q^2}{p^2} \right)^i \frac{ \Gamma(\alpha_1+i)}{ \Gamma(\alpha_1)\; i!} \nonumber \\ & = & \frac{1}{(p^2)^{\alpha_1}} \sum_{i=0}^\infty \sum_{k=0}^i (-1)^i \frac{ (2p\cdot Q)^k\; (Q^2)^{i-k} }{(p^2)^{i}} \frac{ \Gamma(\alpha_1+i)}{ \Gamma(\alpha_1)\; k! \; (i-k)! } \end{eqnarray} The remaining integral over $p$ is of the vacuum bubble type and after applying a bubble tensor reduction to simplify the powers of $(p\cdot Q)$ in the numerator \begin{equation} (p\cdot Q)^{n}\rightarrow \begin{cases} (p^2 Q^2)^{\frac{n}{2}} \; \frac{n!}{2^{n}\left(n/2\right)!} \; \frac{\Gamma\left(D/2\right)} {\Gamma\left(D/2+n/2\right)} & \text{$n$ even},\\ 0 & \text{$n>0$ odd} \end{cases} \end{equation} the bubble integral can be performed using Eq.~(\ref{1loopbubble}). In the resulting expression one of the two sums can be performed using the summation identity \begin{equation} \label{sumidexample1} \sum_{k=0}^n \frac{ (-1)^k\; \Gamma(a+k)}{k!\; (n-k)!\; \Gamma(b+k)} = \frac{ \Gamma(a)\; \Gamma(b-a+n)}{n!\; \Gamma(n+b)\; \Gamma(b-a) } \end{equation} and obtain the well-known result \begin{eqnarray} I(\alpha_1,\alpha_2) &=& i \pi^{(D/2)} (-1)^{-\alpha_1-\alpha_2} (M^2)^{D/2-\alpha_1-\alpha_2} \frac{ \Gamma(D/2-\alpha_1 )} { \Gamma( \alpha_1 ) \; \Gamma( \alpha_2 )}\nonumber\\ & &\qquad\qquad\qquad\qquad\qquad \times\sum_{j=0}^{\infty} \left( \frac{ Q^2}{M^2} \right)^j \frac{ \Gamma(\alpha_1 + j)\; \Gamma(\alpha_1 +\alpha_2 +j-D/2) }{ j!\; \Gamma(D/2+j) } \label{Iexaple00}\\ &=&i \pi^{(D/2)} (-1)^{-\alpha_1-\alpha_2} (M^2)^{D/2-\alpha_1-\alpha_2} \frac{ \Gamma(D/2-\alpha_1 )\; \Gamma(\alpha_1+\alpha_2-D/2)}{ \Gamma(D/2)\; \Gamma(\alpha_2)}\nonumber\\ & &\qquad\qquad\qquad\qquad\qquad \times {}_2F_1 \left( \begin{array}{c} \alpha_1\; ,\; \alpha_1+\alpha_2-D/2 \\ D/2 \end{array} \Biggr| \frac{ Q^2}{M^2} \right) \label{Iexaple0} \end{eqnarray} where ${}_2F_1$ is the Gauss hypergeometric function. To obtain the proper value on the mass shell $Q^2/M^2 = 1$ within dimensional regularization, one goes on-shell in an interval of the dimension $D$ where the integral is convergent and continues the result in the number of space-time dimensions. Indeed, assuming that $D$ is chosen such that the sum is convergent one can use the Gauss summation identity \[ {}_2F_1 \left( \begin{array}{c} a\; ,\; b \\ c \end{array} \Biggr| 1\right) = \frac{ \Gamma(c)\; \Gamma(c-a-b)}{\Gamma(c-a)\; \Gamma(c-b)} \hspace{1.5cm} (c\neq0,-1,\cdots,\; \Re(c-a-b)>0) \] to obtain the on-shell expression Eq.~(\ref{1looponshell}). Depending on the values of $\alpha_1$, $\alpha_2$, integral $I(\alpha_1,\alpha_2)$ can have specific IR divergences on the mass-shell since $[(k+Q)^2-M^2]$ becomes $(k^2+2Q\cdot k)$. One may use IR power-counting in the small $k$ region to determine whether an IR divergence occurs on the mass-shell. Here $1/(k^2+2Q\cdot k)$ counts for only half the power of an ordinary massless line $1/k^2$. Since such an IR divergence appears only on-shell, it manifests itself as non-convergence of the large mass expansion starting exactly on the mass shell. Using integration-by-parts recurrence relations it is possible to choose an appropriate basis set of primitive integrals for which the large mass expansion is convergent on-shell in $D=4-2\varepsilon$. Nevertheless it is instructive to study the behaviour of IR divergent integrals near the mass-shell. Let us therefore consider integral $I(1,2)$. Using the transformation rule for the hypergeometric function \begin{eqnarray} {}_2F_1\left( \begin{array}{c} a\; ,\; b \\ c \end{array} \Biggr| z \right) & = & \frac{ \Gamma(c)\; \Gamma(c-a-b)}{\Gamma(c-a)\; \Gamma(c-b)} {}_2F_1 \left( \begin{array}{c} a\; ,\; b \\ a+b-c+1 \end{array} \Biggr|1- z \right) \nonumber \\ \label{gausstransform} & & + (1-z)^{c-a-b} \frac{\Gamma(c)\; \Gamma(a+b-c)}{\Gamma(a)\; \Gamma(b)} {}_2F_1 \left( \begin{array}{c} c-a\; ,\; c-b \\ c-a-b+1 \end{array} \Biggr|1- z \right) \end{eqnarray} one easily obtains the behaviour near the mass shell \begin{equation} \label{i12nearshell} I(1,2)= \frac{ i \pi \; \Gamma(1+\varepsilon) }{2\varepsilon \; (M^2)^{1+\varepsilon} } \left[ 1 - (1-x)^{-2\varepsilon} \frac{ \Gamma(1+2\varepsilon)\; \Gamma(1-\varepsilon)}{ \Gamma(1+\varepsilon) } \right] + {\cal O}(1-x) \end{equation} where $x= Q^2/M^2$. On shell the term $(1-x)^{-2\varepsilon}$ is nullified within dimensional regularization and for $x<1$ the threshold term can be expanded in $\varepsilon$ \begin{equation} \label{polesonshell} \frac{-1}{2\varepsilon} \left[ 1- (1-x)^{-2\varepsilon} \right] = - \ln(1-x) + \varepsilon \ln^2(1-x) - \frac{2}{3} \varepsilon^2 \ln^3(1-x) + {\cal O} \left( \varepsilon^3\ln^4(1-x) \right) \end{equation} The logarithms $\ln^i(1-x)$ in Eq.~(\ref{polesonshell}) are translated on-shell into a pole. Such logarithmic divergences appear for certain individual diagrams of the muon decay calculations but they cancel to all orders in $\varepsilon$ when the external leg corrections are included. One may therefore use dimensional regularization and evaluate all integrals at threshold. To make the connection with methods where one expands expressions in $\varepsilon$ at an early stage in the calculation, we also expand the $\Gamma$ functions of Eq.~(\ref{Iexaple0}) in $\varepsilon$ \begin{equation} \label{expandGamma} \Gamma(n+1+\varepsilon) = n!\; \Gamma(1+\varepsilon) \left[ 1+ \varepsilon S_1(n) + \frac{\varepsilon^2}{2}( S_1^2(n)-S_2(n)) + {\cal O}(\varepsilon^3)\right] \end{equation} where the harmonic sums $S$ are defined as \begin{equation} S_k(n) = \sum_{i=1}^{n} \frac{1}{i^k} \end{equation} In this way one obtains for integral $I(1,2)$ \begin{eqnarray} I(1,2)&=& \frac{ -i \pi \;\Gamma(1+\varepsilon) }{ (M^2)^{1+\varepsilon} } \; \sum_{j=0}^{\infty} \frac{x^j}{1+j} \left[ 1+2\varepsilon S_1(j) +2\varepsilon^2 S^2_1(j) + \frac{\varepsilon}{1+j} +{\cal O}(\varepsilon^2) \right]\nonumber\\ &=&\frac{ -i \pi \; \Gamma(1+\varepsilon) }{(M^2)^{1+\varepsilon} } \Biggl[ - \ln(1-x) + \varepsilon \ln^2(1-x) - \frac{2}{3} \varepsilon^2 \ln^3(1-x) \nonumber\\ & &\qquad\qquad\qquad\qquad\qquad - 2\zeta_2 \varepsilon^2 \ln(1-x) + \varepsilon \zeta_2 +{\cal O}(\varepsilon^2) + {\cal O}(1-x) \Biggl] \label{Iablogdivergence} \end{eqnarray} in agreement with Eqs.(\ref{i12nearshell}) and (\ref{polesonshell}). However, one needs to consider the divergent logarithms to all orders in $\varepsilon$ to obtain the proper on-shell value in dimensional regularization. As integration-by-parts recurrence relations are used to reduce the number of different integrals that are required, one may conveniently choose a set of primitive integrals that are free from threshold singularities. These well-behaved primitive integrals can then be calculated via a convergent large mass expansion that can be truncated at a fixed order in $\varepsilon$. This is the general strategy followed in the more complicated multiloop integrals. The only caveat is that, after applying recurrence relations, poles may appear in the coefficients multiplying the primitive integrals and these primitive integrals then have to be evaluated to a higher order in $\varepsilon$ than na\"\i vely expected. It follows from Eq. (\ref{gausstransform}) that integral $I(\alpha_1,\alpha_2)$ has general on-shell divergences of the form \begin{equation} \label{oneloopthreshold} \lim_{x\rightarrow1} \left[(1-x)^{(-m-2\varepsilon)} \right] \end{equation} which are nullified\footnote{Studies of the various multiloop integrals that are needed for the muon-decay calculation reveal a slightly more general threshold behaviour \[ \lim_{x\rightarrow1} \left[(1-x)^{(-m-2n\varepsilon)} \right] \] where $m$ is an integer and $n$ is generally the number of independent loop momenta that when they to zero give rise to an infrared divergence on-shell. The absence of unregularized singularities $(1-x)^{-m}$ for $m>0$ is important for the validity of on-shell recurrence relations.} in $D>4+m$ dimensions. Note that for $x>1$ the terms $+i\epsilon$ in the denominators of the propagators that were implicit in Eq. (\ref{prototypeoneloop}) are restored by the substitution $x\rightarrow x +i\epsilon$ in Eq. (\ref{oneloopthreshold}) to give the contributions from channels that open up above threshold, but for the present work this is not needed. \subsubsection{Example 2} \begin{figure} \begin{center} \begin{picture}(200,70)(0,0) \SetOffset(0,0) \SetWidth{2} \Line(30,30)(45,30) \Line(135,30)(164,30) \CArc(105,30)(30,270,360) \Line(105,0)(105,30) \SetWidth{.5} \Line(45,30)(140,30) \CArc(45,60)(30,270,360) \CArc(105,75)(75,217,270) \CArc(75,30)(30,0,180) \Text(45,50)[m]{\scriptsize $p_1$} \Text(106,54)[m]{\scriptsize $p_2$} \Text(80,48)[m]{\scriptsize $p_3$} \Text(90,35)[m]{\scriptsize $p_5$} \Text(63,6)[m]{\scriptsize $p_7$} \Text(99,16)[m]{\scriptsize $p_9$} \Text(120,35)[m]{\scriptsize $p_8$} \Text(131,2)[m]{\scriptsize $p_{10}$} \Text(150,36)[m]{\scriptsize $Q$} \end{picture} \end{center} \caption{ \label{primitiveB4} Integral $I_B(1,1,1,0,1,0,1,1,1,1,0)$. Thin lines correspond to massless scalar propagators and thick lines correspond to massive scalar propagators with mass $M$. The momenta that flow through the propagators are labeled with the line numbers defined for main topology B, Eq.(\ref{defTopoB}). } \end{figure} Let us now consider the calculation of a primitive integral $I_B(1,1,1,0,1,0,1,1,1,1,0)$ (see Fig.\ref{primitiveB4}) using the large mass expansion as an intermediate representation. The basic steps followed will be similar to the steps in the derivation of Eq.~(\ref{Iexaple0}). However, the sums that the Taylor expansions introduce cannot be reduced at the level of $\Gamma$ functions because insufficient generalizations of Eq.~(\ref{sumidexample1}) are known. One is therefore forced to reduce sums after expanding the $\Gamma$ functions in $\varepsilon$. What makes this approach feasible is that the sums over harmonic functions, $S$, encountered in the present calculation can be reduced to expressions given in terms of a small number of higher order sums. Eventually one is only interested in the value of integrals on-shell. For integrals with more than one massive line one may therefore choose to make a large mass expansion in terms of the mass of only one of the propagators, and to keep the masses of other propagators on-shell, if this leads to less complicated intermediate expressions. The problem of reducing harmonic sums also appears in calculations for deep-inelastic lepton-nucleon scattering, where one deals with a light cone expansion instead of the large mass expansion. Large collections of summation identities of the type that are needed for the present work can be found in the literature \cite{finitesums1,sumsjos}. Although summation reduction algebra is rather involved, relatively compact final expressions for the coefficients of the large mass expansion are obtained. On the mass-shell the sums can be reduced to known mathematical constants. We take the momentum $Q$ off-shell, $ s = Q^2<M^2$, and perform a large mass expansion by expanding the integrand as \begin{eqnarray} \frac{1}% {p_1^2\,p_2^2\, p_3^2\, p_5^2\, p_7^2\, p_8^2\,(p_9^2-M^2)(p_{10}^2-M^2)} &\rightarrow& \frac{1}{p_1^2\,p_2^2\, p_3^2\, p_5^2\, p_7^2} \left[ \frac{1}{ p_8^2\,(p_9^2-M^2) (p_{10}^2-M^2)} \right]_{\{p_2,p_5,p_7,Q\}} \nonumber\\ & & + \left[ \frac{1}{p_1^2\,p_2^2\, p_3^2\, p_5^2\, p_7^2\, p_8^2\, (p_9^2-M^2) (p_{10}^2-M^2)} \right]_{\{Q\}} \label{expandB4} \end{eqnarray} where we use the notation that the terms between square brackets are expanded as an infinite Taylor series around the point where the momenta between curly brackets are zero. Here momentum $p_i$ is understood to be the momentum that flows through line $i$ as defined for main topology B, Eq.~(\ref{defTopoB}). After performing the indicated Taylor expansion in $Q$, the last term on the rhs of (\ref{expandB4}) produces integrals of the 4-loop massive bubble type and therefore does not contribute to the imaginary part. The first term on the rhs of (\ref{expandB4}) produces integrals that are a product of a 3-loop massless propagator-type integral and a 1-loop bubble integral, all coupled by tensor numerators. After the necessary tensor reductions, one can evaluate these integrals in terms of $\Gamma$-functions using Eqs.(\ref{1loopbubble}) and (\ref{1loopmassless}). A representation for the imaginary part of integral $I_B(1,1,1,0,$ $1,0,1,1,1,1,0)$ is obtained, in this way, as a multiple sum over a product of $\Gamma$ functions. After expanding the $\Gamma$ functions in $\varepsilon$ as in Eq.~(\ref{expandGamma}) one can reduce the sums over the functions $S_i(k)$ to a basis of independent nested sums to obtain the off-shell expression \begin{multline} \Im( I_B(1,1,1,0,1,0,1,1,1,1,0) ) = \pi \left( N_\varepsilon \right)^4 \left( \frac{s}{M^2}\right)^\varepsilon \sum_{k=1}^{\infty} \left(\frac{s}{M^2}\right)^k \Biggl\{ \frac{1}{\varepsilon } \biggl[ \frac{1}{k} -\frac{1}{k+1} \biggr] \\ + \biggl[ \frac{16}{k}-\frac{16}{k+1} -\frac{1}{k^2} +4 \frac{S_1(k)}{k} -4\frac{S_1(k+1)}{k+1} -2 \frac{S_1(k)}{k^2} \biggr] + \varepsilon \biggl[ \frac{157-8\zeta_2}{k} + \frac{-157+8\zeta_2}{k+1} \\ -\frac{16}{k^2} +\frac{1}{k^3} -\frac{2}{(k+1)^3} +64\frac{S_1(k)}{k} -64\frac{S_1(k+1)}{k+1} -26\frac{S_1(k)}{k^2} +2\frac{S_1(k)}{k^3} +6\frac{S_2(k)}{k} \\ -6\frac{S_2(k+1)}{k+1} -6\frac{S_2(k)}{k^2} +8\frac{S_1^2(k)}{k} -8 \frac{S_1^2(k+1)}{k+1} -6\frac{S_1^2(k)}{k^2} \biggr] + O( \varepsilon^2 ) \Biggr\} \end{multline} where $N_\varepsilon$ is defined in Eq.~(\ref{defnormloop}). On-shell $s/M^2 = 1$ and the sum over $k$ can be performed analytically to give the following result \begin{multline} \textstyle \Im( I_B(1,1,1,0,1,0,1,1,1,1,0) ) = \pi\, (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon} +\Bigl( 20 - \zeta_2 - 4\zeta_3 \Bigr) + \varepsilon \Bigl( 237 - 24 \zeta_2\\ - 53 \zeta_3 - \frac{67}{2}\zeta_4 \Bigr) + {\cal O}(\varepsilon^2) \biggr] \end{multline} where the various infinite sums that appear are \begin{eqnarray} \textstyle \label{beginsumsex3} \sum_{i=1}^{\infty} \frac{1}{i^k} & = & S_k(\infty) = \zeta_k \hspace{1cm} (k > 1) \\ \textstyle \sum_{i=1}^{\infty} \frac{S_1(i)}{i^2} & \equiv & S_{2,1}(\infty) = 2 \zeta_3 \\ \textstyle \sum_{i=1}^{\infty} \frac{S_1(i)}{i^3} & \equiv & \textstyle S_{3,1}(\infty) = \frac{5}{4}\zeta_4 \\ \textstyle \sum_{i=1}^{\infty} \frac{S_2(i)}{i^2} & \equiv & S_{2,2}(\infty) = \textstyle \frac{7}{4} \zeta_4 \\ \textstyle \label{endsumsex3} \sum_{i=1}^{\infty} \frac{ S^2_1(i)}{i^2} & = & \textstyle \frac{17}{4}\zeta_4 \end{eqnarray} and are known from the literature \cite{sumsjos,infinitesums1,infinitesums2,infinitesums3}. \subsection{Example 3} \begin{figure} \begin{center} \begin{picture}(200,70)(0,0) \SetOffset(0,0) \SetWidth{2} \Line(30,30)(45,30) \Line(105,30)(170,30) \SetWidth{.5} \CArc(45,59)(29,270,350) \Oval(75,30)(24,30)(0) \CArc(92,36)(36,238,350) \CArc(117,46.8)(40.2,238,335) \Text(48,48)[m]{\scriptsize $p_1$} \Text(100,50)[m]{\scriptsize $p_2$} \Text(69,34)[m]{\scriptsize $p_3$} \Text(63,14)[m]{\scriptsize $p_4$} \Text(86,13)[m]{\scriptsize $p_5$} \Text(99,24)[m]{\scriptsize $p_6$} \Text(119,19)[m]{\scriptsize $p_7$} \Text(146,12)[m]{\scriptsize $p_8$} \Text(117,35)[m]{\scriptsize $p_9$} \Text(140,35)[m]{\scriptsize $p_{10}$} \Text(165,35)[m]{\scriptsize $Q$} \end{picture} \end{center} \caption{ \label{primitiveB3A} Integral $I_B(1,1,1,1,1,1,1,1,1,1,1)$. Thin lines correspond to massless scalar propagators and thick lines correspond to massive scalar propagators with mass $M$. The momenta that flow through the propagators are labeled with the line numbers defined for main topology B, Eq.~(\ref{defTopoB}). A factor $p_7\cdot Q$ (line 11) in the numerator is not shown explicitly. } \end{figure} For integral $I_B(1,1,1,1,1,1,1,1,1,1,1)$ we proceed as in the previous example and make a large mass expansion by expanding the integrand \begin{eqnarray} \frac{p_7\cdot Q}{p_1^2\,p_2^2\, p_3^2\, p_4^2\, p_5^2\, p_6^2\, p_7^2\, p_8^2\, (p_9^2-M^2) (p_{10}^2-M^2)}&\rightarrow& \nonumber\\ & &\hspace{-3cm}\null\ \null \frac{p_7\cdot Q}{p_1^2\,p_2^2\, p_3^2\, p_4^2\, p_5^2\, p_6^2\, p_7^2\, p_8^2 }\left[ \frac{1}{(p_9^2-M^2) (p_{10}^2-M^2)} \right]_{\{p_9,p_{10}\}} \nonumber\\ & &\hspace{-3cm} + \frac{1}{p_1^2\,p_2^2\, p_3^2\, p_4^2} \left[ \frac{p_7\cdot Q}{p_5^2\, p_6^2\, p_7^2\, p_8^2\, (p_9^2-M^2) (p_{10}^2-M^2)} \right]_{\{p_2,p_4,Q\}} \nonumber\\ & &\hspace{-3cm} + \frac{p_7\cdot Q}{p_1^2\,p_2^2\, p_3^2\, p_4^2\, p_5^2\, p_7^2} \left[ \frac{1}{p_6^2\,p_8^2\, (p_9^2-M^2) (p_{10}^2-M^2)} \right]_{\{p_2,p_5,p_7,Q\}} \nonumber\\ & &\hspace{-3cm} +\left[ \frac{p_7\cdot Q}{p_1^2\,p_2^2\, p_3^2\, p_4^2\, p_5^2\, p_6^2\, p_7^2\, p_8^2\, (p_9^2-M^2) (p_{10}^2-M^2)} \right]_{\{Q\}} \label{expandBprimitive} \end{eqnarray} The last term is a 4-loop massive bubble and does not contribute to the imaginary part. After performing the indicated Taylor expansions, integrals of the massless propagator and massive bubble types remain that can be evaluated in terms of $\Gamma$-functions using Eqs.~(\ref{1loopbubble})--(\ref{bubblem00}). The $\Gamma$-functions are then expanded in $\varepsilon$ and the resulting sums over the functions $S_i(k)$ may be reduced to a basis of independent nested sums. Adding the imaginary parts of the various contributions in (\ref{expandBprimitive}) gives the off-shell expression \begin{multline} \Im(I_B(1,1,1,1,1,1,1,1,1,1,1)) = \left( N_\varepsilon \right)^4 \frac{\pi }{s} \sum_{k=1}^{\infty} \left(\frac{s}{M^2} \right)^k \left\{ \frac{1}{\varepsilon } \left[-\frac{\zeta_2}{k^2} - \frac{(-1)^k \zeta_2}{2 k^2} +\frac{\zeta_2 S_1(k)}{2 k} \right. \right.\\ \left. - \frac{(-1)^k S_{\tilde{2}}(k)}{k^2} +\frac{S_1(k)}{2 k^3} - \frac{S_2(k)}{4 k^2} +\frac{S_1(k) S_1(k)}{4 k^2} -\frac{S_{1,2}(k)}{k} +\frac{S_1(k) S_2(k)}{2 k} -\frac{S_3(k)}{k} +\frac{2}{k^4} \right]\\ \left. +\frac{1}{\varepsilon} \ln^2\left(\frac{s}{M^2}\right) \left[ \frac{1}{4 k^2} - \frac{S_1(k)}{4 k} \right] +\frac{1}{\varepsilon} \ln\left(\frac{s}{M^2}\right) \left[ - \frac{3}{2 k^3} + \frac{ S_1(k)}{2k^2} +\frac{5 S_2(k)}{4 k} -\frac{S_1(k) S_1(k)}{4 k} \right] + {\cal O}(\varepsilon^0) \right\} \label{offshellIb} \end{multline} where \begin{equation} S_{\tilde{k}}(i) = \sum_{j=1}^{i} (-1)^j \frac{1}{j^k} \end{equation} and for higher order nested sums one defines recursively \begin{equation} S_{k,n,...,m}(i) = \sum_{j=1}^{i} \frac{ S_{n,...,m}(j)}{j^k} , \hspace{1cm} S_{\tilde{k},n,...,m}(i) = \sum_{j=1}^{i} (-1)^j \frac{ S_{n,...,m}(j)}{j^k} \end{equation} where each of the indices $n,...,m$ can have a tilde (\ $\tilde{}$\ ) to indicate an alternating sign component. On-shell $s/M^2 = 1$ and the infinite sum over $k$ can be performed to obtain \begin{equation} \label{RESIb} \textstyle \Im(I_B(1,1,1,1,1,1,1,1,1,1,1)) = \left( N_\varepsilon \right)^4 \frac{\pi}{s} \left\{ -\frac{ 17 \zeta_4}{16 \varepsilon} + {\cal O}(\varepsilon^0) \right\} \hspace{4cm} \end{equation} To illustrate the delicate cancellation for $s/M^2 = 1$ between the various divergent sums in Eq.(\ref{offshellIb}) let us also quote the result for the sum over $k$ from 1 up to $y$ instead of infinity (such that Eq.~(\ref{RESIb}) corresponds to $y \rightarrow \infty $). \begin{multline} \Im(I_B(1,1,1,1,1,1,1,1,1,1,1)) = \left( N_\varepsilon \right)^4 \frac{\pi }{s} \frac{1}{\varepsilon} \left\{ - \zeta_2 S_2(y) -\frac{1}{2} \zeta_2 S_{\tilde{2}}(y) - S_{\tilde{2},\tilde{2}}(y) +3 S_{3,1}(y) \right. \\ \left. +\frac{1}{2}S_{2,2}(y) -S_{2,1,1}(y) +S_1(y) \Bigl[S_{2,1}(y)-2 S_3(y) \Bigr] +\frac{1}{2} S_{1,1}(y) \Bigl[ \zeta_2 - S_2(y) \Bigr] \right\} + {\cal O}(\varepsilon^0) \end{multline} Here one encounters in addition to the sums Eqs. (\ref{beginsumsex3}) -- (\ref{endsumsex3}) the known infinite sums \begin{eqnarray} S_{\tilde{k}}(\infty) & = & \textstyle - \left( 1-\frac{1}{2^{k-1}} \right) \zeta_k \hspace{1cm} (k > 1) \\ S_{\tilde{2},\tilde{2}}(\infty) & = &\textstyle \frac{13}{16}\zeta_4 \\ S_{2,1,1}(\infty) & = & 3\zeta_4 \end{eqnarray} and $S_1(y)$ and $S_{1,1}(y)$ diverge logarithmically for large $y$. \subsubsection{Example 4} As a last example the calculation of integral $I_B(1,1,1,1,1,1,0,0,1,2,0)$ that produces a constant ${\rm Li}_4(1/2)$, where ${\rm Li}_k$ is the Polylogarithm \cite{polylog} is discussed. As before a large mass expansion is performed \begin{eqnarray} \frac{1}{p_1^2\, p_2^2\, p_3^2\, p_4^2\, p_5^2\, p_6^2\, (p_9^2-M^2) (p_{10}^2-M^2)^2 }&\rightarrow& \frac{1}{p_1^2\,p_2^2\, p_3^2\, p_4^2} \left[ \frac{1}{p_5^2\, p_6^2\, (p_9^2-M^2) (p_{10}^2-M^2)^2} \right]_{\{p_2,p_4,Q\}} \nonumber\\ &+& \left[ \frac{1}{p_1^2\, p_2^2\, p_3^2\, p_4^2\, p_5^2\, p_6^2\, (p_9^2-M^2) (p_{10}^2-M^2)^2 } \right]_{\{Q\}} \end{eqnarray} where the last term produces 4-loop massive bubble integrals and does not contribute to the imaginary part. After performing the indicated Taylor expansion there remains a product of a massless two loop propagator integral and a massive 2-loop bubble integral that can be evaluated in terms of $\Gamma$-functions using Eqs.(\ref{1loopbubble})--(\ref{bubblem00}). After expanding the $\Gamma$-functions in $\varepsilon$ and reducing the sums over the functions $S_i(k)$ to a basis of independent nested sums one obtains \begin{multline} \Im( I_B(1,1,1,1,1,1,0,0,1,2,0) ) = \frac{\pi }{s} \left( N_\varepsilon \right)^4 \left( \frac{s}{M^2}\right)^{2 \varepsilon} \sum_{k=1}^{\infty} \left(\frac{s}{M^2}\right)^k \Biggl\{ \frac{1}{\varepsilon } \biggl[ -\frac{2}{k^3} + \frac{ (-1)^k \zeta_2}{k} \\ +2 \frac{(-1)^k S_{\tilde{2}}(k)}{k} \biggr] + \biggl[ -\frac{8}{k^3} + \frac{3}{k^4} + \frac{(-1)^k\, \zeta_2\,}{k^2} -6 \frac{S_1(k)}{k^3} +2 \frac{(-1)^k \, S_{\tilde{2}}(k)}{k^2} + \zeta_2\frac{(-1)^k \, S_1(k)}{k} \\ +2 \frac{(-1)^k\, S_{\tilde{2}}(k)\, S_1(k)}{k} +4 \zeta_2 \frac{(-1)^k}{k} +8 \frac{(-1)^k\, S_{\tilde{2}}(k)}{k} - \zeta_3\frac{ (-1)^k}{k} + \frac{9}{2}\frac{(-1)^k\, S_2(k)\, S_{\tilde{1}}(k)}{k}\\ +2 \frac{(-1)^k\, S_{\tilde{2},1}(k)}{k} + \frac{3}{2} \frac{(-1)^k\, S^3_{\tilde{1}}(k)}{k} -9 \frac{(-1)^k\, S_{\tilde{1},\tilde{1},\tilde{1}}(k)}{k} \biggr] + {\cal O}(\varepsilon^2) \Biggr\} \end{multline} On-shell the infinite sum over $k$ can be performed to obtain \begin{multline} \textstyle \Im( I_B(1,1,1,1,1,1,0,0,1,2,0) ) = \frac{ \pi}{s} (N_\varepsilon)^4 \biggl[ \frac{1}{\varepsilon} \Bigl(\frac{5}{4} \zeta_3 - 3 \zeta_2 \ln(2) \Bigr) + \Bigl( 5\zeta_3 - 12 \zeta_2 \ln(2) \\ \textstyle - \frac{49}{8} \zeta_4 + 6{\rm Li_4}(1/2) + \frac{1}{4} \ln^4(2) \Bigr) + {\cal O}(\varepsilon) \biggr] \end{multline} for which all necessary infinite sums are known \cite{sumsjos,infinitesums1,infinitesums2,infinitesums3} \begin{eqnarray} S_{\tilde{1}}(\infty) & = & -\ln(2) \\ S_{\tilde{1},1}(\infty) & = & \textstyle -\frac{1}{2}\zeta_2 +\frac{1}{2}\ln^2(2) \\ S_{\tilde{1},\tilde{2}}(\infty) & = & \textstyle \frac{13}{8}\zeta_3 - \zeta_2 \ln(2) \\ S_{\tilde{2},2}(\infty) & = & \textstyle - 4\, {\rm Li}_4(1/2) - \frac{1}{6}\ln^4 (2) +\zeta_2 \ln^2(2) -\frac{7}{2}\zeta_3 \ln(2) +\frac{51}{16} \zeta_4 \\ S_{\tilde{1},\tilde{3}}(\infty) & = & \textstyle 2\, {\rm Li}_4(1/2) + \frac{1}{12} \ln^4 (2) -\frac{1}{2} \zeta_2 \ln^2(2) + \frac{3}{4} \zeta_3 \ln(2) -\frac{1}{2} \zeta_4 \\ S_{\tilde{1},1,\tilde{2}}(\infty) & = & \textstyle 2 \, {\rm Li}_4(1/2) +\frac{1}{12} \ln^4 (2) + \frac{1}{8} \zeta_3 \ln(2) -\frac{1}{2} \zeta_4 \\ S_{\tilde{1},\tilde{2},1}(\infty) & = & \textstyle 3\, {\rm Li}_4(1/2) +\frac{1}{8} \ln^4 (2) -\frac{3}{4}\zeta_2 \ln^2 (2) +\frac{5}{8} \zeta_3 \ln(2) -\frac{9}{16} \zeta_4 \\ S_{\tilde{1},\tilde{1},\tilde{1},\tilde{1}}(\infty) & = & \textstyle \frac{1}{24} \ln^4(2) +\frac{1}{4} \zeta_2 \ln^2 (2) +\frac{1}{4}\zeta_3 \ln(2) +\frac{9}{16} \zeta_4 \end{eqnarray} \section{The Theoretical Uncertainty} \label{sec:TheoUncert} With the incorporation of the 2-loop QED corrections in $\Delta q$ the largest missing theoretical contributions come from three possible sources. The first is the hadronic uncertainty \cite{muonhad} conservatively estimated to introduce an error of $2\times10^{-8}$ in $\Delta q$ which introduces a relative error of $10^{-8}$ in the extracted value of $G_F$. By examining the logarithms that appear at tree- and 1-loop levels one would expect the leading unknown 2-loop QED corrections to be proportional to\\ $(\alpha/\pi)^2(m_e^2/m_\mu^2)\ln^2(m_e^2/m_\mu^2)=1.4\times10^{-8}$ which would need a coefficient of roughly 140 to introduce a 1\,ppm error in $G_F$. At tree- and 1-loop level the coefficient of the leading logarithm is 12. Allowing for a coefficient as large as 24 gives an an estimate of the theoretical error from this source of $1.7\times10^{-7}$ in the value of $\delta G_F/G_F$. The 3-loop QED corrections may be estimated in the same way as the 2-loop corrections were, before they had been actually calculated, by assuming them to the equal to the known leading logarithm of next order. In this case that translates a relative error of $1.4\times10^{-7}$ in $G_F$. Overall the theoretical uncertainty arising from missing higher-order corrections should not exceed a few parts in $10^{7}$. \section{Experimental Uncertainties} \label{sec:ExpUncert} On the experimental side the accuracy of the known values for the muon lifetime, $\tau_\mu$, the muon mass $m_\mu$ and the muon neutrino mass can each exert a significant effect on the extracted value of $G_F$. From Eqs.(\ref{eq:Gamma0}) and (\ref{eq:Deltaq0}), the change in $G_F$ in response to changes in these quantities is given by \begin{equation} \frac{\delta G_F}{G_F}=-\frac{1}{2}\frac{\delta\tau_\mu}{\tau_\mu} -\frac{5}{2}\frac{\delta m_\mu}{m_\mu} +4\frac{m_{\nu_\mu}^2}{m_\mu^2}. \end{equation} The measured value of the muon lifetime, $\tau_\mu=(2.19703\pm0.00004)\,\mu$s \cite{PDG} is currently the source of the dominant error. Experiments are planned to reduce this error to $\pm4$\,ps Brookhaven National Laboratory \cite{BNL} and $\pm2$\,ps at the Paul Scherrer Institute \cite{PSI}. A new measurement is also expected at the Rutherford-Appleton Laboratory \cite{RAL}. It is therefore likely that uncertainty on $G_F$ coming from this source will be reduced to somewhere in the range 0.5--1\,ppm. The measured value of $m_\mu$ expressed in unified atomic mass units is \cite{PDG} \[ m_\mu=(0.113428913\pm0.000000017)\,{\rm u} \] corresponding to an accuracy of 0.15\,ppm. A new measurement of Planck's constant, $h$, \cite{Planck} means that this can now be converted to units of MeV without introducing additional errors. The current error on $m_\mu$ leads to a 0.38\,ppm uncertainty on the extracted value of $G_F$. Some reduction can be expected in the error on the mass whose effect would then be insignificant. If $m_{\nu_\mu}$ is assumed to be non-zero then setting it to the current upper bound of $m_{\nu_\mu}\le170$\,keV shifts the extracted value of $G_F$ by 10\,ppm. This upper bound is expected to be reduced to some where below 30\,keV by studying the decays of pions in in flight at the Brookhaven muon storage ring \cite{Prisca}. This would affect the extracted value of $\delta G_F/G_F$ at the level of 0.3\,ppm. \section{Weak Corrections to the Muon lifetime} \label{sec:WeakCorr} The weak corrections to the muon lifetime may be encapsulated in a quantity $\Delta r$ defined by the relation \begin{equation} \frac{G_F}{\sqrt{2}}=\frac{g^2}{8 M_W^2}\left(1+\Delta r\right) \label{eq:GFdef} \end{equation} As with Eq.(\ref{eq:alphadef}) the parameters on the left-hand side of Eq.(\ref{eq:GFdef}) represent experimentally-measurable quantities and those on the right-hand side are the renormalized parameters in whatever renormalization scheme has been chosen. Thus $g$ and $M_W^2$ are the renormalized $SU(2)_L$ coupling constant and square of the renormalized $W$ boson mass respectively. $\Delta r$ will be defined through Eq.(\ref{eq:GFdef}) in such a way that that Eq.(\ref{eq:QEDcorr}) is true exactly. In an analogous way to $\Delta q$, $\Delta r$ can be expressed as a power series in $\alpha_r$ \begin{equation} \Delta r=\sum_{i=0}^\infty\Delta r^{(i)} \end{equation} where, once again, the index $i$ indicates the power of $\alpha_r$ that appears in $\Delta r^{(i)}$. Sirlin \cite{Sirlin78,Sirlin80} has described a strategy that, starting from the full electroweak theory, makes the separation of contributions to $\Delta q$ and $\Delta r$ automatic at least up to ${\cal O}(\alpha m_\mu^2/M_W^2)$. In diagrams exhibiting IR divergences, the photon propagator is replaced by \begin{equation} \frac{1}{k^2}\longrightarrow\left\{\frac{1}{k^2}-\frac{1}{k^2 -\Lambda^2}\right\} +\frac{1}{k^2-\Lambda^2}. \label{eq:photonsplit} \end{equation} where it generally convenient to take $\Lambda=M_W$. The term in curly brackets is simply the original photon propagator with a Pauli-Villars regulator. It has the same IR behaviour and gives contributions that are identical to those of Fermi theory up to ${\cal O}(\alpha m_\mu^2/\Lambda^2)$ and thus are contained in $\Delta q$. The second term in (\ref{eq:photonsplit}) gives contributions that retain the original UV behaviour but are free from IR singularities and therefore belong in $\Delta r$. It should be noted that, in contrast to neutral currents, it is not generally the case that charged current processes can be separated into QED and weak parts in a gauge-invariant manner. Sirlin \cite{Sirlin84} has also discussed in general how the strategy may be applied at the 2-loop level. Note that the photons in the second term on the right-hand side of (\ref{eq:photonsplit}) have a mass larger than the muon mass such that in contributions to $\Delta r$ no ``photons" of this type can appear in the final state. This means that one does not need to perform many body phase space integrals for the calculation of $\Delta r$ and in practice the matrix element can be evaluated even at zero external momenta. In contrast the QED corrections to muon decay in the Fermi theory do involve many body phase space integrals, but these corrections contain only one mass scale, ignoring the electron and neutrino masses, and the relatively simple Fermi contact interaction. It is known that \begin{equation} \Delta r^{(0)}=\frac{3m_\mu^2}{10M_W^2} +{\cal O}\left(\frac{m_e^2}{m_\mu^2}\right) \end{equation} that is due to $W$ propagator effects and shifts the extracted value of $G_F$ by 0.52\,ppm. A factor $(1+\Delta r^{(0)})$ is traditionally included in Eq.(\ref{eq:QEDcorr}) that is used to define $G_F$ and up to now this has not been of any significance. It is, however, clearly inappropriate in the definition of $G_F$ as it does not arise from the Fermi theory Lagrangian (\ref{eq:FullLagrangian}) and is most properly included with the weak corrections. It is striking that at the 1\,ppm level experiments performed at low energy are sensitive to the finite range effects of the $W$ boson. \subsection{1-loop Electroweak Corrections to $\Delta r$} The 1-loop corrections to the muon lifetime were first calculated in the Standard Model of electroweak interactions by Sirlin\cite{Sirlin80}. In a general renormalization scheme $\Delta r^{(1)}$ can be written \cite{ZMass4} \begin{eqnarray} \Delta r^{(1)}&=&-\frac{\Pi_{WW}^{(1)}(0)}{M_W^2} +\frac{g^2}{16\pi^2}\left\{ 4\left(\Delta-\ln\frac{M_Z^2}{\mu^2}\right) +6+\left(4+c_\theta^2-\frac{6c_\theta^2}{s_\theta^2}\right) \ln\frac{1}{c_\theta^2}\right\}\nonumber\\ & &\qquad\qquad\qquad\qquad+\frac{g^2}{16\pi^2} \frac{5c_\theta^4-3s_\theta^4}{2s_\theta^2}\ln\frac{1}{c_\theta^2} -\frac{\delta M_W^{2(1)}}{M_W^2}+2\frac{\delta g^{(1)}}{g} \label{eq:Deltar1}\\ &=&-\frac{\Pi_{WW}^{(1)}(0)}{M_W^2} +\frac{g^2}{16\pi^2}\left\{ 4\left(\Delta-\ln\frac{M_Z^2}{\mu^2}\right) +\frac{12s_\theta^2-7}{2s_\theta^2}\ln\frac{1}{c_\theta^2}+6\right\} \nonumber\\ & &\qquad\qquad\qquad\qquad -\frac{\delta M_W^{2(1)}}{M_W^2}+2\frac{\delta g^{(1)}}{g} \end{eqnarray} where all parameters are renormalized parameters in whatever renormalization scheme has been chosen. $\Pi_{WW}^{(1)}(0)$ is the 1-loop $W$ boson self-energy and $s_\theta$ and $c_\theta$ are, respectively the sine and cosine of the weak mixing angle, $\theta_W$, defined as in Ref.\cite{MaldeStuart1} so as to diagonalize the mass matrix of renormalized $Z$ and photon fields. In Eq.(\ref{eq:Deltar1}) the term in curly brackets comes from vertex corrections and the following one from box diagram. Ultraviolet divergences cancel in the overall expression. \subsection{2-loop Electroweak Corrections to $\Delta r$} Certain classes of contributions to $\Delta r$ are simply related to the corresponding ones in $\delta\rho$ because they only enter through the self-energies of the $Z$ and $W$ bosons. In such cases \begin{equation} \Delta r=-\frac{c_\theta^2}{s_\theta^2}\delta\rho. \label{eq:DelrDelrho} \end{equation} Contributions to $\delta\rho^{(2)}$ in the limit of large Higgs mass, $M_H$, were obtained by van~der~Bij and Veltman \cite{BijVelt}. Subsequently van~der~Bij and Hoogeveen \cite{Frank} calculated corrections to $\delta\rho^{(2)}$ arising from a heavy fermion doublet. Their results can be used to obtain the well-known asymptotic expression \begin{equation} \delta\rho^{(2)}=3(2\pi^2-19)x_f^2 \label{eq:Frank} \end{equation} for large top quark mass, $m_t$, and for which \[ x_t=\frac{\alpha}{16\pi s_\theta^2}\frac{m_t^2}{M_W^2}. \] It was this quadratic $m_t$-dependence in $\delta\rho$ that provided some of the strongest constraints on the mass of the top quark before it was directly observed. Thus it was the measurement of $G_F$ was that allowed the top mass to be successfully predicted from precision electroweak data. Consoli, Hollik and Jegerlehner \cite{ConsHollJege} showed how to combine (\ref{eq:Frank}) with 1PR corrections to obtain the asymptotic dependence of $M_W$ on $m_t$. The analogous result for $\Delta r^{(2)}$ as defined via Eqs.(\ref{eq:QEDcorr}) and (\ref{eq:GFdef}) is \begin{equation} \Delta r^{(2)}=\frac{9c_\theta^4}{s_\theta^4}x_f^2 -\frac{3c_\theta^2}{s_\theta^2}(2\pi^2-19)x_f^2. \end{equation} At 2-loop order the contributions that are quartic in $m_t$ are generated via the Yukawa couplings to fermions and therefore come entirely from the scalar sector of the theory. These corrections are given, keeping the full $M_H$ dependence, in Refs.\cite{Barbieri1,Barbieri2,Fleischer1,Fleischer2,DegrFancGamb}. Since it turns out that $m_t$ is roughly the same order as $M_Z$ it is to be expected that the ${\cal O}(\alpha^2 m_t^2 M_Z^2)$ contributions to $\Delta r^{(2)}$ could be similar in size to those that are quartic in $m_t$ and has been borne out by direct calculation \cite{DegrFancGamb}. These subleading corrections do not now just come from $W$ and $Z$ self-energy diagrams alone and so are not simply related to $\delta\rho$. The exact sensitivity of $\Delta r^{(2)}$ to $M_H$, without making an expansion in $m_t$ has been studied in Ref.\cite{BaubWeig}. None of the corrections mentioned above encounters IR problems. All contribute solely to $\Delta r$ and do not require the invocation of (\ref{eq:photonsplit}) in order to separate their QED and weak corrections. Recently the ${\cal O}(N_f\alpha^2)$ corrections to muon decay have been calculated \cite{MaldeStuart2}. These are all 2-loop corrections containing a massless fermion loop and (\ref{eq:photonsplit}) must be applied. In this case, however, the separation is particularly natural. IR divergent contributions come from the Feynman diagrams shown in Figs.~\ref{fig:VertexDiags} and \ref{fig:BoxDiagrams}. IR divergent external leg corrections are shown in Fig. \ref{fig:VertexDiags}(a)--(d) with the `$\times$' in Figs.\ref{fig:VertexDiags}(b) and (d) representing the fermionic contribution to the 1-loop photon 2-point counterterm, $(q^2 g_{\mu\nu}-q_\mu q_\nu)\;2\delta e^{(1f)}/e$. These can be split into QED and weak contributions using (\ref{eq:photonsplit}) in the manner explicitly described in the appendix of Ref.\cite{Sirlin84}. The QED part is already included in Eq.(\ref{eq:Deltaq2}) for $\Delta q^{(2)}$. Remarkably for the weak part, the dependence on the separation mass, $\Lambda$, cancels in any renormalization scheme between the pairs of diagrams Fig.\ref{fig:VertexDiags}(a) with Fig.\ref{fig:VertexDiags}(c) and Fig.\ref{fig:VertexDiags}(b) with Fig.\ref{fig:VertexDiags}(d). The combined weak parts of the Feynman diagrams contains a simple pole at $D=4$ that eventually cancels with other divergences from the weak sector \cite{MaldeStuart2}. The box diagrams, Fig.\ref{fig:BoxDiagrams}, both vanish identically due to a conspiracy in the $\gamma$-matrix algebra of the fermion currents. The general result is proportional to products of left-handed couplings with right-handed couplings; the latter being zero for the $W$ boson. In a general renormalization scheme the ${\cal O}(N_f\alpha^2)$ box diagrams containing a virtual photon and a counterterm insertion on the $W$ propagator produce contributions proportional to the IR divergent 1-loop box diagram, \begin{equation} \begin{picture}(40,60)(0,0) \ArrowLine(0,0)(5,15) \ArrowLine(5,15)(5,45) \ArrowLine(5,45)(0,60) \ArrowLine(40,60)(35,45) \ArrowLine(35,45)(35,15) \Vertex(35,15){1} \ArrowLine(35,15)(40,30) \Photon(5,15)(35,15){-2}{5.5} \Text(21.5,41)[t]{${\scriptstyle W}$} \Photon(5,45)(35,45){2}{5.5} \Text(21.5,11)[t]{${\scriptstyle \gamma}$} \SetWidth{1} \Line(1.5,41.5)(8.5,48.5) \Line(1.5,48.5)(8.5,41.5) \end{picture} \quad\raisebox{30pt}{+}\quad \begin{picture}(40,60)(0,0) \ArrowLine(0,0)(5,15) \Vertex(5,15){1} \ArrowLine(5,15)(5,45) \Vertex(5,45){1} \ArrowLine(5,45)(0,60) \ArrowLine(40,60)(35,45) \Vertex(35,45){1} \ArrowLine(35,45)(35,15) \Vertex(35,15){1} \ArrowLine(35,15)(40,30) \Photon(5,15)(35,15){-2}{5.5} \Text(11.5,41)[t]{${\scriptstyle W}$} \Text(28.5,41)[t]{${\scriptstyle W}$} \Photon(5,45)(35,45){2}{5.5} \Text(21.5,11)[t]{${\scriptstyle \gamma}$} \SetWidth{1} \Line(16.5,41.5)(23.5,48.5) \Line(16.5,48.5)(23.5,41.5) \end{picture} \quad\raisebox{30pt}{+}\quad \begin{picture}(40,60)(0,0) \ArrowLine(0,0)(5,15) \Vertex(5,15){1} \ArrowLine(5,15)(5,45) \Vertex(5,45){1} \ArrowLine(5,45)(0,60) \ArrowLine(40,60)(35,45) \Vertex(35,45){1} \ArrowLine(35,45)(35,15) \Vertex(35,15){1} \ArrowLine(35,15)(40,30) \Photon(5,15)(35,15){-2}{5.5} \Text(21.5,41)[t]{${\scriptstyle W}$} \Photon(5,45)(35,45){2}{5.5} \Text(21.5,11)[t]{${\scriptstyle \gamma}$} \SetWidth{1} \Line(31.5,41.5)(38.5,48.5) \Line(31.5,48.5)(38.5,41.5) \end{picture} \quad\raisebox{30pt}% {$=\left(2\frac{{\displaystyle\delta g^{(1f)}}}{{\displaystyle g}} -\frac{{\displaystyle\delta M_W^{2(1f)}}}{{\displaystyle M_W^2}} \right)$}\quad \begin{picture}(40,60)(0,0) \ArrowLine(0,0)(5,15) \Vertex(5,15){1} \ArrowLine(5,15)(5,45) \Vertex(5,45){1} \ArrowLine(5,45)(0,60) \ArrowLine(40,60)(35,45) \Vertex(35,45){1} \ArrowLine(35,45)(35,15) \Vertex(35,15){1} \ArrowLine(35,15)(40,30) \Photon(5,15)(35,15){-2}{5.5} \Text(21.5,41)[t]{${\scriptstyle W}$} \Photon(5,45)(35,45){2}{5.5} \Text(21.5,11)[t]{${\scriptstyle \gamma}$} \end{picture} \label{eq:boxCTs} \end{equation} and diagrams would need to be treated by applying (\ref{eq:photonsplit}). However in the $\overline{{\rm MS}}$ renormalization scheme adopted here the sum of the counterterms that appear on the right hand side of Eq.(\ref{eq:boxCTs}) vanishes. The superscript ${}^{(1f)}$ denotes the 1-loop contribution to the counterterm from a light fermion species \cite{MaldeStuart1,MaldeStuart2}. \begin{figure} \begin{center} \begin{picture}(75,120)(0,0) \ArrowLine(0,0)(5.25,15) \Vertex(5.25,15){1} \Text(6,0)[bl]{$\mu^-$} \ArrowLine(5.25,15)(15.75,45) \Vertex(15.75,45){1} \ArrowLine(15.75,45)(21,60) \Vertex(21,60){1} \ArrowLine(21,60)(0,120) \Text(15,90)[l]{$\nu_\mu$} \PhotonArc(10.5,30)(15.89,178.60,250.71){2}{2.5} \Text(0,13)[r]{$\gamma$} \ArrowArc(-4.39,35.21)(5,-100.35,61.77) \Vertex(-5.29,30.29){1} \ArrowArc(-4.39,35.21)(5,61.77,-100.35) \Vertex(-2.02,39.62){1} \PhotonArc(10.5,30)(15.89,70.71,142.83){2}{2.5} \Text(13,53)[r]{$\gamma$} \Photon(21,60)(53,60){3}{4} \Text(37,66)[b]{$W$} \ArrowLine(56,120)(53,60) \Vertex(53,60){1} \Text(52,90)[r]{$\bar\nu_e$} \ArrowLine(53,60)(74,120) \Text(67,90)[l]{$e^-$} \Text(-10,65)[bl]{1} \Text(-10,60)[bl]{--} \Text(-10,53)[bl]{2} \Text(37,-3)[b]{(a)} \end{picture} \hfill \begin{picture}(75,120)(0,0) \ArrowLine(0,0)(5.25,15) \Vertex(5.25,15){1} \Text(6,0)[bl]{$\mu^-$} \ArrowLine(5.25,15)(15.75,45) \Vertex(15.75,45){1} \ArrowLine(15.75,45)(21,60) \Vertex(21,60){1} \ArrowLine(21,60)(0,120) \Text(15,90)[l]{$\nu_\mu$} \PhotonArc(10.5,30)(15.89,70.71,250.71){2}{5.5} \Text(0,13)[r]{$\gamma$} \Text(13,53)[r]{$\gamma$} \Photon(21,60)(53,60){3}{4} \Text(37,66)[b]{$W$} \ArrowLine(55,120)(53,60) \Vertex(53,60){1} \Text(52,90)[r]{$\bar\nu_e$} \ArrowLine(53,60)(74,120) \Text(67,90)[l]{$e^-$} \SetWidth{1} \Line(-6.56,39.72)(-2.22,30.71) \Line(-8.90,33.04)(0.11,37.38) \Text(-10,65)[bl]{1} \Text(-10,60)[bl]{--} \Text(-10,53)[bl]{2} \Text(37,-3)[b]{(b)} \end{picture} \hfill \begin{picture}(75,120)(0,0) \ArrowLine(0,0)(21,60) \Vertex(21,60){1} \ArrowLine(21,60)(0,120) \Text(15,90)[l]{$\nu_\mu$} \Photon(21,60)(53,60){3}{4} \Text(37,66)[b]{$W$} \ArrowLine(56,120)(53,60) \Vertex(53,60){1} \Text(52,90)[r]{$\bar\nu_e$} \ArrowLine(53,60)(58.25,75) \Vertex(58.25,75){1} \Text(15,30)[l]{$\mu^-$} \ArrowLine(58.25,75)(68.75,105) \Vertex(68.75,105){1} \ArrowLine(68.75,105)(74,120) \PhotonArc(63.5,90)(15.89,250.71,322.82){2}{2.5} \Text(77,97)[bl]{$\gamma$} \ArrowArc(78.39,84.79)(5,-100.35,61.77) \Vertex(77.49,79.87){1} \ArrowArc(78.39,84.79)(5,61.77,-100.35) \Vertex(80.76,89.20){1} \PhotonArc(63.5,90)(15.89,-1.4,70.71){2}{2.5} \Text(64,72)[tl]{$\gamma$} \Text(74,113)[l]{$e^-$} \Text(-3,65)[bl]{1} \Text(-3,60)[bl]{--} \Text(-3,53)[bl]{2} \Text(37,-3)[b]{(c)} \end{picture} \hfill \begin{picture}(75,120)(0,0) \ArrowLine(0,0)(21,60) \Vertex(21,60){1} \ArrowLine(21,60)(0,120) \Text(15,90)[l]{$\nu_\mu$} \Text(78,97)[bl]{$\gamma$} \PhotonArc(63.5,90)(15.89,250.71,70.71){2}{5.5} \Text(64,72)[tl]{$\gamma$} \Photon(21,60)(53,60){3}{4} \Text(37,66)[b]{$W$} \ArrowLine(55,120)(53,60) \Vertex(53,60){1} \Text(52,90)[r]{$\bar\nu_e$} \ArrowLine(53,60)(58.25,75) \Vertex(58.25,75){1} \Text(15,30)[l]{$\mu^-$} \ArrowLine(58.25,75)(68.75,105) \Vertex(68.75,105){1} \ArrowLine(68.75,105)(74,120) \Text(74,113)[l]{$e^-$} \SetWidth{1} \Line(76.22,89.30)(80.56,80.29) \Line(73.88,82.62)(82.89,86.96) \Text(-3,65)[bl]{1} \Text(-3,60)[bl]{--} \Text(-3,53)[bl]{2} \Text(37,-3)[b]{(d)} \end{picture} \caption{Infrared divergent external leg corrections contributing to muon decay at ${\cal O}(N_f\alpha^2)$. \label{fig:VertexDiags} } \end{center} \end{figure} \begin{figure} \begin{center} \null\hfill \begin{picture}(80,120)(0,0) \ArrowLine(0,0)(10,30) \Vertex(10,30){1} \ArrowLine(10,30)(10,90) \Vertex(10,90){1} \ArrowLine(10,90)(0,120) \Text(7,15)[l]{$\mu^-$} \Text(8,104)[l]{$\nu_\mu$} \ArrowLine(80,120)(70,90) \Vertex(70,90){1} \ArrowLine(70,90)(70,30) \Vertex(70,30){1} \ArrowLine(70,30)(80,60) \Text(78,106)[l]{$\bar\nu_e$} \Text(78,46)[l]{$e^-$} \Photon(10,30)(33,30){-4}{2} \Vertex(33,30){1} \Photon(47,30)(70,30){4}{2} \Vertex(47,30){1} \ArrowArc(40,30)(7,180,0) \ArrowArc(40,30)(7,0,180) \Text(43,82)[t]{$W$} \Photon(10,90)(70,90){4}{5.5} \Text(21.5,37)[b]{$\gamma$} \Text(58.5,37)[b]{$\gamma$} \Text(40,0)[b]{(a)} \end{picture} \hfill \begin{picture}(80,120)(0,0) \ArrowLine(0,0)(10,30) \Vertex(10,30){1} \ArrowLine(10,30)(10,90) \Vertex(10,90){1} \ArrowLine(10,90)(0,120) \Text(7,15)[l]{$\mu^-$} \Text(8,104)[l]{$\nu_\mu$} \ArrowLine(80,120)(70,90) \Vertex(70,90){1} \ArrowLine(70,90)(70,30) \Vertex(70,30){1} \ArrowLine(70,30)(80,60) \Text(78,106)[l]{$\bar\nu_e$} \Text(78,46)[l]{$e^-$} \Photon(10,30)(70,30){-4}{5.5} \Text(21.5,37)[b]{$\gamma$} \Text(58.5,37)[b]{$\gamma$} \Text(43,82)[t]{$W$} \Photon(10,90)(70,90){4}{5.5} \SetWidth{1} \Line(35,35)(45,25) \Line(35,25)(45,35) \Text(40,0)[b]{(b)} \end{picture} \hfill\null \caption{The infrared divergent ${\cal O}(N_f\alpha^2)$ box diagrams occurring in muon decay at ${\cal O}(N_f\alpha^2)$. \label{fig:BoxDiagrams} } \end{center} \end{figure} \subsection{$G_F$ in the Analysis of Electroweak Data} There is an important proviso that must be remembered when using (\ref{eq:photonsplit}) to split up the radiative corrections. The resulting sets of QED and weak corrections must be calculated in a consistent renormalization scheme. In the present paper, the $\overline{{\rm MS}}$ scheme with $\mu=m_\mu$ was used for the QED corrections but in other circumstances this need not be the appropriate choice. In the analysis of electroweak data the numerical values of the renormalized parameters of the theory are initially obtained from a set of simultaneous equations for the physical observables as calculated from the full electroweak Lagrangian. To 1-loop \cite{ZMass4,Z0Pole} in the Standard Model \begin{eqnarray} \sqrt{4\pi\alpha}&=&e\left( 1+\frac{1}{2}\Pi_{\gamma\gamma}^{(1)\prime}(0) +\frac{s_\theta}{c_\theta}\frac{\Pi_{Z\gamma}^{(1)}(0)}{M_Z^2} +s_\theta^2\frac{\delta g^{(1)}}{g} +c_\theta^2\frac{\delta g^{\prime(1)}}{g^\prime}\right), \label{eq:alphaexp}\\ \tau_\mu^{-1}&=&\frac{m_\mu^5}{192\pi^3} \left\{\frac{\sqrt{2}g^2}{8M_W^2} (1+\Delta r^{(0)}+\Delta r^{(1)})\right\}^2 (1+\Delta q^{(0)}+\Delta q^{(1)}), \label{eq:taumuexp}\\ s_p&=&M_Z^2+\delta M_Z^2+\Pi_{ZZ}^{(1)}(M_Z^2). \label{eq:spexp} \end{eqnarray} where $\Pi_{Z\gamma}^{(1)}(0)$ is the transverse part of the 1-loop $Z$-$\gamma$ mixing and $\delta g^{\prime(1)}$ is the 1-loop $U(1)$ coupling constant counterterm. $s_p$ is the position of the pole on the complex plane associated with the unstable $Z^0$ boson. A form similar to Eq.(\ref{eq:alphaexp}) can be found in Ref.\cite{Sirlin84}, Eq.(11a). The quantities on the left hand side of Eqs.(\ref{eq:alphaexp})--(\ref{eq:spexp}) are the physical observables whose values are obtained by experiment and those on the right hand side are the renormalized parameters and counterterms in the particular renormalization scheme that has been chosen which, for the purposes of this discussion, will be assumed to be the $\overline{{\rm MS}}$ scheme. The subscript ${}_r$ has been dropped for the renormalized parameters. For the analysis of electroweak data obtained near the $Z^0$ resonance it is convenient and natural to take for the 't~Hooft mass, $\mu=M_Z$. Upon solving the equations Eqs.(\ref{eq:alphaexp})--(\ref{eq:spexp}) one finds $\alpha_r\sim1/128$ and thus the running of $\alpha$ is automatically incorporated in a self-consistent manner at tree-level which avoids the need to resum large logarithms as is required when the on-shell renormalization scheme is used. It also has the consequence that the quantity in curly brackets in Eq.(\ref{eq:taumuexp}) must also be evaluated at $\mu=M_Z$. This can be implemented by substituting $\alpha_r=\alpha_e(M_Z^2)$ in the formulas (\ref{eq:Deltaq1}) and (\ref{eq:Deltaq2}) for $\Delta q^{(1)}$ and $\Delta q^{(2)}$. The net result is that the quantity in curly brackets effectively defines a running Fermi coupling constant, $G_F(\mu)$ for which $G_F(M_Z)=1.16639\times10^{-5}\,{\rm GeV}^{-2}$ very close to quoted value obtained for $\mu=m_\mu$ but including only 1-loop QED corrections. \section*{Acknowledgments} RGS wishes to thank the Max-Planck-Institute f\"ur Physik, Munich, for hospitality while part of this work was carried out. Helpful and informative discussions with K. Melnikov and A. Sirlin are gratefully acknowledged. This work was supported in part by the US Department of Energy. The work of TR was also supported in part by BMBF under contract No. 057KA92P and DFG Forschergruppe under contract KU 502/8-1. \newpage
\section{Introduction} \label{sec:intro} The central and complex role of galactic gas in the star formation, dynamical, and chemical evolution of galaxies is well established. Evidence is mounting that, at the present epoch, multiphase gaseous halos are a physical extension of their host galaxy's interstellar medium (ISM); their physical extent, spatial distribution, ionization conditions, and chemical enrichment are intimately linked to the energy density rate infused into the galaxy's ISM by stellar winds and ionizing radiation, and by supernovae shock waves (e.g.\ \cite{dahlem98}, and references therein). One long--standing question is how the halos and ISM of earlier epoch galaxies compare or relate, in an evolutionary sense, to those of the present epoch. Normal ($\sim L^{\ast}$) galaxies at intermediate redshifts ($0.5 \leq z \leq 1.0$) are seen to give rise to low ionization {{\rm Mg}\kern 0.1em{\sc ii}~$\lambda\lambda 2796, 2803$} absorption with $W_{r}(2796) \geq 0.3$~{\AA} out to projected distances of $\sim 40h^{-1}$~kpc (e.g.\ \cite{steidel95}). A key question is whether low ionization gas at large galactocentric distances is due to infall (i.e.\ satellite accretion, minor mergers, intragroup or intergalactic infall), or to energetic processes in the ISM (galactic fountains, chimneys). Using high resolution {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} profiles, Churchill, Steidel, \& Vogt (1996\nocite{csv96}) found no suggestive trends between low ionization gas properties and galaxy properties at $\left< z \right> = 0.7$. A next logical step toward addressing this question is to explore the high ionization gas in {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} absorption selected galaxies at these redshifts. The {{\rm C}\kern 0.1em{\sc iv}~$\lambda\lambda 1548, 1550$} doublet is a sensitive probe of higher ionization gas. Using {\hbox{{\rm C}\kern 0.1em{\sc iv}}} and {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}}, Bergeron et~al.\ (1994\nocite{bergeron94}) inferred multiphase ionization structure around a $z=0.79$ galaxy. Churchill \& Charlton (1999\nocite{q1206}) incorporated the {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} kinematics and found high metallicity, multiphase absorption in a possible group of three galaxies at $z=0.93$. In a survey of the 3C~336 field (Q~$1622+238$), Steidel et~al.\ (1997\nocite{3c336paper}) reported that $W_{r}(1548)/W_{r}(2796)$ appeared to be correlated with galaxy--QSO impact parameter, as expected if halo gas density decreases with galactocentric distance. In this {\it Letter}, we present a study of {\hbox{{\rm C}\kern 0.1em{\sc iv}}} absorption associated with $0.4 \leq z \leq 1.4$ {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} absorption--selected galaxies using Faint Object Spectrograph (FOS) data available in the {\it Hubble Space Telescope\/} ({\it HST}) Archive. We compare the {\hbox{{\rm C}\kern 0.1em{\sc iv}}} strengths to the {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} strengths and kinematics and (when available) to the galaxy properties. \begin{figure*}[t] \plotfiddle{f1.eps}{4.7in}{0.}{78.}{78.}{-306}{-80} \vglue -0.1in \protect\caption {\footnotesize The normalized spectroscopic data with the zero point given by the horizontal dotted line. The {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} velocity scale is $500$~{\hbox{km~s$^{-1}$}} and the {\hbox{{\rm C}\kern 0.1em{\sc iv}}} velocity scale is $3000$~{\hbox{km~s$^{-1}$}}. --- (upper sub--panels) The HIRES {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} $\lambda 2796$ transition is plotted in order of kinematic spread, from upper left to lower right. --- (lower sub--panels) The corresponding FOS/{\it HST\/} spectrum is plotted with ticks marking the expected location of the {{\rm C}\kern 0.1em{\sc iv}~$\lambda\lambda 1548, 1550$} doublet, based upon Voigt profile fits to {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}}. The labels ``D'', ``L'', and ``Bl'' denote detection, limit, and blend, respectively. \label{fig:fig1} \vglue -0.2in } \end{figure*} \section{The Data} \label{sec:data} The {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} absorbers are selected from the HIRES/Keck sample of Churchill (1997\nocite{thesis}) and Churchill et~al.\ (1999a\nocite{weak}). The HIRES resolution is $\sim 6$~{\hbox{km~s$^{-1}$}} (\cite{vogt94}). The data were processed using IRAF\footnote{IRAF is distributed by the National Optical Astronomy Observat-- \\ ories, which are operated by AURA, Inc., under contract to the NSF.}, as described in Churchill et~al.\ (1999a\nocite{weak}). The redshifts of the individual {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} sub--components were obtained using {\sc minfit} (\cite{thesis}), a Voigt profile (VP) fitter that uses $\chi ^{2}$ minimization. For $36$ of the {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} absorbers, FOS/{\it HST\/} (resolution $\sim 230$~{\hbox{km~s$^{-1}$}}) spectra covering {\hbox{{\rm C}\kern 0.1em{\sc iv}}} were available from the {\it HST\/} Archive. For four absorbers, {\hbox{{\rm C}\kern 0.1em{\sc iv}}} was taken from ground--based spectra of Steidel \& Sargent (1992\nocite{ss92}), and Sargent, Boksenberg, \& Steidel (1988\nocite{sbs88}). The FOS spectra were processed using the techniques of the {\it HST\/} QSO Absorption Line Key Project (\cite{dpsKP}; \cite{KP13}). For 13 systems, the absorbing galaxy impact parameters, rest--frame $K$ and $B$ luminosities, and $B-K$ colors are available from Steidel, Dickinson, \& Persson (1994\nocite{sdp94}), Churchill et~al.\ (1996\nocite{csv96}), and Steidel et~al.\ (1997\nocite{3c336paper}). We will present a more detailed account in a companion paper (\cite{archive}). \section{Results} \label{sec:results} In Figure~\ref{fig:fig1}, we present the {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} and {\hbox{{\rm C}\kern 0.1em{\sc iv}}} data for each of the 40 systems (note that the velocity scale for {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} is 500~{\hbox{km~s$^{-1}$}} and for {\hbox{{\rm C}\kern 0.1em{\sc iv}}} is 3000~{\hbox{km~s$^{-1}$}}). Ticks above the HIRES spectra give the velocities of the multiple VP {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} sub--components and ticks above the FOS data give the expected location of these components for both members of the {\hbox{{\rm C}\kern 0.1em{\sc iv}}} doublet. The {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} profiles are shown in order of increasing kinematic spread from the upper left to lower right. The kinematic spread is the second velocity moment of the apparent optical depth of the {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} $\lambda 2796$ profile, given by $ \omega _{v}^{2} = \int \tau _{a}(v) v^{2}dv / \int \tau _{a}(v)dv$, where $\tau _{a}(v) = \ln [I_{c}(v)/I(v)]$, and $I(v)$ and $I_{c}(v)$ are the observed flux and the fitted continuum flux at velocity $v$, respectively. The zero--point velocity is given by the optical depth median of the {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} $\lambda 2796$ profile. The kinematic spreads range from a few to $\sim 120$~{\hbox{km~s$^{-1}$}}. \begin{figure*}[b] \plotfiddle{f2.eps}{0.in}{0.}{65.}{65.}{264}{60} \parbox{3.5in}{\phantom{dummy}} \hfill \parbox{3.5in}{\protect\caption{\vglue -1.8in \footnotesize {\sc Fig}.~~2. --- (a) $W_{r}(2796)$ vs.\ $W_{r}(1548)$. --- (b) $W_{r}(2796)$ vs.\ the {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} kinematic spread, $\omega _{v}$. --- (c) $W_{r}(1548)$ vs.\ $\omega _{v}$. Errors are on the order of the data point sizes. Filled points are DLAs and candidate DLAs. The dotted and dashed curves (panel c) are from linear fits, excluding the upper limits (see text). \label{fig:fig2}}} \end{figure*} In Figure~\ref{fig:fig2}$a$, we present $W_{r}(1548)$ vs.\ $W_{r}(2796)$, which exhibits considerable scatter. Solid data points are damped {\hbox{{\rm Ly}\kern 0.1em$\alpha$}} absorbers (DLAs) and candidate DLAs, based upon $W_{r}({\hbox{{\rm Ly}\kern 0.1em$\alpha$}}) \geq 8$~{\AA} or $W_{r}({\hbox{{\rm Mg}\kern 0.1em{\sc ii}}}~\lambda 2796) \simeq W_{r}({\hbox{{\rm Fe}\kern 0.1em{\sc ii}}}~\lambda 2600) \ga 1.0$~{\AA} (\cite{boisse}). \begin{figure*}[t] \vglue -0.1in \plotfiddle{f3a.eps}{1.9in}{0.}{75.}{75.}{-289}{-69} \plotfiddle{f3b.eps}{0.0in}{0.}{75.}{75.}{-26}{-44} \protect\caption{ \footnotesize (a) $W_{r}(1548)/W_{r}(2796)$ vs.\ galaxy impact parameter, $D$. --- (b) $W_{r}(1548)/W_{r}(2796)$ vs.\ galaxy rest--frame color, $B-K$. Errors are on the order of the data point sizes. \label{fig:fig3}} \vglue -0.2in \end{figure*} As seen in Figure~\ref{fig:fig2}$b$, $W_{r}(2796)$ correlates with $\omega _{v}$. Significant scatter arises because $\omega _{v}$ is sensitive to the line--of--sight chance presence and equivalent width distribution of the smaller $W_{r} (2796)$, ``outlying velocity'' clouds (see \cite{kinematicpaper}). The DLAs define a ``saturation line''; profiles with $W_{r}(2796) > 0.3$~{\AA} along this line have saturated cores. As seen in Figure~\ref{fig:fig2}$c$, there is a tight correlation between $\omega _{v}$ and $W_{r}(1548)$. A Spearman--Kendall test, incorporating limits (\cite{eric}), yielded a greater than 99.99\% confidence level. A weighted least--squares fit to the data (dotted line through the origin and with upper limits excluded) yielded a slope of $\omega _{v} \simeq 65$~{\hbox{km~s$^{-1}$}} per $1$~{\AA} of $W_{r}(1548)$. The data exhibit a scatter of $\sigma _{W_{r}}(1548) = 0.22$~{\AA} about the fit. An essentially identical maximum likelihood fit is shown as a dashed line. Over the interval $35 \leq \omega _{v} \leq 55$~{\hbox{km~s$^{-1}$}}, there are four absorbers that lack the higher ionization phase typical for their {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} kinematic spread; they are ``{\hbox{{\rm C}\kern 0.1em{\sc iv}}} deficient''. These systems, Q~0117+213 at $z=1.0479$, Q~1317+277 at $z=0.6606$, Q~0117+213 at $z=0.7290$, and Q~1329+274 at $z=0.8936$ (in order of decreasing $\omega _{v}$), lie $3.3$, $2.9$ $2.5$, and $2.3\sigma$ from the correlation line, respectively. As compared to other DLAs with similar $\omega _{v}$, the DLA at $z=0.6561$ in the field of Q~$1622+238$ (3C~336) also appears to have a slight {\hbox{{\rm C}\kern 0.1em{\sc iv}}} deficiency. \section{Discussion} Three important observational facts are: (1) the scatter in $W_{r}(1548)$ vs.\ $W_{r}(2796)$ indicates that the strength of {\hbox{{\rm C}\kern 0.1em{\sc iv}}} absorption is not driven by the strong ``central'' {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} component that dominates $W_{r}(2796)$, (2) $\omega_{v}$ is sensitive to the presence of small $W_{r}(2796)$, outlying velocity clouds, and (3) the scatter of $W_{r}(2796)$ vs.\ $\omega_{v}$ is large, whereas the scatter of $W_{r}(1548)$ vs.\ $\omega_{v}$ about the correlation line is only $0.22$~{\AA}. These facts imply that, independent of the overall {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} line--of--sight kinematics, the existence and global dynamics of smaller, kinematic ``outliers'' are intimately linked to the presence and physical conditions of a higher ionization phase. It would appear that {\hbox{{\rm C}\kern 0.1em{\sc iv}}} is governed by the same physical processes that give rise to kinematic outlying {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} clouds. The {\hbox{{\rm C}\kern 0.1em{\sc iv}}} absorption could arise due to the ionization balance in the {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} clouds, ionization structure in the clouds, or due to multiphase structure. In most cases, multiphase structure is the likely explanation because the small $W_{r}(2796)$, single phase, outlying velocity clouds with $b\simeq 6$~{\hbox{km~s$^{-1}$}} cannot produce the large observed $W_{r}(1548)$, as shown by Churchill et~al.\ (1999a\nocite{weak}) [also see Churchill \& Charlton (1999\nocite{q1206})]. If the {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} clouds are accreted by infall and/or minor mergers, such that the energetics originated gravitationally (e.g.\ \cite{mo96}), multiphase structure could arise from shock heating and/or merger induced star formation (e.g.\ \cite{hernquist95}). If, on the other hand, the absorbing gas is mechanically produced by winds from massive stars, OB associations, or from galactic fountains and chimneys, a dynamic multiphase structure could arise from shock heated ascending material that forms a high ionization layer (corona) and supports a descending lower ionization layer, which then breaks into cool, infalling clouds [see Avillez (1999\nocite{avillez}), and references therein]. Both scenarios imply a link between galaxy star formation histories, in particular multiple episodes of elevated star formation [but not necessarily bursting (\cite{dahlem98})], and the presence of outlying velocity {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} clouds and strong {\hbox{{\rm C}\kern 0.1em{\sc iv}}}. Infall models predict increasing cloud densities with decreasing galactocentric distance (e.g.\ \cite{mo96}), resulting in an ionization gradient (clouds further out are more highly ionized). In Figure~\ref{fig:fig3}$a$, we show $W_{r}(1548)/W_{r}(2796)$ vs.\ impact parameter for 13 absorbing galaxies. There is no \pagebreak obvious evidence for an ionization gradient (95\% confidence). However, $W_{r}(1548)/W_{r}(2796)$ could be sensitive to halo mass (e.g. \cite{mo96}), to the presence of satellite galaxies (\cite{york86}), or to the sampling of discrete clouds over a range of galactocentric distances along the line of sight. In Figure~\ref{fig:fig3}$b$, we show $W_{r}(1548)/W_{r}(2796)$ vs.\ galaxy $B-K$ color for 11 galaxies. There is a suggested trend ($2\sigma$) for red galaxies (those dominated by late--type stellar populations) to have small $W_{r}(1548)/W_{r}(2796)$. If such a trend is confirmed in a larger data sample, it would not be incompatable with a dynamical multiphase scenario in which absorbing gas properties are linked to the host galaxy stellar populations, and therefore, star formation history. The tight correlation between {\hbox{{\rm Mg}\kern 0.1em{\sc ii}}} kinematics and {\hbox{{\rm C}\kern 0.1em{\sc iv}}} absorption may imply a self--regulating process involving both ionization conditions and kinematics in the halos of higher redshift, $\sim L^{\ast}$ galaxies (e.g. \cite{lb92}), as explored by Norman \& Ikeuchi (1989\nocite{norman}) and Li \& Ikeuchi (1992\nocite{li}). Perhaps outflow energetics from supernovae during periods of elevated star formation are balanced by the galactic gravitational potential well, resulting in a fairly narrow range of kinematic and multiphase ionization conditions. Such a balance might set up a high ionization Galactic--like ``corona'' (\cite{savage97}) in proportion to the kinematics of gravitationally bound, cooling material. All this would imply that galaxy ``coronae'' have been in place since $z \sim 1$, that their nature primarily depends upon the host galaxy's star formation history, and therefore morphology, environment, and stellar populations (as seen locally, e.g.\ \cite{dahlem98}). Detailed observations of the stellar content and galactic morphologies and space--based UV high--resolution spectroscopy of high ionization absorption lines, would be central to establishing the interactive cycles between stars and gas in higher redshift galaxies. \vglue 0.1in Support for this work was provided by the NSF (AST--9617185), and NASA (NAG 5--6399 and AR--07983.01--96A) the latter from the Space Telescope Science Institute, which is operated by ARUA, Inc., under NASA contract NAS5--26555 BTJ acknowledges support from NOAO, which is operated by ARUA, Inc., under cooperative agreement with the NSF.